Next Article in Journal
Chlamydial Infection-Dependent Synthesis of Sphingomyelin as a Novel Anti-Chlamydial Target of Ceramide Mimetic Compounds
Next Article in Special Issue
Anti-SOD1 Nanobodies That Stabilize Misfolded SOD1 Proteins Also Promote Neurite Outgrowth in Mutant SOD1 Human Neurons
Previous Article in Journal
MicroRNA Changes Up to 24 h following Induced Hypoglycemia in Type 2 Diabetes
Previous Article in Special Issue
Neurodegenerative Disorder Risk in Krabbe Disease Carriers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Overview of the Molecular Cues and Their Intracellular Signaling Shared by Cancer and the Nervous System: From Neurotransmitters to Synaptic Proteins, Anatomy of an All-Inclusive Cooperation

by
Marco Arese
1,2,*,†,
Federico Bussolino
1,2,†,
Margherita Pergolizzi
1,2 and
Laura Bizzozero
1,2
1
Department of Oncology, University of Torino, 10060 Candiolo, TO, Italy
2
Candiolo Cancer Institute, FPO-IRCCS, 10060 Candiolo, TO, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(23), 14695; https://doi.org/10.3390/ijms232314695
Submission received: 18 October 2022 / Revised: 18 November 2022 / Accepted: 18 November 2022 / Published: 24 November 2022

Abstract

:
We propose an overview of the molecular cues and their intracellular signaling involved in the crosstalk between cancer and the nervous system. While “cancer neuroscience” as a field is still in its infancy, the relation between cancer and the nervous system has been known for a long time, and a huge body of experimental data provides evidence that tumor–nervous system connections are widespread. They encompass different mechanisms at different tumor progression steps, are multifaceted, and display some intriguing analogies with the nervous system’s physiological processes. Overall, we can say that many of the paradigmatic “hallmarks of cancer” depicted by Weinberg and Hanahan are affected by the nervous system in a variety of manners.

Graphical Abstract

1. Introduction

Tumors are not only composed of transformed cells; they also contain stable cell types that collectively form a complex biosystem. Endothelial cells, immunocompetent cells, cancer-associated fibroblasts (CAFs), neurons, nerves, and other elements with their products (e.g., the extracellular matrix) influence the transformed cells within the tumor microenvironment (TME) and in such a way affect tumor biology [1].
Targeting the TME rather the tumor cells alone has significant clinical repercussions. Indeed, drugs targeting angiogenesis, CAFs, immunocompetent cells, and some related signaling pathways are in clinical trials (reviewed in [2]). Altogether, it appears that we will better understand and tackle many open questions related to cancer only with a detailed knowledge of microenvironmental factors, their sources, and their impacts. Here, we treat the nervous system as one of these sources.
The concept that the nervous system affects the progression of cancer was initially put forward by the Greek physician Galen, who recognized that “melancholic” women were more prone to developing breast cancer than other women (reviewed in [3,4]). The most recent indications from the burgeoning field of cancer neuroscience [5] suggest that nerves are significant players in the tumor microenvironment, engendering a selective pressure on cancer cells by modulating a complex network of mediators related to tumor progression.
There has been a considerable increase in the published research on cancer–nervous system relations in the last three years, while many excellent reviews in the last year alone [6,7,8,9,10,11] have approached the field from different perspectives.
Nerve and cancer cells relate in many ways, some of which are known only at the descriptive level. Taking into account the “manifest” of cancer neuroscience [5], we can cite the four main interfaces through which cancer and the nervous system influence each other: (1) Paracrine signaling between nerve cells and cancer cells, such as neuronal-activity-dependent neurotransmitter or growth factor release, regulates cancer growth in a variety of tissues. The influence of neurons on malignant cells may be mediated by other cell types in the tumor microenvironment. Paracrine factors derived from cancer remodel the nervous system to promote increased neural activity in the tumor microenvironment. (2) Synaptic communication between neurons and brain cancer cells (e.g., malignant glioma), through voltage-regulated mechanisms, can regulate cancer growth. (3) Cancer-related circulating factors can influence nervous system functions, such as sleep, whereas the nervous system can influence cancer progression via circulating factors, such as hormones and progenitor cells, or altered immune system function. (4) Cancer treatments frequently result in nervous system toxicity, ranging from peripheral neuropathy to cognitive impairment (Figure 1).
Here, we review the molecular links between nerve and cancer cells that promote cancer progression, mainly those classified in the just-cited categories 1 and 2. Sometimes these cues have been “evolutionarily co-opted” by tumor cells from the nervous system and can have at least one function in a “tumor autonomous” manner. Obviously, even in this last category, the presence of neuron-related proteins in cancer cells strongly predicts an ability to communicate with neurons through common signaling systems. The signaling pathways of each molecular cue that we consider are briefly overviewed. Functionally, we pay particular attention to alterations in proliferative, apoptotic, survival, migratory/metastatic, or immune-regulating pathways. In more detail, our review considers: (a) discrete nervous system/cancer interaction sites, such as in the perineural invasion (PNI) process and the perineural niche, the sympathetic nervous system (SNS) and catecholamines, the parasympathetic nervous system (PSNS) and acetylcholine (Ach), and the tumor/nervous system interface in central nervous system tumors and electrochemical communication as well as (b) molecules that have widespread activity in various types of cancer and contexts: neurotransmitters/neuropeptides/amino acids/growth factors (dopamine, gamma-aminobutyric acid, serotonin, glutamate, neuropeptides, serine, and nerve growth factor (NGF)) and axon guidance/synaptic proteins. On this background, the perineural niche, with its ever-expanding role in the paracrine interactions in PNI, and the novel remarkable findings on the electrochemical interactions in brain tumors receive particular attention.

2. Perineural Invasion and the Perineural Niche

Perineural invasion (PNI) is a very well known process by pathologists and one the most actively studied events in cancer–nerve relations at the molecular level. Nonetheless, our current understanding of PNI is very limited and is still highly debated. Macroscopically, it happens as cancer cells infiltrate the surrounding nerve endings, invading them through the destruction of the perineurium and entry via the perforating vessels [12].
While virtually all peripheral cancer types display PNI, the process has mostly been studied for head and neck, prostatic, and pancreatic cancers. For many of these, PNI is linked with considerable morbidity, poor outcomes, and decreased survival (reviewed in [13]).
The discrete microenvironment surrounding an invaded nerve, also called the perineural niche, is formed by neural cells, supporting cells, inflammatory cells, extracellular matrix, blood vessels, and immune components. The discovery of the innervated niche as a new specialized microenvironment has opened exciting new avenues of research into bidirectional neural–cancer interactions [14]. The migration of cancer cells along nerves is supported by multiple growth factors and chemokines secreted by these cell types. Nerve injury and repair, in a continuous process, seem to guide this migration (also called tracking) through the secretion of growth factors and chemokines simulating damaged and dying nerves [15]. Amit et al. [15] proposed that PNI occurs in seven stages: tumor cell survival, the formation of a neural steady state, an inflammatory reaction, the recruitment of tumor cells to nerves, the neogenesis of nerves, the adhesion of tumor cells and neurolemma, and nerve invasion. This theory is supported by data from in vitro studies, as the coculture of human prostate cancer (PCa) cells, stromal cells, and mouse dorsal root ganglia increases tumor colony and neurite outgrowth [16].
The molecular characterization of the perineural niche is a burgeoning field of research [17] that would require a whole review by itself. Here, we concentrate on key cellular and molecular aspects of neurogenesis and invasion, including the newly studied role of autophagy as well as new perspectives in therapy.
Most studies emphasize the importance of neural components in tumors, making it critical to investigate the mechanism of nerve emergence in cancer. Strikingly, neurogenesis is not only a local process. Neural progenitors from the brain that express doublecortin (DCX+) can migrate to prostate tumors and initiate neurogenesis, leading to tumor growth and metastasis [18,19]. High expression of the transcription factor Snail in PCa cells improves adherence to nerve cells and promotes neurite outgrowth [19]. As a result, the neurogenesis program is crucial in prostate tumor–nerve crosstalk. A recent single-cell sequencing study discovered that Neurexin 1 and Neuroligin 1 (see below), two genes encoding neuronal synaptic molecules, are strongly linked to PCa progression [20]. Other neurotrophic factors, such as nerve growth factor (NGF) and brain-derived neurotrophic factor (BDNF), have been linked to PNI [21]. PNI in PCa is thought to be characterized by N-CAM upregulation in nerves. Li et al. proposed that cancer cells induce N-CAM upregulation in nerves via a paracrine loop that accelerates cancer cell migration towards nerves via the nuclear factor kappa B (NFkB) pathway [22]. NFkB is involved in many different types of cancer. In PCa, NFkB nuclear translocation has been found to be critical for inhibiting apoptosis and increased tumor cell proliferation in PNI lesions. Semaphorin 4F (SEMA4F), a member of the semaphorin family of axon guidance molecules (see below), is another molecule that correlates with the NFkB pathway in PNI. The SEMA4F expression level has been linked to nerve density and PNI diameter, in addition to the antiapoptotic effect induced by interacting with the NFkB pathway, as the overexpression of SEMA4F can induce neurogenesis [23]. Another semaphorin family member, semaphorin 3C (SEMA3C), has been shown to be involved in PNI in PCa. SEMA3C is activated by monoamine oxidase A (MAOA) in a Twist-1-dependent manner and interacts with its receptor and coreceptor PlexinA2 as well as neuropilin-1 (NRP1), stimulating c-MET. This SEMA3C/PlexinA2/NRP1-cMET axis promotes PNI in PCa and demonstrates the close relationship between PNI and nerve-related signals [24,25]. Many other molecular cues promoting aggressive behavior have been identified in the perineural niche. In pancreatic cancer, the cell adhesion molecule L1CAM and a metalloproteinase mediate paracrine communication between Schwann cells and cancer cells, thus promoting cancer invasion into nerves [26]. In head and neck cancer, the neuropeptide galanin (GAL) activates its receptor, GALR2, on cancer cells and stimulates the transcription of cyclooxygenase-2 and GAL itself. In turn, the cyclooxygenase-2 product prostaglandin E2 promotes cancer invasion, while the GAL released by cancer stimulates neurite outgrowth, promoting PNI [27]. In the same tumor model, the NGF-TrkA axis triggers the epithelial–mesenchymal transition (EMT) and confers resistance to the epidermal growth factor receptor tyrosine kinase (EGFR) inhibitor erlotinib [28]. Moreover, in PCa, nerve-produced glial cell line derived neurotrophic factor receptor alpha (GFRα1) activates the receptor tyrosine kinase RET, thus promoting transformation-associated phenotypes [29].
From a therapeutic perspective targeting the events at the perineural niche, two approaches are worth mentioning here. One is targeting NGF and its receptors. In prostate cancer, this may reduce nerve infiltration, inhibit tumor cell growth, and alleviate cancer-related pain (reviewed in [30]). Another possible strategy is to target the interaction of the nerve and immune cells in the tumor microenvironment. It has been reported that tumor-associated nerves in prostatic cancer have high levels of programmed cell death ligand-1 (PD-L1) expression and inhibit the function of immune cells, providing a new perspective for the use of immune checkpoint inhibitors [31].
A novel and highly investigated player in nerve cancer crosstalk is autophagy. Autophagy is a conserved cellular self-degradation process that is critical for balancing energy sources during neuronal development and nutrient stress [32,33]. Autophagy has recently been identified as a double-edged sword that plays an important role in the innervated niche [34]. The question of whether upregulated autophagy is beneficial or harmful to tumor progression in the innervated niche is still being debated. Autophagy stimulates tumor progression in part because of its ability to maintain innervated niche homeostasis and even promote innervated niche development [35,36]. In contrast, dysfunctional autophagy may result in axon breakdown, dendritic degradation, somal stress, and glial cell death, ultimately leading to innervated niche disorder [37]. As more is learned about autophagy in the innervated niche, new treatment strategies can be developed. Changing the innervated niche by regulating the autophagy pathway has emerged as a novel therapeutic opportunity for cancer treatment and a novel window for drug repurposing in this context. Following an escalating era for drug discovery, repurposing old drugs to treat new indications is becoming an increasingly appealing proposition because it is a time-saving and cost-effective method with high success rates [38,39]. A large number of preclinical trials have shown that multiple noncancer drugs (antipsychotics, cardiovascular drugs, etc.) have off-label antitumor effects [40] Based on the autophagy-mediated innervated niche, repurposing drugs such as beta-adrenergic antagonists is a promising pharmacological approach in cancer neuroscience [41,42]. Hence, it is now recognized that the innervated niche is a complex and heterogeneous site that can influence cancer development [43]. Finally, the role of a specific type of autophagy in glioblastoma progression should be cited here. Pericytes (PCs), contractile perivascular cells, are seized by glioblastoma (GB) to aid tumor progression. The work of Valdor et al. [44] showed that chaperone-mediated autophagy (CMA), a process that transports specific cytosolic proteins to lysosomes for degradation and has pro-oncogenic effects, is significantly upregulated in PCs in response to the oxidative burst that occurs after PC-GB cell interactions. The upregulation of CMA activity in PCs by GB is required for their effective interaction with GB cells, which aids tumor growth. These findings point to abnormal CMA upregulation as a mechanism by which GB cells elicit the immunosuppressive function of PCs and stabilize the GB-PC interactions required for tumor cell survival.
A schematic representation of the mediators and intracellular pathways of PNI is presented in Figure 2.

