ArticlePDF AvailableLiterature Review

Central modulaton of pain

Authors:

Abstract and Figures

It has long been appreciated that the experience of pain is highly variable between individuals. Pain results from activation of sensory receptors specialized to detect actual or impending tissue damage (i.e., nociceptors). However, a direct correlation between activation of nociceptors and the sensory experience of pain is not always apparent. Even in cases in which the severity of injury appears similar, individual pain experiences may vary dramatically. Emotional state, degree of anxiety, attention and distraction, past experiences, memories, and many other factors can either enhance or diminish the pain experience. Here, we review evidence for "top-down" modulatory circuits that profoundly change the sensory experience of pain.
Content may be subject to copyright.
Review series
The Journal of Clinical Investigation      http://www.jci.org      Volume 120      Number 11      November 2010  3779
Central modulation of pain
Michael H. Ossipov, Gregory O. Dussor, and Frank Porreca
Department of Pharmacology, University of Arizona, Tucson, Arizona, USA.
It has long been appreciated that the experience of pain is highly variable between individuals. Pain results from
activation of sensory receptors specialized to detect actual or impending tissue damage (i.e., nociceptors). However,
a direct correlation between activation of nociceptors and the sensory experience of pain is not always apparent.
Even in cases in which the severity of injury appears similar, individual pain experiences may vary dramatically.
Emotional state, degree of anxiety, attention and distraction, past experiences, memories, and many other factors
can either enhance or diminish the pain experience. Here, we review evidence for “top-down” modulatory circuits
that profoundly change the sensory experience of pain.
Existence of an endogenous pain inhibitory system
Early evidence for pain modulatory mechanisms came from obser-
vations of H.K. Beecher, who noted a remarkable attenuation of 
pain experienced by soldiers in combat situations (1). Analogous 
observations have been seen in others, including athletes that con-
tinue competition despite significant injuries (see ref. 2). Beecher, a 
physician who served with the US Army during the Second World 
War, observed that as many as three-quarters of badly wounded 
soldiers reported no to moderate pain and did not want pain relief 
medication (1). This observation was striking, because the wounds 
were not trivial but consisted of compound fractures of long bones 
or penetrating wounds of the abdomen, thorax, or cranium. More-
over, only individuals who were clearly alert, responsive, and not 
in shock were included in his report (1), leading to the conclusion 
that “strong emotions” block pain (1).
The existence of endogenous mechanisms that diminish pain 
through net “inhibition” is now generally accepted. Pain modu-
lation likely exists in the form of a descending pain modulatory 
circuit with inputs that arise in multiple areas, including the hypo-
thalamus, the amygdala, and the rostral anterior cingulate cortex 
(rACC), feeding to the midbrain periaqueductal gray region (PAG), 
and with outputs from the PAG to the medulla. Neurons within 
the nucleus raphe magnus and nucleus reticularis gigantocellu-
laris, which are included within the rostral ventromedial medulla 
(RVM), have been shown to project to the spinal or medullary dor-
sal horns to directly or indirectly enhance or diminish nocicep-
tive traffic, changing the experience of pain (3). This descending 
modulatory circuit is an “opioid-sensitive” circuit (see below) and 
relevant to human experience in many settings, including in states 
of chronic pain, and in the actions of pain-relieving drugs, includ-
ing opiates, cannabinoids, NSAIDs, and serotonin/norepineph-
rine reuptake blockers that mimic, in part, the actions of opiates 
(Figure 1). While the precise mechanisms by which drugs produce 
pain relief is not entirely understood, strong evidence supports the 
actions of these drugs through the pain modulatory circuit or by 
mimicking the consequence of activation of this descending cir-
cuit at the level of the spinal cord.
“Top-down” modulatory pathways have been shown to under-
lie the robust and clinically important phenomenon of placebo 
analgesia, which can be demonstrated in approximately one-third 
of the population (4). Patients that had undergone  removal of 
impacted molars and who were expecting  an analgesic showed 
reduced pain scores after placebo injection (5). Placebo respond-
ers that blindly received the opiate antagonist naloxone indicated 
pain levels similar to those of the nonresponders, indicating that 
placebo analgesia required activation of endogenous opioid-medi-
ated inhibition (5). Neuroimaging  techniques have now estab-
lished that the placebo response is likely mediated by activation 
of pain inhibitory systems, originating from cortical and subcorti-
cal regions (6, 7). Human imaging studies with [11C]-carfentanil 
revealed that placebo analgesia was related to activation of μ-opi-
oid receptors in the rACC, the pregenual cingulate cortex (pCC), 
the dorsolateral prefrontal cortex, and the anterior insular cortex 
(7). Changes in regional blood flow revealed that expectation of 
placebo analgesia activated a neural network from the rACC to 
include subcortical regions known to be active in opioid-mediated 
antinociception, such as the PAG (6). Increased regional cerebral 
blood flow to these sites was associated with a greater placebo 
response, leading to the suggestion that individual variations in 
placebo responses may be linked to differences in either concentra-
tion or function of μ-opioid receptors (6).
Imaging studies have led to the suggestion of a “pain matrix,” 
brain areas that are consistently activated by noxious stimuli. 
These areas often include, but are not restricted to, the rACC, 
pCC, somatosensory cortex 1 and 2, the insula, amygdala and 
thalamus, and the PAG (8). Interestingly, these regions demon-
strate overlap among brain sites activated by opioids and those 
that are activated by placebo analgesia, and imaging studies sug-
gest that coupling between the rACC and the PAG is mediated 
through endogenous opioidergic signaling and is essential to 
both opioid-induced analgesia and placebo-mediated analgesia 
(9). It should be noted that the concept of a pain matrix is not 
meant to suggest a rigid regulatory pathway but rather concep-
tually represents a collection of brain regions that are involved 
in neurological functions, including cognition, emotion, moti-
vation, and  sensation  as  well  as  pain. These regions, acting 
together in the context of modulation of nociception, appear to 
give rise to the experience of pain (10). It is noted that analgesic 
drugs as well as expectation, distraction, emotional context, and 
other factors engage several nodes of the pain matrix to change 
the pain experience.
Engagement of descending modulation can facilitate, as well as 
inhibit, pain. The term “nocebo” has been introduced to describe 
an effect opposite to that of the placebo, indicated by expectation 
of a worsening outcome in response to a treatment (11). For exam-
ple, patients who were expecting pain relief with a NSAID and 
were then told they were to receive a drug that was hyperalgesic 
Conflict of interest: The authors have declared that no conflict of interest exists.
Citation for this article:J Clin Invest. 2010;120(11):3779–3787. doi:10.1172/JCI43766.
review series
3780 The Journal of Clinical Investigation      http://www.jci.org      Volume 120      Number 11      November 2010
responded with enhanced pain (12). When subjects were told ver-
bally, nonverbally through the application of conditioning stim-
uli, or both ways that enhanced pain was to be expected, it was 
found that expectation of pain resulted in pain to nonpainful 
stimuli as well as enhanced pain in response to noxious stimuli 
(13). In order to isolate the effect of expectancy in an imaging 
study, subjects were presented with visual cues indicating that 
either a high or low noxious thermal stimulus would be applied 
but then were actually presented with the high stimulus (14). 
This procedure revealed changes in the ipsilateral caudal ACC, 
the head of the caudate, the cerebellum, and the contralateral 
cuneiform nucleus (nCF), suggesting that increased pain expec-
tancy activates a pain network that modulates afferent input at 
the level of the nCF (14).
Neuroanatomical and electrophysiological evidence
of endogenous pain inhibition
Although the existence of pain modulatory systems had been sur-
mised for many decades, it was not until electrical stimulation or 
microinjection of opiates into specific brain regions that the impor-
tance and clinical significance of such systems was appreciated. In 
what may have been the first demonstration of a brain site-specific 
action for the antinociceptive effect of morphine, Tsou and Jang 
surmised that since morphine blocks pain  at doses that do not 
affect other sensory modalities, it was likely working through a site 
specific for pain control. Thus, they microinjected morphine into 
several regions of the rabbit brain and discovered that a profound 
antinociceptive effect occurred only when morphine was applied 
into the PAG (15). Reynolds found that electrical stimulation of the 
ventrolateral PAG of the rat produced an antinociception so power-
ful that a laparotomy could be performed in a fully conscious rat, 
without observable signs of distress to the animal (16).
Electrical stimulation of the PAG was rapidly adapted to humans 
in efforts to relieve intractable pain (17–19). While PAG stimulation 
has been largely discontinued because of side effects, such as anxiety, 
distress, and, in some instances, development of migraine-like head-
ache (20), deep brain stimulation aimed at other regions remains an 
approach that might control otherwise intractable pain (21). Criti-
cally, the reversal of intractable pain by stimulation of the PAG was 
blocked by naloxone, indicating the activation of an endogenous 
opioidergic pain inhibitory system (17). These early studies were 
not rigorous, placebo-controlled double-blind trials, and, as a con-
sequence, the possibility of placebo analgesia cannot be disregarded. 
Even so, the existence of a placebo effect, as discussed above, is likely 
dependent on activation of pain modulatory circuits.
Preclinical studies  have attempted to delineate the sites  and 
pathways that compose the endogenous pain inhibitory circuit. 
Considerable overlap has been found between sites that produce 
antinociception with either  electrical stimulation or morphine 
Figure 1
Schematic representation of pain modularity circuitry. Nociceptive
inputs enter the spinal dorsal horn through primary afferent fibers that
synapse onto transmission neurons. The projection fibers ascend
through the contralateral spinothalamic tract. Ascending projections
target the thalamus, and collateral projections also target mesence-
phalic nuclei, including the DRt, the RVM, and the midbrain PAG.
Descending projections from the DRt are a critical component of the
DNIC pathway. Rostral projections from the thalamus target areas that
include cortical sites and the amygdala. The lateral capsular part of
the CeA (“nociceptive amygdala”) receives nociceptive inputs from the
brainstem and spinal cord. Inputs from the thalamus and cortex enter
through the lateral (LA) and basolateral (BLA) amygdala. The CeA
sends outputs to cortical sites and the thalamus, in which cognitive and
conscious perceptions of pain are integrated. Descending pain modu-
lation is mediated through projections to the PAG, which also receives
inputs from other sites, including the hypothalamus (data not shown),
and communicates with the RVM as well as other medullary nuclei that
send descending projections to the spinal dorsal horn through the DLF.
The noradrenergic locus coeruleus (LC) receives inputs from the PAG,
communicates with the RVM, and sends descending noradrenergic
inhibitory projections to the spinal cord. Antinociceptive and prono-
ciceptive spinopetal projections from the RVM positively and nega-
tively modulate nociceptive inputs and provide for an endogenous pain
regulatory system. Ascending (red) and descending (green) tracts are
shown schematically. Areas labeled “i–iv” in the small diagram cor-
respond with labeled details of the larger diagram.
review series
The Journal of Clinical Investigation      http://www.jci.org      Volume 120      Number 11      November 2010  3781
microinjection (22–26). Both stimulation-produced antinocicep-
tion (SPA) and antinociception from morphine microinjection 
into supraspinal loci are reversed by naloxone, further implicating 
the activation of endogenous opioidergic systems in these phe-
nomena (27). These studies have revealed descending pain inhibi-
tory projections to the level of the spinal cord, either directly or 
indirectly from the PAG. Surgical disruption of the dorsolateral 
funiculus (DLF) abolished supraspinally mediated antinociception 
(26), and anterograde and retrograde tracing studies revealed that 
the RVM sends spinopetal projections through this tract (28–30). 
The precise site of projection of these fibers and their role in inhi-
bition or facilitation remains unclear.
The amygdala in descending modulation
Human imaging studies reveal connections linking the PAG to 
the amygdala and cortical sites (2, 31). These studies suggest that 
interactions between the prefrontal cortex and the amygdala pro-
vide emotional-affective modulation of cognitive  functions in 
pain, driving tasks such as decision making, assessment of risk/
reward versus pain, or punishment avoidance (32). The amygdala 
plays important roles in emotional responses, stress, and anxiety 
and is believed to be a critical component of the pain matrix. This 
region may contribute significantly to the integration of pain and 
associated responses such as fear and anxiety.
