ArticlePDF AvailableLiterature Review

Abstract and Figures

Neuropeptides are small regulatory molecules found throughout the body, most notably in the nervous, cardiovascular, and gastrointestinal systems. They serve as neurotransmitters or hormones in the regulation of diverse physiological processes. Cancer cells escape normal growth control mechanisms by altering their expression of growth factors, receptors, or intracellular signals, and neuropeptides have recently been recognized as mitogens in cancer growth and development. Many neuropeptides and their receptors exist in multiple subtypes, coupling with different downstream signaling pathways and playing distinct roles in cancer progression. The consideration of neuropeptide/receptor systems as anticancer targets is already leading to new biological and diagnostic knowledge that has the potential to enhance the understanding and treatment of cancer. In this review, recent discoveries regarding neuropeptides in a wide range of cancers, emphasizing their mechanisms of action, signaling cascades, regulation, and therapeutic potential, are discussed. Current technologies used to manipulate and analyze neuropeptides/receptors are described. Applications of neuropeptide analogs and their receptor inhibitors in translational studies and radio‐oncology are rapidly increasing, and the possibility for their integration into therapeutic trials and clinical treatment appears promising.
This content is subject to copyright. Terms and conditions apply.
The role neuropeptides play in cancer
initiation, progression, metastasis, and
therapeutic resistance is poorly under-
stood. Studies have indicated paradoxical
roles for diverse neuropeptides in cancer
processes, raising the question of whether
they serve primarily pro- or anticancer
functions. How neuropeptide expression
and their functions change during cancer
progression, and how their ectopic expres-
sion may influence caner-associated
behavioral changes are discussed.
Y. Wu, A. Berisha,
J. C. Borniger* ........................... 2200111
Neuropeptides in Cancer: Friend and
Foe?
Review
2200111 (1 of 20) © 2022 Wiley-VCH GmbH
www.advanced-bio.com
Review
Neuropeptides in Cancer: Friend and Foe?
Yue Wu, Adrian Berisha, and Jeremy C. Borniger*
DOI: 10.1002/adbi.202200111
highly dynamic network of various cell
types (e.g., cancer cells, endothelial cells,
immune cells, fibroblasts) that exert dier-
ential eects on neuronal activity within
the local environment. Thus, complex
signaling between cancer and stromal cells
in the TME results in altered firing rates
of local nerves. Reciprocally, increased
neuronal activity and subsequent release
of classical neurotransmitters and/or
neuropeptides within the TME results in
enhanced cancer progression and sub-
sequent metastasis in various preclinical
models and clinical studies.[–]
Nerves interact with multiple stromal
cells in the TME where they indirectly
promote tumor growth, progression, and
subsequent metastasis.[] Several studies
have demonstrated that sympathetic nerve
innervation and subsequent release of the
neurotransmitter norepinephrine (NE) is
increased in breast tumors.[,,] However,
recent work has demonstrated that there
is an inverse relationship between tumor
weight and norepinephrine content (i.e.,
larger the size of the tumor, the lower
the levels of NE), despite increased innervation of sympathetic
nerves within the TME.[] These findings may be attributed
to a relatively unexplored phenomenon: neuropeptide release
within the TME. Neuropeptides are molecular messengers that
regulate a variety of functions in the central and peripheral
nervous systems via binding G-protein-coupled receptors
(GPCRs) on target cells. One of the many functions of neuro-
peptides is serving as growth factors for normal cells via the
activation of the heterotrimeric G protein Gq and subsequent
synthesis of second messengers and engagement of tyrosine
phosphorylation cascades. Neuropeptides act as neuromodu-
lators, in that they can alter the response of neurons to neu-
rotransmitters and other circulating signals, such as leptin,
ghrelin, glucose, and insulin—all of which are aected during
cancer progression.[–] For example, both neuropeptide Y
(NPY) and hypocretin/orexin neurons appear to mediate some
of the orexigenic eects of centrally administered ghrelin as
administration of NPY receptor antagonists attenuated ghrelin-
induced feeding. Similarly, ghrelin-induced feeding was sup-
pressed in orexin knockout mice, indicating that ghrelin may
stimulate feeding through both the orexin and NPY systems.[]
However, neuropeptide signaling is exploited within cancer
cells and many studies demonstrate the contribution of neuro-
peptides in tumor cell proliferation and migration,[] with many
major neuropeptides also potentially contributing to cancer
processes (see Table 1). Additionally, neuropeptides exert direct
Neuropeptides are small regulatory molecules found throughout the body,
most notably in the nervous, cardiovascular, and gastrointestinal systems.
They serve as neurotransmitters or hormones in the regulation of diverse
physiological processes. Cancer cells escape normal growth control mecha-
nisms by altering their expression of growth factors, receptors, or intracel-
lular signals, and neuropeptides have recently been recognized as mitogens
in cancer growth and development. Many neuropeptides and their receptors
exist in multiple subtypes, coupling with dierent downstream signaling
pathways and playing distinct roles in cancer progression. The consideration
of neuropeptide/receptor systems as anticancer targets is already leading to
new biological and diagnostic knowledge that has the potential to enhance
the understanding and treatment of cancer. In this review, recent discoveries
regarding neuropeptides in a wide range of cancers, emphasizing their
mechanisms of action, signaling cascades, regulation, and therapeutic
potential are discussed. Current technologies used to manipulate and analyze
neuropeptides/receptors are described. Applications of neuropeptide analogs
and their receptor inhibitors in translational studies and radio-oncology are
rapidly increasing, and the possibility for their integration into therapeutic
trials and clinical treatment appears promising.
Y. Wu, A. Berisha, J. C. Borniger
Cold Spring Harbor Laboratory
One Bungtown Rd, Cold Spring Harbor, NY 11724, USA
E-mail: bornige@cshl.edu
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/adbi.202200111.
1. Introduction
Cancer cells often co-opt neural pathways in the early stages
of tumorigenesis resulting in enhanced pro-tumor signaling
(e.g., survival pathways, resistance to apoptosis). Further, the
invasion of cancer cells in or around pre-existing and/or newly
emerging nerves provides an additional route by which cancer
cells can metastasize even in the absence of lymphatic and vas-
cular invasion.[,] Subsequently, the invasion of cancer cells
along nerves and/or within dierent nerve layers (i.e., endoneu-
rium, perineurium, epineurium), termed perineural invasion
(PNI), often precipitates cancer associated-pain observed in
various types of malignancies.[–] These changes within the
tumor microenvironment (TME) drive hypersensitivity of sen-
sory neurons that reside in dorsal root ganglia (DRGs) which
are primarily responsible for relaying nociceptive sensation to
upstream brain regions. In addition, the TME encompasses a
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (2 of 20)
www.advanced-bio.com
eects on cells that comprise the TME including endothelial
cells and immune cells.[–] Neuropeptide action on immune
cells results in immune cell-derived cytokine release with far-
reaching systemic neuroendocrine and metabolic changes via
interactions with the central nervous system and peripheral
organs. An example of this is demonstrated when discussing
the immunomodulatory eects of the endogenous neuropep-
tide vasoactive intestinal peptide (VIP). For example, VIP and
other neuropeptides can reduce the production of inflamma-
tory cytokines and chemokines while stimulating the produc-
tion of anti-inflammatory cytokines (e.g., IL-, IL-Ra).[]
VIP also down regulates the expression of other inflammatory
mediators such as inducible nitric oxide synthase and cycloox-
ygenase  (COX) and subsequent release of nitric oxide and
prostaglandin E by macrophages, dendritic cells (DCs), and
microglia.[,] VIP also functions as a macrophage-deactivating
factor as VIP has been shown to inhibit in vitro and in vivo
production of the proinflammatory cytokines tumor necrosis
factor alpha (TNF alpha), interleukin- (IL-), interleukin-
(IL-), and of nitric oxide (NO) and stimulates the production
of the anti-inflammatory cytokine interleukin- (IL-).[,]
Additionally, several neuropeptides (e.g., substance P and
neuropeptide Y) are implicated in inflammation, develop-
ment, and tissue repair-common pathways exploited during
the initial stages of tumorigenesis.[] Interestingly, several
neuropeptides are ectopically expressed within various
clinically aggressive solid tumors including breast, pancreatic,
lung, and gastrointestinal cancers.[,–] Several nonpituitary
tumors (e.g., small cell lung carcinomas, carcinoid tumors,
medullary thyroid carcinomas, breast, kidney, colon, prostate)
also ectopically express and secrete the neuropeptide adreno-
corticotropic hormone (ACTH) resulting in the onset and
subsequent diagnosis of ectopic ACTH syndrome observed in
–% of cases of Cushing’s syndrome (CS).[,] The eleva-
tion in ACTH is attributed to cancer-induced amplification of
its precursor peptide [i.e., proopiomelanocortin (POMC)] that
is normally present within the cells from which the cancer
originated.[] This observation demonstrates the ability of
cancer cells to ectopically express neuropeptides and increase
the concentration of neuropeptides that are normally present
within a narrow window in the cells that undergo transfor-
mation. Therefore, neuropeptides in the TME may originate
from nerves themselves or from several dierent forms of
tumors that ectopically express neuropeptides. Interestingly,
Table 1. Major neuropeptides with potential roles in cancer processes.
Neuropeptide and
receptors
Abundant regions Coreleased
neurotransmitters
Physiological functions Ref.
Neuropeptide Y
(NPY, 36 aa); Y1, Y2, Y4,
Y5, y6 receptors
CNS: hypothalamus, amygdala, hippocampus,
nucleus of the solitary tract, locus coeruleus,
nucleus accumbens, cerebral cortex
PNS: adrenal medulla, liver, heart, spleen,
vascular endothelial cells
Norepinephrine,
epinephrine,
GABA, serotonin
CNS: food intake, energy homeostasis,
circadian rhythm, cognition, stress response
PNS: vasoconstrictor
[39–42]
Substance P (SP, 11 aa);
NK1 receptor
CNS: dorsal horn, hypothalamus, Amygdala
PNS: vascular endothelial cells, muscle cells,
immune cells, gastrointestinal and genitourinary
tracts, lung, thyroid gland, tumor cells
Serotonin,
dopamine,
norepinephrine,
acetylcholine,
GABA, glutamate,
epinephrine
CNS: pain, stress, anxiety, neurogenic inflammation,
neuronal survival, and degeneration
PNS: vomiting reflex, defensive behavior,
cardiovascular system, salivary secretion, smooth
muscle contraction, vasodilation, chronic
inflammation
[42–45]
Neurotensin (NTS, 13
aa); NTS1-3 receptors
CNS: spinal cord, trigeminal nucleus, central
amygdaloid nucleus, anterior pituitary, median
eminence, preoptic, and basal hypothalamic areas
PNS: gastrointestinal tract, heart, adrenals,
pancreas, respiratory tract
Dopamine, glycine,
acetylcholine,
norepinephrine
CNS: sensory and motor functions, temperature
regulation, neuroendocrine control of the pituitary,
blood flow, and blood pressure
PNS: modulation of vascular smooth muscle
activity, gastrointestinal motility, and
pancreaticobiliary secretions
[46–49]
Orexin/Hypocretin
(OX, 33 aa/ 28 aa);
OX1/2 receptors
CNS: perifornical, lateral, and posterior hypothalamus
PNS: gastrointestinal tract, kidney, adrenal gland,
pancreas, placenta, stomach, ileum, colon,
colorectal epithelial cells
Glutamate,
dynorphin
CNS: maintain the wakefulness, food consumption,
energy homeostasis, reward seeking, stress,
motivation, drug addictions
PNS: blood pressure, heart rate, gut mobility,
hormone secretion, lipolysis, reproduction
[50–52]
Somatostatin (SST, 14
aa/ 28 aa); SSTR1-5
receptors
CNS: cortex, hypothalamus, brainstem, spinal cord
PNS: gastrointestinal tract, lung, pancreas,
nerves of the heart, thyroid, skin, eye, and thymus
Acetylcholine,
norepinephrine
CNS: control of pituitary hormone secretion,
regulation of locomotor activity and cognitive
functions
PNS: inhibition of intestinal
motility and absorption, vascular contractility,
cell proliferation
[53–55]
Vasoactive intestinal
peptide (VIP, 28 aa);
VPAC1/2 receptors
CNS: suprachiasmatic nuclei of the hypothalamus,
hippocampus, cortex
PNS: gastrointestinal tract, respiratory tract,
pancreas, urinary tract, immune, and endocrine cells
Acetylcholine,
serotonin
CNS: neuroprotective eects
PNS: regulation of intestinal motility
[56–58]
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (3 of 20)
www.advanced-bio.com
the first evidence for the production of VIP by cells of the
immune system was identified in rat peritoneal, intestinal, and
lung mast cells using radioimmunoassay and immunohisto-
chemical studies.[] Additionally, immune cells also produce
neuropeptides (e.g., VIP) in response to antigenic and inflam-
matory signals and can act in an autocrine/paracrine manner
via specific receptors expressed on immune cells.[,] Thus, it
is likely that neuropeptides play a key role in this context (i.e.,
low or high neurotransmitter release) where they can alter
the response of neurons to cancer derived signals. Addition-
ally, it is possible that the increase in tumor volume, despite
a reduction in neurotransmitter release, is achieved because
of the actions of local neuropeptides. A reduction in neuro-
transmitter release and corresponding increases in neuropep-
tide release may be employed by local nerves as a strategy to
optimize eciency in the TME. Important questions remain,
for instance, what are the relative ratios of neurotransmitters
and neuropeptides within the TME? Do neurotransmitters and
neuropeptides work synergistically or antagonistically within
tumors? What are the major sources of neuropeptides in the
TME? Below, we discuss current evidence for the role neuro-
peptides play in local crosstalk within the TME, as well as distal
eects this signaling may have on systemic physiological and
behavioral processes.
2. Characteristic Features of Neuropeptides
Changes in the properties of local nerves within the TME may
have long-lasting implications on subsequent release of neu-
ropeptides. Secretory organelles that undergo exocytosis are
often categorized into two general categories based on size:
the large dense-cored vesicles (LDVs) > nm in diameter
and the smaller vesicles in the range of –nm, classically
referred to as “synaptic vesicles.”[] Classic small molecule
neurotransmitters (e.g., acetylcholine, GABA, glutamate) exert
rapid and short-lasting eects on target tissues, whereas neu-
ropeptides exert a response of longer duration, in a similar
fashion to circulating peptide hormones (e.g., altered gene
expression on the order of minutes to hours). Therefore, both
neuropeptides and peptide hormones can act as autocrine,
paracrine, and endocrine signals in the body.[] However, the
key distinguishing feature between neuropeptides and peptide
hormones—at “normal” physiological conditions—is that neu-
ropeptides are synthesized and secreted by neurons, whereas
peptide hormones are synthesized and secreted from endo-
crine cells. In addition, neuropeptides act on neural substrates,
whereas peptide hormones act on a variety of target organs
and glands. Although neuropeptides and neurotransmitters
are often colocalized in synaptic vesicles, this is not always the
case. Smaller synaptic vesicles contain classical neurotransmit-
ters exclusively, whereas the LDVs contain both neuropeptides
and neurotransmitters.
2.1. Frequency-Dependent Release and Calcium Dynamics
A discussion on how the firing rates (i.e., frequency of
action potentials) of dierent neurons impacts the release
of neuropeptides and neurotransmitters is highly relevant to
our discussion on cancer as the LDVs containing both neuro-
peptides and neurotransmitters require higher firing rates or
“bursting activity” to become “active” (i.e., release neuropep-
tides).[–] In addition, neuropeptides released in the local
environment in response to increased neuronal discharge have
long-lasting eects on nearby neurons, such as increased excit-
ability,[] which may further drive cancer progression. Thus,
the principle of frequency-dependent release is highly impor-
tant in the context of cancer cells because the same neuro-
peptides that are released in response to tumorigenesis and
subsequent PNI may not be released—or released at very low
levels—under steady-state conditions.[] This phenomenon
becomes more evident when looking at two classical examples
of physiological and pathological conditions that can induce the
release of neuropeptides that are otherwise not released. The
first example assessed the changes in firing rates of a subset
of hypothalamic neurons (i.e., oxytocinergic neurons) at basal
conditions and then during prolonged sucking of maternal
nipples from pups, promoting the subsequent secretion of
milk.[] Oxytocinergic neurons were observed to fire at Hz
(.–. Hz) at baseline with prolonged stimulation by ten
or more pups increasing the firing rates of these neurons to
–Hz followed by strong milk ejection from the mammary
gland within – s.[] Therefore, the release of the neuropep-
tide oxytocin can be modulated by external stimulation. Similarly,
the release of endogenous opioid peptides in response to severe
pain exerts analgesic eects in preclinical models.[] Interest-
ingly, intracranial, or intraspinal electrical stimulation is used
to provide relief for patients suering from chronic pain,[,]
raising the possibility that modulation of neuropeptide release
may be a useful therapeutic intervention. Further expanding
on this concept, increased electrical stimulation of isolated rat
neurohypophyses (i.e., posterior pituitary gland) resulted in
increased neuropeptide release. More specifically, Racke et al.[]
demonstrated that electrical stimulation of the rat neurohypo-
physes with a frequency of  Hz caused the release of about
 µU vasopressin and  µU oxytocin. On the other hand, elec-
trical stimulation of the rat neurohypophyses with a frequency
of Hz caused the release of  µU vasopressin and oxytocin,
highlighting the frequency-dependent release of certain neu-
ropeptides. Interestingly, a single burst stimulation with a fre-
quency of Hz resulted in increased vasopressin release from
the rat posterior pituitary gland compared to constant stimula-
tion with a frequency of  Hz–indicating that phasic burst
stimulation may be more eective than tonic constant-frequency
stimulation in triggering neuropeptide release.[,]
Given the fundamental role of calcium (Ca+) ions in exocy-
tosis, several studies have provided evidence that the release of
neuropeptides and classical neurotransmitters exhibit dierent
Ca+ dynamics. Neuropeptide release is triggered by small ele-
vations in Ca+ concentrations in the bulk cytoplasm (i.e., con-
tinued firing of neurons) as opposed to the higher elevations of
Ca+ concentrations (i.e., individual action potentials) required
for the secretion of amino acids and classical neurotransmitters
near the active sites.[] These changes in the amount of Ca+
required for exocytosis may be due to the location of LDVs and
synaptic vesicles relative to the site of Ca+ release. Paradoxi-
cally, classical neurotransmitter release is thought to occur very
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (4 of 20)
www.advanced-bio.com
close to the site of Ca+ entry despite the higher levels of Ca+
required for exocytosis, whereas neuropeptide release occurs at
a distal location from the site of Ca+ entry despite the lower
levels of Ca+ needed for neuropeptide release (see Figure 1).[]
In accordance with these findings, Huang and Neher[] dem-
onstrated that Ca+ dependent exocytosis of LDVs can occur
at Ca+ concentrations that are at least ten times lower than
those required for neurotransmitter and amino acid release
from nerve terminals, indicating that neurotransmitters and
neuropeptides are not necessarily released in response to the
same stimuli despite their colocalization. For example, several
studies indicate that the ratio of neurotransmitters to neuro-
peptides may further dier in response to certain stimuli. A
study by Franck et al.[] demonstrated that the release of the
neuropeptide substance P (SP) from rat spinal cord slices was
increased by electrical stimulation of frequencies in the range
of – Hz. Additionally, the release of the classical small-
molecule neurotransmitter -hydroxytyptamine (-HT; sero-
tonin) per pulse remained constant, indicating that coexisting
neurotransmitters and neuropeptides may be released in dif-
ferent proportions depending on the stimulation frequency.
