Skip to main content
Pain Research and Treatment logoLink to Pain Research and Treatment
. 2012 Apr 24;2012:265605. doi: 10.1155/2012/265605

Drug Management of Visceral Pain: Concepts from Basic Research

Mellar P Davis 1, 2,2,*
PMCID: PMC3348642  PMID: 22619712

Abstract

Visceral pain is experienced by 40% of the population, and 28% of cancer patients suffer from pain arising from intra- abdominal metastasis or from treatment. Neuroanatomy of visceral nociception and neurotransmitters, receptors, and ion channels that modulate visceral pain are qualitatively or quantitatively different from those that modulate somatic and neuropathic pain. Visceral pain should be recognized as distinct pain phenotype. TRPV1, Na 1.8, and ASIC3 ion channels and peripheral kappa opioid receptors are important mediators of visceral pain. Mu agonists, gabapentinoids, and GABAB agonists reduce pain by binding to central receptors and channels. Combinations of analgesics and adjuvants in animal models have supra-additive antinociception and should be considered in clinical trials. This paper will discuss the neuroanatomy, receptors, ion channels, and neurotransmitters important to visceral pain and provide a basic science rationale for analgesic trials and management.

1. Introduction

Normal individuals do not perceive signals emanating from their intestinal tract; however, enteric and extrinsic visceral afferents become hypersensitive in pain-processing disorders such as functional bowel syndromes or in diseases associated with inflammation such as inflammatory bowel disease and pancreatitis. Both inflammatory bowel disease and cancer-related metastases to viscera may produce persistent pain despite resolution of the underlying disease state [1]. Unexplained abdominal pain accounts for 40% of gastroenterology practice in the United Kingdom. Most abdominal pain is due to functional gastrointestinal disorders; irritable bowel syndrome, and functional dyspepsia [2]. Ten to 40% of the normal population will have complaints of abdominal cramps or pain. Most use over-the-counter medications usually antispasmodics and or antacids [3]. Inflammatory bowel disease causes visceral hypersensitivity, intestinal stenosis, anorectal urgency, fistula, and abscess. One-third to half of individuals with inflammatory bowel will have disabling visceral symptoms such as pain or colic or symptoms that resemble the irritable bowel syndrome [4, 5]. Individuals with Crohn's disease in remission frequently have irritable bowel syndrome symptoms as a result of persistent visceral hypersensitivity which may mislead both patients and clinicians to believe that the Crohn's disease is active [69]. Hence, distinguishing functional from organic visceral pain can be a challenge. The duration of pain is longer with functional bowel disorders, whereas organic etiologies produce nocturnal pain and are frequently associated with weight loss and constitutional symptoms [3].

Visceral pain accounts for 28% of cancer-related pain. It is often accompanied by other pains such as neuropathic or somatic pain. Visceral pain in cancer patients may also be the result of treatment complications or comorbid diseases [10]. Causes of cancer-related visceral pain include hepatic metastases with extension to the hepatic capsule, biliary obstruction, pancreatitis as well as pancreatic primaries and peripancreatic nodal enlargement, retroperitoneal adenopathy from metastases, and visceral organ obstruction such as small bowel or colon obstruction, mesenteric infiltration, and peritoneal implants of cancer. Complications include not only intestinal obstruction by perforation and intussusceptions, but also visceral pain is described as pressure-like, intermittently squeezing or cramp, not well localized, vague in character, and difficult for patients to describe. Visceral pain is frequently accompanied by nausea, vomiting, and sweating. Pain, particularly if severe, is often referred to distant somatic (superficial) sites [11]. Treatment recommendations for visceral pain have been the same for somatic pain, yet visceral pain processing is distinctly different from somatic nociception and as a result should perhaps be treated differently from somatic pain (Table 1).

Table 1.

Distinctive features of visceral spinal afferents relative to somatic.

(1) Dual extrinsic afferent system
(2) Convergence of afferents on somatic and other visceral afferents within the spinal cord
(3) Widely overlapping receptive fields
(4) Dichotomization of fibers where a single neuron innervates two viscera
(5) Collateral activation of autonomic and enteric nervous system
(6) Larger cell bodies within dorsal root ganglia
(7) Wide overlapping receptor fields
(8) Lack of specialized nerve terminals
(9) First-order afferents arborize over several spinal segments
(10) Greater expression of transient receptor potential (TRPV1), sodium (Na 1.8), acid (ASIC3) ion channels, and calcitonin gene-related protein (CGRP)
(11) Limited number of stimulus responses (distension, ischemia, and inflammation)
(12) More discrete location of first-order terminals within the spinal cord (superficial dorsal horn, lamina V, and central)
(13) Afferents ascend with parasympathetic and sympathetic neuronal projections
(14) Viscerovisceral hyperalgesia and hypersensitivity
(15) Visceromotor responses and referred pain to somatic sites innervated by the samespinal cord level
(16) Second-order afferents ascend in the dorsal column
(17) Nonsomatotopically arranged input in dorsal column and central lateral thalamus unlike the lateral spinothalamics
(18) Poor representation in S1 cortex
(19) Greater emotional and autonomic responses to pain than somatic pain

2. Neuroanatomy

2.1. Vagus Innervation

This sensory system of the gastrointestinal tract consists of intrinsic (enteric) sensory afferents and extrinsic (vagus, spinal, and pelvic) afferents. The intrinsic system functions independently of the CNS. Neurons are directly exposed mechanical forces and the chemical environment which is unlike somatic afferents neurons. Enterochromaffin cells within the mucosa and enteroendocrine cells release serotonin, cholecystokinin, orexin, and leptin which modulates and regulates motor activity [12]. The submucosal enteric plexus and myenteric plexus have a high degree of synaptic interactions which can be either inhibitory or stimulatory for the purpose of regulating gastrointestinal motility and peristalsis. Both plexuses received input from parasympathetic and sympathetic efferents. Vagal afferents are largely made up of neurons which interact with the submucosal and myenteric plexus and allow “crosstalk” between intrinsic and extrinsic systems [12]. The intrinsic system contains ganglia and interstitial cells of Cajal which act as pacemakers and gauges for muscle tension. Certain ion channels such as the transient receptor potential (TRP) group of ion channels which have a major role in pain processing are not found on intrinsic enteric neurons [12].

The vagus receives input largely from the mucosa and muscle enteric afferents as do the pelvic extrinsic afferents. Cell bodies of the vagus are found in the jugular and nodose ganglia; both ganglia are morphologically, biochemically, and functionally different from one another. Expression of ion channels and signal transduction are distinctly different between the two ganglia [12]. Vagal sensory afferents outnumber efferents 10 to 1 in both ganglia. Both project to the nucleus tractus solitarius [13]. From the nucleus tractus solitarius, second-order afferents project to the parabrachial nucleus and thalamus. Vagal neurons process physiologic information such as the nature and content of luminal constituents and presence or absence of motor activity and in response generate nonpain but potentially noxious symptoms such as nausea, vomiting, and early satiety [13]. Vagal afferent terminals are found in close proximity to enterochromaffin cells and are responsive to regional serotonin by way of certain serotonin (5HT1 5HT 3, 5HT4, and 5HT7) receptors. Ghrelin, released from gastric endocrine cells, binds to receptors on the vagus and nodose ganglia. Ghrelin blocks mucosal and muscle enteric afferents which are mechanically sensitive influencing motility, satiety, gastric emptying, and proximal stomach relaxation. The subjective result is increased appetite and meal volume [14, 15]. Stomach distention, cholecystokinin, and leptin (which bind to receptors on the vagus and nodose ganglia) produce satiety and reduced nutritional intake [1618]. Also found on intrinsic enteric neurons are mu and kappa opioid receptors. Activation of mu receptors inhibits peristalsis, stimulates circular muscle segmentation, reduces intraluminal secretion, and increases fluid absorption from the bowel lumen [19, 20]. The use of mu agonists for visceral pain will therefore have an adverse effect on gastrointestinal motility and may increase patient symptom burden. Vagus muscle mechanicoreceptors are low threshold which elicit maximum firing rates within physiologic distention pressures and differ from extrinsic spinal afferents [21]. The mechanicoreceptors consist of “in-series-tension” receptors between longitudinal and circular muscle and intraganglionic laminar endings which are intimately arrayed around enteric ganglia. Both in series-tension receptors and laminar endings contain nonencapsulated low-threshold sensors with primitive endings responsive to nonnoxious physiologic stimuli [12, 21, 22].

2.2. Spinal Afferents

Extrinsic spinal visceral afferents located in the serosa and mesentery ascend to the thoracolumbar dorsal root ganglia with parasympathetic and sympathetic fibers [21]. A large battery of pain mediators (protons, prostaglandins, interleukins anti-inflammatory cytokines, and bradykinin) are released as a result of visceral inflammation, hollow organ distention or distortion, and ischemia. These mediators bind to various ion channels or receptors found on spinal afferents. Certain channels such as the transient receptor potential ion channels (TRP), acids sensing ion channels (ASIC), and sodium channels (Na) are activated resulting in an action potential. Hence, spinal visceral afferents have promiscuous chemosensitivity [13, 21, 23]. Visceral spinal afferents have a distinctive large cell body relative to somatic afferents visualized within the dorsal root ganglia. Visceral afferents makeup 7% of sensory cell bodies within the dorsal root ganglion [13, 23]. A significant proportion of visceral spinal afferents are high-threshold or silent nociceptors which respond selectively to noxious stimuli. Spinal visceral afferent action potentials propagate centrifugal and centripetal along collaterals resulting in the release of neurotransmitters, mainly calcitonin gene-related protein (CGRP) and substance P (SP) from varicose nerve terminals which in turn modulate blood flow and enteric motor reflexes; the end result is a neurogenic inflammatory response [13, 23]. Visceral spinal afferents converge on somatic afferents within the dorsal root ganglia and dorsal horn which is thought to be the mechanism for somatic referred pain [24, 25]. Convergence also occurs among visceral afferents within the same dorsal horn segment resulting in a viscerovisceral convergence and hyperalgesia [25]. In animal models, visceral pain causes a visceromotor response observed in anterior abdominal wall musculature as a result of convergence which is used as an objective measure of nociception. This can be elicited by experimental colitis, acetic acid peritonitis, or colorectal distention. Visceral convergence leads to cross-organ sensitization in which disease in one visceral organ increases the pain arising from another visceral organ innervated by the same spinal segment [22, 24, 25]. As an example, in mice and rats, experimental colitis increases bladder sensitivity and contractions [26]. Clinical individuals with irritable bowel syndrome have a greater incidence of urinary bladder sensitivity compared with the normal population [27]. Another mechanism proposed for viscerovisceral sensitivity is dichotomized sensory fibers in which one afferent neuron innervates two visceral organs [28, 29]. Another proposed mechanism is generation of retrograde action potentials via viscerovisceral convergence leading to upregulation of neurotransmitters and ion channels in the second visceral organ [30, 31]. Finally, visceral afferents travel with parasympathetic and sympathetic fibers which activate afferents within autonomic ganglia by way of collaterals resulting in an autonomic and motor response in the second visceral organ [32]. Spinal visceral afferents have large overlapping receptive fields relative to somatic afferents and sparse innervation relative to somatic tissues. Because of the anatomical differences and convergence, visceral pain is poorly localized, diffuse, and frequently identified by its somatic referral pattern. Viscerovisceral hypersensitivity can lead to diagnostic dilemmas as to where the pathology is occurring or misdiagnosis of the origin of visceral pain [12]. Somatic convergence can lead to inhibitory or excitatory responses. A counterirritation can produce an inhibitory response at either somatic or visceral site. Inhibitory responses from convergence can last as long as one second and may account for the colicky nature of visceral pain [33]. Patients often use counterirritation at referred somatic sites that reduce visceral pain [33].

Second-order visceral afferents ascend the thoracolumbar spinal cord through the dorsal column to gracile and cuneate nuclei. As a result, an anatomical cervical midline myelotomy interrupts dorsal column secondary afferent neurotransmission and reduces visceral pain and nocifensive responses in animal with pancreatic and duodenal pain [3436]. Third-order afferents ascend to the ventral posterolateral and central lateral nucleus of the thalamus [37, 38]. The central lateral nucleus is the major site for medial afferents traffic important to emotive pain responses. Visceral afferents are highly represented in the medial pain matrix (insular cortex, prefrontal cortex, and anterior cingulate gyrus) which is responsible for the large emotional response to visceral pain. Visceral afferents are sparsely represented in S1 cortex; hence, visceral pain is poorly localizable by patients. The anterior cingulate gyrus is critical to downward modulating pain through the periaqueductal gray and rostral ventromedial medulla (RVM) [39]. The RVM has been demonstrated to exert descending facilitatory influences in visceral pain process and is a critical compound to the maintenance of pain from visceral inflammation [40, 41].

3. Important Channels, Receptors, and Mediators of Visceral Pain

Several channels are important to visceral pain: transient receptor potential vanilloid-1 (TRPV-1), ASIC3 channels and sodium channels (Na) particularly those that are tetrodotoxin resistant (NA 1.8 and 1.9), and calcium channels. Certain receptors downmodulate pain: the gamma aminobutyric acid-B (GABA-B) channels, kappa and mu opioid receptors, and somatostatin receptors. These channels and receptors are potential targets for novel analgesics to treat visceral pain.

3.1. Transient Receptor Potential Ion Channels

The TRP family of ion channels is TRPV1, TRPV2, TRPV3, TRPV4, TRPM 8, and TRPA1. These channels are, in general, thermoreceptors found on poorly myelinated and nonmyelinated afferents arising from the dorsal root ganglia, nodose ganglia, and the CNS [4246]. TRP channels also respond to nonnoxious heat (TRPV3, TRPV4, and TRPM8), capsaicin (TRPV1), methanol (TRPM8) camphor (TRPV3), mustard (TRPA1), protons (TRPV1) and the endogenous cannabinoid, and anandamide (TRPV1) [46]. These channels are calcium permeable, nonselective channels with 4 identical subunits and a central pore. Each subunit consists of a 6 transmembrane (TM) segments with an amphipathic region between TM5 and TM6 that allows cations through. The cytoplasmic N and C terminal tail of the subunits can be modulated by various kinases through phosphorylation. Phosphorylation influences channel endocytosis and expression. Chemical activators may be direct channel activators or may allosterically modify channels to reduce conduction thresholds [46, 47]. The TRP family is subject to multiple single-nucleotide polymorphisms which influence channel activation. These channels are posttranslationally modified by multiple kinases: protein kinase A, protein kinase C, calcium calmodulin kinase II, and the tyrosine kinase src which prevents channel internalization and desensitization, amplifying channel conductance. Multiple pain mediators (certain inflammatory cytokines (ILB), bradykinin, proteases, and ATP) stimulate kinase activation and channel phosphorylation [46, 48, 49].

TRPV1 channels are highly expressed on visceral spinal afferents and colocalize with nerve growth factor receptor (trk-A). They are selectively expressed on peptidergic neurons (neurons which express CGRP and SP) [42, 45]. TRPV1-activated afferents release inflammatory peptides (CGRP and SP to produce neurogenic inflammation, glutamate. TRPV1 plays a significant role in gastrointestinal inflammation, hypersensitization, but also function [5052]. TRPV1 is expressed in bladder afferents and plays a role in micturition as well as pathologic states of interstitial cystitis [53]. Inflammatory states such as Crohn's colitis upregulate TPRV1 through kinases. Nerve growth factor and prostaglandins upregulate TRPV1 through the same mechanism [5457]. TRPV1 has been demonstrated to be upregulated in upper gastrointestinal disorders such as gastroesophageal reflux disease and pancreatitis [58, 59]. TRPV1 remains upregulated not only in those with inflammatory bowel disease in remission but also with irritable bowel type symptoms [60]. TRPV1 is also upregulated in the CNS; inhibitors in development may need to cross the blood-brain barrier to be fully effective [45].

TRPA1 is highly expressed on primary visceral afferents and is important in visceral hypersensitivity [61]. TRPA1 is upregulated in experimental colitis and deletion of TRPA1 expression in a mouse gene “knocke-out” model reduced colitis-induced mechanical hypersensitivity [62].

