Skip to content
Publicly Available Published by De Gruyter June 8, 2023

The role of pain modulation pathway and related brain regions in pain

  • Dandan Yao , Yeru Chen and Gang Chen EMAIL logo

Abstract

Pain is a multifaceted process that encompasses unpleasant sensory and emotional experiences. The essence of the pain process is aversion, or perceived negative emotion. Central sensitization plays a significant role in initiating and perpetuating of chronic pain. Melzack proposed the concept of the “pain matrix”, in which brain regions associated with pain form an interconnected network, rather than being controlled by a singular brain region. This review aims to investigate distinct brain regions involved in pain and their interconnections. In addition, it also sheds light on the reciprocal connectivity between the ascending and descending pathways that participate in pain modulation. We review the involvement of various brain areas during pain and focus on understanding the connections among them, which can contribute to a better understanding of pain mechanisms and provide opportunities for further research on therapies for improved pain management.

1 Introduction

Pain is a complex and multifactorial subjective experience, as defined by the International Association for the Study of Pain (IASP): “An unpleasant sensory and emotional experience associated with, or resembling that associated with, actual or potential tissue damage.” (Raja et al. 2020) When pain persists or recurs for more than three months, it is categorized as chronic pain (Cavalcanti et al. 2021). Chronic pain often presents as a persistent and unavoidable stressor, leading to emotional changes that contribute to its aversive nature. Additionally, a reciprocal relationship exists between pain and affective states, where emotional experiences can exacerbate pain symptoms and perpetuate a vicious cycle between sensory and affective dimensions (Wiech and Tracey 2009). Based on anatomical, physiological, and psychological substrates, pain can be categorized into four main components: nociception, pain perception, suffering, and pain behaviors (Loeser and Melzack 1999). Moreover, there are three main types of pain: nociceptive, neuropathic, and nociplastic, as defined by the IASP (Cohen et al. 2021).

Furthermore, Melzack proposed the concept of the “pain matrix”, which emphasizes that the pain experience arises from the coordinated activity of various brain areas, rather than a single pain center (Melzack 1999). He suggested that the body’s anatomical foundation consists of a neuronal network that extends across extensive brain regions (neuromatrix), generating distinctive patterns of nerve impulses corresponding to different bodily sensations (Melzack 2001). These findings suggest that pain involves the collaborative engagement of interconnected brain structures, potentially encoding sensory and emotional experiences through specific neural pathways.

The “pain matrix” can be anatomically and functionally divided into medial and lateral ascending pathways (Figure 1) (De Ridder et al. 2021; May 2008, 2011; Peyron et al. 2000; Schnitzler and Ploner 2000). The lateral ascending pathway, composed of the thalamus, primary somatosensory cortex (S1), secondary somatosensory cortex (S2), and insular cortex (IC), primarily contributes to the sensory discriminative aspects of pain. Conversely, the medial ascending pathway is believed to process the affective dimensions of pain and encompasses the parabrachial nucleus (PB), prefrontal cortex (PFC), anterior cingulate cortex (ACC), and amygdala (AMG). Moreover, descending pathway, originating in the PFC, ACC, AMG, hypothalamus, and the periaqueductal gray (PAG), modulates nociceptive signal transmission by relaying signals through brainstem nuclei in the PAG and medulla that project to the spinal cord (Cohen and Mao 2014; Millan 2002; Zhuo and Gebhart 1997).

Figure 1: 
Anatomy of the pain neuromatrix. Peripheral nociceptive afferent information projects from the spinal dorsal horn to different targets in the brain. In the ascending pain modulation, the spinothalamic pathway projects to the somatosensory cortex and insula to participate in the sensory-discriminative aspects of pain, while the emotional-affective aspects of pain involve projections from the spino-parabrachio-amygdaloid to prefrontal cortex and limbic system. There is also descending pain modulation mediated by higher structure that project via the PAG to the RVM and finally efferent from the SDH. Abbreviations: SDH, spinal dorsal horn; PB, parabrachial nucleus; S1, primary somatosensory cortex; S2, secondary somatosensory cortex; AMG, amygdala; PFC, prefrontal cortex; ACC, anterior cingulate cortex; PAG, periaqueductal gray; RVM, rostral ventromedial medulla.
Figure 1:

Anatomy of the pain neuromatrix. Peripheral nociceptive afferent information projects from the spinal dorsal horn to different targets in the brain. In the ascending pain modulation, the spinothalamic pathway projects to the somatosensory cortex and insula to participate in the sensory-discriminative aspects of pain, while the emotional-affective aspects of pain involve projections from the spino-parabrachio-amygdaloid to prefrontal cortex and limbic system. There is also descending pain modulation mediated by higher structure that project via the PAG to the RVM and finally efferent from the SDH. Abbreviations: SDH, spinal dorsal horn; PB, parabrachial nucleus; S1, primary somatosensory cortex; S2, secondary somatosensory cortex; AMG, amygdala; PFC, prefrontal cortex; ACC, anterior cingulate cortex; PAG, periaqueductal gray; RVM, rostral ventromedial medulla.

It has been suggested that the imbalance between the ascending and descending pain modulation circuits is thought to be associated with chronic pain in pathological states (Apkarian et al. 2005; Apkarian et al. 2011). Thus, these results provide strong evidence supporting the notion that pain results from the interaction of ascending and descending pathways. And previous evidence has underscored the significance of central and peripheral sensitization in the initiation and maintenance of chronic pain. In this review, we focus on the pain matrix, explore the pain-related brain regions, and elucidate the neural mechanisms governing the ascending and descending pathways involved in pain regulation.

2 The lateral ascending pathway

The sensory discriminative aspects of pain encompass the perception of pain characteristics such as intensity, location, and quality, which are typically transmitted through the spinothalamic-cortical pathway, also known as the lateral pain pathway. Changes in the function of sensory afferent nerves can activate or deactivate the activity in the sensory discrimination dimension, consequently influencing pain processing (Hsieh et al. 2015; Ploner et al. 2000). In patients experiencing severe pain without depression, the plasticity of the pain neural network appears to be primarily confined to the sensory-discriminative aspects of pain (Joo et al. 2021). Furthermore, the abnormal processing of sensory information may also be related to aberrant connectivity from the somatosensory cortex to frontal cortex (Ren et al. 2019).

2.1 Thalamus

The thalamus, a global hub with extensive connections to the somatosensory cortex, transmits sensory information to the cortex and integrates information processing between cortical regions, and is widely recognized as a key region for pain transmission and regulation (Goadsby et al. 2017; Hwang et al. 2017; Lenz et al. 2004; Todd 2010). Approximately 35 % of the spinothalamic neurons terminate in the ventral posterior thalamic nucleus, projecting to S1 and S2, whereas about 25 % target the medial thalamus and then send their axons to the ACC and PFC (Kuner and Kuner 2021). Moreover, the connections between the mediodorsal thalamus (MD) and mPFC are likely to influence emotional and cognitive functions, whereas enhanced synaptic strength in the MD and ACC may contribute to the affective and motivational aspects of pain (Kong et al. 2018; Meda et al. 2019; Thompson and Neugebauer 2019).

One study demonstrated that cerebrovascular accidents in the ventral posterolateral thalamus can result in thermal hyperalgesia and mechanical allodynia, while another study suggested that the lateral thalamus plays a role in the processing of neuropathic pain, especially in thermal hyperalgesia (Nagasaka et al. 2017; Saadé et al. 2006). In line with this, the activation of GABAergic neurons in the ventrobasal thalamus or projections from the thalamic reticular nucleus to the ventrobasal thalamus has shown analgesia effects in chronic inflammatory pain (Zhang et al. 2017). A meta-analysis has indicated that the thalamus is active in both experimentally induced and chronic neuropathic pain conditions (Friebel et al. 2011). Moreover, recent studies have revealed a significant relationship between increased glutamate levels and decreased N-acetylaspartate levels in the thalamus and the development of mechanical allodynia induced by nerve injury, and the enhanced functional connectivity between the thalamus and insula contributes to the occurrence of neuropathic pain (Gustin et al. 2014; Wang et al. 2020).

Low-frequency oscillatory activity is a fundamental functional property of the brain and thalamus. It is noteworthy an early study has found that the abnormal distribution and coherence of thalamocortical low-frequency oscillations are related to the dysregulation of various chronic pain conditions (Llinás et al. 1999). Similarly, an fMRI study demonstrated the involvement of low-frequency oscillations in the thalamocortical network in ascending pain modulation and highlighted the role of abnormal connectivity within these networks in chronic neuropathic pain (Alshelh et al. 2016).

2.2 The primary somatosensory cortex (S1) and secondary somatosensory cortex (S2)

The S1 and S2 have been reported to contribute to the processing of nociceptive information, including the discrimination and localization of pain intensity (Bushnell et al. 1999; Chen et al. 2002). Activation of S1 and S2, as well as IC, has been associated with cold and thermal perception in the neocortex of mice (Beukema et al. 2018; Bokiniec et al. 2018). Furthermore, the involvement of S1 and S2 in the induction of relief from phantom limb pain has been confirmed using noninvasive brain stimulation techniques (Kikkert et al. 2019). A magnetoencephalography study has recorded a significant positive correlation between the intensity of pain stimulation and the activation of the contralateral S1 and bilateral S2 (Timmermann et al. 2001).

Several studies have proposed that various forms of abnormal plasticity in the S1 contribute to chronic pain. Notably, blocking S1 synaptic plasticity alone is insufficient to inhibit pain or even affect pain behavior following injury, indicating that the S1 may affect other pain-related regions that collectively result in chronic pain (Kim et al. 2016, 2017). A previous study demonstrated that optogenetic activation gamma oscillation in the S1 induced pain hypersensitivity and aversive avoidance behaviors, and S1 activity could also further lead to enhanced activation in downstream pain region such as the IC, ACC and AMG (Tan et al. 2019). However, the involvement of the S1 in pain-related negative emotions may not be independent but rather associated with its downstream brain regions. Chemogenic activation of S1 layer II/III neurons significantly reduced mechanical sensitivity and thermal latency (Okada et al. 2021). Interestingly, both excitatory and inhibitory neurons in the S1 layer II/III exhibited increased activity in chronic pain models, but the increase in inhibition was insufficient to counteract the increase in excitation and alleviate chronic pain (Eto et al. 2012).

The S2 is reported to be the final pathway responsible for integrating multimodal sensory information and facilitating the conversion of primary nociceptive signals from the periphery into acute pain perception (Worthen et al. 2011). A recent study discovered increased neural activity in the S2 in mice with spared nerve injury (Inami et al. 2019). Additionally, it has been reported that the S2 receives inputs from thalamic neurons, and a decrease in excitatory tone within the S2 was observed to modulate comorbid pain in depression, but not pain induced by inflammation or nerve injury (Zhu et al. 2019). However, the specific function of the S2 in pain processing requires further elucidation.

2.3 Insular cortex (IC)

The insular, located deep within the lateral fissure of the human brain, functions as a pivotal cortical hub and plays crucial roles in sensory processing, visceral responses, emotional guidance of social behavior, and perceptual self-awareness (Benarroch 2019; Craig 2003; Giesecke et al. 2004; Lu et al. 2016; Wang et al. 2021). Anatomically, the insula has different roles in pain processing: the posterior insular (PI) is primarily involved in the sensory-discriminative aspects of nociceptive inputs, whereas the anterior insular (AI) is more implicated in its affective dimensions (Craig 2002; Singer et al. 2009; Tracey 2005). Moreover, a recent study has shown that the diversity connectivity within the insular promotes to the division of insular functions to some extent (Tian and Zalesky 2018). The IC receives most of the spinothalamic inputs, and the PI and AI have distinct connections to the ascending spinothalamic pathways (Craig 2014; Dum et al. 2009).

Human electroencephalography studies have demonstrated that the IC is activated by acute or chronic pain states, and low-intensity electrical stimulation of the human IC can produce somatosensory sensations and visceral responses, further supporting the critical roles of the IC in pain and sensory perception (Mazzola et al. 2009; Ostrowsky et al. 2002; Ploghaus et al. 1999). Previous studies have found that excitatory patterns and changes are associated with the IC neuroplasticity after nerve injury (Ching et al. 2018; Han et al. 2016). Furthermore, peripheral nerve injury can active the IC and affect its the volume or thickness. A previous meta-analysis has shown that spinal cord injury can lead to volume loss in the gray matter in the IC, particularly in the left IC (Nardone et al. 2018; Wang et al. 2019).

The right AI has been reported to play an important role in large-scale switching, and this structure also contributes to neural networks disruption in chronic pain (Cottam et al. 2018). Additionally, recent study has revealed a relationship between the resting-state functional heterogeneity of the right IC and pain sensitivity (Veréb et al. 2021). Another study has suggested that morphological and functional network properties of the PI also participate in pain sensitivity (Neumann et al. 2021). Interestingly, a recent review found that neuropathic pain affects the AI more than the PI (Chao et al. 2018). Furthermore, the PI is also involved in the maintenance of persistent pain and contributes to the management of chronic pain (Benison et al. 2011; Segerdahl et al. 2015). These results indicate that distinct sites within the IC may be related to different aspects of pain processing.

3 The medial ascending pathways

The medial pain pathway, an essential neural circuitry involved in pain processing, encompasses a network of brain regions that contribute to the affective and emotional aspects of the pain experience. It is important to note that the affective components of pain are distinct from the sensory components, and addressing the affective aspects can further alleviate chronic pain (Zhou et al. 2018). The PB receives neural inputs from the superficial dorsal horn and projects diffusely to the limbic system, including the AMG and PFC, which are responsible for emotional processing. Overall, the spinal-parabrachial pathway may play a critical role in the emotional and autonomic components of pain. Further exploration and understanding of this pathway hold promise for advancing our comprehension of pain perception and developing potential therapeutic interventions.

3.1 Parabrachial nucleus (PB)

The PB, situated at the junction of the midbrain and pons, is widely recognized as a pivotal relay station responsible for receiving nociceptive information from the spinal cord and projecting it to the AMG and other brain regions, thereby converting sensory input into the emotional-affective components of pain. Efferent projections from PB neurons to the ventromedial hypothalamus (VMH) and PAG are responsible for mediating escape behavior, whereas projections to the bed nucleus stria terminalis (BNST) or AMG are involved in aversive learning to noxious stimulation (Chiang et al. 2020). Moreover, PB neurons that project to the thalamic paraventricular (PVT) induce negative emotion-related behaviors, without affecting nociceptive information processing (Zhu et al. 2022). Notably, Matusmoto et al. reported increased spontaneous and evoked activity of PB neurons in inflammatory pain models (Matsumoto et al. 1996), whereas hyperactivity of PB neurons in chronic trigeminal pain was observed solely in response to noxious thermal stimuli (Uddin et al. 2018). These finding underscore the crucial role of PB as a critical node in the pathogenesis of chronic pain and pain processing.