Electrophysiological studies in animals demonstrated that neu-
rons of the central nucleus of the amygdala (CeA) showed exci-
tation with noxious stimulation of the knee joint or deep tissue 
(33) and enhanced responses after peripheral (34) or visceral (35) 
inflammation. Sensitization of CeA neurons, mediated through 
metabotropic glutamate receptors, represents neuroplastic chang-
es that appear to promote chronic pain (36, 37). Administration 
of a corticotropin-releasing factor (CRF1) receptor antagonist 
into the CeA of rats inhibited both nociceptive responses as well 
as anxiety-like behaviors (38). Hemispheric lateralization of the 
role of the amygdala in pain processing has been recently dem-
onstrated, since, although both the left and right CeA  showed 
responses to brief noxious stimuli, only the right CeA respond-
ed with enhancement of firing and increased receptive field size 
after either ipsilateral or contralateral peripheral inflammation 
(39). Moreover, peripheral inflammation produced activation of 
extracellular signal-regulated kinase cascade only in the right CeA, 
regardless of site of inflammation (40), and blockade of activity of 
this kinase in the right CeA, but not the left CeA, blocked behav-
ioral signs of enhanced inflammatory pain (40).
The RVM and descending modulation: ON and OFF cells
Electrophysiologic studies and lesioning experiments have revealed 
that the RVM receives neuronal inputs from the PAG and is likely 
to be the final common relay in descending inhibition of noci-
ception from supraspinal sites (41). The microinjection of lido-
caine into the RVM abolished antinociception arising from elec-
trical stimulation of the PAG (42). Descending projections from 
the RVM course through the DLF to the spinal dorsal horn and 
form synaptic connections with primary afferent terminals and 
second- and third-order neurons that transmit nociceptive signals 
to supraspinal sites as well as with interneurons and thus are well 
situated to modulate nociceptive inputs (43–45).
Important insights into the nature of descending modulatory cir-
cuitry came from studies by Fields and colleagues, in which activity 
of neurons in the RVM were paired with a behavior elicited by a nox-
ious stimulus (i.e., the tail-flick response to noxious heat) in lightly 
anesthetized rats (41, 46, 47). These studies led to the identification 
of a population of RVM neurons that increase firing just prior to 
the initiation of the nociceptive reflex (i.e., “on-cells”), and another 
population of neurons was found to decrease firing just prior to the 
tail-flick (i.e., “off-cells”) (48–50). Activity of other “neutral” cells did 
not correlate with nociceptive stimuli. Both the off-cells and on-cells 
were found to project to the spinal dorsal horn, indicating that they 
may exert modulatory influences on nociceptive inputs (51–52).
This dichotomy in neuronal function is consistent with bidirec-
tional pain modulation. Studies performed with electrical stimula-
tion or microinjection of glutamate into the RVM revealed a bipha-
sic function of the RVM, with regard to pain modulation (53–56). 
Low intensities of stimulation inhibited  nociceptive responses, 
whereas higher levels of stimulation enhanced nociception (53–57). 
This biphasic role of the RVM in pain modulation was shown with 
electrophysiologic responses of spinal cord neurons or with behav-
ioral responses and occurred  with either cutaneous  or visceral 
stimuli (53–57). The electrophysiologic characteristics of the on-
cells are consistent with a pronociceptive function. For example, 
prolonged delivery of a noxious thermal stimulus increased the 
firing rate of RVM on-cells as well as enhanced the intensity of 
the nociceptive response in rats (58). Both enhanced nociceptive 
responses and increased on-cell activity were abolished by lido-
caine microinjected into the RVM (58). Hyperalgesia caused by nal-
oxone-precipitated withdrawal was accompanied by increased on-
cell activity (59). Finally, the subdermal injection of formalin into 
a hind paw of a rat produced exaggerated behavioral responses as 
well as increased responses of on-cells of the RVM (60).
The activation of descending inhibitory pathways that project to 
the spinal and medullary dorsal horns has led to the question of 
the nature of these projections. Opioids administered systemically 
or into the PAG result in increased activity of off-cells through 
disinhibition, and it is believed that activation of off-cells is “nec-
essary and sufficient” for analgesia (61–63). In contrast, the on-
cells are the only population of cells in the RVM directly inhibited 
by opioids, suggesting that these cells likely express the μ-opioid 
receptors (64). These cells are also activated by cholecystokinin 
(CCK) via a CCK2 receptor (48, 64, 65). Neuroanatomical stud-
ies reported a high degree of colocalization of CCK2 receptors 
with μ-opioid receptors on RVM neurons, presumed to be pain 
facilitation cells that may correspond with on-cells (66). Addition-
ally, most (i.e., >60%) off-cells, on-cells, and neutral cells have been 
shown to express glutamate decarboxylase (67). However, the role 
of GABA in the function of these cells remains unclear.
Descending serotonergic pathways
Early studies with  available  serotonergic  antagonists  blocked 
SPA initiated from the RVM (68), leading to the suggestion that 
descending inhibition of pain was mediated through serotonergic 
neurons projecting from the RVM through the DLF (29). Stimula-
tion of the PAG or RVM was found to cause release of serotonin 
in the spinal cord (69), and intrathecal administration of 5-HT 
agonists elicited antinociception (70), whereas intrathecal 5-HT 
antagonists attenuated SPA from the RVM (71). Retrograde label-
ing studies demonstrated the presence of serotonergic projections 
to the spinal dorsal horn arising from the nucleus raphe magnus, 
which is a midline structure within the RVM as well as the nucle-
us paragigantocellularis and the ventral portion of the nucleus 
gigantocellularis (72). Together, such studies led to the reason-
review series
3782 The Journal of Clinical Investigation      http://www.jci.org      Volume 120      Number 11      November 2010
able assumption that the RVM provided descending serotonergic 
pain modulation from the RVM. However, attempts to determine 
whether either the on-cells or off-cells of the RVM are serotonergic 
led to the realization that other, nonserotonergic neurons from 
the RVM may modulate pain (73). In a study of 25 identified RVM 
neurons, none  of the on-cells (i.e.,  8 neurons) or off-cells  (i.e., 
9 neurons) expressed 5-HT, and only 4 out of 8 neutral cells were 
labeled with 5-HT (73). Moreover, only 20% of RVM neurons were 
found to be serotonergic (74), and most of the spinal projections 
from the RVM are either glycinergic or  GABAergic. It has been 
argued that serotonergic RVM neurons are neither on-cells nor off-
cells but that they can modulate the activities of these neurons (see 
refs. 75 and 76). However, a recent study, in which descending sero-
tonergic RVM neurons were selectively ablated through the use of 
shRNA plasmids and electroporation, demonstrated that descend-
ing serotonergic projections from the RVM are important for facili-
tation of pain in inflammatory or neuropathic pain states, although 
they are not necessary for opioid-mediated inhibition of acute 
pain (77). Electrophysiologic studies suggested that GABAergic 
and glycinergic projections from the RVM mediate antinocicep-
tion. In addition to the descending serotonergic populations that 
are activated, the diversity of subtypes of the 5-HT receptors and 
the complex anatomy of the spinal dorsal horn complicates inter-
pretation of the role of serotonin in pain modulation.
The effect of spinal serotonin can be either inhibitory or facilita-
tory, depending on the receptor subtype activated (78–82). Spinal 
administration of an antagonist of the inhibitory 5-HT7 receptor 
blocked the antinociceptive effect of morphine microinjected into 
the RVM, whereas pharmacological antagonism of the facilitatory 
5-HT3 receptor blocked hyperalgesia induced by CCK adminis-
tered into  the  RVM(79).  Further,  systemic  administration  of 
5-HT7 agonists blocked capsaicin-induced hyperalgesia in mice, 
whereas 5-HT7 antagonists elicited mechanical hypersensitivity 
(83). The 5-HT7 receptor has been identified in the dorsal root 
ganglion and  on  central  terminals of primary  afferent  fibers 
(84, 85) as well as on GABAergic interneurons in the dorsal horn 
of the spinal cord (84), which is consistent with a role in pain 
modulation (83). Although these observations indicate an impor-
tant serotonergic role for  pain modulation, the  precise spinal 
mechanisms involved remain unclear.
Noradrenergic systems and pain modulation
Electrical stimulation of the PAG or RVM to elicit antinociception 
increases measured norepinephrine levels in the cerebrospinal fluid, 
andthis effectwas blocked by spinal adrenergic antagonists(69, 86–88). 
These findings suggest a strong contribution of norepinephrine in 
antinociception associated with descending inhibition. While nei-
ther the PAG nor the RVM contain noradrenergic neurons, both 
regions communicate with noradrenergic sites important to pain 
modulation, including the A5 (locus coeruleus), A6, and A7 (Köl-
liker-Füse) nuclei (89–91). These noradrenergic nuclei are a major 
source of direct noradrenergic projections to the spinal cord (3, 92) 
and likely may serve to ultimately inhibit the response of presynap-
tic and postsynaptic spinal pain transmission neurons (3, 92).
Numerous studies have demonstrated that activation of spinal 
α2-adrenergic receptors exerts a strong antinociceptive effect (93–95). 
Spinal clonidine blocked thermal and capsaicin-induced pain in 
healthy human volunteers (96). PAG activation resulted in inhibition 
of the nociceptive responses of dorsal neurons mediated through 
activation of spinal α2receptors (97). Activation of α2-adrenergic 
receptors has been shown to inhibit nociceptive transmission at 
the level of the spinal cord through presynaptic activity, inhibiting 
release of excitatory neurotransmitters from primary afferent ter-
minals, as well as through postsynaptic sites (93). Recordings per-
formed on spinal cord slices revealed that activation of α2-adrenergic 
receptors hyperpolarized neurons and was thus inhibitory. Recently, 
it has also been demonstrated that activation of α1-adrenergic recep-
tors caused depolarization of GABA interneurons (98), demonstrat-
ing an additional mechanism of enhancing inhibition. Activation 
of spinal α1-adrenergic receptors also enhances responses of dorsal 
horn neurons to noxious inputs (97).
Descending modulation and stress-induced analgesia
The mechanisms mediating the suppression of pain by stress have 
been intensively studied. Watkins and colleagues (99) found that 
stress induced by brief foot shock of  the forepaws of rats pro-
duced antinociception as measured in the tail-flick test. Lesions 
of the DLF made rostral to the entry zone of the peripheral nerves 
of the forelimbs, which kept intact any direct spinal communica-
tions between forelimb and tail, abolished stress-induced analge-
sia (SIA), indicating that supraspinal sites were necessary to acti-
vate a spinopetal pain inhibitory circuit (99). Additionally, it was 
found that antinociception induced by brief shock of the fore-
paws was abolished by systemic and intrathecal naloxone, indi-
cating the activation of endogenous opioidergic pain inhibitory 
systems (99). Stress induced by foot shock reduced firing of RVM 
on-cells and increased that of off-cells, consistent with opioider-
gic endogenous pain modulatory systems (100). SIA is associated 
with elevated PAG levels of β-endorphin (101), and microinjection 
of μ-opioid receptor antagonists into the PAG or RVM abolished 
SIA (102–104). Opioid microinjection into the amygdala elicits 
antinociception that is blocked by lidocaine in either the PAG or 
RVM (105). These and other studies led to the conclusion that SIA 
can be opioid sensitive and mediated through descending inhibi-
tory pathways from amygdala, the PAG, and through RVM projec-
tions to the spinal cord (106).
