You are on page 1of 25

B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

a v a i l a b l e a t w w w. s c i e n c e d i r e c t . c o m

w w w. e l s e v i e r. c o m / l o c a t e / b r a i n r e s r e v

Review

Immune and inflammatory mechanisms in neuropathic pain

Gila Moalem, David J. Tracey ⁎


School of Medical Sciences, University of New South Wales, Sydney, NSW 2052, Australia

A R T I C LE I N FO AB S T R A C T

Article history: Tissue damage, inflammation or injury of the nervous system may result in chronic
Accepted 17 November 2005 neuropathic pain characterised by increased sensitivity to painful stimuli (hyperalgesia), the
Available online 4 January 2006 perception of innocuous stimuli as painful (allodynia) and spontaneous pain. Neuropathic
pain has been described in about 1% of the US population, is often severely debilitating and
Keywords: largely resistant to treatment. Animal models of peripheral neuropathic pain are now
Neuropathic pain available in which the mechanisms underlying hyperalgesia and allodynia due to nerve
Nerve injury injury or nerve inflammation can be analysed. Recently, it has become clear that
Immunity inflammatory and immune mechanisms both in the periphery and the central nervous
Inflammation system play an important role in neuropathic pain. Infiltration of inflammatory cells, as well
as activation of resident immune cells in response to nervous system damage, leads to
Abbreviations: subsequent production and secretion of various inflammatory mediators. These mediators
5HT, 5-hydroxytryptamine, promote neuroimmune activation and can sensitise primary afferent neurones and
serotonin contribute to pain hypersensitivity. Inflammatory cells such as mast cells, neutrophils,
8R, 15S-diHETE, (8R, 15S)- macrophages and T lymphocytes have all been implicated, as have immune-like glial cells
dihydroxyeicosa-(5E-9,11,13Z)- such as microglia and astrocytes. In addition, the immune response plays an important role
tetraenoic acid in demyelinating neuropathies such as multiple sclerosis (MS), in which pain is a common
ADP, adenosine diphosphate symptom, and an animal model of MS-related pain has recently been demonstrated. Here,
AMP, adenosine monophosphate we will briefly review some of the milestones in research that have led to an increased
ATP, adenosine triphosphate awareness of the contribution of immune and inflammatory systems to neuropathic pain
BDNF, brain-derived neurotrophic and then review in more detail the role of immune cells and inflammatory mediators.
factor © 2005 Elsevier B.V. All rights reserved.
CCR2, chemokine (C–C motif)
receptor 2
CGRP, calcitonin-gene-related
peptide
COX-1, cyclooxygenase-1
COX-2, cyclooxygenase-2
CX3CR1, G-protein-coupled
chemokine receptor for fractalkine
DRG, dorsal root ganglion
GABA, gamma aminobutyric acid
GBS, Guillain–Barré syndrome
GDNF, glial-cell-line-derived
neurotrophic factor

⁎ Corresponding author. Fax: +61 2 9385 8016.


E-mail address: d.tracey@unsw.edu.au (D.J. Tracey).

0165-0173/$ – see front matter © 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.brainresrev.2005.11.004
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 241

IL-1, interleukin-1
L-NAME, NG-nitro-L-arginine methyl
ester
LTB4, leukotriene B4
MCP-1, monocyte chemoattractant
protein-1
MHC, major histocompatibility
complex
MIP-1α, macrophage inflammatory
protein-1α
MRP-14, cytosolic-calcium-binding
protein of 14 kDa
MS, multiple sclerosis
NGF, nerve growth factor
NO, nitric oxide
NOS, nitric oxide synthase
NSAID, non-steroidal
anti-inflammatory drug
NT-3, Neurotrophin-3
NT-4/5, Neurotrophin-4/5
P2X, ATP-gated ion channels
P2Y, G-protein-coupled ion channels
activated by purine/pyrimidine
nucleotides
PAR-2, protease-activated receptor-2
PGE2, Prostaglandin E2
PGI2, Prostaglandin I2
TCA, tricyclic antidepressant
TEMPOL, 4-hydroxy-2,2,6,6-
tetramethylpiperidine-1-oxyl
Th1 cells, T helper 1 cells
Th2 cells, T helper 2 cells
TLR4, Toll-like receptor 4
TNF, tumour necrosis factor
TrkA, tyrosine kinase receptor A
TrkB, tyrosine kinase receptor B
TrkC, tyrosine kinase receptor C
TRPV1, transient receptor potential
(receptor) vanilloid 1

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
1.1. Chemical environment of sensory neurones contributes to neuropathic pain . . . . . . . . . . . . . . . . . 242
1.2. Inflammatory models of neuropathic pain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
1.3. Neuropathic pain in autoimmune diseases of the nervous system . . . . . . . . . . . . . . . . . . . . . . . 243
1.4. Direct and indirect actions of mediators released by immune cells . . . . . . . . . . . . . . . . . . . . . . . 243
2. Immune cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
2.1. Mast cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
2.2. Neutrophils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
2.3. Macrophages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
2.4. Lymphocytes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
2.5. Glia and Schwann cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
3. Mediators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
3.1. Bradykinin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
3.2. ATP and adenosine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
3.3. Serotonin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
242 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

3.4. Eicosanoids. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250


3.5. Cytokines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
3.5.1. Interleukin-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
3.5.2. Interleukin-6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
3.5.3. Tumour necrosis factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
3.5.4. Interleukin-10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
3.6. Neurotrophins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
3.7. Nitric oxide and reactive oxygen species . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
4. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255

1. Introduction would like to know more about these and other aspects of
neuropathic pain, several recent reviews are available,
Neuropathic pain is caused by lesion or inflammation of dealing with immune and glial mechanisms (Watkins and
the nervous system and is relatively common, with an Maier, 2002, 2005; DeLeo et al., 2004; McMahon et al., 2005;
incidence estimated at 0.6% to 1.5% in the US population Tsuda et al., 2005), inflammatory aspects (Woolf and
(Warfield and Fausett, 2002). It is often severely debilitating Costigan, 1999; DeLeo and Yezierski, 2001; DeLeo et al.,
and largely resistant to treatment (Harden and Cohen, 2004; Sommer and Kress, 2004) and cytokines (Sommer and
2003) mainly because the underlying mechanisms are still Kress, 2004).
poorly understood. Symptoms of neuropathic pain may
include allodynia (pain resulting from a stimulus that is 1.1. Chemical environment of sensory neurones
normally non-painful), hyperalgesia (an excessive response contributes to neuropathic pain
to painful stimuli) and spontaneous pain. The study of
neuropathic pain can be traced back to Weir Mitchell and At the time that the first animal models of pain due to
his classic work on nerve injuries from the American Civil nerve injury were developed, it was widely believed that
War (Mitchell, 1872). Since that time, a great deal has been injury or loss of myelinated and unmyelinated sensory
written on neuropathic pain and its possible causes, but it axons in the sciatic nerve was the key factor in producing
was only with the development of animals models of pain symptoms of neuropathic pain. However, some studies in
due to nerve injury that some real progress was made in the early 1990s suggested that other factors might also be
understanding some of the mechanisms involved. For involved. For example, Maves showed that chromic gut
recent reviews of these mechanisms, see Julius and alone placed next to the sciatic nerve induced thermal
Basbaum (2001) and Scholz and Woolf (2002). The first hyperalgesia without any loss of myelinated axons (Maves
widely used animal model of neuropathic pain was chronic et al., 1993). Of course, it had long been known that C-
constriction injury of the rat sciatic nerve (Bennett and Xie, polymodal nociceptors are sensitised or activated by
1988), in which the sciatic nerve is encircled by four chemical stimuli, including inflammatory mediators (Wood
ligatures of chromic gut. This leads to the cardinal and Docherty, 1997), but these effects were thought of as
symptoms of neuropathic pain—hyperalgesia, allodynia being responsible for inflammatory pain, and little atten-
and apparent spontaneous pain. Other widely used animal tion was given to the idea that immune or inflammatory
models include partial ligation of the sciatic nerve (Seltzer cells or their mediators might play a role in neuropathic
et al., 1990) and section of one or more of the spinal pain until the advent of useful animal models. It was also
nerves which contribute to the sciatic (Kim and Chung, known that nerve injury resulted in activation of mast cells
1992). Research on these animal models of peripheral (Olsson, 1967) and recruitment of neutrophils and macro-
neuropathic pain (the pain syndrome resulting from phages (Perry et al., 1987). However, it was not until 1993
lesions of the peripheral nervous system) has made it that ‘acute changes in the endoneurial microenvironment’
clear that a number of mechanisms are involved, including were explicitly linked with the initial development of
ectopic excitability of sensory neurones, altered gene hyperalgesia (Frisen et al., 1993; Sommer et al., 1993).
expression of sensory neurones and sensitisation of Further work confirmed suggestions that the factors
neurones in the dorsal horn of the spinal cord (Scholz responsible for neuropathic hyperalgesia included Wallerian
and Woolf, 2002; Woolf, 2004). However, there is increasing degeneration and macrophage activation (Sommer et al.,
evidence that inflammatory and immune mechanisms also 1995). At around the same time, Clatworthy followed up the
play a role, and we will begin by looking at a brief history observations of Maves et al. and showed that suppression
of how this evidence has accumulated. We will then look of the inflammatory response in the injured sciatic nerve
in more detail at the evidence for involvement of immune blocked the development of hyperalgesia, while enhancing
cells and inflammatory mediators in neuropathic pain. the inflammatory response augmented the hyperalgesia
However, the review does not provide exhaustive coverage (Clatworthy et al., 1995). Evidence on the involvement of
of all immune and inflammatory mechanisms that con- immune cells and inflammatory mediators in hyperalgesia
tribute to pain following nerve injury. For readers who was reviewed at this time, and it was suggested that these
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 243

mediators (and cytokines in particular) were a probable link hyperalgesia are most likely the combined result of nerve
between the activation or recruitment of immune cells to injury itself (e.g. ectopic firing of injured sensory axons) and
the injured nerve and the development of hyperalgesia and the release of algesic mediators by immune cells activated
neuropathic pain (Tracey and Walker, 1995; Watkins et al., near the site of injury. In fact, neuropathic pain may develop
1995). Glial cells supporting neurones in the spinal cord without any injury of sensory axons. This was shown in rats
(Meller et al., 1994) and Schwann cells supporting axons in by induction of neuropathic hyperalgesia by placing segments
the peripheral nerve (Constable et al., 1994; Wagner and of chromic gut next to the sciatic nerve (Maves et al., 1993) or
Myers, 1996a) were added to the list of cell types which by lesion of the ventral roots of spinal nerves (Li et al., 2002;
might contribute to neuropathic pain. In fact, glial cells and Sheth et al., 2002). In both cases, typical signs of neuropathic
Schwann cells may also play an active role in the immune pain were elicited without damage to sensory axons, implying
process by releasing mediators, including cytokines, and that sensitisation or activation of sensory fibres by inflam-
acting as antigen-presenting cells following nerve injury matory mediators is sufficient to cause neuropathic pain. As a
(Argall et al., 1992; Bergsteinsdottir et al., 1992; Constable et result of observations like these, inflammatory models of
al., 1994). These ideas were not widely accepted at the neuropathic pain have been developed in which carrageenan
time, but, gradually, evidence accumulated to give them or complete Freund's adjuvant (Eliav et al., 1999) or zymosan
further support. For example, tumour necrosis factor (TNF) (Chacur et al., 2001) is applied around the intact sciatic nerve
and interleukins 1 and 6 (IL-1, IL-6) had already been of the rat. These models should help us to understand the role
shown to induce acute or short-term hyperalgesia (Ferreira of inflammatory and immune mechanisms in neuropathic
et al., 1988; Cunha et al., 1992) but were now implicated pain.
directly in neuropathic pain, involving chronic hyperalgesia
and allodynia (Wagner and Myers, 1996b; DeLeo et al., 1997; 1.3. Neuropathic pain in autoimmune diseases of the
Arruda et al., 1998). nervous system
Another example is the contribution of nerve growth
factor (NGF), which shares many attributes with typical Autoimmune diseases of the nervous system including
cytokines (Bonini et al., 2003). NGF is best known for its multiple sclerosis (MS) and Guillain–Barré syndrome (GBS)
role in protecting some neuronal types from cell death are among the most common disabling neurologic diseases,
during development, but Levine and his colleagues showed which cause not only demyelination with impaired motor
that the N-terminal octapeptide of NGF elicited hyperalge- function, but also abnormal sensory phenomena including
sia in the rat when applied to injured tissue (Taiwo et al., chronic neuropathic pain. MS is an autoimmune demyelinat-
1991). These experiments were provoked by apparent ing disease of the central nervous system that causes
similarities between the NGF fragment and bradykinin—an relapsing and chronic neurological impairment. Pain is
inflammatory mediator known to activate or sensitise experienced by approximately 65% of patients with MS at
nociceptors and to elicit hyperalgesia (Dray and Perkins, some time during the course of their disease (Kerns et al.,
1993; Dray, 1997). By this stage, it was known that the level 2002). GBS is an acute inflammatory demyelinating neuropa-
of NGF rises substantially in inflamed tissue and in injured thy caused by an autoimmune attack on the peripheral
nerve (Heumann et al., 1987; Weskamp and Otten, 1987), nervous system and characterised by motor disorders such
and a causal relationship between tissue levels of NGF and as weakness or paralysis and variable sensory disturbances
inflammatory hyperalgesia was confirmed by showing that (Hughes et al., 1999). Pain is a common symptom of GBS
anti-NGF antibodies reduced inflammatory hyperalgesia occurring in 70–90% of cases (Pentland and Donald, 1994;
(Lewin et al., 1994; Woolf et al., 1994). NGF was later Moulin et al., 1997). Immune cells including macrophages and
shown to contribute to neuropathic hyperalgesia as well T cells are believed to play a critical role in the pathogenesis of
(Herzberg et al., 1997; Theodosiou et al., 1999). The both MS and GBS and may contribute to the related pain.
underlying mechanism is not clear, but it is agreed that Studying neuropathic pain in these autoimmune syndromes
NGF has a cytokine-like action on inflammatory cells has been hampered by the lack of proper animal models for
(Otten, 1991) including mast cells, basophils, neutrophils autoimmunity-related pain. However, recent study in mice
and lymphocytes. Furthermore, NGF is produced and has shown that transient demyelination of peripheral affer-
released by leukocytes such as macrophages and T ents without axonal loss results in neuropathic pain behav-
lymphocytes (Vega et al., 2003). NGF is a potent degranu- iour (Wallace et al., 2003), and a suitable animal model of MS-
lator of mast cells (Pearce and Thompson, 1986), which related pain has recently been developed using both active
appear to play an important role in inflammatory (Woolf et and passive induction of experimental autoimmune enceph-
al., 1996) as well as neuropathic hyperalgesia (Zuo et al., alomyelitis (Aicher et al., 2004).
2003). It seems likely that mediators released by activated
mast cells contribute to hyperalgesia following nerve injury 1.4. Direct and indirect actions of mediators released by
— histamine has been strongly implicated (Zuo et al., 2003), immune cells
but several other mediators may also be involved.
While there is good evidence that inflammatory mediators
1.2. Inflammatory models of neuropathic pain contribute to neuropathic pain, it is not clear just how or
where these mediators act. Do they act directly on receptors
It is now apparent that early animal models of neuropathic located on the cell membrane of the neurone or do they
pain suffer from the disadvantage that symptoms such as activate non-neuronal cells (such as mast cells or Schwann
244 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

cells), which then release a secondary mediator that exerts a (Sawynok and Liu, 2003); it also plays a role in neuropathic
direct action on neuronal receptors? Are the neuronal pain (Barclay et al., 2002; Jarvis et al., 2002). It may well be
receptors for chemical mediators located on the nerve that ATP contributes to neuropathic pain by its action on
terminals or cell bodies of sensory neurones, even though axonal receptors.
these may be many centimetres from the lesion site? The
answer to the first question is that both direct and indirect
mechanisms are likely to operate. Some mediators (such as 2. Immune cells
prostaglandin E2) probably sensitise nociceptors directly by
acting on receptors on the neuronal cell membrane (Pitchford Several inflammatory and immune-like glial cells have been
and Levine, 1991). While IL-1 may also have a direct action on implicated in the pathogenesis of neuropathic pain. They
nociceptors (Sommer and Kress, 2004), it has additional include mast cells, neutrophils, macrophages and T cells in
indirect actions by eliciting the production of prostaglandins the peripheral nervous system (Fig. 1) and microglia (Fig. 2)
(Cunha et al., 1992) or even more indirectly via neural and astrocytes in the central nervous system. Table 1
pathways from vagal afferents to the brainstem and then to summarises immune cells and the mediators they release
the spinal cord (Watkins et al., 1995). The answer to the second with their most significant actions.
question is that receptors for inflammatory mediators may be
located not only on the neuronal cell body and peripheral 2.1. Mast cells
nerve terminals, but also on the axon itself.
Recent work by Grafe and his colleagues makes it clear Mast cells are not only critical effector cells in allergic
that some of the neuronal receptors are located on the axon disorders but are also important initiators and effectors of
itself (Irnich et al., 2002; Lang et al., 2003; Moalem et al., innate immunity (Galli et al., 2005). There is a resident
2005). This makes it easier to understand how mediators population of mast cells in the peripheral nerve (Olsson,
released by immune cells clustered around the nerve lesion 1968), and these mast cells are degranulated at the site of a
can influence the excitability of sensory neurones — since nerve lesion (Olsson, 1967; Zuo et al., 2003). The granules
these mediators would be released close to axonal receptors contain mediators such as histamine, proteases and cytokines
and would not need to diffuse distally to nerve terminals or (Metcalfe et al., 1997; Galli et al., 2005), several of which are
proximally to cell bodies. For example, it was shown that capable of sensitising or activating neurones. For example,
ATP increased the excitability of unmyelinated C-fibres in an histamine sensitises visceral nociceptors (Koda and Mizu-
in vitro preparation of an isolated segment of the sural mura, 2002). Although it elicits itch in normal skin, histamine
nerve, in which no cell bodies or nerve terminals were application to the skin of patients with neuropathic pain did
present. Tissue levels of ATP are raised in inflamed tissue, not evoke an itch but instead induced a severe increase in
and ATP activates nociceptors causing pain and hyperalgesia spontaneous burning pain (Baron et al., 2001). Activation of