Through the use of single-cell RNA sequencing analysis, researchers identified 13 types of glutamatergic neurons and sparsely distributed GABAergic neurons within the PB, with the majority of PB neurons expressing multiple neuropeptides and receptors for Calcitonin gene-related peptide (CGRP) (Pauli et al. 2022). Inhibiting the PB has shown potential in attenuating chronic pain, with projections from GABAergic neurons in the central amygdala (CeA) to PB playing a major role as inhibitory inputs, thereby revealing that the activation of CeA-PB projections is involved in alleviating pain and negative affect-related behaviors (Hogri et al. 2022; Raver et al. 2020). Despite GABAergic neurons are scarce and dispersed in the PB compared to glutamatergic neurons, the activation of local GABAergic neurons in the PB is sufficient to achieve direct monosynaptic innervation and functionally inhibit glutamatergic neurons (Sun et al. 2020a). CGRP is encoded by Calca, a calcitonin-related mRNA splice variant, and the strong expression of μ-opioid receptors in the PB is predominantly co-localized with CGRP/Calca neurons (Huang et al. 2021; Rosenfeld et al. 1983). These studies suggest that opioid withdrawal may induce hyperactivity in CGRP/Calca neurons, potentially contributing to the experience of negative emotions associated with opioid withdrawal and opioid-induced respiratory depression.

3.2 Amygdala (AMG)

The AMG, an almond-shaped nuclear complex in the limbic system, was first proposed by Burdach in the early 19th century (LeDoux 2007). Given its well-established role in affective processing and recent findings demonstrating its involvement in nociceptive processing, the AMG is recognized as a unique site that influences pain modulation (Neugebauer et al. 2004; Veinante et al. 2013). Dysfunction of the AMG has been associated with an increased risk of chronic pain development, as decreased AMG volume and increased intra-corticolimbic white matter connections are independent risk factors for pain persistence (Vachon-Presseau et al. 2016). Recent studies on fear learning in humans and animals have consistently suggested that anxiety results from abnormalities in the mechanisms responsible for regulating conditioned fear (Pare and Duvarci 2012). Remarkably, AMG plasticity is believed to play a crucial role in emotion modulation, particularly in fear condition.

Multiple nuclei in the AMG participate in pain-related functions, including the central nucleus (CeA), the basolateral complex (BLA), and the intercalated cell clusters (ITC) (Sah et al. 2003; Thompson and Neugebauer 2017). The CeA, which is responsible for the primary function of AMG outputs, receives nociceptive inputs directly from PB via the spino-parabrachio-amygdaloid pathway and higher processed information inputs indirectly from the thalamus and cortex via the BLA (Neugebauer 2020). The two excitatory inputs from PB and BLA converge in CeA, where they integrate and regulate emotional responses. An optogenetic study has recently shown that activation of excitatory projections from PB to CeA induces negative emotional behaviors in rats, including aversion, anxiety, and depression, whereas the excitatory projections from BLA to CeA counteract these behaviors (Cai et al. 2018). Given that the BLA is a key site of convergence for negative and positive behavior, BLA-CeA projections also contribute to aversive behaviors such as anxiogenesis and defensive responses (Kim et al. 2016, 2017). Additionally, pain significantly increased the synaptic strength of PB inputs in the lateral division of CeA cells when receiving only PB inputs, while it selectively enhanced the synaptic strength of BLA inputs but not PB inputs when receiving both BLA and PB inputs (Ge et al. 2022). Exploring how AMG circuits, particularly those within the CeA, modulate behavioral responses to positive emotions (reward) and negative emotions (aversion, fear, and anxiety) would be of great interest.

The CeA neurons are functionally and neurochemically heterogeneous, consisting of distinct populations of GABAergic neurons that express a variety of neuropeptides, including CGRP, corticotropin releasing factor (CRF), protein kinase C-delta (PKCδ), and somatostatin (SOM) (Neugebauer et al. 2020). These distinct neuronal populations regulate various behaviors, such as anxiety, depression and fear. Notably, blocking the CeA-CGRP receptor (Shinohara et al. 2017) and optogenetically inhibiting CeA-CRF neuron (Mazzitelli et al. 2021, 2022) have demonstrated the potential to alleviate pain and emotional-affective responses. It has been shown that CGRP1 receptor blockade reduces CeA-CRF neuronal activity in chronic neuropathic pain, and that the reduction in CeA-CRF neuronal activity is critical for the facilitated behavioral responses observed with CGRP1 receptor antagonism (Presto and Neugebauer 2022). Recent studies have been shown that activation of CeA-PKC-δ neurons increases anxiolytic effects in mice (Cai et al. 2014; Griessner et al. 2021), whereas activation of CeA-SOM neurons induces the appearance of anxiety-like behaviors (Ahrens et al. 2018). In contrast, Wilson et al. identified that activation of CeA-SOM neurons contributed to pain relief, whereas increased activity of CeA-PKCδ neurons amplified pain-related behaviors, and that both of the two distinct neurons were recipients of inputs from PB (Wilson et al. 2019). An agent-based model has demonstrated that the relative proportion of PKCδ and SOM neurons is a key parameter in predicting pain and this parameter exhibits significant left-right lateralization (Miller Neilan et al. 2021). Interestingly, studies on the hemispheric lateralization of pain-related amygdala function have suggested that the right CeA plays a predominant role in facilitating pain-related behavior, compared with the left CeA (Allen et al. 2021; Sadler et al. 2017).

3.3 Prefrontal cortex (PFC)

Growing evidence from rodent and human studies has confirmed that the PFC plays a critical role in integrating sensory, cognitive, and emotional information related to pain experiences (Bushnell et al. 2013; Hashmi et al. 2013; Maihöfner and Handwerker 2005; Seminowicz and Moayedi 2017). Meanwhile, numerous neuroimaging studies have focused on functional localization and subdivision of the human PFC, suggesting its constituents as the medial PFC (mPFC), dorsolateral PFC (dlPFC), ventrolateral PFC, and orbitofrontal cortex (OFC) (Carlén 2017). Notably, the subregion that is related to pain is localized in the mPFC (Apkarian et al. 2011). The mPFC consolidates pain-related sensory characteristics and affective information, evaluates motivational factors, and computes action progress through motor circuits, to relieve or aggravate pain. Nerve injury results in significant alterations tactile stimuli processing in the ventral mPFC and indicates that ERK phosphorylation participates in pain perception, particularly in neurons located in layers II–III and V–VI (Devoize et al. 2011). Similarly, dysregulation of glutamatergic inputs in neuropathic pain may result in mPFC inactivation and potentially impair PFC-dependent cognitive tasks (Kelly and Martina 2018). Neurophysiological recordings of the mPFC and the mediodorsal thalamus connectivity are relevant to deficits in pain-related working memory (Cardoso-Cruz et al. 2013). According to anatomical and physiological function, the mPFC in rodent consists of the anterior cingulate cortex (ACC), the prelimbic cortex (PL), and the infralimbic cortex (IL) (Laubach et al. 2018).

3.3.1 The anterior cingulate cortex (ACC)

There is converging evidence demonstrating that ACC is critically important for the sensory perception and emotional responses (Barthas et al. 2015; Bushnell et al. 2013; Fuchs et al. 2014; Johansen and Fields 2004; Sellmeijer et al. 2018). Mediodorsal thalamic projections to ACC are involved in chronic pain-related aversion by inhibiting subcortically projecting ACC neurons, whereas BLA-ACC input mitigates aversion (Meda et al. 2019). Additionally, transplanting medial ganglionic eminence cells to the rostral ACC enhances GABAergic inhibitory control and alleviates long-lasting pain-induced aversiveness, but not in the posterior ACC (Juarez-Salinas et al. 2019). Optogenetic activation of the ACC-spinal cord pathway is involved in neuropathic pain hypersensitivity behaviors, whereas its inhibition induces analgesic effects, and the modulation of this pathway is independent of RVM (Chen et al. 2018).

Chronic pain-induced long-lasting anxiety is known to depend on presynaptic plasticity. Early study has demonstrated that long-term potentiation (LTP) of synaptic transmission in the ACC could result in exaggerated chronic pain responses (Li et al. 2010). Blocking hyperpolarization-activated cyclic nucleotide-regulated (HCN) channels in the ACC to erase presynaptic LTP has been shown to significantly reduce anxiety-like behaviors induced by chronic pain (Koga et al. 2015). Another study reported that in neuropathic pain, abnormal HCN channel function of cortical output neurons in the ACC layer V could be restored by type 7 serotonin receptors (5-HT7) (Santello and Nevian 2015). Furthermore, downregulation of mGluR1 has been shown to alleviate the neuronal hyperexcitability of layer V pyramidal neurons in the ACC, which is inhibited by HCN channels (Gao et al. 2016).

A long-standing hypothesis suggests that females may exhibit heightened sensitivity to pain and associated unpleasant emotions compared to males (Pieretti et al. 2016). Evidence suggests that sex differences in excitatory glutamatergic neurons within the ACC play a crucial role in the sensory and affective aspects of pain (Jarrin et al. 2020). Several studies have suggested that women with chronic pain exhibit aberrant ACC circuitry compared to men (Monroe et al. 2018; Osborne et al. 2021). Similarly, neuroimaging data has also identified sex differences in the functional and structural connectivity between the ACC and the descending antinociceptive pathway (Wang et al. 2014). These findings suggest that female brain circuits are more extensively involved in the pain modulatory systems that mediate the habituation of pain.

3.3.2 The prelimbic and infralimbic (PL/IL)

More recently, a comparison of PL and IL projections in rodents and humans found that PL and IL are distributed very differently in the brain. Specifically, the PL primarily projects to limbic system, which is involved in cognition-related function, whereas the IL projects to autonomic/visceral-related sites that affect visceromotor activity (Vertes 2004, 2006). Importantly, complicated goal-directed behaviors require the integration of cognitive and visceromotor activities, which involves the interactions between PL and IL. It has been demonstrated that there is a direct reciprocal layer V–VI connection in PL and IL, and this connection is essential for new learning and fear extinction (Marek et al. 2018; Mukherjee and Caroni 2019).

The PL is located in the medial wall of the PFC in rodents, and consolidates action-outcome relationships and affects action-outcome-driven behavior (Autry and Monteggia 2012; Woon et al. 2020). A recent study found that inactivation of the PL, but not the IL, impairs the acquisition and expression of conditioned place avoidance, indicating that the PL contributes to the aversion dimension of pain and environmental context (Jiang et al. 2014). Optogenetic activation of PL excitatory neurons can attenuate thermal hyperalgesia and anxiety-like behavior in a chronic inflammatory pain model (Wang et al. 2015), while selective silencing of PL excitatory neurons reverses pain-related working memory deficits (Cardoso-Cruz et al. 2019). Furthermore, increasing the basal firing of neurons in the PL with low-frequency optogenetic stimulation scaled up prefrontal control and inhibited pain (Dale et al. 2018). Another study shows that the PL is capable of modulating both sensory and emotional responses to neuropathic pain, which is mediated by GABAergic inputs (Zhang et al. 2015). Interestingly, nerve injury increases the excitability of layer V parvalbumin-positive inhibitory neurons in the PL, consequently leading to a decrease in the excitability of layer V pyramidal neurons, with this effect observed exclusively in male mice and not in females (Jones and Sheets 2020).

Under inflammatory pain conditions, the deactivation of PL cortex cells has been associated with reduced levels of glutamate (and conversely increased GABA) in the PL/IL cortex (Luongo et al. 2013) Overexpression of mGluR1, but not mGluR5, has been recorded in the PL-IL in inflammatory pain, indicating that mGluR1 is involved in the connection between BLA and PL-IL neurons under inflammatory pain (Kelly et al. 2016; Luongo et al. 2013). Another study suggested that the IL cortex facilitates nociceptive behavior through a descending pathway via mGluR5 signaling in normal and arthritic rats, whereas mGluR1 activation is effective only in arthritic rats (David-Pereira et al. 2016; David-Pereira et al. 2017). Furthermore, rescuing impaired endocannabinoid signaling in the mPFC activates mGluR5 function to restore IL outputs, thereby inhibiting pain and attenuating pain-related cognitive function, whereas mGluR1 can reverse the inhibition of neuronal spontaneous firing and overcome pain-related cognitive deficits (Kiritoshi et al. 2016).

Optogenetic inhibition of the IL cortex has been shown to alleviate anxiety-like behavior, while the distinct IL outputs, including the lateral septum and the CeA, exert antagonistic effects in the modulation of anxiety (Chen et al. 2021). Knockdown of mTOR in the IL led to robust depression-like behaviors, accompanied by decreased expression of BDNF bilaterally in both the PL and IL (Garro-Martínez et al. 2021). It has also been reported that peripheral inflammation decreases BDNF levels in the IL, while supplementing BDNF in the IL has been found to alleviate pain and accelerate recovery from inflammatory pain (Yue et al. 2017).

4 The descending pain modulation

The establishment and maintenance of chronic pain involves the alterations in pain modulation pathways. The process of afferent somatosensory information is altered by pathways that descend from the brain to the spinal dorsal horn (SDH), thereby altering the perception and response to somatosensory stimuli and resulting in pain changes. Previous studies have elucidated the integration of descending inhibition and facilitation in the transmission of nociceptive information. The descending facilitation pathway is usually not significantly involved in nociception under normal conditions. Thus, it would be a valuable research hotspot in chronic pain, as it aims to restore the balance between the two pathways by reducing descending facilitation and enhancing the descending inhibition (Kwon et al. 2014). The descending pain modulation system also provides a mechanism for pain can be modulated in cortical and subcortical regions (Bourne et al. 2014). Many complex neural circuits in the brain are involved in the descending pathway, and the PAG-RVM-SDH pathway is considered the most important and classical neural pathway of the descending pain system (François et al. 2017; Heinricher et al. 2009; Lau and Vaughan 2014; Millan 2002; Ossipov et al. 2010).

