Central sensitization: Implications for the diagnosis and treatment of pain



26.02.2018

Clifford J Woolf, Program in Neurobiology and FM Kirby Neurobiology Center, Children’s Hospital Boston, and Department of Neurobiology, Harvard Medical School, Boston, Ma.

Abstract

Nociceptor inputs can trigger a prolonged but reversible increase in the excitability and synaptic efficacy of neurons in central nociceptive pathways, the phenomenon of central sensitization. Central sensitization manifests as pain hypersensitivity, particularly dynamic tactile allodynia, secondary punctate or pressure hyperalgesia, aftersensations, and enhanced temporal summation. It can be readily and rapidly elicited in human volunteers by diverse experimental noxious conditioning stimuli to skin, muscles or viscera, and in addition to producing pain hypersensitivity, results in secondary changes in brain activity that can be detected by electrophysiological or imaging techniques. Studies in clinical cohorts reveal changes in pain sensitivity that have been interpreted as revealing an important contribution of central sensitization to the pain phenotype in patients with fibromyalgia, osteoarthritis, musculoskeletal disorders with generalized pain hypersensitivity, headache, temporomandibular joint disorders, dental pain, neuropathic pain, visceral pain hypersensitivity disorders and postsurgical pain. The comorbidity of those pain hypersensitivity syndromes that present in the absence of inflammation or a neural lesion, their similar pattern of clinical presentation and response to centrally acting analgesics, may reflect a commonality of central sensitization to their pathophysiology. An important question that still needs to be determined is whether there are individuals with a higher inherited propensity for developing central sensitization than others, and if so, whether this conveys an increased risk both of developing conditions with pain hypersensitivity, and their chronification. Diagnostic criteria to establish the presence of central sensitization in patients will greatly assist the phenotyping of patients for choosing treatments that produce analgesia by normalizing hyperexcitable central neural activity. We have certainly come a long way since the first discovery of activity-dependent synaptic plasticity in the spinal cord and the revelation that it occurs and produces pain hypersensitivity in patients. Nevertheless, discovering the genetic and environmental contributors to and objective biomarkers of central sensitization will be highly beneficial, as will additional treatment options to prevent or reduce this prevalent and promiscuous form of pain plasticity.

Introduction

In 1983 I published a study indicating that many features of the pain hypersensitivity accompanying peripheral tissue injury or inflammation were the direct result of an augmentation of sensory signaling in the central nervous system [255]. A central amplification during angina pectoris had been postulated exactly 100 years before by W Allen Sturge MD, who in an 1883 paper in Brain envisaged a possible central nervous system “commotion …… passed up from below” that somehow contributed to the clinical features of ischemic cardiac pain. However, the importance of this clinical insight lay largely dormant for a century, except for one human volunteer study on secondary hyperalgesia that was recognized by the authors as suggestive of a possible central contribution to the spread of pain sensitivity [101]. What I found in a preclinical study on stimulus-response relations in the spinal cord was that the afferent activity induced by peripheral injury triggered a long-lasting increase in the excitability of spinal cord neurons, profoundly changing the gain of the somatosensory system [255]. This central facilitation manifested as a reduction in threshold (allodynia), an increase in responsiveness and prolonged aftereffects to noxious stimuli (hyperalgesia), and a receptive field expansion that enabled input from non-injured tissue to produce pain (secondary hyperalgesia) [255256; 273; 51; 268].

I have recently reviewed the circumstances surrounding the discovery of the activity-dependent synaptic plasticity in the spinal cord that generates post-injury pain hypersensitivity [259], and that became termed “central sensitization” [272], as well as the current state of understanding of the cellular and molecular mechanisms responsible for this form of neuronal plasticity [147]. What I would like to specifically address in this review are the clinical implications of the phenomenon. What has central sensitization taught us about the nature and mechanisms of pain in patients, and what are the implications of central sensitization for pain diagnosis and therapy? Before doing this though, it is important first to understand exactly what central sensitization represents, how it has changed our general understanding of pain mechanisms, as well as reviewing the substantial data on central sensitization derived from studies on experimental pain in human volunteers.

What is central sensitization?

Prior to the discovery of central sensitization, the prevailing view on pain processing in the central nervous system was of a largely passive neural relay that conveyed by encoded action potentials, information on the onset, duration, intensity, location and quality of peripheral noxious stimuli, much like a telephone wire, from one site to another. More specifically, the CNS pathway was seen to constitute particular anatomical connections in the spinal cord, brain stem, thalamus and cortex (the “pain pathway”), linking the sensory inflow generated in high threshold primary afferents with those parts of the cortex that lead to the conscious awareness of painful sensations. The spinal gate control theory by Melzack and Wall in 1965 had highlighted that this sensory relay system could be modulated in the spinal cord by inhibitory controls [163], and considerable progress had been made by the early 1980’s in identifying such inhibitory circuits [18]. Indeed this, together the discovery of enkephalins and endorphins [109; 98], diffuse noxious inhibitory controls [150], transcutaneous nerve stimulation [224], and the rediscovery of acupuncture [25], generated a much greater emphasis at that time on endogenous inhibitory controls than on those factors that might increase excitation, and thereby produce pain hypersensitivity. However, there was one exception, that which related to the discovery of peripheral sensitization in the 1970’s [178]. Work by Iggo [112; 28] and Perl [33; 177; 20] had identified specific high threshold sensory neurons tuned to respond only to noxious stimuli, hence their name nociceptors [265], a term first coined by Sherrington based on his studies on noxious stimulus evoked flexion reflexes. Furthermore, first Perl then others showed that nociceptor peripheral terminals could become “sensitized” after injury, reducing their threshold, mainly to heat stimuli, and only within the site of injury where the terminal was exposed to inflammatory modulators, the zone of primary hyperalgesia [178; 146; 138; 41; 23]. While this phenomenon clearly is a very important contributor to inflammatory pain hypersensitivity [22], it cannot account for dynamic tactile allodynia, the temporal summation of pain, or secondary hyperalgesia. Some other explanation was needed as the neurobiological basis for these symptoms, which turned out to be increased synaptic function triggered within the CNS by nociceptive inputs [257; 237; 268].

The realization that synapses were subject to a form of use-dependent plasticity that could increase their strength or efficacy had steadily gained ground by the early 1980’s. The phenomenon had first been described in the CNS as short lasting a post-tetanic potentiation of mono synaptic IA synaptic input to motor neurons by Lloyd in 1949 [155], one that could spread to other synapses on motor neurons [21]. This was followed by the discovery of windup in dorsal horn neurons by Mendell and Wall in 1965 [164], where repeated low frequency stimulation of a nerve at constant C-fiber strength were found to elicit a progressive increase in action potential firing over the course of the stimulus. A transformative breakthrough was the first description of long term potentiation (LTP) in the hippocampus by Bliss and Lomo in 1973, where a brief high frequency coincident input produced a persistent increase in synaptic efficacy, opening the door for an extensive and still ongoing study into the molecular mechanisms of synaptic plasticity. LTP was first recorded in the spinal cord in 1993 [182], where it represents a particular component of the general phenomenon of central sensitization [113; 122; 114]. In 1976 Kandel and colleagues described a sensitization of the gill withdrawal reflex in the sea snail Aplysia, which was associated with a facilitation of the synapse between sensory and motor neurons [29]. However, this data was interpreted as reflecting memory and learning rather than an invertebrate model of pain hypersensitivity, although of course the two phenomena converge in this, and other model systems, although there are differences too [274; 122].

What I found in my original 1983 and subsequent pre-clinical studies with colleagues at University College London, was that a brief (~10–20 second), low frequency (1–10Hz) burst of action potentials into the CNS generated by electrical stimulation or natural activation of nociceptors increased synaptic efficacy in nociceptive neurons in the dorsal horn of the spinal cord and this lasted for tens of minutes after the end of the conditioning stimulus [255; 244; 256; 267; 50; 245; 273; 51; 272; 263; 230; 264]. This phenomenon differed from windup, which represented a progressively increasing output during the course of a train of identical stimuli (technically called homosynaptic potentiation); central sensitization was concerned instead with the facilitation that manifested after the end of the conditioning stimuli, and that once triggered remained autonomous for some time, or only required a very low level of nociceptor input to sustain it. Furthermore, central sensitization represented a condition where input in one set of nociceptor sensory fibers (the conditioning input) amplified subsequent responses to other non stimulated non-nociceptor or nociceptor fibers (the test input; this form of facilitation is termed heterosynaptic potentiation to distinguish it from homosynaptic potentiation where the test and conditioning input are the same) [231]. The classic form of LTP in the hippocampus is homosynaptic with changes in efficacy restricted to activated synapses, a convergent plasticity, and while this is a feature of some aspects of central sensitization [190], most of its clinically relevant attributes relate to its divergent heterosynaptic components [147]. The underlying neurobiological basis for central sensitization is that for most central circuits, the receptive field properties of neurons defined by the firing of action potentials is only the “tip of the iceberg”. Most of the synaptic input to neurons is subthreshold [262263], acting subliminally either because synaptic input is too weak or membrane excitability is restrained by inhibitory inputs. Increasing synaptic strength by a presynaptic increase in an excitatory transmitter release or in the post synaptic response to the transmitter [264; 271; 231; 129; 247; 130; 133; 154; 46; 100; 151152; 227] or by reducing inhibition [208; 168; 12; 103; 180; 165; 226] or increasing membrane excitability can recruit these normally subthreshold inputs to suprathreshold action potentials, producing profound changes in functional properties [270]. More recently it has become appreciated that in addition to activity-dependent synaptic plasticity, changes in microglia, astrocytes, gap junctions, membrane excitability and gene transcription all can contribute to the maintenance of central sensitization [234; 189; 43; 48; 88; 104; 205; 44; 47; 186]. Figures 1 and ​and22summarize sensory processing under normal circumstances and the changes that result from induction of central sensitization.

Figure 1
Normal sensation. The somatosensory system is organized such that the highly specialized primary sensory neurons that encode low intensity stimuli only activate those central pathways that lead to innocuous sensations, while high intensity stimuli that ...





Figure 2
Central sensitization. With the induction of central sensitization in somatosensory pathways with increases in synaptic efficacy and reductions in inhibition, a central amplification occurs enhancing the pain response to noxious stimuli in amplitude, ...






An important implication of these early basic science studies was the possibility that the pain we experience might not necessarily reflect the presence of a peripheral noxious stimulus. We learn from our everyday experience interfacing with the external environment to interpret pain as reflecting the presence of a peripheral damaging stimulus, and indeed this is critical to its protective function. Central sensitization introduces another dimension, one where the CNS can change, distort or amplify pain, increasing its degree, duration, and spatial extent in a manner that no longer directly reflects the specific qualities of peripheral noxious stimuli, but rather the particular functional states of circuits in the CNS. With the discovery of central sensitization, pain conceptually at least had become “centralized” instead of being exclusively peripherally driven. In this sense central sensitization represents an uncoupling of the clear stimulus response relationship that defines nociceptive pain. Nociceptive pain reflects the perception of noxious stimuli. In the absence of such potentially damaging stimuli there is no nociceptive pain. However, after the discovery of central sensitization it became clear that a noxious stimulus while sufficient was not necessary to produce pain. If the gain of neurons in the “pain pathway” in the CNS was increased, they could now begin to be activated by low threshold, innocuous inputs. In consequence pain, could in these circumstances become the equivalent of an illusory perception, a sensation that has the exact quality of that evoked by a real noxious stimulus but which occurs in the absence of such an injurious stimulus. This does not mean the pain is not real, just that it is not activated by noxious stimuli. Such pain can no longer be termed nociceptive, but rather reflects a state of induced pain hypersensitivity, with almost precisely the same “symptom” profile to that found in many clinical conditions. This raised the immediate obvious question, was central sensitization a contributor to clinical pain hypersensitivity?

These notions were generally not very well received initially, particularly by physicians who believed that pain in the absence of pathology was simply due to individuals seeking work or insurance-related compensation, opioid drug seekers, and patients with psychiatric disturbances; i.e. malingerers, liars and hysterics. That a central amplification of pain might be a “real” neurobiological phenomenon, one that contributes to diverse clinical pain conditions, seemed to them to be unlikely, and most clinicians preferred to use loose diagnostic labels like psychosomatic or somatoform disorder to define pain conditions they did not understand. We can now 30 years later, based on data from many studies in human volunteers and patients, address whether central sensitization, defined operationally as an amplification of neural signaling within the CNS that elicits pain hypersensitivity, is a real phenomenon or not, and can assess its relative contribution to inflammatory, neuropathic and dysfunctional pain disorders in patients [258; 53].

Central sensitization in human volunteers

The first clear demonstration of central sensitization in human volunteers came from a psychophysical study by LaMotte and colleagues on the secondary cutaneous hyperalgesia that is elicited by intradermal capsaicin injection (which activates the TRPV1 receptor). They found intense localized pain lasting minutes at the injection site, followed immediately by three zones of hyperalgesia; a small zone of heat hyperalgesia close to the injection site lasting 1–2 hours, an intermediate zone of dynamic tactile allodynia spreading beyond the area of heat hyperalgesia and lasting several hours, and the largest zone to pinprick, way outside of the injection site, which remained present for up to 24 hours [145]. The investigators then showed that the secondary mechanical hyperalgesia required sensory inflow to the CNS because local anesthesia prior to the capsaicin injection blocked it. In addition because the pain sensitivity crossed a tight band that prevented circulation in the skin, they concluded that it was not due to a local spread of the capsaicin or any peripheral inflammatory mediator. An even more direct demonstration that activity dependent central sensitization was responsible for tactile allodynia and secondary hyperalgesia in humans came from a second study by La Motte, this time with Torebjork in 1992 [233]. They again used intradermal injection of capsaicin to induce an area of tactile allodynia that lasted for 2 hours. Nerve block experiments revealed that while the capsaicin and heat pain was carried by C fibers, the mechanical allodynia was transferred to the CNS by low threshold myelinated fibers. The most elegant part of the study was their finding that electrical intraneural stimulation of single Aβ mechanoreceptive fibers that elicited a non painful tactile sensation before the capsaicin injection, began to produce pain if the fibers’ receptive field fell within the zone of secondary mechanical hyperalgesia. Lidocaine anesthesia of the cutaneous innervation territory of the stimulated fiber did not reverse the pain, showing this was not peripheral in origin. They concluded that the pain evoked by stroking the skin area surrounding a painful intradermal injection of capsaicin “is due to reversible changes in the central processing of mechanoreceptive input from myelinated fibres which normally evoke non-painful tactile sensations”.

