You are on page 1of 12

Journal of Electromyography and Kinesiology 14 (2004) 109–120

www.elsevier.com/locate/jelekin

Review

Biomechanics of musculoskeletal pain: dynamics of the neuromatrix


Partap S. Khalsa 
Department of Biomedical Engineering, State University of New York at Stony Brook, HSC T18-030 Stony Brook, NY 11794-8181, USA

Abstract

Pain, due to mechanical stimuli, is a normal, indeed healthy, response of animals to potential or actual damage to tissues.
Mammals in general, and humans in particular, have evolved a highly sophisticated system of pain perception, which is character-
ized in humans by complementary but distinct neural processing of the intensity and location of a noxious stimulus, and a moti-
vational/emotional or affective response to the stimulus. The peripheral and central neurons that comprise this system, which has
been called the ‘neuromatrix’, dynamically (temporally) respond and adapt to noxious biomechanical stimuli. However, pheno-
typic variability of the neuromatrix can be large, which can result in a host of musculoskeletal conditions that are characterized
by altered pain perception, which can and often does alter the course of the condition. This neural plasticity has been well recog-
nized in the central nervous system, but it has only more recently become known that peripheral nociceptors also adapt to their
altered extracellular matrix environment. This work reviews the biomechanics of pain focusing on the relevant stimulus that initi-
ates responses by nociceptors to the cognitive perception of pain.
# 2003 Elsevier Ltd. All rights reserved.

Keywords: Pain; Perception; Adequate stimulus; Musculoskeletal

1. Introduction nated (or ‘fine’) peripheral afferents, specifically C-noci-


ceptors, due to aberrations of the gene that encodes for
Pain, due to mechanical stimuli, is a normal, indeed receptors for nerve growth factor (NGF), esp. trkA for
healthy, response of animals to potential or actual HSAN IV [52] (Table 1). In contrast, individuals with
damage to tissues [96]. Mammals in general, and congenital indifference to pain have normal sensation,
humans in particular, have evolved a highly sophisti- including responses to noxious stimuli, but do not have
cated system of pain perception [84,111,148], which in the affective response to noxious stimuli [14,49,70].
humans is characterized by complementary but distinct Individuals with lesions affecting the limbic system may
neural processing of the intensity and location of a develop loss of the affective-motivational component of
noxious stimulus [59], and the motivational/emotional pain, but still retain recognition of the intensity and
or affective response to the stimulus [110]. The periph- location of pain [8,114]. The reverse of this has also
eral and central neurons that comprise this system, been seen in patients with lesions of the somatosensory
which Melzack has termed the ‘neuromatrix’ [82,83,85], cortex, who can only vaguely report where the noxious
dynamically (temporally) respond and adapt to nox- stimulus is occurring, but know that it’s unpleasant or
ious biomechanical stimuli [22] (Fig. 1). The impor- painful [109].
tance of these divisions within pain perception is The dynamic and adaptive response of the neuroma-
illuminated by considering how humans with congeni- trix to noxious mechanical stimuli is well exemplified
tal insensitivity to pain respond to noxious stimuli by individuals with fibromyalgia syndrome (FMS), for
(reviewed in [100]). During their lifetimes, these indivi- whom there is a growing consensus that this is prim-
duals, with one of five types of hereditary sensory rad- arily a disorder of the central nervous system (reviewed
icular neuropathy (HSAN I–V), accumulate a plethora in [42,137]). Patients with FMS demonstrate many
of ‘painless injuries’. HSAN causes a loss of unmyeli- characteristics of central sensitization [27], which is
characterized by increased neural responses of the

Tel.: +1-631-444-2457; fax: +1-631-444-6646. second-order nociceptive neurons (with their cell bodies
E-mail address: partap.khalsa@stonybrook.edu (P.S. Khalsa). in the dorsal horn of the spinal cord) to primary
1050-6411/$ - see front matter # 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jelekin.2003.09.020
110 P.S. Khalsa / Journal of Electromyography and Kinesiology 14 (2004) 109–120

Fig. 1. The ‘neuromatrix’ as illustrated by R. Melzack, that consists of central and peripheral neurons acting as parallel sub-neural networks that
comprise three components: sensory/discrimate (S), emotional/affective (A), and evaluative/cognitive (E). The perception of pain, and the motor
and behavioral responses of humans to pain, is influenced by the entire neuromatrix. (From Melzack[83].)

nociceptor inputs from skin and especially muscle nesium ion block on membrane channels allowing a
[152]. FMS patients have greatly enhanced pain calcium ion influx. Using a NMDA antagonist, for
responses during temporal summation or ‘wind-up’ example ketamine, blocks the wind-up in FMS patients
[112,138,139], which is characterized by increasing pain but not their initial pain [139].
to constant mechanical stimuli repeated at short inter- The transmission by second-order nociceptive neu-
vals (on the order of 3 s). Further, they also demon- rons (with their cell-bodies principally in laminae I and
strate sensitization to the initial pain [139]. The II of the dorsal horn) to the thalamus (and from there
mechanism by which wind-up occurs is due to a rela- to the higher level brain centers, including the somato-
tive depolarization of the second-order neurons that
sensory cortex) is dynamically modulated by neurons
have N-methyl-D-aspartate (NMDA) receptors
descending from the midbrain (esp. ventrolateral peria-
[24,30,101]. The repeated stimulation removes a mag-

Table 1
Hereditary sensory radicular neuropathies (HSAN) have been divided
into five different types (I-V), which are characterized by varying sen-
sory deficits, alterations in the deep tendon reflexes, types of tissue
damages that occur due to the form of HSAN, and what neurons are
affected. (Table adapted from Indo et al., Nat. Genetics 1996.)

HSAN Sensory Reflexes Tissue Neurons


Type Deficits Damage Affected
I Distal loss of Absent/Weak Extremities All (c-fibers
all pain ulcerations more than
Painless others)
Injury
II Same as I Same as I Same as I Myelinated
fibers
III Diffuse pain Same as I Corneal ulcer All but Ad
insensitivity Painless
injury
IV Same as III Weak/ Normal Ulceration, C & Ad
pain, self-
mutilation Fig. 2. Descending antinociceptive mechanisms and their activation
V Distal pain Normal Save as IV C by environmental sources of danger. 5HT—serotonergic, NA—nora-
insensitivity drengergic, RVM—rostral ventro-medial medulla, PAG—periaque-
ductal grey, DR—dorsal raphe nucleus. (From Harris[44].)
P.S. Khalsa / Journal of Electromyography and Kinesiology 14 (2004) 109–120 111

