Skip to main content
Physiological Reviews logoLink to Physiological Reviews
. 2018 Aug 15;98(4):2287–2316. doi: 10.1152/physrev.00035.2017

Neuroimmune Communication in Health and Disease

Colin Reardon 1, Kaitlin Murray 1, Alan E Lomax 1
PMCID: PMC6170975  PMID: 30109819

Abstract

The immune and nervous systems are tightly integrated, with each system capable of influencing the other to respond to infectious or inflammatory perturbations of homeostasis. Recent studies demonstrating the ability of neural stimulation to significantly reduce the severity of immunopathology and consequently reduce mortality have led to a resurgence in the field of neuroimmunology. Highlighting the tight integration of the nervous and immune systems, afferent neurons can be activated by a diverse range of substances from bacterial-derived products to cytokines released by host cells. While activation of vagal afferents by these substances dominates the literature, additional sensory neurons are responsive as well. It is becoming increasingly clear that although the cholinergic anti-inflammatory pathway has become the predominant model, a multitude of functional circuits exist through which neuronal messengers can influence immunological outcomes. These include pathways whereby efferent signaling occurs independent of the vagus nerve through sympathetic neurons. To receive input from the nervous system, immune cells including B and T cells, macrophages, and professional antigen presenting cells express specific neurotransmitter receptors that affect immune cell function. Specialized immune cell populations not only express neurotransmitter receptors, but express the enzymatic machinery required to produce neurotransmitters, such as acetylcholine, allowing them to act as signaling intermediaries. Although elegant experiments have begun to decipher some of these interactions, integration of these molecules, cells, and anatomy into defined neuroimmune circuits in health and disease is in its infancy. This review describes these circuits and highlights continued challenges and opportunities for the field.

I. INTRODUCTION

The nervous and immune systems act together as an integrated physiological system to monitor and respond to infection and inflammation. The concept of neuroimmune communication is not new, with many of the symptoms of inflammation arising from the effects of inflammatory mediators on the nervous system (196), and the detection of acetylcholine released from the spleen 90 yr ago (61). Several prominent studies have resulted in a new appreciation for the innervation of lymphoid organs and the functional consequences of neuronal activation for the immune system (2, 89, 224, 259). Perhaps more intriguingly, immune cells can produce neurotransmitters, functioning as a nonneuronal source of these molecules, with release dependent on signals from the innervation or the local tissue milieu (214, 224).

Communication between the immune and nervous systems is bidirectional, with neuronal signaling activated by exposure to pathogens or inflammation and immune cell function effected by neurotransmitters (2, 31, 214, 224, 259, 265). While there are a number of adaptive and maladaptive physiological responses to inflammation, ranging from a classical sickness response to altered satiety (62), many different stimuli that activate afferent pathways can lead to immunomodulation by autonomic neurons.

Following elegant studies documenting the existence of an anti-inflammatory reflex, there has been a resurgence in interest in the ability of the nervous system to regulate immune function. Demonstrating the power of this pathway, electrical stimulation to the efferent arm of this reflex can significantly reduce morbidity and mortality in a mouse model of septic shock (31, 265). These protective effects appear to be conserved in diverse immunopathologies with recent studies documenting therapeutic efficacy in preclinical models of septic shock, postoperative ileitis, rheumatoid arthritis (RA), inflammatory bowel disease (IBD), and renal ischemia reperfusion injury (31, 93, 119, 122, 126, 138). Such success has also led to the development of electrical nerve stimulators for the treatment of chronic inflammatory conditions. The rapid advances in this field and development of neurostimulators have resulted in numerous clinical trials for diseases ranging from IBD to RA (28, 138). Despite promising early preclinical and open label clinical trial results, new preclinical discoveries continue to highlight that there are a considerable number of unknowns. Factors that could conceivably impact the efficacy of electroceuticals include interspecies and individual variation in neural circuits and the effect of chronic inflammation on peripheral neurons and glia. While the vast majority of studies on the neuroimmune reflex arc have been conducted in mice and rats, it is unknown how applicable these circuits will be to humans.

The field of neuroimmune communication has grown at an exponential rate in recent years. With this rapid growth, there have been a tremendous number of advances and several new controversies that have developed. This review provides a contextual background, highlights these recent advances, and discusses some of the current challenges and controversies in the field.

II. DETECTION OF PATHOGENS, IMMUNE ACTIVATION, AND INFLAMMATION BY THE NERVOUS SYSTEM

How the nervous system becomes activated by bacteria and the immune system remains hotly debated. The source of this controversy likely stems from the use of different animal models influencing experimental outcomes that are then generalized to multiple pathways. In general, the proposed mechanisms of communication from the periphery to the central nervous system (CNS) include a humoral route or direct detection by peripheral afferent neurons. Each of these possibilities will be described in detail.

A. Circumventricular Organs

The humoral route of CNS activation by peripheral inflammation or infection involves the passage of bacterial-derived substances or host-derived cytokines to the brain (11, 12). This transmission of signals from the blood to the brain predominantly occurs in parts of the brain that lack a blood-brain barrier (BBB): the circumventricular organs (CVO). In most parts of the brain, exposure of the CNS to these substances is highly regulated due to the lack of fenestrated capillaries in the majority of CNS vascular beds (279). This BBB is further enhanced by the formation of tight junction proteins between adjacent endothelial cells (11). The consequence of this is reduced paracellular permeability, with blood to cerebrospinal fluid (CSF) transport occurring primarily by transcytosis (11, 12, 279). It is important to note that while transcytosis is the predominant mechanism, endothelial cells in the brain have reduced uptake compared with the periphery. While dissolved gasses are freely exchanged at this surface, this structural arrangement in conjunction with specific receptors and transporter proteins in the luminal and abluminal endothelial membranes permits controlled uptake of small hydrophobic molecules, nutrients, and specific peptides or proteins (11). Complementing the actions of the endothelial barrier, it is well established that astrocytes project feet toward the endothelium, adding another layer of resistance to small molecules transiting from the blood to the brain (12). While this highly selective permeability of this barrier ensures that neurons are protected from potentially neurotoxic substances, these structural features would prevent the central nuclei of the autonomic nervous system from monitoring the periphery. The autonomic nervous system can monitor hormones and metabolites in the periphery that are carried in the blood due to a structural adaptation in endothelium of CVO. In contrast to the highly selective nature of the endothelial cells that provide the BBB, the endothelium of CVO in the CNS allows for detection of substances circulating in the blood. These endothelial cells are highly fenestrated and lack the tight intercellular junctions associated with the endothelium that comprise the BBB (100). The eight CVO organs are divided into two discrete groups by functionality: sensory and secretory. The sensory organs include the subfornical organ (SFO), organum vasculosum, lamina terminalis (OVLT), and the area postrema (AP). The AP is of particular interest in the context of neuroimmune regulation due to the nature of its axonal projections. Neurons originating in the AP innervate the nucleus tractus solitarius (NTS) and the dorsal motor nucleus (DMN) of the vagus nerve, among other targets (206). This anatomy would allow for activation of neurons in the AP and regulation of vagal efferent (parasympathetic) signaling to alter physiology potentially including immune function.

B. Afferent Neural Inputs to Immunomodulatory Circuits

Peripheral primary afferent neurons express receptors for microbial constituents and inflammatory products. These neurons therefore have been proposed as a hardwired neural connection that underpins signaling of inflammation or infection from the periphery to the CNS. Two main populations of neurons comprise this pathway: vagal and spinal afferent neurons. Activation of vagal afferent pathways has been most thoroughly studied, so the mechanisms of vagal afferent neuronal activation will be described in most detail.

As the longest cranial nerve in the body, the vagus nerve innervates a diverse range of tissues including the heart, lung, and gastrointestinal tract. The vagus nerve comprises 70–80% afferent axons, with the cell bodies residing in the nodose and jugular ganglia. It is important to note that unique stimuli can trigger vagal afferent neurons in different tissues, ranging from nutrients in the intestinal tract (212) to mechanosensitive receptors in the lung (234). Alterations of these signals during inflammation at any of these sites could conceivably have profound implications in the activation of the anti-inflammatory reflex arcs. Activation of these afferent fibers results in synaptic transmission in the NTS, where information is integrated and sent to the DMN of the vagus, where the cell bodies of cholinergic parasympathetic preganglionic neurons reside. The axons of efferent neurons in the vagus nerve belong to these parasympathetic preganglionic neurons and will henceforth be referred to as vagal efferent neurons. These vagal efferent axons are thought to be an important target of electroceutical treatments that stimulate vagal axons to suppress inflammation (78).

C. Direct Activation of Vagal Afferent Neurons

1. Pathogen-associated molecular patterns

Direct activation of afferent fibers by pathogens has long been suggested to be an important mechanism of relaying peripheral information to the CNS. In brief, detection of pathogens in the body can occur through host receptors selective for highly conserved repeating motifs common to pathogens. These molecules are termed pathogen-associated molecular patterns (PAMPs) and range from lipopolysaccharide (LPS) in the cell wall of gram-negative bacteria, to unmethylated cytosine and guanine (CpG) in DNA. These motifs serve as ligands for specific Toll-like receptors (TLR), with binding inducing intracellular signaling and gene transcription events. Although TLR expression is considered a feature of innate immune cells, vagal afferent neurons also express TLR that facilitate the detection of PAMPs. While TLR3, 4, 7, and 9 are expressed on spinal afferent neurons and in the dorsal root ganglia (159, 208, 255), only TLR4, a selective receptor for LPS, has been detected in vagal afferent neurons from the nodose ganglia (NG) (65, 113). This section describes the evidence for expression of functional TLR on vagal afferents and evidence of functionality.

Vagal afferents are critical to the peripheral detection of LPS and the communication of this message to the CNS. Prior to the discovery of TLR4 and the development of tools to localize cells expressing this receptor (112, 204), afferent neurons were reported to be activated by LPS (92, 146). Peripheral administration of LPS induced sickness behaviors and neural activation in the NTS that was abrogated in mice subjected to subdiaphragmatic vagotomy. In support of these data that suggested LPS-induced activation of vagal afferents, freshly isolated NG and cultured NG neurons from rats express TLR4 mRNA and protein (65, 113). LPS stimulation can exert a variety of effects on vagal afferents including reduced sensitivity to other physiological stimuli. For example, LPS stimulation increased SOCS3 expression in the NG, preventing leptin-mediated STAT3 activation (65). Despite this ability of LPS to activate and modulate vagal afferent signaling, the mechanisms of LPS-induced neural activation have been elusive. Cultured neurons from the NG stimulated with LPS appear to be polymodal chemosensitive afferents that respond to a variety of nociceptive stimuli. Neural activation determined by Ca2+ imaging and patch-clamp experiments further revealed TLR4 independent activation by LPS using TLR4-deficient cells. This TLR4 independent response to LPS was found to be dependent on TRPV1 channel expression and was significantly reduced in TRPV1 knockout (KO) neurons, or wild-type (WT) cells pretreated with a selective TRPV1 blocker. It is however unclear how LPS activates TRPV1 or the ramifications of this in the afferent pathway. Together, these data illustrate the ability of afferent neurons to detect and respond to infection, a key requirement for the nervous system to modulate immune function.

2. Danger-associated molecular patterns

Inflammation and infection can also be detected after the release host-derived intracellular substances. As these substances are typically released as a consequence of inflammation or necrotic cell death, they have been termed danger-associated molecular patterns (DAMPs), or “alarmins” (154, 233). There is a diverse array of substances that serve as DAMPs, including ATP, DNA binding proteins such as high mobility group box 1 (HMGB1), and the alarmin cytokine interleukin (IL)-33 (47, 154, 154). In this section, we describe some of these molecules and the effect they exert on primary afferent neurons (summarized in TABLE 1).

Table 1.

Expression of cytokine, DAMP, and PAMP receptors by neurons and associated cells

Ligand Receptors Cell Type Tissues
ATP P2X1 Neurons NG (134, 266)
P2X2 Neurons NG (142, 266)
SGC NG (79)
P2X3 Neurons DRG (152)
NG (142, 266)
P2X4 Neurons NG (142, 266)
P2X7 SGC NG (79)
CpG TLR9 Neurons DRG (150, 208)
dsRNA TLR3 Neurons DRG (208)
HMGB1 RAGE Neurons DRG (229, 263)
IL-33 ST2/IL-1RAcP Neurons DRG (157)
NG (246)
LPS TLR4 Neurons NG (65, 113)
TRPV1 DRG (208, 255)
ssRNA TLR7 Neurons DRG (159, 208)
TNF-α TNF-α receptor 1/2 Neurons DRG (115)
VAN (107)
NTS (107)
DMN (107)

CpG, unmethylated cytosine-phosphate-guanine dinucleotides; HMGB1, high mobility group box 1; IL, interleukin; LPS, lipopolysaccharide; TNF-α, tumor necrosis factor-α; SGC, satellite glial cells; NG, nodose ganglia; DRG, dorsal root ganglia; VAN, vagal afferent neurons; NTS, nucleus tractus solitarius; DMN, dorsal motor nucleus of the vagus.

3. ATP

Like all DAMPs, ATP released from host cells can occur as a consequence of inflammation or necrosis (83). It is also critical to note that ATP is also released as a neurotransmitter within the autonomic nervous system (40) and can be released from epithelial and endothelial cells (134). In the context of inflammation, ATP can be released through Pannexin-1 (Panx1) channels (50, 230), from a variety of cell types including intestinal epithelial cells, keratinocytes, and neutrophils (73). Release of ATP by Panx1 can occur following activation mechanical stress during inflammation (14, 73), or through activation by caspase 3/7-dependent cleavage (50, 230). It is uncertain if distension and stretch during edema that can occur during inflammation is sufficient to induce Panx1-mediated ATP release. Panx1 activation can also occur by caspase 3/7-mediated cleavage during pyroptosis. This nonapoptotic form of programmed cell death occurring during inflammation results in activation of caspase 1 and consequently caspase 7 (174). As such, multiple inflammatory stimuli appear capable of inducing Panx1 activation and extracellular release of ATP that could be detected by afferent neurons.

Extracellular ATP is capable of binding to purinergic (P2) receptors of two discrete types: P2X and P2Y receptors. Depending on the cell type they are expressed on, ATP initiates signaling that results in a range of responses including chemotaxis of neutrophils (51) to activation of neuronal signaling (40). Vagal afferent neurons have been characterized to be responsive to ATP and express the purinergic receptor subunits P2X2 and P2X3. These early experiments demonstrated afferent neuronal activation in response to ATP, or the P2X agonist α,β-methylene-ATP (276), that was blocked with the P2X2/3 receptor-selective antagonist A-317491 (142). Single-cell RT-PCR conducted on neurons from the NG indicated P2X2 and P2X3 expression compared with the predominant expression of P2X2 in jugular neurons of guinea pig (142). Indicative of species-specific differences, although adult rat NG express P2X1, P2X2, P2X3, and P2X4, P2X3 was the predominant receptor subunit (266). These receptors were functional, with application of extracellular ATP activating currents in whole cell patch-clamp experiments (266). As these different receptor subunits engender unique current responses, these data would suggest that ATP could stimulate different patterns of neural activation in unique vagal afferent fibers. In mice, as in other animals, these receptor subunits form heteromeric receptors as evidenced by genetic P2X3 deficiency only reduced ATP evoked responses with slower desensitization compared with WT neurons (241). As a heteromeric receptor with functional changes depending on subunit expression, it would be interesting to consider if peripheral inflammation alters P2X subunit expression. In support of this possibility, P2X2 and P2X3 expression increased in rat NG after myocardial ischemia (267), a trigger of inflammation and necrosis (86). These heteromeric receptors are not restricted to vagal afferents and have also been found on sensory afferents such as the DRG (152). P2X receptors can also be found on neurons that may participate in immunomodulatory reflexes including spinal afferent and enteric afferent neurons (220). These findings suggest the peripheral nervous system possesses the capability to detect DAMP that are released during inflammation.

D. Cytokines

Afferent innervation has long been proposed to be involved in the monitoring of peripheral immune status. Activation of the immune system following infection, or exposure to PAMPs such as LPS, typically leads to the production of proinflammatory cytokines. These cytokines, the expression of their receptors, and the consequence on afferent signaling will be described.

1. Tumor necrosis factor-α

Activation of the immune system results in the production and release of proinflammatory cytokines, including tumor necrosis factor (TNF)-α. This cytokine is produced by a plethora of immune cells including macrophages, T cells, and mast cells following exposure to stimuli such as LPS, or other PAMPs. Detection of TNF-α occurs by binding to TNF-α receptor 1 (TNFR1) or TNF-α receptor 2 (TNFR2). While TNFR1 is expressed constitutively by a wide variety of cells throughout the body, TNFR2 is expressed on selected cell types including macrophages/monocytes, microglia (33), and neurons in the central (274) and peripheral nervous system (115). In addition to differences in receptor expression profiles, the downstream signaling elicited by ligand binding is unique to each receptor. Following binding to TNFR1, a complex series of intracellular signaling events occurs resulting in activation of IKK and phosphorylation of IκBα. This phosphorylation event signals for ubiquitination and proteasome-mediated destruction of this inhibitory protein, allowing for NF-κB activation (35).

Vagal afferent neurons that project into the NTS of the brain stem constitutively express TNFR1, and TNF-α can activate these neurons or enhance the response to otherwise innocuous stimuli (108), stimulation with ATP (222) or capsaicin (115). Expression of TNFR1 was identified in the NTS by immunohistochemistry, with vagotomy rostral to the NG ablating TNFR1 immunoreactivity, suggesting that the immunoreactivity was derived from the central terminals of vagal afferent neurons. This TNFR1 immunoreactivity was also found in the cell bodies in the celiac ganglia but not the distal vagal trunk (107). It is important to note that the identity of TNFR1+ cells was not confirmed to be neurons, and as such, expression of TNFR1 by microglia (141), astrocytes, or oligodendrocytes (72) in the NTS cannot be excluded. RNA-Seq data from isolated cells of the cerebral cortex indicates higher basal expression of TNFR1 in microglia and astrocytes compared with neurons (277). These data demonstrate that caution should be taken in interpreting that TNFR1+ cells in the NTS are neurons unless other cell types are clearly excluded. Functionally, TNF-α has been shown to activate vagal afferent neurons in the gastrointestinal tract, increasing sensitivity to innocuous stimuli resulting in a vagal-gastroinhibitory reflex and gastric stasis (75). Vagal afferents are therefore capable of detecting and responding to TNF-α, resulting in alterations to host physiology.

2. IL-1β

Similar to TNF-α, IL-1β is produced by a multitude of cells during inflammation or cellular stress. Neurons in the NG that project into the NTS express IL-1β receptor and are activated following intravenous injection of recombinant IL-1β (74, 99). Although these studies suggest neuronal activation of neurons by nuclear c-Fos immunoreactivity, it is conceivable that given the pleiotropic nature of IL-1β, and the multitude of cells that express IL-1βR, activation could be through a paracrine fashion. In support of this contention, activation of vagal afferents was shown to be dependent on IL-1β-induced prostaglandin synthesis (99) from adjacent vascular endothelium (171). Endothelium-derived PGE2 appears to act presynaptically to augment spontaneous excitatory synaptic input to the NTS. These data suggest that IL-1β stimulation of vagal afferents can act through an indirect mechanism. There is evidence of direct neural activation and sensitization of nociceptive terminal of spinal afferents to other stimuli within the skin by IL-1β (23). Together these findings indicated IL-1β can cause neuronal activation or increase the sensitivity of primary afferents through direct and indirect mechanisms serving to detect inflammation in peripheral tissues.

3. IL-33

Identified as a mediator in T helper 2 (Th2) responses including allergy and immunity to extracellular parasites, IL-33 has emerged as another cytokine that can fulfill a role as an alarmin molecule (154). There is constitutive IL-33 expression in a wide variety of cells including endothelial, epithelial, and lymph node stromal cells in humans and mice (182). Under basal conditions, IL-33 is localized to the nucleus and is not detected in the cytoplasm or extracellular space. This basal expression can be increased during inflammation, with increased nuclear IL-33 in alveolar epithelial cells following allergic lung inflammation (104), intestinal nematode infection (275), mouse models of colonic inflammation, and patients with IBD (198, 235).

This cytokine functions as an alarmin, with the release of full-length IL-33 from the nucleus during necrosis, tissue damage, or cellular stress. Similar to IL-1β, IL-33 can be cleaved by activated caspase 3 or caspase 9, typically activated during immunologically quiescent apoptosis. This proteolytic cleavage produces noninflammatory fragments of this cytokine (154). The full-length bioactive IL-33 elicits intracellular signaling through the ST2 receptor, that is expressed by a variety of cell types including vagal afferent (246) and DRG neurons (157). Although functional studies were not performed on vagal afferents, it is intriguing to consider that ST2 on vagal afferents could be another mechanism to detect IL-33 released during necrotic cell death. In primary afferents innervating the skin, IL-33 release triggered by the poison ivy antigen, urushiol, resulted in itch through ST2-induced activation, and IL-33 induced Ca2+ signaling in DRG neurons. At an organism level, urushiol challenge elicited skin inflammation and pruritic behaviors that were exacerbated with coinjection of IL-33, mimicked by injection of IL-33 alone, and blocked with IL-33 neutralizing antibodies (157). Together these data indicate that afferent neurons express this functional receptor, with activation inducing neuronal activation and behavior changes.

E. Indirect Activation of Vagal Afferents in the Neuroimmune Reflex

A multitude of noninflammatory stimuli can activate vagal afferent neurons, resulting in vagal efferent outflow and consequently inhibition of immune cell activation. In the gastrointestinal tract, ingestion of nutrients activates a specialized subset of the epithelial cells lining the intestinal tract. These enteroendocrine cells (EEC) are activated by sugars, dipeptides, and long-chain fatty acids (144, 211). In response to luminal nutrients, these EEC release hormones, including cholecystokinin (CCK), that activate vagal afferent nerves (211). Elegant studies have demonstrated synapses between EEC and the underlying afferent innervation and that these interactions occur throughout the gastrointestinal tract including in the colon (26, 27). The contribution of EEC to neuroimmune regulation has been suggested with eternal high-fat diet to induce CCK release from EEC, leading to activation of CCK receptors on vagal afferents (164, 165). While CCK actions on vagal afferent neurons is usually associated with satiation, it has also been demonstrated that vagal afferent detection of CCK increased immunomodulatory efferent signaling to the spleen, reducing production of TNF-α and IL-6 production during a model of septic shock (165). Although there are undoubtedly several other mechanisms at work in this model, including the sequestering of endotoxin by increased triacylglycerol-rich lipoproteins in serum (166), these studies highlight that activation of vagal afferents by a diverse array of noninflammatory stimuli can be sufficient to activate an anti-inflammatory reflex.

1. Bitter taste receptors

The family of bitter taste receptors comprises closely related family members encoded by ~28 human and 36 mouse T2R genes with unique sensitivities to select ligands (7, 8). The expression of bitter taste receptors in cells and tissues not involved in taste sensation has been revolutionized by the demonstration that bacterial quorum molecules can serve as ligands for these receptors, aiding host detection of bacteria (251). In the lung, the bacterial pathogen Pseudomonas aeruginosa produces acyl-homoserine lactone as a quorum sensing molecule to establish bacterial population dynamics (197, 240). These molecules are sensed by the host through activation of host T2R, increasing the local immune response against this pathogen (148). Intriguingly, bitter taste receptors have been found to be expressed on EEC and other specialized epithelial cells (205). In the gastrointestinal tract, a wide variety of T2R genes are expressed in a location-specific manner in the stomach antrum and fundus, duodenum (270), ileum, and colon (19, 205). It remains to be determined if these receptors can detect bacterial quorum-sensing molecules at the intestinal mucosal surface and participate in host defense, or if these compounds can directly or indirectly initiate neuroimmune signaling.

