Primary Hyperalgesia

2013 ◽  
pp. 3181-3182
Keyword(s):  
1997 ◽  
Vol 20 (3) ◽  
pp. 471-471 ◽  
Author(s):  
William D. Willis

Intradermal capsaicin in humans causes pain, primary hyperalgesia, and secondary mechanical hyperalgesia and allodynia. Parallel changes occur in the responses of primate spinothalamic tract cells and in rat behavior. Neurotransmitters that trigger secondary mechanical hyperalgesia and allodynia include excitatory amino acids and substance P. Secondary mechanical allodynia is actively maintained by central mechanisms. Our group has investigated mechanisms of central sensitization of nociceptive neurons by examining the responses to intradermal injection of capsaicin. These experiments are pertinent to issues raised by coderre & katz (sect. 2).


2019 ◽  
Vol 131 (2) ◽  
pp. 356-368 ◽  
Author(s):  
E. Martin ◽  
C. Narjoz ◽  
X. Decleves ◽  
L. Labat ◽  
C. Lambert ◽  
...  

Abstract Editor’s Perspective What We Already Know about This Topic What This Article Tells Us That Is New Background Central pain sensitization is often refractory to drug treatment. Dextromethorphan, an N-methyl-d-aspartate receptor antagonist, is antihyperalgesic in preclinical pain models. The hypothesis is that dextromethorphan is also antihyperalgesic in humans. Methods This randomized, double-blind, placebo-controlled, crossover study explores the antihyperalgesic effect of single and repeated 30-mg dose of oral dextromethorphan in 20 volunteers, using the freeze-injury pain model. This model leads to development of primary and secondary hyperalgesia, which develops away from the site of injury and is associated with central sensitization and activation of N-methyl-d-aspartate receptor in the spinal cord. The primary outcome was antihyperalgesia calculated with the area under the curve of the percentage change in mechanical pain threshold (electronic von Frey) on the area of secondary hyperalgesia. The secondary outcomes were mechanical pain threshold on the area of primary hyperalgesia and cognitive (reaction time) effect. Results Single 30-mg results are reported. Antihyperalgesia (% · min) is significantly higher on the area of secondary hyperalgesia with dextromethorphan than placebo (median [interquartile range]: 3,029 [746; 6,195] vs. 710 [–3,248; 4,439], P = 0.009, Hedge’s g = 0.8, 95% CI [0.1; 1.4]). On primary hyperalgesia area, mechanical pain threshold 2 h after drug intake is significantly higher with dextromethorphan (P = 0.011, Hedge’s g = 0.63, 95% CI [0.01; 1.25]). No difference in antinociception is observed after thermal painful stimuli on healthy skin between groups. Reaction time (ms) is shorter with placebo than with dextromethorphan (median [interquartile range]: 21.6 [–37.4; 0.1] vs. –1.2 [–24.3; 15.4], P = 0.015, Hedge’s g = 0.75, 95% CI [0.12; 1.39]). Nonserious adverse events occurrence (15%, 3 of 20 volunteers) was similar in both groups. Conclusions This study shows that low-dose (30-mg) dextromethorphan is antihyperalgesic in humans on the areas of primary and secondary hyperalgesia and reverses peripheral and central neuronal sensitization. Because dextromethorphan had no intrinsic antinociceptive effect in acute pain on healthy skin, N-methyl-d-aspartate receptor may need to be sensitized by pain for dextromethorphan to be effective.


Author(s):  
Meyer Richard A. ◽  
Raja Srinivasa N. ◽  
Campbell James N.

1991 ◽  
Vol 66 (1) ◽  
pp. 212-227 ◽  
Author(s):  
T. K. Baumann ◽  
D. A. Simone ◽  
C. N. Shain ◽  
R. H. LaMotte

