GLIAL CELLS RESPONSES: IN OPIOID WITHDRAWAL SYNDROME
HTML Full TextGLIAL CELLS RESPONSES: IN OPIOID WITHDRAWAL SYNDROME
Gurpreet Bawa
Department of Pharmacology, Rayat and Bahra Institute of Phamacy, Sahauran, Kharar, District Mohali-140 104, Punjab, India
ABSTRACT: Drug addiction represents one of the major medical, social, and economic burdens of human behavior. Opioids are powerful relievers; use of opioids for the treatment of pain has been associated with the potential disadvantages including development of tolerance, dependence. There are seven stages involved in opioid addiction. Identification of glial-mediated mechanism inducing opioid side effects includes cytokine receptor, κ-opioid receptor, NMDA receptor and toll like receptor (TLR). Glial activation through TLR leading to the release of proinflammatory cytokines acting on neurons which is important in the complex syndrome of opioid dependence and withdrawal. Moreover, newer agents targeting these glial cell activation such as AV411, AV33, SLC022 and older agents for other diseases conditions such as minocyclline, pentoxifylline, all show varied but promising profiles for providing significant relief from opioid side effects.
Keywords: |
Toll like receptors, Glial cells, Glutamate receptors, Minocyclline, AV411
INTRODUCTION: Opioids are standard drugs used to manage severe pain and are the most commonly used psychoactive substances across the world 1, 2, 3. It is kind of chronic relapsing brain diseases characterized by the loss of control over intake 4, 5, 6. Acute morphine along with other opioid withdrawal proceeds through a number of stages. There are seven stages which are involved in withdrawal syndrome.
Various receptors are involved in opioid withdrawal like mu, kappa, delta, glutamate and toll like receptors (TLR). There are various types of TLR in which some are play important role in opioid withdrawal.
Recent studies reveal that TLRs, including TLR2 AND TLR4, a key link between the innate immune system and CNS 7, 8.
Glial cells are important for structural and metabolic maintenance of the nervous system, there are numerous reports demonstrating the ability of glial to respond to and send signals to neurons and synapses in the central and peripheral nervous system (CNS and PNS) 9.
Opioid induced proinflammatory glial activation has been inferred from:
a) Morphine induced upregulation of mircoglial 10, 11 and astrocytes 11, 12.
b) Morphine induced upregulation and/ or released proinflammatory cytokines 11, 13, 14, 15, 16, 17, 18.
c) Enhanced morphine analgesia by co-administering the microglial attenuators minocycline 10, 13, 14, 15 or AV411 11 and the astrocytes inhibitor fluorocitrate 12.
d) Enhanced morphine analgesia by blocking proinflammatory cytokine action 13, 14, 19.
e) Opioid induced selective activation of microglial p38 MAPK and associated enhanced morphine analgesia 10.
As such, opioid induced proinlammatory glial activation through TLR is characterized by a cellular phenotype of enhanced reactivity and propensity to proinflammation in response to exposure of glial to opioids 13, 14, 15.
The research on glial cells, glial cells supply nerve fibres with energy rich metabolic products and glial cells assist in the repair of injured nerves.
Stages involved in Opioid withdrawal:
- Stage I: Six to fourteen hours after last dose: Drug craving, anxiety, irritability, perspiration, and mild to moderate dysphoria.
- Stage II: Fourteen to eighteen hours after last dose: Yawning, heavy perspiration, mild depression lacrimation, crying, running nose, dysphoria, also intensification of the above symptoms. "yen sleep" (a waking trance-like state)
- Stage III: Sixteen to twenty-four hours after last dose: Rhinorrhea (runny nose) and increase in other of the above dilated pupils, piloerection (gooseflesh), muscle twitches, hot flashes, cold flashes, aching bones and muscles, loss of appetite and the beginning of intestinal cramping.
- Stage IV: Twenty-four to thirty-six hours after last dose: Increase in all of the above including severe cramping and involuntary leg movements ("kicking the habit"), loose stool, insomnia, elevation of blood pressure, moderate elevation in body temperature, increase in frequency of breathing and tidal volume, tachycardia (elevated pulse), restlessness, nausea.
- Stage V: Thirty-six to seventy-two hours after last dose: Increase in the above, fetal position, vomiting, free and frequent liquid diarrhea, which sometimes can accelerate the time of passage of food from mouth to out of system to an hour or less, involuntary ejaculation, which is often painful, saturation of bedding materials with bodily fluids, weight loss of two to five kilos per 24 hours, increased white cell count and other blood changes.
- Stage VI: After completion of above: Recovery of appetite ("the chucks"), and normal bowel function, beginning of transition to post-acute and chronic symptoms that are mainly psychological but that may also include increased sensitivity to pain, hypertension, colitis or other gastrointestinal afflictions related to motility, and problems with weight control in either direction (Chan et al., 1999) 20.
Receptors involved in Opioid withdrawal syndrome: Opioid receptors are a group of G proteins coupled receptors with opioids as ligands. Opiate receptors are distributed widely in the brain and are found in spinal cord and digestive tract 21, 22, 23.
The four major subgroups of opiate receptors are delta (δ), kappa (κ), mu (µ) and nociception and each is involved in controlling different function in the brain24. For example: opiates and endorphins block pain signal by binding to the µ receptor site. The δ receptor in the brain is involved in pain relief, antidepressant effects and physical dependence.
The κ receptor in the brain and spinal cord are linked with sedation, spinal analgesia and pupil constriction. The function of the µ receptor in the brain and spinal cord are physical dependence, respiratory depression, euphoria, pupil constriction and supraspinal analgesia.
Nociception receptors in the brain and spinal cord are involved with appetite, depression, anxiety and the development of tolerance to µ agonists (Table 1).
TABLE 1: RECEPTORS INVOLVED IN OPIOID WITHDRAWAL SYNDROME
Receptors | Subtypes | Location25,26 | Function 25, 26 |
Delta ( δ)
DOP OP1 |
Δ1 δ2 | Brain
Pontine nuclei Amygdale Olfactory bulbs Deep cortex Peripheral sensory neurons |
Analgesia
Antidepressant effects Physical dependence |
Kappa (κ)
KOP OP2 |
κ1 κ2 κ3 | Brain
Hypothalamus Periqueductal gray Claustrum Substantia gelatinosa Peripheral sensory neurons |
Analgesia
Sedation Miosis Inhibition of ADH release Dysphoria |
Mu (µ)
MOP OP3 |
µ1 µ2 µ3 | brain
cortex thalamus striosomes periqueductal gray intestinal tract |
µ1:- analgesia
physical dependence µ2:- respiratory depression miosis euphoria reduced GI motility physical dependence µ3:- possible vasodilation |
Involvement of Glutamate receptors in Opioid withdrawal Syndrome:
- Types of glutamate receptors: Glutamate is one of the most abundant excitatory neurotransmitters in the central nervous system. Once released into the synaptic cleft, glutamate can bind to its receptors and exert its effect. According to pharmacological and molecular biological classification, glutamate receptors can be divided into two categories, ionotropic glutamate receptors (iGluRs) and metabotropic glutamate receptors (mGluRs).
- The role of glutamate and its receptor in opioid addiction: It is known that certain glutamatergic projection could be impacted by addictive drugs. Data indicate that the activation of glutamatergic efferent fibers from the amygdala and prefrontal cortex is critical in the expression of addictive behaviors 27, 28, 29, 30, 31, 32. The alteration of glutamate-mediated transmission, especially the increase of glutamatergic transmission in the NAc may promote the seeking and relapse from abused drugs 33, 34, 35, 36. Microinjection of glutamate into the VTA can increase exploratory motor behaviors and the release of DA in NAc and mPFC. These results suggest that glutamate plays an important role in opioids dependence.
- The role of iGluRs in opioid addiction 37. NMDA receptors in rat brain following long-term treatment with morphine. Then it was shown that the expression of NMDA receptors was upregulated in morphine-dependent rat brains 38. Recently, more studies showed that glutamatergic signal transduction can regulate drug effects, resulting in drug tolerance and dependence, including the tolerance and dependence of opioids. Glutamatergic afferents play a key role in regulating the firing of the VTA neurons. Activation of glutamatergic afferents and the VTA infusion of glutamate receptor agonists increase the firing rates of dopaminergic neurons and induce burst firing in vivo 39, 40, 41.