F. Inflammation-Induced Changes to Afferent Signaling: Satellite Glial Cells

An additional confounding variable that should be considered in neuroimmune communication is the role of satellite glial cells (SGC). Similar to glia in the CNS, these supporting cells encase neurons within sensory ganglia including the DRG and the vagal afferents of the NG. The SGC from the NG respond to proinflammatory cytokines induced by systemic inflammation by becoming activated as evidenced by increased GFAP expression and Ca2+ signaling in response to ATP (79). This increased responsiveness to ATP was due to inflammation-induced expression of the P2X2 and P2X7 receptors, pannexin-1 activation, and increased glia-neuron dye coupling (79). Colonic inflammation also increases glia-neuron dye coupling, spontaneous activity, and hyperexcitability of DRG neurons. This increased neuronal activity in isolated DRG was prevented by the addition of gap junction blockers, suggesting that inflammation-driven glial-neuronal coupling was responsible for aberrant neuronal function (116). This activation of DRG satellite glial cells has been established as a mechanism of inflammation-induced visceral hyperalgesia, that can be reduced with pharmacological blockade of gap junctions (116, 268). These data suggest that SGC in the DRG and NG can increase neuronal signaling, although it is unknown how the activation of SGC during chronic inflammation effects afferent signaling in the anti-inflammatory reflex arc.

G. Other Afferent Neurons: Spinal Afferent Neurons

Spinal afferent neurons, with cell bodies in DRG and trigeminal ganglia, also respond to bacterial products, inflammatory mediators, DAMPs, and PAMPs and transmit information to the CNS. While the primary roles of these neurons are the transfer of afferent and nociceptive information to the brain, spinal afferent neurons may also be important mediators of neural immunoregulation. Activation of the peripheral terminals of spinal afferent neurons leads to the peripheral release of neuropeptides such as substance P and calcitonin gene-related peptide (CGRP) via axon reflexes. These neuropeptides cause vasodilation and lymphocyte extravasation, a process known as neurogenic inflammation. There is also evidence that CGRP can directly suppress macrophage activation (55). Consistent with this, ablation of Nav1.8-expressing afferent neurons increased inflammation in response to Staphylococcus aureus infection (272). The majority of the central terminals of spinal afferent neurons synapse on populations of interneurons in the dorsal horn, which project to higher brain centers and can result in the perception of pain. However, there are well-established local interneuronal projections from the dorsal horn to sympathetic preganglionic neurons within the intermediolateral cell column of the spinal cord (68). An interneuron pathway between spinal afferent and preganglionic splenic sympathetic neurons is particularly intriguing (44). Such a connection could serve as a neuroanatomical substrate of a possible spinal afferent-sympathetic immunomodulatory reflex pathway. This pathway appears to become particularly significant following thoracic spinal cord injury, partly due to the loss of descending inhibitory pathways (257, 278). Therefore, spinal afferent pathways may participate in immunomodulation by release of peripheral immunomodulatory peptides and by activation of autonomic anti-inflammatory pathways.

III. CIRCUITRY OF THE CHOLINERGIC REFLEX ARC

In the cholinergic anti-inflammatory pathway (CAIP), following detection of inflammation by afferent neurons, this information is coordinated in the CNS. In the brain stem, the central terminals of vagal afferent neurons, and those of the neurons in sensory CVOs innervate the NTS. Release of glutamate from these afferent fibers activates neurons within the NTS (5, 145, 248) that project into the DMN, initiating the efferent arm of the anti-inflammatory pathway. This basic functional circuit was established by observing neuronal activation in the NTS, coupled with surgical ablation and neurostimulation techniques. Activation of NTS neurons indicated by increased c-fos expression has been observed during infection with bacteria, nematodes (46), or immune activation elicited by injection of PAMPs such as LPS (156).

The CAIP can also be activated by intracerebroventricular (icv) injection of muscarinic receptor agonists. Injection of muscarine or CNI-1493, an M1 agonist, activates the efferent arm of the anti-inflammatory reflex to reduce carrageen-induced paw inflammation and edema (30). Although it is uncertain how central administration of these muscarinic agonists elicits vagal efferent activity, it has been proposed that neurons within the NTS or DMN express muscarinic receptors, or that these compounds act indirectly by stimulating neurons at other sites that project to the DMN (200). While choline acetyltransferase (ChAT) activity (105) and functional muscarinic receptors are expressed in the NTS (60), vagal preganglionic motor neurons in the DMN do not express muscarinic receptors (111). This would suggest that muscarinic agonists such as CNI1493 could activate interneurons or glia and consequently induce vagus efferent signaling, or alter the response to glutamate. Indirect modulation of sensory inputs has been suggested as there appears to be a lack of tonic ACh signaling. As evidence of this, while injection of ACh into the NTS induces hypotension and bradycardia that can be prevented by muscarinic receptor antagonists, injection of muscarinic receptor antagonists alone fail to induce hypertension (256). It is worth noting that while the anti-inflammatory effects of CNI1493 were abrogated by atropine pretreatment, it is uncertain if central muscarinic signaling occurs during a reflex initiated by peripheral LPS or inflammation.

The role of the vagus nerve in providing protection during systemic inflammation was suggested by experiments using surgical vagotomy. Cutting of the vagus before induction of endotoxemia or polymicrobial sepsis significantly increased mobility and mortality compared with sham-operated animals (31). As the vagus nerve comprises 70–80% afferent neurons, it was not clear if vagal afferents, efferents, or both were part of the functional circuitry. Elegant experiments using electrical vagal neural stimulation (VNS) coupled with vagotomy were central to establishing the importance of the vagal efferent neurons. Application of VNS to the peripheral end of the cut nerve following vagotomy provided protection from endotoxemia, indicating that the efferent innervation of target tissues was responsible for protection (31). These vagal efferents indirectly innervate tissues including the spleen, a site identified as the predominant source of proinflammatory cytokines including TNF-α in mice during endotoxemia (120). These data were interpreted as VNS activation of efferent vagal nerve activity reduces macrophage activation and TNF-α production in the spleen, consequently reducing mortality. Despite this inhibition of aberrant immune responses by the efferent pathway, VNS can also elicit afferent vagal nerve activation, which functions as a potent inhibitor of LPS-induced TNF-α production through an unknown circuit (193a). This study highlights the importance of carefully defining the circuitry being stimulated and could serve as an underlying basis of some of the controversies in the CAIP.

The anatomical connection between the vagus nerve and the sympathetic innervation of the spleen is a highly contentious issue. On the basis of prior neuroanatomical tracing studies, a disynaptic connection was proposed such that the preganglionic parasympathetic axons synapsed onto postganglionic sympathetic neurons in the celiac ganglia (18, 223). These tyrosine hydroxylase (TH)-expressing neurons that produce norepinephrine (NE) are proposed to innervate the spleen. At present, there are several lines of conflicting neuroanatomical evidence that refute or support this pathway. In mice, the injection of cholera toxin B (CTB) into the spleen resulted in labeling of a discrete population of neurons within the DMN that were not found in animals subject to vagotomy (42). In contrast, studies in rats with injection of DiI into the DMN, which contains the cell bodies of parasympathetic preganglionic neurons, and fast blue into the spleen, which contains sympathetic nerve terminals, did not reveal overlap of tracers in neurons in the CG (34). However, analysis of the injected rats revealed abundant fast blue labeling of postganglionic neurons in the suprarenal ganglion without labeling within the CG. Vagal terminals were rare in the suprarenal ganglion, and synaptic contacts between the vagal terminals and the postganglionic sympathetic neurons were not observed (34). These data suggest a complete lack of synaptic connections between parasympathetic preganglionic neurons and sympathetic neurons that innervate the spleen. Electrical stimulation of the efferent vagus in the rat further support this lack of a neural connection from the vagus to the spleen, with electrical VNS unable to elicit electrical activity in sympathetic neurons that projected into the spleen (169). These studies raise the possibility that there are significant neuroanatomical differences in mice compared with rats, or more likely that there are other neural circuits and sympathetic ganglia activated during VNS. As evidence for this multiple circuit hypothesis, injection of pseudorabies virus (PRV) into the spleen results in labeling of the preganglionic sympathetic motor neurons in the spinal cord concurrent with labeling of the DMN. Highlighting that two pathways could innervate the spleen, vagotomy was shown to prevent viral labeling in the DMN, but not the spinal cord (39). In direct opposition to these data, injection of the same PRV strain into the spleen labeled neurons in the intermediolateral nucleus of the spinal cord, but not the DMN (43, 44). Species variation should also not be discounted as functional evidence of vagal-sympathetic anti-inflammatory connections was derived from studies of mice, whereas the neuroanatomical studies were predominantly performed in rats. Discrepancies such as these highlight the continued need to perform comprehensive neuroanatomical and circuitry analysis in a variety of model organisms. The ability to translate these findings to the treatment of inflammation clinically will depend on deciphering not only the neural circuitry but also accounting for the potential existence of multiple immunomodulatory pathways and interspecies variations.

A. Innervation of the Spleen

The innervation of the spleen is critical to the anti-inflammatory reflex in models of sepsis and provides further information about the circuitry involved. The presence of axons in the spleen has long been described in mice (217), rats, cats (25), and humans (109), although the neurotransmitters released and the consequences of the splenic innervation have been contentious.

In mice and rats, splenic innervation is provided by sympathetic fibers that express TH and produce NE. With few exceptions in the literature suggesting sparse innervation by cholinergic fibers, TH+ fibers are readily detected in spleen of these animals (15, 186, 223). Lack of ChAT+ innervation has been established using both immunohistochemistry and lack of green fluorescence protein (GFP)+ axons in ChAT-GFP reporter mice (185). While the cholinergic innervation of the spleen has been reported in humans, it is important to note that this innervation was identified on the basis of acetylcholinesterase (AChE), which is often expressed postsynaptically, and not ChAT expression (140). In support of this lack of cholinergic innervation, vesicular ACh transporter expression was not detected in nerve fibers or bundles by immunohistochemistry on tissue sections of human spleen (110). With the limited number of neuroanatomical studies conducted, it is still not clear if the splenic innervation in humans and non-human primates is similar to other animals. In contrast, the sympathetic innervation of the spleen in mice and rats has been extensively described and characterized. The majority of splenic TH+ axons are associated with the vasculature, with isolated fibers branching off and projecting into the white pulp (80, 81). These fibers in the white pulp extend towards macrophages and the T- and B-cell zones, but are absent from germinal centers. Development of this innervation and the local guidance cues received by neurons innervation appears to be dynamic, with reinnervation of the spleen following chemical sympathectomy (161). While it is unknown what factors control these processes, cytokines such as IL-17A could have a significant role. IL-17A induces the outgrowth of sympathetic neurons in the colon (48) and cultures from the SMG (54), although this has not been shown in secondary lymphoid organs, including the spleen. This role is also complicated by the ability of neurons to direct formation of lymphoid tissues during development and infection. Neuronal release of retinoic acid activates lymphoid stromal cells to induce the recruitment of lymphoid tissue inducer cells during development (258). This interplay between neurons and stromal cells is not restricted to development and occurs during the formation of lymphoid aggregates (tertiary lymphoid tissues) during inflammation (192). These studies suggest that the integration of nervous and immune system begins early during development and continues to shape lymphoid organ function throughout life. It is tempting to further speculate that disruption of these early processes could induce alterations in lymphoid structure and functional changes.

Sympathetic innervation of the spleen in humans and non-human primates is considerably less well characterized. For example, while electron microscopy analysis of human spleen identified neuronal fibers associated with the vasculature and trabeculae, no innervation of the white pulp was observed (109). Other studies have confirmed that as in the mouse, sympathetic neurons innervate of lymph nodes along the vasculature and the parenchyma, terminating near lymphocytes (110, 239). Performing these types of comparative anatomical studies would be significant opportunities for discovery and would likely aid the development of new therapeutics. Given the difficulties in obtaining a healthy human spleen, non-human primate models could conceivably provide significant insight.

IV. NONNEURONAL SOURCES OF NEUROTRANSMITTERS

A. T Cells as Non-neuronal Sources of ACh

Despite the controversies over the precise pathways involved, it is generally accepted that activation of vagal efferent axons in mice and rats induces release of NE, resulting in ACh release in the spleen from non-neuronal sources. Electrical stimulation of the spleen has long been known to induce the release of ACh even in mice and rats despite the lack of cholinergic innervation. The source of this ACh is from non-neuronal sources, primarily specialized subsets of B and T lymphocytes (69, 214, 224). Production of ACh had been described in transformed human and mouse lymphocyte cell lines, and in primary rat B and T cells (218). These findings suggested that given the appropriate stimuli, lymphocytes could produce and release ACh to regulate physiological processes.

In the CAIP, T cells in the spleen are critical, serving as the source of ACh released in response to neuronal NE. As evidence of this, mice lacking T cells, or recipients of transferred T cells with short hairpin RNA mediated knockdown of ChAT, were not protected from endotoxemia by VNS. T-cell expression of ChAT and production of ACh were validated using a ChAT-GFP reporter mouse where the ChAT promoter drives GFP expression (247). Analysis of the spleens and lymph nodes of these mice demonstrated a small subpopulation of CD4+ T cells were GFP+ and made ACh as determined by mass spectrometry. These CD4+ ChAT-GFP were found to express the β2-adrenergic receptor (β2AR) and respond to NE by releasing ACh, suggestive of the ability of ChAT+ T cells to respond to activation of the sympathetic innervation (224).

Subsequent studies determined that despite this novel ability, CD4+ ChAT-GFP+ T cells were not a novel T helper subset, or restricted to a specific lineage. Although flow cytometric immunophenotyping indicated ChAT-GFP+ T cells were enriched for Foxp3+ regulatory T cells, ChAT expression is not restricted to this or other Th lineages (214). Differentiation of naive CD4+ ChAT-GFP- T cells under conditions to induce Th1, Th2, Th17, or Treg T cells in vitro failed to significantly induce ChAT-GFP expression (214). More recent studies have indicated ChAT is also increased in CD4+ Th17 T cells in the small intestine (69). Curiously, this endogenous population of ChAT+ T cells in the lamina propria was not observed by confocal imaging of the intestine of a reporter mouse line where Cre recombinase was expressed under the control of the ChAT promoter allowing for tdTomato expression (91). Although it is tempting to consider that different signals and stimuli experienced by T cells in the intestine versus ex vivo could explain these differences, it is also possible that ChAT+ T cells are not a unique subset and do not fit into pre-established Th lineages.

Supporting this contention, a genetic fate map approach revealed that ChAT expression is transient, with CD4+ T cells capable of losing and re-expressing ChAT if restimulated through the T-cell receptor (214). The commensal bacteria also have a unique role in inducing expression of ChAT in T cells. Antibiotic-mediated depletion of the microbiota by treatment of mice from weaning to adulthood with a cocktail of broad-spectrum antibiotics significantly reduced the number of ChAT+ T cells (214). These cells not only respond to the microbiota, but appear to regulate it as well. Recent studies further suggest that CD4+ ChAT+ T cells function to alter diversity of the microbiota through altering antimicrobial peptide production (69). Together, these data suggest that ChAT+ T cells could be part of a dynamic host-commensal interaction.

These ChAT+ T cells have also been proposed to be part of an alternative pathway for a non-neural mechanism of vagal communication. In this proposed model, stimulation of vagal efferents elicits migration of ChAT+ T cells to the spleen from the intestinal tract. Once stimulated, the ChAT+ T cells are proposed to migrate to the spleen following an as-yet-undefined signal for homing, while continually releasing ACh. In the spleen, ACh derived from these cells then acts on α7R on splenic sympathetic terminals to induce the release of NE, inhibiting macrophage activation (170). In this model, it is uncertain how a neuronal signal could stimulate a T cell in a secondary lymphoid organ to home to the spleen. The local environment of the PP typically programs expression of adhesion molecules and chemokine receptors that will induce homing to the intestinal mucosa (180). Finally, it would seem unlikely that α7R expressed on neurons is required in the transmission of signaling for immune regulation. As evidence of this, the CAIP was abrogated in septic shock (193) and acute kidney injury (97), only in bone marrow chimeric mice lacking the α7R on cells of hematopoietic origin, but not on neurons or other radio-resistant cells. These data suggest that within the spleen, the model proposed by Tracey and colleagues (FIGURE 1), whereby T cells releasing ACh act on target immune cells through activation of α7R, remains the most realistic interpretation of the current experimental data. Critically, this does not preclude the existence of other neural immune circuits within the spleen.

FIGURE 1.

FIGURE 1.

The proposed circuitry of the cholinergic anti-inflammatory pathway. The cholinergic anti-inflammatory pathway is proposed to begin with the detection of inflammation in the periphery. This information is relayed by vagal afferent neurons to the nucleus tractus solitarius (NTS) resulting in vagal efferent (parasympathetic) outflow to sympathetic ganglia that innervate the spleen and release norepinephrine (NE). This NE activates β2-adrenergic receptor (β2AR) on choline acetyltransferase (ChAT+) T cells that function as signaling intermediaries causing ACh to be synthesized and released. Adjacent macrophages express the nicotinic α7-receptor (α7R), which when activated by T cell-derived ACh reduces NF-κB signaling and prevents tumor necrosis factor (TNF)-α production. The alternative sympathetic efferent arm is also depicted whereby the efferent arm comprises pre- and postganglionic sympathetic neurons independent of the involvement of parasympathetic neurons. DAMPs, danger-associated molecular patterns; PAMPs, pathogen-associated molecular patterns; DMN, dorsal motor nucleus of the vagus.

B. Other ACh-Producing Immune Cells

Surprisingly, ChAT-GFP reporter mice revealed that other immune cells can produce ACh. B cells in the spleen and lymph nodes are by far a larger population of ChAT-GFP+ cells compared with CD4+ T cells. In the spleen, we identified that marginal zone and follicular B cells express ChAT at a higher frequency and total number compared with T-cell population (214). The largest proportion of ChAT+ B cells was found within the B1 B cells population in the peritoneal cavity. The function of these various ChAT+ B cell types is currently unknown, although the signals the B cells respond to are vastly different from the ChAT+ T cells. Analysis of the neurotransmitter receptor expression identified that ChAT-GFP+ sorted B cells were significantly enriched in peptidergic neurotransmitter receptors and lacked adrenergic receptors (214). These data suggest that ACh from B and T cells fulfill unique physiological roles in response to specific stimuli. Similar to T cells, ChAT expression in B cells can be induced by the commensal microbiota and could suggest a unique role for these cells in host commensal interactions. As with T cells, ChAT expression in B cells was significantly reduced in mice treated with an antibiotic cocktail, and could be rapidly induced by TLR agonists. These data suggest that TLR agonists derived from the commensal microbiota regulate ChAT expression in lymphocytes (214). Further supporting the role for ChAT induction by the microbiota, B cells isolated from the spleen of ChAT-GFP mice at E18.5 were GFP- unless stimulated with TLR agonists. These data suggest that colonization of animals with the commensal microbiota at birth would appear to exert a unique influence over the expression of ChAT in immune cells. Although a small number of splenic macrophages (214) and dendritic cells (DC) express ChAT (214, 227), a recent publication did not confirm ChAT expression in DC within the Peyer’s patch and mesenteric lymph nodes (69). It is an intriguing and unexplored possibility that these differences could reflect unique populations of DC in these tissues, or that differences in the microbiota in the mouse colonies of the two laboratories. These many questions represent a significant need to better understand not only the regulation of ChAT in non-neuronal cells, but also the stimuli required for ACh release.

C. Immune Cell Production of Noncholinergic Neurotransmitters

Immune cells have been reported to synthesize and release neurotransmitters other than ACh. These include catecholamines such as NE and dopamine, to molecules such as GABA described below.

D. Norepinephrine

As opposed to simply serving as a target for NE, immune cells have been suggested to produce NE as well. Stimulation of rat neutrophils and alveolar macrophages with LPS not only elicited release of NE and epinephrine, but increased transcription of tyrosine hydroxylase and dopamine β-hydroxylase (84). As these genes encode enzymes required for the synthesis of catecholamines, these data suggested catecholamine biosynthesis and release is triggered by PAMPs. More importantly, Flierl et al. (84) demonstrated that macrophage-derived catecholamines exacerbated tissue damage in a model of lung injury in rats via activation of α2 adrenoceptors on alveolar macrophages. Together, these data suggest that phagocytes can synthesize and release catecholamines in response to PAMPs, and that autocrine signaling via α2 adrenoceptors may have proinflammatory effects. Although intriguing, TH staining in immune cells within the spleen or other secondary lymphoid organs has not been reported (185). It is important to note that there are four splice variants of TH in mice, raising the possibility that either immune cells express a unique isoform of TH not detectable by some antibodies or that macrophages cannot produce this neurotransmitter. An inability of macrophages to produce NE is supported by studies using a fate-map reporter system. Mice expressing Cre recombinase under the control of the TH promoter, and possessing a lox-stop-lox tdTomato allele were found to have tdTomato+ neurons, but not macrophages (89). These data would suggest that macrophages are not a large source of catecholamines, or that expression of the biosynthetic machinery occurs under specific circumstances.

E. Dopamine

Non-neuronal cells including B and T cells produce, take up, and respond to dopamine (151). T cells isolated from CSF and single-cell propagation of B and T cell clones contained and released dopamine upon T-cell stimulation (16). These results from long-term propagation cultures were early evidence that lymphocytes could produce dopamine, as opposed to simply taking up and storing this molecule. Subsequent studies have indicated that the enzymatic machinery required for synthesis, TH, and dopamine β-hydroxylase are expressed by lymphocytes and phagocytes (84). Again, given the requirement for TH as the rate-limiting enzyme for catecholamine synthesis, lack of TH immunoreactivity and reporter gene expression in mice would seem to contradict these findings. Uptake from the extracellular environment has also been observed with increased intracellular dopamine content after culture in media supplemented with this neurotransmitter (16). In support of this ability, human peripheral blood lymphocytes express the plasma membrane dopamine transporter DAT, and uptake extracellular 3H-labeled dopamine in a DAT-dependent manner (168). These results suggest that if the genes encoding the biosynthetic pathway are not expressed by lymphocytes, these cells could potentially function as a delivery vehicle for dopamine.

F. GABA

Although considered the main inhibitory neurotransmitter in the CNS, GABA is also present in the periphery in a variety of tissues (98) and immune cells (20, 71). Human lymphocytes express glutamate decarboxylase isoforms that synthesize GABA, the proteins involved in GABA vesicular loading, transport, and GABA receptors (71). Lymphocytes not only take up extracellular GABA; these cells also express GABAA receptors, with GABA inducing currents in patch-clamped lymphocytes and inhibition of phytohaemagglutinin-induced proliferation (71). Prior studies in mice suggested that GABA inhibition of proliferation, that could block immune responses in vivo, is predominantly through GABAA receptors (250). Although it was presumed that the GABA responsive population was predominantly T cells, enrichment of peripheral blood lymphocytes for specific populations was not performed (71).