Fig. 1 – Peripheral nerve injury induces activation of resident immune cells as well as recruitment of inflammatory cells to the
nerve. Injury of a peripheral nerve initiates an inflammatory cascade in which mast cells residing in the nerve are the first to be
activated. They release mediators such as histamine (hist) and TNF, which sensitise nociceptors and contribute to the
recruitment of neutrophils and macrophages. Mediators released by neutrophils (including the chemokine MIP-1α and the
cytokine IL-1β) assist in recruitment of macrophages. Both neutrophils and macrophages in the nerve produce and secrete
mediators such as TNF and PGE2 that can further sensitise nociceptors. Nerve injury also initiates Schwann cell
de-differentiation and the release of several algesic mediators such as pro-inflammatory cytokines, NGF, PGE2 and ATP. These
initial events promote the recruitment of T cells, which can secrete a variety of cytokines depending on their subtype. This
cocktail of mediators serves as a mechanism of enhanced inflammatory response in the injured nerve and contributes to
neuropathic pain.
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 245

mast cells with cromoglycate strongly suppressed the devel-


opment of neuropathic pain and reduced the numbers of
neutrophils and macrophages at the lesion site (Zuo et al.,
2003). Treatment of the nerve-injured rats with H1 and H2
histamine receptor antagonists also alleviated neuropathic
hyperalgesia, although to a lesser extent than stabilisation
with cromoglycate. This suggests that nerve injury induces
degranulation of mast cells and release of histamine, which
acts on histamine receptors on endothelial cells to induce
recruitment of leukocytes such as neutrophils and macro-
phages (Yamaki et al., 1998). Both of these cell types contribute
to neuropathic pain (Liu et al., 2000b; Perkins and Tracey,
2000). The question of how mast cells are activated by nerve
injury is still open, but it could be brought about by increased
levels of adenosine (Sawynok et al., 2000) or bradykinin
(McLean et al., 2000). Other mast cell mediators are probably
also involved in the initiation of neuropathic pain. One
example is tryptase, a trypsin-like serine protease that is
specific to mast cells. One of the most prominent mast cell
mediators, it is associated with the mast cell granules and
released by degranulation. Tryptase acts on the protease-
activated receptor-2 (PAR-2), which is known to be present on
primary sensory neurones and to play a role in inflammation.
Fig. 2 – Peripheral nerve injury induces glial activation in the Recently, it has been shown to trigger inflammatory hyper-
dorsal horn of the spinal cord. Injury to a peripheral nerve algesia and nociceptive behaviour in rats (Kawabata et al.,
initiates increased release of neurotransmitters such as 2001; Vergnolle et al., 2001).
glutamate, substance P and possibly ATP from the central
terminals of primary afferents. These neurotransmitters can 2.2. Neutrophils
activate both second order neurones and glia. For example,
ATP can activate glial cells through binding to their P2X4 Neutrophils (or polymorphonuclear leukocytes) are an essen-
receptors, and fractalkine released by second order neurones tial part of the innate immune system. Their cytoplasm
can activate glia through binding to their CX3CR1 receptors. contains granules and secretory vesicles that can release a
The TLR4 receptor is involved in glial activation, but the wide range of microbicidal effector molecules including
nature of its specific ligand in the spinal cord is unclear. bactericidal proteins and cytokines, proteinases and reactive
These triggers cause production and release of inflammatory oxygen species (Witko-Sarsat et al., 2000; Faurschou and
mediators including pro-inflammatory cytokines (such as Borregaard, 2003). The first steps in migration towards an
TNF and IL-1β), glutamate (Glu), prostaglandins (PGs) and inflammatory focus are adhesion to the surface of the vascular
nitric oxide (NO) from glial cells. These agents are then endothelium (mediated by selectins) followed by rolling and
capable of further enhancing glial activation and production transendothelial migration. Neutrophils then migrate towards
of inflammatory mediators that sensitise dorsal horn the inflammatory site up a gradient of chemoattractants.
neurones, thereby contributing to neuropathic pain. These include formyl peptides released by bacteria or dying
cells and chemokines released by immune cells (Witko-Sarsat
et al., 2000). Neutrophils contribute to inflammatory hyper-
mast cells may also cause secretion of mediators without algesia (Levine et al., 1984, 1985; Bennett et al., 1998b) by
overt degranulation (Theoharides and Cochrane, 2004) by release of the 15-lipoxygenase product 8R, 15S-diHETE (Levine
synthesis of lipid mediators such as prostaglandin D2 or by et al., 1985; White et al., 1990); their release of cytokines such
transcription, translation and secretion of a wide range of as TNF (Witko-Sarsat et al., 2000) may also contribute.
cytokines and chemokines (Mekori and Metcalfe, 2000). Neutrophils are not observed in normal nerves but are found
Activation and degranulation of mast cells therefore release in significant numbers at the site of injury in injured
some mediators (such as serotonin in the rat) which can elicit peripheral nerves (Perry et al., 1987; Clatworthy et al., 1995;
hyperalgesia by a direct action on the nociceptor (Rueff and Perkins and Tracey, 2000; Zuo et al., 2003). Work in our
Dray, 1993; Sommer, 2004). Some mast cell mediators exert a laboratory showed that depletion of circulating neutrophils at
broader range of actions by initiating an inflammatory the time of nerve injury significantly attenuated the induction
cascade (Fig. 1). For example, release of histamine, leuko- of hyperalgesia. However, depletion of neutrophils at day
trienes and chemokines contributes to recruitment of leuko- 8 post-injury did not alleviate hyperalgesia after its normal
cytes including neutrophils and macrophages (Gaboury et al., induction since by this time neutrophil numbers at the site of
1995; Malaviya and Abraham, 2000). In fact, recent work in our nerve injury have declined to negligible levels (Zuo et al., 2003).
laboratory suggests that degranulation of mast cells plays a It seems likely that nerve injury induces an inflammatory
key role in the initiation of neuropathic pain since, in rats cascade initiated by mast cell activation. Activated mast cells
subjected to partial ligation of the sciatic nerve, stabilisation of release mediators that promote migration of neutrophils from
246 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

Table 1 – Immune cells, mediators released and actions Table 1 (continued)


Cell type Mediators Actions Cell type Mediators Actions

Mast cells Cytokines and Recruit leukocytes; Serotonin (in Sensitises nociceptors
chemokines (Galli IL-1β facilitates the murine rodents) (Lang et al., 1990;
et al., 2005; Mekori release of CGRP from (Galli et al., 2005) Rueff and Dray, 1993)
and Metcalfe, 2000) nociceptors (Fukuoka and sensory axons
et al., 1994) and elicits (Moalem et al., 2005)
hyperalgesia (Ferreira
et al., 1988; Follenfant Neutrophils 8R, 15S-diHETE Sensitises nociceptors
et al., 1989; Sung et al., (Levine et al., 1986) (White et al., 1990)
2004; Zelenka et al.,
2005); TNF sensitises Cytokines See above entry for
C-fibres (Junger and (Witko-Sarsat et al., cytokines and
Sorkin, 2000; Sorkin 2000) and chemokines chemokines
et al., 1997) and DRG (Scapini et al., 2000)
neurones (Schäfers et
al., 2003a,b) and Defensins (Faurschou Recruit monocytes,
produces hyperalgesia and Borregaard, 2003) T-cells (Faurschou and
(Cunha et al., 1992; Borregaard, 2003;
Schäfers et al., 2003a,b; Welling et al., 1998)
Zelenka et al., 2005).
Chemokines produce Reactive oxygen Sensitise neurons in
excitatory effects on species (Witko-Sarsat PNS and CNS (Aley et
DRG neurons and et al., 2000) al., 1998; Kim et al.,
produce allodynia 2004; Liu et al., 2000a,b;
(Oh et al., 2001) Meller and Gebhart,
1993; Twining et al.,
Histamine Sensitises nociceptors 2004)
(Mekori and (Koda and Mizumura,
Metcalfe, 2000) 2002), recruits Opioid peptides Anti-nociceptive (Brack
leukocytes (Gaboury (Brack et al., 2004) et al., 2004)
et al., 1995; Yamaki
et al., 1998); enhances Macrophages Cytokines (Nathan, See above entry for
neuropathic pain in 1987) and cytokines and
patients (Baron et al., chemokines chemokines
2001)
Prostaglandins PGE2, PGI2 sensitise
Leukotriene B4 Recruits neutrophils (PGE2, PGI2) nociceptors (Taiwo and
(Galli et al., 2005; (Bennett et al., (Nathan, 1987) Levine, 1989), PGE2 acts
Mekori and 1998a,b; Levine et al., on spinal neurons (Baba
Metcalfe, 2000) 1984) et al., 2001)

Mast cell protease Activates PAR-2 Reactive oxygen See above entry for
(e.g. tryptase) receptor (Kawabata species (Nathan, 1987) reactive oxygen species
(Metcalfe et al., et al., 2001; Vergnolle
1997) et al., 2001) T lymphocytes Pro-inflammatory Increase
cytokines (Mosmann hyperalgesia and
Nerve growth Sensitises nociceptors et al., 1986; Sad et allodynia (Moalem
factor (Bonini et (Bennett et al., al., 1995) et al., 2004)
al., 2003) 1998a,b; Bennett,
2001; Koltzenburg et Anti-inflammatory Decrease hyperalgesia
al., 1999a,b), indirect cytokines (Mosmann and allodynia (Moalem
actions via immune et al., 1986; Sad et al., et al., 2004)
cells and sympathetic 1995)
neurones (Bennett et
al., 1998a,b; Lewin et Glia and Cytokines and See above entry for
al., 1994; Woolf et al., Schwann chemokines cytokines and
1996) cells (Hanisch, 2002) chemokines

Prostaglandins PGE2 sensitises Excitatory amino Modulate synaptic


(PGD2, PGE2) (Galli nociceptors (Taiwo acids (Araque et al., transmission and
et al., 2005; Mekori and Levine, 1989), 2001; Watkins et al., neuronal excitability
and Metcalfe, 2000) depolarises spinal 2001) (Araque et al., 2001)
neurons (Baba et al.,
2001)
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 247

Table 1 (continued) genetic defect which delays the recruitment of macrophages


Cell type Mediators Actions (Perry and Brown, 1992; Araki et al., 2004). Hyperalgesia is
also delayed in these mice, suggesting that macrophages
Nitric oxide (Minghetti Activates guanylyl
contribute to pain resulting from nerve injury (Myers et al.,
and Levi, 1998) cyclase, sensitises
neurons in CNS (Luo 1996). This suggestion was confirmed by the demonstration
and Cizkova, 2000; that depletion of macrophages in nerve-injured rats allevi-
Minghetti and Levi, ated neuropathic hyperalgesia (Liu et al., 2000b). However,
1998) another study found only a limited role of macrophages in
the generation of mechanical allodynia following nerve
Prostaglandins See above entries
injury (Rutkowski et al., 2000). Furthermore, in a model of
(Minghetti and Levi, for prostaglandins
neuropathic pain, the development of mechanical allodynia
1998; Muja and
DeVries, 2004) was totally abrogated in mice with a knockout for the
chemokine receptor CCR2, involved in monocyte recruitment
ATP (Anderson et al., Activates nociceptors (Abbadie et al., 2003). It is noteworthy that macrophages also
2004; Liu et al., 2005) (Hamilton and invade the dorsal root ganglia following peripheral nerve
McMahon, 2000), lesion and may contribute to neuropathic pain by release of
increases sensory
excitatory agents that generate ectopic activity in sensory
C-fibre excitability
(Irnich et al., 2002;
neurones (Hu and McLachlan, 2002).
Moalem et al., 2005) What are the signals that recruit macrophages? Essential
roles are played by monocyte chemoattractant protein-1,
MCP-1; macrophage inflammatory protein-1α, MIP-1α; and
small blood vessels into inflamed tissue (Gaboury et al., 1995), the IL-1β (Perrin et al., 2005); the latter two are known to be
and neutrophils in turn release mediators such as chemokines released by neutrophils (Witko-Sarsat et al., 2000). Release of
and defensins which are chemotactic for macrophages (Well- defensin by neutrophils may also be involved (Welling et al.,
ing et al., 1998; Scapini et al., 2000) (Fig. 1). Finally, it is worth 1998). Which are the mediators that are released by macro-
noting that neutrophils may also have an anti-inflammatory phages to cause hyperalgesia? Macrophages secrete prosta-
and anti-nociceptive role. MRP-14 is a calcium-binding protein glandins, including prostaglandin E2 and I2 (Nathan, 1987),
that forms a significant proportion of the cytoplasmic protein which sensitise primary afferents directly. Prostaglandin
in neutrophils. It de-activates macrophages in vitro and release by macrophages is strongly implicated in neuropathic
suppresses inflammatory pain in mice (Giorgi et al., 1998). pain since inhibition of cyclooxygenase, an enzyme respon-
Under inflammatory conditions, neutrophils may secrete sible for prostaglandin synthesis, relieves hyperalgesia in
opioid peptides, which bind to opioid receptors on peripheral nerve-injured rats (Syriatowicz et al., 1999), and cyclooxygen-
sensory neurones and mediate anti-nociception (Brack et al., ase-2 is upregulated in macrophages in the injured nerve (Ma
2004). The evidence for an anti-nociceptive role for neutro- and Eisenach, 2002, 2003b). Other algesic mediators that are
phils appears to contradict the pro-nociceptive role outlined released by macrophages (Nathan, 1987) and most likely
above. However, neuropathic pain and inflammatory pain contribute to neuropathic pain include reactive oxygen
differ in some respects, even though they share some basic species and the cytokines TNF (Sommer et al., 1998b), IL-1β
mechanisms. In particular, a tertiary peripheral site of opioid and IL-6 (Sommer and Kress, 2004) (Fig. 1).
analgesia is initiated by acute inflammation which may not be
activated in neuropathic pain (Bridges et al., 2001). This 2.4. Lymphocytes
difference may help to explain why opioids are less effective
in relieving neuropathic pain than inflammatory pain (Delle- Lymphocytes can be classed as B lymphocytes, responsible
mijn, 1999). for antibody production, T lymphocytes, which are the
mediators of cellular immunity, and natural killer cells.
2.3. Macrophages Both T cells and natural killer cells are found at the site of
the nerve lesion in rat models of neuropathic pain (Cui et
Macrophages phagocytose foreign particles such as microbes al., 2000). T cells are also found in increased numbers in
but also have an important role in removing injured or dead the dorsal root ganglia and spinal cord after lesions of the
tissue. In the context of this review, they play a vital part in peripheral nerve, leading to the suggestion that they may
phagocytosing the dead or dying remnants of injured play a role in generating neuropathic pain (Hu and
Schwann cells and axotomised axons in Wallerian degener- McLachlan, 2002; Sweitzer et al., 2002). This suggestion
ation (Brück, 1997). There is a resident population of has now been confirmed by the finding that congenitally
macrophages in the peripheral nerve and dorsal root ganglia; athymic nude rats, which lack mature T cells, develop
their equivalent in the CNS is the microglia. In peripheral significantly less mechanical allodynia and thermal hyper-
nerve, the resident macrophages are assisted with clearance algesia when subjected to sciatic nerve injury than their
of cellular debris by an influx of haematogenous macro- heterozygous littermates (Moalem et al., 2004). T cells are
phages (Bendszus and Stoll, 2003; Mueller et al., 2003). rather heterogeneous, they can be divided into CD4+
Clearance of debris is a prerequisite to regeneration in (helper) and CD8+ (cytotoxic) cells, and each of these
peripheral nerves (Perry and Brown, 1992; Correale and Villa, subpopulations can be further divided into type 1 and
2004), and Wallerian degeneration is delayed in mice with a type 2 subsets according to their cytokine secretion profile
248 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

(Mosmann et al., 1986; Sad et al., 1995). The functions of T neuropathic pain inhibits the development of hyperalgesia
helper 1 (Th1) and Th2 cells correlate well with their and allodynia (Raghavendra et al., 2003; Ledeboer et al.,
distinctive cytokines. Th1 cells produce IL-2 and interferon- 2005). Glia are activated by several mediators including ATP,
gamma and are involved in cell-mediated inflammatory bradykinin, substance P and prostaglandins (Hosli and Hosli,
reactions, while Th2 cells produce IL-4, IL-5, IL-6, IL-9, IL-10 1993; Watkins et al., 2001). ATP is known to activate
and IL-13, are involved in antibody and allergic responses nociceptors in the periphery by an action on P2X3 receptors
and inhibit synthesis of pro-inflammatory cytokines by Th1 (Cook et al., 1997), but, in the CNS, it activates P2X4 receptors
cells (Mosmann and Sad, 1996). Th1 and Th2 cytokines are that are selectively expressed by activated microglia and
mutually inhibitory for the functions of the reciprocal appear to be required for neuropathic pain (Tsuda et al.,
phenotype. Recent work by Moalem et al. has demonstrated 2005). Recent evidence suggests that activation of microglia
that passive transfer of type 1 T cells, which produce pro- by fractalkine (Milligan et al., 2004; Verge et al., 2004) and by
inflammatory cytokines, into nude rats increased their the Toll-like receptor 4 (TLR4) (Tanga et al., 2005) is also
sensitivity to noxious stimuli to a level comparable with involved in the initiation of neuropathic pain (Fig. 2).
that of heterozygous rats. By contrast, passive transfer of Fractalkine is a chemokine expressed on the surface of
type 2 T cells, which release anti-inflammatory cytokines, spinal neurones, which activates the CX3CR1 receptor on
into heterozygous rats reduced their pain sensitivity microglia, while the Toll-like receptors recognise invariant
(Moalem et al., 2004). These findings indicate that T cells molecular structures of pathogens and specific ligands
take part in the course of neuropathic pain (Fig. 1) and released after nerve injury.
suggest that Th1 and Th2 cells have opposite effects on Although we have some idea of how glial cells are
neuropathic pain as a result of the distinct sets of activated, there is little concrete evidence on how activated
cytokines they release. The role of B cells in neuropathic glial cells regulate neuronal function to bring about symptoms
pain remains to be elucidated. of neuropathic pain. Activated glia release a number of
mediators that excite spinal neurones (Watkins et al., 2001),
2.5. Glia and Schwann cells including prostanoids and nitric oxide (Minghetti and Levi,
1998), cytokines and chemokines (Hanisch, 2002), excitatory
Glial cells surround and support neurones in the nervous amino acids (Araque et al., 2001) and ATP (Anderson et al.,
system and are 10 to 50 times as numerous as neurones. In 2004). However, the role and relative importance of these glial
the last few years, it has become clear that glia are by no mediators in neuropathic pain still need to be established.
means passive bystanders in the nervous system but have Interestingly, microglia have been recently implicated in
important metabolic and immune functions. In the periph- mirror-image pain, pain arising from sites contralateral to
eral nervous system, glia are represented by Schwann cells, the site of pathology (Koltzenburg et al., 1999b; Twining et al.,
while, in the CNS, there are three types of glial cell that can 2004). The mechanisms underlying the expansion of pain to
be distinguished — the astrocytes and oligodendrocytes the contralateral site are not clear. However, one possible
(macroglia) and the microglia. Microglia form a population mechanism is through spinal gap junction activation. Recent
of resident macrophages in the CNS, constituting about 5– study has demonstrated that treatment with a gap junction
10% of glial cells in the CNS. In the resting state, they have a decoupler reverses neuropathy-induced mirror-image pain
small soma with fine processes, but, when activated by (Spataro et al., 2004). Thus, strong activation of glia at one site
stimuli such as trauma, ischemia or inflammation, they take can lead to gap junctional propagation of calcium waves that
on an amoeboid shape and function as phagocytes (Vilhardt, activates distant glia, leading to release of pain-enhancing
2005). At this stage, they express the major histocompatibil- substances.
ity complex (MHC), which has a role in presenting antigen to Schwann cells in the peripheral nervous system provide
T lymphocytes, and release mediators that include pro- the myelin sheath in myelinated axons and also envelop
inflammatory cytokines (Piehl and Lidman, 2001). Astrocytes small bundles of unmyelinated axons (Remak bundles). The
represent the largest cell population in the CNS. They closely primary role of Schwann cells was long thought to be the
interact with neurones to provide active maintenance of provision of an insulating sheath around myelinated axons
homeostasis by regulating extracellular ion and neurotrans- (facilitating conduction), maintenance of suitable environ-
mitter concentrations and extracellular pH in their micro- ment for peripheral axons, and guiding regenerating axons
environment. The relative contribution of microglia and back to their targets after injury. In recent years, an immune
astrocytes to neuropathic pain is still under investigation. function for Schwann cells has received increasing recogni-
However, recent studies suggest that microglia are more tion since they interact with the immune system in T-cell-
important for the initiation, while astrocytes are more mediated immune responses by expressing MHC class II
important for the maintenance of neuropathic pain (Colburn molecules (Bergsteinsdottir et al., 1992). Schwann cells
et al., 1999; Raghavendra et al., 2003). In this review, we will release several algesic mediators, including TNF (Murwani
use the term glia for both microglia and astrocytes in the et al., 1996; Wagner and Myers, 1996a; Shamash et al., 2002),
CNS. Evidence has been accumulating since the early 1990s IL-1 (Bergsteinsdottir et al., 1991; Shamash et al., 2002), IL-6
that spinal glia contribute to neuropathic pain (DeLeo and (Bolin et al., 1995), NGF (Matsuoka et al., 1991), prostaglandin
Colburn, 1999; Watkins et al., 2001; McMahon et al., 2005; E2 (Muja and DeVries, 2004) and ATP (Liu et al., 2005).
Tsuda et al., 2005). Microglia are activated by peripheral Schwann cells also express a number of ion channels and
nerve injury (Colburn et al., 1999; Tanga et al., 2004), and receptors for mediators including glutamate (Liu and Ben-
inhibition of microglial activation in the several models of nett, 2003), ATP (Grafe et al., 1999; Irnich et al., 2001; Liu et
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 249

