13
OVERVIEW Role of Astrocytes in Pain C.-Y. Chiang B. J. Sessle J. O. Dostrovsky Received: 30 January 2012 / Revised: 18 April 2012 / Accepted: 7 May 2012 / Published online: 26 May 2012 Ó Springer Science+Business Media, LLC 2012 Abstract Over the last decade, a series of studies has demonstrated that glia in the central nervous system play roles in many aspects of neuronal functioning including pain processing. Peripheral tissue damage or inflammation initiates signals that alter the function of the glial cells (microglia and astrocytes in particular), which in turn release factors that regulate nociceptive neuronal excit- ability. Like immune cells, these glial cells not only react at sites of central and/or peripheral nervous system damage but also exert their action at remote sites from the focus of injury or disease. As well as extensive evidence of mi- croglial involvement in various pain states, there is also documentation that astrocytes are involved, sometimes seemingly playing a more dominant role than microglia. The interactions between astrocytes, microglia and neurons are now recognized as fundamental mechanisms underlying acute and chronic pain states. This review focuses on recent advances in understanding of the role of astrocytes in pain states. Keywords Astrocyte Intracellular calcium Gliotransmitter Chemokine Inflammatory pain Neuropathic pain Introduction It has been clearly demonstrated that glia in the central nervous system play roles in many aspects of neuronal functioning including pain processing. Peripheral tissue damage or inflammation initiates signals that alter the function of the glial cells (microglia and astrocytes in particular), causing the release of factors that regulate nociceptive neuronal excitability. The aim of this article is to focus on the recent advances in the role of astrocytes in pain at the spinal cord and medullary levels. Overview of Pain Research Starting from the 1920s, electrophysiological studies have revealed that peripheral nerve fibers are composed of fast conducting thick myelinated Aa and Ab fibers, and slow conducting thin myelinated Ac and Ad and non-myelinated C fibers; normally, many of these Ad and C fibers are involved in pain and aversive reactions. Until fairly recently, research focused on the normal functional char- acteristics of the nociceptors in the peripheral nervous system and the nociceptive neurons in the central nervous system (CNS). Nociceptive neurons are found primarily in the superficial laminae I–II and deep laminae V–VI of the Special Issue: Leif Hertz/overview. C.-Y. Chiang (&) B. J. Sessle J. O. Dostrovsky Department of Oral Physiology, Faculty of Dentistry, University of Toronto, 124 Edward Street, Toronto, ON M5G 1G6, Canada e-mail: [email protected] B. J. Sessle e-mail: [email protected] J. O. Dostrovsky e-mail: [email protected] B. J. Sessle J. O. Dostrovsky Department of Physiology, Faculty of Medicine, University of Toronto, 1 King’s Circle, MSB, Toronto, ON M5S 1A8, Canada 123 Neurochem Res (2012) 37:2419–2431 DOI 10.1007/s11064-012-0801-6

Role of Astrocytes in Pain - Dr. T. Howard Black · To explore the role of astrocytes in pain, two strategies ... (ATP) can readily elicit ... release probability and insertion of

  • Upload
    dangtu

  • View
    213

  • Download
    0

Embed Size (px)

Citation preview

OVERVIEW

Role of Astrocytes in Pain

C.-Y. Chiang • B. J. Sessle • J. O. Dostrovsky

Received: 30 January 2012 / Revised: 18 April 2012 / Accepted: 7 May 2012 / Published online: 26 May 2012

� Springer Science+Business Media, LLC 2012

Abstract Over the last decade, a series of studies has

demonstrated that glia in the central nervous system play

roles in many aspects of neuronal functioning including

pain processing. Peripheral tissue damage or inflammation

initiates signals that alter the function of the glial cells

(microglia and astrocytes in particular), which in turn

release factors that regulate nociceptive neuronal excit-

ability. Like immune cells, these glial cells not only react at

sites of central and/or peripheral nervous system damage

but also exert their action at remote sites from the focus of

injury or disease. As well as extensive evidence of mi-

croglial involvement in various pain states, there is also

documentation that astrocytes are involved, sometimes

seemingly playing a more dominant role than microglia.

The interactions between astrocytes, microglia and neurons

are now recognized as fundamental mechanisms

underlying acute and chronic pain states. This review

focuses on recent advances in understanding of the role of

astrocytes in pain states.

Keywords Astrocyte � Intracellular calcium �Gliotransmitter � Chemokine � Inflammatory pain �Neuropathic pain

Introduction

It has been clearly demonstrated that glia in the central

nervous system play roles in many aspects of neuronal

functioning including pain processing. Peripheral tissue

damage or inflammation initiates signals that alter the

function of the glial cells (microglia and astrocytes in

particular), causing the release of factors that regulate

nociceptive neuronal excitability. The aim of this article is

to focus on the recent advances in the role of astrocytes in

pain at the spinal cord and medullary levels.

Overview of Pain Research

Starting from the 1920s, electrophysiological studies have

revealed that peripheral nerve fibers are composed of fast

conducting thick myelinated Aa and Ab fibers, and slow

conducting thin myelinated Ac and Ad and non-myelinated

C fibers; normally, many of these Ad and C fibers are

involved in pain and aversive reactions. Until fairly

recently, research focused on the normal functional char-

acteristics of the nociceptors in the peripheral nervous

system and the nociceptive neurons in the central nervous

system (CNS). Nociceptive neurons are found primarily in

the superficial laminae I–II and deep laminae V–VI of the

Special Issue: Leif Hertz/overview.

C.-Y. Chiang (&) � B. J. Sessle � J. O. Dostrovsky

Department of Oral Physiology, Faculty of Dentistry,

University of Toronto, 124 Edward Street,

Toronto, ON M5G 1G6, Canada

e-mail: [email protected]

B. J. Sessle

e-mail: [email protected]

J. O. Dostrovsky

e-mail: [email protected]

B. J. Sessle � J. O. Dostrovsky

Department of Physiology, Faculty of Medicine,

University of Toronto, 1 King’s Circle,

MSB, Toronto, ON M5S 1A8, Canada

123

Neurochem Res (2012) 37:2419–2431

DOI 10.1007/s11064-012-0801-6

spinal and medullary dorsal horns, and also in some regions

of the thalamus and cortex. In addition, there have been

many studies of the endogenous analgesic descending

system, which modulates nociceptive inputs particularly at

the spinal and medullary levels [1].

During the 1960–1980s, the discovery in the brain of

specific receptors of amino acids (GABA, glutamate, etc.)

and peptides/proteins (opioid peptides, neurotrophic fac-

tors, kinases, etc.) has tremendously enriched and pro-

moted research into mechanisms underlying various pain

states. Following tissue damage and the activation of

nociceptive afferent endings in the tissue, the local com-

plex process of inflammation which is associated with

release of serotonin, bradykinin, prostaglandins, growth

factors and cytokines etc. can also lead to the increased

responsiveness of the peripheral nociceptor endings; this

sensitization of nociceptive afferent fiber endings (periph-

eral sensitization) is presumed to be partially responsible

for the increased pain sensitivity (primary hyperalgesia) in

injured tissue [2]. Congruent with this peripheral process, a

prolonged nociceptive neuronal hypersensitivity in associ-

ation with an increased synaptic activity in the CNS (e.g.,

the spinal and medullary dorsal horns) has been docu-

mented as increases in spontaneous activity, mechanore-

ceptive field size and responses to mechanical and thermal

stimuli and decreases in activation threshold that are

coherent with the development of hyperalgesia (an exag-

gerated response to noxious stimulation) and allodynia

(pain produced by a normally non-noxious stimulus, e.g.,

touch) [3–5]. This phenomenon of central neuronal

hypersensitivity is termed ‘‘central sensitization’’ and is

now widely accepted as an essential mechanism underlying

various pain states, including acute and chronic inflam-

matory pain and neuropathic pain [6–8].

Despite the remarkable advances in our understanding

of the underlying neuronal mechanisms, most novel drugs

derived from these cellular and molecular investigations

have proven to be of limited clinical effectiveness in the

treatment of pain conditions, suggesting that other impor-

tant factors e.g., neuron-glia (or neuroimmune) interactions

may also be involved [8–14]. Indeed, concomitant with the

studies on neuronal mechanisms, a tremendous develop-

ment of in vitro studies on glia has also been carried out by

means of various neurobiological techniques as well as the

latest in vivo non-invasive methods such as positron

emission tomography, magnetic resonance imaging scan-

ning, magnetic resonance spectroscopy etc. Astrocytes (as

well as microglia and the oligodendrocytes, which are not

included in this review) have been comprehensively

investigated; their functions cover a wide range of

dynamic structure–function characteristics: water and ion

balance, metabolic specialization and cerebral oxidative

metabolism, cell–cell communication, and roles of astro-

cytes in various diseases. The prominent features of

astrocytes relevant to neurotransmission and global brain

function have been recently reviewed in depth [10, 11, 15–

23]; in particular, the studies of Hertz et al. [24, 25] have

played a prominent role in elucidating many of the meta-

bolic and physiological processes of astrocytes. Some

features of astrocytes are briefly summarized in Box 1.

In the early 1990s, pioneering work [26, 27] demon-

strated that the NMDA receptor antagonist MK801 could

block sciatic chronic constriction injury-induced hyperal-

gesia and increased glial fibrillary acidic protein (GFAP)

expression (which is used as a marker of enhanced

astrocytic activity, although the function of GFAP

remains unclear in relation to the activity state of astro-

cytes). In addition, it was shown that intraplantar zymo-

san-induced thermal and mechanical hyperalgesia and

mechanical allodynia were blocked by fluorocitrate, a

selective inhibitor of aconitase which is an enzyme in the

astrocytic Krebs cycle [28, 29]. These findings revealed a

predictive relationship between glial activation and

exaggerated pain states. Subsequently, this analgesic

effect of fluorocitrate (or fluoroacetate) has been repeat-

edly verified in various experimental pain models of acute

and chronic inflammatory pain as well as neuropathic pain

(see below), and was followed in the last decade by the

discovery of a remarkable series of complicated mecha-

nisms related to neuron-glial interactions involved in pain

processing [8, 10, 13, 30–39]. The remainder of this

article will especially focus on recent advances in the role

of astrocytes in pain at the spinal and medullary dorsal

horns, leaving those of satellite glial cells (the astrocyte-

like cells in the spinal dorsal root/trigeminal ganglion) to

other specific reviews [40–43].

