You are on page 1of 34

Central Nervous System Pain Pathways

Oxford Handbooks Online


Central Nervous System Pain Pathways  
Andrew J. Todd and Fan Wang
The Oxford Handbook of the Neurobiology of Pain
Edited by John Wood

Subject: Neuroscience, Molecular and Cellular Systems Online Publication Date: Jul 2018
DOI: 10.1093/oxfordhb/9780190860509.013.5

Abstract and Keywords

Nociceptive primary afferents detect stimuli that are normally perceived as painful, and
these afferents form synapses in the dorsal horn of the spinal cord and the spinal
trigeminal nucleus. Here they are involved in highly complex neuronal circuits involving
projection neurons belonging to the anterolateral tract (ALT) and interneurons, which
modulate the incoming sensory information. The ALT neurons convey somatosensory
information to a variety of brain regions that are involved in the various aspects of the
pain experience. A spinothalamic-cortical pathway provides input to several regions of
the cerebral cortex, including the first and second somatosensory areas (S1, S2), the
insula and the cingluate cortex. These regions are thought be responsible for the sensory-
discriminative aspects of pain (S1), pain-related learning (S2), the autonomic and
motivational responses (insula), and the negative affect (cingulate). Another ascending
system, The spinoparabrachial-limbic pathway targets a variety of brain regions,
including the amygdala, and is likely involved in the affective component of pain. A
descending system that includes the limbic system, the periaqueductal gray matter of the
midbrain, the locus coeruleus, and the rostral ventral medulla, can suppress pain, and
this operates partly through the monoamine transmitters noradrenaline and serotonin
which are released in the spinal and trigeminal dorsal horn.

Keywords: spinal dorsal horn, neuronal circuit, anterolateral tract neuron, spinothalamo-cortical pathway,
spinoparabrachial-limbic pathway

Basic Organization of Pain Pathways

Page 1 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Primary afferent neurons can respond to a variety of stimuli that affect the body surface
or internal tissues and organs. Many of these primary afferents are tuned to detect
stimuli that cause actual or potential tissue damage, and these are known as nociceptors.
Primary afferents from the trunk and limbs enter the spinal cord through the dorsal roots,
while those from the head are conveyed via the trigeminal (fifth cranial) nerve to
trigeminal nuclei in the brainstem. All of these afferents, which use glutamate as their
principal fast transmitter, form excitatory synapses on neurons within the spinal dorsal
horn and trigeminal nuclei. Both of these regions contain large numbers of neurons. Most
of these are interneurons, with axons that arborize locally, giving rise to complex circuits
that process somatosensory information. This information is then conveyed to the brain
via projection neurons. These are relatively large cells with axons that enter the white
matter and are organized into ascending tracts that terminate in various somatosensory
brain regions. In addition to this ascending system, there are also descending pathways,
which originate from the brainstem and from cortical regions (primarily somatosensory
cortex), and terminate diffusely within the spinal dorsal horn and trigeminal nuclei.

The spinal dorsal horn can be divided into six parallel laminae based on neuronal size and
packing density (Rexed, 1952), and this scheme has been used to define the regions
targeted by different types of primary afferent (Figure 1), as well as to define specific
populations of spinal cord neurons. The dorsal horn of the spinal cord is somatotopically
organized, with the body surface being mapped in a two-dimensional pattern on the
rostrocaudal and mediolateral axes. The dorsoventral axis (i.e., across the laminae)
reflects a modality-specific pattern that is determined by the termination of different
types of primary afferent (e.g., nociceptors, thermoreceptors, and low-threshold
mechanoreceptors). Most nociceptive afferents terminate in the superficial part of the
spinal dorsal horn (laminae I and II), or in the most caudal of the trigeminal nuclei, which
is known as the spinal trigeminal nucleus caudalis (SpVc). The SpVc nucleus has a similar
lamination pattern to the spinal dorsal horn, and is also known as the trigeminal or
medullary dorsal horn.

Over 50 years ago, Melzack and Wall (1965) proposed that neuronal circuits within the
dorsal horn could “gate” the sensory information conveyed by nociceptive primary
afferents, and thus modulate the perception of pain. More recent studies have revealed
that the circuitry within this region is highly complex, and we are still far from a complete
understanding of how this is organized. However, it is clear that the spinal dorsal horn
and SpVc play a major role in modulating pain in both normal and pathological states.
Indeed, much of the information transmitted by projection neurons has already
undergone significant processing. The spinal and medullary dorsal horns are likely to
provide important targets for new pain therapies, as they contain numerous receptors
and signaling pathways. This chapter will summarize what we know about the anatomy of
pain pathways from the periphery to higher brain centers, with detailed emphasis on the
organization of neuronal populations and circuits in the spinal cord and SpVc.

Page 2 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Primary Afferent Input to the Spinal and


Trigeminal Dorsal Horn
The organization of primary afferents is dealt with elsewhere, and here we are concerned
with the central projections of these neurons. Early electrophysiological classification
schemes for primary afferents have recently been expanded by transcriptomic
approaches (Li et al., 2016; Usoskin et al., 2015), which have revealed several distinctive
afferent classes, particularly among the fine-diameter (unmyelinated C, and thinly
myelinated Aδ) afferents, most of which function as nociceptors, thermoreceptors, or
pruritoceptors (“itch detectors”). Although the focus of the chapter is on pain, the low-
threshold mechanoreceptors (LTMRs) are also important, because of their likely
involvement in touch-evoked pain (tactile allodynia), which is a feature of certain chronic
pain conditions (Campbell, Raja, Meyer, & Mackinnon, 1988).

The primary afferent input to the spinal dorsal horn is represented schematically in
Figure 1. Although our understanding of the central projections of myelinated afferents
has come mainly from studies in which individual afferents were labelled in combined
electrophysiological/anatomical studies (Brown, 1981), there have been very few studies
of this type for unmyelinated (C) fibers (Sugiura, Lee, & Perl, 1986). Much of what we
know about these is based on the identification of primary afferent terminals by means of
neurochemical markers.

Myelinated LTMRs (A-


LTMRs) arborize in the
deeper part of the dorsal
horn, in a region that
extends from the inner
part of lamina II (lamina
Click to view larger
IIi) to lamina V, and each
Figure 1 Central projections of different classes of
of the different functional
primary afferent. populations has a specific
The major classes of primary afferent that have been laminar distribution
identified in recent transcriptomic studies (Li et al., pattern (Abraira & Ginty,
2016; Usoskin et al., 2015) are listed, together with
what is currently known about their functions and
2013). Some C fibers
their central terminations in the spinal dorsal horn. respond to non-noxious
Note that pruritoceptive afferents may also function
hair movement as well as
as nociceptors, and that the C-MrgD nociceptors may
act as pruritoceptors. Afferents that respond to cooling (C-LTMRs), and
innocuous cooling or warming appear to be poorly these terminate in lamina
represented in these studies, and are therefore not
included in this scheme.
IIi (Seal et al., 2009).
Myelinated nociceptors,
most of which are Aδ fibers, convey “fast pain.” Many of these have a compact
termination zone in lamina I and the outer part of lamina II (lamina IIo) (Light & Perl,
1979), although others appear to arborize diffusely throughout laminae I–V (Boada &
Page 3 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Woodbury, 2008; Woodbury & Koerber, 2003). The remaining C fibers consist mainly of
nociceptors and thermoreceptors, with some also responding to pruritic stimuli. Although
several neurochemically distinct classes of primary afferent have been revealed in
transcriptomic studies, we do not yet fully understand how these map onto functionally
defined populations. Many C fibers (and some Aδ nociceptors) express neuropeptides,
and one well-characterized population consists of cells that contain calcitonin gene-
related peptide (CGRP), substance P, and galanin. These are thought to function as
nociceptors, and terminate mainly in laminae I and IIo, with some branches penetrating
deeply. These afferents express the transient receptor potential channel vanilloid 1
(TRPV1), and are required for perception of heat pain (Cavanaugh et al., 2009). Another
major class of C nociceptors is defined by expression of the Mas-related G protein-
coupled receptor D (MrgD). These afferents, which lack neuropeptides and do not
express TRPV1 in the mouse, have a very compact termination pattern in the middle part
of lamina II, and are needed for the normal perception of mechanical pain (Cavanaugh et
al., 2009; Zylka, Rice, & Anderson, 2005). Two further groups of C fibers that have been
recognized are: (1) those defined by co-expression of the neuropeptides somatostatin and
natriuretic polypeptide B (NPPB), and (2) those that express the Mas-related G protein-
coupled receptors MrgA3/MrgC11. There is evidence that both of these populations
function as pruritoceptors (L. Han et al., 2013; Huang et al., 2018), and both arborize in
the middle part of lamina II.

Since all primary afferents are glutamatergic, they require vesicular glutamate
transporters (VGLUTs) to enrich the amino acid in their synaptic vesicles. These
transporters are differentially expressed among primary afferents. A-LTMRs express
VGLUT1, and account for ~60% of the VGLUT1-immunoreactive boutons in the dorsal
horn, with the remainder being corticospinal tract terminals (Abraira et al., 2017; Todd et
al., 2003). VGLUT2 is expressed by Aδ nociceptors and most C fibers, although the level
of VGLUT2 protein that can be detected in their central terminals with
immunocytochemistry is generally very low (Brumovsky, Watanabe, & Hokfelt, 2007; Todd
et al., 2003). C-LTMRs express VGLUT3, and since they are the major source of this
transporter, antibodies against VGLUT3 can be used to reveal their central terminals,
which are located in the innermost part of lamina II (Seal et al., 2009).

Central terminals of primary afferents possess a variety of receptors that are likely to be
important in regulating their function through presynaptic modulatory mechanisms.
These include ionotropic glutamate receptors, GABAA and GABAB receptors, and
purinergic (P2X3) receptors. They also express receptors for various neuropeptides,
including μ, δ, and κ opioid receptors (Todd & Koerber, 2013).

As stated before, all primary afferents give rise to boutons that form glutamatergic
synapses on dorsal horn neurons. A characteristic feature of some classes of primary
afferent is that their central terminals form complex arrangements known as synaptic
glomeruli. These contain a central bouton (the primary afferent terminal), surrounded by
several dendritic and/or axonal profiles. The dendritic components may be dendritic
spines or shafts, and are postsynaptic to the primary afferent terminal, although they may

Page 4 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

form reciprocal (axodendritic/dendroaxonic) synaptic arrangements. The peripheral


axons in the synaptic glomerulus, which are GABAergic, are presynaptic to the primary
afferent terminal at axoaxonic synapses (Todd, 1996), which are the substrate for
classical presynaptic inhibition. In addition, there can be triadic arrangements, in which a
peripheral axon is presynaptic both to the central (primary afferent) bouton and a
dendrite, with the latter also receiving a synapse from the primary afferent. Two distinct
types of synaptic glomerulus (types I and II) have been identified in rodent dorsal horn,
and these differ in both the appearance of the central (primary afferent) bouton and the
arrangement of peripheral profiles (Ribeiro-da-Silva & Coimbra, 1982) (Figure 2). These
two types are thought to be associated with C-MrgD nociceptors and Aδ D-hair afferents,
respectively. Other classes of primary afferent can also be associated with synaptic
glomeruli. These include some of the substance P/CGRP-containing afferents, as well as
those that express NPPB/somatostatin (Ribeiro-da-Silva, Tagari, & Cuello, 1989; Salio,
Ferrini, Muthuraju, & Merighi, 2014). Other primary afferents form simpler synaptic
arrangements, but they may still be subject to presynaptic inhibition mediated by
axoaxonic synapses.