4.1 Periaqueductal gray (PAG)

The midbrain PAG, an evolutionarily conserved region, plays significant roles in pain, anxiety, depression, and fear (Lau and Vaughan 2014; Tovote et al. 2016). Studies have demonstrated that the ventrolateral PAG (vlPAG) is an important part of the neural circuits that mediate pain regulation and serve as the primary site of endogenous opioid analgesia, and electrical stimulation of the vlPAG could produce a potent analgesic effect, while the dorsolateral/lateral PAG (dl/lPAG) prompts active defensive responses, such as jumping and running (Lanius et al. 2018; McPherson and Ingram 2022). A previous study has confirmed that the vlPAG comprises distinct subpopulations of neurons that modulate pain transmission. For example, persistent inflammation increases presynaptic GABA release from vlPAG neurons in female rats induced by CFA injection (Tonsfeldt et al. 2016). It has been shown that the activation of vlPAG glutamatergic neurons, in contrast to vlPAG GABA neurons, exerts analgesic and anxiogenic effects (Samineni et al. 2017; Zhu et al. 2019). Another study suggested that selective stimulation of dopaminergic neurons within the vlPAG/dorsal raphe produces antinociception without anxiogenic behavioral effects (Taylor et al. 2019).

Interestingly, in neuropathic pain, the activation of the mPFC/vlPAG pathway has a descending inhibitory effect (exerts analgesic and anxiolytic effects) and enhances the activity of GABAergic neurons in this region, while silencing the mPFC/vlPAG pathway may contribute to the establishment of pain and the emergence of anxiety-like behaviors (Yin et al. 2020). A previous study has suggested that nerve injury increases GABAergic interneuron inputs from the BLA to the mPFC due to the reduction in endocannabinoid regulation (Huang et al. 2019). These enhanced synaptic connections mediate projections in the mPFC/vlPAG, which contribute to hyperalgesia after peripheral nerve injury. Furthermore, Liang et al. proposed a novel neural pathway in which the activation of glutamatergic neurons in the posterior PVT-CeA-vlPAG pathway contributes to descending pain facilitation (Liang et al. 2020). Moreover, the activation of glutamatergic neurons projecting from the ACC to vlPAG also induces hyperalgesia and anxiety-like behaviors in nerve-injured mice (Zhu et al. 2021).

Output projections are another important target for investigation and can help us understand how different types of vlPAG neurons transmit to downstream targets and how they change in pathological states. The vlPAG-RVM pathway is widely recognized as the primary control pathway for effectively modulating spinal nociceptive processing. Additionally, other projection regions may also contribute to anti-nociception, thereby expanding our understanding of PAG subpopulations in the descending pathway. Recent studies have focused on dopaminergic neurons from the vlPAG to BNST promoting anti-nociception, and interestingly, the BNST GABAergic neurons that project to the vlPAG is implicated in feeding behavior (Hao et al. 2019; Yu et al. 2021). In addition, projections to the central medial thalamic nucleus may also participate in the transmission and regulation of neuropathic pain (Sun et al. 2020b).

4.2 Rostral ventromedial medulla (RVM)

The RVM acts as a relay station in the descending pathway, receiving descending projections from PAG, thalamus, PB and other encephalic regions, and subsequently projecting to the SDH, thereby playing a key role in the modulation of pain transmission in the central nervous system (Millan 2002; Tang et al. 2021). Based on their physiological characteristics and roles in pain regulation, RVM neurons are classified into three types: on-cell, off-cell, and neutral-cell (Bagley and Ingram 2020; Chen and Heinricher 2022; Heinricher et al. 2009; Ossipov et al. 2014). Electrophysiological studies in rodents have identified that on-cell and off-cell activities are responsible for the pronociceptive and antinociceptive outputs of the RVM, respectively (Heinricher et al. 1994; Heinricher et al. 2009; Morgan and Fields 1994). On-cell activity results in enhanced nociception, and selective activation of on-cell is sufficient to induce hyperalgesia, conversely, selective blockade of on-cell activity can prevent hyperalgesia (Martenson et al. 2009; Neubert et al. 2004). The imbalance between the outputs of the on-cell and off-cell populations can amplify or suppress nociception and pain behaviors. Similarly, mu-opioid receptor agonists can directly inhibit on-cell activity in the RVM and remove the inhibitory effect on off-cell, thereby inhibiting nociceptive transmission in the SDH (Fields 2000; Heinricher et al. 2009). Neutral-cells have showed almost no nociceptive responses, as shown by the absence of alterations in firing patterns during nociceptive reflexes, and it remains an open question whether neutral-cells have a role in pain modulation.

A recent study has shown that the RVM is activated 30 min after exposure to noxious stimuli, suggesting that the PAG-RVM pathway is involved in the early development of pain (Tobaldini et al. 2019). Consistently, the analgesic effect of sciatic nerve stimulation, which activates the RVM, takes effect on post-injury day 1 in a neuropathic pain model (Wong et al. 2022). It is well known that RVM-SDH descending pain modulation pathway releases either serotonin or the inhibitory neurotransmitters GABA and/or glycine. An optogenetic study has shown that activation and inhibition of RVM-SDH GABAergic neurons facilitate and attenuate mechanical pain, respectively, providing evidence that altering the activity of inhibitory RVM inputs into the SDH can achieve bidirectional control (François et al. 2017). Additionally, a previous study found that stimulation of RVM-SDH is inhibitory via GABAergic and glycinergic neurotransmission (no excitatory or serotonergic currents), and kappa-opioid receptor agonists inhibited the GABA/glycine terminals of RVM projections to the SDH, resulting in analgesic effects (Otsu and Aubrey 2022). Interestingly, kappa-opioid receptor agonists appear to exert pronociceptive effects on kappa-opioid receptor-containing RVM neurons that project to the SDH (Nguyen et al. 2022). These findings suggest that descending pain modulation can be considered as a key mechanism for opioid-induced analgesia, providing an opportunity for the endogenous opioid system to regulate pain.

5 Pain pathway is a complex neural network across various brain regions and processes

Pain is a complex experience that engages a network of brain regions, collectively known as the pain matrix, which are responsible for the integration of sensory, emotional, and cognitive processes. Two primary neural circuits contribute to pain processing: the ascending pain pathway and the descending pain pathway. The ascending pathway conveys nociceptive information from the peripheral to higher brain regions, while the descending pathway modulates nociceptive signals as they ascend. It is crucial to recognize that the pain pathway is not a linear, simplistic route, but rather a complex neural network with numerous connections across various brain regions and processes. Pain is not controlled by a single brain region alone, nor is a single brain region responsible for encoding only one aspect of pain. Instead, pain perception and processing involve complex interactions between various brain regions and neural pathways.

It is noteworthy that the thalamus receives nociceptive information from the brainstem and transmits pain signals to the somatosensory cortex and insular, thereby contributing to the perception of pain, including its location, intensity, and quality. Optogenetic inhibition of the S1/ACC pathway has been demonstrated to effectively alleviate pain-related affective symptoms (Singh et al. 2020). Furthermore, the IC is reciprocally connected with the S2, AMG, and ACC, and it has been suggested to play a crucial role in pain processing, particularly in pain intensity coding (Coghill et al. 1999). A recent review indicated that the functional connectivity between the IC and the S1/S2 cortex is significantly reduced in neuropathic pain (Yalcin et al. 2014). Additionally, other findings demonstrated that the IC-BLA pathway modulates both the somatosensory and aversive components of pain, while the MD-BLA pathway is only involved in modulating the aversive aspect of pain (Meng et al. 2022). Consistently, studies have suggested that the IC-BLA pathway is closely associated with the onset and development of empathic pain (Zhang et al. 2022). Furthermore, the ACC is functionally connected to the anterior insular and enables the integration of sensory input into motor, motivational, and emotional responses by switching between large-scale networks to facilitate attention access (Galhardoni et al. 2019).

Additionally, nociceptive information related to the emotional and cognitive aspects of pain ascends via the spino-parabrachio-amygdaloid pathway to the prefrontal cortex and limbic system. Previous studies have suggested that projections from the PB-CeA pathway participate in the nociceptive-emotional connection and it’s tightening in chronic pain (Kato et al. 2018). However, recent research has shown that the CeA-PB pathway efficiently modulates PB responses to nociceptive inputs, and dysregulation of this inhibition amplifies PB neuronal activity, contributing to the pathogenesis of chronic pain (Raver et al. 2020). Meanwhile, activation of AMG-PB projections reduces behaviors associated with negative emotions and enhances behaviors associated with positive effects (Hogri et al. 2022). Subsequent investigations have revealed that the PB projects directly to the IC and IL/PL, which are responsible for visceral sensory and motor cortical areas, respectively (Chiang et al. 2019; Saper and Loewy 2016). Individuals with chronic low back pain exhibit greater resting-state functional connectivity and lower effective connectivity of the AMG-mPFC pathway (Mao et al. 2022). Previous research has demonstrated that in a rat model of arthritis pain, hyperactivity of BLA neurons results not only in the emotional-affective components of pain but also in pain-related decision-making impairments through BLA-mPFC connectivity (Ji et al. 2010). The ACC-BLA circuit has been reported to play a critical role in fear response (Allsop et al. 2018; Jhang et al. 2018). Furthermore, previous studies have demonstrated that the targeting of the PL and IL cortex by the BLA has differential implications for fear conditioning within mPFC networks, whereby BLA inputs to IL may enhance extinction while BLA inputs to PL may facilitate fear expression (Jhang et al. 2018; Senn et al. 2014). Interestingly, another study showed that projection from IL, but not PL, to the BLA was sufficient to attenuate anxiety-like behaviors (Bloodgood et al. 2018; Chen et al. 2021).

The descending pain modulation pathway, also known as the antinociceptive pathway, affects the transmission of information in the brain through the activation of the on or off cells. Moreover, the descending pathway is capable of releasing endogenous opioids, which play an important role in the endogenous modulation of nociceptive transmission in the central nervous system. The PAG serves as a pivotal relay station that integrates sensory and motor functions by receiving inputs from higher centers, such as the AMG, PFC, and ACC, and projecting to the RVM, which ultimately targets the SDH to control the transmission of nociceptive signal (Donaldson and Lumb 2017; Koutsikou et al. 2015). The descend pathway enables the brain to exert top-down control over pain perception. Previous studies have shown that increased ACC projections to the dl/lPAG modulate reflexive and active avoidance behaviors to pain, whereas decreased inputs from the PL to vlPAG are critical for the development and progression of chronic pain (Drake et al. 2021; Lee et al. 2022). These findings suggest that the projections from the PL and ACC to different subregions of PAG exert opposing effects on pain and defensive behavior. Furthermore, the ACC-PAG-RVM pathway appears to contribute to placebo analgesia, which involves the endogenous opioid system (Eippert et al. 2009; Fields 2004). Furthermore, the BLA-mPFC-PAG pathway is involved in the development of mechanical and thermal hypersensitivity responses after peripheral nerve injury, whereas the PFC-AMG-PAG pathway is shown to mediate fear-conditioned analgesia (Butler et al. 2011; Huang et al. 2019).

Notably, the PAG not only receives direct inputs from the spinomesencephalic tract but also receives ascending inputs from the superficial and deep dorsal horns through the PB, contributing to nociception, analgesia, and aversive behaviors (Roeder et al. 2016; Willis and Westlund 1997). Moerover, the PAG transmits signals to forebrain sites via t he PB to modulate ascending nociceptive and/or visceral information (Krout et al. 1998). Interestingly, vPAG-CeA projections are modulated by serotonin and are involved in the response to fear learning (Hon et al. 2022). Overall, the processing of pain involves a complex network of brain regions and pathways, which work together to perform pain perception, assessment, and regulation to allow the body to be protected and respond appropriately when faced with painful stimuli.

6 Conclusions

Multiple brain regions are involved in the establishment and maintenance of chronic pain. Indeed, the persistence of chronic pain in the brain can lead to changes in brain areas involved in emotion and cognition, resulting in symptoms such as depression, anxiety, memory impairment. Central and peripheral sensitization, as well as alterations in pain modulation pathways, is typically associated with chronic pain. Thus, the overall experience of pain is the result of a collection of network activity generated by distinct regions in the brain together with ascending and descending pathways.

This review provides an in-depth analysis of the functional connectivity between various brain regions associated with pain, revealing the nociceptive and affective networks related to pain perception. Pain is mediated by at least two ascending excitatory pathway and one descending pain modulation pathway (De Ridder and Vanneste 2017). The medial and lateral ascending pain pathways function simultaneously and can be modulated individually without interfering with other pathways, although they typically work together (Bushnell et al. 2013; Frot et al. 2008). The cerebral cortex receives ascending nociceptive inputs from the SDH and subsequently projects them to distinct brain regions involved in pain regulation, which is essential for the control of pain threshold and its emotional components.

In recent years, brain functional imaging techniques have been employed to study the mechanisms of chronic pain, providing profound insight into changes in pain matrix during pain processing (Frøkjær et al. 2018; Morton et al. 2016). With the maturity of optogenetics and chemogenetics techniques, as well as the emerging single-cell transcriptomic and epigenomic technologies, we now have the possibility to further investigate pain-related neural circuits and cell-targeted therapy in specific brain regions. Pain management remains to be a huge challenge, and only by thoroughly understanding its mechanisms and investigating innovative therapeutic targets and treatment can we alleviate discomfort and improve the well-being of patients. Therefore, we look forward to developing more effective pain management programs based on the pathophysiology of pain through multidisciplinary collaboration in the future.


Corresponding author: Gang Chen, Department of Anesthesiology, School of Medicine, Shaoxing University, Shaoxing, Zhejiang, China; and Department of Anesthesiology, Sir Run Run Shaw Hospital, School of Medicine, Zhejiang University, Hangzhou, Zhejiang 310058, China, E-mail:

Funding source: Zhejiang Provincial Department of Medicine and Health Science and Technology

Award Identifier / Grant number: No. YH42021010

Award Identifier / Grant number: No.82001424

Award Identifier / Grant number: No.82171176

  1. Author contributions: Gang Chen had the idea for the article; Dandan Yao and Yeru Chen drafted the manuscript. All authors read and approved the final manuscript.

  2. Research funding: This research was supported by the National Natural Science Foundation of China (No.82171176 and No.82001424), and the Zhejiang Provincial Department of Medicine and Health Science and Technology (No. YH42021010).

