Next Article in Journal
Enhanced Clearance of Neurotoxic Misfolded Proteins by the Natural Compound Berberine and Its Derivatives
Previous Article in Journal
Unique Gut Microbiome Signatures Depict Diet-Versus Genetically Induced Obesity in Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Medial Prefrontal Cortex as a Central Hub for Mental Comorbidities Associated with Chronic Pain

Institute of Physiology, Medical University of Innsbruck, 6020 Innsbruck, Austria
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(10), 3440; https://doi.org/10.3390/ijms21103440
Submission received: 15 April 2020 / Revised: 6 May 2020 / Accepted: 7 May 2020 / Published: 13 May 2020
(This article belongs to the Section Molecular Neurobiology)

Abstract

:
Chronic pain patients frequently develop and suffer from mental comorbidities such as depressive mood, impaired cognition, and other significant constraints of daily life, which can only insufficiently be overcome by medication. The emotional and cognitive components of pain are processed by the medial prefrontal cortex, which comprises the anterior cingulate cortex, the prelimbic, and the infralimbic cortex. All three subregions are significantly affected by chronic pain: magnetic resonance imaging has revealed gray matter loss in all these areas in chronic pain conditions. While the anterior cingulate cortex appears hyperactive, prelimbic, and infralimbic regions show reduced activity. The medial prefrontal cortex receives ascending, nociceptive input, but also exerts important top-down control of pain sensation: its projections are the main cortical input of the periaqueductal gray, which is part of the descending inhibitory pain control system at the spinal level. A multitude of neurotransmitter systems contributes to the fine-tuning of the local circuitry, of which cholinergic and GABAergic signaling are particularly emerging as relevant components of affective pain processing within the prefrontal cortex. Accordingly, factors such as distraction, positive mood, and anticipation of pain relief such as placebo can ameliorate pain by affecting mPFC function, making this cortical area a promising target region for medical as well as psychosocial interventions for pain therapy.

1. Introduction

Functional magnetic resonance imaging (fMRI) studies in both humans and rodents have shown that there is not a single brain structure responsible for pain perception, but that different cortical, subcortical, and associative brain regions are activated during a painful stimulus [1]. Due to its complexity and interconnectedness, some researchers refer to this network of regions as the “pain matrix” [2]. Most commonly activated brain regions in response to pain are the thalamus, the somatosensory cortices, the insular and the prefrontal including the anterior cingulate cortex (ACC), the lentiform nucleus, cerebellum, amygdala, and the nucleus accumbens [1,3,4,5,6]. The thalamus relays nociceptive information to the insula, ACC and somatosensory cortices to process sensory-discriminative properties, including the location, quality, and intensity of a pain stimulus, while other brain regions such as the prefrontal cortex, insular cortex, limbic system, cerebellum or periaqueductal gray (PAG) play a role in affective components of the pain sensation and the formation of pain memory [7,8,9,10,11,12,13,14].
Its predominant role in aversive learning and memory puts the medial prefrontal cortex (mPFC) in a pole position as a hub for the development of mental comorbidities associated with chronic pain. Neurons within the mPFC respond to noxious stimuli [15] and electrical stimulation of mPFC inhibits nociceptive responses [16]. Human brain fMRI studies show activation of mPFC during the perception of acute pain stimuli and functional deactivation in chronic pain patients (for review see [17]). Affective and cognitive components of pain sensations are processed by the mPFC [6,18]. Furthermore, the basis of placebo induced analgesia or expectancy of pain relief is based on the ACC-PFC-PAG descending pain pathway, which influences endogenous opioid activity (reviewed by [6]). Resting-state network studies in chronic pain patients report disrupted network properties within the interacting brain regions, the default mode network (DMN). This also includes reduced connectivity between the mPFC and the posterior regions of the DMN, as well as increased connectivity between the mPFC and the insular cortex, which is correlated to pain intensity [19]. In rodents, the mPFC is primarily known for its prominent role in attention, working memory, and goal-directed behavior [20]. It provides top-down regulation of sensory and affective processes, including inhibition of both sensory and affective pain signals by descending projections to the various brain and spinal cord regions: its projections are the main cortical input of the PAG, which is part of the descending inhibitory pain control system at the spinal level [21,22,23]. In both human subjects and rodent models, the mPFC undergoes structural as well as functional changes in chronic pain states [24,25,26,27,28,29,30,31], which is reflected by cognitive deficits and decreased attention (for review see [32]). This review aims to dissect the main local circuitry, including transmitter systems, together with input and output connectivity of mPFC subregions and their importance for affective pain processing and the mental comorbidities frequently associated with chronic pain disorders.

2. Input Systems to mPFC

Neuroanatomical tracing studies reveal distinct mPFC subregions, however, nomenclature and conventions on how to apply them differ between stereotaxic atlases [33,34,35]. Here, we address the prelimbic cortex (PrL), the infralimbic cortex (IL), and the anterior cingulate cortex (ACC) as mPFC subregions, which correspond to the human pregenual and subgenual ACC (Brodmann Areas 32 and 25), and the midcingulate cortex (Brodmann Area 24), respectively [35]. Incoming signals are received by the ACC indirectly via the medial thalamus as well as the amygdala, primary somatosensory cortex, and insular cortices [36,37,38,39,40,41]. Projections from the agranular insular cortex, basal forebrain nuclei, basal nuclei of the amygdala, midline thalamic nuclei (including the paraventricular nucleus), the hippocampus CA1 region and subiculum, ventral tegmental area, PAG, dorsal raphe nuclei, and locus coeruleus target both PrL and IL [42]. Virtually no input to the two regions is documented from primary motor and somatosensory cortices, as well as the dorsal striatum and nucleus accumbens. However, while the PrL receives inputs mainly from anterior forebrain regions, the IL is preferentially targeted by projections from anterior hypothalamus, medial and lateral septum, substantia innominata, as well as the dorsolateral tegmentum [42]. Cholinergic input to the mPFC arises from the basal forebrain (BF) and is associated with attentional tasks and cognitive performance [20]. The vertical limb (VDB) and the hind limb (HDB) of the diagonal band of Broca (DB) are generally accepted as cholinergic input regions to the mPFC [43,44,45,46,47]. Recent studies indicate cholinergic projections from the medial septum (MS) to the mPFC which might contribute to the cognitive and emotional impairments induced by pain [48,49,50]. MS inactivation concurrently reduces the activation of the prefrontal cortex and inhibition of cholinergic MS neurons projecting to the ACC alleviates pain-induced anxiety [50,51,52].

3. Neuron Types and Circuitry in mPFC Cortical Layers

Within the mPFC, five cellular layers can be distinguished [53]: excitatory pyramidal neurons in layers 2/3 (L2/3) and L5 make up ~80% of neurons and (mainly inhibitory) interneurons the remaining ~20% [54]. The mPFC receives inputs and sends reciprocal projections back to almost all areas of the sensory cortex, various motor structures, and subcortical regions, including the hippocampus and amygdala [55]. The majority of inputs are received in superficial L2/3, whereas outputs from layers L2/3 are shared between the prefrontal subregions and also terminate in deeper layers L5 and L6, from where efferent projections are sent to various subcortical regions [54,56].

3.1. Pyramidal Neurons

Pyramidal neurons differ between superficial and deep layers regarding morphological features, molecular markers as well as responsiveness to sensory stimuli; they receive unique synaptic inputs but also possess unique action potential (AP) firing patterns and physiological phenotypes [57,58,59,60,61]. In superficial layers L1 and L2/3, pyramidal neurons and interneurons can reciprocally connect in every possible way [62]. Different layers are tightly coupled to specific functions and differently affected in mental disorders, including schizophrenia, bipolar disorder, stress-induced depression, and substance abuse [63,64,65,66,67]. Pyramidal neurons in L2/3 are nociceptive and show regular spiking, intermediate firing, or intrinsic bursting [68,69]. Excitatory pyramidal neurons in layers L2/3 and L5 respond to noxious mechanical and thermal stimuli with an increase in firing rates and are related to the unpleasantness induced by acute pain [69,70,71,72,73,74]. Pain conditions including peripheral nerve injury induce long-lasting synaptic plasticity by different variations of long-term potentiation (LTP) or long-term depression (LTD) to increase excitatory synaptic transmission resulting in hyperexcitability and disinhibition of the ACC, whereas inhibitory transmission is reduced [68,75,76,77,78,79,80].

3.2. Inhibitory Interneurons

Inhibitory GABAergic interneurons make up 10–20% of all mPFC neurons and are found in all layers [81]. GABAergic interneurons are targeted by thalamocortical projections [82], and as they are divided based on neuropeptide markers together with their layer-specific axonal projections (L1, L2/3, L5, and white matter-projecting interneurons [83]), they may play differential roles in maintaining the excitation/inhibition balance in different mPFC layers and underlie previously described laminar differences in pain processing [84].

3.2.1. Parvalbumin Interneurons

Fast-spiking parvalbumin (FS-PV+) interneurons uniformly show increased and sustained firing during goal-driven attentional processing, correlating to the level of attention. Successful allocation of attention is characterized by strong synchronization of FS-PV+ neurons, increased gamma oscillations, and phase locking of pyramidal firing. FS-PV+ neurons thus act as a functional unit coordinating the activity in the local mPFC circuit during goal-driven attentional processing [85]. PV+ interneurons are critically important for fear conditioning and the reward system, and successful memory consolidation also requires coherent hippocampal-neocortical communication mediated by PV+ cells [86,87]. Optogenetic activation of PV+ interneurons in neuropathic mice causes an increase in mechanical hypersensitivity, suggesting that these cells may be directly involved in sensory processing of chronic pain [88]. Bilateral morphological changes in PV+ interneurons and layer L5/6 IL pyramidal neurons are correlated with neuropathic pain-related cognitive impairments [89] (Figure 1).

3.2.2. CCK Interneurons

Cholecystokinin (CCK) interneurons comprise a larger proportion of the mPFC interneurons compared to PV+ interneurons, targeting a wide range of neuronal subtypes with a distinct connectivity pattern and CCK+, but not PV+ interneurons play a critical role in the retrieval of working memory [91]. Although their functional importance is less understood, they are associated with emotional and cognitive symptoms of major depressive disorder [92].

3.2.3. Somatostatin Interneurons

Somatostatin interneurons facilitate hippocampal-prefrontal synchrony and prefrontal spatial encoding and contribute to cognitive performance. In the prefrontal cortex, they control affective state discrimination [93,94].

3.2.4. VIP Interneurons

GABAergic vasoactive intestinal peptide (VIP) neurons inhibit PV+ and somatostatin-expressing interneurons, thereby disinhibiting pyramidal neurons [95,96,97].

3.3. Excitatory Interneurons

Neocortical choline acetyltransferase (ChAT)-expressing interneurons are a subclass of VIP+ interneurons of which circuit and behavioral functions are largely unknown. ChAT-VIP+ neurons directly excite neighboring neurons in several layers through the fast synaptic transmission of acetylcholine (ACh) in mPFC and control attention behavior [98].

3.4. Modulation by Glial Cells

The modulation of electrical and synaptic neuronal activity by glial cells in pain states is well accepted, whereas the role of microglia and astrocytes in the mPFC is only starting to gain interest [99,100,101]. The majority but not all studies support an upregulation of microglia or microglial markers, that might at least in part be induced by interleukin 1 beta (IL-1β) [102,103,104,105,106,107,108,109,110,111,112]. Inhibition of microglial activation in neuropathic pain models alleviates mechanical allodynia in nerve ligated mice and affective components of pain-like behavior [103,107]. Interestingly, under healthy conditions, the nociceptor transducer channel TRPV1 is specifically expressed in microglia in the ACC, where it controls microglial activation and indirectly enhances glutamatergic transmission in neurons [113]. In addition, astrocyte marker expression increases in inflammatory and neuropathic pain models at least partially as a consequence of enhanced glutamate levels in the ACC [107,114,115,116,117,118,119,120]. Astrocyte numbers appear to be relatively constant, however, complex changes in glutamate transporter GLT1 and ionotropic as well as metabotropic glutamate receptor subunits take place in astrocytes upon peripheral painful events such as formalin injection or paclitaxel neuropathy [120,121]. The selective ablation of astrocytes inhibits long-term facilitation associated with inflammatory pain and relieves escape avoidance behavior, which stresses their functional importance for pain processing in the ACC [117,118].

4. Local Circuits and Differences Between Cortical Layers

4.1. Layer L1

Like in other neocortical areas, layer L1 contains only a few neurons, which are almost exclusively GABAergic interneurons, predominantly late-spiking, continuously adapting and inhibiting pyramidal neurons [57,122,123,124,125]. L1 interneurons are a major target of neuromodulatory inputs that can influence circuit computations by shifting the balance between excitation and inhibition [126]. They express a mixed population of nicotinic receptors, and are densely innervated by both extracortical and recently discovered intracortical cholinergic inputs [98,124]. Even though L1 is largely devoid of neuronal cell bodies, it is rich in spiny apical arborizations of dendrites originating in deeper cortical layers. As such, L1 has been identified as a prominent feedback pathway receiving glutamatergic projections from higher cortical areas as well as various thalamic nuclei [125,127]. Thalamic inputs in the PFC preferentially target L1 neurons and trigger feed-forward inhibition in neighboring interneurons as well as L2/3 pyramidal cells [128,129]. They may, therefore, act as a potential gate of prefrontal pain processing.
In the primary somatosensory cortex (S1) of neuropathic mice, dendritic apical branches of pyramidal neurons show increased spine motility and apical tufts exhibit increased dendritic Ca2+ spikes within days after surgery and this is preserved for almost a month [130,131]. Similar processes apply to the mPFC, and chronic stress exposure induces short- and long-term synaptic changes, in particular atrophy and reduction in spine density of L5 pyramidal cell dendrites [132,133]. Likewise, apical dendrites of both L2/3 and L5 mPFC pyramidal neurons undergo alterations in dendritic morphology and length after nerve injury [84]. Chronic constriction injury reduces expression of the hyperpolarization-activated and cyclic nucleotide-gated HCN1 channel in layer L1 of the ACC [134].

4.2. Layer 2/3

Pyramidal neurons of layer L2 show a unique morphological phenotype with a wide span of apical and a particularly narrow span of basal dendrites [135]. Dendritic spines of L2/3 pyramidal neurons receive mostly excitatory glutamatergic projections from the midline thalamic nucleus, ventral hippocampus, basolateral amygdala (BLA), as well as the contralateral mPFC, but only L3 pyramidal neurons generate action potential bursts in response to depolarizing currents [58,135,136]. L2/3 pyramidal neurons belong to the group of intratelencephalic neurons, targeting the ipsilateral perirhinal cortex, amygdala, and striatum as well as the contralateral striatum and cortex [137,138]. They also influence the local circuitry by providing monosynaptic excitation of L5 pyramidal neurons, as well as by targeting GABAergic interneurons, highlighting their importance in balancing excitatory and inhibitory transmission [139]. Increased NMDA/AMPA ratio, as well as higher complexity and spine density of basal dendrites in ACC/PrL L2/3 pyramidal neurons, is observed one week after surgery and can be linked to sprouting of noradrenergic fibers in L2/3 and subsequent alpha2 induced modulation of Ih currents [25,140]. Increased input resistance and depolarized membrane potential, together with enhanced glutamatergic release are reported in PrL L2/3, whereas IL shows increased length and complexity of only L2/3 but not L5 pyramidal neuron dendrites, suggesting regional and laminar dichotomy [75,84,141,142,143,144]. The amount of excitatory/inhibitory synaptic connections is bidirectionally diminished, resulting in a misbalanced excitatory/inhibitory ratio, and apoptotic loss of subpopulations of inhibitory neurons has been described [79,145].
The ACC is the most dorsal subregion of the mPFC and often studied separately due to its prominent role in the affective component of pain processing [146]. Neuropathy enhances presynaptic glutamate release, postsynaptic AMPA mediated response as well as excitability of ACC layer L2/3 pyramidal neurons, in particular those which are intermediately firing [68,73,75,144]. Apart from augmented glutamatergic transmission, prolonged nociceptive sensitization during the initial induction phase can weaken inhibitory control in layer L2/3 as indicated by reduced frequency of miniature and spontaneous inhibitory postsynaptic currents, suggesting presynaptic GABAergic plasticity [147]. In addition, long-term depression (LTD) can be impaired [148,149]. Lastly, ex vivo findings are supported by in vivo electrophysiological measurements, as long-term changes in spontaneous membrane oscillations are observed in the ACC layer L2/3 after nerve injury [150].

4.3. Layer 5

L5 pyramidal neurons differ from L2/3 pyramidal neurons by a thick band of thalamocortical fibers present in deep L3, as well as by their larger cell bodies [135,138,151]. In contrast to sensory cortices, thalamic afferents in deeper cortical layers are sparse. However, long apical dendrites of L5 pyramidal neurons climb towards layer L1 and form highly branched apical tufts allowing them to be accessed by thalamic inputs [151]. L5 pyramidal neurons are diverse in terms of their morphological and electrophysiological features [135]. In contrast to layer L2/3, which has long been regarded as the input layer of the PFC, layer L5 represents a typical output layer of the cortex with two projection-specific subpopulations: intratelencephalic neurons, also present and already described in layer L2/3, and pyramidal tract neurons, which project mainly to the ipsilateral striatum, thalamus, or brainstem [126,152]. The two neuron subpopulations differ in their intrinsic electrophysiological properties, ion channel composition, and responsiveness to acetylcholine, noradrenaline, and adenosine [126,152].
Evidence for input-specificity of synaptic alterations onto L5 pyramidal and interneurons is emerging from rodent models of sustained pain [153,154,155,156]. ACC L5 pyramidal neurons increase their excitability following nerve injury by modifications of pacemaker ion channels of the HCN family through glutamatergic signaling via metabotropic mGluR1 [74,79,134,154,157]. Even though L5 pyramidal neurons projecting to the contralateral ACC show barely any Ih current and express remarkably low levels of HCN1, they still display hyperactivity after nerve injury [152,154,156]. The downregulation of voltage-gated potassium channels of the Kv2 family, which can be rescued by mGlur1 antagonists, further contributes to this effect [154]. Layer L5 pyramidal neurons are tightly regulated by local inhibitory networks, and an excitatory/inhibitory imbalance may account for the observed changes. Even though the excitability of ACC L5 fast-spiking (FS) interneurons is unaltered, the loss of local synaptic connectivity between pyramidal neurons and these interneurons, together with an imbalance in excitation-inhibition favor excitation in neuropathic pain [79,156,158].
In contrast to superficial ACC/PrL and deep IL, pyramidal neurons of PrL layer L5 develop reduced excitability in chronic pain models [68,73,79,84,88,159,160,161]. This is at least in part mediated by enhanced feed-forward inhibition of L5 pyramidal neurons by PV+ interneurons across different layers [88,160]. A long-range BLA to mPFC to PAG to spinal cord pathway links nerve injury to the enforcement of the BLA-driven input onto PrL L5 PV+ interneurons, followed by reduction of PrL L5 pyramidal neuron output to ventrolateral PAG (vlPAG), which compromises the descending modulatory pain control towards the spinal cord [161]. Such changes can also be driven by reduced excitatory input and can be linked to inputs originating in both the mediodorsal thalamus and the ventral hippocampus [153,162]. They stress the differential roles of neuron populations in different layers and mPFC subregions in the processing of chronic pain.

