Next Article in Journal
BrTCP7 Transcription Factor Is Associated with MeJA-Promoted Leaf Senescence by Activating the Expression of BrOPR3 and BrRCCR
Next Article in Special Issue
Calcium as a Key Player in Arrhythmogenic Cardiomyopathy: Adhesion Disorder or Intracellular Alteration?
Previous Article in Journal
Cell Interactions in Biliary Diseases: Clues from Pathophysiology and Repair Mechanisms to Foster Early Assessment
Previous Article in Special Issue
Endothelium-Dependent Hyperpolarization (EDH) in Diabetes: Mechanistic Insights and Therapeutic Implications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Endothelial Ca2+ Signaling, Angiogenesis and Vasculogenesis: Just What It Takes to Make a Blood Vessel

1
Laboratory of General Physiology, Department of Biology and Biotechnology “L. Spallanzani”, University of Pavia, 27100 Pavia, Italy
2
Research Centre, Salahaddin University-Erbil, Erbil 44001, Iraq
3
Department of Pathological Analysis, College of Science, Knowledge University, Erbil 074016, Iraq
4
Department of Medicine and Health Sciences “Vincenzo Tiberio”, University of Molise, 86100 Campobasso, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2019, 20(16), 3962; https://doi.org/10.3390/ijms20163962
Submission received: 22 July 2019 / Revised: 9 August 2019 / Accepted: 13 August 2019 / Published: 14 August 2019
(This article belongs to the Special Issue Calcium Signaling in Human Health and Diseases 2.0)

Abstract

:
It has long been known that endothelial Ca2+ signals drive angiogenesis by recruiting multiple Ca2+-sensitive decoders in response to pro-angiogenic cues, such as vascular endothelial growth factor, basic fibroblast growth factor, stromal derived factor-1α and angiopoietins. Recently, it was shown that intracellular Ca2+ signaling also drives vasculogenesis by stimulation proliferation, tube formation and neovessel formation in endothelial progenitor cells. Herein, we survey how growth factors, chemokines and angiogenic modulators use endothelial Ca2+ signaling to regulate angiogenesis and vasculogenesis. The endothelial Ca2+ response to pro-angiogenic cues may adopt different waveforms, ranging from Ca2+ transients or biphasic Ca2+ signals to repetitive Ca2+ oscillations, and is mainly driven by endogenous Ca2+ release through inositol-1,4,5-trisphosphate receptors and by store-operated Ca2+ entry through Orai1 channels. Lysosomal Ca2+ release through nicotinic acid adenine dinucleotide phosphate-gated two-pore channels is, however, emerging as a crucial pro-angiogenic pathway, which sustains intracellular Ca2+ mobilization. Understanding how endothelial Ca2+ signaling regulates angiogenesis and vasculogenesis could shed light on alternative strategies to induce therapeutic angiogenesis or interfere with the aberrant vascularization featuring cancer and intraocular disorders.

1. Introduction

The vasculature is a highly branched, tree-like, tubular network, which encompasses a system of hierarchically organized arteries, veins and interconnecting capillary beds that is optimized to provide all tissues with crucial nutrients and oxygen and remove their catabolic waste [1]. Endothelial cells line the interior surface of blood vessels, also known as tunica intima, and dictate vascular branching and morphogenesis, lumen formation, and vessel wall assembly [2]. The vasculature reaches into every organ of the vertebrate body, except the avascular cornea and the cartilage [1]. Therefore, vascular endothelium constitutes a systematically disseminated tissue that weights ≈ 1 kg in a normoweight adult, with a large proportion (>600 g) lining the capillary district, and covers a surface area of 4000–6000 m2. It has been estimated that, if lined end-to-end, the ≈10 trillion (1013) interconnected endothelial cells that form the lumen of blood vessels would wrap more than four times around the circumference of the earth [3,4]. Rather than being an inert barrier that merely regulates the exchange of solutes between circulating blood and surrounding tissues, the endothelial monolayer is critical to maintain cardiovascular homeostasis, by finely tuning a wealth of critical processes, including adjustment of vascular tone according to local energy requirements, hemostasis and coagulation, and inflammation [4,5]. In addition to providing the building blocks of the vascular transport network, endothelial cells deliver paracrine (also termed angiocrine) signals to induce growth, differentiation, or repair processes in the surrounding tissues and/or to modulate their functions [2,6]. Distinct mechanisms contribute to establish and maintain a functional vascular network in the developing embryo as well as in adult tissues [1]. Vascular development is initiated by mesoderm-derived endothelial progenitor cells (EPC), which migrate into the yolk sac and coalesce to form primitive vascular channels, a process termed vasculogenesis. Thereafter, angiogenic remodeling allows this primitive vascular labyrinth to expand into a hierarchical tree-like network of arteries, arterioles, capillary beds, venules and veins. Sprouting angiogenesis consists in the budding of neovessels from pre-existing capillaries and is driven by a decrease in local oxygen tension [1,7]. This is a complex multistep process during which a leading tip cell spearheads a new sprout by migrating outward from the parental vessel towards an angiogenic signal. Subsequently, trailing endothelial stalk cells start to proliferate, thereby supporting sprout elongation, generating the trunk of new capillaries and maintaining connection with the parental vessel. Tip cells emerging from neighbouring sprouts then anastomose to build functional vessel loops, allowing the initiation of blood flow, which also contributes to arterial-venous specification of endothelial cells [1,7]. The deposition of extracellular matrix and the recruitment of mural cells, such as pericytes and smooth muscle cells, promotes vessel maturation and stabilizes new connections. The sprouting process thus iterates until angiogenic cues cease and quiescence is re-established upon restoration of oxygen supply [1,7]. Alternately, microvascular growth can be accomplished through non-sprouting or intussusceptive angiogenesis, which consists of capillary splitting by the insertion of newly formed columns of interstitial tissue, known as pillars or posts, into the vascular lumen [8]. Although sprouting angiogenesis has long been regarded as the main mechanism responsible for vascular growth and remodeling during postnatal life, it has now been established that EPC actively contribute to neovessel formation also in the adult [9,10,11]. EPC are mobilized in circulation from stem cell niches located either in the bone marrow or in the endothelial intima of existing blood vessels to replace injured/senescent endothelial cells. In addition, EPC are massively released upon an ischemic insult, home to the hypoxic site and restore the injured vascular network by stimulating local angiogenesis in a paracrine manner or by physically engrafting within neovessels [9,10].
Under non-pathological conditions, the angiogenic switch is tightly controlled by a complex balance between stimulatory and inhibitory signaling molecules. When angiogenic inducers, including vascular endothelial growth factor (VEGF), basic fibroblast growth factor (bFGF), platelet derived growth factor (PDGF), stromal derived factor-1α (SDF-1α) and angiopoietin-1 (ANG-1)/Tie2, are produced in excess of anti-angiogenic factors, the balance is tipped towards neovessel formation. On the other hand, when the local concentration of anti-angiogenic factors, including thrombospondin-1, endostatin, tumstatin, vasohibin, and C-X-C motif chemokine 10 (CXCL10), overwhelms that of the stimulators, the angiogenic switch is turned off [7,12]. Insufficient vessel growth, malformation and regression contribute to numerous diseases, ranging from acute myocardial infarction, hindlimb ischemia and stroke to pre-eclampsia and neurodegeneration. Conversely, excessive/aberrant vessel growth ultimately results in tumorigenesis, intraocular disorders and inflammatory disease [7,13]. Therefore, unravelling the intracellular signaling pathways that drive vascular growth and remodeling is imperative to design efficient therapeutic strategies to treat, e.g., ischemic pathologies and cancer [7,14,15]. It has long been known that an increase in intracellular Ca2+ concentration ([Ca2+]i) within vascular endothelial cells plays a crucial role in angiogenesis [16,17,18]. Likewise, recent work demonstrated that intracellular Ca2+ signals also drive vasculogenesis by stimulating EPC to undergo proliferation, migration, and tube formation both in vitro and in vivo [19,20]. Herein, we discuss the basic mechanisms of pro-angiogenic Ca2+ signaling in vascular endothelial cells and circulating EPC. First, we survey the distinct Ca2+ signatures evoked by growth factors and chemokines in endothelial cells and the Ca2+-dependent decoders that translate endothelial Ca2+ waves into a pro-angiogenic response. We will briefly mention specific members of the Transient Receptor Potential (TRP) Canonical (TRPC) sub-family, as the role of TRP channels in angiogenesis has been extensively covered in two recent, comprehensive review articles [21,22]. Then, we describe the Ca2+ signaling toolkit recruited by VEGF, insulin-like growth factor 2 (IGF2) and SDF-1α in circulating EPC. The majority of the studies reviewed in the present article deal with endothelial Ca2+ signals arising under non-pathological conditions both in vitro and in vivo. Remodeling of the Ca2+ handling machinery by cardiovascular and oncological diseases is a less known facet of endothelial physiology and is addressed where appropriate.

2. Growth Factors and Chemokines Induce Pro-Angiogenic Ca2+ Signals in Vascular Endothelial Cells

An increase in [Ca2+]i has long been recognized as a key pro-angiogenic pathway that, lying at the intersection of multiple signaling cascades, is recruited by distinct mitogens to promote and modulate endothelial cell fate [23,24]. Growth factors and chemokines induce the angiogenic switch through an increase in [Ca2+]i that stimulates endothelial cell proliferation, adhesion, migration and bidimensional tube formation [16,25,26,27]. Furthermore, laminar shear stress that arises as a result of increased collateral blood flow during arteriogenesis, may also induce pro-angiogenic Ca2+ signals to favor vascular remodeling [28,29]. In addition, vascular endothelial cells may receive Ca2+-related pro-angiogenic inputs also by vasoactive and inflammatory mediators, including thrombin [30,31], ATP [32,33,34], ADP [34], and acetylcholine [35,36], and pleiotropic hormones, such as erythropoietin [37,38]. Finally, mechanical injury of the vascular intima may induce intracellular Ca2+ waves both at the edge of the injured area and at more remote sites, which are likely to promote vascular repair by inducing endothelial cell proliferation and migration [39,40]. Distinct intracellular Ca2+ signatures have been detected depending on the nature and strength of the extracellular stimulus, on the vascular bed and the species, and on the components of the multifaceted endothelial Ca2+ toolkit recruited by pro-angiogenic cues [18,25,41].

2.1. The Endothelial Ca2+ Toolkit Recruited by Growth Factors and Chemokines to Stimulate Angiogenesis

Mammalian cells, including vascular endothelial cells, impinge on two sources to generate the Ca2+ response to extracellular stimuli (Figure 1): endogenous Ca2+ mobilization and extracellular Ca2+ entry [18,42]. The recovery of [Ca2+]i to pre-stimulation levels is driven by a sophisticated network of Ca2+ pumps and transporters, such as Sarco-Endoplasmic Reticulum Ca2+-ATPase (SERCA), which sequesters cytosolic Ca2+ into the endoplasmic reticulum (ER), Plasma Membrane Ca2+-ATPase (PMCA) and Na+/Ca2+ exchanger (NCX), which clear Ca2+ across the plasma membrane (Figure 1) [43,44,45,46,47,48,49,50]. In addition, mitochondria were shown to buffer the influx of Ca2+ through store-operated channels, thereby redirecting entering Ca2+ to the ER via the mitochondrial NCX in absence of the Ca2+-releasing second messenger inositol 1,4,5-trisphosphate (InsP3) [51,52,53].

2.1.1. The Onset of Pro-Angiogenic Ca2+ Signals: PLCβ and PLCγ

Pro-angiogenic cues bind to specific receptor tyrosine kinases (RTK) and Gq/11-protein coupled receptors (Gq/11PCR) which, respectively, recognize growth factors and a wide assortment of chemokines, autacoids and hormones [41,54,55]. RTK and Gq/11PCR, in turn, trigger an increase in endothelial [Ca2+]i by recruiting, respectively, phospholipase C-γ (PLCγ1-2) and phospholipase C-β (PLCβ1-4) (Figure 1). Although both PLCγ1 and PLCγ2 are present in vascular endothelial cells [56,57], PLCγ1 is regarded as the main transducer of RTK activity [58,59] as PLCγ2 role is hitherto restricted to blood lineage cells [60]. PLCβ1-4 are also readily detectable throughout vascular endothelium [56], albeit PLCβ1 is absent in human umbilical vein endothelial cells (HUVEC) [57], which represent one of the most widespread models to investigate endothelial Ca2+ signals [18]. It has been reported that Gαq monomers activate PLCβ isotypes according to the following rank order: PLCβ1 ≥ PLCβ3 > PLCβ2, while PLCβ4 recruitment is heavily limited by ribonucleotides, such as GTP-γ-S [60,61]. The role played by the distinct PLCβ isotypes in the endothelial Ca2+ response to Gq/11PCR has not been carefully dissected, but preliminary evidence suggested the involvement of PLCβ1 [62] and PLCβ3 [63]. In addition, PLCβ1-3, but not PLCβ4, are also sensitive to Gβγ dimers, although they display high affinity only towards PLCβ2 [60]. As a consequence, pro-angiogenic Ca2+ signals may also be induced by Gi/oPCR, such as P2Y12 [64,65], sphingosine-1 phosphate (S1P) receptor 1 (S1R1) [66], and C-X-C chemokine receptor type 4 (CXCR4) [67].
Once engaged by extracellular stimulation, PLCβ and PLCγ isoenzymes catalyze the hydrolysis of phosphatidylinositol 4,5-bisphosphate (PIP2), a minor membrane phospholipid, into the two intracellular second messengers, InsP3 and diacylglycerol (DAG) (Figure 1) [42]. As mentioned above, PLCγ1 is the main PLC isozyme whereby growth factors, such as VEGF, the master regulator of angiogenesis, regulate endothelial cell proliferation, migration, and tube formation [58,59,68]. For instance, targeted next generation sequence recently identified a PLCγ1 with a recurrent nonsynonymous mutation (R707Q) in primary cardiac angiosarcomas [69], a rare set of tumors triggered by aberrant proliferation and migration of coronary endothelial cells. Expression studies revealed that the PLCγ1-R707Q mutant was constitutively active in HUVEC, thereby enhancing InsP3 synthesis and the Ca2+-dependent activation of calcineurin and increasing endothelial cell migration and invasiveness [69]. Surprisingly, a recent series of studies demonstrated that VEGF-induced InsP3 production and downstream pro-angiogenic effects are also mediated by PLCβ3 [70,71]. Furthermore, genetic silencing of PLCβ3 impaired proliferation, migration and tubulogenesis in HUVEC grown in the presence of EGM-2 [72], an endothelial growth medium enriched with multiple growth factors. It has, indeed, been shown that the RTK VEGF receptor 2 (VEGFR2), which is the main signaling VEGF receptor in vascular endothelial cells [24], may recruit PLCβ3 through phosphorylation at serine 537 and 1105 (S537 and S1105) [70]. PLC activation leads to a pro-angiogenic increase in [Ca2+]i that can be initiated by both InsP3 [58,68,69,70,73] and DAG [74,75,76,77,78]. In the following section, we illustrate how InsP3 production results in pro-angiogenic Ca2+ signals in vascular endothelial cells, whereas the mechanism of action of DAG will be briefly described in Section 2.1.3.

2.1.2. Endogenous Ca2+ Release Induced by Pro-Angiogenic Cues: InsP3 Receptors (InsP3R), Ryanodine Receptors (RyR) and Two-Pore Channels (TPC)

The ER represents the largest endogenous Ca2+ store in vascular endothelial cells by containing approximately 75% of the intracellular Ca2+ reservoir (Figure 1) [79]. The ER Ca2+ concentration ([Ca2+]ER) is maintained at around 100-500 μM by SERCA2b [43,47,49,50,80], an ER-specific pump displaying a high affinity for Ca2+, but low Ca2+ transport capacity [18]. Vascular endothelial cells also express SERCA3 [81,82], however, its expression decreases during proliferation in culture [81]. In addition, Ca2+ sequestration in endothelial ER vesicles is impaired by genetic silencing of SERCA2b, but not SERCA3 [48].

2.1.2.1. InsP3R

InsP3 evokes endogenous Ca2+ release by binding to the non-selective cation channels, InsP3R [18,83,84], which are located in the ER membrane (Figure 1). The resulting opening of InsP3R leads to the efflux of intraluminal Ca2+ along the electrochemical gradient between the ER and the cytosol, where [Ca2+]i quickly raises from around 100 nM up to 1 μM [18,83]. InsP3R modulate a wealth of processes, including proliferation, migration, gene expression, fluid secretion, synaptic plasticity, contraction, and membrane excitability [85]. Three InsP3R isoforms, i.e., InsP3R1-3, have been described in mammalian cells [86], all of which are present throughout vascular endothelium [18], although their pattern of expression may vary depending on the species and the vascular bed [81,83,87,88,89,90,91]. We refer the readers to a recent review, which provides a comprehensive overview of endothelial InsP3R [83]. Herein, we just recall that InsP3R are actually primed by InsP3 to respond to stimulation to ambient Ca2+, which gates InsP3R at low concentrations (≈50–200 nM), but inhibits their opening at substantially higher levels [86]. The bell-shaped dependence of InsP3R on surrounding Ca2+ has been also demonstrated in vascular endothelial cells [92]. The InsP3R subtypes differ in their affinity for InsP3 (InsP3R2 > InsP3R1 > InsP3R3) and in their sensitivity to Ca2+-induced inhibition, as InsP3R3 is less prone to close in the presence of high Ca2+ concentration [93]. As a consequence of their peculiar sensitivity to InsP3 and Ca2+, InsP3R1 and InsP3R2 are more tailored to generate long-lasting oscillations in [Ca2+]i, while InsP3R3 rather instigates monophasic Ca2+ signals [83,93]. No study has hitherto evaluated the correlation between the pattern of InsP3R endowed to a given endothelial cell and the induction of intracellular Ca2+ oscillations by pro-angiogenic cues. However, two recent studies demonstrated that acetylcholine induces repetitive Ca2+ transients in mouse brain endothelial cells, which express both InsP3R1 and InsP3R2 [89], but a monophasic increase in [Ca2+]i in human cerebrovascular endothelial cells, in which InsP3R3 is the most abundant isoform [88]. It is, finally, worth recalling here that early work provided the evidence that InsP3Rs may be located also on the plasma membrane, and therefore mediate extracellular Ca2+ entry, in bovine aortic endothelial cells (BAEC) [94] and HUVEC [95]. This observation has been subsequently confirmed in other cell types and increases the versatility of InsP3 signaling in mammalian cells [96].