Another early study, this time by Koltzenburg and Torebjork, using mustard oil (which activates TRPA1) as the pain conditioning stimulus, together again with differential nerve blocks, confirmed that brush-evoked mechanical allodynia was mediated by low threshold Aβ fibers that normally encode non-painful tactile sensations [140]. Unlike after capsaicin, however, the mustard oil evoked tactile allodynia required an ongoing low level input from C-nociceptors to sustain it, indicating that different sensory fibers may have different central actions, some short and others long lasting, and indeed further studies have shown differences in the duration of tactile allodynia after capsaicin and mustard oil [139], the significance of which was not appreciated a the time because it was not clear then that these irritants acted on quite different TRP receptors.

That central sensitization could cause a spread of pain sensitivity across peripheral nerve territories, the neurological dogma for diagnosing a disease of the central rather than peripheral nervous system, was shown by Max and colleagues using the intradermal capsaicin model in volunteers together with radial or ulnar nerve blocks to clearly identify individual nerve territory [192]. Complementing this, a study comparing skin hyperaemia induced by a skin burn injury found that the skin blood flow changes induced by the injury had disappeared by the time secondary mechanical hyperalgesia peaked, and the two were not correlated in time or space, supporting the conclusion that peripheral mechanisms do not contribute to secondary hyperalgesia [198]. Even more dramatic perhaps, was the relatively recent demonstration that intradermal capsaicin induces contralateral hyperalgesia and allodynia that is delayed in its manifestation and reduced in extent compared to the ipsilateral secondary hyperalgesia, but present in a majority of subjects [206], a form perhaps of “tertiary hyperalgesia” that cannot be peripheral in origin. What pain sensitivity we feel then, can be determined by the state of excitability of neurons in the CNS.

Central amplification of Aδ nociceptor fiber test input following a C-fiber conditioning input was shown to contribute to pinprick/punctate secondary hyperalgesia, again using the intradermal capsaicin model [279], underscoring the different identity of the afferent signals that elicit central sensitization as a conditioning stimulus (C-fibers) from those that elicit allodynia (Aβ) or hyperalgesia (Aδ), a further clear manifestation of heterosynaptic facilitation. In a similar vein, another study found that pin prick hyperalgesia induced in response to intradermal capsaicin was actually mediated by capsaicin-insensitive afferents, showing that the test and conditioning inputs in this setting are quite different [87], while the secondary hyperalgesia elicited by intradermal capsaicin was shown by yet other investigators, to be restricted to mechanical stimuli, with no correlation between the magnitude of capsaicin evoked pain and the extent of punctate or tactile secondary hyperalgesia [237]. Furthermore, temporal summation to pin prick in the zone of capsaicin injection (as model of homosynaptic facilitation/windup) was mechanistically independent of the development of secondary hyperalgesia, because while the gain of the stimulus-response relationship in the zone of secondary was increased, that of the windup was not changed, even though the actual pain was enhanced [158]. A similar conclusion was made after a study where repeated intradermal capsaicin injections were reported to produce a progressively diminishing pain, presumably due to desensitization, while the allodynia and punctate hyperalgesia continued to increase [254]. Two more recent studies using high frequency stimulation as the conditioning input to mimic conditions that elicit LTP, found that while changes in the conditioned site (homotopic site) do occur, they are accompanied by a development of pain hypersensitivity in the adjacent non-stimulated heterotopic site (reduction in threshold, pain evoked by light tactile stimuli, and exaggerated response to suprathreshold pinprick stimuli [136; 240], and both sets of investigators concluded that heterosynaptic facilitation predominates in this model of central sensitization, exactly as it does for the low frequency conditioning inputs that mimic the natural firing range of nociceptors. Generalizing, it seems clear that heterosynaptic changes are a major feature of the presentation of central sensitization.

Apart from changes in subjective pain measures, the consequences of central sensitization can also be detected using objective biomarkers. These include long-term changes in nociceptive withdrawal reflexes [24] and increases in cortical event related potential amplitudes [240]. Magnetic source imaging reveals an increase in the excitability of neurons in the somatosensory cortex evoked by low threshold Aβ stimulation within the capsaicin-induced zone of secondary hyperalgesia [17], while magnetoencephalography detects changes in the patterns of cerebral processing [159] and functional MRI, changes in BOLD signals in the cortex, both during secondary hyperalgesia [16]. Another MRI study found changes in the brainstem that are apparently specific to central sensitization, in addition to the changes in the primary somatosensory cortex that are related to the intensity of pain [153].

While most studies have looked at the effects of skin conditioning stimuli on skin pain sensitivity, experimental muscle pain produced by hypertonic saline injections produces long lasting changes in thermal sensitivity in the area of referred pain [203], while sustained nociceptive stimulation of myofascial trigger points induces a wide spread central sensitization [273; 275]. Interestingly, in preclinical models, muscle and joint conditioning afferents have a longer lasting action in producing central sensitization than those from skin [244]. A reverse approach has shown that cutaneous capsaicin increases myofascial trigger point pressure sensitivity in segmentally related muscles [211]. Conditioning nociceptive stimuli originating in viscera, such as exposure of the lower esophagus to acid, also induces central sensitization, leading to viscerovisceral (pain hypersensitivity in the upper esophagus) and viscerosomatic hypersensitivity (allodynia on the chest wall) [193] that can be captured by esophageal evoked potentials [194], and is associated with increased temporal summation [196]. A recent study has replicated this esophageal model of central sensitization using acid and capsaicin infusions, showing also thermal and mechanical pain hypersensitivity in the rectum after the esophageal stimulation [27], indicating how widespread the effects of central sensitization are in the gastro intestinal tract. These changes may be mechanistically related to widespread clinical pain syndromes [95].

One emerging area of considerable interest is the utility of experimental central sensitization in human volunteers to test efficacy in centrally acting drugs. Drugs with efficacy in preclinical models, such as NMDA receptor antagonists [271] can be tested in Phase 1b human proof of principle studies [212]. Ketamine inhibits central temporal summation [8] and secondary mechanical hyperalgesia [142] evoked by repetitive nociceptive electrical stimulation in humans as well as primary and secondary hyperalgesia after an experimental burn injury [116], visceral conditioning inputs [253; 251] and topical [6] or intradermal [204] capsaicin, but not A delta mediated nociceptive pain [181]. Ketamine’s action on experimental pain can be detected by fMRI [210]. Similar activity is found for i.v. dextromethophan [115]. Collectively these data strongly support a role for the NMDA receptor in acute activity-dependent central sensitization [147]. However, the trials also indicate the lack of therapeutic index between reducing central sensitization and inducing psychotomimetic side effects. Another class of drugs that has been extensively studied in human experimental models of central sensitization is the gabapentanoids. Oral gabapentin at doses similar to that used for chronic neuropathic pain when given to human volunteers reduced tactile allodynia and decreased mechanical secondary hyperalgesia elicited by intradermal capsaicin [92]. Even single administration of gabapentin had an antihyperalgesic effect on capsaicin-induced secondary hyperalgesia and reduced fMRI signatures of central sensitization [110]. In another study gabapentin, interestingly reduced cutaneous evoked central sensitization but not muscle pain [201]. Two studies have looked at pregabalin’s efficacy in experimental human central sensitization, one evoked by electrical stimuli [49] and the other by intradermal capsaicin [246]. Both of these double blind studies demonstrated efficacy for pregabalin in terms of experimental tactile allodynia and secondary hyperalgesia. These data suggest that a major component of gabapentin or pregabalin’s mechanism of action is a reduction of central sensitization [238]. Many other centrally acting drugs with analgesic efficacy in patients reduce central sensitization preclinically, including duloxetine, milnacipran and lamotrigene [118; 170; 15] but have not been tested for this action in humans. Drugs that have failed to show efficacy in human studies of activity-dependent central sensitization are NK1 receptor antagonists [252] [49] and COX-2 inhibitors [35; 250; 49]. A COX-2 inhibitor does have efficacy though if the central sensitization is triggered by peripheral inflammation [225], as predicted by preclinical models [189].

Interestingly, while gender has been described as important for differences in nociceptive pain sensitivity, a study on the secondary hyperalgesia induced by heat and capsaicin did not reveal a gender difference [119]. Nevertheless, recent data show that pain sensitivity including secondary hyperalgesia and brush evoked allodynia is heritable, with an estimated 50% genetic contribution to the pain variance [172]. The genetic polymorphisms involved in the differential susceptibility to secondary hyperalgesia have not been comprehensively investigated, although some candidates are beginning to be identified in studies of experimental central sensitization [228]. This is an area that requires major research.

The following conclusions can be made from this survey of the published studies of experimental pain hypersensitivity in human volunteers. Central sensitization is a robust phenomenon, readily induced in human volunteers in response to diverse ways of activating nociceptors (electrical stimulation, capsaicin, mustard oil, acid, heat burn, UV burn, hypertonic saline). Generally this activity-dependent plasticity manifests immediately, but its effects persist for many hours beyond the inducing conditioning stimulus, eventually returning, however, back to baseline, indicating its usual full reversibility. The phenomenon can be elicited by conditioning skin, muscle or visceral organs, and typically presents as dynamic tactile allodynia and punctate hyperalgesia but also enhanced pressure, and in some cases, thermal sensitivity, spreading from the conditioning site to neighboring non stimulated sites, and even to very remote regions. Although there is a homosynaptic (homotopic) aspect to the phenomenon, its major manifestation is heterosynaptic (heterotopic), and for this reason and its reversibility, it is perhaps inaccurate to equate central sensitization with the LTP like phenomena in the cortex that are specifically associated with long term memory. Because central sensitization can be induced in almost all subjects and detected using subjective and objective outcome measures and is sensitive to pharmacological interventions, it is a useful tool for determining the activity of drugs on centrally driven pain hypersensitivity.

Globally, the data obtained in human volunteer studies demonstrate that induction of use-dependent central facilitation in nociceptive central pathways increases pain sensitivity and may, therefore, contribute to clinical pain syndromes. Experimental studies in human volunteers are necessarily restricted to using non-injurious conditioning inputs, and therefore are limited to studying only the activity-dependent components of pain hypersensitivity elicited by sensory inputs, and not those transcription-dependent and structural changes that manifest after inflammation or nerve injury, which may have different mechanisms, time courses and presentations [269; 171; 121; 160; 261; 189; 123; 229; 97; 53; 242]. The limited experience with more severe human experimental injury indicates that central sensitization also contributes to the late hyperalgesia present in this model [58; 176].

Central sensitization and the clinical pain phenotype

What features of the clinical phenotype may be contributed to, or generated exclusively by central sensitization? While the human experimental studies reviewed above indicate that if a patient has dynamic tactile allodynia, secondary punctuate/pressure hyperalgesia, temporal summation and sensory aftereffects, central sensitization may well be involved. Any sensory experience greater in amplitude, duration and spatial extent than that would be expected from a defined peripheral input under normal circumstances, qualifies as potentially reflecting a central amplification due to increased excitation or reduced inhibition. These changes could include a reduction in threshold, exaggerated response to a noxious stimulus, pain after the end of a stimulus, and a spread of sensitivity to normal tissue. However, because we cannot directly measure sensory inflow, and because peripheral changes can contribute to sensory amplification, as with peripheral sensitization, pain hypersensitivity by itself is not enough to make an irrefutable diagnosis of central sensitization. A further complication is that because peripheral input commonly is the trigger of central sensitization, a reduction in pain sensitivity produced by targeting a peripheral trigger with a local anesthetic does not exclude central amplification, but may rather indicate a role of peripheral input in maintaining it [140]. Nevertheless, there are some features of patient’s symptoms which are more likely to indicate central rather than peripheral contribution to pain hypersensitivity. These include pain mediated by low threshold Aβ fibers (determined by nerve block or electrical stimulation), a spread of pain sensitivity to areas with no demonstrable pathology, aftersensations, enhances temporal summation, and the maintenance of pain by low frequency stimuli that normally do not evoke any ongoing pain. To assess how central sensitization may present in patients, we need a detailed phenotyping of different patient cohorts to capture exactly what changes in sensitivity occur, where and when [93; 188; 86; 9; 11; 197; 55]. Ideally this should be combined with objective measures of central activity, such as fMRI, so that clear diagnostic criteria for determining the presence of central sensitization in patients can be established. The utility of diagnostic criteria for the presence of central sensitization would not only be insight into the pathophysiological mechanisms responsible for producing pain, but more so in defining potential treatment strategies. If a particular patient’s pain is primarily the result of abnormal activity in nociceptors, as in patients with primary erythromelalgia [74], the optimal therapy required is likely to be different from a patient whose tactile allodynia and secondary hyperalgesia are entirely maintained by central sensitization due to changes in synaptic efficacy in the spinal cord. This is the rationale for a mechanism-based approach to the diagnosis and treatment of pain [266; 258]. Indeed response to a trial treatment, such as to the NMDA receptor antagonist ketamine, can itself be a potential diagnostic for the presence central sensitization.

To which clinical syndromes does central sensitization contribute?

Given the caveats about the lack of absolute diagnostic criteria for identifying the presence of central sensitization in patients, a fairly large number of studies have nevertheless putatively identified this phenomenon as contributing to patients’ pain phenotype. I will briefly review these, based on disease.

Rheumatoid arthritis (RA)

Patients with RA, the prototypic inflammatory joint disease, have extra-articular tenderness which is correlated with the extent of joint disease [141] but whether this is the result of peripheral or central sensitization has not been studied. A study on juvenile chronic arthritis reported enhanced sensitivity to noxious stimuli both at joints and in remote areas in patients with and without active disease, suggesting the possibility that the disease when active sets up a state of autonomous central sensitization [107]

Osteoarthritis (OA)

This degenerative joint disease with characteristic destruction of cartilage and alteration in bone is a very common cause of chronic pain, particularly in the elderly. The degree of pain does not always correlate with the extent of joint damage or presence of active inflammation raising the possibility that there may be a central component to the pain [26]. Supporting this is the enhanced degree and duration of pain and secondary hyperalgesia evoked by intramuscular injection of hypertonic saline in patients with OA compared to controls [13]. Patients with high pre-operative pain and a low pain threshold have a higher risk of persistent pain after total knee replacement for OA, which was interpreted as reflecting central sensitization [157]. Another study on 62 patients showed that pain of central neural origin (widespread reduced pressure pain thresholds) negatively impacted on knee functional capacity [117]. OA patients have a lower pain threshold and have punctate hyperalgesia in areas of referred pain, which is associated with greater activation in the brainstem as detected by fMRI, representing a possible biomarker for central changes [99]. The centrally acting amine uptake inhibitor duloxetine which reduces central sensitization in preclinical models [124; 15], significantly reduced pain more than placebo in an RCT in 231 patients with knee OA pain [45], indicating that drigs that target central sensitization are efficacious in this patient population. In a recent phenotyping study in 48 patients with painful knee OA and 24 age matched controls, the patients had reduced pressure pain thresholds both at the joint and in remote areas, and increased temporal summation. While the degree of sensitization correlated with the pain, it did not correlate with radiological findings, leading to the conclusion that central sensitization is an important contributor to knee OA pain [7]. Collectively, these data intriguingly suggest that the pain of OA, a peripheral pathology, has an important central component, and this is clearly deserving of more study to understand its extent, mechanism and therapeutic implications.