4.4. Layer 6

In acute brain slices, L6 is easy to distinguish from L5 since its pyramidal neurons possess smaller cell bodies [135,137,163]. More than a third of L6 pyramidal neurons extend large apical dendrites into layer L1 where they receive extracortical input. In contrast to L5, the dendrites of the L6 pyramidal neurons are not oriented perpendicularly towards the brain surface, but rather are oblique or inverted [135]. L6 receives dense cholinergic innervation, which drives a surprisingly strong nicotinic receptor-dependent excitation of pyramidal neurons [164,165]. This distinguishes them from other pyramidal neurons within the mPFC, making them particularly susceptible to the influence of cholinergic projections or local interneurons (reviewed by [18]). A potential role of layer L6 in chronic pain is supported by the activity-dependent increase in p-ERK 1/2 expression not only in L2/3 and L5, but also in layer L6 of the mPFC two weeks following induction of trigeminal neuropathy [166]. The majority of L6 pyramidal neurons give rise to corticothalamic projections and form a corticothalamic feedback pathway that may influence the reliability of thalamic responses to nociceptive stimuli [42]. Other prominent projections target the lateral hypothalamus, the dorsal striatum, thalamic nuclei, but also cortical microcircuits [16,167,168,169].

5. Specific Roles of mPFC Subregions

A multitude of preclinical studies address the role of the mPFC in general for acute, inflammatory, or neuropathic pain. For example, activation of pyramidal neurons in the mPFC attenuates visceral pain [170]. Inflammatory or neuropathic pain leads to impaired performance in the novel object recognition task, cognitive impairment, anxiety, and depression-like behavior as well as reduced sociability [166,171,172,173]. These functional deficits are associated with the deregulated activity or expression of neurotransmitters, their receptors, or regulatory enzymes within the mPFC [30,104,174,175,176,177,178,179,180].

5.1. The Anterior Cingulate Cortex (ACC)

The anterior cingulate cortex is critically important for the affective component of the pain sensation and pain-related mood disorders [181,182,183,184]. Lesions of the ACC alleviate the affective aspects of pain, decrease escape/avoidance behaviors, and decrease the floating time in the forced swim test in neuropathic pain models [185,186,187,188]. In particular, the rostral ACC appears to be important for processing aversive stimuli and posterior lesions abolish the anti-aversive effect of rostral ACC lesion [189]. Although a dual role of ACC in pain sensitization and affect has been corroborated by differential effects observed with ACC microinjections vs. systemic administration of drugs such as morphine, gabapentin, or T-type calcium channel inhibitors, the mechanistic details regarding its modulatory effect on mechanical and thermal sensitivity, spontaneous pain behavior, inflammatory pain, and formalin-induced mechanical hypersensitivity are not yet fully consistent [190,191,192,193]. Different neuronal subpopulations may be relevant in this context. Optogenetic strategies targeting individual neuronal subpopulations help to dissect the local circuits within the ACC [184,189,194,195,196]. Optogenetic stimulation of Thy-1 expressing ACC neurons enhances anxiety and depressive-like behaviors in mice and in turn, bilateral optogenetic inhibition of ACC results in anti-aversive and anti-anxiodepressive effects [188,197]. Optogenetic silencing of excitatory CaMKII positive neurons induces enhanced exploration in the least preferred compartment in a CPP paradigm, and this is not observed by activation of the inhibitory PV+ neurons [198]. Besides their local mPFC connectivity, ACC neurons synapse onto neurons in the neighboring midcingulate and retrosplenial cortices [199]. Pyramidal neurons located in deeper layers project to the hypothalamus, amygdala, PAG, the contralateral hemisphere or other supraspinal areas as well as directly to the spinal dorsal horn [134,156,183]. Direct excitatory projections to neurons in laminae I-III of the spinal dorsal horn are activated in models of neuropathic and cancer pain [200,201,202,203]. Electrical or chemical activation of the ACC facilitates the tail-flick reflex, supporting the existence of descending facilitation [204].
Although there is a general consent that ACC exhibits hyperexcitability in neuropathic pain conditions, there are studies indicating that this generalization may not be absolute. Projections from the mediodorsal thalamus to layer L5 ACC pyramidal neurons elicit both excitation and inhibition, the ratio of which would favor inhibition in a subpopulation of pyramidal neurons that project to subcortical areas [156]. The existence of two different subpopulations of pyramidal neurons in the ACC L5 has been previously reported and the inhibition of the neurons that project to the contralateral side to a nerve injury promotes analgesia without influencing aversive behavior [134]. Inhibiting the excitatory ACC neuronal subpopulation results in aversive behavior during neuropathic pain [156].

5.2. The Prelimbic (PrL) and Infralimbic (IL) Cortices

Only recently, the specificity of mPFC subregions is receiving more attention [170,205,206,207,208,209]. In particular, electrophysiological recordings of L5 pyramidal cells reveal relevant functional differences between PrL and IL: the GABAA antagonist bicuculline modulates neuron activity in PrL, whereas cannabinoid receptor CB1 and mGLuR5 are relevant components in IL [22,27,210].
Prelimbic cortex neurons show increased firing rates after noxious stimulation, while maintained nociceptive input suppresses both basal spontaneous and pain-evoked activity, which can be ascribed to increased gamma-aminobutyric acid (GABA) and decreased glutamate in PrL/IL [211,212,213,214,215,216]. Neuropathic pain causes disruption of working memory and PrL neuronal firing activity, which is reversed by selective inhibition of PrL pyramidal cells using optogenetic interventions [217]. Already one week after induction of neuropathic pain, L5 pyramidal cells of PrL/ACC show reduced responses to excitatory glutamatergic inputs, without changes in intrinsic excitability [153]. This functional deactivation of PrL pyramidal cells may be attributed to a feed-forward inhibition mediated by PV+ GABAergic interneurons, and modulation of these neurons [88]. Anxiety-like behaviors develop at later stages [218]. In neuropathic pain models, metabotropic glutamate receptors mGluR5 and mGluR7 are upregulated in PrL, and pharmacological interventions suggest opposing roles of these two receptors: while mGluR5 aggravates comorbid pain and negative symptoms, mGluR7 improves cognitive performance and restores excitation/inhibition balance at PrL neurons [219,220]. Optogenetic inhibition of the PrL, or one of its projection targets, the nucleus accumbens, augments sensory and affective symptoms of acute pain, and also increases nociceptive sensitivity and aversive responses in neuropathic pain models [221]. Altogether, deactivation of PrL appears to be associated with affective deficits associated with persistent pain and inhibition of hyperactivity of amygdala neurons reverses the deactivation of PrL and concomitant cognitive deficits [26]. This suggests a general shut-down of PrL as an important pathogenetic mechanism that can be overcome by low-frequency electrical stimulation of the PrL cortex or optogenetic activation of PrL layer L5 neurons to relieve both sensory and affective signatures of acute and sustained pain [211,212,222].
In the infralimbic cortex, both mGluR1 and mGluR5 activation induces pronociceptive behavior in monoarthritic rats, which is based on increased activation of dorsal reticular nucleus neurons [223,224]. Peripheral inflammation decreases BDNF levels in the IL but not PrL, and infusion of BDNF into the IL alleviates inflammatory pain and accelerates long-term recovery from inflammatory pain [225]. Attentional deficits are more prevalent in male mice subjected to spared nerve injury (SNI) and correlated to loss of parvalbumin in L5/6 of IL, suggesting a potential role of sex differences in affective pain processing for this subregion [89].

6. Pharmacology of mPFC Neurotransmitter Systems

6.1. Glutamate

Glutamate serves as the primary excitatory neurotransmitter and as a key neuromodulator to control synapse and circuit function over a wide range of spatial and temporal scales in mammalian brains. This functional diversity is implemented by two receptor families, the ionotropic glutamate receptors (iGluRs) of the AMPA or NMDA type and the metabotropic glutamate receptors (mGluRs) [226]. Increased activity of the mPFC is mainly associated with pain-alleviating effects in chronic pain models and infusion of local anesthetic results in nociception, suggesting tonic analgesic activity of these brain regions [223]. Reduced glutamate in the mPFC is associated with emotional and cognitive dysregulation in people with chronic pain [227]. Correspondingly, microinfusion of glutamatergic agonists as well as a partial NMDA receptor agonist in the PFC or even introducing NMDA precursors to the drinking water reduces pain-like behavior, improves the performance in cognitive tasks and increases social interaction behavior in neuropathic rats [173,228]. Direct injection of AMPA receptor positive allosteric modulators in the PFC has an antinociceptive effect and increases the analgesic effect of morphine [229].
A mechanistic approach towards these pharmacological observations is however complicated by the complexity of circuits and differential contributions of glutamatergic receptors at different neuron populations resulting in diverging reports on the contribution and regulation of iGluRs and mGluRs in neuropathic pain models. Injection of the iGluR agonist AMPA into the PrL/IL increases allodynia, and AMPA antagonists inhibit allodynia [230]. Nerve injuries increase AMPA type iGluR (GluA1) accumulation in postsynaptic densities of ACC layer L5 pyramidal neurons, and GluA1 phosphorylation by PKA and redistribution from the cytosol to the membrane plays a critical role in LTP induction and ACC sensitization as well as mechanical hypersensitivity [75,158,200,231,232]. Not surprisingly, GluA1 inhibition or disruption of GluA1 phosphorylation in the ACC ameliorates mechanical hypersensitivity resulting from nerve injury [158,200]. PKMzeta, an important factor in neuronal synaptic potentiation, regulates AMPA receptor trafficking in ACC layer L2/3 pyramidal neurons and inhibition of PKMzeta reverses the increased amplitudes of synaptic currents and attenuates mechanical hypersensitivity [233,234]. Since GluA1 is a downstream target of PKMzeta, and PKMzeta inhibition alleviates inflammation-induced mechanical hypersensitivity as well as anxiety-related behaviors, presumably by normalizing GluA1 protein levels, this is suggested to underlie the pain-relieving effects of electroacupuncture [235].
In contrast, the role of NMDA type iGluRs in mPFC is controversial: NMDA or NMDA receptor agonists in the PrL induce antinociception in neuropathic pain models, whereas an NMDA antagonist reduces fear-induced anti-nociception [236,237,238]. Peripheral inflammation increases glutamatergic presynaptic release, miniature excitatory postsynaptic current (mEPSC) frequencies, and neuronal sensitization via iGluRs containing the NMDA type subunit NR2B; and the increased probability of glutamate release is necessary for the development of inflammatory pain [231,239,240,241,242,243,244]. Additionally, NMDA iGluR subunits NR2B and NR2A are closely related to aversion induced by inflammatory pain [62,245,246,247]. Several enzymes and kinases such as AC1/cAMP/PKA, ERK, p38 MAPK, or modulators such as Zif268/Egr1 affect the circuits that are relevant for aversive reactions in the ACC [248,249,250,251]. Interestingly, caveolin-1 is involved in the ERK/CREB signaling pathway and NR2B membrane trafficking [252,253]. NR2B is upregulated via BDNF/mTOR signaling and TrkB or mTOR inhibition can prevent ACC hyperexcitability, as well as aversive behavior and cold hypersensitivity [247,254,255]. NR2B in the rostral ACC is involved in neuropathic pain-induced aversion as a downstream target of BDNF, and antagonizing TrkB/mTOR augments mechanical hypersensitivity [256,257]. Together, NR2B inhibition and/or downregulation alleviate hypersensitivity and affective behavior and relieve mechanical and thermal hypersensitivity in rodent models of neuropathic and bone cancer pain [62,232,239,245,246,247,256,258]. However, this is in conflict with the overall analgesic effect reported from NMDA agonists, which improve pain-like behavior and cognitive performance in neuropathic rats [236,237,238]. Similar inconsistencies are emerging for metabotropic mGLuR1 receptors, which are upregulated in L5 pyramidal ACC neurons, as a result of a nerve injury, and induce neuronal excitation via the PKC signaling pathway. mGluR1 knockdown selectively attenuates thermal, however, not mechanical hypersensitivity or escape/avoidance behavior [134]. Acid-sensing ion channel 1a (ASIC1a) may act as a modulator of cortical LTP downstream of mGluR1 since ASIC1a inhibition abolishes thermal and mechanical hypersensitivity, however, does not affect aversive behavior [259]. In contrast to mGluR1, mGluR5 is associated with depolarization-induced suppression of excitation and exhibits anxiolytic effects [260].
Several pharmacological tools targeting glutamatergic signaling induce antinociceptive and/or anxiolytic effects in inflammation, presumably by inhibiting the upregulation of GluR1, NR2A, and NR2B in the ACC, as well as by reducing microglia and astrocyte activation and the induction of proinflammatory cytokines [261,262,263]. However, since these potential drugs are so far not directly applied to the ACC, accounting for other effects is challenging and more specific approaches are necessary to target specific neuron populations and their locations within defined brain regions to provide more mechanistic understanding. As a novel and possibly more selective approach intracranial injection of miR-539 mimics reduces NR2B expression and attenuates mechanical hypersensitivity, indicating that microRNAs may be of potential use to target specific genes in neuronal subpopulations in the brain for therapeutic purposes [232].

6.2. Gamma-Aminobutyric Acid (GABA)

In the mammalian brain, GABA is synthesized primarily from glutamate, loaded into synaptic vesicles by a vesicular neurotransmitter transporter (VGAT) and released from synaptic terminals mainly of local inhibitory neurons in a calcium-dependent manner. Ionotropic GABAA receptors carry chloride, however, whereas GABAA receptors normally are composed of combinations of several subunit types, GABAA-ρ receptors are composed of only ρ-subunits [264]. Metabotropic GABAB receptors are associated with G proteins (for review see [265]). Chronic pain leads to an increased input onto mPFC GABAergic interneurons that can inhibit pyramidal neurons, which in turn shuts down the descending inhibitory efferent pathways [26,88,160]. An example is the input from the thalamic paraventricular nucleus to inhibitory neurons of the mPFC, which is enhanced in a rat model of visceral pain [170]. Correspondingly, inhibition of this input attenuates pain through descending pain modulation. Inflammatory pain induces a BLA driven increase in levels of GABA in the PrL, resulting in reduced excitability of pyramidal neurons, and application of the GABAA receptor blocker bicuculline reverses inflammatory mechanical hypersensitivity [211]. Additionally, neuropathy deactivates PrL pyramidal neurons as a result of a feed-forward inhibition via FS-PV+ GABAergic interneurons, and optogenetic inhibition or activation of inhibitory circuits in the mPFC alleviates pain or increases the nocifensive behavior [88,160].
Neuropathic pain enhances the expression levels of GABA transporter-1 (GAT-1), GABAA subunits β2, β3, δ, and γ2, as well as GABAB subunit β2 in the ACC [266,267], and GABAA and GABAB receptors appear to target different components of neuropathic pain. Inhibition of GABAB in the ACC increases sensitivity to mechanical stimuli in naive animals, whereas activation ameliorates nerve injury-induced mechanical hypersensitivity [268]. Surprisingly, activation of GABAA receptors in PrL and IL increases mechanical but not thermal pain sensitivity [269]. In the rostral ACC, microinjection of GABAA but not GABAB agonists attenuates escape/avoidance behavior induced by nerve injury or peripheral inflammation, however without affecting mechanical hypersensitivity [270,271]. PrL GABAA receptor agonists impair the acquisition of formalin-induced aversive behavior, which suggests a specific role of PrL for affective pain processing [272].
As a consequence of peripheral inflammation, altered synaptic inhibitory transmission develops due to impairment of the GABAergic presynaptic vesicle membrane fusion machinery, which may explain decreased GABA transmission [147]. Finally, apoptosis of inhibitory GABAergic interneurons may occur as a consequence of peripheral nerve lesion and could add to the reduction in inhibitory/excitatory ACC connections as well as the gray matter loss associated with SNI-induced mechanical hypersensitivity [79,145,273].

6.3. Endogenous Opioids

mPFC disinhibition can also be triggered by the activation of opioid receptors expressed on GABAergic interneurons. Clinical studies show decreased opioid binding in the contralateral PFC of patients with peripheral neuropathic pain [274]. This indicates increased endogenous opioid release, which leads to receptor internalization or even neuronal loss [275]. In a rat model of neuropathic pain, morphine in the ventrolateral PFC alleviates mechanical allodynia in a dose-dependent manner, and this effect is blocked by the opioid receptor antagonist naloxone [276]. The opioid-induced inhibition in the PFC is mediated mainly via µ-opioid receptors, but expression of kappa-opioid receptors and prodynorphin mRNA is increased in the PFC of neuropathic mice and thus kappa-opioid receptors may be relevant likewise [277].

6.4. Noradrenaline (NA)

The majority of noradrenergic terminals in the cortex originate in the locus coeruleus in the midbrain. NA release from neurons arriving from locus coeruleus at the mPFC suppresses nociceptive-evoked activity in mPFC via alpha2B adrenergic receptors to attenuate neuropathic mechanical hypersensitivity [278,279,280]. Interestingly, neuropathic pain induces noradrenergic impairment and locus coeruleus neurons exhibit increased bursting activity, increased expression of enzymes of noradrenergic metabolism and transport as well as increased expression and sensitivity of α2-adrenoreceptors [281]. HCN channels are regulated by α2-adrenergic receptors, which highlights the role of these channels and the lowering of cAMP in reducing persistent firing and hyperexcitability of layer L2/3 mPFC pyramidal neurons [72]. In contrast, proalgesic effects appear to be mediated by alpha1 receptors [282]. Released noradrenaline induces persistent firing in pyramidal neurons of the PFC via α1-adrenoreceptors that are located on glutamatergic terminals in the PFC and facilitate fast synaptic transmission [283].

6.5. Dopamine (DA)

GABAergic neurons of the mPFC are largely innervated by dopaminergic fibers, and the availability of the D2/D3 receptor subtypes in the PFC correlates with the efficacy of placebo analgesia, suggesting a role of the dopaminergic neurotransmitter system in pain processing [284]. Manipulating DA levels in the mPFC can modulate nociceptive responses in a differential, dose-dependent manner. For example, low DA levels in the mPFC are correlated with increased pain sensitivity and exploratory behaviors, while high DA levels reduce pain sensitivity and facilitate stress-related responses, such as escape from threat [285]. Correspondingly, high-frequency stimulation of the dopaminergic input as well as microinjection of a selective dopamine D2 agonist into the ACC/PrL cause long-lasting suppression of nociceptive responses in the PFC [286]. The antinociceptive effect of DA is blocked by the administration of the D2 receptor antagonist sulpiride which suggests D2 specific nociceptive signaling in the PFC, but also highlights the importance of deeper cortical layers, where D2 receptors are mainly expressed [287].