However, preclinical studies have also revealed that some aspects 
of SIA are not sensitive to naloxone and therefore are likely to be 
mediated via different mechanisms. Recent studies have revealed a 
role of endogenous cannabinoids in SIA and in descending modu-
latory pathways. Inhibition of RVM activity by microinjection of 
muscimol abolished antinociception induced by systemic injec-
tion of the cannabinoid agonist WIN55,212-2 (107). Moreover, 
WIN55,212-2 increased RVM off-cell activity and reduced firing 
of the RVM on-cells, analogous to  the effect of morphine, but 
these effects were not blocked by naloxone, indicating that these 
effects are mediated specifically through cannabinoid receptors 
(107). Studies with a CB1 antagonist revealed that tonic release of 
endogenous cannabinoids increases off-cell activity and diminish-
es on-cell firing and may modulate baseline nociceptive thresholds 
through regulation of RVM activity (107), mechanisms that could 
also underlie opioid-insensitive SIA. Opioid-insensitive SIA was 
abolished by systemic administration of CB1, but not CB2, antag-
onists (108). Microinjection of the CB1 antagonist rimonabant 
into the dorsolateral PAG abolished such antinociception, further 
suggesting that  SIA is mediated by  endogenous  cannabinoids 
(108). Opioid-insensitive SIA was associated with elevated levels 
of endogenous cannabinoids in the PAG, and SIA was enhanced 
by microinjection of inhibitors of monoacylglycerol lipase, which 
hydrolyzes the endogenous cannabinoid 2-arachidonoylglycerol 
review series
The Journal of Clinical Investigation      http://www.jci.org      Volume 120      Number 11      November 2010  3783
(108). Finally, microinjection of a CB1 antagonist into the RVM 
blocked SIA, whereas inhibition of hydrolysis of endogenous can-
nabinoids in the RVM enhanced SIA (109). These studies indicate 
that endogenous cannabinoids, like opioids, regulate pain sensitiv-
ity in response to environmental conditions through descending 
pathways (109). As SIA produces many generalized effects, includ-
ing release  of stress hormones, multiple physiological  actions 
result that may contribute to antinociceptive effects and to pain.
Descending facilitation and experimental chronic pain
Increased descending facilitation in experimental chronic pain mod-
els has been demonstrated; but, to date, how this mechanism par-
ticipates in clinical conditions has not been determined. Emerging 
preclinical evidence suggests that activation of putative pain facili-
tation cells maintains descending facilitation and promotes neuro-
pathic pain. The microinjection of lidocaine into the RVM of rats 
with peripheral nerve injury abolished behavioral signs of enhanced 
abnormal pain  (110–112).  Moreover, the  surgical  disruption of 
the DLF ipsilateral but not contralateral to nerve injury abolished 
behavioral signs of enhanced abnormal pain but did not alter nor-
mal responses in sham-operated animals (111, 113). Microinjection 
of CCK into the RVM produced behavioral evidence of enhanced 
nociception that was blocked by lesion of the DLF (112, 114) and 
markedly increased on-cell activity (115).  Accordingly, microin-
jection of the CCK2 antagonist, L365,260, into the RVM reversed 
behavioral signs of neuropathic pain in nerve-injured rats (112).
Microinjection of the potent μ-opioid agonist dermorphin, con-
jugated to the ribosome-inactivating protein saporin, to rats with 
peripheral nerve injury produced a selective knockdown of RVM 
neurons that express the μ-opioid receptor, along with a rever-
sal of behavioral signs of neuropathic pain (111, 116). Addition-
ally, the selective knockdown of CCK2-expressing neurons using 
CCK-saporin resulted in a substantial reduction in RVM neurons 
expressing the μ-opioid receptor, whereas the knockdown of RVM 
neurons expressing μ-opioid receptors with the dermorphin-sapo-
rin conjugate resulted in a substantial reduction in numbers of 
neurons expressing CCK2 (66). Both of these manipulations abol-
ished behavioral and neurochemical signs of neuropathic pain in 
rats with spinal nerve ligation (66). Furthermore, a recent study 
demonstrated that blockade of RVM activity by microinjection of 
lidocaine elicited reward in models of neuropathic pain, suggest-
ing that descending facilitation also likely contributes to tonic-
aversive (i.e., stimulus-independent) aspects of such pain (117).
Activation of descending facilitation  after  peripheral  nerve 
injury has been associated with pronociceptive changes in the spi-
nal cord. Peripheral nerve injury resulted in enhanced capsaicin-
evoked release of CGRP from primary afferent fibers in spinal cord 
sections and upregulation of spinal dynorphin  to pathological 
levels (111, 118, 119). Manipulations that abolished descending 
facilitation, such as DLF lesions or dermorphin-saporin conjugate 
given into the RVM, also abolished dynorphin upregulation and 
enhanced release of CGRP (111, 118, 119). Recent studies revealed 
that increased concentrations of spinal dynorphin can stimulate 
neurons through increased calcium influx, unexpectedly mediated 
through the bradykinin receptors (120). Blockade of spinal brady-
kinin receptors inhibited behavioral signs of neuropathic pain, vis-
ceral pain, and diminished central sensitization (121–123). Collec-
tively, these studies suggest that descending facilitation represents 
an important mechanism that likely contributes to maintenance 
of central sensitization after peripheral nerve injury (78, 82).
Diffuse noxious inhibitory controls modulation of pain
The concept of diffuse noxious inhibitory control (DNIC) was for-
mulated from observations made with recordings of spinal dorsal 
horn units in anesthetized rats in response to peripheral stimuli 
applied to various parts of the body or electrical stimulation of 
peripheral nerves (124, 125). It was found that peripheral noxious 
stimuli suppressed the neuronal responses of convergent dorsal 
horn units to either electrical stimulation of C-fibers or applica-
tion of noxious heat (124, 125). This inhibitory effect could be 
evoked from noxious stimuli applied to various parts of the body 
and thus was diffuse in nature (124, 125). Importantly, DNIC was 
not demonstrated in dorsal horn units that responded solely to 
noxious, proprioceptive, or innocuous inputs, indicating a require-
ment for convergent neurons receiving both noxious and innocu-
ous stimuli (125). In addition, DNIC was abolished by spinal cord 
section and was diminished by systemic naloxone administration 
(124–126). Visceral pain induced by i.p. injection of phenylben-
zoquinone inhibited vocalization induced  by electrical stimuli 
applied to the tail, and this inhibition was also dose-dependently 
reversed by systemic naloxone (127). Observations that DNIC was 
diminished by electrolytic lesion (128) or lidocaine microinjection 
(129) of the nucleus raphe magnus suggested that there is a contri-
bution from this site to DNIC. However, other studies established 
that lesions of the RVM or the PAG did not block DNIC (130) and 
that DNIC was integrated at the level of the dorsal reticular nucle-
us (130). The dorsal reticular nucleus (DRt) receives nociceptive 
inputs from spinal projections and communicates with the PAG 
and RVM as well as the thalamus and amygdala and sends pain 
modulatory projections to the spinal cord (131–133). Moreover, 
the DRt sends and receives projections from cortical sites as well, 
and a single DRt neuron can project to different CNS sites, thus 
potentially modulating pain through several mechanisms (134). 
The DRt, along with the PAG and the RVM, form parts of a spinal-
supraspinal-spinal feedback loop that modulates pain (134, 135).
Loss of DNIC and chronic pain
These observations  suggest that many chronic pain syndromes 
may be due in part to a loss of DNIC (136). Loss of DNIC could 
manifest as enhanced through either the loss of endogenous inhibi-
tory control or an enhancement of pain facilitation (136). In one 
recent study, patients with irritable bowel syndrome (IBS) or tem-
poromandibular disorder (TMD) or healthy volunteers received an 
experimental pain stimulus in the form of increasing heat applied 
by a probe placed on the palm and a conditioning pain stimulus in 
the form of a foot-bath of noxious-cold water (137). The control 
group demonstrated decreased sensitivity to the noxious thermal 
stimulus when the foot was immersed in cold water, indicating 
active DNIC, whereas not only was DNIC absent in the patients 
with IBS or TMJ, but they showed enhanced sensitivity to the noci-
ceptive stimulus (137). The authors concluded from these data that 
chronic pain could be caused in part by a deficient pain inhibitory 
system (137). Deficits in DNIC have been demonstrated in patients 
with a number of chronic pain syndromes, including, for example, 
osteoarthritis of the knee (138), chronic pancreatitis (139), rheuma-
toid arthritis (140), and long-term trapezius myalgia (141).
Additionally, evidence is growing that a loss of DNIC suggests 
that deficits in endogenous pain modulation may underlie chron-
ic tension-type headache (CTT) as well. In  one study with CTT 
patients and unafflicted volunteers, a training stimulus of noxious 
thermal heat was applied to the thigh and an electrocutaneous 
review series
3784 The Journal of Clinical Investigation      http://www.jci.org      Volume 120      Number 11      November 2010
noxious stimulus was applied to either the forearm or the temple 
(142). The control group demonstrated decreased pain perception 
with the conditioning stimulus, whereas the CTT patients did not, 
indicating a deficit of DNIC (142). Similarly, a more recent study 
using pain from an occlusion cuff and temporal summation from 
repeated pulses from a pressure algometer demonstrated deficien-
cy in DNIC in CTT patients (143). In a recent study performed 
with rats, it was shown that persistent morphine exposure resulted 
in increased sensitivity to sensory thresholds and loss of DNIC in 
trigeminal neurons sensitive to dural stimulation (144). Here, it 
was hypothesized that loss of DNIC may contribute to develop-
ment of medication-overuse headache (144). The DNIC paradigm 
has been used as a clinical tool to predict who might be at risk for 
enhanced postsurgical pain (145, 146).
Descending modulation and pain-relieving drugs
The existence of a descending pain modulatory system provides 
many targets for the development of analgesic drugs or adjuncts 
that enhance the effects of existing analgesics (Figure 2). Opi-
oids act throughout the neuraxis and can relieve pain through 
activities at cortical and subcortical sites, at which affective and 
somatosensory aspects of the pain experience can be modified, as 
well as by activating descending pain inhibitory circuits. Activa-
tion of descending noradrenergic projections from the locus coe-
ruleus and other noradrenergic sites, described above, produces 
antinociception. Accordingly, α2-adrenergic  receptor agonists 
have been shown to produce antinociception as well as to poten-
tiate the antinociceptive effect of opioids (94, 147). Moreover, 
by increasing spinal noradrenergic activity, tricyclic antidepres-
sants and other selective noradrenergic reuptake inhibitors, such 
as duloxetine, enhance the analgesic effect of opioids and show 
clinical efficacy against neuropathic pain (148). It was recently 
also shown that the clinical efficacy of gabapentin may be due to 
its activation of descending noradrenergic systems and release of 
norepinephrine in the spinal cord (149). The COX inhibitors exert 
an analgesic effect by inhibition of PGE2 synthesis, thus reducing 
peripheral and central sensitization. Recent studies also indicate 
that inhibition of COX in the PAG promotes an opioid-mediated 
descending pain inhibition (150).
Summary
The advent of neuroimaging studies and technological advances 
allowing increased spatial and temporal resolution has contrib-
uted greatly to our changing perceptions of how pain is integrated 
and modulated in the central nervous system. From early animal 
studies that described a linear system of pain modulation from 
the PAG through the RVM and descending to the spinal cord, we 
now envision a complex pain matrix that includes important corti-
cal regions and elements of the limbic system as well as midbrain 
and medullary sites. These structures that likely participate in pain 
modulation reflect interacting brain regions that participate in 
pain processing as well as autonomic regulation and sensory and 
emotional management. The concept of top-down pain modula-
tion system accounts for or contributes to pain relief, as seen with 
Figure 2
Schematic representation of bulbospinal pain inhibition and potential
targets of analgesic activity. (A) Descending pain inhibition from the
PAG can be initiated by electrical stimulation or direct microinjection
of opioids. Recent evidence also indicates a role for COX inhibitors in
the PAG as well. Opioids and cannabinoids inhibit pain by enhancing
the baseline firing rate of off-cells and eliminating the off-cell pause in
response to nociceptive stimuli. Inhibition of on-cell activity may abolish
enhanced pain states. The on-cells and off-cells might correlate with
pain facilitatory (+) and inhibitory (–) neurons in the RVM, respectively.