2.1.2.2. RyR

InsP3-induced intracellular Ca2+ release may be amplified by the recruitment of adjoining RyR through the process of Ca2+-induced Ca2+ release [18]. RyR mainly modulate muscle contraction, although they have also been involved in synaptic transmission in the brain, insulin release from pancreatic β-cells and control of vascular tone in endothelial cells [97]. Of the three RyR isoforms, i.e., RyR1-3, expressed in mammalian cells [42], vascular endothelial cells mainly express RyR3 [98]. It should, however, be pointed out that several studies failed to detect the endothelial expression of RyR [87,88,89] and that single-cell RT-PCR revealed that RyR3 was present in a limited percentage (5–25%) of the analyzed endothelial cells [98]. Consistently, although RyR-mediated endogenous Ca2+ release has been reported [99,100,101], the role of RyR in endothelial Ca2+ signaling is usually regarded as inconsistent or not as necessary as that played by InsP3R [39,84,102,103]. However, a recent investigation suggested that RyR may contribute to VEGF-induced Ca2+ signals in human aortic endothelial cells (HAEC) [48].

2.1.2.3. TPC

While the ER is the largest endogenous Ca2+ store recruited by endothelial mitogens, emerging evidence hinted at the acidic vesicles of the endolysosomal (EL) system as an alternative Ca2+ reservoir establishing a Ca2+-mediated cross-talk with the ER to generate pro-angiogenic outputs [41,104]. The EL Ca2+ store is targeted by the latest addition to the family of Ca2+-releasing messengers, namely nicotinic acid adenine dinucleotide phosphate (NAADP), that release EL Ca2+ by activating TPC 1 and 2 (TPC1-2) (Figure 1), which are novel members of the superfamily of voltage-gated Ca2+ channels [105]. NAADP and TPC1-2 regulate a growing number of functions, including autophagy, nutrient sensing, membrane trafficking, exocytosis, fertilization and embryogenesis, proliferation and synaptic transmission [105,106,107,108]. NAADP synthesis has been attributed to the multifunctional enzyme CD38, although NAADP levels are increased, rather than decreased, by genetic suppression of CD38 in several murine tissues [109]. However, CD38 is expressed in endothelial cells [110] and NAADP is synthesized in response to endothelial autacoids, such as histamine [111]. In addition, NAADP supports the angiogenic Ca2+ response to VEGF in HUVEC [104]. It has been proposed that NAADP induces local EL Ca2+ release through TPC1-2 that is, in turn, amplified into a regenerative Ca2+ wave by the Ca2+-dependent recruitment of juxtaposed (<30 nm) InsP3R and RyR at quasi-synaptic ER-EL junctions [112,113]. The so-called “trigger hypothesis”, a term introduced by Antony Galione [112], has been invoked to describe the role played by NAADP in endothelial Ca2+ signaling [88,104,111,114]. However, this Ca2+-mediated cross-talk is bidirectional, as InsP3R at the ER-EL interface may in turn refill acidic vesicles with Ca2+, thereby modulating the EL Ca2+ content [115,116]. Whether this reciprocal EL-to-ER Ca2+ shuttle exists and plays an angiogenic role in vascular endothelial cells is still unclear.

2.1.3. Extracellular Ca2+ Entry Induced by Pro-Angiogenic Cues in Vascular Endothelial Cells: Store-Operated Ca2+ Entry (SOCE), STIM1 and Orai1, and TRPC

SOCE is the most widespread Ca2+ entry pathway in both excitable and non-excitable cells [117,118] and regulates a plethora of cellular functions, including refilling of the ER Ca2+ store, gene expression, cell cycle regulation, cytoskeletal remodeling, NO release and cyclic AMP production [117,119]. SOCE also represents the main Ca2+-permeable pathway, which mediates extracellular Ca2+ entry in response to external autacoids and angiogenic cues in vascular endothelial cells [15,120,121]. Endothelial SOCE is activated upon depletion of the InsP3-sensitive ER Ca2+ pool and may, therefore, be activated by RTK, such as VEGFR2 [48,122,123,124]. The molecular mechanisms underlying endothelial SOCE vary depending on the species, the vascular district and the activating stimulus, as recently reviewed in [15,22,120,121]. Therefore, herein we mainly refer to SOCE activation by pro-angiogenic cues in vascular endothelial cells (Figure 1).

2.1.3.1. SOCE: STIM1 and Orai1

Early work carried out by Mohamed Trebak and coworkers revealed that genetic silencing of STIM1 and Orai1 abolish SOCE in HUVEC (Figure 1) [122]. This finding was later confirmed by at least other three independent studies [123,125,126]. STIM1 provides the ER Ca2+ sensor, which detects ER Ca2+ depletion in response to InsP3-synthesizing stimuli, such as growth factors and chemokines. Then, STIM1 undergoes a complex conformational remodeling which results in the relocation of STIM1 oligomers within the peripheral cisternae of the ER. Herein, STIM1 oligomers assembly into localized clusters, known as puncta, which are ≈20 nm apart from the plasma membrane, thereby tethering and gating Orai1, which constitutes the pore-forming subunit of store-operated channels [15,122,123]. Orai1 channels are hexameric, bind to STIM1 proteins in a 1:2 ratio, and mediate a highly Ca2+-permeable channel, known as Ca2+-release activated Ca2+ current (ICRAC) [117]. The ICRAC exhibits distinguishing biophysical features, including extremely low unitary conductance (10–25 fS for Ca2+), inwardly-rectifying current-to-voltage (I-V) relationship, very positive reversal potential (Erev ≈ +60 mV), and high selectivity for Ca2+ over monovalent cations (PCa/PNa > 1.000) [117]. An ICRAC-like current has been recorded in vascular endothelial cells in response to massive ER Ca2+ depletion [122,127,128] and, as expected, it is inhibited upon genetic deletion of STIM1 and Orai1 [122]. Nevertheless, VEGF-evoked ICRAC is too tiny to be detected by conventional recording systems (i.e., it falls below the pA range) and is yet to be measured [123].

2.1.3.2. SOCE: STIM2 and Orai2-3

In addition to STIM1 and Orai1, vascular endothelial cells express their paralogues, STIM2 and Ora1-3 [88,89,122,123,129,130]. Endothelial STIM2 regulates basal Ca2+ entry and ER Ca2+ loading by regulating resting Ca2+ permeability [131], whereas it is unknown whether it also controls the ICRAC arising in response to physiological stimulation, as shown elsewhere [132,133]. The sole exception is currently provided by a human brain microvascular endothelial cell line, hCMEC/D3, in which STIM2 is the only STIM isoform expressed [88]. Interestingly, STIM2 deletion impairs proliferation in HUVEC although its involvement in SOCE was not investigated [122]. Orai2, in turn, is the most likely candidate to mediate SOCE in the mouse cerebrovascular endothelial cell line, bEND5, in which Orai1 and Orai3 are absent [89]. In addition, Orai2 could serve as a negative modulator of Orai1 and SOCE in a bovine brain capillary endothelial cell line, t-BBEC117, in which Orai2 up-regulation during the G2/M phase may retard the proliferation rate [129]. Finally, the endothelial Orai3 is not sensitive to STIM proteins, but to the arachidonic acid-derived metabolite, leukotriene C4 [134].

2.1.3.3. SOCE: TRPC1 and TRPC4

The molecular make-up of the endothelial SOCE is, however, a matter of intense controversy [22,120,121,135]. It has long been known that multiple members of the TRPC sub-family of non-selective cation channels may also support SOCE in response to pharmacological, i.e., upon SERCA inhibition, and physiological, i.e., by InsP3-synthesizing autacoids, stimuli [21,22]. The TRPC sub-family encompasses seven members (TRPC1-TRPC7) that, with the notable exception of the pseudogene TRPC2 in humans, enable Na+ and Ca2+ entry downstream of PLC activation [22,136]. TRPC channels control a multitude of functions, including gene expression, cell motility and adhesion, membrane excitability, angiogenesis and NO release [136,137,138]. Classically, endothelial TRPC3 and TRPC6 mediate store-independent Ca2+ influx following PLCγ1 activation [21,22]. For instance, TRPC3 and TRPC6 serve as second messengers-operated Ca2+-permeable channels directly gated by DAG (Figure 1) [139]. Intriguingly, TRPC1 and TRPC4 were shown to interact with STIM1 and mediate SOCE in human and mouse lung endothelial cells [140], whereas SOCE was severely down-regulated in cultured aortic endothelial cells isolated from TRPC4-deficient mice [141]. Likewise, a role for endothelial SOCE has been proposed for TRPC1 [124,142] and TRPC4 [143,144] by other independent studies (Figure 1). TRPC channels show a distinct repertoire of biophysical properties as compared to Orai1, which result in a store-operated current, also termed ISOC, which is featured by: 1000 times higher unitary conductance (in the pS range), linear I-V relationship and Erev of ≈0 mV as TRPC channels do not discriminate between Na+, K+ and Ca2+ [145,146]. Quite surprisingly, however, electrophysiological recordings revealed that the endothelial TRPC1 and TRPC4 channels mediate either an ICRAC-like current [141] or a less Ca2+-selective current with biophysical properties intermediate between ICRAC and ISOC [142,143,147,148]. Alternately, endothelial TRPC channels may line the pore of an ISOC-like conductance [124,149,150,151]. A recent series of studies sought to solve this conundrum by suggesting that one TRPC1 and two TRPC4 subunits assemble into a supermolecular ternary complex in vascular endothelial cells [152,153]. The sensitivity of this TRPC1/TRPC4 complex to ER Ca2+ loading is conferred by STIM1 [154], which could interact with both TRPC1 and TRPC4 [152], and by the interaction between TRPC4 and protein 4.1 and between protein 4.1 and spectrin [142,143]. In addition, Orai1 may constitutively interact with TRPC4, and potentially with TRPC1 upon ER Ca2+ depletion, thereby increasing the Ca2+-selectivity and conferring the ICRAC-like fingerprints to the endothelial store-operated current [153,155]. While this information fulfills the goal to complete our description of the mechanisms that underpin the endothelial SOCE, only limited evidence has been provided to support the role of store-operated TRPC1 and TRPC4 in the Ca2+ response to pro-angiogenic cues.

3. Pro-Angiogenic Ca2+ Signals in Vascular Endothelial Cells

The Ca2+ toolkit described in Paragraph 2 may be differently recruited by distinct growth factors and cytokines to induce spatio-temporal Ca2+ signals, which are precisely tailored to regulate specific phases of the angiogenic process. The endothelial Ca2+ response to pro-angiogenic cues usually consists of a transient or biphasic increase in [Ca2+]i or in repetitive Ca2+ spikes. These Ca2+ signatures are mainly driven by endogenous Ca2+ release and SOCE, although the contribution of TRPC channels has been reported. The pro-angiogenic Ca2+ signal can be modulated by concomitant changes in the endothelial membrane potential (VM) due to the recruitment of Ca2+-dependent conductances. Novel high-resolution and high-speed imaging techniques confirmed that VEGF-induced endothelial Ca2+ signals arise in vivo and drive endothelial cell sprouting and migration.

3.1. VEGF-Induced Intracellular Ca2+ Signals in Vascular Endothelial Cells

The VEGF superfamily of growth factors include VEGF-A (which comprises four isoforms: VEGF-A121, VEGF-A145, VEGF-A165, VEGF-A189), VEGF-B, VEGF-C, VEGF-D, VEGF-E (encoded by Orf virus), VEGF-F (Vammin, isolated from Vipera ammodytes venom), and placenta growth factor (PlGF) [24,156]. While VEGF-C and VEGF-D mainly promote development of lymphatic vessels, VEGF-A165 (commonly termed VEGF) is the master regulator of angiogenesis in peripheral circulation as well as in most pathologies associated to aberrant vascular growth, such as cancer and blinding eye disorders [24,156]. VEGF isoforms stimulate vascular and lymphatic endothelial cells by binding to their high affinity cognate receptors, which include the RTK VEGFR1, VEGFR2, and VEGFR3 and the VEGF co-receptors neuropilin 1 and 2 (NRP1 and NRP2, respectively) and heparin sulfate proteoglycans. VEGFR1 and VEGFR2 are mainly expressed in vascular endothelium, while VEGFR3 is restricted to lymphatic endothelial cells. VEGFR2 [also known as KDR (kinase insert domain receptor, human) and Flk1 (fetal liver kinase-1, mouse)] is the main receptor isoform which transduces VEGF signaling in vascular endothelial cells, while VEGFR1 (also termed Fms-like tyrosine kinase 1, Flt1) may exist in a soluble form (sFlt1) which presents a higher affinity for VEGF than VEGFR2 and is, therefore, able to inhibit angiogenesis [24,156]. When VEGF binds to VEGFR2, the receptor undergoes dimerization and auto- or trans-phosphorylation of tyrosine residues on the receptor dimer as well as on downstream mediators of the pro-angiogenic signal. These include PLCγ1 and the RAS/RAF/extracellular signal-regulated kinases (ERK)/mitogen-activated protein kinase (MAPK) pathway, which promotes vascular development and arteriogenesis; the phosphoinositide 3-kinases (PI3K)/AKT pathway, which supports endothelial cell survival and limits apoptosis; endothelial nitric oxide (NO) synthase (eNOS), which stimulates endothelial cell proliferation and migration and drives the increase in capillary permeability; and SRC and small GTPases, which regulate endothelial junctions and endothelial permeability and regulate endothelial cell shape, cell migration and polarization [24,135,157,158].
An increase in [Ca2+]i is regarded as a crucial signal whereby VEGF stimulates vascular endothelial cells to undergo cell fate specification, proliferation, migration, tubulogenesis and neovessel formation [21,24]. The first evidence about the pro-angiogenic role of endothelial Ca2+ signaling dates back to thirty years ago, when Criscuolo and coworkers demonstrated that the tumor-secreted vascular permeability factor, subsequently identified as VEGF by Napoleone Ferrara [156], caused a biphasic increase in [Ca2+]i in several types of endothelial cells, including HUVEC [159]. A subsequent study revealed that the endothelial Ca2+ response to VEGF was mediated by VEGFR2 [160]. The majority of the work elucidating the relationship between VEGF, Ca2+ signaling and angiogenesis has been carried out in HUVEC. In the next chapters, therefore, we first illustrate the mechanisms whereby VEGF induces pro-angiogenic Ca2+ signals in HUVEC and then focus our attention on other endothelial cells types.

3.2. VEGF-Induced Intracellular Ca2+ Signals in HUVEC

The typical Ca2+ response to VEGF in HUVEC consists in a biphasic elevation in [Ca2+]i as originally reported in [73] and subsequently confirmed in [122,123,161,162]. This pattern of signaling comprises an initial Ca2+ peak, which is due to InsP3-dependent Ca2+ release from the ER, followed by a prolonged plateau phase, which is maintained by the interaction between STIM1 and Orai1, i.e., by SOCE activation (Figure 1 and Table 1) [73,122,123]. Notably, genetic deletion (through a small interfering RNA) and pharmacological blockade (with carboxyamidotriazole and S66) of Orai1 prevents HUVEC proliferation, migration and tube formation [73,123]. In addition, VEGFR2 and Orai1 are clustered at restricted sites within the plasma membrane, a mechanism that could remarkably improve the efficiency of VEGF signaling in these cells [123]. It has also been proposed that plasma membrane InsP3R contribute to VEGF-induced Ca2+ entry in HUVEC, but the evidence in favor of this hypothesis is only correlative [95]. Conversely, strong evidence suggest that VEGF-induced extracellular Ca2+ entry in HUVEC may be sustained by the store-independent channels (Figure 1 and Table 1), TRPC3 [77,163] and TRPC6 [76]. TRPC3, in turn, could recruit the reverse (Ca2+-entry mode) of NCX1 [163]. Finally, TRPC1 could also contribute to VEGF-induced Ca2+ plateau in HUVEC [124,164]. TRPC1 is physically retained in close associated with VEGFR2 by Klotho protein, but how VEGFR2 activation results in TRPC1-mediated Ca2+ entry is unclear [164].
More recently, it was found that NAADP supports VEGF-induced InsP3-dependent Ca2+ release in HUVEC (Figure 1 and Table 1) [104]. Accordingly, genetic deletion of TPC1 reduces VEGF-induced intracellular Ca2+ signaling, in vitro tubulogenesis and neovessel formation in vivo [104]. As mentioned in Section 2.1.2.3, NAADP-dependent Ca2+ release from acidic stores could provide the local pulse of Ca2+ necessary for InsP3R activation by InsP3, but this model remains to be confirmed in HUVEC. A recent investigation compared the Ca2+ responses to VEGF-A165 and VEGF-A121 in HUVEC and found that their Ca2+ sensitivity to VEGF-A165 is remarkably higher due to the more efficient recruitment of PLCγ1 [161]. In addition, when VEGFR2 availability is compromised, as observed upon genetic silencing of the dual specificity tyrosine phosphorylation-regulated kinase A (DYRKA), VEGF-induced InsP3-dependent ER Ca2+ release and the subsequent engagement of Ca2+-sensitive decoders is hampered [58]. VEGF-induced intracellular Ca2+ signals have also been recorded in the HUVEC-derived cell line, EA.hy926, in which VEGF triggers a transient increase in [Ca2+]i [165]. This Ca2+ signal is mediated by InsP3R (Table 1) and requires the concomitant production of the gasotransmitter hydrogen sulphide (H2S). Accordingly, the pharmacological blockade of H2S synthesis inhibits the Ca2+ response to VEGF as well as VEGF-induced EA.hy926 cell proliferation and migration [165]. H2S is a known modulator of the endothelial Ca2+ toolkit [166,167,168] and regulates angiogenesis [169], but the mechanistic link between VEGF, H2S and endothelial Ca2+ signaling deserves further investigation.