6.6. Serotonin (5-HT)

Exposure to a novel environment reduces formalin-induced acute nociceptive behavior, which is accompanied by reduced 5-HT and DA metabolites in the mPFC [288]. Correspondingly, the analgesic serotonin receptor 5-HT1A agonist befiradol reduces the activity of dorsal raphe serotonergic neurons and increases the discharge rates of mPFC pyramidal neurons, which is paralleled by decreased extracellular serotonin and increased dopamine in the mPFC [289]. L5 ACC pyramidal neurons express high levels of HCN channels, which develop dysfunction in neuropathic pain models [72]. Serotonin neurotransmission via 5-HT7 restores HCN function and normalizes mechanical hypersensitivity induced by nerve injury [157,290].

6.7. Acetylcholine (ACh)

ACh exerts its neuromodulatory function by both synaptic as well as volume transmission [291]. ACh release can be phasic within seconds or tonic lasting for minutes, which would induce short-term or long-term modulations of the mPFC activity depending on the attentional demands [291,292,293]. Two types of ACh receptors (AChRs), the ionotropic nicotinic (nAChRs) and metabotropic muscarinic (mAChRs) receptors, bind ACh. nAChRs are pentameric ligand-gated ion channels that open upon ACh binding, allowing a direct cationic inward current to depolarize the membrane. Depending on subunit composition they show different affinities for ACh binding, a timescale of evoked currents, as well as calcium conductance [294,295,296,297]. mAChRs are G-protein coupled receptors that interact either with Gq/11 proteins (M1, M3, and M5 mAChRs) or Gi/o proteins (M2 and M5 mAChRs) inhibiting adenylate cyclase [298,299].
Cholinergic modulation of the mPFC has been implicated in various aspects of behavior and cognition, particularly arousal, learning, and attention [20,293]. Cortical gamma band oscillations (20-80 Hz) are a representation of cortical synchrony underlying memory formation, and they encode tonic pain perception in the mPFC [300,301]. These oscillations can be induced by cholinergic agonists, suggesting the role of cholinergic signaling in oscillatory activity patterns [302]. While cholinergic signaling under physiological conditions increases the excitability of mPFC neurons, impaired cholinergic signaling appears to contribute to the deactivation of the mPFC and, therefore, appear to be responsible for the deficits found in chronic pain patients. Accordingly, cholinergic modulation of PrL layer L5 pyramidal cells is impaired after nerve injury and this appears to be related to reduced surface expression of M1 mAChRs in PrL layer L5 (Table 1) [155].
Chemogenetic inhibition of medial septum cholinergic neurons abolishes the hyperexcitability of rostral ACC pyramidal neurons, reduces CPA, and alleviates inflammation-induced mechanical and thermal hypersensitivity, whereas activation of the same medial septum neurons abolishes hypersensitivity by activating the ventral hippocampus [50]. Interestingly, the M1-induced increase in mechanical thresholds observed in injured rats, is also enhancing GABAA mediated transmission in the ACC [303,304]. Although the importance of ACh and cholinergic signaling in the mPFC has been clinically recognized, since for example the anti-depressive effect of the muscarinic antagonist scopolamine is mediated by PrL/IL M1 receptors, neither the (micro-)circuitry of cholinergic control nor the mechanisms underlying the SNI-induced downregulation of muscarinic receptors at PrL pyramidal neurons are fully understood so far [155,293,305].

6.8. Endocannabinoids

Mice with induced osteoarthritis exhibit elevated levels of the endocannabinoid 2-arachidonoylglycerol in the PFC [308]. As previously discussed, the excitability of PFC glutamatergic neurons is inversely related to amygdala activity during painful experiences. Coactivation of CB1 cannabinoid receptors and mGluR5 in the IL of arthritic rats inhibits the pain-related increase of amygdala output neuron activity [22]. Correspondingly, systemic application of the endocannabinoid palmitoylethanolamide reverses changes in the glutamatergic synapses in the mPFC, and resolves pain and depressive-like symptoms associated with neuropathic pain [104]. In complete Freund’s adjuvant (CFA) treated mice, this process is suppressed and mGluR5 protein levels are reduced, whereas injection of a positive allosteric modulator of GluR5 into ACC restores depolarization-induced suppression of excitation and exhibits antinociceptive and anxiolytic effects [260]. In particular, endocannabinoids acting on CB1 receptors are critically involved in a long-range brain circuit targeting PAG, which alters pain behaviors by reducing descending noradrenergic and serotoninergic modulation of spinal pain signals [161]. Interestingly, TRPV1 channels, which can be activated by endocannabinoids, modulate pain processing in the brain [309]. Neuropathic pain induces upregulation of TRPV1 together with fatty acid amide hydrolase (FAAH), and single acute microinjection of combined TRPV1/FAAH antagonists into PrL/IL restores aberrant cortical neuron activity [174], whereas microinjection of the TRPV1 agonist capsaicin decreases allodynia [230].

7. mPFC Output Regions and Descending Modulation

Common brain regions targeted by projections from both PrL and IL are medial thalamic nuclei, the anterior piriform cortex, olfactory forebrain areas, and the orbitomedial, insular, and entorhinal cortices, as well as the midbrain dopaminergic ventral tegmental area (VTA) and medial regions of the periaqueductal gray (PAG) [310]. In addition, dense reciprocal connections exist between the PrL and IL, as well as the ACC [42,311,312]. However, PrL and IL differ in their specific connections to certain subregions and nuclei within those brain areas [310]. Only PrL connects to the agranular insular cortex, the olfactory tubercle, and the nucleus accumbens, the paraventricular, mediodorsal, and reuniens nuclei of the thalamus, the capsular part of the central nucleus and the basolateral nucleus of the amygdala, as well as the dorsal and median raphe nuclei of the brainstem. The lateral septum, the bed nucleus of the stria terminalis and preoptic nuclei, the substantia innominate and the endopiriform nuclei of the basal forebrain, the medial, basomedial, central, and cortical nuclei of the amygdala, different hypothalamic nuclei, including the supramammillary nuclei, as well as the parabrachial and solitary nuclei of the brainstem are exclusively targeted by IL cortex [310,313]. Augmented synaptic connections mediate a feed-forward inhibition of projections from the PFC to the ventrolateral PAG region and its downstream target regions, and these connections modulate pain behaviors by weakening the descending noradrenergic and serotoninergic inhibitory modulation of spinal pain signals [161]. Regarding projections to the nucleus accumbens, neurons from the PrL project to both the core and shell subregions, whereas IL projections only innervate the medial portion of the shell subregion [310,314]. This difference in output systems might be functionally relevant, constituting a dichotomous “Go/Stop” system for reward-seeking and fear-related behaviors [315].

8. Summary

In contrast to acute pain, chronic pain serves no biological or protective function but rather represents a disease by itself, and in many cases effective treatments for chronic pain are not available. Chronic pain reduces the quality of life of patients and leads to comorbidities such as anxiety, depression, insomnia, and cognitive disturbance. The relevant affective and cognitive components of the pain sensation are differentially processed in the different subregions of the mPFC by a complex array of inputs targeting the local circuits via multiple neurotransmitter systems (Table 1). Subregion-specific alterations in feed-forward inhibition of pyramidal neurons, and morphological changes of both interneurons and pyramidal neurons arise in different layers of those subregions and contribute to ACC hyperexcitability in contrast to a partial PrL shutdown (Figure 2). The cholinergic system appears particularly relevant since cholinergic projections from the basal forebrain contribute to the cognitive and emotional impairments induced by pain, and inhibition of cholinergic input from the medial septum projecting to the ACC alleviates pain-induced anxiety. While cholinergic signaling under physiological conditions increases excitability in the mPFC, impaired cholinergic signaling appears to contribute to the deactivation of the mPFC and, therefore, may be partially relevant for the deficits found in chronic pain patients. Furthermore, the basis of placebo induced analgesia or expectancy of pain relief is based on the ACC-PFC-PAG descending pain pathways. These predominant roles in affective pain processing put the medial prefrontal cortex in a pole position as a hub for the development of mental comorbidities associated with chronic pain and a relevant target region for therapeutic interventions.

Author Contributions

K.K.K., M.M., T.K., and M.K. wrote the manuscript. All authors have approved the final version of the manuscript.

Funding

The study was supported by the Austrian Science Fund (FWF) grants P30809 to K.K., and DK-SPIN W1206-06 and P253450 to M.K.

Acknowledgments

Open Access Funding by the Austrian Science Fund (FWF).

Conflicts of Interest

The authors report no conflicts of interest.