queductal gray), which synapse in the rostral medulla sciatic nerve ligation (SNL), gene expression in neural
(e.g., nucleus raphe magnus), which in turn synapse on soma of the ipsilateral dorsal root ganglia (DRG) was
the second-order nociceptive neurons [6,7] (and differentially compared to the contralateral DRG.
reviewed in [11]) (Fig. 2). The neurons in this anti- Wang et al. [153] found that 88 genes were up-regu-
nociceptive or negative feedback system utilize a var- lated (i.e., increased) more than two-fold, with three
iety of neurotransmitters (reviewed in [36]) for synaptic neuropeptide encoding genes (NPY, Galanin, and VIP)
transmission of their signals including non-peptides having increases of 45-, 18-, and 14-fold, respectively.
(serotonin [119], norepinephrine [157], gama-aminobu- Up-regulation of genes occurred in the following broad
tryic acid (GABA—an inhibitory amino acid categories (with the number of specific genes in par-
[44,116,156]), and opioid peptides (endorphins [149], enthesis): cell cycle- and cell death-related (7), neuroin-
enkephalins [135], and dynorphins [76]). The presence flammatory and immune activation (22), ion channel
of this descending antinociceptive system raises the (10), transcription (3), and tissue maintenance/remo-
epistemological question of why is it there in the first deling/plasticity (23). Notably, these DRG up-regu-
place. Bolles and Fanselow [12] published the first lation changes were not found in the dorsal spinal cord
model that hypothesized that it was an evolutionary [144]. Focusing on several immediate early genes
development in mammals that enabled them to (IEG), Delander et al. [25] also used a SNL model and
respond to real danger (i.e., life-threatening situations) found that c-fos, NGFI-A, and jun B were transiently
without activating the recuperative processes involved up-regulated within hours, though c-jun mRNA
with dealing with noxious stimuli. Their seminal work expression was increased during the entire 4-week per-
has subsequently been further augmented and elabo- iod of the study. Not only were there profound chan-
rated by others, which essentially validated their ges in gene expression of DRG soma of injured
hypothesis (reviewed in [44,130]). neurons, SNL also produced profound changes in the
Segmental levels of the spinal cord themselves pos- expression and distribution of voltage-gated sodium
sess some antinociceptive functions using GABAnergic channels in uninjured unmyelinated axons (i.e., C
interneurons [77,146]. In cat and monkey laminae I and fibers) [41]. This suggests that aberrant responses in
II of the lumbar dorsal horn, interneurons immunor- uninjured C fibers appear to be necessary for persist-
eactive to GABA were found to have presynaptic and ence of neuropathic pain.
postsynaptic influences on primary high threshold The mechanosensitive terminal endings of nocicep-
mechanoreceptor (HTM) Ad terminal endings [2]. tors that directly transduce noxious mechanical stimuli
GABAnergic interneurons also synapse on Ad and C are affected by, and also secrete, a variety of neuropep-
terminal endings that express metabotropic glutamate tides (reviewed in [33,43,117]). The basal lamina of
receptors [145]. However, using the spinally-mediated ‘free’ nerve endings of nociceptors (Group IV or C
tail-flick test in rats combined with selective cord trans- afferents in muscle and skin, respectively) are exposed
ection, antinociception to paw shock in the forelimb to the interstial fluid only at certain regions and other-
was shown to be controlled by supraspinal descending wise are covered by Schwan cell cytoplasm [97,150].
control, whereas in the hindlimb it was controlled by Near these exposed areas are vesicles containing neuro-
lumbar intraspinal mechanisms [154]. In addition to peptides, including substance P (SP), somatostatin
GABAnergic interneurons themselves, intraspinal anti- (SOM), and calcitonin-gene related peptide (CGRP),
nociception is also strongly mediated by neurons which are secreted when the nociceptor is stimulated
expressing opioid receptors (e.g., l and j or l and d [4,99]. These areas are the putative locations for a var-
opioid receptors) [155]. The intensity of the noxious iety membrane receptors and ion channels [46,47,97],
stimulus also appears to play a role in whether anti- though these receptors and channels also appear else-
nociception is dominated by intraspinal or supraspinal where in the neurons.
mechanisms. More intense electric (3.0 vs 1.0 mA) tail-
shock to rats produced intraspinal antinociception
[80,81]. 2. Background of primary muscle nociceptors
The time frames, in which neurons change in
response to noxious stimuli, range from milliseconds to Historically, there are two important conceptual
weeks (and even months). Longer term, and more per- developments that began the scientific, neurophysiolo-
sistent, changes in neural responsiveness all require gical examination of pain. First, in 1900, Sherrington
changes in gene expression, transcription, and protein [127] formally defined the difference between the per-
expression. The magnitude of these changes has ception of pain and the neurons which transduce nox-
recently been illuminated using a combination of DNA ious stimuli:
microarray, quantitative real-time PCR, and immuno-
histochemistry to examine differential gene expression . . . the skin is provided with a set of nerve-endings
in a rat, chronic neuropathic pain model [153]. Using whose specific office it is to be amenable to stimuli
112 P.S. Khalsa / Journal of Electromyography and Kinesiology 14 (2004) 109–120

that do the skin injury, stimuli that in continuing that the central nervous profoundly influenced the per-
to act would injure it still further. These nerve- ception of pain, and this theory contravened much of
endings when still connected with the sensorium the previous three centuries of scientific thought and
(using that term simply to mean the neural investigation, which originated with the philosopher,
machinery to which consciousness is adjunct) on physicist, and mathematician Descartes (Fig. 3) [83].
excitation evoke skin pain. . . . For physiological The majority of modern basic science studies investi-
reference, therefore, they are, it seems to me, both gating pain and primary nociceptors have focused on
on this ground and on others which need not be cutaneous nociceptors because the skin and its inner-
entered upon here, preferably termed nocicipient, a vation are easily accessed. These studies have been
name which has the advantage of greater objec- helpful in understanding the neurophysiology of noci-
tivity. ceptors, but are not necessarily as clinically relevant as
investigations of muscle nociceptors. Further, the struc-
A few years later, he described a relationship tural components and their organization are pro-
between noxious stimuli and behavioral reflexes (e.g., foundly different in muscle than skin [35], and hence
withdrawing a limb) to reduce further tissue damage the mechanical states encoded by muscle nociceptors
and to avoid further pain [128]. Second, in 1906, he may be quite different from those encoded by
formalized the concept of the ‘adequate stimulus’ cutaneous nociceptors.
whereby specific types of neurons preferentially
respond to specific types of stimuli [129]. Thus, our
current concept of muscle pain developing from the 3. Anatomy and physiology of muscle nociceptors
activation of primary nociceptive afferents innervating
muscle came from his insights almost a century ago. Sensory information from muscles is carried by affer-
Another major historical development was the devel- ent neurons that are typically categorized as belonging
opment almost 40 years ago of the gate control theory to groups I–IV based upon their conduction velocities
of pain by Melzack and Wall [86]. Arguably, the main (CVs). Lloyd [75] originated this nomenclature specifi-
contribution of their theory was the inclusion of the cally for afferent neurons from muscles (and their ten-
brain as an influence on the transmission of nociceptive dons and fascia). Groups I and II afferents are thickly
signals from the spinal cord [83], rather than the more myelinated and, respectively, have large diameter axons
well known concept of gating in the dorsal horn of (12–20 and 2–16 lm) and conduct impulses at fast
transmission of small diameter afferents (i.e., nocicep- rates (79–114 and 30–80 m/s). Group I afferents
tive afferents) by large diameter afferents (i.e., mechan- mostly, but not exclusively [5], innervate muscle spin-
oreceptive afferents), which had been proposed earlier dles and tendon organs, while Group II afferents
by Noordenbos [105]. Their ‘revolutionary’ idea was innervate spindle secondary endings and spray (Ruffini)
endings [98]. Noxious stimuli in muscle are transduced
by small diameter afferents with thinly myelinated
axons (group III: CV ranging from 2 to 30 m/s) or
unmyelinated axons (group IV: CV < 2 m/s) [51,106].
However, groups III and IV afferents are not exclus-
ively nociceptive, and some of these afferents transduce
non-noxious mechanical and thermal stimuli. Using
early histological methods [48], light and electron
microscopy [136], as well as three-dimensional recon-
struction of series of semi- and ultrathin sections [46],
both groups III and IV afferents have been found to
terminate as ‘free’ or ‘non-corpuscular’ receptive end-
ings. In other connective tissues (e.g., skin, ligament, or
joint capsule), afferents with these same CVs and also
terminating as ‘free endings’ are termed Ad and C
fibers, respectively. Despite their small diameter axons
and slow CVs, group IV afferents supply 40–60% of the
total sensory innervation to muscle [71].
The receptive endings of muscle nociceptors, imaged
by light and electron microscopy, are found in all types
of tissues within muscles: connective tissue, extra- and
intra-fusal muscle fibers, adventitia of arterioles and
Fig. 3. Descartes’ concept of pain transmission. venules, tendon, and fat cells [136]. However, the
P.S. Khalsa / Journal of Electromyography and Kinesiology 14 (2004) 109–120 113

majority of nociceptors, expressing SP and/or CGRP, by proteins that span the lipid bi-layer that forms the
appear to be associated with capillaries and arterioles cell plasma membrane. Conformational changes in the
[118]. A single axon will typically branch, forming 2–4 proteins result in opening and closing of these channels
separate receptive endings that can often innervate dif- (also called ‘gating’), and when the channels are open,
ferent histological structures within the muscle [13,97]. ions can pass into or out of the cell through diffusion.
Messlinger [97] has termed this multiple innervation of Change in ionic concentration at the membrane level
different histological structures as polytopographical produces a change in the electric ‘potential’ (or volt-
innervation and speculates that it may be related to the age), which is also termed the ‘receptor potential’. If
well-documented phenomena of polymodality of noci- sufficiently large, the receptor potential will depolarize
ceptors (cf. [89]). The term ‘polymodal nociceptor’ was the axon to threshold and evoke an ‘action potential’
first introduced by Bessou and Perl [10] to describe that is actively propagated along the length of the
feline, cutaneous nociceptors that had high thresholds axon.
to mechanical stimuli and were also activated by nox- Some nociceptors have cationic channels that are
v
ious heat (> 45 C) as well as irritant chemicals. The gated by mechanical loads[20,32], but the probability
same type of afferents were subsequently found in (or sensitivity to) gating can be influenced by various
human, cutaneous neurons using percutaneous micro- ligands (Fig. 4). Bradykinin (BK) binds to either a B1
neurography [147], and were also found to be present or a B2 membrane-bound receptor, which then initiates
in canine muscle [69]. an intracellular cascade that ends in increased protein
Activation of a peripheral nociceptor by an adequate kinase C (PKC) concentration that increases the sensi-
stimulus depends upon the presence of specific ion tivity of sodium (Na+) channels[31]. Prostaglandin E2
channels and receptors in the plasma membrane of the (PGE2) binds to an EP1 or EP2 receptor, and via the
receptive ending[37] (Fig. 4). Ion channels are formed cyclic adenosine monophosphate (cAMP) pathway,