  3. Conflict of interests: The authors declare no competing interests.

References

Ahrens, S., Wu, M.V., Furlan, A., Hwang, G.R., Paik, R., Li, H., Penzo, M.A., Tollkuhn, J., and Li, B. (2018). A central extended amygdala circuit that modulates anxiety. J. Neurosci. 38: 5567–5583, https://doi.org/10.1523/jneurosci.0705-18.2018.Search in Google Scholar PubMed PubMed Central

Allen, H.N., Bobnar, H.J., and Kolber, B.J. (2021). Left and right hemispheric lateralization of the amygdala in pain. Prog. Neurobiol. 196: 101891, https://doi.org/10.1016/j.pneurobio.2020.101891.Search in Google Scholar PubMed PubMed Central

Allsop, S.A., Wichmann, R., Mills, F., Burgos-Robles, A., Chang, C.J., Felix-Ortiz, A.C., Vienne, A., Beyeler, A., Izadmehr, E.M., Glober, G., et al.. (2018). Corticoamygdala transfer of socially derived information gates observational learning. Cell 173: 1329–1342, https://doi.org/10.1016/j.cell.2018.04.004.Search in Google Scholar PubMed PubMed Central

Alshelh, Z., Di Pietro, F., Youssef, A.M., Reeves, J.M., Macey, P.M., Vickers, E.R., Peck, C.C., Murray, G.M., and Henderson, L.A. (2016). Chronic neuropathic pain: it’s about the rhythm. J. Neurosci. 36: 1008–1018, https://doi.org/10.1523/jneurosci.2768-15.2016.Search in Google Scholar

Apkarian, A.V., Bushnell, M.C., Treede, R.D., and Zubieta, J.K. (2005). Human brain mechanisms of pain perception and regulation in health and disease. Eur. J. Pain 9: 463–484, https://doi.org/10.1016/j.ejpain.2004.11.001.Search in Google Scholar PubMed

Apkarian, V.A., Hashmi, J.A., and Baliki, M.N. (2011). Pain and the brain: specificity and plasticity of the brain in clinical chronic pain. Pain 152: S49–S64, https://doi.org/10.1016/j.pain.2010.11.010.Search in Google Scholar PubMed PubMed Central

Autry, A.E. and Monteggia, L.M. (2012). Brain-derived neurotrophic factor and neuropsychiatric disorders. Pharmacol. Rev. 64: 238–258, https://doi.org/10.1124/pr.111.005108.Search in Google Scholar PubMed PubMed Central

Bagley, E.E. and Ingram, S.L. (2020). Endogenous opioid peptides in the descending pain modulatory circuit. Neuropharmacology 173: 108131, https://doi.org/10.1016/j.neuropharm.2020.108131.Search in Google Scholar PubMed PubMed Central

Barthas, F., Sellmeijer, J., Hugel, S., Waltisperger, E., Barrot, M., and Yalcin, I. (2015). The anterior cingulate cortex is a critical hub for pain-induced depression. Biol. Psychiatry 77: 236–245, https://doi.org/10.1016/j.biopsych.2014.08.004.Search in Google Scholar PubMed

Benarroch, E.E. (2019). Insular cortex: functional complexity and clinical correlations. Neurology 93: 932–938, https://doi.org/10.1212/wnl.0000000000008525.Search in Google Scholar

Benison, A.M., Chumachenko, S., Harrison, J.A., Maier, S.F., Falci, S.P., Watkins, L.R., and Barth, D.S. (2011). Caudal granular insular cortex is sufficient and necessary for the long-term maintenance of allodynic behavior in the rat attributable to mononeuropathy. J. Neurosci. 31: 6317–6328, https://doi.org/10.1523/jneurosci.0076-11.2011.Search in Google Scholar PubMed PubMed Central

Beukema, P., Cecil, K.L., Peterson, E., Mann, V.R., Matsushita, M., Takashima, Y., Navlakha, S., and Barth, A.L. (2018). TrpM8-mediated somatosensation in mouse neocortex. J. Comp. Neurol. 526: 1444–1456, https://doi.org/10.1002/cne.24418.Search in Google Scholar PubMed PubMed Central

Bloodgood, D.W., Sugam, J.A., Holmes, A., and Kash, T.L. (2018). Fear extinction requires infralimbic cortex projections to the basolateral amygdala. Transl. Psychiatry 8: 60, https://doi.org/10.1038/s41398-018-0106-x.Search in Google Scholar PubMed PubMed Central

Bokiniec, P., Zampieri, N., Lewin, G.R., and Poulet, J.F. (2018). The neural circuits of thermal perception. Curr. Opin. Neurobiol. 52: 98–106, https://doi.org/10.1016/j.conb.2018.04.006.Search in Google Scholar PubMed PubMed Central

Bourne, S., Machado, A.G., and Nagel, S.J. (2014). Basic anatomy and physiology of pain pathways. Neurosurg. Clin. N. Am. 25: 629–638, https://doi.org/10.1016/j.nec.2014.06.001.Search in Google Scholar PubMed

Bushnell, M.C., Ceko, M., and Low, L.A. (2013). Cognitive and emotional control of pain and its disruption in chronic pain. Nat. Rev. Neurosci. 14: 502–511, https://doi.org/10.1038/nrn3516.Search in Google Scholar PubMed PubMed Central

Bushnell, M.C., Duncan, G.H., Hofbauer, R.K., Ha, B., Chen, J.I., and Carrier, B. (1999). Pain perception: is there a role for primary somatosensory cortex? Proc. Natl. Acad. Sci. U. S. A. 96: 7705–7709, https://doi.org/10.1073/pnas.96.14.7705.Search in Google Scholar PubMed PubMed Central

Butler, R.K., Nilsson-Todd, L., Cleren, C., Léna, I., Garcia, R., and Finn, D.P. (2011). Molecular and electrophysiological changes in the prefrontal cortex-amygdala-dorsal periaqueductal grey pathway during persistent pain state and fear-conditioned analgesia. Physiol. Behav. 104: 1075–1081, https://doi.org/10.1016/j.physbeh.2011.05.028.Search in Google Scholar PubMed

Cai, H., Haubensak, W., Anthony, T.E., and Anderson, D.J. (2014). Central amygdala PKC-δ(+) neurons mediate the influence of multiple anorexigenic signals. Nat. Neurosci. 17: 1240–1248, https://doi.org/10.1038/nn.3767.Search in Google Scholar PubMed PubMed Central

Cai, Y.Q., Wang, W., Paulucci-Holthauzen, A., and Pan, Z.Z. (2018). Brain circuits mediating opposing effects on emotion and pain. J. Neurosci. 38: 6340–6349, https://doi.org/10.1523/jneurosci.2780-17.2018.Search in Google Scholar PubMed PubMed Central

Cardoso-Cruz, H., Paiva, P., Monteiro, C., and Galhardo, V. (2019). Selective optogenetic inhibition of medial prefrontal glutamatergic neurons reverses working memory deficits induced by neuropathic pain. Pain 160: 805–823, https://doi.org/10.1097/j.pain.0000000000001457.Search in Google Scholar PubMed

Cardoso-Cruz, H., Sousa, M., Vieira, J.B., Lima, D., and Galhardo, V. (2013). Prefrontal cortex and mediodorsal thalamus reduced connectivity is associated with spatial working memory impairment in rats with inflammatory pain. Pain 154: 2397–2406, https://doi.org/10.1016/j.pain.2013.07.020.Search in Google Scholar PubMed

Carlén, M. (2017). What constitutes the prefrontal cortex? Science 358: 478–482, https://doi.org/10.1126/science.aan8868.Search in Google Scholar PubMed

Cavalcanti, M.R.M., Passos, F.R.S., Monteiro, B.S., Gandhi, S.R., Heimfarth, L., Lima, B.S., Nascimento, Y.M., Duarte, M.C., Araujo, A.A.S., Menezes, I.R.A., et al.. (2021). HPLC-DAD-UV analysis, anti-inflammatory and anti-neuropathic effects of methanolic extract of Sideritis bilgeriana (lamiaceae) by NF-κB, TNF-α, IL-1β and IL-6 involvement. J. Ethnopharmacol. 265: 113338, https://doi.org/10.1016/j.jep.2020.113338.Search in Google Scholar PubMed

Chao, T.H., Chen, J.H., and Yen, C.T. (2018). Plasticity changes in forebrain activity and functional connectivity during neuropathic pain development in rats with sciatic spared nerve injury. Mol. Brain 11: 55, https://doi.org/10.1186/s13041-018-0398-z.Search in Google Scholar PubMed PubMed Central

Chen, J.I., Ha, B., Bushnell, M.C., Pike, B., and Duncan, G.H. (2002). Differentiating noxious- and innocuous-related activation of human somatosensory cortices using temporal analysis of fMRI. J. Neurophysiol. 88: 464–474, https://doi.org/10.1152/jn.2002.88.1.464.Search in Google Scholar PubMed

Chen, Q. and Heinricher, M.M. (2022). Shifting the balance: how top-down and bottom-up input modulate pain via the rostral ventromedial medulla. Front. Pain Res. (Lausanne) 3: 932476, https://doi.org/10.3389/fpain.2022.932476.Search in Google Scholar PubMed PubMed Central

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.. (2018). Top-down descending facilitation of spinal sensory excitatory transmission from the anterior cingulate cortex. Nat. Commun. 9: 1886, https://doi.org/10.1038/s41467-018-04309-2.Search in Google Scholar PubMed PubMed Central

Chen, Y.H., Wu, J.L., Hu, N.Y., Zhuang, J.P., Li, W.P., Zhang, S.R., Li, X.W., Yang, J.M., and Gao, T.M. (2021). Distinct projections from the infralimbic cortex exert opposing effects in modulating anxiety and fear. J. Clin. Invest. 131: 2–4, https://doi.org/10.1172/jci145692.Search in Google Scholar PubMed PubMed Central

Chiang, M.C., Bowen, A., Schier, L.A., Tupone, D., Uddin, O., and Heinricher, M.M. (2019). Parabrachial complex: a hub for pain and aversion. J. Neurosci. 39: 8225–8230, https://doi.org/10.1523/jneurosci.1162-19.2019.Search in Google Scholar PubMed PubMed Central

Chiang, M.C., Nguyen, E.K., Canto-Bustos, M., Papale, A.E., Oswald, A.M., and Ross, S.E. (2020). Divergent neural pathways emanating from the lateral parabrachial nucleus mediate distinct components of the pain response. Neuron 106: 927–939, https://doi.org/10.1016/j.neuron.2020.03.014.Search in Google Scholar PubMed

Ching, Y.Y., Wang, C., Tay, T., Loke, Y.M., Tang, P.H., Sng, B.L., and Zhou, J. (2018). Altered sensory insular connectivity in chronic postsurgical pain patients. Front. Hum. Neurosci. 12: 483, https://doi.org/10.3389/fnhum.2018.00483.Search in Google Scholar PubMed PubMed Central

Coghill, R.C., Sang, C.N., Maisog, J.M., and Iadarola, M.J. (1999). Pain intensity processing within the human brain: a bilateral, distributed mechanism. J. Neurophysiol. 82: 1934–1943, https://doi.org/10.1152/jn.1999.82.4.1934.Search in Google Scholar PubMed

Cohen, S.P. and Mao, J. (2014). Neuropathic pain: mechanisms and their clinical implications. Brit. Med. J. 348: f7656, https://doi.org/10.1136/bmj.f7656.Search in Google Scholar PubMed

Cohen, S.P., Vase, L., and Hooten, W.M. (2021). Chronic pain: an update on burden, best practices, and new advances. Lancet 397: 2082–2097, https://doi.org/10.1016/s0140-6736(21)00393-7.Search in Google Scholar

Cottam, W.J., Iwabuchi, S.J., Drabek, M.M., Reckziegel, D., and Auer, D.P. (2018). Altered connectivity of the right anterior insula drives the pain connectome changes in chronic knee osteoarthritis. Pain 159: 929–938, https://doi.org/10.1097/j.pain.0000000000001209.Search in Google Scholar PubMed PubMed Central

Craig, A.D. (2002). How do you feel? Interoception: the sense of the physiological condition of the body. Nat. Rev. Neurosci. 3: 655–666, https://doi.org/10.1038/nrn894.Search in Google Scholar PubMed

Craig, A.D. (2003). Interoception: the sense of the physiological condition of the body. Curr. Opin. Neurobiol. 13: 500–505, https://doi.org/10.1016/s0959-4388(03)00090-4.Search in Google Scholar PubMed

Craig, A.D. (2014). Topographically organized projection to posterior insular cortex from the posterior portion of the ventral medial nucleus in the long-tailed macaque monkey. J. Comp. Neurol. 522: 36–63, https://doi.org/10.1002/cne.23425.Search in Google Scholar PubMed PubMed Central

Dale, J., Zhou, H., Zhang, Q., Martinez, E., Hu, S., Liu, K., Urien, L., Chen, Z., and Wang, J. (2018). Scaling up cortical control inhibits pain. Cell Rep. 23: 1301–1313, https://doi.org/10.1016/j.celrep.2018.03.139.Search in Google Scholar PubMed PubMed Central

David-Pereira, A., Puga, S., Gonçalves, S., Amorim, D., Silva, C., Pertovaara, A., Almeida, A., and Pinto-Ribeiro, F. (2016). Metabotropic glutamate 5 receptor in the infralimbic cortex contributes to descending pain facilitation in healthy and arthritic animals. Neuroscience 312: 108–119, https://doi.org/10.1016/j.neuroscience.2015.10.060.Search in Google Scholar PubMed

David-Pereira, A., Sagalajev, B., Wei, H., Almeida, A., Pertovaara, A., and Pinto-Ribeiro, F. (2017). The medullary dorsal reticular nucleus as a relay for descending pronociception induced by the mGluR5 in the rat infralimbic cortex. Neuroscience 349: 341–354, https://doi.org/10.1016/j.neuroscience.2017.02.046.Search in Google Scholar PubMed

De Ridder, D., Adhia, D., and Vanneste, S. (2021). The anatomy of pain and suffering in the brain and its clinical implications. Neurosci. Biobehav. Rev. 130: 125–146, https://doi.org/10.1016/j.neubiorev.2021.08.013.Search in Google Scholar PubMed

De Ridder, D. and Vanneste, S. (2017). Occipital nerve field transcranial direct current stimulation normalizes imbalance between pain detecting and pain inhibitory pathways in fibromyalgia. Neurotherapeutics 14: 484–501, https://doi.org/10.1007/s13311-016-0493-8.Search in Google Scholar PubMed PubMed Central

Devoize, L., Alvarez, P., Monconduit, L., and Dallel, R. (2011). Representation of dynamic mechanical allodynia in the ventral medial prefrontal cortex of trigeminal neuropathic rats. Eur. J. Pain 15: 676–682, https://doi.org/10.1016/j.ejpain.2010.11.017.Search in Google Scholar PubMed

Donaldson, L.F. and Lumb, B.M. (2017). Top-down control of pain. J. Physiol. 595: 4139–4140, https://doi.org/10.1113/jp273361.Search in Google Scholar PubMed PubMed Central