Click to view larger

Page 5 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Figure 2 Two types of synaptic glomerulus in the


superficial dorsal horn of the rat.

The terminals of some types of primary afferent form


central boutons of synaptic glomeruli within the
dorsal horn. A: Central axons of type I glomeruli (C)
are derived from non-peptidergic C nociceptors, and
are typically surrounded by vesicle-containing
dendrites (V), dendrites that lack vesicles (D), and
peripheral axons (A). The peripheral axon is
presynaptic at an axoaxonic synapse with the central
bouton. The central bouton forms synapses with both
types of dendrite, and can receive dendroaxonic
synapses from the vesicle-containing dendrites. In
some cases, reciprocal axodendritic synapses are
present, as shown in the inset. B: Central axons of
type II glomeruli (C) are thought to originate from Aδ
D-hair endings, and are presynaptic to dendrites (D),
most of which lack vesicles. These glomeruli are
often surrounded by several peripheral axons (A),
which are presynaptic to the C bouton, and can form
triadic synapses involving the C bouton and a
dendrite. Arrows in both parts indicate synapses.
Scale bar: 1 μm.

Source: Todd (1996). Reproduced with permission of


John Wiley & Sons.

Dorsal Horn Interneurons


Interneurons are thought to account for ~99% of all of the nerve cells in the spinal dorsal
horn (Abraira & Ginty, 2013; Todd, 2010, 2017). Interneurons are found in all dorsal horn
laminae, but here we will focus on those in laminae I–III, as these have been extensively
studied and are likely to play important roles in pain mechanisms. All interneurons so far
identified appear to give rise to local axonal projections that arborize within the same
spinal cord segment that contains the cell body. However, retrograde labelling studies
have shown that a significant proportion of the interneurons in this region also have
axons that project for several segments (Bice & Beal, 1997). Until recently, the
postsynaptic targets of these “propriospinal” axons was unknown, but we have found that
at least some of them terminate in the lateral spinal nucleus, a collection of neurons
located in the dorsal part of the lateral white matter (Gutierrez-Mecinas, Polgár, Bell,
Herau, & Todd, 2018). Spinal interneurons are anatomically and physiologically
heterogeneous, and there have been numerous attempts to assign them to specific
functional populations. It is essential to be able to define these populations, so that we
can investigate their contributions to synaptic circuits and their roles in somatosensory
processing.

Page 6 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Classification of Interneurons in Laminae I–III

Excitatory and Inhibitory Interneurons

A basic functional distinction can be made between excitatory interneurons, which use
glutamate as their principal neurotransmitter, and inhibitory interneurons, which are
GABAergic and/or glycinergic. Immunocytochemistry to reveal GABA and glycine has
been used to identify the inhibitory interneurons, but this is technically difficult, and a
more convenient way is to reveal the transcription factor Pax2, which is expressed by all
of these cells (Foster et al., 2015; Larsson, 2017). Quantitative studies in the mouse
reveal that ~25% of neurons in laminae I–II and 40% of those in lamina III are inhibitory,
and these are all thought to be interneurons (Polgár, Durrieux, Hughes, & Todd, 2013).
The remaining neurons are assumed to be glutamatergic, and these will include both
projection neurons (see following discussion) and excitatory interneurons. Axons
belonging to either inhibitory or excitatory interneurons can be identified based on their
expression of vesicular transporters: the vesicular GABA transporter (VGAT) and
VGLUT2, respectively. This method can be applied to neurons that have undergone whole-
cell patch-clamp recording if an appropriate tracer (e.g., Neurobiotin) has been included
in the recording pipette, and this allows the neurotransmitter phenotype of
electrophysiologically characterized neurons to be determined (Yasaka, Tiong, Hughes,
Riddell, & Todd, 2010).

Page 7 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Morphological and Electrophysiological Classification of


Interneurons

In many parts of the central nervous system (CNS), the dendritic morphology of neurons
is closely related to their function; therefore, numerous studies have attempted to identify
morphological classes among the interneurons in laminae I–III. Initially, these were
carried out with Golgi staining, but more recently they have been performed on neurons
that were labelled during whole-cell patch-clamp recording experiments, thus allowing
correlations between structure and function to be investigated. The most widely accepted
classification scheme based on these approaches has been that of Grudt and Perl (2002),
who identified four major types of neuron in lamina II: islet, vertical, radial, and central
cells (Figure 3). Islet cells have axonal and dendritic trees that are highly elongated along
the rostrocaudal axis; vertical cells have a conical dendritic tree that extends ventrally
from the soma; radial cells have radiating dendrites that form a relatively compact
dendritic tree; central cells are similar in appearance to islet cells but with much shorter
dendritic trees. There were also physiological differences between these cells, since
radial and vertical cells often showed delayed action potential firing in response to
injection of depolarizing current, whereas islet cells fired continuously (tonically)
throughout the current injection. Central cells were further subdivided, based on their
firing pattern and the presence or absence of an A-type potassium current (IA current).
However, although other studies have identified cells of these morphological types, a
limitation of this scheme is that a significant proportion of cells cannot be assigned to any
of these classes.

Yasaka et al. (2010)


recorded from a large
sample of lamina II
interneurons and
compared morphology and
firing patterns with
Click to view larger neurotransmitter
Figure 3 Morphological classification of lamina II phenotypes. They reported
neurons.
that all islet cells were
Four main morphological types of interneuron have
inhibitory, whereas radial
been identified in lamina II: islet cells, central cells,
radial cells, and vertical cells. Examples of each type cells and most vertical
are shown here (scale bar: 100 μm). cells were excitatory.
Source: Yasaka et al. (2010). Reproduced with However neurons that
permission from Wolters Kluwer Health, Inc.
were classified as central
cells could be either
inhibitory or excitatory, and there were many “morphologically unclassified” neurons
among both transmitter types. They also reported that firing pattern was related to
function, since most excitatory cells showed firing patterns associated with IA current
(delayed, gap, or reluctant firing), whereas these were seldom seen in inhibitory neurons.
These findings suggest that, although there is a correlation between morphology and
Page 8 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

function, neither morphology nor firing pattern can provide a comprehensive


classification scheme for interneurons in lamina II. Even less is known about interneurons
in other laminae, but it seems likely that a similar degree of caution is needed when
defining populations based on these parameters.

Neurochemical Classification of Interneurons

The dorsal horn contains a wide variety of neuropeptides, neuropeptide receptors, and
other proteins that are differentially expressed by populations of neurons in laminae I–III.
Some of these molecules are preferentially expressed by either excitatory or inhibitory
interneurons (Todd, 2017; Todd & Spike, 1993). For example, several neuropeptides,
including somatostatin, substance P, gastrin-releasing peptide (GRP), neurotensin,
neurokinin B (NKB), and cholecystokinin, are found mainly or exclusively in excitatory
neurons. In contrast, neuropeptide Y (NPY), galanin, and nociceptin are restricted to
inhibitory interneurons, while the opioid peptides enkephalin and dynorphin are found in
both types of interneuron. Calcium-binding proteins are also highly expressed in this
region, with calbindin and calretinin being found predominantly in excitatory neurons,
and parvalbumin mainly in inhibitory cells.

These neurochemical markers often show a non-uniform laminar distribution, which


suggests that they may be expressed by specific subpopulations among the interneurons
in laminae I–III. We have recently confirmed this by showing that non-overlapping
neurochemical populations can be found within both major classes (Boyle et al., 2017;
Gutierrez-Mecinas et al., 2017; Gutierrez-Mecinas, Furuta, Watanabe, & Todd, 2016). In
laminae I–III of the rat dorsal horn, we identified four largely separate classes of
inhibitory interneuron, based on expression of NPY, galanin, parvalbumin, and the
neuronal form of nitric oxide synthase (nNOS) (Polgár, Sardella, et al., 2013). The
galanin-containing cells also express dynorphin, although this is also found in some
excitatory interneurons. We have since found a very similar pattern in the mouse dorsal
horn, although there is significant overlap between the neurons that express nNOS and
those with galanin/dynorphin. Between them, we found that these four populations
accounted for ~75% of the inhibitory interneurons in laminae I–II of the mouse (Boyle et
al., 2017), and the remaining 25% appear to correspond to the inhibitory interneurons
that express the calcium-binding protein calretinin (Smith et al., 2015). Among the
excitatory interneurons, we have recently shown that four neuropeptides—substance P,
GRP, neurotensin, and NKB—are expressed in largely separate subpopulations, which
account for nearly two-thirds of the excitatory neurons in laminae I–II (Gutierrez-Mecinas
et al., 2017). Another peptide, somatostatin, is also restricted to excitatory interneurons,
but this is far more widely expressed. The excitatory interneurons that contain
neurotensin or NKB largely correspond to a previously defined class of neurons: those
that express the γ isoform of protein kinase C (PKCγ). The relative sizes of the different
neurochemical populations of excitatory and inhibitory interneurons within the superficial
laminae are shown in Figure 4.

Page 9 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

There is some evidence


relating neurochemistry to
morphology. Among the
inhibitory interneurons,
those that contain
Click to view larger
parvalbumin or calretinin
Figure 4 Neurochemically defined populations often show islet
among the excitatory and inhibitory interneurons in
the superficial dorsal horn of the mouse spinal cord. morphology (Hughes et al.,
The pie charts show the relative sizes of the
2012; Smith et al., 2015),
neurochemical populations among the excitatory and whereas those that
inhibitory interneurons in laminae I–II of the mouse
express NPY, galanin/
dorsal horn. For the inhibitory interneurons, the
unmarked wedges indicate overlap between the NOS dynorphin, or nNOS are
and Galanin/Dynorphin populations (purple wedge) never islet cells and are
and between the Galanin/Dynorphin and NPY
populations (brown wedge). Note that a substantial
morphologically
proportion of the excitatory interneurons is not heterogeneous (Todd,
accounted for by the expression of the four
2017). Less is known about
neuropeptides shown. NKB: neurokinin B, NTS:
neurotensin; GRP: gastrin-releasing peptide; SP: the morphology of
substance P; CR: calretinin; NPY: neuropeptide Y; neurochemically defined
Gal: galanin; Dyn: dynorphin; NOS: neuronal nitric
oxide synthase; PV: parvalbumin.
excitatory interneuron
populations, although both
Source (part b): Boyle et al. (2017).
the PKCγ cells (which
include neurotensin and
NKB populations) and the GRP cells appear to have central or “unclassified” morphology.
We have recently found that many of the substance P-expressing neurons correspond to
radial cells, but interestingly, none of these four populations (substance P-, GRP-,
neurotensin-, or NKB-expressing cells) seem to include vertical cells (Bell, Dickie,
Iwagaki, Polgár, & Todd, unpublished data).