At the level of the spinal cord, opioids can inhibit transmitter release
from primary afferent terminals as well as activity of pain transmission
neurons. Norepinephrine (NE) release from spinopetal noradrenergic
fibers from medullary sites also inhibits pain transmission. Tricyclic
antidepressants (TCAs) and other norepinephrine reuptake inhibitors
enhance the antinociceptive effect of opioids by increasing the avail-
ability of spinal norepinephrine (box). Areas labeled “i–iii” in the small
diagram correspond with labeled details of the larger diagram. α2A,
α2-adrenergic receptor; DRG, dorsal root ganglion; SNRI, serotonin/
norepinephrine reuptake inhibitor; SP, substance P. (B) Mice deficient
in dopamine β-hydroxylase that do not produce norepinephrine show
a diminished antinociceptive effect of morphine compared with control
animals, suggesting that the presence of norepinephrine, presum-
ably released in the spinal cord, is required for the full expression of
morphine antinociception. The dashed line represents the 50% effect,
and the corresponding dose is the ED50 (that is the dose producing a
50% effect). % MPE, percentage maximal possible effect. ***P < 0.001
compared with the control group. Error bars represent SEM. Copyright
National Academy of Sciences, USA (151).
review series
The Journal of Clinical Investigation      http://www.jci.org      Volume 120      Number 11      November 2010  3785
the placebo effect, stress, DNIC, and the actions of pain-relieving 
drugs, such as opioids, NSAIDs, reuptake blockers, and possibly 
gabapentinoids. These modulatory pathways help to explain how 
personal experience and emotional state as well as societal beliefs 
may alter the experience of pain. Clinical evidence supports the 
emerging view that dysfunctions of descending modulatory path-
ways, resulting in reduced inhibition/enhanced facilitation (e.g., 
loss of DNIC), can result in the enhanced pain observed in many 
chronic pain conditions. While not yet clinically proven, enhanced 
descending facilitation may also play an important role in main-
taining chronic pain. Increased knowledge of the components 
of these clinically validated pain modulatory circuits may offer 
approaches to develop improved pain therapy.
Acknowledgments
We thank Ian Meng, University of New England, for helpful com-
ments on the manuscript. This work was supported by NIH grants 
DA012656, DA023513, and NS066958.
Address correspondence to: Frank Porreca, Department of Pharma-
cology, University of Arizona, Tucson, Arizona 85724, USA. Phone: 
520.626.7421; Fax: 520.626.4182; E-mail: frankp@u.arizona.edu.
  1. Beecher HK. Pain in men wounded in battle. Ann
Surg. 1946;123(1):96–105.
2. Bingel  U,  Tracey I.  Imaging  CNS  modulation 
of pain  in  humans.  Physiology (Bethesda).  2008; 
23(3):371–380.
  3. Fields HL, Basbaum AI, Heinricher  MM. Central 
nervous system mechanisms of pain modulation. 
In: McMahon S,  Koltzenburg M, eds. Textbook of
Pain. 5th ed. Burlington, Massachusetts, USA: Else-
vier Health Sciences; 2005:125–142.
  4. Beecher HK. The powerful placebo. J Am Med Assoc. 
1955;159(17):1602–1606.
  5. Levine JD, Gordon NC, Fields HL. The mechanism 
of placebo analgesia. Lancet. 1978;2(8091):654–657.
  6. Petrovic P, Kalso E, Petersson KM, Ingvar M. Placebo 
and opioid analgesia-- imaging a shared neuronal 
network. Science. 2002;295(5560):1737–1740.
  7. Zubieta JK, et al. Placebo effects mediated by endog-
enous opioid  activity  on  mu-opioid  receptors. 
J Neurosci. 2005;25(34):7754–7762.
  8. Tracey I, Mantyh PW. The cerebral signature for 
pain perception and its modulation. Neuron. 2007; 
55(3):377–391.
9. Eippert  F,  et  al.  Activation  of  the  opioidergic 
descending pain control system underlies placebo 
analgesia. Neuron. 2009;63(4):533–543.
10. Tracey I,  Johns E.  The  pain matrix:  reloaded or 
reborn as we image tonic pain using arterial spin 
labelling. Pain. 2010;148(3):359–360.
  11. Benedetti F, Lanotte M, Lopiano L, Colloca L. When 
words are painful: unraveling the mechanisms of the 
nocebo effect. NeuroScience. 2007;147(2):260–271.
  12. Bened etti  F,   Pollo   A,  Lo piano   L,  Lan otte  M , 
Vighetti S, Rainero I. Conscious expectation and 
unconscious conditioning in analgesic, motor, and 
hormonal placebo/nocebo  responses. J Neurosci. 
2003;23(10):4315–4323.
  13. Colloca L, Benedetti F. Nocebo hyperalgesia: how anx-
iety is turned into pain. Curr Opin Anaesthesiol. 2007; 
20(5):435–439.
14. Keltner JR, Furst A, Fan  C, Redfern R,  Inglis B, 
Fields  HL.  Isolating  the  modulatory  effect  of 
expectation on  pain  transmission: a functional 
magnetic resonance imaging study. J Neurosci. 2006; 
26(16):4437–4443.
  15. Tsou K, Jang CS. Studies on the  site of analgesic 
action of morphine by intracerebral micro-injection. 
Sci Sin. 1964;13:1099–1109.
  16. Reynolds DV. Surgery in the rat during electrical 
analgesia induced by focal brain stimulation. Science. 
1969;164(878):444–445.
  17. Hosobuchi Y, Adams JE, Linchitz R. Pain relief by 
electrical stimulation of the central gray matter in 
humans and its reversal by naloxone. Science. 1977; 
197(4299):183–186.
  18. Richardson DE, Akil H. Pain reduction by electrical 
brain stimulation in man. Part 1: Acute adminis-
tration in periaqueductal and periventricular sites. 
J Neurosurg. 1977;47(2):178–183.
  19. Richardson DE, Akil H. Long term results of peri-
ventricular gray self-stimulation. Neurosurgery. 1977; 
1(2):199–202.
 20. Raskin NH,  Hosobuchi  Y,  Lamb  S.  Headache 
may arise from perturbation  of brain.  Headache. 
1987;27(8):416–420.
  21. Richardson DE. Intracranial stimulation for  the 
control  of  chronic  pain.  Clin Neurosurg.  1983; 
31:316–322.
  22. Yeung JC,  Yaksh TL,  Rudy  TA. Co ncurrent  map-
ping of brain sites for sensitivity to the direct appli-
cation of morphine and focal electrical stimulation 
in the production of antinociception in the rat. 
Pain. 1977;4(1):23–40.
  23. Yaksh TL, Rudy TA. Narcotic analgestics: CNS sites 
and mechanisms of action as revealed by intracere-
bral injection techniques. Pain. 1978;4(4):299–359.
  24. Lewis VA, Gebhart GF. Evaluation of the periaque-
ductal central gray (PAG) as a morphine-specific locus 
of action and examination of morphine-induced and 
stimulation-produced analgesia at coincident PAG 
loci. Brain Res. 1977;124(2):283–303.
25. Lewis VA,  Gebhart  GF.  Morphine-induced  and 
stimulation-produced analgesias  at  coincident 
periaqueductal central gray loci: evaluation of anal-
gesic congruence, tolerance, and cross-tolerance. 
Exp Neurol. 1977;57(3):934–955.
  26. Basbaum AI, Clanton CH, Fields HL.  Opiate and 
stimulus-produced analgesia: functional anatomy 
of a medullospinal pathway. Proc Natl Acad Sci U S A. 
1976;73(12):4685–4688.
  27. Akil H, Mayer DJ,  Liebeskind JC. Antagonism of 
stimulation-produced analgesia by naloxone, a nar-
cotic antagonist. Science. 1976;191(4230):961–962.
28. Basbaum AI, Clanton CH, Fields HL. Three  bul-
bospinal pathways from the rostral medulla of the 
cat: an autoradiographic study of pain modulating 
systems. J Comp Neurol. 1978;178(2):209–224.
  29. Basbaum AI, Fields HL. Endogenous pain control 
mechanisms:review and hypothesis. Ann Neurol. 1978; 
4(5):451–462.
  30. Basbaum AI, Fields HL. Endogenous pain control 
systems: brainstem spinal pathways and endorphin 
circuitry. Annu Rev Neurosci. 1984;7:309–338.
  31. Bingel U, Schoell E, Buchel C. Imaging pain modu-
lation in health and disease. Curr Opin Neurol. 2007; 
20(4):424–431.
  32. Neugebauer V, Galhardo V, Maione S, Mackey SC. 
Forebrain pain mechanisms. Brain Res Rev. 2009; 
60(1):226–242.
33. Neugebauer V,  Li  W. Processing  of  nociceptive 
mechanical and  thermal information  in central 
amygdala neurons with knee-joint input. J Ne uro-
physiol. 2002;87(1):103–112.
  34. Neugebauer V, Li W. Differential sensitization of 
amygdala neurons to afferent inputs in a model of 
arthritic pain. J Neurophysiol. 2003;89(2):716–727.
35. Han JS, Neugebauer V. Synaptic plasticity in the 
amygdala in a visceral pain model in rats. Neurosci
Lett. 2004;361(1–3):254–257.
  36. Han JS, Bird GC, Neugebauer V. Enhanced group 
III mGluR-mediated inhibition of pain-related syn-
aptic plasticity in the amygdala. Neuropharmacology. 
2004;46(7):918–926.
  37. Li W, Neugebauer V. Differential roles of mGluR1 
and mGluR5  in brief and prolonged nociceptive 
processing in central amygdala neurons. J Neuro-
physiol. 2004;91(1):13–24.
  38. Ji G, Fu Y, Ruppert KA, Neugebauer V. Pain-related 
anxiety-like behavior requires CRF1 receptors in 
the amygdala. Mol Pain. 2007;3:13.
  39. Ji G, Neugebauer V. Hemispheric lateralization of 
pain processing by amygdala neurons. J Neurophysiol. 
2009;102(4):2253–2264.
  40. Carrasquillo Y, Gereau  RWt. Hemispheric l ateral-
ization of a molecular signal for pain modulation 
in the amygdala. Mol Pain. 2008;4:24.
  41. Fields HL, Anderson SD, Clanton  CH, Basbaum 
AI. Nucleus raphe magnus: a common  mediator 
of opiate- and stimulus- produced analgesia. Trans
Am Neurol Assoc. 1976;101:208–210.
  42. Sandkuhler J, Gebhart GF. Relative contributions 
of the nucleus raphe magnus and adjacent medul-
lary reticular formation to the inhibition by stimu-
lation in the periaqueductal gray of a spinal noci-
ceptive reflex in the pentobarbital- anesthetized rat. 
Brain Res. 1984;305(1):77–87.
  43. Abols IA, Basbaum AI. Afferent connections of the 
rostral medulla of the cat: a  neural substrate for 
midbrain-medullary interactions in  the modula-
tion of pain. J Comp Neurol. 1981;201(2):285–297.
44. Glazer EJ,  Basbaum  AI.  Immunohistochemical 
localization of  leucine-enkephalin in  the spinal 
cord of the  cat: enkephalin-containing marginal 
neurons  and  pain  modulation.  J Comp Neur ol. 
1981;196(3):377–389.
  45. Basbaum AI. Descending control of pain transmis-
sion: possible serotonergic-enkephalinergic inter-
actions. Adv Exp Med Biol. 1981;133:177–189.
 46. Anderson SD, Basbaum  AI, Fields HL. Response 
of  medullary  raphe  neurons  to  peripheral 
stimulation and  to  systemic opiates. Brain Res. 
1977;123(2):363–368.
  47. Fields HL, Basbaum AI,  Clanton CH, Anderson 
SD. Nucleus  raphe  magnus  inhibition  of  spi-
nal  cord  dorsa l  horn  neuron s.  Brain Res.  1977; 
126(3):441–453.
  48. Heinricher MM, Barbaro NM, Fields HL. Putative 
nociceptive modulating  neurons  in  the  rostral 
ventromedial medulla of the rat: firing of on- and 
off-cells is related to nociceptive responsiveness. 
Somatosens Mot Res. 1989;6(4):427–439.
  49. Potrebic SB, Mason P, Fields HL. The density and 
distribution  of  serotonergic  appositions  onto 
identified neurons in the rat rostral ventromedial 
medulla. J Neurosci. 1995;15(5 pt 1):3273–3283.
  50. Fields HL, Bry J, Hentall  I, Zorman G. The  activ-
ity of  neurons in  the  rostral medulla  of  the rat 
during withdrawal from noxious heat. J Neurosci. 
1983;3(12):2545–2552.
51. Fields HL,  Malick  A ,  Burstein  R.  Dorsal  horn 
proje ctio n  targ ets  of   ON  an d  OFF  c ells  i n  the 
rostral  ventromedial  medulla.  J Neurophysiol. 