3.3. VEGF-Induced Intracellular Ca2+ Signals in Other Vascular Endothelial Cell Types: In Vitro and In Vivo Evidences

Apart from HUVEC, VEGF-induced intracellular Ca2+ signals have been widely characterized in peripheral and pulmonary circulation. For instance, endothelial Ca2+ signaling drives VEGF-induced retinal angiogenesis, whose dysregulation could lead to several sight-threatening diseases, including diabetic retinopathy, age-related macular degeneration, and retinopathy of prematurity [170]. VEGF triggers a biphasic increase in [Ca2+]i in bovine retinal endothelial cells (BREC). As reported in HUVEC, the initial Ca2+ peak is patterned by InsP3-mediated ER Ca2+ release, whereas the plateau phase depends on Ca2+ entry through a yet to be defined Ca2+-permeable route. The pharmacological inhibition of InsP3-dependent Ca2+ mobilization prevents VEGF-induced BREC proliferation, migration, tubulogenesis and sprout formation [170]. Of note, genetic deletion of TRPC4 impairs VEGF-induced migration and tube formation in human retinal microvascular endothelial cells (HRMEC), but it is not clear whether TRPC4 is sustained in a store-dependent manner [171]. VEGF induces a biphasic Ca2+ signal also in the intact endothelium of ovine uterine vasculature [172], although it is not clear whether this Ca2+ response drives VEGF-induced angiogenesis and NO-dependent vasodilation during pregnancy [173]. Surprisingly, pre-treatment with VEGF inhibits ATP-induced intracellular Ca2+ burst and NO production [172,174], thereby mimicking the reduction in uterine vascular resistance that causes pre-eclampsia. An increase in local levels of VEGF has indeed been reported in pre-eclampsia and could be responsible for the local hypertension that presents significant risks of death for both mother and child [175]. VEGF-induced intracellular Ca2+ signals were also observed in mouse aortic endothelial cells (MAEC), in which the Ca2+ response is abolished by genetic deletion of the Ca2+-dependent tyrosine kinase Pyk2 [59]. Likewise, VEGF-induced migration, in vitro tubulogenesis and actin cytoskeleton remodeling are attenuated in MEAC deficient of Pyk2 [59]. An InsP3-dependent pro-angiogenic Ca2+ response to VEGF was also recorded in human dermal microvascular endothelial cells (HDMEC) [176], HAEC [177], human pulmonary artery endothelial cells [178], bovine choroidal endothelial cells [179], porcine aorta endothelial cells (PAEC) [25,180] and coronary venules endothelial cells [181]. VEGF-induced InsP3-dependent Ca2+ release is further supported by RyR and Orai1-dependent extracellular Ca2+ entry in HAEC (Table 1) [48]. VEGF has also been shown to induce endogenous Ca2+ release and extracellular Ca2+ entry, presumably through Orai1, in MAEC [50]. In addition, it has been suggested that VEGF underlies the spontaneous Ca2+ oscillations driving in vitro tubulogenesis in mouse yolk sac endothelial cells [182]. Finally, a heteromeric channel, which is likely to comprise TRPC3 and TRPC6 subunits, could mediate VEGF-induced extracellular Ca2+ entry in HDMEC [74,75].
A sophisticated tool to investigate how the endothelial Ca2+ toolkit shapes the pro-angiogenic Ca2+ response to VEGF is provided by computational modeling. A rule-based modeling approach utilizing the programming language BioNetGent recently confirmed that the extent of VEGF-induced Ca2+ entry during the plateau phase is finely tuned by the ICRAC. Likewise, a reduction in the rate of Ca2+ clearing through either SERCA or PMCA could further sustain the plateau amplitude [183].
Three recent studies have shed novel light on the mechanism whereby VEGF-induced endothelial Ca2+ signals drive angiogenesis. An investigation carried out on PAEC revealed that, as long as VEGF concentration is increased from the low to high nanomolar range, the percentage of cells displaying transient or repetitive Ca2+ spikes decreases, while the fraction of cells showing a biphasic Ca2+ signal increases [25]. As described in more detail in Section 4.3 and Section 4.5, intracellular Ca2+ oscillations selectively drive proliferation, while the persistent plateau phase is crucial for migration [25]. In addition, high-speed, three-dimensional (3D) time-lapse imaging disclosed repetitive intracellular Ca2+ waves induced by VEGF in both stalk and tip cells sprouting from dorsal aorta and posterior cardinal vein in zebrafish [184]. VEGFR2 and VEGFR3 mediate the onset of the intracellular Ca2+ oscillations in the endothelial cells budding from dorsal aorta and posterior cardinal vein, respectively. Intriguingly, the endothelial Ca2+ waves spread along the dorsal aorta until they become restricted to the selected tip cell and, thereafter, to the stalk cells that trail behind, while Dll4/Notch signaling is responsible for suppressing the Ca2+ spikes in endothelial cells in close proximity to tip cells during budding [184]. VEGF-induced intracellular Ca2+ oscillations in tip cells are sustained by InsP3-dependent Ca2+ release and SOCE and maintain Dll4/Notch signaling to coordinate vascular morphogenesis (Table 1) [26]. These reports, therefore, confirmed for the first time that endothelial Ca2+ signaling supports the pro-angiogenic response to VEGF in vivo and pave the way for novel investigations employing super resolution imaging techniques to visualize endothelial Ca2+ signals in physiologically relevant contexts.

3.4. Modulation of VEGF-Induced Endothelial Ca2+ Signals

The endothelial Ca2+ response to VEGF undergoes complex modulation by a plethora of regulatory mechanisms. For instance, the shape of VEGF-induced endothelial Ca2+ signals and their impact on the angiogenic process may be modulated by S-glutathiolation of SERCA2b in HAEC [47]. A number of investigations revealed that VEGF (as well as physiological doses of exogenous hydrogen peroxide or H2O2) stimulated NADPH to generate H2O2, which in turn boosts SERCA2b activity by adding S-glutathione adducts at cysteine 674 (C674) [47]. Likewise, VEGF-induced eNOS activation may lead to SERCA2b C674 S-glutathiolation and increase SERCA2-dependent ER Ca2+ uptake [48]. Of note, adduction of S-glutathione at C674 favors VEGF-induced Ca2+ release through RyR and Orai1, thereby promoting HAEC migration and tubular network formation [47,48]. It has been further reported that, under hypoxic conditions, the increase in reactive oxygen and nitrogen species also leads to SERCA2b S-glutathiolation and endothelial tube formation [49]. In agreement with these observations, mouse hind limb ischemia induces the formation of S-glutathione adducts on endothelial SERCA2b, which is indispensable to support blood flow recovery [50]. Similar to HAEC, VEGF-induced endogenous Ca2+ release and extracellular Ca2+ entry, as well as cell migration, are impaired in mouse cardiac endothelial cells devoid of half C674, which prevents SERCA2b S-glutathiolation [50].
The angiogenic activity of VEGF may also be regulated by PMCA. PMCA isoforms are encoded by four genes (PMCA1-4) [18], of which PMCA1 and PMCA4 are the most abundant in vascular endothelial cells [185,186]. Recent investigations demonstrated that PMCA4 activity results in a low Ca2+ microdomain that tethers the Ca2+-dependent phosphatase calcineurin beneath the plasma membrane, thereby preventing its activation by VEGF [46,187]. As a consequence, VEGF fails to induce migration and tube formation, although not proliferation, in HUVEC [46].
Besides SERCA2b and PMCA4, VEGF-induced endothelial Ca2+ signals and pro-angiogenic activity may be finely tuned by intermediate-conductance Ca2+-activated K+ (IKCa) channels [188]. The IKCa channel, encoded by the IKCa1 gene, is up-regulated in HUVEC pre-treated with VEGF for 48 h and the pharmacological inhibition of IKCa currents prevents capillary formation in in vivo Matrigel plug assays [188]. IKCa activation leads to the hyperpolarization of endothelial VM, which could enhance extracellular Ca2+ entry during the plateau phase of the Ca2+ response to VEGF [189,190]. However, it remains to be elucidated whether VEGF actually recruits IKCa channels and, vice versa, whether IKCa channels sustain VEGF-induced extracellular Ca2+ influx during the angiogenic activity. It should, however, be pointed out that VEGF inhibits InsP3-dependent ER Ca2+ mobilization and the Ca2+-dependent recruitment of IKCa channels in mouse pressurized resistance arteries [191].
The Ca2+ response to VEGF is also sensitive to angiogenesis inhibitors, such as angiostatin and endostatin, which are necessary for vessel pruning and regression of selected vascular branches [12]. Angiostatin comprises of the first four-kringle domains of plasminogen, which is a quite abundant protein in wound microenvironment and may be cleaved into an anti-angiogenic peptide by matrix metalloproteases (MMP) [12]. Similarly, endostatin is a 20 kDa fragment of the endothelial cell basement membrane component type XVIII collagen. Endostatin is readily cleaved from the extracellular matrix by MMP to inhibit endothelial cell proliferation and migration and inducing apoptosis [12]. Early work revealed that angiostatin and endostatin prevent VEGF-induced endogenous Ca2+ release by inducing InsP3-mediated ER Ca2+ release and activating a Ca2+-entry pathway in BAEC (Table 1) [192]. It is, therefore, conceivable that their anti-angiogenic action is, at least partly, effected by depletion of the same InsP3-sensitive ER Ca2+ store targeted by VEGF [192]. Accordingly, if the ER Ca2+ store has been previously depleted by angiostatin and endostatin, VEGF will fail to elicit a pro-angiogenic increase in endothelial [Ca2+]i and to activate SOCE. The ER microenvironment is not homogenous and comprises of multiple sub-sections, each performing distinct tasks [193,194]. Therefore, the ER Ca2+ pool targeted by angiostatin and endostatin could not be coupled to pro-angiogenic Ca2+-dependent decoders. Nevertheless, the fall in [Ca2+]ER could be propagated to the ER Ca2+ pool engaged by VEGF through the mechanism of Ca2+ tunneling, thereby impairing the subsequent Ca2+ response to pro-angiogenic stimulation [195]. In addition, as EL Ca2+ mobilization has also been involved in VEGF-induced Ca2+ signaling, as discussed in Section 3.2, it would be interesting to assess whether angiostatin and endostatin also affect NAADP-induced endogenous Ca2+ mobilization. This finding was not confirmed in the intact endothelium of freshly isolated bovine coronary arteries challenged with bradykinin upon 1 h exposure to endostatin [196]. It would be interesting to assess whether endostatin and angiostatin selectively affect the VEGF-sensitive compartment of the InsP3-releasable ER Ca2+ pool. In addition, computational simulations recently investigated how the angiogenesis inhibitor, thombospondin-1, affects VEGFR2 signaling [197]. Thrombosponin-1 is a matricellular protein which binds to the extracellular matrix and inhibits endothelial cell proliferation and migration besides promoting endothelial apoptosis and repressing VEGF activity [12]. In silico investigations suggest that thrombospondin-1 binds to the endothelial receptor CD47, which in turn induces VEGR2 degradation and impairs VEGF-induced recruitment of Ca2+-sensitive pro-angiogenic pathways [197]. Finally, VEGF-induced endothelial Ca2+ signals are sensitive to Sprouty4 [198], a membrane-bound inhibitor of the ERK pathway. Sprouty4 prevents PIP2 hydrolysis by PLCγ1, thereby suppressing the synthesis of both InsP3 and DAG and dampening protein kinase C (PKC) activation, which contributes to the recruiting of ERK along with the increase in [Ca2+]i [198].

3.5. Endothelial Ca2+ Signals Induced by bFGF, Epidermal Growth Factor (EGF), PDGF, and SDF-1α

As expected by their ability to bind to RTK coupled to the PLCγ1/InsP3 signaling axis, many other growth factors can induce pro-angiogenic Ca2+ signals in vascular endothelial cells [41,199,200]. For instance, bFGF, also known as FGF-2, triggers an increase in [Ca2+]i in a variety of vascular endothelial cells, including HUVEC [63,199,201], SV40-transfected human corneal endothelial cells (SV40-HCEC) [202], and BAEC [192,200]. The Ca2+ response to bFGF is initiated by FGF receptor-1 (FGFR-1) [179,203], requires PLCγ1 activation [201], although the contribution of PLCβ3 has also been reported [63], and is essentially mediated by extracellular Ca2+ entry [200]. It has been suggested that bFGF-induced extracellular Ca2+ entry is mediated by arachidonic acid through a yet to be clearly identified Ca2+-permeable route [204]. The available evidence hints at a heteromeric channel formed by TRPC1 and TRPC4 in BAEC [203,205]. Of note, the pharmacological inhibition of bFGF-induced Ca2+ influx with carboxyamidotriazole and of arachidonic acid production by means of multiple inhibitors of phospholipase A2 and DAG lipase suppress BAEC proliferation [204]. Likewise, insulin-like growth factor-I (IGF-I) induces extracellular Ca2+ entry through the same pathway as that recruited by bFGF. Nevertheless, the physiological outcome of the Ca2+ response to IGF-I has not been evaluated [200]. Similar to VEGF, bFGF-induced Ca2+ entry could be modulated by the concomitant recruitment of large-conductance Ca2+-activated K+ (BKCa) channels [206,207] and IKCa channels [188]. The regulatory role of endothelial VM hyperpolarization is highlighted by the inhibition of bFGF-induced proliferation upon blockade of either BKCa [206] or IKCa channels [188]. Furthermore, bFGF increases the amplitude of the inwardly rectifying K+ current (KIR) in HUVEC, thereby boosting VM hyperpolarization and cell proliferation [208]. Also acidic FGF (aFGF), alternately known as FGF-1, has been shown to promote extracellular Ca2+ entry in SV40-HCEC [202]. This investigation suggested that SV40-HCEC express L-type voltage-gated Ca2+ channels (VGCC) and that voltage-gated Ca2+ entry is induced by aFGF via FGFR-1 activation. However, pharmacological blockade of L-type VGCC also leads to SOCE inhibition, so that the exact mechanism(s) responsible for the Ca2+ response to aFGF remains unclear.
Early studies revealed that EGF is able to trigger a biphasic increase in [Ca2+]i in HUVEC [199] and intracellular Ca2+ oscillations in rat cardiac microvascular endothelial cells (CMEC) [209], while it fails to elevate the [Ca2+]i in ovine uterine endothelial cells [173]. Pharmacological manipulation demonstrated that EGF-induced repetitive Ca2+ spikes are patterned by the dynamic interplay of InsP3-induced ER Ca2+ release and SOCE, while RyR are not involved (Table 1). This investigation suggested that the oscillatory Ca2+ signal is required for rat CMEC proliferation as each cytosolic Ca2+ spike causes an increase in nuclear Ca2+ concentration [209]. As in other cell types, the endothelial Ca2+ response to EGF is likely to be mediated by Erb1, the classical EGF receptor (EGFR) [210]. PDGF, in turn, has been shown to induce multiple intracellular Ca2+ waveforms in PAEC: either a monophasic increase in [Ca2+]i, a biphasic Ca2+ signal or repetitive Ca2+ spikes [211]. PDGF-induced intracellular Ca2+ signals are initiated by PDGF β-receptor and are maintained by extracellular Ca2+ entry [211]. Nevertheless, the molecular machinery responsible for the Ca2+ response to PDGF remains to be elucidated. Finally, the chemokine SDF-1α plays a crucial role during angiogenesis (and vasculogenesis) by inducing proliferation and migration in vascular endothelial cells through the Gi-protein coupled receptor C-X-C chemokine receptor type 4 (CXCR4) [212]. In addition, SDF-1α promotes reendothelialization of denudated arteries upon vascular injury [213]. SDF-1α has long been known to induce biphasic Ca2+ signals in HUVEC [214] and HRMEC [215]. Similar to PDGF, it is still unclear how SDF-1α increases the [Ca2+]i, although it has been shown that extracellular Ca2+ entry is involved in the onset of the signal and in SDF-1α -induced migration in HRMEC [215]. Moreover, the Ca2+ response to SDF-1α is finely tuned by BKCa channels, which support SDF-1α -induced proliferation and migration in HUVEC [67].