Temporomandibular disorders (TMD)

Unlike OA, the pathophysiology of this syndrome is much less well understood. However, TMD has been found to be associated with an increase in generalized pain sensitivity after isometric contraction of the orofacial muscles [166], and widespread bilateral mechanical [78] and thermal [175] pain sensitivity is reported in women with myofascial TMD compared to age matched controls, which was interpreted as suggesting widespread central sensitization. In addition, a greater referred pain is elicited from the more frequent trigger points that are found in these patients, than in controls [77].

As for other types of facial pain, mechanical allodynia is a major feature of periradicular inflammation (periradicular periodontitis) with reduced threshold also in contalateral non inflamed teeth, reflecting central sensitization [132]. After a third molar extraction evidence for central sensitization could be detected for at least a week (enhanced response to repetitive intraoral pinprick and electrical stimulation, aftersensations and extraoral hyperalgesia) [126].

Fibromyalgia (FM)

One of the first suggestions that fibromyalgia patients may have generalized central sensitization came from a psychophysical study that identified widespread reduction in thermal and mechanical pain thresholds, as well as greater cerebral laser evoked potentials [90], a finding replicated soon after [156]. Another early small study using ketamine, showed an NMDA-dependent component to fibromyalgia and suggested that tender points may represent secondary hyperalgesia due to central sensitization [209]. Supporting this, Arendt-Nielson and colleagues found in small study that fibromyalgia patients had lower pressure thresholds and increased temporal summation to muscle stimulation, and that intramuscular hypertonic saline injections provoked a longer lasting and more widespread pain. In a related study, they found that the referred pain, temporal summation, muscular hyperalgesia and muscle pain in fibromyalgia patients were all attenuated by ketamine [96]. In 2001 Staud and Price begun a series of studies on fibromyalgia, first showing temporal summation and after sensations of the pain elicited by repetitive cutaneous thermal stimuli and repetitive mechanical stimuli to muscles [221]. In a second study they found that temporal summation occurred at substantially lower forces and at a lower frequency of stimulation in fibromyalgia patients than in control subjects, and that painful after sensations were greater in amplitude and more prolonged [215]. The enhanced experimental pain in fibromyalgia patients was shown to contribute to the variance of the clinical pain [220]. These investigators then showed that the maintenance of experimentally induced pain in fibromyalgia patients requires significantly less frequent stimulation than in normal controls, and concluded that this heightened sensitivity to very low frequency inputs contributes to the persistent pain in these patients [218]. A later study showed that the temporal summation of pain and its maintenance were widespread, and could be equally elicited from hands or feet, leading to the conclusion that central sensitization in these patients was generalized across the neuraxis [219]. In an fMRI study they then found a stimulus and frequency dependent activation in several brain regions in fibromyalgia patients and controls, including ipsilateral and contralateral thalamus, medial thalamus, S1, bilateral S2, mid- and posterior insula, rostral and mid-anterior cingulate cortex. The stimulus temperatures necessary to evoke equivalent levels of brain activity were, however, significantly less in fibromyalgia patients, suggesting that the enhanced neural mechanisms in fibromyalgia are not the result of selective enhancement at cortical levels [216]. The Staud and Price group then designed experiments to see if peripheral sensitization may contribute to the enhanced temporal summation of thermal pain in fibromyalgia patients and concluded that it does not, based on thermal thresholds [214]. Recently they have found using local anesthetic injections though, that peripheral input from muscle appears to be important in maintaining central sensitization in FM patients [217]. This would mean that fibromyalgia may have both peripheral and central contributions, whose extent may vary from patient to patient. Certainly muscle afferents seem to have a potent capacity in preclinical [244] and experimental human studies [275] to induce central sensitization.

A quantitative sensory testing study in 85 fibromyalgia patients and 40 matched controls found that the patients had altered heat and cold thresholds and a reduced tolerance for pain, as well as a reduced nociceptive reflex threshold, a measure of central excitability [65]. The latter finding was sufficiently different from controls that the authors suggest it could be used as a diagnostic measure of central sensitization, identifying patients for whom centrally acting drugs may be particularly beneficial. Other studies have confirmed the increased generalized sensitivity in FM patients to pressure and thermal stimuli [179]; [94; 173] and to electrical stimulation of skin and muscle, with enhanced cortical evoked potentials [66]. The data overall seem to support a major role for central sensitization in the generation of the symptoms of FM, and the success of centrally acting treatments, such as pregabalin or duloxetine in treating these conditions, may reflect a reduction in central sensitization in these patients.

Miscellaneous Musculoskeletal Disorders

Chronic neck pain resulting from whiplash is associated with lowered pain thresholds in uninjured tissue [57; 222]. Injection of local anesthetic into myofascial trigger points in these patients results in an immediate increase in range of motion and elevation in pressure pain thresholds, which was felt to reflect dynamic maintenance of central sensitization by afferent triggers [85]. Patients with shoulder impingement syndrome also show widespread muscle sensitivity and an increased number of trigger points [105]. A widespread (bilateral) mechanical pain hypersensitivity is observed in patients with unilateral epicondylalgia (tennis elbow) interpreted as indicating central sensitization, possible induced by a peripheral trigger [75]. Similar generalized deep tissue hyperalgesia can also be demonstrated in patients with chronic radiating low back pain with intervertebral disc herniation [173]. Collectively these data indicate that diverse musculoskeletal disorders are characterized by a spread of pain sensitivity to deep uninjured tissue and that low level peripheral inputs may maintain this.

Headache

The first intimation that headaches have an important component mediated by central sensitization came from a study of spontaneous tension-type headaches which found that even in the absence of headache pericranial muscle tenderness was increased in patients compared to control subjects. During headache, muscle tenderness increased and thermal pain threshold decreased in the temporal region, but remained normal in the hand which was interpreted as indicating that segmental central sensitization contributed to pain in frequent sufferers of tension-type headache [120]. This was then followed by the observation by Bernstein and colleagues that cutaneous allodynia developed in 79% of patients during migraine attacks in, and sometimes beyond the area of referred pain [3637]. This finding has been repeated in several studies since then [161; 135; 207; 52]. While cephalic and extracephalic allodynia are well described, spontaneous body pain and allodynia has also been reported as preceding migraine attacks [56]. Laser evoked cutaneous pain thresholds are reduced during migraine attacks and cortical evoked potentials increased [62]. No change in heat pain thresholds are found in chronic tension-type headache, but there is pericranial tenderness [63; 80] and hyperalgesia of neck shoulder muscles [81]. Nociceptive input from muscles has been suggested to contribute to the induction central sensitization in tension-type headache [79], much as has been suggested for FM. In patients with cluster headaches the nociceptive flexion reflex threshold is reduced on the symptomatic side [191]. In a population study on primary headaches in 523 patients, evidence for pain hypersensitivity was found in those with tension type pain, with a greater disturbance in individuals with chronic or more frequent headaches, implying that central sensitization may contribute to the chronification of headache [30], something that is supported by epidemiological data [31]. In a longitudinal prospective study on whether increased pain sensitivity is a cause or an effect, a study in 100 individuals found that subjects had normal thresholds prior to the development of headache, but this decreased in those who then developed chronic tension-type headache, suggesting that the pain hypersensitivity is a consequence of frequent tension-type headaches, and not a predictor or risk factor [32], a finding interpreted as showing that central sensitization plays a role in the chronification of tension-type headaches. Interestingly, a study in patients with either chronic migraine and chronic tension-type headache, found in both cohorts reduced threshold for pressure, pinprick, blink, and the nociceptive flexion reflex, as well as higher windup ratios [83], possibly reflecting a common role for central sensitization in the chronification of different types of headache.

Neuropathic pain

The first demonstration of a likely contribution of central sensitization to neuropathic pain came from a study by Campbell and colleagues, who showed that an ischemic conduction block of large myelinated fibers specifically reduced dynamic tactile allodynia [42], a finding that was soon replicated [140]. Since then careful phenotyping studies of conditions like carpal tunnel syndrome have revealed enhanced bilateral sensitivity and an extraterritorial spread of symptoms in patients with unilateral or single nerve entrapment, supporting a contribution of central sensitization [61; 76; 82; 278]. Furthermore, ketamine reduces established peripheral neuropathic pain [125] and chronic phantom limb pain [73] indicating that ongoing activity- and NMDA receptor-dependent synaptic plasticity may contribute to maintaining neuropathic pain. That tricyclic antidepressants, dual uptake inhibitors and calcium channel alpha(2)-delta ligands, all centrally acting drugs that normalize enhanced neural activity, are the current first line treatments for neuropathic pain [72], reinforces the importance of the central component of the pain and its suitability as a target for treatment.

Complex regional pain syndrome (CRPS)

A prominent feature of chronic CRPS1 is tactile hyperesthesia and pressure hyperalgesia [241], which can be registered as enhanced S1 activation by a neuromagnetometer [243]. There is also thermal hyperalgesia in acute CRPS1 patients, which on the side ipsilateral to the diseased limb, may have a peripheral component due to ongoing aseptic inflammation, but the presence of contralateral hypersensitivity in the absence of any inflammatory changes points to an involvement of the CNS [108]. In a small randomized placebo controlled trial intravenous ketamine reduced CRPS pain [200].

Post-surgical pain

This is a very heterogenous group comprising acute postoperative pain and persistent pain of multiple causes, including surgical induced neuropathic pain [131; 1]. In the acute phase, incisional pain is associated with a secondary punctate hyperalgesia that is ketamine sensitive [223], with no spread in thermal sensitivity [143] indicating induction of central sensitization. Considerable controversy exists over whether pre-emptive treatment targeting central sensitization is superior to postoperative treatment in treating either the acute postoperative pain or its transition to chronic pain [260; 68; 128; 60; 149; 7071; 102; 45; 54; 236]. Surprisingly, because of numerous technical problems related to the design, conduct and interpretation of such studies, this turns out to be a difficult issue to resolve [167; 134]. This is not the place to review the full literature on pre-emptive analgesia, however my personal take on the available data is that there appears to be a small signal for pre- vs. postoperative analgesic treatment in some settings, but it is likely not generally clinically relevant. It seems clearly important though, that patients have full analgesia established on recovery from a general anesthetic or adequate regional anesthesia during surgery, and that this be maintained until surgical healing is well advanced [19; 277; 14]. The treatment plan for controlling postoperative pain can potentially include drugs with action on central sensitization such as ketamine [184], pregabalin [162; 34], gabapentin [202] and duloxetine [106], which in the limited number of trials currently available show some efficacy, but more RCT are required to assess their utility in treating acute postoperative pain or in reducing the risk of developing chronic pain [59].

Visceral Pain Hypersensitivity Syndromes

Pain hypersensitivity is a feature of several common disorders of the gastro-intestinal tract including irritable bowel syndrome, non-cardiac chest pain and chronic pancreatitis, that all appear to have a central sensitization component. A majority of IBS patients have both rectal and somatic hypersensitivity [249]. Repetitive sigmoid stimulation in patients with IBS induces rectal hyperalgesia and viscerosomatic referral [169]. Local rectal anesthesia reduces rectal and somatic pain in irritable bowel syndrome patients, supporting the possibility that visceral hyperalgesia and secondary cutaneous hyperalgesia in irritable bowel syndrome is the result of central sensitization dynamically maintained by input from the GIT. Patients with non cardiac chest pain have esophageal hypersensitivity [195], with a reduced tolerance to repeated distension, increased size of referred pain and a greater propensity to show secondary hyperalgesia after acid infusion in their lower esophagus [69], all interpreted as reflecting the consequence of central sensitization. Chronic pancreatitis is associated with generalized deep pressure hyperalgesia [39; 174] and patients display greater degree and spatial extent secondary hyperalgesia elicited by repetitive experimental stimulation, suggesting enhanced central sensitization [67], that is reduced by a thorascopic splanchnic denervation [38], which may reflect that visceral input from the pancreas maintains the central sensitization.

In the urological tract, pain hypersensitivity is a feature of interstitial cystitis, chronic prostatitis, endometriosis, and vulvodynia, conditions whose pathophysiology and etiology is however, poorly understood. Although central sensitization has been hypothesized to contribute [137], not much data is available and few studies have been performed. Men with chronic prostatitis have though heightened pain sensitivity in the perineum [276; 239], while women with vulvodynia have an enhanced post capsaicin allodynia and secondary hyperalgesia compared to controls [84].

Co-morbidity of pain conditions characterized by pain hypersensitivity

Pain can be defined as nociceptive when it is generated by noxious stimuli, inflammatory when produced by tissue injury and/or immune cell activation, and neuropathic, when it is due to a lesion of the nervous system. What about pain conditions though, where there is no noxious stimulus, inflammation or damage to the nervous system? There are several common syndromes that present with pain hypersensitivity but no clear etiological factor, i.e. are considered “unexplained” and which might actually reflect not peripheral pathology but a primary dysfunction of the nervous system. These include fibromyalgia, tension-type headache, temporomandibular joint disease and irritable bowel syndrome, all of which may have a specific contribution to their phenotype by central sensitization, as detailed above. If a heightened sensitivity of the CNS or an increased propensity to develop central sensitization is a common feature of these syndromes, one would expect that there may be increased co-occurrence or comorbidity of the different conditions. It is also possible that an enhanced capacity to produce or maintain central sensitization is the primary defect in some of these syndromes.

In a study on almost 4,000 twins for comorbidity of chronic fatigue, low back pain, irritable bowel syndrome, chronic tension type headache, temporomandibular joint disease, major depression, panic attacks and post-traumatic stress disorder, associations were found that far exceeded those expected by chance, and the conclusion was that these conditions share a common etiology [199]. Another large epidemiological study on 44,000 individuals including twins for comorbidity with chronic widespread pain, found co-occurrence with chronic fatigue, joint pain, depressive symptoms, and irritable bowel syndrome, leading to the conclusion that associations between chronic widespread pain and its comorbidities may include genetic factors [127]. Yet another study on 2299 subjects for four unexplained syndromes; chronic wide spread pain, chronic orofacial pain, irritable bowel and chronic fatigue, again found that the occurrence of multiple syndromes was greater than expected by chance [2]. These epidemiological findings strongly suggest that there may be a common mechanistic basis for these diverse conditions, and that it may have a hereditary component.