To explore the role of astrocytes in pain, two strategies

have been usually adopted: (1) most studies have used

conventional approaches by which the inflammatory and/or

injury-evoked increased afferent impulses initiate signals

that alter the function of glial cells in the peripheral or

central nervous system, and when recruited, these glial

cells in turn modulate neuronal function. By means of

multiple technologies including electrophysiological, neu-

rochemical, immunocytochemical, optical recording toge-

ther with Ca2? sensitive dyes, and behavioral tests

performed in wild and gene-knockout rodents, these studies

have systematically investigated the individual changes in

astrocytes, microglia, neurons, transmitters, mediators, and

their interactions in different pain states. (2) A few studies

have used specific chemicals to interfere with or block one

of the essential functions of astrocytes (but not microglia or

neurons) to demonstrate their involvement in different pain

states.

2420 Neurochem Res (2012) 37:2419–2431

123

Glia-Neuron Interactions in Pain States

Role of Astrocytes in Different Pain States and Their

Activation and Time Courses Relative to Microglia

Although the features underlying the processes and time

courses of astrocytic and microglial activation in inflam-

matory and neuropathic pain are variable and dependent on

the different experimental pain models [44], a general

scheme still can be outlined.

1. Acute inflammatory pain: Following subcutaneous

injection of the inflammatory irritant formalin, first

GFAP-immunoreactive astrocytes are detected in the

spinal dorsal horn, then OX42-immunoreactive

microglia are observed a few minutes later, and Fos/

NeuN-immunoreactive (neuron marker) neurons are

found slightly later [45]. Similarly, a prompt (within

15 min) onset of central sensitization in the medullary

dorsal horn induced by application of the inflammatory

irritant mustard oil to the tooth pulp can be blocked by

pre-emptive intrathecal superfusion of the astrocytic

aconitase inhibitor fluoroacetate or the NMDA recep-

tor antagonist MK-801 [46, 47]. Following paw

incision, GFAP expression in laminae I-II and ionized

calcium binding adaptor protein (Iba-1; microglia

marker) in deep laminae of the spinal lumbar 5

segment are increased on day 1, and then in the entire

dorsal horn by day 4, and dissipate by day 7 in parallel

with the accompanying allodynia [48].

2. Chronic inflammatory pain: In the acute phase

0.5–24 h after injection of Complete Freund’s Adju-

vant (an immunopotentiator) into a jaw muscle (mas-

seter), significant increases occur in NMDA receptor

NR1 serine 896 phosphorylation and GFAP levels in

the medullary dorsal horn (also known as trigeminal

subnucleus caudalis) and its junctional region with the

more rostral subnucleus interpolaris (the interpolaris/

caudalis transition zone) which is an important struc-

ture for processing orofacial deep as well as cutaneous

nociceptive inputs; local injection of fluorocitrate or

the microglial inhibitor minocycline at day 1 into the

transition zone significantly attenuates the masseter

hyperalgesia that occurs bilaterally in this pain model

[49]. Also, following Complete Freund’s Adjuvant-

induced peripheral inflammation, a significant increase

in microglial markers (macrophage antigen complex-1,

Box 1 Functional features of astrocytes

1. The membrane of astrocytes expresses a variety of neurotransmitter receptors including N-methyl D-aspartate (NMDA) and non-NMDA,

metabotropic glutamatergic, neurotrophic tyrosine kinase, neurokinin-1, purinergic, adrenergic, cytokine, and aquaporin receptors. Most

metabotropic receptors when activated may induce the phospholipase-dependent accumulation of inositol 1, 4, 5-triphosphate that

stimulates the release of Ca2? from intracellular inositol 1, 4, 5-triphosphate-sensitive internal stores [16, 118, 142]

2. Glutamate and adenosine-50-triphosphate (ATP) can readily elicit astroglial transients and oscillatory waves of cytoplasmic Ca2?, which

may propagate to adjoining and/or distant astrocytes and neurons through two pathways: (1) the cytoplasmic diffusion of Ca2?-mobilizing

second messengers inositol 1, 4, 5- triphosphate through the gap junction channels between astroglia, and (2) extracellular diffusion of ATP

that is released through astrocytic connexins (Cx) e.g., Cx-43 and pannexin-1 hemichannels and/or vesicular exocytosis and then acts on

nearby glia and neurons to modulate neurotransmission [15, 30, 119, 120, 143]. A disruption of astrocytic networks (Cx-30-/- and Cx-

43-/-) causes combined pre- and post-synaptic alterations including enhanced neuronal excitability, release probability and insertion of

postsynaptic a-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors [62]

3. In exaggerated pain states reflected in hyperalgesia and allodynia, the activated astrocytes (or properly termed ‘‘reactive astrocytes’’) may

release gliotransmitters such as glutamate, ATP, D-serine, or neurotrophic factors, cytokines (e.g., interleukin (IL)-1beta) and chemokine

chemoattractant ligands (CCL)-2. Notably, glutamate is a potent excitatory neurotransmitter, and ATP through its metabolic product

adenosine indirectly inhibits the function of local cells i.e., neurons and glial cells [16, 18, 22, 30–34, 144–148]

4. Astrocytes readily uptake K? in the extracellular space through Na?-K?-Cl- cotransporters (NKCC). They also regulate NKCC in cerebral

microvessel endothelial cells by secreting interleukin-6, thus participating in the maintenance of cerebral ionic homeostasis across the

blood- brain barrier [149, 150]. Astrocytes also efficiently uptake extrasynaptic glutamate by Na?-dependent electrogenic uptake systems.

An increase in the intracellular concentration of Na? stimulates glycolysis in astrocytes i.e., glucose utilization and lactate production

mediated by an activation of Na?/K? ATPase [25, 151]

5. Glutamate, while being an important excitatory neurotransmitter, is also an important metabolic agent in the astrocytes where it stimulates

glycolysis, i.e., glucose utilization and lactate production which are then degraded to carbon dioxide and water. Of the extrasynaptic

glutamate uptaken by astrocytes, about one-third enters the tricarboxylic acid cycle (Krebs cycle) to generate de novo glutamate, and two-

thirds transform to glutamine (a glutamate precursor) which is released into the extracellular space and then is taken up by glutamatergic

presynaptic terminals to replenish the glutamate pool for neurotransmission. This process is termed the ‘‘glutamate-glutamine shuttle (or

cycle)’’ [24]

6. Astrocytes bidirectionally exchange information with the pre- and post-synaptic neuronal elements, responding to synaptic activity and, in

turn, regulating synaptic transmission (thereby also being an integral part of the ‘‘tripartite synapse’’) and their ramified end-feet closely

contact local vessels [152]. Such a well-organized complex structure can modulate uptake of K?, uptake and release of glucose-derived

transmitters (glutamate and GABA) and various neuroactive molecules to monitor nearby neuronal excitability as well as glial functions and

particularly, to play a critical role in coupling blood flow to meet functional demands of neural activity [16, 17, 21, 22]

Neurochem Res (2012) 37:2419–2431 2421

123

OX-42 and CD14) as well as an enhanced expression

of proinflammatory cytokines such as IL-1b, IL-6 and

tumor necrosis factor (TNFa) occurs in the spinal cord,

brainstem and forebrain during all phases (acute,

subacute and chronic) of inflammation, whereas up-

regulation of astroglial markers (GFAP and S100b, a

calcium-binding peptide) is observed only at the

subacute and chronic phases of inflammation, thus

indicating that microglial activation precedes astro-

cytic activation [50]. Consistently, an early expression

of phosphorylated p38 mitogen-activated protein

kinase (MAPK) in the microglia and late induction

of phosphorylated nuclear factor (NF)jB in astrocytes

in the medullary dorsal horn both play an important

role in trigeminal neuropathic pain resulting from mal-

positioned dental implants [51]; both p38 MAPK and

phosphorylated extracellular signal-regulated protein

kinase (ERK) (but not c-Jun N-terminal kinase, JNK)

are mostly up-regulated and co-localized with neurons

and microglia, but rarely with astrocytes, during the

3–10 days following occlusal interference [52 and

personal communication]. Also, GFAP co-localized

with glutamine synthetase antibody-immunoreactive

astrocytes is significantly increased for 7–14 days after

pulp exposure in a chronic dental pulpitis pain model

[53].

3. Neuropathic pain: In a neuropathic pain model

involving lumbar 5 spinal nerve transection in adult

rats, the spinal S100b mRNA and protein are steadily

increased from day 4 to 28, implicating the late

involvement of astrocytes [54]. In the same pain

model, mechanical allodynia is induced from day 1 to

7 which correlates with Iba-1 increases in the spinal

lumbar 5 segment, whereas GFAP increases from day

4 to 7, thus suggesting that astrocytes (as reflected in

GFAP expression) play a role in the maintenance of

chronic pain while microglia activation (reflected in

Iba-1 expression) closely correlates with the early

phase in this neuropathic pain model [48]. Also, in rats

with injury to the inferior alveolar nerve, GFAP-

labelled astrocytes and nociceptive neuronal excitabil-

ity reflecting central sensitization increases in the

ipsilateral medullary dorsal horn and is associated with

nociceptive behaviour at postoperative day 7; all these

changes can be prevented following intrathecal fluo-

roacetate administration [55]. Similarly, in rats with

upper cervical nerve injury, GFAP-labelled astrocytes

(most also showed glutamine synthetase immunoreac-

tivity) and ERK-immunoreactive neurons occur on day

7 in superficial laminae of both medullary and upper

cervical dorsal horns in association with behavioral

allodynia and heat hyperalgesia that can be prevented

by post-operative administration of fluoroacetate or the

MEK1/2 inhibitor PD98059 [56]. Interestingly, this

spinal/medullary microglial and astrocytic response

profile after nerve injury may vary either during

development of animals [57] or depending on the

lesion site (e.g., sciatic nerve versus infraorbital nerve

chronic constriction injury) in adult animals [58].

Clarifying the temporal profiles of microglial and

astrocytic activation in various pain models will be

helpful in determining the optimal time for drug

administration to counteract the glial changes contrib-

uting to the pain states.

Another important finding is that none of the glial

modulatory drugs affected normal pain processing, indi-

cating that glial activation may be only involved in path-

ological pain states [35, 43]. This leads to the next

consideration of whether direct activation of glia can

generate nociceptive neuronal hypersensitivity associated

with mechanical and thermal hyperalgesia and tactile

allodynia.