3.2. Acid Sensing Ion Channels

ASIC family of ion channels is directly gated by protons as “chemoelectrical transducers.” Tissue acidosis with ischemia, fractures, tumor, and incisions activates ASICs at a pH below 7.4, particularly below 7.0 [63, 64]. ASIC undergoes sustained depolarization with tissue acidosis resulting in nonadapting pain [6365]. The channel structure consists of 6 subunits derived from 4 different genes. Each subunit is identified by the intracellular C and extracellular N terminus amino acid composition (ASIC 1a, 1b, 2a, 2b, 3, and 4) [66]. Each subtype has 2 hydrophobic transmembrane domains which flank a large extracellular loop. Although ASIC channels mainly gate protons, they are also permeable to sodium. ASIC subtypes form channel dimers which influence channel expression. Homomeric ASIC 1a dimers conduct calcium [66]. Subtypes 1, 2, and 3 are expressed on intrinsic enteric neurons, and subtypes 3 are particularly expressed on spinal extrinsic afferents [6769]. ASIC3 is up-regulated by inflammatory cytokines, metal ions, and certain neuropeptides (FMRT amide) [70]. Such neuropeptides are increased in the spinal cord and dorsal root ganglia with visceral inflammation. The endogenous kappa opioid receptor agonist, dynorphin, enhances ASIC activity and prevents desensitization [64]. Nonsteroidal anti-inflammatory drugs (NSAIDs) such as diclofenac inhibit ASIC3. Toxins from the spider, Psalmopoeus cambridgei, and the sea anemone, Anthopleura elegantissima, contain ASIC inhibitors. The potassium sparing diuretic, amiloride, also inhibits multiple ASIC subtypes [7174].

ASIC channels are more important to visceral mechanicoreceptors sensory function than cutaneous afferent function [69, 75]. ASIC3 is also an important mediator of cardiac pain from coronary ischemia, cough, and bronchospasm from airway acidification [76, 77].

3.3. Voltage-Gated Sodium Channels

Sodium-gated channels can be divided into those which are tetrodotoxin sensitive (Na 1.1–1.7) and resistant (Na 1.8 and 1.9) [78]. Tetrodotoxin-resistant channels and Na 1.7 are mainly expressed on nociceptor sensory afferents [79]. These membrane glycoproteins have large alpha subunits and smaller beta subunits which form 9 recognizable subsets [80]. These channels have 4 domains made up of individual 6 transmembrane segments (S1-6). The sodium pore is between the fifth and sixth segments. Na 1.8 is the most important sodium channel visceral pain and hypersensitivity [8082]. In experimental colitis, Na 1.8 expression is increased threefold in dorsal root ganglia within that T 9-13 segment of the mouse. There appears to be a selective expression of this channel and not other sodium channels [83, 84]. Na 1.8 channels are also important to bladder hypersensitivity caused by chemical irritants in experimental animals. Na 1.8 is not important to pain of pancreatitis [85, 86]. In the gene “knock-out” (Na 1.8 null) mouse model, experimental jejunitis failed to produce the expected hypersensitivity as measured by visceromotor responses [81, 82]. Both nerve growth factor and prostaglandins (PGE2) increased expression of tetrodotoxin-resistant sodium channels [8789]. As a result, NSAIDs indirectly block sodium channel neurotransmission by preventing PGE2-induced upregulation of sodium channels [89]. Sodium channels can be up-regulated by phosphorylation of their intracellular C terminus [90, 91]. There are no selective Na 1.8 inhibitors which could potentially target visceral pain. There are a group of peptides, (conotoxins) that are a source of sodium channel blockers being developed, one of which has been approved for spinal analgesia [92].

3.4. Voltage-Gated Calcium Channels

Voltage-gated calcium channels have an alpha-1 subunit which forms the channel pore and alpha-2 delta subunit which facilitates alpha-1 traffic to membrane surfaces [93]. There are 10 different alpha-1 subunits and 4 alpha-2 delta subtypes. Alpha-2 delta subtypes 1 and 2 bind pregabalin and gabapentin. Subtype 1 is found within the medial pain matrix (anterior cingulate gyrus, insular cortex, and amygdala) [94, 95]. Alpha-2 delta subunit 1 is also expressed within the dorsal root ganglia, spinal cord, and smooth muscle of the small bowel associated with N-type calcium channels [94]. Subtype 2 is found in the periaqueductal gray, spinal cord, and diffusely throughout the CNS but not in colon or small bowel smooth muscle [95, 96]. Pregabalin and gabapentin bind to alpha-2 delta subunits in the cytoplasm and prevent expression of calcium channels on membranes [97, 98]. By binding and preventing expression, calcium conduction is blocked, and SP, CGRP, and glutamate cannot be released from primary afferent neurons [93]. Prevention of neurotransmitter release by pregabalin and gabapentin occurs only in pathologic states when calcium channels are being up-regulated and activated [99, 100]. Pregabalin and gabapentin are predominantly central acting analgesic [101103]. Gabapentinoids have been shown to reduce visceral hypersensitivity in experimental animals as well as symptoms of irritable bowel syndrome in humans [104109]. Low-dose gabapentin and morphine reduced writhing in rats given intraperitoneal acetic acid at doses which were found to be ineffective as single agents [110]. Not only do gabapentinoids reduce nociceptive neurotransmission centrally but also improve bowel compliance to distention perhaps through blocking alpha-2 delta subunits in smooth muscle [105, 111].

Other calcium channels may also be involved in visceral hypersensitivity. Activation of T-type calcium channels subtype Ca (v) 3.2 on primary spinal visceral afferents has been associated with irritable bowel-like symptoms in an animal model. Symptom behaviors resolved when T-type calcium channels were blocked [112]. In experimental bowel ischemia, mesenteric afferent neurotransmission was blocked by the L-type calcium channel blocker nifedipine [113].

3.5. Opioid Receptors

Kappa and mu opioid receptors are found on visceral afferents [114116]. Distribution along the gastrointestinal tract indicates that endogenous opiate peptides have a modulating function for both gastrointestinal motility and secretory function. Constipating effects and dysmotility with morphine is predominantly due to actions on intrinsic enteric neurons [117]. Kappa receptors have little adverse effect on gastrointestinal motility, but doses are limited by central toxicity such as dysphoria [118120]. The central toxicity of mu receptor agonists includes confusion, myoclonus, nausea, and vomiting which limits titration and adds symptom burden to patients with visceral pain [118]. In addition, certain mu opiates such as the enkephalin derivative DAMGO stimulate mesenteric afferents action potentials which are blocked by the opioid receptor antagonist alvimopan [121]. Mesenteric afferents respond with increasing firing rate when exposed to mu and delta receptor agonists which does not occur with selective kappa receptor agonists [122, 123]. Morphine has been shown to increase the visceral hypersensitivity and viscerosomatic referral in animal models of colitis and irritable bowel syndrome (caused by butyrate) [124].

There are 3 subtypes of kappa receptors which inhibit afferent firing and visceromotor responses to noxious colorectal distention in animal models [125129]. In these animal models, antinociception occurs without change in colon muscle compliance [125]. Kappa opioids in animal models reduced inflammation (colitis) nociceptive responses [130132]. Kappa receptor gene “knocked-out” mice display increasing writhe to acetic acid peritonitis compared with wild-type mice [133]. The kappa receptor agonist, CR 66, in humans selectively reduced visceral pain by esophageal distention pain tolerance but reduced somatic pain thresholds as measured by cutaneous pain [134]. Kappa receptors are important modulators of visceral pain, and it is the peripheral kappa receptor not the central receptor which is important in contradistinction to mu receptor agonists.

Kappa receptors are also found on the vagus and nodose ganglia. The kappa agonist, asimadoline, has been reported to reduce satiety and enhance postprandial gastric volumes [135, 136]. Asimadoline has no influence on motility. Asimadoline reduced pain in a group of patients with irritable bowel syndrome [136]. However, “on-demand” asimadoline was not effective in reducing irritable bowel symptoms and worsened symptoms in diarrhea prone patients [137]. A second peripheral-restricted kappa receptor agonists, fedotozine, reduced nociceptive reflexes triggered by gut distention and reduced pain related to irritable bowel syndrome and nonulcer dyspepsia [138]. Fedotozine reversed ileus caused by acetic acid peritonitis in rats and prevented c-fos expression in the spinal cord, a measure of nociceptive traffic through the dorsal horn. C-fos was also reduced in numerous upstream brain structures demonstrating the importance of blocking nociceptive traffic to prevent neuroplasticity [139]. Fedotozine also did not influence gastrointestinal transient times. Fedotozine was ineffective in reducing symptoms and gastroparesis in diabetic patients [140]. The kappa receptor agonist ADL 10-0101 reduced pain in patients with pancreatitis without influencing heart rate, respiratory rate, oxygen saturation, or causing nausea. Analgesia appeared to be mediated by receptors outside of the CNS [141]. Other peripherally restricted kappa agonists are being developed [142].

DAMGO and morphine are more potent in reducing acetic acid peritonitis in experimental animals when given intracerebroventricular than intraperitoneal. Morphine antinociception was only partially reversed by peripherally restricted mu receptor antagonists [143]. Methylnaltrexone, a peripherally restricted mu receptor antagonist, reverses dysmotility morphine and other mu opioid agonists without reverse seen analgesia [144, 145]. Mu receptor agonists such as fentanyl are reported to reduce visceromotor responses to colorectal distention in mice [146]. Neither morphine, fentanyl, or delta agonists reduced the firing ratio of mesenteric afferents, whereas kappa receptor agonists (U-50,488 and fedotozine) did reduce firing rate suggesting that the opioid-induced sensitization occurs centrally and that mu receptors are either absent on extrinsic spinal afferents or do not influence action potentials [147]. These findings again indicate that central and not peripheral mu receptors are critical to modulating visceral pain; kappa agonists mediate pain relief by binding to peripheral receptors.

There is evidence that fentanyl significantly reduces visceromotor responses to a greater extent when kappa receptors are nonfunctional (kappa receptor gene “knocked-out” mice) [148]. The combination of fentanyl and the kappa receptor agonist spiradoline produced dose-dependent supra-additive antinociception in an animal model involving colorectal distention. However, there appears to be a complex interplay between kappa and mu receptors. Fentanyl nociception was not blocked, but the kappa agonist antinociception was blocked by the mu receptor antagonist (b-FNA) perhaps suggesting an interaction through dimers [149]. There is little data about the benefits of combining kappa and mu receptor agonists in the clinical management of visceral pain. A recent study of human volunteers found that oxycodone which is thought to be a kappa and mu receptor agonist significantly blocked visceral pain better than morphine which has little kappa receptor activity [150]. By speculation, a combination of a mu receptor agonist and a peripherally restricted kappa receptor agonist might selectively improve visceral pain relative to a peripherally restricted kappa agonist or mu agonist alone.

3.6. Gamma Aminobutyric Acid Receptors

Gamma aminobutyric acid receptor-B (GABA-B) is involved in modulating mechanicosensory traffic through the vagus and brainstem pathways. GABA-B receptors block mechanico-sensory input but do not modulate chemosensitivity. GABA-B agonists directly influence sensory input via binding enteric receptors or by way of activating inhibitory interneurons within the CNS [151, 152]. Baclofen, a GABA-B receptor agonist, has been used in animal models to test visceral nociception. GABA-B receptors couple with N and P/Q calcium channels and potassium channels which may be the mechanism of antinociception. Activation of receptors downregulates calcium channels and activates inward rectify potassium channels which improves repolarization of neurons [153155]. Baclofen in male long Evans rats reduced afferent firing from a colorectal distention. This occurs in a dose-dependent manner unrelated to smooth muscle relaxation or improved colon compliance [156, 157]. Visceromotor responses from colorectal distention are reduced by baclofen [157159]. Baclofen also blocks the expression of c-fos in the dorsal horn to experimental colitis and capsaicin-induced bladder irritation [160, 161]. Baclofen reduced presynaptic release of SP from visceral afferents [162]. Volatile anesthetics such as halothane, isoflurane, sevoflurane sevoflurane, and propofol block visceral nociception via augmentation of GABA-B [163, 164]. Baclofen in most clinical studies was intrathecal, and the pain phenotype was largely neuropathic [165169]. Because baclofen crosses the blood-brain barrier, benefits may be mediated centrally. There is some evidence for synergy between morphine and baclofen [166], but very little clinical data are available about the use of baclofen for visceral pain.

3.7. N-Methyl-D-Aspartate Receptors and ATP-Gated Ion Channels

N-Methyl-D-aspartate (NMDA) receptors are calcium channels activated by glutamate and are slowly desensitized once activated. These high-calcium permeable channels generate synaptic neuroplasticity, wide dynamic range neuronal responses, and gene expression within the CNS [170, 171]. NMDA receptor activation has been associated with visceral hyperalgesia [170]. Due to slow inactivation, NMDA receptors produce long-term potentiation (LTP) which generates chronic pain [172]. NMDA receptors are found on enteric extrinsic afferents from colon and bladder; these afferents release CGRP and SP once the receptor is activated [173]. NMDA receptors are also found on postsynaptic, second-order visceral afferents, within the brainstem and cortex [174178]. Receptor activation by colon distention does not require inflammation to be present which is distinctly different from somatic NMDA receptors [179181]. Intravenous but not intrathecal memantine (an NMDA receptor blocker) reduced visceromotor responses to colorectal distention suggesting that peripheral NMDA receptors are important to antinociception [173]. The NMDA receptor blocker, MK-801, prevented dorsal horn expression of c-fos to noxious colorectal distention, and another NMDA receptor blocker, APV, attenuated visceromotor responses to graded colorectal distention in a dose-dependent manner [173]. NMDA receptors within the RVM are important in modulating pain [182].

There are 2 purinergic ion channels, P2X and P2Y. P2X is an ATP-gated channel, and P2Y is a G-protein couple receptor. ATP released by cell damage, from the sympathetic nervous system or extrinsic sensory neurons, binds to P2X. A subset of P2X channels (P2X3) were upregulated in extrinsic afferents neurons by inflammatory bowel disease [183]. ATP released from the epithelial lining of the mucosa binds to several subsets of P2X (P2X3 and P2X2/3) on submucosal sensory neurons. In this way, P2X modulates peristalsis [184186]. P2X, upregulated by interleukin-1b, was found to be a mechanism for postinfectious (jejunitis) hypersensitivity in animal models. Animals not expressing P2X subtype 7 on afferents did not develop postjejunitis visceral hypersensitivity [187].

3.8. Somatostatin and Bradykinin Receptors

Octreotide blocks somatostatin-2 receptors (SST-2). Mesenteric afferent firing rates caused by ramping jejunal pressures in Wistar rats were blunted by octreotide [188]. Reduced firing rates occurred without change in small bowel compliance. Octreotide also blocked bradykinin activation of mesenteric afferents and improved symptom behaviors of irritable bowel syndrome in animals [188190]. Somatostatin influences motility and increases gastrointestinal transit time [191]. Octreotide anecdotally reduced the abdominal pain of Menetrier's disease and pain related to visceral cancer metastases [192, 193]. Octreotide as an adjuvant to morphine reduced pain in a cohort of individuals with gastrointestinal cancers [194]. Octreotide relieved severe refractory epigastric pain in individuals with weight loss and functional bowel disorders [195]. Unfortunately, octreotide did not reduce pain from pancreatitis [196].

Bradykinin activates mesenteric afferents by way of B2 receptors and by stimulating prostaglandin production [197, 198]. The herbal preparation STW-5 (Ibergast Steigerwald Gmb H, Darmstadt, Germany) dosedependently blocks bradykinin-induced afferents discharges from mesenteric sensory neurons [199201]. Ibergast was effective in reducing symptoms from functional dyspepsia as reported in a meta-analysis [202]. Ibergast also alters smooth muscle contraction and motility [203, 204].