al., 2005) and IL-1 (Skundric et al., 1997). Given the close may also bind non-specifically to the cell surface and elicit
relationship between Schwann cells and peripheral axons degranulation (Griesbacher and Rainer, 1999).
and the presence of axonal receptors for mediators such as Injury of the sciatic nerve in rats led to upregulation of B2
ATP (Irnich et al., 2001, 2002), it seems very likely that and B1 receptors on lumbar dorsal root ganglia (Eckert et al.,
Schwann cells contribute to neuropathic pain by sensitising 1999; Levy and Zochodne, 2000), and in rats with a chronic
or activating the axons of nociceptors (Fig. 1). However, it constriction injury of the sciatic nerve, constant infusion of B2
has not yet been possible to obtain direct evidence for this or B1 receptor antagonists reduced hyperalgesia (Yamaguchi-
as selective inhibition or depletion of Schwann cells in an Sase et al., 2003). A recent study in nerve-injured mice found
intact animal has not yet been achieved. that B2 receptors were mainly expressed by the cell bodies of
unmyelinated axons and that their expression was drastically
decreased after nerve injury. By contrast, B1 receptors were
newly expressed on the cell bodies of myelinated axons,
3. Mediators suggesting that nerve injury induced a switch in the type of
sensory neurone and the type of bradykinin receptor involved
The mediators released by inflammatory and immune cells in nociception (Rashid et al., 2004). Bradykinin and its B2
may act directly to sensitise or activate neurones (usually receptor have recently been implicated in central pain
nociceptors in the periphery or dorsal horn neurones in the transmission. This work showed that the B2 receptor is
spinal cord). Alternatively, they may act on a non-neuronal expressed by dorsal horn neurones and is involved in central
cell, which on activation releases another mediator that does sensitisation evoked by primary afferents (Wang et al., 2005).
act directly on the neurone. These mediators form a long and Activation of B2 receptors on glial cells in the CNS could play a
increasing list that includes bradykinin, ATP and adenosine, role in neuropathic pain, but there is no evidence for this as yet.
serotonin, eicosanoids, cytokines, neurotrophins and reactive
oxygen species (Dray, 1995). This list is not exhaustive, and 3.2. ATP and adenosine
some inflammatory mediators such as protons and histamine
(see the Mast cells section) are not reviewed here. Purines such as ATP and adenosine are well known for their
role in energy metabolism but also have widespread effects on
3.1. Bradykinin neurones and immune cells. ATP is a classical neurotrans-
mitter but is also released by non-neuronal cells and from
Bradykinin and kallidin are peptides that are formed from disrupted cells in injured tissue (Cook and McCleskey, 2002).
precursors in the blood and other tissues. In the blood, Adenosine is a neuromodulator that is produced in the
bradykinin is formed from high molecular weight kininogen extracellular space by degradation of extracellular ATP and
as part of the clotting cascade, while in other tissues kallidin is also transported from the cytoplasm into the interstitial
(lysyl-bradykinin) is formed from a low molecular weight space by transport proteins (Linden, 1994). Inosine is produced
kininogen. At least two subtypes of bradykinin receptors (B1 by deamination of adenosine. It may reach millimolar levels in
and B2) have been characterised on the basis of in vivo and in ischemic tissue and can degranulate mast cells (Jin et al.,
vitro functional studies, using selective agonists and antago- 1997).
nists of these receptors (Regoli et al., 1998). Bradykinin In 1977, Bleehen and Keele reported that ATP, ADP, AMP
produces pain hypersensitivity by sensitising the peripheral and adenosine all produced pain when applied to a blister base
terminals of nociceptors and by potentiating glutamatergic in man (Bleehen and Keele, 1977). Since then, it has been
synaptic transmission in the spinal cord (Wang et al., 2005). found that adenyl compounds and other purines activate
Bradykinin also sensitises nociceptors by disinhibition of the nociceptors and elicit pain by their actions on purinergic
TRPV1 receptor (Stucky et al., 1998; Di Marzo et al., 2002). B2 receptors (Burnstock and Wood, 1996; Sawynok, 1998; Hamil-
receptors are constitutive and are found on a number of cell ton and McMahon, 2000). Receptors for purines have been
types including sensory neurones and microglia (Dray, 1997). classified into adenosine (P1) receptors and nucleotide (P2)
Increased expression of B1 receptors plays a prominent role in receptors (Ralevic and Burnstock, 1998). P2 receptors can be
inflammatory hyperalgesia (Dray and Perkins, 1993; Dray, subdivided into P2X receptors, which are ATP-gated ion
1997), and this upregulation is facilitated by cytokines such as channels (Chizh and Illes, 2001), and P2Y receptors, which
IL-1β (Marceau, 1995). Bradykinin acts mainly on the B2 are G-protein coupled receptors activated by purine or
receptor, while des-Arg9-bradykinin, the major metabolite of pyrimidine nucleotides such as ATP or UTP (von Kugelgen
bradykinin, has high affinity for the B1 receptor but little and Wetter, 2000; Burnstock and Knight, 2004).
affinity for the B2 receptor (Dray and Perkins, 1993; Khasar et Initial experiments on the role of ATP in neuropathic pain
al., 1995). Both B1 and B2 receptors have been shown to be implicated P2X3 receptors, although results were apparently
involved in peripheral inflammatory hyperalgesia (Dray and contradictory. Immunocytochemical studies in rats and
Perkins, 1993; Dray, 1997). Bradykinin stimulates macrophages humans found that P2X3 receptor expression was significantly
to release TNF and IL-1 (Tiffany and Burch, 1989), as well as reduced after axotomy (Bradbury et al., 1998; Yiangou et al.,
factors that are chemotactic for neutrophils and monocytes 2000) or partial nerve ligation (Kage et al., 2002), while the
(Sato et al., 1996). Bradykinin also elicits the release of number of P2X3 immunoreactive neurones in the DRG or
histamine from mast cells but does so by non-specific binding trigeminal ganglia was increased after chronic constriction
to the cell surface rather than to B1 or B2 receptors (Reissmann injury of peripheral nerves (Eriksson et al., 1998; Novakovic et
et al., 2000). In fact, peptide antagonists of B1 or B2 receptors al., 1999). This apparent discrepancy may be due to the
250 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

different types of nerve injury, leading to downregulation of RNAs for 5HT1B, 5HT1D, 5HT2A, 5HT2C, 5HT3 and 5HT7 receptors
P2X3 receptors in injured cells but upregulation in intact cells have been found in dorsal root ganglia (Pierce et al., 1996), and
(Tsuzuki et al., 2001; Kennedy et al., 2003). Electrophysiological 5HT receptors are also found in the cytoplasm of Schwann
recordings from DRG neurones supported the idea that cells (Yoder et al., 1997).
axotomy results in an increase in purinergic sensitivity There is a large projection of serotonergic neurones from
(Chen et al., 2001), most likely mediated by P2X receptors the raphe nuclei to laminae 1 and 2 of the spinal cord, and
(Zhou et al., 2001). While these results are suggestive, they do these serotonergic axons contribute to the descending inhibi-
not directly address the issue of neuropathic pain. This was tion of pain (Millan, 2002). While serotonin suppresses
done by reducing the level of expression of P2X3 subunits in nociception at the level of the spinal cord, it enhances it in
DRG cells with intrathecal administration of antisense oligo- the periphery, where serotonin sensitises nociceptors to
nucleotides. In the partial nerve ligation model of neuropathic thermal stimuli and to bradykinin application (Lang et al.,
pain, this inhibited the initiation of mechanical hyperalgesia 1990; Rueff and Dray, 1993). Serotonin also increases the
and reversed established hyperalgesia (Barclay et al., 2002). excitability of sensory C-fibres in isolated segments of
This result was confirmed using a novel non-nucleotide peripheral nerve (Moalem et al., 2005). In the periphery,
antagonist of P2X3 and P2X2/3 activation. Blocking these serotonin is not released as a neurotransmitter by the axon
receptors in the chronic constriction model of neuropathic terminals of serotonergic neurones but is an inflammatory
pain attenuated both thermal and mechanical allodynia mediator released by platelets and murine mast cells (seroto-
(Jarvis et al., 2002). nin is not released by human mast cells). When injected into
P2X4 and P2X7 receptors have also been implicated in the rat's paw, it induces a hyperalgesia thought to be due to an
neuropathic pain. P2X4 receptors are upregulated in spinal action on 5HT1A receptors (Taiwo and Levine, 1992; Hong and
microglia following nerve injury, and pharmacological block- Abbott, 1994).
ade or inhibition of P2X4 receptors reversed the accompa- Work in our laboratory suggested that serotonin con-
nying allodynia (Tsuda et al., 2003; Inoue et al., 2004). P2X7 tributes to established mechanical hyperalgesia in the
receptors are expressed by immune cells, including mast partial ligation model of neuropathic pain. Subcutaneous
cells, macrophages and T lymphocytes. They have a key role injection of a 5HT2A receptor blocker or a 5HT3 receptor
in the secretion of IL-1β. Recently, it was shown that blocker alleviated hyperalgesia in a dose-dependent fashion
disruption of the P2X7 purinoreceptor gene abolishes neuro- when injected into the hindpaw of the nerve-injured limb.
pathic pain and that the receptor is upregulated in human Injection of the blockers into the contralateral hindpaw did
DRGs and injured nerves from neuropathic pain patients not relieve hyperalgesia, suggesting that the effect was
(Chessell et al., 2005). mediated locally rather than systemically (Theodosiou et
P2Y1 and P2Y2 receptors are found in sensory neurones al., 1999). Other research on the involvement of serotonin in
including nociceptors (Molliver et al., 2002; Xiao et al., 2002; neuropathic pain has focussed on its role in the descending
Stucky et al., 2004) and in Schwann cells (Mayer et al., 1998). In inhibition of pain. Tricyclic antidepressants (TCAs) are
normal rats, P2Y1 mRNA is found in about 20% of DRG widely used to treat neuropathic pain in patients and inhibit
neurones (large as well as small), but this percentage increases reuptake of monoamines such as serotonin and noradren-
to about 70% after section of the sciatic nerve, suggesting a aline into nerve terminals. This is thought to increase the
possible role in neuropathic pain (Xiao et al., 2002). Intrathecal efficacy of descending monoaminergic pathways that inhibit
administration of P2Y receptor agonists appears to have an pain. Some TCAs are effective in alleviating neuropathic
inhibitory effect on neuropathic pain, acting via P2Y2 and/or pain in animal models, but selective inhibitors of serotonin
P2Y4 receptors and P2Y6 receptors (Okada et al., 2002). reuptake were not effective in animal models or in patients
Adenosine elicits cutaneous pain and hyperalgesia when (Max et al., 1992; Jett et al., 1997). However, fenfluramine,
applied peripherally (Bleehen and Keele, 1977; Taiwo and which not only inhibits serotonin reuptake but also elicits
Levine, 1990a). It also activates axonal receptors on human C- releases of serotonin from presynaptic terminals, produced a
fibres (Irnich et al., 2002; Lang et al., 2002). However, its role in significant reduction in mechanical allodynia in rats with
the CNS is primarily inhibitory (see Sawynok and Liu, 2003 for tight ligation of the L5/L6 spinal nerves (Wang et al., 1999).
review). Thus, in rats with spinal nerve ligation, spinal Serotonin reuptake is mediated by the serotonin transporter
administration of adenosine resulted in a dose-dependent (5HTT), and mice deficient for the serotonin transporter did
reduction in tactile allodynia (Lavand'homme and Eisenach, not develop thermal hyperalgesia following chronic constric-
1999). Inhibition of adenosine kinase increases the availability tion of the sciatic nerve (Vogel et al., 2003). Whether this was
of extracellular adenosine in the spinal cord, and spinal due to altered tissue levels of serotonin or downregulation of
administration of adenosine kinase inhibitors relieves tactile 5HT receptors was not clear.
allodynia or suppresses neuronal responses in nerve-injured
rats (Suzuki et al., 2001; Zhu et al., 2001). These inhibitors may 3.4. Eicosanoids
therefore hold some promise for the treatment of patients
with neuropathic pain. Arachidonic acid is a polyunsaturated fatty acid that is found
in the cell membrane. It is the major precursor of the
3.3. Serotonin eicosanoids, and, once it has been freed from the cell
membrane by phospholipase A2, it is converted by enzymes
Serotonin (5-hydroxytryptamine, 5HT) is a neurotransmitter in the arachidonate cascade to compounds that include
synthesised and released by neurones in the CNS. Messenger prostaglandins, thromboxanes and leukotrienes.
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 251

Arachidonic acid is metabolised to prostanoids by the in the dorsal horn of the spinal cord and thalamus following
cyclooxygenase pathway and to leukotrienes by the lipox- peripheral nerve injury (Zhao et al., 2000), and the resulting
ygenase pathway (Wolfe and Horrocks, 1994; Smith et al., allodynia is attenuated by intrathecal injection of ketorolac, a
2000). It may also be converted to isoprostanes by peroxida- COX-1 preferring inhibitor (Ma et al., 2002). COX-1 is expressed
tion; the isoprostanes are inflammatory mediators that constitutively in glial cells and motor neurones of normal rats.
augment nociception (Evans et al., 2000). It is also upregulated after ligation of L5–L6 spinal nerves so
Prostaglandins PGE2 and PGI2 induce hyperalgesia in the that the number of cells immunoreactive for COX-1 is
periphery (Taiwo and Levine, 1990b) where they act directly increased in the superficial laminae of the dorsal horn at 4
on the terminals of nociceptors (Taiwo and Levine, 1989) days after spinal nerve ligation and remains high for at least 2
and probably on macrophages as well (Ma and Eisenach, weeks (Zhu and Eisenach, 2003). How do increased levels of
2003a). More recently, it has been established that prosta- prostaglandins in the spinal cord mediate neuropathic pain?
glandins also contribute to nociception at the level of the There is evidence for at least two mechanisms. The first is that
spinal cord (Malmberg and Yaksh, 1992; Vanegas and Schaible, PGE2 depolarises wide dynamic range neurones in the deep
2001). dorsal horn (Baba et al., 2001), and the second is that PGE2 also
There are two primary isoforms of cyclooxygenase, COX-1 blocks glycinergic inhibition of neurones in the superficial
and COX-2. COX-1 is expressed constitutively in most tissues laminae of the dorsal horn (Ahmadi et al., 2002; Harvey et al.,
and has a homeostatic or housekeeping role, while COX-2 2004).
expression is usually low but can be induced by factors such as Given the evidence for a role of cyclooxygenases in the
inflammatory cytokines in cells including mast cells, macro- development of neuropathic pain and the efficacy of cycloox-
phages (Smith et al., 2000) and spinal neurones (Samad et al., ygenase inhibitors in treating inflammatory pain, it is
2002). In fact, COX-1 may also be induced in some conditions surprising that COX inhibitors are relatively ineffective in the
such as spinal cord injury (Schwab et al., 2000; Samad et al., treatment of neuropathic pain in patients (Namaka et al.,
2002). Prostaglandins exert their effects by actions on G- 2004). This may be due to the predominant role of prostaglan-
protein-coupled receptors including the EP receptors (EP1–4 for dins in the development rather than maintenance of neuro-
PGE2) and the IP receptor for PGI2 or prostacyclin (Hata and pathic pain or that higher effective concentrations of COX
Breyer, 2004). inhibitors are achieved in experimental animals than in
Work in our laboratory showed that mechanical hyper- patients.
algesia in nerve-injured rats was alleviated for up to 10 days by There are three mammalian lipoxygenases, which catalyse
subcutaneous injection of indomethacin (a classic inhibitor of the insertion of oxygen at positions 5, 12 and 15 of arachidonic
cyclooxygenase) into the affected hindpaw. Subcutaneous acid. These are therefore referred to as 5-, 12- and 15-
injection of selective COX-2 inhibitors or an EP1 receptor lipoxygenases. The 5-lipoxygenases form leukotrienes, and
blocker relieved thermal as well as mechanical hyperalgesia, leukotriene B4 (LTB4) produces hyperalgesia by eliciting the
but with a shorter timecourse. We concluded that increased release of a mediator from neutrophils (Levine et al., 1984,
expression of prostaglandins in the region of the nerve lesion 1985), which was identified as the 15-lipoxygenase product
contributed to neuropathic pain (Syriatowicz et al., 1999). It (8R, 15S)-dihydroxyeicosa-(5E-9,11,13Z)-tetraenoic acid (8R,
was then shown in several animal models of neuropathic pain 15S-diHETE) (Levine et al., 1986). This was shown to sensitise
that the number of COX-2 immunoreactive cells was dramat- nociceptors in the rat and may contribute to the role of
ically increased in the region of the nerve lesion, that most of neutrophils in neuropathic pain (see the Neutrophils section).
these cells were infiltrating macrophages (Ma and Eisenach, Nerve growth factor elicits hyperalgesia when injected into
2002, 2003b) and that this was the case in human patients as the rat paw, and inhibition of 5-lipoxygenase inhibits release
well as nerve-injured rats (Durrenberger et al., 2004). In fact, of LTB4, accumulation of neutrophils and the associated
the increase in COX-2 expression appears to be biphasic, with hyperalgesia (Amann et al., 1996; Bennett et al., 1998b). It is
an early increase in Schwann cells (1 day after spinal nerve likely that NGF produces hyperalgesia by inducing the release
injury) and an increase in macrophage expression at about 7– of LTB4 from mast cells, leading in turn to recruitment of
14 days (Takahashi et al., 2004). As one might expect, neutrophils (Bennett et al., 1998b), which then release algesic
increased levels of PGE2 are found in the injured nerves mediators and chemokines. Neutrophils and nerve growth
(Schäfers et al., 2004). Furthermore, cells immunoreactive for factor have both been directly implicated in the genesis of
EP receptors are found in the injured nerve, but not in normal neuropathic pain (Herzberg et al., 1997; Theodosiou et al., 1999;
intact nerves. Once again, many of the immunoreactive cells Perkins and Tracey, 2000), and so it is likely that LTB4 and 8R,
prove to be macrophages (Ma and Eisenach, 2003a). These 15S-diHETE are also involved.
observations, based on several animal models of sciatic nerve
injury, support the idea that upregulation of COX-2 and EP 3.5. Cytokines
receptors in the injured nerve contributes to neuropathic pain.
However, in the spared nerve injury model (Decosterd and Cytokines are small proteins that mediate interactions
Woolf, 2000), where degenerating axons distal to the lesion do between cells over relatively short distances. They are mostly
not mingle with intact axons, treatment with the COX-2 involved in responses to disease or infection. Many of them
inhibitor rofecoxib had no effect on the development of are referred to as interleukins, denoting a mediator released
allodynia and hyperalgesia (Broom et al., 2004). by one leukocyte and acting on another, but they are
Increased expression of cyclooxygenase after nerve injury synthesised by most cell types. Several are pro-inflammato-
is not restricted to the nerve itself. COX-2 is also upregulated ry, such as IL-1β, IL-6 and TNF, while others such as IL-10 are
252 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