Direct Activation of Astrocytes in the Dorsal Horn

by i.t Exogenous Chemicals

Given such a complicated neuron-glia interaction and that

both elements share most of the same receptors in the

spinal and medullary dorsal horns, it is challenging to

selectively target a set of astrocytes or microglia in in vivo

conditions. One approach was provided by Watkins’ group

who showed that i.t. application of an immune stimulus

(HIV envelope glycoprotein, gp120) produces robust

thermal hyperalgesia and mechanical allodynia which are

blocked by spinal pretreatment with fluorocitrate and CNI-

1493; the latter compound is thought to preferentially

disrupt the function of microglia and astrocytes, because

they possess CXCR4 or CCR5 chemokine receptors which

can bind to gp120 and then initiate a cascade of mediators

and release of proinflammatory cytokines, nitric oxide and

prostaglandins to activate neurons as well as glial cells

[59].

Regarding astrocyte activation, a recent study has

demonstrated astrocyte-to-neuron responses in rat spinal

cord slices mediated by gliotransmitters (glutamate), i.e., in

experiments involving patch-clamp and confocal micro-

scopic calcium imaging techniques, application of ben-

zoylbenzoyl ATP (a potent P2X7 receptor agonist) was

shown to trigger Ca2? elevations in astrocytes through

purinergic receptors and induce slow inward currents in

lamina II neurons which were blocked by d-AP5 (amino-5-

phosphonovaleric acid, a competitive NMDA receptor

antagonist) [60]. These findings suggest that astrocyte-

released glutamate evokes NMDA receptor-mediated epi-

sodes of synchronous activity in groups of superficial

2422 Neurochem Res (2012) 37:2419–2431

123

laminae I–II neurons. In accordance with these findings,

the development of thermal hyperalgesia and mechanical

allodynia in a model of inflammation produced by in-

traplantar zymosan is also accompanied by a significant

increase of spontaneous slow inward currents in dorsal

horn neurons. It thus appears that NMDA receptor-medi-

ated astrocyte-to-neuron signaling may contribute to the

control of central sensitization in pathological pain states

[60]. Another study showed that spinal intrathecal injection

of astrocytes, which had been briefly (15 min) pretreated

by TNFa, induced a substantial decrease in paw withdrawal

thresholds, indicating the development of mechanical

allodynia. This allodynia was prevented when the astrocyte

cultures had been pre-treated with a JNK inhibitor or by a

MCP-1 (also termed chemotactic cytokine ligand 2

[CCL2]) neutralizing antibody. Finally, pretreatment of

astrocytes with MCP-1 siRNA attenuated astrocyte-

induced mechanical allodynia [61].

These findings demonstrate that direct activation of

either astrocytes or microglia not only produces exagger-

ated pain states including neuronal hypersensitivity asso-

ciated with behavioral hyperalgesia and allodynia, but also

suggest that they are involved in a particular phase of the

pathological pain process [48, 54]. Thus, astrocytes may

play a role in the initiation of acute inflammatory pain [45–

47, 60] and the maintenance of chronic inflammatory and

neuropathic pain [48, 50, 51, 53–56, 61], and microglia are

involved in early or all phases of pathological pain [48, 50,

51, 59]. However, these findings are not conclusive and

further stringent validation is still needed in additional

experiments such as using antisense, transgenic or gene-

knocked animals as previously reported [61–67]. For

instance, a recent study showed that liposaccharide-

induced hypersensitivity thought to be mediated by toll-

like receptors can be observed in wild-type but not P2X7

knock-out mice [68], indicating that P2X7 receptors are

essential elements for immune antigen-induced neuronal

hypersensitivity.

Signaling Pathways Underlying Inflammatory

and Neuropathic Pain States

Peripheral inflammation and nerve injury activate several

signaling pathways (e.g., protein kinases A and C, calcium/

calmodulin-dependent protein kinase) in primary sensory

and dorsal horn neurons that then activate MAPKs

including p38 MAPK, ERK, and JNK in spinal and med-

ullary dorsal horn neurons and microglia or astrocytes or

both, leading to the glial production of pro-inflammatory

mediators (e.g., IL-1b, TNFa) that sensitize dorsal horn

neurons as well as induce behavioral hyperalgesia and

allodynia [32, 56, 69]. For instance, after nerve injury the

activation of NR2B-containing NMDA receptor-mediated

astrocytic JNK activation in the spinal dorsal horn releases

IL-1b which involves positive feedback mechanisms to

enhance and prolong neuropathic pain [70]. Also, local

injection of Complete Freund’s Adjuvant into the masseter

muscle caused an upregulation of GFAP expression

(astrocytes), IL-1b and phosphorylation of serine 896 of

the NR1 subunit of the NMDA receptor in the trigeminal

interpolaris/caudalis transition zone which is an important

structure for processing orofacial deep as well as cutaneous

nociceptive inputs; these changes can be significantly

reduced by pre-emptive local injection of lidocaine into the

masseter muscle [71]. A recent study has demonstrated that

Janus kinase-signal transducers and activators of the tran-

scription 3 signaling pathway are critical transducers of

astrocyte proliferation (a critical process in reactive as-

trogliosis) and are involved in the maintenance of tactile

allodynia [72].

The proinflammatory cytokines IL-1b, IL-6 and TNFaand reactive oxygen species play a central role in the

pathogenesis of various pain states. Activated astrocytes

can release pro-inflammatory cytokines (e.g., IL-1b),

which powerfully modulate synaptic transmission in the

spinal cord by enhancing excitatory synaptic transmission

via phosphorylation of the NMDA receptor NR-1 subunit

[73] and suppress inhibitory synaptic transmission [74].

Although IL-1b may also be expressed in microglia as well

as astrocytes in some brain areas, this significant role of

astrocytic IL-1b in the spinal cord is supported by a recent

report that an up-regulation of IL-1b occurs in astrocytes in

the medullary dorsal horn by 30 min after Complete Fre-

und’s Adjuvant-induced inflammation in the masseter

muscle [75] and IL-1b up-regulation has been also

observed in the medullary dorsal horn in animals with

hyperalgesia induced by mental nerve transection [76].

Cytokines including IL-1b also exert their actions partially

through the activation of the transcription factor NFjB in

astrocytes, which in turn regulates the transcription of

many inflammatory mediators (including cytokines and

chemokines) thus causing a positive feedback loop [77]. A

recent study using transgenic mice has reported that a

functional inactivation of the NFjB pathway in GFAP-

labelled cells in the spinal cord leads to a reduction in pain

behavior and inflammation produced by sciatic nerve

chronic constriction injury [78].

A short exposure of astrocyte cultures to pro-inflam-

matory TNFa dramatically increases the expression and

release of MCP-1 in a JNK-dependent manner, and

accordingly, i.t. administration of TNFa induces MCP-1

expression in spinal dorsal horn astrocytes in association

with behavioral allodynia that can be reversed by a MCP-1

neutralizing antibody [61]. TNFa enhances spontaneous

release of neurotransmitters from primary afferent fibers by

modulation of tetrodotoxin-sensitive sodium channels

Neurochem Res (2012) 37:2419–2431 2423

123

following sciatic nerve transection [79] and also increases

sEPSC frequency in spinal outer lamina II neurons, an

effect that may involve TRPV1 activation since it can be

abolished in TRPV1 knock-out mice [80]. TNFa is mainly

formed in and released from microglia mediated by phos-

phorylation of p38 MAPK following activation of toll-like

receptors [81] and/or P2X7 and P2Y12 receptors in

microglia upon tissue inflammation and/or nerve injury

[33, 82–85]. Our recent study has also demonstrated that

P2X7 receptor activation is also involved in the central

sensitization in the medullary dorsal horn produced by

mustard oil application to the tooth pulp [86]. In addition,

in primary cultures of spinal cord astrocytes, application of

ATP through its action on astrocytic P2Y1 receptors can

mobilize a release of arachidonic acid and prostaglandin

E2. Inhibition of this prostaglandin E2 release can reduce

behavioral signs of pain after spinal cord injury [87].

Recent studies have particularly emphasized the impor-

tance of chemokines in inflammatory and neuropathic pain

states. The chemokine MCP-1 has been implicated in neu-

ron- and astrocyte-to-microglia signaling. MCP-1 is

released from the central terminals of nociceptors after

peripheral injury. Whereas the activation of microglia sur-

rounding MCP-1-expressing spinal cord dorsal horn neu-

rons peaks by day 14 after a sciatic nerve chronic

constriction injury, astrocyte activation becomes detectable

later, progresses more slowly and also remains increased

until the end of the observation period (150 days) [88],

suggesting that astrocyte activation plays an important role

in maintaining persistent pain states. Recent evidence also

indicates that nerve injury and inflammation activate the

JNK in spinal astrocytes, leading to a substantial increase in

the expression and release of pro-inflammatory chemokines

(e.g., MCP-1) in the spinal dorsal horn which enhance and

prolong persistent pain states. For example, MCP-1 rapidly

induces central sensitization by increasing the activity of

NMDA receptors in dorsal horn neurons [for reviews see

10, 33] and enables T-leukocytes and macrophage migra-

tion into the parenchyma of the CNS, which propagates the

immune response and further induces microglial and

astrocytic immuno-competence [11]. Furthermore, the

matrix metalloproteinase-9-induced pro-IL-1b cleavage

leads to p38 MAPK activation in microglia during the onset

and early stages of neuropathic pain, whereas matrix

metalloproteinase-2-induced pro-IL-1b cleavage leads to

astrocyte activation during the ongoing and later stages of

neuropathic pain [89]. Therefore, spinal and medullary i.t.

administration of inhibitors of the IL-1b, TNFa, JNK,

MCP-1, matrix metalloproteinase-2 or NFjB signaling can

each attenuate processes related to inflammatory, neuro-

pathic pain [90, for more references, see 33].

Microglial activation utilizes two signaling pathways

that also involve astrocytic processes [36]: (1) Through

their pattern recognition receptors, such as toll-like recep-

tors, microglia can detect and differentiate viral, bacterial

and fungal structures and others. Activation of toll-like

receptor-4 in microglia initiates immune-like processes,

such as release of pro-inflammatory cytokines and che-

mokines [32, 33, 48, 91], leading to a positive excitatory

feedback loop in the pain processes and a decrease in

opioid analgesic efficacy as well as phagocytosis and the

release of anti-inflammatory factors e.g., IL-10 [8, 10, 12,

14]. Importantly, selective acute antagonism of the toll-like

receptor-4 results in reversal of neuropathic pain as well as

potentiation of opioid analgesia [14, 92]. Like microglia,

activation of astrocytes also leads to an inflammatory

response producing diverse inflammatory mediators

(TNFa, IL-1b, prostaglandins, NO, and reactive oxygen

species), all of which serve immune surveillance functions.