It was also demonstrated that MK801, an NMDA receptor antagonist, completely blocked the withdrawal symptoms induced by glutamate and naloxone 42, 43. MK-801 can also decrease morphine dependence, which may be related to the downregulation of NMDA receptors 37, 38. Western blot and microdialysis results showed that chronic intermittent use of morphine, cocaine and other addictive drugs can increase the level of glutamate in the VTA, upregulating the expression of AMPA receptor subtypes GluR1 and NMDAR1 in the VTA 44, 45.
Co-application of opioids and NMDA receptor competitive or non-competitive antagonists can block the pain tolerance and physical dependence of opioid and the drug-seeking behaviors 46.
Glutamatergic signal transduction mediated by NMDA receptors was involved in the formation and maintenance of morphine dependence in human 47. So that glutamate released in the central nervous system plays an important role in opioid withdrawal behaviors and that the iGluRs are involved in the process.
- The role of mGluRs in opioid addiction: Metabotropic glutamate receptors, which mediate slow glutamate neurotransmission, are located throughout the limbic and cortical brain regions implicated in drug addiction. There is significant pharmacological and behavioral evidence that group I mGluRs are widely distributed in the projection neurons and intermediate neurons in the shell and core of NAc, providing the morphological evidence for their regulation and therapic effects in reward-related behaviors and drug addiction 48, 49.
3-[(2- methyl-1, 3-thiazol-4-yl)ethynyl] pyridine (MTEP), an mGluR5 antagonist, dose-dependently inhibited morphine withdrawal symptoms induced by naloxone 50. DA and glutamate play critical roles in the induction of LTP in the NAc through the activation of D1 dopamine receptors and group I mGluRs 51. It was reported that the number and function of group II mGluRs upregulated the formation of opioids withdrawal.
Group II mGluRs (mGluR2 and mGluR3) were involved in the negative regulation in the brain reward circuit and the formation of conditional offensive responses in drug dependence and withdrawal. However, the role of group III mGluRs in drug dependence is still not well investigated. Further investigation is required to fully understand the role of mGluRs in the pathological process of drug addiction.
Toll like Receptors: TLRs, a family of evolutionarily conserved pathogen recognition receptors, play a pivotal role in innate immunity. TLR family consists of 13 mammalian members. The cytoplasmic portions of TLRs show high similarity to that of the interleukin-1 receptor (IL-1R) family and are now called the Toll/IL-1 receptor (TIR) domain. A TIR domain is required to initiate intracellular signaling. The extracellular regions of TLRs and IL-1R are markedly different. Whereas IL-1R possesses an Ig-like domain, TLRs contain leucine-rich repeats in their extracellular domains.
TLRs are pattern recognition receptors that sense a wide range of microorganisms, such as bacteria, fungi, protozoa, and viruses. Each TLR has its own intrinsic signaling pathway and induces specific biological responses against microorganisms such as dendritic cell maturation, cytokine production, and the development of adaptive immunity 52.
Glial Cells: Glial cells, also called neuroglia or simply glia are non-neuronal cells that maintain homeostasis, form myelin and provide support and protection for neurons in the brain. Glial cells comprise 90% of the human brain 53.
Types of Glial Cells: Glia are divided into two subtypes, the microglia which function largely as scavengers to engulf apoptotic cell debris and the macroglia comprised of oligodendrocytes that myelinate axons, and astrocytes 54.
- Microglia: Microglia’s are specialized macrophages capable of phagocytosis that protect neurons of the central nervous system 55. They are derived from hematopoietic precursors rather than ectodermal tissue; they are commonly categorized as such because of their supportive role to neurons. These cells comprise approximately 15% of the total cells of the central nervous system. They are found in all regions of the brain and spinal cord. Microglial cells are small relative to macroglial cells, with changing shapes and oblong nuclei. They are mobile within the brain and multiply when the brain is damaged. In the healthy central nervous system, microglia processes constantly sample all aspects of their environment (neurons, macroglia and blood vessels) (Table 2).
TABLE 2: MACROGLIA
Location | Name | Description |
CNS | Astrocytes | The most abundant type of macroglial cell, astrocytes (also called astroglia) has numerous projections that anchor neurons to their blood supply. They regulate external chemical environment of neurons by removing excess ions, notably potassium, and recycling neurotransmitters released during synaptic transmission. The current theory suggests that astrocytes may be the predominant "building blocks" of the blood–brain barrier. Astrocytes may regulate vasoconstriction and vasodilation by producing substances such as arachidonic acid, whose metabolites are vasoactive.
Astrocytes signal each other using calcium. The gap junctions (also known as electrical synapses) between astrocytes allow the messenger moleculeIP3 to diffuse from one astrocyte to another. IP3 activates calcium channels on cellular organelles, releasing calcium into the cytoplasm. This calcium may stimulate the production of more IP3. The net effect is a calcium wave that propagates from cell to cell. Extracellular release of ATP, and consequent activation of purinergic receptors on other astrocytes, may also mediate calcium waves in some cases. In general, there are two types of astrocytes, protoplasmic and fibrous, similar in function but distinct in morphology and distribution. Protoplasmic astrocytes have short, thick, highly branched processes and are typically found in gray matter. Fibrous astrocytes have long, thin, less branched processes and are more commonly found in white matter. It has recently been shown that astrocyte activity is linked to blood flow in the brain, and that this is what is actually being measured in fMRI 56. They also have been involved in neuronal circuits playing an inhibitory role after sensing changes in extracellular calcium 57. |
CNS | Oligodendrocytes | Oligodendrocytes are cells that coat axons in the central nervous system (CNS) with their cell membrane forming a specialized membrane differentiation called myelin, producing the so-called myelin sheath. The myelin sheath provides insulation to the axon that allows electrical signals to propagate more efficiently 58. |
CNS | Ependymal cells | Ependymal cells, also named ependymocytes, line the cavities of the CNS and make up the walls of the ventricles. These cells create and secrete cerebrospinal fluid (CSF) and beat their cilia to help circulate the CSF and make up the Blood-CSF barrier. They are also thought to act as neural stem cells 59. |
CNS | Radial glia | Radial glia cells arise from neuroepithelial cells after the onset of neurogenesis. Their differentiation abilities are more restricted than those of neuroepithelial cells. In the developing nervous system, radial glia functions both as neuronal progenitors and as a scaffold upon which newborn neurons migrate. In the mature brain, the cerebellum and retina retain characteristic radial glial cells. In the cerebellum, these are Bergmann glia, which regulate synaptic plasticity. In the retina, the radial Müller cell is the principal glial cell, and participates in a bidirectional communication with neurons 60. |
PNS | Schwann cells | Similar in function to oligodendrocytes, Schwann cells provide myelination to axons in the peripheral nervous system (PNS). They also havephagocytotic activity and clear cellular debris that allows for regrowth of PNS neurons 61. |
PNS | Satellite cells | Satellite glial cells are small cells that surround neurons in sensory, sympathetic and parasympathetic ganglia 62. These cells help regulate the external chemical environment. Like astrocytes, they are interconnected by gap junctions and respond to ATP by elevating intracellular concentration of calcium ions. They are highly sensitive to injury and inflammation, and appear to contribute to pathological states, such as chronic pain 63. |
PNS | Enteric glial cells | Are found in the intrinsic ganglia of the digestive system. They are thought to have many roles in the enteric system, some related to homeostasis and muscular digestive processes 64. |
Glial in the Central Nervous System: In the central nervous system (CNS), consisting of the brain and spinal cord, the major glial types are astrocytes and oligodendrocytes. The astrocytes, which are more numerous, have many radiating processes that interweave in complex and intimate ways between neuronal cell bodies and fibres.
Some astrocyte processes contact blood vessels and may control the blood-brain barrier which protects the CNS from unwanted substances in the general circulation. Others form cuffs or veils around individual synapses, and synaptic transmission can be modified by signals between nerve terminals and these glial elements.
They also have high affinity uptake sites for major brain neurotransmitters that help to remove excess transmitter following release from nerve terminals. Together this provides compelling evidence that gial cells are directly involved in information processing in the brain. Astrocytes also help to control the levels of potassium in the extracellular space and have major roles in CNS development.