Functions of Dorsal Horn Interneurons

Early insight into the role of inhibitory interneurons came from studies involving spinal
administration of GABAA and glycine receptor antagonists (Sivilotti & Woolf, 1994; Yaksh,
1989), since these interneurons are the main source of GABAergic and glycinergic
synapses within the dorsal horn. These studies showed that blocking inhibitory synaptic
transmission resulted in exaggerated pain responses, and led to the view that inhibitory
interneurons are involved in suppressing both acute and pathological pain (Sandkuhler,
2009; Zeilhofer, Wildner, & Yevenes, 2012). More recently, molecular-genetic approaches
have been used to target specific populations of interneurons and reveal their function.
Foster et al. (2015) used intraspinal injection of viral vectors to manipulate the activity of
glycinergic neurons, based on selective expression of the neuronal glycine transporter
(GlyT2) by these cells. GlyT2-expressing cells account for a large proportion of dorsal
horn inhibitory interneurons: ~20% of those in laminae I–II and ~90% of those in deeper
laminae. Ablation or inactivation of these cells resulted in hypersensitivity to both

Page 10 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

mechanical and thermal stimuli, together with apparent signs of spontaneous itching. In
contrast, activation of the cells suppressed the responses to acute noxious stimuli, as well
as reducing signs of neuropathic pain, and itch behaviors following injection of
pruritogens. Other studies have investigated the roles of smaller populations of inhibitory
interneurons. Parvalbumin-expressing cells are likely to have a role in regulating tactile
input, since they receive synapses from A-LTMRs, and their axons give rise to axoaxonic
synapses on these afferents (Hughes et al., 2012). Consistent with this suggestion,
Petitjean et al. (2015) showed that ablation of parvalbumin-expressing dorsal horn
interneurons caused mechanical allodynia, whereas activating them increased withdrawal
thresholds for mechanical (but not thermal) noxious stimuli and reduced mechanical
allodynia after peripheral nerve injury. There is conflicting evidence concerning the role
of the dynorphin-expressing inhibitory interneurons. Kardon et al. (2014) reported that
these cells, together with those that expressed nNOS, were selectively lost in mice
lacking the transcription factor Bhlhb5. Since these knockout mice develop exaggerated
itch (Ross et al., 2010), they suggested that the dynorphin/galanin and/or nNOS
populations may play a role in suppressing pruritogen-evoked itch. However, Duan et al.
(2014) ablated dynorphin cells and observed an increase in response to noxious
mechanical (but not thermal) stimuli and no change in itch behavior. We have recently
shown that chemogenetic activation of dynorphin cells suppresses itch, whereas
activation of the nNOS cells increased the thresholds for both mechanical and heat pain
(Huang et al., 2018). This suggests that the dynorphin neurons are anti-pruritic, while the
nNOS cells are anti-nociceptive. Cui et al. (2016) identified a population of inhibitory
interneurons that were mainly located in lamina III, and were defined by the early
expression of the receptor tyrosine kinase RET. They showed that ablating these cells
resulted in mechanical allodynia and increased responses to mechanical and thermal
noxious stimuli, as well as increased hyperalgesia in inflammatory and neuropathic
models. In contrast, activating them reduced acute pain and hyperalgesia. François et al.
(2017) reported that Penk1+ (enkephalin expressing) GABAergic interneurons gate
nociceptive sensory transmission through coordinated GABA- and enkephalin-mediated
presynaptic inhibition of sensory afferents.

Several studies have investigated the functions of excitatory interneurons. Mice lacking
PKCγ show normal acute pain thresholds but fail to develop mechanical or thermal
allodynia after nerve injury (Malmberg, Chen, Tonegawa, & Basbaum, 1997), which led to
the suggestion that the PKCγ-expressing excitatory interneurons in laminae II–III are
required for neuropathic pain. Duan et al. (2014) ablated cells that expressed
somatostatin, and found that this dramatically reduced acute mechanical pain, as well as
the mechanical allodynia seen in both neuropathic and inflammatory models. Consistent
with these findings, Christensen et al. (2016) showed that activating somatostatin cells
resulted in pain-like behavior, and confirmed that inactivating them reduced mechanical
pain and allodynia, although they also saw a slight increase in latency of withdrawal to
noxious heat. Somatostatin appears to be expressed by a large and heterogeneous
population of excitatory interneurons in laminae I–II, and these findings demonstrate an
important role for these cells in pain evoked by mechanical stimuli. Peirs et al. (2015)

Page 11 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

showed that transient expression of VGLUT3 defined a large population of excitatory


interneurons in lamina III, which overlapped with PKCγ-expressing neurons. They showed
that deletion of VGLUT3 from these cells attenuated acute mechanical pain and reduced
mechanical allodynia caused by nerve injury or inflammation, whereas chemogenetic
activation of the cells evoked mechanical allodynia but no change in thermal pain
thresholds. The receptor for GRP (GRPR) is expressed by a population of excitatory
interneurons in laminae I–IIo (Sun & Chen, 2007), and there are several lines of evidence
that implicate these cells in the itch elicited by injection of pruritogens (Sun & Chen,
2007; Sun et al., 2009). Mice lacking GRPR, and those in which the GRPR-expressing cells
have been ablated, show reduced responses to pruritogens, while intrathecal GRPR
agonists evoke itch behaviors.

Taken together, the results of these studies indicate that several populations of dorsal
horn interneurons are involved in pain and itch mechanisms, and that these are likely to
have complex and overlapping functions.

Spinal Projection Neurons and Ascending


Tracts
Spinal cord neurons with axons that reach the brain (projection neurons) can be revealed
in anatomical studies by injection of retrograde tracers into regions of the brain to which
they project (Figure 5). Major targets of dorsal horn projection neurons include the
thalamus, the periaqueductal gray (PAG) matter of the midbrain, the lateral parabrachial
area (LPb) in the pons, and various nuclei in the medulla. Dorsal horn neurons projecting
to these brain regions are concentrated in lamina I and scattered throughout the deep
dorsal horn (III–VI), but are largely absent from lamina II. Anatomical and
electrophysiological studies have shown that there is extensive collateralization of axons,
with many cells sending axons to several different brain nuclei (McMahon & Wall, 1985;
Spike, Puskar, Andrew, & Todd, 2003). The projection is mainly to the contralateral side,
although some dorsal horn neurons appear to project symmetrically to both sides of the
brain (Spike et al., 2003), especially the projection from trigeminal SpVc neurons to LPb,
which appears to be bilateral (Saito et al., 2017). Axons of dorsal horn projection neurons
cross the midline and ascend in the white matter of the lateral or ventral funiculus.
Although this system is often referred to as the anterolateral tract (ALT), based on its
location in the anterolateral white matter in the human spinal cord, in rodents at least
some of it occupies the dorsal part of the lateral funiculus (Gutierrez-Mecinas et al.,
2018; McMahon & Wall, 1983). This difference in position may result from the presence
of a large lateral corticospinal tract in humans, but not rodents, which would displace the
ALT ventrally.

Page 12 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Quantitative studies in the lumbar enlargement of the rat spinal cord have shown that
nearly all of the ALT neurons in lamina I send axons to the LPb, with ~30% targeting the
PAG, but surprisingly only ~5% of these cells have axons that reach the thalamus (Todd,
2010). The situation is very different in the cervical enlargement, where the lamina I
spinothalamic component is much larger (~40% of all projection neurons) (Polgár,
Wright, & Todd, 2010), and this appears to be similar to the relative size of the
spinothalamic projection from lamina I in both lumbar and cervical enlargements of the
cat and monkey (Zhang & Craig, 1997; Zhang, Han, & Craig, 1996). The anatomy and the
functions of the spinoparabrachial and spinothalamic pathways are discussed later in this
chapter.

Many of the lamina I ALT


neurons (80% in rat, 90%
in mouse) express the
neurokinin 1 receptor
(NK1r), which is a target
for substance P released
from nociceptive primary
afferents and some
excitatory interneurons
(Cameron et al., 2015;
Click to view larger Mantyh et al., 1995; Todd,
Figure 5 Anterolateral tract (ALT) projection 2010). These cells are
neurons in the rat lumbar enlargement. known to be densely
(A) shows a transverse section from the L4 segment innervated by substance P-
of a rat that had received injections of cholera toxin
B subunit (CTb) into the caudal ventrolateral
containing primary
medulla, and of Fluorogold into the lateral afferents (Todd et al.,
parabrachial area. The section was immunostained to 2002), indicating that they
reveal CTb (red), Fluorogold (green), and the
neuronal marker NeuN (blue). The locations of can be activated both by
Rexed’s laminae are indicated. Tracer injections into glutamatergic synaptic
these two sites can label virtually all of the projection
neurons in lamina I, as well as scattered cells
input and by substance P
throughout the deep dorsal horn (laminae III–VI). acting through volume
Note that some cells have taken up both tracers, and transmission. Not all
therefore appear yellow. Scale bar = 100 μm. (B)
Quantitative data showing the approximate numbers lamina I ALT neurons
of neurons in laminae I and III in the rat L4 spinal express the NK1r, and
segment that can be retrogradely labeled from
various brain regions. LSN: lateral spinal nucleus;
among those that lack the
PAG: periaqueductal grey matter; LPb: lateral receptor, there is a sparse
parabrachial area; NTS: nucleus of the solitary tract; population of very large
CVLM: caudal ventrolateral medulla.
neurons (giant cells) that
Source: Todd (2010). Reproduced with permission
from Springer Nature.
appear to lack significant
primary afferent input
(Polgár, Al-Khater, Shehab, Watanabe, & Todd, 2008).

Page 13 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Both electrophysiological studies and anatomical studies involving the activity-dependent


transcription factor Fos have shown that lamina I projection neurons respond to a variety
of noxious and pruritic stimuli (Akiyama, Curtis, Nguyen, Carstens, & Carstens, 2016;
Andrew, 2009; Bester, Chapman, Besson, & Bernard, 2000; Moser & Giesler, 2014; Todd
et al., 2002). Less is known about ALT projection neurons in deeper laminae, although
one type has been characterized in some detail. This consists of large neurons in lamina
III or IV that have long dorsal dendrites extending into lamina I (Todd, 2010). These cells
are also densely innervated by substance P-expressing nociceptive afferents, and in the
rat (but not the mouse), they all express the NK1r (Cameron et al., 2015). They have also
been shown to respond to noxious stimuli (Polgár, Campbell, MacIntyre, Watanabe, &
Todd, 2007). Not surprisingly, ablation of NK1r-expressing neurons, including ALT cells in
laminae I and III–IV, led to a substantial reduction in hyperalgesia in both neuropathic
and inflammatory pain models (Nichols et al., 1999), although acute pain thresholds were
apparently unchanged. This suggests that the NK1r-expressing ALT projection neurons
are involved in pathological pain, but are not required for detection of acute pain.

Based on the findings in patients with anterolateral cordotomy, it is thought that the ALT
is responsible for perception of stimuli perceived as pain and itch, as well as temperature
sensation. However, it is not the only ascending system to originate from the dorsal horn.
Two additional pathways have been identified in other mammalian species: the
postsynaptic dorsal column pathway (PSDC) and the spinocervical tract. Both of these
have been shown to originate from cells in the deeper laminae of the dorsal horn, but
much less is known about their functions (Abraira & Ginty, 2013; Brown, 1981).

Neuronal Circuits in the Dorsal Horn


Our knowledge about the synaptic connections between different neuronal components in
the dorsal horn has come from two types of experimental approach. Firstly,
immunocytochemical studies have been able to demonstrate contacts (and in some cases
synapses) between axons and cell bodies or dendrites of neurochemically defined
neuronal populations. Secondly, whole-cell patch-clamp recordings either made from
pairs of neurons, or else combined with dorsal root stimulation, have been used to
identify monosynaptic connections. Here, we will restrict ourselves to the synaptic
circuits in which both components are well characterized. A diagram illustrating some of
the circuits described here is shown in Figure 6.