G. Vasoactive Intestinal Peptide

In addition to neurons and endocrine cells, immune cells produce and respond to vasoactive intestinal peptide (VIP). Both CD4+ and CD8+ T lymphocytes express mRNA encoding VIP and secrete VIP in response to antigenic stimulation (66). Receptors for VIP are expressed on almost all immune cell types (67). VIP binds three metabotropic receptors: VPAC1, VPAC2, and PAC1 (90), which are coupled to Gαs, leading to the generation of cAMP and activation of protein kinase A. In addition, in monocytes and macrophages, PAC1 receptors also couple to phospholipase C and protein kinase C activation. VIP has anti-inflammatory effects on innate immune cells, enhances generation of Treg cells, and skews CD4+ T-cell differentiation towards a Th2 phenotype (90). These attributes engender a restitutive effect in vivo, with VIP significantly attenuating production of Th1-driven induced colitis (1). This ability of immune cells to produce and respond to VIP demonstrates the broad array immune functions of classical neurotransmitters and peptide hormones.

H. The Interface Between Neurons and Immune Cells

The physical interface and mechanism(s) of communication between the nervous and immune system are a topic of active investigation. Synapse-like interactions between axons and lymphocytes have been observed in the spleen and are characterized by lymphocytes within 5–15 nm of varicosities containing NE (82). Despite this seminal work, the frequency of interactions between ChAT+ B and T cells has only recently been assessed. Confocal microscopy, and CLARITY imaging of spleens from ChAT-GFP reporter mice, has shown very few ChAT-GFP+ cells are found in close proximity to TH+ axons (185). These data suggest that interactions between sympathetic axons and ChAT+ lymphocytes in the spleen are rare, or transient in nature. It is unknown if adhesion molecules maintain this interaction between lymphocytes and axons, or how long these cells interact. These questions and the functional significance of this interaction remain unexplored and are exciting areas of further study.

V. MODULATION OF IMMUNE FUNCTION BY ACETYLCHOLINE

Inhibition of immune responses through an anti-inflammatory reflex arc requires that the neuronal signaling is detected by the target immune cells and impinges on immunological processes. For the anti-inflammatory reflex, identification of the mechanism of silencing macrophage activation has proved challenging. The following sections serve to review the current state of these efforts and identify remaining opportunities and challenges to further advance the field.

A. Target Cells and Mechanisms of ACh-Induced Regulation

Reduced morbidity and mortality afforded by VNS during endotoxemia occurs by ACh-mediated inhibition of signaling in immune cells. In mouse models of endotoxemia, proinflammatory cytokine production is the basis for multiorgan system dysfunction. While injection of LPS or polymicrobial sepsis can activate a number of immune cells, splenic macrophages are implicated as the predominant source of TNF-α (119). Although macrophages express both muscarinic and nicotinic acetylcholine receptors (summarized in TABLES 2 and 3), the nicotinic alpha 7 receptor (α7R) is required for ACh-induced inhibition of TNF-α production. Experiments using the nicotinic receptor antagonist α-bungarotoxin in vitro reduced the ability of ACh to reduce LPS-induced TNF-α production. Further confirming the requirement of α7 in VNS-mediated protection during septic shock, mice lacking the α7R were not protected from LPS-induced TNF-α production with VNS (265). This mechanism of α7R-dependent inhibition has been expanded to numerous types of immune cells including neutrophils and dendritic cells. At this point, it is uncertain how many other immune cell populations express α7R, as there are few antibodies validated in KO mice. Use of fluorescent reporter protein expression to mark cells expressing α7R appears to be restricted to a fate map approach. Such an approach that uses mice with Cre recombinase expression controlled by the α7R promoter and a LoxP-STOP-LoxP-YFP allele has identified 20–25% of myeloid and lymphoid cells in the adult bone marrow, spleen and Peyer’s patches expressed α7R. As Cre-mediated excision of the STOP is irreversible, it is uncertain if these cells expressed α7R at some point or continue to express α7R. It would be interesting to use this approach together with a mouse designed to express a fluorescent protein under the direct control of the α7R promoter in a fate map approach. These types of studies could aid our understanding of α7R expression in these populations of cells and help to determine if expression occurs during immune cell development or can be triggered in the periphery.

Table 2.

Expression and function of muscarinic receptors on immune cells

Primary Cell Type Receptors Species Tissue/Source Function
B cell M1 Human PBMC ND (249)
M2 Human PBMC ND (249)
M3 Human PBMC ND (249)
M4 Human PBMC ND (249)
M5 Human PBMC ND (249)
T cell M1 Human Spleen ↑TNF-α, IFN-γ, IL-6 (130)
Mouse Spleen ↑IL-10, ↑IL-17, ↓IFN-γ (209)
M2 Human PBMC ND (53)
M3 Mouse Spleen T cell activation-host defense (63)
MLN CD4+ T cell activation, cytokine production, Ca2+ signaling (63)
Human PBMC ND (57, 131)
M4 Human PBMC ND (53, 57)
Mouse Spleen ↑IL-10, ↑IL-17, ↓IFN-γ (209)
M5 Human Spleen ↑IL-6, ↑IFN-γ, TNF-α (130)
Human PBMC ND (57)
Mouse Spleen ↑IL-10, ↑IL-17, ↓IFN-γ (209)
Neutrophils M1 Human Peripheral blood ND (13, 207)
M2 Human Peripheral blood ND (207)
M3 Human Peripheral blood ↑NETosis (45), ND (13, 207)
M4 Human Peripheral blood ND (13)
M5 Human Peripheral blood ND (13)
Macrophages M1 Human Alveolar ND (207)
Mouse Peritoneal ND (132)
M2 Human Alveolar ND (207)
Mouse Peritoneal ND (132)
M3 Guinea pig Alveolar ↑Inflammation, host defense, ↑IL-8 (203)
Human Alveolar ND (207)
Mouse Peritoneal ND (132)
M4 Mouse Peritoneal ND (132)
M5 Mouse Peritoneal ND (132)
Dendritic cells M1 Mouse Bone marrow derived ND (132)
M2 Mouse Bone marrow derived ND (132)
M3 Human Nasal mucosa ↑OX40L (158)
Human PBMC derived Differentiation: ↑TNF-α, ↑HLA-DR; LPS stimulation: ↓TNF-α, ↓HLA-DR (227)
Mouse Bone marrow derived ND (132)
M4 Human PBMC derived Differentiation: ↑TNF-α, ↑HLA-DR; LPS stimulation: ↓TNF-α, ↓HLA-DR (227)
M5 Human PBMC derived Differentiation: ↑TNF-α, ↑HLA-DR; LPS stimulation: ↓TNF-α, ↓HLA-DR (227)

IL, interleukin; LPS, lipopolysaccharide; TNF-α, tumor necrosis factor-α; IFN-γ, interferon-γ; ND, no difference; HLA-DR, human leukocyte antigen-antigen D related; PBMC, peripheral blood mononuclear cells; MLN, mesenteric lymph nodes.

Table 3.

Expression and function of nicotinic receptors on immune cells

Primary Cell Type Receptors Species Source Function
B cell α2 Mouse Spleen ND (132)
α3 Human PBMC ND (231)*
Mouse Spleen ND (133)
α4 Mouse Spleen ↓Antigen specific Ab (238)
α5 Human PBMC ND (231)
Mouse Spleen ND (132)
α6 Human PBMC ND (231)*
Mouse Spleen ND (132)
α7 Human PBMC ND (231)*
Mouse Spleen ND (132, 238), ↓Antigen specific Ab (88)
α9 Human PBMC ND (231)
Mouse Spleen ND (132)
α10 Human PBMC ND (231)
Mouse Spleen ND (132)
β2 Mouse Spleen ↓Antigen specific Ab (238)
T cells α2 Mouse Spleen ND (132)
α3 Human PBMC ND (231)*
Mouse Spleen ND (132)
α5 Human PBMC ND
Mouse Spleen ND (132)
α6 Human PBMC ND (231)*
α7 Human PBMC ND (231)*
Mouse Spleen ND (132), ↓Proinflammatory cytokines (88)
α9 Human PBMC ND
Mouse Spleen ND (132)
α10 Human PBMC ND (231)
Mouse Spleen ND (132)
Neutrophils α2 Mouse Peritoneal ↑Ca2+ signaling (226)†
α3β2 Human PBMC ↑Reactive oxygen intermediates, ↑IL-8 production (121)
Mouse Peritoneal ↑Ca2+ signaling (226)†
α4 Human PBMC ↑Reactive oxygen intermediates, ↑IL-8 production (121)
Mouse Peritoneal ↑Ca2+ signaling (226)†
α5 Mouse Peritoneal ↑Ca2+ signaling (226)†
α6 Mouse Peritoneal ↑Ca2+ signaling (226)†
α7 Human PBMC ↑NETosis (147)
Mouse Peritoneal Regulation of inflammatory responses, ↑Ca2+ signaling (226)†
α9 Mouse Peritoneal ↑Ca2+ signaling (226)†
Macrophages α2 Mouse Peritoneal ND (132)
α4β2 Human Alveoli ↓IL-6, IL-12, and TNF-α
Peritoneal and mucosa ↑Phagocytosis, ↓NF-κB signaling, ↓ TNF-α, ↓CCL2, ↑IL-10 (259)
Mouse Peritoneal ND
α5 Mouse Peritoneal ND (132)
α6 Mouse Peritoneal ND (132)
α7 Human PBMC ↓NF-κB activation, ↓TNF-α (265)
Mouse Peritoneal/spleen ND (132), ↓NF-κB activation, ↓TNF-α (265); ↓ATP-induced IL-1β cleavage (163)
Muscularis externa ↓IL-1α, IL-1β, IL-6, and CCL2, ↑ATP-induced Ca2+ transients (173)
α10 Mouse Bone marrow derived ND (132)
Dendritic cells α2 Mouse Bone marrow derived ND
α5 Mouse Bone marrow derived ND
α6 Mouse Bone marrow derived ND (132)
α7 Mouse Spleen ↓IL-12, ↓activation of T cells, ↓proinflammatory cytokine release (184)
Bone marrow derived ND
α10 Mouse Bone marrow derived ND (132)
*

Mixed population of cells used, impossible to determine precise immune cell phenotype. †Specific receptor not determined. Ab, antibody; IL, interleukin; TNF-α, tumor necrosis factor-α; ND, no difference; PBMC, peripheral blood mononuclear cells.

B. Intracellular Mechanisms of ACh-Mediated Control

Inhibition of TNF-α production by ACh-elicited α7R-dependent signaling has been suggested to involve a wide range of mechanisms from sequestering of signaling proteins, to induction of microRNAs. These multiple proposed mechanisms are likely a consequence of the α7R being a homopentamer of α7 subunits appearing to be an ion channel, with ligand binding also activating intracellular signaling cascades.

As TNF-α production is dependent on NF-κB-mediated signaling and gene transcription, the focus has been on α7R-induced disruptions to this signaling cascade. A key role for the signal transducer of activated T cells 3 (STAT3) has been implicated, in studies with conflicting proposed mechanisms. Nicotine, a nonselective agonist of nicotinic receptors, significantly increased STAT3 phosphorylation while reducing LPS-induced TNF-α production in macrophages (64). Inhibition of TNF-α production by nicotine was dependent on α7R, as pretreatment with selective antagonists restored TNF-α production. These data suggested that α7R activation increases STAT3 phosphorylation, consequently blocking NF-κB-dependent gene expression. Increased STAT3 phosphorylation was also observed in vivo in intestinal muscularis macrophages following VNS (64). VNS-induced activation of STAT3 was crucial to limiting postoperative ileitis, as conditional ablation of STAT3 in macrophages abrogated the protective effect of this treatment. Stimulation of α7R was suggested to increase STAT3 phosphorylation by recruitment and activation of Janus associated kinase-2 (JAK2). Currently, the mechanism responsible for this α7R-induced aggregation of JAK2 is unknown.

In contrast to these findings, α7R agonists prevented LPS-induced STAT3 phosphorylation by JAK2 in the THP-1 macrophage cell line. Inhibition of STAT3 phosphorylation by treatment with α7R agonists or JAK2 inhibitors, siRNA targeting of STAT3, or expression of a STAT3 phosphorylation mutant prevented TNF-α production. These studies suggest that LPS-induced STAT3 phosphorylation is required for optimal NF-κB signaling (202). It is important to note in these studies that the α7R agonist choline reduced TNF-α production in splenocytes and the RAW264.7 cell line at concentrations that did not reduce STAT3 phosphorylation. In support of this proposed mechanism, the nicotinic receptor agonist GTS-21 reduced IL-6-induced JAK2 activation and STAT3 phosphorylation (49). Although induction of STAT3 signaling is not the canonical response to LPS, there is a precedence of LPS-induced JAK2 activation in the RAW264.7 cell line (136). At present, it is not possible to reconcile these proposed roles of STAT3 in α7R-depenant signal transduction, although it is possible that there are multiple downstream signaling pathways induced depending on the strength of the stimulation.

As another alternative, α7R agonists have also been implicated in altering posttranslational processes that would reduce TNF-α release. TNF-α requires cleavage and activation by the enzyme TNF-α converting enzyme (TACE) to become biologically active. Stimuli that reduce TACE expression or activity are therefore able to indirectly block the release of biologically active TNF-α (35). Stimulation through the α7R increased expression of the microRNA mir124, which targeted TACE mRNA, reduced TACE expression, and consequently decreased TNF-α processing (245). Although this study would not account for prior observations of reduced NF-κB DNA binding upon α7R stimulation, it nonetheless highlights that alternative pathways should be explored in immune regulation following α7R stimulation.

Other mechanisms could include a range of cellular responses ranging from activation of kinases to secondary messengers. Activation of T cells through nicotinic receptors can result in intracellular signaling events characterized by kinase activation. T cells treated with nicotine activated T-cell phosphorylation cascade proteins, and this increase was dependent on the SRC family kinases Lck and Fyn leading to Ca2+ mobilization (213). While proposed that the α7R and TCR were physically associated, it is unknown if this is a direct or indirect interaction, or how nicotine signaling through α7R induces activation of the T-cell signaling kinases. As T-cell signaling is a highly coordinated process that dictates T-cell function, alterations can have significant ramifications on T-cell-dependent immune function (37).

With the (conflicting) results in the literature, and lack of studies on the structure and signaling elicited by α7R in immune cells, perhaps it would be prudent to examine signaling in neuronal cells. Activation of the α7R following ligand binding allows for opening of the ion channel and influx of sodium or calcium ions, yet it is uncertain if this occurs in non-neuronal cells. Although whole cell patch-clamping experiments have failed to detect α7-induced currents in lymphocytes (262), single-channel openings have been observed in monocyte-derived macrophages following stimulation with nicotine, ACh, and ACh with or without α7R-positive allosteric modulators (9). These studies illustrated that while inward currents can be observed, this occurred with agonist concentrations that are insufficient to prevent TNF-α production. These data indicate that stimulation of immune cells with α7R agonists elicits multiple cell signaling events and that mechanisms other than ion channel opening are critical for regulating the response to DAMPs.

Other mechanisms of α7R elicited signal transduction that have also been described in neurons. Stimulation of hippocampal neurons with α7R agonists increases adenylate cyclase isoform 1, increasing intracellular cAMP (52). Activation of the membrane-associated adenylate cyclase isoform 6 has also been reported following stimulation of PC12 cell line with α7R agonists. While it is by no means certain that a signaling axis in one cell type will be conserved in immune cells, it is worth noting that increased cAMP can inhibit the activation of a variety of immune cells. In T cells, increased cAMP production prevents signaling from the T-cell receptor, followed by increased PKA-induced phosphorylation of Lck at tyrosine 505 (225, 260, 261). This phosphorylation blocks T-cell activation, preventing differentiation, cytokine production, and proliferation. Inhibition of macrophage activation by cAMP-elicited signaling can also occur through induction of c-Fos binding to NF-κB p50. This sequestration prevents NF-κB-mediated gene transcription reducing production of proinflammatory cytokines (137). It is also important to note that increased cAMP in immune cells does not always lead to immune suppression. In T cells, increased cAMP and downstream signaling cascades with prolonged exposure to IL-1β enhances differentiation into IL-17A producing Th17 T cells (153). Stimulation of α7R should not always be equated with immune inhibition, and the signaling pathways and mechanisms of action require substantial investigation in well-defined primary immune cells.

C. Activation of Other ACh Receptors on Immune Cells

Immune cells have been reported to express a wide array of muscarinic and nicotinic acetylcholine receptors that enhance or decrease the activation. Despite this reported expression of nicotinic and muscarinic receptors, and an ability to respond to these agonists, this expression is not reflected in databases such as the immunological genome project (106). At this point it is unknown if this lack of expression is due to purely technical or biological reasons. For example, is lack of receptor mRNA due to processing induced transcriptional changes that can accompany isolation, or are receptor positive cells a small fraction of the overall population. Despite these concerns, we have included literature that identifies expression by a combination of techniques including functional assays and summarized these findings for muscarinic receptors in TABLE 2 and nicotinic receptors in TABLE 3.

1. Muscarinic receptors

Various immune cells have been found to express muscarinic receptors, with selective agonists impacting cellular functions (TABLE 2). As professional antigen presenting cells that have a unique ability to stimulate naive T cells bearing an appropriate TCR for an antigen, DC are fundamental to immune responses. Human DC derived from blood CD14+ monocytes express muscarinic receptor types M3, M4, and M5. Stimulation during differentiation into DC, or stimulation of mature DC cultures with the muscarinic receptor agonist carbachol increased HLA-DR and TNF-α expression. While suggestive of a proinflammatory state, expression of costimulatory proteins such as CD80 expression were unaltered (227). Muscarinic receptor stimulation may therefore regulate specific aspects of DC function and depend on the local environmental stimuli. In support of this, carbachol decreased LPS-induced HLA-DR, IL-12, and TNF-α expression (227), suggesting a highly complex signaling network, and an ability to either enhance or reduce T cell responses depending on coexposure to DAMPs and muscarinic agonists. Complicating our understanding of the effects of muscarinic receptors on adaptive immune function in vivo, T cells have also been identified to express muscarinic receptors that modulate cell function. The M3 receptor is expressed on CD4+ T cells and is required to mount an immune response directed against helminths and enteric bacterial pathogens. Mice deficient in M3 receptor (M3R) are more susceptible to infection with Nippostrongylus brasiliensis and Salmonella enterica serovar Typhimurium (63). Deficiency in M3R reduces T-cell Ca2+ signaling and activation, culminating in decreased IL-13 and interferon (IFN)-γ production during N. brasiliensis and S. typhimurium infection, respectively. T-cell intrinsic M3R signaling is crucial to this response, as adoptive transfer of WT T cells but not M3R−/− T cells from previously infected donor mice into M3R−/− hosts significantly reduced parasite burden after challenge with N. brasiliensis. These data would suggest that ACh signaling through the T cell M3R, and not host cell M3R is responsible for this phenotype (63).

In addition to T cells, muscarinic receptors are also expressed on B cells. Although radiolabeled 3-quinuclidinyl benzilate bound mouse splenocytes and enriched B cells (6), binding was not observed with N-methylscopolamine to human B cells (77). As both agents bind to muscarinic receptors, these data have been interpreted as mouse B cells, but not human B cells are sensitive to muscarinic receptor agonists. More recent immunocytochemistry studies have suggested surface expression of M1-M5 receptor subtypes on human B cells (249). The immunological ramifications of stimulation through these receptors are not well characterized in the steady state or during infection.

Similar to these cells of the adaptive immune system, innate immune cells express muscarinic receptors and have been implicated in host defense and immune function. Deficiency in M3R significantly increases susceptibility to the mouse enteric pathogen Citrobacter rodentium, increasing bacterial burden and reducing clearance. Increased susceptibility in this model of enteric bacterial infection is due to reduced macrophage function and mucus production, and not a T-cell intrinsic defect (176). Expression of M1–3 receptors is not limited to macrophages in the intestine, with expression also found in the lung macrophages of mice and humans (102, 207). Stimulation of alveolar macrophages with ACh induced in vitro migration, that was blocked by cotreatment with the M3 receptor antagonist tiotropium (38). This role for ACh activation of M3R to increase inflammation or host defense has also been suggested in studies where tiotropium administration reduced lung inflammation induced by repeated instillation of LPS as a model of chronic obstructive pulmonary disorder (203). It is important to note, however, that the cell population targeted by ACh in this study was not elucidated. Functionally, stimulation through these receptors have been suggested to aid host defense by increasing macrophage production of leukotriene B4, a chemotactic signal for neutrophils (207). It is however important to note that increased leukotriene B4 production was only observed in sputum cells from patients with chronic lung pathology, and not samples from control subjects. Despite clear functional evidence of the role of muscarinic receptors in macrophage function, it is interesting to note that expression may be linked to differentiation within tissues. As evidence of this possibility, circulating monocytes that differentiated into macrophages after recruitment do not express any of the five muscarinic receptor subtypes (102). These findings could suggest that muscarinic receptor expression by macrophages could be tissue specific, occurring in response to local signals from tissue.

Other innate immune cells including neutrophils in peripheral blood express M1R and M2R, with varying reports of M3R (207) as well as M4 and M5 (13) expression. Neutrophils protect against pathogens through a variety of mechanisms including phagocytosis and production of free radicals such as superoxide and peroxynitrate. It is unclear if these fundamental immune functions are enhanced or reduced following stimulation by muscarinic receptor-selective agonists. In addition to phagocytosis and free radical production, neutrophils can also produce neutrophil extracellular traps (NETs) to capture and to kill extracellular pathogens (36). NETs are globular structures comprised of neutrophil DNA, histones, and other host cell proteins (36). Release of NETs has been observed following stimulation with ACh through the M3 receptor (45). While tempting to conclude that this response could aid host defense during infection, NETs release can also result in immunopathology (45, 59, 187). These studies should illustrate that despite the tremendous advances in the role of muscarinic receptors in immune modulation, there are significant gaps in our knowledge.

2. Nicotinic receptors

In addition to muscarinic receptors, numerous studies have identified nicotinic ACh receptor expression on immune cells. These data are summarized in TABLE 3, although it is important to note that for some of these receptors, the functional implications and the precise cell types remain unknown.

Nicotinic receptors containing α4/β2 subunits have been implicated in immunomodulation during intestinal inflammation. Activation of α4/β2 receptors by VNS initiated an anti-inflammatory effect, increasing phagocytic activity and IL-10 production while reducing NF-κB signaling in lamina propria macrophages (259). Thus VNS-induced immunosuppression in the intestinal tract can occur through receptors other than α7R. Activation through nicotinic receptors on macrophages however should not be equated to general host protection, as stimulation of nicotinic receptors in lung alveolar macrophages with nicotine or the nicotinic agonist dimethylphenylpiperazinium reduces bactericidal activity and proinflammatory cytokine production (172).

Similar reductions in macrophages phagocytosis and host defense against bacterial pathogens have been described with cigarette smoke. While nicotine has been implied as the agent of action, cigarette smoke comprises an estimated 7,357 chemical compounds, and as such, it is difficult to determine if the observed effects are due to the presence of multiple agonists of different receptors (221). It is also possible that while alveolar macrophage function is significantly reduced, other populations of macrophages and inflammatory monocytes may not express the same neurotransmitter receptors and may respond in a different manner. In support of this supposition, expression of surface proteins and receptors such as mouse mannose receptor are differentially expressed by alveolar macrophages compared with other macrophage populations (117). Macrophage function and responses to agonists may therefore depend on local tissue environment.