References

  1. Apkarian, A.V.; Bushnell, M.C.; Treede, R.D.; Zubieta, J.K. Human brain mechanisms of pain perception and regulation in health and disease. Eur. J. Pain 2005, 9, 463–484. [Google Scholar] [CrossRef] [PubMed]
  2. Tracey, I.; Johns, E. The pain matrix: Reloaded or reborn as we image tonic pain using arterial spin labelling. Pain 2010, 148, 359–360. [Google Scholar] [CrossRef] [PubMed]
  3. Bernard, J.F.; Bester, H.; Besson, J.M. Involvement of the spino-parabrachio -amygdaloid and -hypothalamic pathways in the autonomic and affective emotional aspects of pain. Prog. Brain Res. 1996, 107, 243–255. [Google Scholar]
  4. Becerra, L.; Breiter, H.C.; Wise, R.; Gonzalez, R.G.; Borsook, D. Reward circuitry activation by noxious thermal stimuli. Neuron 2001, 32, 927–946. [Google Scholar] [CrossRef] [Green Version]
  5. Baliki, M.N.; Geha, P.Y.; Fields, H.L.; Apkarian, A.V. Predicting value of pain and analgesia: Nucleus accumbens response to noxious stimuli changes in the presence of chronic pain. Neuron 2010, 66, 149–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Bushnell, M.C.; Ceko, M.; Low, L.A. Cognitive and emotional control of pain and its disruption in chronic pain. Nat. Rev. Neurosci. 2013, 14, 502–511. [Google Scholar] [CrossRef] [Green Version]
  7. Rainville, P.; Duncan, G.H.; Price, D.D.; Carrier, B.; Bushnell, M.C. Pain affect encoded in human anterior cingulate but not somatosensory cortex. Science 1997, 277, 968–971. [Google Scholar] [CrossRef] [Green Version]
  8. Bushnell, M.C.; Duncan, G.H.; Hofbauer, R.K.; Ha, B.; Chen, J.I.; Carrier, B. Pain perception: Is there a role for primary somatosensory cortex? Proc. Natl. Acad. Sci. USA 1999, 96, 7705–7709. [Google Scholar] [CrossRef] [Green Version]
  9. Coghill, R.C.; Sang, C.N.; Maisog, J.M.; Iadarola, M.J. Pain intensity processing within the human brain: A bilateral, distributed mechanism. J. Neurophysiol. 1999, 82, 1934–1943. [Google Scholar] [CrossRef]
  10. Runyan, J.D.; Moore, A.N.; Dash, P.K. A role for prefrontal cortex in memory storage for trace fear conditioning. J. Neurosci. 2004, 24, 1288–1295. [Google Scholar] [CrossRef] [Green Version]
  11. Hennig, K.M.; Colombani, J.; Neufeld, T.P. TOR coordinates bulk and targeted endocytosis in the Drosophila melanogaster fat body to regulate cell growth. J. Cell Biol. 2006, 173, 963–974. [Google Scholar] [CrossRef] [PubMed]
  12. Moulton, E.A.; Schmahmann, J.D.; Becerra, L.; Borsook, D. The cerebellum and pain: Passive integrator or active participator? Brain Res. Rev. 2010, 65, 14–27. [Google Scholar] [CrossRef] [Green Version]
  13. Westlund, K.N. Pain Pathways: Peripheral, Spinal, Ascending, and Descending Pathways. In Practical Management of Pain, 5th ed.; Mosby: Philadelphia, PA, USA, 2014. [Google Scholar]
  14. Foilb, A.R.; Flyer-Adams, J.G.; Maier, S.F.; Christianson, J.P. Posterior insular cortex is necessary for conditioned inhibition of fear. Neurobiol. Learn. Mem. 2016, 134, 317–327. [Google Scholar] [CrossRef] [Green Version]
  15. Condes-Lara, M.; Omana Zapata, I.; Leon-Olea, M.; Sanchez-Alvarez, M. Dorsal raphe and nociceptive stimulations evoke convergent responses on the thalamic centralis lateralis and medial prefrontal cortex neurons. Brain Res. 1989, 499, 145–152. [Google Scholar] [CrossRef]
  16. Hardy, S.G. Analgesia elicited by prefrontal stimulation. Brain Res. 1985, 339, 281–284. [Google Scholar] [CrossRef]
  17. Seminowicz, D.A.; Moayedi, M. The Dorsolateral Prefrontal Cortex in Acute and Chronic Pain. J. Pain 2017, 18, 1027–1035. [Google Scholar] [CrossRef] [PubMed]
  18. Naser, P.V.; Kuner, R. Molecular, Cellular and Circuit Basis of Cholinergic Modulation of Pain. Neuroscience 2018, 387, 135–148. [Google Scholar] [CrossRef] [PubMed]
  19. Baliki, M.N.; Geha, P.Y.; Apkarian, A.V.; Chialvo, D.R. Beyond feeling: Chronic pain hurts the brain, disrupting the default-mode network dynamics. J. Neurosci. 2008, 28, 1398–1403. [Google Scholar] [CrossRef] [Green Version]
  20. Bloem, B.; Poorthuis, R.B.; Mansvelder, H.D. Cholinergic modulation of the medial prefrontal cortex: The role of nicotinic receptors in attention and regulation of neuronal activity. Front. Neural. Circuits 2014, 8, 17. [Google Scholar] [CrossRef] [Green Version]
  21. Fields, H.L.; Bry, J.; Hentall, I.; Zorman, G. The activity of neurons in the rostral medulla of the rat during withdrawal from noxious heat. J. Neurosci. 1983, 3, 2545–2552. [Google Scholar] [CrossRef]
  22. Ji, G.; Neugebauer, V. CB1 augments mGluR5 function in medial prefrontal cortical neurons to inhibit amygdala hyperactivity in an arthritis pain model. Eur. J. Neurosci. 2014, 39, 455–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Lee, M.; Manders, T.R.; Eberle, S.E.; Su, C.; D’Amour, J.; Yang, R.; Lin, H.Y.; Deisseroth, K.; Froemke, R.C.; Wang, J. Activation of corticostriatal circuitry relieves chronic neuropathic pain. J. Neurosci. 2015, 35, 5247–5259. [Google Scholar] [CrossRef] [PubMed]
  24. Apkarian, A.V.; Sosa, Y.; Sonty, S.; Levy, R.M.; Harden, R.N.; Parrish, T.B.; Gitelman, D.R. Chronic back pain is associated with decreased prefrontal and thalamic gray matter density. J. Neurosci. 2004, 24, 10410–10415. [Google Scholar] [CrossRef] [PubMed]
  25. Metz, A.E.; Yau, H.J.; Centeno, M.V.; Apkarian, A.V.; Martina, M. Morphological and functional reorganization of rat medial prefrontal cortex in neuropathic pain. Proc. Natl. Acad. Sci. USA 2009, 106, 2423–2428. [Google Scholar] [CrossRef] [Green Version]
  26. Ji, G.; Sun, H.; Fu, Y.; Li, Z.; Pais-Vieira, M.; Galhardo, V.; Neugebauer, V. Cognitive impairment in pain through amygdala-driven prefrontal cortical deactivation. J. Neurosci. 2010, 30, 5451–5464. [Google Scholar] [CrossRef] [PubMed]
  27. Ji, G.; Neugebauer, V. Pain-related deactivation of medial prefrontal cortical neurons involves mGluR1 and GABA(A) receptors. J. Neurophysiol. 2011, 106, 2642–2652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Moayedi, M.; Weissman-Fogel, I.; Crawley, A.P.; Goldberg, M.B.; Freeman, B.V.; Tenenbaum, H.C.; Davis, K.D. Contribution of chronic pain and neuroticism to abnormal forebrain gray matter in patients with temporomandibular disorder. Neuroimage 2011, 55, 277–286. [Google Scholar] [CrossRef] [PubMed]
  29. Obara, I.; Goulding, S.P.; Gould, A.T.; Lominac, K.D.; Hu, J.H.; Zhang, P.W.; von Jonquieres, G.; Dehoff, M.; Xiao, B.; Seeburg, P.H.; et al. Homers at the Interface between Reward and Pain. Front. Psychiatry 2013, 4, 39. [Google Scholar] [CrossRef] [Green Version]
  30. Hung, K.L.; Wang, S.J.; Wang, Y.C.; Chiang, T.R.; Wang, C.C. Upregulation of presynaptic proteins and protein kinases associated with enhanced glutamate release from axonal terminals (synaptosomes) of the medial prefrontal cortex in rats with neuropathic pain. Pain 2014, 155, 377–387. [Google Scholar] [CrossRef]
  31. Kucyi, A.; Moayedi, M.; Weissman-Fogel, I.; Goldberg, M.B.; Freeman, B.V.; Tenenbaum, H.C.; Davis, K.D. Enhanced medial prefrontal-default mode network functional connectivity in chronic pain and its association with pain rumination. J. Neurosci. 2014, 34, 3969–3975. [Google Scholar] [CrossRef] [Green Version]
  32. Moriarty, O.; McGuire, B.E.; Finn, D.P. The effect of pain on cognitive function: A review of clinical and preclinical research. Prog. Neurobiol. 2011, 93, 385–404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Paxinos, G.; Watson, C. The Rat Brain in Stereotaxic Coordinates; Elsevier/Academic Press: New York, NY, USA, 2007. [Google Scholar]
  34. Franklin, K.; Paxinos, G. The Mouse Brain in Stereotaxic Coordinates; Elsevier/Academic Press: New York, NY, USA, 2008. [Google Scholar]
  35. Laubach, M.; Amarante, L.M.; Swanson, K.; White, S.R. What, If Anything, Is Rodent Prefrontal Cortex? eNeuro 2018, 5. [Google Scholar] [CrossRef] [PubMed]
  36. Hsu, M.M.; Shyu, B.C. Electrophysiological study of the connection between medial thalamus and anterior cingulate cortex in the rat. Neuroreport 1997, 8, 2701–2707. [Google Scholar] [CrossRef]
  37. Kung, J.C.; Shyu, B.C. Potentiation of local field potentials in the anterior cingulate cortex evoked by the stimulation of the medial thalamic nuclei in rats. Brain Res. 2002, 953, 37–44. [Google Scholar] [CrossRef]
  38. Yang, J.W.; Shih, H.C.; Shyu, B.C. Intracortical circuits in rat anterior cingulate cortex are activated by nociceptive inputs mediated by medial thalamus. J. Neurophysiol. 2006, 96, 3409–3422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Eto, K.; Wake, H.; Watanabe, M.; Ishibashi, H.; Noda, M.; Yanagawa, Y.; Nabekura, J. Inter-regional contribution of enhanced activity of the primary somatosensory cortex to the anterior cingulate cortex accelerates chronic pain behavior. J. Neurosci. 2011, 31, 7631–7636. [Google Scholar] [CrossRef] [Green Version]
  40. Delevich, K.; Tucciarone, J.; Huang, Z.J.; Li, B. The mediodorsal thalamus drives feedforward inhibition in the anterior cingulate cortex via parvalbumin interneurons. J. Neurosci. 2015, 35, 5743–5753. [Google Scholar] [CrossRef]
  41. Han, S.; Soleiman, M.T.; Soden, M.E.; Zweifel, L.S.; Palmiter, R.D. Elucidating an Affective Pain Circuit that Creates a Threat Memory. Cell 2015, 162, 363–374. [Google Scholar] [CrossRef] [Green Version]
  42. Hoover, W.B.; Vertes, R.P. Anatomical analysis of afferent projections to the medial prefrontal cortex in the rat. Brain Struct. Funct. 2007, 212, 149–179. [Google Scholar] [CrossRef]
  43. Woolf, N.J.; Eckenstein, F.; Butcher, L.L. Cholinergic systems in the rat brain: I. projections to the limbic telencephalon. Brain Res. Bull. 1984, 13, 751–784. [Google Scholar] [CrossRef]
  44. Bloem, B.; Schoppink, L.; Rotaru, D.C.; Faiz, A.; Hendriks, P.; Mansvelder, H.D.; Van de Berg, W.D.; Wouterlood, F.G. Topographic mapping between basal forebrain cholinergic neurons and the medial prefrontal cortex in mice. J. Neurosci. 2014, 34, 16234–16246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Gielow, M.R.; Zaborszky, L. The Input-Output Relationship of the Cholinergic Basal Forebrain. Cell Rep. 2017, 18, 1817–1830. [Google Scholar] [CrossRef] [PubMed]
  46. Chaves-Coira, I.; Martin-Cortecero, J.; Nunez, A.; Rodrigo-Angulo, M.L. Basal Forebrain Nuclei Display Distinct Projecting Pathways and Functional Circuits to Sensory Primary and Prefrontal Cortices in the Rat. Front. Neuroanat. 2018, 12, 69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Ahrlund-Richter, S.; Xuan, Y.; Van Lunteren, J.A.; Kim, H.; Ortiz, C.; Pollak Dorocic, I.; Meletis, K.; Carlen, M. A whole-brain atlas of monosynaptic input targeting four different cell types in the medial prefrontal cortex of the mouse. Nat. Neurosci. 2019, 22, 657–668. [Google Scholar] [CrossRef] [PubMed]
  48. Knox, D.; Keller, S.M. Cholinergic neuronal lesions in the medial septum and vertical limb of the diagonal bands of Broca induce contextual fear memory generalization and impair acquisition of fear extinction. Hippocampus 2016, 26, 718–726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Ang, S.T.; Ariffin, M.Z.; Khanna, S. The forebrain medial septal region and nociception. Neurobiol. Learn. Mem. 2017, 138, 238–251. [Google Scholar] [CrossRef]
  50. Jiang, Y.Y.; Shao, S.; Zhang, Y.; Zheng, J.; Chen, X.; Cui, S.; Liu, F.Y.; Wan, Y.; Yi, M. Neural pathways in medial septal cholinergic modulation of chronic pain: Distinct contribution of the anterior cingulate cortex and ventral hippocampus. Pain 2018, 159, 1550–1561. [Google Scholar] [CrossRef]
  51. Ang, S.T.; Lee, A.T.; Foo, F.C.; Ng, L.; Low, C.M.; Khanna, S. GABAergic neurons of the medial septum play a nodal role in facilitation of nociception-induced affect. Sci. Rep. 2015, 5, 15419. [Google Scholar] [CrossRef] [Green Version]
  52. Ariffin, M.Z.; Ibrahim, K.M.; Lee, A.T.; Lee, R.Z.; Poon, S.Y.; Thong, H.K.; Liu, E.H.C.; Low, C.M.; Khanna, S. Forebrain medial septum sustains experimental neuropathic pain. Sci. Rep. 2018, 8, 11892. [Google Scholar] [CrossRef] [Green Version]
  53. Van De Werd, H.J.; Rajkowska, G.; Evers, P.; Uylings, H.B. Cytoarchitectonic and chemoarchitectonic characterization of the prefrontal cortical areas in the mouse. Brain Struct. Funct. 2010, 214, 339–353. [Google Scholar] [CrossRef] [Green Version]
  54. Fuster, J. The Prefrontal Cortex; Academic Press: London, UK, 2015. [Google Scholar]
  55. Wood, J.N.; Grafman, J. Human prefrontal cortex: Processing and representational perspectives. Nat. Rev. Neurosci. 2003, 4, 139–147. [Google Scholar] [CrossRef] [PubMed]
  56. Uylings, H.B.; Groenewegen, H.J.; Kolb, B. Do rats have a prefrontal cortex? Behav. Brain Res. 2003, 146, 3–17. [Google Scholar] [CrossRef]
  57. Li, X.Y.; Chen, T.; Descalzi, G.; Koga, K.; Qiu, S.; Zhuo, M. Characterization of neuronal intrinsic properties and synaptic transmission in layer I of anterior cingulate cortex from adult mice. Mol. Pain 2012, 8, 53. [Google Scholar] [CrossRef] [Green Version]
  58. Little, J.P.; Carter, A.G. Subcellular synaptic connectivity of layer 2 pyramidal neurons in the medial prefrontal cortex. J. Neurosci. 2012, 32, 12808–12819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. McGarry, L.M.; Carter, A.G. Inhibitory Gating of Basolateral Amygdala Inputs to the Prefrontal Cortex. J. Neurosci. 2016, 36, 9391–9406. [Google Scholar] [CrossRef] [PubMed]
  60. Anastasiades, P.G.; Marlin, J.J.; Carter, A.G. Cell-Type Specificity of Callosally Evoked Excitation and Feedforward Inhibition in the Prefrontal Cortex. Cell Rep. 2018, 22, 679–692. [Google Scholar] [CrossRef] [Green Version]
  61. Collins, D.P.; Anastasiades, P.G.; Marlin, J.J.; Carter, A.G. Reciprocal Circuits Linking the Prefrontal Cortex with Dorsal and Ventral Thalamic Nuclei. Neuron 2018, 98, 366–379 e364. [Google Scholar] [CrossRef] [Green Version]
  62. Wu, L.J.; Li, X.; Chen, T.; Ren, M.; Zhuo, M. Characterization of intracortical synaptic connections in the mouse anterior cingulate cortex using dual patch clamp recording. Mol. Brain 2009, 2, 32. [Google Scholar] [CrossRef] [Green Version]
  63. Beneyto, M.; Meador-Woodruff, J.H. Lamina-specific abnormalities of NMDA receptor-associated postsynaptic protein transcripts in the prefrontal cortex in schizophrenia and bipolar disorder. Neuropsychopharmacology 2008, 33, 2175–2186. [Google Scholar] [CrossRef]
  64. Arion, D.; Corradi, J.P.; Tang, S.; Datta, D.; Boothe, F.; He, A.; Cacace, A.M.; Zaczek, R.; Albright, C.F.; Tseng, G.; et al. Distinctive transcriptome alterations of prefrontal pyramidal neurons in schizophrenia and schizoaffective disorder. Mol. Psychiatry 2015, 20, 1397–1405. [Google Scholar] [CrossRef]
  65. Shrestha, P.; Mousa, A.; Heintz, N. Layer 2/3 pyramidal cells in the medial prefrontal cortex moderate stress induced depressive behaviors. Elife 2015, 4. [Google Scholar] [CrossRef]
  66. Varodayan, F.P.; Sidhu, H.; Kreifeldt, M.; Roberto, M.; Contet, C. Morphological and functional evidence of increased excitatory signaling in the prelimbic cortex during ethanol withdrawal. Neuropharmacology 2018, 133, 470–480. [Google Scholar] [CrossRef] [PubMed]
  67. Bhattacherjee, A.; Djekidel, M.N.; Chen, R.; Chen, W.; Tuesta, L.M.; Zhang, Y. Cell type-specific transcriptional programs in mouse prefrontal cortex during adolescence and addiction. Nat. Commun. 2019, 10, 4169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Cao, X.Y.; Xu, H.; Wu, L.J.; Li, X.Y.; Chen, T.; Zhuo, M. Characterization of intrinsic properties of cingulate pyramidal neurons in adult mice after nerve injury. Mol. Pain 2009, 5, 73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Koga, K.; Li, X.; Chen, T.; Steenland, H.W.; Descalzi, G.; Zhuo, M. In vivo whole-cell patch-clamp recording of sensory synaptic responses of cingulate pyramidal neurons to noxious mechanical stimuli in adult mice. Mol. Pain 2010, 6, 62. [Google Scholar] [CrossRef] [Green Version]
  70. Wang, J.Y.; Luo, F.; Zhang, H.T.; Chang, J.Y.; Woodward, D.J.; Han, J.S. Nociceptive responses of anterior cingulate cortical ensembles in behaving rats. Beijing Da Xue Xue Bao Yi Xue Ban 2004, 36, 47–51. [Google Scholar]
  71. Becerra, L.; Navratilova, E.; Porreca, F.; Borsook, D. Analogous responses in the nucleus accumbens and cingulate cortex to pain onset (aversion) and offset (relief) in rats and humans. J. Neurophysiol. 2013, 110, 1221–1226. [Google Scholar] [CrossRef] [Green Version]
  72. Cordeiro Matos, S.; Zhang, Z.; Seguela, P. Peripheral Neuropathy Induces HCN Channel Dysfunction in Pyramidal Neurons of the Medial Prefrontal Cortex. J. Neurosci. 2015, 35, 13244–13256. [Google Scholar] [CrossRef] [Green Version]
  73. Yang, Z.; Tan, Q.; Cheng, D.; Zhang, L.; Zhang, J.; Gu, E.W.; Fang, W.; Lu, X.; Liu, X. The Changes of Intrinsic Excitability of Pyramidal Neurons in Anterior Cingulate Cortex in Neuropathic Pain. Front. Cell Neurosci. 2018, 12, 436. [Google Scholar] [CrossRef]
  74. Zhao, R.; Zhou, H.; Huang, L.; Xie, Z.; Wang, J.; Gan, W.B.; Yang, G. Neuropathic Pain Causes Pyramidal Neuronal Hyperactivity in the Anterior Cingulate Cortex. Front. Cell Neurosci. 2018, 12, 107. [Google Scholar] [CrossRef] [Green Version]