3. The Sympathetic Nervous System and Adrenergic Signaling

The SNS regulates gene expression and cellular function throughout the body by releasing two catecholamines: norepinephrine (NE), which is secreted by SNS nerve fibers innervating target tissues, and epinephrine (E), which is a product of the adrenal gland and is released into the blood [46].
Studies of the SNS’s effects on cancer biology were originally driven by clinical observations suggesting a correlation between stress and cancer progression. Chronic stress is linked with an increased frequency of cancer [47] and is associated with poorer clinical consequences [48]. Different experimental models show that stress and increased adrenergic activity promote the progression of a range of malignancies, including prostate, pancreatic, and ovarian cancer [49,50,51,52].
Altogether, E and NE deeply affect multiple biological characteristics of cancers and their microenvironments, including resistance to apoptosis, proliferation, cell survival, invasion, metastasis, and angiogenesis [53,54,55]
E and NE signal through alpha (α)- and beta (β)-adrenergic receptors (α and β-ARs), which are G-protein-coupled receptors (GPCRs) that are broadly expressed in mammalian tissues. In particular, β2-AR are expressed on many tumor cells of epithelial and lymphoid origin. β-ARs signal through the Gαs–cyclic AMP (cAMP) stimulation of mitogen-activated protein kinases (MAPK), protein kinase A (PKA), and exchange protein activated by cAMP (EPAC). The activation of β2-ARs promotes tumor growth through vascular endothelial growth factor (VEGF) and metalloproteases, thus potentiating angiogenesis and metastasization in ovarian, lung, and breast cancer [50].
In esophageal squamous cell carcinoma, β-ARs transactivate the extracellular signal-regulated kinase (ERK)/cyclooxygenase 2 (COX2) signaling pathway, facilitating cell proliferation. In addition, the NE-β-AR axis stimulates hepatocellular carcinoma resistance to anoikis (i.e., the induction of apoptosis in cells upon the loss of adhesion) and invasion through EGFR transactivation [56]. In colon cancer, NE promotes cell proliferation and metastasis by activating the cAMP response element-binding protein-1–microRNA 373 (CREB1-miR-373) axis [57]. Recently [58], the non-receptor tyrosine kinase Src was identified as a key switch that is activated in response to beta-adrenergic signaling in cancer cells.
In pancreatic cancer, NE increases cell viability and invasion and prevents apoptosis in a Notch-1-dependent manner [59]. In the same model, catecholamines promote β2 adrenergic-dependent PDAC (pancreatic ductal adenocarcinoma) development, NGF secretion, and pancreatic nerve density. These findings suggest that catecholamines drive an upregulation of neurotrophins, which in turn increases sympathetic innervation and local NE accumulation [55]. In prostate cancer, NE reactivates the cell cycling of dormant disseminated tumor cells (DTC) through both direct actions and indirectly on adjacent cells in the bone marrow niche [60]. In a lung tumorigenesis model, NE prompts the phosphorylation of L-type voltage-dependent calcium channels (VDCC) through a β-adrenergic receptor-PKA pathway. VDCC triggers calcium mobilization, thereby causing the activation of IGF-1R via the exocytosis of insulin-like growth factor 2 (IGF2) [61].
In the prostate TME, NE inhibits oxidative phosphorylation in endothelial cells while promoting an “angiogenic switch” and cancer progression [62,63].
In an apparent contrast to the data above, exercise-induced catecholamines activate the Hippo tumor suppressor pathway in breast cancer [64]. In this context, both E and NE inhibited breast cancer cell viability as well as tumor growth in vivo. However, it is essential to distinguish between the acute and rapidly falling levels of catecholamines in response to exercise and chronic stress, which causes constant high levels of catecholamines.
The modulation of the tumor immune microenvironment by β2-adrenergic signaling is a key event in the tumor–nerve relations. Indeed, E and NE profoundly affect inflammatory immune cells in cancers [65,66,67]. For example, the activation of β-AR signaling enhances the infiltration of macrophages into the tumor parenchyma and, in so doing, induces a prometastatic gene expression signature in primary breast cancer [68]. Moreover, adrenergic signaling influences the immunosuppressive capacity of myeloid-derived suppressor cells in cancer [69]. Furthermore, the vagus nerve modulation of cytotoxic T cells in the spleen has been shown to contribute to colorectal cancer [9,70].
In melanoma, nonspecific β blockers decrease the recurrence and improve overall survival for metastatic disease treated with immunotherapy. New data show how β blockade improves cytotoxic therapies by regulating the immune response. Concurrent irradiation and beta-blockade also reduce the abundance of immunosuppressive M2-polarized macrophages and regulatory T cells, both of which have been shown to impair cytotoxic T-cell migration and the physical interaction with malignant cells in lung cancer. These findings are consistent with the widespread adrenergic and sensory innervation in lymph nodes as well as the control of T cell egress from lymph nodes via beta 2 adrenergic receptor signaling. Given that a significant portion of the immune response to immune checkpoint blockade occurs outside of the tumor site (e.g., draining lymphoid tissue), the modulation of the neural signaling that controls T-cell retention or release can have significant effects on both the native and immune checkpoint-blockade-induced immune responses [7].

4. The Parasympathetic Nervous System and Acetylcholine

Neurons of the PSNS typically release Ach as their primary neurotransmitter. Data show that while adrenergic signals consistently exert protumorigenic effects, cholinergic (i.e., from Ach) signals play different roles in cancer progression depending on the tumor type. Moreover, the context-dependent effects of parasympathetic signals are less understood at the molecular level, perhaps because of a paucity of tools to precisely target parasympathetic nerves [14].
Two classes of Ach receptors have been isolated: the nicotinic acetylcholine receptors (nAchRs) and the muscarinic receptors (mAchRs). nAchRs are targeted by NNK, a potent tobacco-originated carcinogen [71,72]. They signal through Ca2+ flows inside the cell and enhance the epithelial–mesenchymal transition, proliferation, differentiation, angiogenesis, migration, and invasion [73,74]. nAchR crosstalks with other neurotransmitter receptors, activating multiple secondary pathways, including protein kinase C (PKC), ERK1/2, COX2, prostaglandin E2 (PGE2), CREB, Src, Akt serine-threonine kinase, and the Ras-RAF1-MAPK cascade [75]. nAchRs can also prevent drug-induced apoptosis by upregulating survivin, the apoptosis suppressor gene BCL-2, and nuclear factor NF-κB in breast cancer and pleural mesothelioma.
mAchRs also modulate the progression of different cancers. Ach enhances non-small-cell lung carcinoma (NSCLC) cell proliferation through M3R (a subtype of mAchR)-mediated stimulation of Akt and MAPK [76]. In colon cancer [76] and gastric cancer [77,78], Ach activates M3R, promoting tumor progression, invasion, and metastasis. In contrast, PDAC tumors respond differently to cholinergic stimulation. In these tumors, a muscarinic agonist inhibits pancreatic cancer progression by inhibiting MAPK and phosphoinositide 3-kinase (PI3K)/Akt signaling [79]. Accordingly, resection of the vagus nerve (which carries parasympathetic and sensory inputs) accelerates cancer progression [79,80]. In breast cancer, cholinergic nerves also display an inhibitory role: in one instance, constitutive increases in cholinergic activity by the exogenous expression of long-acting bacterial sodium channels were found to prevent cancer growth [81].
Ach displays an anti-inflammatory activity, and the vagus nerve has a unique role in inhibiting proinflammatory cytokine production [82,83]. Most inflammatory cells express mAChRs and nAChRs and respond to autocrine and paracrine stimuli [82].
One final comment is that, while less diffuse than its adrenergic counterpart, cholinergic signaling also has a role in the neuroregulation of the tumor immune microenvironment. In a mouse model of pancreatic cancer, muscarinic receptor activation lowers TNFα levels in the blood and the spleen, indicating that cholinergic signaling generates an anticancer immune microenvironment [79]. In pancreatic cancer, PNI is associated with both acetylcholine levels and reduced CD8+ T and Th1 cell abundance, while a bilateral vagotomy in mice increased CD8+ T cells in the tumors and resulted in longer survival [7,84].

5. The Tumor/Nervous Interface in the Central Nervous System Tumors and Electrochemical Communication

As could probably be anticipated from all the evidence outlined above, tumor–nervous system interactions can also be observed in primary and metastatic tumors of the central nervous system, where neurons account for approximately half of all cells. Interestingly, two recent publications demonstrated that gliomas (brain tumors derived from glial cells) grow by forming excitatory synapses with neurons in the brain [85,86]. Even more suggestively, cancer cells totally unrelated to the nervous system in origin, i.e., breast cancer metastases, form excitatory synapses with glutamatergic neurons in the brain [87]. These are very remarkable discoveries because they put the arguably most peculiar and functionally advanced neuronal structure, the synapse, in a shared context with tumor cells, further highlighting the cleverness of evolution, which uses its best “molecular solutions” in all possible settings, thus saving energy.
The issue of electrical neuronal input in facilitating cancer proliferation, particularly in brain tumors, deserves further analysis. Glioma cells are affected by electrical activity in the same way as the normal human nervous system. Gliomas, in turn, release growth stimuli into the surrounding tissue, including glutamate, increasing neuronal activity in the microenvironment. Given the presence of calcium permeable AMPA-receptor dependent synapses in high-grade gliomas, glioma-derived glutamate contributes to a bidirectional positive feedback loop of tumor–nerve hyperactivity and tumor proliferation [86].
The electrical integration of brain tumors with neuronal networks causes some clinical manifestations of gliomas, such as seizures, to be reconsidered. Seizures can directly result from glutamate secreting tumors, augmenting neural hyperactivity, despite being commonly regarded as a byproduct of the tumor’s mass effect or surrounding edema. Indeed, glioma subtypes with higher synaptogenic potential have been linked to increased seizure frequency and tumor invasion [88], and recurrent postoperative seizures have been linked to glioblastoma tumor recurrence. [89] These findings raise the question of whether antiepileptic drugs can be used to treat cancer. Preliminary research has looked into various antiepileptic drugs in glioblastoma, but more research is needed. Much remains unknown about how electrical activity from the nervous system influences cancer growth. Although the precise roles of electrical input are still unknown, the current understanding of electrical networking is consistent with the idea that tumors are driven not only by cell intrinsic processes but also by the intricately connected components of their surrounding microenvironment (reviewed in [7]).

6. Molecules with a Widespread Activity in Various Types of Cancer: Neurotransmitters, Amino Acids, Growth Factors, Axon Guidance, and Synaptic Proteins

6.1. Dopamine

Dopamine is a catecholamine family member and a precursor of NE and E. It regulates several key functions from cardiovascular activity to behavior. In peripheral tissues, dopamine is produced by sympathetic neurons, neuroendocrine cells, and the adrenal medulla (reviewed in [90]). Dopamine binds to five different G-protein-coupled receptors (DRs).
In general, there is mixed evidence regarding the influence of dopamine on cancer, suggesting that its role might be dependent on the tumor type, the expressed receptors, and the experimental conditions.
Dopamine or DR agonists inhibit the growth of gastric cancer, breast cancer, and sarcoma [91]. However, in glioblastoma, the inhibition of a DR subtype inhibits the proliferation and survival of cancer stem cells [92]. In pancreatic cancer, the inhibition of another DR subtype reduces cell migration and proliferation and decelerates the growth of experimental tumors in mice [93]. In yet another tumor model, no effect was detected on the proliferation and invasion of colon cancer cells upon DR modulation [94].
A more consolidated mechanism of the tumor-suppressive effect of dopamine may be associated with decreased angiogenesis [95]. Interestingly, dopamine prevents the release of endothelial progenitor cells from the bone marrow by blocking VEGFA-mediated ERK1/2 phosphorylation [96].
Finally, it is plausible that the antitumor effects of dopamine occur by the modulation of various immune-competent cells within the TME [97]. Dopamine also inhibits the function of a subtype of myeloid-derived suppressor cells through D1-like receptors [98] and regulates the activity of peritoneal macrophages and regulatory T lymphocytes [99].

6.2. Gamma-Aminobutyric Acid

Gamma-aminobutyric acid (GABA) is the main inhibitory neurotransmitter in the central nervous system. GABA activity in cancer is multifaceted and appears to be heavily context-dependent. Importantly, the GABA cell content is augmented in several types of human tumors, including glioma, gastric cancer, and breast cancer [100,101,102].
GABA binds two classes of cellular receptors: (a) the ionotropic chloride channels GABAA and GABAC receptors, and (b) the metabotropic GABAB receptors, which are related to the catecholaminergic receptors.
Many tumor tissues express GABA receptors [103,104,105], and with a few exceptions, the GABAA receptor mediates the GABA-induced proliferation of cancer cells, while the GABAB receptor mediates the GABA-induced inhibition of cancer cell growth [105].
The GABAA receptor is upregulated in breast, pancreatic, and prostate cancer [106,107,108]. In pancreatic cancer, GABAA activation increases intracellular Ca2+ levels and stimulates the MAPK/ERK cascade and tumor growth [109]. Muscimol, a GABAA receptor agonist, increases the proliferation of gastric cancer cells by activating the MAPK cascade.
The GABAB receptor is expressed at low levels in pancreatic and liver cancer [110,111]. GABA or GABAB agonists decrease stomach carcinogenesis in rats [112], inhibit experimental metastasis formation in colon cancer [113,114], and decrease cell growth in human hepatocellular carcinoma both in vitro and in vivo [110]. The activation of GABAB receptors strongly inhibits adrenergic-induced ERK1/2 phosphorylation and prevents DNA synthesis and cell migration [115]. On the other hand, GABA itself or baclofen (a GABAB agonist) promote the invasion of prostate cancer cells by the transactivation of EGFR [116,117].
Altogether, a complex connection between GABA and cancer seems to exist, with the differential activities probably mediated by the cancer-specific expression and signaling of its receptors.
Importantly, GABA is present in the TME, raising the likelihood that it might target the infiltrated immune cells [107].
Globally, the GABAergic system may directly or indirectly modulate cancer biology and therefore be a potential therapeutic target. Drug repurposing is within reach, as GABA is broadly used in therapy, for example, as an antidiabetic agent [118].

6.3. Serotonin

Serotonin (5-hydroxytryptamine or 5-HT) is a monoamine neurotransmitter with a wide array of functions, from cognitive and behavioral functions to various peripheral roles in the intestine, blood vessels, and immune system [119].
5-HT is a potent mitogen for many types of nontumoral cells [120,121] as well as for tumor cells from the lung, pancreas, colon, bladder, and prostate [122,123,124,125,126]. However, the serotonin-induced signaling pathways in cancer are complex and only partially understood (reviewed in [127]).
The wide activities of 5-HT are mediated by 5-HT receptors (5-HT1–7) that are extensively distributed in the human body. Two subtypes of receptors, 5-HT1A and 5-HT1B, are overexpressed in high-grade prostate cancer cells [128], while antagonists of these receptors induce apoptosis [123,124].
Epithelial tumors often display abnormal 5-HT signaling. Indeed, 5-HT regulates epithelial homeostasis in the mammary gland, lung, pancreas, and prostate [127,129]. In particular, a series of analogues of the potent 5-HT1B/1D serotonin receptor agonist 5-nonyloxytryptamine (5-NT) induce apoptosis and interfere with signaling pathways that regulate protein translation and survival, such as the Akt/mTOR pathway, in triple-negative breast cancer cells [130]. Human pancreatic cancer tissues contain high amounts of 5-HT and its receptor, HTR2B. Under these circumstances, tumor glycolysis under metabolic stress is allowed and tumor growth is promoted [131].
Finally, platelet-released 5-HT stimulates tumor growth, angiogenesis, and metastatic spread [132]. Remarkably these activities of 5-HT appear to be mediated by endothelial nitric oxide synthase and p-ERK1/2 [133].
Globally, while data suggest that 5-HT signaling is involved in cancer biology, further studies are necessary to clarify which contexts will benefit from a 5-HT-targeted therapy.

6.4. Glutamate

The amino acid glutamate is a fundamental excitatory neurotransmitter in the mammalian CNS that is involved in a wide array of functions from learning to affective behavior. Glutamate is also involved in the central biosynthetic and bioenergetics pathways.
Abnormal glutamate signaling contributes to tumorigenesis in a variety of cancers [134,135], including colorectal [136], glioma [137], gastric [138], oral squamous cell carcinoma [139], and melanoma [140].
Glutamate binds two classes of receptors: ionotropic and metabotropic. Three classes of ionotropic glutamate receptors are known: AMPA, NMDA, and kainate receptors. Conversely, metabotropic receptors (mGluR) are GPCRs constituting a large family containing several subgroups.
NMDA receptor expression has a prognostic value in oral squamous cell carcinoma; in particular, the overexpression of NMDAR1 correlates with the TN stage, the size of the primary tumor, and the metastatic spread [141].
mGluR is also involved in cancer. In colon cancer, high expression of mGluR4 is related to poor prognosis [142], while the activation of mGluR4 protects cells from 5-fluorouracil (5-FU) toxicity; conversely, in the same model, reduced mGluR4 expression or its drug-induced inactivation lead to cell death [136]. In medulloblastoma [143], however, mGluR4 expression is inversely correlated with cancer aggressiveness. The glutamate-mGluR axis can lead to the phosphorylation of the p110β subunit of phosphoinositide 3-kinase, facilitating prostate cancer progression [144]. Thus, the clinical implication of glutamate receptor expression may vary among the various tumor types.
While in general the data accumulated so far suggest the opportunity for the development of therapeutic agents targeting glutamate signaling, more studies are requested regarding the function of glutamate in the immune TME, a key target of cancer therapy.