Fig. 4. Schematic overview of the receptor molecules in the membrane of a nociceptive ending. For reasons of brevity, only four points will be
addressed: (1) Nociceptive nerve endings possess a special type of Na+ channel that is tetrodotoxin-resistant, meaning that it cannot be blocked by
the neurotoxin tetrodotoxin. Sensitization of the ending is accompanied by the synthesis of larger amounts of the membrane protein that forms
the tetrodotoxin-resistant channel. This increases the excitability of the nociceptor. (2) In intact tissue, bradykinin (BK) excites nociceptors by
binding to the B2 receptor molecule; the sensitized ending, however is activated by the BK through B1 receptors. Binding of BK to the specific
receptor molecules is followed by the activation of an intracellular cascade that leads to the synthesis of increased amounts of protein kinase C
(PKC), which enhances the sensitivity of Na+ channels to stimulus-induced depolarization. (3) Sensitization of the nociceptor also is caused by the
binding of other endogenous algesic substances (such as PGE2, adenosine, and serotonin) to specific membrane receptors whose activation leads
via cyclic adenosine monophosphate (cAMP) to the formation of protein kinase A (PKA), which increases the sensitivity of the tetrodotoxin
(TTX)-resistant Na+ channels. Binding of opioids to peripheral opioid receptors inhibits the cyclic adenosine monophosphate cascade, thereby
counteracting the sensitization. (4) Other types of Na+ channels are open after binding of adenosine triphosphate (ATP) to purinergic membrane
receptors or after binding of protons (H+). The latter type is important for pain from inflamed tissue because inflammation is associated with a
lowered pH. This channel is probably also opened by capsaicin, the active ingredient of the red pepper. G, G protein. After binding of the ligand
to the membrane receptor, the G protein changes the intracellular metabolism. PLC, phospholipase C; DAG, diacyl glycerin; IP3, inositol tripho-
sphate (From: Graven-Nielsen & Mense[43].)
114 P.S. Khalsa / Journal of Electromyography and Kinesiology 14 (2004) 109–120

increases non-specific cation channel opening[65]. Simi- sumed to be the cause of muscle tenderness and hyper-
larly, serotonin (5-HT) and adenosine utilize the cAMP algesia[89]. Following the administration of BK,
pathway to affect protein kinase A (PKA), which sensi- muscle nociceptors will respond to non-noxious mech-
tizes a Na+ channel that is tetrodotoxin (TTX)-resist- anical stimuli, which is termed allodynia (i.e., their
ant. A mechanical stimulus sufficient to injure muscle mechanical thresholds have been lowered)[34,93].
cells (disrupt their cell membranes) will cause them to Intriguingly, Mense and Meyer[93] did not find that
release their stores of adenosine tri-phosphate (ATP), nociceptors were sensitized equally to all types of
which can bind to a purinergic receptor P2X and mechanical stimuli; rather, some afferents could be sen-
directly gate a Na+ channel[15]. Acid sensing ion chan- sitized to stretch but not to contraction and others to
nels (ASIC) are cation channels (Na+) that are gated other combinations of mechanical stimuli. Administra-
by protons (H+) or increased pH of the extra-cellular tion of PGE2 and 5-HT further sensitized the nocicep-
matrix[151], and can be sensitized by neuropeptide tors to effects of BK[87].
SF[28]. Similarly, the now well-known vanilloid recep- Ischemic contraction of muscle activates nociceptors
tor, which responds to capsaicin and heat[16,17] can be that would not otherwise respond to muscle contrac-
sensitized by BK, NGF, and ATP[29]. Finally, the ter- tions of similar magnitude[56,88,94]. However,
minal ending of a nociceptor can be de-sensitized (anti- ischemia alone is not an effective stimulus for muscle
nociception) by the binding of endogenous opioids nociceptors[94], and in humans does not produce pain
released by lymphocytes, which have migrated to the unless it persists for long duration[74]. Muscle
inflamed tissue[142,143]. Hence, there are a wide range mechano-nociceptors that do respond to contraction
of influences on the terminal endings of nociceptors do not show an increase in discharge rate during
that can directly activate and/or sensitize its neural ischemic contraction[94]. Thus, nociceptor activation
response, and may cause primary nociceptors to func- during ischemic contraction appears to affect a sub-
tion as pattern generators (i.e., more than just relaying population of nociceptors that may be responding to
the intensity of a noxious stimulus)[117].
changes in pO2, pH, BK, and/or PGE2[89].
Following an acute mechanical trauma to a muscle
(e.g., blunt blow or over-stretch), the resultant pain can
4. Muscle nociceptor response to stimuli be attributed to two different, though related, mechan-
isms. First, the initial pain would undoubtedly be due
To examine the types of stimuli to which in vivo
to the direct activation of mechanically sensitive groups
muscle nociceptors respond, electrophysiological
III and IV afferents with mechanically gated channels
recordings have been made from single peripheral affer-
in their receptive endings[38]. While the neural respon-
ents from cat and rat muscles. Nociceptors were found
ses of low threshold mechanoreceptors (predominantly
to respond to single and/or combinations of natural
group II, but some group III afferents) would tend to
stimuli including: noxious (tissue damaging) mechan-
ical stimuli[56,57,92], temperature[92], BK[93], quickly saturate during such a noxious mechanical
PGE2[123], 5-HT[87], thrombaxane A2[58], ATP[23], stimulus, they would certainly contribute to a percep-
and leukotrienes[90]. In human psychophysics experi- tual localization. Conjointly, the nociceptors innervat-
ments, combinations, but not isolated peptides, of ing the injured tissue would vigorously respond during
CGRP and either SP or neurokinin A injected into the time that their mechanically gated channels were
skeletal muscle produced pain[108]. In skin, nocicep- open. In skin, spatial populations of mechano-nocicep-
tors have been shown to respond to protons (H+ ions, tors have been shown to encode the location and inten-
low pH)[140,141] as well as SP[91]. Further, Heppel- sity of a noxious indentation, as well as provide some
mann et al.[46] have found that group III afferents with information about the geometry of the indenter[63].
receptive endings terminating in dense connective tissue Second, mechanical stimulation of the nociceptors
were more likely to have high mechanical thresholds would cause them to secrete, via the axon reflex, SP,
while those that terminated in soft, more vascularized, CGRP, and SOM into the extracellular matrix, which
connective tissue tended to have lower mechanical would produce a neurogenic inflammation. SP causes
thresholds. Thus, single nociceptors with multiple ter- in an increased permeability of local blood vessels,
minal branches may be influenced solely or pre- which may be assisted by damage to the vessels from
dominantly by one type of noxious stimulus (e.g., the noxious mechanical stimulus. Extravasation of
mechanical or endogenous chemical) because of their blood plasma results in increased concentration of BK
location within a muscle. (from kallidin in the blood), PGE2 (and other prosta-
The sensitivity of muscle nociceptors to noxious glandins) from endothelial cells, and 5-HT from plate-
mechanical stimuli can be increased by the presence of lets. The net effect would be a reduction in threshold
different endogenous substances[55] (and reviewed and increase in sensitivity of mechano-nociceptors
comprehensively by Mense[89]). This mechanism is pre- (producing allodynia and hyperalgesia, respectively), as
P.S. Khalsa / Journal of Electromyography and Kinesiology 14 (2004) 109–120 115