3.6. Endothelial Ca2+ Signals Induced by Angiopoietins (ANG)

While growth factors, such as VEGF, bFGF and PDGF, are required to promote sprouting angiogenesis, the subsequent maturation step is finely tuned by ANG [7]. The human ANG family comprises of three ligands, ANG-1, ANG-2 and ANG-4, and two receptors, Tie1 and Tie2. ANG-1, which is the physiological ligand of the RTK Tie2, maintains endothelial quiescence and prevents apoptosis; in addition, ANG-1 favors vessel stabilization by recruiting pericytes and smooth muscle cells and inducing the basement membrane deposition [216]. ANG-2, in turn, may function as a competitive antagonist or agonist of ANG-1 in a context-dependent manner. For instance, ANG-2 interferes with ANG-1-Tie signaling in resting endothelium to destabilize the vasculature, while it stimulates Tie2 in already activated or stressed endothelial cells [216]. Finally, Tie1 is an orphan receptor which could reduce Tie2 phosphorylation and downstream signaling [7]. A recent investigation revealed that both ANG-1 and ANG-2 induce biphasic Ca2+ signals in HUVEC by selectively promoting ER Ca2+ release through InsP3R and RyR (Table 1) [217]. Of note, ANG-1-induced cell migration is sustained by InsP3R, while ANG-2 requires both InsP3R and RyR. However, while ANG-1 induces in vitro tubulogenesis in a Ca2+-dependent manner, intracellular Ca2+ signaling is dispensable for ANG-2-induced tube formation [217]. Intriguingly, the lysosomal Ca2+ pool, which is involved in VEGF-induced Ca2+ signals [104], is not involved in the Ca2+ response to ANG-1 and ANG-2 [217]. It is, therefore, evident that distinct components of the endothelial Ca2+ toolkit may be recruited by the pro-angiogenic cues that regulate different steps of the angiogenic process. However, while an increase in [Ca2+]i underlies the angiogenic effect of ANG, Tie2 activity is negatively modulated in a Ca2+-dependent manner. Accordingly, pre-treating HUVEC with the Ca2+-ionophore ionomycin promotes calmodulin (CaM) binding to the COOH-terminal tail of Tie2, which results in receptor dephosphorylation and aberrant in vivo angiogenesis [218]. Future work will have to challenge the hypothesis that the Ca2+-dependent regulation of Tie2 activity is accomplished by Ca2+-permeable routes specifically tailored to suit this function.

4. The Ca2+-Dependent Decoders of Angiogenesis

Following the increase in [Ca2+]i, multiple Ca2+-dependent decoders translate endothelial Ca2+ signals into a pro-angiogenic outcome. These include, but are not limited to (Figure 2, Figure 3 and Figure 4): ERK 1/2 [77,104,219], PI3K/Akt [37,104,177], Ca2+/CaM-dependent protein kinase II (CaMKII) [220,221], calpain [222], myosin light chain kinase (MLCK) [25,223], Pyk2 [59,224], eNOS [73,104,225,226], and the transcription factors cAMP responsive element binding protein (CREB), nuclear factor of activated T-cells (NFAT) and nuclear factor kappa enhancer binding protein (NF-κB) [25,45,58,227,228,229]. VEGF has been the most frequent growth factor exploited to investigate how endothelial Ca2+ signaling regulates angiogenesis [24]. However, recent reports demonstrated that also non-canonical WNT signaling stimulates vessel remodeling in a Ca2+-dependent manner [230,231].

4.1. The ERK 1/2 Pathway

The ERK 1/2 phosphorylation cascade represents the most widespread signal transduction pathway whereby VEGF stimulates endothelial cell proliferation, migration and survival, promotes vascular homeostasis and regulates arterial specification [24]. Early work demonstrated that VEGF-induced increase in endothelial [Ca2+]i stimulates the Ca2+-dependent PKCβ2 (cPKCβ2), which in turn recruits the downstream RAF1–MEK–ERK1/2 cascade more efficiently than Ras (Figure 2) [24,232,233]. A number of subsequent reports confirmed that VEGF requires intracellular Ca2+ signaling to engage the ERK1/2 pathway in vascular endothelial cells [26,77,104,171,198,199,234]. The Ca2+-dependent recruitment of the ERK 1/2 pathway may be sustained by endogenous Ca2+ release through InsP3R [26,177] and TPC2 [104] and by extracellular Ca2+ entry through TRPC3 and NCX1 [77], TRPC4 [171] and the SOCE pathway [26].

4.2. The PI3K/Akt Pathway

The serine/threonine Akt1-4 kinases regulate cell proliferation, survival and resistance to apoptosis. More specifically, Akt1 is necessary to promote adult and pathological angiogenesis as well as to regulate vascular development and metabolism [24,235]. Canonically, VEGFR-2 activates PI3K via either Src and vascular endothelial cadherin [236] or the RTK Axl [237], thereby generating phosphatidylinositol-3,4,5-trisphosphate (PIP3), the lipid second messenger responsible for Akt activation. However, VEGF may engage the PI3K/Akt pathway also through an increase in [Ca2+]i, as demonstrated in HUVEC [104] and HCAEC [177]. Notably, TPC2 [104], InsP3R [170,177] and TRPC4 [171] may deliver the Ca2+ signal required for PI3K/Akt activation by VEGF. As discussed in [41], endothelial Ca2+ signaling could activate the small GTPase Ras, which in turn recruits PI3K (as well as ERK1/2) [238]. Furthermore, an increase in [Ca2+]i could directly recruit the PI3K class II α-isoform [239]. However, the Ca2+-sensitive decoder coupling endothelial Ca2+ signaling to PI3K activations remains to be elucidated.

4.3. Calcineurin and NFAT

VEGF-induced endothelial Ca2+ signaling may also recruit the Ca2+-sensitive transcription factor, NFAT (Figure 2), which regulates a transcription program crucial for endothelial cell proliferation, migration, and neovessel formation [24,240]. Furthermore, NFAT is indispensable for patterning the developing vasculature [241]. Five NFAT isoforms have been described: NFAT1-NFAT5 [240]. NFAT is selectively activated by store-operated Ca2+ influx through Orai1 channels [242,243]. Orai1-mediated extracellular Ca2+ entry engages the Ca2+/CaM-dependent phosphatase calcineurin, which dephosphorylates multiple phosphoserines in NFAT1-NFAT4 (Figure 2), thereby promoting its nuclear translocation [24,240,242,243]. Accordingly, it has recently been demonstrated that Orai1 stimulates NFAT nuclear translocation also in vascular endothelial cells [244]. A multitude of genes involved in blood vessel formation are under transcriptional control by NFAT [58,245], including regulator of calcineurin 1 (RCAN1) [246] and the transcription factors, early growth response (EGR)-1 [247], EGR-3 [248] and NR4A [245], which could drive secondary gene regulatory events involved in the response to VEGF. Conversely, RCAN1 encodes for a protein known as Down syndrome critical region 1 (DSCR1), which inhibits calcineurin activity [249], thereby hampering VEGF-induced gene expression [176], migration and angiogenesis [250,251]. In addition, DSCR1 finely tunes tubular morphogenesis by promoting VEGFR2 internalization and dampening VEGF-induced cytoskeletal reorganization and cell polarity during endothelial migration [251]. As mentioned in Section 3.3, a recent investigation demonstrated that low nanomolar doses of VEGF stimulate PAEC to undergo proliferation by selectively inducing the nuclear translocation of NFAT2 [25], the NFAT isoform responsible for cell proliferation [252]. This report failed to detect NFAT2 nuclear translocation in migrating cells when VEGF was applied at higher doses and induced a biphasic Ca2+ signal [25]. This finding is somehow unexpected as it has long been known that NFAT2 also controls endothelial cell migration and tube formation [253]. Solving this discrepancy will certainly deserve future studies.

4.4. CaMKII

CaMKII is an established decoder of intracellular Ca2+ signals in brain [254] and heart [255], in which it integrates repetitive oscillations in [Ca2+]i to, respectively, induce long-term potentiation and regulate cardiac contractility. Emerging evidence has postulated a key role for endothelial CaMKII in the control of endothelial cell proliferation, migration and permeability (Figure 3) [256]. Of the four mammalian CaMKII isoforms (α,β, γ, and δ), CaMKIIα and CaMKIIβ are restricted to the brain, while CaMKIIγ and CaMKIIδ are expressed in peripheral tissues and in vascular endothelial cells [256]. Nevertheless, CaMKIIα/β heteromultimers were detected at the myo-endothelial projections of native mouse mesenteric artery endothelial cells, where they could modulate vascular tone by inhibiting InsP3-dependent Ca2+ pulsars [257]. Early work showed that CaMKII mediates VEGF-induced proliferation, migration, tubulogenesis and sprout formation in BREC [170]. Accordingly, a subsequent report confirmed that bFGF, IGF-1, and hepatocyte growth factor (HGF) induce BREC to undergo in vitro angiogenesis by recruiting CaMKII [220]. Of note, CaMKII stimulates BREC by engaging intermediate kinases to phosphorylate multiple effectors, including Akt, JNK, Src and FAK [220]. Finally, genetic deletion of CaMKIIγ and CaMKIIδ interferes with choroidal neovascularization and hypoxia-induced angiogenesis in vivo [220]. Therefore, distinct CaMKII isoforms could fulfill different functions in vascular endothelial cells: CaMKIIα and CaMKIIβ regulate the vascular tone, while CaMKIIγ and CaMKIIδ promote angiogenesis. CaMKII has also been shown to promote coronary angiogenesis by inducing capillary growth following repeated transient ischemia in mice [221]. Hypoxia indeed causes CaMKII activation, which in turn drives mouse CMEC proliferation and migration [221]. The emerging role of the endothelial CaMKII is further highlighted by a recent investigation, which showed that Notch signaling necessary for vascular remodeling and embryonic survival is compromised in a transgenic mice expressing an oxidation-resistant CaMKIIδ and lacking Regulator of G protein signaling 6 [258].

4.5. MLCK, Calpain and Proline-Rich Tyrosine Kinase-2 (Pyk2)

Endothelial Ca2+ signals drive cell migration by regulating the activity of the Ca2+-sensitive effectors MLCK and calpain [259,260,261]. Directional migration of stalk cells towards the source of growth factors requires the establishment of a front-to-rear polarity along their axis of movement. Migrating cells extend spike-like filopodia or broad lamellipodia, which are driven by actin polymerization, at the leading edge; these protrusions are stabilized through adhesion to the extracellular matrix or to adjoining cells via transmembrane receptors coupled to actin cytoskeleton. These adhesions serve as traction points over which cells move in the direction of the chemotactic stimulus. Thereafter, contraction of the actomyosin network causes dismantling of cell adhesions and retraction of the tail at the rear end and pulls the cell body forward [262]. As widely described elsewhere [212,261], an increase in [Ca2+]i is required to activate MLCK, which induces myosin II-based actomyosin contraction, and the Ca2+-dependent protease, calpain [263] that effects the cleavage of focal adhesion proteins, including focal adhesion kinases, integrins, talin and vinculin. In addition, intracellular Ca2+ signals recruit the Ca2+-sensitive tyrosine kinase Pyk2, which sustains cytoskeletal reorganization at nascent adhesion sites and VEGF-induced eNOS activation [59,224]. Surprisingly, VEGF-induced intracellular Ca2+ release is compromised in Pyk2-deficient mouse aortic endothelial cells [59], thereby suggesting that Pyk2-dependent phosphorylation tunes the endothelial Ca2+ toolkit.
Similar to other cell types, migrating endothelial cells present a remarkable Ca2+ gradient along their axis of movement, as cytosolic Ca2+ is higher at the rear edge due to the higher PMCA activity at the leading edge [223,264]. However, local Ca2+ pulses selectively arise at the front rear of migrating endothelial cells uniformly exposed to bFGF, as recently shown in HUVEC [260]; bFGF-induced local Ca2+ flickers are driven by local PLCγ1 activation and InsP3-dependent Ca2+ release followed by SOCE activation (Table 1) and are necessary to engage MLCK (and, probably, calpain) in the front of migrating cells [223,263]. Notably, SOCE activation at the leading edge is boosted by the local decrease in [Ca2+]ER, which favors STIM1 redistribution at ER-plasma membrane junction at the front of migrating cells [223]. These findings are somehow different as compared to those reported in PAEC, in which MLCK is recruited by a biphasic Ca2+ signal [25]. This discrepancy might reflect differences in the vascular bed (umbilical vein vs. aorta) or species (human vs. porcine). Nevertheless, repeated Ca2+ flickers spatially restricted at the rear part of the cell are regarded as a hallmark of migrating cells and could also be sustained by extracellular Ca2+ entry through the stretch-sensitive TRP Melastatin 7 (TRPM7) [259]. Of note, shear stress may induce Ca2+ entry and downstream recruitment of calpain at focal adhesions of migrating HUVEC [263,265], although TRPM7 silencing does not affect migration [266] and calpain activation [267].
Table 1. The endothelial Ca2+ toolkit: Recruitment by pro- and anti-angiogenic cues.
Table 1. The endothelial Ca2+ toolkit: Recruitment by pro- and anti-angiogenic cues.
SignalECInsP3RRyRTPCSTIM1/Ora1TRPC1/TRPC4TRPC3/TRPC6Ref.
VEGFHUVECYesN.I.YesYesYesYes[73,76,77,104,122,123,124,163,164]
VEGFEA.hy926YesN.I.N.I.NoNoNo[165]
VEGFHAECYesYesN.I.YesN.I.N.I.[48]
VEGFZebrafish tip cellsYesN.I.YesN.I.N.I.N.I.[26]
bFGFHUVECYesN.I.N.I.YesN.I.N.I[203,205,223,260]
EGFCMECYesNoN.I.YesN.I.N.I.[209]
ANGHUVECYesYesNoNoNoNo[217]
AngiostatinBAECYesN.I.N.I.N.I.N.I.N.I.[192]
EndostatinBAECYesN.I.N.I.N.I.N.I.N.I.[192]
Abbreviations: BAEC = bovine aortic endothelial cells, CMEC = cardiac microvascular endothelial cells; HAEC = human aortic endothelial cells; HUVEC = human umbilical vein endothelial cells. Only investigations assessing more than one pro-angiogenic Ca2+ entry/release pathway were described.

4.6. eNOS

It has long been known that NO plays a crucial role in VEGF-induced angiogenesis by stimulating endothelial cell proliferation, migration, substrate adhesion, hyperpermeability and tube formation [59,234,268,269]. Endothelial eNOS is mainly sequestered within Ω-shaped invaginations of the plasma membrane, known as caveolae, which are enriched in cholesterol and are coated on their cytosolic surface by caveolin-1 (Cav1) [270]. In quiescent cells, eNOS activation is hindered by the physical interaction with Cav1, which prevents CaM association to the CaM-binding sequence located between the NH2-terminal oxygenase and the inhibitory COOH-terminal reductase domain. However, an increase in [Ca2+]i displaces Cav1 from eNOS, thereby relieving this tonic inhibition and inducing NO release (Figure 4) [270].
Alternately, eNOS activity can be stimulated by phosphorylation of at least three residues (S615, S633, and S1177) located within the autoinhibitory reductase domain, which prevents eNOS activation in the absence of a Ca2+ rise [270]. A number of kinases stimulate eNOS, including Akt, protein kinase A (PKA), PKC, CaMKII, AMP-activated kinase (AMPK), and Pyk2 [59,270,271]. Early work demonstrated that VEGF induces immediate eNOS activation through an increase in [Ca2+]i followed by delayed NO synthesis which requires eNOS phosphorylation by Akt and PKC [225,226]. The heat shock protein 90 is required to recruit Akt to eNOS upon VEGF stimulation and sustain NO release after the increase in [Ca2+]i [225]. In addition, VEGF may phosphorylate eNOS and promote NO-dependent angiogenesis by engaging AMPK. Of note, VEGF-induced AMPK activation is sustained by the Ca2+/CaM-dependent kinase kinase [271]. Whereas the role of Ca2+ in the immediate eNOS activation has been long established, InsP3-dependent ER Ca2+ release has been regarded as the main pathway responsible for VEGF-induced endothelial NO release [225,226,272]. However, VEGF-induced eNOS may also be recruited by Ca2+ release through TPC2 [104] and by SOCE [272], which represents the most widespread pathway to induce endothelial NO release [120].

4.7. Non-Canonical Wnt/Ca2+ Signaling Pathway

Non-canonical Wnt/Ca2+ signaling pathway plays a pivotal role during early development by controlling cell proliferation, migration, polarity and cell fate specification. In addition, dysregulation of the Wnt/Ca2+ signaling pathway may also lead to neoplastic transformation [273]. Non-canonical Wnt ligands, such as Wnt5a, bind to specific cell surface Frizzled (Fz) receptors, i.e., Fz2-Fz6, which are Gq/11PCR able to recruit PLCβ and induce InsP3-dependent Ca2+ release [273]. The following increase in [Ca2+]i stimulates calcineurin and CaMKII to, respectively, engage NFAT and NF-κB, thereby inducing gene expression [273]. Recent investigations provided the evidence that Wnt5a controls vascular morphogenesis by fine-tuning the angiogenic process [230,231,274,275,276]. Wnt5a may be secreted by vascular endothelial cells [276] or by angiogenic myeloid cells, such as monocytes [277] and macrophages [278]. For instance, autocrine Wnt5a stimulates HUVEC to undergo angiogenesis by activating CaMKII [279]. Likewise, monocyte-derived Wnt5a recruits NF-κB in HDMEC to induce migration, tube formation and neovessel formation in vivo [277]. Furthermore, autocrine Wnt5a/Ca2+ signaling induces NFAT-dependent gene expression to protect HUVEC from apoptosis, prevent vascular regression in retina and drive vascular morphogenesis in aorta [231]. Finally, the Wnt signaling mediator, Secreted frizzle-related protein 2 (SFRP2), activates NFAT to promote migration and tube formation in HCAEC and boosts angiogenesis in a mouse model of angiosarcoma [230]. Of note, recombinant Wnt5s induces intracellular Ca2+ signals in HDMEC, although the involvement of InsP3-dependent ER Ca2+ release remains to be elucidated.