anti-inflammatory. Exogenous administration of these pro- 3.5.2. Interleukin-6


inflammatory cytokines elicits pain and hyperalgesia (Som- IL-6 is synthesised by many cell types, including mast cells,
mer and Kress, 2004), but they do not appear to be involved monocytes, lymphocytes, neurones and glial cells. Once IL-6
in ‘normal’ pain (Watkins et al., 2001). In fact, these three has bound to its receptor IL-6R, it initiates two major
pro-inflammatory cytokines induce the production of each intracellular cascades—one involving Janus kinases and
other, and they act synergistically (Watkins et al., 1999). This STAT factors, the other using the Ras-dependent MAP kinase
leads to the possibility of positive feedback, which could lead pathway (De Jongh et al., 2003). Injury of the sciatic nerve
to chronic pain if not appropriately regulated. Other cyto- induces upregulation of IL-6 and its receptor in the region of
kines such as IL-4 and IL-10 are anti-inflammatory and the lesion, where it appears in macrophages and Schwann
suppress genes that code for IL-1, TNF and chemokines. The cells (Bolin et al., 1995; Kurek et al., 1996; Ito et al., 1998; Grothe
algesic effects of pro-inflammatory cytokines are often et al., 2000). Upregulation of IL-6 in these macrophages
indirect, so that they may not act directly on the nociceptor appears to be induced by PGE2 (Ma and Quirion, 2005). Sciatic
but instead induce the expression of agents (such as PGE2) nerve injury also increases IL-6 levels in the DRG and spinal
that themselves sensitise nociceptors. In inflammation and cord, particularly in the superficial laminae of the dorsal horn
Wallerian degeneration, cytokine production is organised (Murphy et al., 1995; DeLeo et al., 1996; Lee et al., 2004).
into a sequence with TNF playing a leading role (Shamash et Exogenous IL-6 induces thermal hyperalgesia when injected
al., 2002; Cunha et al., 2005). This sequence appears to be one into the lateral cerebral ventricles of the rat (Oka et al., 1995)
of the mechanisms underlying neuropathic pain (for reviews, and increases the heat-evoked release of CGRP from cutane-
see DeLeo and Colburn, 1999; Sommer, 2001; Sorkin, 2002; ous nociceptors (Opree and Kress, 2000; Obreja et al., 2002).
Sommer and Kress, 2004). Chemokines are small chemotactic However, there is little evidence that IL-6 plays a role in
cytokines that are important for leukocyte migration and normal nociception. One study did report that intraplantar
recruitment to damaged sites. Recently, they have been injection of IL-6 induced mechanical hyperalgesia in naive rats
shown to contribute directly to nociception by producing (Cunha et al., 1992). However, this finding has not been
excitatory effects on DRG neurones and inducing allodynia reproduced and later work found that intraplantar IL-6 had no
after injection into the rat's paw (Oh et al., 2001). The effect on mechanical thresholds (Czlonkowski et al., 1993) or
contribution of chemokines to pain has been reviewed produced a thermal hypoalgesia (Flatters et al., 2004). In fact,
recently (Abbadie, 2005; White et al., 2005). peripheral IL-6 inhibited heat responses of nociceptors in an in
vitro preparation.
3.5.1. Interleukin-1 There is limited evidence that IL-6 contributes to the
Binding of IL-1β to its receptor IL1-RI on the cell surface development of neuropathic pain. In the in vivo spinal cord,
initiates several signalling events, such as translocation of NF- peripheral IL-6 inhibited the responses of dorsal horn neu-
κB into the nucleus (Dinarello, 1999). NF-κB then upregulates rones to thermal and mechanical stimuli; after spinal nerve
transcription of several genes, including COX-2, inducible ligation, peripheral IL-6 released the inhibition of mechanical
nitric oxide synthase, TNF, IL-1β and IL-6 (Pahl, 1999; Tegeder responses (Flatters et al., 2004). In IL-6 knockout mice,
et al., 2004). IL-1β may act directly as well as indirectly on nociceptive responses to thermal and mechanical stimuli are
nociceptors. Thus, Fukuoka and colleagues found that IL-1β the same as in wild-type mice (Bianchi et al., 1999; Murphy et
facilitated release of CGRP from nociceptors in an in vitro al., 1999), but the development of mechanical allodynia after
nerve-skin preparation. The latency of the effect was too short spinal nerve lesion is delayed in IL-6 knockouts by comparison
for upregulation of receptors or changes in gene expression, with wild-type mice (Ramer et al., 1998). It is not clear whether
and the cell bodies were not present, suggesting a direct thermal hyperalgesia induced by a spinal nerve lesion is
sensitisation of nociceptors (Fukuoka et al., 1994; Sommer and reduced in IL-6 knockouts (Murphy et al., 1999) or not (Ramer
Kress, 2004). IL-1β elicits hyperalgesia when injected periph- et al., 1998).
erally into the rat paw (Ferreira et al., 1988; Follenfant et al.,
1989), intraneurally into rat sciatic nerve (Zelenka et al., 2005), 3.5.3. Tumour necrosis factor
intrathecally in the rat spinal cord (Sung et al., 2004) or Tumour necrosis factor (TNF, TNFSF2, formerly TNFα) is a
centrally into various regions of the brain (see Bianchi et al., member of a large superfamily of proteins, which have an
1998 for review). In fact, Watkins and her colleagues have unusual trifold symmetry. There is an equally large super-
shown that IL-1β may induce an ‘illness’ hyperalgesia via a family of receptors; the receptors activated by TNF are the
kind of supraspinal loop involving sensory fibres in the constitutively expressed TNFR1 (TNFRSF1A, p55-R) and the
vagus nerve, the nucleus of the solitary tract, nucleus raphe inducible TNFR2 (p75-R) (Locksley et al., 2001). TNFR1 is linked
magnus and projections to the spinal cord (Watkins et al., to pathways for cell death, whereas TNFR2 is not. However,
1994; Watkins and Maier, 1999). IL-1 is implicated in activation of either receptor results in p38 MAP kinase
neuropathic pain since IL-1α and IL-1β are both upregulated signalling (Schäfers et al., 2003b), translocation of NFκB to
in injured peripheral nerve (Gillen et al., 1998; Okamoto et the nucleus and activation of COX-2-dependent prostanoid
al., 2001; Shamash et al., 2002). Neuropathic pain is release, as described above for IL-1 (Dinarello, 1999). TNF is
alleviated in nerve-injured mice by administration of neu- constitutively expressed in cutaneous mast cells (Walsh et al.,
tralizing antibodies to the IL-1 receptor (Sommer et al., 1999; 1991), but, in injury or inflammation, it may be released by
Schäfers et al., 2001) and in nerve-injured rats by intrathecal other cells including neutrophils and macrophages. Intraplan-
IL-1 receptor antagonist in combination with soluble TNF tar injection of TNF into the rat's paw induces mechanical
receptor (Sweitzer et al., 2001). hyperalgesia (Cunha et al., 1992). This can be attributed to
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 253

sensitisation of C-fibres since subcutaneous injection in the Glial-cell-line-derived neurotrophic factor (GDNF) also has
distribution of the sural nerve lowered the thresholds of neurotrophic properties but is structurally unrelated to the
cutaneous C-fibres to mechanical stimulation and led to dimeric neurotrophin family (Lewin and Barde, 1996).
ongoing activity in some of these fibres (Junger and Sorkin, The neurotrophins act on a trio of tyrosine kinase (Trk)
2000). In fact, topical application of TNF to a restricted portion receptors U TrkA, TrkB and TrkC, which primarily bind NGF,
of the sciatic nerve led to ectopic firing of Aδ- and C-fibres BDNF and NT-4/5, and NT-3 respectively. There is also a low
(Sorkin et al., 1997), and intraneural injection of TNF into rat affinity or pan-neurotrophin receptor, p75NTR, which acti-
sciatic nerve at physiological doses induced thermal hyper- vates a separate set of signalling pathways that interact with
algesia and mechanical allodynia (Zelenka et al., 2005), those activated by Trk receptors (Huang and Reichardt, 2003;
suggesting the presence of TNF receptors on the axons. An Nykjaer et al., 2005). During embryonic development, all
alternative possibility is insertion of the TNF trimer into the small-diameter sensory neurones express TrkA, the high
cell membrane of the axon (Baldwin et al., 1996), creating a affinity receptor for NGF. Most of these neurones will become
pore that is permeable to sodium ions and may be voltage- nociceptors, with thinly myelinated axons (Aδ fibres) or
dependent (Kagan et al., 1992). unmyelinated axons (C-fibres). After birth, the expression of
Injury of the sciatic nerve leads to upregulation of TNF and TrkA is downregulated so that TrkA is only expressed by those
its receptors in the nerve (George et al., 1999; Shubayev and small sensory neurones which are peptidergic, whereas the
Myers, 2000; George et al., 2005); this upregulation is found non-peptidergic remainder (IB4 binding neurones) now
mainly in Schwann cells and endothelial cells (Wagner and expresses receptor components for another neurotrophin,
Myers, 1996a). Nerve injury also leads to increased TNF GDNF (Bennett, 2001).
expression in the dorsal horn of the spinal cord and in the Neurotrophins are synthesised and released by a several
locus coeruleus and hippocampus (Ignatowski et al., 1999). An cell types of immune cells, including mast cells and lympho-
important role for TNF in neuropathic pain is indicated by cytes (Moalem et al., 2000; Bonini et al., 2003). BDNF is
reducing TNF levels in nerve-injured rats or mice (Sommer, constitutively expressed by peptidergic nociceptors (Thomp-
2001). Inhibiting TNF synthesis with thalidomide or treatment son et al., 1999).
with anti-TNF neutralising antibodies at the time of nerve Nerve injury produces marked changes in expression of
injury blocked the development of hyperalgesia and allodynia neurotrophins and their receptors, primarily in Schwann cells
in these animal models (Sommer et al., 1998a,b, 2001a). Treat- (Frostick et al., 1998). Schwann cells in the intact peripheral
ment with etanercept, a recombinant TNF receptor (p75)-Fc nerve synthesise NGF, BDNF and GDNF, and this synthesis is
fusion protein that acts as a TNF antagonist, reversed estab- dramatically upregulated by nerve injury (Lindholm et al.,
lished hyperalgesia in mice with a chronic constriction injury 1987; Meyer et al., 1992; Hammarberg et al., 1996). BDNF
of the sciatic nerve (Sommer et al., 2001b), and endoneurial synthesis is also upregulated in sensory neurones following
injection of neutralising antibodies against TNF receptors axotomy (Tonra et al., 1998). In the intact sciatic nerve, mRNAs
showed that neuropathic hyperalgesia is dependent on the for the neurotrophin receptors TrkB and TrkC are expressed,
TNFR1 receptor (Sommer et al., 1998b). Furthermore, applica- but TrkA and p75NTR mRNAs are undetectable. Levels of
tion of exogenous TNF to the DRG increased the sensitivity of p75NTR mRNA are markedly increased in Schwann cells after
injured and adjacent uninjured primary sensory neurones and nerve injury, but TrkA levels remain undetectable (Funakoshi
elicited faster onset of allodynia and spontaneous pain et al., 1993).
behaviour after spinal nerve ligation (Schäfers et al., 2003a). It is now recognised that some neurotrophins (in particular,
NGF and BDNF) play a significant role in nociception (Bennett,
3.5.4. Interleukin-10 2001) so that NGF sensitises nociceptors in the periphery
IL-10 is generally regarded as anti-inflammatory. Its expres- (Koltzenburg et al., 1999a) while BDNF increases the respon-
sion increases gradually over at least 6 weeks following nerve siveness of dorsal horn neurones in the spinal cord (Pezet et
injury (Okamoto et al., 2001), and treatment with a single dose al., 2002). Intraplantar or systemic injection of NGF induces
of IL-10 at the site of a chronic constriction injury significantly hyperalgesia (Lewin et al., 1994; Andreev et al., 1995; Theodo-
reduces hyperalgesia, probably due in part to suppression of siou et al., 1999). Tissue levels of NGF are increased by
TNF expression and macrophage recruitment (Wagner et al., inflammation, and the hyperalgesia resulting from inflamma-
1998). These results are consistent with findings that thalid- tion is prevented by administration of anti-NGF neutralising
omide (an inhibitor of TNF synthesis) decreases TNF levels, antibodies or a TrkA-IgG fusion molecule (Woolf et al., 1994;
increases endoneurial IL-10 levels, and alleviates hyperalgesia McMahon et al., 1995). Some of the hyperalgesic action of NGF
in rats with a CCI nerve lesion (Sommer et al., 1998a; George et is exerted directly on sensory neurones, but indirect actions
al., 2000). Recent work confirms the efficacy of IL-10 in alleviat- via mast cells, neutrophils and sympathetic neurones also
ing neuropathic pain, using virally driven spinal production of make a significant contribution (Lewin et al., 1994; Woolf et al.,
IL-10 in neuropathic pain models (Milligan et al., 2005a,b). 1996; Bennett et al., 1998b).
NGF contributes to the development and maintenance of
3.6. Neurotrophins neuropathic pain in animal models. In rats with a chronic
constriction injury of the sciatic nerve, the onset of hyper-
The neurotrophins are dimeric proteins that are essential for algesia was delayed by application of anti-serum to NGF at
the normal development of the vertebrate nervous system. the site of injury (Herzberg et al., 1997), and thermal
This family includes nerve growth factor (NGF), brain-derived hyperalgesia was reduced or even reversed by injection of
neurotrophic factor (BDNF), neurotrophin (NT)-3 and NT-4/5. anti-serum into the medial dorsal surface of the nerve-injured
254 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

hindlimb (Ro et al., 1999). Established hyperalgesia (thermal thecal GDNF prevents and reverses hyperalgesia following
and mechanical) was reduced by systemic injection of anti- nerve injury (Boucher et al., 2000; Nagano et al., 2003; Wang et
serum to NGF in rats with partial ligation of the sciatic nerve al., 2003). It was suggested that this inhibition of hyperalgesia
(Theodosiou et al., 1999). As mentioned earlier, neuropathic is due to suppression of Nav1.3 expression by injured fibres
pain can develop without injury of sensory neurones and can (Boucher et al., 2000), but other factors such as downregulation
be produced by transection of ventral roots (Li et al., 2002; of the P2X3 purinergic receptor may also be involved (Wang et
Sheth et al., 2002). NGF plays a significant role in this model of al., 2003). Likewise, systemic administration of artemin, a
neuropathic pain too since treatment with NGF anti-serum member of the GDNF family, reverses the behavioural
dramatically attenuated the mechanical allodynia (Li et al., manifestations of neuropathic pain as well as the associated
2003). Just how NGF produces neuropathic hyperalgesia is not morphological and biochemical changes produced by nerve
yet clear. Expression of NGF was increased in small sensory injury (Gardell et al., 2003).
neurones, DRG satellite neurones and keratinocytes in target
tissues of the injured nerve in the ventral root transection 3.7. Nitric oxide and reactive oxygen species
model (Li et al., 2003), and these increased levels of
endogenous NGF probably increase the sensitivity of primary Reactive oxygen species such as nitric oxide and superoxide
afferent nociceptors to thermal and chemical stimuli (Bennett play important roles in inflammatory and immune responses,
et al., 1998a). It is also possible that increased levels of NGF including defense mechanisms against invading microbes
increase the sensitivity of dorsal horn neurones by releasing (Guzik et al., 2003; Roos et al., 2003). They are released by a
larger amounts of neuromodulators such as substance P and number of cell types, including neutrophils (Witko-Sarsat et
CGRP from the central terminals of primary afferents (Ma et al., 2000) and macrophages (Bosca et al., 2005) as well as
al., 1995). astrocytes (Hu et al., 1995) and microglia (Vilhardt, 2005). Nitric
BDNF also contributes to neuropathic pain. It is co- oxide is a diffusible free radical that is synthesised by three
expressed with substance P in 20–30% of sensory neurones distinct nitric oxide synthases (NOS), of which the neuronal
in the dorsal root ganglia, primarily in the peptidergic and endothelial forms (nNOS and eNOS) are constitutive,
subgroup (Luo et al., 2001). BDNF is anterogradely transported while the inducible form (iNOS) is upregulated in immune
from the cell bodies of these sensory neurones in the dorsal cells. Once released, nitric oxide can react with superoxide
root ganglia to their terminals in the dorsal horn (Zhou and radicals to form peroxynitrite, which is toxic and may cause
Rush, 1996), and this anterograde transport is increased by tissue damage.
nerve injury (Tonra et al., 1998). Intrathecal administration of The role of nitric oxide in nociception was reviewed
BDNF induces thermal hyperalgesia and tactile allodynia in recently (Luo and Cizkova, 2000). It elicits pain when injected
mice (Groth and Aanonsen, 2002; Yajima et al., 2005), while into the skin of human subjects (Holthusen and Arndt, 1994,
intrathecal injection of antisense oligonucleotides for BDNF or 1995) and contributes to peripheral hyperalgesia in the skin
its receptor, TrkB, attenuates inflammatory hyperalgesia (Aley et al., 1998) and joints (Lawand et al., 1997), probably by
(Groth and Aanonsen, 2002). In animal models of neuropathic contributing to PGE2-induced sensitisation of primary affer-
pain, nerve injury increased levels of BDNF in the dorsal horn ents (Aley et al., 1998). It may also inhibit inflammatory pain
(Yajima et al., 2005), and these levels were related to the extent (Duarte et al., 1992), although the basis for these opposing
of thermal hyperalgesia (Miletic and Miletic, 2002). The actions is not clear (Levy and Zochodne, 2004). Nitric oxide is
thermal hyperalgesia and tactile allodynia induced by nerve also implicated in central mechanisms of hyperalgesia (Meller
injury were completely suppressed by repeated intrathecal et al., 1992; Lin et al., 1999a) where nNOS and nitric oxide form
injection of an antibody to BDNF or a TrkB/Fc chimera protein part of a second messenger cascade involving cyclic GMP and
which sequesters endogenous BDNF (Yajima et al., 2002, 2005). partly responsible for sensitisation of spinal neurones (Meller
Furthermore, heterozygous BDNF knockout mice developed and Gebhart, 1993; Luo and Cizkova, 2000; Riedel and Neeck,
significantly less hyperalgesia following nerve injury than 2001; Sung et al., 2004). Nitric oxide may also contribute to
wild-type mice (Yajima et al., 2005). However, experiments in sensitisation of central neurones by disinhibition (Lin et al.,
which over-expression of BDNF in the spinal cord was 1999b). However, nitric oxide can contribute to central
generated by a recombinant adeno-associated viral vector mechanisms of analgesia as well as nociception since it
showed an amelioration of neuropathic pain resulting from a enhances opioid analgesia in animal models (Kawabata et al.,
chronic constriction injury of the sciatic nerve (Eaton et al., 1993) and in patients (Lauretti et al., 2002). Furthermore,
2002). It is difficult to reconcile this result with evidence cited NSAIDs linked to an NO-releasing moiety are more effective in
above that BDNF promotes neuropathic pain, although there is alleviating inflammatory hyperalgesia than NSAIDs alone,
independent evidence that BDNF exerts anti-nociceptive although the rationale is not clearly understood (Keeble and
effects (Siuciak et al., 1994). The authors suggest that BDNF Moore, 2002).
may alleviate neuropathic pain by preventing the loss of Nitric oxide is implicated in neuropathic pain (see Levy and
GABAergic interneurones or enhancing GABA synthesis in the Zochodne, 2004 for review). In rats with a chronic constriction
dorsal horn (Eaton et al., 2002). injury of the sciatic nerve, iNOS is induced in macrophages
Other neurotrophins appear to inhibit neuropathic pain. and Schwann cells at the injury site and distal to it (Levy and
NT-3 suppresses thermal but not mechanical hyperalgesia Zochodne, 1998; Levy et al., 1999). In the same model of
following a chronic constriction injury of the sciatic nerve; this neuropathic pain, eNOS could be demonstrated in endbulb-
suppression is associated with decreased expression of the like structures of injured peripheral axons, while treatment
TRPV1 vanilloid receptor (Wilson-Gerwing et al., 2005). Intra- with a non-specific NOS inhibitor (L -NAME) alleviated
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 255