well as activating previously ‘silent’ or insensitive four weeks. There are no published reports of direct
mechano-nociceptors[115,124,125,126]. neurophysiological tests of nociceptor responses to
Sympathetic efferents are known to influence the either phosphate or creatine kinase. Thus, these sub-
development of some types of pain (for review stances may or may not be causally related to delayed
see[113]). However, the interaction between sympath- onset muscle pain. Finally, the role of muscle inflam-
etic efferents and muscle nociceptors is not certain for a mation due to eccentric contractions as a cause of
number of reasons. First, most studies reported in the delayed onset muscle pain is unclear[134]. There is no
literature have examined the interaction between sym- doubt that inflammatory mediators can activate muscle
pathetic efferents and cutaneous nociceptors[113] rather nociceptors and produce pain (as described earlier in
than muscle nociceptors, and it is not known if results this article). However, the administration of anti-
from the cutaneous nociceptors can be appropriately inflammatory medication is ineffective in relieving
extrapolated to muscle nociceptors. Second, increases delayed onset muscle pain[102]. Hence, the hypothesis
in neural response by cutaneous nociceptors during that delayed onset muscle pain is a result of inflam-
sympathetic efferent stimulation have been produced by mation is yet unproven.
electrical stimulation of the sympathetic trunk, but this Muscle cramps are defined as painful, involuntary
is not a physiological mechanism (cf.[120]). Third, contractions of muscle[26]. While muscle cramps are
injection of inflammatory substances have produced extremely common[72], the mechanisms for the result-
increased nociceptive responses (cf.[50,67]), but again ant pain due to the cramp are still largely
these are not necessarily physiological stimuli. Fourth, unknown[89]. Presumably, groups III and IV nocicep-
it seems unlikely that sympathetic efferents would play tors sensitive to muscle contraction are activated as has
any significant role in acute muscle pain. Rather, their been observed in single unit recordings from electrically
role would be plausible in persistent or chronic muscle stimulated cat muscle[51,56,57,68,92,94,106,107]. How-
pain since they have potent roles as inflammatory med- ever, the relevant mechanical stimulus for causing
iators[66,73] and tissue compliance mediators via vaso- nociceptor activation during cramp is unknown.
motor control [95] among others. Mense[89] has speculated that shearing forces may
Delayed onset muscle pain is most commonly asso- directly activate muscle nociceptors during a cramp
ciated with the performance of muscle contractions where only part of the muscle is contracting at a high
where the external force is greater than that generated frequency.
by the muscle itself[54]. This results in muscle lengthen-
ing during the contraction and is commonly termed an
‘eccentric contraction’. Clinical trials have clearly 5. Adequate mechanical stimulus to activate muscle
demonstrated that greater pain is developed from nociceptors
eccentric than concentric or isometric contrac-
tions[103]. Eccentric exercise in humans increases inor- Natural, mechanical stimuli (other than muscle
ganic phosphate[1] and creatine kinase[104] cramp) that provoke the neural responses of muscle
concentrations in muscle. Preventive administration of nociceptors are compression and/or stretch. Using hin-
L-carnitine before eccentric exercise results in signifi- dlimb muscles in the cat, Bessou and Laporte[9] were
cantly decreased pain levels[40] compared to controls. the first to systematically examine the responses of
L-carnitine has vasodilatory effects and presumably flu- group III and IV afferents to these stimuli. They were
shes endogenously released nociceptor ligands from the followed by Paintal[106] who stimulated group III
muscle. Contrary to popular opinion, muscle lactate muscle afferents by also squeezing, stretching, and
(or lactic acid) does not appear to be directly involved prodding the cat hindlimb muscles. These investiga-
in muscle soreness following rigorous exercise[68]. Lac- tions showed that some, but not all, group III and IV
tate concentrations rapidly decrease to resting levels in afferents respond to ‘natural’ mechanical stimuli. They
less than one hour[122] following exercise, and unless also established that some group III afferents respon-
they are activating a secondary agent, are unlikely to ded to innocuous (i.e., non-noxious) mechanical loads,
be involved in development of pain that commonly and, hence, functioned as mechanoreceptors rather
occurs 12–24 h later. Similarly, decreases in muscle pH than nociceptors. Compression and/or stretch of mus-
following exercise are relatively transient and return to cles has also been used to examine the role of group III
resting levels generally in less than 30 min[3,18,122]. and IV afferents as ‘ergoreceptors’ (i.e., neural recep-
However, a recent rat model of long-lasting hyper- tors that reflexively affect respiration and cardiovascu-
algesia has been developed that created hyperalgesia by lar output)[56,57,68]. Group III afferents were found to
injections, separated by two or five days apart, of low be more responsive to static contraction than to rhyth-
pH saline (4.0–6.0 pH) into the gastrocnemius mus- mic contraction in the cat hindlimb, and group III
cle[133]. The lowered pH was transient, lasting less afferents that were associated with cardiovascular
than 20 min, but the subsequent hyperalgesia lasted for response to hindlimb muscle stimulation acted more as
116 P.S. Khalsa / Journal of Electromyography and Kinesiology 14 (2004) 109–120

mechanoreceptors than as nociceptors. Group III affer- developed by the compression[63]. Using an isolated
ents are also sensitized by muscle fatigue, though they rat gracilis muscle–nerve preparation to examine puta-
probably function only in a secondary fashion to tive group III & IV mechano-nociceptors, the neural
mediate the initial decline in motoneuron rate observed response to noxious indentation was significantly and
during fatiguing maximum voluntary muscular con- substantially more correlated with compressive stress
traction[45]. Group III and IV afferents mediate the than force, strain, or displacement[38]. However, dur-
clasp-knife reflex whereby vigorous mechanical stimu- ing uniaxial stretch, group III and IV afferents equally
lation (i.e., stretch) of a muscle inhibits nearby extensor encoded stress and strain[39], suggesting that different
muscles and excites nearby flexor muscles[21,121]. mechanisms may be involved for compression versus
The techniques of microneurography and intraneural stretch.
microstimulation have been employed to examine in Another unexplored area in mechanical activation of
humans how muscle nociceptors encode deep
nociceptors is the physical process whereby the load is
pain[79,131]. Briefly, the technique involved inserting a
transferred from the extracellular matrix (specifically
thin diameter recording electrode through the skin of
collagen fibrils) to the plasma membrane and/or the
an alert, human volunteer, and advancing the electrode
ion channels responsible for producing the receptor
until its tip penetrated a nerve. The electrode was
manipulated incrementally until the neural responses of potential. Ingber[53] has proposed that the cytoskele-
single afferents could be discriminated. Putative noci- ton of cells acts like a ‘tensegrity’ model. In this model,
ceptors were identified on the basis of their conduction deformation of a cell results in mechanical stresses that
velocities. Discriminated, single, groups III or IV affer- would distribute nearly uniformly over the entire
ents were electrically stimulated and the reported pain- plasma membrane. It is also known that transmem-
ful area in the muscle was superficially mapped on the brane proteins called integrins can connect the cell
skin. Then, the same area was compressed with a cylin- cytoskeleton with the extracellular matrix[19], and rat
der at intensities reported as being ‘painful’ by the cutaneous mechanoreceptors and nociceptors express
human volunteers, and the neural responses of the integrin a2b1, the receptor for collagen[64]. Further,
identified afferents were recorded. These studies showed diffusion of a function-blocking monoclonal antibody
that painful (and noxious) mechanical stimuli in to integrin a2b1 into the receptive fields of all types of
human muscles are carried by group III and IV affer- cutaneous mechanoreceptors and nociceptors signifi-
ents in the same fashion as noxious mechanical stimuli cantly and substantially reduced their neural responses
in animal models. to compression[60]. Finally, in high threshold, but not
low threshold, mechanosensitive ion channels in cul-
tured rat DRG neurons, their response to pressure was
6. Relevant mechanical stimulus significantly reduced in the presence of cytoskeleton-
disrupting agents, colchicine and cytochalasin D[20].
An important question to determine is what is the Taken together, these studies all suggest that the in
‘relevant’ mechanical stimulus that activates muscle vivo response of individual mechano-nociceptors is
nociceptors[78,89,132]. During compression or stretch,
strongly dependent on their physical connection to the
mechanically sensitive neurons (e.g., nociceptors) do
extracellular matrix[64].
not experience the global load or deformation, which
we can relatively easily measure. Rather, they experi-
ence the change in the local, internally developed stres-
ses (related to force) and/or strains (related to 7. Conclusion
deformation) in their immediate environs in the mus-
Pain perception in humans is a highly complex sys-
cle[61]. The mechanical state developed at a receptive
tem that integrates noxious stimuli from the tissue with
ending during compression is further influenced by the
the overall status (physical, emotional, and mental) of
substrate beneath the tissue[62]. If the substrate is hard
(e.g., bone), then compressive stresses and strains will the individual. Unlike the five primary senses, pain is
primarily influence the nociceptor. However, if the sub- strongly influenced by positive and negative feedback
strate is compliant (e.g., muscle or fat), then both com- systems within the brain, spinal cord, and the primary
pressive and tensile stresses and strains will develop and nociceptor. Nonetheless, within the pain perceptive sys-
can influence the nociceptor. In rat hairy skin, it was tem, there is a sensory/discriminant component for
found that cutaneous nociceptors were more sensitive which the intensity of an acute noxious stimulus is
to tensile than to compressive stress[62]. Hence, during essentially proportional to the perceived pain. As we
noxious compression of skin overlying soft tissues, the continue to better understand the complexity of this
response of the cutaneous nociceptors would be expec- system, we will be better able to create treatments to
ted to be substantially driven by the tensile loading prevent and treat patients suffering from pain.
P.S. Khalsa / Journal of Electromyography and Kinesiology 14 (2004) 109–120 117