Smaller studies have found comorbidity between fibromyalgia and the following conditions: migraine in females but not males [111], primary headache [64], chronic fatigue symptom [89], systemic lupus erythematosus [213], irritable bowel syndrome [144], rheumatoid arthritis [183], the premenstrual syndrome [3], chronic urticaria [235] and cervical myofascial pain syndrome [40]. Comorbidity has been shown also for back pain and temporomandibular disorders [248], migraine and temporomandibular disorders [91], irritable bowel syndrome and functional dyspepsia, fibromyalgia and chronic pelvic pain [185], and finally between migraine and irritable bowel syndrome, chronic fatigue and fibromyalgia [232]. There is also an overlap between urological disorders like chronic pelvic pain, interstitial cystitis, painful bladder syndrome, chronic prostatitis and vulvodynia with fibromyalgia, chronic fatigue, temporomandibular disorders and irritable bowel syndrome [187], and more specifically between vulvodynia, fibromyalgia and irritable bowel syndrome [10].

The overwhelming conclusion from these diverse epidemiological studies is that chronic pain hypersensitivity in the absence of inflammation or nerve damage results in apparently phenotypically different syndromes depending on the tissue/organs affected. However, the overall similarity of the sensitivity changes may reflect a common contribution of central sensitization, and this may account for the unexpectedly high comorbid rate of the apparently different different syndromes. To test if there are indeed central sensitization syndromes, we will need a clear set of diagnostic criteria and biomarkers for the phenomenon. If this hypothesis is correct, the implications may be that treatment strategies targeted at normalizing hyperexcitability in the CNS may have a shared efficacy for the different manifestations of the central sensitization syndrome.

Conclusions

Clinical pain is not simply the consequence of a “switching on” of the “pain system” in the periphery by a particular pathology, but instead reflects to a substantial extent, the state of excitability of central nociceptive circuits. The induction of activity-dependent increases in synaptic function in these circuits, triggered and maintained by dynamic nociceptor inputs, shifts the sensitivity of the pain system such that normally innocuous inputs can activate it and the perceptual responses to noxious inputs are exaggerated, prolonged and spread widely. These sensory changes represent the manifestation of central sensitization, and extensive experimental medicine and clinical investigations over the past twenty years, have revealed it to be an important component of the pain hypersensitivity present many patients. While considerable progress has been made in teasing out the cellular and molecular mechanism responsible [148], much remains still to be learned, particularly which genetic and environmental contributors increase the risk of developing central sensitization in particular systems, exactly what triggers and sustains the phenomenon, and what is responsible in some individuals for its persistence. Nevertheless, the identification of the contribution of central sensitization to many “unexplained” clinical pain conditions has both provided a mechanistic explanation, and offered a therapeutic target.

Acknowledgments

Supported by research funds from the NIH. I thank all my colleagues whose work has over the past 25 years contributed to the study of central sensitization but particularly Alban Latremoliere for his careful reading of the MS and Christian von Hehn for making the two figures.

Footnotes

There is no conflict of interest.