Therefore, it is likely that repeated stimulation of neurons
at higher frequencies is required for neuropeptide release,
whereas individual action potentials are not sucient. Thus, it
is essential to discern the dierential release of neuropeptides
and neurotransmitters during the early stages of cancer initia-
tion and progression as this will provide additional insight into
the molecular and cellular changes sustaining cancer progres-
sion and may provide further therapeutic targets for various
malignancies.
This phenomenon of cancer cells altering the recruitment
and organization of nerves within the TME has been widely
investigated. Cancer cells often co-opt neuronal programs
through the release of growth factors [e.g., nerve growth factor
(NGF), brain-derived neurotrophic factor (BDNF)] that pro-
mote innervation of the TME.[–] Similar to neuropeptides,
growth factors (e.g., neurotrophins) are proteinaceous signaling
molecules that are produced and secreted by cells, including
neurons. Although growth factors are synthesized, expressed,
and released by neurons they are not generally considered
neuropeptides.[] However, several growth factors may fulfill
the criteria of neuropeptides (e.g., BDNF).[] Studies inves-
tigating the direct impact of cancer cells on the electrophysi-
ological properties of local nerves is, to a large extent, lacking.
This is largely due to an absence of electrophysiological tech-
niques that can detect changes in electrophysiological proper-
ties at single-cell resolution within co-culture in vitro assays,
and especially in vivo. In addition, a lack of understanding
of the cancer cell-neuron connections in the periphery (i.e.,
chemical or electrical in nature) contributes to this absence of
experimental evidence. Despite these limitations, recent work
by Kovacs and Vermeer[] demonstrated that high-grade serous
ovarian carcinoma (HGSOC)—densely innervated by TRPV+
sensory nerves—are more electrically conductive and exhibit
a higher mean spike amplitude than benign/normal ovarian
tissue. These findings are consistent with the hypothesis that
nerves within the TME increase their electrical activity and may
reflect increased numbers of sensory nerves within the TME
of HGSOC. As we alluded to previously, neuropeptide release
is frequency dependent; nerves with higher firing rates release
an abundance of neuropeptides when compared to nerves that
exhibit lower firing rates. Thus, combining the principle of
frequency-dependent release of neuropeptides with the experi-
mental observation that malignant tissues are more electrically
conductive than benign/normal ovarian tissue, it is likely that
neuropeptides are aberrantly released within the TME during
cancer progression as the local nerves may not exhibit the firing
rates required to elicit the release of neuropeptides during
normal, steady-state conditions. However, the study by Kovacs
and Vermeer[] did not measure the change in the firing rates
of neurons when cocultured with cancer cells. They specifi-
cally measured the changes in mean spike amplitude which can
dier independent of changes to the firing rates of neurons
(i.e., higher spike amplitude may not correspond to increased
firing rates of neurons). Additionally, the electrical activity of
malignant, benign, and normal tissues was measured using a
multielectrode array consisting of a  ×  electrode grid (i.e.,
 electrodes) which may not provide high enough spatial reso-
lution to measure the changes in the firing rates of individual
neurons. Building on these findings, future studies will need
to assess the changes in the firing rates of neurons that are
cocultured with cancer cells, and expand these approaches in
vivo. Although logical reasoning and emerging evidence sug-
gests that cancer cells increase the activity of local nerves within
the TME, more studies need to be conducted across various
malignancies to fully establish this phenomenon as a potential
mechanism by which nerves and cancer cells communicate.
Further studies will need to investigate the specific neuropep-
tides that are dierentially expressed and secreted from nerves
Figure 1. Eects of neuropeptides in the tumor microenvironment (TME).
(1) Cancer cells release axon guidance cues, growth factors, and proin-
flammatory cytokines into the TME. The actions of the axon guidance
cues and growth factors results in the (2) recruitment of local nerves and/
or emergence of new nerves into the TME. Once nerves integrate into the
TME, the interaction between cancer cells and local nerves results in an
increase in the firing rates of these nerves with subsequent (4) small ele-
vations in cytoplasmic calcium concentrations required for the (5) release
of neuropeptides from local nerves. Newly released neuropeptides, in
turn, signal back onto cancer cells, driving (6) tumor progression and
initiating the early stages of (7) perineural invasion and (8) subsequent
metastasis along nerve fibers.
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (5 of 20)
www.advanced-bio.com
in malignant and normal/benign tissues. Moreover, it is likely
that the mode or speed of transportation of these neuropeptides
along axons is impacted during cancer progression which may
contribute to ectopic release of these neuropeptides within the
TME.
2.2. Recruitment, Docking, and Fusion of Neuropeptides
In addition to the physical dierences and frequency-
dependent release of these two classes of vesicles, they also
dier in their recruitment, docking, and fusion to the plasma
membrane.[] For example, synaptic vesicles are often found in
clusters restricted to small sections of the presynaptic plasma
membrane termed the “active zone,” which lies at the inter-
face between the presynaptic terminal and the synaptic cleft.
These active zones contain evolutionarily conserved protein
complexes that primarily function to dock synaptic vesicles for
exocytosis and recruit voltage-gated Ca+ channels to the pre-
synaptic membrane.[] However, LDVs are not located at these
active zones but are instead scattered diusely around the nerve
terminal (i.e., axons) and are found in the soma and dendritic
compartments as well.[–] Huang and Neher demonstrated
that LDVs are also located at ectopic sites outside of the active
zones, where neuropeptides are released extrasynaptically at
the somata of dorsal root ganglion (DRG) neurons.[] The dif-
ferences in the location of release between neuropeptides (i.e.,
soma, dendrites) and neurotransmitters (i.e., nerve endings,
synapses) may be partly due to the fact that neuropeptides are
exclusively produced in the cell soma without local synthesis
in nerve endings— in contrast to the classical neurotransmit-
ters which are synthesized within the soma and locally in nerve
endings. As a result of this extra synaptic release of neuropep-
tides, neuropeptides diuse over short distances within the
local environment and longer distances where they can act
as hormones in dierent brain regions—in sharp contrast to
the spatially restricted actions of neurotransmitters.[,] The
ability of neuropeptides to persist in extracellular fluid and dif-
fuse over long distances is largely due to absence of any known
reuptake mechanisms for neuropeptides. On the other hand,
neurotransmitters often have various reuptake mechanisms
which explains why their eects are primarily short-term and
spatially limited. Interestingly, local release of neuropeptides
within the TME sustain the proliferation of solid tumors via
autocrine/paracrine loops.[,,,] In addition, it is possible
that the long-distance eects of these TME-derived neuropep-
tides on distinct brain regions contribute to many of the cancer-
related neurological disorders often observed in patients (e.g.,
cognitive impairment, sleep disruption, depression, and pain).
2.3. Plasticity in Neuropeptide Expression
Neuropeptides are capable of modulating neurotransmitter
release, and reciprocally, neurotransmitters can also aect
the synthesis and subsequent release of neuropeptides. For
example, somatostatin (also known as somatotropin-release
inhibiting factor (SRIF)—a neuropeptide whose expression
is widely distributed in the central and peripheral nervous
systems—increases the release of certain neurotransmitters
such as - hydroxytryptamine (-HT; serotonin) and dopamine
but can also decrease the release of gamma-aminobutyric acid
(GABA) and norepinephrine.[,] Reciprocally, the release of
SRIF is aected by -HT and GABA, suggesting that neuro-
transmitter release also aects the release of neuropeptides.[]
Neuropeptides exhibit additional plasticity in their expression.
Some neuropeptides are abundantly expressed under steady-
state circumstances which indicates that these subsets of neu-
ropeptides are functionally available at any time. The second
type are neuropeptides that are normally expressed at low
levels and are only elevated in response to stimuli (e.g., nerve
injury, inflammation, tumorigenesis), such as adrenocortico-
tropic hormone (ACTH). Neuropeptides that are expressed
early during development are often downregulated postna-
tally.[] This phenomenon constitutes the third type of neuro-
peptide plasticity and is relevant as cancer cells co-opt many of
the neural pathways that are active during development (e.g.,
nerve regeneration, tissue repair), indicating that some of the
same neuropeptides that are expressed during development are
potentially expressed during the early stages of tumorigenesis.
One example highlighting this phenomenon is demonstrated
by Cohen et al.[] when investigating neuropeptide Y (NPY)
expression in the adrenal gland. NPY is expressed in a subset of
cells within the adrenal medulla, the inner part of the adrenal
gland. Cohen et al.[] demonstrated that there is a biphasic
pattern of expression of NPY mRNA during development of
the human adrenal medulla. More specifically, NPY mRNA
is detectable between . weeks of gestational age through
 weeks of gestational age and is then not detectable until
 months after birth. However, when analyzing NPY mRNA
expression in neuroblastoma tumors—which often arise in
the adrenal medulla—/ (%) neuroblastoma tumors
expressed NPY mRNA.[] Thus, this study demonstrates that
certain neuropeptides present during embryonic development
(e.g., NPY) also play a role in the development of tumors from
the same cell of origin (e.g., neuroblastomas). Therefore, tumor
cells may retain the molecular mechanisms that mediate dier-
entiation of the embryonal cells from which they arise.
A major stimulus that promotes neuropeptide release are
cytokines (e.g., induced by nerve injury). Although controlled,
regulated release of these cytokines is beneficial and impor-
tant for remodeling after injury, chronic release is adverse. For
example, proinflammatory cytokines [e.g., interleukin (IL)-,
-, - and TNF-a] are expressed during the first phase of nerve
injury[] and regulate the inflammatory response through the
recruitment of macrophages.[] However, chronic release of
TNF-a has been demonstrated to reproduce the nociceptive
behaviors and endoneurial pathology found following experi-
mental nerve injuries, implicating TNF-a in the pathologies
of neuropathic pain.[–] Cytokines are widely known for
their eects on various solid tumors where they stimulate
neurotransmitter and neuropeptide synthesis and subsequent
release in the TME. A study by Freidin etal.[] demonstrates
the role of cytokines in regulating neuropeptide expression in
sympathetic neurons—which are commonly associated with
aggressive cancers (e.g., breast, prostate, pancreatic), increased
mortality, and disease recurrence.[] Specifically, exposure of cul-
tured explants of the rat superior cervical ganglion—harboring
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (6 of 20)
www.advanced-bio.com
sympathetic nerve cell bodies—to the proinflammatory cytokine
interleukin-β (IL-β) increased the release of the neuropeptide
SP by sympathetic neurons. IL-β and other proinflammatory
cytokines are elevated within tumors where they predominantly
exert pro-tumor eects.[] IL-β is produced and secreted by
various cells that comprise the TME including immune cells,
fibroblasts, and cancer cells.[] Thus, it is likely that a portion
of the SP synthesized and released within tumors (e.g., breast)
is induced by IL-β, which itself is derived from the stromal
cells that encompass the TME and subsequent binding to its
receptors on sympathetic nerves. Increased expression of SP
within the TME is important because SP expression has been
shown to promote cancer initiation, progression, and mediate
perineural invasion in pancreatic cancer.[–] Thus, this
interplay between SP and IL-β and plasticity in neuropeptide
expression in the TME is highly relevant to our discussion on
cancer and should be investigated further to provide insight
into the role of neuropeptides in the TME.
3. Neuropeptides and Their Receptors in Cancer
3.1. Neuropeptide Y
NPY is one of the most prevalent peptides in the mammalian
central and peripheral nervous systems. It is composed of 
amino acids, exerting context-dependent neurotransmitter and
neuromodulator functions throughout the human body. In
addition to NPY, peptide YY (PYY), and pancreatic polypeptide
(PP) are also native ligands of the NPY receptor family, medi-
ating their eects via corresponding receptors with dierent
potencies. NPY receptors are monomer proteins that belong to
the class A G-protein coupled receptors (GPCRs) coupled with
Gi/o. So far, seven Y receptors have been discovered in verte-
brates,[,] of which five (Y, Y, Y, Y, and Y) are present
in mammals and four (Y, Y, Y, and Y) are functional in
humans.[] The Y receptor, on the other hand, is active in rab-
bits and mice.[] NPY and its receptors (Y, Y, Y) contribute
to a wide range of physiological activities as a multiligand/
receptor system that is broadly distributed throughout the body.
In the CNS, they are located in the spinal cord and many brain
regions responsible for food intake, energy homeostasis, circa-
dian rhythms, cognition, and stress responses, among other
functions.[] In the periphery, NPY is mostly expressed in the
sympathetic nervous system, where it regulates cardiovascular
and other processes in conjunction with norepinephrine and
adenosine triphosphate (ATP).
NPY receptors have recently received substantial interest due
to their over-expression in a variety of human malignancies,
including breast carcinomas,[] neuroblastomas,[] Ewing sar-
comas,[] renal cell carcinomas, ovarian cancers, nephroblas-
toma, and gastrointestinal stromal tumors,[] among others.
Reubi et al. used autoradiography to identify NPY receptor
expression in human breast carcinomas, providing initial evi-
dence that NPY may play a role in human cancer.[] In this
study, they discovered that the Y receptor was significantly
overexpressed in % primary breast carcinomas and % of
lymph node metastases, whereas normal breast tissues pref-
erentially expressed the Y receptor subtype. Other studies
have also shown that the NPY receptor subtypes exhibit a
dierentiation-specific expression pattern,[] with expression
shifting from Y to Y in breast tissue undergoing neoplastic
transformation.[] Glioblastomas, a deadly form of brain
cancer, not only produce Y at a high frequency but also have
extraordinarily high Y densities. Moreover, all major glioma
types tested were infiltrated with intratumoral nerve fibers
containing NPY.[] Therefore, Y receptors in tumor cells are
activated by endogenous NPY, triggering functional alterations
to tumor growth.[] Aside from Y and Y, the Y receptor
subtype seems to be critical for cell motility via RhoA activa-
tion.[] Strongly inducible Y expression was found in migra-
tory cells of neuroblastoma[] and liver cancer.[] Together,
it appears that NPY receptors follow four general principles:
) NPY receptors are mainly expressed in specific endocrine
tumors, epithelial malignancies, and embryonal tumors[];
) tumor cells express predominantly Y and/or Y subtypes[];
) Y receptors are mainly involved in the modulation of cancer
cell proliferation, whereas Y receptor activation appears to pro-
mote angiogenesis[]; and ) Y receptors regulate cell prolifer-
ation,[] chemotactic migration, invasion,[,] and contribute
to chemoresistance.[] These properties make NPY receptors
particularly appealing as potential targets for cancer imaging
and treatment.
In most cases, NPY and its receptors serve a pro-oncogenic
role, particularly in neuroendocrine tumors, breast and pros-
tate cancers. However, this eect is not universal, as NPY can
exert tumor suppressive eects in cholangiocarcinoma[] and
hepatocellular carcinoma,[] and it can dramatically inhibit
estrogen-induced cell proliferation via the Y receptor in the
human breast cancer cell line MCF-.[] Generally, NPY pro-
motes or inhibits tumor growth through paracrine,[] auto-
crine, and endocrine[] mechanisms. NPY is primarily
generated in nerve fibers and neuroendocrine cells, then trans-
ported to intratumor nerve fibers or secreted by tumor cells
themselves to activate tumoral NPY receptors. Released NPY
can bind to NPY receptors on tumor cells and tumor blood ves-
sels, thus triggering paracrine and autocrine eects on tumor
cell metabolism and tumoral blood supply. Due to rapid deg-
radation of NPY in the blood stream, endocrine stimulation of
NPY receptors by circulating NPY appears less likely to have
a general function in tumors.[] Furthermore, NPY and par-
ticular Y receptors are expressed in a variety of cell types in the
tumor microenvironment (TME), including immune cells[]
and epithelial cells of tumor-associated blood vessels.[] Addi-
tionally, NPY is engaged in the activation of cancer-associated
fibroblasts (CAFs).[,] Upon paracrine and autocrine regu-
lation, NPY/Y receptors can promote cell proliferation,[,]
mitogenesis, extracellular matrix invasion, migration,[]
and angiogenesis[] in the TME by interacting with multiple
signaling pathways. Calcium/calmodulin-dependent kinase
II (CaMKII), protein kinase C (PKC), and mitogen-activated
protein kinase (MEK/)[,] are major pathways involved,
as illustrated in Figure 2. For example, when NPY binds to
its receptor, the small G-protein Ras is activated. In turn, Ras
recruits and promotes Raf, a serine/threonine protein kinase
that activates MEK, which phosphorylates the MAPK ERK/.
This ERK/MAPK signaling pathway plays a crucial role in
tumor survival and development.[]
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (7 of 20)
www.advanced-bio.com
3.2. Substance P
Substance P (SP) is one of the most important members of the
tachykinins (TKs), which includes an evolutionarily conserved
family of small bio-active peptides. TKs are broadly distrib-
uted in mammalian central and peripheral nervous systems,
participating in the regulation of various physiological pro-
cesses including pain, stress, inflammation, wound healing,
leukocyte tracking, microvasculature permeability, cell sur-
vival, and degeneration,[,] among other processes. Recent
studies demonstrate that TKs play a role in dierent steps of
carcinogenesis, such as mitogenesis, angiogenesis, cell migra-
tion, metastasis, and other growth-related events.[] Three G
protein-coupled receptors (GPCRs), termed NK-, NK- and
NK-, mediate the biological actions of three native ligands: SP,
neurokinin A and neurokinin B. SP is a more potent agonist at
neurokinin- receptors (NKRs) than the other two ligands.[]
Both SP and NKR are overexpressed in multiple cancer types
including breast, ovarian, pancreas, prostate, head and neck
cancers, gliomas, and astrocytomas.[,,–] Specifically,
SP is expressed in many cells of the TME, including epithelial
cells, immune cells, cancer stem cells, and fibroblasts.[] Based
on current evidence, SP expression and distribution in cancer
cells can be characterized as: ) having higher expression of
SP in malignant cells than in normal cells within the same
tissue[,]; ) SP is found both in the cytoplasm and is even
more strongly expressed in the nuclei,[] suggesting its poten-
tial role as a genetic neuromodulator or epigenetic factor (dis-
cussed in Section .)[,,]; ) SP and its interaction with
the NKR have a critical role in the tumor microenvironment.
In a large majority of tumors, NKR are found in the intra- and
peritumoral blood vessels,[] where the SP is released from
sensory nerve terminals. SP/NKR activation enhances the pro-
liferation of cultured endothelial cells in vitro and the growth
of capillary vessels in vivo, which in turn facilitates angiogen-
esis during tumor progression; ) SP is also located in dierent
body fluid compartments, such as blood, cerebrospinal fluid,
and breast milk, likely because SP can form a complex with
fibronectin to protect it from peptidase digestion and increase
its half-life.[]
More recently, information on the underlying mechanisms
of SP/NKR complex function in cancer cell proliferation,
migration, angiogenesis, and maintenance of a favorable TME
has become available. In general, SP promotes tumor growth
and development mainly through four mechanisms: ) via
an autocrine pathway, by which SP is released from primary
tumors[]; ) through the paracrine pathway, where SP is
secreted from tumor cells targeting surrounding stromal cells
and epithelial cells of intratumor blood vessels; ) by means of
the peripheral nervous system, in which direct interactions are
provoked between sensory nerve terminals and tumor cells[];
and ) via an endocrine mechanism, related to the organism’s
emotional state, by which SP reaches the whole body through
the bloodstream.[,] Upon SP binding to NKR, multiple
Figure 2. Primary signaling pathways activated by neuropeptides and their receptors involved in cancer progression. The interaction of neuropeptides
with their GPCRs regulates the Ras/MEK, PI3K/Akt, intracellular Ca2+ mobilization, and NF-kB pathways for cell survival and proliferation; cAMP/
PKA/CREB pathway for invasion and drug resistance; Src/ROCK and JAK/STAT pathways for angiogenesis and metastasis; Wnt/ β-catenin pathway
for cancer cell stemness and malignant progression. As a result, neuropeptides may stimulate or inhibit these signaling cascades, causing cancer cell
proliferation, migration, or apoptosis.