5. The Role of Ca2+ Signaling in Vasculogenesis

While the role of endothelial Ca2+ signaling in angiogenesis has long been recognized, recent reports provided the evidence that an increase in [Ca2+]i plays a crucial role also in vasculogenesis [19,41]. It should, however, be pointed out that the term EPC actually encompasses a large repertoire of different cell types that present distinct phenotypes and stimulate angiogenesis through diverse mechanisms [9,10,280]. A recent Consensus Statement on Nomenclature proposed to abandon the term EPC in favor of a terminology that is more appropriate to define the two broad categories of cell types involved in neovessel formation [280]. These include hematopoietic and endothelial progenitors, which are known, respectively, as myeloid angiogenic cells (MAC) and endothelial colony forming cells (ECFC). MAC, also termed circulating angiogenic cells (CAC) or “early” EPC, are actually hematopoietic progenitors which are liberated from the bone marrow and stimulate capillary sprouting in a paracrine manner. Conversely, ECFC, also known as blood outgrowth endothelial cells (BOEC) or late outgrowth EPC, truly belong to the endothelial lineage, reside in vascular stem cell niches, show high clonal potential, form bidimensional capillary networks in vitro and integrate into host vasculature in vivo [280,281,282]. Accordingly, ECFC may be instrumental to vascular reconstruction after an ischemic insult [280,281,282] or mediate the angiogenic switch in growing tumors [283,284]. Therefore, herein we mainly focus on the role of intracellular Ca2+ signaling in driving ECFC’s angiogenic activity. Most of the findings described in the following chapters were carried out in circulating ECFC, i.e., ECFC isolated from peripheral blood, unless otherwise stated.

5.1. The Ca2+ Toolkit in Human ECFC: Endogenous Ca2+ Release and Extracellular Ca2+ Entry

The network of Ca2+-clearing systems that contribute to maintain the resting [Ca2+]i in ECFC and to remove cytosolic Ca2+ upon chemical stimulation is similar to that described in vascular endothelial cells (Figure 5) [19,20,41]. Briefly, an in-depth transcriptomic analysis revealed that circulating ECFC express SERCA2b, PMCA1b and PMCA4b, while they are devoid of the endothelial NCX isoforms, such as NCX1.3 and NCX1.7 [285]. The Ca2+ response to pro-angiogenic growth factors and chemokines is shaped by both endogenous Ca2+ release and extracellular Ca2+ entry (Figure 5). Intracellular Ca2+ release is sustained by ER Ca2+ release through InsP3R1-3 [286], but not RyR [287,288], and EL Ca2+ mobilization through TPC1 [287]. InsP3 is synthesized following PLCβ2 activation [289], while it is unclear which PLCγ isoform is expressed in ECFC. The main pathway responsible for extracellular Ca2+ entry is provided by SOCE [135,288], which is mediated by the interplay between STIM1, Orai1 and TRPC1 [288,290]. Quite surprisingly, circulating ECFC lack DAG-gated TRPC3 and TRPC6 [288], although the former is abundantly expressed in umbilical cord blood (UCB)-derived ECFC [291]. Pro-angiogenic Ca2+ signals may be delivered by TRP Vanilloid 4 (TRPV4) [292], which is gated by arachidonic acid [287]. It is, however, still unknown whether TRPV4 is also engaged by growth factors or chemokines.

5.2. IGF-2 and SDF-1α Stimulate ECFC Homing to Hypoxic Tissues through an Increase in [Ca2+]i

Ischemic tissues liberate multiple chemokines, such as IGF-2 and SDF-1α, which recruit circulating ECFC or promote ECFC mobilization in peripheral circulation and stimulate their physical engraftment within nascent neovessels [289,293]. Early work showed that IGF-2 binds to IGF receptor 2 (IGFR2) to activate PLCβ2 and induce InsP3-depedendent ER Ca2+ release in UCB-derived ECFC [289]. The ensuing increase in [Ca2+]i is necessary to instigate UCB-derived ECFC to migrate and to adhere to a fibronectin matrix in vitro. Furthermore, IGF-2-evoked Ca2+ signaling drives UCB-derived ECFC homing to damaged vessels and engraftment into nascent vasculature in vivo, thereby promoting vascular reconstruction in a murine model of hindlimb ischemia [289]. Likewise, SDF-1α elicits a biphasic increase in [Ca2+]i in circulating ECFC [294,295]. The Ca2+ response to SDF-1α is triggered by CXCR4, initiated by InsP3-dependent ER Ca2+ mobilization and sustained by SOCE. The interaction between InsP3Rs and SOCE recruits the PI3K/Akt and ERK1/2 signaling pathways, which in turn regulate ECFC migration in vitro and neovessel formation in vivo [294].

5.3. VEGF and NAADP Stimulate Proliferation and Tube Formation through an Increase in [Ca2+]i in ECFC

VEGF is massively secreted by ischemic tissues to stimulate local endothelial cells to undergo angiogenesis and circulating ECFC to integrate within the damaged vascular network [296,297]. It has been shown that VEGF binds to VEGFR2 to induce repetitive oscillations in [Ca2+]i in circulating [286] and UCB-derived [291] ECFC. VEGF-induced intracellular Ca2+ oscillations in circulating ECFC are shaped by rhythmic ER Ca2+ release through InsP3Rs followed by SERCA-mediated Ca2+ sequestration into ER lumen. SOCE is required to sustain the spiking response by refilling the ER with Ca2+ during prolonged stimulation [286]. Notably, VEGF-induced intracellular Ca2+ oscillations are more robust and frequent in UCB-derived ECFC [282,291] and this feature could result in their higher sensitivity to VEGF [298]. Unlike their circulating counterparts, UCB-derived ECFC express TRPC3 and TRPC3-mediated extracellular Ca2+ entry triggers the dynamic interplay between InsP3 and SOCE (Figure 5) [291]. It has, therefore, been proposed that TRPC3 over-expression could represent an alternative strategy to rejuvenate the reparative phenotype of aging ECFC and improve the outcome of therapeutic angiogenesis in cardiovascular patients [41,139]. VEGF-induced intracellular Ca2+ oscillations stimulate proliferation and in vitro tubulogenesis by promoting the nuclear translocation of NF-κB [286]. The intermediate Ca2+-sensitive decoder that translates the spiking Ca2+ signal into NF-κB activation is likely to be provided by CaMKII, which controls ECFC proliferation [299] and is sensitive to endothelial Ca2+ oscillations [300]. Notably, VEGF-induced SOCE is dramatically enhanced at day 0, but not at days 7 and 180, from the event in circulating ECFC harvested from peripheral blood of subjects who underwent acute myocardial infarction. Furthermore, ECFC frequency is higher within the first 24 h after an acute myocardial infarction as compared to healthy individuals, although their angiogenic activity is not altered [301]. Therefore, it is tempting to speculate that SOCE up-regulation results in the higher ECFC frequency reported in infarcted patients [301].
In addition to VEGF, exogenous delivery of NAADP through a liposomal formulation has been shown to stimulate ECFC proliferation [302]. NAADP acts by releasing EL Ca2+ through TPC1 (Figure 5) [302,303] and preliminary evidence suggested that lysosomal Ca2+ release could, in turn, be amplified by InsP3R [303]. While this observation is yet to be fully confirmed, the alternative hypothesis that InsP3-dependent Ca2+ release is necessary to refill the lysosomal Ca2+ pool cannot be ruled out [113,116]. In addition, NAADP could interact with InsP3 to initiate the rhythmic Ca2+ response to VEGF also in circulating ECFC.

5.4. VEGF-Induced Intracellular Ca2+ Oscillations Are Down-Regulated in Tumor-Derived ECFC

It has long been known that the Ca2+ toolkit is dramatically remodeled in tumor endothelial cells, thereby exacerbating the angiogenic switch and conferring resistance to anti-cancer treatments, as widely discussed elsewhere [32,304,305,306,307]. Likewise, InsP3-dependent ER Ca2+ release and SOCE are altered in tumor-derived ECFC, although their dysregulation depends on the tumor type [19,20]. For instance, the intracellular Ca2+ response to InsP3 production is impaired in renal cellular carcinoma (RCC)-derived ECFC due to the reduction in [Ca2+]ER and InsP3R transcripts [285,290]. As a consequence, VEGF fails to elicit intracellular Ca2+ signals in RCC-derived ECFC, although VEGFR2 is normally expressed and SOCE is up-regulated due to the over-expression of STIM1, Orai1 and TRPC1 [290,308]. VEGF-induced oscillations in [Ca2+]i are down-regulated and do not stimulate either proliferation or tube formation also in breast cancer (BC)-derived ECFC [309]. As reported in RCC-derived ECFC [290], the ER is less prone to release Ca2+ [309,310], although there is no change in InsP3R expression and SOCE is not impaired in BC-derived ECFC [309]. Furthermore, VEGF does not trigger sizeable Ca2+ signals in infantile hemangioma (IH)-derived ECFC [311], whereas the spiking Ca2+ response arises but does not stimulate any angiogenic activity in primary myelofibrosis (PMF)-derived ECFC [308]. It has, therefore, been suggested that VEGF is actually unable to promote an angiogenic behavior in tumor-derived ECFC due to the derangement of their Ca2+ signaling machinery [20,283,310]. This feature could explain the resistance (primary or secondary) to anti-VEGF drugs that has long been reported in RCC, BC and PMF patients and highlights the urgency of identifying novel targets to interfere with angiogenesis in cancer patients [283,312]. Of note, pharmacological and genetic deletion of SOCE inhibits proliferation and tube formation in RCC- [290], BCC- [309], and IH-derived [311] ECFC. The growth factor(s) responsible for SOCE activation in tumor-derived is (are) still unknown, but it is worth noting that SOCE is constitutively active in IH-derived ECFC due to the partial depletion of the ER Ca2+ content [311].

5.5. The Ca2+ Toolkit in Rodent MAC

The pro-angiogenic role of intracellular Ca2+ signaling has also been evaluated in rat and mouse bone marrow-derived MAC, which stimulate angiogenesis by delivering paracrine signals, but do not physically integrate within neovessels. Early investigations demonstrated that SOCE is abated by the genetic deletion of Stim1 and TRPC1 in rat MAC, whereas the involvement of Orai1 was not investigated [313,314,315]. The genetic and pharmacological ablation of SOCE machinery suppresses proliferation, migration, and in vitro tubulogenesis in rat MAC [313,314,315]. Likewise, SOCE is expressed and drives proliferation, migration and tube formation also in murine MAC, which express STIM1, Orai1 and TRPC1 [316,317]. Although the molecular composition of SOCE was not dissected in this report, SOCE supports VEGF-induced repetitive Ca2+ spikes [316], as observed in human ECFC [286]. Moreover, STIM1, Orai1 and TRPC1 are down-regulated in MAC deriving from atherosclerotic mice; as a consequence, SOCE, VEGF-induced intracellular Ca2+ oscillations, proliferation and migration are attenuated [316]. The Ca2+-dependent decoder of VEGF-induced intracellular Ca2+ oscillations in MAC is represented by eNOS [316,317], although it is likely that future investigations reveal additional Ca2+-sensitive effectors.

6. Conclusions

Endothelial Ca2+ signaling is instrumental in translating pro-angiogenic cues into an effective stimulus to induce proliferation, migration, tube formation and neovessel formation. Nevertheless, the components of the Ca2+ handling machinery underlying the Ca2+ response to growth factors and chemokines may be differently assorted depending on stimulus, vascular bed and species. Shedding light on this crucial aspect of endothelial Ca2+ signaling requires more investigations on growth factors other than VEGF, which has largely been exploited to understand how an increase in [Ca2+]i regulates angiogenesis and vasculogenesis. The available evidence hints at InsP3Rs and SOCE as the major driver of pro-angiogenic Ca2+ signals in both vascular endothelial cells and ECFC/MAC. Nevertheless, the role of NAADP and TPC1-2 is rapidly emerging and it remains to understand whether extracellular Ca2+ entry through DAG-gated channels, i.e., TRPC3 and TRPC6, during the plateau phase observed in some cell types, such as HUVEC, is merely redundant or is required to recruit Ca2+-sensitive decoders that are not targeted by SOCE. An additional feature that should be taken into account is the dose-response relationship of each growth factor as the Ca2+ signature and Ca2+-dependent pro-angiogenic activity could vary depending on the dose, as reported for VEGF in PAEC.
Understanding how endothelial Ca2+ signaling controls angiogenesis and vasculogenesis is indispensable to design alternative strategies to induce therapeutic angiogenesis in ischemic disorders and interfere with the vascular network in solid malignancies. For instance, it has been demonstrated that genetic or pharmacological manipulation of the endothelial Ca2+ toolkit improves angiogenesis in vivo. VEGF-induced extracellular Ca2+ influx was enhanced by engineering HDMEC with an adenovirus encoding for TRPC6, thereby enhancing their proliferation rate [74]. Similarly, hypoxic preconditioning increased TRPC4 levels and boosted migration in human pulmonary artery endothelial cells [318]. These pieces of evidence suggest that it is therapeutically feasible to exploit endothelial Ca2+ signalling to stimulate therapeutic angiogenesis. Of note, it has been suggested that autologous or endogenous ECFCs, which represent the most suitable EPC subtype for regenerative therapy of ischemic disorders, could be redirected towards a more reparative phenotype by intervening on their Ca2+ toolkit. It has been suggested that ECFCs’ angiogenic activity could be improved by lentivirus-mediated expression of TRPC3, subsequent expansion in vitro and reinoculation into the ischemic tissue [139,319]. Moreover, a recent investigation demonstrated that intramyocardial injection of the secretome collected from hypoxic UCB-derived mesenchymal stem cells induced angiogenesis by promoting intracellular Ca2+ oscillations in resident ECFC in a murine model of AMI [320]. This evidence provided the proof of concept that it is therapeutically feasible to induce cardiac revascularization and partially rescue cardiac function by exploiting intracellular Ca2+ signalling in resident ECFC. An alternative, but not mutually exclusive mechanism, consists in taking advantage from small molecule drug to stimulate angiogenesis by enhancing endothelial Ca2+ signaling. For instance, inhibiting PMCA4 with the small molecule aurintricarboxylic acid enhanced VEGF-induced calcineurin activation, thereby boosting HUVEC migration and tube formation as well as neovessel formation in a murine model of hindlimb ischemia [45]. Conversely, the pharmacological blockade of Orai1 with carboxyamidotriazole, which targets both vascular endothelial cells [73] and tumor-derived ECFC [290], has been probed in phase I-III clinical trials launched towards several malignancies [304,312]. Although more specific Orai1 blockers are certainly required to avoid off-target effects, targeting endothelial SOCE could represent an alternative strategy to fight cancer. Likewise, the pharmacological blockade of TPC1 with NED-19 is coming of age as an efficient means to retard tumor vascularization [104,321]. Alternately, it has been suggested that endothelial Ca2+ signaling could be exploited to favor normalization of tumor vasculature [304], thereby improving the therapeutic outcome of classical anticancer treatments, such as radiotherapy and chemotherapy. For instance, stimulation of purinergic P2X7 and P2Y11 receptors inhibits migration in breast tumor-derived endothelial cells and induces vessel normalization in a Ca2+-dependent manner [33].

Author Contributions

Writing—original draft preparation, F.M.; writing—review and editing, S.N., M.S., P.F., G.G.; supervision, F.M., G.G.