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

1. Aasvang EK, Brandsborg B, Jensen TS, Kehlet H. Heterogeneous sensory processing in persistent postherniotomy pain. Pain. 2010;150(2):237–242. [PubMed]
2. Aggarwal VR, McBeth J, Zakrzewska JM, Lunt M, Macfarlane GJ. The epidemiology of chronic syndromes that are frequently unexplained: do they have common associated factors? Int J Epidemiol. 2006;35(2):468–476. [PubMed]
3. Amital D, Herskovitz C, Fostick L, Silberman A, Doron Y, Zohar J, Itsekson A, Zolti M, Rubinow A, Amital H. The premenstrual syndrome and fibromyalgia--similarities and common features. Clin Rev Allergy Immunol. 2010;38(2–3):107–115. [PubMed]
4. Amr YM, Yousef AA. Evaluation of efficacy of the perioperative administration of Venlafaxine or gabapentin on acute and chronic postmastectomy pain. Clin J Pain. 2010;26(5):381–385. [PubMed]
5. Amr YM, Yousef AA, Alzeftawy AE, Messbah WI, Saber AM. Effect of preincisional epidural fentanyl and bupivacaine on postthoracotomy pain and pulmonary function. Ann Thorac Surg. 2010;89(2):381–385.[PubMed]
6. Andersen OK, Felsby S, Nicolaisen L, Bjerring P, Jensen TS, Arendt-Nielsen L. The effect of Ketamine on stimulation of primary and secondary hyperalgesic areas induced by capsaicin--a double-blind, placebo-controlled, human experimental study. Pain. 1996;66(1):51–62. [PubMed]
7. Arendt-Nielsen L, Nie H, Laursen MB, Laursen BS, Madeleine P, Simonsen OH, Graven-Nielsen T. Sensitization in patients with painful knee osteoarthritis. Pain. 2010;149(3):573–581. [PubMed]
8. Arendt-Nielsen L, Petersen-Felix S, Fischer M, Bak P, Bjerring P, Zbinden AM. The effect of N-methyl-D-aspartate antagonist (ketamine) on single and repeated nociceptive stimuli: a placebo-controlled experimental human study. Anesth Analg. 1995;81(1):63–68. [PubMed]
9. Arning K, Baron R. Evaluation of symptom heterogeneity in neuropathic pain using assessments of sensory functions. Neurotherapeutics. 2009;6(4):738–748. [PMC free article] [PubMed]
10. Arnold LD, Bachmann GA, Rosen R, Kelly S, Rhoads GG. Vulvodynia: characteristics and associations with comorbidities and quality of life. Obstet Gynecol. 2006;107(3):617–624.[PMC free article] [PubMed]
11. Attal N, Bouhassira D, Gautron M, Vaillant JN, Mitry E, Lepere C, Rougier P, Guirimand F. Thermal hyperalgesia as a marker of oxaliplatin neurotoxicity: a prospective quantified sensory assessment study. Pain. 2009;144(3):245–252. [PubMed]
12. Baba H, Ji RR, Kohno T, Moore KA, Ataka T, Wakai A, Okamoto M, Woolf CJ. Removal of GABAergic inhibition facilitates polysynaptic A fiber-mediated excitatory transmission to the superficial spinal dorsal horn. Mol Cell Neurosci. 2003;24(3):818–830. [PubMed]
13. Bajaj P, Graven-Nielsen T, Arendt-Nielsen L. Osteoarthritis and its association with muscle hyperalgesia: an experimental controlled study. Pain. 2001;93(2):107–114. [PubMed]
14. Bamigboye AA, Hofmeyr GJ. Local anaesthetic wound infiltration and abdominal nerves block during caesarean section for postoperative pain relief. Cochrane Database Syst Rev. 2009;(3):CD006954.[PubMed]
15. Bardin L, Gregoire S, Aliaga M, Malfetes N, Vitton O, Ladure P, Newman-Tancredi A, Depoortere R. Comparison of milnacipran, duloxetine and pregabalin in the formalin pain test and in a model of stress-induced ultrasonic vocalizations in rats. Neurosci Res. 2010;66(2):135–140. [PubMed]
16. Baron R, Baron Y, Disbrow E, Roberts TP. Brain processing of capsaicin-induced secondary hyperalgesia: a functional MRI study. Neurology. 1999;53(3):548–557. [PubMed]
17. Baron R, Baron Y, Disbrow E, Roberts TP. Activation of the somatosensory cortex during Abeta-fiber mediated hyperalgesia. A MSI study. Brain Res. 2000;871(1):75–82. [PubMed]
18. Basbaum AI, Fields HL. Endogenous pain control systems: brainstem spinal pathways and endorphin circuitry. Annu Rev Neurosci. 1984;7:309–338. [PubMed]
19. Bell RF, Dahl JB, Moore RA, Kalso E. Perioperative ketamine for acute postoperative pain. Cochrane Database Syst Rev. 2006;(1):CD004603. [PubMed]
20. Bessou P, Perl ER. Response of cutaneous sensory units with unmyelinated fibers to noxious stimuli. J Neurophysiol. 1969;32(6):1025–1043. [PubMed]
21. Beswick FB, Evanson JM. The heterosynaptic activation of motoneurones during post-tetanic potentiation. J Physiol. 1955;128(1):89–98. [PMC free article] [PubMed]
22. Binshtok AM, Wang H, Zimmermann K, Amaya F, Vardeh D, Shi L, Brenner GJ, Ji RR, Bean BP, Woolf CJ, Samad TA. Nociceptors are interleukin-1beta sensors. J Neurosci. 2008;28(52):14062–14073.[PMC free article] [PubMed]
23. Bishop T, Marchand F, Young AR, Lewin GR, McMahon SB. Ultraviolet-B-induced mechanical hyperalgesia: A role for peripheral sensitisation. Pain. 2010 [PubMed]
24. Biurrun Manresa JA, Morch CD, Andersen OK. Long-term facilitation of nociceptive withdrawal reflexes following low-frequency conditioning electrical stimulation: A new model for central sensitization in humans. Eur J Pain. 2010 [PubMed]
25. Bonica JJ. Anesthesiology in the People’s Republic of China. Anesthesiology. 1974;40(2):175–186.[PubMed]
26. Bradley LA, Kersh BC, DeBerry JJ, Deutsch G, Alarcon GA, McLain DA. Lessons from fibromyalgia: abnormal pain sensitivity in knee osteoarthritis. Novartis Found Symp. 2004;260:258–270. discussion 270–259. [PubMed]
27. Brock C, Andresen T, Frokjaer JB, Gale J, Olesen AE, Arendt-Nielsen L, Drewes AM. Central pain mechanisms following combined acid and capsaicin perfusion of the human oesophagus. Eur J Pain. 2010;14(3):273–281. [PubMed]
28. Brown AG, Iggo A. A quantitative study of cutaneous receptors and afferent fibres in the cat and rabbit. J Physiol. 1967;193(3):707–733. [PMC free article] [PubMed]
29. Brunelli M, Castellucci V, Kandel ER. Synaptic facilitation and behavioral sensitization in Aplysia: possible role of serotonin and cyclic AMP. Science. 1976;194(4270):1178–1181. [PubMed]
30. Buchgreitz L, Lyngberg AC, Bendtsen L, Jensen R. Frequency of headache is related to sensitization: a population study. Pain. 2006;123(1–2):19–27. [PubMed]
31. Buchgreitz L, Lyngberg AC, Bendtsen L, Jensen R. Increased prevalence of tension-type headache over a 12-year period is related to increased pain sensitivity. A population study. Cephalalgia. 2007;27(2):145–152. [PubMed]
32. Buchgreitz L, Lyngberg AC, Bendtsen L, Jensen R. Increased pain sensitivity is not a risk factor but a consequence of frequent headache: a population-based follow-up study. Pain. 2008;137(3):623–630.[PubMed]
33. Burgess PR, Perl ER. Myelinated afferent fibres responding specifically to noxious stimulation of the skin. J Physiol. 1967;190(3):541–562. [PMC free article] [PubMed]
34. Burke SM, Shorten GD. Perioperative pregabalin improves pain and functional outcomes 3 months after lumbar discectomy. Anesth Analg. 2010;110(4):1180–1185. [PubMed]
35. Burns D, Hill L, Essandoh M, Jarzembowski TM, Schuler HG, Janicki PK. Effect of valdecoxib pretreatment on pain and secondary hyperalgesia: a randomized controlled trial in healthy volunteers [ISRCTN05282752, NCT00260325] BMC Anesthesiol. 2006;6:3. [PMC free article] [PubMed]
36. Burstein R, Cutrer MF, Yarnitsky D. The development of cutaneous allodynia during a migraine attack clinical evidence for the sequential recruitment of spinal and supraspinal nociceptive neurons in migraine. Brain. 2000;123 ( Pt 8):1703–1709. [PubMed]
37. Burstein R, Yarnitsky D, Goor-Aryeh I, Ransil BJ, Bajwa ZH. An association between migraine and cutaneous allodynia. Ann Neurol. 2000;47(5):614–624. [PubMed]
38. Buscher HC, van Goor H, Wilder-Smith OH. Effect of thoracoscopic splanchnic denervation on pain processing in chronic pancreatitis patients. Eur J Pain. 2007;11(4):437–443. [PubMed]
39. Buscher HC, Wilder-Smith OH, van Goor H. Chronic pancreatitis patients show hyperalgesia of central origin: a pilot study. Eur J Pain. 2006;10(4):363–370. [PubMed]
40. Cakit BD, Taskin S, Nacir B, Unlu I, Genc H, Erdem HR. Comorbidity of fibromyalgia and cervical myofascial pain syndrome. Clin Rheumatol. 2010;29(4):405–411. [PubMed]
41. Campbell JN, Khan AA, Meyer RA, Raja SN. Responses to heat of C-fiber nociceptors in monkey are altered by injury in the receptive field but not by adjacent injury. Pain. 1988;32(3):327–332. [PubMed]
42. Campbell JN, Raja SN, Meyer RA, Mackinnon SE. Myelinated afferents signal the hyperalgesia associated with nerve injury. Pain. 1988;32(1):89–94. [PubMed]
43. Chacur M, Lambertz D, Hoheisel U, Mense S. Role of spinal microglia in myositis-induced central sensitisation: an immunohistochemical and behavioural study in rats. Eur J Pain. 2009;13(9):915–923.[PubMed]
44. Chang YW, Waxman SG. Minocycline Attenuates Mechanical Allodynia and Central Sensitization Following Peripheral Second-Degree Burn Injury. J Pain. 2010 [PubMed]
45. Chappell AS, Ossanna MJ, Liu-Seifert H, Iyengar S, Skljarevski V, Li LC, Bennett RM, Collins H. Duloxetine, a centrally acting analgesic, in the treatment of patients with osteoarthritis knee pain: a 13-week, randomized, placebo-controlled trial. Pain. 2009;146(3):253–260. [PubMed]
46. Chen Z, Muscoli C, Doyle T, Bryant L, Cuzzocrea S, Mollace V, Mastroianni R, Masini E, Salvemini D. NMDA-receptor activation and nitroxidative regulation of the glutamatergic pathway during nociceptive processing. Pain. 2010;149(1):100–106. [PMC free article] [PubMed]
47. Chiang CY, Li Z, Dostrovsky JO, Sessle BJ. Central sensitization in medullary dorsal horn involves gap junctions and hemichannels. Neuroreport. 2010;21(3):233–237. [PMC free article] [PubMed]
48. Chiechio S, Zammataro M, Morales ME, Busceti CL, Drago F, Gereau RWt, Copani A, Nicoletti F. Epigenetic modulation of mGlu2 receptors by histone deacetylase inhibitors in the treatment of inflammatory pain. Mol Pharmacol. 2009;75(5):1014–1020. [PubMed]
49. Chizh BA, Gohring M, Troster A, Quartey GK, Schmelz M, Koppert W. Effects of oral pregabalin and aprepitant on pain and central sensitization in the electrical hyperalgesia model in human volunteers. Br J Anaesth. 2007;98(2):246–254. [PubMed]
50. Cook AJ, Woolf CJ, Wall PD. Prolonged C-fibre mediated facilitation of the flexion reflex in the rat is not due to changes in afferent terminal or motoneurone excitability. Neurosci Lett. 1986;70(1):91–96.[PubMed]
51. Cook AJ, Woolf CJ, Wall PD, McMahon SB. Dynamic receptive field plasticity in rat spinal cord dorsal horn following C-primary afferent input. Nature. 1987;325(7000):151–153. [PubMed]
52. Cooke L, Eliasziw M, Becker WJ. Cutaneous allodynia in transformed migraine patients. Headache. 2007;47(4):531–539. [PubMed]
53. Costigan M, Scholz J, Woolf CJ. Neuropathic pain: a maladaptive response of the nervous system to damage. Annu Rev Neurosci. 2009;32:1–32. [PMC free article] [PubMed]
54. Coughlin SM, Karanicolas PJ, Emmerton-Coughlin HM, Kanbur B, Kanbur S, Colquhoun PH. Better late than never? Impact of local analgesia timing on postoperative pain in laparoscopic surgery: a systematic review and metaanalysis. Surg Endosc. 2010 [PubMed]
55. Cruccu G, Sommer C, Anand P, Attal N, Baron R, Garcia-Larrea L, Haanpaa M, Jensen TS, Serra J, Treede RD. EFNS guidelines on neuropathic pain assessment: revised 2009. Eur J Neurol. 2010;17(8):1010–1018. [PubMed]
56. Cuadrado ML, Young WB, Fernandez-de-las-Penas C, Arias JA, Pareja JA. Migrainous corpalgia: body pain and allodynia associated with migraine attacks. Cephalalgia. 2008;28(1):87–91. [PubMed]
57. Curatolo M, Petersen-Felix S, Arendt-Nielsen L, Giani C, Zbinden AM, Radanov BP. Central hypersensitivity in chronic pain after whiplash injury. Clin J Pain. 2001;17(4):306–315. [PubMed]
58. Dahl JB, Brennum J, Arendt-Nielsen L, Jensen TS, Kehlet H. The effect of pre- versus postinjury infiltration with lidocaine on thermal and mechanical hyperalgesia after heat injury to the skin. Pain. 1993;53(1):43–51. [PubMed]
59. Dahl JB, Mathiesen O, Kehlet H. An expert opinion on postoperative pain management, with special reference to new developments. Expert Opin Pharmacother. 2010 [PubMed]
60. Dahl JB, Moiniche S. Pre-emptive analgesia. Br Med Bull. 2004;71:13–27. [PubMed]
61. de la Llave-Rincon AI, Fernandez-de-las-Penas C, Fernandez-Carnero J, Padua L, Arendt-Nielsen L, Pareja JA. Bilateral hand/wrist heat and cold hyperalgesia, but not hypoesthesia, in unilateral carpal tunnel syndrome. Exp Brain Res. 2009;198(4):455–463. [PubMed]
62. de Tommaso M, Guido M, Libro G, Losito L, Sciruicchio V, Monetti C, Puca F. Abnormal brain processing of cutaneous pain in migraine patients during the attack. Neurosci Lett. 2002;333(1):29–32.[PubMed]
63. de Tommaso M, Libro G, Guido M, Sciruicchio V, Losito L, Puca F. Heat pain thresholds and cerebral event-related potentials following painful CO2 laser stimulation in chronic tension-type headache. Pain. 2003;104(1–2):111–119. [PubMed]
64. de Tommaso M, Sardaro M, Serpino C, Costantini F, Vecchio E, Prudenzano MP, Lamberti P, Livrea P. Fibromyalgia comorbidity in primary headaches. Cephalalgia. 2009;29(4):453–464. [PubMed]
65. Desmeules JA, Cedraschi C, Rapiti E, Baumgartner E, Finckh A, Cohen P, Dayer P, Vischer TL. Neurophysiologic evidence for a central sensitization in patients with fibromyalgia. Arthritis Rheum. 2003;48(5):1420–1429. [PubMed]
66. Diers M, Koeppe C, Yilmaz P, Thieme K, Markela-Lerenc J, Schiltenwolf M, van Ackern K, Flor H. Pain ratings and somatosensory evoked responses to repetitive intramuscular and intracutaneous stimulation in fibromyalgia syndrome. J Clin Neurophysiol. 2008;25(3):153–160. [PubMed]
67. Dimcevski G, Staahl C, Andersen SD, Thorsgaard N, Funch-Jensen P, Arendt-Nielsen L, Drewes AM. Assessment of experimental pain from skin, muscle, and esophagus in patients with chronic pancreatitis. Pancreas. 2007;35(1):22–29. [PubMed]
68. Dirks J, Moiniche S, Hilsted KL, Dahl JB. Mechanisms of postoperative pain: clinical indications for a contribution of central neuronal sensitization. Anesthesiology. 2002;97(6):1591–1596. [PubMed]
69. Drewes AM, Pedersen J, Reddy H, Rasmussen K, Funch-Jensen P, Arendt-Nielsen L, Gregersen H. Central sensitization in patients with non-cardiac chest pain: a clinical experimental study. Scand J Gastroenterol. 2006;41(6):640–649. [PubMed]
70. Duale C, Sibaud F, Guastella V, Vallet L, Gimbert YA, Taheri H, Filaire M, Schoeffler P, Dubray C. Perioperative ketamine does not prevent chronic pain after thoracotomy. Eur J Pain. 2009;13(5):497–505.[PubMed]
71. Dullenkopf A, Muller R, Dillmann F, Wiedemeier P, Hegi TR, Gautschi S. An intraoperative pre-incision single dose of intravenous ketamine does not have an effect on postoperative analgesic requirements under clinical conditions. Anaesth Intensive Care. 2009;37(5):753–757. [PubMed]