  75. Xu, H.; Wu, L.J.; Wang, H.; Zhang, X.; Vadakkan, K.I.; Kim, S.S.; Steenland, H.W.; Zhuo, M. Presynaptic and postsynaptic amplifications of neuropathic pain in the anterior cingulate cortex. J. Neurosci. 2008, 28, 7445–7453. [Google Scholar] [CrossRef] [PubMed]
  76. Zhuo, M. Cortical excitation and chronic pain. Trends Neurosci. 2008, 31, 199–207. [Google Scholar] [CrossRef] [PubMed]
  77. Gong, K.R.; Cao, F.L.; He, Y.; Gao, C.Y.; Wang, D.D.; Li, H.; Zhang, F.K.; An, Y.Y.; Lin, Q.; Chen, J. Enhanced excitatory and reduced inhibitory synaptic transmission contribute to persistent pain-induced neuronal hyper-responsiveness in anterior cingulate cortex. Neuroscience 2010, 171, 1314–1325. [Google Scholar] [CrossRef] [PubMed]
  78. Niikura, K.; Furuya, M.; Narita, M.; Torigoe, K.; Kobayashi, Y.; Takemura, Y.; Yamazaki, M.; Horiuchi, H.; Enomoto, T.; Iseki, M.; et al. Enhancement of glutamatergic transmission in the cingulate cortex in response to mild noxious stimuli under a neuropathic pain-like state. Synapse 2011, 65, 424–432. [Google Scholar] [CrossRef]
  79. Blom, S.M.; Pfister, J.P.; Santello, M.; Senn, W.; Nevian, T. Nerve injury-induced neuropathic pain causes disinhibition of the anterior cingulate cortex. J. Neurosci. 2014, 34, 5754–5764. [Google Scholar] [CrossRef] [Green Version]
  80. Da Silva, J.T.; Seminowicz, D.A. Neuroimaging of pain in animal models: A review of recent literature. Pain Rep. 2019, 4, e732. [Google Scholar] [CrossRef]
  81. Rymar, V.V.; Sadikot, A.F. Laminar fate of cortical GABAergic interneurons is dependent on both birthdate and phenotype. J. Comp. Neurol. 2007, 501, 369–380. [Google Scholar] [CrossRef]
  82. Beierlein, M.; Connors, B.W. Short-term dynamics of thalamocortical and intracortical synapses onto layer 6 neurons in neocortex. J. Neurophysiol. 2002, 88, 1924–1932. [Google Scholar] [CrossRef]
  83. Ding, C.; Emmenegger, V.; Schaffrath, K.; Feldmeyer, D. Layer-specific inhibitory microcircuits of layer 6 interneurons in rat prefrontal cortex. bioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  84. Mitric, M.; Seewald, A.; Moschetti, G.; Sacerdote, P.; Ferraguti, F.; Kummer, K.K.; Kress, M. Layer- and subregion-specific electrophysiological and morphological changes of the medial prefrontal cortex in a mouse model of neuropathic pain. Sci. Rep. 2019, 9, 9479. [Google Scholar] [CrossRef]
  85. Kim, H.; Ahrlund-Richter, S.; Wang, X.; Deisseroth, K.; Carlen, M. Prefrontal Parvalbumin Neurons in Control of Attention. Cell 2016, 164, 208–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Sparta, D.R.; Hovelso, N.; Mason, A.O.; Kantak, P.A.; Ung, R.L.; Decot, H.K.; Stuber, G.D. Activation of prefrontal cortical parvalbumin interneurons facilitates extinction of reward-seeking behavior. J. Neurosci. 2014, 34, 3699–3705. [Google Scholar] [CrossRef] [Green Version]
  87. Marek, R.; Jin, J.; Goode, T.D.; Giustino, T.F.; Wang, Q.; Acca, G.M.; Holehonnur, R.; Ploski, J.E.; Fitzgerald, P.J.; Lynagh, T.; et al. Hippocampus-driven feed-forward inhibition of the prefrontal cortex mediates relapse of extinguished fear. Nat. Neurosci. 2018, 21, 384–392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Zhang, Z.; Gadotti, V.M.; Chen, L.; Souza, I.A.; Stemkowski, P.L.; Zamponi, G.W. Role of Prelimbic GABAergic Circuits in Sensory and Emotional Aspects of Neuropathic Pain. Cell Rep. 2015, 12, 752–759. [Google Scholar] [CrossRef] [Green Version]
  89. Shiers, S.; Pradhan, G.; Mwirigi, J.; Mejia, G.; Ahmad, A.; Kroener, S.; Price, T. Neuropathic Pain Creates an Enduring Prefrontal Cortex Dysfunction Corrected by the Type II Diabetic Drug Metformin but Not by Gabapentin. J. Neurosci. 2018, 38, 7337–7350. [Google Scholar] [CrossRef] [PubMed]
  90. Evans, M.D.; Dumitrescu, A.S.; Kruijssen, D.L.H.; Taylor, S.E.; Grubb, M.S. Rapid Modulation of Axon Initial Segment Length Influences Repetitive Spike Firing. Cell Rep. 2015, 13, 1233–1245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Nguyen, R.; Venkatesan, S.; Binko, M.; Bang, J.Y.; Cajanding, J.D.; Briggs, C.; Sargin, D.; Imayoshi, I.; Lambe, E.K.; Kim, J.C. Cholecystokinin-Expressing Interneurons of the Medial Prefrontal Cortex Mediate Working Memory Retrieval. J. Neurosci. 2020, 40, 2314–2331. [Google Scholar] [CrossRef]
  92. Czeh, B.; Vardya, I.; Varga, Z.; Febbraro, F.; Csabai, D.; Martis, L.S.; Hojgaard, K.; Henningsen, K.; Bouzinova, E.V.; Miseta, A.; et al. Long-Term Stress Disrupts the Structural and Functional Integrity of GABAergic Neuronal Networks in the Medial Prefrontal Cortex of Rats. Front. Cell Neurosci. 2018, 12, 148. [Google Scholar] [CrossRef]
  93. Abbas, A.I.; Sundiang, M.J.M.; Henoch, B.; Morton, M.P.; Bolkan, S.S.; Park, A.J.; Harris, A.Z.; Kellendonk, C.; Gordon, J.A. Somatostatin Interneurons Facilitate Hippocampal-Prefrontal Synchrony and Prefrontal Spatial Encoding. Neuron 2018, 100, 926–939 e923. [Google Scholar] [CrossRef] [Green Version]
  94. Scheggia, D.; Manago, F.; Maltese, F.; Bruni, S.; Nigro, M.; Dautan, D.; Latuske, P.; Contarini, G.; Gomez-Gonzalo, M.; Requie, L.M.; et al. Somatostatin interneurons in the prefrontal cortex control affective state discrimination in mice. Nat. Neurosci. 2020, 23, 47–60. [Google Scholar] [CrossRef]
  95. Lee, S.; Kruglikov, I.; Huang, Z.J.; Fishell, G.; Rudy, B. A disinhibitory circuit mediates motor integration in the somatosensory cortex. Nat. Neurosci. 2013, 16, 1662–1670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Pi, H.J.; Hangya, B.; Kvitsiani, D.; Sanders, J.I.; Huang, Z.J.; Kepecs, A. Cortical interneurons that specialize in disinhibitory control. Nature 2013, 503, 521–524. [Google Scholar] [CrossRef] [PubMed]
  97. Karnani, M.M.; Jackson, J.; Ayzenshtat, I.; Hamzehei Sichani, A.; Manoocheri, K.; Kim, S.; Yuste, R. Opening Holes in the Blanket of Inhibition: Localized Lateral Disinhibition by VIP Interneurons. J. Neurosci. 2016, 36, 3471–3480. [Google Scholar] [CrossRef] [PubMed]
  98. Obermayer, J.; Luchicchi, A.; Heistek, T.S.; de Kloet, S.F.; Terra, H.; Bruinsma, B.; Mnie-Filali, O.; Kortleven, C.; Galakhova, A.A.; Khalil, A.J.; et al. Prefrontal cortical ChAT-VIP interneurons provide local excitation by cholinergic synaptic transmission and control attention. Nat. Commun. 2019, 10, 5280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Chiang, C.Y.; Sessle, B.J.; Dostrovsky, J.O. Role of astrocytes in pain. Neurochem. Res. 2012, 37, 2419–2431. [Google Scholar] [CrossRef]
  100. Fiore, N.T.; Austin, P.J. Are the emergence of affective disturbances in neuropathic pain states contingent on supraspinal neuroinflammation? Brain Behav. Immun. 2016, 56, 397–411. [Google Scholar] [CrossRef]
  101. Tsuda, M.; Koga, K.; Chen, T.; Zhuo, M. Neuronal and microglial mechanisms for neuropathic pain in the spinal dorsal horn and anterior cingulate cortex. J. Neurochem. 2017, 141, 486–498. [Google Scholar] [CrossRef] [Green Version]
  102. Zhang, F.; Vadakkan, K.I.; Kim, S.S.; Wu, L.J.; Shang, Y.; Zhuo, M. Selective activation of microglia in spinal cord but not higher cortical regions following nerve injury in adult mouse. Mol. Pain 2008, 4, 15. [Google Scholar] [CrossRef] [Green Version]
  103. Galan-Arriero, I.; Avila-Martin, G.; Ferrer-Donato, A.; Gomez-Soriano, J.; Bravo-Esteban, E.; Taylor, J. Oral administration of the p38alpha MAPK inhibitor, UR13870, inhibits affective pain behavior after spinal cord injury. Pain 2014, 155, 2188–2198. [Google Scholar] [CrossRef]
  104. Guida, F.; Luongo, L.; Marmo, F.; Romano, R.; Iannotta, M.; Napolitano, F.; Belardo, C.; Marabese, I.; D’Aniello, A.; De Gregorio, D.; et al. Palmitoylethanolamide reduces pain-related behaviors and restores glutamatergic synapses homeostasis in the medial prefrontal cortex of neuropathic mice. Mol. Brain 2015, 8, 47. [Google Scholar] [CrossRef]
  105. Gui, W.S.; Wei, X.; Mai, C.L.; Murugan, M.; Wu, L.J.; Xin, W.J.; Zhou, L.J.; Liu, X.G. Interleukin-1beta overproduction is a common cause for neuropathic pain, memory deficit, and depression following peripheral nerve injury in rodents. Mol. Pain 2016, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Li, Z.; Wei, H.; Piirainen, S.; Chen, Z.; Kalso, E.; Pertovaara, A.; Tian, L. Spinal versus brain microglial and macrophage activation traits determine the differential neuroinflammatory responses and analgesic effect of minocycline in chronic neuropathic pain. Brain Behav. Immun. 2016, 58, 107–117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Miyamoto, K.; Kume, K.; Ohsawa, M. Role of microglia in mechanical allodynia in the anterior cingulate cortex. J. Pharmacol. Sci. 2017, 134, 158–165. [Google Scholar] [CrossRef] [PubMed]
  108. Xu, N.; Tang, X.H.; Pan, W.; Xie, Z.M.; Zhang, G.F.; Ji, M.H.; Yang, J.J.; Zhou, M.T.; Zhou, Z.Q. Spared Nerve Injury Increases the Expression of Microglia M1 Markers in the Prefrontal Cortex of Rats and Provokes Depression-Like Behaviors. Front. Neurosci. 2017, 11, 209. [Google Scholar] [CrossRef] [PubMed]
  109. Barcelon, E.E.; Cho, W.H.; Jun, S.B.; Lee, S.J. Brain Microglial Activation in Chronic Pain-Associated Affective Disorder. Front. Neurosci. 2019, 13, 213. [Google Scholar] [CrossRef] [Green Version]
  110. Diaz, A.F.; Polo, S.; Gallardo, N.; Leanez, S.; Pol, O. Analgesic and Antidepressant Effects of Oltipraz on Neuropathic Pain in Mice by Modulating Microglial Activation. J. Clin. Med. 2019, 8, 890. [Google Scholar] [CrossRef] [Green Version]
  111. Fiore, N.T.; Austin, P.J. Peripheral Nerve Injury Triggers Neuroinflammation in the Medial Prefrontal Cortex and Ventral Hippocampus in a Subgroup of Rats with Coincident Affective Behavioural Changes. Neuroscience 2019, 416, 147–167. [Google Scholar] [CrossRef]
  112. Georgieva, M.; Wei, Y.; Dumitrascuta, M.; Pertwee, R.; Finnerup, N.B.; Huang, W. Fatty acid suppression of glial activation prevents central neuropathic pain after spinal cord injury. Pain 2019, 160, 2724–2742. [Google Scholar] [CrossRef]
  113. Marrone, M.C.; Morabito, A.; Giustizieri, M.; Chiurchiu, V.; Leuti, A.; Mattioli, M.; Marinelli, S.; Riganti, L.; Lombardi, M.; Murana, E.; et al. TRPV1 channels are critical brain inflammation detectors and neuropathic pain biomarkers in mice. Nat. Commun. 2017, 8, 15292. [Google Scholar] [CrossRef]
  114. Narita, M.; Kuzumaki, N.; Narita, M.; Kaneko, C.; Hareyama, N.; Miyatake, M.; Shindo, K.; Miyoshi, K.; Nakajima, M.; Nagumo, Y.; et al. Chronic pain-induced emotional dysfunction is associated with astrogliosis due to cortical delta-opioid receptor dysfunction. J. Neurochem. 2006, 97, 1369–1378. [Google Scholar] [CrossRef]
  115. Kuzumaki, N.; Narita, M.; Narita, M.; Hareyama, N.; Niikura, K.; Nagumo, Y.; Nozaki, H.; Amano, T.; Suzuki, T. Chronic pain-induced astrocyte activation in the cingulate cortex with no change in neural or glial differentiation from neural stem cells in mice. Neurosci. Lett. 2007, 415, 22–27. [Google Scholar] [CrossRef] [PubMed]
  116. Lu, Y.; Zhu, L.; Gao, Y.J. Pain-related aversion induces astrocytic reaction and proinflammatory cytokine expression in the anterior cingulate cortex in rats. Brain Res. Bull. 2011, 84, 178–182. [Google Scholar] [CrossRef] [PubMed]
  117. Chen, F.L.; Dong, Y.L.; Zhang, Z.J.; Cao, D.L.; Xu, J.; Hui, J.; Zhu, L.; Gao, Y.J. Activation of astrocytes in the anterior cingulate cortex contributes to the affective component of pain in an inflammatory pain model. Brain Res. Bull. 2012, 87, 60–66. [Google Scholar] [CrossRef]
  118. Ikeda, H.; Mochizuki, K.; Murase, K. Astrocytes are involved in long-term facilitation of neuronal excitation in the anterior cingulate cortex of mice with inflammatory pain. Pain 2013, 154, 2836–2843. [Google Scholar] [CrossRef] [PubMed]
  119. Yamashita, A.; Hamada, A.; Suhara, Y.; Kawabe, R.; Yanase, M.; Kuzumaki, N.; Narita, M.; Matsui, R.; Okano, H.; Narita, M. Astrocytic activation in the anterior cingulate cortex is critical for sleep disorder under neuropathic pain. Synapse 2014, 68, 235–247. [Google Scholar] [CrossRef] [PubMed]
  120. Masocha, W. Astrocyte activation in the anterior cingulate cortex and altered glutamatergic gene expression during paclitaxel-induced neuropathic pain in mice. PeerJ 2015, 3, e1350. [Google Scholar] [CrossRef] [Green Version]
  121. Alotaibi, G.; Rahman, S. Effects of glial glutamate transporter activator in formalin-induced pain behaviour in mice. Eur. J. Pain 2019, 23, 765–783. [Google Scholar] [CrossRef]
  122. Hestrin, S.; Armstrong, W.E. Morphology and physiology of cortical neurons in layer I. J. Neurosci. 1996, 16, 5290–5300. [Google Scholar] [CrossRef]
  123. Zhou, F.M.; Hablitz, J.J. Morphological properties of intracellularly labeled layer I neurons in rat neocortex. J. Comp. Neurol. 1996, 376, 198–213. [Google Scholar] [CrossRef]
  124. Poorthuis, R.B.; Muhammad, K.; Wang, M.; Verhoog, M.B.; Junek, S.; Wrana, A.; Mansvelder, H.D.; Letzkus, J.J. Rapid Neuromodulation of Layer 1 Interneurons in Human Neocortex. Cell Rep. 2018, 23, 951–958. [Google Scholar] [CrossRef] [Green Version]
  125. Schuman, B.; Machold, R.P.; Hashikawa, Y.; Fuzik, J.; Fishell, G.J.; Rudy, B. Four Unique Interneuron Populations Reside in Neocortical Layer 1. J. Neurosci. 2019, 39, 125–139. [Google Scholar] [CrossRef] [Green Version]
  126. Dembrow, N.; Johnston, D. Subcircuit-specific neuromodulation in the prefrontal cortex. Front. Neural. Circuits 2014, 8, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Tremblay, R.; Lee, S.; Rudy, B. GABAergic Interneurons in the Neocortex: From Cellular Properties to Circuits. Neuron 2016, 91, 260–292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Cruikshank, S.J.; Lewis, T.J.; Connors, B.W. Synaptic basis for intense thalamocortical activation of feedforward inhibitory cells in neocortex. Nat. Neurosci. 2007, 10, 462–468. [Google Scholar] [CrossRef] [PubMed]
  129. McMahon, S.B.; Koltzenburg, M.; Tracey, I.; Turk, D. Wall and Melzack’s Textbook of Pain; Saunders: Philadelphia, PA, USA, 2013. [Google Scholar]
  130. Kim, S.K.; Kato, G.; Ishikawa, T.; Nabekura, J. Phase-specific plasticity of synaptic structures in the somatosensory cortex of living mice during neuropathic pain. Mol. Pain 2011, 7, 87. [Google Scholar] [CrossRef] [Green Version]
  131. Cichon, J.; Blanck, T.J.J.; Gan, W.B.; Yang, G. Activation of cortical somatostatin interneurons prevents the development of neuropathic pain. Nat. Neurosci. 2017, 20, 1122–1132. [Google Scholar] [CrossRef] [Green Version]
  132. Liu, R.J.; Aghajanian, G.K. Stress blunts serotonin- and hypocretin-evoked EPSCs in prefrontal cortex: Role of corticosterone-mediated apical dendritic atrophy. Proc. Natl. Acad. Sci. USA 2008, 105, 359–364. [Google Scholar] [CrossRef] [Green Version]
  133. Negron-Oyarzo, I.; Dagnino-Subiabre, A.; Munoz Carvajal, P. Synaptic Impairment in Layer 1 of the Prefrontal Cortex Induced by Repeated Stress During Adolescence is Reversed in Adulthood. Front. Cell Neurosci. 2015, 9, 442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Gao, S.H.; Wen, H.Z.; Shen, L.L.; Zhao, Y.D.; Ruan, H.Z. Activation of mGluR1 contributes to neuronal hyperexcitability in the rat anterior cingulate cortex via inhibition of HCN channels. Neuropharmacology 2016, 105, 361–377. [Google Scholar] [CrossRef]
  135. Van Aerde, K.I.; Feldmeyer, D. Morphological and physiological characterization of pyramidal neuron subtypes in rat medial prefrontal cortex. Cereb. Cortex 2015, 25, 788–805. [Google Scholar] [CrossRef] [Green Version]
  136. Yuste, R. Dendritic spines and distributed circuits. Neuron 2011, 71, 772–781. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Gabbott, P.L.; Warner, T.A.; Jays, P.R.; Salway, P.; Busby, S.J. Prefrontal cortex in the rat: Projections to subcortical autonomic, motor, and limbic centers. J. Comp. Neurol. 2005, 492, 145–177. [Google Scholar] [CrossRef] [PubMed]
  138. Hirai, Y.; Morishima, M.; Karube, F.; Kawaguchi, Y. Specialized cortical subnetworks differentially connect frontal cortex to parahippocampal areas. J. Neurosci. 2012, 32, 1898–1913. [Google Scholar] [CrossRef] [PubMed]
  139. Quiquempoix, M.; Fayad, S.L.; Boutourlinsky, K.; Leresche, N.; Lambert, R.C.; Bessaih, T. Layer 2/3 Pyramidal Neurons Control the Gain of Cortical Output. Cell Rep. 2018, 24, 2799–2807 e2794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Cordeiro Matos, S.; Zamfir, M.; Longo, G.; Ribeiro-da-Silva, A.; Seguela, P. Noradrenergic fiber sprouting and altered transduction in neuropathic prefrontal cortex. Brain Struct. Funct. 2018, 223, 1149–1164. [Google Scholar] [CrossRef] [PubMed]
  141. Toyoda, H.; Zhao, M.G.; Ulzhofer, B.; Wu, L.J.; Xu, H.; Seeburg, P.H.; Sprengel, R.; Kuner, R.; Zhuo, M. Roles of the AMPA receptor subunit GluA1 but not GluA2 in synaptic potentiation and activation of ERK in the anterior cingulate cortex. Mol. Pain 2009, 5, 46. [Google Scholar] [CrossRef] [Green Version]
  142. Li, X.Y.; Wang, N.; Wang, Y.J.; Zuo, Z.X.; Koga, K.; Luo, F.; Zhuo, M. Long-term temporal imprecision of information coding in the anterior cingulate cortex of mice with peripheral inflammation or nerve injury. J. Neurosci. 2014, 34, 10675–10687. [Google Scholar] [CrossRef] [Green Version]