1. A local cutaneous injury can produce primary hyperalgesia within the injured area and secondary hyperalgesia in the normal surrounding skin. An intradermal injection of capsaicin in humans causes intense pain and hyperalgesia to heat and to mechanical stimuli in the surrounding skin. Psychophysical studies in humans supported the conclusions that the hyperalgesia was predominantly the secondary type and depended on one set of neurons sensitizing another (“neurogenic hyperalgesia”) and that the latter set of neurons is located in the central and not the peripheral nervous system. To further test this hypothesis and to search for peripheral neural mechanisms contributing to the pain and neurogenic hyperalgesia from a local injury, we performed neurophysiological experiments in the monkey (Macaca fascicularis) and recorded the responses of cutaneous primary afferent fibers to an intradermal injection of capsaicin and to mechanical and heat stimuli delivered before and after the injection. 2. Most C- and A-fiber mechanoheat-sensitive nociceptive afferent fibers (CMHs and AMHs, respectively) responded too weakly or transiently to capsaicin to account quantitatively for the magnitude of capsaicin pain. Of the known primary afferents tested with capsaicin injections, only the responses of heat-selective nociceptors could potentially account for the pain measured psychophysically in the human. In addition, a novel type of primary afferent--tentatively termed “chemonociceptive”--may have contributed as well. 3. Nociceptive fibers did not become sensitized to either mechanical or heat stimulation after an injection of capsaicin either outside, adjacent to, or inside the receptive field (RF); any changes that occurred could not explain the hyperalgesia to mechanical or heat stimuli observed in humans. 4. The depressed responsiveness ("desensitization") of both myelinated and unmyelinated nociceptive fibers in the monkey to heat and/or mechanical stimulation of the injection site after capsaicin was injected inside their RFs correlated with the analgesia observed at the capsaicin injection site in the human. 5. Capsaicin, topically applied to the RF in a vehicle of dimethyl sulfoxide or alcohol, excited CMHs and AMHs and enhanced the responses of some of these fibers to heat and/or to stroking the skin. In some cases, similar results were produced by the vehicle alone. However, capsaicin and not the vehicle lowered the thresholds of some CMHs to heat. Thus the sensitization of CMHs contributes to the primary hyperalgesia known to occur within the area of skin directly exposed to topically applied capsaicin.(ABSTRACT TRUNCATED AT 400 WORDS)


2002 ◽  
Vol 97 (3) ◽  
pp. 550-559 ◽  
Author(s):  
Mikito Kawamata ◽  
Hiroaki Watanabe ◽  
Kohki Nishikawa ◽  
Toshiyuki Takahashi ◽  
Yuji Kozuka ◽  
...  

Background To determine the mechanisms of postoperative pain, the effects of local anesthesia on development and maintenance of surgical incision-induced hyperalgesia were evaluated in a crossover, double-blinded, placebo-controlled human study using 17 subjects. Methods An experimental 4-mm-long incision through skin, fascia, and muscle was made in the volar forearm of each subject. In experiment 1, 1% lidocaine or saline in a volume of 0.2 ml was subcutaneously injected into the incision site pretraumatically and posttraumatically. In experiment 2, a 5-cm-long strip of skin was subcutaneously injected with 0.2 ml of 1% lidocaine near the incision site pretraumatically and posttraumatically. Flare, spontaneous pain, and primary and secondary hyperalgesia to punctate mechanical stimuli were assessed after the incision had been made. Results Pretraumatic lidocaine injection prevented the occurrence of spontaneous pain and development of flare formation that was found surrounding the incision site immediately (1 min) after the incision had been made. The lidocaine suppressed primary hyperalgesia more effectively than did posttraumatic block, but only for the first 4 h after the incision. The preincision block prevented development of secondary hyperalgesia, whereas posttraumatic block did not significantly affect the fully developed secondary hyperalgesia. The area of flare formation and the area of secondary hyperalgesia did not extend over the strip of the skin that had been pretraumatically anesthetized, whereas the posttraumatic block did not significantly reduce the area of fully developed secondary hyperalgesia. Conclusions Pretraumatic injection of lidocaine reduces primary hyperalgesia more effectively than does posttraumatic injection, but only for a short period after incision. The spread of secondary hyperalgesia is mediated peripheral nerve fibers, but when secondary hyperalgesia has fully developed, it becomes less dependent on or even independent of peripheral neural activity originating from the injured site.


2007 ◽  
Vol 85 (6) ◽  
pp. 613-620 ◽  
Author(s):  
Lisa C. Loram ◽  
Andreas C. Themistocleous ◽  
Linda G. Fick ◽  
Peter R. Kamerman

We characterized the time course of inflammatory cytokine release at the site of injury and in plasma after surgery on the rat tail. Anesthetized Sprague–Dawley rats had a 20 mm long incision made through the skin and fascia of their tails. Control rats were anesthetized, but no incision was made. Blood and tissue samples were taken 2 h and 1, 2, 4, and 8 days after surgery and analysed by ELISA for interleukin-1β (IL-1β), interleukin-6 (IL-6), tumor necrosis factor-α (TNF-α), and cytokine-induced neutrophil chemoattractant-1 (CINC-1). In another group of rats, daily behavioral measurements were made of the rats’ responses to a blunt noxious mechanical stimulus (4 Newtons) applied to their tails. Primary hyperalgesia developed within 2 h of surgery and lasted for 6 days. The tissue concentrations of IL-1β, IL-6, and CINC-1 increased within 24 h of surgery, and TNF-α concentration increased within 48 h of surgery. Thereafter, cytokine concentrations remained elevated for 4 (IL-1β and IL-6) to 8 days (CINC-1, TNF-α) after surgery. Control animals did not develop hyperalgesia and no changes in cytokines concentrations were detected. Thus, in our model of postoperative pain, secretion of inflammatory cytokines IL-1β, IL-6, TNF-α, and CINC-1 was not essential for the initiation of postoperative hyperalgesia.