4. Protein Kinases

Protein kinase A phosphorylates NMDA receptor subunit NR-1 and the transcription factor cyclic AMP (cAMP) response element binding protein (CREB). CREB upregulates the receptor for SP (neurokinin-1 receptor) [205, 206]. Colonic distention induced protein kinase A phosphorylation of NMDA receptors in the second-order postsynaptic dorsal column neurons. This was be blocked by the NMDA receptor antagonist, CNQX [37]. In the same manner, protein kinase C is upregulated in the gracile nuclei and in postsynaptic dorsal column neurons with visceral stimulation. Protein kinase C phosphorylates a second glutamate receptor, AMPA [207]. Formalin-induced colitis increased the activation of protein kinase C through attachment to neuronal membranes in the dorsal horn [208]. The mitogenic activated protein kinase family of kinases regulates neuron responses and neuroplasticity. Within this family, the extracellular signaling-related kinases (ERK 1, 2) are important to visceral pain. Activation of ERK has been associated with visceral hyperalgesia [209]. ERKs couple to multiple receptors including NMDA and neurokinin-1 [210, 211]. Increased expression ERK 1, 2 in the L6-S1 mouse dorsal horn was associated with a leftward shift in visceromotor responses to colorectal distention. Blocking this kinase attenuated visceromotor responses [212].

5. Pharmacologic Management of Visceral Pain

Clinical trials of specific analgesics for visceral pain are sparsely published. Visceral pain is included with somatic and nociceptive pain in most clinical trials; as a result, it is difficult to determine the appropriate drug choices for visceral pain as a phenotype. The fact that there are some differences between somatic and visceral pain neurotransmission, neurotransmitters, channels, and receptors suggests that there may in fact be real differences in responses to analgesics. The use of potent opioids for inflammatory bowel disease has been associated with higher morbidity and mortality which is not reported for various types of somatic pain [213]. This may reflect the fact that patients on opioids are sicker than those who are not but this may also be related to the opioid. In a similar fashion, certain NSAIDs have been associated with poorer outcomes in inflammatory bowel disease but not for somatic pain [214, 215]. On the other hand, octreotide improves colic and symptoms from bowel obstruction better than anticholinergics; potent opioids actually may worsen colic [216, 217].

It is unlikely that a single analgesic or targeted agent will significantly reduce most visceral pains since multiple neurotransmitters, channels, and receptors are responsible for visceral pain. Analgesics combinations are anticipated to be better than single analgesics [218, 219]. Analgesics which are ineffective as single agents or ineffective at a particular dose may have supra-additive analgesia for visceral pain when combined with a second analgesic. Hence, single-drug trials may not predict the merit of an analgesic supra-additive analgesic combinations cannot be predicted based on single-drug activity. Analgesic dose response relationships for visceral pain may be distinctly different from somatic pain [220]. Opioids for somatic pain may in fact not work well for visceral pain and visa versa. Kappa agonists may be an example [150, 221].

Gastrointestinal motility has been assumed to be a surrogate for gastrointestinal symptoms [222]. For instance, it has been assumed that early satiety is related to reduced gastric emptying or poor proximal gastric compliance. However, kappa receptors improve satiety without changing gastrointestinal motility. Morphine and its metabolite, morphine-6-glucuronide, reduce visceral pain but increase circular muscle contraction, whereas somatostatin reduces visceral pain and reduces smooth muscle contraction [223226]. TRPV1, Na 1.8, ASIC3 targeted analgesics, and kappa receptor antagonists may have less adverse effects on gastrointestinal motility than somatostatin or mu agonists though future trials will need to confirm this speculation.

Another point is that selective visceral blocks and procedures (celiac, hypogastric, and splanchnic blocks, spinal cord stimulators, radiofrequency ablation, and midline myelotomy) needed to complement medical management [35, 227229]. Multimodality approach for pain management requires a multidisciplinary group of experts. There is a tendency to overly rely on analgesics and refer individuals to other disciplines late when pain is poorly controlled; the cancer is widespread and patient tolerance to procedures is poor [230].

There is little doubt that central sensitization plays a role in maintaining visceral pain [79, 182, 231233]. As a result, uncontrolled acute visceral pain will likely lead to central neuroplasticity and chronic pain despite the resolution of underlying cause for pain. Certain analgesics will need to penetrate the CNS to fully and effectively relieve pain [182].

5.1. NSAIDs and Paracetamol

Paracetamol is a weak cyclooxygenase-2 inhibitor and a selective cyclooxygenase 3 inhibitor. It also increases brainstem serotonin neurotransmission, redirects beta-endorphin, and inhibits 5-HT 3 receptors which are pronociceptive [234237]. Paracetamol is commonly used for pain, but little is known about visceral pain responses since most studies have not focused on visceral pain. In an acetic acid peritonitis model, combinations of ketoprofen plus paracetamol were synergistic at a dose ratio of 3:1 (paracetamol to ketoprofen) [238]. Several other studies involving animal models of visceral pain have confirmed benefits of NSAIDs plus paracetamol combinations [239, 240]. In a qualitative systematic review of paracetamol plus NSAIDs combinations, combinations were superior to paracetamol alone in 85% and NSAIDs alone in 64% of studies. Ibuprofen was the most commonly used in the 21 studies reviewed. Drug combinations reduced pain by 35% and analgesic supplemented by 38%. This combination was superior in gynecology surgical studies (83%) indirectly confirming benefits for visceral pain [241].

NSAIDs are effective in reducing cancer pain in the dose-dependent fashion [242]. However, a major systematic review was not selectively focused on visceral pain [243]. Multiple studies have found that NSAIDs are effective for biliary colic and are as effective as opioids with fewer side effects [244247]. One of the mechanisms by which NSAIDs work in renal or biliary colic may involve acetylcholine blockade. Diclofenac blocks acetylcholine-induced smooth muscle contraction [248]. NSAIDs were superior to anticholinergics in relieving biliary colic with a number 3 needed to treat and prevented progression to cholecystitis better than anticholinergic with a number 3 needed to treat [249, 250]. In the same vein, NSAIDs are superior to anticholinergics in relieving renal colic [251]. In a meta-analysis, NSAIDs were superior to opioids in reducing renal colic [252]. Both NSAIDs and opioids provide effective analgesia in acute renal colic, but opioids are associated with a higher incidence of adverse events, particularly vomiting [252].

Nine percent of individuals with inflammatory bowel disease treated with NSAIDs deteriorate clinically and improve once the NSAID is discontinued. Loss of prostaglandin by the NSAID leads to microvascular dysfunction and sustained inflammation [215, 253, 254].

5.2. Opioids

Opioids reduce pain; however, poor coping skills, depression, and catastrophization correlate better with dose than the degree pathology [255, 256]. Opioids add a symptom burden associated with dysmotility [19, 257, 258]. Proactive use of laxatives and bowel softeners are needed. In addition, morphine can produce visceral hyperalgesia [124, 259] which is suppressed by gabapentinoids [260]. This may explain the advantages of this combination when treating visceral pain [219]. Features of hyperalgesia are increased pain intensity extent of pain area, or radiation with reduced responsiveness to opioid analgesia [261]. Opioid-induced hyperalgesia mimics pain associated with progression of the underlying pathological condition. Physicians unaware of this phenomenon increase the opioid dose only to worsen visceral pain. The addition of adjuvant analgesics or opioid rotation is preferred in managing opioid-induced hyperalgesia [262, 263]. Opioid functional selectivity occurs in multiple downstream signaling pathways (beta arrestin 1 and 2, kinases, and G proteins) caused by unique opioid-receptor conformation changes [264]. Functional selectivity is also determined by expression of certain proteins within the vicinity of the receptor. Opioid rotation will alter downstream signaling and reset the analgesic response. Paradoxically, opioid dose reduction may improve pain control [265].

Sustained release oxycodone significantly reduces chronic cancer-related visceral pain. In a multicenter prospective observational study involving 350 individuals with visceral pain, oxycodone reduced pain severity from a mean of 7 to a mean of 2.4 on a numerical rating scale (0 no pain, 10 severe pain) over a two-week period of time. Less than 5% of individuals continued to have severe pain at the end of 2 weeks [266]. A small randomized trial found that morphine was as effective as oxycodone in relieving pain related to pancreatic cancer [267]. Intranasal fentanyl was effective in managing acute visceral pain in the emergency department [268]. At the present time, there is little published data to give direction to the choice of opioids for visceral pain.

Combinations of analgesics which include opioids have been reported in the management of visceral pain. In an animal model, blocking TRPV1 enhanced opioid antinociception [269]. In a mouse model, the triple combination of fentanyl, trazodone, and paracetamol reduced writhing to acetic acid at doses lower than those seen with single drugs [270]. In the rat colorectal distention model, ketorolac alone was ineffective, and morphine modestly effective but the combination of ketorolac and morphine produced synergistic antinociception at dose ratios of 10:1 and 20:1 [271]. Combinations of tramadol plus morphine and tramadol plus fentanyl produced synergistic antinociception in the acetic acid writhe model but not with somatic pain (hot plate test) [272].

5.3. Adjuvant Analgesics

There is preclinical and some clinical evidence that adjuvant analgesics should be used early for visceral pain. The combination of a gabapentinoid plus opioid is a reasonable choice [109, 219]. Octreotide reduces visceral hyperalgesia which is a logical choice to use with morphine in malignant-related bowel obstruction [189]. STW-5 blocks bradykinin receptors and pain from jejunal distention and could potentially improve opioid analgesia [200]. Baclofen reduces visceral nociception in multiple models which makes it attractive to use with opioids and nonopioid analgesics (NSAIDS or paracetamol) [156]. NMDA receptor antagonists reduced visceral pain in a different manner advanced somatic pain and may improve the opioid analgesia for visceral pain differently than from somatic or neuropathic pain [173]. Combinations of NSAIDs, paracetamol, and opioids are likely to be at least additive if not synergistic [241]. Dexmedetomidine, an alpha-2 antagonist, inhibited visceromotor responses to colorectal distention [273]. In the same manner, tricyclic antidepressants and serotonin norepinephrine reuptake inhibitors could improve visceral pain by increasing norepinephrine neurotransmission. In addition, tricyclic antidepressants inhibit sodium channels, NMDA receptors, P2X channels, and prostaglandins [274]. Combinations of antidepressants and opioids may improve visceral pain in a similar fashion as the combination does for neuropathic pain. Overall, the clinical evidence for drug combinations is weak and requires prospective clinical trials.

6. Summary

Visceral pain is mediated by unique peripheral and central pathways. Several ion channels TRPV1, Na 1.8 and ASIC3, and the kappa opioid receptor appear are particularly important to modulating pain. There are qualitative and quantitative differences in pain processing from somatic pain such that visceral pain should be considered a distinct phenotype. Drug development and treatment paradigms have to take this into consideration in cohort and randomized trials. This has been done successfully for the irritable bowel syndrome. Targeted agents are being developed to ion channels; it is unlikely that these targeted agents will produce less dysmotility than mu opioid receptor agonists. Ion channel blockers are likely to have only a modest benefit due to the multiplicity of ion channels involved in visceral pain and will have to be combined with other analgesics. There is evidence that targeted agents for TRPV1 channels improve opioid responses. Adjuvant analgesics such as the gabapentinoids are effective as single agents and should be considered early in the course of opioid therapy to prevent opioid-induced hyperalgesia and minimize opioid side effects by being “opioid sparing.” Peripherally restricted kappa agonists are in development and should be combined with other analgesics and adjuvants in hope of improving pain control. Interventional procedures should be considered early to avoid chronic visceral pain poorly responsive to analgesics. Treatment algorithms have been developed for neuropathic pain based on multiple randomized trials. The same approach should be taken in the management of visceral pain.