hyperalgesia and blocked ectopic mechanosensitivity of S., Bayne, E.K., DeMartino, J.A., MacIntyre, D.E., Forrest, M.J.,
injured A-fibres (Levy et al., 2000). As is the case for 2003. Impaired neuropathic pain responses in mice lacking the
chemokine receptor CCR2. Proc. Natl. Acad. Sci. U. S. A. 100,
inflammatory hyperalgesia, nitric oxide also plays a role in
7947–7952.
central mechanisms of neuropathic pain so that, in nerve-
Ahmadi, S., Lippross, S., Neuhuber, W.L., Zeilhofer, H.U., 2002. PGE
injured rats, intrathecal delivery of the NOS inhibitor L-NAME (2) selectively blocks inhibitory glycinergic neurotransmission
produced a dose-dependent reduction of thermal hyperalgesia onto rat superficial dorsal horn neurons. Nat. Neurosci. 5,
(Lui and Lee, 2004). 34–40.
Other reactive oxygen species are implicated in neuro- Aicher, S.A., Silverman, M.B., Winkler, C.W., Bebo Jr., B.F., 2004.
pathic pain. While nitric oxide synthase catalyses the Hyperalgesia in an animal model of multiple sclerosis. Pain
110, 560–570.
production of nitric oxide, xanthine oxidase catalyses the
Aley, K.O., McCarter, G., Levine, J.D., 1998. Nitric oxide signaling in
production of another free radical, superoxide. Under physi- pain and nociceptor sensitization in the rat. J. Neurosci. 18,
ological conditions, superoxide and nitric oxide react to form 7008–7014.
peroxynitrite, which is a powerful oxidant and potent Amann, R., Schuligoi, R., Lanz, I., Peskar, B.A., 1996. Effect of a
cytotoxin. Xanthine oxidase levels are increased in the injured 5-lipoxygenase inhibitor on nerve growth factor-induced
sciatic nerve (Khalil et al., 1999), and treatment with superox- thermal hyperalgesia in the rat. Eur. J. Pharmacol. 306,
89–91.
ide dismutase or TEMPOL (both of which degrade the
Anderson, C.M., Bergher, J.P., Swanson, R.A., 2004. ATP-induced
superoxide radical) alleviates thermal hyperalgesia resulting
ATP release from astrocytes. J. Neurochem. 88, 246–256.
from a chronic constriction of the rat sciatic nerve (Tal, 1996; Andreev, N., Dimitrieva, N., Koltzenburg, M., McMahon, S.B., 1995.
Khalil et al., 1999). Reactive oxygen species also contribute to Peripheral administration of nerve growth factor in the adult
mechanical allodynia, which is relieved by superoxide dis- rat produces a thermal hyperalgesia that requires the
mutase in an inflammatory model of neuropathic pain presence of sympathetic post-ganglionic neurones. Pain 63,
(Twining et al., 2004) and by potent scavengers of reactive 109–115.
Araki, T., Sasaki, Y., Milbrandt, J., 2004. Increased nuclear NAD
oxygen species in the spinal nerve ligation model (Kim et al.,
biosynthesis and SIRT1 activation prevent axonal
2004). Peroxynitrite also contributes to neuropathic hyperal-
degeneration. Science 305, 1010–1013.
gesia: after partial ligation of the sciatic nerve, peroxynitrite Araque, A., Carmignoto, G., Haydon, P.G., 2001. Dynamic signaling
levels are increased in macrophages and Schwann cells in the between astrocytes and neurons. Annu. Rev. Physiol. 63,
injured nerve (as indicated by immunoreactivity for nitrotyr- 795–813.
osine), while treatment with uric acid, a scavenger of Argall, K.G., Armati, P.J., Pollard, J.D., Watson, E., Bonner, J., 1992.
peroxynitrite, decreased nitrotyrosine formation and alleviat- Interactions between CD4+ T-cells and rat Schwann cells in
vitro. 1. Antigen presentation by Lewis rat Schwann cells to P2-
ed thermal hyperalgesia (Liu et al., 2000a).
specific CD4+ T-cell lines. J. Neuroimmunol. 40, 1–18.
Arruda, J.L., Colburn, R.W., Rickman, A.J., Rutkowski, M.D., DeLeo,
J.A., 1998. Increase of interleukin-6 mRNA in the spinal cord
4. Conclusions following peripheral nerve injury in the rat: potential role
of IL-6 in neuropathic pain. Brain Res. Mol. Brain Res. 62,
Our understanding of the involvement of immune cells and 228–235.
Baba, H., Kohno, T., Moore, K.A., Woolf, C.J., 2001. Direct activation
inflammatory mediators in neuropathic pain has greatly
of rat spinal dorsal horn neurons by prostaglandin E2.
increased in the last decade. Work on animal models of
J. Neurosci. 21, 1750–1756.
nerve injury has implicated complex neuronal changes in Baldwin, R.L., Stolowitz, M.L., Hood, L., Wisnieski, B.J., 1996.
response to the trauma. However, recent studies have Structural changes of tumor necrosis factor alpha associated
emphasised the important contribution of inflammatory with membrane insertion and channel formation. Proc. Natl.
cells and immune-like glial cells, as well as their mediators, Acad. Sci. U. S. A. 93, 1021–1026.
to neuropathic pain. This review summarises current views on Barclay, J., Patel, S., Dorn, G., Wotherspoon, G., Moffatt, S., Eunson,
L., Abdel'al, S., Natt, F., Hall, J., Winter, J., Bevan, S., Wishart, W.,
immune and inflammatory modulation of the neural response
Fox, A., Ganju, P., 2002. Functional downregulation of P2X3
to injury, leading to chronic pain. It highlights activation of
receptor subunit in rat sensory neurons reveals a significant
inflammatory cells in the peripheral nervous system and of role in chronic neuropathic and inflammatory pain. J. Neurosci.
glial cells in the spinal cord, and their subsequent production 22, 8139–8147.
of inflammatory mediators and cytokines that can sensitise Baron, R., Schwarz, K., Kleinert, A., Schattschneider, J., Wasner, G.,
neurones and contribute to pain hypersensitivity. Our current 2001. Histamine-induced itch converts into pain in
knowledge and continuing progress in the field of neuroim- neuropathic hyperalgesia. NeuroReport 12, 3475–3478.
Bendszus, M., Stoll, G., 2003. Caught in the act: in vivo mapping of
munology will ultimately lead to new possibilities for thera-
macrophage infiltration in nerve injury by magnetic resonance
peutic intervention in neuropathic pain, based on targeting imaging. J. Neurosci. 23, 10892–10896.
the inflammatory response that follows nerve trauma or Bennett, D.L., 2001. Neurotrophic factors: important regulators of
injury in the peripheral or central nervous system. nociceptive function. Neuroscientist 7, 13–17.
Bennett, G.J., Xie, Y.K., 1988. A peripheral mononeuropathy in rat
that produces disorders of pain sensation like those seen in
REFERENCES man. Pain 33, 87–107.
Bennett, D.L., Koltzenburg, M., Priestley, J.V., Shelton, D.L.,
McMahon, S.B., 1998a. Endogenous nerve growth factor
Abbadie, C., 2005. Chemokines, chemokine receptors and pain. regulates the sensitivity of nociceptors in the adult rat. Eur. J.
Trends Immunol. 26, 529–534. Neurosci. 10, 1282–1291.
Abbadie, C., Lindia, J.A., Cumiskey, A.M., Peterson, L.B., Mudgett, J. Bennett, G., al-Rashed, S., Hoult, J.R., Brain, S.D., 1998b. Nerve
256 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

growth factor induced hyperalgesia in the rat hind paw is of peri-axonal inflammation in the development of thermal
dependent on circulating neutrophils. Pain 77, 315–322. hyperalgesia and guarding behavior in a rat model of
Bergsteinsdottir, K., Kingston, A., Mirsky, R., Jessen, K.R., 1991. Rat neuropathic pain. Neurosci. Lett. 184, 5–8.
Schwann cells produce interleukin-1. J. Neuroimmunol. 34, Colburn, R.W., Rickman, A.J., DeLeo, J.A., 1999. The effect of site
15–23. and type of nerve injury on spinal glial activation and
Bergsteinsdottir, K., Kingston, A., Jessen, K.R., 1992. Rat Schwann neuropathic pain behavior. Exp. Neurol. 157, 289–304.
cells can be induced to express major histocompatibility Constable, A.L., Armati, P.J., Toyka, K.V., Hartung, H.P., 1994.
complex class II molecules in vivo. J. Neurocytol. 21, 382–390. Production of prostanoids by Lewis rat Schwann cells in vitro.
Bianchi, M., Dib, B., Panerai, A.E., 1998. Interleukin-1 and Brain Res. 635, 75–80.
nociception in the rat. J. Neurosci. Res. 53, 645–650. Cook, S.P., McCleskey, E.W., 2002. Cell damage excites nociceptors
Bianchi, M., Maggi, R., Pimpinelli, F., Rubino, T., Parolaro, D., Poli, through release of cytosolic ATP. Pain 95, 41–47.
V., Ciliberto, G., Panerai, A.E., Sacerdote, P., 1999. Presence of a Cook, S.P., Vulchanova, L., Hargreaves, K.M., Elde, R., McCleskey, E.
reduced opioid response in interleukin-6 knock out mice. Eur. J. W., 1997. Distinct ATP receptors on pain-sensing and stretch-
Neurosci. 11, 1501–1507. sensing neurons. Nature 387, 505–508.
Bleehen, T., Keele, C.A., 1977. Observations on the algogenic Correale, J., Villa, A., 2004. The neuroprotective role of
actions of adenosine compounds on the human blister base inflammation in nervous system injuries. J. Neurol. 251,
preparation. Pain 3, 367–377. 1304–1316.
Bolin, L.M., Verity, A.N., Silver, J.E., Shooter, E.M., Abrams, J.S., Cui, J.G., Holmin, S., Mathiesen, T., Meyerson, B.A., Linderoth, B.,
1995. Interleukin-6 production by Schwann cells and induction 2000. Possible role of inflammatory mediators in tactile
in sciatic nerve injury. J. Neurochem. 64, 850–858. hypersensitivity in rat models of mononeuropathy. Pain 88,
Bonini, S., Rasi, G., Bracci-Laudiero, M.L., Procoli, A., Aloe, L., 2003. 239–248.
Nerve growth factor: neurotrophin or cytokine? Int. Arch. Cunha, F.Q., Poole, S., Lorenzetti, B.B., Ferreira, S.H., 1992. The
Allergy Immunol. 131, 80–84. pivotal role of tumour necrosis factor alpha in the
Bosca, L., Zeini, M., Traves, P.G., Hortelano, S., 2005. Nitric oxide development of inflammatory hyperalgesia. Br. J. Pharmacol.
and cell viability in inflammatory cells: a role for NO in 107, 660–664.
macrophage function and fate. Toxicology 208, 249–258. Cunha, T.M., Verri Jr., W.A., Silva, J.S., Poole, S., Cunha, F.Q.,
Boucher, T.J., Okuse, K., Bennett, D.L., Munson, J.B., Wood, J.N., Ferreira, S.H., 2005. A cascade of cytokines mediates
McMahon, S.B., 2000. Potent analgesic effects of GDNF in mechanical inflammatory hypernociception in mice. Proc.
neuropathic pain states. Science 290, 124–127. Natl. Acad Sci. U. S. A. 102, 1755–1760.
Brack, A., Rittner, H.L., Machelska, H., Leder, K., Mousa, S.A., Czlonkowski, A., Stein, C., Herz, A., 1993. Peripheral mechanisms
Schafer, M., Stein, C., 2004. Control of inflammatory pain by of opioid antinociception in inflammation: involvement of
chemokine-mediated recruitment of opioid-containing cytokines. Eur. J. Pharmacol. 242, 229–235.
polymorphonuclear cells. Pain 112, 229–238. Decosterd, I., Woolf, C.J., 2000. Spared nerve injury: an animal
Bradbury, E.J., Burnstock, G., McMahon, S.B., 1998. The expression model of persistent peripheral neuropathic pain. Pain 87,
of P2X3 purinoreceptors in sensory neurons: effects of axotomy 149–158.
and glial-derived neurotrophic factor. Mol. Cell. Neurosci. 12, DeLeo, J.A., Colburn, R.W., 1999. Proinflammatory cytokines
256–268. and glial cell: their role in neuropathic pain. In: Watkins,
Bridges, D., Thompson, S.W., Rice, A.S., 2001. Mechanisms of L.R., Maier, S.F. (Eds.), Cytokines and Pain. Birkhäuser,
neuropathic pain. Br. J. Anaesth. 87, 12–26. Basel, pp. 159–181.
Broom, D.C., Samad, T.A., Kohno, T., Tegeder, I., Geisslinger, G., DeLeo, J.A., Yezierski, R.P., 2001. The role of neuroinflammation
Woolf, C.J., 2004. Cyclooxygenase 2 expression in the spared and neuroimmune activation in persistent pain. Pain 90, 1–6.
nerve injury model of neuropathic pain. Neuroscience 124, DeLeo, J.A., Colburn, R.W., Nichols, M., Malhotra, A., 1996.
891–900. Interleukin-6-mediated hyperalgesia/allodynia and increased
Brück, W., 1997. The role of macrophages in Wallerian spinal IL-6 expression in a rat mononeuropathy model. J.
degeneration. Brain Pathol. 7, 741–752. Interferon Cytokine Res. 16, 695–700.
Burnstock, G., Knight, G.E., 2004. Cellular distribution and DeLeo, J.A., Colburn, R.W., Rickman, A.J., 1997. Cytokine and
functions of P2 receptor subtypes in different systems. Int. Rev. growth factor immunohistochemical spinal profiles in two
Cytol. 240, 31–304. animal models of mononeuropathy. Brain Res. 759, 50–57.
Burnstock, G., Wood, J.N., 1996. Purinergic receptors: their role in DeLeo, J.A., Tanga, F.Y., Tawfik, V.L., 2004. Neuroimmune
nociception and primary afferent neurotransmission. Curr. activation and neuroinflammation in chronic pain and opioid
Opin. Neurobiol. 6, 526–532. tolerance/hyperalgesia. Neuroscientist 10, 40–52.
Chacur, M., Milligan, E.D., Gazda, L.S., Armstrong, C., Wang, H., De Jongh, R.F., Vissers, K.C., Meert, T.F., Booij, L.H., De Deyne, C.S.,
Tracey, K.J., Maier, S.F., Watkins, L.R., 2001. A new model of Heylen, R.J., 2003. The role of interleukin-6 in nociception and
sciatic inflammatory neuritis (SIN): induction of unilateral pain. Anesth. Analg. 96, 1096–1103 (table of contents).
and bilateral mechanical allodynia following acute Dellemijn, P., 1999. Are opioids effective in relieving neuropathic
unilateral peri-sciatic immune activation in rats. Pain 94, pain? Pain 80, 453–462.
231–244. Di Marzo, V., Blumberg, P.M., Szallasi, A., 2002. Endovanilloid
Chen, Y., Zhang, Y.H., Zhao, Z.Q., 2001. Novel purinergic sensitivity signaling in pain. Curr. Opin. Neurobiol. 12, 372–379.
develops in injured sensory axons following sciatic nerve Dinarello, C.A., 1999. Overview of inflammatory cytokines and
transection in rat. Brain Res. 911, 168–172. their role in pain. In: Watkins, L.R., Maier, S.F. (Eds.), Cytokines
Chessell, I.P., Hatcher, J.P., Bountra, C., Michel, A.D., Hughes, J.P., and Pain. Birkhäuser Verlag, Basel, pp. 1–19.
Green, P., Egerton, J., Murfin, M., Richardson, J., Peck, W.L., Dray, A., 1995. Inflammatory mediators of pain. Br. J. Anaesth. 75,
Grahames, C.B., Casula, M.A., Yiangou, Y., Birch, R., Anand, P., 125–131.
Buell, G.N., 2005. Disruption of the P2X(7) purinoceptor gene Dray, A., 1997. Kinins and their receptors in hyperalgesia. Can. J.
abolishes chronic inflammatory and neuropathic pain. Pain Physiol. Pharmacol. 75, 704–712.
114, 386–396. Dray, A., Perkins, M., 1993. Bradykinin and inflammatory pain.
Chizh, B.A., Illes, P., 2001. P2X receptors and nociception. Trends Neurosci. 16, 99–104.
Pharmacol. Rev. 53, 553–568. Duarte, I.D., dos Santos, I.R., Lorenzetti, B.B., Ferreira, S.H., 1992.
Clatworthy, A.L., Illich, P.A., Castro, G.A., Walters, E.T., 1995. Role Analgesia by direct antagonism of nociceptor sensitization
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 257

involves the arginine-nitric oxide-cGMP pathway. Eur. J. neuropathy decreases endoneurial tumor necrosis
Pharmacol. 217, 225–227. factor-alpha, increases interleukin-10 and has long-term
Durrenberger, P.F., Facer, P., Gray, R.A., Chessell, I.P., Naylor, A., effects on spinal cord dorsal horn met-enkephalin. Pain 88,
Bountra, C., Banati, R.B., Birch, R., Anand, P., 2004. 267–275.
Cyclooxygenase-2 (Cox-2) in injured human nerve and a rat George, A., Buehl, A., Sommer, C., 2005. Tumor necrosis factor
model of nerve injury. J. Peripher. Nerv. Syst. 9, 15–25. receptor 1 and 2 proteins are differentially regulated during
Eaton, M.J., Blits, B., Ruitenberg, M.J., Verhaagen, J., Oudega, M., Wallerian degeneration of mouse sciatic nerve. Exp. Neurol.
2002. Amelioration of chronic neuropathic pain after partial 192, 163–166.
nerve injury by adeno-associated viral (AAV) vector-mediated Gillen, C., Jander, S., Stoll, G., 1998. Sequential expression of mRNA
over-expression of BDNF in the rat spinal cord. Gene Ther. 9, for proinflammatory cytokines and interleukin-10 in the rat
1387–1395. peripheral nervous system: comparison between immune-
Eckert, A., Segond von Banchet, G., Sopper, S., Petersen, M., 1999. mediated demyelination and Wallerian degeneration.
Spatio-temporal pattern of induction of bradykinin receptors J. Neurosci. Res. 51, 489–496.
and inflammation in rat dorsal root ganglia after unilateral Giorgi, R., Pagano, R.L., Dias, M.A., Aguiar-Passeti, T., Sorg, C.,
nerve ligation. Pain 83, 487–497. Mariano, M., 1998. Antinociceptive effect of the calcium-
Eliav, E., Herzberg, U., Ruda, M.A., Bennett, G.J., 1999. Neuropathic binding protein MRP-14 and the role played by neutrophils
pain from an experimental neuritis of the rat sciatic nerve. Pain on the control of inflammatory pain. J. Leukocyte Biol. 64,
83, 169–182. 214–220.
Eriksson, J., Bongenhielm, U., Kidd, E., Matthews, B., Fried, K., Grafe, P., Mayer, C., Takigawa, T., Kamleiter, M., Sanchez-
1998. Distribution of P2X3 receptors in the rat trigeminal Brandelik, R., 1999. Confocal calcium imaging reveals an
ganglion after inferior alveolar nerve injury. Neurosci. Lett. ionotropic P2 nucleotide receptor in the paranodal
254, 37–40. membrane of rat Schwann cells. J. Physiol. 515 (Pt 2),
Evans, A.R., Junger, H., Southall, M.D., Nicol, G.D., Sorkin, L.S., 377–383.
Broome, J.T., Bailey, T.W., Vasko, M.R., 2000. Isoprostanes, Griesbacher, T., Rainer, I., 1999. 5-hydroxytryptamine release from
novel eicosanoids that produce nociception and sensitize rat skin mast cells in vivo induced by peptide but not by
sensory neurons. J. Pharmacol. Exp. Ther. 293, 912–920. nonpeptide ligands for bradykinin receptors.
Faurschou, M., Borregaard, N., 2003. Neutrophil granules and Immunopharmacology 43, 195–201.
secretory vesicles in inflammation. Microbes Infect. 5, Groth, R., Aanonsen, L., 2002. Spinal brain-derived neurotrophic
1317–1327. factor (BDNF) produces hyperalgesia in normal mice while
Ferreira, S.H., Lorenzetti, B.B., Bristow, A.F., Poole, S., 1988. antisense directed against either BDNF or trkB, prevent
Interleukin-1 beta as a potent hyperalgesic agent antagonized inflammation-induced hyperalgesia. Pain 100, 171–181.
by a tripeptide analogue. Nature 334, 698–700. Grothe, C., Heese, K., Meisinger, C., Wewetzer, K., Kunz, D.,
Flatters, S.J., Fox, A.J., Dickenson, A.H., 2004. Nerve injury alters the Cattini, P., Otten, U., 2000. Expression of interleukin-6 and its
effects of interleukin-6 on nociceptive transmission in receptor in the sciatic nerve and cultured Schwann cells:
peripheral afferents. Eur. J. Pharmacol. 484, 183–191. relation to 18-kD fibroblast growth factor-2. Brain Res. 885,
Follenfant, R.L., Nakamura-Craig, M., Henderson, B., Higgs, G.A., 172–181.
1989. Inhibition by neuropeptides of interleukin-1 beta- Guzik, T.J., Korbut, R., Adamek-Guzik, T., 2003. Nitric oxide and
induced, prostaglandin-independent hyperalgesia. Br. J. superoxide in inflammation and immune regulation. J. Physiol.
Pharmacol. 98, 41–43. Pharmacol. 54, 469–487.
Frisen, J., Risling, M., Fried, K., 1993. Distribution and axonal Hamilton, S.G., McMahon, S.B., 2000. ATP as a peripheral mediator
relations of macrophages in a neuroma. Neuroscience 55, of pain. J. Auton. Nerv. Syst. 81, 187–194.
1003–1013. Hammarberg, H., Piehl, F., Cullheim, S., Fjell, J., Hokfelt, T., Fried,
Frostick, S.P., Yin, Q., Kemp, G.J., 1998. Schwann cells, K., 1996. GDNF mRNA in Schwann cells and DRG satellite cells
neurotrophic factors, and peripheral nerve regeneration. after chronic sciatic nerve injury. NeuroReport 7, 857–860.
Microsurgery 18, 397–405. Hanisch, U.K., 2002. Microglia as a source and target of cytokines.
Fukuoka, H., Kawatani, M., Hisamitsu, T., Takeshige, C., 1994. Glia 40, 140–155.
Cutaneous hyperalgesia induced by peripheral injection of Harden, N., Cohen, M., 2003. Unmet needs in the management of
interleukin-1 beta in the rat. Brain Res. 657, 133–140. neuropathic pain. J. Pain Symptom Manage 25, S12–S17.
Funakoshi, H., Frisen, J., Barbany, G., Timmusk, T., Zachrisson, O., Harvey, R.J., Depner, U.B., Wassle, H., Ahmadi, S., Heindl, C.,
Verge, V.M., Persson, H., 1993. Differential expression of Reinold, H., Smart, T.G., Harvey, K., Schutz, B., Abo-Salem, O.M.,
mRNAs for neurotrophins and their receptors after axotomy of Zimmer, A., Poisbeau, P., Welzl, H., Wolfer, D.P., Betz, H.,
the sciatic nerve. J. Cell Biol. 123, 455–465. Zeilhofer, H.U., Muller, U., 2004. GlyR alpha3: an essential target
Gaboury, J.P., Johnston, B., Niu, X.F., Kubes, P., 1995. Mechanisms for spinal PGE2-mediated inflammatory pain sensitization.
underlying acute mast cell-induced leukocyte rolling and Science 304, 884–887.
adhesion in vivo. J. Immunol. 154, 804–813. Hata, A.N., Breyer, R.M., 2004. Pharmacology and signaling of
Galli, S.J., Nakae, S., Tsai, M., 2005. Mast cells in the development of prostaglandin receptors: multiple roles in inflammation and
adaptive immune responses. Nat. Immunol. 6, 135–142. immune modulation. Pharmacol. Ther. 103, 147–166.
Gardell, L.R., Wang, R., Ehrenfels, C., Ossipov, M.H., Rossomando, Herzberg, U., Eliav, E., Dorsey, J.M., Gracely, R.H., Kopin, I.J.,
A.J., Miller, S., Buckley, C., Cai, A.K., Tse, A., Foley, S.F., Gong, B., 1997. NGF involvement in pain induced by chronic
Walus, L., Carmillo, P., Worley, D., Huang, C., Engber, T., constriction injury of the rat sciatic nerve. NeuroReport 8,
Pepinsky, B., Cate, R.L., Vanderah, T.W., Lai, J., Sah, D.W., 1613–1618.
Porreca, F., 2003. Multiple actions of systemic artemin in Heumann, R., Korsching, S., Bandtlow, C., Thoenen, H., 1987.
experimental neuropathy. Nat. Med. 9, 1383–1389. Changes of nerve growth factor synthesis in nonneuronal cells
George, A., Schmidt, C., Weishaupt, A., Toyka, K.V., Sommer, C., in response to sciatic nerve transection. J. Cell Biol. 104,
1999. Serial determination of tumor necrosis factor-alpha 1623–1631.
content in rat sciatic nerve after chronic constriction injury. Holthusen, H., Arndt, J.O., 1994. Nitric oxide evokes pain in
Exp. Neurol. 160, 124–132. humans on intracutaneous injection. Neurosci. Lett. 165,
George, A., Marziniak, M., Schäfers, M., Toyka, K.V., Sommer, 71–74.
C., 2000. Thalidomide treatment in chronic constrictive Holthusen, H., Arndt, J.O., 1995. Nitric oxide evokes pain at
258 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