1995;74(4):1742–1759.
 52. Vanegas H,  Barbaro  NM,  Fields  HL.  Tail-Flick 
Related Activity in Medullospinal Neurons. B rain
Research. 1984;321(1):135–141.
53. Urban MO,  Gebhart  GF. Supraspinal  contribu-
tions to hyperalgesia. Proc Natl Acad Sci U S A. 1999; 
96(14):7687–7692.
review series
3786 The Journal of Clinical Investigation      http://www.jci.org      Volume 120      Number 11      November 2010
  54. Zhuo M, Gebhart GF. Characterization of descend-
ing facilitation and inhibition of spinal nociceptive 
transmission from the nuclei reticularis gigantocel-
lularis and gigantocellularis pars alpha in the rat. 
J Neurophysiol. 1992;67(6):1599–1614.
55. Zhuo M,  Gebhart  GF.  Biphasic  modulation  of 
spinal nociceptive transmission from the medul-
lary raphe  nuclei in  the rat.  J Neurophysiol.  1997; 
78(2):746–758.
  56. Zhuo M, Gebhart GF. Facilitation and attenuation 
of a  visceral  nociceptive ref lex from the  rostro-
ventral medulla in the  rat. Gastroenterology. 2002; 
122(4):1007–1019.
  57. Zhuo M, Gebhart GF. Characterization of descend-
ing inhibition  and  facilitation  from  the  nuclei 
reticularis gigantocellularis and gigantocellularis 
pars alpha in the rat. Pain. 1990;42(3):337–350.
58. Morgan MM, Fields HL. Pronounced  changes in 
the activity of nociceptive modulatory neurons in 
the rostral ventromedial medulla in  response to 
prolonged thermal noxious stimuli. J Neurophysiol. 
1994;72(3):1161–1170.
  59. Kaplan H, Fields HL. Hyperalgesia during acute opi-
oid abstinence: evidence for a nociceptive facilitat-
ing function of the rostral ventromedial medulla. 
J Neurosci. 1991;11(5):1433–1439.
60. Vanegas H, Schaible HG. Descending control of 
persistent pain: inhibitory or facilitatory? Brain Res
Brain Res Rev. 2004;46(3):295–309.
61. Fang FG,  Haws  CM,  Drasner K, Williamson  A, 
Fields HL.  Opioid peptides  (DAGO-enkephalin, 
dynorphin A(1-13), BAM 22P) microinjected into 
the rat brainstem:  comparison of their antinoci-
ceptive  effect and their effect on neuronal  firing 
in the  rostral  ventromedial  medulla.  Brain Res. 
1989;501(1):116–128.
  62. Jensen TS, Yaksh TL. Comparison of the antinoci-
ceptive effect of morphine and glutamate at coinci-
dental sites in the periaqueductal gray and medial 
medulla in rats. Brain Res. 1989;476(1):1–9.
  63. Cheng ZF,  Fields HL , Heinrich er MM.  Morphine 
microinjected into the periaqueductal gray has dif-
ferential effects on 3 classes of medullary neurons. 
Brain Res. 1986;375(1):57–65.
  64. Heinriche r MM,  Morga n MM,  Fiel ds HL.  Dire ct 
and indirect  actions of  morphine  on medullary 
neurons that modulate nociception. NeuroScience. 
1992;48(3):533–543.
  65. Heinricher MM, Morgan MM, Tortorici V, Fields 
HL. Disinhibition  of  off-cells  and  antinocicep-
tion produced  by  an  opioid  action  within  the 
rostral ventromedial medulla. NeuroScience. 1994; 
63(1):279–288.
  66. Zhang W, et al. Neuropathic pain is maintained by 
brainstem neurons co-expressing opioid and chole-
cystokinin receptors. Brain. 2009;132(pt 3):778–787.
  67. Winkler CW, Hermes S M, Chavkin CI, Dr ake CT, 
Morrison SF,  Aicher SA. Kappa  opioid receptor 
(KOR) and GAD67 immunoreactivity are found in 
OFF and NEUTRAL cells in the rostral ventromedi-
al medulla. J Neurophysiol. 2006;96(6):3465–3473.
  68. Oliveras JL, Hosobuchi Y, Guilbaud G, Besson JM. 
Analgesic electrical stimulation of the feline nucle-
us raphe magnus: development of tolerance and its 
reversal by 5-HTP. Brain Res. 1978;146(2):404–409.
  69. Cui M, Feng Y, McAdoo DJ, Willis WD. Periaqueduc-
tal gray stimulation-induced inhibition of nocicep-
tive dorsal horn neurons  in rats is associated with 
the release of norepinephrine, serotonin, and amino 
acids. J Pharmacol Exp Ther. 1999;289(2):868–876.
70. Yaksh TL, Wilson PR. Spinal  serotonin terminal 
system mediates antinociception. J Pharmacol Exp
Ther. 1979;208(3):446–453.
  71. Jensen TS, Yaksh TL. Spinal monoamine and opi-
ate systems  partly mediate the  antinociceptive 
effects produced by glutamate at brainstem sites. 
Brain Res. 1984;321(2):287–297.
72. Kwiat GC, Basbaum AI. The  origin of brainstem 
noradrenergic and serotonergic projections to the 
spinal cord dorsal horn in the  rat. Somatosens Mot
Res. 1992;9(2):157–173.
  73. Potrebic SB, Fields HL, Mason P. Serotonin immu-
noreactivity is contained in one physiological cell 
class in the rat rostral ventromedial medulla. J Neu-
rosci. 1994;14(3 pt 2):1655–1665.
  74. Moore RY. The anatomy of central serotonin neuron 
systems in the rat brain. In: Jacobs BL, Gelperin A, 
eds. Serotonin Neurotransmission And Behavior. Cam-
bridge, Massachusetts, USA: MIT Press; 1981:35–71.
75. Foo H, Mason P. Brainstem  modulation of pain 
during sleep and  waking.  Sleep Med Rev.  2003; 
7(2):145–154.
76. Mason P. Contributions of the medullary  raphe 
and ventromedial reticular region to  pain modu-
lation and other homeostatic functions. Annu Rev
Neurosci. 2001;24:737–777.
  77. Wei F, et  al. Molecular  depletion of descending 
serotonin unmasks its novel facilitatory role in the 
development of persistent  pain. J Neurosci. 2010; 
30(25):8624–8636.
  78. Suzuki R, Rygh LJ, Dickenson AH. Bad news from 
the brain: descending 5-HT pathways that control 
spinal pain processing. Trends Pharmacol Sci. 2004; 
25(12):613–617.
79. Dogrul A, Ossipov  MH,  Porreca F. Differential 
mediation of descending pain facilitation and inhi-
bition by spinal 5HT-3 and 5HT-7 receptors. Brain
Res. 2009;1280:52–59.
  80. Sasaki M,  Obata H, K awahara K,  Saito S, G oto F. 
Peripheral 5-HT2A receptor  antagonism attenu-
ates primary thermal hyperalgesia and secondary 
mechanical allodynia after thermal injury in rats. 
Pain. 2006;122(1–2):130–136.
  81. Green GM, Scarth J, Dickenson  A. An excitator y 
role for 5-HT in spinal inf lammatory nociceptive 
transmission; state-dependent actions via dorsal 
horn 5-HT(3) receptors  in the  anaesthetized rat. 
Pain. 2000;89(1):81–88.
  82. Rahman W, Bauer CS, Bannister K, Vonsy JL, Dol-
phin AC, Dickenson AH. Descending serotonergic 
facilitation and the antinociceptive effects of pre-
gabalin in a rat model  of osteoarthritic pain. M ol
Pain. 2009;5:45.
  83. Brenchat A, et al. 5-HT7 receptor activation inhibits 
mechanical hypersensitivity secondary to capsaicin 
sensitization in mice. Pain. 2009;141(3):239–247.
84. Doly S, Fischer J,  Brisorgueil MJ,  Verge D,  Con-
rath  M.  Pre-  and  postsynaptic  localization  of 
the  5-HT7  receptor  in  rat  dorsal  spinal  cord: 
immunocytochemical  evidence.  J Comp Neurol. 
2005;490(3):256–269.
  85. Pierce PA, Xie GX, Levine JD, Peroutka SJ. 5-Hydroxy-
tryptamine receptor subtype  messenger RNAs in 
rat peripheral  sensory and  sympathetic ganglia: 
a polymerase chain reaction study.  NeuroScience.  
1996;70(2):553–559.
  86. Hammond  DL, Tyc e GM,  Yaksh T L. Ef flux  of 5-
hydroxytryptamine and noradrenaline into spinal 
cord superfusates  during stimulation of  the rat 
medulla. J Physiol. 1985;359:151–162.
  87. Barbaro NM, Hammond DL, Fields HL. Effects of 
intrathecally administered methysergide and yohim-
bine on microstimulation-produced antinocicep-
tion in the rat. Brain Res. 1985;343(2):223–229.
  88. Hammond DL, Yaksh TL. Antagonism of  stimu-
lation-produced antinociception  by  intrathecal 
administration of methysergide or phentolamine. 
Brain Res. 1984;298(2):329–337.
  89. Bajic D, Proudfit HK. Projections of neurons in the 
periaqueductal gray to pontine and medullary cate-
cholamine cell groups involved in the modulation of 
nociception. J Comp Neurol. 1999;405(3):359–379.
  90. Holden JE, Proudfit HK. Enkephalin neurons that 
project to the A7 catecholamine cell group are locat-
ed in nuclei that modulate nociception: ventrome-
dial medulla. NeuroScience. 1998;83(3):929–947.
  91. Yeomans DC,  Proudfit  HK. Projections  of  sub-
stance P-immunoreactive neurons  located in the 
ventromedial medulla  to  the  A7  noradrenergic 
nucleus of the rat demonstrated using retrograde 
tracing combined with immunocytochemistry. 
Brain Research. 1990;532(1–2):329–332.
  92. Proudfit H. The behavioural pharmacology of  the 
noradrenergic system. In: Guilbaud G, ed. Towards The
Use Of Noradrenergic Agonists For The Treatment Of Pain. 
Amsterdam, Nederland: Elsevier; 1992:119–136.
  93. Pertovaara A. Noradrenergic pain modulation. Prog
Neurobiol. 2006;80(2):53–83.
  94. Ossipov MH, Harris S, Lloyd P, Messineo E. An iso-
bolographic analysis of the antinociceptive effect 
of systemically  and  intrathecally  administered 
combinations of clonidine and opiates. J Pharmacol
Exp Ther. 1990;255(3):1107–1116.
  95. Ossipov MH, Harris S, Lloyd P, Messineo E, Lin BS, 
Bagley J. Antinociceptive interaction between opi-
oids and medetomidine: systemic additivity and spi-
nal synergy. Anesthesiology. 1990;73(6):1227–1235.
  96. Eisenach JC, Hood DD, Curry R. Intrathecal, but not 
intravenous, clonidine reduces experimental thermal 
or capsaicin-induced pain and hyperalgesia in nor-
mal volunteers. Anesth Analg. 1998;87(3):591–596.
  97. Budai D, Harasawa I, Fields HL. Midbrain periaq-
ueductal gray (PAG) inhibits nociceptive inputs to 
sacral dorsal horn nociceptive neurons through 
alpha2-adrenergic  receptors.  J Neurophysiol. 
1998;80(5):2244–2254.
98. Gassner M, Ruscheweyh R,  Sandkuhler J.  Direct 
excitation of  spinal GABAergic  interneurons by 
noradrenaline. Pain. 2009;145(1–2):204–210.
99. Watkins LR,  Mayer DJ.  Organization of endog-
enous opiate and nonopiate pain control systems. 
Science. 1982;216(4551):1185–1192.
100. Friederi ch  MW,  Wal ker  JM .  The   effec t  of  fo ot-
shock  on  the  noxious-evoked activity  of  neu-
rons in the rostral ventral  medulla. Brain Res Bull. 
1990;24(4):605–608.
101. Nakagawasai O, et al. Changes in beta-endorphin 
and stress-induced analgesia in mice after exposure 
to forced walking stress. Methods Find Exp Clin Phar-
macol. 1999;21(7):471–476.