6.5. Serine

A very recent publication showed that the amino acid serine, secreted by sensory and sympathetic neurons, stimulates pancreatic cancer metabolism in nutrient-lacking conditions [145]. These notable findings widen the array of cellular phenotypes affected by tumor–nervous system connections and put serine on the list of small molecules of neuronal origin that exert a selective pressure on cancer cells.

6.6. Neuropeptides

Neuropeptides are peptides that neurons use to communicate with the environment. Different neuropeptides modulate a wide range of brain activities but can also function peripherally as endocrine/paracrine factors regulating various physiological processes [146].
Substance P (SP) [147] and neuropeptide Y (NPY) [148], with their receptors of the GPCR superfamily, are the neuropeptides that have been studied the most in cancers. SP modulates the apoptosis, proliferation, migration, angiogenesis, and invasion of many cancers, including colon, pancreatic, melanoma, osteosarcoma, gastric, glioma, larynx, and lung carcinoma [147,149]. The biological activity of SP is primarily mediated by the neurokinin-1 (NK-1) receptor and two second messengers: diacylglicerol (DAG) and inositol 1,4,5-triphosphate (IP3) [149]. Interestingly, the inhibition of the NK-1 receptor with specific pharmacological antagonists results in pronounced antitumor effects [150]. Other specific roles of SP/NK1 in cancer can be described. In breast cancer, the miRNA MiR-34b/c-5p and its target NK-1 receptor affect cell proliferation and apoptosis. SP rescues the effects of miR-34b/c-5p overexpression or NK1 receptor silencing on cell proliferation and apoptosis in in vitro and in vivo assays [151]. In glioma, SP, through the NK-1 receptor, controls cancer cell proliferation by activating c-myc, MAPK, activator protein 1, and ERK 1/2. In glioma cells, SP promotes glycogen degradation, which is essential for glycolysis. By contrast, antagonists of the NK-1 receptor stop the proliferation of tumor cells and the degradation of glycogen and stimulate apoptosis [152].
NPY is a growth-enhancing factor in numerous cancers, including Ewing sarcoma, neuroblastoma, and breast and prostate cancer [148,153]. NPY modulates not only tumor growth and vascularization but also the metastatic and chemoresistant phenotypes of cancer cells [154,155]. NPY, which is abundantly produced by enteric neurons, also regulates carcinogenesis via two systems: first by the downregulation of microRNA-375-dependent apoptosis in intestinal epithelial cells and second by enhanced proliferation (via PI3-K/pAkt) [156].
While the mechanisms through which neuropeptides regulate tumor progression is only partially known, the existing evidence supports the possibility of therapeutically targeting neuropeptides and their receptors.

6.7. Nerve Growth Factor

A nervous growth factor that deserves particular mention is NGF because of its widespread activity in cancer biology. Solid cancers respond to NGF of autocrine/paracrine origin. TrkA and p75NTR are both receptors for NGF that are dysregulated in breast cancer, prostate cancer, and melanoma. The binding of these receptors results in downstream PI3K pathway, MAPK pathway, and c-Jun N-terminal kinase (JNK) pathway activation, which drives increased tumor survival, proliferation, and invasion [157].
Many solid tumors, including gastric cancer, also release NGF in an effort to influence the surrounding neurons and the TME [78]. In breast cancer, the levels of NGF correlate with both the density of tumor innervation and tumor aggression. This supports the hypothesis that tumor-released NGF stimulates neuron growth and the innervation of the tumor. The newly formed axons, in turn, release tumor growth factors, thereby driving tumor aggression. Remarkably, NGF-induced neurons often express tyrosine hydroxylase, further supporting the existence of a positive pressure of sympathetic adrenergic signaling on cancer evolution through increased tumor aggressivity and treatment resistance [158].

6.8. Axon Guidance Molecules and Neuroligin

In this paragraph we initially cover the role in cancer of three main “classic” families of axon guidance molecules (netrins, ephrins, and semaphorins). Many excellent reviews have covered these molecules in the context of cancer [159,160,161,162], so we mainly outline the latest findings. Finally, we report on our newer findings on the tumor autonomous role of the synaptic protein neuroligin.

6.8.1. Axon Guidance Molecules

Netrins are soluble proteins that act through their receptors of the DCC (deleted in colorectal cancer) or UNC5 families and determine the direction of neuronal neurites [163]. Netrins have, however, been found to play a role in tumor biology [160,164], where in particular DCC and UNC5 positively regulate apoptosis in the absence of netrin-1 but negatively regulate apoptosis in the presence of netrin-1 [165]. Reduced netrin-1 expression is present in tumors of the prostate and the nervous system [166,167]. In intestinal cancer, DCC suppresses tumor growth by inducing cancer cell death when netrin-1 expression is low [168]. However, netrin-1 and its receptor DCC also control the progression of melanoma, indicating that the therapeutic targeting of this signaling axis may be a viable method for melanoma treatment [169]. Interestingly, it has also been shown that netrin-1 is elevated in cancer-associated fibroblasts (CAFs) in addition to cancer cells [170]. In addition to its direct impact on cancer cells, netrin-1 inhibition may hence inhibit the crosstalk between pro-neoplastic CAFs and cancer cells, reducing malignant plasticity.
Mechanistically, it was discovered that a portion of colorectal cancers exhibit a netrin-1 increase concurrent with transcription factor NF-kB activation and that netrin-1 is a direct transcriptional target of NF-kB [171]. Additionally, colorectal malignancies in the context of inflammatory bowel disorders (IBD) exhibit an overexpression of netrin-1. Numerous data point to a close connection between chronic inflammation and tumorigenesis, primarily via NF-kB activation. Hence, it was suggested [171] that an NF-kB-mediated increase in netrin-1 expression in IBD patients may have an impact on the promotion and progression of colorectal tumors. Moreover, netrin-1 upregulates YAP (an oncoprotein located in the cytoplasm in an inactive form that, when activated, translocates to the nucleus and activates the transcription of genes responsible for cell division and apoptosis) levels in the nucleus, while inactivating netrin-1, DCC, or UNC-5B decreases YAP protein levels [172].
The Eph receptors, with their ligands the Ephrins, are a large family of transmembrane proteins that were originally implicated in many neurological functions [173,174]. Eph/ephrin signaling is exceptionally complicated because of its bidirectional nature (with Ephs and ephrins interacting “in trans” from two different cells), and as a consequence Eph/ephrin have been shown to coordinate a plethora of tumor-related processes from angiogenesis, inflammation, vascular infiltration, and metastasis to tumor immunity [175,176,177]. To name a few examples, increased Eph/ephrin signaling is associated with tumor progression in melanoma [178], breast [179], prostate [180], and colorectal cancer [175]. The Eph family member EPHA2 has recently been proposed as a negative prognostic factor in triple-negative breast cancer (TNBC) patients; moreover, EPHA2 may act as a precision therapeutic target for this challenging condition [181]. Finally, EphB2 signaling is increased in Barrett’s esophagus and esophageal adenocarcinoma and regulates the MYC oncogene [182]
Mechanistically, after an EPH-ephrin engagement, a complex molecular pathway activates signaling in both cells that are involved. EPHs and ephrins engage with different molecular cascades to further transmit the message into the cytoplasm. Forward signaling involves interactions between EPHs and Src family kinases, which regulate the development of synapses; Rho GTPases, which stabilize junctions; and interactions between ephexins (GEFs that can activate Rho GTPases) and the ERK/MAPK pathway, which promote cell proliferation. Additionally, FAK and the JAK/STAT pathway interact with EPHs, which modifies cell adhesion. The interactions between ephrins and Src, Erk, Rac, paxillin, and p75, which result in integrin-dependent cell adhesion, as well as with Src, Grb4, PTP-BL, and PDZ-RGS3, which control a variety of actions, including cell adhesion, migration, and proliferation, are among the mechanisms implicated in reverse signaling. Phosphoinositide 3-kinase is also known as PI3K Janus kinase, JAK RGS3 stands for regulator of G-protein signaling 3, FAK stands for focal adhesion kinase, and STAT stands for signal transducer and activator of transcription [175].
Semaphorins are a large family of secreted, transmembrane, or GPI-anchored proteins (SEMAs) signaling through their receptors, neuropilins, and plexins. As for many other cues considered here, semaphorins were originally described as guidance molecules for neurons but were then found to be released by several types of cancer cells [183,184] and by cells in the tumor microenvironment [185]. Semaphorins impact many cancer hallmarks: cancer cell proliferation, angiogenesis, immune response, invasion, and metastasis [186].
Regarding the inflammatory and immunological roles of semaphorins, their receptors are abundantly expressed in inflammatory cells, and tumor cells produce their ligands, resulting in intense local and systemic signaling crosstalk. Additionally, many semaphorins regulate immune cells that are both innate and antigen-specific. Notably, semaphorin signals that function as inhibitors of the immune system’s ability to fight cancer are frequently dysregulated in human malignancies and could potentially be targeted for treatment [187].
To illustrate the complexities of the biological activities and signaling of the semaphorin family members in cancer with an example, we take the case of breast cancer, based on the excellent work of Butti and colleagues [188]. Sema3A binds to the neuropilin 1 (NRP1) receptor, ultimately inhibiting tumor growth and angiogenesis. Sema3B binds to NRP1 and inhibits PI3K/Akt signaling, causing apoptosis. Full-length Sema3C interacts with NRP2 on lymphatic endothelial cells in tumors, inhibiting VEGF-C-dependent ERK1/2 and Akt signaling and suppressing lymphangiogenesis and metastasis. Sema4D binds to plexin-B1 and activates ErbB2, phosphorylating plexin-B1. By activating RhoA GTPase, phosphorylated plexin-B1 promotes migration. Through ErbB2-dependent MAPK signaling, cleaved p61-Sema3E binds to plexin-D1 to promote metastasis. Sema3E binds to plexin-D1 and disrupts the interaction between plexin-D1 and NR4A, which is known to induce caspase-9-mediated apoptosis [188].
Overall, notwithstanding their complicated signaling mechanism, research in the field of semaphorins is active in order to develop anticancer strategies [186,189].

6.8.2. Neuroligins

Synaptic proteins involved in cancer include postsynaptic scaffold proteins, receptors, and ion channels and have been very recently reviewed [6]. Moreover, it is particularly worth noting the extensive expression of components of the synaptic machinery, such as synaptophysin, in some non-neural cancers [190]. This is easily noticeable, for example, in small-cell lung carcinoma, a very aggressive type of lung cancer, which displays many similarities with neuronal cells, and some forms of colon and prostate cancers [190]. This attractive biological phenomenon could reflect the selection of cancer cells endowed with the molecular tools needed to interact with the nervous system.
Here, we only consider a novel synaptic protein that we and others have discovered to be involved in cancer: neuroligin 1.
The family of neuroligins is composed of transmembrane postsynaptic proteins of the CNS that function in the fine-tuning of synaptic activity. We have previously shown that a member of this family, neuroligin 1 (NLGN1), is expressed in endothelial cells and modulates angiogenesis [191,192,193,194]. While a database study by Qian et al. [195] showed that NLGN1 is expressed in a subset of colon tumors and has negative prognostic implications, very recently we demonstrated that NLGN1 is expressed in human colorectal tumors, in particular in groups of aggressively migrating cells); single tumor cells from primary tumors (an event called tumor budding), which represents a negative prognostic factor; and in circulating tumor cells as either single cells or aggregates (that are referred to as ‘emboli’ or ‘microemboli’). We discovered that NLGN1 enhances CRC cell extravasation/lung invasion, differential organ metastasization in two mice models, and CRC cell transendothelial migration across an endothelial monolayer in vitro. In CRC cell lines, NLGN1 upregulates mesenchymal markers and WNT target genes, enhances APC (adenomatous polyposis coli, a CRC tumor suppressor that is mutated in up to 80% of CRC samples) localization to the cell membrane, co-immunoprecipitates with some isoforms of this protein, and induces beta-catenin translocation to the nucleus. NLGN1 appears as the first eminently synaptic neuronal protein that modulates CRC aggressiveness by affecting a crucial pathogenetic route of this disease [196].
Before this report, we demonstrated that NLGN1 is expressed by prostatic and pancreatic cancer tissues in distinct stages and tumor districts using immunohistochemistry on human tissue microarrays. Then, using in vitro and in vivo tests, we were able to show that NLGN1 stimulates cancer cells to invade and move along nerves. We investigated a potential cooperation between NLGN1 and the neurotrophic factor glial cell line derived neurotrophic factor (GDNF) because this factor is known to play a function in tumor–nerve interactions. We discovered that the NLGN1-induced in vitro cancer cell invasion of nerves was fully prevented by suppressing GDNF activity with a specific antibody. Finally, we showed that GDNF significantly activates cofilin, a cytoskeletal regulating protein, in the presence of NLGN1, changing filopodia dynamics. Here, it was demonstrated that NLGN1 interacts with one of the most prevalent neurotrophic factors in the nervous system, potentially opening new treatment avenues [197].
Another neuroligin isoform (NLGN3) has been involved in glioma growth. This protein is structurally and functionally related to NLGN1 (72% aa identity and 81% conservation), as they are both present at excitatory synapses. The protease ADAM10 (a disintegrin and Metalloprotease 10) is stimulated by neuronal activity to promote NLGN3 extracellular domain shedding into the tumor microenvironment, which promotes glioma growth [198]. A clinical trial (#NCT04295759, https://clinicaltrials.gov/ (accessed on 17 October 2022)) is underway to test the efficacy of the medication INCB7839 in inhibiting ADAM10 in gliomas. It is important to note that in the brain ADAM-10 similarly sheds the NLGN1 ectodomain in an activity-dependent manner [199]. Therefore, future studies may look at whether soluble NLGN1 can function as a mediator of CRC growth and metastasization.