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (8 of 20)
www.advanced-bio.com
signaling pathways such as phosphoinositide -kinase (PIK)
and mitogen-activated protein kinases (MAPKs) will be acti-
vated, followed by a variety of downstream eectors involved
in protein synthesis, cell cycle progression, gene transcription,
and expression (see Figure).
Tumor progression mediated by SP/NKR occurs primarily
through four major signaling pathways: Wnt/β-catenin, MAPK,
PIK/Akt, and mTOR.[] ) The Wnt signaling cascade is a
crucial transduction pathway that regulates cell survival, pro-
liferation, migration, dierentiation, and organelle dynamics
through the activation of β-catenin.[] Tumorigenesis, TME
remodeling and epithelial-mesenchymal transition (EMT) are
enhanced by Wnt signaling pathway through fine crosstalk
between the tumor and infiltrating immune cells.[] Several
studies demonstrate that pharmacological inhibition of NKR
by antagonists results in suppression of canonical Wnt sign-
aling in colon cancer[,] and hepatoblastoma cell lines.[]
This, in turn, downregulates Wnt-associated protein activity,
including cyclin D, VEGF, c-Myc, and LEF-, thereby leading
to cell cycle arrest, apoptosis, and reduced angiogenesis and
invasion of cancer cells. ) The MAPK signaling pathway is
another key transduction cascade module regulating essential
cellular processes, such as proliferation, dierentiation, migra-
tion, development, inflammation, and apoptosis.[] As two
major components in the MAPK family, phosphorylated ERK
and p have been reported to participate in cancer progression
and invasion.[,] In human astrocytoma cells, SP induces
the accumulation of IL-, IL-, and expression of c-Myc or Src
via p and ERK/ signaling. NKR antagonists can attenuate
this process by reducing SP-induced proinflammatory cytokine
synthesis.[–] SP further activates inflammatory pathways
in macrophages and fibroblasts via ERK/p MAPK-mediated
NF-kB activation, further enhancing EMT and MMP release,
a process critical for cancer invasion and metastasis.[] Addi-
tionally, SP triggers MMP- and MMP- overexpression via
activation of ERK/, JNK, and Akt pathways, leading to cell
proliferation and ECM invasion in breast cancer cells.[,]
) The PIK/AKT signaling pathway plays a crucial role in
cell proliferation and apoptosis resistance by interacting with
the p tumor-suppressor pathway in malignant cells.[]
SP can additionally induce “M-like” macrophage polariza-
tion via PIK/Akt/mTOR signaling, subsequently stimulating
angiogenesis and cell proliferation.[] Building on these find-
ings, Javid et al. reported anticancer activity of the only FDA
approved NKR antagonist, aprepitant, in an esophageal squa-
mous cancer cell line. Aprepitant has eective potency in
inhibiting malignant cell proliferation, inducing apoptosis and
suppressing activation of the PIK/Akt axis and its downstream
eectors (e.g., NF-κB).[] ) Aberrant mTOR signaling caused
by genetic alterations from dierent signaling cascades is also
commonly observed in various types of cancer, regulating cell
proliferation, immune cell dierentiation, and tumor metabo-
lism.[] Rapamycin is a specific inhibitor of mTOR pathway,
inhibiting mTOR function by blocking the cell cycle at the
G/S phase. Several studies have shown that SP/NKR can
activate mTOR in cancer cells. Mayordomo et al. treated breast
cancer cells with a neutralizing anti-SP antibody and observed
cell cycle blockade and inhibition of mTOR signaling as well
as decreased Her and EGFR expression.[] Together, these
studies demonstrate a pleiotropic role for SP in cancer initia-
tion, progression, metastasis, and resistance to treatment.
3.3. Neurotensin (NTS)
NTS is a tridecapeptide located primarily in the brain and intes-
tine, where it functions as both a neuropeptide and an endo-
crine hormone, respectively. Centrally, it is secreted by neurons
and acts as a signaling molecule to neighboring cells, control-
ling body temperature, sensory and motor functions, pituitary
neuroendocrine functions, blood flow and pressure, sleep,
and other homeostatic functions. In the periphery, NTS is
released by endocrine cells in the gastrointestinal tract, where
it acts as a local hormone for the digestive system. Besides its
canonical role in neural regulation and gut motility, emerging
studies support a new role for NTS in a variety of malignancies,
including pituitary adenoma, glioma, head and neck tumors,
pancreatic, colon, prostate, lung, and breast cancers.[] NTS
function is mediated by neurotensin receptors, which are com-
prised of three subtypes (NTSR, , ). NTSR and NTSR are
typical (with seven transmembrane domains) GPCRs, whereas
NTR is a single transmembrane receptor belonging to the
Sortilin family. NTSR has mainly been identified in the CNS
and has a lower anity for NTS, while NTSR and NTSR are
found in a variety of human peripheral tissues and cancer cells.
Interestingly, in some cases, these three receptors can interact
with each other by forming functional heterodimers to finely
adjust their downstream functions.[,]
NTS is produced by multiple cancers, and expression of
NTSRs is also very common (e.g., in % of breast cancer,
% of small-cell lung cancer, –% of pancreatic cancer,
–% melanoma).[] Of note, there may be significant dier-
ences between NTSR expression due to dierent cancer types/
stages and detection methods, such as RT-qPCR, western blot,
northern blot, and autoradiography.[] Generally, NTS may
influence tumors in a conventional endocrine manner, with
endogenous NTS production leading to enhanced tumor devel-
opment via NTR signaling. Furthermore, some tumors have
been found to not only contain NTS receptors but also generate
NTS peptide, resulting in autocrine and/or paracrine actions.
Based on current studies, NTS/NTSRs functions will be dis-
cussed in two types of malignancies: neuroendocrine tumors
and epithelial-derived tumors.
. Neuroendocrine tumors arise in the pituitary, prostate, gas-
trointestinal tract, lung, and thyroid gland, containing spe-
cialized cells with both characteristics of endocrine cells and
nerves. In the androgen-dependent human prostate cancer
cell line LNCap, for example, NTS secretion and NTSR over-
expression were observed under androgen deprivation, fol-
lowed by enhanced neuroendocrine dierentiation, cell pro-
liferation, and invasion.[] Androgen-dependent PC cells
are not able to produce NTS, however NTSR expression sig-
nificantly increased when cells were supplemented with very
low concentrations of NTS (.–n, near physiologic post-
prandial blood concentrations).[] Peripheral activated B lym-
phocytes may be a possible source of circulating NTS.[,]
These results support an endocrine, autocrine/paracrine
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (9 of 20)
www.advanced-bio.com
role for NTS/NTSR activation. Similar autocrine/paracrine
eects also occur in small-cell lung cancer[] and pituitary
adenoma.[] In both cell culture and xenograft models,
higher amounts of NTS secretion and NTSR expression
were characteristic of malignant rather than normal cells,
while NTS-induced cell proliferation and migration can
be dramatically attenuated by SR (an FDA-approved
selective antagonist to NTSR).[] Furthermore, NTS ex-
pression (not NTSR) is associated with significantly worse
survival in melanoma patients.[] Together, NTS, NTSR,
and NTSR are found in neuroendocrine tumor cells, where
they contribute to tumor development via autocrine/parac-
rine or endocrine pathways.
. NTS/NTSR overexpression and a typical autocrine/paracrine
stimulation are also observed in epithelial-derived cancers.
In breast and colon cancers, for instance, NTS appears in
neoplastic cells and induces NTSR overexpression, which
promotes tumor growth and invasion, a process that can be
inhibited by the NTSR antagonist SR.[,] Moreover,
both NTSR and NTSR are found in pancreatic cancers in
the presence of NTS,[,] activating mitogenic signaling
pathways to promote cancer initiation and development.
Interestingly, NTS-induced cell migratory ability depends
on dierent ECM environments; intrinsic cell mobility is
increased, while collective cell migration is reduced.[] In
conclusion, NTS and NTSRs are typically observed in epithe-
lial-derived malignancies, where they promote cancer pro-
gression primarily via autocrine/paracrine pathways.
The influence of NTS/NTSR signaling on cancer progres-
sion can be loosely organized into four stages[] (Figure 3).
) Premalignancy is the first stage, in which various inflam-
matory cytokines, MMPs and ROS are induced under chronic
inflammation. As a result, NTS can act here as a proinflam-
matory neuropeptide, and NTSR is significantly upregulated
during chronic inflammation.[] Within this microenviron-
ment, DNA damage leads to mutations, putatively promoting
carcinogenesis via genomic instability.[] This genomic
damage may result in morphological lesions that are known to
be precursors to certain malignancies. ) Genetic disturbances
and carcinogen stimulation cause epithelial disruption and
facilitate a transition to the malignant state. In this stage, cells
grow uncontrollably, but the cancerous carcinoma is still small
and only localized to one area. Autocrine/paracrine/endocrine-
activated NTSR and NTSR participate in cell survival and
proliferation through the PKC/MAPK/ERK and Ras/PIK/AKT
pathways.[] ) When the tumor is larger, the stroma is altered
due to the loss of the epithelial barrier (i.e., basement mem-
brane), resulting in an invasive state. More signaling cascades
are involved in progression of the malignancy toward a more
aggressive state, such as small RhoGTPases, IL-/CXCR, and
JAK/STAT pathways.[] These can induce stromal damage
and cancer cell infiltration into adjacent tissues or lymph
nodes. ) Enhancers from multiple signaling pathways induce
angiogenesis and further contribute to distant metastasis.
This happens when cancer cells break o a tumor and move
through the body in the bloodstream or lymph system. In the
last two stages, cancer cells in the tumor core are undergoing
hypoxia and are surrounded by a very acidic microenvironment.
NTSR signaling promotes a metastatic phenotype in pancre-
atic cancer cells by inducing localized extracellular acidifica-
tion in normoxic cells, preceding acidosis caused by hypoxia,
and switching to glycolysis. Intriguingly, an NTS analog Lys-
ψ-LysNT exhibits rapid and robust alkalinization to counteract
the acidic microenvironment via activating sodium/proton
exchanger  activity.[] Despite advances in understanding
Figure 3. Neurotensin system (NTS/NTSR) influences on cancer progression. Chronic inflammation generates cytokines, MMPs, and ROS, which
cause genomic damage and promote a premalignant state. Gene dysregulation and carcinogens stimulate NTS/NTSR signaling pathway, causing
epithelial damage and facilitating the transition to a malignant state. Loss of the epithelial barrier alters the stroma, giving rise to an invasive state.
NTS/NTSR system might be further activated through autocrine, paracrine, or endocrine pathways, leading to tumor microenvironment remodeling
and angiogenesis typical of the highly invasive stage. Cellular enhancers, an acidic environment, and hypoxia all contribute to tumor development and
distal metastasis.
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (10 of 20)
www.advanced-bio.com
of the NTS/NTSR system in cancer over the last two decades,
further findings are needed for the development of targeted
therapeutics.
3.4. Orexins/Hypocretins
The orexins/hypocretins (Orexin-A/B, also known as Hypo-
cretin-/) are hypothalamic neuropeptides that regulate sleep/
arousal, food intake, energy homeostasis, stress, reward, and
drug addiction, among other functions. However, orexins are
not limited to the central nervous system, a small fraction is
also produced by the enteric nervous system and secretory tis-
sues in periphery, where they participate in the regulation of
blood pressure, heart rate, gut motility, and hormone secretion.
The two subtypes of orexin (OX), orexin-A (OxA), and orexin-B
(OxB), orchestrate their diverse functions by binding and acti-
vating two GPCRs, OXR, and OXR. OxA has a greater anity
binding to OXR, whereas both OxA and OxB have a similar
anity to OXR. Although orexin-expressing neurons project
widely throughout the CNS, these two receptors are distributed
in dierent brain regions,[] reflecting their dierential physi-
ological functions. Recently, another exciting aspect of orexins
and their receptors has emerged, which is their potential roles
in inflammation and cancer.
Although multiple studies have demonstrated that orexins
induce apoptosis resulting in significant reduction in cancer
cell growth, their proliferative or inhibitory eects are highly
cell-type dependent. On one hand, both orexin isoforms (OxA
and OxB) were found to consistently promote apoptosis and
suppress cell growth in multiple colon cancer cell lines.[]
OXR was only expressed in malignant (and not adjacent
normal) cells as revealed by RT-qPCR and immunohistochem-
istry.[,] In most of the colon cancer cell lines expressing
both receptors, orexin treatment significantly induces apop-
tosis and inhibits growth. Only HCT- cells, which do not
express OXR, did not undergo apoptosis upon treatment
with OxA.[] Similar results were obtained from in vivo colon
tumor xenografts, where dramatic reductions in tumor sizes
were observed for all cell lines upon OxA treatment, excepting
HCT- cells.[] Another study from Wen et al. indicates that
OxA can also reduce cell viability and promote apoptosis in
HCT- cells, but the cells subsequently upregulate autophagy
processes to resist apoptosis,[] likely attenuating the eects of
orexin on cell death. Additionally, dramatic reductions in cell
viability was found in OxA treated rat glioma cells with an IC
in the nanomolar range. These results reflect an overall anti-
cancer eect mediated by OXR.
On the other hand, treatment of gastric cancer cells with
OxA leads to overexpression of OXR, but paradoxically results
in enhanced cell survival and proliferation through activa-
tion of Akt signaling and inhibition of caspase- activity via
ERK/.[,] The same proliferative eects from orexin has
additionally appeared in adrenal cancers.[,] Despite the rel-
atively consistent negative or positive influences in the above
malignancies, orexins have a mixed eect in prostate cancer.
OXR was upregulated upon orexin supplement both in cell
cultures[,] and clinical samples,[] however this was also
associated with increased cell growth in nondierentiated cells
and apoptosis in dierentiated cells. This indicates that there
might be a transformational regulation of orexin receptor
expression during malignancy, and therapeutic treatment
windows may need to take into account receptor expression
dynamics in early versus late-stage prostate cancers.[]
Theoretically, orexins may exert their actions via autocrine,
paracrine, and/or endocrine pathways in order to regulate phys-
iological functions.[] However, it is notable that OxA has never
been detected in tumors,[] suggesting that OXR present in
tumoral tissue was not activated by endogenous OxA. In other
words, the orexin system does not exert its functions via an
autocrine mode based on current studies. Also, the endocrine
pathway is also less plausible as the concentration of circulating
orexins is about  p, which is too low to activate OXR in
tumors. Therefore, both endogenous orexins and OXR have
no impact on tumor growth, and they can only aect malig-
nancy in the presence of exogenous orexins.[] Despite the lack
of orexins within tumors, OxA expression can be detected in
“fiber-like” stroma surrounding prostate tumors and follicular
exocrine epithelia,[] indicating a possible impact on tumor
growth through stromal cells within the TME. Therefore, the
orexin/OXR system probably acts through paracrine eects on
cancerous cells. As described above, their activation can trigger
multiple signaling pathways for the modulation of cell survival,
proliferation, and apoptosis, including cAMP, JNK, p/MAPK,
ERK/, and PIK/Akt. Generally, the cAMP pathway activates
relevant enzymes and regulates gene expression, JNK, ERK/,
and PIK/Akt activation triggers cell proliferation, cell cycle
progression, and apoptosis suppression, while p/MAPK par-
ticipates in programmed cell death.[,] More recently, Voisin
et al. revealed that the orexin/OXR system induced a mito-
chondrial apoptosis response, which is driven by a new sign-
aling pathway including immunoreceptor tyrosine-based motifs
(ITIM) and the tyrosine-protein phosphatase nonreceptor type
(SHP) both in Chinese hamster ovary (CHO) cells and colon
cancer cells.[] In conclusion, recent studies on orexin sign-
aling provides a wide range of possibilities in the treatment of
certain cancers; however, we are still at an early stage in this
novel field and much future work is required.
3.5. Other Neuropeptides
Multiple other neuropeptides, including bombesin-like pep-
tides, gastrin, cholecystokinin (CCK), arginine vasopressin
(AVP), somatostatin (SST), vasoactive intestinal polypeptide
(VIP), and others have also been implicated in cancer growth
and metastasis.[] For unclear reasons, several of them failed
to attract further research interest beyond initial reports, as the
goal of neuropeptide research shifted from basic to clinical.
SST and VIP are two with potentially significant therapeutic
opportunities.
SST, also known as SRIF, is a family of cyclic peptides
mainly produced by normal endocrine, immune, and neuronal
cells, and by certain tumors.[,] SSTs have broad inhibitory
eects on both endocrine and exocrine secretion. SST recep-
tors (SSTR) are widely distributed in healthy tissues and are
involved in various processes, such as inhibiting hormone
secretion, reducing cell proliferation, and promoting apoptosis,
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (11 of 20)
www.advanced-bio.com
which implies a potential inhibitory role for these peptides in
cancer growth. Indeed, preclinical studies showed that SST–
SSTR interactions have tumor-suppressor activity in certain
cancers.[] Direct antitumoral activity is mediated through
receptors on tumor cells. SSTRs are connected with G proteins,
and through them somatostatin inhibits adenyl cyclase (Gi),
activates potassium/calcium channels, stimulates protein phos-
phatases, and intracellular tyrosine phosphatase. In addition,
somatostatin has an indirect antiproliferative eect on tumors
by the inhibition of growth factors and angiogenesis.[,]
Vasoactive intestinal peptide (VIP) is a neuropeptide con-
sisting of  amino acids, which is widely produced in both the
central and peripheral nervous system, as well as in endocrine
and immune cells.[] Besides its canonical neurotransmitter
and neurotrophic eects, VIP also acts as neuroprotective factor
in the CNS via a cAMP-dependent direct pathway, or indi-
rectly via the actions of astrocytes/Treg cells.[] Three types of
VIP receptors (VIPRs) (i.e., VPAC, VPAC, and PAC) have
been identified, all of which belong to the Class B G protein-
coupled receptors. VIPRs are ubiquitously expressed in human
malignancies, where VIP/VIPRs function in an autocrine
fashion where the tumor both possesses VIPRs and secretes
VIP. VPAC is specifically overexpressed in most epithelial
malignancies, including colon, breast, lung, prostate cancers,
as well as in brain tumors like gliomas, while VPAC recep-
tors are only found in some leiomyomas and gastrointestinal
stromal tumors.[,] However, their eects on cancer growth
are bidirectional. In some tumors, activation of these receptors
has growth and invasion stimulatory eects through activation
of NF-kB signaling, increasing expression of cyclin D, MMPs,
VEGF, and decreasing cell adhesion.[,] Their growth inhibi-
tory eects were also reported in glioma cells,[] renal cancer
cells,[] retinoblastoma,[] and human small-cell lung cancer
in vitro and in vivo.[,] These opposing eects are caused by
dierent receptor subtypes coupling to dierent signaling path-
ways,[,,] specific in vitro cell culture settings with calorie
restriction,[] or via unknown mechanisms.[]
Table 2 summarizes the roles of the above-mentioned neuro-
peptides and receptors in cancer, as well as their mechanisms
of action and related signaling networks. The following sections
discuss potential regulatory mechanisms of neuropeptide/
receptor systems, especially their epigenetic regulation using
SST and VIP as examples, as well as their possible therapeutic
applications in cancer.