  143. Hubbard, C.S.; Khan, S.A.; Xu, S.; Cha, M.; Masri, R.; Seminowicz, D.A. Behavioral, metabolic and functional brain changes in a rat model of chronic neuropathic pain: A longitudinal MRI study. Neuroimage 2015, 107, 333–344. [Google Scholar] [CrossRef]
  144. Wu, X.B.; He, L.N.; Jiang, B.C.; Wang, X.; Lu, Y.; Gao, Y.J. Increased CXCL13 and CXCR5 in Anterior Cingulate Cortex Contributes to Neuropathic Pain-Related Conditioned Place Aversion. Neurosci. Bull. 2019, 35, 613–623. [Google Scholar] [CrossRef]
  145. Tan, W.; Yao, W.L.; Zhang, B.; Hu, R.; Wan, L.; Zhang, C.H.; Zhu, C. Neuronal loss in anterior cingulate cortex in spared nerve injury model of neuropathic pain. Neurochem. Int. 2018, 118, 127–133. [Google Scholar] [CrossRef]
  146. Singer, T.; Seymour, B.; O’Doherty, J.; Kaube, H.; Dolan, R.J.; Frith, C.D. Empathy for pain involves the affective but not sensory components of pain. Science 2004, 303, 1157–1162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Koga, K.; Shimoyama, S.; Yamada, A.; Furukawa, T.; Nikaido, Y.; Furue, H.; Nakamura, K.; Ueno, S. Chronic inflammatory pain induced GABAergic synaptic plasticity in the adult mouse anterior cingulate cortex. Mol. Pain 2018, 14, 1744806918783478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Chiou, C.S.; Huang, C.C.; Liang, Y.C.; Tsai, Y.C.; Hsu, K.S. Impairment of long-term depression in the anterior cingulate cortex of mice with bone cancer pain. Pain 2012, 153, 2097–2108. [Google Scholar] [CrossRef] [PubMed]
  149. Kang, S.J.; Liu, M.G.; Chen, T.; Ko, H.G.; Baek, G.C.; Lee, H.R.; Lee, K.; Collingridge, G.L.; Kaang, B.K.; Zhuo, M. Plasticity of metabotropic glutamate receptor-dependent long-term depression in the anterior cingulate cortex after amputation. J. Neurosci. 2012, 32, 11318–11329. [Google Scholar] [CrossRef] [Green Version]
  150. Ning, L.; Ma, L.Q.; Wang, Z.R.; Wang, Y.W. Chronic constriction injury induced long-term changes in spontaneous membrane-potential oscillations in anterior cingulate cortical neurons in vivo. Pain Physician 2013, 16, E577–E589. [Google Scholar]
  151. Cruikshank, S.J.; Ahmed, O.J.; Stevens, T.R.; Patrick, S.L.; Gonzalez, A.N.; Elmaleh, M.; Connors, B.W. Thalamic control of layer 1 circuits in prefrontal cortex. J. Neurosci. 2012, 32, 17813–17823. [Google Scholar] [CrossRef] [Green Version]
  152. Dembrow, N.C.; Chitwood, R.A.; Johnston, D. Projection-specific neuromodulation of medial prefrontal cortex neurons. J. Neurosci. 2010, 30, 16922–16937. [Google Scholar] [CrossRef] [Green Version]
  153. Kelly, C.J.; Huang, M.; Meltzer, H.; Martina, M. Reduced Glutamatergic Currents and Dendritic Branching of Layer 5 Pyramidal Cells Contribute to Medial Prefrontal Cortex Deactivation in a Rat Model of Neuropathic Pain. Front. Cell Neurosci. 2016, 10, 133. [Google Scholar] [CrossRef] [Green Version]
  154. Gao, S.H.; Shen, L.L.; Wen, H.Z.; Zhao, Y.D.; Ruan, H.Z. Inhibition of Metabotropic Glutamate Receptor Subtype 1 Alters the Excitability of the Commissural Pyramidal Neuron in the Rat Anterior Cingulate Cortex after Chronic Constriction Injury to the Sciatic Nerve. Anesthesiology 2017, 127, 515–533. [Google Scholar] [CrossRef]
  155. Radzicki, D.; Pollema-Mays, S.L.; Sanz-Clemente, A.; Martina, M. Loss of M1 Receptor Dependent Cholinergic Excitation Contributes to mPFC Deactivation in Neuropathic Pain. J. Neurosci. 2017, 37, 2292–2304. [Google Scholar] [CrossRef] [Green Version]
  156. Meda, K.S.; Patel, T.; Braz, J.M.; Malik, R.; Turner, M.L.; Seifikar, H.; Basbaum, A.I.; Sohal, V.S. Microcircuit Mechanisms through which Mediodorsal Thalamic Input to Anterior Cingulate Cortex Exacerbates Pain-Related Aversion. Neuron 2019, 102, 944–959. [Google Scholar] [CrossRef] [PubMed]
  157. Santello, M.; Nevian, T. Dysfunction of cortical dendritic integration in neuropathic pain reversed by serotoninergic neuromodulation. Neuron 2015, 86, 233–246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Chen, T.; Wang, W.; Dong, Y.L.; Zhang, M.M.; Wang, J.; Koga, K.; Liao, Y.H.; Li, J.L.; Budisantoso, T.; Shigemoto, R.; et al. Postsynaptic insertion of AMPA receptor onto cortical pyramidal neurons in the anterior cingulate cortex after peripheral nerve injury. Mol. Brain 2014, 7, 76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Wu, X.B.; Liang, B.; Gao, Y.J. The increase of intrinsic excitability of layer V pyramidal cells in the prelimbic medial prefrontal cortex of adult mice after peripheral inflammation. Neurosci. Lett. 2016, 611, 40–45. [Google Scholar] [CrossRef] [PubMed]
  160. Cheriyan, J.; Sheets, P.L. Altered Excitability and Local Connectivity of mPFC-PAG Neurons in a Mouse Model of Neuropathic Pain. J. Neurosci. 2018, 38, 4829–4839. [Google Scholar] [CrossRef]
  161. Huang, J.; Gadotti, V.M.; Chen, L.; Souza, I.A.; Huang, S.; Wang, D.; Ramakrishnan, C.; Deisseroth, K.; Zhang, Z.; Zamponi, G.W. A neuronal circuit for activating descending modulation of neuropathic pain. Nat. Neurosci. 2019, 22, 1659–1668. [Google Scholar] [CrossRef]
  162. Kelly, C.J.; Martina, M. Circuit-selective properties of glutamatergic inputs to the rat prelimbic cortex and their alterations in neuropathic pain. Brain Struct. Funct. 2018, 223, 2627–2639. [Google Scholar] [CrossRef]
  163. Van Eden, C.G.; Uylings, H.B. Cytoarchitectonic development of the prefrontal cortex in the rat. J. Comp. Neurol. 1985, 241, 253–267. [Google Scholar] [CrossRef]
  164. Bailey, C.D.; De Biasi, M.; Fletcher, P.J.; Lambe, E.K. The nicotinic acetylcholine receptor alpha5 subunit plays a key role in attention circuitry and accuracy. J. Neurosci. 2010, 30, 9241–9252. [Google Scholar] [CrossRef] [Green Version]
  165. Sparks, D.W.; Tian, M.K.; Sargin, D.; Venkatesan, S.; Intson, K.; Lambe, E.K. Opposing Cholinergic and Serotonergic Modulation of Layer 6 in Prefrontal Cortex. Front. Neural. Circuits 2017, 11, 107. [Google Scholar] [CrossRef] [Green Version]
  166. Devoize, L.; Alvarez, P.; Monconduit, L.; Dallel, R. Representation of dynamic mechanical allodynia in the ventral medial prefrontal cortex of trigeminal neuropathic rats. Eur. J. Pain 2011, 15, 676–682. [Google Scholar] [CrossRef] [PubMed]
  167. Guillery, R.W.; Sherman, S.M. Thalamic relay functions and their role in corticocortical communication: Generalizations from the visual system. Neuron 2002, 33, 163–175. [Google Scholar] [CrossRef] [Green Version]
  168. Zikopoulos, B.; Barbas, H. Prefrontal projections to the thalamic reticular nucleus form a unique circuit for attentional mechanisms. J. Neurosci. 2006, 26, 7348–7361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Thomson, A.M. Neocortical layer, 6, a review. Front. Neuroanat. 2010, 4, 13. [Google Scholar] [CrossRef] [PubMed]
  170. Jurik, A.; Auffenberg, E.; Klein, S.; Deussing, J.M.; Schmid, R.M.; Wotjak, C.T.; Thoeringer, C.K. Roles of prefrontal cortex and paraventricular thalamus in affective and mechanical components of visceral nociception. Pain 2015, 156, 2479–2491. [Google Scholar] [CrossRef]
  171. Leite-Almeida, H.; Guimaraes, M.R.; Cerqueira, J.J.; Ribeiro-Costa, N.; Anjos-Martins, H.; Sousa, N.; Almeida, A. Asymmetric c-fos expression in the ventral orbital cortex is associated with impaired reversal learning in a right-sided neuropathy. Mol. Pain 2014, 10, 41. [Google Scholar] [CrossRef] [Green Version]
  172. Moriarty, O.; Gorman, C.L.; McGowan, F.; Ford, G.K.; Roche, M.; Thompson, K.; Dockery, P.; McGuire, B.E.; Finn, D.P. Impaired recognition memory and cognitive flexibility in the rat L5-L6 spinal nerve ligation model of neuropathic pain. Scand. J. Pain 2016, 10, 61–73. [Google Scholar] [CrossRef] [Green Version]
  173. Palazzo, E.; Luongo, L.; Guida, F.; Marabese, I.; Romano, R.; Iannotta, M.; Rossi, F.; D’Aniello, A.; Stella, L.; Marmo, F.; et al. D-Aspartate drinking solution alleviates pain and cognitive impairment in neuropathic mice. Amino. Acids 2016, 48, 1553–1567. [Google Scholar] [CrossRef]
  174. De Novellis, V.; Vita, D.; Gatta, L.; Luongo, L.; Bellini, G.; De Chiaro, M.; Marabese, I.; Siniscalco, D.; Boccella, S.; Piscitelli, F.; et al. The blockade of the transient receptor potential vanilloid type 1 and fatty acid amide hydrolase decreases symptoms and central sequelae in the medial prefrontal cortex of neuropathic rats. Mol. Pain 2011, 7, 7. [Google Scholar] [CrossRef] [Green Version]
  175. Cardoso-Cruz, H.; Sousa, M.; Vieira, J.B.; Lima, D.; Galhardo, V. Prefrontal cortex and mediodorsal thalamus reduced connectivity is associated with spatial working memory impairment in rats with inflammatory pain. Pain 2013, 154, 2397–2406. [Google Scholar] [CrossRef]
  176. Lee, H.; Im, J.; Won, H.; Nam, W.; Kim, Y.O.; Lee, S.W.; Lee, S.; Cho, I.H.; Kim, H.K.; Kwon, J.T.; et al. Effects of tianeptine on symptoms of fibromyalgia via BDNF signaling in a fibromyalgia animal model. Korean J. Physiol. Pharmacol. 2017, 21, 361–370. [Google Scholar] [CrossRef] [Green Version]
  177. Brederson, J.D.; Chu, K.L.; Xu, J.; Nikkel, A.L.; Markosyan, S.; Jarvis, M.F.; Edelmayer, R.; Bitner, R.S.; McGaraughty, S. Characterization and comparison of rat monosodium iodoacetate and medial meniscal tear models of osteoarthritic pain. J. Orthop. Res. 2018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Lin, H.C.; Yen, C.T. Differential Expression of Phosphorylated ERK and c-Fos of Limbic Cortices Activities in Response to Tactile Allodynia of Neuropathic Rats. Chin. J. Physiol. 2018, 61, 240–251. [Google Scholar] [CrossRef] [PubMed]
  179. Pan, W.; Zhang, G.F.; Li, H.H.; Ji, M.H.; Zhou, Z.Q.; Li, K.Y.; Yang, J.J. Ketamine differentially restores diverse alterations of neuroligins in brain regions in a rat model of neuropathic pain-induced depression. Neuroreport 2018, 29, 863–869. [Google Scholar] [CrossRef] [PubMed]
  180. Fang, X.; Yang, C.; Li, S.; Zhan, G.; Zhang, J.; Huang, N.; Du, X.; Xu, H.; Hashimoto, K.; Luo, A. Brain-derived neurotrophic factor-TrkB signaling in the medial prefrontal cortex plays a role in the anhedonia-like phenotype after spared nerve injury. Eur. Arch. Psychiatry Clin. Neurosci. 2020, 270, 195–205. [Google Scholar] [CrossRef] [Green Version]
  181. Vogt, B.A. Pain and emotion interactions in subregions of the cingulate gyrus. Nat. Rev. Neurosci. 2005, 6, 533–544. [Google Scholar] [CrossRef] [Green Version]
  182. Koga, K.; Descalzi, G.; Chen, T.; Ko, H.G.; Lu, J.; Li, S.; Son, J.; Kim, T.; Kwak, C.; Huganir, R.L.; et al. Coexistence of two forms of LTP in ACC provides a synaptic mechanism for the interactions between anxiety and chronic pain. Neuron 2015, 85, 377–389. [Google Scholar] [CrossRef] [Green Version]
  183. Bliss, T.V.; Collingridge, G.L.; Kaang, B.K.; Zhuo, M. Synaptic plasticity in the anterior cingulate cortex in acute and chronic pain. Nat. Rev. Neurosci. 2016, 17, 485–496. [Google Scholar] [CrossRef]
  184. Chen, T.; Taniguchi, W.; Chen, Q.Y.; Tozaki-Saitoh, H.; Song, Q.; Liu, R.H.; Koga, K.; Matsuda, T.; Kaito-Sugimura, Y.; Wang, J.; et al. Top-down descending facilitation of spinal sensory excitatory transmission from the anterior cingulate cortex. Nat. Commun. 2018, 9, 1886. [Google Scholar] [CrossRef] [Green Version]
  185. LaGraize, S.C.; Labuda, C.J.; Rutledge, M.A.; Jackson, R.L.; Fuchs, P.N. Differential effect of anterior cingulate cortex lesion on mechanical hypersensitivity and escape/avoidance behavior in an animal model of neuropathic pain. Exp. Neurol. 2004, 188, 139–148. [Google Scholar] [CrossRef]
  186. LaBuda, C.J.; Fuchs, P.N. Attenuation of negative pain affect produced by unilateral spinal nerve injury in the rat following anterior cingulate cortex activation. Neuroscience 2005, 136, 311–322. [Google Scholar] [CrossRef] [PubMed]
  187. Qu, C.; King, T.; Okun, A.; Lai, J.; Fields, H.L.; Porreca, F. Lesion of the rostral anterior cingulate cortex eliminates the aversiveness of spontaneous neuropathic pain following partial or complete axotomy. Pain 2011, 152, 1641–1648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Barthas, F.; Sellmeijer, J.; Hugel, S.; Waltisperger, E.; Barrot, M.; Yalcin, I. The anterior cingulate cortex is a critical hub for pain-induced depression. Biol. Psychiatry 2015, 77, 236–245. [Google Scholar] [CrossRef]
  189. Juarez-Salinas, D.L.; Braz, J.M.; Etlin, A.; Gee, S.; Sohal, V.; Basbaum, A.I. GABAergic cell transplants in the anterior cingulate cortex reduce neuropathic pain aversiveness. Brain 2019, 142, 2655–2669. [Google Scholar] [CrossRef]
  190. LaGraize, S.C.; Borzan, J.; Peng, Y.B.; Fuchs, P.N. Selective regulation of pain affect following activation of the opioid anterior cingulate cortex system. Exp. Neurol. 2006, 197, 22–30. [Google Scholar] [CrossRef] [PubMed]
  191. Shen, F.Y.; Chen, Z.Y.; Zhong, W.; Ma, L.Q.; Chen, C.; Yang, Z.J.; Xie, W.L.; Wang, Y.W. Alleviation of neuropathic pain by regulating T-type calcium channels in rat anterior cingulate cortex. Mol. Pain 2015, 11, 7. [Google Scholar] [CrossRef] [Green Version]
  192. Bannister, K.; Qu, C.; Navratilova, E.; Oyarzo, J.; Xie, J.Y.; King, T.; Dickenson, A.H.; Porreca, F. Multiple sites and actions of gabapentin-induced relief of ongoing experimental neuropathic pain. Pain 2017, 158, 2386–2395. [Google Scholar] [CrossRef]
  193. Gomtsian, L.; Bannister, K.; Eyde, N.; Robles, D.; Dickenson, A.H.; Porreca, F.; Navratilova, E. Morphine effects within the rodent anterior cingulate cortex and rostral ventromedial medulla reveal separable modulation of affective and sensory qualities of acute or chronic pain. Pain 2018, 159, 2512–2521. [Google Scholar] [CrossRef]
  194. Ren, L.Y.; Lu, Z.M.; Liu, M.G.; Yu, Y.Q.; Li, Z.; Shang, G.W.; Chen, J. Distinct roles of the anterior cingulate cortex in spinal and supraspinal bee venom-induced pain behaviors. Neuroscience 2008, 153, 268–278. [Google Scholar] [CrossRef]
  195. Gu, L.; Uhelski, M.L.; Anand, S.; Romero-Ortega, M.; Kim, Y.T.; Fuchs, P.N.; Mohanty, S.K. Pain inhibition by optogenetic activation of specific anterior cingulate cortical neurons. PLoS ONE 2015, 10, e0117746. [Google Scholar] [CrossRef]
  196. Kang, S.J.; Kwak, C.; Lee, J.; Sim, S.E.; Shim, J.; Choi, T.; Collingridge, G.L.; Zhuo, M.; Kaang, B.K. Bidirectional modulation of hyperalgesia via the specific control of excitatory and inhibitory neuronal activity in the ACC. Mol. Brain 2015, 8, 81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Sellmeijer, J.; Mathis, V.; Hugel, S.; Li, X.H.; Song, Q.; Chen, Q.Y.; Barthas, F.; Lutz, P.E.; Karatas, M.; Luthi, A.; et al. Hyperactivity of Anterior Cingulate Cortex Areas 24a/24b Drives Chronic Pain-Induced Anxiodepressive-like Consequences. J. Neurosci. 2018, 38, 3102–3115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Kang, S.J.; Kim, S.; Lee, J.; Kwak, C.; Lee, K.; Zhuo, M.; Kaang, B.K. Inhibition of anterior cingulate cortex excitatory neuronal activity induces conditioned place preference in a mouse model of chronic inflammatory pain. Korean J. Physiol. Pharmacol. 2017, 21, 487–493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Vogt, B.A.; Paxinos, G. Cytoarchitecture of mouse and rat cingulate cortex with human homologies. Brain Struct. Funct. 2014, 219, 185–192. [Google Scholar] [CrossRef] [PubMed]
  200. Chen, T.; Koga, K.; Descalzi, G.; Qiu, S.; Wang, J.; Zhang, L.S.; Zhang, Z.J.; He, X.B.; Qin, X.; Xu, F.Q.; et al. Postsynaptic potentiation of corticospinal projecting neurons in the anterior cingulate cortex after nerve injury. Mol. Pain 2014, 10, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Chiou, C.S.; Chen, C.C.; Tsai, T.C.; Huang, C.C.; Chou, D.; Hsu, K.S. Alleviating Bone Cancer-induced Mechanical Hypersensitivity by Inhibiting Neuronal Activity in the Anterior Cingulate Cortex. Anesthesiology 2016, 125, 779–792. [Google Scholar] [CrossRef]
  202. Moon, H.C.; Heo, W.I.; Kim, Y.J.; Lee, D.; Won, S.Y.; Kim, H.R.; Ha, S.M.; Lee, Y.J.; Park, Y.S. Optical inactivation of the anterior cingulate cortex modulate descending pain pathway in a rat model of trigeminal neuropathic pain created via chronic constriction injury of the infraorbital nerve. J. Pain Res. 2017, 10, 2355–2364. [Google Scholar] [CrossRef] [Green Version]
  203. Chen, Q.Y.; Chen, T.; Zhou, L.J.; Liu, X.G.; Zhuo, M. Heterosynaptic long-term potentiation from the anterior cingulate cortex to spinal cord in adult rats. Mol. Pain 2018, 14, 1744806918798406. [Google Scholar] [CrossRef]
  204. Calejesan, A.A.; Kim, S.J.; Zhuo, M. Descending facilitatory modulation of a behavioral nociceptive response by stimulation in the adult rat anterior cingulate cortex. Eur. J. Pain 2000, 4, 83–96. [Google Scholar] [CrossRef]
  205. Traub, R.J.; Silva, E.; Gebhart, G.F.; Solodkin, A. Noxious colorectal distention induced-c-Fos protein in limbic brain structures in the rat. Neurosci. Lett. 1996, 215, 165–168. [Google Scholar] [CrossRef]
  206. Gibney, S.M.; Gosselin, R.D.; Dinan, T.G.; Cryan, J.F. Colorectal distension-induced prefrontal cortex activation in the Wistar-Kyoto rat: Implications for irritable bowel syndrome. Neuroscience 2010, 165, 675–683. [Google Scholar] [CrossRef]