1999 ◽  
Vol 13 (suppl a) ◽  
pp. 37A-41A ◽  
Author(s):  
GF Gebhart

Hyperalgesia has long been recognized clinically as a consequence of tissue injury. Primary hyperalgesia (arising from the site of injury) is generally considered to be due to sensitization of sensory receptors (eg, nociceptors) and perhaps activation of so-called ‘silent nociceptors’ by mediators released, synthesized or attracted to the site of tissue injury. Key questions associated with understanding visceral hyperalgesia relate to whether the viscera are innervated by nociceptors (ie, sensory receptors that respond selectively to noxious intensities of stimulation), whether visceral receptors and/or afferent fibres sensitize after tissue injury and whether silent nociceptors exist in the viscera. Studies in nonhuman animals have revealed that hollow organs such as the esophagus, gall bladder, stomach, urinary bladder, colon and uterus are innervated by mechanically sensitive receptors with low or high thresholds for response. Accordingly, it appears that the viscera are innervated by nociceptors, although the issue is far from settled. One characteristic of cutaneous nociceptors is their ability to be sensitized when tissue is injured. Mechanosensitive visceral receptors also sensitize when the viscera are experimentally inflamed, but both visceral receptors with low thresholds and those with high thresholds for response are sensitized. Moreover, it is often not appreciated that visceral receptors are likely polymodal rather than unimodal – that is, mechanically sensitive visceral receptors typically are also sensitive to chemical and/or thermal stimuli. In this sense, visceral receptors may be considered evolutionarily older than more highly developed, specialized cutaneous receptors. Finally, there are mechanically insensitive receptors that innervate the viscera and, when tissue is injured, develop spontaneous activity and acquire sensitivity to mechanical stimuli. In the aggregrate, visceral receptors change their behaviour in the presence of tissue injury and, along with activated mechanically insensitive receptors, increase the afferent barrage into the spinal cord, contributing to the development of visceral hyperalgesia.


Author(s):  
Taylor Follansbee ◽  
Mirela Iodi Carstens ◽  
E. Carstens

Pain is defined as “An unpleasant sensory and emotional experience associated with, or resembling that associated with, actual or potential tissue damage,” while itch can be defined as “an unpleasant sensation that evokes the desire to scratch.” These sensations are normally elicited by noxious or pruritic stimuli that excite peripheral sensory neurons connected to spinal circuits and ascending pathways involved in sensory discrimination, emotional aversiveness, and respective motor responses. Specialized molecular receptors expressed by cutaneous nerve endings transduce stimuli into action potentials conducted by C- and Aδ-fiber nociceptors and pruriceptors into the outer lamina of the dorsal horn of the spinal cord. Here, neurons selectively activated by nociceptors, or by convergent input from nociceptors, pruriceptors, and often mechanoreceptors, transmit signals to ascending spinothalamic and spinoparabrachial pathways. The spinal circuitry for itch requires interneurons expressing gastrin-releasing peptide and its receptor, while spinal pain circuitry involves other excitatory neuropeptides; both itch and pain are transmitted by ascending pathways that express the receptor for substance P. Spinal itch- and pain-transmitting circuitry is segmentally modulated by inhibitory interneurons expressing dynorphin, GABA, and glycine, which mediate the antinociceptive and antipruritic effects of noxious counterstimulation. Spinal circuits are also under descending modulation from the brainstem rostral ventromedial medulla. Opioids like morphine inhibit spinal pain-transmitting circuits segmentally and via descending inhibitory pathways, while having the opposite effect on itch. The supraspinal targets of ascending pain and itch pathways exhibit extensive overlap and include the somatosensory thalamus, parabrachial nucleus, amygdala, periaqueductal gray, and somatosensory, anterior cingulate, insular, and supplementary motor cortical areas. Following tissue injury, enhanced pain is evoked near the injury (primary hyperalgesia) due to release of inflammatory mediators that sensitize nociceptors. Within a larger surrounding area of secondary hyperalgesia, innocuous mechanical stimuli elicit pain (allodynia) due to central sensitization of pain pathways. Pruriceptors can also become sensitized in pathophysiological conditions, such as dermatitis. Under chronic itch conditions, low-threshold tactile stimulation can elicit itch (alloknesis), presumably due to central sensitization of itch pathways, although this has not been extensively studied. There is considerable overlap in pain- and itch-signaling pathways and it remains unclear how these sensations are discriminated. Specificity theory states that itch and pain are separate sensations with their own distinct pathways (“labeled lines”). Selectivity theory is similar but incorporates the observation that pruriceptive neurons are also excited by algogenic stimuli that inhibit spinal itch transmission. In contrast, intensity theory states that itch is signaled by low firing rates, and pain by high firing rates, in a common sensory pathway. Finally, the spatial contrast theory proposes that itch is elicited by focal activation of a few nociceptors while activation of more nociceptors over a larger area elicits pain. There is evidence supporting each theory, and it remains to be determined how the nervous system distinguishes between pain and itch.


Sign in / Sign up

Export Citation Format

Share Document