Page 14 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Click to view larger


Figure 6 A diagram showing some of the synaptic
circuits that have been identified in the spinal dorsal
horn.

In this diagram, inhibitory interneurons are shown in


pink, excitatory interneurons in green, projection
neurons in blue, and primary afferents in brown.
Three ALT projection neurons are shown: cells in
lamina I and III that express the neurokinin 1
receptor (NK1r) and a giant lamina I cell. Peptidergic
nociceptors (noci) densely innervate the NK1r+ ALT
neurons, and the lamina III cells also receive input
from A-LTMRs. The lamina III ALT neurons are
selectively innervated by NPY+ inhibitory
interneurons and dynorphin+ (DYN) excitatory
interneurons. The giant ALT cells apparently do not
receive primary afferent input, but are densely
innervated by nNOS+ inhibitory interneurons and by
unknown class(es) of excitatory (glutamatergic)
interneuron (GLU). A circuit involving three types of
excitatory interneuron (PKCγ+, transient central,
and vertical cells) is thought to convey input from A-
LTMRs to lamina I ALT neurons. Relatively little is
known about the primary afferent input to the
vertical cells, but this is thought to include Aδ
nociceptors and probably A-LTMRs. Inhibitory
interneurons that express parvalbumin (PV) are
innervated by A-LTMRs and generate axoaxonic
synapses on these afferents (note that for simplicity,
only one of these axoaxonic synapses is shown).

Synaptic Inputs to Projection Neurons

As we noted, NK1r-expressing ALT neurons in lamina I, as well as those in laminae III–IV,


receive a dense synaptic input from substance P-expressing primary afferents, which are
likely to correspond to peptidergic C fibers. It has been estimated that this input accounts
for over half of the excitatory synaptic input to these neurons (Polgár, Al Ghamdi, & Todd,
2010). This presumably represents a powerful monosynaptic input from nociceptive
primary afferents. The lamina III–IV ALT neurons receive some direct input from
myelinated primary afferents in the deeper laminae, and these are thought to be A-
LTMRs. However, despite having dendrites that pass through the superficial laminae, they

Page 15 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

do not seem to receive significant input from non-peptidergic C nociceptors (Todd, 2010).
This indicates that synaptic connectivity is selectively organized, and does not depend
only on the proximity to specific types of axon.

All of the ALT neurons receive numerous contacts from boutons with strong VGLUT2
immunoreactivity, which are thought to originate from local excitatory interneurons. In
the case of lamina I giant cells, these appear to represent the main (or exclusive) source
of excitatory synaptic input (Polgár et al., 2008). At present, we have limited information
about the types of excitatory interneuron that are presynaptic to projection neurons, but
these are likely to include vertical cells in lamina II (Cordero-Erausquin et al., 2009; Lu &
Perl, 2005). It has been shown that 60% of the VGLUT2-immunoreactive boutons that
synapse on lamina III ALT cells in the rat contain preprodynorphhin, which is only present
in ~5% of all VGLUT2 boutons in this region (Baseer et al., 2012). This therefore suggests
that these cells receive a highly selective input from dynorphin-expressing excitatory
interneurons.

ALT projection neurons also receive synapses from local inhibitory interneurons, and
again there is some evidence that this is organized in a selective manner (Todd, 2010).
The lamina III ALT cells are densely innervated by axons that contain high levels of NPY,
and these may well originate from a specific subset of NPY-expressing GABAergic
interneurons (Polgár, Sardella, Watanabe, & Todd, 2011). In contrast, the lamina I giant
cells are densely innervated by axons that originate from nNOS-expressing inhibitory
interneurons (Ganley et al., 2015).

Synaptic Inputs to Interneurons

It is very likely that all dorsal horn interneurons receive direct synaptic input from
primary afferents, but relatively little is known about the precise organization of these
inputs. Excitatory interneurons of the vertical and radial types are innervated by both Aδ
and C afferents, including TRPV1-expressing nociceptors (Grudt & Perl, 2002; Uta et al.,
2010; Yasaka et al., 2007). Although much of the Aδ input to vertical cells is likely to
originate from nociceptors, some may be from A-LTMRs (Aδ D-hair afferents), since the
dendrites of these cells often extend ventrally into lamina III, and have been shown to
receive contacts from myelinated primary afferents in this region (Yasaka et al., 2014).
PKCγ-expressing neurons in laminae IIi and III are innervated by Aβ afferents, which are
presumably A-LTMRs (Lu et al., 2013). Among the inhibitory interneurons, islet cells have
been shown to receive monosynaptic input from C fibers, but not from myelinated
afferents.

Interestingly, recording studies show that only around 10% of randomly recorded pairs of
interneurons are synaptically linked, which suggests that synapses between different
types of interneuron are also arranged according to specific rules (Lu & Perl, 2005).
Much of what we know about the synaptic connections between interneurons is based on
a series of elegant studies by Perl, Lu, and colleagues (Lu et al., 2013; Lu & Perl, 2003,

Page 16 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

2005; Zheng, Lu, & Perl, 2010). They reported that islet cells were reciprocally connected
with another population of inhibitory interneurons that were identified by the presence of
green fluorescent protein (GFP) in mice in which GFP was expressed under control of the
Prion promoter (PrP). These PrP-GFP cells, which were subsequently shown to
correspond to the nNOS and galanin/dynorphin populations (Iwagaki, Garzillo, Polgár,
Riddell, & Todd, 2013), were presynaptic to vertical cells. Both islet cells (inhibitory) and
PKCγ neurons (excitatory) were shown to be presynaptic to a class of interneurons named
“transient central cells,” and these formed excitatory synapses on vertical cells. Among
the various synaptic connections described here, a serial circuit involving excitatory
synapses between Aβ afferents, PKCγ cells, transient central cells, vertical cells, and
lamina I projection neurons was identified (Figure 6). It was suggested that this pathway,
which was normally under powerful feed-forward inhibitory control, provided a route
through which A-LTMR input could activate nociceptive projection neurons, and thus
contributed to tactile allodynia (Lu et al., 2013). However, since vertical cells may also
receive direct input from A-LTMRs (see preceding discussion), it is likely that more than
one mechanism contributes to allodynia.

These studies have provided major insights into the synaptic circuitry of the dorsal horn.
However, certain caveats should be borne in mind. For example, the interneuron
populations described in these studies may not be completely homogeneous. In addition,
it is likely that both the synaptic inputs and outputs for each class of neuron are more
extensive than those that have so far been identified.

Synaptic Inputs to Primary Afferents

As we noted, primary afferent boutons can receive axoaxonic synapses, which mediate
GABAergic presynaptic inhibition. Although in most cases, the source of these synapses is
not known, Hughes et al. (2012) demonstrated that parvalbumin-expressing inhibitory
interneurons form axoaxonic synapses on the central terminals of presumed Aδ D-hair
afferents. This input was highly selective, because ~75% of the parvalbumin-
immunoreactive boutons in lamina IIi were involved in axoaxonic synapses. François et al.
(2017) showed that Penk1+ (enkephalinergic) and also GABAergic interneurons provide
presynaptic inhibitory inputs to primary afferents. However, the relationship between
parvalbumin+ and the Penk1+ interneurons is unclear (Francois et al., 2017). Zhang et
al. (2015) have suggested that primary afferents also receive direct descending
GABAergic/enkephalinergic inputs.

Neuropeptide Signaling in the Dorsal Horn

Fast synaptic transmission by glutamate, GABA, and glycine is likely to be responsible for
most of the somatosensory information processing in the dorsal horn. However, the

Page 17 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

presence of numerous peptides and their receptors indicates that neuropeptide signaling
is also likely to be important.

Substance P, which can be released from nociceptive primary afferents (and to a lesser
extent, from excitatory interneurons), will act on NK1 receptors that are mainly located
on ALT projection neurons, resulting in depolarization of these cells. This mechanism is
likely to contribute to pain perception, because mice lacking either the neuropeptide or
its receptor show reductions in pain behavior (Cao et al., 1998; De Felipe et al., 1998).
Somatostatin will also be released by both primary afferents and excitatory interneurons,
and will act on the somatostatin 2a (sst2A) receptor, which is expressed on some primary
afferent terminals and on many inhibitory interneurons, including the nNOS and galanin/
dynorphin populations (Todd, 2017). Somatostatin will cause hyperpolarization of these
interneurons, and it has been proposed that the resulting disinhibition of the nNOS and/
or dynorphin/galanin cells contributes to the itch evoked by intrathecal administration of
somatostatin or its analogues (Huang et al., 2018; Kardon et al., 2014). As described
here, GRP released by a population of excitatory interneurons is also thought to play an
important role in itch. Neurotensin and NKB are also expressed by excitatory
interneurons, but, although the corresponding receptors are present in the dorsal horn,
we know little about the functions of these peptides.

NPY and galanin are both expressed by inhibitory interneurons. Galanin is also present in
some nociceptive primary afferents, while NPY is upregulated in A-LTMRs following
peripheral nerve injury (Wakisaka, Kajander, & Bennett, 1991). Receptors for both
peptides are expressed by large numbers of excitatory dorsal horn neurons, and in each
case, release of the peptide is likely to suppress nociceptive transmission (Todd, 2017). At
least in the case of NPY, there is evidence that signaling may be more important in
pathological pain states (Solway, Bose, Corder, Donahue, & Taylor, 2011). The opioid
peptides dynorphin and enkephalin are found in both inhibitory and excitatory cells, and
all three classes of opioid receptor (μ, δ, κ) are expressed by primary afferents, with μ
and κ receptors also being present on some dorsal horn neurons. The δ and μ receptors
are thought to have complementary roles in anti-nociception (Scherrer et al., 2009), while
the κ receptor appears to have an anti-pruritic action (Huang et al., 2018; Kardon et al.,
2014).

Supraspinal Pain Pathways

Page 18 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Parallel Pain Processing pathways

As we mentioned, the second-order spinal ALT neurons or trigeminal medullary projection


neurons send information to multiple higher brain areas such that temperature, itch, and
pain sensation are processed by a large, distributed neural network. Although the exact
connectivity of this large network remains poorly understood, it is generally recognized
that there are two main pain processing streams: the spinothalamo-cortical pathway that
processes the discriminative aspects of pain, and the spinoparabrachial-limbic pathway
that processes the affective-motivational aspects of pain (summarized in Figure 7)
(Gauriau & Bernard, 2002; Hunt & Mantyh, 2001). However, it should be noted that such
a separation is an oversimplified view, since the medial and central thalamic nuclei that
receive inputs from spinothalamic neurons are connected to limbic cortical and
subcortical regions, and several cortical areas receive inputs from both the spinothalamic
and the spinoparabrachial pathways. Furthermore, the two processing streams interact
with each other through cortical–cortical and cortical–limbic connections. In addition,
second-order ALT neurons also have projections to various brainstem nuclei such as the
PAG and reticular formation (Figure 7). Next, we discuss the two main pathways in more
detail.

Click to view larger


Figure 7 Schematic diagram of neural pain
pathways.

ACC, anterior cingulate cortex; PFC, prefrontal


cortex; S1, primary somatosensory cortex; S2,
secondary somatosensory cortex; NAc, nucleus
accumbens; BNST, bed nucleus stria terminalis; PAG,
periaqueductal gray; RVM, rostroventral medulla;
TG, trigeminal ganglion; DRG, dorsal root ganglion.