4. Neuropeptides in Cancer-Associated Behavioral
Problems
Acting as endocrine hormones, neuromodulators, or neuro-
transmitters, neuropeptides regulate a wide variety of physi-
ological processes and behaviors in neurons and non-neuronal
tissues, such as growth, learning and memory, food intake, tem-
perature management, and autonomic responses.[] Therefore,
Table 2. Functions, action modes, and signaling pathways of neuropeptides and their receptors in cancer.
Neuropeptide/
receptor(s)
Cancer types Functions in cancer Action modes in cancer Signaling pathways Ref.
Neuropeptide Y NPY/
Y1R, Y2R
Breast, ovarian, prostate
cancers, pituitary tumor,
adrenocortical lesions,
pheocromocytoma,
neuroblastoma,
gastroenteropancreatic tumors
Cell proliferation, mitogenesis,
matrix invasion, migration,
angiogenesis
Promotion/Suppression of
cancer growth through
autocrine and paracrine
mechanisms
MAPK/ERK, PI3K/Akt [115, 122, 124,
128, 134, 136]
Substance P SP/NK1R Breast, ovarian, pancreas, thyroid,
prostate, lung, oral, head, and
neck cancers, glioma/
astrocytoma, melanoma,
leukemia, retinoblastoma,
larynx carcinoma
Cell proliferation, antiapoptosis,
migration (invasion, infiltration
and metastasis), chronic
inflammation, neoangiogenesis
Promotion of cancer
development through
autocrine, paracrine, endocrine
pathways, or peripheral
nervous system
Wnt/β-catenin, MAPK/
ERK, PI3K/Akt, mTOR
[105, 108, 140,
144]
Neurotensin NTS/
NTR1, NTR3
Pancreatic, colon, prostate, lung,
liver, breast cancers, pituitary
adenoma, glioma, head, and
neck tumors
Cell proliferation, survival,
EMT, migration, invasion,
neoangiogenesis
Stimulation of cancer growth
via endocrine, autocrine, and
paracrine pathways
MAPK/ERK, PI3K/Akt,
FAK/Src, Ca2+
mobilization, JAK/STAT3
[174, 192, 227,
228]
Orexin/Hypocretin
OX/OX1R, OX2R
Colon, pancreas, prostate, gastric
cancers, neuroblastoma
Induction of apoptosis and
inflammation, increased
cell proliferation
Inhibition/Promotion of cancer
growth via paracrine pathway
cAMP, JNK, p38/MAPK,
ERK1/2 ITIM/SHP2
[209, 229–231]
Somatostatin SST/
SST1R-5R
Breast, colon, prostate,
gastrointestinal, small cell lung
cancers
Inhibition of hormone
secretion and cell proliferation,
cell-cycle arrest, induction of
apoptosis, tumor-imaging
Inhibition of cancer growth via
autocrine, paracrine, endocrine
and exocrine pathways
MAPK/ERK, PI3K/Akt [215, 232–234]
Vasoactive Intestinal
Polypeptide VIP/
VPAC1R, VPAC2R
Breast, lung, prostate, pancreas
cancers, glioma, astrocytoma
Activation (common)/
inhibition (less common) of
tumor growth, angiogenesis,
tumor-imaging
Stimulation/Inhibition of
cancer growth via endocrine,
autocrine, and paracrine
pathways
cAMP/PKA/ERK, PI3K/
Akt, PLC/Ca2+,CREB,
NF-κB,
[221, 235, 236]
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (12 of 20)
www.advanced-bio.com
it is critical that we understand how neuropeptides can integrate
the functions of the brain and the systems of the body. Major
advances in the last  years have identified the importance of
neuropeptides such as those released by hypothalamic neurons.
The hypothalamus is a phylogenetically conserved and impor-
tant brain regions responsible for many basic behavioral and
physiological functions including feeding and digestion, fluid
and electrolyte balance, energy metabolism, metabolic control,
reproduction, thermoregulation, and sleep-wake cycles.[,]
As such, the hypothalamus is a critical component of the neu-
roendocrine system in which hypothalamic neurons control
the secretion of hormones in the anterior pituitary gland and
consequently in its target endocrine organs in the periphery.
Peripheral endocrine secretions, in turn, influence neuronal
function in the central nervous system in a predominantly neg-
ative feedback loop. This principle is important because most
hypothalamic and pituitary hormones are neuropeptides, high-
lighting how neuropeptides can integrate the functions of the
brain and the systems of the body.
In addition to normal neuroendocrine secretion, neuropep-
tides can also be produced by tumor cells and this may result
in direct and indirect cancer-associated disorders beyond their
actions within the TME (see Figure4). A brief understanding of
how the brain and body are integrated via the neuroendocrine
system and how tumor-derived neuropeptides are relevant to
this discussion can be achieved through examining the eects
of the pituitary peptide adrenocorticotropic hormone (ACTH).
As part of the hypothalamic-pituitary-adrenal (HPA) axis, ACTH
acts downstream corticotropin-releasing hormone (CRH) on the
adrenal cortex to control the subsequent release of glucocorti-
coids (e.g., cortisol, corticosterone). Glucocorticoid release in
turn exerts a multitude of eects on host physiology (e.g., immu-
nosuppression, integrated stress responses, hyperglycemia) but
importantly, also results in feedback inhibition on the hypothal-
amus and pituitary gland to regulate the magnitude and dura-
tion of glucocorticoid release. Although cortisol is essential to
life, chronically high concentrations of cortisol can damage the
various tissues throughout the body. Common manifestations
include diabetes, hypertension, sudden mood changes, muscle
weakness, and irregular menstrual cycles. The main causes of
Cushing’s syndrome (i.e., chronically elevated cortisol levels)
are ACTH-dependent (when the body makes too much ACTH,
which in turn increases cortisol production), ACTH-independent
(when the adrenal is making too much cortisol and ACTH is
therefore low), and iatrogenic (when the patient is taking long-
term administration of corticosteroid drugs) in nature. Pituitary
adenomas are the most prevalent type of ACTH-dependent dis-
ease, accounting for over % of all occurrences. In this condi-
tion, a small tumor causes increased ACTH production. Ectopic
ACTH-producing tumors comprise the other type of ACTH-
dependent disease. In this rare condition, tumors outside of the
pituitary are generating too much ACTH, most commonly in the
Figure 4. Indirect and direct eects of TME-derived neuropeptides on cognition relevant to cancer. A) Physiological and psychological stress induced
by the presence of a peripheral tumor results in the engagement of the hypothalamic-pituitary-adrenal (HPA) axis and subsequent ACTH release by
the anterior pituitary gland. In addition, ectopic ACTH expression by cancer cells may contribute to increased systemic ACTH levels and subsequent
elevations in systemic cortisol. The consequences include a deregulated negative feedback loop on the HPA axis and subsequent cognitive impairment
attributed mainly to elevated cortisol. These changes can also result in indirect eects on functional capacity, organ function, and mobility. B) Tumor-
derived substance P is secreted by cancer cells in the TME where it enters systemic circulation and eventually the brain parenchyma where it binds its
receptors (e.g., NK-1) which are abundant in the limbic system. Activation of these receptors in the limbic system engages neuronal circuits that result
in behavioral changes including depression, stress, and anxiety. These changes in behavioral states result in enhanced production of substance P by the
limbic system which enter systemic circulation via circumventricular organs and influence the TME in a way that putatively favors tumor progression
(i.e., tumor cell proliferation, angiogenesis, metastasis).
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (13 of 20)
www.advanced-bio.com
lung and thymus gland, but have also been found in the thyroid,
ovary, adrenal gland, and liver. In the ACTH-independent cases,
either both adrenal glands are hyperactive or there is an adrenal
tumor that is making too much cortisol (about %). Although
most of these tumors are benign, they can rarely progress into
adrenal cancers.
However, ACTH also influences complex behavioral pro-
cesses including learning and memory,[–] highlighting
the ability of neuropeptides to travel long-distances within the
CNS where they have direct eects on neurons that are inte-
grated within distinct neuronal circuits. Interestingly, a study
by Heesen et al. demonstrates that hyperactivation of the HPA
axis is related to increased cognitive impairment in multiple
sclerosis.[] Given that cancer patients experience excess levels
of stress, it is conceivable that a similar alteration in the HPA
axis is present in cancer patients. Classic work on frog sympa-
thetic ganglia also demonstrates this ability of neuropeptides
to act on cells micrometers away from their release site–which
may or may not be released at the synaptic site (i.e., they are
ectopically expressed).[] Additionally, ACTH is produced by
other brain regions and peripheral tissues (e.g., gastrointestinal
tract, lymphocytes), where if present in abundant concentra-
tions, can travel to the brain parenchyma and engage in “short-
loop” negative feedback on the hypothalamus with downstream
consequences on systemic physiology and behavior.[,–]
As alluded to previously, neuropeptides influence cognitive
function due to their actions in cortical and subcortical brain
regions.[] Furthering our discussion using the neuropep-
tide ACTH as an example, several seminal papers have dem-
onstrated the role of ACTH in adaptive behavior in preclinical
models.[–] Many of these behavioral deficits induced by
removal of the anterior pituitary gland can be corrected by
administration of ACTH, highlighting its dramatic eects on
cognitive function. Similarly, administration of CRH directly
into the ventricular system in the brain also results in behav-
ioral activation including anxiety and stress responses due to
elevated plasma norepinephrine (NE) and epinephrine.[–]
This demonstrates that neuropeptide eects in the CNS can
extend to the periphery aecting a myriad of physiological out-
puts (e.g., functional capacity, organ function, mobility). Impor-
tantly, the eects of neuropeptides on distinct neuronal circuits
within the CNS is largely due to the ability of neuropeptides
to diuse long distances within the extracellular compartment
and cerebrospinal fluid.[]
Aside from tumor-released neuropeptides, the TME is another
significant source of neuropeptide synthesis. Similarly, it is con-
ceivable that TME-derived neuropeptides (e.g., NPY, SP) have
this same capacity to exert eects on neuronal circuitry directly
or indirectly through altering the actions of other host factors.
Interestingly, SP itself, whose expression is increased in breast
cancer cells, has been shown to directly result in blood brain bar-
rier (BBB) impairment.[] The disruption in the BBB is mainly
achieved through SP-mediated changes in the tight junction pro-
teins ZO- and claudin- in brain microvascular endothelial cells.
The ability of neuropeptides to infiltrate the brain parenchyma
where they exert eects on neuronal circuitry is largely attributed
to the ability of neuropeptides to bypass the BBB via circum-
ventricular organs-regions of the brain that are characterized by
the lack of a BBB and therefore allow for direct communication
between the brain and circulatory system.[] Additionally, these
circumventricular organs (e.g., median eminence) permit hypo-
thalamic neuropeptides to leave the brain without disrupting
the BBB where they can influence systemic physiology.[] In
addition, the presence of peripheral tumors has been shown to
disrupt the integrity of the BBB during cancer progression, pre-
senting an additional route by which TME-derived neuropeptides
can exert eects on neuronal circuitry.[,]
Additionally, SP is also highly relevant to the discussion of
cancer-induced neurological dysfunction (e.g., depression,
stress, anxiety). It is widely recognized that cancer and depres-
sion frequently accompany one another, and depressive symp-
toms can be exacerbated by disease severity and other cancer
associated symptoms (e.g., pain, fatigue). The limbic system
is a brain structure that includes the hypothalamus, amygdala,
and hippocampus. As a result, the limbic system is involved
in the processing of emotion, memory and behavior through
the coordinated activity of interconnected cortical and subcor-
tical brain structures. SP and its receptor NK- are expressed in
the limbic system and their involvement has been implicated
in neurological disorders including depression, stress, and
anxiety.[–] Interestingly, the NK- receptors expressed in the
limbic system are similar to the NK- receptors overexpressed
in human cancer cells and patient tissue samples.[] As men-
tioned previously, SP is also highly expressed in several tumors
including breast. Within the TME, SP is released by nerves and
can enter the systemic circulation to influence systemic physi-
ology including CNS function. SP can enter the brain paren-
chyma, via the aforementioned routes, where it exerts its eects
on neuronal circuits and influences subsequent behavior rel-
evant to cancer (i.e., depression, anxiety, stress) see Figure .
Additionally, the onset of depression can increase the produc-
tion of substance P from the limbic system, where it can enter
systemic circulation, via circumventricular organs, and influ-
ence the TME with consequences such as increased tumor
progression.[,] Thus, TME-derived neuropeptides (e.g., SP)
may have the capacity to increase tumor progression locally and
influence neuronal function distally in the CNS with eects on
cognition and behavior relevant to cancer.
5. Regulation and Therapeutic Applications of
Neuropeptide/Receptor Systems in Cancer
5.1. Regulation of Neuropeptide/Receptor Systems
Although a few neuropeptide receptors are intrinsically
expressed in both healthy and neoplastic cells for physiological
uses, it is more common that they are only present in malig-
nant tissues under inflammation or after exogenous neuro-
peptide treatment. In some cases, neuropeptide receptors shift
their expression from one subtype to the other. For example,
Swift et al. observed that when prostate cancer cell lines become
more aggressive and less dierentiated, elevated NTR expres-
sion shifts toward NTR.[] The appearance of neuropeptide
receptors indicates the stemness and programmed dierentia-
tion abilities of these cells.
External control of receptor function arises from the addition
of agonists or antagonists similar to endogenous ligands. A broad
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (14 of 20)
www.advanced-bio.com
range of neuropeptide mimics and inhibitors have been devel-
oped, with a few already approved and therapeutically accessible
on the market, e.g., the NTS/NTSR antagonist SR; the
SP/NKR antagonist aprepitant. SR (meclinertant) was the
first nonpeptide selective antagonist developed for NTSR. It is
used in scientific research to investigate the interaction between
NTS and other neurotransmitters in the brain, and exhibits anxi-
olytic, antiaddictive, and memory-impairing eects in animal
studies. Aprepitant (Emend, Merck) is the first agent available in
the new drug class of NKR antagonists, clinically used to treat
chemotherapy-induced nausea and vomiting.
Neuropeptide receptor expression is regulated by signal
transduction pathways like the Wnt/β-catenin signaling
pathway, which governs downstream cascades regulating tran-
scriptional control of proliferation and dierentiation. Upon
Wnt signaling, nuclear translocalization of β-catenin activates
the NTSR promoter by creating a complex with transcription
factor Tcf.[] This provides a direct link between develop-
mental and oncogenic processes and neuropeptide function.
Epigenetic modifications are further critical in regulating gene
expression and controlling cancer progression.[] Epigenetic
changes are heritable throughout cell divisions, but they are
reversible since the DNA sequence is not changed. DNA meth-
ylation, histone modifications, and microRNA expression are
common types of epigenetic regulatory mechanisms, resulting in
inaccessible DNA or nascent transcript degradation preventing
the production of mature proteins. A decade ago, DNA meth-
ylation was found in the promoter region of neurotensin gene,
which induces severe reduction of transcription and expres-
sion in the human liver cancer cell line Hep G.[] Expression
of the somatostatin receptor (SSTR) was also downregulated in
neuroendocrine tumors[] and gastric cancer cells[] due to
promoter methylation. The other example is from the pituitary
adenylate cyclase activating polypeptide (PACAP), which belongs
to the VIP/PACAP/secretin family and shares % homology
with VIP. In cervical cancer, PACAP is considered as a methyla-
tion biomarker for early cancer detection.[] In addition to DNA
methylation, histone acetylation is also involved in regulating
SSTR.[] Furthermore, microRNAs can mediate the expression
of some neuropeptide receptors in certain disease models. Jo et al.
demonstrated that SSTR expression was upregulated by
miRNA (miR--p) transfection, and this miRNA enhanced the
SST analog (octreotide)-induced reduction in cell proliferation
in both D/D cultures of neuroendocrine cells.[] In cell-based
models of the chronic bladder pain syndrome, long-term sub-
stance P treatment reduced NKR mRNA levels while increasing
regulatory miR-b and miR-.[] These studies indicate
potential uses of epigenetic drugs (like HATs/HDACs)[]
to modulate the expression of certain neuropeptide receptors and
their therapeutic opportunities.
5.2. Targeting Neuropeptide Receptors for Cancer
Molecular Therapy
Targeted therapies generally take advantage of biological mole-
cules that are uniquely expressed or significantly overexpressed
in tumors. A plausible approach is to target neuropeptides,
growth factors, and the signaling pathways that mediate their
mitogenic eects with antibodies, antagonists, and/or selective
inhibitors. Regarding neuropeptide receptors, the presence of a
distinct cell surface receptor in cancer cells verses their normal
counterparts is a molecular hallmark in targeted therapy. So
far, several of them have been widely used for diagnosis and
therapy, including SST receptors, HER, integrin receptors,
EGFR, and VIP receptors.[] The basic principle is to use neu-
ropeptide analogs targeting certain receptors, improving tumor
visualization or blocking proliferative signaling. We discuss
three types of cancer-specific molecular therapies using neuro-
peptide agonists/antagonists, cytotoxic peptide conjugates, and
radiolabeled ligand conjugates below.
. Agonists/Antagonist-based cancer therapy. Although tumor-in-
hibitory neuropeptides (such as SST and OX) have distinct in-
hibitory eects on tumors, their therapeutic potential is limited
due to their relatively short plasma half-life. Thus, metabolically
stable compounds with native ligand-like properties are ex-
pected to show an improved clinical profile. Metabolic stability
can be accomplished by either shrinking the size to eliminate
catalytic sites, generating cyclic structures, or incorporating
modifications (e.g., D-amino acids or N-methylated amino ac-
ids). A wide variety of agonists/antagonists targeting neuropep-
tide receptors are reviewed recently,[] including NPYR,[,]
NKR,[] SSTR,[,] and VIPR.[]
. Cytotoxic peptide conjugate-based cancer therapy. Conven-
tional tumor chemotherapy is restricted by multidrug resist-
ance, toxicity to normal cells, and a lack of tumor selectivity.
More targeted delivery of chemotherapeutic drugs to cancer
cells can result in higher drug concentrations in tumors, while
lowering toxicity to normal cells. Owing to the high-anity
of peptides and high-expression of receptors on tumor cell
surfaces, tumors can be targeted by cytotoxic neuropeptide
conjugates. VIP-ellipticine conjugates, for example, act as
VPAC receptor agonists, binding to VPAC receptors with
high anity while retaining antiproliferative eect. In one
study, they were absorbed by cancer cells expressing the
VPAC receptor and then catalyzed by proteolytic enzymes,
resulting in ellipticine release and cellular cytotoxicity.[]
. Peptide-receptor radionuclide therapy (PRRT). Radiolabeled-
neuropeptides (e.g., SST and VIP) or analogs are used in
tumor imaging. PRRT is based on the presence of high lev-
els of receptors in tumors and their ability to form ligand–
receptor complexes, allowing radiopharmaceuticals to be
internalized and accumulate inside tumors. Notably, there
are many studies on SSTR-based radiotherapy,[] whereas
reports on VIPR-based radiotherapy for human tumors are
lacking due to inadequate specificity.[,]
In conclusion, these cancer molecular therapy approaches
suggest that neuropeptide receptor-based imaging and treat-
ment will be crucial for early cancer detection and control.
6. Current Research Strategies and
Future Directions
Notable strategies in neuropeptide research can be summa-
rized into three major areas based on structure, location, and
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (15 of 20)
www.advanced-bio.com
function[] (see Figure 5). ) Mass spectrometry (MS) and
MALDI-TOF are usually used to determine peptide sequence,
posttranslational modifications, and structural information.
Spectroscopy and nuclear magnetic resonance (NMR) give
information on protein conformations and folding patterns.