nociceptors of the paravascular tissue and veins in humans. Kawabata, A., Kawao, N., Kuroda, R., Tanaka, A., Itoh, H.,
J. Physiol. 487 (Pt 1), 253–258. Nishikawa, H., 2001. Peripheral PAR-2 triggers thermal
Hong, Y., Abbott, F.V., 1994. Behavioural effects of intraplantar hyperalgesia and nociceptive responses in rats. NeuroReport
injection of inflammatory mediators in the rat. Neuroscience 12, 715–719.
63, 827–836. Keeble, J.E., Moore, P.K., 2002. Pharmacology and potential
Hosli, E., Hosli, L., 1993. Receptors for neurotransmitters on therapeutic applications of nitric oxide-releasing non-steroidal
astrocytes in the mammalian central nervous system. Prog. anti-inflammatory and related nitric oxide-donating drugs. Br.
Neurobiol. 40, 477–506. J. Pharmacol. 137, 295–310.
Hu, P., McLachlan, E.M., 2002. Macrophage and lymphocyte Kennedy, C., Assis, T.S., Currie, A.J., Rowan, E.G., 2003. Crossing the
invasion of dorsal root ganglia after peripheral nerve lesions in pain barrier: P2 receptors as targets for novel analgesics.
the rat. Neuroscience 112, 23–38. J. Physiol. 553, 683–694.
Hu, S., Sheng, W.S., Peterson, P.K., Chao, C.C., 1995. Differential Kerns, R.D., Kassirer, M., Otis, J., 2002. Pain in multiple sclerosis: a
regulation by cytokines of human astrocyte nitric oxide biopsychosocial perspective. J. Rehabil. Res. Dev. 39, 225–232.
production. Glia 15, 491–494. Khalil, Z., Liu, T., Helme, R.D., 1999. Free radicals contribute to the
Huang, E.J., Reichardt, L.F., 2003. Trk receptors: roles in neuronal reduction in peripheral vascular responses and the
signal transduction. Annu. Rev. Biochem. 72, 609–642. maintenance of thermal hyperalgesia in rats with chronic
Hughes, R.A., Hadden, R.D., Gregson, N.A., Smith, K.J., 1999. constriction injury. Pain 79, 31–37.
Pathogenesis of Guillain–Barre syndrome. J. Neuroimmunol. Khasar, S.G., Miao, F.J., Levine, J.D., 1995. Inflammation modulates
100, 74–97. the contribution of receptor-subtypes to bradykinin-induced
Ignatowski, T.A., Covey, W.C., Knight, P.R., Severin, C.M., Nickola, hyperalgesia in the rat. Neuroscience 69, 685–690.
T.J., Spengler, R.N., 1999. Brain-derived TNFalpha mediates Kim, S.H., Chung, J.M., 1992. An experimental model for peripheral
neuropathic pain. Brain Res. 841, 70–77. neuropathy produced by segmental spinal nerve ligation in the
Inoue, K., Tsuda, M., Koizumi, S., 2004. ATP-and adenosine- rat. Pain 50, 355–363.
mediated signaling in the central nervous system: chronic pain Kim, H.K., Park, S.K., Zhou, J.L., Taglialatela, G., Chung, K.,
and microglia: involvement of the ATP receptor P2X4. Coggeshall, R.E., Chung, J.M., 2004. Reactive oxygen species
J. Pharmacol. Sci. 94, 112–114. (ROS) play an important role in a rat model of neuropathic pain.
Irnich, D., Burgstahler, R., Bostock, H., Grafe, P., 2001. ATP affects Pain 111, 116–124.
both axons and Schwann cells of unmyelinated C fibres. Pain Koda, H., Mizumura, K., 2002. Sensitization to mechanical
92, 343–350. stimulation by inflammatory mediators and by mild burn in
Irnich, D., Tracey, D.J., Polten, J., Burgstahler, R., Grafe, P., 2002. ATP canine visceral nociceptors in vitro. J. Neurophysiol. 87,
stimulates peripheral axons in human, rat and mouse— 2043–2051.
differential involvement of A(2B) adenosine and P2X purinergic Koltzenburg, M., Bennett, D.L., Shelton, D.L., McMahon, S.B., 1999a.
receptors. Neuroscience 110, 123–129. Neutralization of endogenous NGF prevents the sensitization
Ito, Y., Yamamoto, M., Li, M., Doyu, M., Tanaka, F., Mutch, T., of nociceptors supplying inflamed skin. Eur. J. Neurosci. 11,
Mitsuma, T., Sobue, G., 1998. Differential temporal expression 1698–1704.
of mRNAs for ciliary neurotrophic factor (CNTF), leukemia Koltzenburg, M., Wall, P.D., McMahon, S.B., 1999b. Does the right
inhibitory factor (LIF), interleukin-6 (IL-6), and their receptors side know what the left is doing? Trends Neurosci. 22, 122–127.
(CNTFR alpha, LIFR beta, IL-6R alpha and gp130) in injured Kurek, J.B., Austin, L., Cheema, S.S., Bartlett, P.F., Murphy, M., 1996.
peripheral nerves. Brain Res. 793, 321–327. Up-regulation of leukaemia inhibitory factor and interleukin-6
Jarvis, M.F., Burgard, E.C., McGaraughty, S., Honore, P., Lynch, K., in transected sciatic nerve and muscle following denervation.
Brennan, T.J., Subieta, A., Van Biesen, T., Cartmell, J., Bianchi, Neuromuscul. Disord. 6, 105–114.
B., Niforatos, W., Kage, K., Yu, H., Mikusa, J., Wismer, C.T., Lang, E., Novak, A., Reeh, P.W., Handwerker, H.O., 1990.
Zhu, C.Z., Chu, K., Lee, C.H., Stewart, A.O., Polakowski, J., Cox, Chemosensitivity of fine afferents from rat skin in vitro.
B.F., Kowaluk, E., Williams, M., Sullivan, J., Faltynek, C., 2002. J. Neurophysiol. 63, 887–901.
A-317491, a novel potent and selective non-nucleotide Lang, P.M., Tracey, D.J., Irnich, D., Sippel, W., Grafe, P., 2002.
antagonist of P2X3 and P2X2/3 receptors, reduces chronic Activation of adenosine and P2Y receptors by ATP in human
inflammatory and neuropathic pain in the rat. Proc. Natl. peripheral nerve. Naunyn-Schmiedeberg's Arch. Pharmacol.
Acad. Sci. U. S. A. 99, 17179–17184. 366, 449–457.
Jett, M.F., McGuirk, J., Waligora, D., Hunter, J.C., 1997. The effects of Lang, P.M., Burgstahler, R., Sippel, W., Irnich, D., Schlotter-Weigel,
mexiletine, desipramine and fluoxetine in rat models involving B., Grafe, P., 2003. Characterization of neuronal nicotinic
central sensitization. Pain 69, 161–169. acetylcholine receptors in the membrane of unmyelinated
Jin, X., Shepherd, R.K., Duling, B.R., Linden, J., 1997. Inosine binds human C-fiber axons by in vitro studies. J. Neurophysiol. 90,
to A3 adenosine receptors and stimulates mast cell 3295–3303.
degranulation. J. Clin. Invest. 100, 2849–2857. Lauretti, G.R., Perez, M.V., Reis, M.P., Pereira, N.L., 2002. Double-
Julius, D., Basbaum, A.I., 2001. Molecular mechanisms of blind evaluation of transdermal nitroglycerine as adjuvant to
nociception. Nature 413, 203–210. oral morphine for cancer pain management. J. Clin. Anesth. 14,
Junger, H., Sorkin, L.S., 2000. Nociceptive and inflammatory effects 83–86.
of subcutaneous TNFalpha. Pain 85, 145–151. Lavand'homme, P.M., Eisenach, J.C., 1999. Exogenous and
Kagan, B.L., Baldwin, R.L., Munoz, D., Wisnieski, B.J., 1992. endogenous adenosine enhance the spinal antiallodynic
Formation of ion-permeable channels by tumor necrosis effects of morphine in a rat model of neuropathic pain. Pain
factor-alpha. Science 255, 1427–1430. 80, 31–36.
Kage, K., Niforatos, W., Zhu, C.Z., Lynch, K.J., Honore, P., Jarvis, Lawand, N.B., Willis, W.D., Westlund, K.N., 1997. Blockade of joint
M.F., 2002. Alteration of dorsal root ganglion P2X3 receptor inflammation and secondary hyperalgesia by L-NAME, a nitric
expression and function following spinal nerve ligation in oxide synthase inhibitor. NeuroReport 8, 895–899.
the rat. Exp. Brain Res. 147, 511–519. Ledeboer, A., Sloane, E.M., Milligan, E.D., Frank, M.G., Mahony,
Kawabata, A., Umeda, N., Takagi, H., 1993. L-arginine exerts a dual J.H., Maier, S.F., Watkins, L.R., 2005. Minocycline attenuates
role in nociceptive processing in the brain: involvement of the mechanical allodynia and proinflammatory cytokine
kyotorphin-Met-enkephalin pathway and NO-cyclic GMP expression in rat models of pain facilitation. Pain 115,
pathway. Br. J. Pharmacol. 109, 73–79. 71–83.
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 259

Lee, H.L., Lee, K.M., Son, S.J., Hwang, S.H., Cho, H.J., 2004. Temporal receptor superfamilies: integrating mammalian biology. Cell
expression of cytokines and their receptors mRNAs in a 104, 487–501.
neuropathic pain model. NeuroReport 15, 2807–2811. Lui, P.W., Lee, C.H., 2004. Preemptive effects of intrathecal
Levine, J.D., Lau, W., Kwiat, G., Goetzl, E.J., 1984. Leukotriene B4 cyclooxygenase inhibitor or nitric oxide synthase inhibitor on
produces hyperalgesia that is dependent on thermal hypersensitivity following peripheral nerve injury.
polymorphonuclear leukocytes. Science 225, 743–745. Life Sci. 75, 2527–2538.
Levine, J.D., Gooding, J., Donatoni, P., Borden, L., Goetzl, E.J., 1985. Luo, Z.D., Cizkova, D., 2000. The role of nitric oxide in nociception.
The role of the polymorphonuclear leukocyte in hyperalgesia. Curr. Rev. Pain 4, 459–466.
J. Neurosci. 5, 3025–3029. Luo, X.G., Rush, R.A., Zhou, X.F., 2001. Ultrastructural localization
Levine, J.D., Lam, D., Taiwo, Y.O., Donatoni, P., Goetzl, E.J., 1986. of brain-derived neurotrophic factor in rat primary sensory
Hyperalgesic properties of 15-lipoxygenase products of neurons. Neurosci. Res. 39, 377–384.
arachidonic acid. Proc. Natl. Acad. Sci. U. S. A. 83, 5331–5334. Ma, W., Eisenach, J.C., 2002. Morphological and pharmacological
Levy, D., Zochodne, D.W., 1998. Local nitric oxide synthase evidence for the role of peripheral prostaglandins in the
activity in a model of neuropathic pain. Eur. J. Neurosci. 10, pathogenesis of neuropathic pain. Eur. J. Neurosci. 15,
1846–1855. 1037–1047.
Levy, D., Zochodne, D.W., 2000. Increased mRNA expression of the Ma, W., Eisenach, J.C., 2003a. Four PGE2 EP receptors are
B1 and B2 bradykinin receptors and antinociceptive effects of up-regulated in injured nerve following partial sciatic nerve
their antagonists in an animal model of neuropathic pain. Pain ligation. Exp. Neurol. 183, 581–592.
86, 265–271. Ma, W., Eisenach, J.C., 2003b. Cyclooxygenase 2 in infiltrating
Levy, D., Zochodne, D.W., 2004. NO pain: potential roles of nitric inflammatory cells in injured nerve is universally up-regulated
oxide in neuropathic pain. Pain Pract. 4, 11–18. following various types of peripheral nerve injury.
Levy, D., Hoke, A., Zochodne, D.W., 1999. Local expression of Neuroscience 121, 691–704.
inducible nitric oxide synthase in an animal model of Ma, W., Quirion, R., 2005. Up-regulation of interleukin-6 induced
neuropathic pain. Neurosci. Lett. 260, 207–209. by prostaglandin E2 from invading macrophages following
Levy, D., Tal, M., Hoke, A., Zochodne, D.W., 2000. Transient action nerve injury: an in vivo and in vitro study. J. Neurochem. 93,
of the endothelial constitutive nitric oxide synthase (ecNOS) 664–673.
mediates the development of thermal hypersensitivity Ma, W., Ribeiro-da-Silva, A., Noel, G., Julien, J.P., Cuello, A.C., 1995.
following peripheral nerve injury. Eur. J. Neurosci. 12, Ectopic substance P and calcitonin gene-related peptide
2323–2332. immunoreactive fibres in the spinal cord of transgenic mice
Lewin, G.R., Barde, Y.A., 1996. Physiology of the neurotrophins. over-expressing nerve growth factor. Eur. J. Neurosci. 7,
Annu. Rev. Neurosci. 19, 289–317. 2021–2035.
Lewin, G.R., Rueff, A., Mendell, L.M., 1994. Peripheral and central Ma, W., Du, W., Eisenach, J.C., 2002. Role for both spinal cord
mechanisms of NGF-induced hyperalgesia. Eur. J. Neurosci. 6, COX-1 and COX-2 in maintenance of mechanical
1903–1912. hypersensitivity following peripheral nerve injury. Brain Res.
Li, L., Xian, C.J., Zhong, J.H., Zhou, X.F., 2002. Effect of lumbar 5 937, 94–99.
ventral root transection on pain behaviors: a novel rat model Malaviya, R., Abraham, S.N., 2000. Role of mast cell leukotrienes in
for neuropathic pain without axotomy of primary sensory neutrophil recruitment and bacterial clearance in infectious
neurons. Exp. Neurol. 175, 23–34. peritonitis. J. Leukocyte Biol. 67, 841–846.
Li, L., Xian, C.J., Zhong, J.H., Zhou, X.F., 2003. Lumbar 5 ventral root Malmberg, A.B., Yaksh, T.L., 1992. Hyperalgesia mediated by spinal
transection-induced upregulation of nerve growth factor in glutamate or substance P receptor blocked by spinal
sensory neurons and their target tissues: a mechanism in cyclooxygenase inhibition. Science 257, 1276–1279.
neuropathic pain. Mol. Cell. Neurosci. 23, 232–250. Marceau, F., 1995. Kinin B1 receptors: a review.
Lin, Q., Palecek, J., Paleckova, V., Peng, Y.B., Wu, J., Cui, M., Willis, Immunopharmacology 30, 1–26.
W.D., 1999a. Nitric oxide mediates the central sensitization of Matsuoka, I., Meyer, M., Thoenen, H., 1991. Cell-type-specific
primate spinothalamic tract neurons. J. Neurophysiol. 81, regulation of nerve growth factor (NGF) synthesis in non-
1075–1085. neuronal cells: comparison of Schwann cells with other cell
Lin, Q., Wu, J., Peng, Y.B., Cui, M., Willis, W.D., 1999b. Nitric oxide- types. J. Neurosci. 11, 3165–3177.
mediated spinal disinhibition contributes to the sensitization Maves, T.J., Pechman, P.S., Gebhart, G.F., Meller, S.T., 1993. Possible
of primate spinothalamic tract neurons. J. Neurophysiol. 81, chemical contribution from chromic gut sutures produces
1086–1094. disorders of pain sensation like those seen in man. Pain 54,
Linden, J., 1994. Purinergic systems. In: Siegel, G.J., et al. (Eds.), 57–69.
Basic Neurochemistry: Molecular, Cellular, and Medical Max, M.B., Lynch, S.A., Muir, J., Shoaf, S.E., Smoller, B., Dubner, R.,
Aspects. Lippincott-Raven, Philadelphia, pp. 401–416. 1992. Effects of desipramine, amitriptyline, and fluoxetine on
Lindholm, D., Heumann, R., Meyer, M., Thoenen, H., 1987. pain in diabetic neuropathy. N. Engl. J. Med. 326, 1250–1256.
Interleukin-1 regulates synthesis of nerve growth factor in Mayer, C., Quasthoff, S., Grafe, P., 1998. Differences in the
non-neuronal cells of rat sciatic nerve. Nature 330, sensitivity to purinergic stimulation of myelinating and non-
658–659. myelinating Schwann cells in peripheral human and rat nerve.
Liu, G.J., Bennett, M.R., 2003. ATP secretion from nerve trunks and Glia 23, 374–382.
Schwann cells mediated by glutamate. NeuroReport 14, McLean, P.G., Ahluwalia, A., Perretti, M., 2000. Association between
2079–2083. kinin B(1) receptor expression and leukocyte trafficking across
Liu, T., Knight, K.R., Tracey, D.J., 2000a. Hyperalgesia due to nerve mouse mesenteric postcapillary venules. J. Exp. Med. 192,
injury—Role of peroxynitrite. Neuroscience 97, 125–131. 367–380.
Liu, T., van Rooijen, N., Tracey, D.J., 2000b. Depletion of McMahon, S.B., Bennett, D.L., Priestley, J.V., Shelton, D.L., 1995. The
macrophages reduces axonal degeneration and hyperalgesia biological effects of endogenous nerve growth factor on adult
following nerve injury. Pain 86, 25–32. sensory neurons revealed by a trkA-IgG fusion molecule. Nat.
Liu, G.J., Werry, E.L., Bennett, M.R., 2005. Secretion of ATP from Med. 1, 774–780.
Schwann cells in response to uridine triphosphate. Eur. J. McMahon, S.B., Cafferty, W.B., Marchand, F., 2005. Immune and
Neurosci. 21, 151–160. glial cell factors as pain mediators and modulators. Exp.
Locksley, R.M., Killeen, N., Lenardo, M.J., 2001. The TNF and TNF Neurol. 192, 444–462.
260 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