  207. Wang, Z.; Bradesi, S.; Charles, J.R.; Pang, R.D.; Maarek, J.M.; Mayer, E.A.; Holschneider, D.P. Functional brain activation during retrieval of visceral pain-conditioned passive avoidance in the rat. Pain 2011, 152, 2746–2756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Felice, V.D.; Gibney, S.M.; Gosselin, R.D.; Dinan, T.G.; O’Mahony, S.M.; Cryan, J.F. Differential activation of the prefrontal cortex and amygdala following psychological stress and colorectal distension in the maternally separated rat. Neuroscience 2014, 267, 252–262. [Google Scholar] [CrossRef] [PubMed]
  209. Wang, H.; Tao, X.; Huang, S.T.; Wu, L.; Tang, H.L.; Song, Y.; Zhang, G.; Zhang, Y.M. Chronic Stress Is Associated with Pain Precipitation and Elevation in DeltaFosb Expression. Front. Pharmacol. 2016, 7, 138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Kiritoshi, T.; Neugebauer, V. Group II mGluRs modulate baseline and arthritis pain-related synaptic transmission in the rat medial prefrontal cortex. Neuropharmacology 2015, 95, 388–394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. Luongo, L.; de Novellis, V.; Gatta, L.; Palazzo, E.; Vita, D.; Guida, F.; Giordano, C.; Siniscalco, D.; Marabese, I.; De Chiaro, M.; et al. Role of metabotropic glutamate receptor 1 in the basolateral amygdala-driven prefrontal cortical deactivation in inflammatory pain in the rat. Neuropharmacology 2013, 66, 317–329. [Google Scholar] [CrossRef]
  212. Wang, G.Q.; Cen, C.; Li, C.; Cao, S.; Wang, N.; Zhou, Z.; Liu, X.M.; Xu, Y.; Tian, N.X.; Zhang, Y.; et al. Deactivation of excitatory neurons in the prelimbic cortex via Cdk5 promotes pain sensation and anxiety. Nat. Commun. 2015, 6, 7660. [Google Scholar] [CrossRef]
  213. Li, A.L.; Yang, X.; Chiao, J.C.; Peng, Y.B. Reduced local field potential power in the medial prefrontal cortex by noxious stimuli. Brain Res. Bull. 2016, 127, 92–99. [Google Scholar] [CrossRef]
  214. Dale, J.; Zhou, H.; Zhang, Q.; Martinez, E.; Hu, S.; Liu, K.; Urien, L.; Chen, Z.; Wang, J. Scaling Up Cortical Control Inhibits Pain. Cell Rep. 2018, 23, 1301–1313. [Google Scholar] [CrossRef]
  215. Fu, B.; Wen, S.N.; Wang, B.; Wang, K.; Zhang, J.Y.; Liu, S.J. Acute and chronic pain affects local field potential of the medial prefrontal cortex in different band neural oscillations. Mol. Pain 2018, 14, 1744806918785686. [Google Scholar] [CrossRef] [Green Version]
  216. Ririe, D.G.; Boada, M.D.; MacGregor, M.K.; Martin, S.J.; Strassburg, T.J.; Kim, S.A.; Eisenach, J.C.; Martin, T.J. Incisional Nociceptive Input Impairs Attention-related Behavior and Is Associated with Reduced Neuronal Activity in the Prefrontal Cortex in Rats. Anesthesiology 2018, 129, 778–790. [Google Scholar] [CrossRef] [PubMed]
  217. Cardoso-Cruz, H.; Paiva, P.; Monteiro, C.; Galhardo, V. Selective optogenetic inhibition of medial prefrontal glutamatergic neurons reverses working memory deficits induced by neuropathic pain. Pain 2019, 160, 805–823. [Google Scholar] [CrossRef] [PubMed]
  218. Sang, K.; Bao, C.; Xin, Y.; Hu, S.; Gao, X.; Wang, Y.; Bodner, M.; Zhou, Y.D.; Dong, X.W. Plastic change of prefrontal cortex mediates anxiety-like behaviors associated with chronic pain in neuropathic rats. Mol. Pain 2018, 14, 1744806918783931. [Google Scholar] [CrossRef]
  219. Palazzo, E.; Romano, R.; Luongo, L.; Boccella, S.; De Gregorio, D.; Giordano, M.E.; Rossi, F.; Marabese, I.; Scafuro, M.A.; de Novellis, V.; et al. MMPIP, an mGluR7-selective negative allosteric modulator, alleviates pain and normalizes affective and cognitive behavior in neuropathic mice. Pain 2015, 156, 1060–1073. [Google Scholar] [CrossRef] [PubMed]
  220. Chung, G.; Kim, C.Y.; Yun, Y.C.; Yoon, S.H.; Kim, M.H.; Kim, Y.K.; Kim, S.J. Upregulation of prefrontal metabotropic glutamate receptor 5 mediates neuropathic pain and negative mood symptoms after spinal nerve injury in rats. Sci. Rep. 2017, 7, 9743. [Google Scholar] [CrossRef]
  221. Zhou, H.; Martinez, E.; Lin, H.H.; Yang, R.; Dale, J.A.; Liu, K.; Huang, D.; Wang, J. Inhibition of the Prefrontal Projection to the Nucleus Accumbens Enhances Pain Sensitivity and Affect. Front. Cell Neurosci. 2018, 12, 240. [Google Scholar] [CrossRef] [PubMed]
  222. Zhou, H.; Zhang, Q.; Martinez, E.; Dale, J.; Robinson, E.; Huang, D.; Wang, J. A novel neuromodulation strategy to enhance the prefrontal control to treat pain. Mol. Pain 2019, 15, 1744806919845739. [Google Scholar] [CrossRef] [Green Version]
  223. David-Pereira, A.; Puga, S.; Goncalves, S.; Amorim, D.; Silva, C.; Pertovaara, A.; Almeida, A.; Pinto-Ribeiro, F. Metabotropic glutamate 5 receptor in the infralimbic cortex contributes to descending pain facilitation in healthy and arthritic animals. Neuroscience 2016, 312, 108–119. [Google Scholar] [CrossRef] [Green Version]
  224. David-Pereira, A.; Sagalajev, B.; Wei, H.; Almeida, A.; Pertovaara, A.; Pinto-Ribeiro, F. The medullary dorsal reticular nucleus as a relay for descending pronociception induced by the mGluR5 in the rat infralimbic cortex. Neuroscience 2017, 349, 341–354. [Google Scholar] [CrossRef]
  225. Yue, L.; Ma, L.Y.; Cui, S.; Liu, F.Y.; Yi, M.; Wan, Y. Brain-derived neurotrophic factor in the infralimbic cortex alleviates inflammatory pain. Neurosci. Lett. 2017, 655, 7–13. [Google Scholar] [CrossRef]
  226. Reiner, A.; Levitz, J. Glutamatergic Signaling in the Central Nervous System: Ionotropic and Metabotropic Receptors in Concert. Neuron 2018, 98, 1080–1098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Baer, M.R.; Gao, X.Z.; Preisler, H.D. Sustained viability of acute myelogenous leukemia cells in liquid suspension culture correlates with occurrence of differentiation. Leuk. Res. 1988, 12, 647–650. [Google Scholar] [CrossRef]
  228. Millecamps, M.; Etienne, M.; Jourdan, D.; Eschalier, A.; Ardid, D. Decrease in non-selective, non-sustained attention induced by a chronic visceral inflammatory state as a new pain evaluation in rats. Pain 2004, 109, 214–224. [Google Scholar] [CrossRef] [PubMed]
  229. Sun, Y.; Liu, K.; Martinez, E.; Dale, J.; Huang, D.; Wang, J. AMPAkines and morphine provide complementary analgesia. Behav. Brain Res. 2017, 334, 1–5. [Google Scholar] [CrossRef]
  230. Giordano, C.; Cristino, L.; Luongo, L.; Siniscalco, D.; Petrosino, S.; Piscitelli, F.; Marabese, I.; Gatta, L.; Rossi, F.; Imperatore, R.; et al. TRPV1-dependent and -independent alterations in the limbic cortex of neuropathic mice: Impact on glial caspases and pain perception. Cereb. Cortex 2012, 22, 2495–2518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Wei, F.; Qiu, C.S.; Kim, S.J.; Muglia, L.; Maas, J.W.; Pineda, V.V.; Xu, H.M.; Chen, Z.F.; Storm, D.R.; Muglia, L.J.; et al. Genetic elimination of behavioral sensitization in mice lacking calmodulin-stimulated adenylyl cyclases. Neuron 2002, 36, 713–726. [Google Scholar] [CrossRef] [Green Version]
  232. Ding, M.; Shen, W.; Hu, Y. The Role of miR-539 in the Anterior Cingulate Cortex in Chronic Neuropathic Pain. Pain Med. 2017, 18, 2433–2442. [Google Scholar] [CrossRef]
  233. Li, X.Y.; Ko, H.G.; Chen, T.; Descalzi, G.; Koga, K.; Wang, H.; Kim, S.S.; Shang, Y.; Kwak, C.; Park, S.W.; et al. Alleviating neuropathic pain hypersensitivity by inhibiting PKMzeta in the anterior cingulate cortex. Science 2010, 330, 1400–1404. [Google Scholar] [CrossRef]
  234. Ko, H.G.; Ye, S.; Han, D.H.; Park, P.; Lim, C.S.; Lee, K.; Zhuo, M.; Kaang, B.K. Transcription-independent expression of PKMzeta in the anterior cingulate cortex contributes to chronically maintained neuropathic pain. Mol. Pain 2018, 14, 1744806918783943. [Google Scholar] [CrossRef] [Green Version]
  235. Du, J.; Fang, J.; Wen, C.; Shao, X.; Liang, Y.; Fang, J. The Effect of Electroacupuncture on PKMzeta in the ACC in Regulating Anxiety-Like Behaviors in Rats Experiencing Chronic Inflammatory Pain. Neural. Plast. 2017, 2017, 3728752. [Google Scholar] [CrossRef] [Green Version]
  236. Millecamps, M.; Centeno, M.V.; Berra, H.H.; Rudick, C.N.; Lavarello, S.; Tkatch, T.; Apkarian, A.V. D-cycloserine reduces neuropathic pain behavior through limbic NMDA-mediated circuitry. Pain 2007, 132, 108–123. [Google Scholar] [CrossRef] [Green Version]
  237. De Freitas, R.L.; Salgado-Rohner, C.J.; Biagioni, A.F.; Medeiros, P.; Hallak, J.E.; Crippa, J.A.; Coimbra, N.C. NMDA and AMPA/kainate glutamatergic receptors in the prelimbic medial prefrontal cortex modulate the elaborated defensive behavior and innate fear-induced antinociception elicited by GABAA receptor blockade in the medial hypothalamus. Cereb. Cortex 2014, 24, 1518–1528. [Google Scholar] [CrossRef] [Green Version]
  238. Ghoreishi-Haack, N.; Priebe, J.M.; Aguado, J.D.; Colechio, E.M.; Burgdorf, J.S.; Bowers, M.S.; Cearley, C.N.; Khan, M.A.; Moskal, J.R. NYX-2925 Is a Novel N-Methyl-d-Aspartate Receptor Modulator that Induces Rapid and Long-Lasting Analgesia in Rat Models of Neuropathic Pain. J. Pharmacol. Exp. Ther. 2018, 366, 485–497. [Google Scholar] [CrossRef] [Green Version]
  239. Wu, L.J.; Toyoda, H.; Zhao, M.G.; Lee, Y.S.; Tang, J.; Ko, S.W.; Jia, Y.H.; Shum, F.W.; Zerbinatti, C.V.; Bu, G.; et al. Upregulation of forebrain NMDA NR2B receptors contributes to behavioral sensitization after inflammation. J. Neurosci. 2005, 25, 11107–11116. [Google Scholar] [CrossRef] [PubMed]
  240. Zhao, M.G.; Ko, S.W.; Wu, L.J.; Toyoda, H.; Xu, H.; Quan, J.; Li, J.; Jia, Y.; Ren, M.; Xu, Z.C.; et al. Enhanced presynaptic neurotransmitter release in the anterior cingulate cortex of mice with chronic pain. J. Neurosci. 2006, 26, 8923–8930. [Google Scholar] [CrossRef] [PubMed]
  241. Wu, L.J.; Steenland, H.W.; Kim, S.S.; Isiegas, C.; Abel, T.; Kaang, B.K.; Zhuo, M. Enhancement of presynaptic glutamate release and persistent inflammatory pain by increasing neuronal cAMP in the anterior cingulate cortex. Mol. Pain 2008, 4, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Toyoda, H.; Zhao, M.G.; Zhuo, M. Enhanced quantal release of excitatory transmitter in anterior cingulate cortex of adult mice with chronic pain. Mol. Pain 2009, 5, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Bie, B.; Brown, D.L.; Naguib, M. Increased synaptic GluR1 subunits in the anterior cingulate cortex of rats with peripheral inflammation. Eur. J. Pharmacol. 2011, 653, 26–31. [Google Scholar] [CrossRef]
  244. Wang, H.; Xu, H.; Wu, L.J.; Kim, S.S.; Chen, T.; Koga, K.; Descalzi, G.; Gong, B.; Vadakkan, K.I.; Zhang, X.; et al. Identification of an adenylyl cyclase inhibitor for treating neuropathic and inflammatory pain. Sci. Transl. Med. 2011, 3, 65ra63. [Google Scholar] [CrossRef] [Green Version]
  245. Lei, L.G.; Sun, S.; Gao, Y.J.; Zhao, Z.Q.; Zhang, Y.Q. NMDA receptors in the anterior cingulate cortex mediate pain-related aversion. Exp. Neurol. 2004, 189, 413–421. [Google Scholar] [CrossRef]
  246. Guo, S.G.; Lv, X.H.; Guan, S.H.; Li, H.L.; Qiao, Y.; Feng, H.; Cong, L.; Wang, G.M. Silencing the NR2B gene in rat ACC neurons by lentivirus-delivered shRNA alleviates pain-related aversion. Int. J. Clin. Exp. Med. 2015, 8, 6725–6734. [Google Scholar]
  247. Zhang, Y.; Ji, F.; Wang, G.; He, D.; Yang, L.; Zhang, M. BDNF Activates mTOR to Upregulate NR2B Expression in the Rostral Anterior Cingulate Cortex Required for Inflammatory Pain-Related Aversion in Rats. Neurochem. Res. 2018, 43, 681–691. [Google Scholar] [CrossRef] [PubMed]
  248. Ko, S.W.; Vadakkan, K.I.; Ao, H.; Gallitano-Mendel, A.; Wei, F.; Milbrandt, J.; Zhuo, M. Selective contribution of Egr1 (zif/268) to persistent inflammatory pain. J. Pain 2005, 6, 12–20. [Google Scholar] [CrossRef] [PubMed]
  249. Cao, H.; Gao, Y.J.; Ren, W.H.; Li, T.T.; Duan, K.Z.; Cui, Y.H.; Cao, X.H.; Zhao, Z.Q.; Ji, R.R.; Zhang, Y.Q. Activation of extracellular signal-regulated kinase in the anterior cingulate cortex contributes to the induction and expression of affective pain. J. Neurosci. 2009, 29, 3307–3321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Cao, H.; Zang, K.K.; Han, M.; Zhao, Z.Q.; Wu, G.C.; Zhang, Y.Q. Inhibition of p38 mitogen-activated protein kinase activation in the rostral anterior cingulate cortex attenuates pain-related negative emotion in rats. Brain Res. Bull. 2014, 107, 79–88. [Google Scholar] [CrossRef]
  251. Kang, W.B.; Yang, Q.; Guo, Y.Y.; Wang, L.; Wang, D.S.; Cheng, Q.; Li, X.M.; Tang, J.; Zhao, J.N.; Liu, G.; et al. Analgesic effects of adenylyl cyclase inhibitor NB001 on bone cancer pain in a mouse model. Mol. Pain 2016, 12. [Google Scholar] [CrossRef] [Green Version]
  252. Wei, F.; Zhuo, M. Activation of Erk in the anterior cingulate cortex during the induction and expression of chronic pain. Mol. Pain 2008, 4, 28. [Google Scholar] [CrossRef] [Green Version]
  253. Yang, J.X.; Hua, L.; Li, Y.Q.; Jiang, Y.Y.; Han, D.; Liu, H.; Tang, Q.Q.; Yang, X.N.; Yin, C.; Hao, L.Y.; et al. Caveolin-1 in the anterior cingulate cortex modulates chronic neuropathic pain via regulation of NMDA receptor 2B subunit. J. Neurosci. 2015, 35, 36–52. [Google Scholar] [CrossRef] [Green Version]
  254. Thibault, K.; Lin, W.K.; Rancillac, A.; Fan, M.; Snollaerts, T.; Sordoillet, V.; Hamon, M.; Smith, G.M.; Lenkei, Z.; Pezet, S. BDNF-dependent plasticity induced by peripheral inflammation in the primary sensory and the cingulate cortex triggers cold allodynia and reveals a major role for endogenous BDNF as a tuner of the affective aspect of pain. J. Neurosci. 2014, 34, 14739–14751. [Google Scholar] [CrossRef] [Green Version]
  255. Lu, B.; Jiang, J.; Sun, J.; Xiao, C.; Meng, B.; Zheng, J.; Li, X.; Wang, R.; Wu, G.; Chen, J. Inhibition of mammalian target of rapamycin activation in the rostral anterior cingulate cortex attenuates pain-related aversion in rats. Behav. Brain Res. 2016, 310, 51–58. [Google Scholar] [CrossRef]
  256. Zhang, L.; Wang, G.; Ma, J.; Liu, C.; Liu, X.; Zhan, Y.; Zhang, M. Brain-derived neurotrophic factor (BDNF) in the rostral anterior cingulate cortex (rACC) contributes to neuropathic spontaneous pain-related aversion via NR2B receptors. Brain Res. Bull. 2016, 127, 56–65. [Google Scholar] [CrossRef] [PubMed]
  257. Um, S.W.; Kim, M.J.; Leem, J.W.; Bai, S.J.; Lee, B.H. Pain-Relieving Effects of mTOR Inhibitor in the Anterior Cingulate Cortex of Neuropathic Rats. Mol. Neurobiol. 2019, 56, 2482–2494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Xu, Y.; Wang, G.; Zou, X.; Yang, Z.; Wang, Q.; Feng, H.; Zhang, M. siRNA-mediated downregulation of GluN2B in the rostral anterior cingulate cortex attenuates mechanical allodynia and thermal hyperalgesia in a rat model of pain associated with bone cancer. Exp. Ther. Med. 2016, 11, 221–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Li, H.S.; Su, X.Y.; Song, X.L.; Qi, X.; Li, Y.; Wang, R.Q.; Maximyuk, O.; Krishtal, O.; Wang, T.; Fang, H.; et al. Protein Kinase C Lambda Mediates Acid-Sensing Ion Channel 1a-Dependent Cortical Synaptic Plasticity and Pain Hypersensitivity. J. Neurosci. 2019, 39, 5773–5793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Guo, B.; Wang, J.; Yao, H.; Ren, K.; Chen, J.; Yang, J.; Cai, G.; Liu, H.; Fan, Y.; Wang, W.; et al. Chronic Inflammatory Pain Impairs mGluR5-Mediated Depolarization-Induced Suppression of Excitation in the Anterior Cingulate Cortex. Cereb. Cortex 2018, 28, 2118–2130. [Google Scholar] [CrossRef]
  261. Sun, T.; Wang, J.; Li, X.; Li, Y.J.; Feng, D.; Shi, W.L.; Zhao, M.G.; Wang, J.B.; Wu, Y.M. Gastrodin relieved complete Freund’s adjuvant-induced spontaneous pain by inhibiting inflammatory response. Int. Immunopharmacol. 2016, 41, 66–73. [Google Scholar] [CrossRef]
  262. Li, S.; Han, J.; Wang, D.S.; Yang, Q.; Feng, B.; Kang, W.B.; Yang, L.; Liu, G.; Zhao, M.G. Sinomenine attenuates chronic inflammatory pain in mice. Metab. Brain Dis. 2017, 32, 211–219. [Google Scholar] [CrossRef]
  263. Fan, Y.F.; Guan, S.Y.; Luo, L.; Li, Y.J.; Yang, L.; Zhou, X.X.; Guo, G.D.; Zhao, M.G.; Yang, Q.; Liu, G. Tetrahydroxystilbene glucoside relieves the chronic inflammatory pain by inhibiting neuronal apoptosis, microglia activation, and GluN2B overexpression in anterior cingulate cortex. Mol. Pain 2018, 14, 1744806918814367. [Google Scholar] [CrossRef]
  264. Olsen, R.W.; Sieghart, W. International Union of Pharmacology. LXX. Subtypes of gamma-aminobutyric acid(A) receptors: Classification on the basis of subunit composition, pharmacology, and function. Update. Pharmacol. Rev. 2008, 60, 243–260. [Google Scholar] [CrossRef] [Green Version]
  265. Owens, D.F.; Kriegstein, A.R. Is there more to GABA than synaptic inhibition? Nat. Rev. Neurosci. 2002, 3, 715–727. [Google Scholar] [CrossRef]
  266. Masocha, W. Comprehensive analysis of the GABAergic system gene expression profile in the anterior cingulate cortex of mice with Paclitaxel-induced neuropathic pain. Gene. Expr. 2015, 16, 145–153. [Google Scholar] [CrossRef] [PubMed]
  267. Nashawi, H.; Masocha, W.; Edafiogho, I.O.; Kombian, S.B. Paclitaxel Causes Electrophysiological Changes in the Anterior Cingulate Cortex via Modulation of the gamma-Aminobutyric Acid-ergic System. Med. Princ. Pract. 2016, 25, 423–428. [Google Scholar] [CrossRef] [PubMed]
  268. Migita, K.; Matsuzaki, Y.; Koga, K.; Matsumoto, T.; Mishima, K.; Hara, S.; Honda, K. Involvement of GABAB receptor in the antihypersensitive effect in anterior cingulate cortex of partial sciatic nerve ligation model. J. Pharmacol. Sci. 2018, 137, 233–236. [Google Scholar] [CrossRef] [PubMed]