Oligodendrocytes form one of the most highly specialized cellular structures in the body, the myelin sheath, which forms electrical insulation around nerve fibres thereby making rapid transmission of electrical signals in the brain possible. The CNS also contains microglia, resident, macrophage-like cells that originate from blood monocytes rather than the neurectoderm.
Glial in the Peripheral Nervous System: In the peripheral nervous system (PNS), the major glial cells are Schwann cells. They ensheath all axons in peripheral nerves and are found in two types, myelinating and non-myelinating. The myelinating Schwann cells form insulating sheaths around axons that are comparable in structure and function to those made by oligodendrocytes in the CNS. The non-myelinating cells show similarities with astrocytes and are likely to have metabolic and mechanical support functions. There is evidence that Schwann cells are indispensable for neuronal survival during development, and in damaged nerves Schwann cells control successful regeneration and restoration of function.
Olfactory ensheathing cells represent a special category of glia that resembles non-myelinating Schwann cells and associate with both the CNS and PNS part of the primary olfactory axons. Another important category of PNS glia is the enteric glia. They are found in the autonomic ganglia of the gut (the enteric nervous system). Unlike other parts of the PNS, the enteric system has complex synaptic interactions and high integrative capacity, and the enteric glia is remarkably like astrocytes in structure and biochemistry. The cell bodies of other autonomic ganglia and sensory ganglia are enveloped by simpler satellite glial cells, while the synapses between nerve terminals and skeletal muscle are covered by terminal glia, also called teloglia or perisynaptic glia. They help to maintain a stability of the neuromuscular junction and regulate synaptic transmission 65.
Role of Microglia in Opioid Withdrawal: Microglial cells like other cells of immune system lineage display functional opioid receptors 66, 67. Overstimulation of these receptors can lead to apoptosis of microglial cells 68. Methamphetamine administration to rats at doses that induce dependence causes activation of microglial cells in the striatum. Since the activation of microglia follows a course similar to the neurotoxicity caused by METH, it has been suggested that METH neurotoxicity might be at least in part mediated by the METH-activated microglia 69, 70. Also attenuation of microglial activation mediates tolerance to the neurotoxicity of METH in the striatum71. Microglia activity might also be involved in the mediation of the behavioral effects of METH because, in rats, minocycline (an anti-inflammatory known to affect microglia) treatment not only reduced damage to dopaminergic terminals but also significantly attenuated behavioral sensitization caused by repeated administration of METH 72.
Role of Astrocytes in Opioid Withdrawal: Astrocytes are the most abundant glial cell type in the central nervous system73. Like other glial cell types, astrocytes were once thought to play only a secondary, non-regulatory and permissive role in nervous function. The cell membranes of astrocytes bear receptors for most neurotransmitters and peptides: glutamate, dopamine, norepinephrine, serotonin, gammaaminobutyric acid, acetylcholine and opioid peptides 74. They also bear in their plasma membranes neurotransmitter 75, 76, 77 and glucose 78 transporters, and aquaporin-4 channels for water transport 79, 80.
Administration of cocaine, amphetamines and most psychostimulants induces activation of astrocytes81, 82, 83. This activation is defined by an increase in the expression of glial fibrillary acidic protein (GFAP), a main component of the cytoskeleton of astrocytes. GFAP is known to be upregulated in response to brain injury and neurotoxicity 84, 85, although changes in GFAP expression are not limited to overt brain injury and many other plastic changes in the neuropil also result in increased GFAP expression 86, 87, 88.For instance, treatment with ethamphetamine (METH, “speed”) results in loss of dopaminergic terminals without detectable loss of neurons 89, but induces astrogliosis with increased GFAP in the striatum, hippocampus and frontal cortex 90.
Chronic treatment with morphine also results in increased GFAP expression or enlarged astrocytes in VTA, NAcc, frontal cortex, locus coeruleus and nucleus of the solitary tract of the rat 91, 92, 93, 94, 95. The morphine-induced increases in GFAP expression and the astroglial activation are probably mediated by a2-adrenoceptors since the antagonist yohimbine inhibits upregulation of a2-adrenoceptors and prevents the increase in GFAP expression caused by chronic morphine treatment 94, 95. The responsiveness of astrocytes of morphine administration 96, 97 that astrocytes might contribute to morphine tolerance and more recent evidence supports that hypothesis 98.
For instance, inactivation of astrocytes by the gliotoxin fluorocitrate attenuates both tolerance to morphine analgesia and morphine-induced increase in GFAP immunostaining 93. Tolerance to morphine has been also related to downregulation of glial glutamate transporters GLT-1 and GLAST in the spinal cord 99 suggesting a link between structural and functional features of astrocytes involved in tolerance to morphine.
The neurotrophic activity of astrocytes may also be relevant to the effects of cocaine in VTA. Glial-derived neurotrophic factor (GDNF) is present in neurons astrocytes and microglial cells, although it seems to be mainly produced by astrocytes 100, 101. GDNF supports the survival and differentiation of dopaminergic cells and protects those cells against METH-induced neurotoxicity as shown in wild-type mice 102 and in heterozygous mice with a partial deletion of the GDNF gene 103. Glial-derived neurotrophic factor, when infused into the VTA reduces the increase in the formation of key proteins induced by cocaine exposure 104. Furthermore mice lacking expression of GDNF display increase behavioral sensitization to cocaine 105 and treatment of mice with the dipeptide Leu-Ileu an inducer of GDNF (and TNF-alpha, also produced by astrocytes) expression blocked the acquisition of METH-induced place preference and sensitization 106, 107.
Tumor necrosis factor alpha, produced by astrocytes and microglial cells, has been shown recently to prevent METH neurotoxicity and dependence in mice possibly through the enhancement of dopamine uptake in the striatum and the prevention of METH-induced increases in extracellular dopamine 108.
Expression of TLRS in CNS Glial cells: Microglia are CNS tissue resident macrophages and act as immune sentinels of the brain. In accordance with this view, primary microglia in vitro constitutively expresses a wide complement of TLRs (TLRs1-9) at varying levels 109, 110. In comparison, primary astrocytes also express a wide variety of TLRs, but at lower levels. Murine astrocytes express TLRs1-9, with particularly high levels of TLR3 111, 112, suggesting that astrocytes may be particularly important for anti-viral responses in the CNS. To date, human astrocytes have been reported to express TLRs1-5 and TLR9, also with particularly high expression of TLR3 109, 113, 114.
The lack of TLR6-8 may be a difference between species or the result of varying isolation and culture conditions. There is also evidence that both oligodendrocytes and neurons can express TLRs, but their role in innate immune responses during CNS 109, 115, 116. Under resting conditions in vivo. Constitutive expression of TLRs is primarily in microglia and largely restricted to the circumventricular organs (CVOs) and meninges, areas with direct access to the circulation, although they may be expressed at lower levels in other regions as well 117, 118, 119.
The levels of TLRs in the CNS can be upregulated by viral and bacterial infection, treatment with TLR stimuli, or CNS autoimmunity 109, 112, 119, 120, providing a mechanism for amplification of inflammatory responses to pathogens infecting the CNS. These stimuli upregulate multiple TLRs in a coordinated fashion, not only the TLR involved in recognition of a particular pathogen or class of pathogens. For example, treatment of astrocytes with the dsRNA synthetic mimic polyinosinic-polycytidylic acid (poly I: C), a viral stimulus, upregulates its own receptor TLR3, and also upregulates TLR2 and TLR4, which are normally used to recognize bacterial product 111.
Similarly, infection of mice with rabies broadly increases CNS expression of TLRs1-4 and 6–9112.The pattern of TLR upregulation is not fixed and varies with the particular pathogen encountered, even with pathogens of a similar class. For example, in contrast to rabies virus, Semliki forest virus infection in the CNS fails to upregulate TLR4 and TLR6, but does increase expression of TLR13 112.
The upregulation of TLRs in the CNS is likely in part due to the infiltration of TLR-expressing inflammatory cells, and in part due to the upregulation of receptor expression on astrocytes and microglia, which occurs in response to a variety of inflammatory stimuli 110, 111.