Funding

This research was funded by the Italian Ministry of Education, University and Research (MIUR): Dipartimenti di Eccellenza Program (2018–2022)—Dept. of Biology and Biotechnology "L. Spallanzani", University of Pavia (F.M.), Fondo Ricerca Giovani from the University of Pavia (F.M.), and the EU Horizon 2020 FETOPEN-2018-2020 Program under Grant Agreement N. 828984 (LION-HEARTED).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Herbert, S.P.; Stainier, D.Y. Molecular control of endothelial cell behaviour during blood vessel morphogenesis. Nat. Rev. Mol. Cell Biol. 2011, 12, 551–564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Ramasamy, S.K.; Kusumbe, A.P.; Adams, R.H. Regulation of tissue morphogenesis by endothelial cell-derived signals. Trends Cell Biol. 2015, 25, 148–157. [Google Scholar] [CrossRef] [PubMed]
  3. Aird, W.C. Spatial and temporal dynamics of the endothelium. J. Thromb. Haemost. 2005, 3, 1392–1406. [Google Scholar] [CrossRef] [PubMed]
  4. Galley, H.F.; Webster, N.R. Physiology of the endothelium. Br. J. Anaesth. 2004, 93, 105–113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Cahill, P.A.; Redmond, E.M. Vascular endothelium—Gatekeeper of vessel health. Atherosclerosis 2016, 248, 97–109. [Google Scholar] [CrossRef]
  6. Rafii, S.; Butler, J.M.; Ding, B.S. Angiocrine functions of organ-specific endothelial cells. Nature 2016, 529, 316–325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Carmeliet, P.; Jain, R.K. Molecular mechanisms and clinical applications of angiogenesis. Nature 2011, 473, 298–307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Ribatti, D.; Crivellato, E. “Sprouting angiogenesis”, a reappraisal. Dev. Biol. 2012, 372, 157–165. [Google Scholar] [CrossRef]
  9. Yoder, M.C. Human endothelial progenitor cells. Cold Spring Harb. Perspect. Med. 2012, 2, a006692. [Google Scholar] [CrossRef]
  10. Yoder, M.C. Endothelial stem and progenitor cells (stem cells): (2017 Grover Conference Series). Pulm. Circ. 2018, 8. [Google Scholar] [CrossRef]
  11. D’Alessio, A.; Moccia, F.; Li, J.H.; Micera, A.; Kyriakides, T.R. Angiogenesis and Vasculogenesis in Health and Disease. BioMed Res. Int. 2015, 2015, 126582. [Google Scholar] [CrossRef] [PubMed]
  12. Wietecha, M.S.; Cerny, W.L.; DiPietro, L.A. Mechanisms of vessel regression: Toward an understanding of the resolution of angiogenesis. Curr. Top. Microbiol. Immunol. 2013, 367, 3–32. [Google Scholar] [CrossRef] [PubMed]
  13. Fischer, C.; Schneider, M.; Carmeliet, P. Principles and therapeutic implications of angiogenesis, vasculogenesis and arteriogenesis. In The Vascular Endothelium II, Handbook of Experimental Pharmacology; Springer: Berlin/Heidelberg, Germany, 2006; Volume 176, pp. 157–212. [Google Scholar]
  14. Potente, M.; Gerhardt, H.; Carmeliet, P. Basic and therapeutic aspects of angiogenesis. Cell 2011, 146, 873–887. [Google Scholar] [CrossRef] [PubMed]
  15. Moccia, F.; Lodola, F.; Dragoni, S.; Bonetti, E.; Bottino, C.; Guerra, G.; Laforenza, U.; Rosti, V.; Tanzi, F. Ca2+ signalling in endothelial progenitor cells: A novel means to improve cell-based therapy and impair tumour vascularisation. Curr. Vasc. Pharmacol. 2014, 12, 87–105. [Google Scholar] [CrossRef] [PubMed]
  16. Moccia, F.; Tanzi, F.; Munaron, L. Endothelial remodelling and intracellular calcium machinery. Curr. Mol. Med. 2014, 14, 457–480. [Google Scholar] [CrossRef] [PubMed]
  17. Munaron, L. Intracellular calcium, endothelial cells and angiogenesis. Recent Pat. Anticancer Drug Discov. 2006, 1, 105–119. [Google Scholar] [CrossRef] [PubMed]
  18. Moccia, F.; Berra-Romani, R.; Tanzi, F. Update on vascular endothelial Ca2+ signalling: A tale of ion channels, pumps and transporters. World J. Biol. Chem. 2012, 3, 127–158. [Google Scholar] [CrossRef]
  19. Moccia, F.; Guerra, G. Ca2+ Signalling in Endothelial Progenitor Cells: Friend or Foe? J. Cell. Physiol. 2016, 231, 314–327. [Google Scholar] [CrossRef]
  20. Moccia, F.; Poletto, V. May the remodeling of the Ca2+ toolkit in endothelial progenitor cells derived from cancer patients suggest alternative targets for anti-angiogenic treatment? Biochim. Biophys. Acta 2015, 1853, 1958–1973. [Google Scholar] [CrossRef]
  21. Smani, T.; Gomez, L.J.; Regodon, S.; Woodard, G.E.; Siegfried, G.; Khatib, A.M.; Rosado, J.A. TRP Channels in Angiogenesis and Other Endothelial Functions. Front. Physiol. 2018, 9, 1731. [Google Scholar] [CrossRef]
  22. Thakore, P.; Earley, S. Transient Receptor Potential Channels and Endothelial Cell Calcium Signaling. Compr. Physiol. 2019, 9, 1249–1277. [Google Scholar] [CrossRef] [PubMed]
  23. Kohn, E.C.; Alessandro, R.; Spoonster, J.; Wersto, R.P.; Liotta, L.A. Angiogenesis: Role of calcium-mediated signal transduction. Proc. Natl. Acad. Sci. USA 1995, 92, 1307–1311. [Google Scholar] [CrossRef] [PubMed]
  24. Simons, M.; Gordon, E.; Claesson-Welsh, L. Mechanisms and regulation of endothelial VEGF receptor signalling. Nat. Rev. Mol. Cell Biol. 2016, 17, 611–625. [Google Scholar] [CrossRef] [PubMed]
  25. Noren, D.P.; Chou, W.H.; Lee, S.H.; Qutub, A.A.; Warmflash, A.; Wagner, D.S.; Popel, A.S.; Levchenko, A. Endothelial cells decode VEGF-mediated Ca2+ signaling patterns to produce distinct functional responses. Sci. Signal. 2016, 9, ra20. [Google Scholar] [CrossRef] [PubMed]
  26. Savage, A.M.; Kurusamy, S.; Chen, Y.; Jiang, Z.; Chhabria, K.; MacDonald, R.B.; Kim, H.R.; Wilson, H.L.; van Eeden, F.J.M.; Armesilla, A.L.; et al. Tmem33 is essential for VEGF-mediated endothelial calcium oscillations and angiogenesis. Nat. Commun. 2019, 10, 732. [Google Scholar] [CrossRef] [PubMed]
  27. Patton, A.M.; Kassis, J.; Doong, H.; Kohn, E.C. Calcium as a molecular target in angiogenesis. Curr. Pharm. Des. 2003, 9, 543–551. [Google Scholar] [CrossRef]
  28. Troidl, C.; Nef, H.; Voss, S.; Schilp, A.; Kostin, S.; Troidl, K.; Szardien, S.; Rolf, A.; Schmitz-Rixen, T.; Schaper, W.; et al. Calcium-dependent signalling is essential during collateral growth in the pig hind limb-ischemia model. J. Mol. Cell. Cardiol. 2010, 49, 142–151. [Google Scholar] [CrossRef]
  29. Troidl, C.; Troidl, K.; Schierling, W.; Cai, W.J.; Nef, H.; Mollmann, H.; Kostin, S.; Schimanski, S.; Hammer, L.; Elsasser, A.; et al. Trpv4 induces collateral vessel growth during regeneration of the arterial circulation. J. Cell. Mol. Med. 2009, 13, 2613–2621. [Google Scholar] [CrossRef]
  30. Andrikopoulos, P.; Kieswich, J.; Harwood, S.M.; Baba, A.; Matsuda, T.; Barbeau, O.; Jones, K.; Eccles, S.A.; Yaqoob, M.M. Endothelial Angiogenesis and Barrier Function in Response to Thrombin Require Ca2+ Influx through the Na+/Ca2+ Exchanger. J. Biol. Chem. 2015, 290, 18412–18428. [Google Scholar] [CrossRef]
  31. Kini, V.; Chavez, A.; Mehta, D. A new role for PTEN in regulating transient receptor potential canonical channel 6-mediated Ca2+ entry, endothelial permeability, and angiogenesis. J. Biol. Chem. 2010, 285, 33082–33091. [Google Scholar] [CrossRef]
  32. Scarpellino, G.; Genova, T.; Avanzato, D.; Bernardini, M.; Bianco, S.; Petrillo, S.; Tolosano, E.; de Almeida Vieira, J.R.; Bussolati, B.; Fiorio Pla, A.; et al. Purinergic Calcium Signals in Tumor-Derived Endothelium. Cancers 2019, 11, 766. [Google Scholar] [CrossRef] [PubMed]
  33. Avanzato, D.; Genova, T.; Fiorio Pla, A.; Bernardini, M.; Bianco, S.; Bussolati, B.; Mancardi, D.; Giraudo, E.; Maione, F.; Cassoni, P.; et al. Activation of P2X7 and P2Y11 purinergic receptors inhibits migration and normalizes tumor-derived endothelial cells via cAMP signaling. Sci. Rep. 2016, 6, 32602. [Google Scholar] [CrossRef] [PubMed]
  34. Lyubchenko, T.; Woodward, H.; Veo, K.D.; Burns, N.; Nijmeh, H.; Liubchenko, G.A.; Stenmark, K.R.; Gerasimovskaya, E.V. P2Y1 and P2Y13 purinergic receptors mediate Ca2+ signaling and proliferative responses in pulmonary artery vasa vasorum endothelial cells. Am. J. Physiol. Cell Physiol. 2011, 300, C266–C275. [Google Scholar] [CrossRef] [PubMed]
  35. Li, X.W.; Wang, H. Non-neuronal nicotinic alpha 7 receptor, a new endothelial target for revascularization. Life Sci. 2006, 78, 1863–1870. [Google Scholar] [CrossRef] [PubMed]
  36. Egleton, R.D.; Brown, K.C.; Dasgupta, P. Angiogenic activity of nicotinic acetylcholine receptors: Implications in tobacco-related vascular diseases. Pharmacol. Ther. 2009, 121, 205–223. [Google Scholar] [CrossRef] [PubMed]
  37. Yu, Y.B.; Su, K.H.; Kou, Y.R.; Guo, B.C.; Lee, K.I.; Wei, J.; Lee, T.S. Role of transient receptor potential vanilloid 1 in regulating erythropoietin-induced activation of endothelial nitric oxide synthase. Acta Physiol. 2017, 219, 465–477. [Google Scholar] [CrossRef] [PubMed]
  38. Maltaneri, R.E.; Schiappacasse, A.; Chamorro, M.E.; Nesse, A.B.; Vittori, D.C. Participation of membrane calcium channels in erythropoietin-induced endothelial cell migration. Eur. J. Cell Biol. 2018, 97, 411–421. [Google Scholar] [CrossRef]
  39. Berra-Romani, R.; Raqeeb, A.; Avelino-Cruz, J.E.; Moccia, F.; Oldani, A.; Speroni, F.; Taglietti, V.; Tanzi, F. Ca2+ signaling in injured in situ endothelium of rat aorta. Cell Calcium 2008, 44, 298–309. [Google Scholar] [CrossRef]
  40. Berra-Romani, R.; Raqeeb, A.; Torres-Jácome, J.; Guzman-Silva, A.; Guerra, G.; Tanzi, F.; Moccia, F. The mechanism of injury-induced intracellular calcium concentration oscillations in the endothelium of excised rat aorta. J. Vasc. Res. 2012, 49, 65–76. [Google Scholar] [CrossRef]
  41. Moccia, F.; Berra-Romani, R.; Rosti, V. Manipulating Intracellular Ca2+ Signals to Stimulate Therapeutic Angiogenesis in Cardiovascular Disorders. Curr. Pharm. Biotechnol. 2018, 19, 686–699. [Google Scholar] [CrossRef]
  42. Berridge, M.J.; Bootman, M.D.; Roderick, H.L. Calcium signalling: Dynamics, homeostasis and remodelling. Nat. Rev. Mol. Cell Biol. 2003, 4, 517–529. [Google Scholar] [CrossRef] [PubMed]
  43. Moccia, F.; Berra-Romani, R.; Baruffi, S.; Spaggiari, S.; Signorelli, S.; Castelli, L.; Magistretti, J.; Taglietti, V.; Tanzi, F. Ca2+ uptake by the endoplasmic reticulum Ca2+-ATPase in rat microvascular endothelial cells. Biochem. J. 2002, 364, 235–244. [Google Scholar] [CrossRef] [PubMed]
  44. Berra-Romani, R.; Raqeeb, A.; Guzman-Silva, A.; Torres-Jacome, J.; Tanzi, F.; Moccia, F. Na+-Ca2+ exchanger contributes to Ca2+ extrusion in ATP-stimulated endothelium of intact rat aorta. Biochem. Biophys. Res. Commun. 2010, 395, 126–130. [Google Scholar] [CrossRef]
  45. Kurusamy, S.; Lopez-Maderuelo, D.; Little, R.; Cadagan, D.; Savage, A.M.; Ihugba, J.C.; Baggott, R.R.; Rowther, F.B.; Martinez-Martinez, S.; Arco, P.G.; et al. Selective inhibition of plasma membrane calcium ATPase 4 improves angiogenesis and vascular reperfusion. J. Mol. Cell. Cardiol. 2017, 109, 38–47. [Google Scholar] [CrossRef]
  46. Baggott, R.R.; Alfranca, A.; Lopez-Maderuelo, D.; Mohamed, T.M.; Escolano, A.; Oller, J.; Ornes, B.C.; Kurusamy, S.; Rowther, F.B.; Brown, J.E.; et al. Plasma membrane calcium ATPase isoform 4 inhibits vascular endothelial growth factor-mediated angiogenesis through interaction with calcineurin. Arterioscler. Thromb. Vasc. Biol. 2014, 34, 2310–2320. [Google Scholar] [CrossRef] [PubMed]
  47. Evangelista, A.M.; Thompson, M.D.; Bolotina, V.M.; Tong, X.; Cohen, R.A. Nox4-and Nox2-dependent oxidant production is required for VEGF-induced SERCA cysteine-674 S-glutathiolation and endothelial cell migration. Free Radic. Biol. Med. 2012, 53, 2327–2334. [Google Scholar] [CrossRef]
  48. Evangelista, A.M.; Thompson, M.D.; Weisbrod, R.M.; Pimental, D.R.; Tong, X.; Bolotina, V.M.; Cohen, R.A. Redox regulation of SERCA2 is required for vascular endothelial growth factor-induced signaling and endothelial cell migration. Antioxid. Redox Signal. 2012, 17, 1099–1108. [Google Scholar] [CrossRef] [PubMed]
  49. Mei, Y.; Thompson, M.D.; Shiraishi, Y.; Cohen, R.A.; Tong, X. Sarcoplasmic/endoplasmic reticulum Ca2+ ATPase C674 promotes ischemia-and hypoxia-induced angiogenesis via coordinated endothelial cell and macrophage function. J. Mol. Cell. Cardiol. 2014, 76, 275–282. [Google Scholar] [CrossRef]
  50. Thompson, M.D.; Mei, Y.; Weisbrod, R.M.; Silver, M.; Shukla, P.C.; Bolotina, V.M.; Cohen, R.A.; Tong, X. Glutathione adducts on sarcoplasmic/endoplasmic reticulum Ca2+ ATPase Cys-674 regulate endothelial cell calcium stores and angiogenic function as well as promote ischemic blood flow recovery. J. Biol. Chem. 2014, 289, 19907–19916. [Google Scholar] [CrossRef]
  51. Malli, R.; Frieden, M.; Osibow, K.; Graier, W.F. Mitochondria efficiently buffer subplasmalemmal Ca2+ elevation during agonist stimulation. J. Biol. Chem. 2003, 278, 10807–10815. [Google Scholar] [CrossRef]
  52. Malli, R.; Frieden, M.; Osibow, K.; Zoratti, C.; Mayer, M.; Demaurex, N.; Graier, W.F. Sustained Ca2+ transfer across mitochondria is Essential for mitochondrial Ca2+ buffering, sore-operated Ca2+ entry, and Ca2+ store refilling. J. Biol. Chem. 2003, 278, 44769–44779. [Google Scholar] [CrossRef] [PubMed]
  53. Malli, R.; Frieden, M.; Trenker, M.; Graier, W.F. The role of mitochondria for Ca2+ refilling of the endoplasmic reticulum. J. Biol. Chem. 2005, 280, 12114–12122. [Google Scholar] [CrossRef] [PubMed]
  54. McCarron, J.G.; Wilson, C.; Heathcote, H.R.; Zhang, X.; Buckley, C.; Lee, M.D. Heterogeneity and emergent behaviour in the vascular endothelium. Curr. Opin. Pharmacol. 2019, 45, 23–32. [Google Scholar] [CrossRef] [PubMed]
  55. McCarron, J.G.; Lee, M.D.; Wilson, C. The Endothelium Solves Problems That Endothelial Cells Do Not Know Exist. Trends Pharmacol. Sci. 2017, 38, 322–338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Beziau, D.M.; Toussaint, F.; Blanchette, A.; Dayeh, N.R.; Charbel, C.; Tardif, J.C.; Dupuis, J.; Ledoux, J. Expression of phosphoinositide-specific phospholipase C isoforms in native endothelial cells. PLoS ONE 2015, 10, e0123769. [Google Scholar] [CrossRef] [PubMed]
  57. Lo Vasco, V.R.; Pacini, L.; Di Raimo, T.; D’Arcangelo, D.; Businaro, R. Expression of phosphoinositide-specific phospholipase C isoforms in human umbilical vein endothelial cells. J. Clin. Pathol. 2011, 64, 911–915. [Google Scholar] [CrossRef] [PubMed]
  58. Rozen, E.J.; Roewenstrunk, J.; Barallobre, M.J.; Di Vona, C.; Jung, C.; Figueiredo, A.F.; Luna, J.; Fillat, C.; Arbones, M.L.; Graupera, M.; et al. DYRK1A Kinase Positively Regulates Angiogenic Responses in Endothelial Cells. Cell Rep. 2018, 23, 1867–1878. [Google Scholar] [CrossRef] [Green Version]