For instance, hours after lipopolysaccharide application to

astrocytic cultures, a prompt increase in CDK11p58

expression in astrocytes promotes the astrocyte-induced

inflammatory response via p38 MAPK and JNK activation

[93]. (2) Microglia-to-neuron or to-astrocyte interactions

through their constitutive chemokine signaling pathways,

i.e., the ligand-receptor pairs CD200-CD200R, and

CX3CL1 (also termed fractalkine)-CX3CR1 normally

promote a ‘‘calming’’ environment. Any disrupted signal-

ing (‘off’ signaling) due to impairment of neuronal integ-

rity, as might be produced for example by nerve injury,

causes glial activation. For example, after spinal nerve

injury, but not after peripheral inflammation, the activated

microglia in the spinal dorsal horn release cathepsin S that

enzymatically cleaves neuronal fractalkine (CX3CL1); this

in turn activates CX3CR1 receptors on microglia, leading

to a further release of proinflammatory cytokines from

microglia, thus establishing a positive feedback loop which

boosts chronic pain states [10]. The neuron-glia interaction

mechanisms underlying inflammatory and neuropathic pain

are diagrammatically shown in Fig. 1. Further clarification

of neuronal mechanisms common to both inflammatory

pain and neuropathic pain promises to shed light on the

mechanisms underlying the transition from acute pain to

persistent pain [39].

Glucose Metabolism in Astrocytes

Glucose is used as an energy substrate by both neurons and

astrocytes but can generate the carbon skeleton of gluta-

mate only in astrocytes, not in neurons, because neurons

lack the enzyme pyruvate carboxylase that allows net

synthesis of the oxaloacetate, a link in the tricarboxylic

acid cycle (also termed Krebs cycle) in astrocytes. As

outlined by Hertz and Zielke [24], the newly produced

oxaloacetate condenses with acetyl coenzyme A derived

2424 Neurochem Res (2012) 37:2419–2431

123

from pyruvate, to generate citrate, which is then metabo-

lized in the tricarboxylic acid cycle to form a-ketoglutarate,

a direct precursor of glutamate.

Fluoroacetate and its metabolite fluorocitrate are potent

inhibitors of glial cells especially astrocytes at low doses

[94, 95]. Fluoroacetate combines with acetyl coenzyme A

to form fluoroacetyl coenzyme A, which can substitute for

acetyl coenzyme A in the tricarboxylic acid cycle and

reacts with citrate synthase to produce fluorocitrate, a

metabolite which then binds very tightly to aconitase,

thereby halting the astrocytic tricarboxylic cycle [96].

Intrathecal administration of low dose fluorocitrate can

block persistent thermal and mechanical hyperalgesia

produced by zymosan [29] and by formalin [97], and these

striking blocking effects have been repeatedly confirmed in

various pain models, such as acute inflammatory pain

induced by capsaicin, mustard oil, snake and scorpion

venoms, acute incision surgery, tetanic stimulation of

peripheral nerves and glycine disinhibition-induced allo-

dynia [47, 98–103]; chronic inflammatory pain induced by

carrageenan, Complete Freund’s Adjuvant and traumatic

dental occlusion [44, 49, 89, 104–106]; nerve injury and

A B

C

Acute inflammatory pain

Peak: 0.5-24 h; Duration: few days

D

ERK/p38

ATPCytokines

IL-1β

p38

Glu

Chronic inflammatory pain

BDNF

ERK

Peak: day 3-14; Duration: 21-28 days

Trk BRp38IL-1β

Glu

SP

ATPCGRP

ATPCCL2

Cytokines

Mature IL-1β

Pro-IL-1βTNFα

NF B

p38

CatSFKN

GluSP

ERK

ERK/p38

Neuropathic pain (maintenance phase)

Peak: day 7-14; Duration: months

BDNF

ATP

FKN

PGE2

COX-2iNOS

NO

CGRPBDNF

CCL2

Caspase-1

MMP2

ATPCCL2

Cytokines

Mature IL-1β

Pro-IL-1βTNFp38

FKN

GluSP

JNK

ERK

Neuropathic pain (initiation+induction phase)

Peak: day 0.5-3; Duration: within a week

ATP

CCR2

PGE2

COX-2iNOS

NO

CGRPBDNF

CCL2

Nav1.3

P2Y12R

Caspase-1TLR4

Glu

ATPCytokines

IL-1β

Glu

Ca2+ Ca2+P2X7R

P2X4R

IL-1R

ERKEnhanced inputs

p38

P2X3R

Ca2+

ATP

GluSP

CGRP

Ca2+Gln PGE2, NO

IL-1β

ATCA

GSGluR

Ca2+ PGE2, NO

ERKCX3CR1

FKN

IL-1β

Glu

GDNFNGF

LPS, OpioidMMP9

LPS, Opioid

TNFα

MMP2

MMP9 MMP2

LPS

CDK11

NO

Presynaptic Postsynaptic

Astrocyte Microglia

/ERK /ERK

Fig. 1 Diagrams showing some neuron-glia interactions in acute

inflammatory pain (a), chronic inflammatory pain (b), and in the

initiation/induction phase (c) and maintenance phase (d) of neuro-

pathic pain in the spinal or trigeminal dorsal horn. Different colorsrepresent different pathways of individual mediators; the different

thickness of arrows/fonts represents their importance in individual

cases. For figure clarity, some receptors have not been linked with

appropriate agonists or mediators, and peptidergic (e.g., substance P

[SP], calcitonin gene-related peptide [CGRP]) mechanisms are

omitted in the diagrams. A adenosine, ATP adenosine-50-triphosphate,

BDNF brain-derived neurotrophic factor, Caspase-1 an enzyme that

proteolytically cleaves other proteins, such as pro-IL-1b into mature

IL-1b, CatS lysosomal cysteine protease Cathepsin S, CCL2 chemo-

kine (C–C motif) ligand 2, also known as monocyte chemotactic

protein-1 (MCP-1), CCR2 receptor for CCL2, CDK11 cyclin-

dependent kinase 11, COX-2 cyclooxygenase-2, CX3CR1 receptor

for Fractalkine, Cytokine immunomodulating agents including IL-1b,

IL-6 and TNFa, ERK extracellular signal-regulated kinase, FKNchemokine Fractalkine or neurotactin, GDNF glial cell line-derived

neurotrophic factor, Gln glutamine, Glu glutamate, GS glutamine

synthetase, IL interleukin, IL-1R interleukin-1 receptor, iNOS cyto-

kine-inducible nitric oxide synthase (NOS-2), JNK c-Jun N-terminal

kinase, LPS lipopolysaccharides, MMP matrix metalloproteinase,

Nav1.3 voltage-gated sodium channel, type III a-isoform, NGF nerve

growth factor, NFjB nuclear factor-jB is a transcription factor, NOnitric oxide, P2XR purinergic 2X receptor, P2YR purinergic 2Y

receptor, P38 p38-mitogen activated protein kinase (MAPK), PGE2prostaglandin E2, R receptor, TCA tricarboxylic acid, TLRs toll-like

receptors, TNF tumor necrosis factor, TrkBR neurotrophic tyrosine

kinase B receptor or BDNF/NT-3 growth factor receptor

Neurochem Res (2012) 37:2419–2431 2425

123

neuropathic pain [33, 55, 56, 65, 107] as well as devel-

opment of morphine tolerance [108, for review, see 8].

Nonetheless, it should be noted that fluorocitrate or fluo-

roacetate, at an appropriate dose, only blocks the exag-

gerated pain state reflected in central sensitization i.e.,

hypersensitivity of nociceptive neurons, but not the normal

nociceptive processing in CNS. Intracerebral application of

higher doses of fluorocitrate ([1 nmol) or fluoroacetate

([1 mmol in our lab) can severely affect astrocytic ultra-

structure and glio-neuronal interactions due to accumula-

tion of citrates in the brain that may chelate free calcium

ions [95].

The Astrocyte-Neuron ‘‘Glutamate-Glutamine Shuttle’’

A major function of astrocytes is the uptake of extracellular

glutamate (and GABA) in the synaptic region through

astrocytic transporters such as glutamate transporter-1, and

glutamate aspartate transporter with the cotransport of Na?/

H? and counter transport of K? [109]. The cytosolic gluta-

mate can be transformed by glutamine synthetase to gluta-

mine (an important precursor for glutamate and GABA),

which is released and then trafficked through transporters

into glutamatergic neuronal terminals to replenish the glu-

tamate transmitter pool [110]. This glutamate-glutamine

shuttle (or cycle) as noted by Hertz and Zielke [24], is a

crucial mechanism in nociceptive neuronal central sensiti-

zation and is specific to astrocytes. Therefore, any chemicals

that specifically disrupt transformation from glutamate to

glutamine in the glutamate-glutamine shuttle e.g., methio-

nine sulfoximine, a selective inhibitor of glutamine syn-

thetase, or that specifically block glutamatergic presynaptic

uptake of glutamine e.g., methylamino-isobutyric acid, a

selective inhibitor of the neuronal system A transporter of

glutamine, may affect the availability of glutamate and

consequently the production of central sensitization in acute

and chronic inflammatory pain states [43, 53, 111, 112]. In

particular, this transporter inhibitor-induced attenuation of

trigeminal central sensitization can be readily restored by

application of exogenous glutamine [112]. It is also note-

worthy that i.t. administration of either of these 2 inhibitors

does not affect the normal basal nociceptive processing of

the dorsal horn neurons [43, 46, 111, 112].

There is also evidence that reduction in the expression

and glutamate uptake activity of glutamate transporters

plays a crucial role in both the induction and maintenance

of neuropathic pain following peripheral nerve injury [113]

and in taxol-induced hyperalgesia [114]. Similarly, glial

glutamate transporters in astrocytes are down-regulated in

spinal pathological pain models and up-regulation or

functional enhancement of these transporters prevents

pathological pain [115]. For example, propentofylline

exerts a protective action against post-ischemic damage in

the CNS, because it causes a potent dose-dependent

induction of glutamate transporter-1 mRNA and protein in

astrocytes and decreases MCP-1 release from astrocytes

[116]. Conversely, i.t. injection of threo-beta-benzylox-

yaspartate (a glutamate uptake blocker) can produce

spontaneous nociceptive behavior [117].