The Spinothalamic-Cortical Pain Discriminative Pathway

Page 19 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Early extensive studies using anterograde and retrograde neuronal tracers in primates
showed that many axons of the ALT nociceptive projection neurons terminate in multiple
nuclei in thalamus, including the ventral posterior complex (with VPL receiving
projections from spinal dorsal horn, and VPM receiving inputs from trigeminal SpVc), the
central nucleus (both centrolateral or CL, and centromedial or CM), the medial dorsal
nucleus (MD), and posterior nuclear group (Po) (Boivie, 1979; Friedman & Murray, 1986;
Gingold, Greenspan, & Apkarian, 1991). Note that Po is not a well-defined region in
primates. In rodents, another area, called PoT (posterior triangular nucleus), which is
located in the caudalmost part of Po, should be added as one of the main targets for ALT
nociceptive axons (Al-Khater, Kerr, & Todd, 2008; Gauriau & Bernard, 2004). Neurons of
these thalamic nuclei in turn project to several cortical areas, including the primary
somatosensory (S1), the secondary somatosensory (S2), the insula, and the cingulate
cortex. The main axonal target of VPM and VPL neurons is S1. Neurons in Po project to
S1, S2, and the insular cortex. CL and CM send axons to S1 and the cingulate cortex,
whereas MD projects to both the insular and cingulate cortex (Burton & Jones, 1976;
Craig, 2014; Friedman & Murray, 1986; Gingold et al., 1991). Generally speaking, S1 is
critically involved in detecting the location, type, and intensity of pain. This is the main
reason that this spinothalamic-cortical pathway is regarded as the sensory-discrimination
pathway. On the other hand, S2 is thought to play a role in recognizing, learning, and
remembering painful events; the insular cortex seems to be important for the autonomous
and motivational responses during pain, and also for learning; while the cingulate cortex
is linked to perceiving the negative affect of pain, and overall integration of pain
perception and action selection (Schnitzler & Ploner, 2000). Thus, this spinothalamic
pathway also contributes to pain learning and pain affect.

Among the thalamic nuclei, VPM/VPL and Po nuclei have been most extensively
characterized. VPM/VPL and Po receive both innocuous tactile inputs (through the dorsal
column pathway) and nociceptive inputs (through the spinothalamic component of the
ALT). Studies in primates have found that neurons in VPM/VPL that responded to tactile
or painful stimuli are anatomically segregated into different clusters, and express
different calcium-binding proteins (Rausell, Bae, Vinuela, Huntley, & Jones, 1992; Rausell
& Jones, 1991). The same anatomical segregation of touch- and pain-processing neurons
may also exist in the primate Po nucleus. Furthermore, the tactile inputs are relayed to
middle layers in S1 by VPM/VPL/Po touch neurons, whereas the noxious inputs from the
spinothalamic tract are relayed to layer 1 in S1 by nociceptive VPM/VPL/Po neurons. In
rodents, VPM/VPL also contain neurons that respond either to innocuous tactile or to
noxious stimuli. The majority of neurons specifically activated in response to noxious
stimuli are found in the medial part of Po (PoM) and in PoT (Frangeul et al., 2014;
Gauriau & Bernard, 2004; Masri et al., 2009). Tactile-responsive rodent VPM/VPL axons
target Layer 4 of S1, whereas noci-responsive axons innervate Layer 1 and Layer 5a of S1
(Pouchelon et al., 2014). Thus, both in primates and in rodents, tactile and pain
information are processed by different layers in S1. Both PoM and PoT neurons also
project axons to multiple layers (including Layer 4) of S2 and insular cortex (Gauriau &

Page 20 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Bernard, 2004; Pouchelon et al., 2014). The central and medial thalamic nuclei (CL, CM,
MD) contain predominantly nociceptive neurons, and many of them receive bilateral
inputs from wide areas of the body (Dong, Ryu, & Wagman, 1978; Iwata et al., 2011).

Recognition of the role of S1 in processing the sensory discriminative aspect of pain


initially came from human studies. Patients with S1 surgically removed or injured had
problems in localizing pain. Patients with S1 lesions reported no sensation of pain but
intact perception of the negative affect and unpleasantness of pain (Ploner, Freund, &
Schnitzler, 1999). The primate S1 is subdivided into Brodmann’s areas 3a, 3b, 1, and 2.
Nociceptive neurons are found in areas 3a, 3b, and 1, and are interspersed among
neurons that respond solely to non-noxious touch stimuli. Area 2 appears to process
tactile and proprioceptive information. Nociceptive S1 neurons in 3b/1 mediate the “first
pain,” i.e., the well localized sensation evoked by activation of Aδ nociceptors.
Nociceptive neurons in area 3a mediate second pain evoked by slowly conducting C
fibers. It is believed that the interactions between 3a and 3b/1 collectively encode the
intensity, duration, and type of pain (Vierck, Whitsel, Favorov, Brown, & Tommerdahl,
2013).

The Spinoparabrachial-Limbic Pain Affective Pathway

The LPb is another major target of spinal ALT projection neurons. Electrophysiological
recordings revealed that noxious-responsive LPb neurons are driven by Aδ and C-
nociceptive inputs and generally have receptive fields covering a large area of the body.
Many LPb neurons respond to both cutaneous and visceral noxious stimuli (Gauriau &
Bernard, 2002). In other words, the spinoparabrachial pathway does not encode precise
locations of noxious stimuli. A subset of LPb neurons respond to warmth, cooling, or itch.

The nociceptive LPb neurons send ascending projections to the central nucleus of
amygdala (CeA), the lateral bed nucleus of stria terminalis (BNST), the midline
paraventricular nucleus of the thalamus (PVT), and the lateral and periventricular region
of hypothalamus (Hyp) (Rodriguez et al., 2017). CeA and BNST neurons are known to
project to the nucleus accumbens (NAc) and to the cingulate and prefrontal cortex. PVT
provides input to the cingulate and insular cortex. CeA, BNST, and the cingulate cortex
(especially the anterior cingulate cortex, ACC) are known to process affective emotional
responses. The lateral-capsular region of CeA is dubbed the “nociceptive amygdala,” as
this region receives purely nociceptive inputs from LPb (Neugebauer, 2015). Activation of
the axons from CGRP-expressing LPb neurons in CeA is sufficient to cause fear/threat
learning (Han, Soleiman, Soden, Zweifel, & Palmiter, 2015). In various preclinical pain
models, it has been found that there was enhanced excitatory transmission at LPb–CeA
synapses, along with increased background and evoked activity, and increased expression
of neuroactivity markers such as Fos and phosphorylated extracellular signal-regulated
kinase (pERK) in CeA (Thompson & Neugebauer, 2017). Interestingly, such neuroplastic
changes in inflammatory or neuropathic pain models were only observed in the right, but

Page 21 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

not the left, CeA (Thompson & Neugebauer, 2017), although the mechanisms underlying
such lateralization remain unclear.

The cingulate cortex receives inputs from amygdala (both the central amygdala as well as
the lateral and basolateral amygdala). Patients with cingulotomy reported that they could
still perceive pain sensations, but such sensation was less unpleasant, and they were less
bothered by the pain (Foltz & White, 1962). These observations have led to the notion
that the main function of this spinoparabrachial pathway is to process pain affect.
However, as mentioned previously, the midline thalamic nuclei also project to cingulate
cortex, and hence the spinothalamo-cortical pathway is also involved in the affective pain
component. The noci-responsive neurons in PVT, Hyp, and the insular are probably
involved in processing the autonomic and motivational aspects of pain.

Notably, trigeminal nociceptive sensory neurons innervating the face and head region
send axon collaterals that synapse directly with nociceptive LPb neurons (Figure 7), and
this is in addition to the canonical projection of these primary afferents to the medullary
SpVc nucleus. Specifically, stimulation of this trigemino-parabrachial monosynaptic
pathway drives robust aversive behavior and distress calls, whereas silencing this direct
pathway reduces craniofacial pain-related behaviors, but not those from elsewhere in the
body (Rodriguez et al., 2017). Thus, noxious stimuli experienced by the head and face
region engage dual pathways (directly or indirectly via SpVc) to activate the limbic
system, which may underlie the heightened affective response to face/head pain
compared to body pain.

LPb neurons also send descending projections to PAG, the rostral ventral medulla (RVM),
and reticular regions of the medulla. We will discuss the potential functions of these
projections later, with other descending pathways.

Descending Pathways
All pain-associated cortical areas mentioned here send descending axons that terminate
in multiple brainstem regions, including the PAG, the pons, the midline raphe nuclei, and
various medullary reticular nuclei. Corticospinal and corticobulbar neurons also
terminate in the spinal and trigeminal dorsal horn. These cortical descending projections
are collectively referred to as the corticofugal pathway. Accumulating evidence indicates
that descending cortical projections (from S1, cingulate and insular cortex) facilitate
sensory transmission. Pain-induced plasticity in these descending cortical neurons plays
important roles in facilitating pain hypersensitivity and/or maintaining chronic pain state
(Chen et al., 2014; Cichon, Blanck, Gan, & Yang, 2017; Kuner & Flor, 2016; Tan et al.,
2017), although much of the detailed circuitry and the underlying mechanisms remain to
be discovered.

Page 22 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

The CeA and hypothalamic nuclei send descending projections to LPb, PAG, and the locus
coeruleus (LC, where norepinephrinergic neurons reside). Neurons in LPb innervate PAG
and RVM, including the raphe nucleus, which contains serotonergic neurons (Krukoff,
Harris, & Jhamandas, 1993). PAG also provide inputs to RVM (Heinricher, Tavares, Leith,
& Lumb, 2009). The LC norepinephrinergic and the RVM serotonergic neurons project
diffusely to the dorsal horn, including its superficial part. Both monoamines are thought
to operate mainly through volume transmission, and therefore the distribution of
monoamine receptors is likely to be the main factor determining their targets. However,
we still know relatively little about the expression of these receptors by specific types of
dorsal horn neuron and primary afferent. In addition to descending monoaminergic
axons, there is also a GABAergic projection from RVM (Antal, Petko, Polgar, Heizmann, &
Storm-Mathisen, 1996). Some of the RVM GABAergic descending neurons co-express
Penk1, and they appear to play an inhibitory role in attenuating pain transmission;
whereas other RVM GABAergic descending neurons that do not express Penk1 seem to
facilitate pain transmission (Francois et al., 2017; Zhang et al., 2015). The pathway from
CeA via PAG to RVM pain-inhibitory neurons is considered the brain’s main descending
pain-suppressing system. The main descending pathways are shown with red arrows in
Figure 7.

References
Abraira, V. E., & Ginty, D. D. (2013). The sensory neurons of touch. Neuron, 79(4), 618–
639. doi:10.1016/j.neuron.2013.07.051

Abraira, V. E., Kuehn, E. D., Chirila, A. M., Springel, M. W., Toliver, A. A., Zimmerman, A.
L., … Ginty, D. D. (2017). The cellular and synaptic architecture of the mechanosensory
dorsal horn. Cell, 168(1–2), 295–310, e219. doi:10.1016/j.cell.2016.12.010

Akiyama, T., Curtis, E., Nguyen, T., Carstens, M. I., & Carstens, E. (2016). Anatomical
evidence of pruriceptive trigeminothalamic and trigeminoparabrachial projection neurons
in mice. Journal of Comparative Neurology, 524(2), 244–256. doi:10.1002/cne.23839

Al-Khater, K. M., Kerr, R., & Todd, A. J. (2008). A quantitative study of spinothalamic
neurons in laminae I, III, and IV in lumbar and cervical segments of the rat spinal cord.
Journal of Comparative Neurology, 511(1), 1–18.