  59. Matsui, A.; Okigaki, M.; Amano, K.; Adachi, Y.; Jin, D.; Takai, S.; Yamashita, T.; Kawashima, S.; Kurihara, T.; Miyazaki, M.; et al. Central role of calcium-dependent tyrosine kinase PYK2 in endothelial nitric oxide synthase-mediated angiogenic response and vascular function. Circulation 2007, 116, 1041–1051. [Google Scholar] [CrossRef]
  60. Suh, P.G.; Park, J.I.; Manzoli, L.; Cocco, L.; Peak, J.C.; Katan, M.; Fukami, K.; Kataoka, T.; Yun, S.; Ryu, S.H. Multiple roles of phosphoinositide-specific phospholipase C isozymes. BMB Rep. 2008, 41, 415–434. [Google Scholar] [CrossRef]
  61. Rhee, S.G. Regulation of phosphoinositide-specific phospholipase C. Annu. Rev. Biochem. 2001, 70, 281–312. [Google Scholar] [CrossRef]
  62. You, J.; Peng, W.; Lin, X.; Huang, Q.L.; Lin, J.Y. PLC/CAMK IV-NF-kappaB involved in the receptor for advanced glycation end products mediated signaling pathway in human endothelial cells. Mol. Cell. Endocrinol. 2010, 320, 111–117. [Google Scholar] [CrossRef] [PubMed]
  63. Zou, Q.Y.; Zhao, Y.J.; Zhou, C.; Liu, A.X.; Zhong, X.Q.; Yan, Q.; Li, Y.; Yi, F.X.; Bird, I.M.; Zheng, J. G Protein alpha Subunit 14 Mediates Fibroblast Growth Factor 2-Induced Cellular Responses in Human Endothelial Cells. J. Cell. Physiol. 2019, 234, 10184–10195. [Google Scholar] [CrossRef] [PubMed]
  64. Korybalska, K.; Rutkowski, R.; Luczak, J.; Czepulis, N.; Karpinski, K.; Witowski, J. The role of purinergic P2Y12 receptor blockers on the angiogenic properties of endothelial cells: An in vitro study. J. Physiol. Pharmacol. 2018, 69. [Google Scholar] [CrossRef]
  65. Gunduz, D.; Tanislav, C.; Schluter, K.D.; Schulz, R.; Hamm, C.; Aslam, M. Effect of ticagrelor on endothelial calcium signalling and barrier function. Thromb. Haemost. 2017, 117, 371–381. [Google Scholar] [CrossRef] [PubMed]
  66. Mehta, D.; Konstantoulaki, M.; Ahmmed, G.U.; Malik, A.B. Sphingosine 1-phosphate-induced mobilization of intracellular Ca2+ mediates rac activation and adherens junction assembly in endothelial cells. J. Biol. Chem. 2005, 280, 17320–17328. [Google Scholar] [CrossRef] [PubMed]
  67. Kuhlmann, C.R.; Schaefer, C.A.; Reinhold, L.; Tillmanns, H.; Erdogan, A. Signalling mechanisms of SDF-induced endothelial cell proliferation and migration. Biochem. Biophys. Res. Commun. 2005, 335, 1107–1114. [Google Scholar] [CrossRef] [PubMed]
  68. Meyer, R.D.; Latz, C.; Rahimi, N. Recruitment and activation of phospholipase Cgamma1 by vascular endothelial growth factor receptor-2 are required for tubulogenesis and differentiation of endothelial cells. J. Biol. Chem. 2003, 278, 16347–16355. [Google Scholar] [CrossRef]
  69. Kunze, K.; Spieker, T.; Gamerdinger, U.; Nau, K.; Berger, J.; Dreyer, T.; Sindermann, J.R.; Hoffmeier, A.; Gattenlohner, S.; Brauninger, A. A recurrent activating PLCG1 mutation in cardiac angiosarcomas increases apoptosis resistance and invasiveness of endothelial cells. Cancer Res. 2014, 74, 6173–6183. [Google Scholar] [CrossRef]
  70. Bhattacharya, R.; Kwon, J.; Li, X.; Wang, E.; Patra, S.; Bida, J.P.; Bajzer, Z.; Claesson-Welsh, L.; Mukhopadhyay, D. Distinct role of PLCbeta3 in VEGF-mediated directional migration and vascular sprouting. J. Cell Sci. 2009, 122, 1025–1034. [Google Scholar] [CrossRef]
  71. Hoeppner, L.H.; Phoenix, K.N.; Clark, K.J.; Bhattacharya, R.; Gong, X.; Sciuto, T.E.; Vohra, P.; Suresh, S.; Bhattacharya, S.; Dvorak, A.M.; et al. Revealing the role of phospholipase Cbeta3 in the regulation of VEGF-induced vascular permeability. Blood 2012, 120, 2167–2173. [Google Scholar] [CrossRef]
  72. Ha, J.M.; Baek, S.H.; Kim, Y.H.; Jin, S.Y.; Lee, H.S.; Kim, S.J.; Shin, H.K.; Lee, D.H.; Song, S.H.; Kim, C.D.; et al. Regulation of retinal angiogenesis by phospholipase C-beta3 signaling pathway. Exp. Mol. Med. 2016, 48, e240. [Google Scholar] [CrossRef]
  73. Faehling, M.; Kroll, J.; Fohr, K.J.; Fellbrich, G.; Mayr, U.; Trischler, G.; Waltenberger, J. Essential role of calcium in vascular endothelial growth factor A-induced signaling: Mechanism of the antiangiogenic effect of carboxyamidotriazole. FASEB J. 2002, 16, 1805–1807. [Google Scholar] [CrossRef] [PubMed]
  74. Hamdollah Zadeh, M.A.; Glass, C.A.; Magnussen, A.; Hancox, J.C.; Bates, D.O. VEGF-mediated elevated intracellular calcium and angiogenesis in human microvascular endothelial cells in vitro are inhibited by dominant negative TRPC6. Microcirculation 2008, 15, 605–614. [Google Scholar] [CrossRef] [PubMed]
  75. Cheng, H.W.; James, A.F.; Foster, R.R.; Hancox, J.C.; Bates, D.O. VEGF activates receptor-operated cation channels in human microvascular endothelial cells. Arterioscler. Thromb. Vasc. Biol. 2006, 26, 1768–1776. [Google Scholar] [CrossRef]
  76. Ge, R.; Tai, Y.; Sun, Y.; Zhou, K.; Yang, S.; Cheng, T.; Zou, Q.; Shen, F.; Wang, Y. Critical role of TRPC6 channels in VEGF-mediated angiogenesis. Cancer Lett. 2009, 283, 43–51. [Google Scholar] [CrossRef] [PubMed]
  77. Andrikopoulos, P.; Eccles, S.A.; Yaqoob, M.M. Coupling between the TRPC3 ion channel and the NCX1 transporter contributed to VEGF-induced ERK1/2 activation and angiogenesis in human primary endothelial cells. Cell. Signal. 2017, 37, 12–30. [Google Scholar] [CrossRef]
  78. Antigny, F.; Girardin, N.; Frieden, M. Transient receptor potential canonical channels are required for in vitro endothelial tube formation. J. Biol. Chem. 2012, 287, 5917–5927. [Google Scholar] [CrossRef]
  79. Wood, P.G.; Gillespie, J.I. Evidence for mitochondrial Ca2+-induced Ca2+ release in permeabilised endothelial cells. Biochem. Biophys. Res. Commun. 1998, 246, 543–548. [Google Scholar] [CrossRef]
  80. Keeley, T.P.; Siow, R.C.M.; Jacob, R.; Mann, G.E. Reduced SERCA activity underlies dysregulation of Ca2+ homeostasis under atmospheric O2 levels. FASEB J. 2018, 32, 2531–2538. [Google Scholar] [CrossRef]
  81. Mountian, I.; Manolopoulos, V.G.; De Smedt, H.; Parys, J.B.; Missiaen, L.; Wuytack, F. Expression patterns of sarco/endoplasmic reticulum Ca2+-ATPase and inositol 1,4,5-trisphosphate receptor isoforms in vascular endothelial cells. Cell Calcium 1999, 25, 371–380. [Google Scholar] [CrossRef]
  82. Estrada, I.A.; Donthamsetty, R.; Debski, P.; Zhou, M.H.; Zhang, S.L.; Yuan, J.X.; Han, W.; Makino, A. STIM1 restores coronary endothelial function in type 1 diabetic mice. Circ. Res. 2012, 111, 1166–1175. [Google Scholar] [CrossRef] [PubMed]
  83. Sun, M.Y.; Geyer, M.; Komarova, Y.A. IP3 receptor signaling and endothelial barrier function. Cell. Mol. Life Sci. 2017, 74, 4189–4207. [Google Scholar] [CrossRef] [PubMed]
  84. Ottolini, M.; Hong, K.; Sonkusare, S.K. Calcium signals that determine vascular resistance. Wiley Interdiscip. Rev. Syst. Biol. Med. 2019, e1448. [Google Scholar] [CrossRef] [PubMed]
  85. Berridge, M.J. Inositol trisphosphate and calcium signalling mechanisms. Biochim. Biophys. Acta 2009, 1793, 933–940. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Prole, D.L.; Taylor, C.W. Structure and Function of IP3 Receptors. Cold Spring Harb. Perspect. Biol. 2019, 11. [Google Scholar] [CrossRef] [PubMed]
  87. Hertle, D.N.; Yeckel, M.F. Distribution of inositol-1,4,5-trisphosphate receptor isotypes and ryanodine receptor isotypes during maturation of the rat hippocampus. Neuroscience 2007, 150, 625–638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Zuccolo, E.; Laforenza, U.; Negri, S.; Botta, L.; Berra-Romani, R.; Faris, P.; Scarpellino, G.; Forcaia, G.; Pellavio, G.; Sancini, G.; et al. Muscarinic M5 receptors trigger acetylcholine-induced Ca2+ signals and nitric oxide release in human brain microvascular endothelial cells. J. Cell. Physiol. 2019, 234, 4540–4562. [Google Scholar] [CrossRef] [PubMed]
  89. Zuccolo, E.; Lim, D.; Kheder, D.A.; Perna, A.; Catarsi, P.; Botta, L.; Rosti, V.; Riboni, L.; Sancini, G.; Tanzi, F.; et al. Acetylcholine induces intracellular Ca2+ oscillations and nitric oxide release in mouse brain endothelial cells. Cell Calcium 2017, 66, 33–47. [Google Scholar] [CrossRef]
  90. Isakson, B.E. Localized expression of an Ins(1,4,5)P3 receptor at the myoendothelial junction selectively regulates heterocellular Ca2+ communication. J. Cell Sci. 2008, 121, 3664–3673. [Google Scholar] [CrossRef]
  91. Grayson, T.H.; Haddock, R.E.; Murray, T.P.; Wojcikiewicz, R.J.; Hill, C.E. Inositol 1,4,5-trisphosphate receptor subtypes are differentially distributed between smooth muscle and endothelial layers of rat arteries. Cell Calcium 2004, 36, 447–458. [Google Scholar] [CrossRef]
  92. Carter, T.D.; Ogden, D. Kinetics of Ca2+ release by InsP3 in pig single aortic endothelial cells: Evidence for an inhibitory role of cytosolic Ca2+ in regulating hormonally evoked Ca2+ spikes. J. Physiol. 1997, 504, 17–33. [Google Scholar] [CrossRef] [PubMed]
  93. Foskett, J.K.; White, C.; Cheung, K.H.; Mak, D.O. Inositol trisphosphate receptor Ca2+ release channels. Physiol. Rev. 2007, 87, 593–658. [Google Scholar] [CrossRef] [PubMed]
  94. Vaca, L.; Kunze, D.L. IP3-activated Ca2+ channels in the plasma membrane of cultured vascular endothelial cells. Am. J. Physiol. 1995, 269, C733–C738. [Google Scholar] [CrossRef] [PubMed]
  95. Garnier-Raveaud, S.; Usson, Y.; Cand, F.; Robert-Nicoud, M.; Verdetti, J.; Faury, G. Identification of membrane calcium channels essential for cytoplasmic and nuclear calcium elevations induced by vascular endothelial growth factor in human endothelial cells. Growth Factors 2001, 19, 35–48. [Google Scholar] [CrossRef] [PubMed]
  96. Dellis, O.; Dedos, S.G.; Tovey, S.C.; Taufiq Ur, R.; Dubel, S.J.; Taylor, C.W. Ca2+ entry through plasma membrane IP3 receptors. Science 2006, 313, 229–233. [Google Scholar] [CrossRef] [PubMed]
  97. Santulli, G.; Lewis, D.; des Georges, A.; Marks, A.R.; Frank, J. Ryanodine Receptor Structure and Function in Health and Disease. Subcell. Biochem. 2018, 87, 329–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Kohler, R.; Brakemeier, S.; Kuhn, M.; Degenhardt, C.; Buhr, H.; Pries, A.; Hoyer, J. Expression of ryanodine receptor type 3 and TRP channels in endothelial cells: Comparison of in situ and cultured human endothelial cells. Cardiovasc. Res. 2001, 51, 160–168. [Google Scholar] [CrossRef]
  99. Paltauf-Doburzynska, J.; Frieden, M.; Spitaler, M.; Graier, W.F. Histamine-induced Ca2+ oscillations in a human endothelial cell line depend on transmembrane ion flux, ryanodine receptors and endoplasmic reticulum Ca2+-ATPase. J. Physiol. 2000, 524, 701–713. [Google Scholar] [CrossRef]
  100. Zhou, L.; Yang, B.; Wang, Y.; Zhang, H.L.; Chen, R.W.; Wang, Y.B. Bradykinin regulates the expression of claudin-5 in brain microvascular endothelial cells via calcium-induced calcium release. J. Neurosci. Res. 2014, 92, 597–606. [Google Scholar] [CrossRef]
  101. Wang, Y.B.; Liu, Y.H. Bradykinin selectively modulates the blood-tumor barrier via calcium-induced calcium release. J. Neurosci. Res. 2009, 87, 660–667. [Google Scholar] [CrossRef]
  102. Kameritsch, P.; Pogoda, K.; Ritter, A.; Munzing, S.; Pohl, U. Gap junctional communication controls the overall endothelial calcium response to vasoactive agonists. Cardiovasc. Res. 2012, 93, 508–515. [Google Scholar] [CrossRef] [PubMed]
  103. Borisova, L.; Wray, S.; Eisner, D.A.; Burdyga, T. How structure, Ca signals, and cellular communications underlie function in precapillary arterioles. Circ. Res. 2009, 105, 803–810. [Google Scholar] [CrossRef] [PubMed]
  104. Favia, A.; Desideri, M.; Gambara, G.; D’Alessio, A.; Ruas, M.; Esposito, B.; Del Bufalo, D.; Parrington, J.; Ziparo, E.; Palombi, F.; et al. VEGF-induced neoangiogenesis is mediated by NAADP and two-pore channel-2-dependent Ca2+ signaling. Proc. Natl. Acad. Sci. USA 2014, 111, E4706–E4715. [Google Scholar] [CrossRef] [PubMed]
  105. Patel, S. Function and dysfunction of two-pore channels. Sci. Signal. 2015, 8, re7. [Google Scholar] [CrossRef] [PubMed]
  106. Moccia, F.; Nusco, G.A.; Lim, D.; Kyozuka, K.; Santella, L. NAADP and InsP3 play distinct roles at fertilization in starfish oocytes. Dev. Biol. 2006, 294, 24–38. [Google Scholar] [CrossRef] [Green Version]
  107. Foster, W.J.; Taylor, H.B.C.; Padamsey, Z.; Jeans, A.F.; Galione, A.; Emptage, N.J. Hippocampal mGluR1-dependent long-term potentiation requires NAADP-mediated acidic store Ca2+ signaling. Sci. Signal. 2018, 11, 9093. [Google Scholar] [CrossRef]
  108. Faris, P.; Pellavio, G.; Ferulli, F.; Di Nezza, F.; Shekha, M.; Lim, D.; Maestri, M.; Guerra, G.; Ambrosone, L.; Pedrazzoli, P.; et al. Nicotinic Acid Adenine Dinucleotide Phosphate (NAADP) Induces Intracellular Ca2+ Release through the Two-Pore Channel TPC1 in Metastatic Colorectal Cancer Cells. Cancers 2019, 11, 542. [Google Scholar] [CrossRef]
  109. Guse, A.H.; Diercks, B.P. Integration of nicotinic acid adenine dinucleotide phosphate (NAADP)-dependent calcium signalling. J. Physiol. 2018, 596, 2735–2743. [Google Scholar] [CrossRef] [Green Version]
  110. Boslett, J.; Hemann, C.; Christofi, F.L.; Zweier, J.L. Characterization of CD38 in the major cell types of the heart: Endothelial cells highly express CD38 with activation by hypoxia-reoxygenation triggering NAD(P)H depletion. Am. J. Physiol. Cell Physiol. 2018, 314, C297–C309. [Google Scholar] [CrossRef]
  111. Esposito, B.; Gambara, G.; Lewis, A.M.; Palombi, F.; D’Alessio, A.; Taylor, L.X.; Genazzani, A.A.; Ziparo, E.; Galione, A.; Churchill, G.C.; et al. NAADP links histamine H1 receptors to secretion of von Willebrand factor in human endothelial cells. Blood 2011, 117, 4968–4977. [Google Scholar] [CrossRef] [Green Version]
  112. Galione, A. A primer of NAADP-mediated Ca2+ signalling: From sea urchin eggs to mammalian cells. Cell Calcium 2015, 58, 27–47. [Google Scholar] [CrossRef] [PubMed]
  113. Faris, P.; Shekha, M.; Montagna, D.; Guerra, G.; Moccia, F. Endolysosomal Ca2+ Signalling and Cancer Hallmarks: Two-Pore Channels on the Move, TRPML1 Lags Behind! Cancers 2018, 11, 27. [Google Scholar] [CrossRef] [PubMed]
  114. Berra-Romani, R.; Faris, P.; Pellavio, G.; Orgiu, M.; Negri, S.; Forcaia, G.; Var-Gaz-Guadarrama, V.; Garcia-Carrasco, M.; Botta, L.; Sancini, G.; et al. Histamine induces intracellular Ca2+ oscillations and nitric oxide release in endothelial cells from brain microvascular circulation. J. Cell. Physiol. 2019. [Google Scholar] [CrossRef] [PubMed]