7. Conclusions

A number of very recent reviews have comprehensively analyzed tumor/nervous system crosstalk from different perspectives [7,14,200,201]. Here, we aimed to comprehensively analyze the shared molecules and signaling pathways between the systems and their biological outcomes. Figure 3 and Table 1 present a quick overview of the molecules that we considered, their signaling, and their biological effects.
Our review has left out some molecular cues (e.g., some members of the neurotrophins and the neuropoietin family) with very specific activities that are presented in other more dedicated publications [202,203,204].
We have collectively shown that many of the original hallmarks of cancer [205] are affected by its interaction with the nervous system. We believe that these include acquired capabilities for sustaining proliferative signaling, resisting cell death, inducing/accessing the vasculature, activating invasion and metastasis, reprogramming cellular metabolism, and avoiding immune destruction. Hence, our review certainly sustains the idea that the molecular tumor–nerve crosstalk may represent a further hallmark of cancer on its own, as recently proposed [10].
As was certainly noticed, except for the case of the adrenergic input on epithelial tumor growth, the effects of the nervous system mediators on cancer biology are extremely multifaceted and complex. In the case of the differential effects of the SNS and PSNS, one can make an interesting parallel with the complementary physiological effects that these two branches of the autonomic nervous system have in the human body.
In general, the effect of a given neural cue is heavily dependent on the considered isoform and the tumor model used. Although at times this may be due to different experimental settings, we believe that this situation mostly reflects the extensive innate inter- and intratumor heterogeneity, both among the transformed cells and their microenvironments. In particular, different regions of tumors display different molecular compositions of the extracellular matrix, different quantities of infiltrating normal cells, and different amounts of blood and lymphatic vessels. In this setting, single tumor cells within a mass are subject to a range of microenvironmental inputs and respond with a variety of phenotypic manifestations. We believe that, as for other microenvironmental cues, the nervous system is heavily involved in this diversity.
Finally, it should be remembered that the goal of innervation is not only to transmit rapid signals between nerves and tissues. Denervation and cross-reinnervation experiments in striated muscle have shown that neuronal signals regulate the phenotypic traits of innervated target tissues [206,207]. This is a key observation for the study of nervous influence on tumors. In fact, if this is also the case for nerve–tumor cell interactions, whereby a selective force is imposed by the nervous system on the phenotypic features of cancer cells, then the stimulation/inhibition of nerves entering tumors, the denervation of tumors, or the modulation of the molecules reviewed here might offer new therapeutic options.