Andrew, D. (2009). Sensitization of lamina I spinoparabrachial neurons parallels heat


hyperalgesia in the chronic constriction injury model of neuropathic pain. Journal of
Physiology, 587(Pt 9), 2005–2017.

Antal, M., Petko, M., Polgar, E., Heizmann, C. W., & Storm-Mathisen, J. (1996). Direct
evidence of an extensive GABAergic innervation of the spinal dorsal horn by fibers
descending from the rostral ventromedial medulla. Neuroscience, 73(2), 509–518. doi:
0306-4522(96)00063-2 [pii]

Page 23 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Baseer, N., Polgar, E., Watanabe, M., Furuta, T., Kaneko, T., & Todd, A. J. (2012).
Projection neurons in lamina III of the rat spinal cord are selectively innervated by local
dynorphin-containing excitatory neurons. Journal of Neuroscience, 32(34), 11854–11863.
doi:32/34/11854 [pii] 10.1523/JNEUROSCI.2707-12.2012

Bester, H., Chapman, V., Besson, J. M., & Bernard, J. F. (2000). Physiological properties of
the lamina I spinoparabrachial neurons in the rat. Journal of Neurophysiology, 83(4),
2239–2259.

Bice, T. N., & Beal, J. A. (1997). Quantitative and neurogenic analysis of the total
population and subpopulations of neurons defined by axon projection in the superficial
dorsal horn of the rat lumbar spinal cord. Journal of Comparative Neurology, 388(4), 550–
564. doi:10.1002/(SICI)1096-9861(19971201)388:4<550::AID-CNE4>3.0.CO;2-1 [pii]

Boada, M. D., & Woodbury, C. J. (2008). Myelinated skin sensory neurons project
extensively throughout adult mouse substantia gelatinosa. Journal of Neuroscience, 28(9),
2006–2014. doi:10.1523/JNEUROSCI.5609-07.2008

Boivie, J. (1979). An anatomical reinvestigation of the termination of the spinothalamic


tract in the monkey. Journal of Comparative Neurology, 186(3), 343–369. doi:10.1002/cne.
901860304

Boyle, K. A., Gutierrez-Mecinas, M., Polgar, E., Mooney, N., O’Connor, E., Furuta, T., …
Todd, A. J. (2017). A quantitative study of neurochemically defined populations of
inhibitory interneurons in the superficial dorsal horn of the mouse spinal cord.
Neuroscience, 363, 120–133. doi:10.1016/j.neuroscience.2017.08.044

Brown, A. G. (1981). Organization in the Spinal Cord: The Anatomy and Physiology of
Identified Neurons. Berlin: Springer-Verlag.

Brumovsky, P., Watanabe, M., & Hokfelt, T. (2007). Expression of the vesicular glutamate
transporters-1 and -2 in adult mouse dorsal root ganglia and spinal cord and their
regulation by nerve injury. Neuroscience, 147(2), 469–490. doi:10.1016/j.neuroscience.
2007.02.068

Burton, H., & Jones, E. G. (1976). The posterior thalamic region and its cortical projection
in New World and Old World monkeys. Journal of Comparative Neurology, 168(2), 249–
301. doi:10.1002/cne.901680204

Cameron, D., Polgar, E., Gutierrez-Mecinas, M., Gomez-Lima, M., Watanabe, M., & Todd,
A. J. (2015). The organization of spinoparabrachial neurons in the mouse. Pain, 156(10),
2061–2071. doi:10.1097/j.pain.0000000000000270

Campbell, J. N., Raja, S. N., Meyer, R. A., & Mackinnon, S. E. (1988). Myelinated afferents
signal the hyperalgesia associated with nerve injury. Pain, 32(1), 89–94.

Page 24 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Cao, Y. Q., Mantyh, P. W., Carlson, E. J., Gillespie, A. M., Epstein, C. J., & Basbaum, A. I.
(1998). Primary afferent tachykinins are required to experience moderate to intense pain.
Nature, 392(6674), 390–394. doi:10.1038/32897

Cavanaugh, D. J., Lee, H., Lo, L., Shields, S. D., Zylka, M. J., Basbaum, A. I., & Anderson,
D. J. (2009). Distinct subsets of unmyelinated primary sensory fibers mediate behavioral
responses to noxious thermal and mechanical stimuli. Proceedings of the National
Academy of Sciences USA, 106(22), 9075–9080. doi:0901507106 [pii] 10.1073/pnas.
0901507106

Chen, T., Koga, K., Descalzi, G., Qiu, S., Wang, J., Zhang, L. S., … Zhuo, M. (2014).
Postsynaptic potentiation of corticospinal projecting neurons in the anterior cingulate
cortex after nerve injury. Molecular Pain, 10, 33. doi:10.1186/1744-8069-10-33

Christensen, A. J., Iyer, S. M., Francois, A., Vyas, S., Ramakrishnan, C., Vesuna, S., …
Delp, S. L. (2016). In vivo interrogation of spinal mechanosensory circuits. Cell Reports,
17(6), 1699–1710. doi:10.1016/j.celrep.2016.10.010

Cichon, J., Blanck, T. J. J., Gan, W. B., & Yang, G. (2017). Activation of cortical somatostatin
interneurons prevents the development of neuropathic pain. Nature Neuroscience, 20(8),
1122–1132. doi:10.1038/nn.4595

Cordero-Erausquin, M., Allard, S., Dolique, T., Bachand, K., Ribeiro-da-Silva, A., & De
Koninck, Y. (2009). Dorsal horn neurons presynaptic to lamina I spinoparabrachial
neurons revealed by transynaptic labeling. Journal of Comparative Neurology, 517(5),
601–615.

Craig, A. D. (2014). Topographically organized projection to posterior insular cortex from


the posterior portion of the ventral medial nucleus in the long-tailed macaque monkey.
Journal of Comparative Neurology, 522(1), 36–63. doi:10.1002/cne.23425

Cui, L., Miao, X., Liang, L., Abdus-Saboor, I., Olson, W., Fleming, M. S., … Luo, W. (2016).
Identification of early RET+ deep dorsal spinal cord interneurons in gating pain.
[Published erratum]. Neuron, 91(6), 1413. doi:10.1016/j.neuron.2016.09.010

De Felipe, C., Herrero, J. F., O’Brien, J. A., Palmer, J. A., Doyle, C. A., Smith, A. J., … Hunt,
S. P. (1998). Altered nociception, analgesia and aggression in mice lacking the receptor
for substance P. Nature, 392(6674), 394–397. doi:10.1038/32904

Dong, W. K., Ryu, H., & Wagman, I. H. (1978). Nociceptive responses of neurons in medial
thalamus and their relationship to spinothalamic pathways. Journal of Neurophysiology,
41(6), 1592–1613. doi:10.1152/jn.1978.41.6.1592

Duan, B., Cheng, L., Bourane, S., Britz, O., Padilla, C., Garcia-Campmany, L., … Ma, Q.
(2014). Identification of spinal circuits transmitting and gating mechanical pain. Cell,
159(6), 1417–1432. doi:10.1016/j.cell.2014.11.003

Page 25 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Foltz, E. L., & White, L. E., Jr. (1962). Pain “relief” by frontal cingulumotomy. Journal of
Neurosurgery, 19, 89–100. doi:10.3171/jns.1962.19.2.0089

Foster, E., Wildner, H., Tudeau, L., Haueter, S., Ralvenius, W. T., Jegen, M., … Zeilhofer, H.
U. (2015). Targeted ablation, silencing, and activation establish glycinergic dorsal horn
neurons as key components of a spinal gate for pain and itch. Neuron, 85(6), 1289–1304.
doi:10.1016/j.neuron.2015.02.028

Francois, A., Low, S. A., Sypek, E. I., Christensen, A. J., Sotoudeh, C., Beier, K. T., …
Scherrer, G. (2017). A brainstem-spinal cord inhibitory circuit for mechanical pain
modulation by GABA and enkephalins. Neuron, 93(4), 822–839 e826. doi:10.1016/
j.neuron.2017.01.008

Frangeul, L., Porrero, C., Garcia-Amado, M., Maimone, B., Maniglier, M., Clasca, F., &
Jabaudon, D. (2014). Specific activation of the paralemniscal pathway during nociception.
European Journal of Neuroscience, 39(9), 1455–1464. doi:10.1111/ejn.12524

Friedman, D. P., & Murray, E. A. (1986). Thalamic connectivity of the second


somatosensory area and neighboring somatosensory fields of the lateral sulcus of the
macaque. Journal of Comparative Neurology, 252(3), 348–373. doi:10.1002/cne.
902520305

Ganley, R. P., Iwagaki, N., Del Rio, P., Baseer, N., Dickie, A. C., Boyle, K. A., … Todd, A. J.
(2015). Inhibitory interneurons that express GFP in the PrP-GFP mouse spinal cord are
morphologically heterogeneous, innervated by several classes of primary afferent and
include lamina I projection neurons among their postsynaptic targets. Journal of
Neuroscience, 35(19), 7626–7642. doi:10.1523/JNEUROSCI.0406-15.2015

Gauriau, C., & Bernard, J. F. (2002). Pain pathways and parabrachial circuits in the rat.
Experimental Physiology, 87(2), 251–258. doi:EPH_2357 [pii]

Gauriau, C., & Bernard, J. F. (2004). Posterior triangular thalamic neurons convey
nociceptive messages to the secondary somatosensory and insular cortices in the rat.
Journal of Neuroscience, 24(3), 752–761. doi:10.1523/JNEUROSCI.3272-03.2004 24/3/752
[pii]

Gingold, S. I., Greenspan, J. D., & Apkarian, A. V. (1991). Anatomic evidence of


nociceptive inputs to primary somatosensory cortex: Relationship between spinothalamic
terminals and thalamocortical cells in squirrel monkeys. Journal of Comparative
Neurology, 308(3), 467–490. doi:10.1002/cne.903080312

Grudt, T. J., & Perl, E. R. (2002). Correlations between neuronal morphology and
electrophysiological features in the rodent superficial dorsal horn. Journal of Physiology,
540(Pt 1), 189–207.

Page 26 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Gutierrez-Mecinas, M., Bell, A. M., Marin, A., Taylor, R., Boyle, K. A., Furuta, T., … Todd,
A. J. (2017). Preprotachykinin A is expressed by a distinct population of excitatory
neurons in the mouse superficial spinal dorsal horn including cells that respond to
noxious and pruritic stimuli. Pain, 158(3), 440–456. doi:10.1097/j.pain.
0000000000000778

Gutierrez-Mecinas, M., Furuta, T., Watanabe, M., & Todd, A. J. (2016). A quantitative
study of neurochemically defined excitatory interneuron populations in laminae I–III of
the mouse spinal cord. Molecular Pain, 12, 1744806916629065. doi:
10.1177/1744806916629065

Gutierrez-Mecinas, M., Polgár, E., Bell, A. M., Herau, M., & Todd, A. J. (2018). Substance
P-expressing excitatory interneurons in the mouse superficial dorsal horn provide a
propriospinal input to the lateral spinal nucleus. Brain Structure & Function, in press.
doi:10.1007/s00429-018-1629-x

Han, L., Ma, C., Liu, Q., Weng, H. J., Cui, Y., Tang, Z., … Dong, X. (2013). A subpopulation
of nociceptors specifically linked to itch. Nature Neuroscience, 16(2), 174–182. doi:
10.1038/nn.3289

Han, S., Soleiman, M. T., Soden, M. E., Zweifel, L. S., & Palmiter, R. D. (2015). Elucidating
an affective pain circuit that creates a threat memory. Cell, 162(2), 363–374. doi:10.1016/
j.cell.2015.05.057

Heinricher, M. M., Tavares, I., Leith, J. L., & Lumb, B. M. (2009). Descending control of
nociception: Specificity, recruitment and plasticity. Brain Research Reviews, 60(1), 214–
225. doi:S0165-0173(08)00147-1 [pii] 10.1016/j.brainresrev.2008.12.009

Huang, J., Polgár, E., Solinski, H. J., Mishra, S., Tseng, P.-Y., Iwagaki, N., … Hoon, M. A.
(2018). Dissecting the molecular and cellular pathway for somatostatin-induced itch.
Nature Neuroscience, 21, 707–716. doi:10.1038/s41593-018-0119-z.