Drake, R.A., Steel, K.A., Apps, R., Lumb, B.M., and Pickering, A.E. (2021). Loss of cortical control over the descending pain modulatory system determines the development of the neuropathic pain state in rats. Elife 10: 4–7, https://doi.org/10.7554/elife.65156.Search in Google Scholar

Dum, R.P., Levinthal, D.J., and Strick, P.L. (2009). The spinothalamic system targets motor and sensory areas in the cerebral cortex of monkeys. J. Neurosci. 29: 14223–14235, https://doi.org/10.1523/jneurosci.3398-09.2009.Search in Google Scholar PubMed PubMed Central

Eippert, F., Bingel, U., Schoell, E.D., Yacubian, J., Klinger, R., Lorenz, J., and Büchel, C. (2009). Activation of the opioidergic descending pain control system underlies placebo analgesia. Neuron 63: 533–543, https://doi.org/10.1016/j.neuron.2009.07.014.Search in Google Scholar PubMed

Eto, K., Ishibashi, H., Yoshimura, T., Watanabe, M., Miyamoto, A., Ikenaka, K., Moorhouse, A.J., and Nabekura, J. (2012). Enhanced GABAergic activity in the mouse primary somatosensory cortex is insufficient to alleviate chronic pain behavior with reduced expression of neuronal potassium-chloride cotransporter. J. Neurosci. 32: 16552–16559, https://doi.org/10.1523/jneurosci.2104-12.2012.Search in Google Scholar

Fields, H. (2004). State-dependent opioid control of pain. Nat. Rev. Neurosci. 5: 565–575, https://doi.org/10.1038/nrn1431.Search in Google Scholar PubMed

Fields, H.L. (2000). Pain modulation: expectation, opioid analgesia and virtual pain. Prog. Brain Res. 122: 245–253, https://doi.org/10.1016/s0079-6123(08)62143-3.Search in Google Scholar PubMed

François, A., Low, S.A., Sypek, E.I., Christensen, A.J., Sotoudeh, C., Beier, K.T., Ramakrishnan, C., Ritola, K.D., Sharif-Naeini, R., Deisseroth, K., et al.. (2017). A Brainstem-spinal cord inhibitory circuit for mechanical pain modulation by GABA and enkephalins. Neuron 93: 822–839, https://doi.org/10.1016/j.neuron.2017.01.008.Search in Google Scholar PubMed PubMed Central

Friebel, U., Eickhoff, S.B., and Lotze, M. (2011). Coordinate-based meta-analysis of experimentally induced and chronic persistent neuropathic pain. Neuroimage 58: 1070–1080, https://doi.org/10.1016/j.neuroimage.2011.07.022.Search in Google Scholar PubMed PubMed Central

Frøkjær, J.B., Olesen, S.S., Graversen, C., Andresen, T., Lelic, D., and Drewes, A.M. (2018). Neuroimaging of the human visceral pain system-a methodological review. Scand. J. Pain 2: 95–104, https://doi.org/10.1016/j.sjpain.2011.02.006.Search in Google Scholar PubMed

Frot, M., Mauguière, F., Magnin, M., and Garcia-Larrea, L. (2008). Parallel processing of nociceptive A-delta inputs in SII and midcingulate cortex in humans. J. Neurosci. 28: 944–952, https://doi.org/10.1523/jneurosci.2934-07.2008.Search in Google Scholar PubMed PubMed Central

Fuchs, P.N., Peng, Y.B., Boyette-Davis, J.A., and Uhelski, M.L. (2014). The anterior cingulate cortex and pain processing. Front. Integr. Neurosci. 8: 35, https://doi.org/10.3389/fnint.2014.00035.Search in Google Scholar PubMed PubMed Central

Galhardoni, R., Aparecida da Silva, V., García-Larrea, L., Dale, C., Baptista, A.F., Barbosa, L.M., Menezes, L.M.B., de Siqueira, Srdt, Valério, F., Rosi, J.Jr., et al.. (2019). Insular and anterior cingulate cortex deep stimulation for central neuropathic pain: disassembling the percept of pain. Neurology 92: e2165–e75, https://doi.org/10.1212/wnl.0000000000007396.Search in Google Scholar

Gao, S.H., Wen, H.Z., Shen, L.L., Zhao, Y.D., and Ruan, H.Z. (2016). Activation of mGluR1 contributes to neuronal hyperexcitability in the rat anterior cingulate cortex via inhibition of HCN channels. Neuropharmacology 105: 361–377, https://doi.org/10.1016/j.neuropharm.2016.01.036.Search in Google Scholar PubMed

Garro-Martínez, E., Fullana, M.N., Florensa-Zanuy, E., Senserrich, J., Paz, V., Ruiz-Bronchal, E., Adell, A., Castro, E., Díaz, Á., Pazos, Á, et al. (2021). mTOR knockdown in the infralimbic cortex evokes a depressive-like state in mouse. Int. J. Mol. Sci. 22: 1–5, https://doi.org/10.3390/ijms22168671.Search in Google Scholar PubMed PubMed Central

Ge, J., Cai, Y., and Pan, Z.Z. (2022). Synaptic plasticity in two cell types of central amygdala for regulation of emotion and pain. Front. Cell Neurosci. 16: 997360, https://doi.org/10.3389/fncel.2022.997360.Search in Google Scholar PubMed PubMed Central

Giesecke, T., Gracely, R.H., Grant, M.A., Nachemson, A., Petzke, F., Williams, D.A., and Clauw, D.J. (2004). Evidence of augmented central pain processing in idiopathic chronic low back pain. Arthritis Rheum 50: 613–623, https://doi.org/10.1002/art.20063.Search in Google Scholar PubMed

Goadsby, P.J., Holland, P.R., Martins-Oliveira, M., Hoffmann, J., Schankin, C., and Akerman, S. (2017). Pathophysiology of migraine: a disorder of sensory processing. Physiol. Rev. 97: 553–622, https://doi.org/10.1152/physrev.00034.2015.Search in Google Scholar PubMed PubMed Central

Griessner, J., Pasieka, M., Böhm, V., Grössl, F., Kaczanowska, J., Pliota, P., Kargl, D., Werner, B., Kaouane, N., Strobelt, S., et al.. (2021). Central amygdala circuit dynamics underlying the benzodiazepine anxiolytic effect. Mol. Psychiatry 26: 534–544, https://doi.org/10.1038/s41380-018-0310-3.Search in Google Scholar PubMed PubMed Central

Gustin, S.M., Wrigley, P.J., Youssef, A.M., McIndoe, L., Wilcox, S.L., Rae, C.D., Edden, R.A.E., Siddall, P.J., and Henderson, L.A. (2014). Thalamic activity and biochemical changes in individuals with neuropathic pain after spinal cord injury. Pain 155: 1027–1036, https://doi.org/10.1016/j.pain.2014.02.008.Search in Google Scholar PubMed PubMed Central

Han, J., Cha, M., Kwon, M., Hong, S.K., Bai, S.J., and Lee, B.H. (2016). In vivo voltage-sensitive dye imaging of the insular cortex in nerve-injured rats. Neurosci. Lett. 634: 146–152, https://doi.org/10.1016/j.neulet.2016.10.015.Search in Google Scholar PubMed

Hao, S., Yang, H., Wang, X., He, Y., Xu, H., Wu, X., Pan, L., Liu, Y., Lou, H., Xu, H., et al.. (2019). The lateral hypothalamic and BNST GABAergic projections to the anterior ventrolateral periaqueductal gray regulate feeding. Cell Rep. 28: 616–624, https://doi.org/10.1016/j.celrep.2019.06.051.Search in Google Scholar PubMed

Hashmi, J.A., Baliki, M.N., Huang, L., Baria, A.T., Torbey, S., Hermann, K.M., Schnitzer, T.J., and Apkarian, A.V. (2013). Shape shifting pain: chronification of back pain shifts brain representation from nociceptive to emotional circuits. Brain 136: 2751–2768, https://doi.org/10.1093/brain/awt211.Search in Google Scholar PubMed PubMed Central

Heinricher, M.M., Morgan, M.M., Tortorici, V., and Fields, H.L. (1994). Disinhibition of off-cells and antinociception produced by an opioid action within the rostral ventromedial medulla. Neuroscience 63: 279–288, https://doi.org/10.1016/0306-4522(94)90022-1.Search in Google Scholar PubMed

Heinricher, M.M., Tavares, I., Leith, J.L., and Lumb, B.M. (2009). Descending control of nociception: specificity, recruitment and plasticity. Brain Res. Rev. 60: 214–225, https://doi.org/10.1016/j.brainresrev.2008.12.009.Search in Google Scholar PubMed PubMed Central

Hogri, R., Teuchmann, H.L., Heinke, B., Holzinger, R., Trofimova, L., and Sandkühler, J. (2022). GABAergic CaMKIIα+ amygdala output attenuates pain and modulates emotional-motivational behavior via parabrachial inhibition. J. Neurosci. 42: 5373–5388, https://doi.org/10.1523/jneurosci.2067-21.2022.Search in Google Scholar PubMed PubMed Central

Hon, O.J., DiBerto, J.F., Mazzone, C.M., Sugam, J., Bloodgood, D.W., Hardaway, J.A., Husain, M., Kendra, A., McCall, N.M., Lopez, A.J., et al.. (2022). Serotonin modulates an inhibitory input to the central amygdala from the ventral periaqueductal gray. Neuropsychopharmacology 47: 2194–2204, https://doi.org/10.1038/s41386-022-01392-4.Search in Google Scholar PubMed PubMed Central

Hsieh, P.C., Tseng, M.T., Chao, C.C., Lin, Y.H., Tseng, W.I., Liu, K.H., Chiang, M.C., and Hsieh, S.T. (2015). Imaging signatures of altered brain responses in small-fiber neuropathy: reduced functional connectivity of the limbic system after peripheral nerve degeneration. Pain 156: 904–916, https://doi.org/10.1097/j.pain.0000000000000128.Search in Google Scholar PubMed

Huang, D., Grady, F.S., Peltekian, L., Laing, J.J., and Geerling, J.C. (2021). Efferent projections of CGRP/Calca-expressing parabrachial neurons in mice. J. Comp. Neurol. 529: 2911–2957, https://doi.org/10.1002/cne.25136.Search in Google Scholar PubMed PubMed Central

Huang, J., Gadotti, V.M., Chen, L., Souza, I.A., Huang, S., Wang, D., Ramakrishnan, C., Deisseroth, K., Zhang, Z., and Zamponi, G.W. (2019). A neuronal circuit for activating descending modulation of neuropathic pain. Nat. Neurosci. 22: 1659–1668, https://doi.org/10.1038/s41593-019-0481-5.Search in Google Scholar PubMed

Hwang, K., Bertolero, M.A., Liu, W.B., and D’Esposito, M. (2017). The Human thalamus is an integrative hub for functional brain networks. J. Neurosci. 37: 5594–5607, https://doi.org/10.1523/jneurosci.0067-17.2017.Search in Google Scholar

Inami, C., Tanihira, H., Kikuta, S., Ogasawara, O., Sobue, K., Kume, K., Osanai, M., and Ohsawa, M. (2019). Visualization of brain activity in a neuropathic pain model using quantitative activity-dependent manganese magnetic resonance imaging. Front. Neural. Circuits 13: 74, https://doi.org/10.3389/fncir.2019.00074.Search in Google Scholar PubMed PubMed Central

Jarrin, S., Pandit, A., Roche, M., and Finn, D.P. (2020). Differential role of anterior cingulate cortical glutamatergic neurons in pain-related aversion learning and nociceptive behaviors in male and female rats. Front. Behav. Neurosci. 14: 139, https://doi.org/10.3389/fnbeh.2020.00139.Search in Google Scholar PubMed PubMed Central

Jhang, J., Lee, H., Kang, M.S., Lee, H.S., Park, H., and Han, J.H. (2018). Anterior cingulate cortex and its input to the basolateral amygdala control innate fear response. Nat. Commun. 9: 2744, https://doi.org/10.1038/s41467-018-05090-y.Search in Google Scholar PubMed PubMed Central

Ji, G., Sun, H., Fu, Y., Li, Z., Pais-Vieira, M., Galhardo, V., and Neugebauer, V. (2010). Cognitive impairment in pain through amygdala-driven prefrontal cortical deactivation. J. Neurosci. 30: 5451–5464, https://doi.org/10.1523/jneurosci.0225-10.2010.Search in Google Scholar

Jiang, Z.C., Pan, Q., Zheng, C., Deng, X.F., Wang, J.Y., and Luo, F. (2014). Inactivation of the prelimbic rather than infralimbic cortex impairs acquisition and expression of formalin-induced conditioned place avoidance. Neurosci. Lett. 569: 89–93, https://doi.org/10.1016/j.neulet.2014.03.074.Search in Google Scholar PubMed PubMed Central

Johansen, J.P. and Fields, H.L. (2004). Glutamatergic activation of anterior cingulate cortex produces an aversive teaching signal. Nat. Neurosci. 7: 398–403, https://doi.org/10.1038/nn1207.Search in Google Scholar PubMed

Jones, A.F. and Sheets, P.L. (2020). Sex-specific disruption of distinct mPFC inhibitory neurons in spared-nerve injury model of neuropathic pain. Cell Rep. 31: 107729, https://doi.org/10.1016/j.celrep.2020.107729.Search in Google Scholar PubMed PubMed Central

Joo, S.Y., Park, C.H., Cho, Y.S., Seo, C.H., and Ohn, S.H. (2021). Plastic changes in pain and motor network induced by chronic burn pain. J. Clin. Med. 10: 6–8, https://doi.org/10.3390/jcm10122592.Search in Google Scholar PubMed PubMed Central

Juarez-Salinas, D.L., Braz, J.M., Etlin, A., Gee, S., Sohal, V., and Basbaum, A.I. (2019). GABAergic cell transplants in the anterior cingulate cortex reduce neuropathic pain aversiveness. Brain 142: 2655–2669, https://doi.org/10.1093/brain/awz203.Search in Google Scholar PubMed PubMed Central

Kato, F., Sugimura, Y.K., and Takahashi, Y. (2018). Pain-associated neural plasticity in the parabrachial to central amygdala circuit : pain changes the brain, and the brain changes the pain. Adv. Exp. Med. Biol. 1099: 157–166, https://doi.org/10.1007/978-981-13-1756-9_14.Search in Google Scholar PubMed

Kelly, C.J., Huang, M., Meltzer, H., and Martina, M. (2016). 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. 10: 133, https://doi.org/10.3389/fncel.2016.00133.Search in Google Scholar PubMed PubMed Central