Astrocytic Ca21 Wave Propagation and Gap Junctions

Another important feature of astrocytes is that glutamate

and ATP can readily elicit astrocytic transients and oscil-

latory waves of cytoplasmic Ca2?, which may propagate to

adjoining and/or distant astrocytes through gap junctions

and hemichannels; this reactive astrocytic network may

release gliotransmitters (glutamate, ATP, D-serine) to

modulate synaptic transmission. The intercellular Ca2?

wave propagation may involve two pathways: (1) the

cytoplasmic diffusion of Ca2?-mobilizing second messen-

ger inositol 1, 4, 5-triphosphate through the gap junction

channels between astrocytes, and/or (2) extracellular dif-

fusion of ATP that is released through Cx-43 and pann-

exin-1 hemichannels and/or vesicular exocytosis [60, 118–

121]. Gap junctions also occur between neurons and

astrocytes that are formed by proteins Cx-30/Cx-43 and

inactivation of these proteins attenuates hippocampal syn-

aptic transmission [62].

Carbenoxolone is a potent blocker of the hemichannel

protein Cx-43 and plasmalemmal channel protein pannex-

in-1 and can produce analgesia in different behavioral pain

models by inhibiting the coupling between astrocytes or

between astrocytes and neurons [118, 120, 122]. In a for-

malin-induced orofacial pain model, the expression of

many heterotypic Cx-43/Cx-32 (astrocyte/neuron) gap

junctions is up-regulated in the medullary dorsal horn

along with nociceptive behavior, and both are significantly

attenuated by carbenoxolone pretreatment [123]. Also, in

an acute inflammatory pain model produced by mustard oil

application to the tooth pulp, medullary superfusion of

carbenoxolone can completely block central sensitization

induced in the medullary dorsal horn [124], suggesting that

spatio-temporal features of central sensitization, such as the

increased mechanoreceptive field size in particular, may

result from the intercellular Ca2? wave-mediated asyn-

chronous excitation of neurons and non-neuronal cells,

mainly involving astrocytic gap junctions and hemichan-

nels [16, 18, 23, 120]. Interestingly, in sciatic inflammatory

neuropathic or chronic constriction injury pain models, a

low dose of carbenoxolone can reverse ‘mirror image’

(contralateral) mechanical allodynia, while leaving ipsi-

lateral mechanical allodynia unaffected, thus suggesting

that gap junctions may be involved in mediating the ‘mirror

2426 Neurochem Res (2012) 37:2419–2431

123

image’ allodynia [125]. In addition, we have observed that

carbenoxolone can equipotently depress bilateral orofacial

allodynia as well as medullary dorsal horn central sensiti-

zation produced by unilateral injury of the infraorbital

nerve [126]. An earlier study has reported that bilateral

allodynia associated with upregulation of GFAP-labelled

satellite glial cells and TNFa expression bilaterally in the

dorsal root ganglia and dorsal horns occurs in an unilateral

nerve injury pain model [127]. These previous findings

indicate that carbenoxolone blocks the maintenance of

neuropathic pain as well as the development of acute

inflammatory pain.

There are also several other interesting findings relevant

to carbenoxolone and gap junctions: (1) In double gene-

knockout mice, a disruption of astrocytic networks

(Cx-30-/- and Cx-43-/-) causes combined pre- and post-

synaptic alterations including enhanced neuronal excit-

ability, release probability and insertion of postsynaptic

AMPA receptors, suggesting that gap junctions may have

inhibitory influences on neurotransmission in normal con-

ditions [62]; (2) Carbenoxolone completely reverses the

inhibitory effects of cromakalim (a KATP channel opener)

on the hyperalgesia and allodynia after sciatic nerve injury

[128]; (3). Carbenoxolone also potently blocks P2X7

receptors [122]. These observations suggest that carben-

oxolone by blocking gap junctions may exert an excitatory

effect under normal conditions but may depress neuro-

transmission in some pain states but confirmation of this

interpretation requires further investigations.

Astrocyte-Neuron ‘‘Serine Shuttle’’

Allodynia and hyperalgesia rely on activation of the

NMDA receptor, a specific subtype of the ionotropic glu-

tamate receptor, which is composed of a voltage-dependent

ion channel non-selective to cations normally blocked by

magnesium, and other (allosteric, glycine etc.) binding

sites. Astrocytes may regulate NMDA receptor activation

through their specific release of D-serine (a co-agonist of

the NMDA receptor), which acts on the strychnine-insen-

sitive glycine binding site on the NMDA receptor [129].

For instance, D-serine is necessary for the acquisition of

specific pain-related negative emotion involving the rostral

anterior cingulate cortex but is not involved in formalin-

induced acute nociceptive behaviors and electric foot

shock-induced conditioned place avoidance [130]. In an

intrathecally strychnine-induced orofacial allodynia model,

fluorocitrate but not minocycline blocks allodynia as well

as reduces the over-expression of GFAP (but not OX-42)

and c-fos in the superficial layers of the medullary dorsal

horn. Intrathecal D-amino acid oxidase, which selectively

induces degradation of D-serine, blocks the allodynia which

can be restored by intrathecal exogenous D-serine; inter-

estingly, this exogenous D-serine also restores the allodynia

blocked by fluorocitrate [131]. Other studies have provided

evidence that in normal rats, intrathecal D-serine signifi-

cantly enhances the C-fiber (but not Ab-fiber) and Ad-fiber

evoked responses in nociceptive wide dynamic range

neurons in the spinal dorsal horn, and co-administering

D-serine with the glycine binding site antagonist 7-chloro-

kynurenic acid completely blocks the facilitation of

D-serine on C-fiber evoked responses [132]. However, a

recent study showed that a D-amino acid oxidase inhibitor

produces analgesia via blockade of spinal hydrogen per-

oxide production rather than by interacting with spinal

D-serine [133].

Concluding Perspectives

Recent studies provide evidence that by means of Ca2?

signaling and gliotransmitters and/or neuroactive mole-

cules, activated astrocytes are involved in key functions of

the brain under physiological conditions, such as astrocyte-

neuron lactate transport which is required for long-term

memory formation [134] and the control of breathing

through their pH-dependent release of ATP [135]; the

physiological functions of astrocytic Ca2? signaling and

gliotransmitters have been recently reviewed [19, 20, 30,

136]. In the case of pain mechanisms, the importance of

astrocytic Ca2? signaling in normal nociceptive processing

remains unclear. For instance, why does a temporary dis-

ruption of the astrocytic metabolic cycle and/or the gen-

eration of astrocytic [Ca2?]i waves only attenuate the

exaggerated pain component but does not affect basal

nociceptive neuronal responses to noxious stimuli and

animal’s nocifensive behavior? Does the mechanism

underlying the initiation, location, duration and integration

of gliotransmitters of astrocytes that have been demon-

strated in vitro also operate in in vivo states? What factors

cause the discrepant results between wild and gene-

knockout rodents? Additional studies of other astrocytic

mechanisms such as those involving the astrocyte-neuron

lactate transport, ion/water balance [137–139], and neuro-

vascular coupling [17, 140] may provide a clearer under-

standing of the role of astrocytes and other non-neural

processes in acute, chronic inflammatory and neuropathic

pain conditions.

Virtually all pharmacological agents developed to alle-

viate acute or chronic pain have targeted neuronal processes

underlying pain. Since astrocytes as well as microglia can

each modulate the central nociceptive mechanisms in

pathological pain states, they may provide new targets for

the development of novel drugs to control pain. However,

this will be especially challenging since glia originate from

Neurochem Res (2012) 37:2419–2431 2427

123

the immune system which is ubiquitously distributed in the

whole body and participates in many functions. Thus side-

effects of glial inhibitors specifically developed to coun-

teract pain may be difficult to avoid and thus specifically

targeting gene therapy-based approaches may be an

appropriate or even obligatory choice [141].

Acknowledgments This work was supported by the NIH Grant DE-

04786 to B.J.S. and CIHR Grant MOP-82831 to J.O.D.

References

1. Perl ER (2007) Ideas about pain, a historical view. Nat Rev

Neurosci 8:71–80

2. Hucho T, Levine JD (2007) Signaling pathways in sensitization:

toward a nociceptor cell biology. Neuron 55:365–376

3. Ren K, Dubner R (1999) Central nervous system plasticity and

persistent pain. J Orofac Pain 13:155–163

4. Dostrovsky JO (2000) Role of thalamus in pain. In: Sandkuhler

J, Bromm B, Gebhart GF (eds) Nervous system plasticity

and chronic pain. Prog Brain Res 129:245–257 (Elsevier,

Amsterdam)

5. Sessle BJ (2005) Trigeminal central sensitization. Rev Analg

8:85–102

6. Ji RR, Kohno T, Moore KA, Woolf CJ (2003) Central sensiti-

zation and LTP: do pain and memory share similar mechanisms?

Trends Neurosci 26:696–705

7. Woolf CJ, Salter MW (2006) Plasticity and pain: role of the

dorsal horn. In: McMahon SB, Koltzenburg M (eds) Wall and

melzack’s textbook of pain, chapter 5, 5th edn. Elsevier, Lon-

don, pp 91–105

8. Milligan ED, Watkins LR (2009) Pathological and protective

roles of glia in chronic pain. Nat Rev Neurosci 10:23–36

9. Scholz J, Woolf CJ (2007) The neuropathic pain triad: neurons,

immune cells and glia. Nat Neurosci 10:1361–1368

10. Abbadie C, Bhangoo S, De Koninck Y, Malcangio M, Melik-

Parsadaniantz S, White FA (2009) Chemokines and pain

mechanisms. Brain Res Rev 60:125–134

11. Romero-Sandoval EA, Horvath RJ, DeLeo JA (2008) Neuro-

immune interactions and pain: focus on glial-modulating targets.

Curr Opin Investig Drugs 9:726–734

12. Buchanan MM, Hutchinson M, Watkins LR, Yin H (2010) Toll-

like receptor 4 in CNS pathologies. J Neurochem 114:13–27

13. Ren K, Dubner R (2010) Interactions between the immune and

nervous systems in pain. Nat Med 16:1267–1276

14. Hutchinson MR, Shavit Y, Grace PM, Rice KC, Maier SF,

Watkins LR (2011) Exploring the neuroimmunopharmacology

of opioids: an integrative review of mechanisms of central

immune signaling and their implications for opioid analgesia.