TLR Signalling mediators Glial cells activation in the CNS: Innate immunity in the cns depends primarily on the function of glial cells, especially in microglia which are important for the activation of adaptive system 121.
- TLR in Microglia: TLR mediated signalling promotes the production of a variety of inflammatory mediators 122, 123. Exogenous and Endogenous TLR ligands activate microglial cells. TLR may mediate different pathway in microglial leading to either neuroprotective or neurotoxic phenotype1 24.However, activated microglia with TLR ligands also produce neurotoxic molecules such as nitric oxide (NO), reactive oxygen species (ROS), peroxynitriate, proinflammatory cytokines (TNF-α, IL-1β) which leads to withdrawal syndrome 125.
TLR appear to activate very similar signaling pathways to IL-1 and some researchers now refer to this pathway as the TLR-IL signaling pathway126. That is, TLRs work through activation of an adaptor protein known as myeloid differentiation factor 88 (MyD88). This factor leads to activation of the IL-1 receptor- associated kinases (IRAKs) and TNF receptor – associated factor-6 (TRAF6), which finally culminates in activation of NF-Κb 127. Other TLR- associated pathways include the JNK and interferon (IFN) pathways. Both TLR2 and TLR4 are important in recognizing endogenous pain-mediating signals. These studies have shown a highly interconnected web of pathway involving TLRs and other well–defined proinflammatory pathways which are associated with glial activation and opioid side effects 128, 129.
Relation between Glial cells and Proinflammatory cytokines: Glial cells of the CNS such as astrocytes, oligodendrocytes and microglia can respond to and also produce many of the proinflammatory cytokines initially attributed to lymphocytes and macrophages. These cytokines include IL-1, IL-2, IL-3, IL-4, IL-6, IL-8, IL-10, IFN-α, IFN-β, IFN-γ, TNF-α, TNF-β, TGF-β and colony sti ulating factor (CSFs) 130. Opioid withdrawal induces glial activation and proinflammatory cytokines (TNF-α) expression in different sites of the brain.
Medications with potential for use as Opioid adjuncts:
- Ibudilast (AV411) is a blood-brain permeable, nonspecific phosphodiesterase inhibitor that acts centrally by attenuation of glial cell activation and reduction of proinflammatory activating factors, such as cytokines (TNF-α, IL-1β), nitric oxide, and chemokines such as monocyte schemo-attractant protein-1 and fractalkines; it increases production of anti-inflammatory IL-10131,132. It was initially used to treat bronchial asthma and poststroke dizziness, which has been attributed to its ability to reduce inflammation and cause vasodilatation 133.
AV411 also reduces mechanical allodynia caused by neuropathic pain as well as noxious neuropathy induced by chemotherapeutic agents (paclitaxel and vincristine); it also reduces morphine tolerance. When administered systemically, AV411 was shown to be distributed to the spinal cord and to attenuate morphine induced glial cell activation in certain brain regions. AV411 has also been shown to inhibit peripheral inflammatory cells 134, 135, which have been suggested as a possible cause of reduced pain perception peripherally.
AV411 also reduces spontaneous opioid withdrawal, protected naloxone-induced morphine withdrawal when given during the period of development of morphine dependence, and simultaneously enhanced analgesic effects 136.These effects were seen with both morphine and oxycodone, with no changes in plasma morphine levels. Another study shows that AV411 decreases a morphine-induced increase of dopamine in the nucleus accumbens, a nucleus known to be associated with morphine-induced drug reward as well as withdrawal 137, 138.
- Dizocilpine (MK-801) is an NMDA-positive glutamate receptor noncompetitive antagonist that attenuates opioid tolerance and does not influence the antinociceptive effects of morphine. When injected intrathecally, it decreases morphine tolerance at the spinal level139
- Propentofylline (SLC022) is an orally available, blood-brain permeable, methylxanthine derivative that acts as a glial inhibitor which attenuates neuropathic pain states as well as chemotherapy-induced painful neuropathy 140.It decreases allodynia, possibly through altering γ-aminobutyric acid (GABA)ergic tone through modulation of glutamic acid decarboxylase in the spinal cord after injury, as well as reducing an injury-induced increased expression of GFAP 141. Intraperitoneal injections of propentofylline attenuated condition-placed preference, a measure of drug reward in animals that were dependent on methamphetamine and morphine; this attenuation is thought to be caused by astrocytic activation 142. Propentofylline also act as a neuroprotective agent in ischemia models 143.
- AV333 is a plasmid that has been shown to be a well-tolerated and effective antineuropathic agent when injected intrathecally. It functions as a glial cell inhibitor and promotes an increase in the amount of the anti-inflammatory cytokine IL-10 in the spinal cord
- Minocycline is a semisynthetic, second-generation broad spectrum, blood-brain barrier permeable tetracycline that has been historically used for its antimicrobial properties. However, it possess neuroprotective effects with reported benefits in experimental models of neurodegenerative disease, traumatic brain injury, and cerebral ischemia. Minocycline’s protective role occurs by suppression of the mitochondrial permeability transition, inhibition of caspace-1 and -3 expressions and inhibition of microglial activation and proliferation 144 via antihyperalgesic and antiallodynic effects 145, 146.
Minocycline inhibits the activation of microglial cells, which are thought to initiate neuropathic pain, thus preventing development of neuropathic pain in animal models. However, once these cells are activated, minocycline does not seem to be as effective in reducing pain states146. Although minocycline enhances the analgesic efficacy of opioids, it may also increase undesirable effects of opioids such as respiratory depression and drug dependence.
Minocycline is a p-glycoprotein (p-gp) inhibitor, and inhibition of p-gp can cause altered pharmacokinetics of opioids, thus leading to increased bioavailability and ultimately an increase in adverse effects.
- Pentoxifylline is an inhibitor of glial activation, nonspecific cytokine synthesis, and phospho-diesterase (PDE) 147. Pentoxifylline inhibit the production of mRNA and protein levels of proinflammatory cytokines such as TNF-α, IL-1β, and IL-6, which were associated with reduced neuropathic pain 148 and inflammatory pain 149. Along with reduction of these cytokines through inhibition of NK-κB, attenuation of pain symptoms has also been shown to be associated with elevated levels of the anti-inflammatory cytokine, IL-10, in the CNS 150.
Similar to AV411’s inhibition of PDE, pentoxifylline reduces cyclic adenosine monophosphate (cAMP) levels, and this in turn results in decreased TNF-α and IL1-β production by microglia. These cytokines cause upregulation of nitric oxide synthase, which increases nitric oxide (NO) levels. NO is known to affect dopamine levels in the mesolimbic system, where increased dopamine levels are associated with opioid reward. Thus, these PDE inhibitors can potentially modulate drug reward and dependence 151.
Another suggested mechanism of attenuating morphine reward relates to a reduced production of adenosine by limited cAMP hydrolization. Adenosine causes inhibition of the inhibitory GABA pathways, which modulate pathways in the ventral tegmental area (VTA) of the mesolimbic system. The VTA contains cells that project to the nucleus accumbens and are a source of dopamine. Therefore, with reduced adenosine levels by PDE inhibitors, there may be attenuation of morphine reward through decreased dopamine in the nucleus accumbens and activation of cells in the VTA as well as glial cells151. In conclusion, pentoxifylline attenuates neuropathic pain states and may also contribute to reduction of morphine-induced tolerance and reward
Research on Glial Cells:
- Glial cells supply nerve fibres with energy-rich metabolic products. Glial cells pass on metabolites to neurons: Around 100 billion neurons in the human brain enable us to think, feel and act. They transmit electrical impulses to remote parts of the brain and body via long nerve fibres known as axons. This communication requires enormous amounts of energy, which the neurons are thought to generate from sugar. Axons are closely associated with glial cells which, on the one hand, surround them with an electrically insulating myelin sheath and, on the other hand support their long-term function. Klaus Armin and his research group from the Max Planck Institute of Experimental Medicine in Göttingen have now discovered a possible mechanism by which these glial cells in the brain can support their associated axons and keep them alive in the long term.
- Glial cells assist in the repair of injured nerves: When a nerve is damaged, glial cells produce the protein neuregulin1 and thereby promote the regeneration of nerve tissue. Unlike the brain and spinal cord, the peripheral nervous system has an astonishing capacity for regeneration following injury. Researchers discovered that, following nerve damage, peripheral glial cells produce the growth factor neuregulin1, which makes an important contribution to the regeneration of damaged nerves.