72. Dworkin RH, O’Connor AB, Audette J, Baron R, Gourlay GK, Haanpaa ML, Kent JL, Krane EJ, Lebel AA, Levy RM, Mackey SC, Mayer J, Miaskowski C, Raja SN, Rice AS, Schmader KE, Stacey B, Stanos S, Treede RD, Turk DC, Walco GA, Wells CD. Recommendations for the pharmacological management of neuropathic pain: an overview and literature update. Mayo Clin Proc. 2010;85(3 Suppl):S3–14.[PMC free article] [PubMed]
73. Eichenberger U, Neff F, Sveticic G, Bjorgo S, Petersen-Felix S, Arendt-Nielsen L, Curatolo M. Chronic phantom limb pain: the effects of calcitonin, ketamine, and their combination on pain and sensory thresholds. Anesth Analg. 2008;106(4):1265–1273. table of contents. [PubMed]
74. Estacion M, Harty TP, Choi JS, Tyrrell L, Dib-Hajj SD, Waxman SG. A sodium channel gene SCN9A polymorphism that increases nociceptor excitability. Ann Neurol. 2009;66(6):862–866. [PubMed]
75. Fernandez-Carnero J, Fernandez-de-Las-Penas C, de la Llave-Rincon AI, Ge HY, Arendt-Nielsen L. Widespread mechanical pain hypersensitivity as sign of central sensitization in unilateral epicondylalgia: a blinded, controlled study. Clin J Pain. 2009;25(7):555–561. [PubMed]
76. Fernandez-de-las-Penas C, de la Llave-Rincon AI, Fernandez-Carnero J, Cuadrado ML, Arendt-Nielsen L, Pareja JA. Bilateral widespread mechanical pain sensitivity in carpal tunnel syndrome: evidence of central processing in unilateral neuropathy. Brain. 2009;132(Pt 6):1472–1479. [PubMed]
77. Fernandez-de-Las-Penas C, Galan-Del-Rio F, Alonso-Blanco C, Jimenez-Garcia R, Arendt-Nielsen L, Svensson P. Referred Pain from Muscle Trigger Points in the Masticatory and Neck-Shoulder Musculature in Women with Temporomandibular Disoders. J Pain. 2010 [PubMed]
78. Fernandez-de-las-Penas C, Galan-del-Rio F, Fernandez-Carnero J, Pesquera J, Arendt-Nielsen L, Svensson P. Bilateral widespread mechanical pain sensitivity in women with myofascial temporomandibular disorder: evidence of impairment in central nociceptive processing. J Pain. 2009;10(11):1170–1178. [PubMed]
79. Fernandez-de-Las-Penas C, Ge HY, Arendt-Nielsen L, Cuadrado ML, Pareja JA. The local and referred pain from myofascial trigger points in the temporalis muscle contributes to pain profile in chronic tension-type headache. Clin J Pain. 2007;23(9):786–792. [PubMed]
80. Fernandez-de-las-Penas C, Ge HY, Cuadrado ML, Madeleine P, Pareja JA, Arendt-Nielsen L. Bilateral pressure pain sensitivity mapping of the temporalis muscle in chronic tension-type headache. Headache. 2008;48(7):1067–1075. [PubMed]
81. Fernandez-de-Las-Penas C, Madeleine P, Caminero A, Cuadrado M, Arendt-Nielsen L, Pareja J. Generalized neck-shoulder hyperalgesia in chronic tension-type headache and unilateral migraine assessed by pressure pain sensitivity topographical maps of the trapezius muscle. Cephalalgia. 2009 [PubMed]
82. Fernandez-de-Las-Penas C, Madeleine P, Martinez-Perez A, Arendt-Nielsen L, Jimenez-Garcia R, Pareja JA. Pressure pain sensitivity topograhical maps reveal bilateral hyperalgesia of the hands in patients with unilateral carpal tunnel syndrome. Arthritis Care Res (Hoboken) 2010 [PubMed]
83. Filatova E, Latysheva N, Kurenkov A. Evidence of persistent central sensitization in chronic headaches: a multi-method study. J Headache Pain. 2008;9(5):295–300. [PMC free article] [PubMed]
84. Foster DC, Dworkin RH, Wood RW. Effects of intradermal foot and forearm capsaicin injections in normal and vulvodynia-afflicted women. Pain. 2005;117(1–2):128–136. [PubMed]
85. Freeman MD, Nystrom A, Centeno C. Chronic whiplash and central sensitization; an evaluation of the role of a myofascial trigger points in pain modulation. J Brachial Plex Peripher Nerve Inj. 2009;4:2.[PMC free article] [PubMed]
86. Freynhagen R, Rolke R, Baron R, Tolle TR, Rutjes AK, Schu S, Treede RD. Pseudoradicular and radicular low-back pain--a disease continuum rather than different entities? Answers from quantitative sensory testing. Pain. 2008;135(1–2):65–74. [PubMed]
87. Fuchs PN, Campbell JN, Meyer RA. Secondary hyperalgesia persists in capsaicin desensitized skin. Pain. 2000;84(2–3):141–149. [PubMed]
88. Gao YJ, Zhang L, Samad OA, Suter MR, Yasuhiko K, Xu ZZ, Park JY, Lind AL, Ma Q, Ji RR. JNK-induced MCP-1 production in spinal cord astrocytes contributes to central sensitization and neuropathic pain. J Neurosci. 2009;29(13):4096–4108. [PMC free article] [PubMed]
89. Geisser ME, Strader Donnell C, Petzke F, Gracely RH, Clauw DJ, Williams DA. Comorbid somatic symptoms and functional status in patients with fibromyalgia and chronic fatigue syndrome: sensory amplification as a common mechanism. Psychosomatics. 2008;49(3):235–242. [PubMed]
90. Gibson SJ, Littlejohn GO, Gorman MM, Helme RD, Granges G. Altered heat pain thresholds and cerebral event-related potentials following painful CO2 laser stimulation in subjects with fibromyalgia syndrome. Pain. 1994;58(2):185–193. [PubMed]
91. Goncalves DA, Bigal ME, Jales LC, Camparis CM, Speciali JG. Headache and symptoms of temporomandibular disorder: an epidemiological study. Headache. 2010;50(2):231–241. [PubMed]
92. Gottrup H, Juhl G, Kristensen AD, Lai R, Chizh BA, Brown J, Bach FW, Jensen TS. Chronic oral gabapentin reduces elements of central sensitization in human experimental hyperalgesia. Anesthesiology. 2004;101(6):1400–1408. [PubMed]
93. Gottrup H, Kristensen AD, Bach FW, Jensen TS. Aftersensations in experimental and clinical hypersensitivity. Pain. 2003;103(1–2):57–64. [PubMed]
94. Granot M, Buskila D, Granovsky Y, Sprecher E, Neumann L, Yarnitsky D. Simultaneous recording of late and ultra-late pain evoked potentials in fibromyalgia. Clin Neurophysiol. 2001;112(10):1881–1887.[PubMed]
95. Graven-Nielsen T, Arendt-Nielsen L. Assessment of mechanisms in localized and widespread musculoskeletal pain. Nat Rev Rheumatol. 2010 [PubMed]
96. Graven-Nielsen T, Aspegren Kendall S, Henriksson KG, Bengtsson M, Sorensen J, Johnson A, Gerdle B, Arendt-Nielsen L. Ketamine reduces muscle pain, temporal summation, and referred pain in fibromyalgia patients. Pain. 2000;85(3):483–491. [PubMed]
97. Griffin RS, Costigan M, Brenner GJ, Ma CH, Scholz J, Moss A, Allchorne AJ, Stahl GL, Woolf CJ. Complement induction in spinal cord microglia results in anaphylatoxin C5a-mediated pain hypersensitivity. J Neurosci. 2007;27(32):8699–8708. [PubMed]
98. Guillemin R, Ling N, Burgus R, Bloom F, Segal D. Characterization of the endorphins, novel hypothalamic and neurohypophysial peptides with opiate-like activity: evidence that they induce profound behavioral changes. Psychoneuroendocrinology. 1977;2(1):59–62. [PubMed]
99. Gwilym SE, Keltner JR, Warnaby CE, Carr AJ, Chizh B, Chessell I, Tracey I. Psychophysical and functional imaging evidence supporting the presence of central sensitization in a cohort of osteoarthritis patients. Arthritis Rheum. 2009;61(9):1226–1234. [PubMed]
100. Han JS, Adwanikar H, Li Z, Ji G, Neugebauer V. Facilitation of synaptic transmission and pain responses by CGRP in the amygdala of normal rats. Mol Pain. 2010;6(1):10. [PMC free article] [PubMed]
101. Hardy JD, Wolff HG, Goodell H. Experimental evidence on the nature of cutaneous hyperalgesia. J Clin Invest. 1950;29(1):115–140. [PMC free article] [PubMed]
102. Hariharan S, Moseley H, Kumar A, Raju S. The effect of preemptive analgesia in postoperative pain relief--a prospective double-blind randomized study. Pain Med. 2009;10(1):49–53. [PubMed]
103. Harvey RJ, Depner UB, Wassle H, Ahmadi S, Heindl C, Reinold H, Smart TG, Harvey K, Schutz B, Abo-Salem OM, Zimmer A, Poisbeau P, Welzl H, Wolfer DP, Betz H, Zeilhofer HU, Muller U. GlyR alpha3: an essential target for spinal PGE2-mediated inflammatory pain sensitization. Science. 2004;304(5672):884–887. [PubMed]
104. Hathway GJ, Vega-Avelaira D, Moss A, Ingram R, Fitzgerald M. Brief, low frequency stimulation of rat peripheral C-fibres evokes prolonged microglial-induced central sensitization in adults but not in neonates. Pain. 2009;144(1–2):110–118. [PMC free article] [PubMed]
105. Hidalgo-Lozano A, Fernandez-de-las-Penas C, Alonso-Blanco C, Ge HY, Arendt-Nielsen L, Arroyo-Morales M. Muscle trigger points and pressure pain hyperalgesia in the shoulder muscles in patients with unilateral shoulder impingement: a blinded, controlled study. Exp Brain Res. 2010;202(4):915–925.[PubMed]
106. Ho KY, Tay W, Yeo MC, Liu H, Yeo SJ, Chia SL, Lo NN. Duloxetine reduces morphine requirements after knee replacement surgery. Br J Anaesth. 2010 [PubMed]
107. Hogeweg JA, Kuis W, Huygen AC, de Jong-de vos van Steenwijk C, Bernards AT, Oostendorp RA, Helders PJ. The pain threshold in juvenile chronic arthritis. Br J Rheumatol. 1995;34(1):61–67. [PubMed]
108. Huge V, Lauchart M, Forderreuther S, Kaufhold W, Valet M, Azad SC, Beyer A, Magerl W. Interaction of hyperalgesia and sensory loss in complex regional pain syndrome type I (CRPS I) PLoS One. 2008;3(7):e2742. [PMC free article] [PubMed]
109. Hughes J, Smith TW, Kosterlitz HW, Fothergill LA, Morgan BA, Morris HR. Identification of two related pentapeptides from the brain with potent opiate agonist activity. Nature. 1975;258(5536):577–580.[PubMed]
110. Iannetti GD, Zambreanu L, Wise RG, Buchanan TJ, Huggins JP, Smart TS, Vennart W, Tracey I. Pharmacological modulation of pain-related brain activity during normal and central sensitization states in humans. Proc Natl Acad Sci U S A. 2005;102(50):18195–18200. [PMC free article] [PubMed]
111. Ifergane G, Buskila D, Simiseshvely N, Zeev K, Cohen H. Prevalence of fibromyalgia syndrome in migraine patients. Cephalalgia. 2006;26(4):451–456. [PubMed]
112. Iggo A. Cutaneous mechanoreceptors with afferent C fibres. J Physiol. 1960;152:337–353.[PMC free article] [PubMed]
113. Ikeda H, Heinke B, Ruscheweyh R, Sandkuhler J. Synaptic plasticity in spinal lamina I projection neurons that mediate hyperalgesia. Science. 2003;299(5610):1237–1240. [PubMed]
114. Ikeda H, Stark J, Fischer H, Wagner M, Drdla R, Jager T, Sandkuhler J. Synaptic amplifier of inflammatory pain in the spinal dorsal horn. Science. 2006;312(5780):1659–1662. [PubMed]
115. Ilkjaer S, Dirks J, Brennum J, Wernberg M, Dahl JB. Effect of systemic N-methyl-D-aspartate receptor antagonist (dextromethorphan) on primary and secondary hyperalgesia in humans. Br J Anaesth. 1997;79(5):600–605. [PubMed]
116. Ilkjaer S, Petersen KL, Brennum J, Wernberg M, Dahl JB. Effect of systemic N-methyl-D-aspartate receptor antagonist (ketamine) on primary and secondary hyperalgesia in humans. Br J Anaesth. 1996;76(6):829–834. [PubMed]
117. Imamura M, Imamura ST, Kaziyama HH, Targino RA, Hsing WT, de Souza LP, Cutait MM, Fregni F, Camanho GL. Impact of nervous system hyperalgesia on pain, disability, and quality of life in patients with knee osteoarthritis: a controlled analysis. Arthritis Rheum. 2008;59(10):1424–1431. [PubMed]
118. Iyengar S, Webster AA, Hemrick-Luecke SK, Xu JY, Simmons RM. Efficacy of duloxetine, a potent and balanced serotonin-norepinephrine reuptake inhibitor in persistent pain models in rats. J Pharmacol Exp Ther. 2004;311(2):576–584. [PubMed]
119. Jensen MT, Petersen KL. Gender differences in pain and secondary hyperalgesia after heat/capsaicin sensitization in healthy volunteers. J Pain. 2006;7(3):211–217. [PubMed]
120. Jensen R. Mechanisms of spontaneous tension-type headaches: an analysis of tenderness, pain thresholds and EMG. Pain. 1996;64(2):251–256. [PubMed]
121. Ji RR, Baba H, Brenner GJ, Woolf CJ. Nociceptive-specific activation of ERK in spinal neurons contributes to pain hypersensitivity. Nat Neurosci. 1999;2(12):1114–1119. [PubMed]
122. Ji RR, Kohno T, Moore KA, Woolf CJ. Central sensitization and LTP: do pain and memory share similar mechanisms? Trends Neurosci. 2003;26(12):696–705. [PubMed]
123. 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. [PubMed]
124. Jones CK, Peters SC, Shannon HE. Efficacy of duloxetine, a potent and balanced serotonergic and noradrenergic reuptake inhibitor, in inflammatory and acute pain models in rodents. J Pharmacol Exp Ther. 2005;312(2):726–732. [PubMed]
125. Jorum E, Warncke T, Stubhaug A. Cold allodynia and hyperalgesia in neuropathic pain: the effect of N-methyl-D-aspartate (NMDA) receptor antagonist ketamine--a double-blind, cross-over comparison with alfentanil and placebo. Pain. 2003;101(3):229–235. [PubMed]
126. Juhl GI, Jensen TS, Norholt SE, Svensson P. Central sensitization phenomena after third molar surgery: a quantitative sensory testing study. Eur J Pain. 2008;12(1):116–127. [PubMed]
127. Kato K, Sullivan PF, Evengard B, Pedersen NL. Chronic widespread pain and its comorbidities: a population-based study. Arch Intern Med. 2006;166(15):1649–1654. [PubMed]
128. Katz J, Cohen L, Schmid R, Chan VW, Wowk A. Postoperative morphine use and hyperalgesia are reduced by preoperative but not intraoperative epidural analgesia: implications for preemptive analgesia and the prevention of central sensitization. Anesthesiology. 2003;98(6):1449–1460. [PubMed]
129. Kawasaki Y, Kohno T, Zhuang ZY, Brenner GJ, Wang H, Van Der Meer C, Befort K, Woolf CJ, Ji RR. Ionotropic and metabotropic receptors, protein kinase A, protein kinase C, and Src contribute to C-fiber-induced ERK activation and cAMP response element-binding protein phosphorylation in dorsal horn neurons, leading to central sensitization. J Neurosci. 2004;24(38):8310–8321. [PubMed]
130. Kawasaki Y, Zhang L, Cheng JK, Ji RR. Cytokine mechanisms of central sensitization: distinct and overlapping role of interleukin-1beta, interleukin-6, and tumor necrosis factor-alpha in regulating synaptic and neuronal activity in the superficial spinal cord. J Neurosci. 2008;28(20):5189–5194. [PMC free article][PubMed]
131. Kehlet H, Jensen TS, Woolf CJ. Persistent postsurgical pain: risk factors and prevention. Lancet. 2006;367(9522):1618–1625. [PubMed]
132. Khan AA, Owatz CB, Schindler WG, Schwartz SA, Keiser K, Hargreaves KM. Measurement of mechanical allodynia and local anesthetic efficacy in patients with irreversible pulpitis and acute periradicular periodontitis. J Endod. 2007;33(7):796–799. [PubMed]
133. Kim YH, Park CK, Back SK, Lee CJ, Hwang SJ, Bae YC, Na HS, Kim JS, Jung SJ, Oh SB. Membrane-delimited coupling of TRPV1 and mGluR5 on presynaptic terminals of nociceptive neurons. J Neurosci. 2009;29(32):10000–10009. [PubMed]
134. Kissin I. Preemptive analgesia: problems with assessment of clinical significance. Methods Mol Biol. 2010;617:475–482. [PubMed]
135. Kitaj MB, Klink M. Pain thresholds in daily transformed migraine versus episodic migraine headache patients. Headache. 2005;45(8):992–998. [PubMed]
136. Klein T, Stahn S, Magerl W, Treede RD. The role of heterosynaptic facilitation in long-term potentiation (LTP) of human pain sensation. Pain. 2008;139(3):507–519. [PubMed]
137. Klumpp DJ, Rudick CN. Summation model of pelvic pain in interstitial cystitis. Nat Clin Pract Urol. 2008;5(9):494–500. [PubMed]
138. Kocher L, Anton F, Reeh PW, Handwerker HO. The effect of carrageenan-induced inflammation on the sensitivity of unmyelinated skin nociceptors in the rat. Pain. 1987;29(3):363–373. [PubMed]