Pharmacol Rev 63:772–810

15. Volterra A, Meldolesi J (2005) Astrocytes, from brain glue to

communication elements: the revolution continues. Nat Rev

Neurosci 6:626–640

16. Haydon PG, Carmignoto G (2006) Astrocyte control of synaptic

transmission and neurovascular coupling. Physiol Rev 86:1009–

1031

17. Iadecola C, Nedergaard M (2007) Glial regulation of the cere-

bral microvasculature. Nat Neurosci 10:1369–1376

18. Fellin T (2009) Communication between neurons and astro-

cytes: relevance to the modulation of synaptic and network

activity. J Neurochem 108:533–544

19. Fiacco TA, Agulhon C, McCarthy KD (2009) Sorting out

astrocyte physiology from pharmacology. Annu Rev Pharmacol

Toxicol 49:151–174

20. Hamilton NB, Attwell D (2010) Do astrocytes really exocytose

neurotransmitters? Nat Rev Neurosci 11:227–238

21. Attwell D, Buchan AM, Charpak S, Lauritzen M, Macvicar BA,

Newman EA (2010) Glial and neuronal control of brain blood

flow. Nature 468:232–243

22. Halassa MM, Haydon PG (2010) Integrated brain circuits:

astrocytic networks modulate neuronal activity and behavior.

Annu Rev Physiol 72:335–355

23. Giaume C, Koulakoff A, Roux L, Holcman D, Rouach N (2010)

Astroglial networks: a step further in neuroglial and gliovascular

interactions. Nat Rev Neurosci 11:87–99

24. Hertz L, Zielke HR (2004) Astrocytic control of glutamatergic

activity: astrocytes as stars of the show. Trends Neurosci

27:735–743

25. Hertz L, Peng L, Dienel GA (2007) Energy metabolism in

astrocytes: high rate of oxidative metabolism and spatiotemporal

dependence on glycolysis/glycogenolysis. J Cereb Blood Flow

Metab 27:219–249

26. Garrison CJ, Dougherty PM, Kajander KC, Carlton SM (1991)

Staining of glial fibrillary acidic protein (GFAP) in lumbar

spinal cord increases following a sciatic nerve constriction

injury. Brain Res 565:1–7

27. Garrison CJ, Dougherty PM, Carlton SM (1994) GFAP

expression in lumbar spinal cord of naive and neuropathic rats

treated with MK-801. Exp Neurol 129:237–243

28. Hassel B, Paulsen RE, Johnson A, Fonnum F (1992) Selective

inhibition of glial cell metabolism by fluorocitrate. Brain Res

249:120–124

29. Meller ST, Dykstra C, Grzybycki D, Murphy S, Gebhart GF

(1994) The possible role of glia in nociceptive processing and

hyperalgesia in the spinal cord of the rat. Neuropharmacology

33:1471–1478

30. Calı C, Marchaland J, Spagnuolo P, Gremion J, Bezzi P (2009)

Regulated exocytosis from astrocytes physiological and patho-

logical related aspects. Int Rev Neurobiol 85:261–293

31. McMahon SB, Malcongio M (2009) Current challenges in glia-

pain biology. Neuron 64:46–54

32. Ji RR, Gereau RW 4th, Malcangio M, Strichartz GR (2009)

MAP kinase and pain. Brain Res Rev 60:135–148

33. Gao YJ, Ji RR (2010) Chemokines, neuronal-glial interactions,

and central processing of neuropathic pain. Pharmacol Ther

126:56–68

34. Parpura V, Zorec R (2010) Gliotransmission: exocytotic release

from astrocytes. Brain Res Rev 63:83–92

35. Watkins LR, Milligan ED, Maier SF (2001) Glial activation:

a driving force for pathological pain. Trends Neurosci 24:

450–455

36. Hanisch UK, Kettenmann H (2007) Microglia: active sensor and

versatile effector cells in the normal and pathologic brain. Nat

Neurosci 10:1387–1394

37. Inoue K, Tsuda M (2009) Microglia and neuropathic pain. Glia

57:1469–1479

38. Jarvis MF (2009) The neural–glial purinergic receptor ensemble

in chronic pain states. Trends Neurosci 33:48–57

39. Xu Q, Yaksh TL (2011) A brief comparison of the pathophys-

iology of inflammatory versus neuropathic pain. Curr Opin

Anaesthesiol 24:400–407

40. Hanani M (2005) Satellite glial cells in sensory ganglia: from

form to function. Brain Res Brain Res Rev 48:457–476

41. Ohara PT, Vit J-P, Bhargava A, Romero M, Sundberg C,

Charles A, Jasmin L (2009) Gliopathic pain: when satellite glial

cells go bad. Neuroscientist 15:450–463

2428 Neurochem Res (2012) 37:2419–2431

123

42. Takeda M, Takahashi M, Matsumoto S (2009) Contribution of

the activation of satellite glia in sensory ganglia to pathological

pain. Neurosci Biobehav Rev 33:784–792

43. Chiang CY, Dostrovsky JO, Iwata K, Sessle BJ (2011) Role of

glia in orofacial pain. Neuroscientist 17:303–320

44. Clark AK, Gentry C, Bradbury EJ, McMahon SB, Malcangio M

(2007) Role of spinal microglia in rat models of peripheral nerve

injury and inflammation. Eur J Pain 11:223–230

45. Qin M, Wang JJ, Cao R, Zhang H, Duan L, Gao B, Xiong YF,

Chen LW, Rao ZR (2006) The lumbar spinal cord glial cells

actively modulate subcutaneous formalin induced hyperalgesia

in the rat. Neurosci Res 55:442–450

46. Chiang CY, Zhang S, Xie YF, Hu JW, Dostrovsky JO, Sessle BJ

(2005) Endogenous ATP involvement in mustard oil-induced

central sensitization in trigeminal subnucleus caudalis (medul-

lary dorsal horn). J Neurophysiol 94:1751–1760

47. Xie YF, Zhang S, Chiang CY, Hu JW, Dostrovsky JO, Sessle BJ

(2007) Involvement of glia in central sensitization in trigeminal

subnucleus caudalis (medullary dorsal horn). Brain Behav Im-

mun 21:634–641

48. Romero-Sandoval A, Chai N, Nutile-McMenemy N, Deleo JA

(2008) A comparison of spinal Iba1 and GFAP expression in

rodent models of acute and chronic pain. Brain Res 1219:116–

126

49. Shimizu K, Guo W, Wang H, Lagraize SC, Zou S, Iwata K, Wei

F, Dubner R, Ren K (2009) Differential involvement of tri-

geminal transition zone and laminated subnucleus caudalis in

orofacial deep and cutaneous hyperalgesia: the effects of inter-

leukin-10 and glial inhibitors. Mol Pain 5:75

50. Raghavendra V, Tanga FY, DeLeo JA (2004) Complete Freunds

adjuvant-induced peripheral inflammation evokes glial activa-

tion and proinflammatory cytokine expression in the CNS. Eur J

Neurosci 20:467–473

51. Lee MK, Han SR, Park MK, Kim MJ, Bae YC, Kim SK, Park

JS, Ahn DK (2011) Behavioral evidence for the differential

regulation of p-p38 MAPK and p-NF-kappa B in rats with tri-

geminal neuropathic pain. Mol Pain 7:57

52. Cao Y, Li K, Fu K, Xie Q (2009) Activation of mitogen-acti-

vated protein kinases in Vsp following occlusal interference.

IADR Pan Asian Pac Fed (PAPF) and IADR Asia/Pac Reg

(APR) Sept 22–24 No 521

53. Tsuboi Y, Iwata K, Dostrovsky JO, Chiang CY, Sessle BJ, Hu

JW (2011) Modulation of astroglial glutamine synthetase

activity affects nociceptive behaviour and central sensitization

of medullary dorsal horn nociceptive neurons in a rat model of

chronic pulpitis. Eur J Neurosci 34:292–302

54. Tanga FY, Raghavendra V, Nutile-McMenemy N, Marks A,

Deleo JA (2006) Role of astrocytic S100beta in behavioral

hypersensitivity in rodent models of neuropathic pain. Neuro-

science 140:1003–1010

55. Okada-Ogawa A, Suzuki I, Sessle BJ, Chiang CY, Salter MW,

Dostrovsky JO, Tsuboi Y, Kondo M, Kitagawa J, Kobayashi A,

Noma N, Imamura Y, Iwata K (2009) Astroglia in medullary

dorsal horn (trigeminal spinal subnucleus caudalis) are involved

in trigeminal neuropathic pain mechanisms. J Neurosci

29:11161–11171

56. Kobayashi A, Shinoda M, Sessle BJ, Honda K, Imamura Y,

Hitomi S, Tsuboi Y, Okada-Ogawa A, Iwata K (2011) Mecha-

nisms involved in extraterritorial facial pain following cervical

spinal nerve injury in rats. Mol Pain 7:12

57. Vega-Avelaira D, Moss A, Fitzgerald M (2007) Age-related

changes in the spinal cord microglial and astrocytic response

profile to nerve injury. Brain Behav Immun 21:617–623

58. Latremoliere A, Mauborgne A, Masson J, Bourgoin S, Kayser

V, Hamon M, Pohl M (2008) Differential implication of pro-

inflammatory cytokine interleukin-6 in the development of

cephalic versus extracephalic neuropathic pain in rats. J Neuro-

sci 28:8489–8501

59. Milligan ED, Mehmert KK, Hinde JL, Harvey LO, Martin D,

Tracey KJ, Maier SF, Watkins LR (2000) Thermal hyperalgesia

and mechanical allodynia produced by intrathecal administra-

tion of the human immunodeficiency virus-1 (HIV-1) envelope

glycoprotein, gp120. Brain Res 861:105–116

60. Bardoni R, Ghirri A, Zonta M, Betelli C, Vitale G, Ruggieri V,

Sandrini M, Carmignoto G (2010) Glutamate-mediated astro-

cyte-to-neuron signalling in the rat dorsal horn. J Physiol

588:831–846

61. Gao YJ, Zhang L, Ji RR (2010) Spinal injection of TNF-a-

activated astrocytes produces persistent pain symptom

mechanical allodynia by releasing monocyte chemoattractant

protein-1. Glia 58:1871–1880

62. Pannasch U, Vargova L, Reingruber J, Ezan P, Holcman D,

Giaume C, Sykova E, Rouach N (2011) Astroglial networks

scale synaptic activity and plasticity. Proc Natl Acad Sci USA

108:8467–8472

63. DeLeo JA, Rutkowski MD, Stalder AK, Campbell IL (2000)

Transgenic expression of TNF by astrocytes increases mechan-

ical allodynia in a mouse neuropathy model. NeuroReport

11:599–602

64. Tsuda M, Shigemoto-Mogami Y, Koizumi S, Mizokoshi A,

Kohsaka S, Salter MW, Inoue K (2003) P2X4 receptors induced

in spinal microglia gate tactile allodynia after nerve injury.