Kelly, C.J. and Martina, M. (2018). Circuit-selective properties of glutamatergic inputs to the rat prelimbic cortex and their alterations in neuropathic pain. Brain Struct. Funct. 223: 2627–2639, https://doi.org/10.1007/s00429-018-1648-7.Search in Google Scholar PubMed PubMed Central

Kikkert, S., Mezue, M., O’Shea, J., Henderson Slater, D., Johansen-Berg, H., Tracey, I., and Makin, T.R. (2019). Neural basis of induced phantom limb pain relief. Ann. Neurol. 85: 59–73, https://doi.org/10.1002/ana.25371.Search in Google Scholar PubMed PubMed Central

Kim, J., Pignatelli, M., Xu, S., Itohara, S., and Tonegawa, S. (2016). Antagonistic negative and positive neurons of the basolateral amygdala. Nat. Neurosci. 19: 1636–1646, https://doi.org/10.1038/nn.4414.Search in Google Scholar PubMed PubMed Central

Kim, J., Zhang, X., Muralidhar, S., LeBlanc, S.A., and Tonegawa, S. (2017). Basolateral to central amygdala neural circuits for appetitive behaviors. Neuron 93: 1464–1479, https://doi.org/10.1016/j.neuron.2017.02.034.Search in Google Scholar

Kim, S.K., Hayashi, H., Ishikawa, T., Shibata, K., Shigetomi, E., Shinozaki, Y., Inada, H., Roh, S.E., Kim, S.J., Lee, G., et al.. (2016). Cortical astrocytes rewire somatosensory cortical circuits for peripheral neuropathic pain. J. Clin. Invest. 126: 1983–1997, https://doi.org/10.1172/jci82859.Search in Google Scholar

Kim, W., Kim, S.K., and Nabekura, J. (2017). Functional and structural plasticity in the primary somatosensory cortex associated with chronic pain. J. Neurochem. 141: 499–506, https://doi.org/10.1111/jnc.14012.Search in Google Scholar

Kiritoshi, T., Ji, G., and Neugebauer, V. (2016). Rescue of impaired mGluR5-driven endocannabinoid signaling restores prefrontal cortical output to inhibit pain in arthritic rats. J. Neurosci. 36: 837–850, https://doi.org/10.1523/jneurosci.4047-15.2016.Search in Google Scholar

Koga, K., Descalzi, G., Chen, T., Ko, H.G., Lu, J., Li, S., Son, J., Kim, T., Kwak, C., Huganir, R.L., et al.. (2015). Coexistence of two forms of LTP in ACC provides a synaptic mechanism for the interactions between anxiety and chronic pain. Neuron 85: 377–389, https://doi.org/10.1016/j.neuron.2015.05.016.Search in Google Scholar

Kong, Q.M., Qiao, H., Liu, C.Z., Zhang, P., Li, K., Wang, L., Li, J.T., Su, Y., Li, K.Q., Yan, C.G., et al.. (2018). Aberrant intrinsic functional connectivity in thalamo-cortical networks in major depressive disorder. CNS Neurosci. Ther. 24: 1063–1072, https://doi.org/10.1111/cns.12831.Search in Google Scholar

Koutsikou, S., Watson, T.C., Crook, J.J., Leith, J.L., Lawrenson, C.L., Apps, R., and Lumb, B.M. (2015). The periaqueductal gray orchestrates sensory and motor circuits at multiple levels of the neuraxis. J. Neurosci. 35: 14132–14147, https://doi.org/10.1523/jneurosci.0261-15.2015.Search in Google Scholar

Krout, K.E., Jansen, A.S., and Loewy, A.D. (1998). Periaqueductal gray matter projection to the parabrachial nucleus in rat. J. Comp. Neurol. 401: 437–454, https://doi.org/10.1002/(sici)1096-9861(19981130)401:4<437::aid-cne2>3.0.co;2-5.10.1002/(SICI)1096-9861(19981130)401:4<437::AID-CNE2>3.0.CO;2-5Search in Google Scholar

Kuner, R. and Kuner, T. (2021). Cellular circuits in the brain and their modulation in acute and chronic pain. Physiol. Rev. 101: 213–258, https://doi.org/10.1152/physrev.00040.2019.Search in Google Scholar

Kwon, M., Altin, M., Duenas, H., and Alev, L. (2014). The role of descending inhibitory pathways on chronic pain modulation and clinical implications. Pain Pract. 14: 656–667, https://doi.org/10.1111/papr.12145.Search in Google Scholar

Lanius, R.A., Boyd, J.E., McKinnon, M.C., Nicholson, A.A., Frewen, P., Vermetten, E., Jetly, R., and Spiegel, D. (2018). A review of the neurobiological basis of trauma-related dissociation and its relation to cannabinoid- and opioid-mediated stress response: a transdiagnostic, translational approach. Curr. Psychiatry Rep. 20: 118, https://doi.org/10.1007/s11920-018-0983-y.Search in Google Scholar PubMed

Lau, B.K. and Vaughan, C.W. (2014). Descending modulation of pain: the GABA disinhibition hypothesis of analgesia. Curr. Opin. Neurobiol. 29: 159–164, https://doi.org/10.1016/j.conb.2014.07.010.Search in Google Scholar PubMed

Laubach, M., Amarante, L.M., Swanson, K., and White, S.R. (2018). What, if anything, is rodent prefrontal cortex? eNeuro 5: 5–6, https://doi.org/10.1523/eneuro.0315-18.2018.Search in Google Scholar PubMed PubMed Central

LeDoux, J. (2007). The amygdala. Curr. Biol. 17: R868–R874, https://doi.org/10.1016/j.cub.2007.08.005.Search in Google Scholar PubMed

Lee, J.Y., You, T., Lee, C.H., Im, G.H., Seo, H., Woo, C.W., and Kim, S.G. (2022). Role of anterior cingulate cortex inputs to periaqueductal gray for pain avoidance. Curr. Biol. 32: 2834–2847, https://doi.org/10.1016/j.cub.2022.04.090.Search in Google Scholar PubMed

Lenz, F.A., Weiss, N., Ohara, S., Lawson, C., and Greenspan, J.D. (2004). The role of the thalamus in pain. Suppl. Clin. Neurophysiol. 57: 50–61, https://doi.org/10.1016/s1567-424x(09)70342-3.Search in Google Scholar PubMed

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.. (2010). Alleviating neuropathic pain hypersensitivity by inhibiting PKMzeta in the anterior cingulate cortex. Science 330: 1400–1404, https://doi.org/10.1126/science.1191792.Search in Google Scholar PubMed

Liang, S.H., Zhao, W.J., Yin, J.B., Chen, Y.B., Li, J.N., Feng, B., Lu, Y.C., Wang, J., Dong, Y.L., and Li, Y.Q. (2020). A neural circuit from thalamic paraventricular nucleus to central amygdala for the facilitation of neuropathic pain. J. Neurosci. 40: 7837–7854, https://doi.org/10.1523/jneurosci.2487-19.2020.Search in Google Scholar

Llinás, R.R., Ribary, U., Jeanmonod, D., Kronberg, E., and Mitra, P.P. (1999). Thalamocortical dysrhythmia: a neurological and neuropsychiatric syndrome characterized by magnetoencephalography. Proc. Natl. Acad. Sci. U. S. A. 96: 15222–15227, https://doi.org/10.1073/pnas.96.26.15222.Search in Google Scholar PubMed PubMed Central

Loeser, J.D. and Melzack, R. (1999). Pain: an overview. Lancet 353: 1607–1609, https://doi.org/10.1016/s0140-6736(99)01311-2.Search in Google Scholar PubMed

Lu, C., Yang, T., Zhao, H., Zhang, M., Meng, F., Fu, H., Xie, Y., and Xu, H. (2016). Insular cortex is critical for the perception, modulation, and chronification of pain. Neurosci. Bull. 32: 191–201, https://doi.org/10.1007/s12264-016-0016-y.Search in Google Scholar PubMed PubMed Central

Luongo, L., de Novellis, V., Gatta, L., Palazzo, E., Vita, D., Guida, F., Giordano, C., Siniscalco, D., Marabese, I., De Chiaro, M., et al.. (2013). Role of metabotropic glutamate receptor 1 in the basolateral amygdala-driven prefrontal cortical deactivation in inflammatory pain in the rat. Neuropharmacology 66: 317–329, https://doi.org/10.1016/j.neuropharm.2012.05.047.Search in Google Scholar PubMed

Maihöfner, C. and Handwerker, H.O. (2005). Differential coding of hyperalgesia in the human brain: a functional MRI study. Neuroimage 28: 996–1006, https://doi.org/10.1016/j.neuroimage.2005.06.049.Search in Google Scholar PubMed

Mao, C.P., Yang, H.J., Yang, Q.X., Sun, H.H., Zhang, G.R., and Zhang, Q.J. (2022). Altered amygdala-prefrontal connectivity in chronic nonspecific low back pain: resting-state fMRI and dynamic causal modelling study. Neuroscience 482: 18–29, https://doi.org/10.1016/j.neuroscience.2021.12.003.Search in Google Scholar PubMed

Marek, R., Xu, L., Sullivan, R.K.P., and Sah, P. (2018). Excitatory connections between the prelimbic and infralimbic medial prefrontal cortex show a role for the prelimbic cortex in fear extinction. Nat. Neurosci. 21: 654–658, https://doi.org/10.1038/s41593-018-0137-x.Search in Google Scholar PubMed

Martenson, M.E., Cetas, J.S., and Heinricher, M.M. (2009). A possible neural basis for stress-induced hyperalgesia. Pain 142: 236–244, https://doi.org/10.1016/j.pain.2009.01.011.Search in Google Scholar PubMed PubMed Central

Matsumoto, N., Bester, H., Menendez, L., Besson, J.M., and Bernard, J.F. (1996). Changes in the responsiveness of parabrachial neurons in the arthritic rat: an electrophysiological study. J. Neurophysiol. 76: 4113–4126, https://doi.org/10.1152/jn.1996.76.6.4113.Search in Google Scholar PubMed

May, A. (2008). Chronic pain may change the structure of the brain. Pain 137: 7–15, https://doi.org/10.1016/j.pain.2008.02.034.Search in Google Scholar PubMed

May, A. (2011). Structural brain imaging: a window into chronic pain. Neuroscientist 17: 209–220, https://doi.org/10.1177/1073858410396220.Search in Google Scholar PubMed

Mazzitelli, M., Marshall, K., Pham, A., Ji, G., and Neugebauer, V. (2021). Optogenetic manipulations of amygdala neurons modulate spinal nociceptive processing and behavior under normal conditions and in an arthritis pain model. Front. Pharmacol. 12: 668337, https://doi.org/10.3389/fphar.2021.668337.Search in Google Scholar PubMed PubMed Central

Mazzitelli, M., Yakhnitsa, V., Neugebauer, B., and Neugebauer, V. (2022). Optogenetic manipulations of CeA-CRF neurons modulate pain- and anxiety-like behaviors in neuropathic pain and control rats. Neuropharmacology 210: 109031, https://doi.org/10.1016/j.neuropharm.2022.109031.Search in Google Scholar PubMed PubMed Central

Mazzola, L., Isnard, J., Peyron, R., Guénot, M., and Mauguière, F. (2009). Somatotopic organization of pain responses to direct electrical stimulation of the human insular cortex. Pain 146: 99–104, https://doi.org/10.1016/j.pain.2009.07.014.Search in Google Scholar PubMed

McPherson, K.B. and Ingram, S.L. (2022). Cellular and circuit diversity determines the impact of endogenous opioids in the descending pain modulatory pathway. Front. Syst. Neurosci. 16: 963812, https://doi.org/10.3389/fnsys.2022.963812.Search in Google Scholar PubMed PubMed Central

Meda, K.S., Patel, T., Braz, J.M., Malik, R., Turner, M.L., Seifikar, H., Basbaum, A.I., and Sohal, V.S. (2019). Microcircuit mechanisms through which mediodorsal thalamic input to anterior cingulate cortex exacerbates pain-related aversion. Neuron 102: 944–959, https://doi.org/10.1016/j.neuron.2019.03.042.Search in Google Scholar PubMed PubMed Central

Melzack, R. (1999). From the gate to the neuromatrix. Pain (Suppl. 6) 3: S121–S26, https://doi.org/10.1016/s0304-3959(99)00145-1.Search in Google Scholar PubMed

Melzack, R. (2001). Pain and the neuromatrix in the brain. J Dent Educ 65: 1378–1382, https://doi.org/10.1002/j.0022-0337.2001.65.12.tb03497.x.Search in Google Scholar

Meng, X., Yue, L., Liu, A., Tao, W., Shi, L., Zhao, W., Wu, Z., Zhang, Z., Wang, L., Zhang, X., et al.. (2022). Distinct basolateral amygdala excitatory inputs mediate the somatosensory and aversive-affective components of pain. J. Biol. Chem. 298: 102207, https://doi.org/10.1016/j.jbc.2022.102207.Search in Google Scholar PubMed PubMed Central

Millan, M.J. (2002). Descending control of pain. Prog. Neurobiol. 66: 355–474, https://doi.org/10.1016/s0301-0082(02)00009-6.Search in Google Scholar PubMed

Miller Neilan, R., Majetic, G., Gil-Silva, M., Adke, A.P., Carrasquillo, Y., and Kolber, B.J. (2021). Agent-based modeling of the central amygdala and pain using cell-type specific physiological parameters. PLoS Comput. Biol. 17: e1009097, https://doi.org/10.1371/journal.pcbi.1009097.Search in Google Scholar PubMed PubMed Central

Monroe, T.B., Fillingim, R.B., Bruehl, S.P., Rogers, B.P., Dietrich, M.S., Gore, J.C., Atalla, S.W., and Cowan, R.L. (2018). Sex differences in brain regions modulating pain among older adults: a cross-sectional resting state functional connectivity study. Pain Med. 19: 1737–1747, https://doi.org/10.1093/pm/pnx084.Search in Google Scholar PubMed PubMed Central

Morgan, M.M. and Fields, H.L. (1994). Pronounced changes in the activity of nociceptive modulatory neurons in the rostral ventromedial medulla in response to prolonged thermal noxious stimuli. J. Neurophysiol. 72: 1161–1170, https://doi.org/10.1152/jn.1994.72.3.1161.Search in Google Scholar PubMed

Morton, D.L., Sandhu, J.S., and Jones, A.K. (2016). Brain imaging of pain: state of the art. J. Pain Res. 9: 613–624, https://doi.org/10.2147/jpr.s60433.Search in Google Scholar PubMed PubMed Central