139. Koltzenburg M, Lundberg LE, Torebjork HE. Dynamic and static components of mechanical hyperalgesia in human hairy skin. Pain. 1992;51(2):207–219. [PubMed]
140. Koltzenburg M, Torebjork HE, Wahren LK. Nociceptor modulated central sensitization causes mechanical hyperalgesia in acute chemogenic and chronic neuropathic pain. Brain. 1994;117 ( Pt 3):579–591. [PubMed]
141. Konttinen YT, Honkanen VE, Gronblad M, Keinonen M, Santavirta N, Santavirta S. The relation of extraarticular tenderness to inflammatory joint disease and personality in patients with rheumatoid arthritis. J Rheumatol. 1992;19(6):851–855. [PubMed]
142. Koppert W, Dern SK, Sittl R, Albrecht S, Schuttler J, Schmelz M. A new model of electrically evoked pain and hyperalgesia in human skin: the effects of intravenous alfentanil, S(+)-ketamine, and lidocaine. Anesthesiology. 2001;95(2):395–402. [PubMed]
143. Kupers R, Schneider FC, Christensen R, Naert A, Husted H, Paulson OB, Kehlet H. No evidence for generalized increased postoperative responsiveness to pain: a combined behavioral and serial functional magnetic resonance imaging study. Anesth Analg. 2009;109(2):600–606. [PubMed]
144. Kurland JE, Coyle WJ, Winkler A, Zable E. Prevalence of irritable bowel syndrome and depression in fibromyalgia. Dig Dis Sci. 2006;51(3):454–460. [PubMed]
145. LaMotte RH, Shain CN, Simone DA, Tsai EF. Neurogenic hyperalgesia: psychophysical studies of underlying mechanisms. J Neurophysiol. 1991;66(1):190–211. [PubMed]
146. LaMotte RH, Thalhammer JG, Torebjork HE, Robinson CJ. Peripheral neural mechanisms of cutaneous hyperalgesia following mild injury by heat. J Neurosci. 1982;2(6):765–781. [PubMed]
147. Latremoliere A, Woolf CJ. Central sensitization: a generator of pain hypersensitivity by central neural plasticity. J Pain. 2009;10(9):895–926. [PMC free article] [PubMed]
148. Latremoliere A, Woolf CJ. Synaptic plasticity and central sensitization: author reply. J Pain. 2010;11(8):801–803. [PubMed]
149. Lavand’homme P, De Kock M. The use of intraoperative epidural or spinal analgesia modulates postoperative hyperalgesia and reduces residual pain after major abdominal surgery. Acta Anaesthesiol Belg. 2006;57(4):373–379. [PubMed]
150. Le Bars D, Chitour D, Kraus E, Clot AM, Dickenson AH, Besson JM. The effect of systemic morphine upon diffuse noxious inhibitory controls (DNIC) in the rat: evidence for a lifting of certain descending inhibitory controls of dorsal horn convergent neurones. Brain Res. 1981;215(1–2):257–274.[PubMed]
151. LeBlanc BW, Iwata M, Mallon AP, Rupasinghe CN, Goebel DJ, Marshall J, Spaller MR, Saab CY. A cyclic peptide targeted against PSD-95 blocks central sensitization and attenuates thermal hyperalgesia. Neuroscience. 2010;167(2):490–500. [PMC free article] [PubMed]
152. Lee KY, Chung K, Chung JM. Involvement of reactive oxygen species in long-term potentiation in the spinal cord dorsal horn. J Neurophysiol. 2010;103(1):382–391. [PMC free article] [PubMed]
153. Lee MC, Zambreanu L, Menon DK, Tracey I. Identifying brain activity specifically related to the maintenance and perceptual consequence of central sensitization in humans. J Neurosci. 2008;28(45):11642–11649. [PubMed]
154. Li J, Baccei ML. Excitatory synapses in the rat superficial dorsal horn are strengthened following peripheral inflammation during early postnatal development. Pain. 2009;143(1–2):56–64. [PubMed]
155. Lloyd DP. Post-tetanic potentiation of response in monosynaptic reflex pathways of the spinal cord. J Gen Physiol. 1949;33(2):147–170. [PMC free article] [PubMed]
156. Lorenz J, Grasedyck K, Bromm B. Middle and long latency somatosensory evoked potentials after painful laser stimulation in patients with fibromyalgia syndrome. Electroencephalogr Clin Neurophysiol. 1996;100(2):165–168. [PubMed]
157. Lundblad H, Kreicbergs A, Jansson KA. Prediction of persistent pain after total knee replacement for osteoarthritis. J Bone Joint Surg Br. 2008;90(2):166–171. [PubMed]
158. Magerl W, Wilk SH, Treede RD. Secondary hyperalgesia and perceptual wind-up following intradermal injection of capsaicin in humans. Pain. 1998;74(2–3):257–268. [PubMed]
159. Maihofner C, Jesberger F, Seifert F, Kaltenhauser M. Cortical processing of mechanical hyperalgesia: a MEG study. Eur J Pain. 2010;14(1):64–70. [PubMed]
160. Mannion RJ, Costigan M, Decosterd I, Amaya F, Ma QP, Holstege JC, Ji RR, Acheson A, Lindsay RM, Wilkinson GA, Woolf CJ. Neurotrophins: peripherally and centrally acting modulators of tactile stimulus-induced inflammatory pain hypersensitivity. Proc Natl Acad Sci U S A. 1999;96(16):9385–9390.[PMC free article] [PubMed]
161. Mathew NT, Kailasam J, Seifert T. Clinical recognition of allodynia in migraine. Neurology. 2004;63(5):848–852. [PubMed]
162. Mathiesen O, Rasmussen ML, Dierking G, Lech K, Hilsted KL, Fomsgaard JS, Lose G, Dahl JB. Pregabalin and dexamethasone in combination with paracetamol for postoperative pain control after abdominal hysterectomy. A randomized clinical trial. Acta Anaesthesiol Scand. 2009;53(2):227–235.[PubMed]
163. Melzack R, Wall PD. Pain mechanisms: a new theory. Science. 1965;150(699):971–979. [PubMed]
164. Mendell LM, Wall PD. Responses of Single Dorsal Cord Cells to Peripheral Cutaneous Unmyelinated Fibres. Nature. 1965;206:97–99. [PubMed]
165. Miraucourt LS, Moisset X, Dallel R, Voisin DL. Glycine inhibitory dysfunction induces a selectively dynamic, morphine-resistant, and neurokinin 1 receptor- independent mechanical allodynia. J Neurosci. 2009;29(8):2519–2527. [PubMed]
166. Mohn C, Vassend O, Knardahl S. Experimental pain sensitivity in women with temporomandibular disorders and pain-free controls: the relationship to orofacial muscular contraction and cardiovascular responses. Clin J Pain. 2008;24(4):343–352. [PubMed]
167. Moiniche S, Kehlet H, Dahl JB. A qualitative and quantitative systematic review of preemptive analgesia for postoperative pain relief: the role of timing of analgesia. Anesthesiology. 2002;96(3):725–741. [PubMed]
168. Moore KA, Kohno T, Karchewski LA, Scholz J, Baba H, Woolf CJ. Partial peripheral nerve injury promotes a selective loss of GABAergic inhibition in the superficial dorsal horn of the spinal cord. J Neurosci. 2002;22(15):6724–6731. [PubMed]
169. Munakata J, Naliboff B, Harraf F, Kodner A, Lembo T, Chang L, Silverman DH, Mayer EA. Repetitive sigmoid stimulation induces rectal hyperalgesia in patients with irritable bowel syndrome. Gastroenterology. 1997;112(1):55–63. [PubMed]
170. Munro G. Pharmacological assessment of the rat formalin test utilizing the clinically used analgesic drugs gabapentin, lamotrigine, morphine, duloxetine, tramadol and ibuprofen: influence of low and high formalin concentrations. Eur J Pharmacol. 2009;605(1–3):95–102. [PubMed]
171. 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. [PubMed]
172. Norbury TA, MacGregor AJ, Urwin J, Spector TD, McMahon SB. Heritability of responses to painful stimuli in women: a classical twin study. Brain. 2007;130(Pt 11):3041–3049. [PubMed]
173. O’Neill S, Manniche C, Graven-Nielsen T, Arendt-Nielsen L. Generalized deep-tissue hyperalgesia in patients with chronic low-back pain. Eur J Pain. 2007;11(4):415–420. [PubMed]
174. Olesen SS, Brock C, Krarup AL, Funch-Jensen P, Arendt-Nielsen L, Wilder-Smith OH, Drewes AM. Descending Inhibitory Pain Modulation Is Impaired in Patients With Chronic Pancreatitis. Clin Gastroenterol Hepatol. 2010;8(8):724–730. [PubMed]
175. Park JW, Clark GT, Kim YK, Chung JW. Analysis of thermal pain sensitivity and psychological profiles in different subgroups of TMD patients. Int J Oral Maxillofac Surg. 2010 [PubMed]
176. Pedersen JL, Crawford ME, Dahl JB, Brennum J, Kehlet H. Effect of preemptive nerve block on inflammation and hyperalgesia after human thermal injury. Anesthesiology. 1996;84(5):1020–1026.[PubMed]
177. Perl ER. Myelinated afferent fibres innervating the primate skin and their response to noxious stimuli. J Physiol. 1968;197(3):593–615. [PMC free article] [PubMed]
178. Perl ER, Kumazawa T, Lynn B, Kenins P. Sensitization of high threshold receptors with unmyelinated (C) afferent fibers. Prog Brain Res. 1976;43:263–277. [PubMed]
179. Petzke F, Clauw DJ, Ambrose K, Khine A, Gracely RH. Increased pain sensitivity in fibromyalgia: effects of stimulus type and mode of presentation. Pain. 2003;105(3):403–413. [PubMed]
180. Prescott SA, Sejnowski TJ, De Koninck Y. Reduction of anion reversal potential subverts the inhibitory control of firing rate in spinal lamina I neurons: towards a biophysical basis for neuropathic pain. Mol Pain. 2006;2:32. [PMC free article] [PubMed]
181. Rabben T. Effects of the NMDA receptor antagonist ketamine in electrically induced A delta-fiber pain. Methods Find Exp Clin Pharmacol. 2000;22(3):185–189. [PubMed]
182. Randic M, Jiang MC, Cerne R. Long-term potentiation and long-term depression of primary afferent neurotransmission in the rat spinal cord. J Neurosci. 1993;13(12):5228–5241. [PubMed]
183. Ranzolin A, Brenol JC, Bredemeier M, Guarienti J, Rizzatti M, Feldman D, Xavier RM. Association of concomitant fibromyalgia with worse disease activity score in 28 joints, health assessment questionnaire, and short form 36 scores in patients with rheumatoid arthritis. Arthritis Rheum. 2009;61(6):794–800. [PubMed]
184. Remerand F, Le Tendre C, Baud A, Couvret C, Pourrat X, Favard L, Laffon M, Fusciardi J. The early and delayed analgesic effects of ketamine after total hip arthroplasty: a prospective, randomized, controlled, double-blind study. Anesth Analg. 2009;109(6):1963–1971. [PubMed]
185. Riedl A, Schmidtmann M, Stengel A, Goebel M, Wisser AS, Klapp BF, Monnikes H. Somatic comorbidities of irritable bowel syndrome: a systematic analysis. J Psychosom Res. 2008;64(6):573–582.[PubMed]
186. Rivera-Arconada I, Lopez-Garcia JA. Changes in membrane excitability and potassium currents in sensitized dorsal horn neurons of mice pups. J Neurosci. 2010;30(15):5376–5383. [PubMed]
187. Rodriguez MA, Afari N, Buchwald DS. Evidence for overlap between urological and nonurological unexplained clinical conditions. J Urol. 2009;182(5):2123–2131. [PMC free article] [PubMed]
188. Rolke R, Baron R, Maier C, Tolle TR, Treede RD, Beyer A, Binder A, Birbaumer N, Birklein F, Botefur IC, Braune S, Flor H, Huge V, Klug R, Landwehrmeyer GB, Magerl W, Maihofner C, Rolko C, Schaub C, Scherens A, Sprenger T, Valet M, Wasserka B. Quantitative sensory testing in the German Research Network on Neuropathic Pain (DFNS): standardized protocol and reference values. Pain. 2006;123(3):231–243. [PubMed]
189. Samad TA, Moore KA, Sapirstein A, Billet S, Allchorne A, Poole S, Bonventre JV, Woolf CJ. Interleukin-1beta-mediated induction of Cox-2 in the CNS contributes to inflammatory pain hypersensitivity. Nature. 2001;410(6827):471–475. [PubMed]
190. Sandkuhler J. Understanding LTP in pain pathways. Mol Pain. 2007;3:9. [PMC free article] [PubMed]
191. Sandrini G, Antonaci F, Lanfranchi S, Milanov I, Danilov A, Nappi G. Asymmetrical reduction of the nociceptive flexion reflex threshold in cluster headache. Cephalalgia. 2000;20(7):647–652. [PubMed]
192. Sang CN, Gracely RH, Max MB, Bennett GJ. Capsaicin-evoked mechanical allodynia and hyperalgesia cross nerve territories. Evidence for a central mechanism. Anesthesiology. 1996;85(3):491–496. [PubMed]
193. Sarkar S, Aziz Q, Woolf CJ, Hobson AR, Thompson DG. Contribution of central sensitisation to the development of non-cardiac chest pain. Lancet. 2000;356(9236):1154–1159. [PubMed]
194. Sarkar S, Hobson AR, Furlong PL, Woolf CJ, Thompson DG, Aziz Q. Central neural mechanisms mediating human visceral hypersensitivity. Am J Physiol Gastrointest Liver Physiol. 2001;281(5):G1196–1202. [PubMed]
195. Sarkar S, Thompson DG, Woolf CJ, Hobson AR, Millane T, Aziz Q. Patients with chest pain and occult gastroesophageal reflux demonstrate visceral pain hypersensitivity which may be partially responsive to acid suppression. Am J Gastroenterol. 2004;99(10):1998–2006. [PubMed]
196. Sarkar S, Woolf CJ, Hobson AR, Thompson DG, Aziz Q. Perceptual wind-up in the human oesophagus is enhanced by central sensitisation. Gut. 2006;55(7):920–925. [PMC free article] [PubMed]
197. Scholz J, Mannion RJ, Hord DE, Griffin RS, Rawal B, Zheng H, Scoffings D, Phillips A, Guo J, Laing RJ, Abdi S, Decosterd I, Woolf CJ. A novel tool for the assessment of pain: validation in low back pain. PLoS Med. 2009;6(4):e1000047. [PMC free article] [PubMed]
198. Schulte H, Sollevi A, Segerdahl M. The distribution of hyperaemia induced by skin burn injury is not correlated with the development of secondary punctate hyperalgesia. J Pain. 2004;5(4):212–217. [PubMed]
199. Schur EA, Afari N, Furberg H, Olarte M, Goldberg J, Sullivan PF, Buchwald D. Feeling bad in more ways than one: comorbidity patterns of medically unexplained and psychiatric conditions. J Gen Intern Med. 2007;22(6):818–821. [PMC free article] [PubMed]
200. Schwartzman RJ, Alexander GM, Grothusen JR, Paylor T, Reichenberger E, Perreault M. Outpatient intravenous ketamine for the treatment of complex regional pain syndrome: a double-blind placebo controlled study. Pain. 2009;147(1–3):107–115. [PubMed]
201. Segerdahl M. Multiple dose gabapentin attenuates cutaneous pain and central sensitisation but not muscle pain in healthy volunteers. Pain. 2006;125(1–2):158–164. [PubMed]
202. Sen H, Sizlan A, Yanarates O, Emirkadi H, Ozkan S, Dagli G, Turan A. A comparison of gabapentin and ketamine in acute and chronic pain after hysterectomy. Anesth Analg. 2009;109(5):1645–1650.[PubMed]
203. Sesana L, Caprioli E, Marcazzan GM. Long period study of outdoor radon concentration in Milan and correlation between its temporal variations and dispersion properties of atmosphere. J Environ Radioact. 2003;65(2):147–160. [PubMed]
204. Sethna NF, Liu M, Gracely R, Bennett GJ, Max MB. Analgesic and cognitive effects of intravenous ketamine-alfentanil combinations versus either drug alone after intradermal capsaicin in normal subjects. Anesth Analg. 1998;86(6):1250–1256. [PubMed]
205. Seybold VS. The role of peptides in central sensitization. Handb Exp Pharmacol. 2009;(194):451–491. [PubMed]
206. Shenker NG, Haigh RC, Mapp PI, Harris N, Blake DR. Contralateral hyperalgesia and allodynia following intradermal capsaicin injection in man. Rheumatology (Oxford) 2008;47(9):1417–1421.[PMC free article] [PubMed]
207. Shibata K, Yamane K, Iwata M. Change of excitability in brainstem and cortical visual processing in migraine exhibiting allodynia. Headache. 2006;46(10):1535–1544. [PubMed]
208. Sivilotti L, Woolf CJ. The contribution of GABAA and glycine receptors to central sensitization: disinhibition and touch-evoked allodynia in the spinal cord. J Neurophysiol. 1994;72(1):169–179.[PubMed]