Hughes, D. I., Sikander, S., Kinnon, C. M., Boyle, K. A., Watanabe, M., Callister, R. J., &
Graham, B. A. (2012). Morphological, neurochemical and electrophysiological features of
parvalbumin-expressing cells: A likely source of axo-axonic inputs in the mouse spinal
dorsal horn. Journal of Physiology, 590(Pt 16), 3927–3951. doi:jphysiol.2012.235655 [pii]
10.1113/jphysiol.2012.235655

Hunt, S. P., & Mantyh, P. W. (2001). The molecular dynamics of pain control. Nature
Reviews Neuroscience, 2(2), 83–91. doi:10.1038/35053509

Iwagaki, N., Garzillo, F., Polgár, E., Riddell, J. S., & Todd, A. J. (2013). Neurochemical
characterisation of lamina II inhibitory interneurons that express GFP in the PrP-GFP
mouse. Molecular Pain, 9(1), 56. doi:10.1186/1744-8069-9-56

Page 27 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Iwata, K., Miyachi, S., Imanishi, M., Tsuboi, Y., Kitagawa, J., Teramoto, K., … Takada, M.
(2011). Ascending multisynaptic pathways from the trigeminal ganglion to the anterior
cingulate cortex. Experimental Neurology, 227(1), 69–78. doi:10.1016/j.expneurol.
2010.09.013

Kardon, A. P., Polgár, E., Hachisuka, J., Snyder, L. M., Cameron, D., Savage, S., … Ross, S.
E. (2014). Dynorphin acts as a neuromodulator to inhibit itch in the dorsal horn of the
spinal cord. Neuron, 82(3), 573–586. doi:10.1016/j.neuron.2014.02.046

Krukoff, T. L., Harris, K. H., & Jhamandas, J. H. (1993). Efferent projections from the
parabrachial nucleus demonstrated with the anterograde tracer Phaseolus vulgaris
leucoagglutinin. Brain Research Bulletin, 30(1–2), 163–172.

Kuner, R., & Flor, H. (2016). Structural plasticity and reorganization in chronic pain.
Nature Reviews Neuroscience, 18(1), 20–30. doi:10.1038/nrn.2016.162

Larsson, M. (2017). Pax2 is persistently expressed by GABAergic neurons throughout the


adult rat dorsal horn. Neuroscience Letters, 638, 96–101. doi:10.1016/j.neulet.
2016.12.015

Li, C. L., Li, K. C., Wu, D., Chen, Y., Luo, H., Zhao, J. R., … Zhang, X. (2016).
Somatosensory neuron types identified by high-coverage single-cell RNA-sequencing and
functional heterogeneity. Cell Research, 26(1), 83–102. doi:10.1038/cr.2015.149

Light, A. R., & Perl, E. R. (1979). Spinal termination of functionally identified primary
afferent neurons with slowly conducting myelinated fibers. Journal of Comparative
Neurology, 186(2), 133–150.

Lu, Y., Dong, H., Gao, Y., Gong, Y., Ren, Y., Gu, N., … Xiong, L. (2013). A feed-forward
spinal cord glycinergic neural circuit gates mechanical allodynia. Journal of Clinical
Investigation, 123(9), 4050–4062. doi:10.1172/JCI70026

Lu, Y., & Perl, E. R. (2003). A specific inhibitory pathway between substantia gelatinosa
neurons receiving direct C-fiber input. Journal of Neuroscience, 23(25), 8752–8758.

Lu, Y., & Perl, E. R. (2005). Modular organization of excitatory circuits between neurons
of the spinal superficial dorsal horn (laminae I and II). Journal of Neuroscience, 25(15),
3900–3907.

Malmberg, A. B., Chen, C., Tonegawa, S., & Basbaum, A. I. (1997). Preserved acute pain
and reduced neuropathic pain in mice lacking PKCgamma. Science, 278(5336), 279–283.

Mantyh, P. W., DeMaster, E., Malhotra, A., Ghilardi, J. R., Rogers, S. D., Mantyh, C. R., …
et al. (1995). Receptor endocytosis and dendrite reshaping in spinal neurons after
somatosensory stimulation. Science, 268(5217), 1629–1632.

Page 28 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Masri, R., Quiton, R. L., Lucas, J. M., Murray, P. D., Thompson, S. M., & Keller, A. (2009).
Zona incerta: A role in central pain. Journal of Neurophysiology, 102(1), 181–191. doi:
10.1152/jn.00152.2009

McMahon, S. B., & Wall, P. D. (1983). A system of rat spinal cord lamina 1 cells projecting
through the contralateral dorsolateral funiculus. Journal of Comparative Neurology,
214(2), 217–223. doi:10.1002/cne.902140209

McMahon, S. B., & Wall, P. D. (1985). Electrophysiological mapping of brainstem


projections of spinal cord lamina I cells in the rat. Brain Research, 333(1), 19–26.

Melzack, R., & Wall, P. D. (1965). Pain mechanisms: A new theory. Science, 150(699), 971–
979.

Moser, H. R., & Giesler, G. J., Jr. (2014). Characterization of pruriceptive


trigeminothalamic tract neurons in rats. Journal of Neurophysiology, 111(8), 1574–1589.
doi:10.1152/jn.00668.2013

Neugebauer, V. (2015). Amygdala pain mechanisms. Handbook of Experimental


Pharmacology, 227, 261–284. doi:10.1007/978-3-662-46450-2_13

Nichols, M. L., Allen, B. J., Rogers, S. D., Ghilardi, J. R., Honore, P., Luger, N. M., …
Mantyh, P. W. (1999). Transmission of chronic nociception by spinal neurons expressing
the substance P receptor. Science, 286(5444), 1558–1561.

Peirs, C., Williams, S. P., Zhao, X., Walsh, C. E., Gedeon, J. Y., Cagle, N. E., … Seal, R. P.
(2015). Dorsal horn circuits for persistent mechanical pain. Neuron, 87(4), 797–812. doi:
10.1016/j.neuron.2015.07.029

Petitjean, H., Pawlowski, S. A., Fraine, S. L., Sharif, B., Hamad, D., Fatima, T., … Sharif-
Naeini, R. (2015). Dorsal horn parvalbumin neurons are gate-keepers of touch-evoked
pain after nerve injury. Cell Reports, 13(6), 1246–1257. doi:10.1016/j.celrep.2015.09.080

Ploner, M., Freund, H. J., & Schnitzler, A. (1999). Pain affect without pain sensation in a
patient with a postcentral lesion. Pain, 81(1–2), 211–214.

Polgár, E., Al-Khater, K. M., Shehab, S., Watanabe, M., & Todd, A. J. (2008). Large
projection neurons in lamina I of the rat spinal cord that lack the neurokinin 1 receptor
are densely innervated by VGLUT2-containing axons and possess GluR4-containing AMPA
receptors. Journal of Neuroscience, 28(49), 13150–13160.

Polgár, E., Al Ghamdi, K. S., & Todd, A. J. (2010). Two populations of neurokinin 1
receptor-expressing projection neurons in lamina I of the rat spinal cord that differ in
AMPA receptor subunit composition and density of excitatory synaptic input.
Neuroscience, 167(4), 1192–1204. doi:S0306-4522(10)00379-9 [pii] 10.1016/
j.neuroscience.2010.03.028

Page 29 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Polgár, E., Campbell, A. D., MacIntyre, L. M., Watanabe, M., & Todd, A. J. (2007).
Phosphorylation of ERK in neurokinin 1 receptor-expressing neurons in laminae III and IV
of the rat spinal dorsal horn following noxious stimulation. Molecular Pain, 3, 4. doi:
1744-8069-3-4 [pii] 10.1186/1744-8069-3-4

Polgár, E., Durrieux, C., Hughes, D. I., & Todd, A. J. (2013). A quantitative study of
inhibitory interneurons in laminae I–III of the mouse spinal dorsal horn. PLoS One, 8(10),
e78309. doi:10.1371/journal.pone.0078309

Polgár, E., Sardella, T. C., Tiong, S. Y., Locke, S., Watanabe, M., & Todd, A. J. (2013).
Functional differences between neurochemically defined populations of inhibitory
interneurons in the rat spinal dorsal horn. Pain, 154(12), 2606–2615. doi:10.1016/j.pain.
2013.05.001

Polgár, E., Sardella, T. C., Watanabe, M., & Todd, A. J. (2011). Quantitative study of NPY-
expressing GABAergic neurons and axons in rat spinal dorsal horn. Journal of
Comparative Neurology, 519(6), 1007–1023. doi:10.1002/cne.22570

Polgár, E., Wright, L. L., & Todd, A. J. (2010). A quantitative study of brainstem
projections from lamina I neurons in the cervical and lumbar enlargement of the rat.
Brain Research, 1308, 58–67.

Pouchelon, G., Gambino, F., Bellone, C., Telley, L., Vitali, I., Luscher, C., … Jabaudon, D.
(2014). Modality-specific thalamocortical inputs instruct the identity of postsynaptic L4
neurons. Nature, 511(7510), 471–474. doi:10.1038/nature13390

Rausell, E., Bae, C. S., Vinuela, A., Huntley, G. W., & Jones, E. G. (1992). Calbindin and
parvalbumin cells in monkey VPL thalamic nucleus: Distribution, laminar cortical
projections, and relations to spinothalamic terminations. Journal of Neuroscience, 12(10),
4088–4111.

Rausell, E., & Jones, E. G. (1991). Chemically distinct compartments of the thalamic VPM
nucleus in monkeys relay principal and spinal trigeminal pathways to different layers of
the somatosensory cortex. Journal of Neuroscience, 11(1), 226–237.

Rexed, B. (1952). The cytoarchitectonic organization of the spinal cord in the cat. Journal
of Comparative Neurology, 96(3), 414–495.