Mukherjee, A. and Caroni, P. (2019). Author correction: infralimbic cortex is required for learning alternatives to prelimbic promoted associations through reciprocal connectivity. Nat. Commun. 10: 3082, https://doi.org/10.1038/s41467-019-11205-w.Search in Google Scholar PubMed PubMed Central

Nagasaka, K., Takashima, I., Matsuda, K., and Higo, N. (2017). Late-onset hypersensitivity after a lesion in the ventral posterolateral nucleus of the thalamus: a macaque model of central post-stroke pain. Sci. Rep. 7: 10316, https://doi.org/10.1038/s41598-017-10679-2.Search in Google Scholar PubMed PubMed Central

Nardone, R., Höller, Y., Sebastianelli, L., Versace, V., Saltuari, L., Brigo, F., Lochner, P., and Trinka, E. (2018). Cortical morphometric changes after spinal cord injury. Brain Res. Bull. 137: 107–119, https://doi.org/10.1016/j.brainresbull.2017.11.013.Search in Google Scholar PubMed

Neubert, M.J., Kincaid, W., and Heinricher, M.M. (2004). Nociceptive facilitating neurons in the rostral ventromedial medulla. Pain 110: 158–165, https://doi.org/10.1016/j.pain.2004.03.017.Search in Google Scholar PubMed

Neugebauer, V. (2020). Amygdala physiology in pain. Handb. Behav. Neurosci. 26: 101–113, https://doi.org/10.1016/b978-0-12-815134-1.00004-0.Search in Google Scholar PubMed PubMed Central

Neugebauer, V., Li, W., Bird, G.C., and Han, J.S. (2004). The amygdala and persistent pain. Neuroscientist 10: 221–234, https://doi.org/10.1177/1073858403261077.Search in Google Scholar PubMed

Neugebauer, V., Mazzitelli, M., Cragg, B., Ji, G., Navratilova, E., and Porreca, F. (2020). Amygdala, neuropeptides, and chronic pain-related affective behaviors. Neuropharmacology 170: 108052, https://doi.org/10.1016/j.neuropharm.2020.108052.Search in Google Scholar PubMed PubMed Central

Neumann, L., Wulms, N., Witte, V., Spisak, T., Zunhammer, M., Bingel, U., and Schmidt-Wilcke, T. (2021). Network properties and regional brain morphology of the insular cortex correlate with individual pain thresholds. Hum. Brain Mapp. 42: 4896–4908, https://doi.org/10.1002/hbm.25588.Search in Google Scholar PubMed PubMed Central

Nguyen, E., Smith, K.M., Cramer, N., Holland, R.A., Bleimeister, I.H., Flores-Felix, K., Silberberg, H., Keller, A., Le Pichon, C.E., and Ross, S.E. (2022). Medullary kappa-opioid receptor neurons inhibit pain and itch through a descending circuit. Brain 145: 2586–2601, https://doi.org/10.1093/brain/awac189.Search in Google Scholar PubMed PubMed Central

Okada, T., Kato, D., Nomura, Y., Obata, N., Quan, X., Morinaga, A., Yano, H., Guo, Z., Aoyama, Y., Tachibana, Y., et al. (2021). Pain induces stable, active microcircuits in the somatosensory cortex that provide a therapeutic target. Sci. Adv. 7: 6–8, https://doi.org/10.1126/sciadv.abd8261.Search in Google Scholar PubMed PubMed Central

Osborne, N.R., Cheng, J.C., Rogachov, A., Kim, J.A., Hemington, K.S., Bosma, R.L., Inman, R.D., and Davis, K.D. (2021). Abnormal subgenual anterior cingulate circuitry is unique to women but not men with chronic pain. Pain 162: 97–108, https://doi.org/10.1097/j.pain.0000000000002016.Search in Google Scholar PubMed

Ossipov, M.H., Dussor, G.O., and Porreca, F. (2010). Central modulation of pain. J. Clin. Invest. 120: 3779–3787, https://doi.org/10.1172/jci43766.Search in Google Scholar

Ossipov, M.H., Morimura, K., and Porreca, F. (2014). Descending pain modulation and chronification of pain. Curr. Opin. Support Palliat. Care 8: 143–151, https://doi.org/10.1097/spc.0000000000000055.Search in Google Scholar

Ostrowsky, K., Magnin, M., Ryvlin, P., Isnard, J., Guenot, M., and Mauguière, F. (2002). Representation of pain and somatic sensation in the human insula: a study of responses to direct electrical cortical stimulation. Cereb. Cortex 12: 376–385, https://doi.org/10.1093/cercor/12.4.376.Search in Google Scholar PubMed

Otsu, Y. and Aubrey, K.R. (2022). Kappa opioids inhibit the GABA/glycine terminals of rostral ventromedial medulla projections in the superficial dorsal horn of the spinal cord. J. Physiol. 600: 4187–4205, https://doi.org/10.1113/jp283021.Search in Google Scholar

Pare, D. and Duvarci, S. (2012). Amygdala microcircuits mediating fear expression and extinction. Curr. Opin. Neurobiol. 22: 717–723, https://doi.org/10.1016/j.conb.2012.02.014.Search in Google Scholar PubMed PubMed Central

Pauli, J.L., Chen, J.Y., Basiri, M.L., Park, S., Carter, M.E., Sanz, E., McKnight, G.S., Stuber, G.D., and Palmiter, R.D. (2022). Molecular and anatomical characterization of parabrachial neurons and their axonal projections. Elife 11: 5–7.10.7554/eLife.81868Search in Google Scholar PubMed PubMed Central

Peyron, R., Laurent, B., and García-Larrea, L. (2000). Functional imaging of brain responses to pain. A review and meta-analysis (2000), Neurophysiol. Clin. 30: 263–288, https://doi.org/10.1016/s0987-7053(00)00227-6.Search in Google Scholar PubMed

Pieretti, S., Di Giannuario, A., Di Giovannandrea, R., Marzoli, F., Piccaro, G., Minosi, P., and Aloisi, A.M. (2016). Gender differences in pain and its relief. Ann. Ist. Super Sanita. 52: 184–189, https://doi.org/10.4415/ANN_16_02_09.Search in Google Scholar PubMed

Ploghaus, A., Tracey, I., Gati, J.S., Clare, S., Menon, R.S., Matthews, P.M., and Rawlins, J.N. (1999). Dissociating pain from its anticipation in the human brain. Science 284: 1979–1981, https://doi.org/10.1126/science.284.5422.1979.Search in Google Scholar PubMed

Ploner, M., Schmitz, F., Freund, H.J., and Schnitzler, A. (2000). Differential organization of touch and pain in human primary somatosensory cortex. J. Neurophysiol. 83: 1770–1776, https://doi.org/10.1152/jn.2000.83.3.1770.Search in Google Scholar PubMed

Presto, P. and Neugebauer, V. (2022). Sex differences in CGRP regulation and function in the amygdala in a rat model of neuropathic pain. Front. Mol. Neurosci. 15: 928587, https://doi.org/10.3389/fnmol.2022.928587.Search in Google Scholar PubMed PubMed Central

Raja, S.N., Carr, D.B., Cohen, M., Finnerup, N.B., Flor, H., Gibson, S., Keefe, F.J., Mogil, J.S., Ringkamp, M., Sluka, K.A., et al.. (2020). The revised International Association for the Study of Pain definition of pain: concepts, challenges, and compromises. Pain 161: 1976–1982, https://doi.org/10.1097/j.pain.0000000000001939.Search in Google Scholar PubMed PubMed Central

Raver, C., Uddin, O., Ji, Y., Li, Y., Cramer, N., Jenne, C., Morales, M., Masri, R., and Keller, A. (2020). An amygdalo-parabrachial pathway regulates pain perception and chronic pain. J. Neurosci. 40: 3424–3442, https://doi.org/10.1523/jneurosci.0075-20.2020.Search in Google Scholar

Ren, J., Xiang, J., Chen, Y., Li, F., Wu, T., and Shi, J. (2019). Abnormal functional connectivity under somatosensory stimulation in migraine: a multi-frequency magnetoencephalography study. J. Headache Pain 20: 3, https://doi.org/10.1186/s10194-019-0958-3.Search in Google Scholar PubMed PubMed Central

Roeder, Z., Chen, Q., Davis, S., Carlson, J.D., Tupone, D., and Heinricher, M.M. (2016). Parabrachial complex links pain transmission to descending pain modulation. Pain 157: 2697–2708, https://doi.org/10.1097/j.pain.0000000000000688.Search in Google Scholar PubMed PubMed Central

Rosenfeld, M.G., Mermod, J.J., Amara, S.G., Swanson, L.W., Sawchenko, P.E., Rivier, J., Vale, W.W., and Evans, R.M. (1983). Production of a novel neuropeptide encoded by the calcitonin gene via tissue-specific RNA processing. Nature 304: 129–135, https://doi.org/10.1038/304129a0.Search in Google Scholar PubMed

Saadé, N.E., Al Amin, H., Abdel Baki, S., Safieh-Garabedian, B., Atweh, S.F., and Jabbur, S.J. (2006). Transient attenuation of neuropathic manifestations in rats following lesion or reversible block of the lateral thalamic somatosensory nuclei. Exp. Neurol. 197: 157–166, https://doi.org/10.1016/j.expneurol.2005.09.005.Search in Google Scholar PubMed

Sadler, K.E., McQuaid, N.A., Cox, A.C., Behun, M.N., Trouten, A.M., and Kolber, B.J. (2017). Divergent functions of the left and right central amygdala in visceral nociception. Pain 158: 747–759, https://doi.org/10.1097/j.pain.0000000000000830.Search in Google Scholar PubMed PubMed Central

Sah, P., Faber, E.S., Lopez De Armentia, M., and Power, J. (2003). The amygdaloid complex: anatomy and physiology. Physiol. Rev. 83: 803–834, https://doi.org/10.1152/physrev.00002.2003.Search in Google Scholar PubMed

Samineni, V.K., Grajales-Reyes, J.G., Copits, B.A., O’Brien, D.E., Trigg, S.L., Gomez, A.M., Bruchas, M.R., and Gereau, R.W.th. (2017). Divergent modulation of nociception by glutamatergic and GABAergic neuronal subpopulations in the periaqueductal gray. eNeuro 4: 4–9, https://doi.org/10.1523/eneuro.0129-16.2017.Search in Google Scholar

Santello, M. and Nevian, T. (2015). Dysfunction of cortical dendritic integration in neuropathic pain reversed by serotoninergic neuromodulation. Neuron 86: 233–246, https://doi.org/10.1016/j.neuron.2015.03.003.Search in Google Scholar PubMed

Saper, C.B. and Loewy, A.D. (2016). Commentary on: efferent connections of the parabrachial nucleus in the rat. C.B. Saper and A.D. Loewy, Brain Research 197: 291–317, 1980. Brain Res. 197: 1645: 15–7. https://doi.org/10.1016/0006-8993(80)91117-8.Search in Google Scholar PubMed

Schnitzler, A. and Ploner, M. (2000). Neurophysiology and functional neuroanatomy of pain perception. J. Clin. Neurophysiol. 17: 592–603, https://doi.org/10.1097/00004691-200011000-00005.Search in Google Scholar PubMed

Segerdahl, A.R., Mezue, M., Okell, T.W., Farrar, J.T., and Tracey, I. (2015). The dorsal posterior insula subserves a fundamental role in human pain. Nat. Neurosci. 18: 499–500, https://doi.org/10.1038/nn.3969.Search in Google Scholar PubMed PubMed Central

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.. (2018). Hyperactivity of anterior cingulate cortex areas 24a/24b drives chronic pain-induced anxiodepressive-like consequences. J. Neurosci. 38: 3102–3115, https://doi.org/10.1523/jneurosci.3195-17.2018.Search in Google Scholar PubMed PubMed Central

Seminowicz, D.A. and Moayedi, M. (2017). The dorsolateral prefrontal cortex in acute and chronic pain. J. Pain 18: 1027–1035, https://doi.org/10.1016/j.jpain.2017.03.008.Search in Google Scholar PubMed PubMed Central

Senn, V., Wolff, S.B., Herry, C., Grenier, F., Ehrlich, I., Gründemann, J., Fadok, J.P., Müller, C., Letzkus, J.J., and Lüthi, A. (2014). Long-range connectivity defines behavioral specificity of amygdala neurons. Neuron 81: 428–437, https://doi.org/10.1016/j.neuron.2013.11.006.Search in Google Scholar PubMed

Shinohara, K., Watabe, A.M., Nagase, M., Okutsu, Y., Takahashi, Y., Kurihara, H., and Kato, F. (2017). Essential role of endogenous calcitonin gene-related peptide in pain-associated plasticity in the central amygdala. Eur. J. Neurosci. 46: 2149–2160, https://doi.org/10.1111/ejn.13662.Search in Google Scholar PubMed PubMed Central

Singer, T., Critchley, H.D., and Preuschoff, K. (2009). A common role of insula in feelings, empathy and uncertainty. Trends Cogn. Sci. 13: 334–340, https://doi.org/10.1016/j.tics.2009.05.001.Search in Google Scholar PubMed

Singh, A., Patel, D., Li, A., Hu, L., Zhang, Q., Liu, Y., Guo, X., Robinson, E., Martinez, E., Doan, L., et al.. (2020). Mapping cortical integration of sensory and affective pain pathways. Curr. Biol. 30: 1703–1715, https://doi.org/10.1016/j.cub.2020.02.091.Search in Google Scholar PubMed PubMed Central

Sun, L., Liu, R., Guo, F., Wen, M.Q., Ma, X.L., Li, K.Y., Sun, H., Xu, C.L., Li, Y.Y., Wu, M.Y., et al.. (2020a). Parabrachial nucleus circuit governs neuropathic pain-like behavior. Nat. Commun. 11: 5974, https://doi.org/10.1038/s41467-020-19767-w.Search in Google Scholar PubMed PubMed Central

Sun, Y., Wang, J., Liang, S.H., Ge, J., Lu, Y.C., Li, J.N., Chen, Y.B., Luo, D.S., Li, H., and Li, Y.Q. (2020b). Involvement of the ventrolateral periaqueductal gray matter-central medial thalamic nucleus-basolateral amygdala pathway in neuropathic pain regulation of rats. Front. Neuroanat. 14: 32, https://doi.org/10.3389/fnana.2020.00032.Search in Google Scholar PubMed PubMed Central

Tan, L.L., Oswald, M.J., Heinl, C., Retana Romero, O.A., Kaushalya, S.K., Monyer, H., and Kuner, R. (2019). Gamma oscillations in somatosensory cortex recruit prefrontal and descending serotonergic pathways in aversion and nociception. Nat. Commun. 10: 983, https://doi.org/10.1038/s41467-019-08873-z.Search in Google Scholar PubMed PubMed Central