The composition of the nicotinic receptor also has drastic implications for ligand binding and functionality. Nicotinic receptors are comprised of five receptor protein subunits that assemble to form a pentameric structure with a ligand-binding site and channel (4, 181). Unlike the homeric α7R, heteromeric receptors, including α4β2, are comprised of different subunits and can adopt a variety of stoichiometries that impact upon function. The α4β2 receptor comprises three α4 and two β2 [(α4)2(β2)3] with an alternative (α4)2(β2)3 configuration that has high agonist sensitivity (181). Receptor configuration is not static, with prolonged nicotinic agonist exposure increasing the proportion of (α4)2(β2)3 receptors (188). It is uncertain if different immune cell populations or macrophages in lung compared with other organs have unique stoichiometries that could impact signaling characteristics.

Indirect effects of acetylcholine on immune cell populations should also be considered in vivo, and significantly complicate interpretation at the tissue or organism level. Despite reduced activity of alveolar macrophages, nicotinic receptor stimulation significantly increased expression of proteins chemotactic for neutrophils. This suggests that while alveolar macrophage function could be reduced, overall host defense as provided by neutrophils might remain intact. Nicotine as a nonselective agonist can also significantly alter the functionality of neutrophils. While leukocytosis has been well documented in tobacco smokers, some studies have identified increased responsiveness to chemokines (254) as well as production of chemokines and reactive oxygen species (ROS) by neutrophils (121, 125). Stimulation of neutrophils from healthy nonsmoker volunteers with nicotine significantly increased IL-8 and superoxide production in vitro (121). It is worth noting that while the dihydrorhodamine 123 (DHR) reagent has become a gold standard in the evaluation of ROS there are several concerns that complicate the use of DHR to accurately determine ROS generation (127). In neutrophils, nicotine has also been reported to enhance release of lysozyme, neutrophil elastase, β-glucuronidase, prostaglandin E2, and the leukotrienes. While increased release of these compounds could be viewed as beneficial for host protection, nicotine treatment of macrophages decreased phagocytosis while having no effect on chemotaxis to chemokines, or cell viability (236). At this time, it is unknown what receptors are required for these increased host-defense responses to occur. Curiously, nicotine has also been reported to reduce the bactericidal activity of human peripheral blood neutrophils. Exposure of cells to high concentrations of nicotine that would be found in chewing tobacco or smokeless tobacco products reduced bacterial killing by directly absorbing superoxide (194). Together, these data indicate the continued need to carefully define the receptors required for these effects of nicotine on the vital functions of innate immune cells.

VI. OTHER NEUROIMMUNE CIRCUITS

Although the vagally activated CAIP has been the focus of intense interest and study with application to regulation of immune activity in a variety of conditions, it is becoming increasingly clear that other reflexes occur.

A. Neural Modulation of Intestinal Immune Function

The interactions between the nervous and immune system are particularly interesting in the gastrointestinal tract. The gastrointestinal tract contains a large component of the immune system in close proximity to the nerve terminals of enteric neurons and adrenergic postganglionic sympathetic neurons with cell bodies located in prevertebral ganglia. As an interface between the body and the environment, the immune system in the gastrointestinal mucosal must balance the requirement to protect against pathogenic microbes, while being tolerant to components of the diet and commensal microbiota. Inappropriate immune reactions to the commensal microbiota, or components originating from the diet have the potential to have significant detrimental effects on the health of an individual. While the precise etiology of IBD has remained elusive, it is generally accepted that aberrant immune system activation is an underlying cause (3, 271). There are an estimated 4–4.5 million people in the United States and Europe with IBD (128). This chronic inflammation presents clinically with abdominal pain, fever, diarrhea, fatigue, and rectal bleeding (189). Treatments typically are designed to reduced inflammation in general and include nonsteroidal anti-inflammatory drugs, biologicals targeted to specific cytokines, or surgical intervention (3, 21). Although monoclonal antibodies to TNF-α have been revolutionary in the treatment of IBD, these therapeutics can lose efficacy over time and increase the risk of infection (3, 22, 189). Such limitations demonstrate the continued need for novel therapeutics that target the underlying mechanisms of inflammation (85).

Several well-characterized spontaneous genetic and induced models of IBD have been developed that serve to interrogate unique aspects of the disease process. While spontaneous disease models typically have increased variability and incomplete penetrance, induced colonic inflammation models have been developed that mimic specific aspects of pathology. Here we briefly describe these models; however, it is well beyond the scope of this review to detail the nuances of each system, and interested readers are referred to one of several excellent reviews specifically on this topic (135, 162, 244). Acute or chronic inflammation can be induced by exposure to chemical irritants such as dextran sodium sulfate (DSS), or through the adoptive transfer of colitogenic T cells (215, 216, 244). DSS is toxic to intestinal epithelial cells (IEC) resulting in a mid-distal colonic ulceration of the mucosa, with inflammation due to monocytes/macrophages and neutrophil recruitment and activation (17, 191, 252), without requiring T cells (129). While this colitis model has provided insights into the mechanisms of innate immune cell activation and healing of the colonic mucosa, the lack of T-cell involvement provides an incomplete picture. Consequently, an induced T-cell-dependent colitis model was developed using adoptive transfer of FACS sorted effector T cells (TE) into immunodeficient (RAG1−/−) hosts (244). It is important to consider that therapeutics that reduce disease severity or do not appear efficacious in one model may only reflect an inability to modulate a specific aspect of disease.

With intestinal inflammation in preclinical models and patients with IBD occurring due to overexuberant immune activation, immune modulation with VNS has emerged as a potential therapeutic. Endogenous vagal nerve signaling has been implicated in neuroimmune regulation of colonic inflammation; with vagotomy significantly increasing the severity of DSS-induced disease (70, 93, 95). Similarly, activation of the efferent arm through central administration of the cholinesterase inhibitor galantamine significantly reduced the severity of DSS-induced colitis in an α7R-dependent manner (126). This role of α7 nicotinic receptors in mediating the vagal anti-inflammatory effects on colitis remains controversial, with evidence in support of (93, 96) and refuting the importance of α7 receptors (70). It is difficult to reconcile these differences in the requirement of α7R in inhibition of DSS-induced colitis. These conflicting results could be due to a range of factors including the differences in commensal microbes or enteric viruses (41) at different institutions. These studies also suggest clear differences in the ability of the CAIP to modulate innate versus adaptive immune cells during colitis. While vagotomy increased DSS-induced colitis severity, no effect of vagotomy was seen in weight loss, pro-inflammatory cytokine production in a T-cell transfer-induced colitis (70). Since the DSS and T-cell transfer models have distinct immunological mechanisms, these results suggest that separate neuroimmune reflexes could inhibit the function of specific types of immune cells in the gut.

Neuronal inhibition of macrophage function that requires the vagus nerve has also been demonstrated in inflammation-induced inhibition of small intestinal motility. Muscularis macrophages, which have been implicated in the pathogenesis of postoperative ileus, are associated with the enteric nervous system and interstitial cells of Cajal (89, 173, 177, 178, 183). These muscularis macrophages are closely juxtaposed to cholinergic axons, evidenced by ChAT staining, and receive inputs from parasympathetic neurons with cell bodies in the DMN (173). Electrical VNS reduced postoperative ileus independently of the spleen, and required expression of α7R on cells of hematopoietic origin that are presumably macrophages (64, 173). These data were interpreted as vagal efferent axons activate cholinergic enteric neurons, which in turn release ACh to inhibit muscularis macrophage activation (FIGURE 2). The circuitry of this neural anti-inflammatory pathway thus appears to differ from the vagal anti-inflammatory pathways that operate during colitis and sepsis. Further suggestive of multiple neuroimmune circuits that could regulate colonic inflammation, a sympathetic neuronal pathway has been described. Elegant experiments using selective sympathectomy in mice before induction of experimental colitis by DSS administration enhanced disease activity index scores, without a concomitant increase in weight loss, or production of proinflammatory cytokines (269). In a similar manner, implantation of a cuff electrode to facilitate stimulation of the superior mesenteric nerve reduced disease activity scores, without reducing the histopathology, endoscopic damage, or cytokine production (269). Although neural stimulation was sufficient to evoke physiological changes, it is uncertain if the selected stimulation parameters were appropriate for regulation of overt inflammation. It is also conceivable that the DSS model of inflammation was simply too severe for this treatment approach. As this exciting technology matures, it may be possible to tailor nerve stimulation parameters to ameliorate specific types of intestinal inflammation. These data highlight the complexity of neural regulation of immune responses in the intestinal tract, and the value of in-depth characterization of the immune cell populations that interact with neurons and are affected by neuronal signaling in situ.

FIGURE 2.

FIGURE 2.

Vagus nerve mediation regulation of intestinal inflammation. Neural inhibition of intestinal inflammation can also occur by a vagal efferent to enteric nervous system (ENS) circuit. Activation of efferent vagus nerve activity induces activation of cholinergic neurons in the ENS and release of ACh in the intestine, blocking macrophage activation in an α7-receptor (α7R)-dependent manner. TNFα, tumor necrosis factor-α.

The influence of vagal immunomodulatory pathways is also evident in the maintenance of immune homeostasis outside the context of inflammatory diseases. Oral tolerance, which prevents local and systemic immune responses to innocuous antigens that have been administered orally, is a critical determinant of health (195). As further evidence of the role of the vagus in intestinal immune function, vagotomy, but not deficiency of α7 nicotinic receptors, prevented the development of oral tolerance to ovalbumin due to a reduction in antigen specific regulatory T cells (70). These findings would suggest that vagal signaling induces a tolerogenic state to antigens derived from diet and microbiota in the intestinal tract.

Communication between the nervous and immune systems may also be involved in host responses to infectious agents in the intestinal tract. The activity of sympathetic neurons innervating the small intestine were found to increase significantly 2 h post-infection with a strain of Salmonella typhimurium unable to invade host cells. These results would suggest that the presence of a bacterial pathogen in the intestinal lumen can be detected, and integrated by the nervous system. As muscularis macrophages in the duodenum, jejunum, and ileum are closely associated with the sympathetic axons of the myenteric ganglia (89, 183), it was proposed that enteric bacterial pathogens in the intestinal lumen could result in neuronal signaling to muscularis macrophages. The functional consequences of sympathetic neural activation were decreased colonic motility, and induction of an alternatively activated (M2) macrophage differentiation program (89). Sympathetic induction of an M2 profile in muscularis macrophages was dependent on expression of β2AR but not α7R (FIGURE 3). Treatment of WT but not β2AR KO cells with NE or salbutamol increased this gene profile. Supporting in vivo data indicated increased expression of genes common to M2 macrophages including arginase and the scavenger receptor Chi3l3 2 h post-infection with S. typhimurium. These exciting data raise the possibility that pathogen detection can occur rapidly and begin to induce a M2 macrophage transcriptional program at sites distant from the area of infection. Although the afferent mechanism was not elucidated, the commensal microbiota was previously shown to induce muscularis macrophages to activate enteric neurons and induce intestinal contractility (183). It is uncertain if S. typhimurium infection can result in macrophage-induced activation of neurons to effect macrophages distant to the site of infection, or if this neuronal induced M2 profile is truly host protective during enteric bacterial challenge.

FIGURE 3.

FIGURE 3.

Sympathetic innervation of the intestine promotes anti-inflammatory macrophage differentiation. Neuroimmune communication in the intestine requires close approximation of sympathetic axons in and macrophages in the muscularis. Activation of sympathetic innervation following infection induced release of norepinephrine (NE) from nerve terminals and induction of an anti-inflammatory macrophage gene profile. β2AR, β2-adrenergic receptor.

It is unclear if only the sympathetic neurons in the myenteric plexus and muscularis macrophages are activated during infection with S. typhimurium or other pathogens. Prior three-dimensional imaging studies of mouse ileum revealed TH+ axons in the submucosa, with some, albeit sparse innervation in the macrophage rich mucosa (87). This sparse innervation in the naive mice does not seem to be static, with DSS-induced inflammation increasing the density of sympathetic axons in the colonic lamina propria (48). As an additional complication, tissue resident macrophages are phenotypically distinct from inflammatory monocyte-derived macrophages that have been recruited to a site of inflammation. It is well appreciated that while tissue resident macrophages in the intestine initially are derived from yolk-sac progenitors, these cells are eventually replaced after weaning by monocytes from the blood throughout life (10) and after inflammation (219). It is unknown how inflammation and changes in macrophage phenotypes would impact the bidirectional macrophage-neuronal communication. As evidence for how different populations of macrophages can uniquely respond to neuronal signals, VNS increases phagocytosis by lamina propria macrophages in a α7R-independent manner (259). It is also uncertain if the enteric glia participate in these reflexes in their capacity as neuronal support cells. Enteric glial cells are tightly integrated with the enteric nervous system, exerting regulatory control over colonic physiology, and are activated in tissues with overt inflammation (232, 237). Activation of these cells has also been suggested to contribute to inflammation during DSS-induced colitis and in patients with IBD (56, 76, 237). There is a significant gap in our knowledge of the influence of inflammation on the neuroimmune circuitry in the intestine, and this represents tremendous opportunity for future studies.

B. Activation of Sciatic-Vagus Adrenal Pathway

The requirement for surgical implantation of electrodes and vagal nerve stimulators has been suggested to limit clinical applications during acute inflammation (253). In response, the ability to activate neuroimmune reflex arcs without device implantation have been championed. Electrical stimulation of the sciatic nerve using electroacupuncture was found to reduce LPS-induced macrophage activation, as indicated by reduced proinflammatory cytokine production in mice (253). Protection afforded by electroacupuncture was voltage dependent and required local nociceptive sensory afferents and an intact sciatic nerve. Functional mapping of this circuitry further identified a requirement for the vagus, as cervical or subdiaphragmatic vagotomy abolished the protective effects. Unlike the vagovagal-splenic CAIP, electroacupuncture-induced protection required the adrenal medullae but not the spleen (253). Functional parasympathetic innervation of the adrenal gland was suggested by prior neuroanatomical tracing in rats, with labeling in the DMN after fast blue injection into the adrenal medulla (58, 190). Although dopamine, norepinephrine, and epinephrine were released with sciatic electrical stimulation, only activation of dopamine D1 receptors were required to inhibit TNF-α production (253). Subsequent studies have demonstrated that D1 receptor activation increases intracellular cAMP, attenuating systemic immune responses (273). Highlighting that this pathway is discrete from the other reflexes, electroacupuncture-induced inhibition of TNF-α production was still observed in α7R knockout mice. From these studies there appear to be other immune modulatory neural circuits, beyond the classically described CAIP that involve the vagus nerve.

C. Neuroimmunomodulation of Renal Injury

Evidence of additional neuronal circuits that can prevent immunopathology have been obtained in models of renal ischemia and reperfusion injury (IRI). Experimentally induced ischemia followed by reperfusion of the kidney results in activation and infiltration of immune cells that contribute to organ damage and physiological dysfunction (29). While activation of the CAIP can prevent renal injury, kidney IRI has provided evidence of alternative functional circuitry regulating immune function (122). In addition to the protective CAIP activated by VNS, activation of afferent vagal and the sympathetic renal nerves was observed. Selective afferent stimulation was equally efficacious in protecting against IRI compared with efferent or intact VNS (122). These studies highlight that while efferent VNS is sufficient to induce protection, it is not necessary.

Prior neuroanatomical studies indicated a group of neurons that reside in the ventrolateral medulla, named C1 neurons, can be activated by inflammation and stress and innervate the DMN of the vagus (2, 101). With this in mind, the contribution of C1 neurons to anti-inflammatory responses in kidney IRI was evaluated using an optogenetic approach. This technique utilizes expression of the light-gated cation channel, Channelrhodopsin-2 (ChR2). Pulsing of neurons expressing this protein with specific wavelengths of light can therefore be used to drive activation (32). With the use of this approach, a protective renal sympathetic circuit was activated in mice by optogenetic activation of adrenergic C1 neurons (2). Similar to the cholinergic anti-inflammatory pathway during sepsis, protective effects required α7 nicotinic, β2 adrenergic receptors, and the spleen. While optogenetic simulation of these catecholaminergic C1 neurons increased action potential discharge in parasympathetic and sympathetic pathways, subdiaphragmatic vagotomy did not abrogate protection from IRI. This suggested that protection was mediated predominantly through activation of sympathetic neurons.

As C1 neurons were known to be involved in the response to stressors, protection against renal IRI could also be activated by brief restraint stress. T cells were implicated as being critical to this pathway, with reduced renal damage occurring following transfer of splenic T cells. This protection afforded by T-cell transfer was further enhanced if donor mice had been subjected to restraint stress, or if CD4+ T cells were exposed to NE in vitro before transfer (2). While this has been interpreted as protection is due to activation of these CD4+ T cells and may even hint at a “memory” effect, it is uncertain if protection is due to a neuronal anti-inflammatory reflex. These data would suggest that once triggered by NE, ChAT+ T cells release ACh and continue to do so following adoptive transfer. Moreover, it is uncertain how many ChAT+ T cells would be transferred in this paradigm, given that these cells comprise 1–4% of the CD4+ T cells in the spleen (214, 224). These data suggest that sympathetic neurons are sufficient to activate the anti-inflammatory pathway via activation of β2 receptors on splenocytes, which in turn leads to release of ACh from CD4+ T cells and the activation of α7 nicotinic receptors (FIGURE 4). This motor pathway is very similar to that described for sympathetic immunomodulation during sepsis. However, although the findings of Abe et al. (2) suggest the improvement of renal ischemia reperfusion injury is due to an anti-inflammatory effect, direct measures of immune cell activation were not reported. Prior studies have noted that VNS affords protection during IRI of the kidney by reducing neutrophil infiltration and increasing M2 macrophage differentiation (122), although the role of these immune cells in this model does not appear to be certain at this point (124). Resolving the mechanism of protection in renal IRI will be crucial in our understanding of a new neuroimmune circuit, or to refine previously proposed models. It will also be important to determine whether stimulation of C1 neurons can suppress inflammation in models of RA and IBD.

FIGURE 4.

FIGURE 4.

A sympathetic pathway in the inhibition of systemic inflammation. C1 neurons residing in the medulla oblongata have been implicated as another coordination center that initiates descending signals to reduce immune activation. It is important to note that the immunomodulatory circuitry of this motor pathway does not require the vagus nerve. NE, norepinephrine; TNFα, tumor necrosis factor-α; β2AR, β2-adrenergic receptor; α7R, α7-receptor.

VII. CLINICAL APPLICATIONS OF VNS

Therapeutic devices that stimulate nerves to alleviate disease have become an intense focus in recent years, captivating the imagination and hopes of the lay public and general scientific community. These hopes have been buoyed by several companies seeking to leverage preclinical data to bring clinical neural stimulators for inflammatory disease to market (78). The abundance of positive results in preclinical models of immunopathologies using electrical VNS have prompted considerable effort in bringing these devices to clinic. Clinically, VNS commonly refers to the surgical implantation of helical or cuff electrodes around the left cervical vagus, with the electrodes connected to a stimulus unit. To date, these trials have yielded promising albeit mixed results. As an example of this promise, pilot studies using VNS for the treatment of IBD in patients reported reduced disease severity. In therapy, naive patients using VNS as an alternative to anti-TNF-α, five of seven patients showed significant improvements in the Crohn’s disease activity score, endoscopic evaluation, reduced serum C-reactive protein, and fecal calprotectin (28). At cessation of the study, four of these five responding patients had entered into remission of disease. While the two patients that were refractive to VNS treatment were removed from the study after 3 mo, this outcome may also provide unique insight into VNS as a therapeutic. It is noted that these patients had the most severe disease of this cohort, suggesting that VNS is potentially better suited for maintenance of remission. It is also worth considering if enough preclinical information is available to understand the functional circuitry, and optimal electroceutical stimulation parameters for immunosuppression.

VNS has also proven efficacious in preventing immune cell activation in response to PAMPs or in patients with RA. In patients with epilepsy that were implanted with a VNS device, active stimulation reduced TNF-α production from isolated leukocytes in response to LPS up to 4 h after VNS. Reduced responsiveness to LPS following VNS was also observed in leukocytes isolated from patients with RA, in addition to reduced disease severity (138). These exciting studies are among the first to demonstrate significant reductions in the clinical severity of RA.

Despite these positive outcomes, the invasive nature of electrode implantation has also led to study of noninvasive techniques for VNS. Preclinical studies transcutaneous vagal nerve stimulation (tcVNS), where the electrode is placed externally on the neck during stimulation, indicated efficacy in the control of inflammation (118). In the clinic, peripheral blood mononuclear cells from healthy volunteers receiving tcVNS produced significantly less proinflammatory cytokines compared with cells from the nonstimulated control group (149). Not all of these “less invasive” VNS approaches are efficacious. In a randomized double-blinded study, transvenous vagal nerve stimulation was safe and well tolerated in healthy human volunteers. Activation of the neural stimulation catheter inserted into the jugular vein failed to protect subjects from increased cytokine production, body temperature, and heart rate elicited by intravenous injection of LPS (139). Moreover, VNS did not reduce cytokine production, or alter neutrophil phagocytosis in leukocytes challenged with LPS ex vivo. This randomized clinical trial used several methods to ensure correct electrode placement, and maximal stimulation without inducing pain or discomfort in subjects, resulting in a wide range of stimulation amplitudes. While the authors note that these are all within the range of voltages used for other preclinical and clinical studies, one cannot help but wonder if the variability of stimulation parameters contributed to the lack of efficacy. It is also conceivable that more than one neural pathway is being stimulated, depending on the electrode placement and stimulation parameters. In support of this contention, vagal nerve stimulation without prior surgical vagotomy or blocking of afferent signaling increases neural activity in the greater splanchnic nerve (199). This neural circuit has previously been shown to augment proinflammatory cytokine production (114). This randomized clinical trial is, however, one of the few to evaluate VNS in healthy human subjects.

VIII. COMPREHENSIVE MAPPING OF FUNCTIONAL CIRCUITS

Application of devices that stimulate neuroimmune reflex arcs to alleviate disease in patients are still in their infancy. At this point, there are few studies that have attempted to document the intrinsic variation in innervation of peripheral organs in outbred animals, let alone in humans afflicted with chronic inflammatory diseases. Of the few studies comparing the innervation between animal species, none has meticulously documented the extent of variation between individual subjects. In target organs such as lymph nodes, axons have been reported in close proximity to lymphocytes in mice, rats, cat (25), and rabbit (179). Unfortunately, these studies did not assess the intersubject variability of innervation or the origin of these neurons. These are important variables to account for in the successful design and implementation of therapies that seek to deliver signals to defined cell populations. Another critical variable that may impact clinical significance is the effect of inflammation on the innervation of the affected organ. Chronic inflammation can significantly alter innervation of effected organs, with immunopathologies such as RA and IBD associated with increased production of cytokines, including IL-17A (3, 175). Sympathetic neurons express IL-17A receptors and IL-17A can promote neurite outgrowth (54, 103). The consequences of this neurite outgrowth would be predicted to include increased sympathetic neural density. In support of this contention that inflammation may alter the innervation, increased sympathetic innervation has been reported in patients with chronic IBD (24, 143), although other studies have reported a decrease in sympathetic innervation (243). Studies from several preclinical colitis models report reduced sympathetic and cholinergic fibers in submucosal (160) and myenteric ganglia (210). Although increased TH immunoreactive fibers are visible in the mucosa, the sparse nature precluded quantitative analysis (210). Similar depletion of nerves, including sympathetic innervation, occurs in the synovium of joints in preclinical models (167) and patients with RA (264). Depletion or a complete absence of sympathetic axons in the spleen has also been observed in spleens of patients with septic shock (110). Together these findings support the need for studies that directly address how inflammation or other preexisting conditions influence the efficacy of neurostimulation devices. Another important consideration is whether inflammation affects the release of neurotransmitters without affecting neuroanatomy (123). These knowledge gaps highlight the continued need for pre-clinical and clinical studies to assess the effect of inflammation on the innervation of various organs. These changes in innervation could be critical factors in the success of a targeted electrical neural stimulation. Consider for instance that increased density could result in increased and unintended effects, while decreased innervation density could conceivably reduce the efficacy. Although the benefit of electroceutical devices is the ability to finely taper the therapy to the individual, increased or decreased neural signaling following inflammation would seem a basic question that should be addressed. Additionally, while stimulation of nerves comprised of multiple axons types, such as the vagus, are likely to yield the same result in different subjects, it is uncertain if next generation devices targeting specific fibers will encounter significant intersubject variability.