Ribeiro-da-Silva, A., & Coimbra, A. (1982). Two types of synaptic glomeruli and their
distribution in laminae I–III of the rat spinal cord. Journal of Comparative Neurology,
209(2), 176–186. doi:10.1002/cne.902090205

Ribeiro-da-Silva, A., Tagari, P., & Cuello, A. C. (1989). Morphological characterization of


substance P-like immunoreactive glomeruli in the superficial dorsal horn of the rat spinal
cord and trigeminal subnucleus caudalis: A quantitative study. Journal of Comparative
Neurology, 281(4), 497–515. doi:10.1002/cne.902810402

Page 30 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Rodriguez, E., Sakurai, K., Xu, J., Chen, Y., Toda, K., Zhao, S., … Wang, F. (2017). A
craniofacial-specific monosynaptic circuit enables heightened affective pain. Nature
Neuroscience, 20(12), 1734–1743. doi:10.1038/s41593-017-0012-1

Ross, S. E., Mardinly, A. R., McCord, A. E., Zurawski, J., Cohen, S., Jung, C., … Greenberg,
M. E. (2010). Loss of inhibitory interneurons in the dorsal spinal cord and elevated itch in
Bhlhb5 mutant mice. Neuron, 65(6), 886–898. doi:S0896-6273(10)00142-X [pii] 10.1016/
j.neuron.2010.02.025

Saito, H., Katagiri, A., Okada, S., Mikuzuki, L., Kubo, A., Suzuki, T., … Iwata, K. (2017).
Ascending projections of nociceptive neurons from trigeminal subnucleus caudalis: A
population approach. Experimental Neurology, 293, 124–136. doi:10.1016/j.expneurol.
2017.03.024

Salio, C., Ferrini, F., Muthuraju, S., & Merighi, A. (2014). Presynaptic modulation of spinal
nociceptive transmission by glial cell line-derived neurotrophic factor (GDNF). Journal of
Neuroscience, 34(41), 13819–13833. doi:10.1523/JNEUROSCI.0808-14.2014

Sandkuhler, J. (2009). Models and mechanisms of hyperalgesia and allodynia.


Physiological Reviews, 89(2), 707–758. doi:89/2/707 [pii] 10.1152/physrev.00025.2008

Scherrer, G., Imamachi, N., Cao, Y. Q., Contet, C., Mennicken, F., O’Donnell, D., …
Basbaum, A. I. (2009). Dissociation of the opioid receptor mechanisms that control
mechanical and heat pain. Cell, 137(6), 1148–1159. doi:10.1016/j.cell.2009.04.019

Schnitzler, A., & Ploner, M. (2000). Neurophysiology and functional neuroanatomy of pain
perception. Journal of Clinical Neurophysiology, 17(6), 592–603.

Seal, R. P., Wang, X., Guan, Y., Raja, S. N., Woodbury, C. J., Basbaum, A. I., & Edwards, R.
H. (2009). Injury-induced mechanical hypersensitivity requires C-low threshold
mechanoreceptors. Nature, 462(7273), 651–655.

Sivilotti, L., & Woolf, C. J. (1994). The contribution of GABAA and glycine receptors to
central sensitization: Disinhibition and touch-evoked allodynia in the spinal cord. Journal
of Neurophysiology, 72(1), 169–179.

Smith, K. M., Boyle, K. A., Madden, J. F., Dickinson, S. A., Jobling, P., Callister, R. J., …
Graham, B. A. (2015). Functional heterogeneity of calretinin-expressing neurons in the
mouse superficial dorsal horn: Implications for spinal pain processing. Journal of
Physiology, 593(19), 4319–4339. doi:10.1113/JP270855

Solway, B., Bose, S. C., Corder, G., Donahue, R. R., & Taylor, B. K. (2011). Tonic inhibition
of chronic pain by neuropeptide Y. Proceedings of the National Academy of Sciences USA,
108(17), 7224–7229. doi:10.1073/pnas.1017719108

Page 31 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Spike, R. C., Puskar, Z., Andrew, D., & Todd, A. J. (2003). A quantitative and morphological
study of projection neurons in lamina I of the rat lumbar spinal cord. European Journal of
Neuroscience, 18(9), 2433–2448.

Sugiura, Y., Lee, C. L., & Perl, E. R. (1986). Central projections of identified, unmyelinated
(C) afferent fibers innervating mammalian skin. Science, 234(4774), 358–361.

Sun, Y. G., & Chen, Z. F. (2007). A gastrin-releasing peptide receptor mediates the itch
sensation in the spinal cord. Nature, 448(7154), 700–703. doi:nature06029 [pii] 10.1038/
nature06029

Sun, Y. G., Zhao, Z. Q., Meng, X. L., Yin, J., Liu, X. Y., & Chen, Z. F. (2009). Cellular basis
of itch sensation. Science, 325(5947), 1531–1534. doi:1174868 [pii] 10.1126/science.
1174868

Tan, L. L., Pelzer, P., Heinl, C., Tang, W., Gangadharan, V., Flor, H., … Kuner, R. (2017). A
pathway from midcingulate cortex to posterior insula gates nociceptive hypersensitivity.
Nature Neuroscience, 20(11), 1591–1601. doi:10.1038/nn.4645

Thompson, J. M., & Neugebauer, V. (2017). Amygdala plasticity and pain. Pain Research &
Management, 2017, 8296501. doi:10.1155/2017/8296501

Todd, A. J. (1996). GABA and glycine in synaptic glomeruli of the rat spinal dorsal horn.
European Journal of Neuroscience, 8(12), 2492–2498.

Todd, A. J. (2010). Neuronal circuitry for pain processing in the dorsal horn. Nature
Reviews Neuroscience, 11(12), 823–836. doi:nrn2947 [pii] 10.1038/nrn2947

Todd, A. J. (2017). Identifying functional populations among the interneurons in laminae


I–III of the spinal dorsal horn. Molecular Pain, 13, 1744806917693003. doi:
10.1177/1744806917693003

Todd, A. J., Hughes, D. I., Polgár, E., Nagy, G. G., Mackie, M., Ottersen, O. P., & Maxwell,
D. J. (2003). The expression of vesicular glutamate transporters VGLUT1 and VGLUT2 in
neurochemically defined axonal populations in the rat spinal cord with emphasis on the
dorsal horn. European Journal of Neuroscience, 17(1), 13–27.

Todd, A. J., & Koerber, H. R. (2013). Neuroanatomical substrates of spinal nociception. In


S. McMahon, M. Koltzenburg, I. Tracey, & D. Turk (Eds.), Wall and Melzack’s Textbook of
Pain (6th ed., pp. 77–93). Edinburgh, UK: Elsevier.

Todd, A. J., Puskar, Z., Spike, R. C., Hughes, C., Watt, C., & Forrest, L. (2002). Projection
neurons in lamina I of rat spinal cord with the neurokinin 1 receptor are selectively
innervated by substance P-containing afferents and respond to noxious stimulation.
Journal of Neuroscience, 22(10), 4103–4113.

Page 32 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Todd, A. J., & Spike, R. C. (1993). The localization of classical transmitters and
neuropeptides within neurons in laminae I–III of the mammalian spinal dorsal horn.
Progress in Neurobiology, 41(5), 609–645.

Usoskin, D., Furlan, A., Islam, S., Abdo, H., Lonnerberg, P., Lou, D., … Ernfors, P. (2015).
Unbiased classification of sensory neuron types by large-scale single-cell RNA
sequencing. Nature Neuroscience, 18(1), 145–153. doi:10.1038/nn.3881

Uta, D., Furue, H., Pickering, A. E., Rashid, M. H., Mizuguchi-Takase, H., Katafuchi, T., …
Yoshimura, M. (2010). TRPA1-expressing primary afferents synapse with a
morphologically identified subclass of substantia gelatinosa neurons in the adult rat
spinal cord. European Journal of Neuroscience, 31(11), 1960–1973. doi:EJN7255 [pii]
10.1111/j.1460-9568.2010.07255.x

Vierck, C. J., Whitsel, B. L., Favorov, O. V., Brown, A. W., & Tommerdahl, M. (2013). Role of
primary somatosensory cortex in the coding of pain. Pain, 154(3), 334–344. doi:10.1016/
j.pain.2012.10.021

Wakisaka, S., Kajander, K. C., & Bennett, G. J. (1991). Increased neuropeptide Y (NPY)-
like immunoreactivity in rat sensory neurons following peripheral axotomy. Neuroscience
Letters, 124(2), 200–203.

Woodbury, C. J., & Koerber, H. R. (2003). Widespread projections from myelinated


nociceptors throughout the substantia gelatinosa provide novel insights into neonatal
hypersensitivity. Journal of Neuroscience, 23(2), 601–610.

Yaksh, T. L. (1989). Behavioral and autonomic correlates of the tactile evoked allodynia
produced by spinal glycine inhibition: Effects of modulatory receptor systems and
excitatory amino acid antagonists. Pain, 37(1), 111–123.

Yasaka, T., Kato, G., Furue, H., Rashid, M. H., Sonohata, M., Tamae, A., … Yoshimura, M.
(2007). Cell-type-specific excitatory and inhibitory circuits involving primary afferents in
the substantia gelatinosa of the rat spinal dorsal horn in vitro. Journal of Physiology,
581(Pt 2), 603–618.

Yasaka, T., Tiong, S. Y., Polgar, E., Watanabe, M., Kumamoto, E., Riddell, J. S., & Todd, A. J.
(2014). A putative relay circuit providing low-threshold mechanoreceptive input to lamina
I projection neurons via vertical cells in lamina II of the rat dorsal horn. Molecular Pain,
10, 3. doi:10.1186/1744-8069-10-3

Yasaka, T., Tiong, S. Y. X., Hughes, D. I., Riddell, J. S., & Todd, A. J. (2010). Populations of
inhibitory and excitatory interneurons in lamina II of the adult rat spinal dorsal horn
revealed by a combined electrophysiological and anatomical approach. Pain, 151, 475–
488.

Page 33 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018


Central Nervous System Pain Pathways

Zeilhofer, H. U., Wildner, H., & Yevenes, G. E. (2012). Fast synaptic inhibition in spinal
sensory processing and pain control. Physiological Reviews, 92(1), 193–235. doi:92/1/193
[pii] 10.1152/physrev.00043.2010

Zhang, E. T., & Craig, A. D. (1997). Morphology and distribution of spinothalamic lamina I
neurons in the monkey. Journal of Neuroscience, 17(9), 3274–3284.

Zhang, E. T., Han, Z. S., & Craig, A. D. (1996). Morphological classes of spinothalamic
lamina I neurons in the cat. Journal of Comparative Neurology, 367(4), 537–549.

Zhang, Y., Zhao, S., Rodriguez, E., Takatoh, J., Han, B. X., Zhou, X., & Wang, F. (2015).
Identifying local and descending inputs for primary sensory neurons. Journal of Clinical
Investigation, 125(10), 3782–3794. doi:10.1172/JCI81156

Zheng, J., Lu, Y., & Perl, E. R. (2010). Inhibitory neurones of the spinal substantia
gelatinosa mediate interaction of signals from primary afferents. Journal of Physiology,
588(Pt 12), 2065–2075. doi:jphysiol.2010.188052 [pii] 10.1113/jphysiol.2010.188052

Zylka, M. J., Rice, F. L., & Anderson, D. J. (2005). Topographically distinct epidermal
nociceptive circuits revealed by axonal tracers targeted to Mrgprd. Neuron, 45(1), 17–25.
doi:S0896627304008037 [pii] 10.1016/j.neuron.2004.12.015

Andrew J. Todd

Andrew J. Todd, Institute of Neuroscience and Psychology, College of Medical,


Veterinary and Life Sciences, University of Glasgow, Glasgow, UK

Fan Wang

Fan Wang, Department of Neurobiology, Duke University School of Medicine,


Durham, NC, United States

Page 34 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). © Oxford University Press, 2018. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy and Legal Notice).

date: 18 July 2018

You might also like