Author Contributions

Conceptualizaton: M.A. and F.B.; writing—original draft preparation: M.A., F.B., M.P. and L.B.; Supervision: M.A. and F.B.; Funding acquisition: M.A. and F.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a Fondazione Umberto Veronesi 2014 Young Investigator Grant to M.A. and AIRC IG grants #11503 and #15238 to M.A. and #22910, #12182, and #18652 to F.B.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lacina, L.; Coma, M.; Dvovránková, B.; Kodet, O.; Melegova, N.; Gal, P.; Smetana, K. Evolution of Cancer progression in the context of Darwinism. Anticancer Res. 2019, 39, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Gal, P.; Varinska, L.; Faber, L.; Novákv, S.; Szabo, P.; Mitrengova, P.; Mirossay, A.; Muvcaji, P.; Smetana, K. How signaling molecules regulate tumor microenvironment: Parallels to wound repair. Molecules 2017, 22, 1818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Fife, A.; Beasley, P.J.; Fertig, D.L. Psychoneuroimmunology and cancer: Historical perspectives and current research. Adv. Neuroimmunol. 1996, 6, 179–190. [Google Scholar] [CrossRef]
  4. Kiecolt-Glaser, J.K.; Robles, T.; Heffner, K.; Loving, T.; Glaser, R. Psycho-oncology and cancer: Psychoneuroimmunology and cancer. Ann. Oncol. 2002, 13, 165–169. [Google Scholar] [CrossRef] [PubMed]
  5. Monje, M.; Borniger, J.C.; D’Silva, N.J.; Deneen, B.; Dirks, P.B.; Fattahi, F. Roadmap for the Emerging Field of Cancer Neuroscience. Cell 2020, 181, 219–222. [Google Scholar] [CrossRef]
  6. Li, J.; Xu, Y.; Zhu, H.; Wang, Y.; Li, P.; Wang, D. The dark side of synaptic proteins in tumours. Br. J. Cancer 2022, 3, 1184–1192. [Google Scholar] [CrossRef]
  7. Shi, D.D.; Guo, J.A.; Hoffman, H.I.; Su, J.; Mino-Kenudson, M.; Barth, J.L.; Schenkel, J.M.; Loeffler, J.S.; Shih, H.A.; Hong, T.S.; et al. Therapeutic avenues for cancer neuroscience: Translational frontiers and clinical opportunities. Lancet Oncol. 2022, 23, e62–e74. [Google Scholar] [CrossRef]
  8. Hernandez, S.; Serrano, A.G.; Solis Soto, L.M. The Role of Nerve Fibers in the Tumor Immune Microenvironment of Solid Tumors. Adv. Biol. 2022, 6, 2200046. [Google Scholar] [CrossRef] [PubMed]
  9. Pan, C.; Winkler, F. Insights and opportunities at the crossroads of cancer and neuroscience. Nat. Cell Biol 2022, 24, 1454–1460. [Google Scholar] [CrossRef] [PubMed]
  10. Amit, M. Cancer Neuroscience: Evidence for a New Hallmark of Cancer. Adv. Biol. 2022, 6, 2200192. [Google Scholar] [CrossRef] [PubMed]
  11. Silverman, D.A.; Martinez, V.K.; Dougherty, P.M.; Myers, J.N.; Calin, G.A.; Amit, M. Cancer-Associated Neurogenesis and Nerve-Cancer Cross-talk. Cancer Res. 2021, 81, 1431–1440. [Google Scholar] [CrossRef] [PubMed]
  12. Kayahara, M.; Nakagawara, H.; Kitagawa, H.; Ohta, T. The nature of neural invasion by pancreatic cancer. Pancreas 2007, 35, 218. [Google Scholar] [CrossRef]
  13. Liebig, C.; Ayala, G.; Wilks, J.A.; Berger, D.H.; Albo, D. Perineural invasion in cancer. Cancer 2009, 115, 3379–3391. [Google Scholar] [CrossRef]
  14. Zahalka, A.H.; Frenette, P.S. Nerves in cancer. Nat. Rev. Cancer 2020, 20, 143–157. [Google Scholar] [CrossRef] [PubMed]
  15. Amit, M.; Na’ara, S.; Gil, Z. Mechanisms of cancer dissemination along nerves. Nat. Rev. 2016, 16, 399. [Google Scholar] [CrossRef]
  16. Ayala, G.E.; Wheeler, T.M.; Shine, H.D.; Schmelz, M.; Frolov, A.; Chakraborty, S.; Rowley, D. In vitro dorsal root ganglia and human prostate cell line interaction: Redefining perineural invasion in prostate cancer. Prostate 2001, 49, 213–223. [Google Scholar] [CrossRef] [PubMed]
  17. Niu, Y.; Förster, S.; Muders, M. The Role of Perineural Invasion in Prostate Cancer and Its Prognostic Significance. Cancers 2022, 14, 4065. [Google Scholar] [CrossRef]
  18. Mauffrey, P.; Tchitchek, N.; Barroca, V.; Bemelmans, A.-P.; Firlej, V.; Allory, Y.; Romeo, P.-H.; Magnon, C. Progenitors from the central nervous system drive neurogenesis in cancer. Nature 2019, 569, 672–678. [Google Scholar] [CrossRef] [PubMed]
  19. Edwards, G.; Campbell, T.; Henderson, V.; Danaher, A.; Wu, D.; Srinivasan, R.; Rezvani, K.; Odero-Marah, V.A. SNAIL Transctiption factor in prostate cancer cells promotes neurite outgrowth. Biochimie 2021, 180, 1–9. [Google Scholar] [CrossRef]
  20. Eksi, S.E.; Chitsazan, A.; Sayar, Z.; Thomas, G.V.; Fields, A.J.; Kopp, R.P.; Spellman, P.T.; Adey, A.C. Epigenetic loss of heterogeneity from low to high grade localized prostate tumours. Nat. Commun. 2021, 12, 7292. [Google Scholar] [CrossRef] [PubMed]
  21. Frunza, A.; Slavescu, D.; Lascar, I. Perineural invasion in head and neck cancers-a review. J. Med. Life 2014, 7, 121. [Google Scholar]
  22. Li, R.; Wheeler, T.; Dai, H.; Ayala, G. Neural cell adhesion molecule is upregulated in nerves with prostate cancer invasion. Hum. Pathol. 2003, 34, 457–461. [Google Scholar] [CrossRef] [PubMed]
  23. Ding, Y.; He, D.; Florentin, D.; Frolov, A.; Hilsenbeck, S.; Ittmann, M.; Kadmon, D.; Miles, B.; Rowley, D.; Ayala, G. Semaphorin 4F as a Critical Regulator of Neuroepithelial Interactions and a Biomarker of Aggressive Prostate CancerSemaphorin 4F and Nerves. Clin. Cancer Res. 2013, 19, 6101–6111. [Google Scholar] [CrossRef] [Green Version]
  24. Yin, L.; Li, J.; Wang, J.; Pu, T.; Wei, J.; Li, Q.; Wu, B.J. MAOA promotes prostate cancer cell perineural invasion through SEMA3C/PlexinA2/NRP1-cMET signaling. Oncogene 2021, 40, 1362–1374. [Google Scholar] [CrossRef] [PubMed]
  25. Adekoya, T.O.; Richardson, R.M. Cytokines and chemokines as mediators of prostate cancer metastasis. Int. J. Mol. Sci. 2020, 21, 4449. [Google Scholar] [CrossRef]
  26. Na’ara, S.; Amit, M.; Gil, Z. L1CAM induces perineural invasion of pancreas cancer cells by upregulation of metalloproteinase expression. Oncogene 2019, 38, 596–608. [Google Scholar] [CrossRef] [PubMed]
  27. Scanlon, C.S.; Banerjee, R.; Inglehart, R.C.; Liu, M.; Russo, N.; Hariharan, A.; Van Tubergen, E.A.; Corson, S.L.; Asangani, I.A.; Mistretta, C.M.; et al. Galanin modulates the neural niche to favour perineural invasion in head and neck cancer. Nat. Commun 2015, 6, 6885. [Google Scholar] [CrossRef] [Green Version]
  28. Lin, C.; Ren, Z.; Yang, X.; Yang, R.; Chen, Y.; Liu, Z.; Dai, Z.; Zhang, Y.; He, Y.; Zhang, C.; et al. Nerve growth factor (NGF)-TrkA axis in head and neck squamous cell carcinoma triggers EMT and confers resistance to the EGFR inhibitor erlotinib. Cancer Lett. 2020, 472, 81–96. [Google Scholar] [CrossRef] [PubMed]
  29. Ban, K.; Feng, S.; Shao, L.; Ittmann, M. RET signaling in prostate cancer. Clin. Cancer Res. 2017, 23, 4885–4896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Jobling, P.; Pundavela, J.; Oliveira, S.M.; Roselli, S.; Walker, M.M.; Hondermarck, H. Nerve-Cancer Cell Cross-talk: A Novel Promoter of Tumor Progression. Cancer Res. 2015, 75, 1777–1781. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Mo, R.-J.; Han, Z.-D.; Liang, Y.-K.; Ye, J.-H.; Wu, S.-L.; Lin, S.X.; Zhang, Y.-Q.; Song, S.-D.; Jiang, F.-N.; Zhong, W.-D.; et al. Expression of PD-L1 in tumor-associated nerves correlates with reduced CD8+ tumor-associated lymphocytes and poor prognosis in prostate cancer. Int. J. Cancer 2019, 144, 3099–3110. [Google Scholar] [CrossRef] [PubMed]
  32. Mizushima, N.; Levine, B.; Cuervo, A.M.; Klionsky, D.J. Autophagy fights disease through cellular self-digestion. Nature 2008, 451, 1069–1075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Morishita, H.; Mizushima, N. Diverse cellular roles of autophagy. Annu. Rev. Cell Dev. Biol. 2019, 35, 453–475. [Google Scholar] [CrossRef]
  34. Cai, Q.; Ganesan, D. Regulation of neuronal autophagy and the implications in neurodegenerative diseases. Neurobiol. Dis. 2021, 4, 105582. [Google Scholar] [CrossRef] [PubMed]
  35. Yazdankhah, M.; Ghosh, S.; Shang, P.; Stepicheva, N.; Hose, S.; Liu, H.; Chamling, X.; Tian, S.; Sullivan, M.L.; Calderon, M.J.; et al. BNIP3L-mediated mitophagy is required for mitochondrial remodeling during the differentiation of optic nerve oligodendrocytes. Autophagy 2021, 17, 3140–3159. [Google Scholar] [CrossRef] [PubMed]
  36. Griffey, C.J.; Yamamoto, A. Macroautophagy in CNS health and disease. Nat. Rev. Neurosci. 2022, 23, 411–427. [Google Scholar] [CrossRef] [PubMed]
  37. Wang, P.; Chen, X.; Wang, Y.; Jia, C.; Liu, X.; Wang, Y.; Wu, H.; Cai, H.; Shen, H.-M.; Le, W. Essential role for autophagy protein VMP1 in maintaining neuronal homeostasis and preventing axonal degeneration. Cell Death Dis. 2021, 12, 116. [Google Scholar] [CrossRef]
  38. Pushpakom, S.; Iorio, F.; Eyers, P.A.; Escott, K.J.; Hopper, S.; Wells, A.; Doig, A.; Guilliams, T.; Latimer, J.; McNamee, C.; et al. Drug repurposing: Progress, challenges and recommendations. Nat. Rev. Drug Discov. 2019, 18, 41–58. [Google Scholar] [CrossRef] [PubMed]
  39. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  40. Regulska, K.; Regulski, M.; Karolak, B.; Murias, M.; Stanisz, B. Can cardiovascular drugs support cancer treatment? The rationale for drug repurposing. Drug Discov. Today 2019, 24, 1059–1065. [Google Scholar] [CrossRef]
  41. Demir, I.E.; Reyes, C.M.; Alrawashdeh, W.; Ceyhan, G.O.; Deborde, S.; Friess, H.; Görgülü, K.; Istvanffy, R.; Jungwirth, D.; Kuner, R.; et al. Future directions in preclinical and translational cancer neuroscience research. Nature 2020, 1, 1027–1031. [Google Scholar] [CrossRef]
  42. Shaw, V.; Srivastava, S.; Srivastava, S.K. Repurposing antipsychotics of the diphenylbutylpiperidine class for cancer therapy. Semin. Cancer Biol. 2021, 1, 75–83. [Google Scholar] [CrossRef]
  43. Shi, J.; Xu, J.; Li, Y.; Li, B.; Ming, H.; Nice, E.C.; Huang, C.; Li, Q.; Wang, C. Drug repurposing in cancer neuroscience: From the viewpoint of the autophagy-mediated innervated niche. Front. Pharmacol. 2022, 13, 990665. [Google Scholar] [CrossRef] [PubMed]
  44. Valdor, R.; Garc’ia-Bernal, D.; Riquelme, D.; Martinez, C.M.; Moraleda, J.M.; Cuervo, A.M.; Macian, F.; Martinez, S. Glioblastoma ablates pericytes antitumor immune function through aberrant up-regulation of chaperone-mediated autophagy. Proc. Natl. Acad. Sci. USA 2019, 116, 20655–20665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Chen, S.-H.; Zhang, B.-Y.; Zhou, B.; Zhu, C.-Z.; Sun, L.-Q.; Feng, Y.-J. Perineural invasion of cancer: A complex crosstalk between cells and molecules in the perineural niche. Am. J. Cancer Res. 2019, 9, 1–21. [Google Scholar] [PubMed]
  46. William Tank, A.; Lee Wong, D. Peripheral and central effects of circulating catecholamines. Compr. Physiol. 2011, 5, 1–15. [Google Scholar]
  47. Lillberg, K.; Verkasalo, P.K.; Kaprio, J.; Teppo, L.; Helenius, H.; Koskenvuo, M. Stressful life events and risk of breast cancer in 10,808 women: A cohort study. Am. J. Epidemiol. 2003, 157, 415–423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Chida, Y.; Hamer, M.; Wardle, J.; Steptoe, A. Do stress-related psychosocial factors contribute to cancer incidence and survival? Nat. Clin. Pract. Oncol. 2008, 5, 466–475. [Google Scholar] [CrossRef] [PubMed]
  49. Partecke, L.I.; Speerforck, S.; Käding, A.; Seubert, F.; Kühn, S.; Lorenz, E.; Schwandke, S.; Sendler, M.; Keßler, W.; Trung, D.N.; et al. Chronic stress increases experimental pancreatic cancer growth, reduces survival and can be antagonised by beta-adrenergic receptor blockade. Pancreatology 2016, 16, 423–433. [Google Scholar] [CrossRef]
  50. Thaker, P.H.; Han, L.Y.; Kamat, A.A.; Arevalo, J.M.; Takahashi, R.; Lu, C.; Jennings, N.B.; Armaiz-Pena, G.; Bankson, J.A.; Ravoori, M.; et al. Chronic stress promotes tumor growth and angiogenesis in a mouse model of ovarian carcinoma. Nat. Med. 2006, 12, 939–944. [Google Scholar] [CrossRef]
  51. Hassan, S.; Karpova, Y.; Baiz, D.; Yancey, D.; Pullikuth, A.; Flores, A.; Register, T.; Cline, J.M.; D’Agostino, R.; Danial, N.; et al. Behavioral stress accelerates prostate cancer development in mice. J. Clin. Investig. 2013, 123, 2972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Schuller, H.M.; Al-Wadei, H.A.; Ullah, M.F.; Plummer, H.K. Regulation of pancreatic cancer by neuropsychological stress responses: A novel target for intervention. Carcinogenesis 2012, 33, 2871. [Google Scholar] [CrossRef] [PubMed]
  53. Magnon, C.; Hall, S.J.; Lin, J.; Xue, X.; Gerber, L.; Freedland, S.J.; Frenette, P.S. Autonomic nerve development contributes to prostate cancer progression. Science 2013, 341, 6142. [Google Scholar] [CrossRef] [Green Version]
  54. Sood, A.K.; Armaiz-Pena, G.N.; Halder, J.; Nick, A.M.; Stone, R.L.; Hu, W. Adrenergic modulation of focal adhesion kinase protects human ovarian cancer cells from anoikis. J. Clin. Investig. 2010, 120, 24579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Renz, B.W.; Takahashi, R.; Tanaka, T.; Macchini, M.; Hayakawa, Y.; Dantes, Z.; Maurer, H.C.; Chen, X.; Jiang, Z.; Westphalen, C.B.; et al. β2 adrenergic-neurotrophin feedforward loop promotes pancreatic cancer. Cancer Cell 2018, 33, 75–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Li, J.; Yang, X.M.; Wang, Y.H.; Feng, M.X.; Liu, X.J.; Zhang, Y.L. Monoamine oxidase A suppresses hepatocellular carcinoma metastasis by inhibiting the adrenergic system and its transactivation of EGFR signaling. J. Hepatol. 2014, 60, 246. [Google Scholar] [CrossRef] [Green Version]
  57. Han, J.; Jiang, Q.; Ma, R.; Zhang, H.; Tong, D.; Tang, K.; Wang, X.; Ni, L.; Miao, J.; Duan, B.; et al. Norepinephrine-CREB1-miR-373 axis promotes progression of colon cancer. Mol. Oncol. 2020, 14, 1059–1073. [Google Scholar] [CrossRef] [Green Version]
  58. Armaiz-Pena, G.N.; Allen, J.K.; Cruz, A.; Stone, R.L.; Nick, A.M.; Lin, Y.G.; Han, L.Y.; Mangala, L.S.; Villares, G.J.; Vivas-Mejia, P.; et al. Src activation by β-adrenoreceptors is a key switch for tumour metastasis. Nat. Commun. 2013, 4, 1403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Qian, W.; Lv, S.; Li, J.; Chen, K.; Jiang, Z.; Cheng, L.; Zhou, C.; Yan, B.; Cao, J.; Ma, Q.; et al. Norepinephrine enhances cell viability and invasion, and inhibits apoptosis of pancreatic cancer cells in a Notch-1-dependent manner. Oncol. Rep. 2018, 40, 3015–3023. [Google Scholar] [CrossRef] [PubMed]
  60. Decker, A.M.; Jung, Y.; Cackowski, F.C.; Yumoto, K.; Wang, J.; Taichman, R.S. Sympathetic signaling reactivates quiescent disseminated prostate cancer cells in the bone marrow. Mol. Cancer Res. 2017, 15, 1644–1655. [Google Scholar] [CrossRef] [Green Version]