CONCLUSION: Opioid addiction, a significant and social problem is complicated by the phenomena of tolerance and dependence. There are various receptors which are involved in opioid withdrawal like glutamate receptors, κ-opioid, and toll like receptor (TLR). Glial activation via TLR increases the proinflammatory cytokines (TNF-α) causes the opioid withdrawal syndrome.
Hence, suppression of glial cells proinflammatory cytokines through TLR can significantly reduce opioid withdrawal syndrome. There are number of exciting directions for the use of glial-modifying agents as opioid adjuncts for the treatment of withdrawal syndrome. It appears that basic and clinical research involving both previously discovered agents such as Minocycline, pentoxifylline as well as newer agents such as AV411 and SLCO22.
REFERENCES:
- Akil H, Lewis JW. Neurotransmitters and pain control, Vol 9. Pain and headache. Karger, Basel 1987; 129–159.
- Van Ree JM, Gerrits MAFM, Vanderschuren LJMJ (1999) Opioids, reward and addiction: an encounter of biology, psychology, and medicine. Pharmacol Rev 51:341–396.
- Hardman JG, Limbird LE, Gilman AG. Goodman and Gilman’s. The Pharmacological basis of therapeutics. McGraw Hill, New York. 2001.
- American Psychiatric Association. Diagnostic and Statistical Manual of Mental disorders, 4th ed. American Psychiatric Press, Washington, DC. 1994.
- Koob G.F, Nestler E.J; “The neurobiology of drug addiction”. J. Neuropsychiatry Clin. Neurosci 1997, 9: 482–497.
- Gass J.T, Olive M.F; “Glutamatergic substrates of drug addiction and alcoholism.Biochem”. Pharmacol 2008, 75: 218–265.
- Kielian T; “Toll-like receptors in central nervous system glial inflammation and homeostasis”. J Neurosci Res 2006, 83:711–730.
- Kim D, Kim M.A, Cho I.H et al; “A critical role of toll-like receptor 2 in nerve injury-induced spinal cord glial cell activation and pain hypersensitivity”. J Biol Chem 2007, 282:14975–14983.
- Keith J. Todd 1, Alexandre Serrano et al. Glial cells in synaptic plasticity. Journal of Physiology - Paris 99 (2006) 75–83.
- Cui Y, Liao X.X, Liu W, et al; “A novel role of minocycline: attenuating morphine antinociceptive tolerance by inhibition of p38 MAPK in the activated spinal microglia”. Brain Behav. Immun 2008, 22: 114–123.
- Hutchinson M.R, Lewis S.S, et al; “Reduction of opioid withdrawal and potentiation of acute opioid analgesia by systemic AV411 (ibudilast)”. Brain Behav. Immun 2009. 23, 240–250.
- Song P, Zhao Z.Q, 2001; “The involvement of glial cells in the development of morphine tolerance”. Neurosci. Res. 39, 281–286.
- Hutchinson M.R, Coats B.D, et al. “Proinflammatory cytokines oppose opioid-induced acute and chronic analgesia”. Brain Behav. Immun 2008a, 22: 1178–1189.
- Hutchinson M.R, Northcutt, A.L, et al. “Minocycline suppresses morphine-induced respiratory depression, suppresses morphine-induced reward, and enhances systemic morphine induced analgesia”. Brain Behav. Immun 2008b, 22: 1248–1256.
- Hutchinson M.R, Zhang, Y, et al. “Nonstereoselective reversal of neuropathic pain by naloxone and naltrexone: involvement of toll-like receptor 4 (TLR4)”. Eur. J. Neurosci 2008c, 28: 20–29.
- Johnston I.N, Milligan E.D, et al. “A role for proinflammatory cytokines and fractalkine in analgesia, tolerance, and subsequent pain facilitation induced by chronic intrathecal morphine”. J. Neurosci 2004, 24, 7353–7365.
- Raghavendra V, Rutkowski M.D, DeLeo J.A, et al. “The role of spinal neuroimmune activation in morphine tolerance/hyperalgesia in neuropathic and shamoperate rats”. J. Neurosci 2002, 22, 9980–9989.
- Raghavendra V, Tanga F.Y, DeLeo J.A; “Attenuation of morphine tolerance, withdrawal-induced hyperalgesia, and associated spinal inflammatory immune responses by propentofylline in rats”. Neuropsychopharmacology 2004, 29, 327–334.
- Shavit Y, Wolf G, Goshen I, Livshits D, Yirmiya R; “Interleukin-1 antagonizes morphine analgesia and underlies morphine tolerance” . Pain 2005 115, 50–59.
- Chan R, Irvine R, White J. "Cardiovascular changes during morphine administration and spontaneous withdrawal in the rat". Eur. J. Pharmacol. 1999368 (1): 25–33.
- Dhawan B.N, Cesselin F, et al; "International Union of Pharmacology. XII. Classification of opioid receptors". Pharmacol. Rev 1996. 48: 567–92.
- Janecka A, Fichna J, Janecki T; "Opioid receptors and their ligands". Curr Top Med Chem 4: 1–17.
- Waldhoer M, Bartlett SE, Whistler JL (2004). "Opioid receptors". Annu. Rev. Biochem.73: 953–90.
- Corbett A.D, Henderson G, McKnight A.T, Paterson S.J. Br. J. Pharmacol 2006. 147: 153- 162.
- Stein C, Schäfer M, Machelska H; “Attacking pain at its source: new perspectives on opioids”. Nature Med 2003, 9:1003-1008.
- Fine Perry G, Russell K, Portenoy; "Chapter 2: The Endogenous Opioid System". A Clinical Guide to Opioid Analgesia 2004. McGraw Hill.
- Pierce R.C, Kalivas P.W; “A circuitry model of the expression of behavioral sensitization to amphetamine-like psychostimulants”. Brain Res. Rev 1997, 25:192–216.
- Goldstein R.A, Volkow N.D; “Drug addiction and its underlying neurobiological basis: neuroimaging evidence for the involvement of the frontal cortex.Am”. J. Psychiatry 2002, 159: 1642–1652.
- McLaughlin J, See R.E; “Selective inactivation of the dorsomedial prefrontal cortex and the basolateral amygdala attenuates conditioned-cued reinstatement of extinguished cocaine-seeking behavior in rats”. Psychopharmacology (Berl) 2003, 168: 57–65.
- See R.., Fuchs R.A, Ledford C.C, McLaughlin J; “Drug addiction, relapse and the amygdale”. Ann. N.Y. Acad. Sci 2003, 985: 294–307.
- Kalivas P.W; “Glutamate system in cocaine addiction”.Curr. Opin. Pharmacol 2004, 4: 23-29.
- Kalivas P.W, LaLumiere R.T, Knackstedt L, Shen H; “Glutamate transmission in addiction”. Neuropharmacolgy 2009, 56: 169–173.
- Zhu H, Rockhold R.W, Ho I.K; “The role of glutamate in physical dependence on opioids”. Jpn. J. Pharmacol 1998, 76: 1–14.
- Watanabe T, Nakagawa T, Yamamoto R, Maeda A, Minami M, Satoh M; “Involvement of glutamate receptors within the central nucleus of the amygdale in naloxone-precipitated morphine withdrawal-induced conditioned place aversion in rats.” Jpn. J. Pharmacol 2002, 88: 399–406.
- Kenny P.J, Markou, A; “The ups and downs of addiction: role of metabotropic glutamate receptors”. Trends Pharmacol. Sci 2004, 25: 265–272.
- Paul J.K, Athina M; “The ups and downs of addiction: role of metabotropic glutamate receptors”. Trends Pharmacol. Sci 2004, 25: 265–272.
- Gudehithlu K.P, Reddy P.L, Bhargava H.N; “Effect of morphine tolerance and abstinence on the binding of [3H] MK-801 to brain regions and spinal cord of the rat”. Brain Res 1994. 639, 269–274.