References

  • 1.Bielefeldt K, Christianson JA, Davis RM. Basic and clinical aspects of visceral sensation: transmission in the CNS. Neurogastroenterology and Motility. 2005;17(4):488–499. doi: 10.1111/j.1365-2982.2005.00671.x. [DOI] [PubMed] [Google Scholar]
  • 2.Locke GR, III, Talley NJ, Fett SL, Zinsmeister AR, Melton LJ., III Prevalence and clinical spectrum of gastroesophageal reflux: a population-based study in Olmsted County, Minnesota. Gastroenterology. 1997;112(5):1448–1456. doi: 10.1016/s0016-5085(97)70025-8. [DOI] [PubMed] [Google Scholar]
  • 3.Makharia GK. Understanding and treating abdominal pain and spasms in organic gastrointestinal diseases: Inflammatory bowel disease and biliary diseases. Journal of Clinical Gastroenterology. 2011;45(supplement 2):S89–S93. doi: 10.1097/MCG.0b013e31821fbd82. [DOI] [PubMed] [Google Scholar]
  • 4.Cross RK, Wilson KT, Binion DG. Narcotic use in patients with Crohn’s disease. American Journal of Gastroenterology. 2005;100(10):2225–2229. doi: 10.1111/j.1572-0241.2005.00256.x. [DOI] [PubMed] [Google Scholar]
  • 5.Edwards JT, Radford-Smith GL, Florin TH. Chronic narcotic use in inflammatory bowel disease patients: prevalence and clinical characteristics. Journal of Gastroenterology and Hepatology. 2001;16(11):1235–1238. doi: 10.1046/j.1440-1746.2001.02468.x. [DOI] [PubMed] [Google Scholar]
  • 6.Long MD, Drossman DA. Inflammatory bowel disease, irritable bowel syndrome, or what: a challenge to the functional-organic dichotomy. American Journal of Gastroenterology. 2010;105(8):1796–1798. doi: 10.1038/ajg.2010.162. [DOI] [PubMed] [Google Scholar]
  • 7.Spiller R, Garsed K. Infection, inflammation, and the irritable bowel syndrome. Digestive and Liver Disease. 2009;41(12):844–849. doi: 10.1016/j.dld.2009.07.007. [DOI] [PubMed] [Google Scholar]
  • 8.Simrén M, Axelsson J, Gillberg R, Abrahamsson H, Svedlund J, Björnsson ES. Quality of life in inflammatory bowel disease in remission: the impact of IBS-like symptoms and associated psychological factors. American Journal of Gastroenterology. 2002;97(2):389–396. doi: 10.1111/j.1572-0241.2002.05475.x. [DOI] [PubMed] [Google Scholar]
  • 9.Ansari R, Attari F, Razjouyan H, et al. Ulcerative colitis and irritable bowel syndrome: relationships with quality of life. European Journal of Gastroenterology and Hepatology. 2008;20(1):46–50. doi: 10.1097/MEG.0b013e3282f16a62. [DOI] [PubMed] [Google Scholar]
  • 10.Caraceni A, Portenoy RK. An international survey of cancer pain characteristics and syndromes. Pain. 1999;82(3):263–274. doi: 10.1016/S0304-3959(99)00073-1. [DOI] [PubMed] [Google Scholar]
  • 11.Christo PJ, Mazloomdoost D. Cancer pain and analgesia. Annals of the New York Academy of Sciences. 2008;1138:278–298. doi: 10.1196/annals.1414.033. [DOI] [PubMed] [Google Scholar]
  • 12.Blackshaw LA, Brookes SJ, Grundy D, Schemann M. Sensory transmission in the gastrointestinal tract. Neurogastroenterology and Motility. 2007;19(1):1–3. doi: 10.1111/j.1365-2982.2006.00871.x. [DOI] [PubMed] [Google Scholar]
  • 13.Grundy D. Signalling the state of the digestive tract. Autonomic Neuroscience. 2006;125(1-2):76–80. doi: 10.1016/j.autneu.2006.01.009. [DOI] [PubMed] [Google Scholar]
  • 14.Cummings DE, Frayo RS, Marmonier C, Aubert R, Chapelot D. Plasma ghrelin levels and hunger scores in humans initiating meals voluntarily without time- and food-related cues. American Journal of Physiology. 2004;287(2 50-2):E297–E304. doi: 10.1152/ajpendo.00582.2003. [DOI] [PubMed] [Google Scholar]
  • 15.Wren AM, Seal LJ, Cohen MA, et al. Ghrelin enhances appetite and increases food intake in humans. Journal of Clinical Endocrinology and Metabolism. 2001;86(12):5992–5995. doi: 10.1210/jcem.86.12.8111. [DOI] [PubMed] [Google Scholar]
  • 16.Deutsch JA, Young WG, Kalogeris TJ. The stomach signals satiety. Science. 1978;201(4351):165–167. doi: 10.1126/science.663647. [DOI] [PubMed] [Google Scholar]
  • 17.Moran TH, Kinzig KP. Gastrointestinal satiety signals. II. Cholecystokinin. American Journal of Physiology. 2004;286(2):G183–G188. doi: 10.1152/ajpgi.00434.2003. [DOI] [PubMed] [Google Scholar]
  • 18.Buyse M, Ovesjö ML, Goïot H, et al. Expression and regulation of leptin receptor proteins in afferent and efferent neurons of the vagus nerve. European Journal of Neuroscience. 2001;14(1):64–72. doi: 10.1046/j.0953-816x.2001.01628.x. [DOI] [PubMed] [Google Scholar]
  • 19.Davis MP. The opioid bowel syndrome: a review of pathophysiology and treatment. Journal of Opioid Management. 2005;1(3):153–161. doi: 10.5055/jom.2005.0035. [DOI] [PubMed] [Google Scholar]
  • 20.Page AJ, O’Donnell TA, Blackshaw LA. Opioid modulation of ferret vagal afferent mechanosensitivity. American Journal of Physiology. 2008;294(4):G963–G970. doi: 10.1152/ajpgi.00562.2007. [DOI] [PubMed] [Google Scholar]
  • 21.Grundy D. What activates visceral afferents? Gut. 2004;53(2):ii5–ii8. doi: 10.1136/gut.2003.033415. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Robinson DR, Gebhart GF. Inside information: the unique features of visceral sensation. Molecular Interventions. 2008;8(5):242–253. doi: 10.1124/mi.8.5.9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Grundy D, Schemann M. Enteric nervous system. Current Opinion in Gastroenterology. 2006;22(2):102–110. doi: 10.1097/01.mog.0000208459.46395.16. [DOI] [PubMed] [Google Scholar]
  • 24.Brumovsky PR, Gebhart GF. Visceral organ cross-sensitization—an integrated perspective. Autonomic Neuroscience. 2010;153(1-2):106–115. doi: 10.1016/j.autneu.2009.07.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Giamberardino MA, Vecchiet L. Visceral pain, referred hyperalgesia and outcome: new concepts. European Journal of Anaesthesiology, Supplement. 1995;12(10):61–66. [PubMed] [Google Scholar]
  • 26.Pezzone MA, Liang R, Fraser MO. A model of neural cross-talk and irritation in the pelvis: implications for the overlap of chronic pelvic pain disorders. Gastroenterology. 2005;128(7):1953–1964. doi: 10.1053/j.gastro.2005.03.008. [DOI] [PubMed] [Google Scholar]
  • 27.Whorwell PJ, McCallum M, Creed FH, Roberts CT. Non-colonic features of irritable bowel syndrome. Gut. 1986;27(1):37–40. doi: 10.1136/gut.27.1.37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Sinclair DC, Weddell G, Feindel WH. Referred pain and associated phenomena. Brain. 1948;71(2):184–211. doi: 10.1093/brain/71.2.184. [DOI] [PubMed] [Google Scholar]
  • 29.Giamberardino MA, Costantini R, Affaitati G, et al. Viscero-visceral hyperalgesia: characterization in different clinical models. Pain. 2010;151(2):307–322. doi: 10.1016/j.pain.2010.06.023. [DOI] [PubMed] [Google Scholar]
  • 30.Ustinova EE, Gutkin DW, Pezzone MA. Sensitization of pelvic nerve afferents and mast cell infiltration in the urinary bladder following chronic colonic irritation is mediated by neuropeptides. American Journal of Physiology. 2007;292(1):F123–F130. doi: 10.1152/ajprenal.00162.2006. [DOI] [PubMed] [Google Scholar]
  • 31.Ustinova EE, Fraser MO, Pezzone MA. Colonic irritation in the rat sensitizes urinary bladder afferents to mechanical and chemical stimuli: an afferent origin of pelvic organ cross-sensitization. American Journal of Physiology. 2006;290(6):F1478–F1487. doi: 10.1152/ajprenal.00395.2005. [DOI] [PubMed] [Google Scholar]
  • 32.Cohen DP, Ikeda SR, Lewis DL. Neuropeptide Y and calcitonin gene-related peptide modulate voltage- gated Ca2+ channels in mature female rat paracervical ganglion neurons. Journal of the Society for Gynecologic Investigation. 1996;3(6):342–349. [PubMed] [Google Scholar]
  • 33.Tattersall JE, Cervero F, Lumb BM. Viscerosomatic neurons in the lower thoracic spinal cord of the cat: excitations and inhibitions evoked by splanchnic and somatic nerve volleys and by stimulation of brain stem nuclei. Journal of Neurophysiology. 1986;56(5):1411–1423. doi: 10.1152/jn.1986.56.5.1411. [DOI] [PubMed] [Google Scholar]
  • 34.Giesler GJ, Jr., Nahin RL, Madsen AM. Postsynaptic dorsal column pathway of the rat. I. Anatomical studies. Journal of Neurophysiology. 1984;51(2):260–275. doi: 10.1152/jn.1984.51.2.260. [DOI] [PubMed] [Google Scholar]
  • 35.Viswanathan A, Burton AW, Rekito A, McCutcheon IE. Commissural myelotomy in the treatment of intractable visceral pain: technique and outcomes. Stereotactic and Functional Neurosurgery. 2010;88(6):374–382. doi: 10.1159/000319041. [DOI] [PubMed] [Google Scholar]
  • 36.Becker R, Sure U, Bertalanffy H, Sindou M. Punctate midline myelotomy: a new approach in the management of visceral pain. Acta Neurochirurgica. 1999;141(8):881–883. doi: 10.1007/s007010050390. [DOI] [PubMed] [Google Scholar]
  • 37.Willis WD, Al-Chaer ED, Quast MJ, Westlund KN. A visceral pain pathway in the dorsal column of the spinal cord. Proceedings of the National Academy of Sciences of the United States of America. 1999;96(14):7675–7679. doi: 10.1073/pnas.96.14.7675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Ren Y, Zhang L, Lu Y, Yang H, Westlun KN. Central lateral thalamic neurons receive noxious visceral mechanical and chemical input in rats. Journal of Neurophysiology. 2009;102(1):244–258. doi: 10.1152/jn.90985.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Cao Z, Wu X, Chen S, et al. Anterior cingulate cortex modulates visceral pain as measured by visceromotor responses in viscerally hypersensitive rats. Gastroenterology. 2008;134(2):535–543. doi: 10.1053/j.gastro.2007.11.057. [DOI] [PubMed] [Google Scholar]
  • 40.Vera-Portocarrero LP, Yie JX, Kowal J, Ossipov MH, King T, Porreca F. Descending facilitation from the rostral ventromedial medulla maintains visceral pain in rats with experimental pancreatitis. Gastroenterology. 2006;130(7):2155–2164. doi: 10.1053/j.gastro.2006.03.025. [DOI] [PubMed] [Google Scholar]
  • 41.Urban MO, Gebhart GF. Supraspinal contributions to hyperalgesia. Proceedings of the National Academy of Sciences of the United States of America. 1999;96(14):7687–7692. doi: 10.1073/pnas.96.14.7687. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Hwang SJ, Valtschanoff JG. Vanilloid receptor VR1-positive afferents are distributed differently at different levels of the rat lumbar spinal cord. Neuroscience Letters. 2003;349(1):41–44. doi: 10.1016/s0304-3940(03)00750-x. [DOI] [PubMed] [Google Scholar]
  • 43.Guo A. Immunocytochemical localization of the vanilloid receptor 1 (VR1): relationship to neliropeptides, the P2X3 purinoceptor and IB4 binding sites. European Journal of Neuroscience. 1999;11(3):946–958. doi: 10.1046/j.1460-9568.1999.00503.x. [DOI] [PubMed] [Google Scholar]
  • 44.Valtschanoff JG, Rustioni A, Guo A, Hwang SJ. Vanilloid receptor VR1 is both presynaptic and postsynaptic in the superficial laminae of the rat dorsal horn. Journal of Comparative Neurology. 2001;436(2):225–235. [PubMed] [Google Scholar]
  • 45.Roberts LA, Connor M. TRPV1 antagonists as a potential treatment for hyperalgesia. Recent Patents on CNS Drug Discovery. 2006;1(1):65–76. doi: 10.2174/157488906775245309. [DOI] [PubMed] [Google Scholar]
  • 46.Messeguer A, Planells-Cases R, Ferrer-Montiel A. Physiology and pharmacology of the vanilloid receptor. Current Neuropharmacology. 2006;4(1):1–15. doi: 10.2174/157015906775202995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Zygmunt PM, Petersson J, Andersson DA, et al. Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamide. Nature. 1999;400(6743):452–457. doi: 10.1038/22761. [DOI] [PubMed] [Google Scholar]
  • 48.Miranda A, Nordstrom E, Mannem A, Smith C, Banerjee B, Sengupta JN. The role of transient receptor potential vanilloid 1 in mechanical and chemical visceral hyperalgesia following experimental colitis. Neuroscience. 2007;148(4):1021–1032. doi: 10.1016/j.neuroscience.2007.05.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Kirkup AJ, Brunsden AM, Grundy D. Receptors and transmission in the brain-gut axis: potential for novel therapies I. Receptors on visceral afferents. American Journal of Physiology. 2001;280(5):G787–G794. doi: 10.1152/ajpgi.2001.280.5.G787. [DOI] [PubMed] [Google Scholar]
  • 50.Jones RCW, Jr., Xu L, Gebhart GF. The mechanosensitivity of mouse colon afferent fibers and their sensitization by inflammatory mediators require transient receptor potential vanilloid 1 and acid-sensing ion channel 3. Journal of Neuroscience. 2005;25(47):10981–10989. doi: 10.1523/JNEUROSCI.0703-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Szallasi A, Cortright DN, Blum CA, Eid SR. The vanilloid receptor TRPV1: 10 years from channel cloning to antagonist proof-of-concept. Nature Reviews Drug Discovery. 2007;6(5):357–372. doi: 10.1038/nrd2280. [DOI] [PubMed] [Google Scholar]
  • 52.Geppetti P, Trevisani M. Activation and sensitisation of the vanilloid receptor: role in gastrointestinal inflammation and function. British Journal of Pharmacology. 2004;141(8):1313–1320. doi: 10.1038/sj.bjp.0705768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Birder LA, Nakamura Y, Kiss S, et al. Altered urinary bladder function in mice lacking the vanilloid receptor TRPV1. Nature Neuroscience. 2002;5(9):856–860. doi: 10.1038/nn902. [DOI] [PubMed] [Google Scholar]
  • 54.Chan CL, Facer P, Davis JB, et al. Sensory fibres expressing capsaicin receptor TRPV1 in patients with rectal hypersensitivity and faecal urgency. The Lancet. 2003;361(9355):385–391. doi: 10.1016/s0140-6736(03)12392-6. [DOI] [PubMed] [Google Scholar]
  • 55.Yiangou Y, Facer P, Dyer NH, et al. Vanilloid receptor 1 immunoreactivity in inflamed human bowel. The Lancet. 2001;357(9265):1338–1339. doi: 10.1016/s0140-6736(00)04503-7. [DOI] [PubMed] [Google Scholar]
  • 56.Ji RR, Samad TA, Jin SX, Schmoll R, Woolf CJ. p38 MAPK activation by NGF in primary sensory neurons after inflammation increases TRPV1 levels and maintains heat hyperalgesia. Neuron. 2002;36(1):57–68. doi: 10.1016/s0896-6273(02)00908-x. [DOI] [PubMed] [Google Scholar]
  • 57.Premkumar LS, Ahern GP. Induction of vanilloid receptor channel activity by protein kinase C. Nature. 2000;408(6815):985–990. doi: 10.1038/35050121. [DOI] [PubMed] [Google Scholar]
  • 58.Matthews PJ, Aziz Q, Facer P, Davis JB, Thompson DG, Anand P. Increased capsaicin receptor TRPV1 nerve fibres in the inflamed human oesophagus. European Journal of Gastroenterology and Hepatology. 2004;16(9):897–902. doi: 10.1097/00042737-200409000-00014. [DOI] [PubMed] [Google Scholar]
  • 59.Adam B, Liebregts T, Gschossmann JM, et al. Severity of mucosal inflammation as a predictor for alterations of visceral sensory function in a rat model. Pain. 2006;123(1-2):179–186. doi: 10.1016/j.pain.2006.02.029. [DOI] [PubMed] [Google Scholar]