References [21] C.L. Cleland, L. Hayward, W.Z. Rymer, Neural mechanisms


underlying the clasp-knife reflex in the cat. II. Stretch-sensitive
[1] R. Aldridge, E.B. Cady, D.A. Jones, G. Obletter, Muscle pain muscular-free nerve endings, J. Neurophysiol 64 (4) (1990)
after exercise is linked with an inorganic phosphate increase as 1319–1330.
shown by 31P NMR, Biosci. Rep 6 (7) (1986) 663–667. [22] T.J. Coderre, J. Katz, A.L. Vaccarino, R. Melzack, Contri-
[2] F.J. Alvarez, A.M. Kavookjian, A.R. Light, Synaptic interac- bution of central neuroplasticity to pathological pain: review of
tions between GABA-immunoreactive profiles and the term- clinical and experimental evidence, Pain 52 (3) (1993) 259–285.
inals of functionally defined myelinated nociceptors in the [23] S.P. Cook, L. Vulchanova, K.M. Hargreaves, R. Elde, E.W.
monkey and cat spinal cord, J. Neurosci. 12 (8) (1992) 2901– McCleskey, Distinct ATP receptors on pain-sensing and
2917. stretch-sensing neurons, Nature 387 (6632) (1997) 505–508.
[3] J. Bangsbo, L. Johansen, T. Graham, B. Saltin, Lactate and [24] S.N. Davies, D. Lodge, Evidence for involvement of N-methy-
H+ effluxes from human skeletal muscles during intense, laspartate receptors in ‘wind-up’ of class 2 neurones in the dor-
dynamic exercise, J. Physiol. (Lond.) 462 (1993) 115–133. sal horn of the rat, Brain Res 424 (2) (1987) 402–406.
[4] L.A. Barber, M.R. Vasko, Activation of protein kinase C aug- [25] G.E. Delander, E. Schott, E. Brodin, B.B. Fredholm, Spinal
ments peptide release from rat sensory neurons, J. Neurochem. expression of mRNA for immediate early genes in a model of
67 (1) (1996) 72–80. chronic pain, Acta Physiol. Scand 161 (4) (1997) 517–525.
[5] D. Barker, M.C. Ip, M.N. Adal, A correlation between the [26] D. Denny-Brown, J.M. Foley, Myokymia and the benign fasci-
receptor population of the cat’s soleus muscle and the afferent culation of muscle cramps, Trans. Assoc. Am. Phys 61 (1948)
fibre-diameter spectrum of the nerve supplying it, in: Sym- 88–96.
posium on Muscle Receptors 1962, pp. 257–262. [27] J.A. Desmeules, C. Cedraschi, E. Rapiti, E. Baumgartner, A.
[6] A.I. Basbaum, H.L. Fields, Endogenous pain control mechan- Cohen, P. Cohen, P. Dayer, T.L. Vischer, Neurophysiologic
isms: review and hypothesis, Ann. Neurol 4 (5) (1978) 451–462. evidence for a central sensitization in patients with fibro-
[7] A.I. Basbaum, H.L. Fields, The origin of descending pathways myalgia, Arthritis Rheum 48 (5) (2003) 1420–1429.
in the dorsolateral funiculus of the spinal cord of the cat and [28] E. Deval, A. Baron, E. Lingueglia, H. Mazarguil, J. Zajac, M.
rat: further studies on the anatomy of pain modulation, J. Lazdunski, Effects of neuropeptide SF and related peptides on
Comp. Neurol. 187 (3) (1979) 513–531. acid sensing ion channel 3 and sensory neuron excitability,
[8] M. Berthier, S. Starkstein, R. Leiguarda, Asymbolia for pain: a Neuropharmacology 44 (5) (2003) 662–671.
sensory-limbic disconnection syndrome, Ann. Neurol 24 (1) [29] V. Di Marzo, P.M. Blumberg, A. Szallasi, Endovanilloid sig-
(1988) 41–49. naling in pain, Curr. Opin. Neurobiol 12 (4) (2002) 372–379.
[9] P. Bessou, Y.A. LaPorte, ctivation of unmyelinated afferent [30] A.H. Dickenson, A.F. Sullivan, Evidence for a role of the
fibers of small caliber originating in muscle (French), J. Physiol. NMDA receptor in the frequency dependent potentiation of
(Paris) 152 (1958) 1587–1590. deep rat dorsal horn nociceptive neurones following C fibre
[10] P. Bessou, E.R. Perl, Response of cutaneous sensory units with stimulation, Neuropharmacology 26 (8) (1987) 1235–1238.
unmyelinated fibers to noxious stimuli, J. Neurophysiol 32 (6) [31] A. Dray, Kinins and their receptors in hyperalgesia, Can. J.
(1969) 1025–1043. Physiol. Pharmacol 75 (6) (1997) 704–712.
[11] R.J. Bodnar, Supraspinal circuitry mediating opioid anti- [32] L.J. Drew, J.N. Wood, P.D. Cesare, istinct mechanosensitive
nociception: antagonist and synergy studies in multiple sites, J. properties of capsaicin-sensitive and -insensitive sensory neu-
Biomed. Sci. 7 (3) (2000) 181–194. rons, J. Neurosci 22 (12) (2002) RC228.
[12] R.C. Bolles, M.S. Fanselow, A perceptual–defensive–recuperat- [33] R. Dubner, K. Ren, Endogenous mechanisms of sensory modu-
ive model of fear and pain, Behav. Brain Sci 3 (1980) 291–323. lation, Pain 6 (Suppl) (1999) S45–S53.
[13] G. Bove, A.R. Light, Unmyelinated nociceptors of rat para- [34] M. Franz, S. Mense, Muscle receptors with group IV afferent
spinal tissue, J. Neurophysiol 73 (1995) 1752–1762. fibres responding to application of bradykinin, Brain Res 92 (3)
[14] D. Bowsher, J. Lahuerta, B. Peach, D. Venn, M. Haywar, J. (1975) 369–383.
Mumford, J. Mumford, C. Haggett, Familial indifference to [35] Y.C. Fung, Biomechanics: mechanical properties of living
pain with somatosensory asymmetry: possible central anomaly, tissues (1993), Springer-Verlag, New York.
Rev. Neurol. (Paris) 158 (2) (2002) 195–202. [36] S. Furst, Transmitters involved in antinociception in the spinal
[15] G. Burnstock, P2X receptors in sensory neurones, Br. J. cord, Brain Res. Bull 48 (2) (1999) 129–141.
Anaesth. 84 (4) (2000) 476–488. [37] J. Garcia-Anoveros, D.P. Corey, The molecules of mechan-
[16] M.J. Caterina, D. Julius, The vanilloid receptor: a molecular osensation, Ann. Rev. Neurosci 20 (1997) 567–594.
gateway to the pain pathway, Ann. Rev. Neurosci. 24 (2001) [38] W. Ge, P.S. Khalsa, Encoding of Compressive stress during
487–517. indentation by group III and IV Muscle mechano-nociceptors
[17] M.J. Caterina, M.A. Schumacher, M. Tominaga, T.A. Rosen, in rat gracilis muscle, J. Neurophysiol. 89 (2) (2003) 785–792.
J.D. Levine, D. Julius, The capsaicin receptor: a heat-activated [39] W. Ge, P.S. Khalsa, Mechanical states encoded by muscle noci-
ion channel in the pain pathway, Nature 389 (6653) (1997) 816– ceptors during tension in isolated rat gracilis muscle. Abstr.
824. Soc. Neurosci. 29 (2003) 62–69.
[18] M.J. Caterina, M.A. Schumacher, M. Tominaga, T.A. Rosen, [40] M.A. Giamberardino, L. Dragani, R. Valente, F. Di Lisa, R.
J.D. Levine, D. Julius, The capsaicin receptor: a heat-activated Vecchiet, L. Vecchiet, Effects of prolonged L-carnitine adminis-
ion channel in the pain pathway [see comments], Nature 389 tration on delayed muscle pain and CK release after eccentric
(6653) (1997) 816–824. effort, Int. J. Sports Med 17 (5) (1996) 320–324.
[19] M.E. Chicurel, C.S. Chen, D.E. Ingber, Cellular control lies in [41] M.S. Gold, D. Weinreich, C.S. Kim, R. Wang, J. Treanor, F.
the balance of forces, Curr. Opin. Cell Biol 10 (2) (1998) 232– Lai, J. Lai, Redistribution of Na(V)1.8 in uninjured axons
239. enables neuropathic pain, J. Neurosci. 23 (1) (2003) 158–166.
[20] H. Cho, J. Shin, C.Y. Shin, S.Y. Lee, U. Oh, Mechanosensitive [42] T. Graven-Nielsen, L. Arendt-Nielsen, Peripheral and central
ion channels in cultured sensory neurons of neonatal rats, J. sensitization in musculoskeletal pain disorders: an experimental
Neurosci. 22 (4) (2002) 1238–1247. approach, Curr. Rheumatol. Rep 4 (4) (2002) 313–321.
118 P.S. Khalsa / Journal of Electromyography and Kinesiology 14 (2004) 109–120