102. Wiedenmayer CP, Barr GA. Mu opioid receptors 
in the ventrolateral  periaqueductal gray mediate 
stress-induced analgesia but not immobility in rat 
pups. Behav Neurosci. 2000;114(1):125–136.
103. Foo  H,  Helmstetter  FJ.  Hypoalgesia elicited by 
a conditioned  stimulus is  blocked by  a mu,  but 
not a delta or a kappa, opioid antagonist injected 
into the rostral ventromedial medulla. Pain. 1999; 
83(3):427–431.
104. Bellgowan PS, Helmstetter FJ. The role of mu and 
kappa opioid receptors within the periaqueductal 
gray in the expression of conditional hypoalgesia. 
Brain Res. 1998;791(1–2):83–89.
105. Helmstetter FJ, Tershner SA, Poore LH, Bellgowan 
PS. Antinociception following opioid stimulation 
of the basolateral amygdala is expressed through 
the periaqueductal gray and rostral ventromedial 
medulla. Brain Res. 1998;779(1–2):104–118.
106. Hopkins E, Spinella M,  Pavlovic ZW, Bodnar RJ. 
Alterations in swim stress-induced analgesia and 
hypothermia following  serotonergic  or  NMDA 
antagonists in the rostral ventromedial medulla of 
rats. Physiol Behav. 1998;64(3):219–225.
107. Meng ID, Manning BH, Martin WJ, Fields HL. An 
analgesia circuit activated by cannabinoids. Nature. 
1998;395(6700):381–383.
108. Hohmann  AG,  et  al.  An  endocannabinoid 
mechanism for stress-induced  analgesia. Nature. 
2005;435(7045):1108–1112.
109. Suplita RL 2nd, Farthing JN, Gutierrez T, Hohm-
ann AG. Inhibition of fatty-acid amide hydrolase 
enhances cannabinoid  stress-induced analgesia: 
sites of action in the dorsolateral periaqueductal 
gray and rostral ventromedial medulla. Neurophar-
review series
The Journal of Clinical Investigation      http://www.jci.org      Volume 120      Number 11      November 2010  3787
macology. 2005;49(8):1201–1209.
110. Pertovaara A, Wei H, Hamalainen MM. Lidocaine 
in the rostroventromedial medulla and the periaq-
ueductal gray attenuates allodynia in neuropathic 
rats. Neuroscience Letters. 1996;218(2):127–130.
111. Burgess  SE,  et  al.  Time-dependent  descending 
facilitation from the rostral ventromedial medulla 
maintains, but does not initiate, neuropathic pain. 
J Neurosci. 2002;22(12):5129–5136.
112. Kovelowski CJ, Ossipov MH, Sun H, Lai J, Malan 
TP, Porreca  F.  Supraspina l cholecy stokinin  may 
drive tonic  descending facilitation  mechanisms 
to maintain neuropathic  pain  in  the  rat.  Pain. 
2000;87(3):265–273.
113. Ossipov MH, Hong Sun T, Malan P Jr, Lai J, Porreca 
F. Mediation of spinal nerve injury induced tactile 
allodynia by  descending  facilitatory  pathways in 
the dorsolateral  funiculus  in rats. Neurosci Lett. 
2000;290(2):129–132.
114. Xie JY, et al. Cholecystokinin in the rostral ventro-
medial medulla mediates opioid-induced hyper-
algesia and  antinociceptive tolerance.  J Neurosci. 
2005;25(2):409–416.
115. Heinricher  MM,  Neubert  MJ.  Neural  basis  for 
the  hyperalgesic  action  of  cholecystokinin  in 
the rostral ventromedial medulla.  J Ne urophysiol. 
2004;92(4):1982–1989.
116. Porreca F, et al.  Inhibition  of neuropathic pain 
by selective ablation of brainstem medullary cells 
expressing  the  mu-opioid  receptor.  J Neurosci. 
2001;21(14):5281–5288.
1 17. Kin g  T,  et   al.  U nmas king   the   toni c-av ersi ve 
state in  neuropathic  pain.  Nat Neurosci.  2009; 
12(11):1364–1366.
118. Gardell LR, et al. Mouse strains that lack spinal dyn-
orphin upregulation after peripheral nerve injury 
do not  develop  neuropathic  pain. NeuroScience.  
2004;123(1):43–52.
119. Gardell   LR,  et   al.  En hance d  evok ed  exc itat ory 
transmitter  release  in  experimental  neuropa-
thy requires  descending  facilitation.  J Neurosci. 
2003;23(23):8370–8379.
120. Lai J, et al. Dynorphin A activates bradykinin recep-
tors to maintain neuropathic pain. Nat Neurosci. 
2006;9(12):1534–1540.
121. Wang H, et  al. Bradykinin produces pain hyper-
sensitivity  by  potentiating  spina l cord  glutama-
tergic synaptic  transmission.  J Neurosci.  2005; 
25(35):7986–7992.
122.Chen  Q,  Vera-Portocarrero  LP,  Ossipov  MH, 
Vardanyan M,  Lai  J,  Porreca  F.  Attenuation of 
persistent experimental pancreatitis pain by a bra-
dykinin b2 receptor antagonist [published online 
ahead of print June 5, 2010]. Pancreas. doi:10.1097/
MPA.0b013e3181df1c90.
123. Lai J, Luo MC, Chen Q, Porreca F. Pronociceptive 
actions of dynorphin via bradykinin receptors. Neu-
rosci Lett. 2008;437(3):175–179.
124. Le B ars  D,   Dicke nson   AH,  Be sson  J M.  Dif fuse 
noxious inhibitory  controls (DNIC). II.  Lack  of 
effect on  non-convergent neurones,  supraspinal 
involvement and  theoretical implications. Pain . 
1979;6(3):305–327.
125. Le B ars  D,   Dicke nson   AH,  Be sson  J M.  Dif fuse 
noxious inhibitory controls (DNIC). I.  Effects on 
dorsal horn convergent neurones in  the rat. Pain. 
1979;6(3):283–304.
126. Le Bars D, Chitour D, Kraus E, Dickenson AH, Bes-
son JM. Effect of  naloxone upon diffuse noxious 
inhibitory controls  (DNIC) in  the rat. Brain Res. 
1981;204(2):387–402.
127. Kraus E, Le Bars D, Besson JM. Behavioral confirma-
tion of “diffuse noxious inhibitory controls” (DNIC) 
and evidence  for  a  role of  endogenous  opiates. 
Brain Res. 1981;206(2):495–499.
128. Le  Bars D, et  al. [Are  bulbo-spinal serotonergic 
systems involved in the  detection of nociceptive 
messages? (author’s transl)]. J Physiol (Paris). 1981;
77(2–3):463–471.
129. Morton CR, Maisch B, Zimmermann M. Diffuse 
noxious inhibitory controls of lumbar spinal neu-
rons involve a supraspinal loop in the cat. Brain Res. 
1987;410(2):347–352.
130. Villanueva L, Le Bars D. The activation of bulbo-
spinal controls by  peripheral nociceptive inputs: 
diffuse noxious inhibitory controls. Biol Res. 1995; 
28(1):113–125.
131. Leite-Almeida H,  Valle-Fernandes A,  Almeida  A. 
Brain projections from the medullary dorsal reticu-
lar nucleus: an anterograde and retrograde tracing 
study in the rat. NeuroScience. 2006;140(2):577–595.
132. Lima D, Almeida A. The medullary dorsal reticular 
nucleus as a pronociceptive centre of the pain con-
trol system. Prog Neurobiol. 2002;66(2):81–108.
133. Almeida A, Tavares I, Lima D, Coimbra A. Descend-
ing projections from the medullary dorsal reticular 
nucleus make synaptic contacts with spinal cord 
lamina I cells projecting to that nucleus: an electron 
microscopic tracer study  in the rat.  NeuroScience.  
1993;55(4):1093–1106.
134. Monconduit  L,  Desbois  C,  Villanueva  L.  The 
integrative role of  the rat  medullary subnucleus 
reticularis dorsalis in nociception. Eur J Neurosci. 
2002;16(5):937–944.
135. Le Bars D. The whole body receptive field of dorsal 
horn multireceptive neurones. Brain Res Brain Res
Rev. 2002;40(1–3):29–44.
136. Villanueva  L.  Diffuse  Noxious  Inhibitory  Con-
trol (DNIC)  as a  tool for exploring  dysfunction 
of endogenous  pain  modulatory  systems.  Pain. 
2009;143(3):161–162.
137. King CD, Wong F, Currie T, Mauderli AP, Fillingim 
RB, Riley JL 3rd. Deficiency in endogenous modu-
lation of prolonged heat pain in patients with Irri-
table Bowel Syndrome and Temporomandibular 
Disorder. Pain. 2009;143(3):172–178.
138. Arendt-Nielsen L, et al.  Sensitization in patients 
with  painful  knee  osteoarthritis.  Pain.  2010; 
149(3):573–581.
139. Olesen SS, et al. Descending inhibitory pain modula-
tion is impaired in patients with chronic pancreatitis. 
Clin Gastroenterol Hepatol. 2010;8(8):724–730.
140. Leffler AS, Kosek E, Lerndal T, Nordmark B, Hans-
son P.  Somatosensory perception and  function 
of diffuse noxious inhibitory controls (DNIC) in 
patients suffering from rheumatoid arthritis. Eur J
Pain. 2002;6(2):161–176.
141. Leffler AS, Hansson P, Kosek  E. Somatosensory 
perception in a remote pain-free area and function 
of diffuse noxious inhibitory controls (DNIC) in 
patients suffering from long-term trapezius myalgia. 
Eur J Pain. 2002;6(2):149–159.
142. Pielsticker A, Haag G, Zaudig M, Lautenbacher S. 
Impairment of pain inhibition in chronic tension-
type headache. Pain. 2005;118(1–2):215–223.
143. Cathcart S, Winefield AH, Lushington K, Rolan P. 
Noxious inhibition  of  temporal  summation  is 
impaired in chronic tension-type headache. Headache. 
2009;50(3):403–412.
144. Okada-Ogawa  A,  Porreca F,  Meng  ID.  Sustained 
morphine-induced sensitization and loss  of dif-
fuse noxious  inhibitory  controls  in  dura-sensi-
tive medullary  dorsal  horn  neurons.  J Neurosci. 
2009;29(50):15828–15835.
145. Landau R, et al. An experimental paradigm for the 
prediction of Post-Operative Pain (PPOP). J Vis Exp. 
2010;(35). pii: 1671. doi:10.3791/1671.
146. Yarnitsky D, et al. Prediction of chronic post-opera-
tive pain:  pre-operative DNIC  testing  identifies 
patients at risk. Pain. 2008;138(1):22–28.
147. Fairbanks CA, Kitto  KF, Nguyen  HO, Stone  LS, 
Wilcox GL. Clonidine and dexmedetomidine pro-
duce antinociceptive synergy in mouse spinal cord. 
Anesthesiology. 2009;110(3):638–647.
148. Maizels  M,  McCarberg  B.  Antidepressants  and 
antiepileptic drugs for  chronic non-cancer  pain. 
Am Fam Physician. 2005;71(3):483–490.
149. Hayashida  K,  DeGoes  S,  Curry  R,  Eisenach JC. 
Gabapentin activates spinal noradrenergic activity 
in rats and humans and reduces  hypersensitivity 
after surgery. Anesthesiology. 2007;106(3):557–562.
150. Conroy JL, et al. Opioids  activate brain analgesic 
circuits through  cytochrome P450/epoxygenase 
signaling. Nat Neurosci. 2010;13(3):284–286.
151. Jasmin L, et al. The NK1 receptor  mediates both 
the hyperalgesia and the resistance to morphine in 
mice lacking noradrenaline. Proc Natl Acad Sci U S A. 
2002;99(2):1029–1034.
... It profoundly impacts individuals' quality of life and places a considerable social and economic burden on societies worldwide [2][3][4][5][6]. It is categorized based on its duration, with acute pain having a short-term nature and often being linked to known causes, such as exposure to heat, extreme cold, chemical irritants [7], and chronic pain which persists or recurs for a minimum of three months [2,[8][9][10], affecting approximately 20 % of the global population [2,11]. Chronic pain's development and progression are influenced by various factors, and among these, cancer pain continues to represent a significant portion [12,13]. ...