  • 60.Akbar A, Yiangou Y, Facer P, et al. Expression of the TRPV1 receptor differs in quiescent inflammatory bowel disease with or without abdominal pain. Gut. 2010;59(6):767–774. doi: 10.1136/gut.2009.194449. [DOI] [PubMed] [Google Scholar]
  • 61.Lapointe TK, Altier C. The role of TRPA1 in visceral inflammation and pain. Channels. 2011;5(6):20–24. doi: 10.4161/chan.5.6.18016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Cattaruzza F, Spreadbury I, Miranda-Morales M, Grady EF, Vanner S, Bunnett NW. Transient receptor potential ankyrin-1 has a major role in mediating visceral pain in mice. American Journal of Physiology. 2010;298(1):G81–G91. doi: 10.1152/ajpgi.00221.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Waldmann R, Champigny G, Bassilana F, Heurteaux C, Lazdunski M. A proton-gated cation channel involved in acid-sensing. Nature. 1997;386(6621):173–177. doi: 10.1038/386173a0. [DOI] [PubMed] [Google Scholar]
  • 64.Deval E, Gasull X, Noël J, et al. Acid-sensing ion channels (ASICs): pharmacology and implication in pain. Pharmacology and Therapeutics. 2010;128(3):549–558. doi: 10.1016/j.pharmthera.2010.08.006. [DOI] [PubMed] [Google Scholar]
  • 65.Yagi J, Wenk HN, Naves LA, McCleskey EW. Sustained currents through ASIC3 ion channels at the modest pH changes that occur during myocardial ischemia. Circulation Research. 2006;99(5):501–509. doi: 10.1161/01.RES.0000238388.79295.4c. [DOI] [PubMed] [Google Scholar]
  • 66.Deval E, Friend V, Thirant C, et al. Regulation of sensory neuron-specific acid-sensing ion channel 3 by the adaptor protein Na+/H+ exchanger regulatory factor-1. Journal of Biological Chemistry. 2006;281(3):1796–1807. doi: 10.1074/jbc.M509669200. [DOI] [PubMed] [Google Scholar]
  • 67.Yiangou Y, Facer P, Smith JA, et al. Increased acid-sensing ion channel ASIC-3 in inflamed human intestine. European Journal of Gastroenterology and Hepatology. 2001;13(8):891–896. doi: 10.1097/00042737-200108000-00003. [DOI] [PubMed] [Google Scholar]
  • 68.Wultsch T, Painsipp E, Shahbazian A, et al. Deletion of the acid-sensing ion channel ASIC3 prevents gastritis-induced acid hyperresponsiveness of the stomach-brainstem axis. Pain. 2008;134(3):245–253. doi: 10.1016/j.pain.2007.04.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Chen CC, Zimmer A, Sun WH, Hall J, Brownstein MJ. A role for ASIC3 in the modulation of high-intensity pain stimuli. Proceedings of the National Academy of Sciences of the United States of America. 2002;99(13):8992–8997. doi: 10.1073/pnas.122245999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Deval E, Noël J, Lay N, et al. ASIC3, a sensor of acidic and primary inflammatory pain. The EMBO Journal. 2008;27(22):3047–3055. doi: 10.1038/emboj.2008.213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Bassilana F, Champigny G, Waldmann R, De Weille JR, Heurteaux C, Lazdunski M. The acid-sensitive ionic channel subunit ASIC and the mammalian degenerin MDEG form a heteromultimeric H+-gated Na+ channel with novel properties. Journal of Biological Chemistry. 1997;272(46):28819–28822. doi: 10.1074/jbc.272.46.28819. [DOI] [PubMed] [Google Scholar]
  • 72.Chen X, Orser BA, MacDonald JF. Design and screening of ASIC inhibitors based on aromatic diamidines for combating neurological disorders. European Journal of Pharmacology. 2010;648(1–3):15–23. doi: 10.1016/j.ejphar.2010.09.005. [DOI] [PubMed] [Google Scholar]
  • 73.Voilley N, de Weille J, Mamet J, Lazdunski M. Nonsteroid anti-inflammatory drugs inhibit both the activity and the inflammation-induced expression of acid-sensing ion channels in nociceptors. Journal of Neuroscience. 2001;21(20):8026–8033. doi: 10.1523/JNEUROSCI.21-20-08026.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Dorofeeva NA, Barygin OI, Staruschenko A, Bolshakov KV, Magazanik LG. Mechanisms of non-steroid anti-inflammatory drugs action on ASICs expressed in hippocampal interneurons. Journal of Neurochemistry. 2008;106(1):429–441. doi: 10.1111/j.1471-4159.2008.05412.x. [DOI] [PubMed] [Google Scholar]
  • 75.Page AJ, Brierley SM, Martin CM, et al. The ion channel ASIC1 contributes to visceral but not cutaneous mechanoreceptor function. Gastroenterology. 2004;127(6):1739–1747. doi: 10.1053/j.gastro.2004.08.061. [DOI] [PubMed] [Google Scholar]
  • 76.Sutherland SP, Benson CJ, Adelman JP, McCleskey EW. Acid-sensing ion channel 3 matches the acid-gated current in cardiac ischemia-sensing neurons. Proceedings of the National Academy of Sciences of the United States of America. 2001;98(2):711–716. doi: 10.1073/pnas.011404498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Canning BJ, Farmer DG, Mori N. Mechanistic studies of acid-evoked coughing in anesthetized guinea pigs. American Journal of Physiology. 2006;291(2):R454–R463. doi: 10.1152/ajpregu.00862.2005. [DOI] [PubMed] [Google Scholar]
  • 78.Catterall WA. Cellular and molecular biology of voltage-gated sodium channels. Physiological Reviews. 1992;72(supplement 4):S15–S48. doi: 10.1152/physrev.1992.72.suppl_4.S15. [DOI] [PubMed] [Google Scholar]
  • 79.Woolf CJ, Costigan M. Transcriptional and posttranslational plasticity and the generation of inflammatory pain. Proceedings of the National Academy of Sciences of the United States of America. 1999;96(14):7723–7730. doi: 10.1073/pnas.96.14.7723. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Qi FH, Zhou YL, Xu GY. Targeting voltage-gated sodium channels for treatment for chronic visceral pain. World Journal of Gastroenterology. 2011;17(19):2357–2364. doi: 10.3748/wjg.v17.i19.2357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Laird JM, Souslova V, Wood JN, Cervero F. Deficits in visceral pain and referred hyperalgesia in Nav1.8 (SNS/PN3)-null mice. Journal of Neuroscience. 2002;22(19):8352–8356. doi: 10.1523/JNEUROSCI.22-19-08352.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Hillsley K, Lin JH, Stanisz A, et al. Dissecting the role of sodium currents in visceral sensory neurons in a model of chronic hyperexcitability using Nav 1.8 and Nav 1.9 null mice. Journal of Physiology. 2006;576(1):257–267. doi: 10.1113/jphysiol.2006.113597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Cervero F, Laird JM. Role of ion channels in mechanisms controlling gastrointestinal pain pathways. Current Opinion in Pharmacology. 2003;3(6):608–612. doi: 10.1016/j.coph.2003.06.007. [DOI] [PubMed] [Google Scholar]
  • 84.Beyak MJ, Vanner S. Inflammation-induced hyperexcitability of nociceptive gastrointestinal DRG neurones: the role of voltage-gated ion channels. Neurogastroenterology and Motility. 2005;17(2):175–186. doi: 10.1111/j.1365-2982.2004.00596.x. [DOI] [PubMed] [Google Scholar]
  • 85.Yoshimura N, Seki S, Novakovic SD, et al. The involvement of the tetrodotoxin-resistant sodium channel Nav1.8 (PN3/SNS) in a rat model of visceral pain. Journal of Neuroscience. 2001;21(21):8690–8696. doi: 10.1523/JNEUROSCI.21-21-08690.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Gebhart GF, Bielefeldt K, Ozaki N. Gastric hyperalgesia and changes in voltage gated sodium channel function in the rat. Gut. 2002;51(1):i15–i18. doi: 10.1136/gut.51.suppl_1.i15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Beyak MJ, Ramji N, Krol KM, Kawaja MD, Vanner SJ. Two TTX-resistant Na+ currents in mouse colonic dorsal root ganglia neurons and their role in colitis-induced hyperexcitability. American Journal of Physiology. 2004;287(4):G845–G855. doi: 10.1152/ajpgi.00154.2004. [DOI] [PubMed] [Google Scholar]
  • 88.Gold MS, Zhang L, Wrigley DL, Traub RJ. Prostaglandin E2 modulates TTX-R /Na in rat colonic sensory neurons. Journal of Neurophysiology. 2002;88(3):1512–1522. doi: 10.1152/jn.2002.88.3.1512. [DOI] [PubMed] [Google Scholar]
  • 89.Ma K, Zhou QH, Chen J, Du DP, Ji Y, Jiang W. TTX-R Na+ current-reduction by celecoxib correlates with changes in PGE2 and CGRP within rat DRG neurons during acute incisional pain. Brain Research. 2008;1209(C):57–64. doi: 10.1016/j.brainres.2008.02.096. [DOI] [PubMed] [Google Scholar]
  • 90.Choi JS, Hudmon A, Waxman SG, Dib-Hajj SD. Calmodulin regulates current density and frequency-dependent inhibition of sodium channel Nav1.8 in DRG neurons. Journal of Neurophysiology. 2006;96(1):97–108. doi: 10.1152/jn.00854.2005. [DOI] [PubMed] [Google Scholar]
  • 91.Villarreal CF, Sachs D, Funez MI, Parada CA, de Queiroz Cunha F, Ferreira SH. The peripheral pro-nociceptive state induced by repetitive inflammatory stimuli involves continuous activation of protein kinase A and protein kinase C epsilon and its NaV1.8 sodium channel functional regulation in the primary sensory neuron. Biochemical Pharmacology. 2009;77(5):867–877. doi: 10.1016/j.bcp.2008.11.015. [DOI] [PubMed] [Google Scholar]
  • 92.Li P, Zhu S. Molecular design of new sodium channel blockers. Biochemical and Biophysical Research Communications. 2011;414(2):321–325. doi: 10.1016/j.bbrc.2011.09.052. [DOI] [PubMed] [Google Scholar]
  • 93.Gale JD, Houghton LA. Alpha 2 delta (α2δ) ligands, gabapentin and pregabalin: what is the evidence for potential use of these ligands in irritable bowel syndrome. Frontiers in Pharmacology. 2011;2:p. 28. doi: 10.3389/fphar.2011.00028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Taylor CP, Garrido R. Immunostaining of rat brain, spinal cord, sensory neurons and skeletal muscle for calcium channel alpha2-delta (α2-δ) type 1 protein. Neuroscience. 2008;155(2):510–521. doi: 10.1016/j.neuroscience.2008.05.053. [DOI] [PubMed] [Google Scholar]
  • 95.Cole RL, Lechner SM, Williams ME, et al. Differential distribution of voltage-gated calcium channel alpha-2 delta (α2δ) subunit mRNA-containing cells in the rat central nervous system and the dorsal root ganglia. Journal of Comparative Neurology. 2005;491(3):246–269. doi: 10.1002/cne.20693. [DOI] [PubMed] [Google Scholar]
  • 96.Gong HC, Hang J, Kohler W, Li L, Su TZ. Tissue-specific expression and gabapentin-binding properties of calcium channel α2δ subunit subtypes. Journal of Membrane Biology. 2001;184(1):35–43. doi: 10.1007/s00232-001-0072-7. [DOI] [PubMed] [Google Scholar]
  • 97.Hendrich J, Van Minh AT, Heblich F, et al. Pharmacological disruption of calcium channel trafficking by the α2δ ligand gabapentin. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(9):3628–3633. doi: 10.1073/pnas.0708930105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Mich PM, Horne WA. Alternative splicing of the Ca2+ channel β4 subunit confers specificity for gabapentin inhibition of Cav2.1 trafficking. Molecular Pharmacology. 2008;74(3):904–912. doi: 10.1124/mol.108.045153. [DOI] [PubMed] [Google Scholar]
  • 99.Dooley DJ, Donovan CM, Pugsley TA. Stimulus-dependent modulation of 3H norepinephrine release from rat neocortical slices by gabapentin and pregabalin. Journal of Pharmacology and Experimental Therapeutics. 2000;295(3):1086–1093. [PubMed] [Google Scholar]
  • 100.Patel MK, Gonzalez MI, Bramwell S, Pinnock RD, Lee K. Gabapentin inhibits excitatory synaptic transmission in the hyperalgesic spinal cord. British Journal of Pharmacology. 2000;130(8):1731–1734. doi: 10.1038/sj.bjp.0703530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Field MJ, Cox PJ, Stott E, et al. Identification of the α2-δ-1 subunit of voltage-calcium calcium channels as a molecular target for pain mediating the analgesic actions of pregabalin. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(46):17537–17542. doi: 10.1073/pnas.0409066103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Field MJ, Holloman EF, McCleary S, Hughes J, Singh L. Evaluation of gabapentin and S-(+)-3-isobutylgaba in a rat model of postoperative pain. Journal of Pharmacology and Experimental Therapeutics. 1997;282(3):1242–1246. [PubMed] [Google Scholar]
  • 103.Field MJ, Oles RJ, Lewis AS, McCleary S, Hughes J, Singh L. Gabapentin (neurontin) and S-(+)-3-isobutylgaba represent a novel class of selective antihyperalgesic agents. British Journal of Pharmacology. 1997;121(8):1513–1522. doi: 10.1038/sj.bjp.0701320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Million M, Wang L, Adelson DW, Roman F, Diop L, Taché Y. Pregabalin decreases visceral pain and prevents spinal neuronal activation in rats. Gut. 2007;56(10):1482–1484. doi: 10.1136/gut.2007.129304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Lee KJ, Kim JH, Cho SW. Gabapentin reduces rectal mechanosensitivity and increases rectal compliance in patients with diarrhoea-predominant irritable bowel syndrome. Alimentary Pharmacology and Therapeutics. 2005;22(10):981–988. doi: 10.1111/j.1365-2036.2005.02685.x. [DOI] [PubMed] [Google Scholar]
  • 106.Houghton LA, Fell C, Whorwell PJ, Jones I, Sudworth DP, Gale JD. Effect of a second-generation α2δ ligand (pregabalin) on visceral sensation in hypersensitive patients with irritable bowel syndrome. Gut. 2007;56(9):1218–1225. doi: 10.1136/gut.2006.110858. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Diop L, Raymond F, Fargeau H, Petoux F, Chovet M, Doherty AM. Pregabalin (Cl-1008) inhibits the trinitrobenzene sulfonic acid-induced chronic colonic allodynia in the rat. Journal of Pharmacology and Experimental Therapeutics. 2002;302(3):1013–1022. doi: 10.1124/jpet.302.3.1013. [DOI] [PubMed] [Google Scholar]
  • 108.Feng Y, Cui M, Willis WD. Gabapentin markedly reduces acetic acid-induced visceral nociception. Anesthesiology. 2003;98(3):729–733. doi: 10.1097/00000542-200303000-00023. [DOI] [PubMed] [Google Scholar]
  • 109.Stepanovíc-Petrovíc RM, Tomíc MA, Vučkovíc SM, et al. The antinociceptive effects of anticonvulsants in a mouse visceral pain model. Anesthesia and Analgesia. 2008;106(6):1897–1903. doi: 10.1213/ane.0b013e318172b993. [DOI] [PubMed] [Google Scholar]
  • 110.Reihanikermani H, Ansari M, Soltani A, Meymandi MS. Amitriptyline pharmacokinetics in experimental spinal cord injury in the rabbit. Indian Journal of Pharmaceutical Sciences. 2008;70(6):782–785. doi: 10.4103/0250-474X.49122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Ravnefjord A, Brusberg M, Larsson H, Lindström E, Martínez V. Effects of pregabalin on visceral pain responses and colonic compliance in rats. British Journal of Pharmacology. 2008;155(3):407–416. doi: 10.1038/bjp.2008.259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Marger F, Gelot A, Alloui A, et al. T-type calcium channels contribute to colonic hypersensitivity in a rat model of irritable bowel syndrome. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(27):11268–11273. doi: 10.1073/pnas.1100869108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Jiang W, Kirkup AJ, Grundy D. Mast cells drive mesenteric afferent signalling during acute intestinal ischaemia. Journal of Physiology. 2011;589(15):3867–3882. doi: 10.1113/jphysiol.2011.209478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Ding YQ, Li JL, Lü BZ, Wang D, Miao-Li Zhang, Li JS. Co-localization of μ-opioid receptor-like immunoreactivity with substance P-LI, calcitonin gene-related peptide-LI and nitric oxide synthase- LI in vagal and glossopharyngeal afferent neurons of the rat. Brain Research. 1998;792(1):149–153. doi: 10.1016/s0006-8993(98)00205-4. [DOI] [PubMed] [Google Scholar]