- Koyuncuoglu, H, Nurten A, Yamanturk, P, Nurten, R; “The importance of the number of NMDA receptors in the development of supersensitivity or tolerance to and dependence on morphine”. Pharmacol. Res 1999, 39: 311–319
- Grace A.A, Bunney B.S; “The control of firing pattern in nigral dopamine neurons: single spike firing”. J. Neurosci.1984, 4: 2866–2876.
- Chergui K, Charlety P.J, Akaoka H, Saunier C.F, Brunet J.L, Buda M, Svensson T.H., Chouvet G; “Tonic activation of NMDA receptors causes spontaneous burst discharge of rat midbrain dopamine neurons in vivo”. Eur. J. Neurosci.1993, 5: 137–144.
- Murase S, Grenhoff J, Chouvet G, Gonon F.G, Svensson T.H; “Prefrontal cortex regulates burst firing and transmitter release in rat mesolimbic dopamine neurons studied in vivo”. Neurosci. Lett.1993, 157: 53–56.
- Tokuyama S, Wakabayashi H, Ho I.K; “Direct evidence for a role of glutamate in the expression of the opioid withdrawal syndrome”. Eur. J. Pharmacol.1996, 295:123–129.
- Tokuyama S, Zh, H, Oh S, Ho I.K, Yamamoto T. “Further evidence for a role of NMDA receptors in the locus coeruleus in the expression of withdrawal syndrome from opioids”. Neurochem. Int. 2001, 39: 103–109.
- Fitzgerald L.W, Ortiz J, Hamedani A.G, Nestler E.J, 1996. “Drugs of abuse and stress increase the expression of GluR1 and NMDAR1 glutamate receptor subunits in the rat ventral tegmental area: common adaptations among cross-sensitizing agents”.J. Neurosci.1996, 16: 274–282.
- Kalivas P.W, Duffy P; “Repeated cocaine administration alters extracellular glutamate in the ventral tegmental area”. J. Neurochem.1998, 70: 1497–1502.
- Nestler E.J, Hyman S.E, Malenka R.C, 2001. Molecular Neuropharmacology: A Foundation for Clinical Neuroscience. McGraw-Hill, New York.
- Bisaga A, Comer S.D, Ward A.S; “The NMDA antagonist memantine attenuates the expression of opioid physical dependence in humans. Psychopharmacology 2001, 157: 1–10
- Mitrano D.A, Smith; “Comparative analysis of the subcellular and subsynaptic localization of mGluR1a and mGluR5 metabotropic glutamate receptors in the shell and core of the nucleus accumbens in rat and monkey”. J. Comp.Neurol.2007, 500: 788–806.
- Mitrano D.A, Arnold C, Smith Y; “Subcellular and subsynaptic localization of group I metabotropic glutamate receptors in the nucleus accumbens of cocainetreated rats”. Neuroscience 2008, 154: 653–666.
- Palucha A, Branski P, Pilc A; “Selective mGlu5 receptor antagonist MTEP attenuates naloxone-induced morphine withdrawal symptoms”. Pol. J. Pharmacol.2004, 56: 863–866.
- Schotanus S.M, Chergui K; “Dopamine D1 receptors and group I metabotropic glutamate receptors contribute to the induction of long-term potentiation in the nucleus accumbens. Neuropharmacology 2008, 54: 837–844.
- Akira S, Uematsu S and Takeuchi, O. Cell 2006. 124: 783–801.
- Jessen, Kristjan R. and Mirsky, Rhona. Glial cells in the enteric nervous system contain glial fibrillary acidic protein Nature1980, 286, 736–737.
- Swaminathan, Nikhil. "Glia—the other brain cells".2011.
- Brodal, 2010. 19.
- Swaminathan N. "Brain-scan mystery solved". Scientific American Mind 2008, 7.
- Torres A. "Extracellular Ca2+ Acts as a Mediator of Communication from Neurons to Glia". Science Signaling 2012 24: 208.
- Baumann, Nicole, Pham-Dinh, Danielle. "Biology of Oligodendrocyte and Myelin in the Mammalian Central Nervous System", Physiological Reviews 2001, 18 (2): 871–927
- Johansson CB, Momma S, Clarke DL, Risling M, Lendahl U, Frisén J. "Identification of a neural stem cell in the adult mammalian central nervous system".Cell 1999, 96 (1): 25–
- Campbell K, Götz M. "Radial glia: multi-purpose cells for vertebrate brain development". Trends Neurosci 2002, 25 (5): 235–8.
- 61.Jessen, K. R. & Mirsky, R. "The origin and development of glial cells in peripheral nerves", Nature Reviews Neuroscience 2005, 6 (9): 671–682
- Hanani, M. Satellite glial cells in sensory ganglia: from form to function. Brain Res. Rev. 48:457–476, 2005
- Ohara PT et al., Evidence for a role of connexin 43 in trigeminal pain using RNA interference in vivo. J Neurophysiol 2008;100:3064–3073
- Bassotti, G. et al, Laboratory Investigation 2007, 87, 628–632.
- Nave, K.-A., &Trapp, B. (Eds.).Glia special issue: myelinating glial cells 2000, 29.Wiley-Liss Inc.
- Bidlack JM. Detection and function of opioid receptors on cells from the immune system. Clin Diagn Lab Immunol 2000; 7(5):719–723.
- Chao CC, Hu S, Shark KB, Sheng WS, Gekker G, Peterson PK. Activation of mu opioid Receptors inhibits microglial cell chemotaxis. J Pharmacol Exp Ther 1997; 281(2):998–1004.
- Hu S, Sheng WS, Lokensgard JR, Peterson PK. Morphine induces apoptosis of human microglia and neurons. Neuropharmacology 2002; 42(6):829–836.
- Thomas DM, Dowgiert J, Geddes TJ, Francescutti-Verbeem D, Liu X, Kuhn DM. Microglial activation is a pharmacologically specific marker for the neurotoxic amphetamines. Neurosci Lett 2004; 367(3):349–354.
- Thomas DM, Walker PD, Benjamins JA, Geddes TJ, Kuhn DM. Methamphetamine neurotoxicity in dopamine nerve endings of the striatum is associated with microglial activation. J Pharmacol Exp Ther 2004; 311(1):1–7.
- Thomas DM, Kuhn DM. Attenuated microglial activation mediates tolerance to the neurotoxic effects of methamphetamine. J Neurochem 2005; 92(4):790–797.
- Zhang L, Kitaichi K, Fujimoto Y, et al. Protective effects of minocycline on behavioral changes and neurotoxicity in mice after administration of methamphetamine. Prog Neuropsychopharmacol Biol Psychiatry 2006; 30(8):1381–1393.
- Pope, A. Neuroglia: Quantitative aspects. In: Schoffeniels, E.; Franck, G.; Hertz, L.; Tower, DB. editors. Dynamic Properties of Glia Cells. London: Pergamon; 1978. p. 13-20.
- Haydon PG, Carmignoto G. Astrocyte control of synaptic transmission and neurovascular coupling. Physiol Rev 2006; 86(3):1009–1031. [PubMed: 16816144]
- Duan S, Anderson CM, Stein BA, Swanson RA. Glutamate induces rapid upregulation of Astrocyte and cell-surface expression of GLAST. J Neurosci 1999; 19(23):10193–10200.
- Furuta A, Rothstein JD, Martin LJ. Glutamate transporter protein subtypes are expressed differentially during rat CNS development. J Neurosci 1997; 18(21):8363–8375.
- Had-Aissouni L, Re DB, Nieoullon A, Kerkerian-Le Goff L. Importance of astrocytic inactivation of synaptically released glutamate for cell survival in the central nervous system--are astrocytes vulnerable to low intracellular glutamate concentrations? J Physiol Paris 2002; 96(3–4):317–322.
- Leino RL, Gerhart DZ, van Bueren AM, McCall AL, Drewes LR. Ultrastructural localization of GLUT 1 and GLUT 3 glucose transporters in rat brain. J Neurosci Res 1997; 49(5):617–626.
- Badaut J, Verbavatz JM, Freund-Mercier MJ, Lasbennes F. Presence of aquaporin-4 and Muscarinic receptors in astrocytes and ependymal cells in rat brain: a clue to a common function? Neurosci Lett 2000; 292(2):75–78.
- Badaut J, Lasbennes F, Magistretti PJ, Regli L. Aquaporins in brain: distribution, physiology, and pathophysiology. J Cereb Blood Flow Metabol 2002; 22(4):367–378.