  61. Jang, H.-J.; Boo, H.-J.; Lee, H.J.; Min, H.-Y.; Lee, H.-Y. Chronic stress facilitates lung tumorigenesis by promoting exocytosis of IGF2 in lung epithelial cells. Cancer Res. 2016, 76, 6607–6619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Zahalka, A.H.; Arnal-Estapé, A.; Maryanovich, M.; Nakahara, F.; Cruz, C.D.; Finley, L.W.; Frenette, P.S. Adrenergic nerves activate an angio-metabolic switch in prostate cancer. Science 2017, 358, 321–326. [Google Scholar] [CrossRef] [PubMed]
  63. Hondermarck, H.; Jobling, P. The sympathetic nervous system drives tumor angiogenesis. Trends Cancer 2018, 4, 93–94. [Google Scholar] [CrossRef] [PubMed]
  64. Dethlefsen, C.; Hansen, L.S.; Lillelund, C.; Andersen, C.; Gehl, J.; Christensen, J.F.; Pedersen, B.K.; Hojman, P. Exercise-induced catecholamines activate the hippo tumor suppressor pathway to reduce risks of breast cancer development. Cancer Res. 2017, 77, 4894–4904. [Google Scholar] [CrossRef] [Green Version]
  65. Song, Y.; Gan, Y.; Wang, Q.; Meng, Z.; Li, G.; Shen, Y.; Wu, Y.; Li, P.; Yao, M.; Gu, J.; et al. Enriching the housing environment for mice enhances their NK cell antitumor immunity via sympathetic nerve-dependent regulation of NKG2D and CCR5. Cancer Res. 2017, 77, 1611–1622. [Google Scholar] [CrossRef] [Green Version]
  66. Kim, T.-H.; Ly, C.; Christodoulides, A.; Nowell, C.J.; Gunning, P.W.; Sloan, E.K.; Rowat, C.A. Stress hormone signaling through β-adrenergic receptors regulates macrophage mechanotype and function. FASEB J. 2019, 33, 3997–4006. [Google Scholar] [CrossRef] [PubMed]
  67. Dimitrov, S.; Lange, T.; Gouttefangeas, C.; Jensen, A.T.; Szczepanski, M.; Lehnnolz, J.; Soekadar, S.; Rammensee, H.-G.; Born, J.; Besedovsky, L. Gαs-coupled receptor signaling and sleep regulate integrin activation of human antigen-specific T cells. J. Exp. Med. 2019, 216, 517–526. [Google Scholar] [CrossRef] [Green Version]
  68. Sloan, E.K.; Priceman, S.J.; Cox, B.F.; Yu, S.; Pimentel, M.A.; Tangkanangnukul, V.; Arevalo, J.M.; Morizono, K.; Karanikolas, B.D.; Wu, L.; et al. The sympathetic nervous system induces a metastatic switch in primary breast cancer. Cancer Res. 2010, 70, 7042–7052. [Google Scholar] [CrossRef] [Green Version]
  69. Mohammadpour, H.; MacDonald, C.R.; Qiao, G.; Chen, M.; Dong, B.; Hylander, B.L.; McCarthy, P.L.; Abrams, S.I.; Repasky, E.A.; Priceman, S.J.; et al. β2 adrenergic receptor-mediated signaling regulates the immunosuppressive potential of myeloid-derived suppressor cells. J. Clin. Investig. 2019, 129, 5537–5552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Dubeykovskaya, Z.; Si, Y.; Chen, X.; Worthley, D.L.; Renz, B.W.; Urbanska, A.M.; Hayakawa, Y.; Xu, T.; Westphalen, C.B.; Dubeykovskiy, A.; et al. Neural innervation stimulates splenic TFF2 to arrest myeloid cell expansion and cancer. Nat. Commun. 2016, 7, 10517. [Google Scholar] [CrossRef] [Green Version]
  71. Nimmakayala, R.K.; Seshacharyulu, P.; Lakshmanan, I.; Rachagani, S.; Chugh, S.; Karmakar, S. Cigarette smoke induces stem cell features of pancreatic cancer cells via PAF1. Gastroenterology 2018, 155, 2862. [Google Scholar] [CrossRef]
  72. Elisia, I.; Cho, B.; Hay, M.; Li, M.Y.; Hofs, E.; Lam, V.; Dyer, R.A.; Lum, J.; Krystal, G. The effect of diet and exercise on tobacco carcinogen-induced lung cancer. Carcinogenesis 2019, 40, 448–460. [Google Scholar] [CrossRef]
  73. Shimizu, R.; Ibaragi, S.; Eguchi, T.; Kuwajima, D.; Kodama, S.; Nishioka, T.; Okui, T.; Obata, K.; Takabatake, K.; Kawai, H.; et al. Nicotine promotes lymph node metastasis and cetuximab resistance in head and neck squamous cell carcinoma. Int. J. Oncol. 2019, 54, 283–294. [Google Scholar] [CrossRef] [Green Version]
  74. Schaal, C.M.; Bora-Singhal, N.; Kumar, D.M.; Chellappan, S.P. Regulation of Sox2 and stemness by nicotine and electronic-cigarettes in non-small cell lung cancer. Mol. Cancer 2018, 17, 149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Li, Z.J.; Cho, C.H. Neurotransmitters, more than meets the eye–neurotransmitters and their perspectives in cancer development and therapy. Eur. J. Pharm. 2011, 2, 667. [Google Scholar] [CrossRef] [PubMed]
  76. Patane, S. M3 muscarinic acetylcholine receptor in cardiology and oncology. Int. J. Cardiol. 2014, 4, 177. [Google Scholar] [CrossRef] [PubMed]
  77. Yu, H.; Xia, H.; Tang, Q.; Xu, H.; Wei, G.; Chen, Y.; Dai, X.; Gong, Q.; Bi, F. Acetylcholine acts through M3 muscarinic receptor to activate the EGFR signaling and promotes gastric cancer cell proliferation. Sci. Rep. 2017, 7, 40802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Hayakawa, Y.; Sakitani, K.; Konishi, M.; Asfaha, S.; Niikura, R.; Tomita, H.; Renz, B.W.; Tailor, Y.; Macchini, M.; Middelhoff, M.; et al. Nerve growth factor promotes gastric tumorigenesis through aberrant cholinergic signaling. Cancer Cell 2017, 31, 21–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Renz, B.W.; Tanaka, T.; Sunagawa, M.; Takahashi, R.; Jiang, Z.; Macchini, M.; Dantes, Z.; Valenti, G.; White, R.A.; Middelhoff, M.A.; et al. Cholinergic signaling via muscarinic receptors directly and indirectly suppresses pancreatic tumorigenesis and cancer stemness. Cancer Discov. 2018, 8, 1458–1473. [Google Scholar] [CrossRef] [Green Version]
  80. Partecke, L.I.; Käding, A.; Trung, D.N.; Diedrich, S.; Sendler, M.; Weiss, F.; Kühn, J.-P.; Mayerle, J.; Beyer, K.; von Bernstorff, W.; et al. Subdiaphragmatic vagotomy promotes tumor growth and reduces survival via TNFα in a murine pancreatic cancer model. Oncotarget 2017, 8, 22501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Kamiya, A.; Hayama, Y.; Kato, S.; Shimomura, A.; Shimomura, T.; Irie, K.; Kaneko, R.; Yanagawa, Y.; Kobayashi, K.; Ochiya, T. Genetic manipulation of autonomic nerve fiber innervation and activity and its effect on breast cancer progression. Nat. Neurosci. 2019, 22, 1289–1305. [Google Scholar] [CrossRef]
  82. Fujii, T.; Mashimo, M.; Moriwaki, Y.; Misawa, H.; Ono, S.; Horiguchi, K. Physiological functions of the cholinergic system in immune cells. J. Pharm. Sci. 2017, 1, 134. [Google Scholar] [CrossRef]
  83. Zila, I.; Mokra, D.; Kopincova, J.; Kolomaznik, M.; Javorka, M.; Calkovska, A. Vagal-immune interactions involved in cholinergic anti-inflammatory pathway. Physiol. Res. 2017, 1, 66. [Google Scholar] [CrossRef]
  84. Yang, M.-W.; Tao, L.-Y.; Jiang, Y.-S.; Yang, J.-Y.; Huo, Y.-M.; Liu, D.-J.; Li, J.; Fu, X.-L.; He, R.; Lin, C.; et al. Perineural invasion reprograms the immune microenvironment through cholinergic signaling in pancreatic ductal adenocarcinoma. Cancer Res. 2020, 80, 1991–2003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Venkataramani, V.; Tanev, D.I.; Strahle, C.; Studier-Fischer, A.; Fankhauser, L.; Kessler, T.; Körber, C.; Kardorff, M.; Ratliff, M.; Xie, R.; et al. Glutamatergic synaptic input to glioma cells drives brain tumour progression. Nature 2019, 573, 532–538. [Google Scholar] [CrossRef] [PubMed]
  86. Venkatesh, H.S.; Morishita, W.; Geraghty, A.C.; Silverbush, D.; Gillespie, S.M.; Arzt, M.; Tam, L.T.; Espenel, C.; Ponnuswami, A.; Ni, L.; et al. Electrical and synaptic integration of glioma into neural circuits. Nature 2019, 573, 539–545. [Google Scholar] [CrossRef]
  87. Zeng, Q.; Michael, I.P.; Zhang, P.; Saghafinia, S.; Knott, G.; Jiao, W.; McCabe, B.D.; Galván, J.A.; Robinson, H.P.; Zlobec, I.; et al. Synaptic proximity enables NMDAR signalling to promote brain metastasis. Nature 2019, 573, 526–531. [Google Scholar] [CrossRef] [PubMed]
  88. John Lin, C.-C.; Yu, K.; Hatcher, A.; Huang, T.-W.; Lee, H.K.; Carlson, J.; Weston, M.C.; Chen, F.; Zhang, Y.; Zhu, W.; et al. Identification of diverse astrocyte populations and their malignant analogs. Nat. Neurosci. 2017, 20, 396–405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Chaichana, K.L.; Parker, S.L.; Olivi, A.; Quiñones-Hinojosa, A. Long-term seizure outcomes in adult patients undergoing primary resection of malignant brain astrocytomas. J. Neurosurg. 2009, 111, 282–292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Rub’i, B.; Maechler, P. Minireview: New roles for peripheral dopamine on metabolic control and tumor growth: Let’s seek the balance. Endocrinology 2010, 151, 5570–5581. [Google Scholar] [CrossRef] [Green Version]
  91. Roney, M.S.I.; Park, S.-K. Antipsychotic dopamine receptor antagonists, cancer, and cancer stem cells. Arch. Pharm. Res. 2018, 41, 384–408. [Google Scholar] [CrossRef]
  92. Dolma, S.; Selvadurai, H.J.; Lan, X.; Lee, L.; Kushida, M.; Voisin, V.; Whetstone, H.; So, M.; Aviv, T.; Park, N.; et al. Inhibition of dopamine receptor D4 impedes autophagic flux, proliferation, and survival of glioblastoma stem cells. Cancer Cell 2016, 29, 859–873. [Google Scholar] [CrossRef] [PubMed]
  93. Jandaghi, P.; Najafabadi, H.S.; Bauer, A.S.; Papadakis, A.I.; Fassan, M.; Hall, A.; Monast, A.; von Knebel Doeberitz, M.; Neoptolemos, J.P.; Costello, E.; et al. Expression of DRD2 is increased in human pancreatic ductal adenocarcinoma and inhibitors slow tumor growth in mice. Gastroenterology 2016, 151, 1218–1231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Sarkar, C.; Chakroborty, D.; Chowdhury, U.R.; Dasgupta, P.S.; Basu, S. Dopamine increases the efficacy of anticancer drugs in breast and colon cancer preclinical models. Clin. Cancer Res. 2008, 14, 2502–2510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Chakroborty, D.; Sarkar, C.; Mitra, R.B.; Banerjee, S.; Dasgupta, P.S.; Basu, S. Depleted dopamine in gastric cancer tissues: Dopamine treatment retards growth of gastric cancer by inhibiting angiogenesis. Clin. Cancer Res. 2004, 10, 4349–4356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Chakroborty, D.; Chowdhury, U.R.; Sarkar, C.; Baral, R.; Dasgupta, P.S.; Basu, S. Dopamine regulates endothelial progenitor cell mobilization from mouse bone marrow in tumor vascularization. J. Clin. Investig. 2008, 118, 1380–1389. [Google Scholar] [CrossRef] [Green Version]
  97. Pinoli, M.; Marino, F.; Cosentino, M. Dopaminergic regulation of innate immunity: A review. J. Neuroimmune Pharmacol. 2017, 12, 602–623. [Google Scholar] [CrossRef]
  98. Wu, J.; Zhang, R.; Tang, N.; Gong, Z.; Zhou, J.; Chen, Y.; Chen, K.; Cai, W. Dopamine inhibits the function of Gr-1+ CD115+ myeloid-derived suppressor cells through D1-like receptors and enhances anti-tumor immunity. J. Leukoc. Biol. 2015, 97, 191–200. [Google Scholar] [CrossRef] [Green Version]
  99. Cosentino, M.; Fietta, A.M.; Ferrari, M.; Rasini, E.; Bombelli, R.; Carcano, E.; Saporiti, F.; Meloni, F.; Marino, F.; Lecchini, S. Human CD4+ CD25+ regulatory T cells selectively express tyrosine hydroxylase and contain endogenous catecholamines subserving an autocrine/paracrine inhibitory functional loop. Blood 2007, 109, 632–642. [Google Scholar] [CrossRef] [Green Version]
  100. Maemura, K.; Shiraishi, N.; Sakagami, K.; Kawakami, K.; Inoue, T.; Murano, M. Proliferative effects of gamma-aminobutyric acid on the gastric cancer cell line are associated with extracellular signal-regulated kinase 1/2 activation. J. Gastroenterol. Hepatol. 2009, 2, 24. [Google Scholar]
  101. Hujber, Z.; Horváth, G.; PetHovári, G.; Krencz, I.; Dankó, T.; Mészáros, K.; Rajnai, H.; Szoboszlai, N.; Leenders, W.P.; Jeney, A.; et al. GABA, glutamine, glutamate oxidation and succinic semialdehyde dehydrogenase expression in human gliomas. J. Exp. Clin. Cancer Res. 2018, 37, 271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Zhang, D.; Li, X.; Yao, Z.; Wei, C.; Ning, N.; Li, J. GABAergic signaling facilitates breast cancer metastasis by promoting ERK1/2-dependent phosphorylation. Cancer Lett. 2014, 348, 100–108. [Google Scholar] [CrossRef]
  103. Kanbara, K.; Otsuki, Y.; Watanabe, M.; Yokoe, S.; Mori, Y.; Asahi, M.; Neo, M. GABA B receptor regulates proliferation in the high-grade chondrosarcoma cell line OUMS-27 via apoptotic pathways. BMC Cancer 2018, 18, 263. [Google Scholar] [CrossRef]
  104. Sung, H.Y.; Yang, S.-D.; Ju, W.; Ahn, J.-H. Aberrant epigenetic regulation of GABRP associates with aggressive phenotype of ovarian cancer. Exp. Mol. Med. 2017, 49, e335. [Google Scholar] [CrossRef] [PubMed]
  105. Zhang, X.; Du, Z.; Liu, J.; He, J. Gamma-aminobutyric acid receptors affect the progression and migration of tumor cells. J. Recept. Signal Transduct. Res. 2014, 1, 34. [Google Scholar]
  106. Gumireddy, K.; Li, A.; Kossenkov, A.V.; Sakurai, M.; Yan, J.; Li, Y.; Xu, H.; Wang, J.; Zhang, P.J.; Zhang, L.; et al. The mRNA-edited form of GABRA3 suppresses GABRA3-mediated Akt activation and breast cancer metastasis. Nat. Commun. 2016, 7, 10715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Jiang, S.H.; Zhu, L.L.; Zhang, M.; Li, R.K.; Yang, Q.; Yan, J.Y.; Zhang, C.; Yang, J.Y.; Dong, F.Y.; Dai, M.; et al. GABRP regulates chemokine signalling, macrophage recruitment and tumour progression in pancreatic cancer through tuning KCNN4-mediated Ca2+ signalling in a GABA-independent manner. Gut 2019, 68, 1551. [Google Scholar] [CrossRef] [PubMed]
  108. Abdul, M.; Mccray, S.D.; Hoosein, N.M. Expression of gamma-aminobutyric acid receptor (subtype A) in prostate cancer. Acta Oncol. 2008, 47, 1546–1550. [Google Scholar] [CrossRef]
  109. Takehara, A.; Hosokawa, M.; Eguchi, H.; Ohigashi, H.; Ishikawa, O.; Nakamura, Y.; Nakagawa, H. γ-aminobutyric acid (GABA) stimulates pancreatic cancer growth through overexpressing GABAA receptor π subunit. Cancer Res. 2007, 67, 9704–9712. [Google Scholar] [CrossRef] [Green Version]
  110. Wang, T.; Huang, W.; Chen, F. Baclofen, a GABAB receptor agonist, inhibits human hepatocellular carcinoma cell growth in vitro and in vivo. Life Sci. 2008, 82, 536–541. [Google Scholar] [CrossRef]
  111. Schuller, H.M. Regulatory role of G protein-coupled receptors in pancreatic cancer development and progression. Curr. Med. Chem. 2018, 25, 2566–2575. [Google Scholar] [CrossRef] [PubMed]
  112. Tatsuta, M.; Iishi, H.; Baba, M.; Nakaizumi, A.; Ichii, M.; Taniguchi, H. Inhibition by γ-Amino-n-butyric Acid and Baclofen of Gastric Carcinogenesis Induced by N′-Methyl-N′-nitro-N′-nitrosoguanidine in Wistar Rats. Cancer Res. 1990, 50, 4931–4934. [Google Scholar] [PubMed]
  113. Joseph, J.; Niggemann, B.; Zaenker, K.S.; Entschladen, F. The neurotransmitter γ-aminobutyric acid is an inhibitory regulator for the migration of SW 480 colon carcinoma cells. Cancer Res. 2002, 62, 6467–6469. [Google Scholar]
  114. Thaker, P.H.; Yokoi, K.; Jennings, N.B.; Li, Y.; Rebhun, R.B.; Rousseau Jr, D.L.; Fan, D.; Sood, A.K. Inhibition of experimental colon cancer metastasis by the GABA-receptor agonist nembutal. Cancer Biol. Ther. 2005, 4, 753–758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Schuller, H.M. Neurotransmission and cancer: Implications for prevention and therapy. Anti-Cancer Drugs 2008, 19, 655–671. [Google Scholar] [CrossRef]