Mekori, Y.A., Metcalfe, D.D., 2000. Mast cells in innate immunity. 1997. Pain in Guillain–Barre syndrome. Neurology 48,
Immunol. Rev. 173, 131–140. 328–331.
Meller, S.T., Gebhart, G.F., 1993. Nitric oxide (NO) and nociceptive Mueller, M., Leonhard, C., Wacker, K., Ringelstein, E.B., Okabe, M.,
processing in the spinal cord. Pain 52, 127–136. Hickey, W.F., Kiefer, R., 2003. Macrophage response to
Meller, S.T., Pechman, P.S., Gebhart, G.F., Maves, T.J., 1992. Nitric peripheral nerve injury: the quantitative contribution of
oxide mediates the thermal hyperalgesia produced in a model resident and hematogenous macrophages. Lab. Invest. 83,
of neuropathic pain in the rat. Neuroscience 50, 7–10. 175–185.
Meller, S.T., Dykstra, C., Grzybycki, D., Murphy, S., Gebhart, G.F., Muja, N., DeVries, G.H., 2004. Prostaglandin E(2) and
1994. The possible role of glia in nociceptive processing and 6-keto-prostaglandin F(1alpha) production is elevated
hyperalgesia in the spinal cord of the rat. Neuropharmacology following traumatic injury to sciatic nerve. Glia 46, 116–129.
33, 1471–1478. Murphy, P.G., Grondin, J., Altares, M., Richardson, P.M., 1995.
Metcalfe, D.D., Baram, D., Mekori, Y.A., 1997. Mast cells. Physiol. Induction of interleukin-6 in axotomized sensory neurons.
Rev. 77, 1033–1079. J. Neurosci. 15, 5130–5138.
Meyer, M., Matsuoka, I., Wetmore, C., Olson, L., Thoenen, H., 1992. Murphy, P.G., Ramer, M.S., Borthwick, L., Gauldie, J., Richardson,
Enhanced synthesis of brain-derived neurotrophic factor in the P.M., Bisby, M.A., 1999. Endogenous interleukin-6 contributes
lesioned peripheral nerve: different mechanisms are to hypersensitivity to cutaneous stimuli and changes in
responsible for the regulation of BDNF and NGF mRNA. J. Cell neuropeptides associated with chronic nerve constriction in
Biol. 119, 45–54. mice. Eur. J. Neurosci. 11, 2243–2253.
Miletic, G., Miletic, V., 2002. Increases in the concentration of brain Murwani, R., Hodgkinson, S., Armati, P., 1996. Tumor necrosis
derived neurotrophic factor in the lumbar spinal dorsal horn factor alpha and interleukin-6 mRNA expression in neonatal
are associated with pain behavior following chronic Lewis rat Schwann cells and a neonatal rat Schwann cell line
constriction injury in rats. Neurosci. Lett. 319, 137–140. following interferon gamma stimulation. J. Neuroimmunol. 71,
Millan, M.J., 2002. Descending control of pain. Prog. Neurobiol. 66, 65–71.
355–474. Myers, R.R., Heckman, H.M., Rodriguez, M., 1996. Reduced
Milligan, E.D., Zapata, V., Chacur, M., Schoeniger, D., Biedenkapp, hyperalgesia in nerve-injured WLD mice: relationship to nerve
J., O'Connor, K.A., Verge, G.M., Chapman, G., Green, P., Foster, fiber phagocytosis, axonal degeneration, and regeneration in
A.C., Naeve, G.S., Maier, S.F., Watkins, L.R., 2004. Evidence normal mice. Exp. Neurol. 141, 94–101.
that exogenous and endogenous fractalkine can induce Nagano, M., Sakai, A., Takahashi, N., Umino, M., Yoshioka, K.,
spinal nociceptive facilitation in rats. Eur. J. Neurosci. 20, Suzuki, H., 2003. Decreased expression of glial cell line-derived
2294–2302. neurotrophic factor signaling in rat models of neuropathic
Milligan, E.D., Langer, S.J., Sloane, E.M., He, L., Wieseler-Frank, J., pain. Br. J. Pharmacol. 140, 1252–1260.
O'Connor, K., Martin, D., Forsayeth, J.R., Maier, S.F., Johnson, K., Namaka, M., Gramlich, C.R., Ruhlen, D., Melanson, M., Sutton, I.,
Chavez, R.A., Leinwand, L.A., Watkins, L.R., 2005a. Controlling Major, J., 2004. A treatment algorithm for neuropathic pain.
pathological pain by adenovirally driven spinal production of Clin. Ther. 26, 951–979.
the anti-inflammatory cytokine, interleukin-10. Eur. J. Nathan, C.F., 1987. Secretory products of macrophages. J. Clin.
Neurosci. 21, 2136–2148. Invest. 79, 319–326.
Milligan, E.D., Sloane, E.M., Langer, S.J., Cruz, P.E., Chacur, M., Novakovic, S.D., Kassotakis, L.C., Oglesby, I.B., Smith, J.A., Eglen,
Spataro, L., Wieseler-Frank, J., Hammack, S.E., Maier, S.F., R.M., Ford, A.P., Hunter, J.C., 1999. Immunocytochemical
Flotte, T.R., Forsayeth, J.R., Leinwand, L.A., Chavez, R., Watkins, localization of P2X3 purinoceptors in sensory neurons in
L.R., 2005b. Controlling neuropathic pain by adeno-associated naive rats and following neuropathic injury. Pain 80, 273–282.
virus driven production of the anti-inflammatory cytokine, Nykjaer, A., Willnow, T.E., Petersen, C.M., 2005. p75(NTR)—Live or
interleukin-10. Mol. Pain 1, 9. let die. Curr. Opin. Neurobiol. 15, 49–57.
Minghetti, L., Levi, G., 1998. Microglia as effector cells in brain Obreja, O., Schmelz, M., Poole, S., Kress, M., 2002. Interleukin-6 in
damage and repair: focus on prostanoids and nitric oxide. Prog. combination with its soluble IL-6 receptor sensitises rat skin
Neurobiol. 54, 99–125. nociceptors to heat, in vivo. Pain 96, 57–62.
Mitchell, S.W., 1872. Injuries of Nerves and Their Consequences. Oh, S.B., Tran, P.B., Gillard, S.E., Hurley, R.W., Hammond, D.L.,
J.B. Lippincott, Philadelphia. Miller, R.J., 2001. Chemokines and glycoprotein120 produce
Moalem, G., Gdalyahu, A., Shani, Y., Otten, U., Lazarovici, P., pain hypersensitivity by directly exciting primary nociceptive
Cohen, I.R., Schwartz, M., 2000. Production of neurotrophins by neurons. J. Neurosci. 21, 5027–5035.
activated T cells: implications for neuroprotective Oka, T., Oka, K., Hosoi, M., Hori, T., 1995. Intracerebroventricular
autoimmunity. J. Autoimmun. 15, 331–345. injection of interleukin-6 induces thermal hyperalgesia in rats.
Moalem, G., Xu, K., Yu, L., 2004. T lymphocytes play a role in Brain Res. 692, 123–128.
neuropathic pain following peripheral nerve injury in rats. Okada, M., Nakagawa, T., Minami, M., Satoh, M., 2002. Analgesic
Neuroscience 129, 767–777. effects of intrathecal administration of P2Y nucleotide receptor
Moalem, G., Grafe, P., Tracey, D.J., 2005. Chemical mediators agonists UTP and UDP in normal and neuropathic pain model
enhance the excitability of unmyelinated sensory axons in rats. J. Pharmacol. Exp. Ther. 303, 66–73.
normal and injured peripheral nerve of the rat. Neuroscience Okamoto, K., Martin, D.P., Schmelzer, J.D., Mitsui, Y., Low, P.A.,
134, 1399–1411. 2001. Pro-and anti-inflammatory cytokine gene expression in
Molliver, D.C., Cook, S.P., Carlsten, J.A., Wright, D.E., McCleskey, E. rat sciatic nerve chronic constriction injury model of
W., 2002. ATP and UTP excite sensory neurons and induce CREB neuropathic pain. Exp. Neurol. 169, 386–391.
phosphorylation through the metabotropic receptor, P2Y2. Eur. Olsson, Y., 1967. Degranulation of mast cells in peripheral nerve
J. Neurosci. 16, 1850–1860. injuries. Acta Neurol. Scand. 43, 365–374.
Mosmann, T.R., Sad, S., 1996. The expanding universe of T-cell Olsson, Y., 1968. Mast cells in the nervous system. Int. Rev. Cytol.
subsets: Th1, Th2 and more. Immunol. Today 17, 138–146. 24, 27–70.
Mosmann, T.R., Cherwinski, H., Bond, M.W., Giedlin, M.A., Opree, A., Kress, M., 2000. Involvement of the proinflammatory
Coffman, R.L., 1986. Two types of murine helper T cell clone. I. cytokines tumor necrosis factor-alpha, IL-1 beta, and IL-6 but
Definition according to profiles of lymphokine activities and not IL-8 in the development of heat hyperalgesia: effects on
secreted proteins. J. Immunol. 136, 2348–2357. heat-evoked calcitonin gene-related peptide release from rat
Moulin, D.E., Hagen, N., Feasby, T.E., Amireh, R., Hahn, A., skin. J. Neurosci. 20, 6289–6293.
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 261

Otten, U., 1991. Nerve growth factor: a signalling protein between differentiation of precursor mouse CD8+ T cells into cytotoxic
the nervous and the immune system. In: Basbaum, A.I., CD8+ T cells secreting Th1 or Th2 cytokines. Immunity 2,
Besson, J.-M.R. (Eds.), Towards A New Pharmacotherapy of 271–279.
Pain. Wiley, Chichester, pp. 353–363. Samad, T.A., Sapirstein, A., Woolf, C.J., 2002. Prostanoids and pain:
Pahl, H.L., 1999. Activators and target genes of Rel/NF-kappaB unraveling mechanisms and revealing therapeutic targets.
transcription factors. Oncogene 18, 6853–6866. Trends Mol. Med. 8, 390–396.
Pearce, F.L., Thompson, H.L., 1986. Some characteristics of Sato, E., Koyama, S., Nomura, H., Kubo, K., Sekiguchi, M., 1996.
histamine secretion from rat peritoneal mast cells stimulated Bradykinin stimulates alveolar macrophages to release
with nerve growth factor. J. Physiol. 372, 379–393. neutrophil, monocyte, and eosinophil chemotactic activity.
Pentland, B., Donald, S.M., 1994. Pain in the Guillain–Barre J. Immunol. 157, 3122–3129.
syndrome: a clinical review. Pain 59, 159–164. Sawynok, J., 1998. Adenosine receptor activation and nociception.
Perkins, N.M., Tracey, D.J., 2000. Hyperalgesia due to nerve injury: Eur. J. Pharmacol. 347, 1–11.
role of neutrophils. Neuroscience 101, 745–757. Sawynok, J., Liu, X.J., 2003. Adenosine in the spinal cord and
Perrin, F.E., Lacroix, S., Aviles-Trigueros, M., David, S., 2005. periphery: release and regulation of pain. Prog. Neurobiol. 69,
Involvement of monocyte chemoattractant protein-1, 313–340.
macrophage inflammatory protein-1{alpha} and interleukin-1 Sawynok, J., Reid, A., Liu, X.J., 2000. Involvement of mast cells,
{beta} in Wallerian degeneration. Brain. sensory afferents and sympathetic mechanisms in paw
Perry, V.H., Brown, M.C., 1992. Role of macrophages in peripheral oedema induced by adenosine A(1) and A(2B/3) receptor
nerve degeneration and repair. BioEssays 14, 401–406. agonists. Eur. J. Pharmacol. 395, 47–50.
Perry, V.H., Brown, M.C., Gordon, S., 1987. The macrophage Scapini, P., Lapinet-Vera, J.A., Gasperini, S., Calzetti, F., Bazzoni, F.,
response to central and peripheral nerve injury. A possible role Cassatella, M.A., 2000. The neutrophil as a cellular source of
for macrophages in regeneration. J. Exp. Med. 165, 1218–1223. chemokines. Immunol. Rev. 177, 195–203.
Pezet, S., Malcangio, M., McMahon, S.B., 2002. BDNF: a Schäfers, M., Brinkhoff, J., Neukirchen, S., Marziniak, M., Sommer,
neuromodulator in nociceptive pathways? Brain Res. Brain C., 2001. Combined epineurial therapy with neutralizing
Res. Rev. 40, 240–249. antibodies to tumor necrosis factor-alpha and interleukin-1
Piehl, F., Lidman, O., 2001. Neuroinflammation in the rat—CNS receptor has an additive effect in reducing neuropathic pain in
cells and their role in the regulation of immune reactions. mice. Neurosci. Lett. 310, 113–116.
Immunol. Rev. 184, 212–225. Schäfers, M., Lee, D.H., Brors, D., Yaksh, T.L., Sorkin, L.S., 2003a.
Pierce, P.A., Xie, G.X., Levine, J.D., Peroutka, S.J., 1996. 5- Increased sensitivity of injured and adjacent uninjured rat
Hydroxytryptamine receptor subtype messenger RNAs in rat primary sensory neurons to exogenous tumor necrosis factor-
peripheral sensory and sympathetic ganglia: a polymerase alpha after spinal nerve ligation. J. Neurosci. 23, 3028–3038.
chain reaction study. Neuroscience 70, 553–559. Schäfers, M., Svensson, C.I., Sommer, C., Sorkin, L.S., 2003b. Tumor
Pitchford, S., Levine, J.D., 1991. Prostaglandins sensitize necrosis factor-alpha induces mechanical allodynia after
nociceptors in cell culture. Neurosci. Lett. 132, 105–108. spinal nerve ligation by activation of p38 MAPK in primary
Raghavendra, V., Tanga, F., DeLeo, J.A., 2003. Inhibition of sensory neurons. J. Neurosci. 23, 2517–2521.
microglial activation attenuates the development but not Schäfers, M., Marziniak, M., Sorkin, L.S., Yaksh, T.L., Sommer,
existing hypersensitivity in a rat model of neuropathy. C., 2004. Cyclooxygenase inhibition in nerve-injury-and
J. Pharmacol. Exp. Ther. 306, 624–630. TNF-induced hyperalgesia in the rat. Exp. Neurol. 185,
Ralevic, V., Burnstock, G., 1998. Receptors for purines and 160–168.
pyrimidines. Pharmacol. Rev. 50, 413–492. Scholz, J., Woolf, C.J., 2002. Can we conquer pain? Nat. Neurosci. 5,
Ramer, M.S., Murphy, P.G., Richardson, P.M., Bisby, M.A., 1998. 1062–1067 (Suppl).
Spinal nerve lesion-induced mechanoallodynia and adrenergic Schwab, J.M., Brechtel, K., Nguyen, T.D., Schluesener, H.J., 2000.
sprouting in sensory ganglia are attenuated in interleukin-6 Persistent accumulation of cyclooxygenase-1 (COX-1)
knockout mice. Pain 78, 115–121. expressing microglia/macrophages and upregulation by
Rashid, M.H., Inoue, M., Matsumoto, M., Ueda, H., 2004. Switching endothelium following spinal cord injury. J. Neuroimmunol.
of bradykinin-mediated nociception following partial sciatic 111, 122–130.
nerve injury in mice. J. Pharmacol. Exp. Ther. 308, 1158–1164. Seltzer, Z., Dubner, R., Shir, Y., 1990. A novel behavioral model of
Regoli, D., Nsa Allogho, S., Rizzi, A., Gobeil, F.J., 1998. Bradykinin neuropathic pain disorders produced in rats by partial sciatic
receptors and their antagonists. Eur. J. Pharmacol. 348, 1–10. nerve injury. Pain 43, 205–218.
Reissmann, S., Pineda, F., Vietinghoff, G., Werner, H., Gera, L., Shamash, S., Reichert, F., Rotshenker, S., 2002. The cytokine
Stewart, J.M., Paegelow, I., 2000. Structure activity relationships network of Wallerian degeneration: tumor necrosis factor-
for bradykinin antagonists on the inhibition of cytokine release alpha, interleukin-1alpha, and interleukin-1beta. J. Neurosci.
and the release of histamine. Peptides 21, 527–533. 22, 3052–3060.
Riedel, W., Neeck, G., 2001. Nociception, pain, and antinociception: Sheth, R.N., Dorsi, M.J., Li, Y., Murinson, B.B., Belzberg, A.J., Griffin,
current concepts. Z. Rheumatol. 60, 404–415. J.W., Meyer, R.A., 2002. Mechanical hyperalgesia after an L5
Ro, L.S., Chen, S.T., Tang, L.M., Jacobs, J.M., 1999. Effect of NGF and ventral rhizotomy or an L5 ganglionectomy in the rat. Pain 96,
anti-NGF on neuropathic pain in rats following chronic 63–72.
constriction injury of the sciatic nerve. Pain 79, 265–274. Shubayev, V.I., Myers, R.R., 2000. Upregulation and interaction of
Roos, D., van Bruggen, R., Meischl, C., 2003. Oxidative killing of TNFalpha and gelatinases A and B in painful peripheral nerve
microbes by neutrophils. Microbes Infect. 5, 1307–1315. injury. Brain Res. 855, 83–89.
Rueff, A., Dray, A., 1993. Pharmacological characterization of the Siuciak, J.A., Altar, C.A., Wiegand, S.J., Lindsay, R.M., 1994.
effects of 5-hydroxytryptamine and different prostaglandins Antinociceptive effect of brain-derived neurotrophic factor and
on peripheral sensory neurons in vitro. Agents Actions 38, neurotrophin-3. Brain Res. 633, 326–330.
C13–C15 (Spec No). Skundric, D.S., Bealmear, B., Lisak, R.P., 1997. Induced upregulation
Rutkowski, M.D., Pahl, J.L., Sweitzer, S., van Rooijen, N., DeLeo, J.A., of IL-1, IL-1RA and IL-1R type I gene expression by Schwann
2000. Limited role of macrophages in generation of nerve cells. J. Neuroimmunol. 74, 9–18.
injury-induced mechanical allodynia. Physiol. Behav. 71, Smith, W.L., DeWitt, D.L., Garavito, R.M., 2000. Cyclooxygenases:
225–235. structural, cellular, and molecular biology. Annu. Rev.
Sad, S., Marcotte, R., Mosmann, T.R., 1995. Cytokine-induced Biochem. 69, 145–182.
262 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