[43] T. Graven-Nielsen, S. Mense, The peripheral apparatus of mus- [64] P.S. Khalsa, C. Zhang, D. Sommerfeldt, M. Hadjiargyrou,
cle pain: evidence from animal and human studies, Clin. J. Pain Expression of integrin alpha2beta1 in axons and receptive end-
17 (1) (2001) 2–10. ings of neurons in rat, hairy skin, Neurosci. Lett 293 (1) (2000)
[44] J.A. Harris, Descending antinociceptive mechanisms in the 13–16.
brainstem: their role in the animal’s defensive system, J. Phy- [65] S.G. Khasar, A.K. Ouseph, B. Chou, T. Ho, P.G. Green, J.D.
siol. (Paris) 90 (1) (1996) 15–25. Levine, Is there more than one prostaglandin E receptor sub-
[45] L. Hayward, U. Wesselmann, W.Z. Rymer, Effects of muscle type mediating hyperalgesia in the rat hindpaw? Neuroscience
fatigue on mechanically sensitive afferents of slow conduction 64 (4) (1995) 1161–1165.
velocity in the cat triceps surae, J. Neurophysiol 65 (2) (1991) [66] L.R. Kim, K. Whelpdale, M. Zurowski, B. Pomeranz, Sym-
360–370. pathetic denervation impairs epidermal healing in cutaneous
[46] B. Heppelmann, K. Messlinger, W.F. Neiss, R.F. Schmidt, wounds, Wound Repair Regen 6 (3) (1998) 194–201.
Serial sectioning, electron microscopy, and three-dimensional [67] E. Kinnman, J.D. Levine, Involvement of the sympathetic post-
reconstruction of fine nerve fibres and other extended objects, J. ganglionic neuron in capsaicin-induced secondary hyperalgesia
Microsc 156 (1989) 163–172. in the rat, Neuroscience 65 (1) (1995) 283–291.
[47] B. Heppelmann, K. Messlinger, W.F. Neiss, R.F. Schmidt, [68] K.D. Kniffki, S. Mense, R.F. Schmidt, Responses of group IV
Ultrastructural three-dimensional reconstruction of group III afferent units from skeletal muscle to stretch, contraction and
and group IV sensory nerve endings (‘free nerve endings’) in the chemical stimulation, Exp. Brain Res 61 (1978) 511–522.
knee joint capsule of the cat: evidence for multiple receptive [69] T. Kumazawa, K. Mizumura, Thin-fibre receptors responding
sites, J. Comp. Neurol 292 (1) (1990) 103–116. to mechanical, chemical, and thermal stimulation in the skeletal
[48] J.C. Hinsey, Some observations on the innervation of skeletal muscle of the dog, J. Physiol 273 (1) (1977) 179–194.
muscle of the cat, J. Comp. Neurol 44 (1927) 87–195. [70] P. Landrieu, G. Said, C. Allaire, Dominantly transmitted con-
[49] E. Hirsch, D. Moye, J.H. Dimon III, Congenital indifference to genital indifference to pain, Ann. Neurol 27 (5) (1990) 574–578.
pain: long-term follow-up of two cases, South. Med. J 88 (8) [71] L.A. Langford, Unmyelinated axon ratios in cat motor,
(1995) 851–857. cutaneous and articular nerves, Neurosci. Lett 40 (1) (1983) 19–
[50] S.J. Hu, J. Zhu, Sympathetic facilitation of sustained discharges 22.
of polymodal nociceptors, Pain 38 (1) (1989) 85–90. [72] R.B.M. Layzer, Muscle pain, cramps, and fatigue, in: A.G.
[51] A.N. Iggo, on-myelinated afferent fibres from mammalian skel- Franzini-Armystrong, C. Franzini-Armystrong (Eds.),
etal muscle, J. Physiol 155 (1960) 52P–53P. McGraw-Hill, New York, 1994, pp. 3462–3497.
[52] Y. Indo, M. Tsuruta, Y. Hayashida, M.A. Karim, K. Ohta, T. [73] J.D. Levine, T.J. Coderre, K. Covinsky, A.I. Basbaum, Neural
Mitsubuchi, H. Mitsubuchi, H. Tonoki, Y. Awaya, I. Matsuda, influences on synovial mast cell density in rat, J. Neurosci. Res
Mutations in the TRKA/NGF receptor gene in patients with 26 (3) (1990) 301–307.
congenital insensitivity to pain with anhidrosis, Nat. Genet 13 [74] T. Lewis, Pain (1944), The Macmillan Company, New York.
(4) (1996) 485–488. [75] D.P.C. Lloyd, Neuron patterns controlling transmission of ipsi-
[53] D.E. Ingber, Tensegrity: the architectural basis of cellular lateral hind limb reflexes in cat, J. Neurophysiol 6 (1943) 293–
mechanotransduction, Ann. Rev. Physiol 59 (1997) 575–599. 315.
[54] D.A. Jones, D.J. Newham, C. Torgan, Mechanical influences [76] L. MacArthur, K. Ren, E. Pfaffenroth, E. Franklin, M.A.
on long-lasting human muscle fatigue and delayed-onset pain, Ruda, Descending modulation of opioid-containing nociceptive
J. Physiol 412 (1989) 415–427. neurons in rats with peripheral inflammation and hyperalgesia,
[55] M.P. Kaufman, K.J. Rybicki, Discharge properties of group III Neuroscience 88 (2) (1999) 499–506.
and IV muscle afferents: their responses to mechanical and [77] M. Malcangio, N.G. Bowery, GABA and its receptors in the
metabolic stimuli, Circ. Res 61 (4) (1987) 60–65. spinal cord, Trends Pharmacol. Sci 17 (12) (1996) 457–462.
[56] M.P. Kaufman, K.J. Rybicki, T.G. Waldrop, G.A. Ordway, [78] P. Marchettini, Muscle pain: animal and human experimental
Effect of ischemia on responses of group III and IV afferents to and clinical studies, Muscle Nerve 16 (1993) 1033–1039.
contraction, J. Appl. Physiol 57 (3) (1984) 644–650. [79] P. Marchettini, D.A. Simone, G. Caputi, J.L. Ochoa, Pain
[57] M.P. Kaufman, T.G. Waldrop, K.J. Rybicki, G.A. Ordway, from excitation of identified muscle nociceptors in humans,
J.H. Mitchell, Effects of static and rhythmic twitch contractions Brain Res 740 (1-2) (1996) 109–116.
on the discharge of group III and IV muscle afferents, J. Appl. [80] M.W. Meagher, P.S. Chen, J.A. Salinas, J.W. Grau, Activation
Physiol 55 (1) (1983) 105–112. of the opioid and nonopioid hypoalgesic systems at the level of
[58] J. Kenagy, J. VanCleave, L. Pazdernik, J.A. Orr, Stimulation the brainstem and spinal cord: does a coulometric relation pre-
of group III and IV afferent nerves from the hindlimb by dict the emergence or form of environmentally induced hypoal-
thromboxane A2, Brain Res 744 (1) (1997) 175–178. gesia?, Behav. Neurosci 107 (3) (1993) 493–505.
[59] D.R. Kenshalo Jr., O. Isensee, Responses of primate SI cortical [81] M.W. Meagher, J.W. Grau, R.A. King, Role of supraspinal
neurons to noxious stimuli, J. Neurophysiol. 50 (6) (1983) systems in environmentally induced antinociception: effect of
1479–1496. spinalization and decerebration on brief shock-induced and
[60] P.S. Khalsa, W. Ge, M. Hadjiargyrou, Integrin alpha beta 1 long shock-induced antinociception, Behav. Neurosci 104 (2)
antibody modulates mechanoreceptor response in rat hairy (1990) 328–338.
skin, Abstr. Soc. Neurosci 27 (2001) 392. [82] R. Melzack, Phantom limbs and the concept of a neuromatrix,
[61] P.S. Khalsa, A.H. Hoffman, P. Grigg, Mechanical states enco- Trends Neurosci 13 (3) (1990) 88–92.
ded by stretch-sensitive neurons in feline joint capsule, J. Neu- [83] R. Melzack, From the gate to the neuromatrix, Pain 6 (Suppl)
rophysiol. 76 (1) (1996) 175–187. (1999) S121–S126.
[62] P.S. Khalsa, R.H. LaMotte, P. Grigg, Tension and com- [84] R. Melzack, K.L. Casey, Sensory, motivational, and central
pression responses of nociceptors in rat hairy skin, J. Neuro- control determinations of pain: a new conceptual model, in:
physiol 77 (7) (1997) 492–505. D.R. Kenshalo (Ed.), Charles C. Thomas, Springfield, 1968,
[63] P.S. Khalsa, C. Zhang, Y.X. Qin, Encoding of location and pp. 432–443.
intensity of noxious indentation into rat skin by spatial popula- [85] R. Melzack, T.J. Coderre, J. Katz, A.L. Vaccarino, Central
tions of cutaneous, mechano-nociceptors, J. Neurophysiol 83 neuroplasticity and pathological pain, Ann. NY Acad. Sci 933
(2000) 3049–3061. (2001) 157–174.
P.S. Khalsa / Journal of Electromyography and Kinesiology 14 (2004) 109–120 119