  116. Azuma, H.; Inamoto, T.; Sakamoto, T.; Kiyama, S.; Ubai, T.; Shinohara, Y.; Maemura, K.; Tsuji, M.; Segawa, N.; Masuda, H.; et al. γ-Aminobutyric acid as a promoting factor of cancer metastasis; induction of matrix metalloproteinase production is potentially its underlying mechanism. Cancer Res. 2003, 63, 8090–8096. [Google Scholar]
  117. Xia, S.; He, C.; Zhu, Y.; Wang, S.; Li, H.; Zhang, Z.; Jiang, X.; Liu, J. GABABR-induced EGFR transactivation promotes migration of human prostate cancer cells. Mol. Pharmacol. 2017, 92, 265–277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Wang, Q.; Ren, L.; Wan, Y.; Prud’homme, G.J. GABAergic regulation of pancreatic islet cells: Physiology and antidiabetic effects. J. Cell. Physiol. 2019, 234, 14432–14444. [Google Scholar] [CrossRef]
  119. Mohammad-Zadeh, L.F.; Moses, L.; Gwaltney-Brant, S.M. Serotonin: A review. J. Vet. Pharmacol. Ther. 2008, 31, 187–199. [Google Scholar] [CrossRef] [PubMed]
  120. Kim, H.; Toyofuku, Y.; Lynn, F.C.; Chak, E.; Uchida, T.; Mizukami, H.; Fujitani, Y.; Kawamori, R.; Miyatsuka, T.; Kosaka, Y.; et al. Serotonin regulates pancreatic beta cell mass during pregnancy. Nat. Med. 2010, 16, 804. [Google Scholar] [CrossRef] [PubMed]
  121. Chabbi-Achengli, Y.; Coudert, A.E.; Callebert, J.; Geoffroy, V.; Côté, F.; Collet, C.; de Vernejoul, M.-C. Decreased osteoclastogenesis in serotonin-deficient mice. Proc. Natl. Acad. Sci. USA 2012, 109, 2567–2572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Siddiqui, E.J.; Shabbir, M.; Mikhailidis, D.P.; Thompson, C.S.; Mumtaz, F.H. The Role of Serotonin (5-Hydroxytryptamine1A and 1B) Receptors in Prostate Cancer Cell Proliferation. J. Urol. 2006, 176, 1648–1653. [Google Scholar] [CrossRef]
  123. Siddiqui, E.J.; Shabbir, M.A.; Mikhailidis, D.P.; Mumtaz, F.H.; Thompson, C.S. The effect of serotonin and serotonin antagonists on bladder cancer cell proliferation. BJU Int. 2006, 97, 634–639. [Google Scholar] [CrossRef] [PubMed]
  124. Liu, Y.; Zhang, H.; Wang, Z.; Wu, P.; Gong, W. 5-Hydroxytryptamine1a receptors on tumour cells induce immune evasion in lung adenocarcinoma patients with depression via autophagy/pSTAT3. Eur. J. Cancer 2019, 1, 114. [Google Scholar] [CrossRef] [PubMed]
  125. Del Bello, F.; Bonifazi, A.; Giorgioni, G.; Quaglia, W.; Amantini, C.; Morelli, M.B.; Santoni, G.; Battiti, F.O.; Vistoli, G.; Cilia, A.; et al. Chemical manipulations on the 1, 4-dioxane ring of 5-HT1A receptor agonists lead to antagonists endowed with antitumor activity in prostate cancer cells. Eur. J. Med. Chem. 2019, 168, 461–473. [Google Scholar] [CrossRef]
  126. Gautam, J.; Banskota, S.; Regmi, S.C.; Ahn, S.; Jeon, Y.H.; Jeong, H.; Kim, S.J.; Nam, T.; Jeong, B.-S.; Kim, J.-A. Tryptophan hydroxylase 1 and 5-HT 7 receptor preferentially expressed in triple-negative breast cancer promote cancer progression through autocrine serotonin signaling. Mol. Cancer 2016, 15, 75. [Google Scholar] [CrossRef]
  127. Sarrouilhe, D.; Mesnil, M. Serotonin and human cancer: A critical view. Biochimie 2019, 161, 46–50. [Google Scholar] [CrossRef]
  128. Dizeyi, N.; Bjartell, A.; Nilsson, E.; Hansson, J.; Gadaleanu, V.; Cross, N.; Abrahamsson, P.-A. Expression of Serotonin Receptors and Role of Serotonin in Human Prostate Cancer Tissue and Cell Lines. Prostate 2004, 59, 328–336. [Google Scholar] [CrossRef]
  129. Sarrouilhe, D.; Clarhaut, J.; Defamie, N.; Mesnil, M. Serotonin and cancer: What is the link? Curr. Mol. Med. 2015, 15, 62–77. [Google Scholar] [CrossRef]
  130. Jose, J.; Tavares, C.D.; Ebelt, N.D.; Lodi, A.; Edupuganti, R.; Xie, X.; Devkota, A.K.; Kaoud, T.S.; Van Den Berg, C.L.; Anslyn, E.V.; et al. Serotonin analogues as inhibitors of breast cancer cell growth. ACS Med. Chem. Lett. 2017, 8, 1072–1076. [Google Scholar] [CrossRef] [PubMed]
  131. Jiang, S.-H.; Li, J.; Dong, F.-Y.; Yang, J.-Y.; Liu, D.-J.; Yang, X.-M.; Wang, Y.-H.; Yang, M.-W.; Fu, X.-L.; Zhang, X.-X.; et al. Increased serotonin signaling contributes to the Warburg effect in pancreatic tumor cells under metabolic stress and promotes growth of pancreatic tumors in mice. Gastroenterology 2017, 153, 277–291. [Google Scholar] [CrossRef] [PubMed]
  132. Mammadova-Bach, E.; Mauler, M.; Braun, A.; Duerschmied, D. Autocrine and paracrine regulatory functions of platelet serotonin. Platelets 2018, 29, 541–548. [Google Scholar] [CrossRef]
  133. Asada, M.; Ebihara, S.; Yamanda, S.; Niu, K.; Okazaki, T.; Sora, I.; Arai, H. Depletion of serotonin and selective inhibition of 2B receptor suppressed tumor angiogenesis by inhibiting endothelial nitric oxide synthase and extracellular signal-regulated kinase 1/2 phosphorylation. Neoplasia 2009, 11, 408–410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Ribeiro, M.P.; Custódio, J.B.; Santos, A.E. Ionotropic glutamate receptor antagonists and cancer therapy: Time to think out of the box? Cancer Chemother. Pharmacol. 2017, 79, 219–225. [Google Scholar] [CrossRef] [PubMed]
  135. Lumeng, J.Y.; Wall, B.A.; Wangari-Talbot, J.; Chen, S. Metabotropic glutamate receptors in cancer. Neuropharmacology 2017, 115, 193–202. [Google Scholar]
  136. Yoo, B.C.; Jeon, E.; Hong, S.-H.; Shin, Y.-K.; Chang, H.J.; Park, J.-G. Metabotropic glutamate receptor 4-mediated 5-fluorouracil resistance in a human colon cancer cell line. Clin. Cancer Res. 2004, 10, 4176–4184. [Google Scholar] [CrossRef] [Green Version]
  137. Groot, J.F.; Piao, Y.; Lu, L.; Fuller, G.N.; Yung, W.K.A. Knockdown of GluR1 expression by RNA interference inhibits glioma proliferation. J. Neuro-Oncol. 2008, 88, 121–133. [Google Scholar] [CrossRef]
  138. Liu, J.-W.; Myoung, S.K.; Nagpal, J.; Yamashita, K.; Poeta, L.; Chang, X.; Lee, J.; Park, H.L.; Jeronimo, C.; Westra, W.H.; et al. Quantitative hypermethylation of NMDAR2B in human gastric cancer. Int. J. Cancer 2007, 121, 1994–2000. [Google Scholar] [CrossRef]
  139. Park, S.-Y.; Lee, S.-A.; Han, I.-H.; Yoo, B.-C.; Lee, S.-H.; Park, J.-Y.; Cha, I.-H.; Kim, J.; Choi, S.-W. Clinical significance of metabotropic glutamate receptor 5 expression in oral squamous cell carcinoma. Oncol. Rep. 2007, 17, 81–87. [Google Scholar] [CrossRef] [Green Version]
  140. Pollock, P.M.; Cohen-Solal, K.; Sood, R.; Namkoong, J.; Martino, J.J.; Koganti, A.; Zhu, H.; Robbins, C.; Makalowska, I.; Shin, S.-S.; et al. Melanoma mouse model implicates metabotropic glutamate signaling in melanocytic neoplasia. Nat. Genet. 2003, 34, 108–112. [Google Scholar] [CrossRef] [PubMed]
  141. Choi, S.-W.; Park, S.-Y.; Hong, S.-P.; Pai, H.; Choi, J.-Y.; Kim, S.-G. The expression of NMDA receptor 1 is associated with clinicopathological parameters and prognosis in the oral squamos cell carcinoma. J. Oral Pathol. Med. 2004, 33, 533–537. [Google Scholar] [CrossRef] [PubMed]
  142. Chang, H.J.; Yoo, B.C.; Lim, S.-B.; Jeong, S.-Y.; Kim, W.H.; Park, J.-G. Metabotropic glutamate receptor 4 expression in colorectal carcinoma and its prognostic significance. Clin. Cancer Res. 2005, 11, 3288–3295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Iacovelli, L.; Arcella, A.; Battaglia, G.; Pazzaglia, S.; Aronica, E.; Spinsanti, P.; Caruso, A.; De Smaele, E.; Saran, A.; Gulino, A.; et al. Pharmacological activation of mGlu4 metabotropic glutamate receptors inhibits the growth of medulloblastomas. J. Neurosci. 2006, 26, 8388–8397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Palamiuc, L.; Emerling, B.M. PSMA brings new flavors to PI3K signaling: A role for glutamate in prostate cancer. J. Exp. Med. 2018, 215, 17. [Google Scholar] [CrossRef] [PubMed]
  145. Banh, R.S.; Biancur, D.E.; Yamamoto, K.; Sohn, A.S.; Walters, B.; Kuljanin, M.; Gikandi, A.; Wang, H.; Mancias, J.D.; Schneider, R.J.; et al. Neurons release serine to support mRNA translation in pancreatic cancer. Cell 2020, 183, 1202–1218. [Google Scholar] [CrossRef] [PubMed]
  146. Kaczy’nska, K.; Zajkac, D.; Wojciechowski, P.; Kogut, E.; Szereda-Przestaszewska, M. Neuropeptides and breathing in health and disease. Pulm. Pharmacol. Ther. 2018, 48, 217–224. [Google Scholar] [CrossRef] [PubMed]
  147. Covenas, R.; Munoz, M. Cancer progression and substance P. Histol. Histopathol. 2014, 1, 29. [Google Scholar]
  148. Tilan, J.; Kitlinska, J. Neuropeptide Y (NPY) in tumor growth and progression: Lessons learned from pediatric oncology. Neuropeptides 2016, 10, 55. [Google Scholar] [CrossRef]
  149. Munoz, M.; Rosso, M.; Covenas, R. The NK-1 receptor: A new target in cancer therapy. Curr. Drug Targets 2011, 1, 12. [Google Scholar] [CrossRef]
  150. Munoz, M.; Rosso, M.; Covenas, R. The NK-1 receptor antagonist L-732,138 induces apoptosis in human gastrointestinal cancer cell lines. Pharm. Rep. 2017, 6, 69. [Google Scholar]
  151. Zhang, L.; Wang, L.; Dong, D.; Wang, Z.; Ji, W.; Yu, M.; Zhang, F.; Niu, R.; Zhou, Y. MiR-34b/c-5p and the neurokinin-1 receptor regulate breast cancer cell proliferation and apoptosis. Cell Prolif. 2019, 52, e12527. [Google Scholar] [CrossRef]
  152. Muñoz, M.; Coveñas, R. Glioma and neurokinin-1 receptor antagonists: A new therapeutic approach. Anti-Cancer Agents Med. Chem. 2019, 19, 92–100. [Google Scholar] [CrossRef] [PubMed]
  153. Li, J.; Tian, Y.; Wu, A. Neuropeptide Y receptors: A promising target for cancer imaging and therapy. Regen. Biomater. 2015, 2, 215–219. [Google Scholar] [CrossRef] [PubMed]
  154. Galli, S.; Naranjo, A.; Van Ryn, C.; Tilan, J.U.; Trinh, E.; Yang, C.; Tsuei, J.; Hong, S.-H.; Wang, H.; Izycka-Swieszewska, E.; et al. Neuropeptide Y as a biomarker and therapeutic target for neuroblastoma. Am. J. Pathol. 2016, 186, 3040–3053. [Google Scholar] [CrossRef] [Green Version]
  155. Czarnecka, M.; Trinh, E.; Lu, C.; Kuan-Celarier, A.; Galli, S.; Hong, S.-H.; Tilan, J.U.; Talisman, N.; Izycka-Swieszewska, E.; Tsuei, J.; et al. Neuropeptide Y receptor Y5 as an inducible pro-survival factor in neuroblastoma: Implications for tumor chemoresistance. Oncogene 2015, 34, 3131–3143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Jeppsson, S.; Srinivasan, S.; Chandrasekharan, B. Neuropeptide Y (NPY) promotes inflammation-induced tumorigenesis by enhancing epithelial cell proliferation. Am. J. Physiol. -Gastrointest. Liver Physiol. 2017, 312, G103–G111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Molloy, N.H.; Read, D.E.; Gorman, A.M. Nerve growth factor in cancer cell death and survival. Cancers 2011, 3, 510–530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Pundavela, J.; Roselli, S.; Faulkner, S.; Attia, J.; Scott, R.J.; Thorne, R.F.; Forbes, J.F.; Bradshaw, R.A.; Walker, M.M.; Jobling, P.; et al. Nerve fibers infiltrate the tumor microenvironment and are associated with nerve growth factor production and lymph node invasion in breast cancer. Mol. Oncol. 2015, 9, 1626–1635. [Google Scholar] [CrossRef] [Green Version]
  159. Ylivinkka, I.; Keski-Oja, J.; Hyytiäinen, M. Netrin-1: A regulator of cancer cell motility? Eur. J. Cell Biol. 2016, 95, 513–520. [Google Scholar] [CrossRef]
  160. Bruikman, C.S.; Zhang, H.; Kemper, A.M.; van Gils, J.M. Netrin family: Role for protein isoforms in cancer. J. Nucleic Acids 2019, 2019, 3947123. [Google Scholar] [CrossRef] [PubMed]
  161. Rehman, M.; Tamagnone, L. Semaphorins in cancer: Biological mechanisms and therapeutic approaches. Semin. Cell Dev. Biol. 2013, 1, 179–189. [Google Scholar] [CrossRef] [PubMed]
  162. Mishra, R.; Kumar, D.; Tomar, D.; Chakraborty, G.; Kumar, S.; Kundu, G.C. The potential of class 3 semaphorins as both targets and therapeutics in cancer. Expert Opin. Ther. Targets 2015, 19, 427–442. [Google Scholar] [CrossRef] [PubMed]
  163. Killeen, M.T.; Sybingco, S.S. Netrin, Slit and Wnt receptors allow axons to choose the axis of migration. Dev. Biol. 2008, 323, 143–151. [Google Scholar] [CrossRef] [Green Version]
  164. Jiang, S.; Richaud, M.; Vieugué, P.; Rama, N.; Delcros, J.-G.; Siouda, M.; Sanada, M.; Redavid, A.; Ducarouge, B.; Hervieu, M.; et al. Targeting netrin-3 in small cell lung cancer and neuroblastoma. EMBO Mol. Med. 2021, 13, e12878. [Google Scholar] [CrossRef] [PubMed]
  165. Arakawa, H. Netrin-1 and its receptors in tumorigenesis. Nat. Rev. 2004, 4, 978–987. [Google Scholar] [CrossRef] [PubMed]
  166. Meyerhardt, J.A.; Caca, K.; Eckstrand, B.C.; Hu, G.; Lengauer, C.; Banavali, S.; Look, A.T.; Fearon, E.R. Netrin-1: Interaction with deleted in colorectal cancer (DCC) and alterations in brain tumors and neuroblastomas. Cell Growth Differ. 1999, 10, 35–42. [Google Scholar]
  167. Latil, A.; Chêne, L.; Cochant-Priollet, B.; Mangin, P.; Fournier, G.; Berthon, P.; Cussenot, O. Quantification of expression of netrins, slits and their receptors in human prostate tumors. Int. J. Cancer 2003, 103, 306–315. [Google Scholar] [CrossRef] [PubMed]
  168. Mehlen, P.; Fearon, E.R. Role of the dependence receptor DCC in colorectal cancer pathogenesis. J. Clin. Oncol. Oncol. 2004, 22, 3420–3428. [Google Scholar] [CrossRef] [PubMed]
  169. Boussouar, A.; Tortereau, A.; Manceau, A.; Paradisi, A.; Gadot, N.; Vial, J.; Neves, D.; Larue, L.; Battistella, M.; Leboeuf, C.; et al. Netrin-1 and Its Receptor DCC Are Causally Implicated in Melanoma ProgressionDCC and Netrin-1 in Melanoma. Cancer Res. 2020, 80, 747–756. [Google Scholar] [CrossRef] [PubMed]
  170. Sung, P.-J.; Rama, N.; Imbach, J.; Fiore, S.; Ducarouge, B.; Neves, D.; Chen, H.-W.; Bernard, D.; Yang, P.-C.; Bernet, A.; et al. Cancer-associated fibroblasts produce netrin-1 to control cancer cell plasticity. Cancer Res. 2019, 79, 3651–3661. [Google Scholar] [CrossRef] [PubMed]
  171. Paradisi, A.; Mehlen, P. Netrin-1, a missing link between chronic inflammation and tumor progression. Cell Cycle 2010, 9, 1253–1262. [Google Scholar] [CrossRef] [Green Version]
  172. Qi, Q.; Li, D.Y.; Luo, H.R.; Guan, K.-L.; Ye, K. Netrin-1 exerts oncogenic activities through enhancing Yes-associated protein stability. Proc. Natl. Acad. Sci. USA 2015, 112, 7255–7260. [Google Scholar] [CrossRef] [Green Version]
  173. Kullander, K.; Klein, R. Mechanisms and functions of Eph and ephrin signalling. Nat. Rev. Mol. Cell Biol. 2002, 3, 475–486. [Google Scholar] [CrossRef] [PubMed]
  174. Cutforth, T.; Moring, L.; Mendelsohn, M.; Nemes, A.; Shah, N.M.; Kim, M.M.; Frisén, J.; Axel, R. Axonal ephrin-As and odorant receptors: Coordinate determination of the olfactory sensory map. Cell 2003, 114, 311–322. [Google Scholar] [CrossRef] [Green Version]
  175. Papadakos, S.P.; Petrogiannopoulos, L.; Pergaris, A.; Theocharis, S. The EPH/Ephrin System in Colorectal Cancer. Int. J. Mol. Sci. 2022, 23, 2761. [Google Scholar] [CrossRef]