Nature 424:778–783

65. Clark AK, Yip PK, Grist J, Gentry C, Staniland AA, Marchand

F, Dehvari M, Wotherspoon G, Winter J, Ullah J, Bevan S,

Malcangio M (2007) Inhibition of spinal microglial cathepsin S

for the reversal of neuropathic pain. Proc. Natl Acad Sci USA

104:10655–10660

66. Staniland AA, Clark AK, Wodarski R, Sasso O, Maione F,

D’Acquisto F, Malcangio M (2010) Reduced inflammatory and

neuropathic pain and decreased spinal microglial response in

fractalkine receptor (CX3CR1) knockout mice. J Neurochem

114:1143–1157

67. Zheng W, Ouyang H, Zheng X, Liu S, Mata M, Fink DJ, Hao S

(2011) Glial TNFa in the spinal cord regulates neuropathic pain

induced by HIV gp120 application in rats. Mol Pain 7:40

68. Clark AK, Staniland AA, Marchand F, Kaan TK, McMahon SB,

Malcangio M (2010) P2X7-dependent release of interleukin-

1beta and nociception in the spinal cord following lipopoly-

saccharide. J Neurosci 30:573–582

69. Nakajima A, Tsuboi Y, Suzuki I, Honda K, Shinoda M, Kondo

M, Matsuura S, Shibuta K, Yasuda M, Shimizu N, Iwata K

(2011) PKCgamma in Vc and C1/C2 is involved in trigeminal

neuropathic pain. J Dent Res 90:777–781

70. Wang W, Mei XP, Wei YY, Zhang MM, Zhang T, Wang W, Xu

LX, Wu SX, Li YQ (2011) Neuronal NR2B-containing NMDA

receptor mediates spinal astrocytic c-Jun N-terminal kinase

activation in a rat model of neuropathic pain. Brain Behav Im-

mun 25:1355–1366

71. Wang H, Guo W, Yang K, Wei F, Dubner R (2010) Contribution

of primary afferent input to trigeminal astroglial hyperactivity,

cytokine induction and NMDA receptor phosphorylation. Open

Pain J 3:144–152

72. Tsuda M, Kohro Y, Yano T, Tsujikawa T, Kitano J, Tozaki-

Saitoh H, Koyanagi S, Ohdo S, Ji RR, Salter MW, Inoue K

(2011) JAK-STAT3 pathway regulates spinal astrocyte prolif-

eration and neuropathic pain maintenance in rats. Brain 134(Pt

4):1127–1139

73. Zhang RX, Li A, Liu B, Wang L, Ren K, Zhang H, Berman BM,

Lao L (2008) IL-1ra alleviates inflammatory hyperalgesia

through preventing phosphorylation of NMDA receptor NR-1

subunit in rats. Pain 135:232–239

Neurochem Res (2012) 37:2419–2431 2429

123

74. Gustafson-Vickers SL, Lu VB, Lai AY, Todd KG, Ballanyi K,

Smith PA (2008) Long-term actions of interleukin-1beta on

delay and tonic firing neurons in rat superficial dorsal horn and

their relevance to central sensitization. Mol Pain 4:63

75. Guo W, Wang H, Watanabe M, Shimizu K, Zou S, LaGraize

SC, Wei F, Dubner R, Ren K (2007) Glial-cytokine-neuronal

interactions underlying the mechanisms of persistent pain.

J Neurosci 27:6006–6018

76. Takahashi K, Watanabe M, Suekawa Y, Ito G, Inubushi T,

Hirose N, Murasaki K, Hiyama S, Uchida T, Tanne K (2011) IL-

1beta in the trigeminal subnucleus caudalis contributes to extra-

territorial allodynia/hyperalgesia following a trigeminal nerve

injury. Eur J Pain 15:467.e1–467.e14

77. Ledeboer A, Gamanos M, Lai W, Martin D, Maier SF, Watkins

LR, Quan N (2005) Involvement of spinal cord nuclear factor

kappaB activation in rat models of proinflammatory cytokine-

mediated pain facilitation. Eur J Neurosci 22:1977–1986

78. Fu ES, Zhang YP, Sagen J, Candiotti KA, Morton PD, Liebl DJ,

Bethea JR, Brambilla R (2010) Transgenic inhibition of glial

NF-kappa B reduces pain behavior and inflammation after

peripheral nerve injury. Pain 148:509–518

79. Spicarova D, Nerandzic V, Palecek J (2011) Modulation of

spinal cord synaptic activity by tumor necrosis factor alpha in a

model of peripheral neuropathy. J Neuroinflammation 8:177

80. Park CK, Lu N, Xu ZZ, Liu T, Serhan CN, Ji RR (2011)

Resolving TRPV1- and TNF-a-mediated spinal cord synaptic

plasticity and inflammatory pain with neuroprotectin D1.

J Neurosci 31(42):15072–15085

81. Tanga FY, Nutile-McMenemy N, DeLeo JA (2005) The CNS

role of Toll-like receptor 4 in innate neuroimmunity and painful

neuropathy. Proc Natl Acad Sci USA 102:5856–5861

82. Kobayashi K, Yamanaka H, Fukuoka T, Dai Y, Obata K,

Noguchi K (2008) P2Y12 receptor upregulation in activated

microglia is a gateway of p38 signaling and neuropathic pain.

J Neurosci 28:2892–2902

83. Tozaki-Saitoh H, Tsuda M, Miyata H, Ueda K, Kohsaka S,

Inoue K (2008) P2Y12 receptors in spinal microglia are required

for neuropathic pain after peripheral nerve injury. J Neurosci

28:4949–4956

84. Surprenant A, North RA (2009) Signaling at purinergic P2X

receptors. Annu Rev Physiol 71:333–359

85. Trang T, Beggs S, Salter MW (2011) ATP receptors gate

microglia signaling in neuropathic pain. Exp Neurol. doi:

10.1016/j.expneurol.2011.11.012

86. Itoh K, Chiang CY, Li Z, Lee JC, Dostrovsky JO, Sessle BJ

(2011) Central sensitization of nociceptive neurons in rat med-

ullary dorsal horn involves purinergic P2X7 receptors. Neuro-

science 192:721–731

87. Xia M, Zhu Y (2011) Signaling pathways of ATP-induced PGE2

release in spinal cord astrocytes are EGFR transactivation-

dependent. Glia 59:664–674

88. Zhang J, De Koninck Y (2006) Spatial and temporal relationship

between monocyte chemoattractant protein-1 expression and

spinal glial activation following peripheral nerve injury. J Neu-

rochem 97:772–783

89. Kawasaki Y, Xu ZZ, Wang X, Park JY, Zhuang ZY, Tan PH,

Gao YJ, Roy K, Corfas G, Lo EH, Ji RR (2008) Distinct roles of

matrix metalloproteases in the early- and late-phase develop-

ment of neuropathic pain. Nat Med 14:331–336

90. Zheng W, Ouyang H, Zheng X, Liu S, Mata M, Fink DJ, Hao S

(2011) Glial TNFa in the spinal cord regulates neuropathic pain

induced by HIV gp120 application in rats. Mol Pain 20(7):40

91. Gao YJ, Ji RR (2010) Targeting astrocyte signaling for chronic

pain. Neurotherapeutics 7:482–493

92. Hutchinson MR, Bland ST, Johnson KW, Rice KC, Maier SF,

Watkins LR (2007) Opioid-induced glial activation:

mechanisms of activation and implications for opioid analgesia,

dependence, and reward. ScientificWorldJournal 7:98–111

93. Liu X, Cheng C, Shao B, Wu X, Ji Y, Liu Y, Lu X, Shen A

(2011) CDK11(p58) promotes rat astrocyte inflammatory

response via activating p38 and JNK pathways induced by

lipopolysaccharide. Neurochem Res. doi:

10.1007/s11064-011-0643-7

94. Wei F, Guo W, Zou S, Ren K, Dubner R (2008) Supraspinal

glial-neuronal interactions contribute to descending pain facili-

tation. J Neurosci 28:10482–10495

95. Fonnum F, Johnsen A, Hassel B (1997) Use of fluorocitrate and

fluoroacetate in the study of brain metabolism. Glia 21:106–113

96. Proudfoot AT, Bradberry SM, Vale JA (2006) Sodium fluoro-

acetate poisoning. Toxicol Rev 25:213–219

97. Watkins LR, Martin D, Ulrich P, Tracey KJ, Maier SF (1997)

Evidence for the involvement of spinal cord glia in subcutane-

ous formalin induced hyperalgesia in the rat. Pain 71:225–235

98. Obata H, Eisenach JC, Hussain H, Bynum T, Vincler M (2006)

Spinal glial activation contributes to postoperative mechanical

hypersensitivity in the rat. J Pain 7:816–822

99. Chen Y, Willcockson HH, Valtschanoff JG (2009) Influence of

the vanilloid receptor TRPV1 on the activation of spinal cord

glia in mouse models of pain. Exp Neurol 220:383–390

100. Chacur M, Gutierrez JM, Milligan ED, Wieseler-Frank J, Britto

LR, Maier SF, Watkins LR, Cury Y (2004) Snake venom

components enhance pain upon subcutaneous injection: an ini-

tial examination of spinal cord mediators. Pain 111:65–76

101. Jiang F, Liu T, Cheng M, Pang XY, Bai ZT, Zhou JJ, Ji YH

(2009) Spinal astrocyte and microglial activation contributes to

rat pain-related behaviors induced by the venom of scorpion

Buthus martensi Karch. Eur J Pharmacol 623:52–64

102. Ying B, Lu N, Zhang YQ, Zhao ZQ (2006) Involvement of

spinal glia in tetanically sciatic stimulation-induced bilateral

mechanical allodynia in rats. Biochem Biophys Res Commun

340:1264–1272

103. Miraucourt LS, Peirs C, Dallel R, Voisin DL (2011) Glycine

inhibitory dysfunction turns touch into pain through astrocyte-

derived D-serine. Pain 152:1340–1348

104. Roberts J, Ossipov MH, Porreca F (2009) Glial activation in the

rostroventromedial medulla promotes descending facilitation to

mediate inflammatory hypersensitivity. Eur J Neurosci

30:229–241

105. Sun S, Cao H, Han M, Li TT, Zhao ZQ, Zhang YQ (2008)

Evidence for suppression of electroacupuncture on spinal glial

activation and behavioral hypersensitivity in a rat model of

monoarthritis. Brain Res Bull 75:83–93

106. Chen J, Zhang J, Zhao Y, Yuan L, Nie X, Li J, Ma Z, Zhang Y,

Wang Q, Chen Y, Jin Y, Rao Z (2007) Hyperalgesia in response

to traumatic occlusion and GFAP expression in rat parabrachial

nucleus: modulation with fluorocitrate. Cell Tissue Res

329:231–237

107. Milligan ED, Twining C, Chacur M, Biedenkapp J, O’Connor K,

Poole S, Tracey K, Martin D, Maier SF, Watkins LR (2003)