... Ziconotide has a molecular formula of C 102 H 172 N 36 O 32 S 7 , a relative molecular weight of 2639.12, a CAS chemical identifier number of 107452-89-1, and an isoelectric point of 11.2. The molecule consists of 25 highly hydrophilic amino acids arranged in the following sequence: H-Cys-Lys-Gly-Lys-Gly-Ala-Lys-Cys-Ser-Arg-Leu-Met-Tyr-Asp-Cys-Cys-Thr-Gly-Ser-Cys-Arg-Ser-Gly-Lys-Cys-NH2 cyclic (1)(2)(3)(4)(5)(6)(7)(8)(9)(10)(11)(12)(13)(14)(15)(16), (8)(9)(10)(11)(12)(13)(14)(15)(16)(17)(18)(19)(20), (15)(16)(17)(18)(19)(20)(21)(22)(23)(24)(25)-tris(disulfide) (Fig. 2) [31][32][33]. These amino play a crucial role in stability of molecule [34] and includes essential amino acids, including four Lysine (Lys) and two Arginine (Arg) residues. ...
... Ziconotide has a molecular formula of C 102 H 172 N 36 O 32 S 7 , a relative molecular weight of 2639.12, a CAS chemical identifier number of 107452-89-1, and an isoelectric point of 11.2. The molecule consists of 25 highly hydrophilic amino acids arranged in the following sequence: H-Cys-Lys-Gly-Lys-Gly-Ala-Lys-Cys-Ser-Arg-Leu-Met-Tyr-Asp-Cys-Cys-Thr-Gly-Ser-Cys-Arg-Ser-Gly-Lys-Cys-NH2 cyclic (1)(2)(3)(4)(5)(6)(7)(8)(9)(10)(11)(12)(13)(14)(15)(16), (8)(9)(10)(11)(12)(13)(14)(15)(16)(17)(18)(19)(20), (15)(16)(17)(18)(19)(20)(21)(22)(23)(24)(25)-tris(disulfide) (Fig. 2) [31][32][33]. These amino play a crucial role in stability of molecule [34] and includes essential amino acids, including four Lysine (Lys) and two Arginine (Arg) residues. ...
Article
Full-text available
Managing severe chronic pain is a challenging task, given the limited effectiveness of available pharmacological and non-pharmacological treatments. This issue continues to be a significant public health concern, requiring a substantial therapeutic response. Ziconotide, a synthetic peptide initially isolated from Conus magus in 1982 and approved by the US Food and Drug Administration and the European Medicines Agency in 2004, is the first-line intrathecal method for individuals experiencing severe chronic pain refractory to other therapeutic measures. Ziconotide produces powerful analgesia by blocking N-type calcium channels in the spinal cord, which inhibits the release of pain-relevant neurotransmitters from the central terminals of primary afferent neurons. However, despite possessing many favorable qualities, including the absence of tolerance development, respiratory depression, and withdrawal symptoms (largely due to the absence of a G-protein mediation mechanism), ziconotide's application is limited due to factors such as intrathecal administration and a narrow therapeutic window resulting from significant dose-related undesired effects of the central nervous system. This review aims to provide a comprehensive and clinically relevant summary of the literatures concerning the pharmacokinetics and metabolism of intrathecal ziconotide. It will also describe strategies intended to enhance clinical efficacy while reducing the incidence of side effects. Additionally, the review will explore the current efforts to refine the structure of ziconotide for better clinical outcomes. Lastly, it will prospect potential developments in the new class of selective N-type voltage-sensitive calcium-channel blockers.
... It also seems that beta oscillations are important for motor function, especially during active movements (during desynchronization and synchronization). One of the main brain inhibitory neurotransmitters associated with pain processing is gamma-aminobutyric acid (GABA) [39]. Beta oscillations are known to be related to GABAergic activity, produced by inhibitory interneurons in the somatosensory cortices [40,41], and again may be related to this state of increased focus in brain activity so as to inhibit surrounding areas. ...
... Beta oscillations are known to be related to GABAergic activity, produced by inhibitory interneurons in the somatosensory cortices [40,41], and again may be related to this state of increased focus in brain activity so as to inhibit surrounding areas. The presence of long-term chronic pain induces changes in the "pain matrix" structures-the thalamus, the anterior cingulate cortex (ACC), the posterior cingulate cortex (PCC), the insula, the amygdala, the primary and secondary somatosensory cortices, and the periaqueductal gray (PAG)-and, consequently, implicates disruption of the descending modulatory pain system and an imbalance between excitation and inhibition of pain signaling [39,42], and thus may disrupt this potential inhibitory activity of beta oscillations. The presence of high EEG beta oscillations can represent an increase in GABAergic activity, more cortical organization, and a better modulatory effect, which may be a result of reduced pain perception [43]. ...
Article
Full-text available
This study aimed to investigate clinical and physiological predictors of brain oscillatory activity in patients with fibromyalgia (FM), assessing resting-state power, event-related desynchronization (ERD), and event-related synchronization (ERS) during tasks. We performed a cross-sectional analysis, including clinical and neurophysiological data from 78 subjects with FM. Multivariate regression models were built to explore predictors of electroencephalography bands. Our findings show a negative correlation between beta oscillations and pain intensity; fibromyalgia duration is positively associated with increased oscillatory power at low frequencies and in the beta band; ERS oscillations in the theta and alpha bands seem to be correlated with better symptoms of FM; fatigue has a signature in the alpha band—a positive relationship in resting-state and a negative relationship in ERS oscillations. Specific neural signatures lead to potential clusters of neural adaptation, in which beta oscillatory activity in the resting state represents a more adaptive activity when pain levels are low and stimulus-evoked oscillations at lower frequencies are likely brain compensatory mechanisms. These neurophysiological changes may help to understand the impact of long-term chronic pain in the central nervous system and the descending inhibitory system in fibromyalgia subjects.
... Not only baricitinib and abrocitinib, but also pregabalin and neurotropin demonstrated a reduction in their therapeutic effects on mechanical alloknesis by inhibiting the descending inhibitory system (Fig. 4b, c, f, and g). The descending inhibitory system can be categorized into 2 main pathways: the noradrenergic pathway, which involves the projection of noradrenalin from the locus coeruleus to the dorsal horn of the spinal cord, and the serotonergic pathway, which projects serotonin from the hypothalamus through the raphe nuclei to the dorsal horn of the spinal cord (32). Within these pathways, there are various receptor types that play a role in their function. ...
... It was reported that α1 and α2 receptors (33) and 5-HT 1 A, 5-HT 1 B, 5-HT 1 D, 5-HT 2 A, 5-HT 3 and 5-HT 4 receptors (34) are expressed in the spinal cord and contribute to the operation of the descending inhibitory system. Historically, the noradrenergic and serotonergic pathways were known to alleviate thermal, mechanical, and neuropathic pain in animal models (32). However, recent research has indicated their involvement in acute (35) and chronic itch (36). ...
Article
Full-text available
Pruritus in the elderly, particularly those cases without skin dryness or other identifiable causes, makes treatment challenging due to the lack of evidence regarding the therapeutic effects of antipruritics. This study proposes an age-related alloknesis mouse model for an evaluation system for such cases, and aimed to investigate the effectiveness and mechanisms of action of several drugs commonly used as antipruritics in Japan, utilizing this model. Mice 69–80 weeks old were used as aged mice, and the level of mechanical alloknesis was counted as the number of scratching behaviours in response to innocuous stimuli. Bepotastine, neurotropin, pregabalin, baricitinib, and abrocitinib were used as antipruritics, and yohimbine and methysergide as inhibitors of the descending inhibitory pathway. The findings suggest that mechanical alloknesis in aged mice is a suitable animal model for assessing pruritus in the elderly without xerosis, and pregabalin, neurotropin, baricitinib, and abrocitinib may be effective antipruritics in the elderly through activating both the noradrenergic and serotonergic descending inhibitory pathways. These findings may be useful for the selection of antipruritics for pruritus in the elderly without skin lesions or dryness.
... Despite the recent advances in cancer pharmacotherapies, traditional pain management still relies on opioids and standard first-and second-line analgesics, which are usually indicated for neuropathic pain conditions [9,30]. Opioid drugs, cannabinoids, and serotonergic agonists can activate the mechanism of the central modulation of pain, providing neuronal activation in brain areas that inhibit or control pain, which are unquestionably critical pharmacological targets for treating chronic pain conditions such as those involving cancer-related pain [18,52,53]. This type of pain continues to be overlooked by healthcare providers, with patients frequently not being asked or not receiving proper education about their pain [54]. ...
Article
Full-text available
Simple Summary Cancer stands as a major public health concern worldwide, and a significant proportion of survivors are affected by recurrent and usually unrelieved cancer pain. The participation of ion channels in cancer pain’s initiation and maintenance is well-described; however, the evidence on the contribution of transient receptor potential vanilloid 4 channels is scarce. The limited studies published to date encountered enhanced TRPV4 expression in cancer-induced bone pain (CIBP), perineural, and orofacial cancer models, increasing pain perception. These outcomes support the hypothesis that the modulation of TRPV4 channels can be targeted as a novel therapeutic approach for treating difficult pain. This review summarizes the role of TRPV4 in cancer etiology and cancer-induced pain mechanisms, as well as the current status of the clinical research. Abstract Despite the unique and complex nature of cancer pain, the activation of different ion channels can be related to the initiation and maintenance of pain. The transient receptor potential vanilloid 4 (TRPV4) is a cation channel broadly expressed in sensory afferent neurons. This channel is activated by multiple stimuli to mediate pain perception associated with inflammatory and neuropathic pain. Here, we focused on summarizing the role of TRPV4 in cancer etiology and cancer-induced pain mechanisms. Many studies revealed that the administration of a TRPV4 antagonist and TRPV4 knockdown diminishes nociception in chemotherapy-induced peripheral neuropathy (CIPN). Although the evidence on TRPV4 channels’ involvement in cancer pain is scarce, the expression of these receptors was reportedly enhanced in cancer-induced bone pain (CIBP), perineural, and orofacial cancer models following the inoculation of tumor cells to the bone marrow cavity, sciatic nerve, and tongue, respectively. Effective pain management is a continuous problem for patients diagnosed with cancer, and current guidelines fail to address a mechanism-based treatment. Therefore, examining new molecules with potential antinociceptive properties targeting TRPV4 modulation would be interesting. Identifying such agents could lead to the development of treatment strategies with improved pain-relieving effects and fewer adverse effects than the currently available analgesics.
... The analgesic effects of serotonin and the role of the descending serotonergic pathway in pain inhibition are well recognized (Martikainen, 2009). Stimulation of the PAG was found to cause serotonin release in the spinal cord and intrathecal administration of 5-HT agonists induced antinociception (Ossipov et al., 2010). Consistent with previous research, this study showed a decrease in DA and serotonin levels in all diabetic groups. ...
Article
Full-text available
Evidence suggests that insulin resistance plays an important role in developing diabetes complications. The association between insulin resistance and pain perception is less well understood. This study aimed to investigate the effects of peripheral insulin deficiency on pain pathways in the brain. Diabetes was induced in 60 male rats with streptozotocin (STZ). Insulin was injected into the left ventricle of the brain by intracerebroventricular (ICV) injection, then pain was induced by subcutaneous injection of 2.5% formalin. Samples were collected at 4 weeks after STZ injection. Dopamine (DA), serotonin, reactive oxygen species (ROS), and mitochondrial glutathione (mGSH) were measured by ELISA, and gene factors were assessed by RT‐qPCR. In diabetic rats, the levels of DA, serotonin, and mGSH decreased in the nuclei of the thalamus, raphe magnus, and periaqueductal gray, and the levels of ROS increased. In addition, the levels of expression of the neuron‐specific enolase and receptor for advanced glycation end genes increased, but the expression of glial fibrillary acidic protein expression was reduced. These results support the findings that insulin has an analgesic effect in non‐diabetic rats, as demonstrated by the formalin test. ICV injection of insulin reduces pain sensation, but this was not observed in diabetic rats, which may be due to cell damage ameliorated by insulin.