X-ray crystallography characterizes structural details with high
spatial resolution. ) Regarding peptide location, antibody-
based assays (e.g., radioimmunoassays, immunohistochem-
istry, immunocytochemistry) and immunoelectron microscopy
are commonly used approaches enabling peptide visualization.
Additionally, in situ hybridization is used for target-specific
expression mapping of neuropeptide coding genes, while
promoter-reporter gene constructs enable transcript detection
in living cells and organisms. Moreover, mass spectrometry
imaging[] is capable of capturing the entire neuropeptidome
without prior knowledge. Bio imaging and microscopy map the
architecture of the nervous system. ) Behavioral studies can
first obtain a general understanding of functions as a judging
potential for disease treatment. Electrophysiology approaches
provide understanding of synaptic mechanisms, followed by
quantitative analyses (e.g., qPCR, western blotting, ELISA,
MS) to imply functions by measuring changes in neuropeptide
levels due to specific behaviors or conditions. Furthermore,
Yulong Li’s group has made significant progress in developing
a number of genetically encoded fluorescent indicators, empha-
sizing how these novel biosensors can be used in uncovering
functional roles of neurotransmitters and neuromodulators in
the nervous system and potentially peripheral malignancies.[]
Neuropeptide/receptor systems are a complicated landscape
that may be brought into better resolution by the combined
eorts of chemists and biologists, as well as a revitalized drive
for drug development. For in vitro screening of potential ago-
nists and antagonists, a combination of recombinant expres-
sion and synthetic biology approaches can be employed for
high-throughput and large-scale synthesis of neuropeptide
analog libraries. Due to the low abundance of neuropeptides
in vivo and their high proteolytic degradation vulnerability,
multiple neuropeptides have been expressed at high levels in
E. coli, e.g., an NPY precursor[] and NTS.[] Additionally,
large-scale peptide production usually relies on solid phase
synthesis, by which diversities of ligands are restricted or time/
cost-consuming to achieve. In such cases, cell-free expression
can be an alternative way to achieve peptide synthesis in large-
scale, high-throughput, and flexible reconstitutions. Compared
to neuropeptides themselves, stable production of their recep-
tors with complete biological functionalities is even more chal-
lenging. As discussed, most neuropeptide receptors are integral
transmembrane GPCRs. GPCRs are one of the most pharma-
cologically successful drug targets, with about % of all com-
mercial drugs and % of newly approved drugs on the market
targeting GPCRs.[] However, their preparation remains dif-
ficult in vitro or in vivo. Many of them are also missing crys-
tallography structure information, which hinders study of the
interactions between peptides and their receptors. Notably, with
the broad application of CRISPR/Cas systems in gene knock-
in/out or creation of transgenic animals, genome-scale CRISPR
interference screening has identified novel determinants of
GPCR-mediated transcriptional signaling. Last year, the launch
of the AlphaFold Protein Structure Database delivered a major
breakthrough in protein structure prediction, which paves a
powerful and rapid supplementary way toward understanding
structure/function relationships other than the traditional
labor-intensive crystallization. Upon noncanonical amino acid
incorporation and click-chemistry conjugation, unbiased nas-
cent peptide labeling became possible both in cell culture and
in animal models, by which newly synthesized peptides can be
visualized through clicking on a fluorophore probe (FUNCAT)
or be pulled down from the endogenous protein pool for fur-
ther analysis though clicking on an anity tag (BONCAT).[]
Targeting the neuropeptide/receptor axis may become a more
widely accepted therapeutic strategy for cancer diagnosis and
therapy in the near future. Monitoring biomarkers for major
tumoral alterations associated with the neuropeptide-receptor
signaling network has great potential and deserves more
investigation.
7. Conclusions
The study of neuropeptides in cancer research is developing
rapidly. Multiple neuropeptides and their receptors are found in
a variety type of cancers, usually exhibiting dierential expres-
sion patterns between malignant and normal cells during
tumor progression. Dierent neuropeptides exert distinct
actions on cancer cells either directly or indirectly through the
tumor microenvironment. These actions are due to receptor
coupling with dierent downstream signaling cascades. Neuro-
peptides may also act in an endocrine fashion to regulate distal
processes in the central nervous system and thereby aect
behaviors. Currently, cancer research on peptides is domi-
nated by two active fields: the first is the search for novel pep-
tide receptors overexpressed in certain cancers; the second is
the discovery of new peptide analogs targeting corresponding
receptors for eective cancer visualization and treatment.
Despite our improved understanding of the mechanisms that
influence neuropeptides and their receptor functions, as well
Figure 5. Typical approaches used in neuropeptide research for deter-
mining structure, localization, and function.
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (16 of 20)
www.advanced-bio.com
as our ability to manipulate their expression in cancer cells,
numerous questions remain unanswered. These include: what
causes the disparity across cancer types, as well as the dis-
crepancies between in vitro and in vivo data? Is neuropeptide
receptor subtype and expression levels comparable in primary
tumors and metastases? Can we take advantage of multiple con-
comitant receptor expression in tumors for the use of a thera-
peutic drug cocktail? It is notable that technological advances in
neuropeptide structural elucidation, localization mapping, and
functional understanding have greatly benefited neuropeptide
research, yet no single approach can oer us all the answers we
seek. The combination of various bioanalytical, bioinformatics,
and molecular neuropharmacological tools will drive neuropep-
tide research to new frontiers with likely benefits toward cancer
elimination.
Acknowledgements
This work was supported by a BBRF NARSAD Young Investigator Grant
(No. 28291), an AACR Grant for Transformative Cancer Research (No.
20-20-26-BORN), and a Pershing Square Foundation Innovation Fund
Award to JCB.
Conflict of Interest
The authors declare no conflict of interest.
Keywords
behavior, cancer, neuropeptide, receptor, signaling pathway
Received: April 16, 2022
Revised: May 31, 2022
Published online:
[1] F. Marchesi, L. Piemonti, A. Mantovani, P. Allavena, Cytokine
Growth Factor Rev. 2010, 21, 77.
[2] C.Liebig, G.Ayala, J. A.Wilks, D. H.Berger, D.Albo, Cancer 2009,
115, 3379.
[3] A. A.Bapat, G.Hostetter, D. D.Von Ho, H.Han, Nat. Rev. Cancer
2011, 11, 695.
[4] Z. Zhu, H. Friess, F. F. Dimola, A. Zimmermann, H. U. Graber,
M.Korc, M. W.Büchler, J. Clin. Oncol. 1999, 17, 2419.
[5] A. H.Zahalka, P. S.Frenette, Nat. Rev. Cancer 2020, 20, 143.
[6] M. Monje, J. C. Borniger, N. J. D’silva, B. Deneen, P. B. Dirks,
F. Fattahi, P. S.Frenette, L.Garzia, D. H. Gutmann, D. Hanahan,
S. L. Hervey-Jumper, H. Hondermarck, J. B. Hurov, A. Kepecs,
S. M. Knox, A. C. Lloyd, C. Magnon, J. L. Saloman, R. A. Segal,
E. K. Sloan, X. Sun, M. D. Taylor, K. J. Tracey, L. C. Trotman,
D. A.Tuveson, T. C.Wang, R. A.White, F.Winkler, Cell 2020, 181, 219.
[7] D. D. Shi, J. A. Guo, H. I.Homan, J. Su, M. Mino-Kenudson,
J. L. Barth, J. M. Schenkel, J. S. Loeer, H. A. Shih, T. S.Hong,
J. Y. Wo, A. J.Aguirre, T. Jacks, L. Zheng, P. Y.Wen, T. C.Wang,
W. L.Hwang, Lancet Oncol. 2022, 23, e62.
[8] E. K. Sloan, S. J. Priceman, B. F. Cox, S. Yu,
M. A. Pimentel, V. Tangkanangnukul, J. M. G. Arevalo,
K.Morizono, B. D. W.Karanikolas, L. Wu, A. K.Sood, S. W.Cole,
Cancer Res. 2010, 70, 7042.
[9] A. Kamiya, Y. Hayama, S. Kato, A. Shimomura, T. Shimomura,
K.Irie, R.Kaneko, Y.Yanagawa, K.Kobayashi, T.Ochiya, Nat. Neu-
rosci. 2019, 22, 1289.
[10] F. Chen, X.Zhuang, L.Lin, P.Yu, Y.Wang, Y. Shi, G.Hu, YuSun,
BMC Med. 2015, 13, 45.
[11] M. J. Szpunar, E. K. Belcher, R. P.Dawes, K. S. Madden, Brain
Behav., Immun. 2016, 53, 223.
[12] Y. Shimizu, N. Nagaya, T. Isobe, M. Imazu, H. Okumura,
H.Hosoda, M. Kojima, K. Kangawa, N. Kohno, Clin. Cancer Res.
2003, 9, 774.
[13] J. M Garcia, M. Garcia-Touza, R. A. Hijazi, G. Taet, D. Epner,
D. Mann, R. G. Smith, G. R. Cunningham, M. Marcelli, J. Clin.
Endocrinol. Metab. 2005, 90, 2920.
[14] I.Wolf, S.Sadetzki, H.Kanety, Y.Kundel, C.Pariente, N. Epstein,
B.Oberman, R.Catane, B.Kaufman, I.Shimon, Cancer 2006, 106,
966.
[15] F. F. Bolukbas, H. Kilic, C. Bolukbas, M. Gumus, M. Horoz,
N. S.Turhal, B.Kavakli, BMC Cancer 2004, 4, 29.
[16] A. M.Wallace, A.Kelly, N.Sattar, C. S.Mcardle, D. C.Mcmillan,
Nutr. Cancer 2002, 44, 157.
[17] G. J. Siegel, B. W. Agrano, R. Wayne Albers, P. B.Molino, in
Basic Neurochemistry: Molecular, Cellular, and Medical Aspects, Vol.
27, 6th ed., 1999, p. 465–466.
[18] K. Toshinai, Y. Date, N. Murakami, M. Shimada, M. S.Mondal,
T. Shimbara, J.-L. Guan, Q.-P. Wang, H. Funahashi, T. Sakurai,
S.Shioda, S. Matsukura, K.Kangawa, M.Nakazato, Endocrinology
2003, 144, 1506.
[19] E.Rozengurt, Trends Endocrinol. Metab. 2002, 13, 128.
[20] I.Berczi, I. M. Chalmers, E.Nagy, R. J.Warrington, Baillieres Clin.
Rheumatol. 1996, 10, 227.
[21] L. Souza-Moreira, J. Campos-Salinas, M. Caro, E. Gonzalez-Rey,
Neuroendocrinology 2011, 94, 89.
[22] C. P.Sung, A. J.Arleth, G. Z.Feuerstein, Circ. Res. 1991, 68, 314.
[23] H. A.Bull, J.Hothersall, N.Chowdhury, J.Cohen, P. M.Dowd, J.
Invest. Dermatol. 1996, 106, 655.
[24] M.Delgado, D.Pozo, D.Ganea, Pharmacol. Rev. 2004, 56, 249.
[25] M.Delgado, G. MJonakait, D.Ganea, Glia 2002, 39, 148.
[26] E.Gonzalez-Rey, M.Delgado, Brain Behav. Immun. 2008, 22, 35.
[27] D. Dewit, P. Gourlet, Z. Amraoui, P. Vertongen, F. Willems,
P.Robberecht, M.Goldman, Immunol. Lett. 1998, 60, 57.
[28] Z.Xin, S.Sriram, J. Neuroimmunol. 1998, 89, 206.
[29] M.Schäer, T. Beiter, H. D.Becker, T. K. Hunt, Arch. Surg. 1998,
133, 1107.
[30] E.Rozengurt, Curr. Opin. Oncol. 1999, 11, 116.
[31] E.Rozengurt, J. H.Walsh, Annu. Rev. Physiol. 2001, 63, 49.
[32] J. CReubi, M.Gugger, B.Waser, J. C.Schaer, Cancer Res. 2001, 61,
4636.
[33] A. M. Isidori, A. Lenzi, Arq. Bras. Endocrinol. Metabol. 2007, 51,
1217.
[34] G. W. Liddle, J. R. Givens, W. E. Nicholson, D. P. Island, Cancer
Res. 1965, 25, 1057.
[35] B. L.Wajchenberg, B. B.Mendonça, B.Liberman, M. A. A.Pereira,
M. A.Kirschner, J. Steroid Biochem. Mol. Biol. 1995, 53, 139.
[36] E.Cutz, W.Chan, N. S.Track, A.Goth, S. I.Said, Nature 1978, 275,
661.
[37] E.Gonzalez-Rey, A.Chorny, M.Delgado, Nat. Rev. Immunol. 2007,
7, 52.
[38] E. Gonzalezrey, M. Delgado, Trends Pharmacol. Sci. 2007, 28,
482.
[39] D. R.Gehlert, Neuropeptides 2004, 38, 135.
[40] T. Pedrazzini, F.Pralong, E. Grouzmann, Cell. Mol. Life Sci. 2003,
60, 350.
[41] S.Perboni, M.Vignoni, A. Inui, in Handbook of Biologically Active
Peptides, 2nd ed. (Ed: A. J. Kastin), Academic Press, Boston, MA
2013, p. 1143–1148.
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (17 of 20)
www.advanced-bio.com
[42] C.Salio, L. Lossi, F.Ferrini, A.Merighi, Cell Tissue Res. 2006, 326,
583.
[43] C. LDevane, Pharmacotherapy 2001, 21, 1061.
[44] T.Hökfelt, B.Pernow, J.Wahren, J. Intern. Med. 2001, 249, 27.
[45] E. Svensson, J. Apergis-Schoute, G.Burnstock, M. P. Nusbaum,
D.Parker, H. B.Schiöth, Front. Neural Circuits 2018, 12, 117.
[46] G. R. Uhl, M. J. Kuhar, S. H. Snyder, Proc. Natl. Acad. Sci. USA
1977, 74, 4059.
[47] W. H. Rostène, M. J. Alexander, Front. Neuroendocrinol. 1997, 18,
115.
[48] W. C.Mustain, P. G.Rychahou, B. M.Evers, Curr. Opin. Endocrinol.
Diabetes Obes. 2011, 18, 75.
[49] O. EOsadchii, Usp. Fiziol. Nauk 2004, 35, 3.
[50] T.Kodadek, D.Cai, Mol. BioSyst. 2010, 6, 1366.
[51] A.Inutsuka, A.Yamanaka, Front. Endocrinol. 2013, 4, 18.
[52] J.Li, Z.Hu, L.deLecea, Br. J. Pharmacol. 2014, 171, 332.
[53] B. H.Shamsi, M.Chatoo, X. K.Xu, X.Xu, X. Q.Chen, Front. Endo-
crinol. 2021, 12, 652363.
[54] P.Barnett, Endocrine 2003, 20, 255.
[55] M.Theodoropoulou, G. K.Stalla, Front. Neuroendocrinol. 2013, 34,
228.
[56] M.Iwasaki, Y.Akiba, J. D.Kaunitz, F1000Res 2019, 8, 1629.
[57] A. J. Harmar, J. Fahrenkrug, I. Gozes, M. Laburthe, V. May,
J. R.Pisegna, D. Vaudry, H.Vaudry, J. A.Waschek, S. I.Said, Br. J.
Pharmacol. 2012, 166, 4.
[58] J.Waschek, Br. J. Pharmacol. 2013, 169, 512.
[59] A. KThureson-Klein, R. L.Klein, Int. Rev. Cytol. 1990, 121, 67.
[60] A. F.Russo, Headache 2017, 57, 37.
[61] A. J.Bean, R. H.Roth, J. Neurosci. 1991, 11, 2694.
[62] A. V.Edwards, P.-O.Andersson, J.Järhult, S. R.Bloom, Q. J. Exp.
Physiol. Cogn. Med. Sci. 1984, 69, 607.
[63] K.Iverfeldt, P.Serfözö, L. DArnesto, T.Bartfai, Acta Physiol. Scand.
1989, 137, 63.
[64] A. W. Duggan, P. J. Hope, B. Jarrott, H.-G. Schaible,
S. M.Fleetwood-Walker, Neuroscience 1990, 35, 195.
[65] T.Hökfelt, Neuron 1991, 7, 867.
[66] J. S.Han, Trends Neurosci. 2003, 26, 17.
[67] Q. Wang, L. Mao, Y. Shi, J. Han, Neuropharmacology 1990, 29,
1123.
[68] K.Kumar, C.Toth, R. K.Nath, Neurosurgery 1997, 40, 736.
[69] K. J. Burchiel, V. C. Anderson, F. D. Brown, R. G. Fessler,
W. A. Friedman, S.Pelofsky, R. L. Weiner, J. Oakley, D. Shatin,
Spine 1996, 21, 2786.
[70] K. Racké, F. Burns, B. Haas, J. Niebauer, E. Pitzius, Naunyn-
Schmiedebergs Arch. Pharmacol. 1989, 339, 617.
[71] M.Cazalis, G.Dayanithi, J. J.Nordmann, J. Physiol. 1985, 369, 45.
[72] H.-C.Tsai, F.Zhang, A.Adamantidis, G. D.Stuber, A.Bonci, L.De
Lecea, K.Deisseroth, Science 2009, 324, 1080.
[73] M. Verhage, H. T. Mcmahon, W. E. J. M.Ghijsen, F. Boomsma,
G.Scholten, V. M.Wiegant, D. G.Nicholls, Neuron 1991, 6, 517.
[74] J.Franck, E.Brodin, G.Fried, J. Neurochem. 1993, 61, 704.
[75] A. N.Van Den Pol, Neuron 2012, 76, 98.
[76] D.Hanahan, R. A.Weinberg, Cell 2011, 144, 646.
[77] A. H. Zahalka, A. Arnal-Estapé, M. Maryanovich, F. Nakahara,
C. D.Cruz, L. W. S.Finley, P. S.Frenette, Science 2017, 358, 321.
[78] B. W. Renz, R. Takahashi, T. Tanaka, M. Macchini, Y. Hayakawa,
Z. Dantes, H. C Maurer, X. Chen, Z. Jiang, C. B Westphalen,
M. Ilmer, G. Valenti, S. K. Mohanta, A. J. R. Habenicht,
M. Middelho, T. Chu, K. Nagar, Y. Tailor, R. Casadei, M. Di
Marco, A. Kleespies, R. A. Friedman, H. Remotti, M. Reichert,
D. L. Worthley, J. Neumann, J. Werner, A. C. Iuga, K. P. Olive,
T. C.Wang, Cancer Cell 2018, 33, 75.
[79] C.-M. Zhao, Y. Hayakawa, Y. Kodama, S. Muthupalani,
C. B. Westphalen, G. T. Andersen, A.Flatberg, H. Johannessen,
R. A.Friedman, B. W. Renz, A. K. Sandvik, V.Beisvag, H.Tomita,
A.Hara, M.Quante, Z. Li, M. D. Gershon, K. Kaneko, J. G. Fox,
T. C.Wang, D.Chen, Sci. Transl. Med. 2014, 6, 250ra115.
[80] L. Dollé, I. El Yazidi-Belkoura, E. Adriaenssens, V. Nurcombe,
H.Hondermarck, Oncogene 2003, 22, 5592.
[81] J. PBurbach, Methods Mol. Biol. 2011, 789, 1.
[82] C.Salio, S. Averill, J. V.Priestley, A.Merighi, Dev. Neurobiol. 2007,
67, 326.
[83] A.Kovacs, D. W.Vermeer, M.Madeo, H. D.Reavis, S. J.Vermeer,
C. S.Williamson, A.Rickel, J.Stamp, C. T.Lucido, J.Cain, M.Bell,
M.Morgan, J.-Y.Yoon, M. A.Mitchell, N.Tulina, S.Stuckelberger,
A. Budina, D. K. Omran, E. Jung, L. E. Schwartz, T. Eichwald,
Z. Hong, J. Weimer, J. E. Hooper, A. K. Godwin, S. Talbot,
R.Drapkin, P. D.Vermeer, bioRxiv 2020, 2020.04.24.058594.