[86] R. Melzack, P.D. Wall, Pain mechanisms: a new theory, A and substance P: effects on nociception and neurogenic
Science 150 (699) (1965) 971–979. inflammation in human skin and temporal muscle, Peptides 12
[87] S. Mense, Sensitization of group IV muscle receptors to brady- (2) (1991) 333–337.
kinin by 5-hydroxytryptamine and prostaglandin E2, Brain Res [109] M. Ploner, H.J. Freund, A. Schnitzler, Pain affect without pain
225 (1) (1981) 95–105. sensation in a patient with a postcentral lesion, Pain 81 (1-2)
[88] S. Mense, Electrophysiology of muscular nociception in (1999) 211–214.
ischemia and oxygen deficiency, Vasa Suppl 32 (1991) 553–561. [110] D.D. Price, Psychological and neural mechanisms of the affect-
[89] S. Mense, Nociception from skeletal muscle in relation to clini- ive dimension of pain, Science 288 (5472) (2000) 1769–1772.
cal muscle pain, Pain 54 (1993) 241–289. [111] D.D. Price, J.J. Barrell, P. Rainville, Integrating experiential-
[90] S. Mense, U. Hoheisel, Influence of leukotriene D4 on the dis- phenomenological methods and neuroscience to study neural
charges of slowly conducting afferent units from normal and mechanisms of pain and consciousness, Conscious. Cogn 11 (4)
inflamed muscle in the rat, Pfluger’s Arch 415 (S1) (1990) R105. (2002) 593–608.
[91] S. Mense, U. Hoheisel, A. Reinert, The possible role of sub- [112] D.D. Price, R. Staud, M.E. Robinson, A.P. Mauderli, R.
stance P in eliciting and modulating deep somatic pain, Prog. Vierck, C.J. Vierck, Enhanced temporal summation of second
Brain Res 110 (1996) 125–135. pain and its central modulation in fibromyalgia patients, Pain
[92] S. Mense, H. Meyer, Different types of slowly conducting affer- 99 (1-2) (2002) 49–59.
ent units in cat skeletal muscle and tendon, J. Physiol. (Lond) [113] S.N. Raja, Role of the sympathetic nervous system in acute
363 (1985) 403–417. pain and inflammation, Ann. Med 27 (2) (1995) 241–246.
[93] S. Mense, H. Meyer, Bradykinin-induced modulation of the [114] V.S. Ramachandran, Consciousness and body image: lessons
response behaviour of different types of feline group III and IV from phantom limbs, Capgras syndrome and pain asymbolia,
muscle receptors, J. Physiol. (Lond) 398 (1988) 49–63. Phil. Trans. R. Soc. Lond B 353 (1377) (1998) 1851–1859.
[94] S. Mense, M. Stahnke, Response in muscle afferent fibres of [115] P.W. Reeh, J. Bayer, L. Kocher, H.O. Handwerker, Sensitiza-
slow conduction velocity to contractions and ischaemia in the tion of nociceptive cutaneous nerve fibers from the rat’s tail by
cat, J. Physiol 342 (1983) 383–397. noxious mechanical stimulation, Exp. Brain Res 65 (3) (1987)
[95] M. Merhi, R.D. Helme, Z. Khalil, Age-related changes in sym- 505–512.
pathetic modulation of sensory nerve activity in rat skin, [116] D.B. Reichling, A.I. Basbaum, Contribution of brainstem
Inflamm. Res 47 (6) (1998) 239–244. GABAergic circuitry to descending antinociceptive controls: I.
[96] H. Merskey, Classification of chronic pain: descriptions of GABA-immunoreactive projection neurons in the periaque-
chronic pain syndromes and definitions of pain terms, Pain ductal gray and nucleus raphe magnus, J. Comp. Neurol. 302
Suppl 3 (1986) S1–S225. (2) (1990) 370–377.
[97] K. Messlinger, Functional morphology of nociceptive and [117] D.B. Reichling, J.D. Levine, The primary afferent nociceptor as
other fine sensory endings (free nerve endings) in different pattern generator, Pain 6 (Suppl) (1999) S103–S109.
tissues, Prog. Brain Res 113 (1996) 273–298. [118] A. Reinert, A. Kaske, S. Mense, Inflammation-induced increase
[98] J.H. Mitchell, R.F. Schmidt, Cardiovascular reflex control by in the density of neuropeptide-immunoreactive nerve endings in
afferent fibers from skeletal muscle receptors III. In: J.T. Shep- rat skeletal muscle, Exp. Brain Res 121 (2) (1998) 174–180.
herd, F.M. Abboud (Eds) Handbook of Physiology, Volume 3, [119] M.H. Roberts, M. Davies, D. Girdlestone, G.A. Foster, Effects
Section 2, The Cardiovascular System. The American Physio- of 5-hydroxytryptamine agonists and antagonists on the
logical Society, Bethesda, Maryland. (1983) 523–658. responses of rat spinal motoneurones to raphe obscurus stimu-
[99] C. Molander, J. Ygge, C.J. Dalsgaard, Substance P-, somatos- lation, Br. J. Pharmacol 95 (2) (1988) 437–448.
tatin- and calcitonin gene-related peptide-like immunoreactivity [120] W.J. Roberts, S.M. Elardo, Sympathetic activation of unmyeli-
and fluoride resistant acid phosphatase-activity in relation nated mechanoreceptors in cat skin, Brain Res 339 (1) (1985)
to retrogradely labeled cutaneous, muscular and visceral 123–125.
primary sensory neurons in the rat, Neurosci. Lett 74 (1) (1987) [121] W.Z. Rymer, J.C. Houk, P.E. Crago, Mechanisms of the clasp-
37–42. knife reflex studied in an animal model, Exp. Brain Res 37
[100] E.M. Nagasako, A.L. Oaklander, R.H. Dworkin, Congenital (1979) 93–113.
insensitivity to pain: an update, Pain 101 (3) (2003) 213–219. [122] K. Sahlin, R.C. Harris, B. Nylind, E. Hultman, Lactate content
[101] V. Neugebauer, T. Lucke, H.G. Schaible, Differential effects of and pH in muscle obtained after dynamic exercise, Pflugers
N-methyl-D-aspartate (NMDA) and non-NMDA receptor Arch 367 (2) (1976) 143–149.
antagonists on the responses of rat spinal neurons with joint [123] H.G. Schaible, R.F. Schmidt, Time course of mechan-
input, Neurosci. Lett 155 (1) (1993) 29–32. osensitivity changes in articular afferents during a developing
[102] D.J. Newham, The consequences of eccentric contractions and experimental arthritis, J. Neurophysiol 60 (6) (1988) 2180–2195.
their relationship to delayed onset muscle pain, Eur. J. Appl. [124] M. Schmelz, R. Schmidt, M. Ringkamp, H.O. Handwerker,
Physiol 57 (3) (1988) 353–359. H.E. Torebjork, Sensitization of insensitive branches of C noci-
[103] D.J. Newham, D.A. Jones, G. Ghosh, P. Aurora, Muscle fati- ceptors in human skin, J. Physiol 480 (Pt 2) (1994) 389–394.
gue and pain after eccentric contractions at long and short [125] R. Schmidt, M. Schmelz, C. Forster, M. Ringkamp, E. Hand-
length, Clin. Sci 74 (5) (1988) 553–557. werker, H. Handwerker, Novel classes of responsive and unre-
[104] D.J. Newham, D.A. Jones, S.E. Tolfree, R.H. Edwards, Skel- sponsive C nociceptors in human skin, J Neurosci. 15 (1) (1995)
etal muscle damage: a study of isotope uptake, enzyme efflux 333–341.
and pain after stepping, Eur. J. Appl. Physiol. Occup. Physiol [126] R. Schmidt, M. Schmelz, H.E. Torebjork, H.O. Handwerker,
55 (1) (1986) 106–112. Mechano-insensitive nociceptors encode pain evoked by tonic
[105] W. Noordenbos. Pain (1959), Elsevier, Amsterdam. pressure to human skin, Neuroscience 98 (4) (2000) 793–800.
[106] A.S. Paintal, Functional analysis of group III afferent fibres of [127] C.S. Sherrington,Cutaneous sensations, in: E.A. Schaefer (Ed.)
mammalian muscles, J. Physiol 152 (1960) 250–270. Textbook of Physiology. Edinburgh and London, 1900, pp.
[107] A.S. Paintal, Participation by pressure-pain receptors in the 920–1001, Macmillan, London.
flexion reflex, J Physiol. (Lond) 156 (1961) 498–514. [128] C.S. Sherrington, Qualitative difference of spinal reflex corre-
[108] U. Pedersen-Bjergaard, L.B. Nielsen, K. Jensen, L. Edvinsson, sponding with qualitative difference of cutaneous stimulus, J.
I. Jansen, J. Olesen, Calcitonin gene-related peptide, neurokinin Physiol 30 (1903) 39–46.
120 P.S. Khalsa / Journal of Electromyography and Kinesiology 14 (2004) 109–120