Tang, J.S., Chiang, C.Y., Dostrovsky, J.O., Yao, D., and Sessle, B.J. (2021). Responses of neurons in rostral ventromedial medulla to nociceptive stimulation of craniofacial region and tail in rats. Brain Res. 1767: 147539, https://doi.org/10.1016/j.brainres.2021.147539.Search in Google Scholar PubMed PubMed Central

Taylor, N.E., Pei, J., Zhang, J., Vlasov, K.Y., Davis, T., Taylor, E., Weng, F.J., Van Dort, C.J., Solt, K., and Brown, E.N. (2019). The role of glutamatergic and dopaminergic neurons in the periaqueductal gray/dorsal raphe: separating analgesia and anxiety. eNeuro 6: 4–10.10.1523/ENEURO.0018-18.2019Search in Google Scholar PubMed PubMed Central

Thompson, J.M. and Neugebauer, V. (2017). Amygdala plasticity and pain. Pain Res. Manag. 2017: 829650110.1155/2017/8296501Search in Google Scholar PubMed PubMed Central

Thompson, J.M. and Neugebauer, V. (2019). Cortico-limbic pain mechanisms. Neurosci. Lett. 702: 15–23, https://doi.org/10.1016/j.neulet.2018.11.037.Search in Google Scholar PubMed PubMed Central

Tian, Y. and Zalesky, A. (2018). Characterizing the functional connectivity diversity of the insula cortex: subregions, diversity curves and behavior. Neuroimage 183: 716–733, https://doi.org/10.1016/j.neuroimage.2018.08.055.Search in Google Scholar PubMed

Timmermann, L., Ploner, M., Haucke, K., Schmitz, F., Baltissen, R., and Schnitzler, A. (2001). Differential coding of pain intensity in the human primary and secondary somatosensory cortex. J. Neurophysiol. 86: 1499–1503, https://doi.org/10.1152/jn.2001.86.3.1499.Search in Google Scholar PubMed

Tobaldini, G., Sardi, N.F., Guilhen, V.A., and Fischer, L. (2019). Pain inhibits pain: an ascending-descending pain modulation pathway linking mesolimbic and classical descending mechanisms. Mol. Neurobiol. 56: 1000–1013, https://doi.org/10.1007/s12035-018-1116-7.Search in Google Scholar PubMed

Todd, A.J. (2010). Neuronal circuitry for pain processing in the dorsal horn. Nat. Rev. Neurosci. 11: 823–836, https://doi.org/10.1038/nrn2947.Search in Google Scholar PubMed PubMed Central

Tonsfeldt, K.J., Suchland, K.L., Beeson, K.A., Lowe, J.D., Li, M.H., and Ingram, S.L. (2016). Sex differences in GABAA signaling in the periaqueductal gray induced by persistent inflammation. J. Neurosci. 36: 1669–1681, https://doi.org/10.1523/jneurosci.1928-15.2016.Search in Google Scholar

Tovote, P., Esposito, M.S., Botta, P., Chaudun, F., Fadok, J.P., Markovic, M., Wolff, S.B., Ramakrishnan, C., Fenno, L., Deisseroth, K., et al.. (2016). Midbrain circuits for defensive behaviour. Nature 534: 206–212, https://doi.org/10.1038/nature17996.Search in Google Scholar PubMed

Tracey, I. (2005). Nociceptive processing in the human brain. Curr. Opin. Neurobiol. 15: 478–487, https://doi.org/10.1016/j.conb.2005.06.010.Search in Google Scholar PubMed

Uddin, O., Studlack, P., Akintola, T., Raver, C., Castro, A., Masri, R., and Keller, A. (2018). Amplified parabrachial nucleus activity in a rat model of trigeminal neuropathic pain. Neurobiol. Pain 3: 22–30, https://doi.org/10.1016/j.ynpai.2018.02.002.Search in Google Scholar PubMed PubMed Central

Vachon-Presseau, E., Tétreault, P., Petre, B., Huang, L., Berger, S.E., Torbey, S., Baria, A.T., Mansour, A.R., Hashmi, J.A., Griffith, J.W., et al.. (2016). Corticolimbic anatomical characteristics predetermine risk for chronic pain. Brain 139: 1958–1970, https://doi.org/10.1093/brain/aww100.Search in Google Scholar PubMed PubMed Central

Veinante, P., Yalcin, I., and Barrot, M. (2013). The amygdala between sensation and affect: a role in pain. J. Mol. Psychiatry 1: 9, https://doi.org/10.1186/2049-9256-1-9.Search in Google Scholar PubMed PubMed Central

Veréb, D., Kincses, B., Spisák, T., Schlitt, F., Szabó, N., Faragó, P., Kocsis, K., Bozsik, B., Tóth, E., Király, A., et al.. (2021). Resting-state functional heterogeneity of the right insula contributes to pain sensitivity. Sci. Rep. 11: 22945, https://doi.org/10.1038/s41598-021-02474-x.Search in Google Scholar PubMed PubMed Central

Vertes, R.P. (2004). Differential projections of the infralimbic and prelimbic cortex in the rat. Synapse 51: 32–58, https://doi.org/10.1002/syn.10279.Search in Google Scholar PubMed

Vertes, R.P. (2006). Interactions among the medial prefrontal cortex, hippocampus and midline thalamus in emotional and cognitive processing in the rat. Neuroscience 142: 1–20, https://doi.org/10.1016/j.neuroscience.2006.06.027.Search in Google Scholar PubMed

Wang, G., Erpelding, N., and Davis, K.D. (2014). Sex differences in connectivity of the subgenual anterior cingulate cortex. Pain 155: 755–763, https://doi.org/10.1016/j.pain.2014.01.005.Search in Google Scholar PubMed

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.. (2015). Deactivation of excitatory neurons in the prelimbic cortex via Cdk5 promotes pain sensation and anxiety. Nat. Commun. 6: 7660, https://doi.org/10.1038/ncomms8660.Search in Google Scholar PubMed PubMed Central

Wang, N., Zhang, Y.H., Wang, J.Y., and Luo, F. (2021). Current understanding of the involvement of the insular cortex in neuropathic pain: a narrative review. Int. J. Mol. Sci. 22: 2648, https://doi.org/10.3390/ijms22052648.Search in Google Scholar PubMed PubMed Central

Wang, W., Tang, S., Li, C., Chen, J., Li, H., Su, Y., and Ning, B. (2019). Specific brain morphometric changes in spinal cord injury: a voxel-based meta-analysis of white and gray matter volume. J. Neurotrauma 36: 2348–2357, https://doi.org/10.1089/neu.2018.6205.Search in Google Scholar PubMed

Wang, Z., Huang, S., Yu, X., Li, L., Yang, M., Liang, S., Liu, W., and Tao, J. (2020). Altered thalamic neurotransmitters metabolism and functional connectivity during the development of chronic constriction injury induced neuropathic pain. Biol. Res. 53: 36, https://doi.org/10.1186/s40659-020-00303-5.Search in Google Scholar PubMed PubMed Central

Wiech, K. and Tracey, I. (2009). The influence of negative emotions on pain: behavioral effects and neural mechanisms. Neuroimage 47: 987–994, https://doi.org/10.1016/j.neuroimage.2009.05.059.Search in Google Scholar PubMed

Willis, W.D. and Westlund, K.N. (1997). Neuroanatomy of the pain system and of the pathways that modulate pain. J. Clin. Neurophysiol. 14: 2–31, https://doi.org/10.1097/00004691-199701000-00002.Search in Google Scholar PubMed PubMed Central

Wilson, T.D., Valdivia, S., Khan, A., Ahn, H.S., Adke, A.P., Martinez Gonzalez, S., Sugimura, Y.K., and Carrasquillo, Y. (2019). Dual and opposing functions of the central amygdala in the modulation of pain. Cell Rep. 29: 332–346, https://doi.org/10.1016/j.celrep.2019.09.011.Search in Google Scholar PubMed PubMed Central

Wong, C.E., Hu, C.Y., Lee, P.H., Huang, C.C., Huang, H.W., Huang, C.Y., Lo, H.T., Liu, W., and Lee, J.S. (2022). Sciatic nerve stimulation alleviates acute neuropathic pain via modulation of neuroinflammation and descending pain inhibition in a rodent model. J. Neuroinflammation 19: 153, https://doi.org/10.1186/s12974-022-02513-y.Search in Google Scholar PubMed PubMed Central

Woon, E.P., Sequeira, M.K., Barbee, B.R., and Gourley, S.L. (2020). Involvement of the rodent prelimbic and medial orbitofrontal cortices in goal-directed action: a brief review. J. Neurosci. Res. 98: 1020–1030, https://doi.org/10.1002/jnr.24567.Search in Google Scholar PubMed PubMed Central

Worthen, S.F., Hobson, A.R., Hall, S.D., Aziz, Q., and Furlong, P.L. (2011). Primary and secondary somatosensory cortex responses to anticipation and pain: a magnetoencephalography study. Eur. J. Neurosci. 33: 946–959, https://doi.org/10.1111/j.1460-9568.2010.07575.x.Search in Google Scholar PubMed

Yalcin, I., Barthas, F., and Barrot, M. (2014). Emotional consequences of neuropathic pain: insight from preclinical studies. Neurosci. Biobehav. Rev. 47: 154–164, https://doi.org/10.1016/j.neubiorev.2014.08.002.Search in Google Scholar PubMed

Yin, J.B., Liang, S.H., Li, F., Zhao, W.J., Bai, Y., Sun, Y., Wu, Z.Y., Ding, T., Sun, Y., Liu, H.X., et al.. (2020). dmPFC-vlPAG projection neurons contribute to pain threshold maintenance and antianxiety behaviors. J. Clin. Invest. 130: 6555–6570, https://doi.org/10.1172/jci127607.Search in Google Scholar

Yu, W., Pati, D., Pina, M.M., Schmidt, K.T., Boyt, K.M., Hunker, A.C., Zweifel, L.S., McElligott, Z.A., and Kash, T.L. (2021). Periaqueductal gray/dorsal raphe dopamine neurons contribute to sex differences in pain-related behaviors. Neuron 109: 1365–1380, https://doi.org/10.1016/j.neuron.2021.03.001.Search in Google Scholar PubMed PubMed Central

Yue, L., Ma, L.Y., Cui, S., Liu, F.Y., Yi, M., and Wan, Y. (2017). Brain-derived neurotrophic factor in the infralimbic cortex alleviates inflammatory pain. Neurosci. Lett. 655: 7–13, https://doi.org/10.1016/j.neulet.2017.06.028.Search in Google Scholar PubMed

Zhang, C., Chen, R.X., Zhang, Y., Wang, J., Liu, F.Y., Cai, J., Liao, F.F., Xu, F.Q., Yi, M., and Wan, Y. (2017). Reduced GABAergic transmission in the ventrobasal thalamus contributes to thermal hyperalgesia in chronic inflammatory pain. Sci. Rep. 7: 41439, https://doi.org/10.1038/srep41439.Search in Google Scholar PubMed PubMed Central

Zhang, M.M., Geng, A.Q., Chen, K., Wang, J., Wang, P., Qiu, X.T., Gu, J.X., Fan, H.W., Zhu, D.Y., Yang, S.M., et al.. (2022). Glutamatergic synapses from the insular cortex to the basolateral amygdala encode observational pain. Neuron 110: 1993–2008, https://doi.org/10.1016/j.neuron.2022.03.030.Search in Google Scholar PubMed

Zhang, Z., Gadotti, V.M., Chen, L., Souza, I.A., Stemkowski, P.L., and Zamponi, G.W. (2015). Role of prelimbic GABAergic circuits in sensory and emotional aspects of neuropathic pain. Cell Rep. 12: 752–759, https://doi.org/10.1016/j.celrep.2015.07.001.Search in Google Scholar PubMed

Zhou, H., Zhang, Q., Martinez, E., Dale, J., Hu, S., Zhang, E., Liu, K., Huang, D., Yang, G., Chen, Z., et al.. (2018). Ketamine reduces aversion in rodent pain models by suppressing hyperactivity of the anterior cingulate cortex. Nat. Commun. 9: 3751, https://doi.org/10.1038/s41467-018-06295-x.Search in Google Scholar PubMed PubMed Central

Zhu, H., Xiang, H.C., Li, H.P., Lin, L.X., Hu, X.F., Zhang, H., Meng, W.Y., Liu, L., Chen, C., Shu, Y., et al.. (2019). Inhibition of GABAergic neurons and excitation of glutamatergic neurons in the ventrolateral periaqueductal gray participate in electroacupuncture analgesia mediated by cannabinoid receptor. Front. Neurosci. 13: 484, https://doi.org/10.3389/fnins.2019.00484.Search in Google Scholar PubMed PubMed Central

Zhu, X., Xu, Y., Shen, Z., Zhang, H., Xiao, S., Zhu, Y., Wu, M., Chen, Y., Wu, Z., Xu, Y., et al.. (2021). Rostral anterior cingulate cortex-ventrolateral periaqueductal gray circuit underlies electroacupuncture to alleviate hyperalgesia but not anxiety-like behaviors in mice with spared nerve injury. Front. Neurosci. 15: 757628, https://doi.org/10.3389/fnins.2021.757628.Search in Google Scholar PubMed PubMed Central

Zhu, X., Zhou, W., Jin, Y., Tang, H., Cao, P., Mao, Y., Xie, W., Zhang, X., Zhao, F., Luo, M.H., et al.. (2019). A central amygdala input to the parafascicular nucleus controls comorbid pain in depression. Cell Rep. 29: 3847–3858, https://doi.org/10.1016/j.celrep.2019.11.003.Search in Google Scholar PubMed PubMed Central

Zhu, Y.B., Wang, Y., Hua, X.X., Xu, L., Liu, M.Z., Zhang, R., Liu, P.F., Li, J.B., Zhang, L., and Mu, D. (2022). PBN-PVT projections modulate negative affective states in mice. Elife 11: 6–9, https://doi.org/10.7554/elife.68372.Search in Google Scholar

Zhuo, M. and Gebhart, G.F. (1997). Biphasic modulation of spinal nociceptive transmission from the medullary raphe nuclei in the rat. J. Neurophysiol. 78: 746–758, https://doi.org/10.1152/jn.1997.78.2.746.Search in Google Scholar PubMed

Received: 2023-03-25
Accepted: 2023-05-18
Published Online: 2023-06-08
Published in Print: 2023-12-15

© 2023 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 27.4.2024 from https://www.degruyter.com/document/doi/10.1515/revneuro-2023-0037/html
Scroll to top button