[84] P.De Camilli, Trends Pharmacol. Sci. 1991, 12, 446.
[85] A.Betz, P. Thakur, H. J.Junge, U.Ashery, J.-S.Rhee, V.Scheuss,
C.Rosenmund, J.Rettig, N.Brose, Neuron 2001, 30, 183.
[86] G.Pelletier, J.Guy, Y. S.Allen, J. M.Polak, Neuropeptides 1984, 4, 319.
[87] J.Guy, H.Vaudry, G.Pelletier, Brain Res. 1982, 239, 265.
[88] G.Pelletier, R.Leclerc, R. Puviani, J. M. Polak, Brain Res. 1981, 210,
356.
[89] F. L. Strand, Neuropeptides: Regulators of Physiological Processes,
The MIT Press, Cambridge, MA 1998.
[90] L.-Y. M.Huang, E.Neher, Neuron 1996, 17, 135.
[91] K.Fuxe, L. F.Agnati, Volume Transmission in the Brain: Novel Mech-
anisms for Neural Transmission, Vol. 1. Raven Press, New York 1991.
[92] E.Rozengurt, Science 1986, 234, 161.
[93] J. P.Hannon, C.Nunn, B.Stolz, C.Bruns, G.Weckbecker, I.Lewis,
T.Troxler, K.Hurth, D.Hoyer, J. Mol. Neurosci. 2002, 18, 15.
[94] I.-S. Selmer, M. Schindler, J. P. Allen, P. P. A Humphrey,
P. C.Emson, Regul. Pept. 2000, 90, 1.
[95] J. Epelbaum, P.Dournaud, M. Fodor, C.Viollet, Crit. Rev. Neuro-
biol. 1994, 8, 25.
[96] T.Hokfelt, C.Broberger, Z. Q.Xu, V.Sergeyev, R.Ubink, M.Diez,
Neuropharmacology 2000, 39, 1337.
[97] P. S.Cohen, M. J.Cooper, L. J.Helman, C. J.Thiele, R. C.Seeger,
M. A.Israel, Cancer Res. 1990, 50, 6055.
[98] S.Rotshenker, J. Neuroinflammation 2011, 8, 109.
[99] M.Liefner, H.Siebert, T.Sachse, U.Michel, G.Kollias, W.Brück, J.
Neuroimmunol. 2000, 108, 147.
[100] R.Wagner, R. R.Myers, Neuroreport 1996, 7, 2897.
[101] A. Hess, R. Axmann, J. Rech, S. Finzel, C. Heindl, S. Kreitz,
M. Sergeeva, M. Saake, M. Garcia, G. Kollias, R. H. Straub,
O.Sporns, A.Doerfler, K. Brune, G. Schett, Proc. Natl. Acad. Sci.
USA. 2011, 108, 3731.
[102] L.Leung, C. M.Cahill, J. Neuroinflammation 2010, 7, 27.
[103] M.Freidin, J. A.Kessler, Proc. Natl. Acad. Sci. USA 1991, 88, 3200.
[104] C.Rébé, F.Ghiringhelli, Cancers 2020, 12, 1791.
[105] R.Covenas, M.Munoz, Histol. Histopathol. 2014, 29, 881.
[106] M. Muñoz, A. González-Ortega, M. Rosso, M. J. Robles-Frias,
A.Carranza, M. V. Salinas-Martín, R. Coveñas, Peptides 2012, 38,
318.
[107] M.Muñoz, R.Coveñas, Peptides 2013, 48, 1.
[108] X.Li, G.Ma, Q.Ma, W.Li, J.Liu, L.Han, W.Duan, Q.Xu, H.Liu,
Z.Wang, Q.Sun, F.Wang, E.Wu, Mol. Cancer Res. 2013, 11, 294.
[109] C. P. Tensen, K. J. A. Cox, J. F. Burke, R. Leurs, R. C. Van Der
Schors, W. P. M.Geraerts, E. Vreugdenhil, H. Van Heerikhuizen,
Eur. J. Neurosci. 1998, 10, 3409.
[110] D.Larhammar, E.Salaneck, Neuropeptides 2004, 38, 141.
[111] X.Pedragosa Badia, J.Stichel, A.Beck-Sickinger, Front. Endocrinol.
2013, 4, 1.
[112] P. Starbäck, A. Wraith, H.Eriksson, D. Larhammar, Biochem. Bio-
phys. Res. Commun. 2000, 277, 264.
[113] P.Shende, D.Desai, Adv. Exp. Med. Biol. 2020, 1237, 37.
[114] J. C.Reubi, M.Gugger, B.Waser, J. C.Schaer, Cancer Res. 2001, 61,
4636.
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (18 of 20)
www.advanced-bio.com
[115] J. Kitlinska, K. Abe, L. Kuo, J. Pons, M. Yu, L. Li, J. Tilan,
L. Everhart, E. W. Lee, Z. Zukowska, J. A. Toretsky, Cancer Res.
2005, 65, 1719.
[116] M.Körner, B.Waser, J. C.Reubi, Clin. Cancer Res. 2008, 14, 5043.
[117] J.Li, Y.Tian, A.Wu, Regener. Biomater. 2015, 2, 215.
[118] D. A.Dimaggio, J. M.Farah, T. C.Westfall, Endocrinology 1994, 134,
719.
[119] M.Körner, J. C.Reubi, J. Neuropathol. Exp. Neurol. 2008, 67, 741.
[120] N.Abualsaud, L.Caprio, S.Galli, E. Krawczyk, L. Alamri, S. Zhu,
G. I.Gallicano, J.Kitlinska, Front. Cell Dev. Biol. 2021, 8, 627090.
[121] P. Dietrich, L. Wormser, V. Fritz, T. Seitz, M. De Maria,
A. Schambony, A. E. Kremer, C. Günther, T. Itzel, W. E.Thasler,
A.Teufel, J.Trebicka, A.Hartmann, M. F.Neurath, S.Von Hörsten,
A. K.Bosserho, C.Hellerbrand, J. Clin. Invest. 2020, 130, 2509.
[122] M.Körner, J. C.Reubi, Peptides 2007, 28, 419.
[123] M.Ruscica, E. Dozio, M. Motta, P.Magni, Curr. Top. Med. Chem.
2007, 7, 1682.
[124] P. J. Medeiros, B. K.Al-Khazraji, N. M. Novielli, L. M. Postovit,
A. F.Chambers, D. N.Jackson, Int. J. Cancer 2012, 131, 276.
[125] S. Demorrow, P. Onori, J. Venter, P. Invernizzi, G. Frampton,
M. White, A. Franchitto, S. Kopriva, F. Bernuzzi, H. Francis,
M. Coufal, S.Glaser, G. Fava, F. Meng, D. Alvaro, G. Carpino,
E.Gaudio, G.Alpini, Am. J. Physiol. Cell Physiol. 2011, 300, C1078.
[126] X.Lv, F.Zhao, X.Huo, W.Tang, B.Hu, X.Gong, J. Yang, Q.Shen,
W.Qin, Med. Oncol. 2016, 33, 70.
[127] H.Amlal, S.Faroqui, A.Balasubramaniam, S.Sheri, Cancer Res.
2006, 66, 3706.
[128] Y.Ding, M.Lee, Y.Gao, P.Bu, C. Coarfa, B.Miles, A. Sreekumar,
C. J.Creighton, G.Ayala, Prostate 2021, 81, 58.
[129] P.Magni, Curr. Protein Pept. Sci. 2003, 4, 45.
[130] M.Dimitrijević, S.Stanojević, Amino Acids 2013, 45, 41.
[131] S. Jeppsson, S. Srinivasan, B. Chandrasekharan, Am. J. Physiol.:
Gastrointest. Liver Physiol. 2017, 312, G103.
[132] P. F. Wong, M. G. Gall, W. W. Bachovchin, G. W. Mccaughan,
F. M.Keane, M. D.Gorrell, Peptides 2016, 75, 80.
[133] E. J.Hamson, F. M.Keane, S.Tholen, O.Schilling, M. D. Gorrell,
Proteomics: Clin. Appl. 2014, 8, 454.
[134] Z. Zukowska-Grojec, E. Karwatowska-Prokopczuk, W. Rose,
J. Rone, S. Movafagh, H. Ji, Y. Yeh, W.-T.Chen, H. K. Kleinman,
E.Grouzmann, D. S.Grant, Circ. Res. 1998, 83, 187.
[135] J.Pons, J. Kitlinska, H. Ji, E. W.Lee, Z.Zukowska, Can. J. Physiol.
Pharmacol. 2003, 81, 177.
[136] Y. Zhang, C.-Y. Liu, W.-C. Chen, Y.-C. Shi, C.-M. Wang, S. Lin,
H.-F.He, Cell Biosci. 2021, 11, 151.
[137] S.Lecat, L.Belemnaba, J.-L.Galzi, B.Bucher, Cell Signal. 2015, 27, 1297.
[138] C.Palma, Curr. Drug Targets 2006, 7, 1043.
[139] R. G. Murthy, B. Y. Reddy, J. E. Ruggiero, P. Rameshwar, Front.
Biosci. 2007, 12, 4779.
[140] H.Javid, F. Mohammadi, E.Zahiri, S. I.Hashemy, J. Physiol. Bio-
chem. 2019, 75, 415.
[141] L.Quartara, C. AMaggi, Neuropeptides 1997, 31, 537.
[142] M.Bigioni, A.Benzo, C.Irrissuto, C. A.Maggi, C.Goso, Anticancer
Drugs 2005, 16, 1083.
[143] L.Debeljuk, Peptides 2006, 27, 736.
[144] F. Esteban, P. Ramos-García, M. Muñoz, M. Á. González-Moles,
Int. J. Environ. Res. Public Health 2021, 19, 375.
[145] M.Słoniecka, P.Danielson, J. Mol. Med. 2019, 97, 1477.
[146] M. Muñoz, A. González-Ortega, R. Coveñas, Invest New Drugs
2012, 30, 529.
[147] D. Singh, D. D. Joshi, M. Hameed, J. Qian, P. Gascón,
P. B. Maloof, A. Mosenthal, P. Rameshwar, Proc. Natl. Acad. Sci.
USA 2000, 97, 388.
[148] S.Brener, M. A.González-Moles, D. Tostes, F. Esteban, J. A. Gil-
Montoya, I.Ruiz-Ávila, M.Bravo, M.Muñoz, Anticancer Res. 2009,
29, 2323.
[149] M. Muñoz, A. Pavón, M. Rosso, M. V. Salinas, A. Pérez,
A.Carranza, A.González-Ortega, Placenta 2010, 31, 649.
[150] M.Muñoz, A.González-Ortega, M. V.Salinas-Martín, A.Carranza,
S.Garcia-Recio, V. Almendro, R. Coveñas, Int. J. Oncol. 2014, 45,
1658.
[151] I. M. Hennig, J. A. Laissue, U. Horisberger, J.-C. Reubi, Int. J.
Cancer 1995, 61, 786.
[152] P. Rameshwar, D. D. Joshi, P. Yadav, J. Qian, P. Gascon,
V. T.Chang, D. Anjaria, J. S. Harrison, X. Song, Blood 2001, 97,
3025.
[153] S.Garcia-Recio, G.Fuster, P.Fernandez-Nogueira, E. M.Pastor-Arroyo,
S. YPark, C.Mayordomo, E.Ametller, M.Mancino, X.Gonzalez-Farre,
H. G.Russnes, P.Engel, D.Costamagna, P. L.Fernandez, P.Gascón,
V.Almendro, Cancer Res. 2013, 73, 6424.
[154] M.Munoz, M.Rosso, R.Covenas, Curr. Drug Targets 2011, 12, 909.
[155] M.Munoz, M.Rosso, R.Covenas, Curr. Med. Chem. 2010, 17, 504.
[156] Y.Komiya, R.Habas, Organogenesis 2008, 4, 68.
[157] S.Patel, A.Alam, R.Pant, S.Chattopadhyay, Front. Immunol. 2019,
10, 2872.
[158] A. Garnier, J. Vykoukal, J. Hubertus, E. Alt, D. Von Schweinitz,
R.Kappler, M.Berger, M.Ilmer, Int. J. Oncol. 2015, 47, 151.
[159] X.-L.Niu, J.-F.Hou, J.-X.Li, Biol. Res. 2018, 51, 14.
[160] A.Plotnikov, E.Zehorai, S. Procaccia, R. Seger, Biochim. Biophys.
Acta 2011, 1813, 1619.
[161] P. P.Roux, J.Blenis, Microbiol. Mol. Biol. Rev. 2004, 68, 320.
[162] Y. J.Guo, W. W.Pan, S.-B.Liu, Z.-F.Shen, Y.Xu, L.-L.Hu, Exp. Ther.
Med. 2020, 19, 1997.
[163] B. L. Fiebich, S. Schleicher, R. D. Butcher, A. Craig, K. Lieb, J.
Immunol. 2000, 165, 5606.
[164] W.Luo, T. R.Sharif, M.Sharif, Cancer Res. 1996, 56, 4983.
[165] K.Yamaguchi, T.Kugimiya, T.Miyazaki, Brain Tumor Pathol. 2005,
22, 1.
[166] S.Singh, S.Kumaravel, S.Dhole, S.Roy, V.Pavan, S.Chakraborty,
Indian J. Surg. Oncol. 2021, 12, 93.
[167] J.Li, Q.Zeng, Y.Zhang, X.Li, H.Hu, X.Miao, W.Yang, W.Zhang,
X.Song, L.Mou, R.Wang, Eur. J. Cell Biol. 2016, 95, 368.
[168] D. Sood, M. Tang-Schomer, D. Pouli, C. Mizzoni, N. Raia,
A. Tai, K.Arkun, J. Wu, L. D. Black, B. Scheer, I.Georgakoudi,
D. A.Steindler, D. L.Kaplan, Nat. Commun. 2019, 10, 4529.
[169] A. G.Abraham, E.O’neill, Biochem. Soc. Trans. 2014, 42, 798.
[170] J. E.Lim, E.Chung, Y.Son, Sci. Rep. 2017, 7, 9417.
[171] H. Javid, J. Asadi, F. Zahedi Avval, A. R. Afshari, S. I. Hashemy,
Mol. Biol. Rep. 2020, 47, 2253.
[172] Z.Zou, T.Tao, H.Li, X.Zhu, Cell Biosci. 2020, 10, 31.
[173] C. Mayordomo, S. García-Recio, E. Ametller, P. Fernández-
Nogueira, E. M. Pastor-Arroyo, L. Vinyals, I. Casas, P. Gascón,
V.Almendro, J. Cell. Physiol. 2012, 227, 1358.
[174] R. E.Carraway, A. M.Plona, Peptides 2006, 27, 2445.
[175] A.Perron, N. Sharif, P.Sarret, T.Stroh, A.Beaudet, Biochem. Bio-
phys. Res. Commun. 2007, 353, 582.
[176] S. Martin, V. Navarro, J. P. Vincent, J. Mazella, Gastroenterology
2002, 123, 1135.
[177] R. M.Myers, J. W.Shearman, M. O.Kitching, A.Ramos-Montoya,
D. E.Neal, S. V.Ley, ACS Chem. Biol. 2009, 4, 503.
[178] I. Sehgal, S. Powers, B. Huntley, G. Powis, M. Pittelkow,
N. J.Maihle, Proc. Natl. Acad. Sci. USA 1994, 91, 4673.
[179] L. Seethalakshmi, S. P. Mitra, P. R. Dobner, M. Menon,
R. E.Carraway, Prostate 1997, 31, 183.
[180] R. M Schimp, A. M. Lhiaubet, M. Vial, W. Rostene, Prog.
Neuroendocrinimmunol. 1992, 5, 126.
[181] S. Saada, P. Marget, A.-L. Fauchais, M.-C. Lise, G. Chemin,
P. Sindou, C. Martel, L.Delpy, E. Vidal, A. Jaccard, D. Troutaud,
F.Lalloué, M.-O.Jauberteau, J. Immunol. 2012, 189, 5293.
[182] T. P. Davis, H. S. Burgess, S. Crowell, T. W. Moody,
A.Culling-Berglund, R. H.Liu, Eur. J. Pharmacol. 1989, 161, 283.
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (19 of 20)
www.advanced-bio.com
[183] R. R. Giorgi, T.Chile, A. R. Bello, R. Reyes, M. A. H. Z.Fortes,
M. C.Machado, V. A. Cescato, N. R.Musolino, M. D.Bronstein,
D. Giannella-Neto, M. L. Corrêa-Giannella, J. Neuroendocrinol.
2008, 20, 1052.
[184] T. W. Moody, J. Chiles, M. Casibang, E. Moody, D. Chan,
T. P.Davis, Peptides 2001, 22, 109.
[185] M. Alifano, M. Loi, S.Camilleri-Broet, S. Dupouy, J. F. Régnard,
P.Forgez, Biochimie 2010, 92, 164.
[186] J. J.Maoret, Y.Anini, C.Rouyer-Fessard, D.Gully, M.Laburthe, Int.
J. Cancer 1999, 80, 448.
[187] K. Yoshinaga, B. M. Evers, M. Izukura, D. Parekh, T. Uchida,
C. M.Townsend, J. C.Thompson, Surg. Oncol. 1992, 1, 127.
[188] T.Mijatovic, P.Gailly, V. Mathieu, N. De Néve, P.Yeaton, R.Kiss,
C.Decaestecker, Cell. Oncol. 2007, 29, 315.
[189] K. Takahashi, S. Ehata, K. Miyauchi, Y. Morishita, K.Miyazawa,
K.Miyazono, Mol. Oncol. 2021, 15, 151.
[190] I.Castagliuolo, C.-C.Wang, L.Valenick, A.Pasha, S.Nikulasson,
R. E.Carraway, C.Pothoulakis, J. Clin. Invest. 1999, 103, 843.
[191] S.Kawanishi, S.Ohnishi, N.Ma, Y.Hiraku, M.Murata, Int. J. Mol.
Sci. 2017, 18, 1808.
[192] Q.Ouyang, JiZhou, W.Yang, H.Cui, M.Xu, L. Yi, Clin. Exp. Phar-
macol. Physiol. 2017, 44, 841.
[193] U.Olszewski, G.Hamilton, Mol. Oncol. 2009, 3, 204.
[194] C.Wang, Q.Wang, B.Ji, Y.Pan, C.Xu, B. Cheng, BoBai, J.Chen,
Front. Mol. Neurosci. 2018, 11, 220.
[195] T.Voisin, A.El Firar, M.Fasseu, C.Rouyer-Fessard, V.Descatoire,
F. Walker, V.Paradis, P.Bedossa, D.Henin, T. Lehy, M.Laburthe,
Cancer Res. 2011, 71, 3341.
[196] P. Rouet-Benzineb, C. Rouyer-Fessard, A. Jarry, V. Avondo,
C.Pouzet, M.Yanagisawa, C. Laboisse, M. Laburthe, T.Voisin, J.
Biol. Chem. 2004, 279, 45875.
[197] J.Wen, Y.Zhao, L.Guo, Int. J. Mol. Med. 2016, 37, 126.
[198] J.Wen, Y.Zhao, Y.Shen, L.Guo, Mol. Med. Rep. 2015, 11, 3439.
[199] Y.Liu, Y.Zhao, S.Ju, L.Guo, Int. J. Mol. Med. 2015, 35, 539.