Better understanding of the functional and anatomical circuitry in humans will be challenging. Functional circuit tracing and an understanding of variance between subjects in and patients could be benefited with analysis of non-human primates. Technologies including optogenetics have been successfully used in the CNS of Rhesus macaques (242). These techniques could be readily adapted for the peripheral nervous system in an effort to map the neuroimmune connection.

IX. CONCLUSION

The immune and nervous systems are tightly integrated physiological systems that are finally being appreciated as such. While immunology has been regarded as separate or discrete from physiology, progress in this field has been transformed by transdisciplinary studies and research teams. This approach has already led to a resurgence in the field and the dawn of therapeutic approaches that may prove effective in the treatment of chronic inflammatory conditions. Development of technologies in neuroscience and seemingly unrelated fields will continue to drive exciting new opportunities for discovery in this field. The recent surge of findings in neuroimmune communication should not be equated to there being nothing left to discover. From the controversies over established reflex arcs, limited neuroanatomical data, continued discoveries of novel circuits, and the challenges in translating this research to humans, this is an exciting time to engage in studies of the communication between nervous and immune system.

GRANTS

This work was supported by National Institutes of Health Office of the Director SPARC Grant 1OT2OD023871–01 and Crohn's and Colitis Canada.

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the authors.

ACKNOWLEDGMENTS

Address for reprint requests and other correspondence: C. Reardon, Assistant Professor, Univ. of California, Davis, VM: Anatomy, Physiology, & Cell Biology, 1089 Veterinary Medicine Dr., VM3B, Rm. 2007, Davis, CA 95616 (e-mail: creardon@ucdavis.edu).

REFERENCES

  • 1.Abad C, Martinez C, Juarranz MG, Arranz A, Leceta J, Delgado M, Gomariz RP. Therapeutic effects of vasoactive intestinal peptide in the trinitrobenzene sulfonic acid mice model of Crohn’s disease. Gastroenterology 124: 961–971, 2003. doi: 10.1053/gast.2003.50141. [DOI] [PubMed] [Google Scholar]
  • 2.Abe C, Inoue T, Inglis MA, Viar KE, Huang L, Ye H, Rosin DL, Stornetta RL, Okusa MD, Guyenet PG. C1 neurons mediate a stress-induced anti-inflammatory reflex in mice. Nat Neurosci 20: 700–707, 2017. doi: 10.1038/nn.4526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Abraham C, Cho JH. Inflammatory bowel disease. N Engl J Med 361: 2066–2078, 2009. doi: 10.1056/NEJMra0804647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Albuquerque EX, Pereira EFR, Alkondon M, Rogers SW. Mammalian nicotinic acetylcholine receptors: from structure to function. Physiol Rev 89: 73–120, 2009. doi: 10.1152/physrev.00015.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Andresen MC, Yang MY. Non-NMDA receptors mediate sensory afferent synaptic transmission in medial nucleus tractus solitarius. Am J Physiol Heart Circ Physiol 259: H1307–H1311, 1990. [DOI] [PubMed] [Google Scholar]
  • 6.Atweh SF, Grayhack JJ, Richman DP. A cholinergic receptor site on murine lymphocytes with novel binding characteristics. Life Sci 35: 2459–2469, 1984. doi: 10.1016/0024-3205(84)90455-7. [DOI] [PubMed] [Google Scholar]
  • 7.Bachmanov AA, Beauchamp GK. Taste receptor genes. Annu Rev Nutr 27: 389–414, 2007. doi: 10.1146/annurev.nutr.26.061505.111329. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Bachmanov AA, Bosak NP, Lin C, Matsumoto I, Ohmoto M, Reed DR, Nelson TM. Genetics of taste receptors. Curr Pharm Des 20: 2669–2683, 2014. doi: 10.2174/13816128113199990566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Báez-Pagán CA, Delgado-Vélez M, Lasalde-Dominicci JA. Activation of the Macrophage α7 Nicotinic Acetylcholine Receptor and Control of Inflammation. J Neuroimmune Pharmacol 10: 468–476, 2015. doi: 10.1007/s11481-015-9601-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Bain CC, Bravo-Blas A, Scott CL, Perdiguero EG, Geissmann F, Henri S, Malissen B, Osborne LC, Artis D, Mowat AM. Constant replenishment from circulating monocytes maintains the macrophage pool in the intestine of adult mice. Nat Immunol 15: 929–937, 2014. doi: 10.1038/ni.2967. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Ballabh P, Braun A, Nedergaard M. The blood-brain barrier: an overview: structure, regulation, and clinical implications. Neurobiol Dis 16: 1–13, 2004. doi: 10.1016/j.nbd.2003.12.016. [DOI] [PubMed] [Google Scholar]
  • 12.Banks WA. The blood-brain barrier in neuroimmunology: tales of separation and assimilation. Brain Behav Immun 44: 1–8, 2015. doi: 10.1016/j.bbi.2014.08.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Bany U, Gajewski M, Ksiezopolska-Pietrzak K, Jóźwicka M, Klimczak E, Ryzewski J, Chwalińska-Sadowska H, Maśliński W. Expression of mRNA encoding muscarinic receptor subtypes in neutrophils of patients with rheumatoid arthritis. Ann N Y Acad Sci 876: 301–304, 1999. doi: 10.1111/j.1749-6632.1999.tb07653.x. [DOI] [PubMed] [Google Scholar]
  • 14.Bao L, Locovei S, Dahl G. Pannexin membrane channels are mechanosensitive conduits for ATP. FEBS Lett 572: 65–68, 2004. doi: 10.1016/j.febslet.2004.07.009. [DOI] [PubMed] [Google Scholar]
  • 15.Bellinger DL, Lorton D, Hamill RW, Felten SY, Felten DL. Acetylcholinesterase staining and choline acetyltransferase activity in the young adult rat spleen: lack of evidence for cholinergic innervation. Brain Behav Immun 7: 191–204, 1993. doi: 10.1006/brbi.1993.1021. [DOI] [PubMed] [Google Scholar]
  • 16.Bergquist J, Tarkowski A, Ekman R, Ewing A. Discovery of endogenous catecholamines in lymphocytes and evidence for catecholamine regulation of lymphocyte function via an autocrine loop. Proc Natl Acad Sci USA 91: 12912–12916, 1994. doi: 10.1073/pnas.91.26.12912. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Berndt BE, Zhang M, Chen G-H, Huffnagle GB, Kao JY. The role of dendritic cells in the development of acute dextran sulfate sodium colitis. J Immunol 179: 6255–6262, 2007. doi: 10.4049/jimmunol.179.9.6255. [DOI] [PubMed] [Google Scholar]
  • 18.Berthoud HR, Powley TL. Characterization of vagal innervation to the rat celiac, suprarenal and mesenteric ganglia. J Auton Nerv Syst 42: 153–169, 1993. doi: 10.1016/0165-1838(93)90046-W. [DOI] [PubMed] [Google Scholar]
  • 19.Bezençon C, le Coutre J, Damak S. Taste-signaling proteins are coexpressed in solitary intestinal epithelial cells. Chem Senses 32: 41–49, 2007. doi: 10.1093/chemse/bjl034. [DOI] [PubMed] [Google Scholar]
  • 20.Bhat R, Axtell R, Mitra A, Miranda M, Lock C, Tsien RW, Steinman L. Inhibitory role for GABA in autoimmune inflammation. Proc Natl Acad Sci USA 107: 2580–2585, 2010. doi: 10.1073/pnas.0915139107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Billiet T, Rutgeerts P, Ferrante M, Van Assche G, Vermeire S. Targeting TNF-α for the treatment of inflammatory bowel disease. Expert Opin Biol Ther 14: 75–101, 2014. doi: 10.1517/14712598.2014.858695. [DOI] [PubMed] [Google Scholar]
  • 22.Billioud V, Ford AC, Tedesco ED, Colombel J-F, Roblin X, Peyrin-Biroulet L. Preoperative use of anti-TNF therapy and postoperative complications in inflammatory bowel diseases: a meta-analysis. J Crohn’s Colitis 7: 853–867, 2013. doi: 10.1016/j.crohns.2013.01.014. [DOI] [PubMed] [Google Scholar]
  • 23.Binshtok AM, Wang H, Zimmermann K, Amaya F, Vardeh D, Shi L, Brenner GJ, Ji R-R, Bean BP, Woolf CJ, Samad TA. Nociceptors are interleukin-1β sensors. J Neurosci 28: 14062–14073, 2008. doi: 10.1523/JNEUROSCI.3795-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Birch D, Knight GE, Boulos PB, Burnstock G. Analysis of innervation of human mesenteric vessels in non-inflamed and inflamed bowel–a confocal and functional study. Neurogastroenterol Motil 20: 660–670, 2008. doi: 10.1111/j.1365-2982.2008.01082.x. [DOI] [PubMed] [Google Scholar]
  • 25.Blue J, Weiss L. Species variation in the structure and function of the marginal zone–an electron microscope study of cat spleen. Am J Anat 161: 169–187, 1981. doi: 10.1002/aja.1001610204. [DOI] [PubMed] [Google Scholar]
  • 26.Bohórquez DV, Samsa LA, Roholt A, Medicetty S, Chandra R, Liddle RA. An enteroendocrine cell-enteric glia connection revealed by 3D electron microscopy. PLoS One 9: e89881, 2014. doi: 10.1371/journal.pone.0089881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Bohórquez DV, Shahid RA, Erdmann A, Kreger AM, Wang Y, Calakos N, Wang F, Liddle RA. Neuroepithelial circuit formed by innervation of sensory enteroendocrine cells. J Clin Invest 125: 782–786, 2015. doi: 10.1172/JCI78361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Bonaz B, Sinniger V, Hoffmann D, Clarençon D, Mathieu N, Dantzer C, Vercueil L, Picq C, Trocmé C, Faure P, Cracowski JL, Pellissier S. Chronic vagus nerve stimulation in Crohn’s disease: a 6-month follow-up pilot study. Neurogastroenterol Motil 28: 948–953, 2016. doi: 10.1111/nmo.12792. [DOI] [PubMed] [Google Scholar]
  • 29.Bonventre JV, Yang L. Cellular pathophysiology of ischemic acute kidney injury. J Clin Invest 121: 4210–4221, 2011. doi: 10.1172/JCI45161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Borovikova LV, Ivanova S, Nardi D, Zhang M, Yang H, Ombrellino M, Tracey KJ. Role of vagus nerve signaling in CNI-1493-mediated suppression of acute inflammation. Auton Neurosci 85: 141–147, 2000. doi: 10.1016/S1566-0702(00)00233-2. [DOI] [PubMed] [Google Scholar]
  • 31.Borovikova LV, Ivanova S, Zhang M, Yang H, Botchkina GI, Watkins LR, Wang H, Abumrad N, Eaton JW, Tracey KJ. Vagus nerve stimulation attenuates the systemic inflammatory response to endotoxin. Nature 405: 458–462, 2000. doi: 10.1038/35013070. [DOI] [PubMed] [Google Scholar]
  • 32.Boyden ES, Zhang F, Bamberg E, Nagel G, Deisseroth K. Millisecond-timescale, genetically targeted optical control of neural activity. Nat Neurosci 8: 1263–1268, 2005. doi: 10.1038/nn1525. [DOI] [PubMed] [Google Scholar]
  • 33.Brambilla R, Ashbaugh JJ, Magliozzi R, Dellarole A, Karmally S, Szymkowski DE, Bethea JR. Inhibition of soluble tumour necrosis factor is therapeutic in experimental autoimmune encephalomyelitis and promotes axon preservation and remyelination. Brain 134: 2736–2754, 2011. doi: 10.1093/brain/awr199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Bratton BO, Martelli D, McKinley MJ, Trevaks D, Anderson CR, McAllen RM. Neural regulation of inflammation: no neural connection from the vagus to splenic sympathetic neurons. Exp Physiol 97: 1180–1185, 2012. doi: 10.1113/expphysiol.2011.061531. [DOI] [PubMed] [Google Scholar]
  • 35.Brenner D, Blaser H, Mak TW. Regulation of tumour necrosis factor signalling: live or let die. Nat Rev Immunol 15: 362–374, 2015. doi: 10.1038/nri3834. [DOI] [PubMed] [Google Scholar]
  • 36.Brinkmann V, Zychlinsky A. Neutrophil extracellular traps: is immunity the second function of chromatin? J Cell Biol 198: 773–783, 2012. doi: 10.1083/jcb.201203170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Brownlie RJ, Zamoyska R. T cell receptor signalling networks: branched, diversified and bounded. Nat Rev Immunol 13: 257–269, 2013. doi: 10.1038/nri3403. [DOI] [PubMed] [Google Scholar]
  • 38.Bühling F, Lieder N, Kühlmann UC, Waldburg N, Welte T. Tiotropium suppresses acetylcholine-induced release of chemotactic mediators in vitro. Respir Med 101: 2386–2394, 2007. doi: 10.1016/j.rmed.2007.06.009. [DOI] [PubMed] [Google Scholar]
  • 39.Buijs RM, van der Vliet J, Garidou M-L, Huitinga I, Escobar C. Spleen vagal denervation inhibits the production of antibodies to circulating antigens. PLoS One 3: e3152, 2008. doi: 10.1371/journal.pone.0003152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Burnstock G. Historical review: ATP as a neurotransmitter. Trends Pharmacol Sci 27: 166–176, 2006. doi: 10.1016/j.tips.2006.01.005. [DOI] [PubMed] [Google Scholar]
  • 41.Cadwell K, Patel KK, Maloney NS, Liu TC, Ng AC, Storer CE, Head RD, Xavier R, Stappenbeck TS, Virgin HW. Virus-plus-susceptibility gene interaction determines Crohn’s disease gene Atg16L1 phenotypes in intestine. Cell 141: 1135–1145, 2010. doi: 10.1016/j.cell.2010.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Cailotto C, Costes LM, van der Vliet J, van Bree SH, van Heerikhuize JJ, Buijs RM, Boeckxstaens GE. Neuroanatomical evidence demonstrating the existence of the vagal anti-inflammatory reflex in the intestine. Neurogastroenterol Motil 24: 191–e93, 2012. doi: 10.1111/j.1365-2982.2011.01824.x. [DOI] [PubMed] [Google Scholar]
  • 43.Cano G, Card JP, Rinaman L, Sved AF. Connections of Barrington’s nucleus to the sympathetic nervous system in rats. J Auton Nerv Syst 79: 117–128, 2000. doi: 10.1016/S0165-1838(99)00101-0. [DOI] [PubMed] [Google Scholar]
  • 44.Cano G, Sved AF, Rinaman L, Rabin BS, Card JP. Characterization of the central nervous system innervation of the rat spleen using viral transneuronal tracing. J Comp Neurol 439: 1–18, 2001. doi: 10.1002/cne.1331. [DOI] [PubMed] [Google Scholar]
  • 45.Carmona-Rivera C, Purmalek MM, Moore E, Waldman M, Walter PJ, Garraffo HM, Phillips KA, Preston KL, Graf J, Kaplan MJ, Grayson PC. A role for muscarinic receptors in neutrophil extracellular trap formation and levamisole-induced autoimmunity. JCI Insight 2: e89780, 2017. doi: 10.1172/jci.insight.89780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Castex N, Fioramonti J, Ducos de Lahitte J, Luffau G, More J, Bueno L. Brain Fos expression and intestinal motor alterations during nematode-induced inflammation in the rat. Am J Physiol Gastrointest Liver Physiol 274: G210–G216, 1998. [DOI] [PubMed] [Google Scholar]
  • 47.Cayrol C, Girard J-P. IL-33: an alarmin cytokine with crucial roles in innate immunity, inflammation and allergy. Curr Opin Immunol 31: 31–37, 2014. doi: 10.1016/j.coi.2014.09.004. [DOI] [PubMed] [Google Scholar]
  • 48.Cervi AL, Moynes DM, Chisholm SP, Nasser Y, Vanner SJ, Lomax AE. A role for interleukin 17A in IBD-related neuroplasticity. Neurogastroenterol Motil 29: e13112, 2017. doi: 10.1111/nmo.13112. [DOI] [PubMed] [Google Scholar]
  • 49.Chatterjee PK, Al-Abed Y, Sherry B, Metz CN. Cholinergic agonists regulate JAK2/STAT3 signaling to suppress endothelial cell activation. Am J Physiol Cell Physiol 297: C1294–C1306, 2009. doi: 10.1152/ajpcell.00160.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Chekeni FB, Elliott MR, Sandilos JK, Walk SF, Kinchen JM, Lazarowski ER, Armstrong AJ, Penuela S, Laird DW, Salvesen GS, Isakson BE, Bayliss DA, Ravichandran KS. Pannexin 1 channels mediate ‘find-me’ signal release and membrane permeability during apoptosis. Nature 467: 863–867, 2010. doi: 10.1038/nature09413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Chen Y, Corriden R, Inoue Y, Yip L, Hashiguchi N, Zinkernagel A, Nizet V, Insel PA, Junger WG. ATP release guides neutrophil chemotaxis via P2Y2 and A3 receptors. Science 314: 1792–1795, 2006. doi: 10.1126/science.1132559. [DOI] [PubMed] [Google Scholar]
  • 52.Cheng Q, Yakel JL. Activation of α7 nicotinic acetylcholine receptors increases intracellular cAMP levels via activation of AC1 in hippocampal neurons. Neuropharmacology 95: 405–414, 2015. doi: 10.1016/j.neuropharm.2015.04.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Cherubini E, Tabbì L, Scozzi D, Mariotta S, Galli E, Carello R, Avitabile S, Tayebati SK, Amenta F, De Vitis C, Mancini R, Ricci A. Modified expression of peripheral blood lymphocyte muscarinic cholinergic receptors in asthmatic children. J Neuroimmunol 284: 37–43, 2015. doi: 10.1016/j.jneuroim.2015.04.016. [DOI] [PubMed] [Google Scholar]
  • 54.Chisholm SP, Cervi AL, Nagpal S, Lomax AE. Interleukin-17A increases neurite outgrowth from adult postganglionic sympathetic neurons. J Neurosci 32: 1146–1155, 2012. doi: 10.1523/JNEUROSCI.5343-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Chiu IM, Heesters BA, Ghasemlou N, Von Hehn CA, Zhao F, Tran J, Wainger B, Strominger A, Muralidharan S, Horswill AR, Bubeck Wardenburg J, Hwang SW, Carroll MC, Woolf CJ. Bacteria activate sensory neurons that modulate pain and inflammation. Nature 501: 52–57, 2013. doi: 10.1038/nature12479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Cirillo C, Sarnelli G, Esposito G, Grosso M, Petruzzelli R, Izzo P, Calì G, D’Armiento FP, Rocco A, Nardone G, Iuvone T, Steardo L, Cuomo R. Increased mucosal nitric oxide production in ulcerative colitis is mediated in part by the enteroglial-derived S100B protein. Neurogastroenterol Motil 21: 1209–e112, 2009. doi: 10.1111/j.1365-2982.2009.01346.x. [DOI] [PubMed] [Google Scholar]
  • 57.Costa P, Auger CB, Traver DJ, Costa LG. Identification of m3, m4 and m5 subtypes of muscarinic receptor mRNA in human blood mononuclear cells. J Neuroimmunol 60: 45–51, 1995. doi: 10.1016/0165-5728(95)00051-3. [DOI] [PubMed] [Google Scholar]
  • 58.Coupland RE, Parker TL, Kesse WK, Mohamed AA. The innervation of the adrenal gland. III. Vagal innervation. J Anat 163: 173–181, 1989. [PMC free article] [PubMed] [Google Scholar]
  • 59.Czaikoski PG, Mota JMSC, Nascimento DC, Sônego F, Castanheira FVES, Melo PH, Scortegagna GT, Silva RL, Barroso-Sousa R, Souto FO, Pazin-Filho A, Figueiredo F, Alves-Filho JC, Cunha FQ. Neutrophil Extracellular Traps Induce Organ Damage during Experimental and Clinical Sepsis. PLoS One 11: e0148142, 2016. doi: 10.1371/journal.pone.0148142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Da Silva LG, Dias ACR, Furlan E, Colombari E. Nitric oxide modulates the cardiovascular effects elicited by acetylcholine in the NTS of awake rats. Am J Physiol Regul Integr Comp Physiol 295: R1774–R1781, 2008. doi: 10.1152/ajpregu.00559.2007. [DOI] [PubMed] [Google Scholar]
  • 61.Dale HH, Dudley HW. The presence of histamine and acetylcholine in the spleen of the ox and the horse. J Physiol 68: 97–123, 1929. doi: 10.1113/jphysiol.1929.sp002598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Dantzer R. Cytokine, sickness behavior, and depression. Immunol Allergy Clin North Am 29: 247–264, 2009. doi: 10.1016/j.iac.2009.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Darby M, Schnoeller C, Vira A, Culley FJ, Bobat S, Logan E, Kirstein F, Wess J, Cunningham AF, Brombacher F, Selkirk ME, Horsnell WGC. The M3 muscarinic receptor is required for optimal adaptive immunity to helminth and bacterial infection. PLoS Pathog 11: e1004636, 2015. [Correction in PLoS Pathog 11: e1004727, 2015.] 10.1371/journal.ppat.1004636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.De Jonge WJ, van der Zanden EP, The FO, Bijlsma MF, van Westerloo DJ, Bennink RJ, Berthoud HR, Uematsu S, Akira S, van den Wijngaard RM, Boeckxstaens GE. Stimulation of the vagus nerve attenuates macrophage activation by activating the Jak2-STAT3 signaling pathway. Nat Immunol 6: 844–851, 2005. [Erratum in Nat Immunol 6: 954, 2005.] doi: 10.1038/ni1229. [DOI] [PubMed] [Google Scholar]
  • 65.De Lartigue G, Barbier de la Serre C, Espero E, Lee J, Raybould HE. Diet-induced obesity leads to the development of leptin resistance in vagal afferent neurons. Am J Physiol Endocrinol Metab 301: E187–E195, 2011. doi: 10.1152/ajpendo.00056.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Delgado M, Ganea D. Vasoactive intestinal peptide and pituitary adenylate cyclase-activating polypeptide inhibit expression of Fas ligand in activated T lymphocytes by regulating c-Myc, NF-kappa B, NF-AT, and early growth factors 2/3. J Immunol 166: 1028–1040, 2001. doi: 10.4049/jimmunol.166.2.1028. [DOI] [PubMed] [Google Scholar]
  • 67.Delgado M, Pozo D, Ganea D. The significance of vasoactive intestinal peptide in immunomodulation. Pharmacol Rev 56: 249–290, 2004. doi: 10.1124/pr.56.2.7. [DOI] [PubMed] [Google Scholar]
  • 68.Deuchars SA, Lall K. Sympathetic Preganglionic Neurons. V. Properties and Inputs. In: Comprehensive Physiology. New York: Wiley, 2011. [DOI] [PubMed] [Google Scholar]
  • 69.Dhawan S, De Palma G, Willemze RA, Hilbers FW, Verseijden C, Luyer MD, Nuding S, Wehkamp J, Souwer Y, de Jong EC, Seppen J, van den Wijngaard RM, Wehner S, Verdu E, Bercik P, de Jonge WJ. Acetylcholine-producing T cells in the intestine regulate antimicrobial peptide expression and microbial diversity. Am J Physiol Gastrointest Liver Physiol 311: G920–G933, 2016. doi: 10.1152/ajpgi.00114.2016. [DOI] [PubMed] [Google Scholar]
  • 70.Di Giovangiulio M, Bosmans G, Meroni E, Stakenborg N, Florens M, Farro G, Gomez-Pinilla PJ, Matteoli G, Boeckxstaens G. Vagotomy affects the development of oral tolerance and increases susceptibility to develop colitis independently of the alpha-7 nicotinic receptor. Mol Med 22: 464–476, 2016. doi: 10.2119/molmed.2016.00062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Dionisio L, De Rosa MJ, Bouzat C, Esandi MC. An intrinsic GABAergic system in human lymphocytes. Neuropharmacology 60: 513–519, 2011. doi: 10.1016/j.neuropharm.2010.11.007. [DOI] [PubMed] [Google Scholar]
  • 72.Dopp JM, Mackenzie-Graham A, Otero GC, Merrill JE. Differential expression, cytokine modulation, and specific functions of type-1 and type-2 tumor necrosis factor receptors in rat glia. J Neuroimmunol 75: 104–112, 1997. doi: 10.1016/S0165-5728(97)00009-X. [DOI] [PubMed] [Google Scholar]
  • 73.Dubyak GR. Both sides now: multiple interactions of ATP with pannexin-1 hemichannels. Focus on “A permeant regulating its permeation pore: inhibition of pannexin 1 channels by ATP”. Am J Physiol Cell Physiol 296: C235–C241, 2009. doi: 10.1152/ajpcell.00639.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Ek M, Kurosawa M, Lundeberg T, Ericsson A. Activation of vagal afferents after intravenous injection of interleukin-1β: role of endogenous prostaglandins. J Neurosci 18: 9471–9479, 1998. doi: 10.1523/JNEUROSCI.18-22-09471.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Emch GS, Hermann GE, Rogers RC. TNF-α activates solitary nucleus neurons responsive to gastric distension. Am J Physiol Gastrointest Liver Physiol 279: G582–G586, 2000. doi: 10.1152/ajpgi.2000.279.3.G582. [DOI] [PubMed] [Google Scholar]
  • 76.Esposito G, Capoccia E, Turco F, Palumbo I, Lu J, Steardo A, Cuomo R, Sarnelli G, Steardo L. Palmitoylethanolamide improves colon inflammation through an enteric glia/toll like receptor 4-dependent PPAR-α activation. Gut 63: 1300–1312, 2014. doi: 10.1136/gutjnl-2013-305005. [DOI] [PubMed] [Google Scholar]
  • 77.Eva C, Ferrero P, Rocca P, Funaro A, Bergamasco B, Ravizza L, Genazzani E. [3H]N-methylscopolamine binding to muscarinic receptors in human peripheral blood lymphocytes: characterization, localization on T-lymphocyte subsets and age-dependent changes. Neuropharmacology 28: 719–726, 1989. doi: 10.1016/0028-3908(89)90157-3. [DOI] [PubMed] [Google Scholar]
  • 78.Famm K, Litt B, Tracey KJ, Boyden ES, Slaoui M. Drug discovery: a jump-start for electroceuticals. Nature 496: 159–161, 2013. [Correction in Nature 496: 300, 2013.] 10.1038/496159a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Feldman-Goriachnik R, Belzer V, Hanani M. Systemic inflammation activates satellite glial cells in the mouse nodose ganglion and alters their functions. Glia 63: 2121–2132, 2015. doi: 10.1002/glia.22881. [DOI] [PubMed] [Google Scholar]
  • 80.Felten DL, Felten SY. Sympathetic noradrenergic innervation of immune organs. Brain Behav Immun 2: 293–300, 1988. doi: 10.1016/0889-1591(88)90031-1. [DOI] [PubMed] [Google Scholar]
  • 81.Felten DL, Felten SY, Bellinger DL, Carlson SL, Ackerman KD, Madden KS, Olschowki JA, Livnat S. Noradrenergic sympathetic neural interactions with the immune system: structure and function. Immunol Rev 100: 225–260, 1987. doi: 10.1111/j.1600-065X.1987.tb00534.x. [DOI] [PubMed] [Google Scholar]
  • 82.Felten SY, Olschowka J. Noradrenergic sympathetic innervation of the spleen. II. Tyrosine hydroxylase (TH)-positive nerve terminals form synapticlike contacts on lymphocytes in the splenic white pulp. J Neurosci Res 18: 37–48, 1987. doi: 10.1002/jnr.490180108. [DOI] [PubMed] [Google Scholar]
  • 83.Ferrari D, Chiozzi P, Falzoni S, Dal Susino M, Melchiorri L, Baricordi OR, Di Virgilio F. Extracellular ATP triggers IL-1 beta release by activating the purinergic P2Z receptor of human macrophages. J Immunol 159: 1451–1458, 1997. [PubMed] [Google Scholar]
  • 84.Flierl MA, Rittirsch D, Nadeau BA, Chen AJ, Sarma JV, Zetoune FS, McGuire SR, List RP, Day DE, Hoesel LM, Gao H, Van Rooijen N, Huber-Lang MS, Neubig RR, Ward PA. Phagocyte-derived catecholamines enhance acute inflammatory injury. Nature 449: 721–725, 2007. doi: 10.1038/nature06185. [DOI] [PubMed] [Google Scholar]
  • 85.Ford AC, Lacy BE, Talley NJ. Irritable Bowel Syndrome. N Engl J Med 376: 2566–2578, 2017. doi: 10.1056/NEJMra1607547. [DOI] [PubMed] [Google Scholar]
  • 86.Frangogiannis NG. The inflammatory response in myocardial injury, repair, and remodelling. Nat Rev Cardiol 11: 255–265, 2014. doi: 10.1038/nrcardio.2014.28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Fu Y-Y, Peng S-J, Lin H-Y, Pasricha PJ, Tang S-C. 3-D imaging and illustration of mouse intestinal neurovascular complex. Am J Physiol Gastrointest Liver Physiol 304: G1–G11, 2013. doi: 10.1152/ajpgi.00209.2012. [DOI] [PubMed] [Google Scholar]
  • 88.Fujii YX, Fujigaya H, Moriwaki Y, Misawa H, Kasahara T, Grando SA, Kawashima K. Enhanced serum antigen-specific IgG1 and proinflammatory cytokine production in nicotinic acetylcholine receptor α7 subunit gene knockout mice. J Neuroimmunol 189: 69–74, 2007. doi: 10.1016/j.jneuroim.2007.07.003. [DOI] [PubMed] [Google Scholar]
  • 89.Gabanyi I, Muller PA, Feighery L, Oliveira TY, Costa-Pinto FA, Mucida D. Neuro-immune Interactions Drive Tissue Programming in Intestinal Macrophages. Cell 164: 378–391, 2016. doi: 10.1016/j.cell.2015.12.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Ganea D, Hooper KM, Kong W. The neuropeptide vasoactive intestinal peptide: direct effects on immune cells and involvement in inflammatory and autoimmune diseases. Acta Physiol (Oxf) 213: 442–452, 2015. doi: 10.1111/apha.12427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Gautron L, Rutkowski JM, Burton MD, Wei W, Wan Y, Elmquist JK. Neuronal and nonneuronal cholinergic structures in the mouse gastrointestinal tract and spleen. J Comp Neurol 521: 3741–3767, 2013. doi: 10.1002/cne.23376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Gaykema RP, Goehler LE, Tilders FJ, Bol JG, McGorry M, Fleshner M, Maier SF, Watkins LR. Bacterial endotoxin induces fos immunoreactivity in primary afferent neurons of the vagus nerve. Neuroimmunomodulation 5: 234–240, 1998. doi: 10.1159/000026343. [DOI] [PubMed] [Google Scholar]
  • 93.Ghia J-E, Blennerhassett P, Collins SM. Vagus nerve integrity and experimental colitis. Am J Physiol Gastrointest Liver Physiol 293: G560–G567, 2007. doi: 10.1152/ajpgi.00098.2007. [DOI] [PubMed] [Google Scholar]
  • 95.Ghia J-E, Blennerhassett P, El-Sharkawy RT, Collins SM. The protective effect of the vagus nerve in a murine model of chronic relapsing colitis. Am J Physiol Gastrointest Liver Physiol 293: G711–G718, 2007. doi: 10.1152/ajpgi.00240.2007. [DOI] [PubMed] [Google Scholar]
  • 96.Ghia JE, Blennerhassett P, Deng Y, Verdu EF, Khan WI, Collins SM. Reactivation of inflammatory bowel disease in a mouse model of depression. Gastroenterology 136: 2280–2288.e4, 2009. doi: 10.1053/j.gastro.2009.02.069. [DOI] [PubMed] [Google Scholar]
  • 97.Gigliotti JC, Huang L, Bajwa A, Ye H, Mace EH, Hossack JA, Kalantari K, Inoue T, Rosin DL, Okusa MD. Ultrasound Modulates the Splenic Neuroimmune Axis in Attenuating AKI. J Am Soc Nephrol 26: 2470–2481, 2015. doi: 10.1681/ASN.2014080769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Gladkevich A, Korf J, Hakobyan VP, Melkonyan KV. The peripheral GABAergic system as a target in endocrine disorders. Auton Neurosci 124: 1–8, 2006. doi: 10.1016/j.autneu.2005.11.002. [DOI] [PubMed] [Google Scholar]
  • 99.Goehler LE, Gaykema RPA, Hammack SE, Maier SF, Watkins LR. Interleukin-1 induces c-Fos immunoreactivity in primary afferent neurons of the vagus nerve. Brain Res 804: 306–310, 1998. doi: 10.1016/S0006-8993(98)00685-4. [DOI] [PubMed] [Google Scholar]
  • 100.Gross PM. Circumventricular organ capillaries. Prog Brain Res 91: 219–233, 1992. doi: 10.1016/S0079-6123(08)62338-9. [DOI] [PubMed] [Google Scholar]
  • 101.Guyenet PG, Stornetta RL, Bochorishvili G, Depuy SD, Burke PG, Abbott SB. C1 neurons: the body’s EMTs. Am J Physiol Regul Integr Comp Physiol 305: R187–R204, 2013. doi: 10.1152/ajpregu.00054.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Gwilt CR, Donnelly LE, Rogers DF. The non-neuronal cholinergic system in the airways: an unappreciated regulatory role in pulmonary inflammation? Pharmacol Ther 115: 208–222, 2007. doi: 10.1016/j.pharmthera.2007.05.007. [DOI] [PubMed] [Google Scholar]
  • 103.Habash T, Saleh A, Roy Chowdhury SK, Smith DR, Fernyhough P. The proinflammatory cytokine, interleukin-17A, augments mitochondrial function and neurite outgrowth of cultured adult sensory neurons derived from normal and diabetic rats. Exp Neurol 273: 177–189, 2015. doi: 10.1016/j.expneurol.2015.08.016. [DOI] [PubMed] [Google Scholar]
  • 104.Hardman CS, Panova V, McKenzie AN. IL-33 citrine reporter mice reveal the temporal and spatial expression of IL-33 during allergic lung inflammation. Eur J Immunol 43: 488–498, 2013. doi: 10.1002/eji.201242863. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Helke CJ, Handelmann GE, Jacobowitz DM. Choline acetyltransferase activity in the nucleus tractus solitarius: regulation by the afferent vagus nerve. Brain Res Bull 10: 433–436, 1983. doi: 10.1016/0361-9230(83)90139-9. [DOI] [PubMed] [Google Scholar]
  • 106.Heng TS, Painter MW, Elpek K, Lukacs-Kornek V, Mauermann N, Turley SJ, Koller D, Kim FS, Wagers AJ, Asinovski N, Davis S, Fassett M, Feuerer M, Gray DHD, Haxhinasto S, Hill JA, Hyatt G, Laplace C, Leatherbee K, Mathis D, Benoist C, Jianu R, Laidlaw DH, Best JA, Knell J, Goldrath AW, Jarjoura J, Sun JC, Zhu Y, Lanier LL, Ergun A, Li Z, Collins JJ, Shinton SA, Hardy RR, Friedline R, Sylvia K, Kang J; Immunological Genome Project Consortium . The Immunological Genome Project: networks of gene expression in immune cells. Nat Immunol 9: 1091–1094, 2008. doi: 10.1038/ni1008-1091. [DOI] [PubMed] [Google Scholar]
  • 107.Hermann GE, Hebert SL, Van Meter MJ, Holmes GM, Rogers RC. TNF alpha-p55 receptors: medullary brainstem immunocytochemical localization in normal and vagus nerve-transected rats. Brain Res 1004: 156–166, 2004. doi: 10.1016/j.brainres.2003.11.078. [DOI] [PubMed] [Google Scholar]
  • 108.Hermann GE, Hebert SL, Van Meter MJ, Holmes GM, Rogers RC. TNF α-p55 receptors: medullary brainstem immunocytochemical localization in normal and vagus nerve-transected rats. Brain Res 1004: 156–166, 2004. doi: 10.1016/j.brainres.2003.11.078. [DOI] [PubMed] [Google Scholar]
  • 109.Heusermann U, Stutte HJ. Electron microscopic studies of the innervation of the human spleen. Cell Tissue Res 184: 225–236, 1977. doi: 10.1007/BF00223070. [DOI] [PubMed] [Google Scholar]
  • 110.Hoover DB, Brown TC, Miller MK, Schweitzer JB, Williams DL. Loss of Sympathetic Nerves in Spleens from Patients with End Stage Sepsis. Front Immunol 8: 1712, 2017. doi: 10.3389/fimmu.2017.01712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Hoover DB, Hancock JC, DePorter TE. Effect of vagotomy on cholinergic parameters in nuclei of rat medulla oblongata. Brain Res Bull 15: 5–11, 1985. doi: 10.1016/0361-9230(85)90054-1. [DOI] [PubMed] [Google Scholar]
  • 112.Hoshino K, Takeuchi O, Kawai T, Sanjo H, Ogawa T, Takeda Y, Takeda K, Akira S. Cutting edge: Toll-like receptor 4 (TLR4)-deficient mice are hyporesponsive to lipopolysaccharide: evidence for TLR4 as the Lps gene product. J Immunol 162: 3749–3752, 1999. [PubMed] [Google Scholar]
  • 113.Hosoi T, Okuma Y, Matsuda T, Nomura Y. Novel pathway for LPS-induced afferent vagus nerve activation: possible role of nodose ganglion. Auton Neurosci 120: 104–107, 2005. doi: 10.1016/j.autneu.2004.11.012. [DOI] [PubMed] [Google Scholar]
  • 114.Hosoi T, Okuma Y, Nomura Y. Electrical stimulation of afferent vagus nerve induces IL-1beta expression in the brain and activates HPA axis. Am J Physiol Regul Integr Comp Physiol 279: R141–R147, 2000. doi: 10.1152/ajpregu.2000.279.1.R141. [DOI] [PubMed] [Google Scholar]
  • 115.Hu Y, Gu Q, Lin R-L, Kryscio R, Lee L-Y. Calcium transient evoked by TRPV1 activators is enhanced by tumor necrosis factor-α in rat pulmonary sensory neurons. Am J Physiol Lung Cell Mol Physiol 299: L483–L492, 2010. doi: 10.1152/ajplung.00111.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Huang TY, Belzer V, Hanani M. Gap junctions in dorsal root ganglia: possible contribution to visceral pain. Eur J Pain 14: 49.e1–49.e9, 2010. doi: 10.1016/j.ejpain.2009.02.005. [DOI] [PubMed] [Google Scholar]
  • 117.Hussell T, Bell TJ. Alveolar macrophages: plasticity in a tissue-specific context. Nat Rev Immunol 14: 81–93, 2014. doi: 10.1038/nri3600. [DOI] [PubMed] [Google Scholar]
  • 118.Huston JM, Gallowitsch-Puerta M, Ochani M, Ochani K, Yuan R, Rosas-Ballina M, Ashok M, Goldstein RS, Chavan S, Pavlov VA, Metz CN, Yang H, Czura CJ, Wang H, Tracey KJ. Transcutaneous vagus nerve stimulation reduces serum high mobility group box 1 levels and improves survival in murine sepsis. Crit Care Med 35: 2762–2768, 2007. [DOI] [PubMed] [Google Scholar]
  • 119.Huston JM, Ochani M, Rosas-Ballina M, Liao H, Ochani K, Pavlov VA, Gallowitsch-Puerta M, Ashok M, Czura CJ, Foxwell B, Tracey KJ, Ulloa L. Splenectomy inactivates the cholinergic anti-inflammatory pathway during lethal endotoxemia and polymicrobial sepsis. J Exp Med 203: 1623–1628, 2006. doi: 10.1084/jem.20052362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Huston JM, Rosas-Ballina M, Xue X, Dowling O, Ochani K, Ochani M, Yeboah MM, Chatterjee PK, Tracey KJ, Metz CN. Cholinergic neural signals to the spleen down-regulate leukocyte trafficking via CD11b. J Immunol 183: 552–559, 2009. doi: 10.4049/jimmunol.0802684. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Iho S, Tanaka Y, Takauji R, Kobayashi C, Muramatsu I, Iwasaki H, Nakamura K, Sasaki Y, Nakao K, Takahashi T. Nicotine induces human neutrophils to produce IL-8 through the generation of peroxynitrite and subsequent activation of NF-kappaB. J Leukoc Biol 74: 942–951, 2003. doi: 10.1189/jlb.1202626. [DOI] [PubMed] [Google Scholar]
  • 122.Inoue T, Abe C, Sung SS, Moscalu S, Jankowski J, Huang L, Ye H, Rosin DL, Guyenet PG, Okusa MD. Vagus nerve stimulation mediates protection from kidney ischemia-reperfusion injury through α7nAChR+ splenocytes. J Clin Invest 126: 1939–1952, 2016. doi: 10.1172/JCI83658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Jacobson K, McHugh K, Collins SM. Experimental colitis alters myenteric nerve function at inflamed and noninflamed sites in the rat. Gastroenterology 109: 718–722, 1995. doi: 10.1016/0016-5085(95)90378-X. [DOI] [PubMed] [Google Scholar]
  • 124.Jang HR, Rabb H. The innate immune response in ischemic acute kidney injury. Clin Immunol 130: 41–50, 2009. doi: 10.1016/j.clim.2008.08.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Jay M, Kojima S, Gillespie MN. Nicotine potentiates superoxide anion generation by human neutrophils. Toxicol Appl Pharmacol 86: 484–487, 1986. doi: 10.1016/0041-008X(86)90376-5. [DOI] [PubMed] [Google Scholar]
  • 126.Ji H, Rabbi MF, Labis B, Pavlov VA, Tracey KJ, Ghia JE. Central cholinergic activation of a vagus nerve-to-spleen circuit alleviates experimental colitis. Mucosal Immunol 7: 335–347, 2014. doi: 10.1038/mi.2013.52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Kalyanaraman B, Darley-Usmar V, Davies KJA, Dennery PA, Forman HJ, Grisham MB, Mann GE, Moore K, Roberts LJ II, Ischiropoulos H. Measuring reactive oxygen and nitrogen species with fluorescent probes: challenges and limitations. Free Radic Biol Med 52: 1–6, 2012. doi: 10.1016/j.freeradbiomed.2011.09.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Kaplan GG. The global burden of IBD: from 2015 to 2025. Nat Rev Gastroenterol Hepatol 12: 720–727, 2015. doi: 10.1038/nrgastro.2015.150. [DOI] [PubMed] [Google Scholar]
  • 129.Katakura K, Lee J, Rachmilewitz D, Li G, Eckmann L, Raz E. Toll-like receptor 9-induced type I IFN protects mice from experimental colitis. J Clin Invest 115: 695–702, 2005. doi: 10.1172/JCI22996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Kawashima K, Fujii T, Moriwaki Y, Misawa H. Critical roles of acetylcholine and the muscarinic and nicotinic acetylcholine receptors in the regulation of immune function. Life Sci 91: 1027–1032, 2012. doi: 10.1016/j.lfs.2012.05.006. [DOI] [PubMed] [Google Scholar]
  • 131.Kawashima K, Fujii T, Moriwaki Y, Misawa H, Horiguchi K. Reconciling neuronally and nonneuronally derived acetylcholine in the regulation of immune function. Ann N Y Acad Sci 1261: 7–17, 2012. doi: 10.1111/j.1749-6632.2012.06516.x. [DOI] [PubMed] [Google Scholar]
  • 132.Kawashima K, Yoshikawa K, Fujii YX, Moriwaki Y, Misawa H. Expression and function of genes encoding cholinergic components in murine immune cells. Life Sci 80: 2314–2319, 2007. doi: 10.1016/j.lfs.2007.02.036. [DOI] [PubMed] [Google Scholar]
  • 133.Kawashima K, Yoshikawa K, Fujii YX, Moriwaki Y, Misawa H. Expression and function of genes encoding cholinergic components in murine immune cells. Life Sci 80: 2314–2319, 2007. doi: 10.1016/j.lfs.2007.02.036. [DOI] [PubMed] [Google Scholar]
  • 134.Khakh BS, North RA. P2X receptors as cell-surface ATP sensors in health and disease. Nature 442: 527–532, 2006. doi: 10.1038/nature04886. [DOI] [PubMed] [Google Scholar]
  • 135.Kiesler P, Fuss IJ, Strober W. Experimental Models of Inflammatory Bowel Diseases. Cell Mol Gastroenterol Hepatol 1: 154–170, 2015. doi: 10.1016/j.jcmgh.2015.01.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Kimura A, Naka T, Muta T, Takeuchi O, Akira S, Kawase I, Kishimoto T. Suppressor of cytokine signaling-1 selectively inhibits LPS-induced IL-6 production by regulating JAK-STAT. Proc Natl Acad Sci USA 102: 17089–17094, 2005. doi: 10.1073/pnas.0508517102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Koga K, Takaesu G, Yoshida R, Nakaya M, Kobayashi T, Kinjyo I, Yoshimura A. Cyclic adenosine monophosphate suppresses the transcription of proinflammatory cytokines via the phosphorylated c-Fos protein. Immunity 30: 372–383, 2009. doi: 10.1016/j.immuni.2008.12.021. [DOI] [PubMed] [Google Scholar]
  • 138.Koopman FA, Chavan SS, Miljko S, Grazio S, Sokolovic S, Schuurman PR, Mehta AD, Levine YA, Faltys M, Zitnik R, Tracey KJ, Tak PP. Vagus nerve stimulation inhibits cytokine production and attenuates disease severity in rheumatoid arthritis. Proc Natl Acad Sci USA 113: 8284–8289, 2016. doi: 10.1073/pnas.1605635113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Kox M, van Eijk LT, Verhaak T, Frenzel T, Kiers HD, Gerretsen J, van der Hoeven JG, Kornet L, Scheiner A, Pickkers P. Transvenous vagus nerve stimulation does not modulate the innate immune response during experimental human endotoxemia: a randomized controlled study. Arthritis Res Ther 17: 150, 2015. doi: 10.1186/s13075-015-0667-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Kudoh G, Hoshi K, Murakami T. Fluorescence microscopic and enzyme histochemical studies of the innervation of the human spleen. Arch Histol Jpn 42: 169–180, 1979. doi: 10.1679/aohc1950.42.169. [DOI] [PubMed] [Google Scholar]
  • 141.Kuno R, Wang J, Kawanokuchi J, Takeuchi H, Mizuno T, Suzumura A. Autocrine activation of microglia by tumor necrosis factor-α. J Neuroimmunol 162: 89–96, 2005. doi: 10.1016/j.jneuroim.2005.01.015. [DOI] [PubMed] [Google Scholar]
  • 142.Kwong K, Kollarik M, Nassenstein C, Ru F, Undem BJ. P2X2 receptors differentiate placodal vs. neural crest C-fiber phenotypes innervating guinea pig lungs and esophagus. Am J Physiol Lung Cell Mol Physiol 295: L858–L865, 2008. doi: 10.1152/ajplung.90360.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Kyösola K, Penttilä O, Salaspuro M. Rectal mucosal adrenergic innervation and enterochromaffin cells in ulcerative colitis and irritable colon. Scand J Gastroenterol 12: 363–367, 1977. doi: 10.3109/00365527709180942. [DOI] [PubMed] [Google Scholar]
  • 144.Latorre R, Sternini C, De Giorgio R, Greenwood-Van Meerveld B. Enteroendocrine cells: a review of their role in brain-gut communication. Neurogastroenterol Motil 28: 620–630, 2016. doi: 10.1111/nmo.12754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Lawrence AJ, Jarrott B. l-Glutamate as a neurotransmitter at baroreceptor afferents: evidence from in vivo microdialysis. Neuroscience 58: 585–591, 1994. doi: 10.1016/0306-4522(94)90083-3. [DOI] [PubMed] [Google Scholar]
  • 146.Layé S, Bluthé RM, Kent S, Combe C, Médina C, Parnet P, Kelley K, Dantzer R. Subdiaphragmatic vagotomy blocks induction of IL-1 beta mRNA in mice brain in response to peripheral LPS. Am J Physiol Regul Integr Comp Physiol 268: R1327–R1331, 1995. [DOI] [PubMed] [Google Scholar]
  • 147.Lee J, Luria A, Rhodes C, Raghu H, Lingampalli N, Sharpe O, Rada B, Sohn DH, Robinson WH, Sokolove J. Nicotine drives neutrophil extracellular traps formation and accelerates collagen-induced arthritis. Rheumatology (Oxford) 56: 644–653, 2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Lee RJ, Xiong G, Kofonow JM, Chen B, Lysenko A, Jiang P, Abraham V, Doghramji L, Adappa ND, Palmer JN, Kennedy DW, Beauchamp GK, Doulias P-T, Ischiropoulos H, Kreindler JL, Reed DR, Cohen NA. T2R38 taste receptor polymorphisms underlie susceptibility to upper respiratory infection. J Clin Invest 122: 4145–4159, 2012. doi: 10.1172/JCI64240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Lerman I, Hauger R, Sorkin L, Proudfoot J, Davis B, Huang A, Lam K, Simon B, Baker DG. Noninvasive Transcutaneous Vagus Nerve Stimulation Decreases Whole Blood Culture-Derived Cytokines and Chemokines: A Randomized, Blinded, Healthy Control Pilot Trial. Neuromodulation 19: 283–290, 2016. doi: 10.1111/ner.12398. [DOI] [PubMed] [Google Scholar]
  • 150.Lerner TN, Ye L, Deisseroth K. Communication in Neural Circuits: Tools, Opportunities, and Challenges. Cell 164: 1136–1150, 2016. doi: 10.1016/j.cell.2016.02.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Levite M. Dopamine and T cells: dopamine receptors and potent effects on T cells, dopamine production in T cells, and abnormalities in the dopaminergic system in T cells in autoimmune, neurological and psychiatric diseases. Acta Physiol (Oxf) 216: 42–89, 2016. doi: 10.1111/apha.12476. [DOI] [PubMed] [Google Scholar]
  • 152.Lewis C, Neidhart S, Holy C, North RA, Buell G, Surprenant A. Coexpression of P2X2 and P2X3 receptor subunits can account for ATP-gated currents in sensory neurons. Nature 377: 432–435, 1995. doi: 10.1038/377432a0. [DOI] [PubMed] [Google Scholar]
  • 153.Li X, Murray F, Koide N, Goldstone J, Dann SM, Chen J, Bertin S, Fu G, Weinstein LS, Chen M, Corr M, Eckmann L, Insel PA, Raz E. Divergent requirement for Gαs and cAMP in the differentiation and inflammatory profile of distinct mouse Th subsets. J Clin Invest 122: 963–973, 2012. doi: 10.1172/JCI59097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Liew FY, Girard J-P, Turnquist HR. Interleukin-33 in health and disease. Nat Rev Immunol 16: 676–689, 2016. doi: 10.1038/nri.2016.95. [DOI] [PubMed] [Google Scholar]
  • 156.Lin HC, Wan FJ, Kang BH, Wu CC, Tseng CJ. Systemic administration of lipopolysaccharide induces release of nitric oxide and glutamate and c-fos expression in the nucleus tractus solitarii of rats. Hypertension 33: 1218–1224, 1999. doi: 10.1161/01.HYP.33.5.1218. [DOI] [PubMed] [Google Scholar]
  • 157.Liu B, Tai Y, Achanta S, Kaelberer MM, Caceres AI, Shao X, Fang J, Jordt S-E. IL-33/ST2 signaling excites sensory neurons and mediates itch response in a mouse model of poison ivy contact allergy. Proc Natl Acad Sci USA 113: E7572–E7579, 2016. doi: 10.1073/pnas.1606608113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Liu T, Xie C, Chen X, Zhao F, Liu AM, Cho DB, Chong J, Yang PC. Role of muscarinic receptor activation in regulating immune cell activity in nasal mucosa. Allergy 65: 969–977, 2010. doi: 10.1111/j.1398-9995.2009.02281.x. [DOI] [PubMed] [Google Scholar]
  • 159.Liu T, Xu Z-Z, Park C-K, Berta T, Ji R-R. Toll-like receptor 7 mediates pruritus. Nat Neurosci 13: 1460–1462, 2010. doi: 10.1038/nn.2683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Lomax AE, O’Reilly M, Neshat S, Vanner SJ. Sympathetic vasoconstrictor regulation of mouse colonic submucosal arterioles is altered in experimental colitis. J Physiol 583: 719–730, 2007. doi: 10.1113/jphysiol.2007.136838. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Lorton D, Hewitt D, Bellinger DL, Felten SY, Felten DL. Noradrenergic reinnervation of the rat spleen following chemical sympathectomy with 6-hydroxydopamine: pattern and time course of reinnervation. Brain Behav Immun 4: 198–222, 1990. doi: 10.1016/0889-1591(90)90023-J. [DOI] [PubMed] [Google Scholar]
  • 162.Low D, Nguyen DD, Mizoguchi E. Animal models of ulcerative colitis and their application in drug research. Drug Des Devel Ther 7: 1341–1357, 2013. doi: 10.2147/DDDT.S40107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Lu B, Kwan K, Levine YA, Olofsson PS, Yang H, Li J, Joshi S, Wang H, Andersson U, Chavan SS, Tracey KJ. α7 Nicotinic acetylcholine receptor signaling inhibits inflammasome activation by preventing mitochondrial DNA release. Mol Med 20: 350–358, 2014. doi: 10.2119/molmed.2013.00117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Lubbers T, De Haan JJ, Hadfoune M, Zhang Y, Luyer MD, Grundy D, Buurman WA, Greve JW. Lipid-enriched enteral nutrition controls the inflammatory response in murine Gram-negative sepsis. Crit Care Med 38: 1996–2002, 2010. doi: 10.1097/CCM.0b013e3181eb90d7. [DOI] [PubMed] [Google Scholar]
  • 165.Luyer MD, Greve JW, Hadfoune M, Jacobs JA, Dejong CH, Buurman WA. Nutritional stimulation of cholecystokinin receptors inhibits inflammation via the vagus nerve. J Exp Med 202: 1023–1029, 2005. doi: 10.1084/jem.20042397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Luyer MD, Jacobs JA, Vreugdenhil AC, Hadfoune M, Dejong CH, Buurman WA, Greve JW. Enteral administration of high-fat nutrition before and directly after hemorrhagic shock reduces endotoxemia and bacterial translocation. Ann Surg 239: 257–264, 2004. doi: 10.1097/01.sla.0000108695.60059.80. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Mapp PI, Walsh DA, Garrett NE, Kidd BL, Cruwys SC, Polak JM, Blake DR. Effect of three animal models of inflammation on nerve fibres in the synovium. Ann Rheum Dis 53: 240–246, 1994. doi: 10.1136/ard.53.4.240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Marazziti D, Baroni S, Catena Dell’Osso M, Masala I, Fabbrini L, Betti L, Giannaccini G, Dell’osso B, Lucacchini A. Presence and characterization of the dopamine transporter in human resting lymphocytes. Neurochem Res 33: 1011–1016, 2008. doi: 10.1007/s11064-007-9541-4. [DOI] [PubMed] [Google Scholar]
  • 169.Martelli D, Farmer DGS, Yao ST. The splanchnic anti-inflammatory pathway: could it be the efferent arm of the inflammatory reflex? Exp Physiol 101: 1245–1252, 2016. doi: 10.1113/EP085559. [DOI] [PubMed] [Google Scholar]
  • 170.Martelli D, McKinley MJ, McAllen RM. The cholinergic anti-inflammatory pathway: a critical review. Auton Neurosci 182: 65–69, 2014. doi: 10.1016/j.autneu.2013.12.007. [DOI] [PubMed] [Google Scholar]
  • 171.Marty V, El Hachmane M, Amédée T. Dual modulation of synaptic transmission in the nucleus tractus solitarius by prostaglandin E2 synthesized downstream of IL-1beta. Eur J Neurosci 27: 3132–3150, 2008. doi: 10.1111/j.1460-9568.2008.06296.x. [DOI] [PubMed] [Google Scholar]
  • 172.Matsunaga K, Klein TW, Friedman H, Yamamoto Y. Involvement of nicotinic acetylcholine receptors in suppression of antimicrobial activity and cytokine responses of alveolar macrophages to Legionella pneumophila infection by nicotine. J Immunol 167: 6518–6524, 2001. doi: 10.4049/jimmunol.167.11.6518. [DOI] [PubMed] [Google Scholar]
  • 173.Matteoli G, Gomez-Pinilla PJ, Nemethova A, Di Giovangiulio M, Cailotto C, van Bree SH, Michel K, Tracey KJ, Schemann M, Boesmans W, Vanden Berghe P, Boeckxstaens GE. A distinct vagal anti-inflammatory pathway modulates intestinal muscularis resident macrophages independent of the spleen. Gut 63: 938–948, 2014. doi: 10.1136/gutjnl-2013-304676. [DOI] [PubMed] [Google Scholar]
  • 174.McIlwain DR, Berger T, Mak TW. Caspase functions in cell death and disease. Cold Spring Harb Perspect Biol 5: a008656, 2013. doi: 10.1101/cshperspect.a008656. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.McInnes IB, Schett G. The pathogenesis of rheumatoid arthritis. N Engl J Med 365: 2205–2219, 2011. doi: 10.1056/NEJMra1004965. [DOI] [PubMed] [Google Scholar]
  • 176.McLean LP, Smith A, Cheung L, Sun R, Grinchuk V, Vanuytsel T, Desai N, Urban JF Jr, Zhao A, Raufman J-P, Shea-Donohue T. Type 3 Muscarinic Receptors Contribute to Clearance of Citrobacter rodentium. Inflamm Bowel Dis 21: 1860–1871, 2015. doi: 10.1097/MIB.0000000000000408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Mikkelsen HB. Interstitial cells of Cajal, macrophages and mast cells in the gut musculature: morphology, distribution, spatial and possible functional interactions. J Cell Mol Med 14: 818–832, 2010. doi: 10.1111/j.1582-4934.2010.01025.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Mikkelsen HB, Thuneberg L, Rumessen JJ, Thorball N. Macrophage-like cells in the muscularis externa of mouse small intestine. Anat Rec 213: 77–86, 1985. doi: 10.1002/ar.1092130111. [DOI] [PubMed] [Google Scholar]
  • 179.Moore RD, Mumaw VR, Schoenberg MD. The Structure of the Spleen and Its Functional Implications. Exp Mol Pathol 3: 31–50, 1964. doi: 10.1016/0014-4800(64)90016-4. [DOI] [PubMed] [Google Scholar]
  • 180.Mora JR, Bono MR, Manjunath N, Weninger W, Cavanagh LL, Rosemblatt M, Von Andrian UH. Selective imprinting of gut-homing T cells by Peyer’s patch dendritic cells. Nature 424: 88–93, 2003. doi: 10.1038/nature01726. [DOI] [PubMed] [Google Scholar]
  • 181.Moroni M, Zwart R, Sher E, Cassels BK, Bermudez I. α4β2 nicotinic receptors with high and low acetylcholine sensitivity: pharmacology, stoichiometry, and sensitivity to long-term exposure to nicotine. Mol Pharmacol 70: 755–768, 2006. doi: 10.1124/mol.106.023044. [DOI] [PubMed] [Google Scholar]
  • 182.Moussion C, Ortega N, Girard J-P. The IL-1-like cytokine IL-33 is constitutively expressed in the nucleus of endothelial cells and epithelial cells in vivo: a novel ‘alarmin’? PLoS One 3: e3331, 2008. doi: 10.1371/journal.pone.0003331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Muller PA, Koscsó B, Rajani GM, Stevanovic K, Berres M-L, Hashimoto D, Mortha A, Leboeuf M, Li X-M, Mucida D, Stanley ER, Dahan S, Margolis KG, Gershon MD, Merad M, Bogunovic M. Crosstalk between muscularis macrophages and enteric neurons regulates gastrointestinal motility. Cell 158: 300–313, 2014. [Erratum in Cell 158: 1210, 2014.] doi: 10.1016/j.cell.2014.04.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Munyaka P, Rabbi MF, Pavlov VA, Tracey KJ, Khafipour E, Ghia JE. Central muscarinic cholinergic activation alters interaction between splenic dendritic cell and CD4+CD25- T cells in experimental colitis. PLoS One 9: e109272, 2014. doi: 10.1371/journal.pone.0109272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Murray K, Godinez DR, Brust-Mascher I, Miller EN, Gareau MG, Reardon C. Neuroanatomy of the spleen: mapping the relationship between sympathetic neurons and lymphocytes. PLoS One 12: e0182416, 2017. doi: 10.1371/journal.pone.0182416. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Nance DM, Burns J. Innervation of the spleen in the rat: evidence for absence of afferent innervation. Brain Behav Immun 3: 281–290, 1989. doi: 10.1016/0889-1591(89)90028-7. [DOI] [PubMed] [Google Scholar]
  • 187.Narasaraju T, Yang E, Samy RP, Ng HH, Poh WP, Liew AA, Phoon MC, van Rooijen N, Chow VT. Excessive neutrophils and neutrophil extracellular traps contribute to acute lung injury of influenza pneumonitis. Am J Pathol 179: 199–210, 2011. doi: 10.1016/j.ajpath.2011.03.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Nelson ME, Kuryatov A, Choi CH, Zhou Y, Lindstrom J. Alternate stoichiometries of α4β2 nicotinic acetylcholine receptors. Mol Pharmacol 63: 332–341, 2003. doi: 10.1124/mol.63.2.332. [DOI] [PubMed] [Google Scholar]
  • 189.Nielsen OH, Ainsworth MA. Tumor necrosis factor inhibitors for inflammatory bowel disease. N Engl J Med 369: 754–762, 2013. doi: 10.1056/NEJMct1209614. [DOI] [PubMed] [Google Scholar]
  • 190.Niijima A. Electrophysiological study on the vagal innervation of the adrenal gland in the rat. J Auton Nerv Syst 41: 87–92, 1992. doi: 10.1016/0165-1838(92)90130-9. [DOI] [PubMed] [Google Scholar]
  • 191.Okayasu I, Hatakeyama S, Yamada M, Ohkusa T, Inagaki Y, Nakaya R. A novel method in the induction of reliable experimental acute and chronic ulcerative colitis in mice. Gastroenterology 98: 694–702, 1990. doi: 10.1016/0016-5085(90)90290-H. [DOI] [PubMed] [Google Scholar]
  • 192.Olivier BJ, Cailotto C, van der Vliet J, Knippenberg M, Greuter MJ, Hilbers FW, Konijn T, Te Velde AA, Nolte MA, Boeckxstaens GE, de Jonge WJ, Mebius RE. Vagal innervation is required for the formation of tertiary lymphoid tissue in colitis. Eur J Immunol 46: 2467–2480, 2016. doi: 10.1002/eji.201646370. [DOI] [PubMed] [Google Scholar]
  • 193.Olofsson PS, Katz DA, Rosas-Ballina M, Levine YA, Ochani M, Valdés-Ferrer SI, Pavlov VA, Tracey KJ, Chavan SS. α7 nicotinic acetylcholine receptor (α7nAChR) expression in bone marrow-derived non-T cells is required for the inflammatory reflex. Mol Med 18: 539–543, 2012. doi: 10.2119/molmed.2011.00405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193a.Olofsson PS, Levine YA, Caravaca A, Chavan SS, Pavlov VA, Faltys M, Tracey KJ. Single-Pulse and Unidirectional Electrical Activation of the Cervical Vagus Nerve Reduces Tumor Necrosis Factor in Endotoxemia. Bioelectron Med 2: 37–42, 2015. doi: 10.15424/bioelectronmed.2015.00006. [DOI] [Google Scholar]
  • 194.Pabst MJ, Pabst KM, Collier JA, Coleman TC, Lemons-Prince ML, Godat MS, Waring MB, Babu JP. Inhibition of neutrophil and monocyte defensive functions by nicotine. J Periodontol 66: 1047–1055, 1995. doi: 10.1902/jop.1995.66.12.1047. [DOI] [PubMed] [Google Scholar]
  • 195.Pabst O, Mowat AM. Oral tolerance to food protein. Mucosal Immunol 5: 232–239, 2012. doi: 10.1038/mi.2012.4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Padro CJ, Sanders VM. Neuroendocrine regulation of inflammation. Semin Immunol 26: 357–368, 2014. doi: 10.1016/j.smim.2014.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Parsek MR, Greenberg EP. Acyl-homoserine lactone quorum sensing in gram-negative bacteria: a signaling mechanism involved in associations with higher organisms. Proc Natl Acad Sci USA 97: 8789–8793, 2000. doi: 10.1073/pnas.97.16.8789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Pastorelli L, Garg RR, Hoang SB, Spina L, Mattioli B, Scarpa M, Fiocchi C, Vecchi M, Pizarro TT. Epithelial-derived IL-33 and its receptor ST2 are dysregulated in ulcerative colitis and in experimental Th1/Th2 driven enteritis. Proc Natl Acad Sci USA 107: 8017–8022, 2010. doi: 10.1073/pnas.0912678107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Patel YA, Saxena T, Bellamkonda RV, Butera RJ. Kilohertz frequency nerve block enhances anti-inflammatory effects of vagus nerve stimulation. Sci Rep 7: 39810, 2017. doi: 10.1038/srep39810. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Pavlov VA, Wang H, Czura CJ, Friedman SG, Tracey KJ. The cholinergic anti-inflammatory pathway: a missing link in neuroimmunomodulation. Mol Med 9: 125–134, 2003. [PMC free article] [PubMed] [Google Scholar]
  • 202.Peña G, Cai B, Liu J, van der Zanden EP, Deitch EA, de Jonge WJ, Ulloa L. Unphosphorylated STAT3 modulates alpha 7 nicotinic receptor signaling and cytokine production in sepsis. Eur J Immunol 40: 2580–2589, 2010. doi: 10.1002/eji.201040540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Pera T, Zuidhof A, Valadas J, Smit M, Schoemaker RG, Gosens R, Maarsingh H, Zaagsma J, Meurs H. Tiotropium inhibits pulmonary inflammation and remodelling in a guinea pig model of COPD. Eur Respir J 38: 789–796, 2011. doi: 10.1183/09031936.00146610. [DOI] [PubMed] [Google Scholar]
  • 204.Poltorak A, He X, Smirnova I, Liu MY, Van Huffel C, Du X, Birdwell D, Alejos E, Silva M, Galanos C, Freudenberg M, Ricciardi-Castagnoli P, Layton B, Beutler B. Defective LPS signaling in C3H/HeJ and C57BL/10ScCr mice: mutations in Tlr4 gene. Science 282: 2085–2088, 1998. doi: 10.1126/science.282.5396.2085. [DOI] [PubMed] [Google Scholar]
  • 205.Prandi S, Bromke M, Hübner S, Voigt A, Boehm U, Meyerhof W, Behrens M. A subset of mouse colonic goblet cells expresses the bitter taste receptor Tas2r131. PLoS One 8: e82820, 2013. doi: 10.1371/journal.pone.0082820. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Price CJ, Hoyda TD, Ferguson AV. The area postrema: a brain monitor and integrator of systemic autonomic state. Neuroscientist 14: 182–194, 2008. doi: 10.1177/1073858407311100. [DOI] [PubMed] [Google Scholar]
  • 207.Profita M, Giorgi RD, Sala A, Bonanno A, Riccobono L, Mirabella F, Gjomarkaj M, Bonsignore G, Bousquet J, Vignola AM. Muscarinic receptors, leukotriene B4 production and neutrophilic inflammation in COPD patients. Allergy 60: 1361–1369, 2005. doi: 10.1111/j.1398-9995.2005.00892.x. [DOI] [PubMed] [Google Scholar]
  • 208.Qi J, Buzas K, Fan H, Cohen JI, Wang K, Mont E, Klinman D, Oppenheim JJ, Howard OMZ. Painful pathways induced by TLR stimulation of dorsal root ganglion neurons. J Immunol 186: 6417–6426, 2011. doi: 10.4049/jimmunol.1001241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Qian J, Galitovskiy V, Chernyavsky AI, Marchenko S, Grando SA. Plasticity of the murine spleen T-cell cholinergic receptors and their role in in vitro differentiation of naïve CD4 T cells toward the Th1, Th2 and Th17 lineages. Genes Immun 12: 222–230, 2011. doi: 10.1038/gene.2010.72. [DOI] [PubMed] [Google Scholar]
  • 210.Rahman AA, Robinson AM, Jovanovska V, Eri R, Nurgali K. Alterations in the distal colon innervation in Winnie mouse model of spontaneous chronic colitis. Cell Tissue Res 362: 497–512, 2015. doi: 10.1007/s00441-015-2251-3. [DOI] [PubMed] [Google Scholar]
  • 211.Raybould HE. Gut chemosensing: interactions between gut endocrine cells and visceral afferents. Auton Neurosci 153: 41–46, 2010. doi: 10.1016/j.autneu.2009.07.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Raybould HE. Visceral perception: sensory transduction in visceral afferents and nutrients. Gut 51, Suppl 1: i11–i14, 2002. doi: 10.1136/gut.51.suppl_1.i11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Razani-Boroujerdi S, Boyd RT, Dávila-García MI, Nandi JS, Mishra NC, Singh SP, Pena-Philippides JC, Langley R, Sopori ML. T cells express α7-nicotinic acetylcholine receptor subunits that require a functional TCR and leukocyte-specific protein tyrosine kinase for nicotine-induced Ca2+ response. J Immunol 179: 2889–2898, 2007. doi: 10.4049/jimmunol.179.5.2889. [DOI] [PubMed] [Google Scholar]
  • 214.Reardon C, Duncan GS, Brüstle A, Brenner D, Tusche MW, Olofsson PS, Rosas-Ballina M, Tracey KJ, Mak TW. Lymphocyte-derived ACh regulates local innate but not adaptive immunity. Proc Natl Acad Sci USA 110: 1410–1415, 2013. [Correction in Proc Natl Acad Sci USA 110: 5269, 2013.] 10.1073/pnas.1221655110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Reardon C, Lechmann M, Brüstle A, Gareau MG, Shuman N, Philpott D, Ziegler SF, Mak TW. Thymic stromal lymphopoetin-induced expression of the endogenous inhibitory enzyme SLPI mediates recovery from colonic inflammation. Immunity 35: 223–235, 2011. doi: 10.1016/j.immuni.2011.05.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Reardon C, Sanchez A, Hogaboam CM, McKay DM. Tapeworm infection reduces epithelial ion transport abnormalities in murine dextran sulfate sodium-induced colitis. Infect Immun 69: 4417–4423, 2001. doi: 10.1128/IAI.69.7.4417-4423.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Reilly FD, McCuskey PA, Miller ML, McCuskey RS, Meineke HA. Innervation of the periarteriolar lymphatic sheath of the spleen. Tissue Cell 11: 121–126, 1979. doi: 10.1016/0040-8166(79)90012-0. [DOI] [PubMed] [Google Scholar]
  • 218.Rinner I, Kawashima K, Schauenstein K. Rat lymphocytes produce and secrete acetylcholine in dependence of differentiation and activation. J Neuroimmunol 81: 31–37, 1998. doi: 10.1016/S0165-5728(97)00155-0. [DOI] [PubMed] [Google Scholar]
  • 219.Rivollier A, He J, Kole A, Valatas V, Kelsall BL. Inflammation switches the differentiation program of Ly6Chi monocytes from anti-inflammatory macrophages to inflammatory dendritic cells in the colon. J Exp Med 209: 139–155, 2012. doi: 10.1084/jem.20101387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Roberts JA, Lukewich MK, Sharkey KA, Furness JB, Mawe GM, Lomax AE. The roles of purinergic signaling during gastrointestinal inflammation. Curr Opin Pharmacol 12: 659–666, 2012. doi: 10.1016/j.coph.2012.09.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Rodgman A, Perfetti TA. The Chemical Components of Tobacco and Tobacco Smoke. Boca Raton, FL: CRC, 2013, p. xciii. doi: 10.1201/b13973 [DOI] [Google Scholar]
  • 222.Rogers RC, Van Meter MJ, Hermann GE. Tumor necrosis factor potentiates central vagal afferent signaling by modulating ryanodine channels. J Neurosci 26: 12642–12646, 2006. doi: 10.1523/JNEUROSCI.3530-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Rosas-Ballina M, Ochani M, Parrish WR, Ochani K, Harris YT, Huston JM, Chavan S, Tracey KJ. Splenic nerve is required for cholinergic antiinflammatory pathway control of TNF in endotoxemia. Proc Natl Acad Sci USA 105: 11008–11013, 2008. doi: 10.1073/pnas.0803237105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Rosas-Ballina M, Olofsson PS, Ochani M, Valdés-Ferrer SI, Levine YA, Reardon C, Tusche MW, Pavlov VA, Andersson U, Chavan S, Mak TW, Tracey KJ. Acetylcholine-synthesizing T cells relay neural signals in a vagus nerve circuit. Science 334: 98–101, 2011. doi: 10.1126/science.1209985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Ruppelt A, Mosenden R, Grönholm M, Aandahl EM, Tobin D, Carlson CR, Abrahamsen H, Herberg FW, Carpén O, Taskén K. Inhibition of T cell activation by cyclic adenosine 5′-monophosphate requires lipid raft targeting of protein kinase A type I by the A-kinase anchoring protein ezrin. J Immunol 179: 5159–5168, 2007. doi: 10.4049/jimmunol.179.8.5159. [DOI] [PubMed] [Google Scholar]
  • 226.Safronova VG, Vulfius CA, Shelukhina IV, Mal’tseva VN, Berezhnov AV, Fedotova EI, Miftahova RG, Kryukova EV, Grinevich AA, Tsetlin VI. Nicotinic receptor involvement in regulation of functions of mouse neutrophils from inflammatory site. Immunobiology 221: 761–772, 2016. doi: 10.1016/j.imbio.2016.01.016. [DOI] [PubMed] [Google Scholar]
  • 227.Salamone G, Lombardi G, Gori S, Nahmod K, Jancic C, Amaral MM, Vermeulen M, Español A, Sales ME, Geffner J. Cholinergic modulation of dendritic cell function. J Neuroimmunol 236: 47–56, 2011. doi: 10.1016/j.jneuroim.2011.05.007. [DOI] [PubMed] [Google Scholar]
  • 229.Saleh A, Smith DR, Tessler L, Mateo AR, Martens C, Schartner E, Van der Ploeg R, Toth C, Zochodne DW, Fernyhough P. Receptor for advanced glycation end-products (RAGE) activates divergent signaling pathways to augment neurite outgrowth of adult sensory neurons. Exp Neurol 249: 149–159, 2013. doi: 10.1016/j.expneurol.2013.08.018. [DOI] [PubMed] [Google Scholar]
  • 230.Sandilos JK, Chiu YH, Chekeni FB, Armstrong AJ, Walk SF, Ravichandran KS, Bayliss DA. Pannexin 1, an ATP release channel, is activated by caspase cleavage of its pore-associated C-terminal autoinhibitory region. J Biol Chem 287: 11303–11311, 2012. doi: 10.1074/jbc.M111.323378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231.Sato KZ, Fujii T, Watanabe Y, Yamada S, Ando T, Kazuko F, Kawashima K. Diversity of mRNA expression for muscarinic acetylcholine receptor subtypes and neuronal nicotinic acetylcholine receptor subunits in human mononuclear leukocytes and leukemic cell lines. Neurosci Lett 266: 17–20, 1999. doi: 10.1016/S0304-3940(99)00259-1. [DOI] [PubMed] [Google Scholar]
  • 232.Savidge TC, Newman P, Pothoulakis C, Ruhl A, Neunlist M, Bourreille A, Hurst R, Sofroniew MV. Enteric glia regulate intestinal barrier function and inflammation via release of S-nitrosoglutathione. Gastroenterology 132: 1344–1358, 2007. doi: 10.1053/j.gastro.2007.01.051. [DOI] [PubMed] [Google Scholar]
  • 233.Schaefer L. Complexity of danger: the diverse nature of damage-associated molecular patterns. J Biol Chem 289: 35237–35245, 2014. doi: 10.1074/jbc.R114.619304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Schelegle ES, Green JF. An overview of the anatomy and physiology of slowly adapting pulmonary stretch receptors. Respir Physiol 125: 17–31, 2001. doi: 10.1016/S0034-5687(00)00202-4. [DOI] [PubMed] [Google Scholar]
  • 235.Sedhom MA, Pichery M, Murdoch JR, Foligné B, Ortega N, Normand S, Mertz K, Sanmugalingam D, Brault L, Grandjean T, Lefrancais E, Fallon PG, Quesniaux V, Peyrin-Biroulet L, Cathomas G, Junt T, Chamaillard M, Girard JP, Ryffel B. Neutralisation of the interleukin-33/ST2 pathway ameliorates experimental colitis through enhancement of mucosal healing in mice. Gut 62: 1714–1723, 2013. doi: 10.1136/gutjnl-2011-301785. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Seow WK, Thong YH, Nelson RD, MacFarlane GD, Herzberg MC. Nicotine-induced release of elastase and eicosanoids by human neutrophils. Inflammation 18: 119–127, 1994. doi: 10.1007/BF01534553. [DOI] [PubMed] [Google Scholar]
  • 237.Sharkey KA. Emerging roles for enteric glia in gastrointestinal disorders. J Clin Invest 125: 918–925, 2015. doi: 10.1172/JCI76303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Skok M, Grailhe R, Changeux J-P. Nicotinic receptors regulate B lymphocyte activation and immune response. Eur J Pharmacol 517: 246–251, 2005. doi: 10.1016/j.ejphar.2005.05.011. [DOI] [PubMed] [Google Scholar]
  • 239.Sloan EK, Nguyen CT, Cox BF, Tarara RP, Capitanio JP, Cole SW. SIV infection decreases sympathetic innervation of primate lymph nodes: the role of neurotrophins. Brain Behav Immun 22: 185–194, 2008. doi: 10.1016/j.bbi.2007.07.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Smith RS, Iglewski BH. Pseudomonas aeruginosa quorum sensing as a potential antimicrobial target. J Clin Invest 112: 1460–1465, 2003. doi: 10.1172/JCI200320364. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Souslova V, Cesare P, Ding Y, Akopian AN, Stanfa L, Suzuki R, Carpenter K, Dickenson A, Boyce S, Hill R, Nebenuis-Oosthuizen D, Smith AJH, Kidd EJ, Wood JN. Warm-coding deficits and aberrant inflammatory pain in mice lacking P2X3 receptors. Nature 407: 1015–1017, 2000. doi: 10.1038/35039526. [DOI] [PubMed] [Google Scholar]
  • 242.Stauffer WR, Lak A, Yang A, Borel M, Paulsen O, Boyden ES, Schultz W. Dopamine Neuron-Specific Optogenetic Stimulation in Rhesus Macaques. Cell 166: 1564–1571.e6, 2016. doi: 10.1016/j.cell.2016.08.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Straub RH, Grum F, Strauch U, Capellino S, Bataille F, Bleich A, Falk W, Schölmerich J, Obermeier F. Anti-inflammatory role of sympathetic nerves in chronic intestinal inflammation. Gut 57: 911–921, 2008. doi: 10.1136/gut.2007.125401. [DOI] [PubMed] [Google Scholar]
  • 244.Strober W, Fuss IJ, Blumberg RS. The immunology of mucosal models of inflammation. Annu Rev Immunol 20: 495–549, 2002. doi: 10.1146/annurev.immunol.20.100301.064816. [DOI] [PubMed] [Google Scholar]
  • 245.Sun Y, Li Q, Gui H, Xu DP, Yang YL, Su DF, Liu X. MicroRNA-124 mediates the cholinergic anti-inflammatory action through inhibiting the production of pro-inflammatory cytokines. Cell Res 23: 1270–1283, 2013. doi: 10.1038/cr.2013.116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Talbot S, Abdulnour RE, Burkett PR, Lee S, Cronin SJ, Pascal MA, Laedermann C, Foster SL, Tran JV, Lai N, Chiu IM, Ghasemlou N, DiBiase M, Roberson D, Von Hehn C, Agac B, Haworth O, Seki H, Penninger JM, Kuchroo VK, Bean BP, Levy BD, Woolf CJ. Silencing Nociceptor Neurons Reduces Allergic Airway Inflammation. Neuron 87: 341–354, 2015. doi: 10.1016/j.neuron.2015.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Tallini YN, Shui B, Greene KS, Deng KY, Doran R, Fisher PJ, Zipfel W, Kotlikoff MI. BAC transgenic mice express enhanced green fluorescent protein in central and peripheral cholinergic neurons. Physiol Genomics 27: 391–397, 2006. doi: 10.1152/physiolgenomics.00092.2006. [DOI] [PubMed] [Google Scholar]
  • 248.Talman WT, Perrone MH, Reis DJ. Evidence for l-glutamate as the neurotransmitter of baroreceptor afferent nerve fibers. Science 209: 813–815, 1980. doi: 10.1126/science.6105709. [DOI] [PubMed] [Google Scholar]
  • 249.Tayebati SK, El-Assouad D, Ricci A, Amenta F. Immunochemical and immunocytochemical characterization of cholinergic markers in human peripheral blood lymphocytes. J Neuroimmunol 132: 147–155, 2002. doi: 10.1016/S0165-5728(02)00325-9. [DOI] [PubMed] [Google Scholar]
  • 250.Tian J, Chau C, Hales TG, Kaufman DL. GABA(A) receptors mediate inhibition of T cell responses. J Neuroimmunol 96: 21–28, 1999. doi: 10.1016/S0165-5728(98)00264-1. [DOI] [PubMed] [Google Scholar]
  • 251.Tizzano M, Gulbransen BD, Vandenbeuch A, Clapp TR, Herman JP, Sibhatu HM, Churchill MEA, Silver WL, Kinnamon SC, Finger TE. Nasal chemosensory cells use bitter taste signaling to detect irritants and bacterial signals. Proc Natl Acad Sci USA 107: 3210–3215, 2010. doi: 10.1073/pnas.0911934107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Tlaskalová-Hogenová H, Tucková L, Stepánková R, Hudcovic T, Palová-Jelínková L, Kozáková H, Rossmann P, Sanchez D, Cinová J, Hrncír T, Kverka M, Frolová L, Uhlig H, Powrie F, Bland P. Involvement of innate immunity in the development of inflammatory and autoimmune diseases. Ann N Y Acad Sci 1051: 787–798, 2005. doi: 10.1196/annals.1361.122. [DOI] [PubMed] [Google Scholar]
  • 253.Torres-Rosas R, Yehia G, Peña G, Mishra P, del Rocio Thompson-Bonilla M, Moreno-Eutimio MA, Arriaga-Pizano LA, Isibasi A, Ulloa L. Dopamine mediates vagal modulation of the immune system by electroacupuncture. Nat Med 20: 291–295, 2014. doi: 10.1038/nm.3479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Totti N III, McCusker KT, Campbell EJ, Griffin GL, Senior RM. Nicotine is chemotactic for neutrophils and enhances neutrophil responsiveness to chemotactic peptides. Science 223: 169–171, 1984. doi: 10.1126/science.6318317. [DOI] [PubMed] [Google Scholar]
  • 255.Tse KH, Chow KB, Leung WK, Wong YH, Wise H. Primary sensory neurons regulate Toll-like receptor-4-dependent activity of glial cells in dorsal root ganglia. Neuroscience 279: 10–22, 2014. doi: 10.1016/j.neuroscience.2014.08.033. [DOI] [PubMed] [Google Scholar]
  • 256.Tsukamoto K, Yin M, Sved AF. Effect of atropine injected into the nucleus tractus solitarius on the regulation of blood pressure. Brain Res 648: 9–15, 1994. doi: 10.1016/0006-8993(94)91898-8. [DOI] [PubMed] [Google Scholar]
  • 257.Ueno M, Ueno-Nakamura Y, Niehaus J, Popovich PG, Yoshida Y. Silencing spinal interneurons inhibits immune suppressive autonomic reflexes caused by spinal cord injury. Nat Neurosci 19: 784–787, 2016. doi: 10.1038/nn.4289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Van de Pavert SA, Olivier BJ, Goverse G, Vondenhoff MF, Greuter M, Beke P, Kusser K, Höpken UE, Lipp M, Niederreither K, Blomhoff R, Sitnik K, Agace WW, Randall TD, de Jonge WJ, Mebius RE. Chemokine CXCL13 is essential for lymph node initiation and is induced by retinoic acid and neuronal stimulation. Nat Immunol 10: 1193–1199, 2009. doi: 10.1038/ni.1789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Van der Zanden EP, Snoek SA, Heinsbroek SE, Stanisor OI, Verseijden C, Boeckxstaens GE, Peppelenbosch MP, Greaves DR, Gordon S, De Jonge WJ. Vagus nerve activity augments intestinal macrophage phagocytosis via nicotinic acetylcholine receptor α4β2. Gastroenterology 137: 1029–1039.e4, 2009. doi: 10.1053/j.gastro.2009.04.057. [DOI] [PubMed] [Google Scholar]
  • 260.Vang T, Abrahamsen H, Myklebust S, Horejsí V, Taskén K. Combined spatial and enzymatic regulation of Csk by cAMP and protein kinase a inhibits T cell receptor signaling. J Biol Chem 278: 17597–17600, 2003. doi: 10.1074/jbc.C300077200. [DOI] [PubMed] [Google Scholar]
  • 261.Vang T, Torgersen KM, Sundvold V, Saxena M, Levy FO, Skålhegg BS, Hansson V, Mustelin T, Taskén K. Activation of the COOH-terminal Src kinase (Csk) by cAMP-dependent protein kinase inhibits signaling through the T cell receptor. J Exp Med 193: 497–507, 2001. doi: 10.1084/jem.193.4.497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.Villiger Y, Szanto I, Jaconi S, Blanchet C, Buisson B, Krause KH, Bertrand D, Romand JA. Expression of an α7 duplicate nicotinic acetylcholine receptor-related protein in human leukocytes. J Neuroimmunol 126: 86–98, 2002. doi: 10.1016/S0165-5728(02)00057-7. [DOI] [PubMed] [Google Scholar]
  • 263.Vincent AM, Perrone L, Sullivan KA, Backus C, Sastry AM, Lastoskie C, Feldman EL. Receptor for advanced glycation end products activation injures primary sensory neurons via oxidative stress. Endocrinology 148: 548–558, 2007. doi: 10.1210/en.2006-0073. [DOI] [PubMed] [Google Scholar]
  • 264.Waldburger J-M, Firestein GS. Regulation of peripheral inflammation by the central nervous system. Curr Rheumatol Rep 12: 370–378, 2010. doi: 10.1007/s11926-010-0124-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Wang H, Yu M, Ochani M, Amella CA, Tanovic M, Susarla S, Li JH, Wang H, Yang H, Ulloa L, Al-Abed Y, Czura CJ, Tracey KJ. Nicotinic acetylcholine receptor alpha7 subunit is an essential regulator of inflammation. Nature 421: 384–388, 2003. doi: 10.1038/nature01339. [DOI] [PubMed] [Google Scholar]
  • 266.Wang L, Feng D, Yan H, Wang Z, Pei L. Comparative analysis of P2X1, P2X2, P2X3, and P2X4 receptor subunits in rat nodose ganglion neurons. PLoS One 9: e96699, 2014. doi: 10.1371/journal.pone.0096699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Wang Y, Li G, Yu K, Liang S, Wan F, Xu C, Gao Y, Liu S, Lin J. Expressions of P2X2 and P2X3 receptors in rat nodose neurons after myocardial ischemia injury. Auton Neurosci 145: 71–75, 2009. doi: 10.1016/j.autneu.2008.11.006. [DOI] [PubMed] [Google Scholar]
  • 268.Warwick RA, Hanani M. The contribution of satellite glial cells to chemotherapy-induced neuropathic pain. Eur J Pain 17: 571–580, 2013. doi: 10.1002/j.1532-2149.2012.00219.x. [DOI] [PubMed] [Google Scholar]
  • 269.Willemze RA, Welting O, van Hamersveld HP, Meijer SL, Folgering JHA, Darwinkel H, Witherington J, Sridhar A, Vervoordeldonk MJ, Seppen J, de Jonge WJ. Neuronal control of experimental colitis occurs via sympathetic intestinal innervation. Neurogastroenterol Motil 30: e13163, 2018. doi: 10.1111/nmo.13163. [DOI] [PubMed] [Google Scholar]
  • 270.Wu SV, Rozengurt N, Yang M, Young SH, Sinnett-Smith J, Rozengurt E. Expression of bitter taste receptors of the T2R family in the gastrointestinal tract and enteroendocrine STC-1 cells. Proc Natl Acad Sci USA 99: 2392–2397, 2002. doi: 10.1073/pnas.042617699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 271.Xavier RJ, Podolsky DK. Unravelling the pathogenesis of inflammatory bowel disease. Nature 448: 427–434, 2007. doi: 10.1038/nature06005. [DOI] [PubMed] [Google Scholar]
  • 272.Xu W, He B, Chiu A, Chadburn A, Shan M, Buldys M, Ding A, Knowles DM, Santini PA, Cerutti A. Epithelial cells trigger frontline immunoglobulin class switching through a pathway regulated by the inhibitor SLPI. Nat Immunol 8: 294–303, 2007. doi: 10.1038/ni1434. [DOI] [PubMed] [Google Scholar]
  • 273.Yan Y, Jiang W, Liu L, Wang X, Ding C, Tian Z, Zhou R. Dopamine controls systemic inflammation through inhibition of NLRP3 inflammasome. Cell 160: 62–73, 2015. doi: 10.1016/j.cell.2014.11.047. [DOI] [PubMed] [Google Scholar]
  • 274.Yang L, Lindholm K, Konishi Y, Li R, Shen Y. Target depletion of distinct tumor necrosis factor receptor subtypes reveals hippocampal neuron death and survival through different signal transduction pathways. J Neurosci 22: 3025–3032, 2002. doi: 10.1523/JNEUROSCI.22-08-03025.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 275.Yasuda K, Muto T, Kawagoe T, Matsumoto M, Sasaki Y, Matsushita K, Taki Y, Futatsugi-Yumikura S, Tsutsui H, Ishii KJ, Yoshimoto T, Akira S, Nakanishi K. Contribution of IL-33-activated type II innate lymphoid cells to pulmonary eosinophilia in intestinal nematode-infected mice. Proc Natl Acad Sci USA 109: 3451–3456, 2012. doi: 10.1073/pnas.1201042109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.Yu S, Undem BJ, Kollarik M. Vagal afferent nerves with nociceptive properties in guinea-pig oesophagus. J Physiol 563: 831–842, 2005. doi: 10.1113/jphysiol.2004.079574. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Zhang Y, Chen K, Sloan SA, Bennett ML, Scholze AR, O’Keeffe S, Phatnani HP, Guarnieri P, Caneda C, Ruderisch N, Deng S, Liddelow SA, Zhang C, Daneman R, Maniatis T, Barres BA, Wu JQ. An RNA-sequencing transcriptome and splicing database of glia, neurons, and vascular cells of the cerebral cortex. J Neurosci 34: 11929–11947, 2014. doi: 10.1523/JNEUROSCI.1860-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Zhang Y, Guan Z, Reader B, Shawler T, Mandrekar-Colucci S, Huang K, Weil Z, Bratasz A, Wells J, Powell ND, Sheridan JF, Whitacre CC, Rabchevsky AG, Nash MS, Popovich PG. Autonomic dysreflexia causes chronic immune suppression after spinal cord injury. J Neurosci 33: 12970–12981, 2013. doi: 10.1523/JNEUROSCI.1974-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 279.Zhao Z, Nelson AR, Betsholtz C, Zlokovic BV. Establishment and Dysfunction of the Blood-Brain Barrier. Cell 163: 1064–1078, 2015. doi: 10.1016/j.cell.2015.10.067. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Physiological Reviews are provided here courtesy of American Physiological Society

RESOURCES