209. Sorensen J, Bengtsson A, Backman E, Henriksson KG, Bengtsson M. Pain analysis in patients with fibromyalgia. Effects of intravenous morphine, lidocaine, and ketamine. Scand J Rheumatol. 1995;24(6):360–365. [PubMed]
210. Sprenger T, Valet M, Woltmann R, Zimmer C, Freynhagen R, Kochs EF, Tolle TR, Wagner KJ. Imaging pain modulation by subanesthetic S-(+)-ketamine. Anesth Analg. 2006;103(3):729–737.[PubMed]
211. Srbely JZ, Dickey JP, Bent LR, Lee D, Lowerison M. Capsaicin-induced central sensitization evokes segmental increases in trigger point sensitivity in humans. J Pain. 2010;11(7):636–643. [PubMed]
212. Staahl C, Olesen AE, Andresen T, Arendt-Nielsen L, Drewes AM. Assessing efficacy of non-opioid analgesics in experimental pain models in healthy volunteers: an updated review. Br J Clin Pharmacol. 2009;68(3):322–341. [PMC free article] [PubMed]
213. Staud R. Are patients with systemic lupus erythematosus at increased risk for fibromyalgia? Curr Rheumatol Rep. 2006;8(6):430–435. [PubMed]
214. Staud R, Bovee CE, Robinson ME, Price DD. Cutaneous C-fiber pain abnormalities of fibromyalgia patients are specifically related to temporal summation. Pain. 2008;139(2):315–323. [PMC free article][PubMed]
215. Staud R, Cannon RC, Mauderli AP, Robinson ME, Price DD, Vierck CJ., Jr Temporal summation of pain from mechanical stimulation of muscle tissue in normal controls and subjects with fibromyalgia syndrome. Pain. 2003;102(1–2):87–95. [PubMed]
216. Staud R, Craggs JG, Perlstein WM, Robinson ME, Price DD. Brain activity associated with slow temporal summation of C-fiber evoked pain in fibromyalgia patients and healthy controls. Eur J Pain. 2008;12(8):1078–1089. [PMC free article] [PubMed]
217. Staud R, Nagel S, Robinson ME, Price DD. Enhanced central pain processing of fibromyalgia patients is maintained by muscle afferent input: a randomized, double-blind, placebo-controlled study. Pain. 2009;145(1–2):96–104. [PMC free article] [PubMed]
218. Staud R, Price DD, Robinson ME, Mauderli AP, Vierck CJ. Maintenance of windup of second pain requires less frequent stimulation in fibromyalgia patients compared to normal controls. Pain. 2004;110(3):689–696. [PubMed]
219. Staud R, Robinson ME, Price DD. Temporal summation of second pain and its maintenance are useful for characterizing widespread central sensitization of fibromyalgia patients. J Pain. 2007;8(11):893–901.[PMC free article] [PubMed]
220. Staud R, Robinson ME, Vierck CJ, Jr, Cannon RC, Mauderli AP, Price DD. Ratings of experimental pain and pain-related negative affect predict clinical pain in patients with fibromyalgia syndrome. Pain. 2003;105(1–2):215–222. [PubMed]
221. Staud R, Vierck CJ, Cannon RL, Mauderli AP, Price DD. Abnormal sensitization and temporal summation of second pain (wind-up) in patients with fibromyalgia syndrome. Pain. 2001;91(1–2):165–175.[PubMed]
222. Sterling M, Jull G, Vicenzino B, Kenardy J. Characterization of acute whiplash-associated disorders. Spine (Phila Pa 1976) 2004;29(2):182–188. [PubMed]
223. Stubhaug A, Breivik H, Eide PK, Kreunen M, Foss A. Mapping of punctuate hyperalgesia around a surgical incision demonstrates that ketamine is a powerful suppressor of central sensitization to pain following surgery. Acta Anaesthesiol Scand. 1997;41(9):1124–1132. [PubMed]
224. Sweet WH. Control of pain by direct electrical stimulation of peripheral nerves. Clin Neurosurg. 1976;23:103–111. [PubMed]
225. Sycha T, Anzenhofer S, Lehr S, Schmetterer L, Chizh B, Eichler HG, Gustorff B. Rofecoxib attenuates both primary and secondary inflammatory hyperalgesia: a randomized, double blinded, placebo controlled crossover trial in the UV-B pain model. Pain. 2005;113(3):316–322. [PubMed]
226. Takazawa T, MacDermott AB. Glycinergic and GABAergic tonic inhibition fine tune inhibitory control in regionally distinct subpopulations of dorsal horn neurons. J Physiol. 2010;588(Pt 14):2571–2587. [PMC free article] [PubMed]
227. Tao YX. Dorsal horn alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor trafficking in inflammatory pain. Anesthesiology. 2010;112(5):1259–1265. [PMC free article] [PubMed]
228. Tegeder I, Adolph J, Schmidt H, Woolf CJ, Geisslinger G, Lotsch J. Reduced hyperalgesia in homozygous carriers of a GTP cyclohydrolase 1 haplotype. Eur J Pain. 2008;12(8):1069–1077. [PubMed]
229. Tegeder I, Costigan M, Griffin RS, Abele A, Belfer I, Schmidt H, Ehnert C, Nejim J, Marian C, Scholz J, Wu T, Allchorne A, Diatchenko L, Binshtok AM, Goldman D, Adolph J, Sama S, Atlas SJ, Carlezon WA, Parsegian A, Lotsch J, Fillingim RB, Maixner W, Geisslinger G, Max MB, Woolf CJ. GTP cyclohydrolase and tetrahydrobiopterin regulate pain sensitivity and persistence. Nat Med. 2006;12(11):1269–1277. [PubMed]
230. Thompson SW, King AE, Woolf CJ. Activity-Dependent Changes in Rat Ventral Horn Neurons in vitro; Summation of Prolonged Afferent Evoked Postsynaptic Depolarizations Produce a d-2-Amino-5-Phosphonovaleric Acid Sensitive Windup. Eur J Neurosci. 1990;2(7):638–649. [PubMed]
231. Thompson SW, Woolf CJ, Sivilotti LG. Small-caliber afferent inputs produce a heterosynaptic facilitation of the synaptic responses evoked by primary afferent A-fibers in the neonatal rat spinal cord in vitro. J Neurophysiol. 1993;69(6):2116–2128. [PubMed]
232. Tietjen GE, Brandes JL, Peterlin BL, Eloff A, Dafer RM, Stein MR, Drexler E, Martin VT, Hutchinson S, Aurora SK, Recober A, Herial NA, Utley C, White L, Khuder SA. Allodynia in migraine: association with comorbid pain conditions. Headache. 2009;49(9):1333–1344. [PubMed]
233. Torebjork HE, Lundberg LE, LaMotte RH. Central changes in processing of mechanoreceptive input in capsaicin-induced secondary hyperalgesia in humans. J Physiol. 1992;448:765–780. [PMC free article][PubMed]
234. Torres Tortosa M, Roman Rico D. Risk of transmission of human immunodeficiency virus type 1 after accidents with needles from drug addicts, occurred in the community. Rev Clin Esp. 1991;189(2):95–96.[PubMed]
235. Torresani C, Bellafiore S, De Panfilis G. Chronic urticaria is usually associated with fibromyalgia syndrome. Acta Derm Venereol. 2009;89(4):389–392. [PubMed]
236. Trabulsi EJ, Patel J, Viscusi ER, Gomella LG, Lallas CD. Preemptive Multimodal Pain Regimen Reduces Opioid Analgesia for Patients Undergoing Robotic-Assisted Laparoscopic Radical Prostatectomy. Urology. 2010 [PubMed]
237. Treede RD, Meyer RA, Raja SN, Campbell JN. Peripheral and central mechanisms of cutaneous hyperalgesia. Prog Neurobiol. 1992;38(4):397–421. [PubMed]
238. Tuchman M, Barrett JA, Donevan S, Hedberg TG, Taylor CP. Central Sensitization and Ca(V)alpha(2)delta Ligands in Chronic Pain Syndromes: Pathologic Processes and Pharmacologic Effect. J Pain. 2010 [PubMed]
239. Turini D, Beneforti P, Spinelli M, Malagutti S, Lazzeri M. Heat/burning sensation induced by topical application of capsaicin on perineal cutaneous area: new approach in diagnosis and treatment of chronic prostatitis/chronic pelvic pain syndrome? Urology. 2006;67(5):910–913. [PubMed]
240. van den Broeke EN, van Rijn CM, Biurrun Manresa JA, Andersen OK, Arendt-Nielsen L, Wilder-Smith OH. Neurophysiological Correlates of Nociceptive Heterosynaptic Long-Term Potentiation in Humans. J Neurophysiol. 2010 [PubMed]
241. Vaneker M, Wilder-Smith OH, Schrombges P, de Man-Hermsen I, Oerlemans HM. Patients initially diagnosed as ‘warm’ or ‘cold’ CRPS 1 show differences in central sensory processing some eight years after diagnosis: a quantitative sensory testing study. Pain. 2005;115(1–2):204–211. [PubMed]
242. Vardeh D, Wang D, Costigan M, Lazarus M, Saper CB, Woolf CJ, Fitzgerald GA, Samad TA. COX2 in CNS neural cells mediates mechanical inflammatory pain hypersensitivity in mice. J Clin Invest. 2009;119(2):287–294. [PMC free article] [PubMed]
243. Vartiainen NV, Kirveskari E, Forss N. Central processing of tactile and nociceptive stimuli in complex regional pain syndrome. Clin Neurophysiol. 2008;119(10):2380–2388. [PubMed]
244. Wall PD, Woolf CJ. Muscle but not cutaneous C-afferent input produces prolonged increases in the excitability of the flexion reflex in the rat. J Physiol. 1984;356:443–458. [PMC free article] [PubMed]
245. Wall PD, Woolf CJ. The brief and the prolonged facilitatory effects of unmyelinated afferent input on the rat spinal cord are independently influenced by peripheral nerve section. Neuroscience. 1986;17(4):1199–1205. [PubMed]
246. Wang H, Bolognese J, Calder N, Baxendale J, Kehler A, Cummings C, Connell J, Herman G. Effect of morphine and pregabalin compared with diphenhydramine hydrochloride and placebo on hyperalgesia and allodynia induced by intradermal capsaicin in healthy male subjects. J Pain. 2008;9(12):1088–1095.[PubMed]
247. Wang H, Kohno T, Amaya F, Brenner GJ, Ito N, Allchorne A, Ji RR, Woolf CJ. Bradykinin produces pain hypersensitivity by potentiating spinal cord glutamatergic synaptic transmission. J Neurosci. 2005;25(35):7986–7992. [PubMed]
248. Wiesinger B, Malker H, Englund E, Wanman A. Back pain in relation to musculoskeletal disorders in the jaw-face: a matched case-control study. Pain. 2007;131(3):311–319. [PubMed]
249. Wilder-Smith CH, Robert-Yap J. Abnormal endogenous pain modulation and somatic and visceral hypersensitivity in female patients with irritable bowel syndrome. World J Gastroenterol. 2007;13(27):3699–3704. [PMC free article] [PubMed]
250. Willert RP, Delaney C, Hobson AR, Thompson DG, Woolf CJ, Aziz Q. Constitutive cyclo-oxygenase-2 does not contribute to the development of human visceral pain hypersensitivity. Eur J Pain. 2006;10(6):487–494. [PubMed]
251. Willert RP, Delaney C, Kelly K, Sharma A, Aziz Q, Hobson AR. Exploring the neurophysiological basis of chest wall allodynia induced by experimental oesophageal acidification - evidence of central sensitization. Neurogastroenterol Motil. 2007;19(4):270–278. [PubMed]
252. Willert RP, Hobson AR, Delaney C, Hicks KJ, Dewit OE, Aziz Q. Neurokinin-1 receptor antagonism in a human model of visceral hypersensitivity. Aliment Pharmacol Ther. 2007;25(3):309–316. [PubMed]
253. Willert RP, Woolf CJ, Hobson AR, Delaney C, Thompson DG, Aziz Q. The development and maintenance of human visceral pain hypersensitivity is dependent on the N-methyl-D-aspartate receptor. Gastroenterology. 2004;126(3):683–692. [PubMed]
254. Witting N, Svensson P, Arendt-Nielsen L, Jensen TS. Repetitive intradermal capsaicin: differential effect on pain and areas of allodynia and punctate hyperalgesia. Somatosens Mot Res. 2000;17(1):5–12.[PubMed]
255. Woolf CJ. Evidence for a central component of post-injury pain hypersensitivity. Nature. 1983;306(5944):686–688. [PubMed]
256. Woolf CJ. Long term alterations in the excitability of the flexion reflex produced by peripheral tissue injury in the chronic decerebrate rat. Pain. 1984;18(4):325–343. [PubMed]
257. Woolf CJ. Generation of acute pain: central mechanisms. Br Med Bull. 1991;47(3):523–533.[PubMed]
258. Woolf CJ. Pain: moving from symptom control toward mechanism-specific pharmacologic management. Ann Intern Med. 2004;140(6):441–451. [PubMed]
259. Woolf CJ. Central sensitization: uncovering the relation between pain and plasticity. Anesthesiology. 2007;106(4):864–867. [PubMed]
260. Woolf CJ, Chong MS. Preemptive analgesia--treating postoperative pain by preventing the establishment of central sensitization. Anesth Analg. 1993;77(2):362–379. [PubMed]
261. Woolf CJ, Costigan M. Transcriptional and posttranslational plasticity and the generation of inflammatory pain. Proc Natl Acad Sci U S A. 1999;96(14):7723–7730. [PMC free article] [PubMed]
262. Woolf CJ, King AE. Physiology and morphology of multireceptive neurons with C-afferent fiber inputs in the deep dorsal horn of the rat lumbar spinal cord. J Neurophysiol. 1987;58(3):460–479.[PubMed]
263. Woolf CJ, King AE. Subthreshold components of the cutaneous mechanoreceptive fields of dorsal horn neurons in the rat lumbar spinal cord. J Neurophysiol. 1989;62(4):907–916. [PubMed]
264. Woolf CJ, King AE. Dynamic alterations in the cutaneous mechanoreceptive fields of dorsal horn neurons in the rat spinal cord. J Neurosci. 1990;10(8):2717–2726. [PubMed]
265. Woolf CJ, Ma Q. Nociceptors--noxious stimulus detectors. Neuron. 2007;55(3):353–364. [PubMed]
266. Woolf CJ, Max MB. Mechanism-based pain diagnosis: issues for analgesic drug development. Anesthesiology. 2001;95(1):241–249. [PubMed]
267. Woolf CJ, McMahon SB. Injury-induced plasticity of the flexor reflex in chronic decerebrate rats. Neuroscience. 1985;16(2):395–404. [PubMed]
268. Woolf CJ, Salter MW. Neuronal plasticity: increasing the gain in pain. Science. 2000;288(5472):1765–1769. [PubMed]
269. Woolf CJ, Shortland P, Coggeshall RE. Peripheral nerve injury triggers central sprouting of myelinated afferents. Nature. 1992;355(6355):75–78. [PubMed]
270. Woolf CJ, Shortland P, Sivilotti LG. Sensitization of high mechanothreshold superficial dorsal horn and flexor motor neurones following chemosensitive primary afferent activation. Pain. 1994;58(2):141–155. [PubMed]
271. Woolf CJ, Thompson SW. The induction and maintenance of central sensitization is dependent on N-methyl-D-aspartic acid receptor activation; implications for the treatment of post-injury pain hypersensitivity states. Pain. 1991;44(3):293–299. [PubMed]
272. Woolf CJ, Thompson SW, King AE. Prolonged primary afferent induced alterations in dorsal horn neurones, an intracellular analysis in vivo and in vitro. J Physiol (Paris) 1988;83(3):255–266. [PubMed]
273. Woolf CJ, Wall PD. Relative effectiveness of C primary afferent fibers of different origins in evoking a prolonged facilitation of the flexor reflex in the rat. J Neurosci. 1986;6(5):1433–1442. [PubMed]
274. Woolf CJ, Walters ET. Common patterns of plasticity contributing to nociceptive sensitization in mammals and Aplysia. Trends Neurosci. 1991;14(2):74–78. [PubMed]
275. Xu YM, Ge HY, Arendt-Nielsen L. Sustained Nociceptive Mechanical Stimulation of Latent Myofascial Trigger Point Induces Central Sensitization in Healthy Subjects. J Pain. 2010 [PubMed]
276. Yang CC, Lee JC, Kromm BG, Ciol MA, Berger RE. Pain sensitization in male chronic pelvic pain syndrome: why are symptoms so difficult to treat? J Urol. 2003;170(3):823–826. discussion 826–827.[PubMed]
277. Zakine J, Samarcq D, Lorne E, Moubarak M, Montravers P, Beloucif S, Dupont H. Postoperative ketamine administration decreases morphine consumption in major abdominal surgery: a prospective, randomized, double-blind, controlled study. Anesth Analg. 2008;106(6):1856–1861. [PubMed]
278. Zanette G, Cacciatori C, Tamburin S. Central sensitization in carpal tunnel syndrome with extraterritorial spread of sensory symptoms. Pain. 2010;148(2):227–236. [PubMed]
279. Ziegler EA, Magerl W, Meyer RA, Treede RD. Secondary hyperalgesia to punctate mechanical stimuli. Central sensitization to A-fibre nociceptor input. Brain. 1999;122 ( Pt 12):2245–2257. [PubMed]