Preprint
Full-text available
Hormonal regulation during food ingestion and its association with pain prompted the investigation of the impact of glucagon-like peptide-1 (GLP-1) on the transient receptor potential vanilloid 1 (TRPV1). Both endogenous and synthetic GLP-1 and an antagonist of GLP-1, exendin 9–39, reduced heat sensitivity in naïve mice. GLP-1-derived peptides (liraglutide, exendin-4, and exendin 9–39) effectively inhibited capsaicin (CAP)-induced currents and calcium responses in cultured sensory neurons and TRPV1-expressing cell lines. Notably, the exendin 9–39 alleviated CAP-induced acute pain, as well as chronic pain induced by complete Freund’s adjuvant (CFA) and spared nerve injury (SNI) in mice, without causing hyperthermia associated with other TRPV1 inhibitors. Electrophysiological analyses revealed that exendin 9–39 binds to the extracellular side of TRPV1, functioning as a noncompetitive inhibitor of CAP. Exendin 9–39 did not affect proton-induced TRPV1 activation, suggesting its selective antagonism. Among exendin 9–39 fragments, exendin 20–29 specifically binds to TRPV1, alleviating pain in both acute and chronic pain models without interfering with GLP-1R function. Our study revealed a novel role for GLP-1 and its derivatives in pain relief, proposing exendin 20–29 as a promising therapeutic candidate.
Chapter
Chronic Musculoskeletal Pain is a prevalent cause of Chronic pain with a significant economic burden. Musculoskeletal pain can be caused by repetitive strain, overuse, work-related musculoskeletal disorders, or disease conditions. It has a variety of causes which involves any condition that affects the muscles, bone, joints, connective tissue and systemic diseases with musculoskeletal manifestations. Pain becomes chronic when it lasts for longer than 3 months and lasts more than the anticipated usual course of an acute disease or injury. The pathophysiology of chronic pain involves a combination of peripheral, central and psychologic mechanisms. It is a complex disease process that has neurologic, psychological and physiologic components. The mechanisms and modulation of pain involves a complicated process involving peripheral nociceptors, spinal cord, central nervous system, ascending and descending pain pathways, pain producing chemical agents, neurotransmitters, immune cells, and sympathetic nervous system. A complicated interaction of sensory, motivational and cognitive processes determines the complex expression of pain behavior. There are psychosocial, cultural and environmental factors involved in addition to biology that affect the complex pain experience. This Biopsychosocial model of pain experience is important in designing effective chronic pain programs and including an interdisciplinary model in the evaluation and treatment of chronic pain.
Article
Full-text available
This work aimed to evaluate the potential role of the 5-HT(7) receptor in nociception secondary to a sensitizing stimulus in mice. For this purpose, the effects of relevant ligands (5-HT(7) receptor agonists: AS-19, MSD-5a, E-55888; 5-HT(7) receptor antagonists: SB-258719, SB-269970; 5-HT(1A) receptor agonist: F-13640; 5-HT(1A) receptor antagonist: WAY-100635) were assessed on capsaicin-induced mechanical hypersensitivity, a pain behavior involving hypersensitivity of dorsal horn neurons (central sensitization). For the 5-HT(7) receptor agonists used, binding profile and intrinsic efficacy to stimulate cAMP formation in HEK-293F cells expressing the human 5-HT(7) receptor were also evaluated. AS-19 and E-55888 were selective for 5-HT(7) receptors. E-55888 was a full agonist whereas AS-19 and MSD-5a behaved as partial agonists, with maximal effects corresponding to 77% and 61%, respectively, of the cAMP response evoked by the full agonist 5-HT. Our in vivo results revealed that systemic administration o
Article
Descending influences on the spinal nociceptive tail-flick (TF) reflex produced by focal electrical stimulation and glutamate microinjection in the nuclei reticularis gigantocellularis (NGC) and gigantocellularis pars alpha (NGCα) were examined and characterized in rats lightly anesthetized with pentobarbital. Both inhibition and facilitation of the TF reflex were produced by electrical stimulation at identical sites in the NGC/NGCα; glutamate microinjection only inhibited the TF reflex. The chronaxie of stimulation for inhibition of the TF reflex was 169 ± 28 μsec. Inhibition of the TF reflex by stimulation was produced throughout the NGC and NGCα; intensities of stimulation for inhibition were least in the ventral NGC and in the NGCα. At threshold intensities of stimulation, inhibition of the TF reflex did not outlast the period of stimulation. Facilitation of the TF reflex was produced at many of the same sites at which stimulation inhibited the TF reflex, but always at lesser intensities of stimulation (mean, 10 μA vs. 43 μA for inhibition, n = 25). Stimulation in the NGC/NGCα at threshold intensities for facilitation or inhibition of the TF reflex did not significantly affect blood pressure. Strength-duration characterization of electrical stimulation and microinjection of glutamate into identical sites in the NGC and NGCα suggest that descending inhibition of the TF reflex results from activation of cell bodies in the NGC and NGCα.
Article
Intensely noxious peripheral stimuli of the anaesthetized rat produce two changes in the activity of convergent dorsal horn units: the segmental neuronal pool is activated, while all other convergent neurones are inhibited. These Diffuse Noxious Inhibitory Controls (DNIC) are highly potent (60-80% inhibition) and suppress all convergent neuronal activity, whether spontaneous or evoked by noxious or nonnoxious stimuli. On the other hand, they have no effect on other dorsal horn cell types, including noxious-only and proprioceptive units. The 'DNIC' circuits include at least one supraspinal relay since DNIC is not seen in spinal animals. Furthermore, they are greatly reduced by lesions of the Nucleus Raphe Magnus (NRM). It has been shown that this nucleus massively projects onto the spinal cord, in particular onto the dorsal horn, and that stimulation of the NRM results in convergent unit inhibition of the same degree of magnitude as with DNIC. The role of serotonergic mechanisms in DNIC can be demonstrated pharmacologically: pCPA pre-treatment (3 daily I.P. injections, 300 mg/kg) or cinanserin (4 mg/kg I.V.) both result in a potent decrease (50-80%). We have proposed that the nociceptive message from the convergent units could result in a contrast between activity of the activated segmental pool and silence of the remaining convergent units. If this hypothesis can be verified, then some raphe nuclei and brain stem serotonergic pathways may function as filters in the detection of nociceptive messages, allowing extraction of information from somatic background activity including the firing from peripheral mechanoreceptors. While superficially paradoxical in fact our hypothesis fits well with the observation of profound analgesia following NRM stimulation: indeed, this hypothetical contrast would be completely eliminated by NRM stimulation since both neuronal pools would then be inhibited.
Article
The present study investigated the role of μ, δ, and κ receptors within the RVM in mediating expression of conditional hypoalgesia (CHA). Five groups of rats with RVM cannulae were given daily sessions of paired or unpaired presentations of an auditory CS (white noise) and foot shock across three consecutive days. On the test day, rats in the Paired condition were injected with the μ antagonist CTAP, the δ antagonist naltrindole, the κ antagonist nor-BNI, or saline. Rats in the Unpaired condition were injected with saline. TFLs were measured before and after injections, as well as during and after presentations of the CS. The results showed that none of the drugs affected baseline TFLs. During CS presentation, rats in the Paired condition injected with saline showed longer TFLs than those in the Unpaired condition given saline, confirming the presence of CHA. Expression of this response was blocked by CTAP, but was unaffected by naltrindole or nor-BNI. These results suggest that μ, but not δ or κ, opioid receptors in the RVM mediate expression of CHA.
Article
The antinociceptive interaction on the tail flick (TF) and hot plate (HP) tests between opioid analgesics and medetomidine after intravenous (iv) or intrathecal administration were examined by isobolographic analysis. Male Sprague-Dawley rats received fixed ratios of medetomidine to morphine, fentanyl, and meperidine of 1:10 and 1:30, 10:1, and 1:3, respectively, by iv administration or 10:1 3:1 and 10:1, and 1:3 by intrathecal administration, respectively. Data were expressed as the percentage maximal possible effect (%MPE). The A50 (dose producing 50% MPE) for each drug or drug combination was determined from the dose-response curve. Isobolographic analysis revealed that the effect of medetomidine combined with fentanyl, morphine, or meperidine was additive after iv administration. The intrathecal administration of combinations of medetomidine with the opioids produced a synergistic antinociceptive effect in the TF but not HP test. These data confirmed that the interaction between medetomidine and opioids in producing antinociception may be additive or synergistic, depending on the route of administration, drug ratio administered, and level of processing of the nociceptive input (i.e., spinal vs. supraspinal). Moreover, these results were consistent with a spinal role for alpha-2 adrenoceptors in mediating antinociception. The authors suggest that the interaction between the opioid and alpha-2 adrenergic receptors occurs within the spinal cord.
Article
Placebos have doubtless been used for centuries by wise physicians as well as by quacks, but it is only recently that recognition of an enquiring kind has been given the clinical circumstance where the use of this tool is essential "... to distinguish pharmacological effects from the effects of suggestion, and... to obtain an unbiased assessment of the result of experiment." It is interesting that Pepper could say as recently as 10 years ago "apparently there has never been a paper published discussing [primarily] the important subject of the placebo." In 1953 Gaddum1 said: Such tablets are sometimes called placebos, but it is better to call them dummies. According to the Shorter Oxford Dictionary the word placebo has been used since 1811 to mean a medicine given more to please than to benefit the patient. Dummy tablets are not particularly noted for the pleasure which they give to their recipients.
Article
HOPKINS, E., M. SPINELLA, Z. W. PAVLOVIC AND R. J. BODNAR. Alterations in swim stress-induced analgesia and hypothermia following serotonergic or NMDA antagonists in the rostral ventromedial medulla of rats. PHYSIOL. BEHAV. 64(3) 219–225, 1998.—Serotonergic, NMDA, or opioid antagonists in the rostral ventromedial medulla (RVM) reduce morphine analgesia elicited from the periaqueductal gray (PAG). Continuous (CCWS) and intermittent (ICWS) cold-water swims elicit respective naltrexone-insensitive and naltrexone-sensitive analgesic responses. CCWS analgesia is reduced by systemic NMDA receptor antagonism and by systemic, but not intrathecal serotonergic antagonism. ICWS analgesia is reduced by both systemic and intrathecal serotonergic antagonism, but unaffected by systemic NMDA antagonism. The present study evaluated whether serotonergic (methysergide: 5–10 μg) or competitive [AP7 (2-amino-7-phosphonoheptanoic acid): 0.01–0.1 μg] or non-competitive [MK-801 (dizocilipine maleate): 0.3–3 μg] NMDA antagonists in the RVM altered CCWS and ICWS analgesia and hypothermia as well as basal nociceptive latencies. Methysergide in the RVM significantly potentiated CCWS, but not ICWS analgesia. In contrast, AP7 in the RVM significantly potentiated ICWS analgesia. Antagonist-induced changes in either hypothermia or basal nociception failed to account for any alterations in stress-induced analgesia. These data suggest that serotonergic, but not NMDA, receptors in the RVM may mediate collateral inhibition between mesencephalic morphine analgesia and naltrexone-insensitive CCWS analgesia.
Article
Evidence exists to indicate that tactile allodynia arising from peripheral nerve injury is integrated predominately at supraspinal, rather than spinal, sites. In the present experiments, the possibility that disruption of descending pathways through the dorsolateral funiculus (DLF) might alter expression of nerve-injury induced tactile allodynia was explored. Male, Sprague–Dawley rats received L5/L6 spinal nerve ligation (SNL). Lesions to the DLF were made ipsilateral or contralateral to SNL. Tactile allodynia was determined by measuring withdrawal thresholds to probing with von Frey filaments. Rats with DLF lesions presented no apparent motor deficits and did not alter sensory threshold in sham-SNL operated rats. DLF lesions made ipsilateral to SNL completely blocked tactile allodynia in SNL rats. Contralateral DLF lesions and sham surgery did not have any effect on SNL-induced allodynia. These results indicate that tactile allodynia after peripheral nerve injury is dependent upon tonic activation of net descending facilitation from supraspinal sites and support the hypothesis of tonic activation of descending facilitation as a basis for chronic pain.