- Itzhak Y, Achat-Mendes C. Methamphetamine and MDMA (ecstasy) neurotoxicity: 'of mice and men'. IUBMB Life 2004; 56(5):249–255.
- 82.Fattore L, Puddu MC, Picciau S, et al. Astroglial in vivo response to cocaine in mouse dentate gyrus: a quantitative and qualitative analysis by confocal microscopy. Neuroscience 2002; 110(1):1–6.
- Hebert MA, O'Callaghan JP. Protein phosphorylation cascades associated with Methamphetamine induced glial activation. Ann N Y Acad Sci 2000; 914:238–262.
- Hill SJ, Barbarese E, McIntosh TK. Regional heterogeneity in the response of astrocytes follows X ing traumatic brain injury in the adult rat. J Neuropathol Exp Neurol 1996; 551221-29X
- Eng LF, Ghirnikar RS. GFAP and astrogliosis. Brain Pathol 1994;4(3):229–237.
- Minn A, Schubert M, Neiss WF, Muller-Hill B. Enhanced GFAP expression in astrocytes of transgenic mice expressing the human brain-specific trypsinogen IV. Glia 1998; 22(4):338-347.
- Steward O, Torre ER, Tomasulo R, Lothman E. Neuronal activity up-regulates astroglial gene expression. Proc Natl Acad Sci U S A 1991; 88(15):6819–6823.
- Steward O, Torre ER, Tomasulo R, Lothman E. Seizures and the regulation of astroglial gene expression. Epilepsy Res Supp 1992; 7:197–209.
- Ricaurte GA, Seiden LS, Schuster CR. Further evidence that amphetamines produce long-lasting dopamine neurochemical deficits by destroying dopamine nerve fibers. Brain Res 1984; 303(2):359– 364.
- Pubill D, Canudas AM, Pallas M, Camins A, Camarasa J, Escubedo E. Different glial response tomethamphetamine- and methylenedioxymethamphetamine- induced neurotoxicity. Naunyn Schmiedebergs Arch Pharmacol 2003; 367(5):490–499.
- Marie-Claire C, Courtin C, Roques BP, Noble F. Cytoskeletal genes regulation by chronic morphine treatment in rat striatum. Neuropsychopharmacology 2004; 29(12):2208–2215.
- Beitner-Johnson D, Guitart X, Nestler EJ. Glial fibrillary acidic protein and the mesolimbic Dopamine system: regulation by chronic morphine and Lewis-Fischer strain differences in the rat ventral tegmental area. J Neurochem 1993;61(5):1766–1773.
- Song P, Zhao ZQ. The involvement of glial cells in the development of morphine tolerance. Neurosci Res 2001; 39(3):281–286.
- Garrido E, Perez-Garcia C, Alguacil LF, Diez-Fernandez C. The alpha2-adrenoceptor
- Antagonist yohimbine reduces glial fibrillary acidic protein upregulation induced by chronic morphine administration. Neurosci Lett 2005; 383(1–2):141–144.
- Alonso E, Garrido E, Diez-Fernandez C, et al. Yohimbine prevents morphine-induced of glial fibrillary acidic protein in brain-stem and alpha2-adrenoceptor gene expression in hippocampus. Neurosci Lett 2007; 412(2):163–167.
- Ronnback L, Hansson E. Are astroglial cells involved in morphine tolerance? Neurochem Res 1988; 13 (2):87–103.
- Ronnback L, Hansson E. Modulation of astrocyte activity--one way to reinforce morphine effects? Cell Mol Biol 1988; 34(4):337–349.
- Narita M, Miyatake M, Suzuki M, Kuzumaki N, Suzuki T. Role of astrocytes in rewarding effects of drugs of abuse. Nihon Shinkei Seishin Yakurigaku Zasshi 2006; 26(1):33–39.
- Mao J, Sung B, Ji RR, Lim G. Chronic morphine induces down-regulation of spinal glutamate transporters: implications in morphine tolerance and abnormal pain sensitivity. J Neurosci 2002; 22 (18):8312–8323. [PubMed: 12223586]
- Appel E, Kolman O, Kazimirsky G, Blumberg PM, Brodie C. Regulation of GDNF expression in cultured astrocytes by inflammatory stimuli. Neuroreport 1997; 8(15):3309–3312.
- Lee YJ, Jin JK, Jeong BH, Carp RI, Kim YS. Increased expression of glial cell line-derived neurotrophic factor (GDNF) in the brains of scrapie-infected mice. Neurosci Lett 2006; 410(3):178–182.
- Cass WA. GDNF selectively protects dopamine neurons over serotonin neurons against the neurotoxic effects of methamphetamine. J Neurosci 1996; 16(24):8132–8139.
- Boger HA, Middaugh LD, Patrick KS, et al. Long-Term consequences of methamphetamine exposure in young adults are exacerbated in glial cell line-derived neurotrophic factor heterozygous mice. J. Neurosci 2007; 27:8816–8825.
- Pierce RC, Bari AA. The role of neurotrophic factors in psychostimulant-induced behavioral and neuronal plasticity. Rev Neurosci 2001; 12(2):95–110.
- Messer CJ, Eisch AJ, Carlezon WA Jr, et al. Role for GDNF in biochemical and behavioral adaptations to drugs of abuse. Neuron 2000; 26(1):247–257.
- Niwa M, Nitta A, Yamada K, Nabeshima T. The roles of glial cell line-derived neurotrophic factor, tumor necrosis factor-alpha, and an inducer of these factors in drug dependence. J Pharmacol Sci 2007; 104(2):116–121.
- Niwa M, Nitta A, Yamada Y, et al. An inducer for glial cell line-derived neurotrophic factor and tumor necrosis factor-alpha protects against methamphetamine-induced rewarding effects and sensitization. Biol Psychiatry 2007; 61(7):890–901.
- Nakajima A, Yamada K, Nagai T, et al. Role of tumor necrosis factor-alpha in Methamphetamine induced drug dependence and neurotoxicity. J Neurosci 2004; 24(9):2212–2225.
- Coyne L; "Target Analysis - Toll-like Receptors”. Pharmaprojects2008, 29: 1-4.
- Kawai T, Akira S; “The roles of TLRs, RLRs and NLRs in pathogen recognition”. Int. Immunol 2009, 21: 317-337.
- Mata-Haro V, Cekic C, Martin M, Chilton PM, Casella CR, Mitchell TC; “The vaccine adjuvant monophosphoryl lipid A as a TRIFbiased agonist of TLR4” . Science 2007, 316: 1628–1632.
- Casella C.R, Mitchell T.C, Putting endotoxin to work for us: “monophosphoryl lipid A as a safe and effective vaccine adjuvant.Cell”. Mol. Life Sci 2008, 65: 3231–3340.
- Lombardi V., Van Overtvelt L, Horiot S, Moussu H, Chabre H, Louise A, Balazuc AM, Mascarell L, Moingeon P; “Toll-like receptor 2 agonist Pam3CSK4 enhances the induction of antigen-specific tolerance via the sublingual route” . Clin. Exp. Allergy 2008 38: 1705-1706.
- Ishii KJ, Akira S; “Toll or Toll-Free Adjuvant Path toward the Optimal Vaccine Development”. J. Clin. Immunol 2007. 27: 363-371.
- Andersen-Nissen E, Smith KD, Bonneau R, Strong RK, Aderem A; “A conserved surface on Toll-like receptor 5 recognizes bacterial flagellin” . J. Exp. Med 2007, 204: 393–403.
- Burdelya LG, Krivokrysenko VI, Tallant TC; “An agonist of Toll like receptor 5 has radioprotective activity in mouse and primate models”. Science 2008, 320: 226–230.
- Huleatt JW, Nakaar V, Desai P; “Potent immunogenicity and efficacy of a universal influenza vaccine candidate comprising a recombinant fusion protein linking influenza M2e to the TLR5 ligand flagellin”. Vaccine 2008, 26: 201–214.
- Gupta AK, Cherman AM, Tyring SK; “Viral and nonviral uses of imiquimod: a review”. J. Cutan. Med. Surg 2004, 8: 338–352.