  • 115.Aicher SA, Goldberg A, Sharma S, Pickel VM. μ-opioid receptors present in vagal afferents and their dendritic targets in the medial nucleus tractus solitarius. Journal of Comparative Neurology. 2000;422(2):181–190. doi: 10.1002/(sici)1096-9861(20000626)422:2<181::aid-cne3>3.0.co;2-g. [DOI] [PubMed] [Google Scholar]
  • 116.Rau KK, Caudle RM, Cooper BY, Johnson RD. Diverse immunocytochemical expression of opioid receptors in electrophysiologically defined cells of rat dorsal root ganglia. Journal of Chemical Neuroanatomy. 2005;29(4):255–264. doi: 10.1016/j.jchemneu.2005.02.002. [DOI] [PubMed] [Google Scholar]
  • 117.Bueno L, Fioramonti J. Action of opiates on gastrointestinal function. Bailliere’s Clinical Gastroenterology. 1988;2(1):123–139. doi: 10.1016/0950-3528(88)90024-3. [DOI] [PubMed] [Google Scholar]
  • 118.Ahlbeck K. Opioids: a two-faced Janus. Current Medical Research and Opinion. 2011;27(2):439–448. doi: 10.1185/03007995.2010.545379. [DOI] [PubMed] [Google Scholar]
  • 119.Rimoy GH, Wright DM, Bhaskar NK, Rubin PC. The cardiovascular and central nervous system effects in the human of U-62066E. A selective opioid receptor agonist. European Journal of Clinical Pharmacology. 1994;46(3):203–207. doi: 10.1007/BF00192549. [DOI] [PubMed] [Google Scholar]
  • 120.Kumor KM, Haertzen CA, Johnson RE. Human psychopharmacology of ketocyclazocine as compared with cyclazocine, morphine and placebo. Journal of Pharmacology and Experimental Therapeutics. 1986;238(3):960–968. [PubMed] [Google Scholar]
  • 121.Grundy D, Booth CE, Winchester W, Hicks GA. Peripheral opiate action on afferent fibres supplying the rat intestine. Neurogastroenterology and Motility. 2004;16(supplement 2):29–37. doi: 10.1111/j.1743-3150.2004.00557.x. [DOI] [PubMed] [Google Scholar]
  • 122.Eastwood C, Grundy D. Opioid-receptor-mediated excitation of rat mesenteric afferent fibres supplying the rat jejunum. Neurogastroenterology and Motility. 2000;12(6):517–522. doi: 10.1046/j.1365-2982.2000.00226.x. [DOI] [PubMed] [Google Scholar]
  • 123.Gray AC, White PJ, Coupar IM. Characterisation of opioid receptors involved in modulating circular and longitudinal muscle contraction in the rat ileum. British Journal of Pharmacology. 2005;144(5):687–694. doi: 10.1038/sj.bjp.0706107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Lian B, Vera-Portocarrero L, King T, Ossipov MH, Porreca F. Opioid-induced latent sensitization in a model of non-inflammatory viscerosomatic hypersensitivity. Brain Research C. 2010;1358:64–70. doi: 10.1016/j.brainres.2010.08.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Su X, Sengupta JN, Gebhart GF. Effects of kappa opioid receptor-selective agonists on responses of pelvic nerve afferents to noxious colorectal distension. Journal of Neurophysiology. 1997;78(2):1003–1012. doi: 10.1152/jn.1997.78.2.1003. [DOI] [PubMed] [Google Scholar]
  • 126.Burton MB, Gebhart GF. Effects of kappa-opioid receptor agonists on responses to colorectal distension in rats with and without acute colonic inflammation. Journal of Pharmacology and Experimental Therapeutics. 1998;285(2):707–715. [PubMed] [Google Scholar]
  • 127.Herrero JF, Headley PM. The effects of sham and full spinalization on the systemic potency of μ- and κ-opioids on spinal nociceptive reflexes in rats. British Journal of Pharmacology. 1991;104(1):166–170. doi: 10.1111/j.1476-5381.1991.tb12402.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Rivière PJ. Peripheral kappa-opioid agonists for visceral pain. British Journal of Pharmacology. 2004;141(8):1331–1334. doi: 10.1038/sj.bjp.0705763. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Kivell B, Prisinzano TE. Kappa opioids and the modulation of pain. Psychopharmacology. 2010;209(2):109–119. doi: 10.1007/s00213-010-1819-6. [DOI] [PubMed] [Google Scholar]
  • 130.Gebhart GF, Su X, Joshi S, Ozaki N, Sengupta JN. Peripheral opioid modulation of visceral pain. Annals of the New York Academy of Sciences. 2000;909:41–50. doi: 10.1111/j.1749-6632.2000.tb06675.x. [DOI] [PubMed] [Google Scholar]
  • 131.Ackley MA, Hurley RW, Virnich DE, Hammond DL. A cellular mechanism for the antinociceptive effect of a kappa opioid receptor agonist. Pain. 2001;91(3):377–388. doi: 10.1016/S0304-3959(00)00464-4. [DOI] [PubMed] [Google Scholar]
  • 132.Sengupta JN, Snider A, Su X, Gebhart GF. Effects of kappa opioids in the inflamed rat colon. Pain. 1999;79(2-3):175–185. doi: 10.1016/s0304-3959(98)00175-4. [DOI] [PubMed] [Google Scholar]
  • 133.Simonin F, Valverde O, Smadja C, et al. Disruption of the κ-opioid receptor gene in mice enhances sensitivity to chemical visceral pain, impairs pharmacological actions of the selective κ-agonist U-50,488H and attenuates morphine withdrawal. The EMBO Journal. 1998;17(4):886–897. doi: 10.1093/emboj/17.4.886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Arendt-Nielsen L, Olesen AE, Staahl C, et al. Analgesic efficacy of peripheral κ-opioid receptor agonist CR665 Compared to oxycodone in a multi-modal, multi-tissue experimental human pain model: selective effect on visceral pain. Anesthesiology. 2009;111(3):616–624. doi: 10.1097/ALN.0b013e3181af6356. [DOI] [PubMed] [Google Scholar]
  • 135.Delgado-Aros S, Chial HJ, Cremonini F, et al. Effects of asimadoline, a κ-opioid agonist, on satiation and postprandial symptoms in health. Alimentary Pharmacology and Therapeutics. 2003;18(5):507–514. doi: 10.1046/j.1365-2036.2003.01670.x. [DOI] [PubMed] [Google Scholar]
  • 136.Camilleri M. Novel pharmacology: asimadoline, a κ-opioid agonist, and visceral sensation. Neurogastroenterology and Motility. 2008;20(9):971–979. doi: 10.1111/j.1365-2982.2008.01183.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Szarka LA, Camilleri M, Burton D, et al. Efficacy of on-demand asimadoline, a peripheral κ-opioid agonist, in females with irritable bowel syndrome. Clinical Gastroenterology and Hepatology. 2007;5(11):1268–1275. doi: 10.1016/j.cgh.2007.07.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Delvaux M. Pharmacology and clinical experience with fedotozine. Expert Opinion on Investigational Drugs. 2001;10(1):97–110. doi: 10.1517/13543784.10.1.97. [DOI] [PubMed] [Google Scholar]
  • 139.Bonaz B, Riviere PJ, Sinniger V, et al. Fedotozine, a kappa-opioid agonist, prevents spinal and supra-spinal Fos expression induced by a noxious visceral stimulus in the rat. Neurogastroenterology and Motility. 2000;12(2):135–147. doi: 10.1046/j.1365-2982.2000.00188.x. [DOI] [PubMed] [Google Scholar]
  • 140.Jones KL, Wishart JM, Berry MK, Abitbol J-L, Horowitz M. Effects of fedotozine on gastric emptying and upper gastrointestinal symptoms in diabetic gastroparesis. Alimentary Pharmacology and Therapeutics. 2000;14(7):937–943. doi: 10.1046/j.1365-2036.2000.00790.x. [DOI] [PubMed] [Google Scholar]
  • 141.Eisenach JC, Carpenter R, Curry R. Analgesia from a peripherally active κ-opioid receptor agonist in patients with chronic pancreatitis. Pain. 2003;101(1-2):89–95. doi: 10.1016/s0304-3959(02)00259-2. [DOI] [PubMed] [Google Scholar]
  • 142.Vanderah TW, Largent-Milnes T, Lai J, et al. Novel d-amino acid tetrapeptides produce potent antinociception by selectively acting at peripheral κ-opioid receptors, European Journal of Pharmacology. 2008;583(1):62–72. doi: 10.1016/j.ejphar.2008.01.011. [DOI] [PubMed] [Google Scholar]
  • 143.Al-Khrasani M, Lackó E, Riba P, et al. The central versus peripheral antinociceptive effects of μ-opioid receptor agonists in the new model of rat visceral pain. Brain Research Bulletin. 2012;87(2-3):238–243. doi: 10.1016/j.brainresbull.2011.10.018. [DOI] [PubMed] [Google Scholar]
  • 144.Greenwood-Van Meerveld B, Gardner CJ, Little PJ, Hicks GA, Dehaven-Hudkins DL. Dehaven-Hudkins DL. Preclinical studies of opioids and opioid antagonists on gastrointestinal function. Neurogastroenterology & Motility. 2004;16(supplement 2):46–53. doi: 10.1111/j.1743-3150.2004.00555.x. [DOI] [PubMed] [Google Scholar]
  • 145.Sandner-Kiesling A, Eisenach JC. Pharmacology of opioid inhibition to noxious uterine cervical distension. Anesthesiology. 2002;97(4):966–971. doi: 10.1097/00000542-200210000-00032. [DOI] [PubMed] [Google Scholar]
  • 146.Larsson M, Arvidsson S, Ekman C, Bayati A. A model for chronic quantitative studies of colorectal sensitivity using balloon distension in conscious mice—effects of opioid receptor agonists. Neurogastroenterology and Motility. 2003;15(4):371–381. doi: 10.1046/j.1365-2982.2003.00418.x. [DOI] [PubMed] [Google Scholar]
  • 147.Sengupta JN, Su X, Gebhart GF. κ, but not μ or δ, opioids attenuate responses to distention of afferent fibers innervating the rat colon. Gastroenterology. 1996;111(4):968–980. doi: 10.1016/s0016-5085(96)70064-1. [DOI] [PubMed] [Google Scholar]
  • 148.Larsson MH, Bayati A, Lindström E, Larsson H. Involvement of κ-opioid receptors in visceral nociception in mice. Neurogastroenterology and Motility. 2008;20(10):1157–1164. doi: 10.1111/j.1365-2982.2008.01161.x. [DOI] [PubMed] [Google Scholar]
  • 149.Briggs SL, Rech RH. Antinociceptive interactions of micro- and kappa-opioid agonists in the colorectal distension assay in rats. Pharmacology Biochemistry and Behavior. 2009;92(2):343–350. doi: 10.1016/j.pbb.2008.12.023. [DOI] [PubMed] [Google Scholar]
  • 150.Staahl C, Christrup LL, Andersen SD, Arendt-Nielsen L, Drewes AM. A comparative study of oxycodone and morphine in a multi-modal, tissue-differentiated experimental pain model. Pain. 2006;123(1-2):28–36. doi: 10.1016/j.pain.2006.02.006. [DOI] [PubMed] [Google Scholar]
  • 151.Partosoedarso ER, Young RL, Blackshaw LA. GABAB receptors on vagal afferent pathways: peripheral and central inhibition. American Journal of Physiology Gastrointestinal and Liver Physiology. 2001;280(4):G658–G668. doi: 10.1152/ajpgi.2001.280.4.G658. [DOI] [PubMed] [Google Scholar]
  • 152.Smid SD, Young RL, Cooper NJ, Blackshaw LA. GABABR expressed on vagal afferent neurones inhibit gastric mechanosensitivity in ferret proximal stomac. American Journal of Physiology Gastrointestinal and Liver Physiology. 2001;281(6):G1494–G1501. doi: 10.1152/ajpgi.2001.281.6.G1494. [DOI] [PubMed] [Google Scholar]
  • 153.Bussières N, El Manira A. GABAB receptor activation inhibits N- and P/Q-type calcium channels in cultured lamprey sensory neurons. Brain Research. 1999;847(2):175–185. doi: 10.1016/s0006-8993(99)02002-8. [DOI] [PubMed] [Google Scholar]
  • 154.Filippov AK, Couve A, Pangalos MN, Walsh FS, Brown DA, Moss SJ. Heteromeric assembly of GABABR1 and GABABR2 receptor subunits inhibits Ca2+ current in sympathetic neurons. Journal of Neuroscience. 2000;20(8):2867–2874. doi: 10.1523/JNEUROSCI.20-08-02867.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Wang X, Lambert NA. GABAB receptors couple to potassium and calcium channels on identified lateral perforant pathway projection neurons. Journal of Neurophysiology. 2000;83(2):1073–1078. doi: 10.1152/jn.2000.83.2.1073. [DOI] [PubMed] [Google Scholar]
  • 156.Sengupta JN, Medda BK, Shaker R. Effect of GABAB receptor agonist on distension-sensitive pelvic nerve afferent fibers innervating rat colon. American Journal of Physiology Gastrointestinal and Liver Physiology. 2002;283(6):G1343–G1351. doi: 10.1152/ajpgi.00124.2002. [DOI] [PubMed] [Google Scholar]
  • 157.Brusberg M, Ravnefjord A, Martinsson R, Larsson H, Martinez V, Lindström E. The GABAB receptor agonist, baclofen, and the positive allosteric modulator, CGP7930, inhibit visceral pain-related responses to colorectal distension in rats. Neuropharmacology. 2009;56:362–367. doi: 10.1016/j.neuropharm.2008.09.006. [DOI] [PubMed] [Google Scholar]
  • 158.Hara K, Saito Y, Kirihara Y, Yamada Y, Sakura S, Kosaka Y. The interaction of antinociceptive effects of morphine and GABA receptor agonists within the rat spinal cord. Anesthesia and Analgesia. 1999;89(2):422–427. doi: 10.1097/00000539-199908000-00032. [DOI] [PubMed] [Google Scholar]
  • 159.Lindström E, Brusberg M, Ravnefjord A, et al. Oral baclofen reduces visceral pain-related pseudo-affective responses to colorectal distension in rats: relation between plasma exposure and efficacy. Scandinavian Journal of Gastroenterology. 2011;46(6):652–662. doi: 10.3109/00365521.2011.560677. [DOI] [PubMed] [Google Scholar]
  • 160.Birder LA, de Groat WC. Contribution of C-fiber afferent nerves and autonomic pathways in the urinary bladder to spinal c-fos expression induced by bladder irritation. Somatosensory and Motor Research. 1998;15(1):5–12. doi: 10.1080/08990229870916. [DOI] [PubMed] [Google Scholar]
  • 161.Lu Y, Westlund KN. Effects of baclofen on colon inflammation-induced Fos, CGRP and SP expression in spinal cord and brainstem. Brain Research. 2001;889(1-2):118–130. doi: 10.1016/s0006-8993(00)03124-3. [DOI] [PubMed] [Google Scholar]
  • 162.Riley RC, Trafton JA, Chi SI, Basbaum AI. Presynaptic regulation of spinal cord tachykinin signaling via GABAB but not GABAA receptor activation. Neuroscience. 2001;103(3):725–737. doi: 10.1016/s0306-4522(00)00571-6. [DOI] [PubMed] [Google Scholar]
  • 163.Wang Y, Wu J, Lin Q, Nauta HJ, Yue Y, Fang L. Effects of general anesthetics on visceral pain transmission in the spinal cord. Molecular Pain. 2008;4, article 50 doi: 10.1186/1744-8069-4-50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Dong XP, Xu TL. The actions of propofol on γ-aminobutyric acid-A and glycine receptors in acutely dissociated spinal dorsal horn neurons of the rat. Anesthesia and Analgesia. 2002;95(4):907–914. doi: 10.1097/00000539-200210000-00021. [DOI] [PubMed] [Google Scholar]
  • 165.Herman RM, D’Luzansky SC, Ippolito R. Intrathecal baclofen suppresses central pain in patients with spinal lesions. A pilot study. Clinical Journal of Pain. 1992;8(4):338–345. [PubMed] [Google Scholar]
  • 166.Slonimski M, Abram SE, Zuniga RE. Intrathecal baclofen in pain management. Regional Anesthesia and Pain Medicine. 2004;29(3):269–276. doi: 10.1016/j.rapm.2004.01.005. [DOI] [PubMed] [Google Scholar]