  269. Geng, K.W.; He, T.; Wang, R.R.; Li, C.L.; Luo, W.J.; Wu, F.F.; Wang, Y.; Li, Z.; Lu, Y.F.; Guan, S.M.; et al. Ethanol Increases Mechanical Pain Sensitivity in Rats via Activation of GABAA Receptors in Medial Prefrontal Cortex. Neurosci. Bull. 2016, 32, 433–444. [Google Scholar] [CrossRef] [Green Version]
  270. Wang, H.; Ren, W.H.; Zhang, Y.Q.; Zhao, Z.Q. GABAergic disinhibition facilitates polysynaptic excitatory transmission in rat anterior cingulate cortex. Biochem. Biophys. Res. Commun. 2005, 338, 1634–1639. [Google Scholar] [CrossRef]
  271. LaGraize, S.C.; Fuchs, P.N. GABAA but not GABAB receptors in the rostral anterior cingulate cortex selectively modulate pain-induced escape/avoidance behavior. Exp. Neurol. 2007, 204, 182–194. [Google Scholar] [CrossRef] [Green Version]
  272. Jiang, Z.C.; Pan, Q.; Zheng, C.; Deng, X.F.; Wang, J.Y.; Luo, F. Inactivation of the prelimbic rather than infralimbic cortex impairs acquisition and expression of formalin-induced conditioned place avoidance. Neurosci. Lett. 2014, 569, 89–93. [Google Scholar] [CrossRef] [Green Version]
  273. Seminowicz, D.A.; Laferriere, A.L.; Millecamps, M.; Yu, J.S.; Coderre, T.J.; Bushnell, M.C. MRI structural brain changes associated with sensory and emotional function in a rat model of long-term neuropathic pain. Neuroimage 2009, 47, 1007–1014. [Google Scholar] [CrossRef] [Green Version]
  274. Maarrawi, J.; Peyron, R.; Mertens, P.; Costes, N.; Magnin, M.; Sindou, M.; Laurent, B.; Garcia-Larrea, L. Differential brain opioid receptor availability in central and peripheral neuropathic pain. Pain 2007, 127, 183–194. [Google Scholar] [CrossRef]
  275. Sprenger, T.; Berthele, A.; Platzer, S.; Boecker, H.; Tolle, T.R. What to learn from in vivo opioidergic brain imaging? Eur. J. Pain 2005, 9, 117–121. [Google Scholar] [CrossRef]
  276. Zhao, M.; Wang, J.Y.; Jia, H.; Tang, J.S. Roles of different subtypes of opioid receptors in mediating the ventrolateral orbital cortex opioid-induced inhibition of mirror-neuropathic pain in the rat. Neuroscience 2007, 144, 1486–1494. [Google Scholar] [CrossRef] [PubMed]
  277. Palmisano, M.; Caputi, F.F.; Mercatelli, D.; Romualdi, P.; Candeletti, S. Dynorphinergic system alterations in the corticostriatal circuitry of neuropathic mice support its role in the negative affective component of pain. Genes Brain Behav. 2018, e12467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Condes-Lara, M. Different direct pathways of locus coeruleus to medial prefrontal cortex and centrolateral thalamic nucleus: Electrical stimulation effects on the evoked responses to nociceptive peripheral stimulation. Eur. J. Pain 1998, 2, 15–23. [Google Scholar] [CrossRef]
  279. Kandel, E.R.; Schwartz, J.H.; Jessel, T.; Siegelbaum, S.A.; Hudspeth, A.J. Principles of Neural Science; McGraw-Hill: New York, NY, USA, 2013. [Google Scholar]
  280. Chu, K.L.; Xu, J.; Frost, J.; Li, L.; Gomez, E.; Dart, M.J.; Jarvis, M.F.; Meyer, M.D.; McGaraughty, S. A selective alpha2 B adrenoceptor agonist (A-1262543) and duloxetine modulate nociceptive neurones in the medial prefrontal cortex, but not in the spinal cord of neuropathic rats. Eur. J. Pain 2015, 19, 649–660. [Google Scholar] [CrossRef] [PubMed]
  281. Alba-Delgado, C.; Llorca-Torralba, M.; Horrillo, I.; Ortega, J.E.; Mico, J.A.; Sanchez-Blazquez, P.; Meana, J.J.; Berrocoso, E. Chronic pain leads to concomitant noradrenergic impairment and mood disorders. Biol. Psychiatry 2013, 73, 54–62. [Google Scholar] [CrossRef]
  282. Kaushal, R.; Taylor, B.K.; Jamal, A.B.; Zhang, L.; Ma, F.; Donahue, R.; Westlund, K.N. GABA-A receptor activity in the noradrenergic locus coeruleus drives trigeminal neuropathic pain in the rat; contribution of NAalpha1 receptors in the medial prefrontal cortex. Neuroscience 2016, 334, 148–159. [Google Scholar] [CrossRef] [Green Version]
  283. Zhang, Z.; Cordeiro Matos, S.; Jego, S.; Adamantidis, A.; Seguela, P. Norepinephrine drives persistent activity in prefrontal cortex via synergistic alpha1 and alpha2 adrenoceptors. PLoS ONE 2013, 8, e66122. [Google Scholar] [CrossRef] [Green Version]
  284. Jarcho, J.M.; Feier, N.A.; Labus, J.S.; Naliboff, B.; Smith, S.R.; Hong, J.Y.; Colloca, L.; Tillisch, K.; Mandelkern, M.A.; Mayer, E.A.; et al. Placebo analgesia: Self-report measures and preliminary evidence of cortical dopamine release associated with placebo response. Neuroimage Clin. 2016, 10, 107–114. [Google Scholar] [CrossRef] [Green Version]
  285. Dent, M.F.; Neill, D.B. Dose-dependent effects of prefrontal dopamine on behavioral state in rats. Behav. Neurosci. 2012, 126, 620–639. [Google Scholar] [CrossRef]
  286. Sogabe, S.; Yagasaki, Y.; Onozawa, K.; Kawakami, Y. Mesocortical dopamine system modulates mechanical nociceptive responses recorded in the rat prefrontal cortex. BMC Neurosci. 2013, 14, 65. [Google Scholar] [CrossRef] [Green Version]
  287. Santana, N.; Mengod, G.; Artigas, F. Quantitative analysis of the expression of dopamine D1 and D2 receptors in pyramidal and GABAergic neurons of the rat prefrontal cortex. Cereb. Cortex 2009, 19, 849–860. [Google Scholar] [CrossRef] [Green Version]
  288. Ford, G.K.; Moriarty, O.; McGuire, B.E.; Finn, D.P. Investigating the effects of distracting stimuli on nociceptive behaviour and associated alterations in brain monoamines in rats. Eur. J. Pain 2008, 12, 970–979. [Google Scholar] [CrossRef] [PubMed]
  289. Llado-Pelfort, L.; Assie, M.B.; Newman-Tancredi, A.; Artigas, F.; Celada, P. In vivo electrophysiological and neurochemical effects of the selective 5-HT1A receptor agonist, F13640, at pre- and postsynaptic 5-HT1A receptors in the rat. Psychopharmacology 2012, 221, 261–272. [Google Scholar] [CrossRef] [PubMed]
  290. Santello, M.; Bisco, A.; Nevian, N.E.; Lacivita, E.; Leopoldo, M.; Nevian, T. The brain-penetrant 5-HT7 receptor agonist LP-211 reduces the sensory and affective components of neuropathic pain. Neurobiol. Dis. 2017, 106, 214–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  291. Sarter, M.; Parikh, V.; Howe, W.M. Phasic acetylcholine release and the volume transmission hypothesis: Time to move on. Nat. Rev. Neurosci. 2009, 10, 383–390. [Google Scholar] [CrossRef] [Green Version]
  292. Parikh, V.; Kozak, R.; Martinez, V.; Sarter, M. Prefrontal acetylcholine release controls cue detection on multiple timescales. Neuron 2007, 56, 141–154. [Google Scholar] [CrossRef] [Green Version]
  293. Hasselmo, M.E.; Sarter, M. Modes and models of forebrain cholinergic neuromodulation of cognition. Neuropsychopharmacology 2011, 36, 52–73. [Google Scholar] [CrossRef] [Green Version]
  294. Alkondon, M.; Albuquerque, E.X. The nicotinic acetylcholine receptor subtypes and their function in the hippocampus and cerebral cortex. Prog. Brain Res. 2004, 145, 109–120. [Google Scholar] [CrossRef]
  295. Fucile, S.; Sucapane, A.; Eusebi, F. Ca2+ permeability through rat cloned alpha9-containing nicotinic acetylcholine receptors. Cell Calcium. 2006, 39, 349–355. [Google Scholar] [CrossRef]
  296. Gotti, C.; Zoli, M.; Clementi, F. Brain nicotinic acetylcholine receptors: Native subtypes and their relevance. Trends Pharmacol. Sci. 2006, 27, 482–491. [Google Scholar] [CrossRef] [PubMed]
  297. Albuquerque, E.X.; Pereira, E.F.; Alkondon, M.; Rogers, S.W. Mammalian nicotinic acetylcholine receptors: From structure to function. Physiol. Rev. 2009, 89, 73–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  298. Brown, D.A. Muscarinic acetylcholine receptors (mAChRs) in the nervous system: Some functions and mechanisms. J. Mol. Neurosci. 2010, 41, 340–346. [Google Scholar] [CrossRef] [PubMed]
  299. Bubser, M.; Byun, N.; Wood, M.R.; Jones, C.K. Muscarinic receptor pharmacology and circuitry for the modulation of cognition. Handb. Exp. Pharmacol. 2012, 121–166. [Google Scholar] [CrossRef]
  300. Zhang, Z.G.; Hu, L.; Hung, Y.S.; Mouraux, A.; Iannetti, G.D. Gamma-band oscillations in the primary somatosensory cortex--a direct and obligatory correlate of subjective pain intensity. J. Neurosci. 2012, 32, 7429–7438. [Google Scholar] [CrossRef] [Green Version]
  301. Schulz, E.; May, E.S.; Postorino, M.; Tiemann, L.; Nickel, M.M.; Witkovsky, V.; Schmidt, P.; Gross, J.; Ploner, M. Prefrontal Gamma Oscillations Encode Tonic Pain in Humans. Cereb. Cortex 2015, 25, 4407–4414. [Google Scholar] [CrossRef] [Green Version]
  302. Pafundo, D.E.; Miyamae, T.; Lewis, D.A.; Gonzalez-Burgos, G. Cholinergic modulation of neuronal excitability and recurrent excitation-inhibition in prefrontal cortex circuits: Implications for gamma oscillations. J. Physiol. 2013, 591, 4725–4748. [Google Scholar] [CrossRef]
  303. Koga, K.; Matsuzaki, Y.; Honda, K.; Eto, F.; Furukawa, T.; Migita, K.; Irie, K.; Mishima, K.; Ueno, S. Activations of muscarinic M1 receptors in the anterior cingulate cortex contribute to the antinociceptive effect via GABAergic transmission. Mol. Pain 2017, 13, 1744806917692330. [Google Scholar] [CrossRef] [Green Version]
  304. Koga, K.; Matsuzaki, Y.; Migita, K.; Shimoyama, S.; Eto, F.; Nakagawa, T.; Matsumoto, T.; Terada, K.; Mishima, K.; Furue, H.; et al. Stimulating muscarinic M1 receptors in the anterior cingulate cortex reduces mechanical hypersensitivity via GABAergic transmission in nerve injury rats. Brain Res. 2019, 1704, 187–195. [Google Scholar] [CrossRef]
  305. Navarria, A.; Wohleb, E.S.; Voleti, B.; Ota, K.T.; Dutheil, S.; Lepack, A.E.; Dwyer, J.M.; Fuchikami, M.; Becker, A.; Drago, F.; et al. Rapid antidepressant actions of scopolamine: Role of medial prefrontal cortex and M1-subtype muscarinic acetylcholine receptors. Neurobiol. Dis. 2015, 82, 254–261. [Google Scholar] [CrossRef] [Green Version]
  306. Masocha, W. Gene expression profile of sodium channel subunits in the anterior cingulate cortex during experimental paclitaxel-induced neuropathic pain in mice. PeerJ 2016, 4, e2702. [Google Scholar] [CrossRef] [Green Version]
  307. Li, T.T.; Ren, W.H.; Xiao, X.; Nan, J.; Cheng, L.Z.; Zhang, X.H.; Zhao, Z.Q.; Zhang, Y.Q. NMDA NR2A and NR2B receptors in the rostral anterior cingulate cortex contribute to pain-related aversion in male rats. Pain 2009, 146, 183–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. La Porta, C.; Bura, S.A.; Llorente-Onaindia, J.; Pastor, A.; Navarrete, F.; Garcia-Gutierrez, M.S.; De la Torre, R.; Manzanares, J.; Monfort, J.; Maldonado, R. Role of the endocannabinoid system in the emotional manifestations of osteoarthritis pain. Pain 2015, 156, 2001–2012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  309. Madasu, M.K.; Roche, M.; Finn, D.P. Supraspinal Transient Receptor Potential Subfamily V Member 1 (TRPV1) in Pain and Psychiatric Disorders. Mod. Trends Pharm. 2015, 30, 80–93. [Google Scholar] [CrossRef] [Green Version]
  310. Vertes, R.P. Differential projections of the infralimbic and prelimbic cortex in the rat. Synapse 2004, 51, 32–58. [Google Scholar] [CrossRef] [PubMed]
  311. Ji, G.; Neugebauer, V. Modulation of medial prefrontal cortical activity using in vivo recordings and optogenetics. Mol. Brain 2012, 5, 36. [Google Scholar] [CrossRef] [Green Version]
  312. Mukherjee, A.; Caroni, P. Infralimbic cortex is required for learning alternatives to prelimbic promoted associations through reciprocal connectivity. Nat. Commun. 2018, 9, 2727. [Google Scholar] [CrossRef]
  313. Gutman, D.A.; Keifer, O.P., Jr.; Magnuson, M.E.; Choi, D.C.; Majeed, W.; Keilholz, S.; Ressler, K.J. A DTI tractography analysis of infralimbic and prelimbic connectivity in the mouse using high-throughput MRI. Neuroimage 2012, 63, 800–811. [Google Scholar] [CrossRef] [Green Version]
  314. Berendse, H.W.; Galis-de Graaf, Y.; Groenewegen, H.J. Topographical organization and relationship with ventral striatal compartments of prefrontal corticostriatal projections in the rat. J. Comp. Neurol. 1992, 316, 314–347. [Google Scholar] [CrossRef]
  315. Moorman, D.E.; James, M.H.; McGlinchey, E.M.; Aston-Jones, G. Differential roles of medial prefrontal subregions in the regulation of drug seeking. Brain Res. 2015, 1628, 130–146. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Subregion-specific modification of mPFC L5 pyramidal neurons by PV+ GABAergic interneurons in pathological pain. In the ACC, the number of synapses formed by PV+ interneurons (PV-IN, purple) is decreased (red crosses), which leads to disinhibition of pyramidal neurons (PN, blue) and hyperexcitability; in the PrL, PV+ interneurons increase their feed-forward inhibition onto pyramidal neurons, thereby decreasing pyramidal neuron excitability; in the IL, PV content is decreased (faint purple) in PV+ interneurons and axon initial segment (AIS) of pyramidal neurons is shortened (red circle), which has been associated with decreased pyramidal neuron excitability [90].
Figure 1. Subregion-specific modification of mPFC L5 pyramidal neurons by PV+ GABAergic interneurons in pathological pain. In the ACC, the number of synapses formed by PV+ interneurons (PV-IN, purple) is decreased (red crosses), which leads to disinhibition of pyramidal neurons (PN, blue) and hyperexcitability; in the PrL, PV+ interneurons increase their feed-forward inhibition onto pyramidal neurons, thereby decreasing pyramidal neuron excitability; in the IL, PV content is decreased (faint purple) in PV+ interneurons and axon initial segment (AIS) of pyramidal neurons is shortened (red circle), which has been associated with decreased pyramidal neuron excitability [90].
Ijms 21 03440 g001
Figure 2. Schematic of pain-induced changes of L2/3 and L5 excitability in the mPFC subregions. Green arrows indicate a net increase in excitability, red arrow indicates a net decrease in excitability, gray minuses indicate no reported excitability changes.
Figure 2. Schematic of pain-induced changes of L2/3 and L5 excitability in the mPFC subregions. Green arrows indicate a net increase in excitability, red arrow indicates a net decrease in excitability, gray minuses indicate no reported excitability changes.
Ijms 21 03440 g002
Table 1. Regulation of neurotransmitter receptors and ion channels in the different subregions and layers of the mPFC.
Table 1. Regulation of neurotransmitter receptors and ion channels in the different subregions and layers of the mPFC.
PalaRegion (Layer)RegulationPain ModelReference
AChR M1PrL/IL (L5)downSNI[155]
AMPA GluA1ACCupCFA-induced chronic inflammatory pain[243]
AMPA GluA1ACCupCFA-induced chronic inflammatory pain[261]
AMPA GluA1ACCupCFA-induced chronic inflammatory pain[262]
AMPA GluA1ACCincreased membrane insertionperipheral nerve ligation[200]
AMPA GluA1ACCincreased membrane insertionperipheral nerve ligation[158]
GABAA β2ACCupPaclitaxel-induced thermal hyperalgesia[266]
GABAA β3ACCupPaclitaxel-induced thermal hyperalgesia[266]
GABAA δACCupPaclitaxel-induced thermal hyperalgesia[266]
GABAB ACCupPaclitaxel-induced thermal hyperalgesia[266]
HCN1ACCdownCCI[134]
HCNACC/PrL (L5)functional reduction of IhSNI[72]
HCNACC/PrL/IL (L2/3)Ih modulation by noradrenergic alpha2 activationSNI[140]
HCNACC (L5)functional reduction of IhCCI[157]
mGluR1ACCupCCI[134]
mGluR5ACCdownCFA-induced chronic inflammatory pain[260]
mGluR5PrLupSNL[220]
mGluR7PrLdownSNI[219]
Nav1.1ACCupPaclitaxel-induced thermal hyperalgesia[306]
Nav1.2ACCupPaclitaxel-induced thermal hyperalgesia[306]
Nav1.6ACCupPaclitaxel-induced thermal hyperalgesia[306]
NaxACCupPaclitaxel-induced thermal hyperalgesia[306]
Nav β1ACCupPaclitaxel-induced thermal hyperalgesia[306]
Nav β3ACCupPaclitaxel-induced thermal hyperalgesia[306]
NMDA NR2AACCupformalin-induced conditioned place avoidance[307]
NMDA NR2AACCupCFA-induced chronic inflammatory pain[261]
NMDA NR2AACCdownCFA-induced chronic inflammatory pain[263]
NMDA NR2BmPFCupSNI[173]
NMDA NR2BACCupCFA-induced chronic inflammatory pain[239]
NMDA NR2BACCupformalin-induced conditioned place avoidance[307]
NMDA NR2BACCupCFA-induced chronic inflammatory pain[261]
NMDA NR2BACCupCFA-induced chronic inflammatory pain[262]
NMDA NR2BACCupCFA-induced chronic inflammatory pain[263]
NMDA NR2BACCupSNI[247]
TRPV1PrL/IL (L2/3)upSNI[174]
TRPV1PrL/IL (L2/3 and L5)upSNI[230]
CCI, chronic constriction injury; CFA, complete Freund’s adjuvant; SNI, spared nerve injury; SNL, spinal nerve ligation.

Share and Cite

MDPI and ACS Style

Kummer, K.K.; Mitrić, M.; Kalpachidou, T.; Kress, M. The Medial Prefrontal Cortex as a Central Hub for Mental Comorbidities Associated with Chronic Pain. Int. J. Mol. Sci. 2020, 21, 3440. https://doi.org/10.3390/ijms21103440

AMA Style

Kummer KK, Mitrić M, Kalpachidou T, Kress M. The Medial Prefrontal Cortex as a Central Hub for Mental Comorbidities Associated with Chronic Pain. International Journal of Molecular Sciences. 2020; 21(10):3440. https://doi.org/10.3390/ijms21103440

Chicago/Turabian Style

Kummer, Kai K., Miodrag Mitrić, Theodora Kalpachidou, and Michaela Kress. 2020. "The Medial Prefrontal Cortex as a Central Hub for Mental Comorbidities Associated with Chronic Pain" International Journal of Molecular Sciences 21, no. 10: 3440. https://doi.org/10.3390/ijms21103440

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop