Next Article in Journal
Challenges and Perspectives in Nutritional Counselling and Nursing: A Narrative Review
Previous Article in Journal
The Increase of Osteoporotic Hip Fractures and Associated One-Year Mortality in Poland: 2008–2015
Previous Article in Special Issue
Neuroprotective Peptides in Retinal Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pituitary Adenylate Cyclase-Activating Polypeptide: 30 Years in Research Spotlight and 600 Million Years in Service

1
Department of Experimental Zoology and Neurobiology, University of Pécs, 7624 Pécs, Hungary
2
Department of Immunology, School of Medicine, Tufts University, Boston, MA 02111, USA
*
Author to whom correspondence should be addressed.
Submission received: 2 August 2019 / Revised: 2 September 2019 / Accepted: 10 September 2019 / Published: 18 September 2019

Abstract

:
Emerging from the depths of evolution, pituitary adenylate cyclase-activating polypeptide (PACAP) and its receptors (i.e., PAC1, VPAC1, VPAC2) are present in multicellular organisms from Tunicates to humans and govern a remarkable number of physiological processes. Consequently, the clinical relevance of PACAP systems spans a multifaceted palette that includes more than 40 disorders. We aimed to present the versatility of PACAP1-38 actions with a focus on three aspects: (1) when PACAP1-38 could be a cause of a malfunction, (2) when PACAP1-38 could be the cure for a malfunction, and (3) when PACAP1-38 could either improve or impair biology. PACAP1-38 is implicated in the pathophysiology of migraine and post-traumatic stress disorder whereas an outstanding protective potential has been established in ischemia and in Alzheimer’s disease. Lastly, PACAP receptors could mediate opposing effects both in cancers and in inflammation. In the light of the above, the duration and concentrations of PACAP agents must be carefully set at any application to avoid unwanted consequences. An enormous amount of data accumulated since its discovery (1989) and the first clinical trials are dated in 2017. Thus in the field of PACAP research: “this is not the end, not even the beginning of the end, but maybe the end of the beginning.”

1. Introduction

Pituitary adenylate cyclase activating polypeptide (PACAP) has been in the spotlight of extensive basic and applied research since its discovery [1,2]. The career of PACAP peptides (PACAP1-38 and PACAP1-27) is reflected by the ca. 3500 publications to date and highlighted in four excellent research articles published in Nature [3,4,5,6].
Three major binding sites have been recognized to mediate PACAP1-38 effects: PAC1, VPAC1, and VPAC2 receptors. The VPAC receptors also bind a PACAP-related peptide, vasoactive intestinal peptide (VIP) with the same high affinity as PACAP [7]. In other aspects, however, both VPAC1 and VPAC2 are unique entities. Their coding sequences show only 55% homology [8]. Several pharmaceutical agonists and antagonists differentiate between the two receptors with specific binding characteristics [7,9] and their tissue distributions show specific differences [10]. The complexity of PACAP receptors is further increased by PAC1 receptor (PAC1-R) isoforms through alternative splicing at the transcript level. In vertebrates, 20 PAC1-R isoforms have been identified [11], primarily in cell lines in vitro, but a number of them have also been verified in vivo with development-related expression patterns [12,13]. The isoforms were shown to affect ligand-binding selectivities and signaling mechanisms, but, in contrast to the large number of isoforms, only a few signal pathways are utilized. Nevertheless, differential expression of PAC1-R isoforms contributes to the differences in action exerted by the peptides.
The PACAP peptides were discovered as neurohormones first. PACAP1-38 was later recognized as a transmitter and neurotrophic mediator. In addition to the nervous system, PACAP and its receptors have been described in almost each bracket of the mammalian organism including various organs (e.g., pancreas, cardiovascular system, testes, gastrointestinal tract, etc.) [14]. Not surprisingly, approximately 40 different pathological conditions have been reported where PACAP1-38 may have clinical relevance (Table 1).
In the present review, we intended to follow the evolution of the PACAP system to elucidate why this highly conserved peptide has been spotted at the roots of nearly all physiological processes. Furthermore, to dissect disparate actions of PACAP1-38, we highlighted the six most extensively studied disorders and organize them according to PACAP1-38 involvement in pathophysiology, in prevention/treatment, and in conditions where its role happens to be ambiguous.

2. Emerging from the Depths of Evolution: PACAP/Glucagon Family

Despite its ubiquitous roles and pleiotropic regulations in various tissues, PACAP1-38 was discovered surprisingly late, as recently as 1989 [1], partly because of its unique characteristics. One of them, its highly basic charge was also fundamental in its discovery. The peptide was plotted as a highly basic releasing hormone “different from any other known releasing hormones,” according to Arimura, its discoverer [15]. Due to its positive charge, the majority of PACAP molecules circulate in ceruloplasmin complex [16], which is a unique feature among other circulating peptide hormones. This is critical to understand its availability in serum, the typical radioimmunoassay (RIA) based publications do not specify the biologically available free peptide levels, which is expected to be significantly lower. The majority of RIA assays measured 5–20 fmol/mL (15–60 pg/mL) total PACAP levels in serum [17]. Unfortunately, the biologically-available free peptide levels are not reported, but estimated to be an order of magnitude lower (0.5–2 fmol/mL or 1.5-6 pg/mL). Other short peptides with characteristics similar to PACAP and also act on G-protein coupled receptors (GPCRs) (e.g., adrenocorticotropic hormone (ACTH), parathyroid hormone, glucagon, gastrin) are different in several aspects. They have clear target organs and/or perform systemic functions in homeostasis. Their total plasma levels are several folds higher (50–150 pg/mL) than PACAP (American College of Physicians, Normal Laboratory Values, www.acponline.org) [18] and are not observed to form ceruloplasmin complexes. The low free peptide levels and the apparent lack of a systemic function strongly suggest that PACAP1-38 is a unique hormone. Most probably its major function is a paracrine operator instead of a systemic "master regulator." In the words of its discoverer: “although PACAP was a hypophysiotropic peptide, it was dissimilar to all other hormones” [15].
Its long evolutionary history also argues that it represents an ancestral form of local regulation with powerful local effects and an array of outcomes. PACAP is a ligand for GPCRs, which themselves evolved even earlier, in paleozoic archaebacteria and early eukaryotes. The extant GPCRs in multicellular organisms trace their origin from unicellular amoeboid eukaryotes with more than 700 hundred million years of history where cyclic-AMP (cAMP) and its GPCR receptor served as a stress signal [19]. The PACAP receptor family (part of the Secretin family) arose from the Adhesion receptor branch that regulate close contact local interactions [20,21]. The ancestral features of the PACAP system, i.e., stress signaling, cAMP pathways, and close contact local regulation, are still parts of its repertoire and represent important clues for understanding its physiology [22,23,24,25,26,27].
The ligand for the PACAP GPCRs, the preproPACAP gene (ADCYAP1) evolved ca. 600 million years ago in Tunicates (Urochordates), which are the ancestors of all vertebrates [132]. PACAP emerged along with other vertebrate landmark features including the notochord and migratory neural crest cells, which further suggests a critical PACAP function in vertebrate biology [133]. Additional exon and gene duplications resulted in several paralogs in the PACAP/glucagon/secretin family that today are present on different chromosomes, but PACAP was the ancestral molecule [14]. GenBank searches of PACAP paralogs (see Figure 1) identified that the closest related hormone is the vasoactive intestinal peptide (VIP) with 19 identical residues compared to the 27 amino acid sequence of the short PACAP isoform, PACAP1-27 (70.4% homology). The other members include the peptide histidine isoleucineamide (PHI) with 11/27 identical residues (40.7% homology), both secretin and GHRH (growth hormone releasing hormone) with 37% homology, glucagon with 8/27 identity (29.6% homology), and PRP (PACAP Related Peptide) with 6/27 identical residues (22.2% homology) [134].
In contrast, comparison of PACAP orthologs in various vertebrate species demonstrated an astonishing level of conservation, as shown in Figure 2. Not a single substitution occurred since mammalian branching ca. 80–100 million years ago. Only a single residue changed since the mammalian-reptile split ca. 250 million years ago (97% conservation) and only four positions were replaced since the urochordate-vertebrate separation ca. 560 million years ago (~80% conservation), which indicates enormous evolutionary pressure and a crucial PACAP function that has yet to be fully understood [10,134,135].

3. PACAP Receptors and Signal Transduction Bias

One of the surprising functional characteristics of PACAP is a wide spectrum of outcomes including potentially opposing effects [136]. This reported signaling bias is based on its complex pathways through multiple receptors, multiple receptor isoforms, and multiple signal transduction mechanisms.
There are three major and several minor receptors that have been shown to bind PACAP isoforms and initiate a variety of intracellular signals. PAC1-R is a Type I receptor pharmacologically, with high affinity for PACAP, but low binding constant for VIP. Type IA receptors bind both PACAP1-27 and PACAP1-38, while Type IB receptors bind only PACAP1-38 with high affinity [137]. Although Type IB receptors have not been studied extensively, the 2018 IUPHAR/BPS database, a comprehensive database for pharmacological receptors [138], posits alternative splicing and G protein subunit/second messenger mechanisms in its specificity. The VPAC1-R and VPAC2-R receptors are Type II receptors with the same high affinities for both PACAP and VIP, comparable with the PAC1 receptor. PAC1-R was reported to display signal transduction bias through several signal pathways, while the Type II receptors use almost exclusively cAMP-PKA activation [14,139]. In addition, the secretin receptor also binds both PACAP and VIP with equal affinity and stimulates the same biological effect to the same extent as secretin, but the latter shows higher affinity [140].
The reported signal transduction bias of the PAC1-R receptor is governed by several splicing variations. The third intracellular domain (ICL3) can integrate three extra mini-exons (HIP, HOP1, and HOP2, a shorter variant of HOP1) in six splicing variations in rat (HIP, HOP1, HOP2, HIPHOP1, HIPHOP2, and the Null-variation) [5,141]. In humans, only HIP and HOP1 was detected with four splice variants [142]. This third variable intracellular domain is the binding site for a variety of G-protein α subunits. Depending on the splicing structure, different Gα isoforms can bind and initiate alternative signal cascades with bias for the adenylate cyclase-cAMP-PKA pathway or polyphosphoinositide-specific phospholipase C-dependent signaling. Another receptor variation in the transmembrane domain II and IV (PACAPR-TM4) activates an L-type calcium channel and calcium signaling [143]. Considering the lack of follow-up and mechanistic studies to explain the robust changes by a single point mutation, the validity of this variant needs further clarifications. In addition, PAC1-R activation can also lead to β-arrestin–mediated receptor internalization and endosomal signaling, which results in sustained Mitogen-activated protein kinase/MAPK/ERK pathway activation [144,145]. Furthermore, a VPAC1/RAMP2 (Receptor Activity Modifying Protein) complex in vitro shows phosphoinositide hydrolysis (unique in Type II receptors), with no change in cyclic AMP stimulation [146]. Other rare intramembrane, extracellular, and intracellular loop variations were also reported, which further increased the pharmacological and signal transduction complexities of PACAP receptors [147,148].
Altogether, more than twenty PACAP receptor variants have been identified in vertebrates [8]. Although these were detected in cell lines at first, several of them have also been verified in vivo. They show differential regulation in development indicating functional significance. [12,13]. The isoforms affect ligand-binding selectivities and may control the signaling bias. In contrast to the large number of receptors and variants, however, only a few signal pathways are activated. However, as the coreceptor composition and signal patterns may be tissue-specific representing tissue-specific transcriptional programs, a few signal pathways may still orchestrate tissue-specific responses. Nevertheless, differential expression of PAC1-R isoforms has been reported that probably contributes to the differences in action exerted by the peptides.

4. When PACAP1-38 Hurts

4.1. Involvement in a Migraine Attack

All primary headaches are considered severe neurological conditions causing poor quality of life and disability. Extensive investigations during the last two decades shifted the view of migraine pathophysiology from vascular mechanisms toward complex neuronal theories [149]. It has been suggested lately that migraine arises from inside the central nervous system and the vascular changes are only an epiphenomenon instead of the underlying mechanism. In our current view, a migraine could be considered as epilepsy of the sensory networks including the trigeminovascular system responsible for the throbbing headache. The central neural components that can trigger the dural-trigeminovascular reaction include multiple brain regions from the medulla up to cortical areas. The list in ascending order is astonishingly complex and includes the trigeminocervical complex, the rostro-ventral medulla, the dorsolateral pons, the peri-aqueductal grey matter, locus coeruleus, and superior salivatory nucleus. These areas are thought to be migraine generators that are modulated by diencephalic areas including hypothalamic as well as thalamic nuclei. Dural trigeminovascular activation, which is evidently causing the symptoms, is induced through the peripheral sphenopalatine or the trigeminal ganglions [149]. Ultimately, their fibers innervate the cranial vessels and release substances (i.e. PACAP1-38, VIP, calcitonin gene-related peptide (CGRP), Substance P, NO), which, in turn, evoke vasodilation, neurogenic inflammation, and mast cell degranulation [150,151,152].
A preponderance of evidence emerged in the last decade that PACAP1-38 is one of the potential inducers of migraine. Functional studies demonstrated that 10 pmol/kg/min PACAP1-38 infusion evoked headache and migraine-like attacks, in both healthy patients and migraineurs [102,103] and focused great attention on PACAP1-38. It is noteworthy that, since these pioneering studies, five clinical trials have been initiated using PACAP(s) as challenge agents to test drugs (e.g. Imigran, Sumatripan) in migraine patients.
PACAP1-38 has been established in practically all aspects of trigeminovascular activation: (i) vasodilation of dural blood vessels [101,103,104], (ii) excitation of the central migraine generator [100], and (iii) mast cell degranulation [153]. One of the first actions attributed to PACAP1-38 after its discovery was a remarkable vasodilator effect [154,155,156]. The expression of all three PACAP1-38 receptors was reported in cranial/cerebral arteries in rats and humans [157,158,159]. The effect of blood-borne PACAP1-38 was a selective and long-term dilation of the external carotid artery and its branches (middle meningeal artery, superficial temporal artery) while the internal carotid artery and its middle cerebral branch remained unaffected. To date, the mechanism of the evoked vasodilation continues to be debated. Although circulating PACAP1-38 can pass the blood-brain barrier (BBB) to a certain extent via protein-transport systems (protein-transport system-6) [160], it is unlikely that PACAP1-38 could penetrate the walls of the above-mentioned big arteries to reach smooth muscle PAC1/VPAC2 receptors. Rather, it could be taken up via capillaries acting from inside out. Therefore, new considerations must be given to some central actions of PACAP1-38 toward inducing migraine. This is supported by the fact that all types of PACAP receptors were expressed by neurons of the trigeminocervical complex and the trigeminal and sphenopalatine ganglia [105,106,161]. Akerman [100] and colleagues’ elegant study showed that PACAP1-38 first evoked vasorelaxation of meningeal vessels through VPAC2 receptors, which is followed by an increase of spontaneous firing of trigeminocervical neurons after a 1.5-hour delay. Furthermore, the sensitivity of these neurons for intracranial dural stimulation also significantly increased. The delayed activation of the trigeminocervical complex was mediated through PAC1 as well as VPAC1-R. These landmark studies established a fundamental role of PACAP1-38 in the etiology of migraine.
PACAP1-38 is produced in the central nervous system in both neural and non-neural sources. PACAP1-38 expressing neurons have been identified in trigeminal and spheno-palatine ganglia that innervate extra-cranial and intracranial vascular structures in the head [157,162]. Furthermore, human mast cells were also reported to release PACAP1-38 [163]. Their role in migraine pathophysiology was corroborated earlier by studies reporting that histamine and prostaglandin I2 secreted by mast cells caused an immediate headache and migraine-like attacks [164]. Interestingly, in addition to PACAP1-38 expression, these special secretory immune cells also express PAC1 and VPAC1 receptors in a resting state and the VPAC2 receptor upon activation [165]. PACAP1-38 binding to its receptors causes degranulation of mast cells and induces an array of actions including skin edema or even dural vessel dilation [150,166,167]. The results showed that administration of PACAP6-38 (a potent mast cell degranulating agent through VPAC2 [168]) had no immediate effect on dural vessel diameter but exerted a delayed yet prolonged vasodilation. This result suggests that mast cells partially contribute to the vascular events prior to migraine, which implicates mast cells in the patho-mechanisms of the disease.
Mast cells, evidently involved in inflammation and an allergic reaction [169,170], also play a role in neurogenic inflammation. Sensory afferent nerve endings including those of the trigeminal nerve, release neuromodulators: calcitonin gene-related peptide (CGRP), substance P (SP), and PACAP1-38. Each of them has been investigated in both migraine and related neurogenic inflammation research, with mixed results [171]. CGRP did not seem to be involved in neurogenic inflammation and it induced neither mast cell degranulation in humans nor plasma protein extravasation [172,173]. Nevertheless, antagonists for CGRP (gepants) have been in the spotlight of pharmaceutical companies seeking migraine therapies. In contrast, SP has a strong degranulating effect on mast cells and is, thus, clearly involved in neurogenic inflammation. However, its failed SP receptor antagonist clinical trials for migraine [174] ruled out any roles in migraine pathophysiology. As discussed above, PACAP1-38 is implicated in neurogenic inflammation in the periphery and in migraine, but its role in dural neurogenic inflammation has yet to be investigated. Nevertheless, PACAP1-38 has a high impact on inflammatory processes in other tissues, which will be discussed later.
Considering the established negative effects of PACAP, systemic PACAP1-38 infusions may have side effects on the cardiovascular system. Involvement of PACAPs in the renin/angiotensin system and the presence of PACAP receptors in the coronary arteries [175,176] point to increased cardiovascular risk upon systemic PACAP treatment. While examining migraine patients, Birk et al. [102] reported that a severe increase in heart rate prevented the use of higher than 15 pmol/kg/min dosage of PACAP1-38 in humans. Seeliger and colleagues did not describe differences in heart rate or blood pressure upon 100 pmol/kg/h PACAP1-27 infusion but reported a significant increase of body temperature and persisting, six-hour long erythema suggesting that higher systemic doses should not be used in human in vivo experiments [177]. Unfortunately, their report did not present any data whether this low PACAP concentration induced a migraine attack or headache, and the Birk team did not mention any sign of skin edema or flushing either, which further underlined the need for standardized reporting criteria in the field. Furthermore, a clinical trial aimed at determining the maximum human dose would be essential and urgently needed.
A current clinical trial (identifier: NCT03238781) [178] aims to evaluate the effect of a PAC1-R antibody in migraine patients. It is important to point out that, according our current knowledge, the migraine-inducing effect of PACAP1-38 is an inside-out action rather than a direct vasodilatory effect, and the passage of antibodies through the BBB is poor. Nevertheless, the antibody could block trigeminal PAC1 receptors since the trigeminal ganglion is located outside of the BBB. Therefore, a systemic antibody approach is feasible [179].

4.2. Contribution to Post-Traumatic Stress Disorder (PTSD)

Although not life-threatening, PTSD is a severe, highly disabling psychiatric disorder and a large economic burden on the patients and on society at large. The symptoms include flashbacks, nightmares, severe anxiety, intrusive thoughts, arousal, and reactive symptoms that completely disrupt the daily lives of the afflicted.
Increased CRH (corticotropin releasing hormone) and ACTH levels in these patients prompted studies that proved the involvement of PACAP1-38 in the long-term stress-axis and a crucial role in the stress response [180]. PACAP1-38 was documented to play an essential role in the emergency response of the autonomic nervous system by stimulating catecholamine secretion from the adrenal medulla through sympatho-adrenal fibers [27,181]. The stress-response related brain regions (i.e., bed nucleus of stria terminalis—BNST, amygdala complex, hippocampus, and medial prefrontal cortex) were shown to express both PACAP1-38 and PAC1-R suggesting a role in stress, anxiety, and fear-related learning [182,183,184,185,186] through PACAP1-38-containing fibers that innervated CRH-expressing neurons in the paraventricular nucleus and BNST [187]. Second, PACAP1-38 was reported to upregulate CRH expression and production [188,189]. Third, although the acute stress response is PACAP1-38-independent, in the sustained stress response (with CRH release) PACAP1-38 signaling is essential [190]. To study anxiety levels, PACAP knockout mice were used. Through an initial screen for anxiety-related behavior, (i.e., open-field test, elevated plus maze, novel-object test) PACAP1-38 KO mice showed a lack of fear, hyperactivity, and increased exploration, which are related to reduced anxiety [191,192]. Another set of experiments aimed to investigate the effect of PACAP1-38 on psychotic behavior. Injecting PACAP1-38 into the paraventricular nucleus, central amygdala or bed nucleus of stria terminalis (BNST) resulted in significantly suppressed exploratory activities, and increased withdrawal coupled with immobility and enhanced startle behavior, respectively [186,187,193]. The findings point to PACAP1-38 contributions to anxiety-like responses through BNST circuits whereas its role in fear manifestation is linked to PAC1-R expression in the amygdala [194,195].
The processes of fear-learning, creating fear-memories and recalling fear have a tremendous impact on the development of an anxiety disorder. Extensive research of the last decade revealed more details about the function of the PACAP system with enhanced fear memories, which could lead to anxiety disorders like PTSD [196,197]. PACAP1-38 was reported to induce both short-term and long-term synaptic plasticity by regulating pre-synaptic and post-synaptic components [198]. For example, expression of activity-regulated cytoskeleton-associated protein (Arc/Arg3.1) improved memory performance in rodents [89,199]. In fact, PACAP1-38-induced memory formation also includes conditioned fear. PACAP1-38 exerts acute and chronic effects on the consolidation process of fear memory in the BNST, amygdala, or prefrontal cortex [199,200,201]. It is noteworthy that spatial memory performance was not affected by PACAP1-38 [200]. Mediation of these effects involves the PAC1-R receptor as an injection of PACAP6-38 or maxadilan into the BNST, which altered the stress-induced behavioral responses in cued fear conditioning [202]. Nonetheless, observations obtained on VPAC2 knockout mice suggest that this receptor might also be involved in fear memory formation. By testing Pavlovian fear conditioning, the VPAC2 knockout mice exhibited normal acquisition but aggravated extinction in terms of contextual and cued fear memory [203].
Apart from animal models, Ressler et al. [4] reported that a SNP, rs2267735 linked to the PAC1-R gene, predicted the development of PTSD in traumatized women. Interestingly, rs2267735 SNP maps the estrogen-response element of the ADCYAP1R1 gene, which indicates that PAC1-R expression is probably estrogen-dependent. Since this breakthrough study, this particular SNP has been confirmed to correlate with (i) increased amygdala reactivity, (ii) increased hippocampus reactivity, and (iii) reduced interaction between the two prominent brain regions [204]. A meta-analysis performed by Lindert et al. [205] also revealed that rs2267735 may increase the risk for PTSD.
Since the ancient PACAP system forms a bridge between the neuro-endocrine CRH/ACTH axis and brain areas in charge of emotional memory, it also links/mediates their pathologies. PACAP1-38 signaling through PAC1-R plays a key role in behavioral reactions to stress. Therefore, any imbalance of the signaling system could be responsible for the abnormal, sustained stress response called PTSD.

5. When PACAP1-38 Rescues

5.1. Alzheimer’s Disease and PACAP1-38

Neurodegenerative diseases are a large and diverse group of incurable and debilitating conditions with progressive neuronal degeneration including Alzheimer’s disease, Parkinson’s disease, prion disease, Huntington’s disease, and spinocerebellar ataxia. Alzheimer’s disease is associated with dementia caused by slow and gradual death of nerve cells. At present, existing medications for Alzheimer’s diseases are very limited, and only treat the symptoms, rather than addressing the cause [206].
An early pioneering study reported that PACAP1-38 protects PC12 cells against β-amyloid-induced toxicity by suppressing the apoptotic machinery [207]. On cultured cortical neurons, Han et al. [83] reported that PACAP1-38 enhanced Sirtuin-3 production, which can protect stress-induced mitochondrial integrity and energy metabolism. PACAP peptides were also shown to increase the activity but not the expression of α-secretase enzymes in HEK 293 cells [208]. Amyloid precursor protein (APP) is processed by α-secretase. The cleavage results in the amyloid-β peptide and the release of the soluble N-terminal APP fragments (APPsα). Secretion of APPsα has proliferative, anti-apoptotic, and neuroprotective effects on neurons and prevents the deposition of β-amyloid peptide in plaques [209]. After intranasal administration of PACAP1-38 for three months in AD transgenic mice, APP processing shifted toward the non-amyloidogenic pathway and, in addition, brain-derived neurotrophic factor (BDNF) expression also increased as another beneficial effect of PACAP1-38 inhalation. Most importantly, the mice exhibited improved cognitive functions [84]. It would have been interesting to observe the anxiety levels of the animals (for explanation, see the PTSD section). Dogrukol-Ak et al. [210] achieved promising results by working out an alternative route to deliver PACAP1-38 into the central nervous system. They developed antisense compounds to selectively block the PTS-6 efflux pump that resulted in the accumulation of PACAP1-38 and 1-27. The increase in PACAP1-38 in such a way improved the learning ability of Alzheimer’s AD transgenic mice. They used the T-maze foot shock avoidance test to examine learning skills in this study. However, using a contextual fear memory paradigm is not an appropriate tool to study the association of PACAP1-38 with Alzheimer’s, even though the model is correct (for explanation, see the PTSD section).
In several studies, PACAP1-38 gene expression, protein levels, and the PAC1-R receptor were downregulated in both mouse models and postmortem human brain samples, which indicates correlations between PACAP1-38 levels and disease progression [82,83,211]. It seems that the decline of the PACAP system negatively correlates with plaque density and the severity of dementia. However, these results have provided only suggestions for the PACAP1-38 contribution to the pathogenesis. To clarify whether the malfunction of the PACAP/PAC1-R system is the reason or the consequence of Alzheimer’s disease, further experiments are warranted.

5.2. Ischemia/Stroke

Significant efforts of PACAP research are dedicated to investigate ischemic injuries. Ischemia is defined as an inadequate blood supply to a given tissue/organ due to blockage of blood vessels, which, in turn, leads to various levels of tissue destruction depending on the conditions. Local disturbance of the bloodstream can be caused by vasoconstriction, embolism, or thrombosis. Hypoxia is not the only challenge. The subsequent reperfusion-induced generation of reactive oxygen species (ROS) is an additional problem [212]. We will summarize potential PACAP1-38 therapy options for ischemic renal, cardiac, and hepatic injuries and review the results from two important areas including the brain and retinal ischemia.
PACAP1-38 has been shown to have a powerful vasodilator effect on cardiac vessels and the demonstrated PAC1-R expression in cardiac myocytes established the foundations for further studies to investigate the protective potential of PACAP1-38 in ischemic conditions [213,214]. A handful of publications showed that PACAP1-38 protected the cells in in vitro ischemia models [215,216] or in oxidative challenge [217]. A couple of papers have been addressed to the PACAP1-38 effect on hepatic ischemic damage as well. It has been established that all three PACAP receptors were upregulated during ischemia. Administration of PACAP1-38 in vivo unquestionably proved that PACAP1-38 ameliorated ischemic injury through multiple pathways: by inhibiting macrophage sequestration, suppressing pro-inflammatory cytokine release, and by attenuating apoptosis [218].
Investigations of PACAP knockout mice in cardiac, hepatic, or renal ischemia/reperfusion provided similar conclusions. The absence of endogenously produced PACAP1-38 resulted in increased vulnerability in ischemia and oxidative stress [219,220,221]. The results indicated a protective role for the endogenous PACAP system in pathophysiological conditions.
The following experiments provided a better understanding of the mechanisms in ischemic pathophysiology and the mechanisms of PACAP1-38 protection. In the heart, Alston et al. elucidated an interesting interaction between PACAP1-38 and locally-released cytokines. Elevated transcription of PACAP1-38 along with VIP in the heart as well as in the stellate ganglion after myocardial infarction was mediated by cytokine-activated gp130 signaling [219]. Those cytokines that activate gp130 (e.g., interleukin-6, corticotrophin-1) are considered as important survival factors for cardiac myocytes [222,223]. Therefore, it is reasonable to assume that the protective effect of the above-mentioned cytokines might be mediated by PACAP1-38. The attenuating effect of PACAP1-38 in the kidney is attributed to the suppression of Toll-like receptor 4/MyD88 signaling, known to be responsible for ischemia-induced inflammation. The protective effect was observed even if PACAP1-38 delivery was delayed by 24 hours [224,225,226]. Liu and colleagues [94] provided a new and curious insight by demonstrating that the protective effect of PACAP1-38 in the liver is accomplished through induction of the Yes-associated protein, which is an oncoprotein important in cell proliferation and regeneration. This protein is elevated in hepatocellular carcinoma [227]. This phenomenon (protecting the liver by inducing a cancer protein) is a fascinating example of the ambiguous functions of PACAP1-38 that will be discussed later.
The ischemia-generated combined anoxia and hypoglycemia develop more severe conditions in neural tissues because the resulting apoptosis and oxidative stress are exacerbated by glutamate-induced excitotoxicity [228,229], with serious consequences [230,231]. Furthermore, neuronal tissues are much more susceptible to ischemic insults due to their poor regenerative capacities. Seeking therapeutic agents, therefore, is a priority in this field. In the present review, we give a general summary of ischemia in the central nervous system and discuss the details of retinal ischemic injury as a well-studied field within the central nervous system.
The retina is an outpost of the central nervous system, but it is exposed to higher risk of ischemia, not only by thrombosis or embolism but through other pathologies including diabetic retinopathy or glaucoma [232]. In both the bilateral common carotid artery occlusion model (BCCAO) and the intraocular hypertension ischemia/reperfusion model, dramatic cell loss was observed and reported. The morphometric analysis revealed that the total thickness of the BCCAO retina appeared to be half of the control retinas caused by severe cell death in the ganglion as well as inner nuclear layer three weeks post-operation [230,233]. Ganglion cell death was prevented, however, by either 10 fM or 10 pM intravitreally injected PACAP1-38, which indicated two distinct PACAP receptors with different affinities. The action of 10 fM PACAP1-38 was inhibited by a MAP kinase inhibitor, which suggested that the higher affinity PACAP receptor signaled through the MAPK/Erk pathway [234]. The results were confirmed by maxadilan, a PAC1-R agonist that exerted the same protective effect in BCCAO retinas, whereas the inhibitor PACAP6-38 prevented it [233,235]. Furthermore, Szabo and coworkers [236] showed that the underlying mechanism of PACAP1-38 protection markedly suppressed inflammatory cytokines, e.g., CNTF (ciliary neurotrophic factor), fractalkine, sICAM (soluble inter-cellular adhesion molecule-1), and IL-1 in the injured retina, similarly to other tissues. Moreover, PACAP1-38 treatment not only inhibits cell loss but enables the functionality of retinal circuitries [91]. An accurate and comprehensive metabolomics analysis characterizing the cell state and metabolic behavior of retinal tissue during the ischemic insult was conducted by D’Alessandro and colleagues [237]. Concomitant treatment with PACAP1-38 had strong and complex effects on the upside-down metabolic state induced by ischemia. First, PACAP1-38 exerted a multi-faceted blockade against glutamate-induced excitotoxicity by reducing glutamate, glutamine, and α-ketoglutarate levels. Second, it could fight oxidative stress, not only by decreasing pro-inflammatory factors, but also normalizing glutathione levels, which is an important free radical scavenger. Third, PACAP1-38 was able to redirect the metabolic flux from the pentose phosphate pathway toward glycolysis, which is a shift often observed in cells undergoing oxidative stress [238]. However, the peptide failed to restore energy metabolism and adenosine tri-phosphate (ATP) production [237]. Fourth, another alternative route for PACAP1-38 to cope with the ischemic challenge is to activate non-neural cell components in the retina, namely, microglia and macrophages. These cells produce anti-inflammatory interleukin-10, which may also contribute to the protective action of PACAP1-38 [239].
In addition to the above retinal ischemia observations, a result from studying brain ischemia further expanded our knowledge on PACAP1-38 functions: inhibition of caspase-3 activity [92] and inhibition of proinflammatory cytokine release [240,241]. Studies on microglia regulation by PACAP1-38 have revealed additional mechanisms in cerebral ischemia. First, PACAP1-38 seems to protect microglia in ischemia by inhibiting a microglial Toll-like receptor, MyD88 and by blocking the tumor necrosis factor α expression and NO production [240,241]. This finding might explain the increased cell number described in the retina by Wada et al. [239]. PACAP1-38 was shown to decrease proinflammatory mediators and, thereby, to resolve inflammatory processes, but also to induce the M2 phenotype transformation of microglia cells [90]. M2 is a preferred phenotype of microglia in neurodegenerative diseases since their presence is associated with regenerative growth, anti-inflammation, tissue repair, extracellular matrix reconstruction, or uptake of apoptotic cells [242]. This phenomenon resembles the role of PACAP1-38 in T-helper1/2 shift in peripheral inflammatory response that will be discussed in the next section. Furthermore, studying brain ischemia revealed that PACAP1-38 affects apurinic/apyrimidinic endonuclease 1/redox factor-1 (APE1) expression and function, which is an essential component of both DNA repair and redox signaling [243]. In addition, these studies also led to another important recognition. PACAP1-38 is a potent inducer of neural progenitors via the PAC1-R in the adult brain [244]. Other studies also support this observation. Under ischemia, the subventricular zone of the hippocampus contained significantly fewer proliferating progenitor cells in PACAP heterozygous knockout mice than in the normal control [95].
The role of VPAC receptors in cardiovascular side effects appears to be rather controversial. On the one hand, VPAC receptors could induce VEGF-mediated endothelial cell proliferation in vitro [245], while, on the other, VPAC2 receptors were reported to increase the severity of ischemic damage, hemorrhage, and, thus, the mortality of type-2 diabetic Goto-Kaizaki rats [246]. It is important to point out, however, that the diabetic condition was coupled with endothelial and vascular malfunctions [247]. Consequently, to elucidate the effects of VPAC2 receptor stimulation in ischemic injuries, this condition should be further investigated in non-diabetic models.
In addition to the direct effects mediated by PACAP receptors, PACAP also acts indirectly through the induction of neurotrophin release (e.g., BDNF), activation of trkB receptors, and attenuation of neuronal growth inhibitory signaling molecules p75NTR and the Nogo receptor [92,248].
In sum, PACAP1-38 could be an effective therapeutic agent for ischemic pathologies in both direct and indirect ways, and with both concomitant and delayed administration. The latter property could confer the peptide powerful potential in the recovery process following ischemic insults. The reported development of a promising new PACAP analog with higher resistance to dipeptidyl peptidase IV and, most importantly, with reduced cardiovascular effects in animal models offers a new avenue toward therapeutic applications [121].

6. When PACAP1-38 Appears to be Janus-Faced

6.1. Role of PACAP1-38 in Inflammation

There are two sets of available experimental data about the role of the PACAP system in inflammation, one confers on PACAP1-38 a broad range of anti-inflammatory effects and the other proves that PACAP1-38 is responsible for inflammation. According to the data comprehensively reviewed by Gomariz et al. [249], VIP/PACAP1-38 exerted purely anti-inflammatory effects on multiple sites of the mammalian immune system. Since the peptides are expressed in the same immune cell populations and target the same effectors, they are often referred to as the VIP-PACAP system sharing overlapping effects. The expression of their receptors, however, show differences. PAC1-Rs are borne exclusively by macrophages [250] whereas VPAC1-Rs are expressed by all immune cells (i.e., macrophages, lymphocytes, monocytes) [251,252,253]. The VPAC2 receptor was inducible in lymphocytes as well as in macrophages upon activation in rat [165], but its expression was constitutive by human lymphocytes [254]. Overall, both peptides regulate cytokine/chemokine production, predominantly by decreasing inflammatory factor expression [249]. They suppress mobility/migration by downregulating adhesion molecules and inhibiting infiltration of neutrophils [255]. Furthermore, VIP/PACAP1-38 control T-cell differentiation by pushing T-lymphocytes toward the T-helper2 fate [256], which is thought to foster anti-inflammatory signals [257]. By preventing T-helper1 cell differentiation and suppression of inflammatory interleukin production, PACAP1-38 treatment resulted in an improvement of multiple sclerosis symptoms [115].
Apart from studies where exogenously administered PACAP1-38 effects were examined, genetic modification of PACAP or PAC1-R genes also earned us valuable information about altered immune functions [258]. In PACAP knockout mice, airway lipopolysaccharide (LPS)-induced inflammation, dextran sodium sulfate-induced colitis, and multiple sclerosis were all markedly aggravated [117]. Likewise, PAC1 and VPAC2 knockout mice exhibited worsening symptoms of LPS-induced endotoxemia and delayed-type hypersensitivity, respectively [259,260]. The common motif observed in all of these immune diseases was enhanced inflammation. However, Jongsma and co-workers reported [261] that PAC1 knockout mice showed an immediate increased reaction to noxious stimuli (i.e., formalin) but appeared to be less sensitive later, which suggests that PAC1 signaling has a biphasic effect in pain transmission. More importantly, the observed decrease in the late phase response indicates that PACAP1-38 is involved in inflammation-induced nociception.
It is important to point out that another set of reports presented evidence that the PACAP system contributes to inflammatory mechanisms. First, PACAP1-38 and its receptors are ubiquitously expressed in the neurons of dorsal root ganglia (sensory afferents), as well as in the spinal horn [262,263,264,265]. Elevated expression of PACAP1-38 in the sensory neurons in inflammatory circumstances implicated PACAP in neurogenic inflammation [64,266,267]. In neurogenic inflammation, sensory fibers release a large set of neuropeptides including calcitonin gene-related peptide (CGRP)/substance P (SP), PACAP1-38/VIP, and neuropeptide Y, which, in turn, cause various inflammatory responses (i.e., mast cells degranulation, vasodilation, plasma protein extravasation) [151,152,166]. The first functional report indicating that PACAP was a potent edema inducer was published by Warren and colleagues [77]. Both intradermal (10−8 M up to 10−6 M) and intravenous (7.5, 15, 30, and 100 pmol/kg/min) application of PACAP1-38 induced concentration-dependent skin edema and erythema in humans [177]. Significantly smaller edema was developed in PACAP knockout animals in neurogenic inflammation. Nevertheless, lacking PACAP1-38 did not influence the symptoms of non-neurogenic inflammation [76]. Investigation of ocular inflammation evoked by either electroconvulsive or endotoxin treatments resulted in the same conclusion. Neuropeptide release from the trigeminal sensory fibers caused the development of neurogenic inflammation. PACAP1-38 induced conjunctival hyperemia and protein extravasation was diminished by L-NAME, which is a nitric oxide synthase inhibitor. The finding demonstrates that NO not only mediates the effect of PACAP1-38 but, in turn, NO effectively induces neuropeptide release from nerve endings [69,268]. The results of a recent paper raise a novel possibility of applying PACAP1-38 eye drops in the dry eye syndrome, since the peptide potently increased lacrimation [269]. In retinal ischemic injury, external application of PACAP1-38 also seemed to be beneficial [99]. Although these reports did not indicate any signs of ocular inflammation upon treatments, reports of PACAP involvement in the pathogenesis of ocular inflammation warrant careful considerations. Chronic urinary inflammation was also accompanied by increased PACAP1-38 expression in both the dorsal root ganglia and the spinal cord. Moreover, both the immunoreactivity levels and the number of PACAP1-38 immunoreactive cells increased [64].
Altogether, PACAP1-38 appears to be involved in three types of inflammations: autoimmune-induced inflammation (e.g., rheumatoid arthritis, collagen-induced arthritis, Crohn’s disease), infection-induced inflammation, and neurogenic inflammation. In respect of the first two groups, PACAP1-38 has a remarkable attenuating effect, whereas it contributes to the pathophysiology of the latter.

6.2. Roles of PACAP in Stem Cell Regulation and Cancer Formation

Stem cells first evolved at the dawn of multicellularity where PACAP had important roles. Its adhesion receptor heritage and paracrine regulation appear to be central to maintain the stem cell niche. Stem/progenitor generating germinal zones of the brain are known as neurogenic niches including the subventricular zone of the lateral ventricles and the subgranular zone of the dentate gyrus. In this case, the extracellular matrix (ECM) and soluble factors maintain “stemness,” which involves self-renewal and multipotency [270,271]. PACAP was shown to upregulate ECM components and ECM-modifying enzymes in adult neural progenitors (aNPCs) that increased their surface adhesion in their niche [26]. These progenitors migrate in the rostral migratory stream (RMS) from the apical subventricular zone (SVZa) to the olfactory bulb in infant mice. PAC1-R co-localized with sites where progenitor adhesion was critical, which includes the progenitor niche (SVZa) and at their destination, but not in the migration process (in the olfactory tract) [272]. PACAP-regulated adhesion is important in stem cell homing as well. PACAP signaling was proved to induce bone marrow-derived stem cells to colonize into ischemic brains [273]. PACAP was also shown to promote neuroblast self-renewal [274] and to protect adult neural progenitors from stress-induced apoptosis [275]. This role, however, also has a Janus-face since experimental evidence demonstrated both pro-apoptotic and anti-apoptotic functions of PACAP1-38 in the postnatal mammalian retina [276].
Furthermore, PACAP1-38 promotes progenitor expansion by supporting proliferation. There are several reports that the PACAP/PAC1-R autocrine system is critical to extend the adult neural progenitor pool in the brain [244] and that appears to be the first step in astrocyte differentiation [277]. Not surprisingly, PACAP also has a role in retinal ontogenesis. It controls progenitor cell proliferation together with Kruppel-like factor 4 [278]. PACAP has a similar effect on olfactory neurogenesis. It promotes proliferation of the basal cell layer as well as the survival of the immature and mature neuronal layers [279]. Moreover, this role extends beyond the central nervous system. PACAP was reported to contribute to the proliferation of hematopoietic progenitor cells in murine bone marrow [280] and promotes proliferation of murine primordial germ cells as well [281].
PACAP also shows its Janus-face at the level of stem cell exit, by initiating stem/progenitor differentiation. PACAP was reported to both promote and block differentiation, depending on the stem cell niche and developmental stage. Astrocyte development, for instance, is induced from neural progenitors, but only after these cells have already generated neurons. This gliogenic switch is regulated by intrinsic and extrinsic factors and one of these positive differentiation factors is the PACAP/PAC1R system through the PACAP-cAMP-Ca(2+)-DREAM signaling cascade [282]. At different times during development, PACAP1-38 exerts different actions. It acts as an antimitotic factor on early neural progenitors directing neuronal differentiation, whereas, on late progenitors, it regulates the generation of oligodendrocytes [274,283]. Moreover, there is evidence that PACAP1-38 is also involved in the regulation of terminally differentiated astrocyte functions e.g., plasticity, glycogen synthesis, gliotransmitter production, etc. [284]. PACAP1-38 has been reported to promote survival, inhibit migration, and activate differentiation and neurite outgrowth of cerebellar granule cell precursors. In cerebellar neuroblasts, PACAP1-38 is a potent inhibitor of the mitochondrial apoptotic pathway [285]. The Lot1 transcription factor was reported as a critical mediator in the PACAP1-38/cyclic AMP differentiation pathway that negatively regulates cerebellar neuronal precursor proliferation [286]. Not surprisingly, in mice lacking pituitary adenylate cyclase-activating polypeptide, cerebellar development is seriously affected [287].
In contrast, PACAP1-38 counteracted hedgehog-dependent motor neuron production in mouse embryonic stem cells when the aNPCs were cultured in the absence of growth factors [288]. PACAP1-38 can also negatively regulate the proliferation of retinal progenitor cells through downregulation of cyclin D1 [289]. In the control of megakaryopoiesis and platelet production, PACAP1-38 and the vasoactive intestinal peptide (VIP) were reported to have a role through their common Gα-coupled receptor VPAC1. Treatment with PACAP1-38, VIP, or the adenylyl cyclase activator forskolin inhibits megakaryocyte differentiation, which results in thrombopathy and thrombocytopenia [290]. Injections with inhibitory anti-PACAP1-38 (PP1A4) or anti-VPAC1 (23A11) antibodies to block the pathway increased platelet numbers. The observation raised the possibility of clinical intervention for platelet recovery after myelosuppressive therapy [290].
Although modern cancer biology considers cancer as a Darwinian and adaptive tissue ecosystem [291], several lines of evidence indicate that most cancers are initiated and maintained by a single or very low number of cancer stem cells [292,293]. Mechanisms of stem cell regulation and dysregulation, therefore, are critical to understand cancer biology. The findings that PACAP1-38 regulates almost every aspect of stem cell physiology suggest that the PACAP1-38 system also contributes to cancer mechanisms. It is not surprising, therefore, that, in a lot of cancers of various pathologies, PACAP1-38 and its receptors are widely expressed and play opposing roles from activation to inhibition.
Studying PACAP and PACAP receptor mechanisms in cancer is in its infancy. In most cases, we can only describe some fundamental observations, in a proof of principle manner. Most cancer cells have greatly dysregulated transcriptional programs and, consequently, many peptides and receptors are atypically expressed.
However, it is not biologically insignificant. We have ample evidence that atypically expressed receptors for a growth factor are the very mechanisms of cancer initiation (e.g., Insulin-like growth factor receptor isoforms in lung cancer and HER2/neu in breast cancer). PACAP receptor overexpression in a stimulated environment with high neuroendocrine input may initiate dedifferentiation and cancer. For instance, prostate has a high neuroendocrine component and neuroendocrine prostate cancers are the most malignant. High PACAP or PACAP receptor levels in blood or urine may serve as future cancer markers.
In vitro studies detected PAC1-R binding sites in a number of cell lines from neural or neuroendocrine tumors. They were found in the human neuroblastoma NB-OK cell line [294,295], oligodendroglioma [296], and the neuroendocrine BON cells [297]. They were also present in non-neural cancer cell lines, like the medullary carcinoma 6/23 cell lines [298] and the rat pancreatic acinar AR4–2J cell line [299]. The Type II binding VPAC2-R gene was expressed in the hypothalamic GnRH neural cell line GT1–7 [300]. PACAP1-38 receptors have also been detected in adrenal pheochromocytoma PC12 cells [301] and adrenocortical NCI-H295 cells [302]. VPAC1-R mRNA was present in breast and intestinal cell lines, while both Type II receptors were expressed in neuroectodermal and pancreatic cell lines [28,303,304,305].
Apart from cultured in vitro systems, a long list of neoplastic tissues also expresses the receptors in vivo. Cancers from the lung [306] and colon [31] were positive for PACAP1-38 receptors. In particular, VPAC1-R was expressed in the mucosa and myenteric neurons. VPAC2 was detected in blood vessels, smooth muscles, and neuroendocrine cells. PAC1-R was found in myenteric neurons [34]. Type I receptors were found in breast cancer [307] and prostate cancer [46,308]. On the other hand, VPAC receptors were found in neural cancers, in human pituitary adenoma [309], and in brain glioma [310]. A large-scale study of human cancer tissues established that almost all human cancers express some forms of PACAP1-38 receptors and usually at higher levels than in normal tissues. Breast cancers (100% receptor incidence), prostate (100%), pancreas (65%), lung (58%), colon (96%), stomach (54%), liver (49%), urinary bladder (100%) carcinomas, lymphomas (58%), and meningiomas (100%) typically express the VPAC1-R. Leiomyomas express the VPAC2 receptor, whereas most paragangliomas, pheochromocytomas, and endometrial carcinomas express the PAC1-R [311].
The expression of PACAP itself has also been reported in human cancers including several neurological tumors. PACAP1-38 mRNA is expressed in most gliomas but only in 20% of meningiomas [296]. Significant PACAP1-38 expression was detected in human neuroblastomas [312,313], which suggests that it could also control neuroblastoma tumor cell proliferation. Most pituitary tumors show high PACAP1-38 expression [314]. Considering that cAMP is a mitogen in pituitary cells, PACAP1-38 may contribute to tumorigenesis [315]. In non-neurological tumors, PACAP1-38 is highly expressed in breast carcinoma [316] and in pancreatic carcinoma and pheochromocytoma [317]. PACAP1-38 has also been reported in ovarian tumors [318], prostate cancer cell lines [319], and cutaneous (Merkel) carcinoma [320].
The ubiquitous expression of the PACAP system in cancers strongly suggests that PACAP1-38 may play a role in the process. In fact, PACAP1-38 is showing its double face again with its capacity to both contribute and suppress malignancy. Astrocytomas show high expression of PACAP1-38 and its receptors. By selecting cell lines with increasing malignancy, PACAP1-38 levels did not change, but PAC1-R and VPAC1-R mRNA levels correlated with malignancy. Treatment by PACAP1-38 (10 pM) for three days resulted in increased proliferation that also correlated with malignancy. The results suggest that PACAP1-38 receptor expression and activity contribute to the malignancy and proliferative potential of astrocytomas [321]. PACAP1-38-27 stimulated colony formation in NCI-H838 cells, whereas the PAC1-R antagonist, PACAP6-38, reduced it, which indicates that PACAP1-38 promotes malignancy [322]. VPAC1 overexpression was linked with poor differentiation of colon cancer and epidermal growth factor receptor activation in cancer cells. In addition, VPAC1 overexpression in both blood vessels and macrophages in the tumor may also contribute to aggressive cancer [32]. An evolutionary stress-related effect of PACAP1-38, apoptosis protection also contributes to malignancy-promotion. Malignant peripheral nerve sheath tumors (MPNSTs) are sarcomas triggered by metabolic stress. Endogenous PACAP1-38-mediated signaling was reported to increase MPNST cell resistance to apoptosis from serum starvation and enhanced malignancy [323].
However, another set of reports demonstrated that PACAP1-38 actually suppresses malignancy. In the human glioblastoma cell line T98G, only the VPAC2 receptor subclass was expressed. VIP, PACAP1-38-27, and PACAP1-38-38 were potent and efficient inhibitors of cell proliferation, which established that PACAP1-38 is a tumor suppressor in this scenario [324]. The VIP peptide was also shown to inhibit human renal cell carcinoma proliferation [47]. Multiple myeloma is a plasma cell tumor in the bone marrow. PACAP1-38 suppressed myeloma-stimulated interleukin 6 (IL-6) secretion by the bone marrow stromal cells. By modifying the bone marrow milieu, PACAP1-38 suppressed myeloma cell growth indirectly [325]. By investigating the effect of PACAP1-38 on human retinoblastoma Y79 cell viability, 1–5 μM potently reduced retinoblastoma cell survival in a dose-dependent manner. In cognate peptidergic GPCR signaling, a 5-uM binding constant is typically not considered specific. The authors also ruled out PAC1 and VPAC receptors as mediators, which suggests novel PACAP1-38 pathways [48]. The significance of the finding is that, in paracrine operations, the local PACAP concentrations can be even higher. Therefore, it is not immaterial what signaling is active under these conditions. Horizontal cells are the cell-of-origin for retinoblastoma, which is a tumorous fatal childhood disease [326]. We found that intravitreal PACAP1-38 administration caused upregulation in PCNA protein levels in a stage-dependent manner and led to an increase in horizontal cells numbers, presumably via PAC1-R [49]. Our unpublished data shows that, prior to PCNA upregulation, PACAP1-38 downregulates Rb protein transcription [327]. The results clearly indicate that PACAP1-38 positively regulates horizontal cell numbers and postulate a novel possibility that PACAP signaling is critical in retinoblastoma.
Altogether, the data demonstrated the ubiquitous presence and activities of the PACAP system in both stem cells and cancer. PACAP was shown to affect every aspect of stem cell biology. It protects stem cells and promotes self-renewal not only in the neurogenic niche but in other niches as well. It promotes progenitor expansion by supporting stem/progenitor proliferation, but in the stem cell exit and differentiation, PACAP shows its double nature. Depending on the stage and tissue context, it can both promote and block differentiation.
The above functions and dysfunctions in stem cell regulation may be the link that associates PACAP to cancer. PACAP and its receptors appear to be expressed in various combinations in almost every cancer investigated, both in vitro in cell lines and in vivo in the cancer tissues. As far as its functions and biological effects are concerned, the confounding data point to both tumor suppression and tumor promotion.
This almost universal ambiguity of the PACAP system may relate to a fundamental feature of this regulation, which is an ancestral stress response with short-term local operations. For instance, most cells respond to damage by apoptosis. As we showed, PACAP protects from apoptosis in almost every context, but the outcomes can be radically different. Protecting damaged stem cells in the retina may heal and rejuvenate, but blocking apoptosis in damaged lung and colon progenitors may generate cancer. The data established the power and the versatility of the PACAP system, but in the field of PACAP research “this is not the end, not even the beginning of the end, but may be the end of the beginning.”

7. Conclusions

The data presented in this review demonstrate the extensive efforts to establish the role of the PACAP system in human biology and to take advantage of these peptides in therapeutic applications. However, despite this level of attention, no large-scale diagnostic or clinical applications are under way. The majority of clinical trials target only a few conditions (i.e., nephrotic syndrome, migraine, rosacea, cluster headache, major depression) and apply PACAP as a challenge agent, investigate its role to initiate disease, or test its involvement. We know only one ongoing clinical trial that attempts to test systemic injections of anti-PAC1-R as a potential tool to treat a migraine (identifier: NCT03238781) [178].
The reason is one of the inherent features of PACAP that it acts through a wide array of actions, which are both positive and negative. The unpredictable severity and number of side effects in both the nervous system and the peripheral organs argue against systemic application [328]. With respect to neurological/neurodegenerative disorders, investigations should focus on local administration. Several methodological approaches have been published that deserve consideration. For instance, several experiments support the idea of intranasal application, which have been already proven to be effective in treating both Alzheimer’s and Huntigton’s mouse models [84,89]. Transplantation of PACAP secreting stem cells has been applied successfully and appeared to be beneficial in the ischemic brain [90]. Specific inhibition of the PTS-6 transporter and consequent blocking of PACAP efflux through the blood-brain barrier is another option for taking advantage of the neuroprotective actions of PACAP [210].
Regardless, the take home message of this review is a warning. The available data strongly suggest that at any applications of PACAP treatments, the duration and concentrations of PACAP agents must be carefully considered to avoid unwanted consequences such as a migraine-attack, anxiety disorder, or even tumor formation.

Author Contributions

V.D. and P.G. wrote the manuscript. A.M. collected references and edited figures, tables, and text. R.G. reviewed the manuscript and provided expertise.

Funding

This research was funded by The National Research, Development, and Innovation Office (KTIA_NAP_20017-1.2.1-NKP-2017-0002 and NKFIH 119 289) and Higher Education Institutional Excellence Programme of the Ministry of Human Capacities in Hungary, within the framework of the 20765-3/2019/FEKUTSTRAT II. 5. thematic programme of the University of Pécs, supported this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miyata, A.; Arimura, A.; Dahl, R.R.; Minamino, N.; Uehara, A.; Jiang, L.; Culler, M.D.; Coy, D.H. Isolation of a novel 38 residue-hypothalamic polypeptide which stimulates adenylate cyclase in pituitary cells. Biochem. Biophys. Res. Commun. 1989, 164, 567–574. [Google Scholar] [CrossRef]
  2. Miyata, A.; Jiang, L.; Dahl, R.D.; Kitada, C.; Kubo, K.; Fujino, M.; Minamino, N.; Arimura, A. Isolation of a neuropeptide corresponding to the N-terminal 27 residues of the pituitary adenylate cyclase activating polypeptide with 38 residues (PACAP38). Biochem. Biophys. Res. Commun. 1990, 170, 643–648. [Google Scholar] [CrossRef]
  3. Krashes, M.J.; Shah, B.P.; Madara, J.C.; Olson, D.P.; Strochlic, D.E.; Garfield, A.S.; Vong, L.; Pei, H.; Watabe-Uchida, M.; Uchida, N.; et al. An excitatory paraventricular nucleus to AgRP neuron circuit that drives hunger. Nature 2014, 507, 238–242. [Google Scholar] [CrossRef]
  4. Ressler, K.J.; Mercer, K.B.; Bradley, B.; Jovanovic, T.; Mahan, A.; Kerley, K.; Norrholm, S.D.; Kilaru, V.; Smith, A.K.; Myers, A.J.; et al. Post-traumatic stress disorder is associated with PACAP and the PAC1 receptor. Nature 2011, 470, 492–497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Spengler, D.; Waeber, C.; Pantaloni, C.; Holsboer, F.; Bockaert, J.; Seeburg, P.H.; Journot, L. Differential signal transduction by five splice variants of the PACAP receptor. Nature 1993, 365, 170–175. [Google Scholar] [PubMed]
  6. Zhong, Y. Mediation of PACAP-like neuropeptide transmission by coactivation of Ras/Raf and cAMP signal transduction pathways in Drosophila. Nature 1995, 375, 588–592. [Google Scholar] [CrossRef]
  7. Harding, S.D.; Sharman, J.L.; Faccenda, E.; Southan, C.; Pawson, A.J.; Ireland, S.; Gray, A.J.G.; Bruce, L.; Alexander, S.P.H.; Anderton, S.; et al. The IUPHAR/BPS Guide to PHARMACOLOGY in 2018: Updates and expansion to encompass the new guide to IMMUNOPHARMACOLOGY. Nucleic Acids Res. 2018, 46, D1091–D1106. [Google Scholar] [CrossRef] [PubMed]
  8. Cunha-Reis, D.; Ribeiro, J.A.; de Almeida, R.F.M.; Sebastiao, A.M. VPAC1 and VPAC2 receptor activation on GABA release from hippocampal nerve terminals involve several different signalling pathways. Br. J. Pharmacol. 2017, 174, 4725–4737. [Google Scholar] [CrossRef]
  9. Tsutsumi, M.; Claus, T.H.; Liang, Y.; Li, Y.; Yang, L.; Zhu, J.; Dela Cruz, F.; Peng, X.; Chen, H.; Yung, S.L.; et al. A potent and highly selective VPAC2 agonist enhances glucose-induced insulin release and glucose disposal: A potential therapy for type 2 diabetes. Diabetes 2002, 51, 1453–1460. [Google Scholar] [CrossRef]
  10. McRory, J.; Sherwood, N.M. Two protochordate genes encode pituitary adenylate cyclase-activating polypeptide and related family members. Endocrinology 1997, 138, 2380–2390. [Google Scholar] [CrossRef]
  11. Blechman, J.; Levkowitz, G. Alternative Splicing of the Pituitary Adenylate Cyclase-Activating Polypeptide Receptor PAC1: Mechanisms of Fine Tuning of Brain Activity. Front. Endocrinol. (Lausanne) 2013, 4, 55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Denes, V.; Czotter, N.; Lakk, M.; Berta, G.; Gabriel, R. PAC1-expressing structures of neural retina alter their PAC1 isoform splicing during postnatal development. Cell Tissue Res. 2014, 355, 279–288. [Google Scholar] [CrossRef] [PubMed]
  13. Lakk, M.; Szabo, B.; Volgyi, B.; Gabriel, R.; Denes, V. Development-related splicing regulates pituitary adenylate cyclase-activating polypeptide (PACAP) receptors in the retina. Invest. Ophthalmol. Vis. Sci. 2012, 53, 7825–7832. [Google Scholar] [CrossRef]
  14. Vaudry, D.; Falluel-Morel, A.; Bourgault, S.; Basille, M.; Burel, D.; Wurtz, O.; Fournier, A.; Chow, B.K.; Hashimoto, H.; Galas, L.; et al. Pituitary adenylate cyclase-activating polypeptide and its receptors: 20 years after the discovery. Pharmacol. Rev. 2009, 61, 283–357. [Google Scholar] [CrossRef]
  15. Arimura, A. PACAP: The road to discovery. Peptides 2007, 28, 1617–1619. [Google Scholar] [CrossRef] [PubMed]
  16. Tams, J.W.; Johnsen, A.H.; Fahrenkrug, J. Identification of pituitary adenylate cyclase-activating polypeptide1-38-binding factor in human plasma, as ceruloplasmin. Biochem. J. 1999, 341, 271–276. [Google Scholar] [CrossRef]
  17. Filipsson, K.; Tornoe, K.; Holst, J.; Ahren, B. Pituitary adenylate cyclase-activating polypeptide stimulates insulin and glucagon secretion in humans. J. Clin. Endocrinol. Metab. 1997, 82, 3093–3098. [Google Scholar]
  18. American College of Physicians, Normal Laboratory Values. Available online: www.acponline.org (accessed on 16 September 2019).
  19. Du, Q.; Kawabe, Y.; Schilde, C.; Chen, Z.H.; Schaap, P. The Evolution of Aggregative Multicellularity and Cell-Cell Communication in the Dictyostelia. J. Mol. Biol. 2015, 427, 3722–3733. [Google Scholar] [CrossRef]
  20. Nordstrom, K.J.; Sallman Almen, M.; Edstam, M.M.; Fredriksson, R.; Schioth, H.B. Independent HHsearch, Needleman—Wunsch-based, and motif analyses reveal the overall hierarchy for most of the G protein-coupled receptor families. Mol. Biol. Evol. 2011, 28, 2471–2480. [Google Scholar] [CrossRef] [PubMed]
  21. Foord, S.M.; Bonner, T.I.; Neubig, R.R.; Rosser, E.M.; Pin, J.P.; Davenport, A.P.; Spedding, M.; Harmar, A.J. International Union of Pharmacology. XLVI. G protein-coupled receptor list. Pharmacol. Rev. 2005, 57, 279–288. [Google Scholar] [CrossRef] [PubMed]
  22. Ferguson, E.V.; Roy, A.; Ciechanski, P.; Wilson, R.J. Stress peptide PACAP stimulates and stabilizes neonatal breathing through distinct mechanisms. Respir. Physiol. Neurobiol. 2013, 187, 217–223. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, L.; Eiden, L.E. Two ancient neuropeptides, PACAP and AVP, modulate motivated behavior at synapses in the extrahypothalamic brain: A study in contrast. Cell Tissue Res. 2019, 375, 103–122. [Google Scholar] [CrossRef] [PubMed]
  24. Jozsa, R.; Nemeth, J.; Tamas, A.; Hollosy, T.; Lubics, A.; Jakab, B.; Olah, A.; Lengvari, I.; Arimura, A.; Reglodi, D. Short-term fasting differentially alters PACAP and VIP levels in the brains of rat and chicken. Ann. N. Y. Acad. Sci. 2006, 1070, 354–358. [Google Scholar] [CrossRef] [PubMed]
  25. Armstrong, B.D.; Abad, C.; Chhith, S.; Cheung-Lau, G.; Hajji, O.E.; Nobuta, H.; Waschek, J.A. Impaired nerve regeneration and enhanced neuroinflammatory response in mice lacking pituitary adenylyl cyclase activating peptide. Neuroscience 2008, 151, 63–73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Waschek, J.A.; Cohen, J.R.; Chi, G.C.; Proszynski, T.J.; Niewiadomski, P. PACAP Promotes Matrix-Driven Adhesion of Cultured Adult Murine Neural Progenitors. ASN Neuro 2017, 9, 1759091417708720. [Google Scholar] [CrossRef]
  27. Hamelink, C.; Tjurmina, O.; Damadzic, R.; Young, W.S.; Weihe, E.; Lee, H.W.; Eiden, L.E. Pituitary adenylate cyclase-activating polypeptide is a sympathoadrenal neurotransmitter involved in catecholamine regulation and glucohomeostasis. Proc. Natl. Acad. Sci. USA 2002, 99, 461–466. [Google Scholar] [CrossRef]
  28. Zibara, K.; Zeidan, A.; Mallah, K.; Kassem, N.; Awad, A.; Mazurier, F.; Badran, B.; El-Zein, N. Signaling pathways activated by PACAP in MCF-7 breast cancer cells. Cell Signal. 2018, 50, 37–47. [Google Scholar] [CrossRef]
  29. Jung, S.; Yi, L.; Jeong, D.; Kim, J.; An, S.; Oh, T.J.; Kim, C.H.; Kim, C.J.; Yang, Y.; Kim, K.I.; et al. The role of ADCYAP1, adenylate cyclase activating polypeptide 1, as a methylation biomarker for the early detection of cervical cancer. Oncol. Rep. 2011, 25, 245–252. [Google Scholar]
  30. Lee, J.H.; Lee, J.Y.; Rho, S.B.; Choi, J.S.; Lee, D.G.; An, S.; Oh, T.; Choi, D.C.; Lee, S.H. PACAP inhibits tumor growth and interferes with clusterin in cervical carcinomas. FEBS Lett. 2014, 588, 4730–4739. [Google Scholar] [CrossRef] [Green Version]
  31. Germano, P.M.; Le, S.V.; Oh, D.S.; Fan, R.; Lieu, S.; Siu, A.; Pisegna, J.R. Differential coupling of the PAC1 SV1 splice variant on human colonic tumors to the activation of intracellular cAMP but not intracellular Ca2+ does not activate tumor proliferation. J. Mol. Neurosci. 2004, 22, 83–92. [Google Scholar] [CrossRef]
  32. Liu, S.; Zeng, Y.; Li, Y.; Guo, W.; Liu, J.; Ouyang, N. VPAC1 overexpression is associated with poor differentiation in colon cancer. Tumour Biol. 2014, 35, 6397–6404. [Google Scholar] [CrossRef] [PubMed]
  33. Le, S.V.; Yamaguchi, D.J.; McArdle, C.A.; Tachiki, K.; Pisegna, J.R.; Germano, P. PAC1 and PACAP expression, signaling, and effect on the growth of HCT8, human colonic tumor cells. Regul. Pept. 2002, 109, 115–125. [Google Scholar] [CrossRef]
  34. Schulz, S.; Rocken, C.; Mawrin, C.; Weise, W.; Hollt, V.; Schulz, S. Immunocytochemical identification of VPAC1, VPAC2, and PAC1 receptors in normal and neoplastic human tissues with subtype-specific antibodies. Clin. Cancer Res. 2004, 10, 8235–8242. [Google Scholar] [CrossRef] [PubMed]
  35. Lauffer, J.M.; Tang, L.H.; Zhang, T.; Hinoue, T.; Rahbar, S.; Odo, M.; Modlin, I.M.; Kidd, M. PACAP mediates the neural proliferative pathway of Mastomys enterochromaffin-like cell transformation. Regul. Pept. 2001, 102, 157–164. [Google Scholar] [CrossRef]
  36. Barbarin, A.; Seite, P.; Godet, J.; Bensalma, S.; Muller, J.M.; Chadeneau, C. Atypical nuclear localization of VIP receptors in glioma cell lines and patients. Biochem. Biophys. Res. Commun. 2014, 454, 524–530. [Google Scholar] [CrossRef] [PubMed]
  37. Cochaud, S.; Meunier, A.C.; Monvoisin, A.; Bensalma, S.; Muller, J.M.; Chadeneau, C. Neuropeptides of the VIP family inhibit glioblastoma cell invasion. J. Neurooncol. 2015, 122, 63–73. [Google Scholar] [CrossRef] [PubMed]
  38. Moody, T.W.; Osefo, N.; Nuche-Berenguer, B.; Ridnour, L.; Wink, D.; Jensen, R.T. Pituitary adenylate cyclase-activating polypeptide causes tyrosine phosphorylation of the epidermal growth factor receptor in lung cancer cells. J. Pharmacol. Exp. Ther. 2012, 341, 873–881. [Google Scholar] [CrossRef] [PubMed]
  39. Giunta, S.; Castorina, A.; Adorno, A.; Mazzone, V.; Carnazza, M.L.; D’Agata, V. PACAP and VIP affect NF1 expression in rat malignant peripheral nerve sheath tumor (MPNST) cells. Neuropeptides 2010, 44, 45–51. [Google Scholar] [CrossRef]
  40. Thouennon, E.; Pierre, A.; Tanguy, Y.; Guillemot, J.; Manecka, D.L.; Guerin, M.; Ouafik, L.; Muresan, M.; Klein, M.; Bertherat, J.; et al. Expression of trophic amidated peptides and their receptors in benign and malignant pheochromocytomas: High expression of adrenomedullin RDC1 receptor and implication in tumoral cell survival. Endocr. Relat. Cancer 2010, 17, 637–651. [Google Scholar] [CrossRef]
  41. Cohen, J.R.; Resnick, D.Z.; Niewiadomski, P.; Dong, H.; Liau, L.M.; Waschek, J.A. Pituitary adenylyl cyclase activating polypeptide inhibits gli1 gene expression and proliferation in primary medulloblastoma derived tumorsphere cultures. BMC Cancer 2010, 10, 676. [Google Scholar] [CrossRef]
  42. Lelievre, V.; Seksenyan, A.; Nobuta, H.; Yong, W.H.; Chhith, S.; Niewiadomski, P.; Cohen, J.R.; Dong, H.; Flores, A.; Liau, L.M.; et al. Disruption of the PACAP gene promotes medulloblastoma in ptc1 mutant mice. Dev. Biol. 2008, 313, 359–370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Zia, H.; Leyton, J.; Casibang, M.; Hau, V.; Brenneman, D.; Fridkin, M.; Gozes, I.; Moody, T.W. (N-stearyl, norleucine17) VIP hybrid inhibits the growth of pancreatic cancer cell lines. Life Sci. 2000, 66, 379–387. [Google Scholar] [CrossRef]
  44. Gruppetta, M.; Formosa, R.; Falzon, S.; Ariff Scicluna, S.; Falzon, E.; Degeatano, J.; Vassallo, J. Expression of cell cycle regulators and biomarkers of proliferation and regrowth in human pituitary adenomas. Pituitary 2017, 20, 358–371. [Google Scholar] [CrossRef] [PubMed]
  45. Fernandez-Martinez, A.B.; Bajo, A.M.; Sanchez-Chapado, M.; Prieto, J.C.; Carmena, M.J. Vasoactive intestinal peptide behaves as a pro-metastatic factor in human prostate cancer cells. Prostate 2009, 69, 774–786. [Google Scholar] [CrossRef] [PubMed]
  46. Mammi, C.; Frajese, G.V.; Vespasiani, G.; Mariani, S.; Gnessi, L.; Farini, D.; Fabbri, A.; Frajese, G.; Moretti, C. PAC1-R null isoform expression in human prostate cancer tissue. Prostate 2006, 66, 514–521. [Google Scholar] [CrossRef] [PubMed]
  47. Vacas, E.; Fernandez-Martinez, A.B.; Bajo, A.M.; Sanchez-Chapado, M.; Schally, A.V.; Prieto, J.C.; Carmena, M.J. Vasoactive intestinal peptide (VIP) inhibits human renal cell carcinoma proliferation. Biochim. Biophys. Acta 2012, 1823, 1676–1685. [Google Scholar] [CrossRef] [Green Version]
  48. Wojcieszak, J.; Zawilska, J.B. PACAP38 and PACAP6-38 exert cytotoxic activity against human retinoblastoma Y79 cells. J. Mol. Neurosci. 2014, 54, 463–468. [Google Scholar] [CrossRef] [PubMed]
  49. Denes, V.; Hideg, O.; Nyisztor, Z.; Lakk, M.; Godri, Z.; Berta, G.; Geck, P.; Gabriel, R. The Neuroprotective Peptide PACAP1-38 Contributes to Horizontal Cell Development in Postnatal Rat Retina. Invest. Ophthalmol. Vis. Sci. 2019, 60, 770–778. [Google Scholar] [CrossRef] [Green Version]
  50. Green, B.D.; Irwin, N.; Cassidy, R.S.; Gault, V.A.; Flatt, P.R. Long-term administration of PACAP receptor antagonist, PACAP(6-27), impairs glucose tolerance and insulin sensitivity in obese diabetic ob/ob mice. Peptides 2006, 27, 2343–2349. [Google Scholar] [CrossRef]
  51. Nakata, M.; Shintani, N.; Hashimoto, H.; Baba, A.; Yada, T. Intra-islet PACAP protects pancreatic beta-cells against glucotoxicity and lipotoxicity. J. Mol. Neurosci. 2010, 42, 404–410. [Google Scholar] [CrossRef]
  52. Sanlioglu, A.D.; Karacay, B.; Balci, M.K.; Griffith, T.S.; Sanlioglu, S. Therapeutic potential of VIP vs PACAP in diabetes. J. Mol. Endocrinol. 2012, 49, R157–R167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Yamamoto, K.; Hashimoto, H.; Tomimoto, S.; Shintani, N.; Miyazaki, J.; Tashiro, F.; Aihara, H.; Nammo, T.; Li, M.; Yamagata, K.; et al. Overexpression of PACAP in transgenic mouse pancreatic beta-cells enhances insulin secretion and ameliorates streptozotocin-induced diabetes. Diabetes 2003, 52, 1155–1162. [Google Scholar] [CrossRef] [PubMed]
  54. Yu, R.; Wang, J.; Li, J.; Wang, Y.; Zhang, H.; Chen, J.; Huang, L.; Liu, X. A novel cyclopeptide from the cyclization of PACAP(1-5) with potent activity towards PAC1 attenuates STZ-induced diabetes. Peptides 2010, 31, 1062–1067. [Google Scholar] [CrossRef]
  55. Amato, R.; Biagioni, M.; Cammalleri, M.; Dal Monte, M.; Casini, G. VEGF as a Survival Factor in Ex Vivo Models of Early Diabetic Retinopathy. Invest. Ophthalmol. Vis. Sci. 2016, 57, 3066–3076. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. D’Amico, A.G.; Maugeri, G.; Rasa, D.M.; Bucolo, C.; Saccone, S.; Federico, C.; Cavallaro, S.; D’Agata, V. Modulation of IL-1beta and VEGF expression in rat diabetic retinopathy after PACAP administration. Peptides 2017, 97, 64–69. [Google Scholar] [CrossRef] [PubMed]
  57. D’Amico, A.G.; Maugeri, G.; Reitano, R.; Bucolo, C.; Saccone, S.; Drago, F.; D’Agata, V. PACAP Modulates Expression of Hypoxia-Inducible Factors in Streptozotocin-Induced Diabetic Rat Retina. J. Mol. Neurosci. 2015, 57, 501–509. [Google Scholar] [CrossRef]
  58. Giunta, S.; Castorina, A.; Bucolo, C.; Magro, G.; Drago, F.; D’Agata, V. Early changes in pituitary adenylate cyclase-activating peptide, vasoactive intestinal peptide and related receptors expression in retina of streptozotocin-induced diabetic rats. Peptides 2012, 37, 32–39. [Google Scholar] [CrossRef]
  59. Maugeri, G.; D’Amico, A.G.; Saccone, S.; Federico, C.; Cavallaro, S.; D’Agata, V. PACAP and VIP Inhibit HIF-1alpha-Mediated VEGF Expression in a Model of Diabetic Macular Edema. J. Cell Physiol. 2017, 232, 1209–1215. [Google Scholar] [CrossRef]
  60. Szabadfi, K.; Atlasz, T.; Kiss, P.; Reglodi, D.; Szabo, A.; Kovacs, K.; Szalontai, B.; Setalo, G., Jr.; Banki, E.; Csanaky, K.; et al. Protective effects of the neuropeptide PACAP in diabetic retinopathy. Cell Tissue Res. 2012, 348, 37–46. [Google Scholar] [CrossRef]
  61. Banki, E.; Degrell, P.; Kiss, P.; Kovacs, K.; Kemeny, A.; Csanaky, K.; Duh, A.; Nagy, D.; Toth, G.; Tamas, A.; et al. Effect of PACAP treatment on kidney morphology and cytokine expression in rat diabetic nephropathy. Peptides 2013, 42, 125–130. [Google Scholar] [CrossRef]
  62. Lauenstein, H.D.; Quarcoo, D.; Plappert, L.; Schleh, C.; Nassimi, M.; Pilzner, C.; Rochlitzer, S.; Brabet, P.; Welte, T.; Hoymann, H.G.; et al. Pituitary adenylate cyclase-activating peptide receptor 1 mediates anti-inflammatory effects in allergic airway inflammation in mice. Clin. Exp. Allergy 2011, 41, 592–601. [Google Scholar] [CrossRef] [PubMed]
  63. Rasbach, E.; Splitthoff, P.; Bonaterra, G.A.; Schwarz, A.; Mey, L.; Schwarzbach, H.; Eiden, L.E.; Weihe, E.; Kinscherf, R. PACAP deficiency aggravates atherosclerosis in ApoE deficient mice. Immunobiology 2019, 224, 124–132. [Google Scholar] [CrossRef] [PubMed]
  64. Vizzard, M.A. Up-regulation of pituitary adenylate cyclase-activating polypeptide in urinary bladder pathways after chronic cystitis. J. Comp. Neurol. 2000, 420, 335–348. [Google Scholar] [CrossRef]
  65. Girard, B.M.; Tooke, K.; Vizzard, M.A. PACAP/Receptor System in Urinary Bladder Dysfunction and Pelvic Pain Following Urinary Bladder Inflammation or Stress. Front. Syst. Neurosci. 2017, 11, 90. [Google Scholar] [CrossRef] [PubMed]
  66. Elekes, K.; Sandor, K.; Moricz, A.; Kereskai, L.; Kemeny, A.; Szoke, E.; Perkecz, A.; Reglodi, D.; Hashimoto, H.; Pinter, E.; et al. Pituitary adenylate cyclase-activating polypeptide plays an anti-inflammatory role in endotoxin-induced airway inflammation: In vivo study with gene-deleted mice. Peptides 2011, 32, 1439–1446. [Google Scholar] [CrossRef] [PubMed]
  67. Bereswill, S.; Escher, U.; Grunau, A.; Kuhl, A.A.; Dunay, I.R.; Tamas, A.; Reglodi, D.; Heimesaat, M.M. Pituitary Adenylate Cyclase-Activating Polypeptide-A Neuropeptide as Novel Treatment Option for Subacute Ileitis in Mice Harboring a Human Gut Microbiota. Front. Immunol. 2019, 10, 554. [Google Scholar] [CrossRef] [PubMed]
  68. Heimesaat, M.M.; Dunay, I.R.; Schulze, S.; Fischer, A.; Grundmann, U.; Alutis, M.; Kuhl, A.A.; Tamas, A.; Toth, G.; Dunay, M.P.; et al. Pituitary adenylate cyclase-activating polypeptide ameliorates experimental acute ileitis and extra-intestinal sequelae. PLoS ONE 2014, 9, e108389. [Google Scholar] [CrossRef] [PubMed]
  69. Wang, Z.Y.; Waldeck, K.; Grundemar, L.; Hakanson, R. Ocular inflammation induced by electroconvulsive treatment: Contribution of nitric oxide and neuropeptides mobilized from C-fibres. Br. J. Pharmacol. 1997, 120, 1491–1496. [Google Scholar] [CrossRef]
  70. Grassel, S.; Muschter, D. Do Neuroendocrine Peptides and Their Receptors Qualify as Novel Therapeutic Targets in Osteoarthritis? Int. J. Mol. Sci. 2018, 19, 367. [Google Scholar] [CrossRef]
  71. Giunta, S.; Castorina, A.; Marzagalli, R.; Szychlinska, M.A.; Pichler, K.; Mobasheri, A.; Musumeci, G. Ameliorative effects of PACAP against cartilage degeneration. Morphological, immunohistochemical and biochemical evidence from in vivo and in vitro models of rat osteoarthritis. Int. J. Mol. Sci. 2015, 16, 5922–5944. [Google Scholar] [CrossRef]
  72. Chen, Y.D.; Zhou, Z.G.; Wang, Z.; Gao, H.K.; Yan, W.W.; Wang, C.; Zhao, G.P.; Peng, X.H. Pituitary adenylate cyclase activating-peptide and its receptor antagonists in development of acute pancreatitis in rats. World J. Gastroenterol. 2005, 11, 538–544. [Google Scholar] [CrossRef] [PubMed]
  73. Hamagami, K.; Sakurai, Y.; Shintani, N.; Higuchi, N.; Ikeda, K.; Hashimoto, H.; Suzuki, A.; Kiyama, H.; Baba, A. Over-expression of pancreatic pituitary adenylate cyclase-activating polypeptide (PACAP) aggravates cerulein-induced acute pancreatitis in mice. J. Pharmacol. Sci. 2009, 110, 451–458. [Google Scholar] [CrossRef] [PubMed]
  74. Sakurai, Y.; Shintani, N.; Arimori, A.; Hamagami, K.; Higuchi, N.; Inoue, H.; Ikeda, K.; Hayata, A.; Hashimoto, H.; Baba, A. Cerulein-induced acute pancreatitis in PACAP knockout mice. J. Mol. Neurosci. 2011, 43, 8–15. [Google Scholar] [CrossRef] [PubMed]
  75. Choi, J.E.; Di Nardo, A. Skin neurogenic inflammation. Semin. Immunopathol. 2018, 40, 249–259. [Google Scholar] [CrossRef] [PubMed]
  76. Helyes, Z.; Kun, J.; Dobrosi, N.; Sandor, K.; Nemeth, J.; Perkecz, A.; Pinter, E.; Szabadfi, K.; Gaszner, B.; Tekus, V.; et al. Pituitary Adenylate Cyclase-Activating Polypeptide Is Upregulated in Murine Skin Inflammation and Mediates Transient Receptor Potential Vanilloid-1-Induced Neurogenic Edema. J. Invest. Dermatol. 2015, 135, 2209–2218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Warren, J.B.; Larkin, S.W.; Coughlan, M.; Kajekar, R.; Williams, T.J. Pituitary adenylate cyclase activating polypeptide is a potent vasodilator and oedema potentiator in rabbit skin in vivo. Br. J. Pharmacol. 1992, 106, 331–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Figueiredo, C.A.; Dusedau, H.P.; Steffen, J.; Gupta, N.; Dunay, M.P.; Toth, G.K.; Reglodi, D.; Heimesaat, M.M.; Dunay, I.R. Immunomodulatory Effects of the Neuropeptide Pituitary Adenylate Cyclase-Activating Polypeptide in Acute Toxoplasmosis. Front. Cell Infect. Microbiol. 2019, 9, 154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Souza, T.M.; Temerozo, J.R.; Giestal-de-Araujo, E.; Bou-Habib, D.C. The effects of neurotrophins and the neuropeptides VIP and PACAP on HIV-1 infection: Histories with opposite ends. Neuroimmunomodulation 2014, 21, 268–282. [Google Scholar] [CrossRef]
  80. Temerozo, J.R.; de Azevedo, S.S.D.; Insuela, D.B.R.; Vieira, R.C.; Ferreira, P.L.C.; Carvalho, V.F.; Bello, G.; Bou-Habib, D.C. The Neuropeptides Vasoactive Intestinal Peptide and Pituitary Adenylate Cyclase-Activating Polypeptide Control HIV-1 Infection in Macrophages Through Activation of Protein Kinases A and C. Front. Immunol. 2018, 9, 1336. [Google Scholar] [CrossRef] [Green Version]
  81. Temerozo, J.R.; Joaquim, R.; Regis, E.G.; Savino, W.; Bou-Habib, D.C. Macrophage Resistance to HIV-1 Infection is Enhanced by the Neuropeptides VIP and PACAP. PLoS ONE 2013, 8, e67701. [Google Scholar] [CrossRef]
  82. Han, P.; Caselli, R.J.; Baxter, L.; Serrano, G.; Yin, J.; Beach, T.G.; Reiman, E.M.; Shi, J. Association of pituitary adenylate cyclase-activating polypeptide with cognitive decline in mild cognitive impairment due to Alzheimer disease. JAMA Neurol 2015, 72, 333–339. [Google Scholar] [CrossRef] [PubMed]
  83. Han, P.; Liang, W.; Baxter, L.C.; Yin, J.; Tang, Z.; Beach, T.G.; Caselli, R.J.; Reiman, E.M.; Shi, J. Pituitary adenylate cyclase-activating polypeptide is reduced in Alzheimer disease. Neurology 2014, 82, 1724–1728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Rat, D.; Schmitt, U.; Tippmann, F.; Dewachter, I.; Theunis, C.; Wieczerzak, E.; Postina, R.; van Leuven, F.; Fahrenholz, F.; Kojro, E. Neuropeptide pituitary adenylate cyclase-activating polypeptide (PACAP) slows down Alzheimer’s disease-like pathology in amyloid precursor protein-transgenic mice. FASEB J. 2011, 25, 3208–3218. [Google Scholar] [CrossRef] [PubMed]
  85. Lohoff, F.W.; Bloch, P.J.; Weller, A.E.; Ferraro, T.N.; Berrettini, W.H. Association analysis of the pituitary adenylate cyclase-activating polypeptide (PACAP/ADCYAP1) gene in bipolar disorder. Psychiatr Genet. 2008, 18, 53–58. [Google Scholar] [CrossRef] [PubMed]
  86. Katayama, T.; Hattori, T.; Yamada, K.; Matsuzaki, S.; Tohyama, M. Role of the PACAP-PAC1-DISC1 and PACAP-PAC1-stathmin1 systems in schizophrenia and bipolar disorder: Novel treatment mechanisms? Pharmacogenomics 2009, 10, 1967–1978. [Google Scholar] [CrossRef] [PubMed]
  87. Bhandare, A.M.; Kapoor, K.; Powell, K.L.; Braine, E.; Casillas-Espinosa, P.; O’Brien, T.J.; Farnham, M.M.J.; Pilowsky, P.M. Inhibition of microglial activation with minocycline at the intrathecal level attenuates sympathoexcitatory and proarrhythmogenic changes in rats with chronic temporal lobe epilepsy. Neuroscience 2017, 350, 23–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Cunha-Reis, D.; Mde, F.A.; Ribeiro, J.A. Endogenous inhibition of hippocampal LTD and depotentiation by vasoactive intestinal peptide VPAC1 receptors. Hippocampus 2014, 24, 1353–1363. [Google Scholar] [CrossRef]
  89. Cabezas-Llobet, N.; Vidal-Sancho, L.; Masana, M.; Fournier, A.; Alberch, J.; Vaudry, D.; Xifro, X. Pituitary Adenylate Cyclase-Activating Polypeptide (PACAP) Enhances Hippocampal Synaptic Plasticity and Improves Memory Performance in Huntington’s Disease. Mol. Neurobiol. 2018, 55, 8263–8277. [Google Scholar] [CrossRef]
  90. Brifault, C.; Gras, M.; Liot, D.; May, V.; Vaudry, D.; Wurtz, O. Delayed pituitary adenylate cyclase-activating polypeptide delivery after brain stroke improves functional recovery by inducing m2 microglia/macrophage polarization. Stroke 2015, 46, 520–528. [Google Scholar] [CrossRef]
  91. Danyadi, B.; Szabadfi, K.; Reglodi, D.; Mihalik, A.; Danyadi, T.; Kovacs, Z.; Batai, I.; Tamas, A.; Kiss, P.; Toth, G.; et al. PACAP application improves functional outcome of chronic retinal ischemic injury in rats-evidence from electroretinographic measurements. J. Mol. Neurosci. 2014, 54, 293–299. [Google Scholar] [CrossRef]
  92. Lazarovici, P.; Cohen, G.; Arien-Zakay, H.; Chen, J.; Zhang, C.; Chopp, M.; Jiang, H. Multimodal neuroprotection induced by PACAP38 in oxygen-glucose deprivation and middle cerebral artery occlusion stroke models. J. Mol. Neurosci. 2012, 48, 526–540. [Google Scholar] [CrossRef] [PubMed]
  93. Lenti, L.; Zimmermann, A.; Kis, D.; Olah, O.; Toth, G.K.; Hegyi, O.; Busija, D.W.; Bari, F.; Domoki, F. PACAP and VIP differentially preserve neurovascular reactivity after global cerebral ischemia in newborn pigs. Brain Res. 2009, 1283, 50–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Liu, Y.; Lu, T.; Zhang, C.; Xue, Z.; Xu, J.; Busuttil, R.W.; Xia, Q.; Xu, N.; Kupiec-Weglinski, J.W.; Ji, H. Pituitary Adenylate Cyclase-activating Polypeptides Prevent Hepatocyte Damage by Promoting Yes-associated Protein in Liver Ischemia-Reperfusion Injury. Transplantation 2019, 103, 1639–1648. [Google Scholar] [CrossRef] [PubMed]
  95. Matsumoto, M.; Nakamachi, T.; Watanabe, J.; Sugiyama, K.; Ohtaki, H.; Murai, N.; Sasaki, S.; Xu, Z.; Hashimoto, H.; Seki, T.; et al. Pituitary Adenylate Cyclase-Activating Polypeptide (PACAP) is Involved in Adult Mouse Hippocampal Neurogenesis after Stroke. J. Mol. Neurosci. 2016, 59, 270–279. [Google Scholar] [CrossRef] [PubMed]
  96. Nakamachi, T.; Tsuchida, M.; Kagami, N.; Yofu, S.; Wada, Y.; Hori, M.; Tsuchikawa, D.; Yoshikawa, A.; Imai, N.; Nakamura, K.; et al. IL-6 and PACAP receptor expression and localization after global brain ischemia in mice. J. Mol. Neurosci. 2012, 48, 518–525. [Google Scholar] [CrossRef] [PubMed]
  97. Riek-Burchardt, M.; Kolodziej, A.; Henrich-Noack, P.; Reymann, K.G.; Hollt, V.; Stumm, R. Differential regulation of CXCL12 and PACAP mRNA expression after focal and global ischemia. Neuropharmacology 2010, 58, 199–207. [Google Scholar] [CrossRef]
  98. Stetler, R.A.; Zhang, F.; Liu, C.; Chen, J. Ischemic tolerance as an active and intrinsic neuroprotective mechanism. Handb. Clin. Neurol 2009, 92, 171–195. [Google Scholar] [PubMed]
  99. Werling, D.; Banks, W.A.; Salameh, T.S.; Kvarik, T.; Kovacs, L.A.; Vaczy, A.; Szabo, E.; Mayer, F.; Varga, R.; Tamas, A.; et al. Passage through the Ocular Barriers and Beneficial Effects in Retinal Ischemia of Topical Application of PACAP1-38 in Rodents. Int. J. Mol. Sci. 2017, 18, 675. [Google Scholar] [CrossRef]
  100. Akerman, S.; Goadsby, P.J. Neuronal PAC1 receptors mediate delayed activation and sensitization of trigeminocervical neurons: Relevance to migraine. Sci. Transl. Med. 2015, 7, 308ra157. [Google Scholar] [CrossRef]
  101. Amin, F.M.; Asghar, M.S.; Guo, S.; Hougaard, A.; Hansen, A.E.; Schytz, H.W.; van der Geest, R.J.; de Koning, P.J.; Larsson, H.B.; Olesen, J.; et al. Headache and prolonged dilatation of the middle meningeal artery by PACAP38 in healthy volunteers. Cephalalgia 2012, 32, 140–149. [Google Scholar] [CrossRef]
  102. Birk, S.; Sitarz, J.T.; Petersen, K.A.; Oturai, P.S.; Kruuse, C.; Fahrenkrug, J.; Olesen, J. The effect of intravenous PACAP38 on cerebral hemodynamics in healthy volunteers. Regul. Pept. 2007, 140, 185–191. [Google Scholar] [CrossRef] [PubMed]
  103. Schytz, H.W.; Birk, S.; Wienecke, T.; Kruuse, C.; Olesen, J.; Ashina, M. PACAP38 induces migraine-like attacks in patients with migraine without aura. Brain 2009, 132, 16–25. [Google Scholar] [CrossRef]
  104. Amin, F.M.; Hougaard, A.; Schytz, H.W.; Asghar, M.S.; Lundholm, E.; Parvaiz, A.I.; de Koning, P.J.; Andersen, M.R.; Larsson, H.B.; Fahrenkrug, J.; et al. Investigation of the pathophysiological mechanisms of migraine attacks induced by pituitary adenylate cyclase-activating polypeptide-38. Brain 2014, 137, 779–794. [Google Scholar] [CrossRef] [PubMed]
  105. Csati, A.; Tajti, J.; Kuris, A.; Tuka, B.; Edvinsson, L.; Warfvinge, K. Distribution of Vasoactive Intestinal Peptide, Pituitary Adenylate Cyclase-Activating Peptide, Nitric Oxide Synthase, and Their Receptors in Human and Rat Sphenopalatine Ganglion. Neuroscience 2012, 202, 158–168. [Google Scholar] [CrossRef] [PubMed]
  106. Joo, K.M.; Chung, Y.H.; Kim, M.K.; Nam, R.H.; Lee, B.L.; Lee, K.H.; Cha, C.I. Distribution of vasoactive intestinal peptide and pituitary adenylate cyclase-activating polypeptide receptors (VPAC(1), VPAC(2), and PAC(1) receptor) in the rat brain. J. Compar. Neurol. 2004, 476, 388–413. [Google Scholar] [CrossRef] [PubMed]
  107. Blackwood, D.H.; Fordyce, A.; Walker, M.T.; St Clair, D.M.; Porteous, D.J.; Muir, W.J. Schizophrenia and affective disorders--cosegregation with a translocation at chromosome 1q42 that directly disrupts brain-expressed genes: Clinical and P300 findings in a family. Am. J. Hum. Genet. 2001, 69, 428–433. [Google Scholar] [CrossRef] [PubMed]
  108. Hashimoto, R.; Hashimoto, H.; Shintani, N.; Ohi, K.; Hori, H.; Saitoh, O.; Kosuga, A.; Tatsumi, M.; Iwata, N.; Ozaki, N.; et al. Possible association between the pituitary adenylate cyclase-activating polypeptide (PACAP) gene and major depressive disorder. Neurosci. Lett. 2010, 468, 300–302. [Google Scholar] [CrossRef]
  109. Pinhasov, A.; Nesher, E.; Gross, M.; Turgeman, G.; Kreinin, A.; Yadid, G. The role of the PACAP signaling system in depression. Curr. Pharm. Des. 2011, 17, 990–1001. [Google Scholar] [CrossRef]
  110. Tohyama, M.; Miyata, S.; Hattori, T.; Shimizu, S.; Matsuzaki, S. Molecular basis of major psychiatric diseases such as schizophrenia and depression. Anat. Sci. Int. 2015, 90, 137–143. [Google Scholar] [CrossRef]
  111. Abad, C.; Jayaram, B.; Becquet, L.; Wang, Y.; O’Dorisio, M.S.; Waschek, J.A.; Tan, Y.V. VPAC1 receptor (Vipr1)-deficient mice exhibit ameliorated experimental autoimmune encephalomyelitis, with specific deficits in the effector stage. J. Neuroinflamm. 2016, 13, 169. [Google Scholar] [CrossRef] [Green Version]
  112. Abad, C.; Waschek, J.A. Immunomodulatory roles of VIP and PACAP in models of multiple sclerosis. Curr. Pharm. Des. 2011, 17, 1025–1035. [Google Scholar] [CrossRef] [PubMed]
  113. Baranowska-Bik, A.; Kochanowski, J.; Uchman, D.; Wolinska-Witort, E.; Kalisz, M.; Martynska, L.; Baranowska, B.; Bik, W. Vasoactive intestinal peptide (VIP) and pituitary adenylate cyclase activating polypeptide (PACAP) in humans with multiple sclerosis. J. Neuroimmunol. 2013, 263, 159–161. [Google Scholar] [CrossRef] [PubMed]
  114. Bonaventura, G.; Iemmolo, R.; D’Amico, A.G.; La Cognata, V.; Costanzo, E.; Zappia, M.; D’Agata, V.; Conforti, F.L.; Aronica, E.; Cavallaro, S. PACAP and PAC1R are differentially expressed in motor cortex of amyotrophic lateral sclerosis patients and support survival of iPSC-derived motor neurons. J. Cell Physiol. 2018, 233, 3343–3351. [Google Scholar] [CrossRef] [PubMed]
  115. Kato, H.; Ito, A.; Kawanokuchi, J.; Jin, S.; Mizuno, T.; Ojika, K.; Ueda, R.; Suzumura, A. Pituitary adenylate cyclase-activating polypeptide (PACAP) ameliorates experimental autoimmune encephalomyelitis by suppressing the functions of antigen presenting cells. Mult. Scler 2004, 10, 651–659. [Google Scholar] [CrossRef] [PubMed]
  116. Maugeri, G.; D’Amico, A.G.; Rasa, D.M.; Federico, C.; Saccone, S.; Morello, G.; La Cognata, V.; Cavallaro, S.; D’Agata, V. Molecular mechanisms involved in the protective effect of pituitary adenylate cyclase-activating polypeptide in an in vitro model of amyotrophic lateral sclerosis. J. Cell Physiol. 2019, 234, 5203–5214. [Google Scholar] [CrossRef] [PubMed]
  117. Tan, Y.V.; Abad, C.; Lopez, R.; Dong, H.; Liu, S.; Lee, A.; Gomariz, R.P.; Leceta, J.; Waschek, J.A. Pituitary adenylyl cyclase-activating polypeptide is an intrinsic regulator of Treg abundance and protects against experimental autoimmune encephalomyelitis. Proc. Natl. Acad. Sci. USA 2009, 106, 2012–2017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Tan, Y.V.; Waschek, J.A. Targeting VIP and PACAP receptor signalling: New therapeutic strategies in multiple sclerosis. ASN Neuro 2011, 3, AN20110024. [Google Scholar] [CrossRef] [PubMed]
  119. Van, C.; Condro, M.C.; Lov, K.; Zhu, R.; Ricaflanca, P.T.; Ko, H.H.; Diep, A.L.; Hoang, A.Q.; Pisegna, J.; Rohrer, H.; et al. Correction to: PACAP/PAC1 Regulation of Inflammation via Catecholaminergic Neurons in a Model of Multiple Sclerosis. J. Mol. Neurosci. 2019, 68, 452. [Google Scholar] [CrossRef] [PubMed]
  120. Brown, D.; Tamas, A.; Reglodi, D.; Tizabi, Y. PACAP protects against inflammatory-mediated toxicity in dopaminergic SH-SY5Y cells: Implication for Parkinson’s disease. Neurotox Res. 2014, 26, 230–239. [Google Scholar] [CrossRef] [PubMed]
  121. Lamine, A.; Letourneau, M.; Doan, N.D.; Maucotel, J.; Couvineau, A.; Vaudry, H.; Chatenet, D.; Vaudry, D.; Fournier, A. Characterizations of a synthetic pituitary adenylate cyclase-activating polypeptide analog displaying potent neuroprotective activity and reduced in vivo cardiovascular side effects in a Parkinson’s disease model. Neuropharmacology 2016, 108, 440–450. [Google Scholar] [CrossRef] [PubMed]
  122. Shivers, K.Y.; Nikolopoulou, A.; Machlovi, S.I.; Vallabhajosula, S.; Figueiredo-Pereira, M.E. PACAP27 prevents Parkinson-like neuronal loss and motor deficits but not microglia activation induced by prostaglandin J2. Biochim. Biophys. Acta 2014, 1842, 1707–1719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Wang, G.; Pan, J.; Tan, Y.Y.; Sun, X.K.; Zhang, Y.F.; Zhou, H.Y.; Ren, R.J.; Wang, X.J.; Chen, S.D. Neuroprotective effects of PACAP27 in mice model of Parkinson’s disease involved in the modulation of K(ATP) subunits and D2 receptors in the striatum. Neuropeptides 2008, 42, 267–276. [Google Scholar] [CrossRef] [PubMed]
  124. Chang, S.C.; Xie, P.; Anton, R.F.; De Vivo, I.; Farrer, L.A.; Kranzler, H.R.; Oslin, D.; Purcell, S.M.; Roberts, A.L.; Smoller, J.W.; et al. No association between ADCYAP1R1 and post-traumatic stress disorder in two independent samples. Mol. Psychiatry 2012, 17, 239–241. [Google Scholar] [CrossRef] [PubMed]
  125. Dias, B.G.; Ressler, K.J. PACAP and the PAC1 receptor in post-traumatic stress disorder. Neuropsychopharmacology 2013, 38, 245–246. [Google Scholar] [CrossRef] [PubMed]
  126. Hashimoto, R.; Hashimoto, H.; Shintani, N.; Chiba, S.; Hattori, S.; Okada, T.; Nakajima, M.; Tanaka, K.; Kawagishi, N.; Nemoto, K.; et al. Pituitary adenylate cyclase-activating polypeptide is associated with schizophrenia. Mol. Psychiatry 2007, 12, 1026–1032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Koga, M.; Ishiguro, H.; Horiuchi, Y.; Inada, T.; Ujike, H.; Itokawa, M.; Otowa, T.; Watanabe, Y.; Someya, T.; Arinami, T. Replication study of association between ADCYAP1 gene polymorphisms and schizophrenia. Psychiatr. Genet. 2010, 20, 123–125. [Google Scholar] [CrossRef] [PubMed]
  128. Hua, R.; Mao, S.S.; Zhang, Y.M.; Chen, F.X.; Zhou, Z.H.; Liu, J.Q. Effects of pituitary adenylate cyclase activating polypeptide on CD4(+)/CD8(+) T cell levels after traumatic brain injury in a rat model. World J. Emerg. Med. 2012, 3, 294–298. [Google Scholar] [CrossRef]
  129. Miyamoto, K.; Tsumuraya, T.; Ohtaki, H.; Dohi, K.; Satoh, K.; Xu, Z.; Tanaka, S.; Murai, N.; Watanabe, J.; Sugiyama, K.; et al. PACAP38 suppresses cortical damage in mice with traumatic brain injury by enhancing antioxidant activity. J. Mol. Neurosci. 2014, 54, 370–379. [Google Scholar] [CrossRef]
  130. Fang, K.M.; Chen, J.K.; Hung, S.C.; Chen, M.C.; Wu, Y.T.; Wu, T.J.; Lin, H.I.; Chen, C.H.; Cheng, H.; Yang, C.S.; et al. Effects of combinatorial treatment with pituitary adenylate cyclase activating peptide and human mesenchymal stem cells on spinal cord tissue repair. PLoS ONE 2010, 5, e15299. [Google Scholar] [CrossRef]
  131. Tsuchida, M.; Nakamachi, T.; Sugiyama, K.; Tsuchikawa, D.; Watanabe, J.; Hori, M.; Yoshikawa, A.; Imai, N.; Kagami, N.; Matkovits, A.; et al. PACAP stimulates functional recovery after spinal cord injury through axonal regeneration. J. Mol. Neurosci. 2014, 54, 380–387. [Google Scholar] [CrossRef]
  132. Lemaire, P. Evolutionary crossroads in developmental biology: The tunicates. Development 2011, 138, 2143–2152. [Google Scholar] [CrossRef] [PubMed]
  133. Jeffery, W.R. Chordate ancestry of the neural crest: New insights from ascidians. Semin Cell Dev. Biol. 2007, 18, 481–491. [Google Scholar] [CrossRef] [PubMed]
  134. Nakamachi, T.; Ohtaki, H.; Seki, T.; Yofu, S.; Kagami, N.; Hashimoto, H.; Shintani, N.; Baba, A.; Mark, L.; Lanekoff, I.; et al. PACAP suppresses dry eye signs by stimulating tear secretion. Nat. Commun. 2016, 7, 12034. [Google Scholar] [CrossRef] [PubMed]
  135. Cardoso, J.C.; Felix, R.C.; Martins, R.S.; Trindade, M.; Fonseca, V.G.; Fuentes, J.; Power, D.M. PACAP system evolution and its role in melanophore function in teleost fish skin. Mol. Cell Endocrinol. 2015, 411, 130–145. [Google Scholar] [CrossRef] [PubMed]
  136. Lu, N.; Zhou, R.; DiCicco-Bloom, E. Opposing mitogenic regulation by PACAP in sympathetic and cerebral cortical precursors correlates with differential expression of PACAP receptor (PAC1-R) isoforms. J. Neurosci. Res. 1998, 53, 651–662. [Google Scholar] [CrossRef]
  137. Robberecht, P.; Woussen-Colle, M.C.; De Neef, P.; Gourlet, P.; Buscail, L.; Vandermeers, A.; Vandermeers-Piret, M.C.; Christophe, J. The two forms of the pituitary adenylate cyclase activating polypeptide (PACAP (1-27) and PACAP (1-38)) interact with distinct receptors on rat pancreatic AR 4-2J cell membranes. FEBS Lett. 1991, 286, 133–136. [Google Scholar] [CrossRef] [Green Version]
  138. IUPHAR/BPS Guide to Pharmacology. Available online: https://www.guidetopharmacology.org/GRAC/FamilyDisplayForward?familyId=67 (accessed on 16 September 2019).
  139. Hirabayashi, T.; Nakamachi, T.; Shioda, S. Discovery of PACAP and its receptors in the brain. J. Headache Pain 2018, 19, 28. [Google Scholar] [CrossRef]
  140. Felley, C.P.; Qian, J.M.; Mantey, S.; Pradhan, T.; Jensen, R.T. Chief cells possess a receptor with high affinity for PACAP and VIP that stimulates pepsinogen release. Am. J. Physiol. 1992, 263, G901–G907. [Google Scholar] [CrossRef] [PubMed]
  141. Harmar, A.J.; Fahrenkrug, J.; Gozes, I.; Laburthe, M.; May, V.; Pisegna, J.R.; Vaudry, D.; Vaudry, H.; Waschek, J.A.; Said, S.I. Pharmacology and functions of receptors for vasoactive intestinal peptide and pituitary adenylate cyclase-activating polypeptide: IUPHAR review 1. Br. J. Pharmacol. 2012, 166, 4–17. [Google Scholar] [CrossRef] [PubMed]
  142. Pisegna, J.R.; Wank, S.A. Cloning and characterization of the signal transduction of four splice variants of the human pituitary adenylate cyclase activating polypeptide receptor. Evidence for dual coupling to adenylate cyclase and phospholipase C. J. Biol. Chem. 1996, 271, 17267–17274. [Google Scholar] [CrossRef] [PubMed]
  143. Chatterjee, T.K.; Sharma, R.V.; Fisher, R.A. Molecular cloning of a novel variant of the pituitary adenylate cyclase-activating polypeptide (PACAP) receptor that stimulates calcium influx by activation of L-type calcium channels. J. Biol. Chem. 1996, 271, 32226–32232. [Google Scholar] [CrossRef] [PubMed]
  144. Merriam, L.A.; Baran, C.N.; Girard, B.M.; Hardwick, J.C.; May, V.; Parsons, R.L. Pituitary adenylate cyclase 1 receptor internalization and endosomal signaling mediate the pituitary adenylate cyclase activating polypeptide-induced increase in guinea pig cardiac neuron excitability. J. Neurosci. 2013, 33, 4614–4622. [Google Scholar] [CrossRef] [PubMed]
  145. May, V.; Buttolph, T.R.; Girard, B.M.; Clason, T.A.; Parsons, R.L. PACAP-induced ERK activation in HEK cells expressing PAC1 receptors involves both receptor internalization and PKC signaling. Am. J. Physiol. Cell Physiol. 2014, 306, C1068–C1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Christopoulos, A.; Christopoulos, G.; Morfis, M.; Udawela, M.; Laburthe, M.; Couvineau, A.; Kuwasako, K.; Tilakaratne, N.; Sexton, P.M. Novel receptor partners and function of receptor activity-modifying proteins. J. Biol. Chem. 2003, 278, 3293–3297. [Google Scholar] [CrossRef] [PubMed]
  147. Dickson, L.; Finlayson, K. VPAC and PAC receptors: From ligands to function. Pharmacol. Ther. 2009, 121, 294–316. [Google Scholar] [CrossRef] [PubMed]
  148. Moody, T.W.; Ito, T.; Osefo, N.; Jensen, R.T. VIP and PACAP: Recent insights into their functions/roles in physiology and disease from molecular and genetic studies. Curr. Opin. Endocrinol. Diabetes Obes. 2011, 18, 61–67. [Google Scholar] [CrossRef] [PubMed]
  149. Puledda, F.; Messina, R.; Goadsby, P.J. An update on migraine: Current understanding and future directions. J. Neurol 2017, 264, 2031–2039. [Google Scholar] [CrossRef]
  150. Baun, M.; Pedersen, M.H.; Olesen, J.; Jansen-Olesen, I. Dural mast cell degranulation is a putative mechanism for headache induced by PACAP-38. Cephalalgia 2012, 32, 337–345. [Google Scholar] [CrossRef]
  151. Carlton, S.M. Nociceptive primary afferents: They have a mind of their own. J. Physiol. 2014, 592, 3403–3411. [Google Scholar] [CrossRef]
  152. Luger, T.A. Neuromediators—A crucial component of the skin immune system. J. Dermatol. Sci. 2002, 30, 87–93. [Google Scholar] [CrossRef]
  153. Jansen-Olesen, I.; Pedersen, S.H. PACAP and its receptors in cranial arteries and mast cells. J. Headache Pain 2018, 19, 16. [Google Scholar] [CrossRef] [PubMed]
  154. Minkes, R.K.; McMahon, T.J.; Hood, J.S.; Murphy, W.A.; Coy, D.H.; McNamara, D.B.; Kadowitz, P.J. Differential effects of PACAP and VIP on the pulmonary and hindquarters vascular beds of the cat. J. Appl. Physiol. 1992, 72, 1212–1217. [Google Scholar] [CrossRef] [PubMed]
  155. Tong, S.; Parfenova, H.; Shibata, M.; Zuckerman, S.; Armstead, W.M.; Leffler, C.W. Pituitary adenylate cyclase-activating polypeptide dilates cerebral arterioles of newborn pigs. Proc. Soc. Exp. Biol. Med. 1993, 203, 343–347. [Google Scholar] [CrossRef] [PubMed]
  156. Ascuitto, R.J.; RossAscuitto, N.T.; Waddell, A.E.; Kadowitz, P.J. Contractile and coronary vascular effects of pituitary adenylate cyclase activating polypeptide in neonatal pig hearts. Cardiovascular Res. 1996, 31, E153–E159. [Google Scholar] [CrossRef]
  157. Edvinsson, L.; Elsas, T.; Suzuki, N.; Shimizu, T.; Lee, T.J. Origin and Co-localization of nitric oxide synthase, CGRP, PACAP, and VIP in the cerebral circulation of the rat. Microsc Res. Tech. 2001, 53, 221–228. [Google Scholar] [CrossRef] [PubMed]
  158. Knutsson, M.; Edvinsson, L. Distribution of mRNA for VIP and PACAP receptors in human cerebral arteries and cranial ganglia. Neuroreport 2002, 13, 507–509. [Google Scholar] [CrossRef]
  159. Boni, L.J.; Ploug, K.B.; Olesen, J.; Jansen-Olesen, I.; Gupta, S. The in vivo effect of VIP, PACAP-38 and PACAP-27 and mRNA expression of their receptors in rat middle meningeal artery. Cephalalgia 2009, 29, 837–847. [Google Scholar] [CrossRef]
  160. Amin, F.M.; Schytz, H.W. Transport of the pituitary adenylate cyclase-activating polypeptide across the blood-brain barrier: Implications for migraine. J. Headache Pain 2018, 19, 35. [Google Scholar] [CrossRef]
  161. Chaudhary, P.; Baumann, T.K. Expression of VPAC2 receptor and PAC1 receptor splice variants in the trigeminal ganglion of the adult rat. Brain Res. Mol. Brain Res. 2002, 104, 137–142. [Google Scholar] [CrossRef]
  162. Hensley, K.; Pretorius, J.; Chan, B.; Page, K.; Liu, H.; Choi, C.; Shi, D.; Xu, C.; Edvinsson, L.; Miller, S. PAC1 receptor mRNA and protein distribution in rat and human trigeminal and sphenopalatine ganglia, spinal trigeminal nucleus and in dura mater. Cephalalgia 2019, 39, 827–840. [Google Scholar] [CrossRef]
  163. Okragly, A.J.; Morin, S.M.; DeRosa, D.; Martin, A.P.; Johnson, K.W.; Johnson, M.P.; Benschop, R.J. Human mast cells release the migraine-inducing factor pituitary adenylate cyclase-activating polypeptide (PACAP). Cephalalgia 2018, 38, 1564–1574. [Google Scholar] [CrossRef] [PubMed]
  164. Conti, P.; D’Ovidio, C.; Conti, C.; Gallenga, C.E.; Lauritano, D.; Caraffa, A.; Kritas, S.K.; Ronconi, G. Progression in migraine: Role of mast cells and pro-inflammatory and anti-inflammatory cytokines. Eur. J. Pharmacol. 2019, 844, 87–94. [Google Scholar] [CrossRef] [PubMed]
  165. Delgado, M.; Pozo, D.; Martinez, C.; Garrido, E.; Leceta, J.; Calvo, J.R.; Gomariz, R.P. Characterization of gene expression of VIP and VIP1-receptor in rat peritoneal lymphocytes and macrophages. Regul. Pept. 1996, 62, 161–166. [Google Scholar] [CrossRef]
  166. Schmidt-Choudhury, A.; Furuta, G.T.; Galli, S.J.; Schmidt, W.E.; Wershil, B.K. Mast cells contribute to PACAP-induced dermal oedema in mice. Regul. Pept. 1999, 82, 65–69. [Google Scholar] [CrossRef]
  167. Bhatt, D.K.; Gupta, S.; Olesen, J.; Jansen-Olesen, I. PACAP-38 infusion causes sustained vasodilation of the middle meningeal artery in the rat: Possible involvement of mast cells. Cephalalgia 2014, 34, 877–886. [Google Scholar] [CrossRef] [PubMed]
  168. Robberecht, P.; Gourlet, P.; De Neef, P.; Woussen-Colle, M.C.; Vandermeers-Piret, M.C.; Vandermeers, A.; Christophe, J. Structural requirements for the occupancy of pituitary adenylate-cyclase-activating-peptide (PACAP) receptors and adenylate cyclase activation in human neuroblastoma NB-OK-1 cell membranes. Discovery of PACAP(6-38) as a potent antagonist. Eur. J. Biochem. 1992, 207, 239–246. [Google Scholar] [CrossRef] [PubMed]
  169. Mukai, K.; Tsai, M.; Saito, H.; Galli, S.J. Mast cells as sources of cytokines, chemokines, and growth factors. Immunol. Rev. 2018, 282, 121–150. [Google Scholar] [CrossRef] [PubMed]
  170. Wernersson, S.; Pejler, G. Mast cell secretory granules: Armed for battle. Nat. Rev. Immunol. 2014, 14, 478–494. [Google Scholar] [CrossRef] [PubMed]
  171. Ramachandran, R. Neurogenic inflammation and its role in migraine. Semin Immunopathol. 2018, 40, 301–314. [Google Scholar] [CrossRef] [PubMed]
  172. Markowitz, S.; Saito, K.; Moskowitz, M.A. Neurogenically mediated leakage of plasma protein occurs from blood vessels in dura mater but not brain. J. Neurosci. 1987, 7, 4129–4136. [Google Scholar] [CrossRef] [Green Version]
  173. Markowitz, S.; Saito, K.; Moskowitz, M.A. Neurogenically mediated plasma extravasation in dura mater: Effect of ergot alkaloids. A possible mechanism of action in vascular headache. Cephalalgia 1988, 8, 83–91. [Google Scholar] [CrossRef] [PubMed]
  174. May, A.; Goadsby, P.J. Substance P receptor antagonists in the therapy of migraine. Expert Opin. Investig. Drugs 2001, 10, 673–678. [Google Scholar] [CrossRef] [PubMed]
  175. Chan, K.Y.; Baun, M.; de Vries, R.; van den Bogaerdt, A.J.; Dirven, C.M.; Danser, A.H.; Jansen-Olesen, I.; Olesen, J.; Villalon, C.M.; MaassenVanDenBrink, A.; et al. Pharmacological characterization of VIP and PACAP receptors in the human meningeal and coronary artery. Cephalalgia 2011, 31, 181–189. [Google Scholar] [CrossRef] [PubMed]
  176. Hautmann, M.; Friis, U.G.; Desch, M.; Todorov, V.; Castrop, H.; Segerer, F.; Otto, C.; Schutz, G.; Schweda, F. Pituitary adenylate cyclase-activating polypeptide stimulates renin secretion via activation of PAC1 receptors. J. Am. Soc. Nephrol. 2007, 18, 1150–1156. [Google Scholar] [CrossRef] [PubMed]
  177. Seeliger, S.; Buddenkotte, J.; Schmidt-Choudhury, A.; Rosignoli, C.; Shpacovitch, V.; von Arnim, U.; Metze, D.; Rukwied, R.; Schmelz, M.; Paus, R.; et al. Pituitary adenylate cyclase activating polypeptide: An important vascular regulator in human skin in vivo. Am. J. Pathol. 2010, 177, 2563–2575. [Google Scholar] [CrossRef] [PubMed]
  178. Clinical trials.gov. Available online: https://clinicaltrials.gov/ct2/show/NCT03238781 (accessed on 16 September 2019).
  179. Eftekhari, S.; Salvatore, C.A.; Johansson, S.; Chen, T.B.; Zeng, Z.; Edvinsson, L. Localization of CGRP, CGRP receptor, PACAP and glutamate in trigeminal ganglion. Relation to the blood-brain barrier. Brain Res. 2015, 1600, 93–109. [Google Scholar] [CrossRef] [PubMed]
  180. Hashimoto, H.; Shintani, N.; Tanida, M.; Hayata, A.; Hashimoto, R.; Baba, A. PACAP is implicated in the stress axes. Curr. Pharm. Des. 2011, 17, 985–989. [Google Scholar] [CrossRef]
  181. Przywara, D.A.; Guo, X.; Angelilli, M.L.; Wakade, T.D.; Wakade, A.R. A non-cholinergic transmitter, pituitary adenylate cyclase-activating polypeptide, utilizes a novel mechanism to evoke catecholamine secretion in rat adrenal chromaffin cells. J. Biol. Chem. 1996, 271, 10545–10550. [Google Scholar] [CrossRef]
  182. Koves, K.; Arimura, A.; Gorcs, T.G.; Somogyvarivigh, A. Comparative Distribution of Immunoreactive Pituitary Adenylate-Cyclase Activating Polypeptide and Vasoactive Intestinal Polypeptide in Rat Forebrain. Neuroendocrinology 1991, 54, 159–169. [Google Scholar] [CrossRef]
  183. Kozicz, T.; Vigh, S.; Arimura, A. The source of origin of PACAP- and VIP-immunoreactive fibers in the laterodorsal division of the bed nucleus of the stria terminalis in the rat. Brain Res. 1998, 810, 211–219. [Google Scholar] [CrossRef]
  184. Otto, C.; Zuschratter, W.; Gass, P.; Schutz, G. Presynaptic localization of the PACAP-typeI-receptor in hippocampal and cerebellar mossy fibres. Mol. Brain Res. 1999, 66, 163–174. [Google Scholar] [CrossRef]
  185. Piggins, H.D.; Stamp, J.A.; Burns, J.; Rusak, B.; Semba, K. Distribution of pituitary adenylate cyclase activating polypeptide (PACAP) immunoreactivity in the hypothalamus and extended amygdala of the rat. J. Comp. Neurol. 1996, 376, 278–294. [Google Scholar] [CrossRef]
  186. Hammack, S.E.; Cheung, J.; Rhodes, K.M.; Schutz, K.C.; Falls, W.A.; Braas, K.M.; May, V. Chronic stress increases pituitary adenylate cyclase-activating peptide (PACAP) and brain-derived neurotrophic factor (BDNF) mRNA expression in the bed nucleus of the stria terminalis (BNST): Roles for PACAP in anxiety-like behavior. Psychoneuroendocrinology 2009, 34, 833–843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Hammack, S.E.; Roman, C.W.; Lezak, K.R.; Kocho-Shellenberg, M.; Grimmig, B.; Falls, W.A.; Braas, K.; May, V. Roles for pituitary adenylate cyclase-activating peptide (PACAP) expression and signaling in the bed nucleus of the stria terminalis (BNST) in mediating the behavioral consequences of chronic stress. J. Mol. Neurosci. 2010, 42, 327–340. [Google Scholar] [CrossRef]
  188. Agarwal, A.; Halvorson, L.M.; Legradi, G. Pituitary adenylate cyclase-activating polypeptide (PACAP) mimics neuroendocrine and behavioral manifestations of stress: Evidence for PKA-mediated expression of the corticotropin-releasing hormone (CRH) gene. Brain Res. Mol. Brain Res. 2005, 138, 45–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Grinevich, V.; Fournier, A.; Pelletier, G. Effects of pituitary adenylate cyclase-activating polypeptide (PACAP) on corticotropin-releasing hormone (CRH) gene expression in the rat hypothalamic paraventricular nucleus. Brain Res. 1997, 773, 190–196. [Google Scholar] [CrossRef]
  190. Stroth, N.; Eiden, L.E. Stress hormone synthesis in mouse hypothalamus and adrenal gland triggered by restraint is dependent on pituitary adenylate cyclase-activating polypeptide signaling. Neuroscience 2010, 165, 1025–1030. [Google Scholar] [CrossRef] [Green Version]
  191. Hashimoto, H.; Shintani, N.; Tanaka, K.; Mori, W.; Hirose, M.; Matsuda, T.; Sakaue, M.; Miyazaki, J.; Niwa, H.; Tashiro, F.; et al. Altered psychomotor behaviors in mice lacking pituitary adenylate cyclase-activating polypeptide (PACAP). Proc. Natl. Acad. Sci. USA 2001, 98, 13355–13360. [Google Scholar] [CrossRef] [Green Version]
  192. Gaszner, B.; Kormos, V.; Kozicz, T.; Hashimoto, H.; Reglodi, D.; Helyes, Z. The behavioral phenotype of pituitary adenylate-cyclase activating polypeptide-deficient mice in anxiety and depression tests is accompanied by blunted c-Fos expression in the bed nucleus of the stria terminalis, central projecting Edinger-Westphal nucleus, ventral lateral septum, and dorsal raphe nucleus. Neuroscience 2012, 202, 283–299. [Google Scholar]
  193. Norrholm, S.D.; Das, M.; Legradi, G. Behavioral effects of local microinfusion of pituitary adenylate cyclase activating polypeptide (PACAP) into the paraventricular nucleus of the hypothalamus (PVN). Regul. Pept. 2005, 128, 33–41. [Google Scholar] [CrossRef] [Green Version]
  194. Walker, D.L.; Davis, M. Double dissociation between the involvement of the bed nucleus of the stria terminalis and the central nucleus of the amygdala in startle increases produced by conditioned versus unconditioned fear. J. Neurosci. 1997, 17, 9375–9383. [Google Scholar] [CrossRef] [PubMed]
  195. Walker, D.L.; Toufexis, D.J.; Davis, M. Role of the bed nucleus of the stria terminalis versus the amygdala in fear, stress, and anxiety. Eur. J. Pharmacol. 2003, 463, 199–216. [Google Scholar] [CrossRef]
  196. King, S.B.; Toufexis, D.J.; Hammack, S.E. Pituitary adenylate cyclase activating polypeptide (PACAP), stress, and sex hormones. Stress 2017, 20, 465–475. [Google Scholar] [CrossRef] [PubMed]
  197. Miles, O.W.; Maren, S. Role of the Bed Nucleus of the Stria Terminalis in PTSD: Insights From Preclinical Models. Front. Behav. Neurosci. 2019, 13, 68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Starr, E.R.; Margiotta, J.F. Pituitary adenylate cyclase activating polypeptide induces long-term, transcription-dependent plasticity and remodeling at autonomic synapses. Mol. Cell Neurosci. 2017, 85, 170–182. [Google Scholar] [CrossRef]
  199. Meloni, E.G.; Kaye, K.T.; Venkataraman, A.; Carlezon, W.A., Jr. PACAP increases Arc/Arg 3.1 expression within the extended amygdala after fear conditioning in rats. Neurobiol. Learn. Mem. 2019, 157, 24–34. [Google Scholar] [CrossRef]
  200. Kirry, A.J.; Herbst, M.R.; Poirier, S.E.; Maskeri, M.M.; Rothwell, A.C.; Twining, R.C.; Gilmartin, M.R. Pituitary adenylate cyclase-activating polypeptide (PACAP) signaling in the prefrontal cortex modulates cued fear learning, but not spatial working memory, in female rats. Neuropharmacology 2018, 133, 145–154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Meloni, E.G.; Venkataraman, A.; Donahue, R.J.; Carlezon, W.A., Jr. Bi-directional effects of pituitary adenylate cyclase-activating polypeptide (PACAP) on fear-related behavior and c-Fos expression after fear conditioning in rats. Psychoneuroendocrinology 2016, 64, 12–21. [Google Scholar] [CrossRef]
  202. Roman, C.W.; Lezak, K.R.; Hartsock, M.J.; Falls, W.A.; Braas, K.M.; Howard, A.B.; Hammack, S.E.; May, V. PAC1 receptor antagonism in the bed nucleus of the stria terminalis (BNST) attenuates the endocrine and behavioral consequences of chronic stress. Psychoneuroendocrinology 2014, 47, 151–165. [Google Scholar] [CrossRef] [Green Version]
  203. Ago, Y.; Hayata, A.; Hashimoto, H. [Pathophysiological implication of the VPAC2 receptor in psychiatric disorders]. Nihon Yakurigaku Zasshi 2018, 151, 249–253. [Google Scholar] [CrossRef] [PubMed]
  204. Stevens, J.S.; Almli, L.M.; Fani, N.; Gutman, D.A.; Bradley, B.; Norrholm, S.D.; Reiser, E.; Ely, T.D.; Dhanani, R.; Glover, E.M.; et al. PACAP receptor gene polymorphism impacts fear responses in the amygdala and hippocampus. Proc. Natl. Acad. Sci. USA 2014, 111, 3158–3163. [Google Scholar] [CrossRef] [Green Version]
  205. Lindert, J.; von Ehrenstein, O.S.; Wehrwein, A.; Brahler, E.; Schafer, I. [Anxiety, Depression and Posttraumatic Stress Disorder in Refugees—A Systematic Review]. Psychother. Psychosom. Med. Psychol. 2018, 68, 22–29. [Google Scholar]
  206. Lao, K.; Ji, N.; Zhang, X.; Qiao, W.; Tang, Z.; Gou, X. Drug development for Alzheimer’s disease: Review. J. Drug Target. 2019, 27, 164–173. [Google Scholar] [CrossRef] [PubMed]
  207. Onoue, S.; Endo, K.; Ohshima, K.; Yajima, T.; Kashimoto, K. The neuropeptide PACAP attenuates beta-amyloid (1-42)-induced toxicity in PC12 cells. Peptides 2002, 23, 1471–1478. [Google Scholar] [CrossRef]
  208. Kojro, E.; Postina, R.; Buro, C.; Meiringer, C.; Gehrig-Burger, K.; Fahrenholz, F. The neuropeptide PACAP promotes the alpha-secretase pathway for processing the Alzheimer amyloid precursor protein. FASEB J. 2006, 20, 512–514. [Google Scholar] [CrossRef] [PubMed]
  209. Sun, X.; Chen, W.D.; Wang, Y.D. beta-Amyloid: The key peptide in the pathogenesis of Alzheimer’s disease. Front. Pharmacol. 2015, 6, 221. [Google Scholar] [CrossRef]
  210. Dogrukol-Ak, D.; Kumar, V.B.; Ryerse, J.S.; Farr, S.A.; Verma, S.; Nonaka, N.; Nakamachi, T.; Ohtaki, H.; Niehoff, M.L.; Edwards, J.C.; et al. Isolation of peptide transport system-6 from brain endothelial cells: Therapeutic effects with antisense inhibition in Alzheimer and stroke models. J. Cereb. Blood Flow Metab. 2009, 29, 411–422. [Google Scholar] [CrossRef] [PubMed]
  211. Wu, Z.L.; Ciallella, J.R.; Flood, D.G.; O’Kane, T.M.; Bozyczko-Coyne, D.; Savage, M.J. Comparative analysis of cortical gene expression in mouse models of Alzheimer’s disease. Neurobiol. Aging 2006, 27, 377–386. [Google Scholar] [CrossRef] [PubMed]
  212. Ferrari, R.; Ceconi, C.; Curello, S.; Cargnoni, A.; De Giuli, F.; Visioli, O. Occurrence of oxidative stress during myocardial reperfusion. Mol. Cell Biochem. 1992, 111, 61–69. [Google Scholar] [CrossRef]
  213. Ross-Ascuitto, N.T.; Ascuitto, R.J.; Ramage, D.; Kydon, D.W.; Coy, D.H.; Kadowitz, P.J. Pituitary adenylate cyclase activating polypeptide: A neuropeptide with potent inotropic and coronary vasodilatory effects in neonatal pig hearts. Pediatr. Res. 1993, 34, 323–328. [Google Scholar] [CrossRef]
  214. Sano, H.; Miyata, A.; Horio, T.; Nishikimi, T.; Matsuo, H.; Kangawa, K. The effect of pituitary adenylate cyclase activating polypeptide on cultured rat cardiocytes as a cardioprotective factor. Regul. Pept. 2002, 109, 107–113. [Google Scholar] [CrossRef]
  215. Racz, B.; Reglodi, D.; Horvath, G.; Szigeti, A.; Balatonyi, B.; Roth, E.; Weber, G.; Alotti, N.; Toth, G.; Gasz, B. Protective effect of PACAP against doxorubicin-induced cell death in cardiomyocyte culture. J. Mol. Neurosci. 2010, 42, 419–427. [Google Scholar] [CrossRef] [PubMed]
  216. Roth, E.; Weber, G.; Kiss, P.; Horvath, G.; Toth, G.; Gasz, B.; Ferencz, A.; Gallyas, F., Jr.; Reglodi, D.; Racz, B. Effects of PACAP and preconditioning against ischemia/reperfusion-induced cardiomyocyte apoptosis in vitro. Ann. N. Y. Acad. Sci. 2009, 1163, 512–516. [Google Scholar] [CrossRef] [PubMed]
  217. Gasz, B.; Racz, B.; Roth, E.; Borsiczky, B.; Ferencz, A.; Tamas, A.; Cserepes, B.; Lubics, A.; Gallyas, F., Jr.; Toth, G.; et al. Pituitary adenylate cyclase activating polypeptide protects cardiomyocytes against oxidative stress-induced apoptosis. Peptides 2006, 27, 87–94. [Google Scholar] [CrossRef] [PubMed]
  218. Ji, H.; Zhang, Y.; Shen, X.D.; Gao, F.; Huang, C.Y.; Abad, C.; Busuttil, R.W.; Waschek, J.A.; Kupiec-Weglinski, J.W. Neuropeptide PACAP in mouse liver ischemia and reperfusion injury: Immunomodulation by the cAMP-PKA pathway. Hepatology 2013, 57, 1225–1237. [Google Scholar] [CrossRef] [PubMed]
  219. Alston, E.N.; Parrish, D.C.; Hasan, W.; Tharp, K.; Pahlmeyer, L.; Habecker, B.A. Cardiac ischemia-reperfusion regulates sympathetic neuropeptide expression through gp130-dependent and independent mechanisms. Neuropeptides 2011, 45, 33–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Mori, H.; Nakamachi, T.; Ohtaki, H.; Yofu, S.; Sato, A.; Endo, K.; Iso, Y.; Suzuki, H.; Takeyama, Y.; Shintani, N.; et al. Cardioprotective effect of endogenous pituitary adenylate cyclase-activating polypeptide on Doxorubicin-induced cardiomyopathy in mice. Circ. J. 2010, 74, 1183–1190. [Google Scholar] [CrossRef] [PubMed]
  221. Szakaly, P.; Laszlo, E.; Kovacs, K.; Racz, B.; Horvath, G.; Ferencz, A.; Lubics, A.; Kiss, P.; Tamas, A.; Brubel, R.; et al. Mice deficient in pituitary adenylate cyclase activating polypeptide (PACAP) show increased susceptibility to in vivo renal ischemia/reperfusion injury. Neuropeptides 2011, 45, 113–121. [Google Scholar] [CrossRef]
  222. Brar, B.K.; Stephanou, A.; Liao, Z.; O’Leary, R.M.; Pennica, D.; Yellon, D.M.; Latchman, D.S. Cardiotrophin-1 can protect cardiac myocytes from injury when added both prior to simulated ischaemia and at reoxygenation. Cardiovasc. Res. 2001, 51, 265–274. [Google Scholar] [CrossRef] [Green Version]
  223. Yoshida, K.; Taga, T.; Saito, M.; Suematsu, S.; Kumanogoh, A.; Tanaka, T.; Fujiwara, H.; Hirata, M.; Yamagami, T.; Nakahata, T.; et al. Targeted disruption of gp130, a common signal transducer for the interleukin 6 family of cytokines, leads to myocardial and hematological disorders. Proc. Natl. Acad. Sci. USA 1996, 93, 407–411. [Google Scholar] [CrossRef]
  224. Khan, A.M.; Li, M.; Abdulnour-Nakhoul, S.; Maderdrut, J.L.; Simon, E.E.; Batuman, V. Delayed administration of pituitary adenylate cyclase-activating polypeptide 38 ameliorates renal ischemia/reperfusion injury in mice by modulating Toll-like receptors. Peptides 2012, 38, 395–403. [Google Scholar] [CrossRef] [PubMed]
  225. Li, M.; Balamuthusamy, S.; Khan, A.M.; Maderdrut, J.L.; Simon, E.E.; Batuman, V. Pituitary adenylate cyclase-activating polypeptide ameliorates cisplatin-induced acute kidney injury. Peptides 2010, 31, 592–602. [Google Scholar] [CrossRef] [PubMed]
  226. Peuskens, J.; Devoitille, J.M.; Kusters, J.; Vandenhoven, G. An open multicentre pilot study examining the safety, efficacy and tolerability of fast titrated (800 mg/day by day 4) quetiapine in the treatment of schizophrenia/schizoaffective disorder. Int. J. Psychiatry Clin. Pract. 2008, 12, 261–267. [Google Scholar] [CrossRef] [PubMed]
  227. Abylkassov, R.; Xie, Y. Role of Yes-associated protein in cancer: An update. Oncol. Lett. 2016, 12, 2277–2282. [Google Scholar] [CrossRef] [PubMed]
  228. Choi, D.W. Methods for antagonizing glutamate neurotoxicity. Cerebrovasc. Brain Metab. Rev. 1990, 2, 105–147. [Google Scholar] [PubMed]
  229. Won, S.J.; Kim, D.Y.; Gwag, B.J. Cellular and molecular pathways of ischemic neuronal death. J. Biochem. Mol. Biol. 2002, 35, 67–86. [Google Scholar] [CrossRef] [PubMed]
  230. Dijk, F.; Kamphuis, W. Ischemia-induced alterations of AMPA-type glutamate receptor subunit. Expression patterns in the rat retina—An immunocytochemical study. Brain Res. 2004, 997, 207–221. [Google Scholar] [CrossRef]
  231. Sucher, N.J.; Lipton, S.A.; Dreyer, E.B. Molecular basis of glutamate toxicity in retinal ganglion cells. Vis. Res. 1997, 37, 3483–3493. [Google Scholar] [CrossRef] [Green Version]
  232. Roy, S.; Kern, T.S.; Song, B.; Stuebe, C. Mechanistic Insights into Pathological Changes in the Diabetic Retina: Implications for Targeting Diabetic Retinopathy. Am. J. Pathol. 2017, 187, 9–19. [Google Scholar] [CrossRef]
  233. Atlasz, T.; Babai, N.; Reglodi, D.; Kiss, P.; Tamas, A.; Bari, F.; Domoki, F.; Gabriel, R. Diazoxide is protective in the rat retina against ischemic injury induced by bilateral carotid occlusion and glutamate-induced degeneration. Neurotox Res. 2007, 12, 105–111. [Google Scholar] [CrossRef]
  234. Seki, S.; Aoki, M.; Hosokawa, T.; Saito, T.; Masuma, R.; Komori, M.; Kurasaki, M. Bisphenol-A suppresses neurite extension due to inhibition of phosphorylation of mitogen-activated protein kinase in PC12 cells. Chem. Biol. Interact. 2011, 194, 23–30. [Google Scholar] [CrossRef] [PubMed]
  235. Vaczy, A.; Reglodi, D.; Somoskeoy, T.; Kovacs, K.; Lokos, E.; Szabo, E.; Tamas, A.; Atlasz, T. The Protective Role of PAC1-Receptor Agonist Maxadilan in BCCAO-Induced Retinal Degeneration. J. Mol. Neurosci. 2016, 60, 186–194. [Google Scholar] [CrossRef] [PubMed]
  236. Szabo, A.; Danyadi, B.; Bognar, E.; Szabadfi, K.; Fabian, E.; Kiss, P.; Mester, L.; Manavalan, S.; Atlasz, T.; Gabriel, R.; et al. Effect of PACAP on MAP kinases, Akt and cytokine expressions in rat retinal hypoperfusion. Neurosci. Lett. 2012, 523, 93–98. [Google Scholar] [CrossRef] [PubMed]
  237. D’Alessandro, A.; Cervia, D.; Catalani, E.; Gevi, F.; Zolla, L.; Casini, G. Protective effects of the neuropeptides PACAP, substance P and the somatostatin analogue octreotide in retinal ischemia: A metabolomic analysis. Mol. Biosyst. 2014, 10, 1290–1304. [Google Scholar] [CrossRef] [PubMed]
  238. Verma, N.; Pink, M.; Boland, S.; Rettenmeier, A.W.; Schmitz-Spanke, S. Benzo[a]pyrene-induced metabolic shift from glycolysis to pentose phosphate pathway in the human bladder cancer cell line RT4. Sci. Rep. 2017, 7, 9773. [Google Scholar] [CrossRef] [PubMed]
  239. Wada, Y.; Nakamachi, T.; Endo, K.; Seki, T.; Ohtaki, H.; Tsuchikawa, D.; Hori, M.; Tsuchida, M.; Yoshikawa, A.; Matkovits, A.; et al. PACAP attenuates NMDA-induced retinal damage in association with modulation of the microglia/macrophage status into an acquired deactivation subtype. J. Mol. Neurosci. 2013, 51, 493–502. [Google Scholar] [CrossRef] [PubMed]
  240. Qiu, Y.; Zeltzer, S.; Zhang, Y.; Wang, F.; Chen, G.H.; Dayrit, J.; Murdock, B.J.; Bhan, U.; Toews, G.B.; Osterholzer, J.J.; et al. Early induction of CCL7 downstream of TLR9 signaling promotes the development of robust immunity to cryptococcal infection. J. Immunol. 2012, 188, 3940–3948. [Google Scholar] [CrossRef]
  241. Suk, K.; Park, J.H.; Lee, W.H. Neuropeptide PACAP inhibits hypoxic activation of brain microglia: A protective mechanism against microglial neurotoxicity in ischemia. Brain Res. 2004, 1026, 151–156. [Google Scholar] [CrossRef]
  242. Tang, Y.; Le, W. Differential Roles of M1 and M2 Microglia in Neurodegenerative Diseases. Mol. Neurobiol. 2016, 53, 1181–1194. [Google Scholar] [CrossRef]
  243. Stetler, R.A.; Gao, Y.; Zukin, R.S.; Vosler, P.S.; Zhang, L.; Zhang, F.; Cao, G.; Bennett, M.V.; Chen, J. Apurinic/apyrimidinic endonuclease APE1 is required for PACAP-induced neuroprotection against global cerebral ischemia. Proc. Natl. Acad. Sci. USA 2010, 107, 3204–3209. [Google Scholar] [CrossRef] [Green Version]
  244. Mercer, A.; Ronnholm, H.; Holmberg, J.; Lundh, H.; Heidrich, J.; Zachrisson, O.; Ossoinak, A.; Frisen, J.; Patrone, C. PACAP promotes neural stem cell proliferation in adult mouse brain. J. Neurosci. Res. 2004, 76, 205–215. [Google Scholar] [CrossRef] [PubMed]
  245. Yang, J.; Shi, Q.D.; Song, T.B.; Feng, G.F.; Zang, W.J.; Zong, C.H.; Chang, L. Vasoactive intestinal peptide increases VEGF expression to promote proliferation of brain vascular endothelial cells via the cAMP/PKA pathway after ischemic insult in vitro. Peptides 2013, 42, 105–111. [Google Scholar] [CrossRef] [PubMed]
  246. Darsalia, V.; Mansouri, S.; Wolbert, P.; Barde, S.; Sjoholm, A.; Patrone, C. The specific VPAC2 agonist Bay 55-9837 increases neuronal damage and hemorrhagic transformation after stroke in type 2 diabetic rats. Neuropeptides 2013, 47, 133–137. [Google Scholar] [CrossRef] [PubMed]
  247. Shi, Y.; Vanhoutte, P.M. Macro- and microvascular endothelial dysfunction in diabetes. J. Diabetes 2017, 9, 434–449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Kaneko, Y.; Tuazon, J.P.; Ji, X.; Borlongan, C.V. Pituitary Adenylate Cyclase Activating Polypeptide Elicits Neuroprotection Against Acute Ischemic Neuronal Cell Death Associated with NMDA Receptors. Cell Physiol. Biochem. 2018, 51, 1982–1995. [Google Scholar] [CrossRef] [PubMed]
  249. Gomariz, R.P.; Juarranz, Y.; Abad, C.; Arranz, A.; Leceta, J.; Martinez, C. VIP-PACAP system in immunity: New insights for multitarget therapy. Ann. N. Y. Acad. Sci. 2006, 1070, 51–74. [Google Scholar] [CrossRef] [PubMed]
  250. Pozo, D.; Delgado, M.; Martinez, C.; Gomariz, R.P.; Guerrero, J.M.; Calvo, J.R. Functional characterization and mRNA expression of pituitary adenylate cyclase activating polypeptide (PACAP) type I receptors in rat peritoneal macrophages. Biochim. Biophys. Acta 1997, 1359, 250–262. [Google Scholar] [CrossRef] [Green Version]
  251. Ottaway, C.A. In vitro alteration of receptors for vasoactive intestinal peptide changes the in vivo localization of mouse T cells. J. Exp. Med. 1984, 160, 1054–1069. [Google Scholar] [CrossRef] [PubMed]
  252. Sakakibara, H.; Shima, K.; Said, S.I. Characterization of vasoactive intestinal peptide receptors on rat alveolar macrophages. Am. J. Physiol. 1994, 267, L256–L262. [Google Scholar] [CrossRef] [PubMed]
  253. Wiik, P.; Opstad, P.K.; Boyum, A. Binding of vasoactive intestinal polypeptide (VIP) by human blood monocytes: Demonstration of specific binding sites. Regul. Pept. 1985, 12, 145–153. [Google Scholar] [CrossRef]
  254. Lara-Marquez, M.L.; O’Dorisio, M.S.; Karacay, B. Vasoactive intestinal peptide (VIP) receptor type 2 (VPAC2) is the predominant receptor expressed in human thymocytes. Ann. N. Y. Acad. Sci. 2000, 921, 45–54. [Google Scholar] [CrossRef] [PubMed]
  255. Martinez, C.; Juarranz, Y.; Abad, C.; Arranz, A.; Miguel, B.G.; Rosignoli, F.; Leceta, J.; Gomariz, R.P. Analysis of the role of the PAC1 receptor in neutrophil recruitment, acute-phase response, and nitric oxide production in septic shock. J. Leukoc Biol. 2005, 77, 729–738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Delgado, M.; Ganea, D. Vasoactive intestinal peptide and pituitary adenylate cyclase-activating polypeptide inhibit interleukin-12 transcription by regulating nuclear factor kappaB and Ets activation. J. Biol. Chem. 1999, 274, 31930–31940. [Google Scholar] [CrossRef] [PubMed]
  257. Chen, C.; Li, K.; Jiang, H.; Song, F.; Gao, H.; Pan, X.; Shi, B.; Bi, Y.; Wang, H.; Wang, H.; et al. Development of T cells carrying two complementary chimeric antigen receptors against glypican-3 and asialoglycoprotein receptor 1 for the treatment of hepatocellular carcinoma. Cancer Immunol. Immunother. 2017, 66, 475–489. [Google Scholar] [CrossRef] [PubMed]
  258. Abad, C.; Tan, Y.V. Immunomodulatory Roles of PACAP and VIP: Lessons from Knockout Mice. J. Mol. Neurosci. 2018, 66, 102–113. [Google Scholar] [CrossRef]
  259. Martinez, C.; Abad, C.; Delgado, M.; Arranz, A.; Juarranz, M.G.; Rodriguez-Henche, N.; Brabet, P.; Leceta, J.; Gomariz, R.P. Anti-inflammatory role in septic shock of pituitary adenylate cyclase-activating polypeptide receptor. Proc. Natl. Acad. Sci. USA 2002, 99, 1053–1058. [Google Scholar] [CrossRef] [Green Version]
  260. Goetzl, E.J.; Voice, J.K.; Shen, S.; Dorsam, G.; Kong, Y.; West, K.M.; Morrison, C.F.; Harmar, A.J. Enhanced delayed-type hypersensitivity and diminished immediate-type hypersensitivity in mice lacking the inducible VPAC(2) receptor for vasoactive intestinal peptide. Proc. Natl. Acad. Sci. USA 2001, 98, 13854–13859. [Google Scholar] [CrossRef]
  261. Jongsma, H.; Pettersson, L.M.; Zhang, Y.; Reimer, M.K.; Kanje, M.; Waldenstrom, A.; Sundler, F.; Danielsen, N. Markedly reduced chronic nociceptive response in mice lacking the PAC1 receptor. Neuroreport 2001, 12, 2215–2219. [Google Scholar] [CrossRef]
  262. Jana, B.; Palus, K.; Meller, K.; Calka, J. Porcine dorsal root ganglia ovarian neurons are affected by long lasting testosterone treatment. Physiol. Res. 2016, 65, 1019–1030. [Google Scholar]
  263. Kozlowska, A.; Mikolajczyk, A.; Adamiak, Z.; Majewski, M. Distribution and chemical coding of sensory neurons innervating the skin of the porcine hindlimb. Neuropeptides 2017, 61, 1–14. [Google Scholar] [CrossRef]
  264. Moller, K.; Zhang, Y.Z.; Hakanson, R.; Luts, A.; Sjolund, B.; Uddman, R.; Sundler, F. Pituitary adenylate cyclase activating peptide is a sensory neuropeptide: Immunocytochemical and immunochemical evidence. Neuroscience 1993, 57, 725–732. [Google Scholar] [CrossRef]
  265. Dun, E.C.; Huang, R.L.; Dun, S.L.; Dun, N.J. Pituitary adenylate cyclase activating polypeptide-immunoreactivity in human spinal cord and dorsal root ganglia. Brain Res. 1996, 721, 233–237. [Google Scholar] [CrossRef]
  266. Zhang, Y.; Danielsen, N.; Sundler, F.; Mulder, H. Pituitary adenylate cyclase-activating peptide is upregulated in sensory neurons by inflammation. Neuroreport 1998, 9, 2833–2836. [Google Scholar] [CrossRef] [PubMed]
  267. Mulder, H.; Jongsma, H.; Zhang, Y.; Gebre-Medhin, S.; Sundler, F.; Danielsen, N. Pituitary adenylate cyclase-activating polypeptide and islet amyloid polypeptide in primary sensory neurons: Functional implications from plasticity in expression on nerve injury and inflammation. Mol. Neurobiol. 1999, 19, 229–253. [Google Scholar] [CrossRef] [PubMed]
  268. Wang, Z.Y.; Alm, P.; Hakanson, R. The contribution of nitric oxide to endotoxin-induced ocular inflammation: Interaction with sensory nerve fibres. Br. J. Pharmacol 1996, 118, 1537–1543. [Google Scholar] [CrossRef] [PubMed]
  269. Shioda, S.; Takenoya, F.; Hirabayashi, T.; Wada, N.; Seki, T.; Nonaka, N.; Nakamachi, T. Effects of PACAP on Dry Eye Symptoms, and Possible Use for Therapeutic Application. J. Mol. Neurosci 2019, 68, 420–426. [Google Scholar] [CrossRef] [PubMed]
  270. Kazanis, I. The subependymal zone neurogenic niche: A beating heart in the centre of the brain: How plastic is adult neurogenesis? Opportunities for therapy and questions to be addressed. Brain 2009, 132, 2909–2921. [Google Scholar] [CrossRef] [PubMed]
  271. Mercier, F. Fractones: Extracellular matrix niche controlling stem cell fate and growth factor activity in the brain in health and disease. Cell Mol. Life Sci. 2016, 73, 4661–4674. [Google Scholar] [CrossRef] [PubMed]
  272. Matsuno, R.; Ohtaki, H.; Nakamachi, T.; Watanabe, J.; Yofu, S.; Hayashi, D.; Takeda, T.; Nonaka, N.; Seki, M.; Nakamura, M.; et al. Distribution and localization of pituitary adenylate cyclase-activating polypeptide-specific receptor (PAC1R) in the rostral migratory stream of the infant mouse brain. Regul. Pept. 2008, 145, 80–87. [Google Scholar] [CrossRef] [PubMed]
  273. Lin, C.H.; Chiu, L.; Lee, H.T.; Chiang, C.W.; Liu, S.P.; Hsu, Y.H.; Lin, S.Z.; Hsu, C.Y.; Hsieh, C.H.; Shyu, W.C. PACAP38/PAC1 signaling induces bone marrow-derived cells homing to ischemic brain. Stem Cells 2015, 33, 1153–1172. [Google Scholar] [CrossRef]
  274. Suh, J.; Lu, N.; Nicot, A.; Tatsuno, I.; DiCicco-Bloom, E. PACAP is an anti-mitogenic signal in developing cerebral cortex. Nat. Neurosci. 2001, 4, 123–124. [Google Scholar] [CrossRef] [PubMed]
  275. Mansouri, S.; Agartz, I.; Ogren, S.O.; Patrone, C.; Lundberg, M. PACAP Protects Adult Neural Stem Cells from the Neurotoxic Effect of Ketamine Associated with Decreased Apoptosis, ER Stress and mTOR Pathway Activation. PLoS ONE 2017, 12, e0170496. [Google Scholar] [CrossRef] [PubMed]
  276. Nyisztor, Z.; Denes, V.; Kovacs-Valasek, A.; Hideg, O.; Berta, G.; Gabriel, R. Pituitary Adenylate Cyclase Activating Polypeptide (PACAP1-38) Exerts Both Pro and Anti-Apoptotic Effects on Postnatal Retinal Development in Rat. Neuroscience 2018, 385, 59–66. [Google Scholar] [CrossRef] [PubMed]
  277. Nishimoto, M.; Furuta, A.; Aoki, S.; Kudo, Y.; Miyakawa, H.; Wada, K. PACAP/PAC1 autocrine system promotes proliferation and astrogenesis in neural progenitor cells. Glia 2007, 55, 317–327. [Google Scholar] [CrossRef] [PubMed]
  278. Njaine, B.; Rocha-Martins, M.; Vieira-Vieira, C.H.; De-Melo, L.D.; Linden, R.; Braas, K.; May, V.; Martins, R.A.; Silveira, M.S. Pleiotropic functions of pituitary adenylyl cyclase-activating polypeptide on retinal ontogenesis: Involvement of KLF4 in the control of progenitor cell proliferation. J. Mol. Neurosci. 2014, 54, 430–442. [Google Scholar] [CrossRef] [PubMed]
  279. Hansel, D.E.; Eipper, B.A.; Ronnett, G.V. Regulation of olfactory neurogenesis by amidated neuropeptides. J. Neurosci. Res. 2001, 66, 1–7. [Google Scholar] [CrossRef] [PubMed]
  280. Xu, Z.; Ohtaki, H.; Watanabe, J.; Miyamoto, K.; Murai, N.; Sasaki, S.; Matsumoto, M.; Hashimoto, H.; Hiraizumi, Y.; Numazawa, S.; et al. Pituitary adenylate cyclase-activating polypeptide (PACAP) contributes to the proliferation of hematopoietic progenitor cells in murine bone marrow via PACAP-specific receptor. Sci. Rep. 2016, 6, 22373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  281. Pesce, M.; Canipari, R.; Ferri, G.L.; Siracusa, G.; De Felici, M. Pituitary adenylate cyclase-activating polypeptide (PACAP) stimulates adenylate cyclase and promotes proliferation of mouse primordial germ cells. Development 1996, 122, 215–221. [Google Scholar] [PubMed]
  282. McManus, M.F.; Chen, L.C.; Vallejo, I.; Vallejo, M. Astroglial differentiation of cortical precursor cells triggered by activation of the cAMP-dependent signaling pathway. J. Neurosci. 1999, 19, 9004–9015. [Google Scholar] [CrossRef] [PubMed]
  283. Lee, M.; Lelievre, V.; Zhao, P.; Torres, M.; Rodriguez, W.; Byun, J.Y.; Doshi, S.; Ioffe, Y.; Gupta, G.; de los Monteros, A.E.; et al. Pituitary adenylyl cyclase-activating polypeptide stimulates DNA synthesis but delays maturation of oligodendrocyte progenitors. J. Neurosci. 2001, 21, 3849–3859. [Google Scholar] [CrossRef] [PubMed]
  284. Masmoudi-Kouki, O.; Gandolfo, P.; Castel, H.; Leprince, J.; Fournier, A.; Dejda, A.; Vaudry, H.; Tonon, M.C. Role of PACAP and VIP in astroglial functions. Peptides 2007, 28, 1753–1760. [Google Scholar] [CrossRef] [PubMed]
  285. Falluel-Morel, A.; Chafai, M.; Vaudry, D.; Basille, M.; Cazillis, M.; Aubert, N.; Louiset, E.; de Jouffrey, S.; Le Bigot, J.F.; Fournier, A.; et al. The neuropeptide pituitary adenylate cyclase-activating polypeptide exerts anti-apoptotic and differentiating effects during neurogenesis: Focus on cerebellar granule neurones and embryonic stem cells. J. Neuroendocrinol. 2007, 19, 321–327. [Google Scholar] [CrossRef] [PubMed]
  286. Fila, T.; Trazzi, S.; Crochemore, C.; Bartesaghi, R.; Ciani, E. Lot1 is a key element of the pituitary adenylate cyclase-activating polypeptide (PACAP)/cyclic AMP pathway that negatively regulates neuronal precursor proliferation. J. Biol. Chem. 2009, 284, 15325–15338. [Google Scholar] [CrossRef] [PubMed]
  287. Allais, A.; Burel, D.; Isaac, E.R.; Gray, S.L.; Basille, M.; Ravni, A.; Sherwood, N.M.; Vaudry, H.; Gonzalez, B.J. Altered cerebellar development in mice lacking pituitary adenylate cyclase-activating polypeptide. Eur. J. Neurosci. 2007, 25, 2604–2618. [Google Scholar] [CrossRef] [PubMed]
  288. Hirose, M.; Niewiadomski, P.; Tse, G.; Chi, G.C.; Dong, H.; Lee, A.; Carpenter, E.M.; Waschek, J.A. Pituitary adenylyl cyclase-activating peptide counteracts hedgehog-dependent motor neuron production in mouse embryonic stem cell cultures. J. Neurosci. Res. 2011, 89, 1363–1374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  289. Njaine, B.; Martins, R.A.; Santiago, M.F.; Linden, R.; Silveira, M.S. Pituitary adenylyl cyclase-activating polypeptide controls the proliferation of retinal progenitor cells through downregulation of cyclin D1. Eur. J. Neurosci. 2010, 32, 311–321. [Google Scholar] [CrossRef] [PubMed]
  290. Freson, K. From patient to mouse to therapy: Role of the neuropeptide pacap in platelet function and formation. Verh K Acad. Geneeskd Belg. 2010, 72, 239–251. [Google Scholar] [PubMed]
  291. Merlo, L.M.; Pepper, J.W.; Reid, B.J.; Maley, C.C. Cancer as an evolutionary and ecological process. Nat. Rev. Cancer 2006, 6, 924–935. [Google Scholar] [CrossRef]
  292. Capp, J.P. Cancer Stem Cells: From Historical Roots to a New Perspective. J. Oncol. 2019, 2019, 5189232. [Google Scholar] [CrossRef]
  293. Wang, X. Stem cells in tissues, organoids, and cancers. Cell Mol. Life Sci. 2019. [Google Scholar] [CrossRef] [PubMed]
  294. Cauvin, A.; Buscail, L.; Gourlet, P.; De Neef, P.; Gossen, D.; Arimura, A.; Miyata, A.; Coy, D.H.; Robberecht, P.; Christophe, J. The novel VIP-like hypothalamic polypeptide PACAP interacts with high affinity receptors in the human neuroblastoma cell line NB-OK. Peptides 1990, 11, 773–777. [Google Scholar] [CrossRef]
  295. Vertongen, P.; De Clerck, P.; Fournet, J.C.; Martelli, H.; Helardot, P.; Devalck, C.; Peeters, T.; Sariban, E.; Robberecht, P. Comparison between vasoactive intestinal polypeptide and pituitary adenylate cyclase activating polypeptide levels in neuroblastoma tumour tissues. Neuropeptides 1997, 31, 409–413. [Google Scholar] [CrossRef]
  296. Jaworski, D.M. Expression of pituitary adenylate cyclase-activating polypeptide (PACAP) and the PACAP-selective receptor in cultured rat astrocytes, human brain tumors, and in response to acute intracranial injury. Cell Tissue Res. 2000, 300, 219–230. [Google Scholar] [CrossRef] [PubMed]
  297. Lieu, S.N.; Oh, D.S.; Pisegna, J.R.; Germano, P.M. Neuroendocrine tumors express PAC1 receptors. Ann. N. Y. Acad. Sci. 2006, 1070, 399–404. [Google Scholar] [CrossRef] [PubMed]
  298. Vertongen, P.; Ciccarelli, E.; Woussen-Colle, M.C.; De Neef, P.; Robberecht, P.; Cauvin, A. Pituitary adenylate cyclase-activating polypeptide receptors of types I and II and glucagon-like peptide-I receptors are expressed in the rat medullary carcinoma of the thyroid cell line 6/23. Endocrinology 1994, 135, 1537–1542. [Google Scholar] [CrossRef] [PubMed]
  299. Buscail, L.; Gourlet, P.; Cauvin, A.; De Neef, P.; Gossen, D.; Arimura, A.; Miyata, A.; Coy, D.H.; Robberecht, P.; Christophe, J. Presence of highly selective receptors for PACAP (pituitary adenylate cyclase activating peptide) in membranes from the rat pancreatic acinar cell line AR 4-2J. FEBS Lett. 1990, 262, 77–81. [Google Scholar] [CrossRef]
  300. Olcese, J.; McArdle, C.A.; Middendorff, R.; Greenland, K. Pituitary adenylate cyclase-activating peptide and vasoactive intestinal peptide receptor expression in immortalized LHRH neurons. J. Neuroendocrinol. 1997, 9, 937–943. [Google Scholar] [CrossRef] [PubMed]
  301. Jamen, F.; Laden, J.C.; Bouschet, T.; Rodriguez-Henche, N.; Bockaert, J.; Brabet, P. Nerve growth factor upregulates the PAC1 promoter by activating the MAP kinase pathway in rat PC12 cells. Ann. N. Y. Acad. Sci. 2000, 921, 390–394. [Google Scholar] [CrossRef] [PubMed]
  302. Haidan, A.; Hilbers, U.; Bornstein, S.R.; Ehrhart-Bornstein, M. Human adrenocortical NCI-H295 cells express VIP receptors. Steroidogenic effect of vasoactive intestinal peptide (VIP). Peptides 1998, 19, 1511–1517. [Google Scholar] [CrossRef]
  303. Dagar, S.; Sekosan, M.; Rubinstein, I.; Onyuksel, H. Detection of VIP receptors in MNU-induced breast cancer in rats: Implications for breast cancer targeting. Breast Cancer Res. Treat. 2001, 65, 49–54. [Google Scholar] [CrossRef]
  304. Madsen, B.; Georg, B.; Madsen, M.W.; Fahrenkrug, J. Estradiol down regulates expression of vasoactive intestinal polypeptide receptor type-1 in breast cancer cell lines. Mol. Cell Endocrinol. 2001, 172, 203–211. [Google Scholar] [CrossRef]
  305. Waschek, J.A.; Richards, M.L.; Bravo, D.T. Differential expression of VIP/PACAP receptor genes in breast, intestinal, and pancreatic cell lines. Cancer Lett. 1995, 92, 143–149. [Google Scholar] [CrossRef]
  306. Busto, R.; Carrero, I.; Guijarro, L.G.; Solano, R.M.; Zapatero, J.; Noguerales, F.; Prieto, J.C. Expression, pharmacological, and functional evidence for PACAP/VIP receptors in human lung. Am. J. Physiol. 1999, 277, L42–L48. [Google Scholar] [CrossRef] [PubMed]
  307. Garcia-Fernandez, M.O.; Collado, B.; Bodega, G.; Cortes, J.; Ruiz-Villaespesa, A.; Carmena, M.J.; Prieto, J.C. Pituitary adenylate cyclase-activating peptide/vasoactive intestinal peptide receptors in human normal mammary gland and breast cancer tissue. Gynecol. Endocrinol. 2005, 20, 327–333. [Google Scholar] [CrossRef] [PubMed]
  308. Moretti, C.; Mammi, C.; Frajese, G.V.; Mariani, S.; Gnessi, L.; Arizzi, M.; Wannenes, F.; Frajese, G. PACAP and type I PACAP receptors in human prostate cancer tissue. Ann. N. Y. Acad. Sci. 2006, 1070, 440–449. [Google Scholar] [CrossRef]
  309. Robberecht, P.; Vertongen, P.; Velkeniers, B.; de Neef, P.; Vergani, P.; Raftopoulos, C.; Brotchi, J.; Hooghe-Peters, E.L.; Christophe, J. Receptors for pituitary adenylate cyclase activating peptides in human pituitary adenomas. J. Clin. Endocrinol. Metab. 1993, 77, 1235–1239. [Google Scholar] [PubMed]
  310. Vertongen, P.; d’Haens, J.; Michotte, A.; Velkeniers, B.; van Rampelbergh, J.; Svoboda, M.; Robberecht, P. Expression of pituitary adenylate cyclase activating polypeptide and receptors in human brain tumors. Peptides 1995, 16, 713–719. [Google Scholar] [CrossRef]
  311. Reubi, J.C.; Laderach, U.; Waser, B.; Gebbers, J.O.; Robberecht, P.; Laissue, J.A. Vasoactive intestinal peptide/pituitary adenylate cyclase-activating peptide receptor subtypes in human tumors and their tissues of origin. Cancer Res. 2000, 60, 3105–3112. [Google Scholar] [PubMed]
  312. Isobe, K.; Kaneko, M.; Kaneko, S.; Nissato, S.; Nanmoku, T.; Takekoshi, K.; Okuda, Y.; Kawakami, Y. Expression of mRNAs for PACAP and its receptor in human neuroblastomas and their relationship to catecholamine synthesis. Regul. Pept. 2004, 123, 29–32. [Google Scholar] [CrossRef]
  313. Waschek, J.A.; Lelievre, V.; Bravo, D.T.; Nguyen, T.; Muller, J.M. Retinoic acid regulation of the VIP and PACAP autocrine ligand and receptor system in human neuroblastoma cell lines. Peptides 1997, 18, 835–841. [Google Scholar] [CrossRef]
  314. Takahashi, K.; Totsune, K.; Murakami, O.; Sone, M.; Itoi, K.; Hayashi, Y.; Ohi, R.; Mouri, T. Pituitary adenylate cyclase activating polypeptide (PACAP)-like immunoreactivity in ganglioneuroblastoma and neuroblastoma. Regul. Pept. 1993, 49, 19–24. [Google Scholar] [CrossRef]
  315. Spada, A.; Lania, A.; Mantovani, S. Cellular abnormalities in pituitary tumors. Metabolism 1996, 45 (Suppl. 1), 46–48. [Google Scholar] [CrossRef]
  316. Garcia-Fernandez, M.O.; Bodega, G.; Ruiz-Villaespesa, A.; Cortes, J.; Prieto, J.C.; Carmena, M.J. PACAP expression and distribution in human breast cancer and healthy tissue. Cancer Lett. 2004, 205, 189–195. [Google Scholar] [CrossRef] [PubMed]
  317. Fahrenkrug, J.; Buhl, T.; Hannibal, J. PreproPACAP-derived peptides occur in VIP-producing tumours and co-exist with VIP. Regul. Pept. 1995, 58, 89–98. [Google Scholar] [CrossRef]
  318. Odum, L.; Fahrenkrug, J. Pituitary adenylate cyclase activating polypeptide (PACAP) in human ovarian cancers. Cancer Lett. 1998, 125, 185–189. [Google Scholar] [CrossRef]
  319. Gutierrez-Canas, I.; Rodriguez-Henche, N.; Bolanos, O.; Carmena, M.J.; Prieto, J.C.; Juarranz, M.G. VIP and PACAP are autocrine factors that protect the androgen-independent prostate cancer cell line PC-3 from apoptosis induced by serum withdrawal. Br. J. Pharmacol. 2003, 139, 1050–1058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  320. Godlewski, J.; Kowalczyk, A.; Kozielec, Z.; Pidsudko, Z.; Kmiec, A.; Siedlecka-Kroplewska, K. Plasticity of neuropeptidergic neoplasm cells in the primary and metastatic Merkel cell carcinoma. Folia Histochem. Cytobiol. 2013, 51, 168–173. [Google Scholar] [CrossRef]
  321. Nakamachi, T.; Sugiyama, K.; Watanabe, J.; Imai, N.; Kagami, N.; Hori, M.; Arata, S.; Shioda, S. Comparison of expression and proliferative effect of pituitary adenylate cyclase-activating polypeptide (PACAP) and its receptors on human astrocytoma cell lines. J. Mol. Neurosci. 2014, 54, 388–394. [Google Scholar] [CrossRef]
  322. Zia, F.; Fagarasan, M.; Bitar, K.; Coy, D.H.; Pisegna, J.R.; Wank, S.A.; Moody, T.W. Pituitary adenylate cyclase activating peptide receptors regulate the growth of non-small cell lung cancer cells. Cancer Res. 1995, 55, 4886–4891. [Google Scholar]
  323. Castorina, A.; Giunta, S.; Scuderi, S.; D’Agata, V. Involvement of PACAP/ADNP signaling in the resistance to cell death in malignant peripheral nerve sheath tumor (MPNST) cells. J. Mol. Neurosci. 2012, 48, 674–683. [Google Scholar] [CrossRef]
  324. Vertongen, P.; Camby, I.; Darro, F.; Kiss, R.; Robberecht, P. VIP and pituitary adenylate cyclase activating polypeptide (PACAP) have an antiproliferative effect on the T98G human glioblastoma cell line through interaction with VIP2 receptor. Neuropeptides 1996, 30, 491–496. [Google Scholar] [CrossRef]
  325. Li, M.; Cortez, S.; Nakamachi, T.; Batuman, V.; Arimura, A. Pituitary adenylate cyclase-activating polypeptide is a potent inhibitor of the growth of light chain-secreting human multiple myeloma cells. Cancer Res. 2006, 66, 8796–8803. [Google Scholar] [CrossRef] [PubMed]
  326. Ajioka, I.; Martins, R.A.; Bayazitov, I.T.; Donovan, S.; Johnson, D.A.; Frase, S.; Cicero, S.A.; Boyd, K.; Zakharenko, S.S.; Dyer, M.A. Differentiated horizontal interneurons clonally expand to form metastatic retinoblastoma in mice. Cell 2007, 131, 378–390. [Google Scholar] [CrossRef]
  327. Denes, V.; Hideg, O.; Nyisztor, Z.; Lakk, M.; Berta, G.; Toth, K.; Gabriel, R. The neuroprotective peptide, PACAP signaling could be critical in retinoblastoma through induction of horizontal cell proliferation. Presented at European Retina Meeting, Helsinki, Finland, 12–14 September 2019. [Google Scholar]
  328. Rubio, G.; Marin, M.; Arias, F.; Lopez-Trabada, J.R.; Iribarren, M.; Alfonso, S.; Prieto, R.; Blanco, A.; Urosa, B.; Montes, V.; et al. Inclusion of Alcoholic Associations Into a Public Treatment Programme for Alcoholism Improves Outcomes During the Treatment and Continuing Care Period: A 6-Year Experience. Alcohol. Alcohol. 2018, 53, 78–88. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Core PACAP1-27 sequence homology within the human PACAP/glucagon family (PACAP paralogs). Peptide names and abbreviations are indicated at left. Amino acids are given in the single letter code, the numbers on the top show their positions. Sequence comparison was performed by BLASTP v.2.9.0 search on Non-redundant GenBank CDS translations + PDB + SwissProt + PIR + PRF databases. The numbers at right indicate percent homologies. Accession numbers: PACAP, (preproprotein) NP_001093203, VIP, AAB22264.1, PHI, 1010243A, Secretin, AAG31443.1, GHRH, AAH99727.1, Glucagon, NP_002045.1, PRP, (preproprotein) NP_001093203 [10].
Figure 1. Core PACAP1-27 sequence homology within the human PACAP/glucagon family (PACAP paralogs). Peptide names and abbreviations are indicated at left. Amino acids are given in the single letter code, the numbers on the top show their positions. Sequence comparison was performed by BLASTP v.2.9.0 search on Non-redundant GenBank CDS translations + PDB + SwissProt + PIR + PRF databases. The numbers at right indicate percent homologies. Accession numbers: PACAP, (preproprotein) NP_001093203, VIP, AAB22264.1, PHI, 1010243A, Secretin, AAG31443.1, GHRH, AAH99727.1, Glucagon, NP_002045.1, PRP, (preproprotein) NP_001093203 [10].
Jcm 08 01488 g001
Figure 2. Comparison of species specific PACAP1-38 amino acid sequences (PACAP orthologs). The species names are indicated on the left. Percent homologies are calculated and shown in the “%” column on the left. Amino acids are given in the single letter code. The numbers on the top show their positions. Sequence comparison was performed by BLASTP v.2.9.0 search on Non-redundant GenBank CDS translations + PDB + SwissProt + PIR + PRF databases. Accession numbers: Human PACAP, (preproprotein) NP_001093203, Mouse, BAA28355, Rat, NP_058685, Sheep, AAB21469.1, Chicken, AAX56089.1, Frog (Xenopus), AAD56956.1, Salmon (Oncorhynchus), P41585.1, Catfish (Ictalurus punctatus), AAK66970.1, and Tunicate (Chelyosoma productum) [10].
Figure 2. Comparison of species specific PACAP1-38 amino acid sequences (PACAP orthologs). The species names are indicated on the left. Percent homologies are calculated and shown in the “%” column on the left. Amino acids are given in the single letter code. The numbers on the top show their positions. Sequence comparison was performed by BLASTP v.2.9.0 search on Non-redundant GenBank CDS translations + PDB + SwissProt + PIR + PRF databases. Accession numbers: Human PACAP, (preproprotein) NP_001093203, Mouse, BAA28355, Rat, NP_058685, Sheep, AAB21469.1, Chicken, AAX56089.1, Frog (Xenopus), AAD56956.1, Salmon (Oncorhynchus), P41585.1, Catfish (Ictalurus punctatus), AAK66970.1, and Tunicate (Chelyosoma productum) [10].
Jcm 08 01488 g002
Table 1. Summary of the most relevant PACAP-related disorders.
Table 1. Summary of the most relevant PACAP-related disorders.
DisorderImplicationPACAP/VIP ReceptorOutcomeLiterature
Cancerpathogenesis/therapeutic for details see Section 6.2
Breast cancerpathogenesisPAC1, VPAC1, VPAC2 [28]
Cervical cancertherapeuticPAC1 [29,30]
Colon cancerpathogenesis/therapeuticPAC1 (Hip), VPAC1 [31,32,33]
Gastric cancerpathogenesisPAC1, VPAC1 [34,35]
Gliomaspathogenesis/therapeuticPAC1, VPAC1, VPAC2 [36,37]
Lung canerpathogenesisPAC1, VPAC1, VPAC2 [38]
Malignant peripheral nerve sheath tumor (MPNST)pathogenesisnot determined [39]
Malignant pheochromocytomaspathogenesisPAC1, VPAC1 [40]
MedulloblastomatherapeuticPAC1 [41,42]
Pancreatic cancerpathogenesisPAC1, VPAC1 [43]
Pituitary adenomas [44]
ProstatepathogenesisPAC1 (null), VPAC1 [45,46]
Renal cell carcinomatherapeuticVPAC1 [47]
RetinoblastomaunclearPAC1 [48,49]
Diabetes (Type II, insulin resistance)therapeuticPAC1, VPAC1, VPAC2- induces insulin secretion,
- PAC1 antagonist impairs glucose tolerance
- PACAP overexpression induces beta cell proliferation
[50,51,52,53,54]
Diabetic retinopathytherapeuticPAC1, VPAC1, VPAC2- increases Akt and ERK1/2 phosphorylation
- reduces the activation of p38 mitogen-activated protein kinase
- reduces the expression of IL-1β in diabetic animals
- downregulates VEGF and VEGF receptors, inversely regulates HIFs: downregulating HIF-1α and HIF-2α while upregulating HIF-3α
- restores both Bcl-2 and p53 mRNA and protein expression
- maintains synapse integrity
[55,56,57,58,59,60]
Diabetic nephropathytherapeuticnot determined- downregulates of several cytokines including CINC-1, TIMP-1, LIX, MIG, s-ICAM[61]
InflammationTherapeutic/pathogenesis for details see Section 6.1
Allergic airway inflammationtherapeuticPAC1, VPAC1, VPAC2 [62]
Atherosclerosistherapeuticnot determined [63]
Chronic cystitis, urinary bladder inflammationpathogenesisPAC1, VPAC1, VPAC2 [64,65]
Endotoxin induced airway inflammationtherapeuticPAC1, VPAC1, VPAC2 [66]
Ileitistherapeuticnot determined [67,68]
Ocular inflammationpathogenesisnot determined [69]
Osteoarthritistherapeuticnot determined [70,71]
Pancreatitispathogenesisnot determined- enhances caerulein-induced pancreatitis via downregulation of RegIII-β
- lack of endogenous PACAP ameliorates symptoms
[72,73,74]
Skin neurogenic inflammationpathogenesisPAC1 [75,76,77]
ToxoplasmosistherapeuticVPAC1, VPAC2 [78]
Human immunodeficiency virus (HIV) infectiontherapeuticPAC1, VPAC1, VPAC2- increases macrophage resistance to HIV-1 replication
- reduces macrophage production of HIV-1
- induces the synthesis of β-chemokines and IL-10
- inhibits NF-kB, and reduced Cyclin D1 levels
[79,80,81]
Neurological/neurodegenerative disordersTherapeutic/pathogenesis
Alzheimer’s diseasestherapeuticPAC1for details, see Section 5.1[82,83,84]
Bipolar disordertherapeuticPAC1- no association with SNPs of PACAP gene
- regulates binding between DISC1 and DISC1-binding zinc-finger protein responsible for neurite outgrowth
[85,86]
EpilepsytherapeuticPAC1, VPAC1- during seizure PACAP is secreted and exert neuroprotective effects by regulating microglial phenotype, microglial interleukin secretion
- inhibits long-term depression and depotentiation
[87,88]
Huntington’s diseasestherapeuticPAC1- improves memory performance[89]
Ischemia/stroke (cardiac, renal, hepatic, retinal, brain)therapeuticPAC1for details, see Section 5.2[90,91,92,93,94,95,96,97,98,99]
MigrainepathogenesisPAC1, VPAC1, VPAC2for details, see Section 4.1[100,101,102,103,104,105,106]
Major depression (MDD)therapeuticnot determinedPACAP regulated DISC1 mutation is linked to major MDD
- SNP3 (rs1893154) of the PACAP gene is significantly correlated with MDD
[107,108,109,110].
Multiple sclerosistherapeuticVPAC1, PAC1- prevents Thelper1 cell differentiation and suppression of inflammatory interleukin production[111,112,113,114,115,116,117,118,119]
Parkinson’s diseasestherapeutic - potent preventive molecule against cell loss and autophagy
- facilitates dopaminergic neurotransmission
- effective PACAP analog was developed with reduced cardiovascular side effects
[120,121,122,123]
Post-traumatic stress disorderpathogenesisPAC1 (VPAC2)for details, see Section 4.2[4,124,125]
SchizophreniatherapeuticPAC1- SNPs of PACAP or PAC1 gene are associated with schizophrenia
- regulates binding between DISC1 and DBZ responsible for neurite outgrowth
[86,110,126,127]
Traumatic injuries
Traumatic brain injurytherapeuticnot determined- attenuates neural injury by increasing level of SOD-2 and GPx-1
- suppresses T-cell response
- decreases inflammatory interleukin levels
[128,129]
Spinal cord injurytherapeuticPAC1- activates axon regeneration through CRMP-2 and activation of glial elements[130,131]
Legends: ERK—extracellular signal-regulated kinases; IL—interleukin; VEGF—vascular endothelial growth factor; HIF—hypoxia-inducible factor; CINC-1—cytokine-induced neutrophil chemoattractant; TIMP-1—metallopeptidase inhibitor 1; MIG—monokine induced by gamma interferon; s-ICAM—soluble intercellular adhesion molecule 1; NF-kB—nuclear factor kappa-light-chain-enhancer of activated B cells; SNP—single nucleotide polymorphism; DISC1—Disrupted-in-Schizophrenia-1; SOD-2—superoxide dismutase-2; GPx-1—glutathione peroxidase-1; CRMP-2—collapsin response mediator protein-2.

Share and Cite

MDPI and ACS Style

Denes, V.; Geck, P.; Mester, A.; Gabriel, R. Pituitary Adenylate Cyclase-Activating Polypeptide: 30 Years in Research Spotlight and 600 Million Years in Service. J. Clin. Med. 2019, 8, 1488. https://0-doi-org.brum.beds.ac.uk/10.3390/jcm8091488

AMA Style

Denes V, Geck P, Mester A, Gabriel R. Pituitary Adenylate Cyclase-Activating Polypeptide: 30 Years in Research Spotlight and 600 Million Years in Service. Journal of Clinical Medicine. 2019; 8(9):1488. https://0-doi-org.brum.beds.ac.uk/10.3390/jcm8091488

Chicago/Turabian Style

Denes, Viktoria, Peter Geck, Adrienn Mester, and Robert Gabriel. 2019. "Pituitary Adenylate Cyclase-Activating Polypeptide: 30 Years in Research Spotlight and 600 Million Years in Service" Journal of Clinical Medicine 8, no. 9: 1488. https://0-doi-org.brum.beds.ac.uk/10.3390/jcm8091488

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

Article Metrics

Back to TopTop