  • 167.Harmer JP, Larson BS. Pain relief from baclofen analgesia in a neuropathic pain patient who failed opioid and pharmacotherapy: case report. Journal of Pain and Palliative Care Pharmacotherapy. 2002;16(3):61–64. doi: 10.1080/j354v16n03_06. [DOI] [PubMed] [Google Scholar]
  • 168.Zuniga RE, Perera S, Abram SE. Intrathecal baclofen: a useful agent in the treatment of well-established complex regional pain syndrome. Regional Anesthesia and Pain Medicine. 2002;27(1):90–93. doi: 10.1053/rapm.2002.29244. [DOI] [PubMed] [Google Scholar]
  • 169.Zuniga RE, Schlicht CR, Abram SE. Intrathecal baclofen is analgesic in patients with chronic pain. Anesthesiology. 2000;92(3):876–880. doi: 10.1097/00000542-200003000-00037. [DOI] [PubMed] [Google Scholar]
  • 170.Mayer EA, Gebhart GF. Basic and clinical aspects of visceral hyperalgesia. Gastroenterology. 1994;107(1):271–293. doi: 10.1016/0016-5085(94)90086-8. [DOI] [PubMed] [Google Scholar]
  • 171.Michaelis EK. Molecular biology of glutamate receptors in the central nervous system and their role in excitotoxicity, oxidative stress and aging. Progress in Neurobiology. 1998;54(4):369–415. doi: 10.1016/s0301-0082(97)00055-5. [DOI] [PubMed] [Google Scholar]
  • 172.Randic M, Jiang MC, Rusin KI, Cerne R, Kolaj M. Interactions between excitatory amino acids and tachykinins and long-term changes of synaptic responses in the rat spinal dorsal horn. Regulatory Peptides. 1993;46(1-2):418–420. doi: 10.1016/0167-0115(93)90106-i. [DOI] [PubMed] [Google Scholar]
  • 173.McRoberts JA, Coutinho SV, Marvizón JC, et al. Role of peripheral N-methyl-D-aspartate (NMDA) receptors in visceral nociception in rats. Gastroenterology. 2001;120(7):1737–1748. doi: 10.1053/gast.2001.24848. [DOI] [PubMed] [Google Scholar]
  • 174.Liu H, Mantyh PW, Basbaum AI. NMDA-receptor regulation of substance P release from primary afferent nociceptors. Nature. 1997;386(6626):721–724. doi: 10.1038/386721a0. [DOI] [PubMed] [Google Scholar]
  • 175.Wang H, Pickel VM. Presence of NMDA-type glutamate receptors in cingulate corticostriatal terminals and their postsynaptic targets. Synapse. 2000;35(4):300–310. doi: 10.1002/(SICI)1098-2396(20000315)35:4<300::AID-SYN8>3.0.CO;2-C. [DOI] [PubMed] [Google Scholar]
  • 176.Wang H, Stradtman GG, III, Wang XJ, Gao WJ. A specialized NMDA receptor function in layer 5 recurrent microcircuitry of the adult rat prefrontal cortex. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(43):16791–16796. doi: 10.1073/pnas.0804318105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Wang H, Gracy KN, Pickel VM. μ-Opioid and NMDA-type glutamate receptors are often colocalized in spiny neurons within patches of the caudate-putamen nucleus. Journal of Comparative Neurology. 1999;412(1):132–146. doi: 10.1002/(sici)1096-9861(19990913)412:1<132::aid-cne10>3.0.co;2-b. [DOI] [PubMed] [Google Scholar]
  • 178.Chinnery RM, Shaw PJ, Ince PG, Johnson M. Autoradiographic distribution of binding sites for the non-NMDA receptor antagonist [3H]CNQX in human motor cortex, brainstem and spinal cord. Brain Research. 1993;630(1-2):75–81. doi: 10.1016/0006-8993(93)90644-3. [DOI] [PubMed] [Google Scholar]
  • 179.Olivar T, Laird JMA. Differential effects of N-methyl-D-receptor blockade on nociceptive somatic and visceral reflexes. Pain. 1999;79(1):67–73. doi: 10.1016/S0304-3959(98)00152-3. [DOI] [PubMed] [Google Scholar]
  • 180.Kolhekar R, Gebhart GF. Modulation of spinal visceral nociceptive transmission by NMDA receptor activation in the rat. Journal of Neurophysiology. 1996;75(6):2344–2353. doi: 10.1152/jn.1996.75.6.2344. [DOI] [PubMed] [Google Scholar]
  • 181.Ji T, Traub RJ. Spinal NMDA receptors contribute to neuronal processing of acute noxious and nonnoxious colorectal stimulation in the rat. Journal of Neurophysiology. 2001;86(4):1783–1791. doi: 10.1152/jn.2001.86.4.1783. [DOI] [PubMed] [Google Scholar]
  • 182.Anand P, Aziz Q, Willert R, van Oudenhove L. Peripheral and central mechanisms of visceral sensitization in man. Neurogastroenterology and Motility. 2007;19(supplement 1):29–46. doi: 10.1111/j.1365-2982.2006.00873.x. [DOI] [PubMed] [Google Scholar]
  • 183.Yiangou Y, Facer P, Baecker PA, et al. ATP-gated ion channel P2X3 is increased in human inflammatory bowel disease. Neurogastroenterology and Motility. 2001;13(4):365–369. doi: 10.1046/j.1365-2982.2001.00276.x. [DOI] [PubMed] [Google Scholar]
  • 184.Bueno L, Fioramonti J, Garcia-Villar R. Pathobiology of visceral pain: molecular mechanisms and therapeutic implications. III. Visceral afferent pathways: a source of new therapeutic targets for abdominal pain. American Journal of Physiology. 2000;278(5):G670–G676. doi: 10.1152/ajpgi.2000.278.5.G670. [DOI] [PubMed] [Google Scholar]
  • 185.Gold MS, Gebhart GF. Nociceptor sensitization in pain pathogenesis. Nature Medicine. 2010;16(11):1248–1257. doi: 10.1038/nm.2235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Shinoda M, Feng B, Gebhart GF. Peripheral and central P2X3 receptor contributions to colon mechanosensitivity and hypersensitivity in the mouse. Gastroenterology. 2009;137(6):2096–2104. doi: 10.1053/j.gastro.2009.06.048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Keating C, Pelegrin P, Martínez CM, Grundy D. P2X7 receptor-dependent intestinal afferent hypersensitivity in a mouse model of postinfectious irritable bowel syndrome. Journal of Immunology. 2011;187(3):1467–1474. doi: 10.4049/jimmunol.1100423. [DOI] [PubMed] [Google Scholar]
  • 188.Booth CE, Kirkup AJ, Hicks GA, Humphrey PP, Grundy D. Somatostatin sst2 receptor-mediated inhibition of mesenteric afferent nerves of the jejunum in the anesthetized rat. Gastroenterology. 2001;121(2):358–369. doi: 10.1053/gast.2001.26335. [DOI] [PubMed] [Google Scholar]
  • 189.Schwetz I, Naliboff B, Munakata J, et al. Anti-hyperalgesic effect of octreotide in patients with irritable bowel syndrome. Alimentary Pharmacology and Therapeutics. 2004;19(1):123–131. doi: 10.1111/j.1365-2036.2004.01774.x. [DOI] [PubMed] [Google Scholar]
  • 190.Klooker TK, Kuiken SD, Lei A, Boeckxstaens GE. Effect of long-term treatment with octreotide on rectal sensitivity in patients with non-constipated irritable bowel syndrome. Alimentary Pharmacology and Therapeutics. 2007;26(4):605–615. doi: 10.1111/j.1365-2036.2007.03398.x. [DOI] [PubMed] [Google Scholar]
  • 191.O'Donnell LJ, Watson AJ, Cameron D, Farthing MJG. Effect of octreotide on mouth-to-caecum transit time in healthy subjects and in the irritable bowel syndrome. Alimentary Pharmacology and Therapeutics. 1990;4(2):177–182. doi: 10.1111/j.1365-2036.1990.tb00463.x. [DOI] [PubMed] [Google Scholar]
  • 192.Rothenberg M, Pai R, Stuart K. Successful use of octreotide to treat ménétrier’s disease: a rare cause of abdominal pain, weight loss, edema, and hypoalbuminemia. Digestive Diseases and Sciences. 2009;54(7):1403–1407. doi: 10.1007/s10620-009-0754-z. [DOI] [PubMed] [Google Scholar]
  • 193.De Conno F, Saita L, Ripamonti C, Ventafridda V. Subcutaneous octreotide in the treatment of pain in advanced cancer patients. Journal of Pain and Symptom Management. 1994;9(1):34–38. doi: 10.1016/0885-3924(94)90144-9. [DOI] [PubMed] [Google Scholar]
  • 194.Befon S, Mystakidou K, Lyra M, Tubanakis N, Vlahos L. Continuous subcutaneous octreotide in gastrointestinal cancer patients: pain control and β-endorphin levels. Anticancer Research C. 2000;20(5):4039–4046. [PubMed] [Google Scholar]
  • 195.Ducrotte P, Maillot C, Leroi AM, Lalaude O, Colin R, Denis P. Octreotide in refractory functional epigastric pain with nutritional impairment—an open study. Alimentary Pharmacology and Therapeutics. 1999;13(7):969–975. doi: 10.1046/j.1365-2036.1999.00541.x. [DOI] [PubMed] [Google Scholar]
  • 196.Malfertheiner P, Mayer D, Büchler M, Domínguez-Muñoz JE, Schiefer B, Ditschuneit H. Treatment of pain in chronic pancreatitis by inhibition of pancreatic secretion with octreotide. Gut. 1995;36(3):450–454. doi: 10.1136/gut.36.3.450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Maubach KA, Grundy D. The role of prostaglandins in the bradykinin-induced activation of serosal afferents of the rat jejunum in vitro. Journal of Physiology. 1999;515, part 1:277–285. doi: 10.1111/j.1469-7793.1999.277ad.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Rang HP, Bevan S, Dray A. Chemical activation of nociceptive peripheral neurones. British Medical Bulletin. 1991;47(3):534–548. doi: 10.1093/oxfordjournals.bmb.a072491. [DOI] [PubMed] [Google Scholar]
  • 199.Mueller MH, Gong Q, Kelber O, et al. A novel herbal preparation desensitizes mesenteric afferents to bradykinin in the rat small intestine. Neurogastroenterology and Motility. 2009;21(4):467–476. doi: 10.1111/j.1365-2982.2008.01232.x. [DOI] [PubMed] [Google Scholar]
  • 200.Müller MH, Liu CY, Glatzle J, et al. STW 5 (Iberogast) reduces afferent sensitivity in the rat small intestine. Phytomedicine. 2006;13(supplement 1):100–106. doi: 10.1016/j.phymed.2006.03.014. [DOI] [PubMed] [Google Scholar]
  • 201.Liu CY, Müller MH, Glatzle J, et al. The herbal preparation STW 5 (Iberogast) desensitizes intestinal afferents in the rat small intestine. Neurogastroenterology and Motility. 2004;16(6):759–764. doi: 10.1111/j.1365-2982.2004.00576.x. [DOI] [PubMed] [Google Scholar]
  • 202.Melzer J, Rösch W, Reichling J, Brignoli R, Saller R. Meta-analysis: phytotherapy of functional dyspepsia with the herbal drug preparation STW 5 (Iberogast) Alimentary Pharmacology and Therapeutics. 2004;20(11-12):1279–1287. doi: 10.1111/j.1365-2036.2004.02275.x. [DOI] [PubMed] [Google Scholar]
  • 203.Sibaev A, Yuece B, Kelber O, et al. STW 5 (Iberogast) and its individual herbal components modulate intestinal electrophysiology of mice. Phytomedicine. 2006;13(supplement 5):80–89. doi: 10.1016/j.phymed.2006.03.015. [DOI] [PubMed] [Google Scholar]
  • 204.Ammon HP, Kelber O, Okpanyi SN. Spasmolytic and tonic effect of Iberogast (STW 5) in intestinal smooth muscle. Phytomedicine. 2006;13(1):67–74. doi: 10.1016/j.phymed.2006.08.004. [DOI] [PubMed] [Google Scholar]
  • 205.Bullock BP, Habener JF. Phosphorylation of the cAMP response element binding protein CREB by cAMP-dependent protein kinase A and glycogen synthase kinase-3 alters DNA- binding affinity, conformation, and increases net charge. Biochemistry. 1998;37(11):3795–3809. doi: 10.1021/bi970982t. [DOI] [PubMed] [Google Scholar]
  • 206.Crump FT, Dillman KS, Craig AM. cAMP-dependent protein kinase mediates activity-regulated synaptic targeting of NMDA receptors. Journal of Neuroscience. 2001;21(14):5079–5088. doi: 10.1523/JNEUROSCI.21-14-05079.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Hughes AS, Averill S, King VR, Molander C, Shortland PJ. Neurochemical characterization of neuronal populations expressing protein kinase C gamma isoform in the spinal cord and gracile nucleus of the rat. Neuroscience. 2008;153(2):507–517. doi: 10.1016/j.neuroscience.2008.01.082. [DOI] [PubMed] [Google Scholar]
  • 208.Zhang Y, Gong K, Zhou W, et al. Involvement of subtypes γ and ε of protein kinase C in colon pain induced by formalin injection. NeuroSignals. 2011;19(3):142–150. doi: 10.1159/000328311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Galan A, Cervero F, Laird JM. Extracellular signaling-regulated kinase-1 and -2 (ERK 1/2) mediate referred hyperalgesia in a murine model of visceral pain. Molecular Brain Research. 2003;116(1-2):126–134. doi: 10.1016/s0169-328x(03)00284-5. [DOI] [PubMed] [Google Scholar]
  • 210.Ji RR, Baba H, Brenner GJ, Woolf CJ. Nociceptive-specific activation of ERK in spinal neurons contributes to pain hypersensitivity. Nature Neuroscience. 1999;2(12):1114–1119. doi: 10.1038/16040. [DOI] [PubMed] [Google Scholar]
  • 211.Karim F, Wang CC, Gereau RW. Metabotropic glutamate receptor subtypes 1 and 5 are activators of extracellular signal-regulated kinase signaling required for inflammatory pain in mice. Journal of Neuroscience. 2001;21(11):3771–3779. doi: 10.1523/JNEUROSCI.21-11-03771.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Lai HH, Qiu CS, Crock LW, Morales ME, Ness TJ, Gereau IV RW. Activation of spinal extracellular signal-regulated kinases (ERK) 1/2 is associated with the development of visceral hyperalgesia of the bladder. Pain. 2011;152(9):2117–2124. doi: 10.1016/j.pain.2011.05.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Lichtenstein GR, Feagan BG, Cohen RD, et al. Serious infections and mortality in association with therapies for Crohn's disease: TREAT registry. Clinical Gastroenterology and Hepatology. 2006;4(5):621–630. doi: 10.1016/j.cgh.2006.03.002. [DOI] [PubMed] [Google Scholar]
  • 214.Bonner GF. Exacerbation of inflammatory bowel disease associated with use of celecoxib. American Journal of Gastroenterology. 2001;96(4):1306–1308. doi: 10.1111/j.1572-0241.2001.03730.x. [DOI] [PubMed] [Google Scholar]
  • 215.Reinisch W, Miehsler W, Dejaco C, et al. An open-label trial of the selective cyclo-oxygenase-2 inhibitor, rofecoxib, in inflammatory bowel disease-associated peripheral arthritis and arthralgia. Alimentary Pharmacology and Therapeutics. 2003;17(11):1371–1380. doi: 10.1046/j.1365-2036.2003.01596.x. [DOI] [PubMed] [Google Scholar]
  • 216.Ripamonti C, Mercadante S, Groff L, Zecca E, De Conno F, Casuccio A. Role of octreotide, scopolamine butylbromide, and hydration in symptom control of patients with inoperable bowel obstruction and nasogastric tubes: a prospective randomized trial. Journal of Pain and Symptom Management. 2000;19(1):23–34. doi: 10.1016/s0885-3924(99)00147-5. [DOI] [PubMed] [Google Scholar]
  • 217.Mercadante S, Ripamonti C, Casuccio A, Zecca E, Groff L. Comparison of octreotide and hyoscine butylbromide in controlling gastrointestinal symptoms due to malignant inoperable bowel obstruction. Supportive Care in Cancer. 2000;8(3):188–191. doi: 10.1007/s005200050283. [DOI] [PubMed] [Google Scholar]
  • 218.Guzman F, Braun C, Lim RK, Potter GD, Rodgers DW. Narcotic and non-narcotic analgesics which block visceral pain evoked by intra-arterial injection of bradykinin and other algesic agents. Arch Int Pharmacodyn Ther. 1964;149:571–588. [PubMed] [Google Scholar]
  • 219.Meymandi MS, Sepehri G. Gabapentin action and interaction on the antinociceptive effect of morphine on visceral pain in mice. European Journal of Anaesthesiology. 2008;25(2):129–134. doi: 10.1017/S0265021507001226. [DOI] [PubMed] [Google Scholar]