Spinal glia and proinflammatory cytokines mediate mirror-

image neuropathic pain in rats. J Neurosci 23:1026–1040

108. Song P, Zhao ZQ (2001) The involvement of glial cells in the

development of morphine tolerance. Neurosci Res 39:281–286

109. Danbolt NC (2001) Glutamate uptake. Progr Neurobiol

65:1–105

110. Kanamori K, Ross BD (2006) Kinetics of glial glutamine efflux

and the mechanism of neuronal uptake studied in vivo in mildly

hyperammonemic rat brain. J Neurochem 99:1103–1113

111. Chiang CY, Wang J, Xie YF, Zhang S, Hu JW, Dostrovsky JO,

Sessle BJ (2007) Astroglial glutamate-glutamine shuttle is

involved in central sensitization of nociceptive neurons in rat

medullary dorsal horn. J Neurosci 27:9068–9076

2430 Neurochem Res (2012) 37:2419–2431

123

112. Chiang CY, Li Z, Dostrovsky JO, Hu JW, Sessle BJ (2008)

Glutamine uptake contributes to central sensitization in the

medullary dorsal horn. NeuroReport 18:1151–1154

113. Sung B, Lim G, Mao J (2003) Altered expression and uptake

activity of spinal glutamate transporters following peripheral

nerve injury contributes to the pathogenesis of neuropathic pain

in rats. J Neurosci 23:2899–2910

114. Weng HR, Aravindan N, Cata JP, Chen JH, Shaw AD,

Dougherty PM (2005) Spinal glial glutamate transporters

downregulate in rats with taxol-induced hyperalgesia. Neurosci

Lett 386:18–22

115. Nakagawa T, Kaneko S (2010) Spinal astrocytes as therapeutic

targets for pathological pain. J Pharmacol Sci 114:347–353

116. Sweitzer S, De Leo JA (2011) Propentofylline: glial modulation,

neuroprotection, and alleviation of chronic pain. Handb Exp

Pharmacol 200:235–250

117. Liaw WJ, Stephens RL Jr, Binns BC, Chu Y, Sepkuty JP, Johns

RA, Rothstein JD, Tao YX (2005) Spinal glutamate uptake is

critical for maintaining normal sensory transmission in rat spinal

cord. Pain 115:60–70

118. Scemes E, Giaume C (2006) Astrocyte calcium waves: what

they are and what they do. Glia 54:716–725

119. Kang J, Kang N, Lovatt D, Torres A, Zhao Z, Lin J, Nedergaard

M (2008) Connexin 43 hemichannels are permeable to ATP.

J Neurosci 28:4702–4711

120. Scemes E, Spray DC, Meda P (2009) Connexins, Pannexins,

innexins: novel roles of ‘‘hemi-channels’’. Pflugers Arch

457:1207–1226

121. Nagy JI, Dudek FE, Rash JE (2004) Update on connexins and

gap junctions in neurons and glia in the mammalian nervous

system. Brain Res Brain Res Rev 47:191–215

122. Suadicani SO, Brosnan CF, Scemes E (2006) P2X7 receptors

mediate ATP release and amplification of astrocytic intercellular

Ca2? signaling. J Neurosci 26:1378–1385

123. Lan L, Yuan H, Duan L, Cao R, Gao B, Shen J, Xiong Y, Chen

LW, Rao ZR (2007) Blocking the glial function suppresses

subcutaneous formalin-induced nociceptive behavior in the rat.

Neurosci Res 57:112–119

124. Chiang CY, Li Z, Dostrovsky JO, Sessle BJ (2010) Central

sensitization in medullary dorsal horn involves gap junctions

and hemichannels. NeuroReport 21:233–237

125. Spataro LE, Sloane EM, Milligan ED, Wieseler-Frank J, Scho-

eniger D, Jekich BM, Barrientos RM, Maier SF, Watkins LR

(2004) Spinal gap junctions: potential involvement in pain

facilitation. J Pain 5:392–405

126. Wang H, Cao Y, Chiang CY, Dostrovsky JO, Sessle BJ (2011)

Carbenoxolone attenuates neuropathic pain behaviour and med-

ullary dorsal horn central sensitization associated with partial

infraorbital nerve transection in rats. Am Neurosci Abstr 73.04

127. Hatashita S, Sekiguchi M, Kobayashi H, Konno S, Kikuchi S

(2008) Contralateral neuropathic pain and neuropathology in

dorsal root ganglion and spinal cord following hemilateral nerve

injury in rats. Spine 33:1344–1351

128. Wu XF, Liu WT, Liu YP, Huang ZJ, Zhang YK, Song XJ (2011)

Reopening of ATP-sensitive potassium channels reduces neu-

ropathic pain and regulates astroglial gap junctions in the rat

spinal cord. Pain 152:2605–2615

129. Oliet SH, Mothet JP (2009) Regulation of N-methyl-D-aspartate

receptors by astrocytic D-serine. Neuroscience 158:275–283

130. Ren WH, Guo JD, Cao H, Wang H, Wang PF, Sha H, Ji RR,

Zhao ZQ, Zhang YQ (2006) Is endogenous D-serine in the

rostral anterior cingulate cortex necessary for pain-related neg-

ative affect? J Neurochem 96:1636–1647

131. Miraucourt LS, Peirs C, Dallel R, Voisin DL (2011) Glycine

inhibitory dysfunction turns touch into pain through astrocyte-

derived D-serine. Pain 152:1340–1348

132. Guo JD, Wang H, Zhang YQ, Zhao ZQ (2006) Distinct effects

of D-serine on spinal nociceptive responses in normal and car-

rageenan-injected rats. Biochem Biophys Res Commun 343:

401–406

133. Lu JM, Gong N, Wang YC, Wang YX (2011) DAAO-mediated

spinal hydrogen peroxide is specifically and largely responsible

for formalin-induced central sensitization-involved pain. Br J

Pharmacol doi. doi:10.1111/j.1476-5381.2011.01680.x

134. Suzuki A, Stern SA, Bozdagi O, Huntley GW, Walker RH,

Magistretti PJ, Alberini CM (2011) Astrocyte-neuron lactate

transport is required for long-term memory formation. Cell

144:810–823

135. Gourine AV, Kasymov V, Marina N, Tang F, Figueiredo MF,

Lane S, Teschemacher AG, Spyer KM, Deisseroth K, Kasparov

S (2010) Astrocytes control breathing through pH-dependent

release of ATP. Science 329:571–575

136. Clarke LE, Attwell D (2012) An astrocyte TRP switch for

inhibition. Nat Neurosci 15:3–4

137. Nesic O, Lee J, Johnson KM, Ye Z, Xu GY, Unabia GC, Wood

TG, McAdoo DJ, Westlund KN, Hulsebosch CE, Regino Perez-

Polo J (2005) Transcriptional profiling of spinal cord injury-

induced central neuropathic pain. J Neurochem 95:998–1014

138. Nesic O, Guest JD, Zivadinovic D, Narayana PA, Herrera JJ,

Grill RJ, Mokkapati VU, Gelman BB, Lee J (2010) Aquaporins

in spinal cord injury: the janus face of aquaporin 4. Neurosci-

ence 168:1019–1035

139. Mulligan SJ, MacVicar BA (2006) VRACs CARVe a path for

novel mechanisms of communication in the CNS. Sci STKE

357:pe42

140. Seo HS, Kim HW, Roh DH, Yoon SY, Kwon YB, Han HJ,

Chung JM, Beitz AJ, Lee JH (2008) A new rat model for

thrombus-induced ischemic pain (TIIP); development of bilat-

eral mechanical allodynia. Pain 139:520–532

141. Heinzmann S, McMahon SB (2011) New molecules for the

treatment of pain. Curr Opin Support Palliat Care 5:111–115

142. Hertz L, Lovatt D, Goldman SA, Nedergaard M (2010)

Adrenoceptors in brain: cellular gene expression and effects on

astrocytic metabolism and [Ca(2?)]i. Neurochem Int 57:

411–420

143. Pelegrin P, Surprenant A (2006) Pannexin-1 mediate large pore

formation and interleukin-1beta release by the ATP-gated P2X7

receptor. EMBO J 25:5071–5082

144. Pezet S, McMahon SB (2006) Neurotrophins: mediators and

modulators of pain. Annu Rev Neurosci 29:507–538

145. Sawynok J (2006) Adenosine and ATP receptors. Handb Exp

Pharmacol 177:309–328

146. Burnstock G, Fredholm BB, Verkhratsky A (2011) Adenosine and

ATP receptors in the brain. Curr Top Med Chem 11:973–1011

147. Choi I-S, Cho J-H, Jang I-S (2011) A1 receptors inhibit gluta-

mate release in rat medullary dorsal horn neurons. NeuroReport

22:711–715

148. Zhang X, Wang J, Zhou Q, Xu Y, Pu S, Wu J, Xue Y, Tian Y,

Lu J, Jiang W, Du D (2011) Brain-derived neurotrophic factor-

activated astrocytes produce mechanical allodynia in neuro-

pathic pain. Neuroscience 199:452–460

149. Pellerin L, Magistretti PJ (1996) Excitatory amino acids stim-

ulate aerobic glycolysis in astrocytes via an activation of the

Na?/K?ATPase. Dev Neurosci 18:336–342

150. Sun D, Lytle C, O’Donnell ME (1997) IL-6 secreted by as-

troglial cells regulates Na-K-Cl cotransport in brain microvessel

endothelial cells. Am J Physiol 272(6 Pt 1):C1829–C1835

151. Jayakumar AR, Norenberg MD (2010) The Na-K-Cl Co-trans-

porter in astrocyte swelling. Metab Brain Dis 25:31–38

152. Perea G, Navarrete M, Araque A (2009) Tripartite synapses:

astrocytes process and control synaptic information. Trends

Neurosci 32:421–431

Neurochem Res (2012) 37:2419–2431 2431

123