- Schon MP, Schon M; “TLR7 and TLR8 as targets in cancer therapy”. Oncogene 2008, 27: 190–199.
- Berman B, Poochareon VN, Villa AM; “Novel Dermatologic Uses of the Immune Response Modifier Imiquimod 5% Cream. Skin Ther” . Lett 2002, 7: 1-6.
- Bailey S.L, Carpentier P.A, McMahon E.J, Begolka W.S, and S. D. Miller S.D; “Innate and adaptive immune responses of the central nervous system,” Critical Reviews in Immunology 2006, 26, 149–188.
- Block M.L, Zecca L. and Hong J.S; “MicroGliamediated neurotoxicity: uncovering the molecular mechanisms,”Nature Reviews Neuroscience 2007, 8: 57–69.
- Konat G.W, Kielian T. and Marriott I; “The role of Toll-like receptors in CNS response to microbial challenge,” Journal of Neurochemistry2006, 99:1–12.
- Chen K, Iribarren P, Hu J.et al; “Activation of Toll-like receptor 2 on microGlia promotes cell uptake of Alzheimer disease-associated amyloid β peptide,” Journal of Biological Chemistry 2006, 281: no. 6, 3651–3659, 2006.
- Xie Z, Wei M, Morgan T.E, et al. “Peroxynitrite mediates neurotoxicity of amyloid β- peptide 1-42- and lipopolysaccharide- activated micro glia,” Journal of Neuroscience 2002, 22, no. 9, 3484–3492.
- Muzio M, Mantovani A; “Toll-like receptors (TLRs) signalling and expression pattern”. J Endotoxin Res 2001, 7:297–300.
- Guo L.H., Schluesener H.J; “The innate immunity of the central nervous system in chronic pain: the role of Toll-like receptors”. Cell Mol Life Sci 2007, 64:1128–1136. This article offers a detailed discussion on TLRs and their role in the development of various types of pain.
- Tanga F.Y, Nutile-McMenemy N, Deleo J.A. “The CNS role of Toll like receptor 4 in innate neuroimmunity and painful neuropathy”. Proc Natl Acad Sci U S A 2005, 102: 5856– 5861.
- Kim D., Kim MA., Cho IH., et al; “A critical role of toll-like receptor 2 in nerve injury-induced spinal cord glial cell activation and pain hypersensitivity” . J Biol Chem 2007, 282:14975–14983.
- Benvensile, E.N. The role of cytokines in multiple sclerosis, autoimmune encephalina and other neurological disorders. In human cytokines: Their role in Research and Therapy. 195-216.
- Avigen, Inc.: Available at www.avigen.com. Accessed November2009.
- Ledeboer A, Hutchinson MR, Watkins LR, Johnson KW: Ibudilast (AV-411). A new class therapeutic candidate for neuropathic pain and opioid withdrawal syndromes. Expert Opin Investig Drugs 2007, 16:935–950. This review details the effectiveness of ibudilast in the management of opioid tolerance and reward.
- Kishi Y, Ohta S, Kasuya N, et al.: Ibudilast: a non-selective PDE inhibitor with multi actions on blood cells and the vascular wall. Cardiovasc Drug Rev 2001, 19:215–225.
- Fujimoto T, Sakoda S, Fujimura H, Yanagihara T: Ibudilast, a phosphodiesterase inhibitor, ameliorates experimental autoimmune encephalomyelitis in Dark August rats. J Neuroimmunol 1999, 95:35–42.
- Feng J, Misu T, Fujihara K, et al.: Ibudilast, a nonselective phosphodiesterase inhibitor, regulates Th1/Th2 balance and NKT cell subset in multiple sclerosis. Mult Scler 2004, 10:494–498
- Hutchinson MR, Zhang Y, Shridhar M, et al.: Evidence that opioids may have toll-like receptor 4 and MD-2 effects. Brain Behav Immun 2010, 24:83–95. This paper summarizes recent information regarding the highly central role of TLR-4 receptors with implications to the current developments in the opioidrelated tolerance and dependence modifying agents.
- Harris GC, Aston-Jones G: Involvement of D2 dopamine receptors in the nucleus accumbens in the opiate withdrawal syndrome. Nature 1994, 371:155–157.
- Bland ST, Hutchinson MR, Maier SF, et al.: The glial activation inhibitor AV411 reduces morphine-induced nucleus accumbens dopamine release. Brain Behav Immun 2009, 23:492–497. This review discusses the mechanism of action of AV411 and its relationship to TLRs.
- Kest B, Mogil JS, Shamgar BE, et al.: The NMDA receptor antagonist MK-801 protects against the development of morphine tolerance after intrathecal administration. Proc West Pharmacol Soc 1993, 36:307–310.
- Sweitzer SM, Pahl JL, DeLeo JA: Propentofylline attenuates vincristine-induced peripheral neuropathy in the rat. Neurosci Lett 2006, 400:258–261.
- Gwak YS, Crown ED, Unabia GC, Hulsebosch CE: Propentofylline attenuates allodynia, glial activation and modulates GABAergic tone after spinal cord injury in the rat. Pain 2008, 138:410–422.
- Narita M, Miyatake M, Narita M, et al.: Direct evidence of astrocytic modulation in the development of rewarding effects induced by drugs of abuse. Neuropsychopharmacology 2006, 31:2476–2488.
- Mika J, Wawrzczak-Bargiela A, Osikowicz M, et al.: Attenuation of morphine tolerance by minocycline and pentoxifylline in naive and neuropathic mice. Brain Behav Immun 2009, 23:75– 84.
- Kim HS, Suh YH: Minocycline and neurodegenerative diseases. Behav Brain Res 2009, 196:168–179.
- Li WW, Setzu A, Zhao C, Franklin RJ: Minocycline-mediated inhibition of microglia activation impairs oligodendrocyte progenitor cell responses and remyelination in a non-immune model of demyelination. J Neuroimmunol 2005, 158:58–66.
- Ledeboer A, Sloane EM,Milligan ED, et al.: Minocycline attenuates mechanical allodynia and proinflammatory cytokine expression in rat models of pain facilitation. Pain 2005, 115:71–83.
- 147. Hutchinson MR, Bland ST, Johnson KW, et al.: Opioid-induced glial activation: mechanisms of activation and implications for opioid analgesia, dependence, and reward. ScientificWorldJournal 2007, 7:98–111. This article discusses TLRs and modifying agents and their relevance to the future of the management of pain. It also includes detailed information related to the research regarding naloxone in the management of opioid side effects.
- Wei T, Sabsovich I, Guo TZ, et al.: Pentoxifylline attenuates nociceptive sensitization and cytokine expression in a tibia fracture rat model of complex regional pain syndrome. Eur J Pain 2009, 13:253–262.
- Dorazil-Dudzik M, Mika J, Schafer MK, et al.: The effects of local pentoxifylline and propentofylline treatment on formalininduced pain and tumor necrosis factor-alpha messenger RNA levels in the inflamed tissue of the rat paw. Anesth Analg 2004, 98:1566–1573.
- Liu J, Feng X, Yu M, et al.: Pentoxifylline attenuates the development of hyperalgesia in a rat model of neuropathic pain. Neurosci Lett 2007, 412:268–272.
- Bland ST, Hutchinson MR, Maier SF, et al.: The glial activation inhibitor AV411 reduces morphine-induced nucleus accumbens dopamine release. Brain Behav Immun 2009, 23:492–497. This review discusses the mechanism of action of AV411 and its relationship to TLRs
How to cite this article:
Bawa G: Glial Cells Responses: In Opioid Withdrawal Syndrome. Int J Pharm Sci Res 2013; 4(6); 2118-2131. doi: 10.13040/IJPSR.0975-8232.4(6).2118-31
All © 2013 are reserved by International Journal of Pharmaceutical Sciences and Research. This Journal licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 3.0 Unported License.
Article Information
10
2118-2131
429KB
1187
English
IJPSR
Gurpreet Bawa
Department of Pharmacology, Rayat and Bahra Institute of Phamacy, Sahauran, Kharar, District Mohali-140 104, Punjab, India
gurpreet_bawa@ymail.com
19 February, 2013
23 April, 2013
18 May, 2013
http://dx.doi.org/10.13040/IJPSR.0975-8232.4(6).2118-31
01 June, 2013