Sommer, C., 2001. Cytokines and neuropathic pain. In: Hansson, trafficking into the central nervous system: potential
P.T., et al. (Eds.), Neuropathic Pain: Pathophysiology and relationship to neuropathic pain. Pain 100, 163–170.
Treatment. IASP Press, Seattle, pp. 37–62. Syriatowicz, J.P., Hu, D., Walker, J.S., Tracey, D.J., 1999.
Sommer, C., 2004. Serotonin in pain and analgesia: actions in the Hyperalgesia due to nerve injury: role of prostaglandins.
periphery. Mol. Neurobiol. 30, 117–125. Neuroscience 94, 587–594.
Sommer, C., Kress, M., 2004. Recent findings on how Taiwo, Y.O., Levine, J.D., 1989. Prostaglandin effects after
proinflammatory cytokines cause pain: peripheral elimination of indirect hyperalgesic mechanisms in the skin of
mechanisms in inflammatory and neuropathic hyperalgesia. the rat. Brain Res. 492, 397–399.
Neurosci. Lett. 361, 184–187. Taiwo, Y.O., Levine, J.D., 1990a. Direct cutaneous hyperalgesia
Sommer, C., Galbraith, J.A., Heckman, H.M., Myers, R.R., 1993. induced by adenosine. Neuroscience 38, 757–762.
Pathology of experimental compression neuropathy Taiwo, Y.O., Levine, J.D., 1990b. Effects of cyclooxygenase products
producing hyperesthesia. J. Neuropathol. Exp. Neurol. 52, of arachidonic acid metabolism on cutaneous nociceptive
223–233. threshold in the rat. Brain Res. 537, 372–374.
Sommer, C., Lalonde, A., Heckman, H.M., Rodriguez, M., Myers, Taiwo, Y.O., Levine, J.D., 1992. Serotonin is a directly-acting
R.R., 1995. Quantitative neuropathology of a focal nerve hyperalgesic agent in the rat. Neuroscience 48, 485–490.
injury causing hyperalgesia. J. Neuropathol. Exp. Neurol. 54, Taiwo, Y.O., Levine, J.D., Burch, R.M., Woo, J.E., Mobley, W.C., 1991.
635–643. Hyperalgesia induced in the rat by the amino-terminal
Sommer, C., Marziniak, M., Myers, R.R., 1998a. The effect of octapeptide of nerve growth factor. Proc. Natl. Acad. Sci. U. S. A.
thalidomide treatment on vascular pathology and 88, 5144–5148.
hyperalgesia caused by chronic constriction injury of rat nerve. Takahashi, M., Kawaguchi, M., Shimada, K., Konishi, N., Furuya, H.,
Pain 74, 83–91. Nakashima, T., 2004. Cyclooxygenase-2 expression in Schwann
Sommer, C., Schmidt, C., George, A., 1998b. Hyperalgesia in cells and macrophages in the sciatic nerve after single spinal
experimental neuropathy is dependent on the TNF receptor 1. nerve injury in rats. Neurosci. Lett. 363, 203–206.
Exp. Neurol. 151, 138–142. Tal, M., 1996. A novel antioxidant alleviates heat hyperalgesia in
Sommer, C., Petrausch, S., Lindenlaub, T., Toyka, K.V., 1999. rats with an experimental painful peripheral neuropathy.
Neutralizing antibodies to interleukin 1-receptor reduce pain NeuroReport 7, 1382–1384.
associated behavior in mice with experimental neuropathy. Tanga, F.Y., Raghavendra, V., DeLeo, J.A., 2004. Quantitative
Neurosci. Lett. 270, 25–28. real-time RT-PCR assessment of spinal microglial and
Sommer, C., Lindenlaub, T., Teuteberg, P., Schäfers, M., Hartung, astrocytic activation markers in a rat model of neuropathic
T., Toyka, K.V., 2001a. Anti-TNF-neutralizing antibodies reduce pain. Neurochem. Int. 45, 397–407.
pain-related behavior in two different mouse models of painful Tanga, F.Y., Nutile-McMenemy, N., Deleo, J.A., 2005. The CNS role
mononeuropathy. Brain Res. 913, 86–89. of Toll-like receptor 4 in innate neuroimmunity and painful
Sommer, C., Schäfers, M., Marziniak, M., Toyka, K.V., 2001b. neuropathy. Proc. Natl. Acad. Sci. U. S. A. 102, 5856–5861.
Etanercept reduces hyperalgesia in experimental painful Tegeder, I., Niederberger, E., Schmidt, R., Kunz, S., Guhring, H.,
neuropathy. J. Peripher. Nerv. Syst. 6, 67–72. Ritzeler, O., Michaelis, M., Geisslinger, G., 2004. Specific
Sorkin, L.S., 2002. Neuroinflammation, cytokines and inhibition of IkappaB kinase reduces hyperalgesia in
neuropathic pain. In: Malmberg, A.B., Chaplan, S.R. (Eds.), inflammatory and neuropathic pain models in rats. J. Neurosci.
Mechanisms and Mediators of Neuropathic Pain. Birkhäuser 24, 1637–1645.
Verlag, Basel, pp. 67–75. Theodosiou, M., Rush, R.A., Zhou, X.F., Hu, D., Walker, J.S., Tracey,
Sorkin, L.S., Xiao, W.H., Wagner, R., Myers, R.R., 1997. Tumour D.J., 1999. Hyperalgesia due to nerve damage: role of nerve
necrosis factor-alpha induces ectopic activity in nociceptive growth factor. Pain 81, 245–255.
primary afferent fibres. Neuroscience 81, 255–262. Theoharides, T.C., Cochrane, D.E., 2004. Critical role of mast cells
Spataro, L.E., Sloane, E.M., Milligan, E.D., Wieseler-Frank, J., in inflammatory diseases and the effect of acute stress.
Schoeniger, D., Jekich, B.M., Barrientos, R.M., Maier, S.F., J. Neuroimmunol. 146, 1–12.
Watkins, L.R., 2004. Spinal gap junctions: potential Thompson, S.W., Bennett, D.L., Kerr, B.J., Bradbury, E.J., McMahon,
involvement in pain facilitation. J. Pain 5, 392–405. S.B., 1999. Brain-derived neurotrophic factor is an endogenous
Stucky, C.L., Abrahams, L.G., Seybold, V.S., 1998. Bradykinin modulator of nociceptive responses in the spinal cord. Proc.
increases the proportion of neonatal rat dorsal root ganglion Natl. Acad. Sci. U. S. A. 96, 7714–7718.
neurons that respond to capsaicin and protons. Neuroscience Tiffany, C.W., Burch, R.M., 1989. Bradykinin stimulates tumor
84, 1257–1265. necrosis factor and interleukin-1 release from macrophages.
Stucky, C.L., Medler, K.A., Molliver, D.C., 2004. The P2Y agonist UTP FEBS Lett. 247, 189–192.
activates cutaneous afferent fibers. Pain 109, 36–44. Tonra, J.R., Curtis, R., Wong, V., Cliffer, K.D., Park, J.S., Timmes, A.,
Sung, C.S., Wen, Z.H., Chang, W.K., Ho, S.T., Tsai, S.K., Chang, Y.C., Nguyen, T., Lindsay, R.M., Acheson, A., DiStefano, P.S., 1998.
Wong, C.S., 2004. Intrathecal interleukin-1beta administration Axotomy upregulates the anterograde transport and
induces thermal hyperalgesia by activating inducible nitric expression of brain-derived neurotrophic factor by sensory
oxide synthase expression in the rat spinal cord. Brain Res. neurons. J. Neurosci. 18, 4374–4383.
1015, 145–153. Tracey, D.J., Walker, J.S., 1995. Pain due to nerve damage: are
Suzuki, R., Stanfa, L.C., Kowaluk, E.A., Williams, M., Jarvis, M.F., inflammatory mediators involved? Inflamm. Res. 44, 407–411.
Dickenson, A.H., 2001. The effect of ABT-702, a novel adenosine Tsuda, M., Shigemoto-Mogami, Y., Koizumi, S., Mizokoshi, A.,
kinase inhibitor, on the responses of spinal neurones following Kohsaka, S., Salter, M.W., Inoue, K., 2003. P2X4 receptors
carrageenan inflammation and peripheral nerve injury. Br. J. induced in spinal microglia gate tactile allodynia after nerve
Pharmacol. 132, 1615–1623. injury. Nature 424, 778–783.
Sweitzer, S., Martin, D., DeLeo, J.A., 2001. Intrathecal Tsuda, M., Inoue, K., Salter, M.W., 2005. Neuropathic pain and
interleukin-1 receptor antagonist in combination with spinal microglia: a big problem from molecules in “small” glia.
soluble tumor necrosis factor receptor exhibits an anti- Trends Neurosci. 28, 101–107.
allodynic action in a rat model of neuropathic pain. Tsuzuki, K., Kondo, E., Fukuoka, T., Yi, D., Tsujino, H., Sakagami,
Neuroscience 103, 529–539. M., Noguchi, K., 2001. Differential regulation of P2X(3) mRNA
Sweitzer, S.M., Hickey, W.F., Rutkowski, M.D., Pahl, J.L., DeLeo, J.A., expression by peripheral nerve injury in intact and injured
2002. Focal peripheral nerve injury induces leukocyte neurons in the rat sensory ganglia. Pain 91, 351–360.
B RA I N RE SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4 263

Twining, C.M., Sloane, E.M., Milligan, E.D., Chacur, M., Martin, D., Watkins, L.R., Maier, S.F., 2002. Beyond neurons: evidence that
Poole, S., Marsh, H., Maier, S.F., Watkins, L.R., 2004. Peri-sciatic immune and glial cells contribute to pathological pain states.
proinflammatory cytokines, reactive oxygen species, and Physiol. Rev. 82, 981–1011.
complement induce mirror-image neuropathic pain in rats. Watkins, L.R., Maier, S.F., 2005. Immune regulation of central
Pain 110, 299–309. nervous system functions: from sickness responses to
Vanegas, H., Schaible, H.G., 2001. Prostaglandins and pathological pain. J. Intern. Med. 257, 139–155.
cyclooxygenases [correction of cycloxygenases] in the spinal Watkins, L.R., Wiertelak, E.P., Goehler, L.E., Smith, K.P., Martin, D.,
cord. Prog. Neurobiol. 64, 327–363. Maier, S.F., 1994. Characterization of cytokine-induced
Vega, J.A., Garcia-Suarez, O., Hannestad, J., Perez-Perez, M., hyperalgesia. Brain Res. 654, 15–26.
Germana, A., 2003. Neurotrophins and the immune system. Watkins, L.R., Maier, S.F., Goehler, L.E., 1995. Immune activation:
J. Anat. 203, 1–19. the role of pro-inflammatory cytokines in inflammation, illness
Verge, G.M., Milligan, E.D., Maier, S.F., Watkins, L.R., Naeve, G.S., responses and pathological pain states. Pain 63, 289–302.
Foster, A.C., 2004. Fractalkine (CX3CL1) and fractalkine receptor Watkins, L.R., Hansen, M.K., Nguyen, K.T., Lee, J.E., Maier, S.F.,
(CX3CR1) distribution in spinal cord and dorsal root ganglia 1999. Dynamic regulation of the proinflammatory cytokine,
under basal and neuropathic pain conditions. Eur. J. Neurosci. interleukin-1beta: molecular biology for non-molecular
20, 1150–1160. biologists. Life Sci. 65, 449–481.
Vergnolle, N., Bunnett, N.W., Sharkey, K.A., Brussee, V., Compton, Watkins, L.R., Milligan, E.D., Maier, S.F., 2001. Glial activation: a
S.J., Grady, E.F., Cirino, G., Gerard, N., Basbaum, A.I., Andrade- driving force for pathological pain. Trends Neurosci. 24,
Gordon, P., Hollenberg, M.D., Wallace, J.L., 2001. Proteinase- 450–455.
activated receptor-2 and hyperalgesia: a novel pain pathway. Welling, M.M., Hiemstra, P.S., van den Barselaar, M.T.,
Nat. Med. 7, 821–826. Paulusma-Annema, A., Nibbering, P.H., Pauwels, E.K.,
Vilhardt, F., 2005. Microglia: phagocyte and glia cell. Int. J. Calame, W., 1998. Antibacterial activity of human neutrophil
Biochem. Cell Biol. 37, 17–21. defensins in experimental infections in mice is
Vogel, C., Mossner, R., Gerlach, M., Heinemann, T., Murphy, D.L., accompanied by increased leukocyte accumulation. J. Clin.
Riederer, P., Lesch, K.P., Sommer, C., 2003. Absence of thermal Invest. 102, 1583–1590.
hyperalgesia in serotonin transporter-deficient mice. Weskamp, G., Otten, U., 1987. An enzyme-linked immunoassay for
J. Neurosci. 23, 708–715. nerve growth factor (NGF): a tool for studying regulatory
von Kugelgen, I., Wetter, A., 2000. Molecular pharmacology of mechanisms involved in NGF production in brain and in
P2Y-receptors. Naunyn-Schmiedeberg's Arch. Pharmacol. 362, peripheral tissues. J. Neurochem. 48, 1779–1786.
310–323. White, D.M., Basbaum, A.I., Goetzl, E.J., Levine, J.D., 1990. The
Wagner, R., Myers, R.R., 1996a. Schwann cells produce tumor 15-lipoxygenase product, 8R,15S-diHETE, stereospecifically
necrosis factor alpha: expression in injured and non-injured sensitizes C-fiber mechanoheat nociceptors in hairy skin of rat.
nerves. Neuroscience 73, 625–629. J. Neurophysiol. 63, 966–970.
Wagner, R., Myers, R.R., 1996b. Endoneurial injection of TNF-alpha White, F.A., Bhangoo, S.K., Miller, R.J., 2005. Chemokines:
produces neuropathic pain behaviors. NeuroReport 7, integrators of pain and inflammation. Nat. Rev., Drug Discov. 4,
2897–2901. 834–844.
Wagner, R., Janjigian, M., Myers, R.R., 1998. Anti-inflammatory Wilson-Gerwing, T.D., Dmyterko, M.V., Zochodne, D.W.,
interleukin-10 therapy in CCI neuropathy decreases thermal Johnston, J.M., Verge, V.M., 2005. Neurotrophin-3 suppresses
hyperalgesia, macrophage recruitment, and endoneurial thermal hyperalgesia associated with neuropathic pain and
TNF-alpha expression. Pain 74, 35–42. attenuates transient receptor potential vanilloid receptor-1
Wallace, V.C., Cottrell, D.F., Brophy, P.J., Fleetwood-Walker, S.M., expression in adult sensory neurons. J. Neurosci. 25, 758–767.
2003. Focal lysolecithin-induced demyelination of peripheral Witko-Sarsat, V., Rieu, P., Descamps-Latscha, B., Lesavre, P.,
afferents results in neuropathic pain behavior that is Halbwachs-Mecarelli, L., 2000. Neutrophils: molecules,
attenuated by cannabinoids. J. Neurosci. 23, 3221–3233. functions and pathophysiological aspects. Lab. Invest. 80,
Walsh, L.J., Trinchieri, G., Waldorf, H.A., Whitaker, D., Murphy, 617–653.
G.F., 1991. Human dermal mast cells contain and release Wolfe, L.S., Horrocks, L.A., 1994. Eicosanoids. In: Siegel, G.J., et al.
tumor necrosis factor alpha, which induces endothelial (Eds.), Basic Neurochemistry : Molecular, Cellular, and Medical
leukocyte adhesion molecule 1. Proc. Natl. Acad. Sci. U. S. A. Aspects. Lippincott-Raven, Philadelphia, pp. 475–490.
88, 4220–4224. Wood, J.N., Docherty, R., 1997. Chemical activators of sensory
Wang, Y.X., Bowersox, S.S., Pettus, M., Gao, D., 1999. neurons. Annu. Rev. Physiol. 59, 457–482.
Antinociceptive properties of fenfluramine, a serotonin Woolf, C.J., 2004. Dissecting out mechanisms responsible for
reuptake inhibitor, in a rat model of neuropathy. J. Pharmacol. peripheral neuropathic pain: implications for diagnosis and
Exp. Ther. 291, 1008–1016. therapy. Life Sci. 74, 2605–2610.
Wang, R., Guo, W., Ossipov, M.H., Vanderah, T.W., Porreca, F., Lai, Woolf, C.J., Costigan, M., 1999. Transcriptional and
J., 2003. Glial cell line-derived neurotrophic factor normalizes posttranslational plasticity and the generation of
neurochemical changes in injured dorsal root ganglion inflammatory pain. Proc. Natl. Acad. Sci. U. S. A. 96, 7723–7730.
neurons and prevents the expression of experimental Woolf, C.J., Safieh-Garabedian, B., Ma, Q.P., Crilly, P., Winter, J.,
neuropathic pain. Neuroscience 121, 815–824. 1994. Nerve growth factor contributes to the generation of
Wang, H., Kohno, T., Amaya, F., Brenner, G.J., Ito, N., inflammatory sensory hypersensitivity. Neuroscience 62,
Allchorne, A., Ji, R.R., Woolf, C.J., 2005. Bradykinin produces 327–331.
pain hypersensitivity by potentiating spinal cord Woolf, C.J., Ma, Q.P., Allchorne, A., Poole, S., 1996. Peripheral cell
glutamatergic synaptic transmission. J. Neurosci. 25, types contributing to the hyperalgesic action of nerve growth
7986–7992. factor in inflammation. J. Neurosci. 16, 2716–2723.
Warfield, C.A., Fausett, H.J., 2002. Manual of Pain Management. Xiao, H.S., Huang, Q.H., Zhang, F.X., Bao, L., Lu, Y.J., Guo, C., Yang,
Lippincott Williams and Wilkins, Philadelphia. L., Huang, W.J., Fu, G., Xu, S.H., Cheng, X.P., Yan, Q., Zhu, Z.D.,
Watkins, L.R., Maier, S.F., 1999. Illness-induced hyperalgesia: Zhang, X., Chen, Z., Han, Z.G., Zhang, X., 2002. Identification of
mediators, mechanisms and implications. In: Watkins, L.R., gene expression profile of dorsal root ganglion in the rat
Maier, S.F. (Eds.), Cytokines and Pain. Birkhäuser Verlag, Basel, peripheral axotomy model of neuropathic pain. Proc. Natl.
pp. 39–57. Acad. Sci. U. S. A. 99, 8360–8365.
264 B RA I N R E SE A R CH RE V I EW S 51 ( 20 0 6 ) 2 4 0–2 6 4

Yajima, Y., Narita, M., Narita, M., Matsumoto, N., Suzuki, T., Zelenka, M., Schäfers, M., Sommer, C., 2005. Intraneural injection
2002. Involvement of a spinal brain-derived neurotrophic of interleukin-1beta and tumor necrosis factor-alpha into rat
factor/full-length TrkB pathway in the development of nerve sciatic nerve at physiological doses induces signs of
injury-induced thermal hyperalgesia in mice. Brain Res. 958, neuropathic pain. Pain 116, 257–263.
338–346. Zhao, Z., Chen, S.R., Eisenach, J.C., Busija, D.W., Pan, H.L., 2000.
Yajima, Y., Narita, M., Usui, A., Kaneko, C., Miyatake, M., Narita, M., Spinal cyclooxygenase-2 is involved in development of
Yamaguchi, T., Tamaki, H., Wachi, H., Seyama, Y., Suzuki, T., allodynia after nerve injury in rats. Neuroscience 97,
2005. Direct evidence for the involvement of brain-derived 743–748.
neurotrophic factor in the development of a neuropathic pain- Zhou, X.F., Rush, R.A., 1996. Endogenous brain-derived
like state in mice. J. Neurochem. 93, 584–594. neurotrophic factor is anterogradely transported in primary
Yamaguchi-Sase, S., Hayashi, I., Okamoto, H., Nara, Y., sensory neurons. Neuroscience 74, 945–953.
Matsuzaki, S., Hoka, S., Majima, M., 2003. Amelioration of Zhou, J., Chung, K., Chung, J.M., 2001. Development of purinergic
hyperalgesia by kinin receptor antagonists or kininogen sensitivity in sensory neurons after peripheral nerve injury in
deficiency in chronic constriction nerve injury in rats. the rat. Brain Res. 915, 161–169.
Inflamm. Res. 52, 164–169. Zhu, X., Eisenach, J.C., 2003. Cyclooxygenase-1 in the spinal cord is
Yamaki, K., Thorlacius, H., Xie, X., Lindbom, L., Hedqvist, P., Raud, altered after peripheral nerve injury. Anesthesiology 99,
J., 1998. Characteristics of histamine-induced leukocyte rolling 1175–1179.
in the undisturbed microcirculation of the rat mesentery. Br. J. Zhu, C.Z., Mikusa, J., Chu, K.L., Cowart, M., Kowaluk, E.A., Jarvis,
Pharmacol. 123, 390–399. M.F., McGaraughty, S., 2001. A-134974: a novel adenosine
Yiangou, Y., Facer, P., Birch, R., Sangameswaran, L., Eglen, R., kinase inhibitor, relieves tactile allodynia via spinal sites of
Anand, P., 2000. P2X3 receptor in injured human sensory action in peripheral nerve injured rats. Brain Res. 905,
neurons. NeuroReport 11, 993–996. 104–110.
Yoder, E.J., Tamir, H., Ellisman, M.H., 1997. Serotonin receptors Zuo, Y., Perkins, N.M., Tracey, D.J., Geczy, C.L., 2003. Inflammation
expressed by myelinating Schwann cells in rat sciatic nerve. and hyperalgesia induced by nerve injury in the rat: a key role
Brain Res. 753, 299–308. of mast cells. Pain 105, 467–479.

You might also like