[129] C.S. Sherrington, Integrative Action of the Nervous System [146] A.J. Todd, Electron microscope study of Golgi-stained cells in
1906, vol. xvi, p. 411, Yale University Press, New Haven and lamina II of the rat spinal dorsal horn, J. Comp. Neurol. 275
London. (1) (1988) 145–157.
[130] B. Siegfried, H.R. Frischknecht, R.L. Nunes de Souza, An [147] H.E. Torebjork, Afferent C units responding to mechanical,
ethological model for the study of activation and interaction of thermal and chemical stimuli in human non-glabrous skin, Acta
pain, memory and defensive systems in the attacked mouse. Physiol. Scand 92 (3) (1974) 374–390.
Role of endogenous opioids, Neurosci. Biobehav. Rev 14 (4) [148] R.D. Treede, D.R. Kenshalo, R.H. Gracely, A.K. Jones, The
(1990) 481–490. cortical representation of pain, Pain 79 (2-3) (1999) 105–111.
[131] D.A. Simone, M. Marchetti, G. Caputi, J.L. Ochoa, Identifi- [149] E. Vasquez, H. Vanegas, The antinociceptive effect of PAG-
cation of muscle afferents subserving sensation of deep pain in microinjected dipyrone in rats is mediated by endogenous
humans, J. Neurophysiol 72 (2) (1994) 883–889. opioids of the rostral ventromedical medulla, Brain Res 854 (1-
[132] D.A. Simone, P. Marchettini, J.L. Ochoa, Primary afferent 2) (2000) 249–252.
nerve fibrs that contribute to muscle pain sensation in humans, [150] M. von During, K.H. Andres, Topography and ultrastructure
Pain Forum 6 (4) (1997) 207–212. of group III and IV nerve terminals of the cat’s gastrocnemius-
[133] K.A. Sluka, A. Kalra, S.A. Moore, Unilateral intramuscular soleus muscle, in: W. Zenker, W.L. Neuhuber (Eds.), Plenum
injections of acidic saline produce a bilateral, long-lasting Press, New York, 1990, pp. 35–41.
hyperalgesia, Muscle Nerve 24 (1) (2001) 37–46. [151] R. Waldmann, G. Champigny, E. Lingueglia, J.R. De Weille,
[134] L.L. Smith, Acute inflammation: the underlying mechanism in C. Heurteaux, M. Lazdunski, H(+)-gated cation channels,
delayed onset muscle soreness? Med. Sci. Sports Exerc 23 (5) Ann. NY Acad. Sci 868 (1999) 67–76.
(1991) 542–551. [152] P.D. Wall, C.J. Woolf, Muscle but not cutaneous C-afferent
[135] J.H. Sohn, B.H. Lee, S.H. Park, J.W. Ryu, B.O. Kim, Y.G. input produces prolonged increases in the excitability of the
Park, Microinjection of opiates into the periaqueductal gray flexion reflex in the rat, J. Physiol 356 (1984) 443–458.
matter attenuates neuropathic pain symptoms in rats, Neurore- [153] H. Wang, H. Sun, K. Della Penna, R.J. Benz, J. Xu, D.L.
port 11 (7) (2000) 1413–1416. Holder, D.J. Holder, K.S. Koblan, Chronic neuropathic pain is
[136] M.J. Stacey, Free nerve endings in skeletal muscle of the cat, J. accompanied by global changes in gene expression and shares
Anat 105 (2) (1969) 231–254. pathobiology with neurodegenerative diseases, Neuroscience
[137] R. Staud, Evidence of involvement of central neural mechan- 114 (3) (2002) 529–546.
isms in generating fibromyalgia pain, Curr. Rheumatol. Rep 4 [154] L.R. Watkins, D.A. Cobelli, D.J. Mayer, Opiate vs non-opiate
(4) (2002) 299–305. footshock induced analgesia (FSIA): descending and intraspinal
[138] R. Staud, R.C. Cannon, A.P. Mauderli, M.E. Robinson, D.D. components, Brain Res 245 (1) (1982) 97–106.
Vierck, C.J. Vierck, Temporal summation of pain from mech- [155] L.R. Watkins, E.P. Wiertelak, J.E. Grisel, L.H. Silbert, S.F.
anical stimulation of muscle tissue in normal controls and sub- Maier, Parallel activation of multiple spinal opiate systems
jects with fibromyalgia syndrome, Pain 102 (1-2) (2003) 87–95. appears to mediate ‘non-opiate’ stress-induced analgesias, Brain
[139] R. Staud, C.J. Vierck, R.L. Cannon, A.P. Mauderli, D.D. Res 594 (1) (1992) 99–108.
Price, Abnormal sensitization and temporal summation of [156] K. Yang, W.L. Ma, Y.P. Feng, Y.X. Dong, Y.Q. Li, Origins of
second pain (wind-up) in patients with fibromyalgia syndrome, GABA(B) receptor-like immunoreactive terminals in the rat
Pain 91 (1-2) (2001) 165–175. spinal dorsal horn, Brain Res. Bull 58 (5) (2002) 499–507.
[140] K.H. Steen, P.W. Reeh, F. Anton, H.O. Handwerker, Protons [157] C. Zhang, S.W. Yang, Y.G. Guo, J.T. Qiao, N. Dafny, Locus
selectively induce lasting excitation and sensitization to mech- coeruleus stimulation modulates the nociceptive response in
anical stimulation of nociceptors in rat skin, in vitro, J. Neu- parafascicular neurons: an analysis of descending and ascend-
rosci 12 (1) (1992) 86–95. ing pathways, Brain Res. Bull 42 (4) (1997) 273–278.
[141] K.H. Steen, A.E. Steen, P.W. Reeh, A dominant role of acid
pH in inflammatory excitation and sensitization of nociceptors
in rat skin, in vitro, J. Neurophysiol. 15 (5) (1995) 3982–3989.
[142] C. Stein, C. Gramsch, A.H. Hassan, R. Przewlocki, C.G.
Peter, K. Peter, A. Herz, Local opioid receptors mediating anti-
nociception in inflammation: endogenous ligands, Prog. Clin. Partap S. Khalsa, D.C., Ph.D., F.A.C.O. is
Biol. Res 328 (1990) 425–427. an Associate Professor of Biomedical Engin-
[143] C. Stein, A.H. Hassan, R. Przewlocki, C. Gramsch, K. Peter, eering, Orthopaedics, and Neurobiology at
A. Herz, Opioids from immunocytes interact with receptors on the State University of New York(SUNY) at
sensory nerves to inhibit nociception in inflammation, Proc. Stony Brook, and currently serves as the
Natl Acad. Sci. USA 87 (15) (1990) 5935–5939. Graduate Program Director for the Depart-
[144] H. Sun, J. Xu, K.B. Della Penna, R.J. Benz, F. Kinose, D.J. ment of Biomedical Engineering. Dr Khalsa
Koblan, K.S. Koblan, D.L. Gerhold, H. Wang, Dorsal horn- has an active research laboratory investigat-
enriched genes identified by DNA microarray, in situ hybridiza- ing neurophysiological mechanisms of
tion and immunohistochemistry, BMC Neurosci 3 (1) (2002) mechanosensory neurons, soft tissue bio-
11. mechanics, spine biomechanics, and mechan-
[145] Y.X. Tao, Y.Q. Li, Z.Q. Zhao, R.A. Johns, Synaptic relation- isms of muscle and low back pain. He maintained, for 17 years, a
ship of the neurons containing a metabotropic glutamate recep- private practice of chiropractic in Massachusetts, is board-certified in
tor, MGluR5, with nociceptive primary afferent and chiropractic orthopaedics, and served terms as the President and
GABAergic terminals in rat spinal superficial laminae, Brain Vice-President of his local chapter of the Massachusetts Chiropractic
Res 875 (1-2) (2000) 138–143. Society.

You might also like