  • 220.Gaensler EA. Quantitative determination of the visceral pain threshold in man; characteristics of visceral pain, effect of inflammation and analgesics on the threshold, and relationship of analgesia to visceral spasm. The Journal of Clinical Investigation. 1951;30(4):406–420. doi: 10.1172/JCI102458. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Olesen AE, Staahl C, Arendt-Nielsen L, Drewes AM. Different effects of morphine and oxycodone in experimentally evoked hyperalgesia: a human translational study. British Journal of Clinical Pharmacology. 2010;70(2):189–200. doi: 10.1111/j.1365-2125.2010.03700.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Nelson K, Walsh D, Sheehan F. Cancer and chemotherapy-related upper gastrointestinal symptoms: the role of abnormal gastric motor function and its evaluation in cancer patients. Supportive Care in Cancer. 2002;10(6):455–461. doi: 10.1007/s00520-002-0340-9. [DOI] [PubMed] [Google Scholar]
  • 223.Iwata H, Tsuchiya S, Nakamura T, Yano S. Morphine leads to contraction of the ileal circular muscle via inhibition of the nitrergic pathway in mice. European Journal of Pharmacology. 2007;574(1):66–70. doi: 10.1016/j.ejphar.2007.06.029. [DOI] [PubMed] [Google Scholar]
  • 224.Iwata H, Tsuchiya S, Ueno K, Nakamura T, Yano S. Morphine-6-glucuronide induces contraction of the ileal circular muscle more potently than morphine in mice. European Journal of Pharmacology. 2008;600(1–3):130–132. doi: 10.1016/j.ejphar.2008.10.009. [DOI] [PubMed] [Google Scholar]
  • 225.Corleto VD, Severi C, Romano G, et al. Somatostatin receptor subtypes mediate contractility on human colonic smooth muscle cells. Neurogastroenterology and Motility. 2006;18(3):217–225. doi: 10.1111/j.1365-2982.2005.00752.x. [DOI] [PubMed] [Google Scholar]
  • 226.Gu Z-F, Corleto VD, Mantey SA, Coy DH, Maton PN, Jensen RT. Somatostatin receptor subtype 3 mediates the inhibitory action of somatostatin on gastric smooth muscle cells. American Journal of Physiology. 1995;268(5, part 1):G739–G748. doi: 10.1152/ajpgi.1995.268.5.G739. [DOI] [PubMed] [Google Scholar]
  • 227.Erdek MA, Halpert DE, Fernández MG, Cohen SP. Assessment of celiac plexus block and neurolysis outcomes and technique in the management of refractory visceral cancer pain. Pain Medicine. 2010;11(1):92–100. doi: 10.1111/j.1526-4637.2009.00756.x. [DOI] [PubMed] [Google Scholar]
  • 228.Nair RT, vanSonnenberg E, Shankar S, et al. Visceral and soft-tissue tumors: radiofrequency and alcohol ablation for pain relief— initial experience. Radiology. 2008;248(3):1067–1076. doi: 10.1148/radiol.2483061883. [DOI] [PubMed] [Google Scholar]
  • 229.Kapural L, Deer T, Yakovlev A, et al. Technical aspects of spinal cord stimulation for managing chronic visceral abdominal pain: the results from the national survey. Pain Medicine. 2010;11(5):685–691. doi: 10.1111/j.1526-4637.2010.00806.x. [DOI] [PubMed] [Google Scholar]
  • 230.Caraceni A, Portenoy RK. Pain management in patients with pancreatic carcinoma. Cancer. 1996;78(3, supplement):639–653. doi: 10.1002/(SICI)1097-0142(19960801)78:3<639::AID-CNCR45>3.0.CO;2-X. [DOI] [PubMed] [Google Scholar]
  • 231.Sarkar S, Aziz Q, Woolf CJ, Hobson AR, Thompson DG. Contribution of central sensitisation to the development of non-cardiac chest pain. The Lancet. 2000;356(9236):1154–1159. doi: 10.1016/S0140-6736(00)02758-6. [DOI] [PubMed] [Google Scholar]
  • 232.Neumann S, Doubell TP, Leslie T, Woolf CJ. Inflammatory pain hypersensitivity mediated by phenotypic switch in myelinated primary sensory neurons. Nature. 1996;384(6607):360–364. doi: 10.1038/384360a0. [DOI] [PubMed] [Google Scholar]
  • 233.Ji RR, Woolf CJ. Neuronal plasticity and signal transduction in nociceptive neurons: Implications for the initiation and maintenance of pathological pain. Neurobiology of Disease. 2001;8(1):1–10. doi: 10.1006/nbdi.2000.0360. [DOI] [PubMed] [Google Scholar]
  • 234.Graham GG, Scott KF. Mechanism of action of paracetamol. American Journal of Therapeutics. 2005;12(1):46–55. doi: 10.1097/00045391-200501000-00008. [DOI] [PubMed] [Google Scholar]
  • 235.Graham GG, Scott KF. Mechanisms of action of paracetamol and related analgesics. Inflammopharmacology. 2003;11(4):401–413. doi: 10.1163/156856003322699573. [DOI] [PubMed] [Google Scholar]
  • 236.Pelissier T, Alloui A, Paeile C, Eschalier A. Evidence of a central antinociceptive effect of paracetamol involving spinal 5HT3 receptors. NeuroReport. 1995;6(11):1546–1548. doi: 10.1097/00001756-199507310-00020. [DOI] [PubMed] [Google Scholar]
  • 237.Pelissier T, Alloui A, Caussade F, et al. Paracetamol exerts a spinal antinociceptive effect involving an indirect interaction with 5-hydroxytryptamine3 receptors: in vivo and in vitro evidence. Journal of Pharmacology and Experimental Therapeutics. 1996;278(1):8–14. [PubMed] [Google Scholar]
  • 238.Qiu HX, Liu J, Kong H, Liu Y, Mei XG. Isobolographic analysis of the antinociceptive interactions between ketoprofen and paracetamol. European Journal of Pharmacology. 2007;557(2-3):141–146. doi: 10.1016/j.ejphar.2006.11.017. [DOI] [PubMed] [Google Scholar]
  • 239.Miranda HF, Puig MM, Prieto JC, Pinardi G. Synergism between paracetamol and nonsteroidal anti-inflammatory drugs in experimental acute pain. Pain. 2006;121(1-2):22–28. doi: 10.1016/j.pain.2005.11.012. [DOI] [PubMed] [Google Scholar]
  • 240.Miranda HF, Prieto JC, Puig MM, Pinardi G. Isobolographic analysis of multimodal analgesia in an animal model of visceral acute pain. Pharmacology Biochemistry and Behavior. 2008;88(4):481–486. doi: 10.1016/j.pbb.2007.10.005. [DOI] [PubMed] [Google Scholar]
  • 241.Ong CK, Seymour RA, Lirk P, Merry AF. Combining paracetamol (acetaminophen) with nonsteroidal antiinflammatory drugs: a qualitative systematic review of analgesic efficacy for acute postoperative pain. Anesthesia and Analgesia. 2010;110(4):1170–1179. doi: 10.1213/ANE.0b013e3181cf9281. [DOI] [PubMed] [Google Scholar]
  • 242.Mercadante S, Casuccio A, Agnello A, Pumo S, Kargar J, Garofalo S. Analgesic effects of nonsteroidal anti-inflammatory drugs in cancer pain due to somatic or visceral mechanisms. Journal of Pain and Symptom Management. 1999;17(5):351–356. doi: 10.1016/s0885-3924(98)00141-9. [DOI] [PubMed] [Google Scholar]
  • 243.McNicol E, Strassels SA, Goudas L, Lau J, Carr DB. NSAIDS or paracetamol, alone or combined with opioids, for cancer pain. Cochrane Database of Systematic Reviews. 2005;(1) doi: 10.1002/14651858.CD005180. Article ID CD005180. [DOI] [PubMed] [Google Scholar]
  • 244.Olsen JC, McGrath NA, Schwarz DG, Cutcliffe BJ, Stern JL. A double-blind randomized clinical trial evaluating the analgesic efficacy of ketorolac versus butorphanol for patients with suspected biliary colic in the emergency department. Academic Emergency Medicine. 2008;15(8):718–722. doi: 10.1111/j.1553-2712.2008.00178.x. [DOI] [PubMed] [Google Scholar]
  • 245.Morgan G. Beneficial effects of NSAIDs in the gastrointestinal tract. European Journal of Gastroenterology and Hepatology. 1999;11(4):393–400. doi: 10.1097/00042737-199904000-00006. [DOI] [PubMed] [Google Scholar]
  • 246.Meyers D, Feldstein DA. Initial treatment of biliary colic: are NSAIDs better than opiates? Wisconsin Medical Journal. 2005;104(4):9–63. [PubMed] [Google Scholar]
  • 247.NSAIDS for renal and biliary colic: intramuscular diclofenac. Drug and Therapeutics Bulletin. 1987;25(22):85–86. [PubMed] [Google Scholar]
  • 248.Kulkarni SK, Patil CS, Jain NK, Singh A. Modulatory effect of diclofenac on antispasmodic effect of pitofenone in cholinergic spasm. Indian Journal of Experimental Biology. 2004;42(6):567–569. [PubMed] [Google Scholar]
  • 249.Akriviadis EA, Hatzigavriel M, Kapnias D, Kirimlidis J, Markantas A, Garyfallos A. Treatment of biliary colic with diclofenac: a randomized, double-blind, placebo-controlled study. Gastroenterology. 1997;113(1):225–231. doi: 10.1016/s0016-5085(97)70099-4. [DOI] [PubMed] [Google Scholar]
  • 250.Al-Waili N, Saloom KY. The analgesic effect of intravenous tenoxicam in symptomatic treatment of biliary colic: a comparison with hyoscine N-butylbromide. European Journal of Medical Research. 1998;3(10):475–479. [PubMed] [Google Scholar]
  • 251.Mora Durban MJ, Extramiana Cameno J, Arrizabalaga Moreno M, et al. Flubiprofen vs dipyrone combined with hyoscine: the analgesic efficacy in renal colic. Archivos Españoles de Urología. 1995;48(9):867–873. [PubMed] [Google Scholar]
  • 252.Holdgate A, Pollock T. Nonsteroidal anti-inflammatory drugs (NSAIDs) versus opioids for acute renal colic. Cochrane Database of Systematic Reviews. 2004;18(2) doi: 10.1002/14651858.CD004137.pub3. Article ID CD004137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Hatoum OA, Miura H, Binion DG. The vascular contribution in the pathogenesis of inflammatory bowel disease. American Journal of Physiology. 2003;285(5):H1791–H1796. doi: 10.1152/ajpheart.00552.2003. [DOI] [PubMed] [Google Scholar]
  • 254.Hatoum OA, Binion DG, Otterson MF, Gutterman DD. Acquired microvascular dysfunction in inflammatory bowel disease: loss of nitric oxide-mediated vasodilation. Gastroenterology. 2003;125(1):58–69. doi: 10.1016/s0016-5085(03)00699-1. [DOI] [PubMed] [Google Scholar]
  • 255.Jensen MK, Thomsen AB, Højsted J. 10-year follow-up of chronic non-malignant pain patients: opioid use, health related quality of life and health care utilization. European Journal of Pain. 2006;10(5):423–433. doi: 10.1016/j.ejpain.2005.06.001. [DOI] [PubMed] [Google Scholar]
  • 256.Burns JW, Bruehl S. Anger management style, opioid analgesic use, and chronic pain severity: a test of the opioid-deficit hypothesis. Journal of Behavioral Medicine. 2005;28(6):555–563. doi: 10.1007/s10865-005-9020-9. [DOI] [PubMed] [Google Scholar]
  • 257.Tuteja AK, Biskupiak J, Stoddard GJ, Lipman AG. Opioid-induced bowel disorders and narcotic bowel syndrome in patients with chronic non-cancer pain. Neurogastroenterology and Motility. 2010;22(4):424–e96. doi: 10.1111/j.1365-2982.2009.01458.x. [DOI] [PubMed] [Google Scholar]
  • 258.Choung RS, Locke GR, III, Zinsmeister AR, Schleck CD, Talley NJ. Opioid bowel dysfunction and narcotic bowel syndrome: a population-based study. American Journal of Gastroenterology. 2009;104(5):1199–1204. doi: 10.1038/ajg.2009.71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Okada-Ogawa A, Porreca F, Meng ID. Sustained morphine-induced sensitization and loss of diffuse noxious inhibitory controls in dura-sensitive medullary dorsal horn neurons. Journal of Neuroscience. 2009;29(50):15828–15835. doi: 10.1523/JNEUROSCI.3623-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.Bannister K, Sikandar S, Bauer CS, Dolphin AC, Porreca F, Dickenson AH. Pregabalin suppresses spinal neuronal hyperexcitability and visceral hypersensitivity in the absence of peripheral pathophysiology. Anesthesiology. 2011;115(1):144–152. doi: 10.1097/ALN.0b013e31821f6545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Célèrier E, Laulin JP, Corcuff JB, Le Moal M, Simonnet G. Progressive enhancement of delayed hyperalgesia induced by repeated heroin administration: a sensitization process. Journal of Neuroscience. 2001;21(11):4074–4080. doi: 10.1523/JNEUROSCI.21-11-04074.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.Koppert W, Schmelz M. The impact of opioid-induced hyperalgesia for postoperative pain. Best Practice and Research. 2007;21(1):65–83. doi: 10.1016/j.bpa.2006.12.004. [DOI] [PubMed] [Google Scholar]
  • 263.Fine PG. Opioid insights: opioid-induced hyperalgesia and opioid rotation. Journal of Pain and Palliative Care Pharmacotherapy. 2004;18(3):75–79. [PubMed] [Google Scholar]
  • 264.Raehal KM, Schmid CL, Groer CE, Bohn LM. Functional selectivity at the μ-opioid receptor: implications for understanding opioid analgesia and tolerance. Pharmacological Reviews. 2011;63(4):1001–1019. doi: 10.1124/pr.111.004598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Angst MS, Clark JD. Opioid-induced hyperalgesia: a qualitative systematic review. Anesthesiology. 2006;104(3):570–587. doi: 10.1097/00000542-200603000-00025. [DOI] [PubMed] [Google Scholar]
  • 266.Liguori S, Gottardi M, Micheletto G, Bruno L. Pharmacological approach to chronic visceral pain. Focus on oxycodone controlled release: an open multicentric study. European Review for Medical and Pharmacological Sciences. 2010;14(3):185–190. [PubMed] [Google Scholar]
  • 267.Mercadante S, Tirelli W, David F, et al. Morphine versus oxycodone in pancreatic cancer pain: a randomized controlled study. Clinical Journal of Pain. 2010;26(9):794–797. doi: 10.1097/ajp.0b013e3181ecd895. [DOI] [PubMed] [Google Scholar]
  • 268.Johnston S, Wilkes GJ, Thompson JA, Ziman M, Brightwell R. Inhaled methoxyflurane and intranasal fentanyl for prehospital management of visceral pain in an Australian ambulance service. Emergency Medicine Journal. 2011;28(1):57–63. doi: 10.1136/emj.2009.078717. [DOI] [PubMed] [Google Scholar]
  • 269.Chen SR, Pan HL. Loss of TRPV1-expressing sensory neurons reduces spinal μ opioid receptors but paradoxically potentiates opioid analgesia. Journal of Neurophysiology. 2006;95(5):3086–3096. doi: 10.1152/jn.01343.2005. [DOI] [PubMed] [Google Scholar]
  • 270.Fernández-Dueñas V, Poveda R, Fernández A, Sánchez S, Planas E, Ciruela F. Fentanyl-trazodone-paracetamol triple drug combination: multimodal analgesia in a mouse model of visceral pain. Pharmacology Biochemistry and Behavior. 2011;98(3):331–336. doi: 10.1016/j.pbb.2011.01.023. [DOI] [PubMed] [Google Scholar]
  • 271.Maves TJ, Pechman PS, Meller ST, Gebhart GF. Ketorolac potentiates morphine antinociception during visceral nociception in the rat. Anesthesiology. 1994;80(5):1094–1101. doi: 10.1097/00000542-199405000-00018. [DOI] [PubMed] [Google Scholar]
  • 272.Romero A, Miranda HF, Puig MM. Analysis of the opioid-opioid combinations according to the nociceptive stimulus in mice. Pharmacological Research. 2010;61(6):511–518. doi: 10.1016/j.phrs.2010.02.011. [DOI] [PubMed] [Google Scholar]
  • 273.Ulger F, Bozkurt A, Bilge SS, et al. The antinociceptive effects of intravenous dexmedetomidine in colorectal distension-induced visceral pain in rats: the role of opioid receptors. Anesthesia and Analgesia. 2009;109(2):616–622. doi: 10.1213/ane.0b013e3181a9fae2. [DOI] [PubMed] [Google Scholar]
  • 274.Micó JA, Ardid D, Berrocoso E, Eschalier A. Antidepressants and pain. Trends in Pharmacological Sciences. 2006;27(7):348–354. doi: 10.1016/j.tips.2006.05.004. [DOI] [PubMed] [Google Scholar]

Articles from Pain Research and Treatment are provided here courtesy of Wiley

RESOURCES