[200] R. Spinazzi, M. Rucinski, G. Neri, L. K. Malendowicz,
G. G.Nussdorfer, J. Clin. Endocrinol. Metab. 2005, 90, 3544.
[201] X.Chang, Y.Zhao, S.Ju, L.Guo, Int. J. Mol. Med. 2014, 34, 1523.
[202] D. Alexandre, C. Hautot, M. Mehio, L. Jeandel, M. Courel,
T.Voisin, A.Couvineau, F.Gobet, J.Leprince, C.Pfister, Y.Anouar,
N.Chartrel, Eur. J. Cancer 2014, 50, 2126.
[203] S. Valiante, G. Liguori, S. Tafuri, L. M. Pavone, R. Campese,
R. Monaco, G. Iachetta, L. Assisi, N. Mirabella, M. Forte,
A.Costagliola, A.Vittoria, Biochem. Biophys. Res. Commun. 2015,
464, 1290.
[204] M. M.Basar, U.Han, M.Cakan, S.Alpcan, H.Basar, Turk. J. Urol.
2013, 39, 78.
[205] N. L Graybill, V. Weissig, SAGE Open Med. 2017, 5,
205031211773577.
[206] E. S.Greene, M.Zampiga, F.Sirri, T.Ohkubo, S. Dridi, Sci. Rep.
2020, 10, 19191.
[207] C.Alain, N.Pascal, G.Valérie, V.Thierry, Molecules 2021, 26, 4849.
[208] S.Valiante, G.Liguori, S.Tafuri, R.Campese, R.Monaco, S.Paino,
V.Laforgia, N.Staiano, A.Vittoria, J. Anat. 2013, 222, 473.
[209] J. P.Kukkonen, Curr. Top. Behav. Neurosci. 2017, 33, 17.
[210] T. Voisin, A. El Firar, C. Rouyer-Fessard, V. Gratio, M.Laburthe,
FASEB J. 2008, 22, 1993.
[211] L. E.Heasley, Oncogene 2001, 20, 1563.
[212] S.Froidevaux, A. N.Eberle, Biopolymers 2002, 66, 161.
[213] G. Weckbecker, I. Lewis, R. Albert, H. A. Schmid, D. Hoyer,
C.Bruns, Nat. Rev. Drug Discovery 2003, 2, 999.
[214] S. Pyronnet, C. Bousquet, S. Najib, R. Azar, H. Laklai, C. Susini,
Mol. Cell. Endocrinol. 2008, 286, 230.
[215] R. Eychenne, C. Bouvry, M. Bourgeois, P. Loyer, E. Benoist,
N.Lepareur, Molecules 2020, 25, 4012.
[216] J.Fahrenkrug, Results Probl. Cell Dier. 2010, 50, 221.
[217] G.Deng, L.Jin, Neurol. Res. 2017, 39, 65.
[218] J. C.Reubi, U.Läderach, B.Waser, J. O.Gebbers, P. Robberecht,
J. A.Laissue, Cancer Res. 2000, 60, 3105.
[219] J. C.Reubi, M.Körner, B.Waser, L.Mazzucchelli, L.Guillou, Eur. J.
Nucl. Med. Mol. Imaging 2004, 31, 803.
[220] B.Tang, X. Yong, R.Xie, Q.-W.Li, S.-M. Yang, Int. J. Oncol. 2014,
44, 1023.
[221] T. W.Moody, B.Nuche-Berenguer, R. T.Jensen, Curr. Opin. Endo-
crinol. Diabetes Obes. 2016, 23, 38.
[222] A. G. D’amico, S. Scuderi, S. Saccone, A. Castorina, F. Drago,
V.D’agata, J. Mol. Neurosci. 2013, 51, 503.
[223] E. Vacas, A. B. Fernández-Martínez, A. M. Bajo, M. Sánchez-
Chapado, A. V. Schally, J. C.Prieto, M. J.Carmena, Biochim. Bio-
phys. Acta 2012, 1823, 1676.
[224] J.Wojcieszak, J. B.Zawilska, J. Mol. Neurosci. 2014, 54, 463.
[225] K.Maruno, A. Absood, S. I.Said, Proc. Natl. Acad. Sci. USA 1998,
95, 14373.
[226] K.Maruno, S. I.Said, Ann. N. Y. Acad. Sci. 1996, 805, 389;.
[227] Y.Ye, P.Liu, Y.Wang, H.Li, F.Wei, Y.Cheng, L.Han, J.Yu, Int Rev.
Immunol. 2016, 35, 340.
[228] Z.Wu, D.Martinez-Fong, J.Tredaniel, P.Forgez, Front. Endocrinol.
2012, 3, 184.
[229] J. P.Kukkonen, P. M.Turunen, Front. Neurol. Neurosci. 2021, 45, 91.
[230] A. Couvineau, S. Dayot, P. Nicole, V. Gratio, V. Rebours,
A.Couvelard, T.Voisin, Front. Endocrinol. 2018, 9, 573.
[231] J. P.Kukkonen, C. S.Leonard, Br. J. Pharmacol. 2014, 171, 314.
[232] P. Msaouel, E. Galanis, M. Koutsilieris, Expert Opin. Invest. Drugs
2009, 18, 1297.
[233] J. C.Reubi, J.Laissue, E.Krenning, S. W. J.Lamberts, J. Steroid Bio-
chem. Mol. Biol. 1992, 43, 27.
[234] M. J.Klomp, S. U.Dalm, M.De Jong, R. A. Feelders, J. Hofland,
L. J.Hofland, Rev. Endocr. Metab. Disord. 2021, 22, 495.
[235] T. W.Moody, J. M.Hill, R. T.Jensen, Peptides 2003, 24, 163.
[236] I.Langer, J.Jeandriens, A.Couvineau, S.Sanmukh, D. Latek, Bio-
medicines 2022, 10, 406.
[237] C. B.Saper, B. B.Lowell, Curr. Biol. 2014, 24, R1111.
[238] S. WRanson, H. W.Magoun, Biochem. Pharmacol. 1939, 41, 56.
[239] I.Izquierdo, R. D.Dias, Behav. Neural Biol. 1983, 38, 144.
[240] C. Nyakas, in Hormones and Brain Function (Ed: K. Lissák),
Springer US, Boston, MA 1973, pp. 83–89.
[241] P. E.Gold, R. L.Delanoy, in Endogenous Peptides and Learning and
Memory Processes (Ed: J. L. Martinez), Academic Press, New York
1981, pp. 79–98.
[242] C. Heesen, S. M. Gold, A. Raji, K. Wiedemann, K. H. Schulz,
Psychoneuroendocrinology 2002, 27, 505.
[243] L. Y.Jan, Y. N.Jan, J. Physiol. 1982, 327, 219.
[244] M. M.Mashaly, J. M.Trout, G. L.Hendricks, 3rd, Poult. Sci. 1993,
72, 1289.
[245] G.Csaba, Orv. Hetil. 2011, 152, 777.
[246] G.Seiden, A.Brodish, Neuroendocrinology 1971, 8, 154.
[247] C. H. Sawyer, M. Kawakami, B. Meyerson, D. I. Whitmoyer,
J. J.Lilley, Brain Res. 1968, 10, 213.
[248] G. W. Bennett, T. M. Ballard, C. D. Watson, K. C. F. Fone, Exp.
Gerontol. 1997, 32, 451.
[249] D.De Wied, Am. J. Physiol. 1964, 207, 255.
[250] D.De Wied, Proc. Soc. Exp. Biol. Med. 1966, 122, 28.
[251] A. J.Dunn, W.Hendrik Gispen, Neurosci. Biobehav. Rev. 1977, 1, 15.
[252] G. F. Koob, in Perspectives on Behavioral Medicine
(Ed: R. B. Williams), Academic Press, New York 1985, pp. 39–52.
[253] M.Irwin, R.Hauger, M.Brown, Endocrinology 1992, 131, 1047.
[254] E. M.Friedman, M. R.Irwin, Ann. N. Y. Acad. Sci. 1995, 771, 396.
[255] M. R.Brown, L. A.Fisher, Brain Res. 1983, 280, 75.
[256] P. L.Rodriguez, S. X.Jiang, Y. G.Fu, S.Avraham, H. K.Avraham,
Int. J. Cancer 2014, 134, 1034.
[257] P. M.Gross, Prog. Brain Res. 1992, 91, 219.
Adv. Biology 2022, 2200111
www.advancedsciencenews.com
© 2022 Wiley-VCH GmbH
2200111 (20 of 20)
www.advanced-bio.com
[258] W. F.Ganong, Clin. Exp. Pharmacol. Physiol. 2000, 27, 422.
[259] H. R. Wardill, K. A. Mander, Y. Z. A. Van Sebille, R. J. Gibson,
R. M.Logan, J. M.Bowen, S. T.Sonis, Int. J. Cancer 2016, 139, 2635.
[260] M. Mampay, M. S.Flint, G. K. Sheridan, Br. J. Pharmacol. 2021,
178, 3977.
[261] M. S. Kramer, N. Cutler, J. Feighner, R. Shrivastava, J. Carman,
J. J. Sramek, S. A. Reines, G.Liu, D. Snavely, E. Wyatt-Knowles,
J. J. Hale, S. G. Mills, M. Maccoss, C. J. Swain, T. Harrison,
R. G.Hill, F. Hefti, E. M. Scolnick, M. A. Cascieri, G. G.Chicchi,
S. Sadowski, A. R. Williams, L. Hewson, D. Smith, E. J. Carlson,
R. J.Hargreaves, N. M. J.Rupniak, Science 1998, 281, 1640.
[262] B.Bondy, T. C.Baghai, C.Minov, C.Schüle, M. J.Schwarz, P.Zwanzger,
R.Rupprecht, H.-J. ÜMöller, Biol. Psychiatry 2003, 53, 538.
[263] K.Ebner, N. M.Rupniak, A.Saria, N.Singewald, Proc. Natl. Acad.
Sci. USA 2004, 101, 4280.
[264] S. L. Swift, J. E. Burns, N. J. Maitland, Cancer Res. 2010, 70,
347.
[265] F. Souazé, V. Viardot-Foucault, N. Roullet, M. Toy-Miou-Leong,
A. Gompel, E. Bruyneel, E. Comperat, M. C. Faux, M. Mareel,
W. Rostène, J.-F. Fléjou, C. Gespach, P. Forgez, Carcinogenesis
2006, 27, 708.
[266] K.Galoian, P.Patel, Biomed. Rep. 2017, 6, 3.
[267] Z.Dong, X.Wang, Q.Zhao, C. M.Townsend Jr., B. M.Evers, Am.
J. Physiol. 1998, 274, G535.
[268] K.Jackson, M.Soutto, D.Peng, T.Hu, D.Marshal, W.El-Rifai, Dig.
Dis. Sci. 2011, 56, 125.
[269] J.-H.Lee, J.-Y. Lee, S. B. Rho, J.-S. Choi, D.-G. Lee, S. An, T.Oh,
D.-C.Choi, S.-H.Lee, FEBS Lett. 2014, 588, 4730.
[270] M. J. Veenstra, P. M. Van Koetsveld, F. Dogan, W. E. Farrell,
R. A. Feelders, S. W. J. Lamberts, W. W. De Herder, G. Vitale,
L. J.Hofland, Oncotarget 2018, 9, 14791.
[271] H.Jo, Y.Park, J. Kim, H.Kwon, T.Kim, J. Lee, J.-C. Pyun, M.Lee,
M.Yun, PLoS One 2020, 15, e0240107.
[272] V. Sanchez Freire, F. C. Burkhard, T. M. Kessler, A. Kuhn,
A.Draeger, K.Monastyrskaya, Am. J. Pathol. 2010, 176, 288.
[273] D. Wu, Ye Qiu, Y. Jiao, Z. Qiu, Da Liu, Front. Oncol. 2020, 10,
560487.
[274] T. W.Moody, Ann. N. Y. Acad. Sci. 2019, 1455, 141.
[275] S. P.Brothers, C.Wahlestedt, EMBO Mol. Med. 2010, 2, 429.
[276] M.Muñoz, R.Coveñas, F.Esteban, M.Redondo, J. Biosci. 2015, 40,
441.
[277] O.Keskin, S.Yalcin, OncoTargets Ther. 2013, 6, 471.
[278] I.Gozes, S.Furman, Curr. Pharm. Des. 2003, 9, 483.
[279] T. W. Moody, G. Czerwinski, N. I. Tarasova, D. L. Moody,
C. J.Michejda, Regul. Pept. 2004, 123, 187.
[280] J. C.Reubi, Endocr. Rev. 2003, 24, 389.
[281] C. Rangger, A. Helbok, M. Ocak, T. Radolf, F. Andreae,
I. J.Virgolini, E. Von Guggenberg, C.Decristoforo, Anticancer Res.
2013, 33, 1537.
[282] K.Delaney, A. R.Buchberger, L.Atkinson, S.Gründer, A.Mousley,
L.Li, J. Exp. Biol. 2018, 221, 151167.
[283] Z.Wu, D.Lin, Y.Li, Nat. Rev. Neurosci. 2022, 23, 257.
[284] M. R. Schiller, A. B. Kohn, L. M. Mende-Mueller, K. Miller,
V. Y. H.Hook, FEBS Lett. 1996, 382, 6.
[285] P. T. FWilliamson, J. F Roth, T. Haddingham, A. Watts, Protein
Expression Purif. 2000, 19, 271.
[286] A. S. Hauser, M. M.Attwood, M. Rask-Andersen, H. B.Schiöth,
D. E.Gloriam, Nat. Rev. Drug Discovery 2017, 16, 829.
[287] B. Alvarez-Castelao, C. T. Schanzenbächer, C. Hanus, C. Glock,
S. Tom Dieck, A. R. Dörrbaum, I. Bartnik, B. Nassim-Assir,
E.Ciirdaeva, A.Mueller, D. C.Dieterich, D. A.Tirrell, J. D.Langer,
E. M.Schuman, Nat. Biotechnol. 2017, 35, 1196.
Y. Wu received her Master’s degree in Pharmaceutical Chemistry from Sun yat-sen University,
China and completed her Ph.D. degree in Synthetic Biology from the University of Queensland,
Australia in 2020. She is currently a postdoctoral fellow at Cold Spring Harbor Laboratory, USA.
Her research interests include exploring interactions between the peripheral nervous system
and cancer, metabolic labeling, detection, and identification of nascent proteome changes in the
nervous system during cancer progression.
A. Berisha was born in Long Island, New York, and is an identical twin and aspiring physician.
Adrian graduated from Stony Brook University with a B.S. in Biology with Honors. He participated
in undergraduate research at the Division of Nephrology and Hypertension at Renaissance School
of Medicine. Adrian is currently employed as a Research Technician in Jeremy Borniger’s lab at
Cold Spring Harbor Laboratory. His primary focus is to investigate the contribution of brainstem
nuclei and the sympathetic nervous system in promoting tumor progression in preclinical models
and identify the neural pathways innervating peripheral tumors.
Adv. Biology 2022, 2200111
... The overexpression of the same receptor in different types of tumors allows for similar antitumor strategies by, for example, administering a peptide receptor antagonist alone or combined with chemotherapy/radiotherapy. The high expression of peptide receptors in tumor cells could also be helpful for cancer diagnosis (tumor biomarker) and treatment after implementing antitumor strategies such as the use of peptide receptor antagonists [2], cytotoxic peptide conjugate-based cancer therapy (e.g., vasoactive intestinal peptide-ellipticine conjugate), or peptide receptor radionuclide therapy (PRRT) [9,10]. For example, the PRRT clinical applications targeting the somatostatin receptor have been reviewed by the group of Christophe M. Deroose from the University Hospitals Leuven (Leuven, Belgium) [11]. ...
... It has been shown that peptides also display an antitumor effect, for example, by blocking cell proliferation and angiogenesis [5,9]. The use of glucagon-like peptide-1 receptor agonists (GLP-1 RAs) to treat hepatocellular carcinoma (HCC) has been reviewed by the group of Georgios Germanidis from the Aristotle University of Thessaloniki (Thessaloniki, Greece) [13]. ...
... For these reasons, the authors suggested obtaining HIF peptide inhibitors that share some of the advantages of HIFs as small chemical inhibitors, such as permeability, pharmacokinetics, oral delivery, and half-life [16]. Peptide cyclization, elimination of catalytic sites, co-administration of peptides with protease inhibitors, covalent modifications in the amino acid sequence, peptide-loaded nanoparticles, absorption enhancers, and the conjugation of peptide drugs to synthetic or natural polymers represent current strategies to enhance their therapeutic potential (i.e., cell-targeting peptides and cell-permeable peptides), stability, and delivery [9,10]. ...
Article
Full-text available
Undoubtedly, much progress has been made in treating cancer over the past few years, but unfortunately, 28 [...]
... Cancers 2023, 15, 1694 2 of 63 exert an antiapoptotic action, and stimulate the growth of blood vessels and lymphangiogenesis. However, some peptides such as neuropeptide Y, orexin, and vasoactive intestinal peptide also exert an anticancer effect [7]. Furthermore, tumor cells release peptides acting through autocrine, paracrine, and endocrine (tumor mass) mechanisms [1][2][3]5,6,[8][9][10][11]. ...
... Compared with healthy individuals, a rise in circulating NKA/SP has been reported in patients suffering from ileal metastatic carcinoid tumors showing cutaneous flushing; the release of both peptides from carcinoid tumors was partially blocked after the administration of a somatostatin analog [164]. The presence of NKA, NKA 3-10 , and NKA [4][5][6][7][8][9][10] has been detected in ileal metastatic carcinoid tumors [165]. Therefore, NKA shows an Nterminal heterogeneity in these tumors, and carcinoid tumors can release different amounts of several tachykinins contributing to individual differences. ...
... This is a crucial point because NKA exerted its physiological actions not exclusively via NK-2R but also via NK-1R. Moreover, it is essential to remark that NKA, NKA 3-10 , and NKA [4][5][6][7][8][9][10] have been reported in ileal metastatic carcinoid tumors; these tumors can release different amounts of several tachykinins contributing to individual differences [165]. Knowing the physiological actions mediated by the unmodified peptide and its fragments is an important point that must be studied in other tumors. ...
Article
Full-text available
Simple Summary Neurokinins A and B, adrenomedullin, adrenomedullin 2, amylin, and calcitonin gene-related peptide are essential in different tumors. These peptides are involved in tumor cell proliferation and migration, metastasis, angiogenesis, and lymphangiogenesis. Accordingly, several antitumor therapeutic strategies, including peptide receptor antagonists, can be developed. This review highlights the essential roles played by both tachykinin and calcitonin/calcitonin gene-related peptide families in cancer progression, which support the application of promising clinical antitumor therapeutic strategies. Abstract The roles played by the peptides belonging to the tachykinin (neurokinin A and B) and calcitonin/calcitonin gene-related peptide (adrenomedullin, adrenomedullin 2, amylin, and calcitonin gene-related peptide (CGRP)) peptide families in cancer development are reviewed. The structure and dynamics of the neurokinin (NK)-2, NK-3, and CGRP receptors are studied together with the intracellular signaling pathways in which they are involved. These peptides play an important role in many cancers, such as breast cancer, colorectal cancer, glioma, lung cancer, neuroblastoma, oral squamous cell carcinoma, phaeochromocytoma, leukemia, bladder cancer, endometrial cancer, Ewing sarcoma, gastric cancer, liver cancer, melanoma, osteosarcoma, ovarian cancer, pancreatic cancer, prostate cancer, renal carcinoma, and thyroid cancer. These peptides are involved in tumor cell proliferation, migration, metastasis, angiogenesis, and lymphangiogenesis. Several antitumor therapeutic strategies, including peptide receptor antagonists, are discussed. The main research lines to be developed in the future are mentioned.