  176. Kakarla, M.; ChallaSivaKanaka, S.; Dufficy, M.F.; Gil, V.; Filipovich, Y.; Vickman, R.; Crawford, S.E.; Hayward, S.W.; Franco, O.E. Ephrin B Activate Src Family Kinases in Fibroblasts Inducing Stromal Remodeling in Prostate Cancer. Cancers 2022, 14, 2336. [Google Scholar] [CrossRef] [PubMed]
  177. Shiuan, E.; Chen, J. Eph receptor tyrosine kinases in tumor immunity. Cancer Res. 2016, 76, 6452–6457. [Google Scholar] [CrossRef] [Green Version]
  178. Broggini, T.; Piffko, A.; Hoffmann, C.J.; Ghori, A.; Harms, C.; Adams, R.H.; Vajkoczy, P.; Czabanka, M. Ephrin-B2-EphB4 communication mediates tumor-endothelial cell interactions during hematogenous spread to spinal bone in a melanoma metastasis model. Oncogene 2020, 39, 7063–7075. [Google Scholar] [CrossRef] [PubMed]
  179. Fox, B.P.; Kandpal, R.P. Invasiveness of breast carcinoma cells and transcript profile: Eph receptors and ephrin ligands as molecular markers of potential diagnostic and prognostic application. Biochem. Biophys. Res. Commun. 2004, 318, 882–892. [Google Scholar] [CrossRef] [PubMed]
  180. Zeng, G.; Hu, Z.; Kinch, M.S.; Pan, C.-X.; Flockhart, D.A.; Kao, C.; Gardner, T.A.; Zhang, S.; Li, L.; Baldridge, L.A.; et al. High-level expression of EphA2 receptor tyrosine kinase in prostatic intraepithelial neoplasia. Am. J. Pathol. 2003, 163, 2271–2276. [Google Scholar] [CrossRef]
  181. Nikas, I.; Giaginis, C.; Petrouska, K.; Alexandrou, P.; Michail, A.; Sarantis, P.; Tsourouflis, G.; Danas, E.; Pergaris, A.; Politis, P.K.; et al. EPHA2, EPHA4, and EPHA7 Expression in Triple-Negative Breast Cancer. Diagnostics 2022, 12, 366. [Google Scholar] [CrossRef] [PubMed]
  182. Venkitachalam, S.; Babu, D.; Ravillah, D.; Katabathula, R.M.; Joseph, P.; Singh, S.; Udhayakumar, B.; Miao, Y.; Martinez-Uribe, O.; Hogue, J.A.; et al. The Ephrin B2 Receptor Tyrosine Kinase is a Regulator of Proto-oncogene MYC and Molecular Programs Central to Barrett’s Neoplasia. Gastroenterology 2022, 163, 1228–1241. [Google Scholar] [CrossRef] [PubMed]
  183. Castro-Rivera, E.; Ran, S.; Brekken, R.A.; Minna, J.D. Semaphorin 3B inhibits the phosphatidylinositol 3-kinase/Akt pathway through neuropilin-1 in lung and breast cancer cells. Cancer Res. 2008, 68, 8295–8303. [Google Scholar] [CrossRef] [Green Version]
  184. Caunt, M.; Mak, J.; Liang, W.-C.; Stawicki, S.; Pan, Q.; Tong, R.K.; Kowalski, J.; Ho, C.; Reslan, H.B.; Ross, J.; et al. Blocking neuropilin-2 function inhibits tumor cell metastasis. Cancer Cell 2008, 13, 331–342. [Google Scholar] [CrossRef] [Green Version]
  185. Leffers, N.; Gooden, M.J.; de Jong, R.A.; Hoogeboom, B.-N.; ten Hoor, K.A.; Hollema, H.; Boezen, H.M.; van der Zee, A.G.; Daemen, T.; Nijman, H.W. Prognostic significance of tumor-infiltrating T-lymphocytes in primary and metastatic lesions of advanced stage ovarian cancer. Cancer Immunol. Immunother. 2009, 58, 449–459. [Google Scholar] [CrossRef] [Green Version]
  186. Fard, D.; Tamagnone, L. Semaphorins in health and disease. CytokineGrowth Factor Rev. 2021, 57, 55–63. [Google Scholar] [CrossRef] [PubMed]
  187. Franzolin, G.; Tamagnone, L. Semaphorin signaling in cancer-associated inflammation. Int. J. Mol. Sci. 2019, 20, 377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Butti, R.; Kumar, T.V.; Nimma, R.; Kundu, G.C. Impact of semaphorin expression on prognostic characteristics in breast cancer. Breast Cancer Targets Ther. 2018, 10, 79. [Google Scholar] [CrossRef] [Green Version]
  189. Toledano, S.; Nir-Zvi, I.; Engelman, R.; Kessler, O.; Neufeld, G. Class-3 semaphorins and their receptors: Potent multifunctional modulators of tumor progression. Int. J. Mol. Sci. 2019, 20, 556. [Google Scholar] [CrossRef] [Green Version]
  190. Hagel, C.; Stavrou, D. Neuronal markers in non-neuronal tissues. Neuronal Act. Tumor Tissue 2007, 39, 64–77. [Google Scholar]
  191. Samarelli, A.V.; Riccitelli, E.; Bizzozero, L.; Silveira, T.N.; Seano, G.; Pergolizzi, M.; Vitagliano, G.; Cascone, I.; Carpentier, G.; Bottos, A.; et al. Neuroligin 1 induces blood vessel maturation by cooperating with the α6 integrin. J. Biol. Chem. 2014, 289, 19466–19476. [Google Scholar] [CrossRef] [PubMed]
  192. Rissone, A.; Foglia, E.; Sangiorgio, L.; Cermenati, S.; Nicoli, S.; Cimbro, S.; Beltrame, M.; Bussolino, F.; Cotelli, F.; Arese, M. The Synaptic Proteins β-Neurexin and Neuroligin Synergize With Extracellular Matrix-Binding Vascular Endothelial Growth Factor A During Zebrafish Vascular Development. Arterioscler. Thromb. Vasc. Biol. 2012, 32, 1563–1572. [Google Scholar] [CrossRef] [Green Version]
  193. Bottos, A.; Rissone, A.; Bussolino, F.; Arese, M. Neurexins and neuroligins: Synapses look out of the nervous system. Cell. Mol. Life Sci. 2011, 68, 2655–2666. [Google Scholar] [CrossRef] [PubMed]
  194. Bottos, A.; Destro, E.; Rissone, A.; Graziano, S.; Cordara, G.; Assenzio, B.; Cera, M.R.; Mascia, L.; Bussolino, F.; Arese, M. The synaptic proteins neurexins and neuroligins are widely expressed in the vascular system and contribute to its functions. Proc. Natl. Acad. Sci. USA 2009, 106, 20782–20787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Yu, Q.; Wang, X.; Yang, Y.; Chi, P.; Huang, J.; Qiu, S.; Zheng, X.; Chen, X. Upregulated NLGN1 predicts poor survival in colorectal cancer. BMC Cancer 2021, 21, 884. [Google Scholar] [CrossRef] [PubMed]
  196. Pergolizzi, M.; Bizzozero, L.; Maione, F.; Maldi, E.; Isella, C.; Macagno, M.; Mariella, E.; Bardelli, A.; Medico, E.; Marchiò, C.; et al. The neuronal protein Neuroligin 1 promotes colorectal cancer progression by modulating the APC/β-catenin pathway. J. Exp. Clin. Cancer Res. 2022, 41, 266. [Google Scholar] [CrossRef] [PubMed]
  197. Bizzozero, L.; Pergolizzi, M.; Pascal, D.; Maldi, E.; Villari, G.; Erriquez, J.; Volante, M.; Serini, G.; Marchiò, C.; Bussolino, F.; et al. Tumoral Neuroligin 1 Promotes Cancer-Nerve Interactions and Synergizes with the Glial Cell Line-Derived Neurotrophic Factor. Cells 2022, 11, 280. [Google Scholar] [CrossRef] [PubMed]
  198. Venkatesh, H.S.; Johung, T.B.; Caretti, V.; Noll, A.; Tang, Y.; Nagaraja, S.; Gibson, E.M.; Mount, C.W.; Polepalli, J.; Mitra, S.S.; et al. Neuronal activity promotes glioma growth through neuroligin-3 secretion. Cell 2015, 161, 803–816. [Google Scholar] [CrossRef] [Green Version]
  199. Suzuki, K.; Hayashi, Y.; Nakahara, S.; Kumazaki, H.; Prox, J.; Horiuchi, K.; Zeng, M.; Tanimura, S.; Nishiyama, Y.; Osawa, S.; et al. Activity-dependent proteolytic cleavage of neuroligin-1. Neuron 2012, 76, 410–422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Hunt, P.J.; Kabotyanski, K.E.; Calin, G.A.; Xie, T.; Myers, J.N.; Amit, M. Interrupting Neuron—Tumor Interactions to Overcome Treatment Resistance. Cancers 2020, 12, 3741. [Google Scholar] [CrossRef]
  201. Jung, E.; Alfonso, J.; Monyer, H.; Wick, W.; Winkler, F. Neuronal signatures in cancer. Int. J. Cancer 2020, 147, 3281–3291. [Google Scholar] [CrossRef] [PubMed]
  202. Arese, M.; Bussolino, F.; Pergolizzi, M.; Bizzozero, L.; Pascal, D. Tumor progression: The neuronal input. Ann. Transl. Med. 2018, 6, 2972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Mancino, M.; Ametller, E.; Gascón, P.; Almendro, V. The neuronal influence on tumor progression. Biochim. Biophys. Acta (BBA)—Rev. Cancer 2011, 1816, 105–118. [Google Scholar]
  204. Jiang, S.-H.; Hu, L.-P.; Wang, X.; Li, J.; Zhang, Z.-G. Neurotransmitters: Emerging targets in cancer. Oncogene 2020, 39, 503–515. [Google Scholar] [CrossRef] [PubMed]
  205. Hanahan, D. Hallmarks of cancer: New dimensions. Cancer Discov. 2022, 12, 31–46. [Google Scholar] [CrossRef] [PubMed]
  206. Mendler, L.; Pinter, S.; Kiricsi, M.; Baka, Z.; Dux, L. Regeneration of reinnervated rat soleus muscle is accompanied by fiber transition toward a faster phenotype. J. Histochem. Cytochem. 2008, 56, 111–123. [Google Scholar] [CrossRef] [PubMed]
  207. Gauthier, G.F.; Burke, R.E.; Lowey, S.; Hobbs, A.W. Myosin isozymes in normal and cross-reinnervated cat skeletal muscle fibers. J. Cell Biol. 1983, 97, 756–771. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The four main areas of cancer/nervous system interactions.
Figure 1. The four main areas of cancer/nervous system interactions.
Ijms 23 14695 g001
Figure 2. Examples of molecular mechanisms involved in PNI/perineural niche that are presented in this review. Major cell types in the perineural niche, as well as various soluble signaling molecules and their receptors, form a complex signal network that allows nerves and tumors to interact. For recent perineural-niche- and PNI-focused reviews, see [17,45].
Figure 2. Examples of molecular mechanisms involved in PNI/perineural niche that are presented in this review. Major cell types in the perineural niche, as well as various soluble signaling molecules and their receptors, form a complex signal network that allows nerves and tumors to interact. For recent perineural-niche- and PNI-focused reviews, see [17,45].
Ijms 23 14695 g002
Figure 3. Outline of the tumor cell/ nervous system interfaces and their biological effects on cancer cells that have been cited in this review. Up arrows indicate the upregulation of a certain pathway/phenotype. Down arrows indicate downregulation. CD: context-dependent, means that variable responses depend on the tumor type and various other experimental conditions. Intra- and intertumor heterogeneity most likely play a role. All other abbreviations can be found in the main text.
Figure 3. Outline of the tumor cell/ nervous system interfaces and their biological effects on cancer cells that have been cited in this review. Up arrows indicate the upregulation of a certain pathway/phenotype. Down arrows indicate downregulation. CD: context-dependent, means that variable responses depend on the tumor type and various other experimental conditions. Intra- and intertumor heterogeneity most likely play a role. All other abbreviations can be found in the main text.
Ijms 23 14695 g003
Table 1. A schematic presentation of the neurotransmitters/amino acids/growth factors considered in this review, with molecular data. More details on the intracellular pathways, including the abbreviations used, can be found in the main text and references.
Table 1. A schematic presentation of the neurotransmitters/amino acids/growth factors considered in this review, with molecular data. More details on the intracellular pathways, including the abbreviations used, can be found in the main text and references.
FamilyMemberCellular ReceptorAffected Intracellular PathwaysRefs
Neurotransmitters/amino acidsNorepinephrine (noradrenaline)α-β adrenergic receptorsGαs–cAMP, MAPK, PKA, EPAC
ERK/cyclooxygenase 2
CREB1-miR-373 axis, Src
PKA, VDCC, Hippo
[50,56,58,61,64]
DopamineDopamine 1–5 (D1–5) receptorsVEGFA-mediated ERK1/2 phosphorylation[95,96]
Serotonin (5-hydroxytryptamine)5-HT1-7 receptorsAkt/Mtor, glycolysis, endothelial nitric oxide synthase, p-ERK1/2[130,131,133]
AcetylcholineMuscarinic acetylcholine receptor, nicotinic acetylcholine receptorCa2+ flux, PKC/ERK1/2, COX2, PGE2, CREB, SRC, AKT, Ras-RAF1, MAPK, survivin, BCL-2, NF-κB, PI3K/AKT[2,5,73,75,76,79]
GlutamateMetabotropic glutamate receptor, NMDA receptor, Kainate receptor, AMPA receptorPI3K (p110)[144]
Gamma-aminobutyric acidGABAB GABAA
GABAc
receptors
Ca2+ flux
MAPK/ERK
[103,109]
SerineAmino acid transportersmRNA translation[145]
NeuropeptidesNeuropeptide YNPY1R, NPY2R, PPYR1, NPY5RPI3-K/pAkt
microRNA-375
[156]
Substance PNK1, NNK2, NNK3 inositol 1,4,5-triphosphate (IP3) DAG, c-myc, mitogen-activated protein kinases, activator protein 1, extracellular signal-regulated kinases 1 and 2, glycogen breakdown[149,152]
Growth factorsNGFTrkA, p75NTRPI3K, MAPK
c-Jun N-terminal kinase
[157]
Axon guidanceNetrin 1DCC, UNC5Netrin 1 is a target of NFkB and induces YAP.[171,172]
EphrinsEphsForward signaling: Src family kinases, Rho GTPases, Ephexins (GEFs that can activate Rho GTPases), ERK/MAPK pathway, which promote cell proliferation. Additionally, FAK and the JAK/STAT pathway interact with EPHs, which modifies cell adhesion.
Reverse signaling: interactions between ephrins and Src, Erk, Rac, T, paxillin, p75, and integrin-dependent cell adhesion as well as with Src, Grb4, PTP-BL, and PDZ-RGS3, which controls a variety of actions, including cell adhesion, migration, and proliferation.
[175]
SemaphorinsNeuropilins, plexinsSema3A: nrp1 PTEN/FOXO 3a-dependent MelCAM (CD146) expression.
Sema3B binds to NRP1 and inhibits PI3K/Akt signaling.
Full-length Sema3C NRP2 inhibits VEGF-C-dependent ERK1/2 and Akt signaling and suppresses lymphangiogenesis and metastasis.
Cleaved Sema3C promotes cancer cell survival.
Sema4D binds to plexin-B1 and activates ErbB2, phosphorylating plexin-B1. By activating RhoA GTPase, phosphorylated plexin-B1 promotes migration. Through ErbB2-dependent MAPK signaling, cleaved p61-Sema3E binds to plexin-D1 to promote metastasis. Sema3E binds to plexin-D1 and disrupts the interactions between plexin-D1 and NR4A, which is known to induce caspase-9-mediated apoptosis.
[188]
Synaptic proteinsNeuroligin 1-Neuroligin 1 recruits APC to the plasma membrane, blocking beta catenin degradation and causing its transfer to the nucleus, where it promotes EMT-linked gene transcription.[196]
Neuroligin 1-Neuroligin 1 synergizes with GDNF to induce cancer cell invasion of nerves and to activate the cytoskeleton-regulating protein cofilin.[199]
Neuroligin 3-NLGN3 extracellular domain shedding by ADAM10 into the tumor microenvironment promotes glioma growth.[198]
For all other synaptic proteins, see [6]. [6]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Arese, M.; Bussolino, F.; Pergolizzi, M.; Bizzozero, L. An Overview of the Molecular Cues and Their Intracellular Signaling Shared by Cancer and the Nervous System: From Neurotransmitters to Synaptic Proteins, Anatomy of an All-Inclusive Cooperation. Int. J. Mol. Sci. 2022, 23, 14695. https://doi.org/10.3390/ijms232314695

AMA Style

Arese M, Bussolino F, Pergolizzi M, Bizzozero L. An Overview of the Molecular Cues and Their Intracellular Signaling Shared by Cancer and the Nervous System: From Neurotransmitters to Synaptic Proteins, Anatomy of an All-Inclusive Cooperation. International Journal of Molecular Sciences. 2022; 23(23):14695. https://doi.org/10.3390/ijms232314695

Chicago/Turabian Style

Arese, Marco, Federico Bussolino, Margherita Pergolizzi, and Laura Bizzozero. 2022. "An Overview of the Molecular Cues and Their Intracellular Signaling Shared by Cancer and the Nervous System: From Neurotransmitters to Synaptic Proteins, Anatomy of an All-Inclusive Cooperation" International Journal of Molecular Sciences 23, no. 23: 14695. https://doi.org/10.3390/ijms232314695

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