... This is because peptides generally show poor bioavailability and a short half-life. However, it is essential to emphasize that some strategies are being developed to increase the therapeutic effects of peptides as well as their stability and delivery (e.g., peptide-loaded nanoparticles, peptide cyclization, conjugation of peptide drugs to natural/synthetic polymers, manipulation of the amino acid sequence, cell-targeting peptides, and cell-penetrating peptides) [5,6]. Some cell-penetrating short peptides show an antitumor action, and, in addition, they can also be used to carry anticancer cargo into tumor cells [7]. ...
... As indicated above, the non-peptide NK-1R antagonist aprepitant (MK-869, L-754,030, Emend) is used in clinical practice as an antiemetic (oral administration; it is safe and well tolerated in general) and exerts, in vitro and in vivo, a broad antitumor effect by inducing apoptosis in many human cancer cells and decreasing the tumor volume or cell count in many types of cancer, such as acute lymphoblastic/myeloid leukemia, breast, prostate, lung, ovarian and cervical cancers, chronic myeloid leukemia, colorectal, esophageal, larynx, pancreatic, urinary bladder and gastric carcinomas, glioblastoma multiforme, hepatoblastoma, melanoma, neuroblastoma, osteosarcoma, retinoblastoma, and rhabdoid tumors [57]. The IC 50 of aprepitant for normal cells (e.g., fibroblasts, lymphocytes, breast epithelial cells) is much higher than that for cancer cells [5,13], which validates the argument for the safety of aprepitant. Although most of the studies published have shown a higher antiproliferative effect of aprepitant against tumor cells than against normal cells, a study has reported that, although the drug exerted an antitumor action, this action was not selective, equally affecting normal and tumor cells [58]. ...
Article
Full-text available
Peptidergic systems show promise as targets for fighting tumors. While some peptides encourage the growth and spread of tumor cells and angiogenic mechanisms, others display antitumor properties. As such, peptide ligands and receptor antagonists could be used as antitumor agents alone or in conjunction with chemotherapy or radiotherapy. Peptide receptor antagonists can counteract the oncogenic effects of specific peptides by inducing apoptosis in various types of tumor cells, hindering cancer cell migration and inhibiting angiogenesis. Peptides and peptide receptor antagonists are not currently used in clinical practice as antitumor agents. Still, aprepitant, a neurokinin 1 receptor antagonist, is a promising candidate due to its ability to promote apoptosis in many cancer cells. However, to utilize aprepitant as an anticancer agent, the dosage must be increased and administered for a more extended period. Moving beyond current protocols for aprepitant’s use as an antiemetic is essential. Additionally, a common anticancer strategy with aprepitant is possible regardless of cancer cell type. Finally, combining aprepitant with chemotherapy or radiotherapy is encouraged.
... Neuropeptide Y (NPY) and its precursor ProNPY are mainly involved in food intake, blood pressure regulation, and energy homeostasis, by stimulating five G proteincoupled receptors (Y1 to Y6) [2]. Effects progression have also been highlighted, with NPY and related peptides being involved in cell proliferation, matrix invasion, and angiogenesis [3]. In the context of PCa, these peptides have been proposed as promising biomarkers to improve PCa diagnosis and prognosis [4,5]. ...
Article
Full-text available
Neuropeptide Y (NPY) and related peptides have been proposed as promising biomarkers for the diagnosis of prostate cancer by previous immunoassays and immunohistochemical studies. In this study, we evaluated the additional value of NPY and related peptides compared with prostate-specific antigen (PSA). We performed a comprehensive analysis of NPY, its precursors, and metabolite concentrations in both plasma and tissue samples from 181 patients using a highly specific liquid chromatography tandem mass spectrometry method. Compared with PSA, NPY and related peptides (NPYs) were less accurate at diagnosing significant prostate cancer. Combinations of NPYs in a stepwise approach did not improve a model that would be beneficial for patients. NPY may be beneficial for patients presenting with a PSA concentration in the gray area between 4 and 9 ng/ml, but the strength of this conclusion is limited. Thus, the use of NPYs as standalone or in combination with other variables, such as PSA, prostate volume, or age, to improve the diagnosis is not supported by our study.
... Although by themselves, these factors may not be critical for cancer cell survival, they were shown to facilitate tumor growth in the context of chemotherapy. Given that neurotransmitters and neuropeptides can activate signaling pathways pertaining to cell proliferation and survival, such as the PI3K, MAPK and Akt pathways, their inhibition should constitute a relevant chemo-or targeted therapy sensitization strategy (132). Table 1 lists some of the reported effect of factors released by nerves within tumors and their reported effects on tumorigenesis. ...
Article
Full-text available
Emerging evidence suggests that nerves within the tumor microenvironment play a crucial role in regulating angiogenesis. Neurotransmitters and neuropeptides released by nerves can interact with nearby blood vessels and tumor cells, influencing their behavior and modulating the angiogenic response. Moreover, nerve-derived signals may activate signaling pathways that enhance the production of pro-angiogenic factors within the tumor microenvironment, further supporting blood vessel growth around tumors. The intricate network of communication between neural constituents and the vascular system accentuates the potential of therapeutically targeting neural-mediated pathways as an innovative strategy to modulate tumor angiogenesis and, consequently, neoplastic proliferation. Hereby, we review studies that evaluate the precise molecular interplay and the potential clinical ramifications of manipulating neural elements for the purpose of anti-angiogenic therapeutics within the scope of cancer treatment.
Article
Full-text available
The substance P (SP)/neurokinin-1 receptor (NK-1R) system is involved in cancer progression. NK-1R, activated by SP, promotes tumor cell proliferation and migration, angiogenesis, the Warburg effect, and the prevention of apoptosis. Tumor cells overexpress NK-1R, which influences their viability. A typical specific anticancer strategy using NK-1R antagonists, irrespective of the tumor type, is possible because these antagonists block all the effects mentioned above mediated by SP on cancer cells. This review will update the information regarding using NK-1R antagonists, particularly Aprepitant, as an anticancer drug. Aprepitant shows a broad-spectrum anticancer effect against many tumor types. Aprepitant alone or in combination therapy with radiotherapy or chemotherapy could reduce the sequelae and increase the cure rate and quality of life of patients with cancer. Current data open the door to new cancer research aimed at antitumor therapeutic strategies using Aprepitant. To achieve this goal, reprofiling the antiemetic Aprepitant as an anticancer drug is urgently needed.
Chapter
Full-text available
Nervous systems were not classically considered to be actively involved in the process of tumorigenesis and cancer metastasis. However, studies over the last decade have demonstrated the presence of neurons and glial cells in the peritumoral regions of many human tumors, and the density of tumor-associated nerves is often correlated with cancer progression and metastatic spread. In general, it has been demonstrated that neuronal activity is widely pro-cancer in the brain, and blocking neural activity usually confers anticancer benefits [1]. For example, increased neuronal excitability has been observed in preclinical models of both pediatric and adult gliomas [2,3,4], and nerve ablation can suppress tumor development in various other malignancies [5]. With the growing global cancer burden, research in the nascent field of cancer neuroscience is quickly becoming intense, attracting interdisciplinary efforts from all over the world. This chapter will focus on the reciprocal cross talk between cancer and the nervous system via direct and indirect pathways. Subsequent influences on host behavior and cancer therapeutic resistance will also be discussed.KeywordsCognitiveBehaviorNervous systemTumorigenesisCross talk
Chapter
The chapters in this book summarize the state of the field as we find it in 2023. The assignment of this last chapter is to look into the future, attempt to guess where cancer neuroscience will go, and suggest the steps that might take the field into the next decade. As shown in this volume, this area of investigation is exploding, making this task difficult. Thus, being predictive is a perilous venture as many of the obvious directions are currently being explored. So, instead of attempting (and failing) to divine the future, here we propose a potential way to reframe the question(s) of nerve-cancer interactions in a manner that might indicate avenues not currently being explored. The following discussion focuses on solid tumors that occur outside of the CNS, but we will note features that are common to CNS tumors where appropriate.KeywordsCancerPeripheral nervous systemTumor microenvironmentImmunosuppressionNeuroimmuneCheckpoint protein
Article
Neuropeptide FF (NPFF) belongs to the RFamide peptide family. NPFF regulates a variety of physiological functions by binding to a G protein-coupled receptor, NPFFR2. Epithelial ovarian cancer (EOC) is a leading cause of death among gynecological malignancies. The pathogenesis of EOC can be regulated by many local factors, including neuropeptides, through an autocrine/paracrine manner. However, to date, the expression and/or function of NPFF/NPFFR2 in EOC is undetermined. In this study, we show that the upregulation of NPFFR2 mRNA was associated with poor overall survival in EOC. The TaqMan probe-based RT-qPCR showed that NPFF and NPFFR2 were expressed in three human EOC cells, CaOV3, OVCAR3, and SKOV3. In comparison, NPFF and NPFFR2 expression levels were higher in SKOV3 cells than in CaOV3 or OVCAR3 cells. Treatment of SKOV3 cells with NPFF did not affect cell viability and proliferation but stimulated cell invasion. NPFF treatment upregulates matrix metalloproteinase-9 (MMP-9) expression. Using the siRNA-mediated knockdown approach, we showed that the stimulatory effect of NPFF on MMP-9 expression was mediated by the NPFFR2. Our results also showed that ERK1/2 signaling was activated in SKOV3 cells in response to the NPFF treatment. In addition, blocking the activation of ERK1/2 signaling abolished the NPFF-induced MMP-9 expression and cell invasion. This study provides evidence that NPFF stimulates EOC cell invasion by upregulating MMP-9 expression through the NPFFR2-mediated ERK1/2 signaling pathway.
Article
Full-text available
Homeostasis of the human immune system is regulated by many cellular components, including two neuropeptides, VIP and PACAP, primary stimuli for three class B G protein-coupled receptors, VPAC1, VPAC2, and PAC1. Vasoactive intestinal peptide (VIP) and pituitary adenylate cyclase-activating polypeptide (PACAP) regulate intestinal motility and secretion and influence the functioning of the endocrine and immune systems. Inhibition of VIP and PACAP receptors is an emerging concept for new pharmacotherapies for chronic inflammation and cancer, while activation of their receptors provides neuroprotection. A small number of known active compounds for these receptors still impose limitations on their use in therapeutics. Recent cryo-EM structures of VPAC1 and PAC1 receptors in their agonist-bound active state have provided insights regarding their mechanism of activation. Here, we describe major molecular switches of VPAC1, VPAC2, and PAC1 that may act as triggers for receptor activation and compare them with similar non-covalent interactions changing upon activation that were observed for other GPCRs. Interhelical interactions in VIP and PACAP receptors that are important for agonist binding and/or activation provide a molecular basis for the design of novel selective drugs demonstrating anti-inflammatory, anti-cancer, and neuroprotective effects. The impact of genetic variants of VIP, PACAP, and their receptors on signalling mediated by endogenous agonists is also described. This sequence diversity resulting from gene splicing has a significant impact on agonist selectivity and potency as well as on the signalling properties of VIP and PACAP receptors.
Article
Full-text available
Head and neck cancer is a growing worldwide public health problem, accounting for approximately 1,500,000 new cases and 500,000 deaths annually. Substance P (SP) is a peptide of the tachykinin family, which has roles related to a large number of physiological mechanisms in humans. The implications of SP in carcinogenesis have recently been reported through the stimulation of the neurokinin 1 receptor (NK1R), or directly, through the effects derived from the constitutive activation of NK1R. Consequently, SP/NK1R seems to play relevant roles in cancer, upregulating cell proliferation, cell migration and chronic inflammation, among other oncogenic actions. Furthermore, there is growing evidence pointing to a central role for SP in tumour progression, singularly so in laryngeal and oral squamous cell carcinomas. The current narrative review of the literature focuses on the relationship between the SP/NK1R system and chronic inflammation and cancer in the head-and-neck region. We described a role for SP/NK1R in the transition from chronic inflammation of the head and neck mucosa, to preneoplastic and neoplastic transformation and progression.
Article
Full-text available
Over 20 years ago, orexin neuropeptides (Orexin-A/hypocretin-1 and Orexin-B/hypocretins-2) produced from the same precursor in hypothalamus were identified. These two neurotransmitters and their receptors (OX1R and OX1R), present in the central and peripheral nervous system, play a major role in wakefulness but also in drug addiction, food consumption, homeostasis, hormone secretion, reproductive function, lipolysis and blood pressure regulation. With respect to these biological functions, orexins were involved in various pathologies encompassing narcolepsy, neurodegenerative diseases, chronic inflammations, metabolic syndrome and cancers. The expression of OX1R in various cancers including colon, pancreas and prostate cancers associated with its ability to induce a proapoptotic activity in tumor cells, suggested that the orexins/OX1R system could have a promising therapeutic role. The present review summarizes the relationship between cancers and orexins/OX1R system as an emerging target.
Article
Full-text available
Neuropeptide Y (NPY), one of the most abundant neuropeptides in the body, is widely expressed in the central and peripheral nervous systems and acts on the cardiovascular, digestive, endocrine, and nervous systems. NPY affects the nutritional and inflammatory microenvironments through its interaction with immune cells, brain-derived trophic factor (BDNF), and angiogenesis promotion to maintain body homeostasis. Additionally, NPY has great potential for therapeutic applications against various diseases, especially as an adjuvant therapy for stem cells. In this review, we discuss the research progress regarding NPY, as well as the current evidence for the regulation of NPY in each microenvironment, and provide prospects for further research on related diseases.
Article
Full-text available
Orexin receptors (OXRs) are promiscuous G-protein-coupled receptors that signal via several G-proteins and, putatively, via other proteins. On which basis the signal pathways are selected and orchestrated is largely unknown. We also have an insufficient understanding of the kind of signaling that is important for specific types of cellular responses. OXRs are able to form complexes with several other G-protein-coupled receptors in vitro, and one possibility is that the complexing partners regulate the use of certain signal transducers. In the central nervous system neurons, the main acute downstream responses of OXR activation are the inhibition of K+ channels and the activation of the Na+/Ca2+ exchanger and non-selective cation channels of unknown identity. The exact nature of the intracellular signal chain between the OXRs and these downstream targets is yet to be elucidated, but the Gq-phospholipase C (PLC) protein kinase C pathway - which is a significant signaling pathway for OXRs in recombinant cells - may be one of the players in neurons. The Gq-PLC pathway may also, under certain circumstances, take the route to diacylglycerol lipase, which leads to the production of the potent endocannabinoid (eCB), 2-arachidonoyl glycerol, and thereby connects orexins with eCB signaling. In addition, OXRs have been studied in the context of neurodegeneration and cancer cell death. Overall, OXR signaling is complex, and it can change depending on the cell type and environment.
Article
Full-text available
People that develop extracranial cancers often display co‐morbid neurological disorders, such as anxiety, depression and cognitive impairment, even before commencement of chemotherapy. This suggests bidirectional crosstalk between non‐CNS tumours and the brain, which can regulate peripheral tumour growth. However, the reciprocal neurological effects of tumour progression on brain homeostasis are not well understood. Here, we review brain regions involved in regulating peripheral tumour development and how they, in turn, are adversely affected by advancing tumour burden. Tumour‐induced activation of the immune system, blood–brain barrier breakdown and chronic neuroinflammation can lead to circadian rhythm dysfunction, sleep disturbances, aberrant glucocorticoid production, decreased hippocampal neurogenesis and dysregulation of neural network activity, resulting in depression and memory impairments. Given that cancer‐related cognitive impairment diminishes patient quality of life, reduces adherence to chemotherapy and worsens cancer prognosis, it is essential that more research is focused at understanding how peripheral tumours affect brain homeostasis.
Article
Full-text available
Somatostatin (SST) and somatostatin receptors (SSTRs) play an important role in the brain and gastrointestinal (GI) system. SST is produced in various organs and cells, and the inhibitory function of somatostatin-containing cells is involved in a range of physiological functions and pathological modifications. The GI system is the largest endocrine organ for digestion and absorption, SST-endocrine cells and neurons in the GI system are a critical effecter to maintain homeostasis via SSTRs 1-5 and co-receptors, while SST-SSTRs are involved in chemo-sensory, mucus, and hormone secretion, motility, inflammation response, itch, and pain via the autocrine, paracrine, endocrine, and exoendocrine pathways. It is also a power inhibitor for tumor cell proliferation, severe inflammation, and post-operation complications, and is a first-line anti-cancer drug in clinical practice. This mini review focuses on the current function of producing SST endocrine cells and local neurons SST-SSTRs in the GI system, discusses new development prognostic markers, phosphate-specific antibodies, and molecular imaging emerging in diagnostics and therapy, and summarizes the mechanism of the SST family in basic research and clinical practice. Understanding of endocrines and neuroendocrines in SST-SSTRs in GI will provide an insight into advanced medicine in basic and clinical research.
Article
Full-text available
Neuropeptide Y (NPY) has been implicated in the regulation of cellular motility under various physiological and pathological conditions, including cancer dissemination. Yet, the exact signaling pathways leading to these effects remain unknown. In a pediatric malignancy, neuroblastoma (NB), high NPY release from tumor tissue associates with metastatic disease. Here, we have shown that NPY stimulates NB cell motility and invasiveness and acts as a chemotactic factor for NB cells. We have also identified the Y5 receptor (Y5R) as the main NPY receptor mediating these actions. In NB tissues and cell cultures, Y5R is highly expressed in migratory cells and accumulates in regions of high RhoA activity and dynamic cytoskeleton remodeling. Y5R stimulation activates RhoA and results in Y5R/RhoA-GTP interactions, as shown by pull-down and proximity ligation assays, respectively. This is the first demonstration of the role for the NPY/Y5R axis in RhoA activation and the subsequent cytoskeleton remodeling facilitating cell movement. These findings implicate Y5R as a target in anti-metastatic therapies for NB and other cancers expressing this receptor.
Article
Neurotransmitters and neuromodulators have a wide range of key roles throughout the nervous system. However, their dynamics in both health and disease have been challenging to assess, owing to the lack of in vivo tools to track them with high spatiotemporal resolution. Thus, developing a platform that enables minimally invasive, large-scale and long-term monitoring of neurotransmitters and neuromodulators with high sensitivity, high molecular specificity and high spatiotemporal resolution has been essential. Here, we review the methods available for monitoring the dynamics of neurotransmitters and neuromodulators. Following a brief summary of non-genetically encoded methods, we focus on recent developments in genetically encoded fluorescent indicators, highlighting how these novel indicators have facilitated advances in our understanding of the functional roles of neurotransmitters and neuromodulators in the nervous system. These studies present a promising outlook for the future development and use of tools to monitor neurotransmitters and neuromodulators.
Article
With increasing attention on the essential roles of the tumour microenvironment in recent years, the nervous system has emerged as a novel and crucial facilitator of cancer growth. In this Review, we describe the foundational, translational, and clinical advances illustrating how nerves contribute to tumour proliferation, stress adaptation, immunomodulation, metastasis, electrical hyperactivity and seizures, and neuropathic pain. Collectively, this expanding knowledge base reveals multiple therapeutic avenues for cancer neuroscience that warrant further exploration in clinical studies. We discuss the available clinical data, including ongoing trials investigating novel agents targeting the tumour–nerve axis, and the therapeutic potential for repurposing existing neuroactive drugs as an anti-cancer approach, particularly in combination with established treatment regimens. Lastly, we discuss the clinical challenges of these treatment strategies and highlight unanswered questions and future directions in the burgeoning field of cancer neuroscience.