  115. Ronco, V.; Potenza, D.M.; Denti, F.; Vullo, S.; Gagliano, G.; Tognolina, M.; Guerra, G.; Pinton, P.; Genazzani, A.A.; Mapelli, L.; et al. A novel Ca2+-mediated cross-talk between endoplasmic reticulum and acidic organelles: Implications for NAADP-dependent Ca2+ signalling. Cell Calcium 2015, 57, 89–100. [Google Scholar] [CrossRef] [PubMed]
  116. Yang, J.; Zhao, Z.; Gu, M.; Feng, X.; Xu, H. Release and uptake mechanisms of vesicular Ca2+ stores. Protein Cell 2018, 10, 8–19. [Google Scholar] [CrossRef] [PubMed]
  117. Prakriya, M.; Lewis, R.S. Store-Operated Calcium Channels. Physiol. Rev. 2015, 95, 1383–1436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Moccia, F.; Zuccolo, E.; Soda, T.; Tanzi, F.; Guerra, G.; Mapelli, L.; Lodola, F.; D’Angelo, E. Stim and Orai proteins in neuronal Ca2+ signaling and excitability. Front. Cell. Neurosci. 2015, 9, 153. [Google Scholar] [CrossRef] [PubMed]
  119. Parekh, A.B. Functional consequences of activating store-operated CRAC channels. Cell Calcium 2007, 42, 111–121. [Google Scholar] [CrossRef]
  120. Blatter, L.A. Tissue Specificity: SOCE: Implications for Ca2+ Handling in Endothelial Cells. Adv. Exp. Med. Biol. 2017, 993, 343–361. [Google Scholar] [CrossRef]
  121. Groschner, K.; Shrestha, N.; Fameli, N. Cardiovascular and Hemostatic Disorders: SOCE in Cardiovascular Cells: Emerging Targets for Therapeutic Intervention. Adv. Exp. Med. Biol. 2017, 993, 473–503. [Google Scholar] [CrossRef]
  122. Abdullaev, I.F.; Bisaillon, J.M.; Potier, M.; Gonzalez, J.C.; Motiani, R.K.; Trebak, M. Stim1 and Orai1 mediate CRAC currents and store-operated calcium entry important for endothelial cell proliferation. Circ. Res. 2008, 103, 1289–1299. [Google Scholar] [CrossRef] [PubMed]
  123. Li, J.; Cubbon, R.M.; Wilson, L.A.; Amer, M.S.; McKeown, L.; Hou, B.; Majeed, Y.; Tumova, S.; Seymour, V.A.L.; Taylor, H.; et al. Orai1 and CRAC channel dependence of VEGF-activated Ca2+ entry and endothelial tube formation. Circ. Res. 2011, 108, 1190–1198. [Google Scholar] [CrossRef] [PubMed]
  124. Jho, D.; Mehta, D.; Ahmmed, G.; Gao, X.P.; Tiruppathi, C.; Broman, M.; Malik, A.B. Angiopoietin-1 opposes VEGF-induced increase in endothelial permeability by inhibiting TRPC1-dependent Ca2 influx. Circ. Res. 2005, 96, 1282–1290. [Google Scholar] [CrossRef] [PubMed]
  125. Antigny, F.; Jousset, H.; Konig, S.; Frieden, M. Thapsigargin activates Ca2+ entry both by store-dependent, STIM1/Orai1-mediated, and store-independent, TRPC3/PLC/PKC-mediated pathways in human endothelial cells. Cell Calcium 2011, 49, 115–127. [Google Scholar] [CrossRef] [PubMed]
  126. Zhou, M.H.; Zheng, H.; Si, H.; Jin, Y.; Peng, J.M.; He, L.; Zhou, Y.; Munoz-Garay, C.; Zawieja, D.C.; Kuo, L.; et al. Stromal interaction molecule 1 (STIM1) and Orai1 mediate histamine-evoked calcium entry and nuclear factor of activated T-cells (NFAT) signaling in human umbilical vein endothelial cells. J. Biol. Chem. 2014, 289, 29446–29456. [Google Scholar] [CrossRef]
  127. Girardin, N.C.; Antigny, F.; Frieden, M. Electrophysiological characterization of store-operated and agonist-induced Ca2+ entry pathways in endothelial cells. Pflug. Arch. 2010, 460, 109–120. [Google Scholar] [CrossRef]
  128. Fasolato, C.; Nilius, B. Store depletion triggers the calcium release-activated calcium current (ICRAC) in macrovascular endothelial cells: A comparison with Jurkat and embryonic kidney cell lines. Pflug. Arch. 1998, 436, 69–74. [Google Scholar] [CrossRef]
  129. Kito, H.; Yamamura, H.; Suzuki, Y.; Yamamura, H.; Ohya, S.; Asai, K.; Imaizumi, Y. Regulation of store-operated Ca2+ entry activity by cell cycle dependent up-regulation of Orai2 in brain capillary endothelial cells. Biochem. Biophys. Res. Commun. 2015, 459, 457–462. [Google Scholar] [CrossRef]
  130. Sachdeva, R.; Fleming, T.; Schumacher, D.; Homberg, S.; Stilz, K.; Mohr, F.; Wagner, A.H.; Tsvilovskyy, V.; Mathar, I.; Freichel, M. Methylglyoxal evokes acute Ca2+ transients in distinct cell types and increases agonist-evoked Ca2+ entry in endothelial cells via CRAC channels. Cell Calcium 2019, 78, 66–75. [Google Scholar] [CrossRef]
  131. Liou, J.; Kim, M.L.; Heo, W.D.; Jones, J.T.; Myers, J.W.; Ferrell, J.E., Jr.; Meyer, T. STIM is a Ca2+ sensor essential for Ca2+-store-depletion-triggered Ca2+ influx. Curr. Biol. 2005, 15, 1235–1241. [Google Scholar] [CrossRef]
  132. Diercks, B.P.; Werner, R.; Weidemuller, P.; Czarniak, F.; Hernandez, L.; Lehmann, C.; Rosche, A.; Kruger, A.; Kaufmann, U.; Vaeth, M.; et al. ORAI1, STIM1/2, and RYR1 shape subsecond Ca2+ microdomains upon T cell activation. Sci. Signal. 2018, 11, 358. [Google Scholar] [CrossRef] [PubMed]
  133. Thiel, M.; Lis, A.; Penner, R. STIM2 drives Ca2+ oscillations through store-operated Ca2+ entry caused by mild store depletion. J. Physiol. 2013, 591, 1433–1445. [Google Scholar] [CrossRef] [PubMed]
  134. Li, J.; Bruns, A.F.; Hou, B.; Rode, B.; Webster, P.J.; Bailey, M.A.; Appleby, H.L.; Moss, N.K.; Ritchie, J.E.; Yuldasheva, N.Y.; et al. Orai3 Surface Accumulation and Calcium Entry Evoked by Vascular Endothelial Growth Factor. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 1987–1994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Moccia, F.; Dragoni, S.; Lodola, F.; Bonetti, E.; Bottino, C.; Guerra, G.; Laforenza, U.; Rosti, V.; Tanzi, F. Store-dependent Ca2+ entry in endothelial progenitor cells as a perspective tool to enhance cell-based therapy and adverse tumour vascularization. Curr. Med. Chem. 2012, 19, 5802–5818. [Google Scholar] [CrossRef] [PubMed]
  136. Earley, S.; Brayden, J.E. Transient receptor potential channels in the vasculature. Physiol. Rev. 2015, 95, 645–690. [Google Scholar] [CrossRef] [PubMed]
  137. Curcic, S.; Schober, R.; Schindl, R.; Groschner, K. TRPC-mediated Ca2+ signaling and control of cellular functions. Semin. Cell Dev. Biol. 2019. [Google Scholar] [CrossRef]
  138. Di Buduo, C.A.; Moccia, F.; Battiston, M.; De Marco, L.; Mazzucato, M.; Moratti, R.; Tanzi, F.; Balduini, A. The importance of calcium in the regulation of megakaryocyte function. Haematologica 2014, 99, 769–778. [Google Scholar] [CrossRef] [Green Version]
  139. Moccia, F.; Lucariello, A.; Guerra, G. TRPC3-mediated Ca2+ signals as a promising strategy to boost therapeutic angiogenesis in failing hearts: The role of autologous endothelial colony forming cells. J. Cell. Physiol. 2018, 233, 3901–3917. [Google Scholar] [CrossRef]
  140. Sundivakkam, P.C.; Freichel, M.; Singh, V.; Yuan, J.P.; Vogel, S.M.; Flockerzi, V.; Malik, A.B.; Tiruppathi, C. The Ca2+ sensor stromal interaction molecule 1 (STIM1) is necessary and sufficient for the store-operated Ca2+ entry function of transient receptor potential canonical (TRPC) 1 and 4 channels in endothelial cells. Mol. Pharmacol. 2012, 81, 510–526. [Google Scholar] [CrossRef]
  141. Freichel, M.; Suh, S.H.; Pfeifer, A.; Schweig, U.; Trost, C.; Weissgerber, P.; Biel, M.; Philipp, S.; Freise, D.; Droogmans, G.; et al. Lack of an endothelial store-operated Ca2+ current impairs agonist-dependent vasorelaxation in TRP4-/-mice. Nat. Cell Biol. 2001, 3, 121–127. [Google Scholar] [CrossRef]
  142. Brough, G.H.; Wu, S.; Cioffi, D.; Moore, T.M.; Li, M.; Dean, N.; Stevens, T. Contribution of endogenously expressed Trp1 to a Ca2+-selective, store-operated Ca2+ entry pathway. FASEB J. 2001, 15, 1727–1738. [Google Scholar] [CrossRef] [PubMed]
  143. Cioffi, D.L.; Wu, S.; Alexeyev, M.; Goodman, S.R.; Zhu, M.X.; Stevens, T. Activation of the endothelial store-operated ISOC Ca2+ channel requires interaction of protein 4.1 with TRPC4. Circ. Res. 2005, 97, 1164–1172. [Google Scholar] [CrossRef] [PubMed]
  144. Tiruppathi, C.; Freichel, M.; Vogel, S.M.; Paria, B.C.; Mehta, D.; Flockerzi, V.; Malik, A.B. Impairment of store-operated Ca2+ entry in TRPC4(-/-) mice interferes with increase in lung microvascular permeability. Circ. Res. 2002, 91, 70–76. [Google Scholar] [CrossRef] [PubMed]
  145. Parekh, A.B.; Putney, J.W., Jr. Store-operated calcium channels. Physiol. Rev. 2005, 85, 757–810. [Google Scholar] [CrossRef] [PubMed]
  146. Albert, A.P.; Large, W.A. Store-operated Ca2+-permeable non-selective cation channels in smooth muscle cells. Cell Calcium 2003, 33, 345–356. [Google Scholar] [CrossRef]
  147. Wu, S.; Cioffi, E.A.; Alvarez, D.; Sayner, S.L.; Chen, H.; Cioffi, D.L.; King, J.; Creighton, J.R.; Townsley, M.; Goodman, S.R.; et al. Essential role of a Ca2+-selective, store-operated current (ISOC) in endothelial cell permeability: Determinants of the vascular leak site. Circ. Res. 2005, 96, 856–863. [Google Scholar] [CrossRef] [PubMed]
  148. Vaca, L.; Kunze, D.L. Depletion of intracellular Ca2+ stores activates a Ca2+-selective channel in vascular endothelium. Am. J. Physiol. 1994, 267, C920–C925. [Google Scholar] [CrossRef] [PubMed]
  149. Groschner, K.; Hingel, S.; Lintschinger, B.; Balzer, M.; Romanin, C.; Zhu, X.; Schreibmayer, W. Trp proteins form store-operated cation channels in human vascular endothelial cells. FEBS Lett. 1998, 437, 101–106. [Google Scholar] [CrossRef]
  150. Encabo, A.; Romanin, C.; Birke, F.W.; Kukovetz, W.R.; Groschner, K. Inhibition of a store-operated Ca2+ entry pathway in human endothelial cells by the isoquinoline derivative LOE 908. Br. J. Pharmacol. 1996, 119, 702–706. [Google Scholar] [CrossRef]
  151. Ma, X.; Cheng, K.T.; Wong, C.O.; O’Neil, R.G.; Birnbaumer, L.; Ambudkar, I.S.; Yao, X. Heteromeric TRPV4-C1 channels contribute to store-operated Ca2+ entry in vascular endothelial cells. Cell Calcium 2011, 50, 502–509. [Google Scholar] [CrossRef]
  152. Cioffi, D.L.; Barry, C.; Stevens, T. Store-operated calcium entry channels in pulmonary endothelium: The emerging story of TRPCS and Orai1. Adv. Exp. Med. Biol. 2010, 661, 137–154. [Google Scholar] [CrossRef] [PubMed]
  153. Cioffi, D.L.; Wu, S.; Chen, H.; Alexeyev, M.; St Croix, C.M.; Pitt, B.R.; Uhlig, S.; Stevens, T. Orai1 determines calcium selectivity of an endogenous TRPC heterotetramer channel. Circ. Res. 2012, 110, 1435–1444. [Google Scholar] [CrossRef] [PubMed]
  154. Vasauskas, A.A.; Chen, H.; Wu, S.; Cioffi, D.L. The serine-threonine phosphatase calcineurin is a regulator of endothelial store-operated calcium entry. Pulm. Circ. 2014, 4, 116–127. [Google Scholar] [CrossRef] [PubMed]
  155. Xu, N.; Cioffi, D.L.; Alexeyev, M.; Rich, T.C.; Stevens, T. Sodium entry through endothelial store-operated calcium entry channels: Regulation by Orai1. Am. J. Physiol. Cell Physiol. 2015, 308, C277–C288. [Google Scholar] [CrossRef] [PubMed]
  156. Apte, R.S.; Chen, D.S.; Ferrara, N. VEGF in Signaling and Disease: Beyond Discovery and Development. Cell 2019, 176, 1248–1264. [Google Scholar] [CrossRef] [Green Version]
  157. Koch, S.; Claesson-Welsh, L. Signal transduction by vascular endothelial growth factor receptors. Cold Spring Harb. Perspect. Med. 2012, 2, a006502. [Google Scholar] [CrossRef]
  158. Mancardi, D.; Pla, A.F.; Moccia, F.; Tanzi, F.; Munaron, L. Old and new gasotransmitters in the cardiovascular system: Focus on the role of nitric oxide and hydrogen sulfide in endothelial cells and cardiomyocytes. Curr. Pharm. Biotechnol. 2011, 12, 1406–1415. [Google Scholar] [CrossRef]
  159. Criscuolo, G.R.; Lelkes, P.I.; Rotrosen, D.; Oldfield, E.H. Cytosolic calcium changes in endothelial cells induced by a protein product of human gliomas containing vascular permeability factor activity. J. Neurosurg. 1989, 71, 884–891. [Google Scholar] [CrossRef] [Green Version]
  160. Cunningham, S.A.; Tran, T.M.; Arrate, M.P.; Bjercke, R.; Brock, T.A. KDR activation is crucial for VEGF165-mediated Ca2+ mobilization in human umbilical vein endothelial cells. Am. J. Physiol. 1999, 276, C176–C181. [Google Scholar] [CrossRef]
  161. Fearnley, G.W.; Bruns, A.F.; Wheatcroft, S.B.; Ponnambalam, S. VEGF-A isoform-specific regulation of calcium ion flux, transcriptional activation and endothelial cell migration. Biol. Open 2015, 4, 731–742. [Google Scholar] [CrossRef] [Green Version]
  162. Dawson, N.S.; Zawieja, D.C.; Wu, M.H.; Granger, H.J. Signaling pathways mediating VEGF165-induced calcium transients and membrane depolarization in human endothelial cells. FASEB J. 2006, 20, 991–993. [Google Scholar] [CrossRef] [PubMed]
  163. Andrikopoulos, P.; Baba, A.; Matsuda, T.; Djamgoz, M.B.; Yaqoob, M.M.; Eccles, S.A. Ca2+ influx through reverse mode Na+/Ca2+ exchange is critical for vascular endothelial growth factor-mediated extracellular signal-regulated kinase (ERK) 1/2 activation and angiogenic functions of human endothelial cells. J. Biol. Chem. 2011, 286, 37919–37931. [Google Scholar] [CrossRef] [PubMed]
  164. Kusaba, T.; Okigaki, M.; Matui, A.; Murakami, M.; Ishikawa, K.; Kimura, T.; Sonomura, K.; Adachi, Y.; Shibuya, M.; Shirayama, T.; et al. Klotho is associated with VEGF receptor-2 and the transient receptor potential canonical-1 Ca2+ channel to maintain endothelial integrity. Proc. Natl. Acad. Sci. USA 2010, 107, 19308–19313. [Google Scholar] [CrossRef] [PubMed]
  165. Potenza, D.M.; Guerra, G.; Avanzato, D.; Poletto, V.; Pareek, S.; Guido, D.; Gallanti, A.; Rosti, V.; Munaron, L.; Tanzi, F.; et al. Hydrogen sulphide triggers VEGF-induced intracellular Ca2+ signals in human endothelial cells but not in their immature progenitors. Cell Calcium 2014, 56, 225–234. [Google Scholar] [CrossRef] [PubMed]
  166. Moccia, F.; Bertoni, G.; Pla, A.F.; Dragoni, S.; Pupo, E.; Merlino, A.; Mancardi, D.; Munaron, L.; Tanzi, F. Hydrogen sulfide regulates intracellular Ca2+ concentration in endothelial cells from excised rat aorta. Curr. Pharm. Biotechnol. 2011, 12, 1416–1426. [Google Scholar] [CrossRef] [PubMed]
  167. Pupo, E.; Pla, A.F.; Avanzato, D.; Moccia, F.; Cruz, J.E.; Tanzi, F.; Merlino, A.; Mancardi, D.; Munaron, L. Hydrogen sulfide promotes calcium signals and migration in tumor-derived endothelial cells. Free Radic. Biol. Med. 2011, 51, 1765–1773. [Google Scholar] [CrossRef] [PubMed]
  168. Munaron, L.; Avanzato, D.; Moccia, F.; Mancardi, D. Hydrogen sulfide as a regulator of calcium channels. Cell Calcium 2013, 53, 77–84. [Google Scholar] [CrossRef]
  169. Altaany, Z.; Moccia, F.; Munaron, L.; Mancardi, D.; Wang, R. Hydrogen sulfide and endothelial dysfunction: Relationship with nitric oxide. Curr. Med. Chem. 2014, 21, 3646–3661. [Google Scholar] [CrossRef]
  170. Banumathi, E.; O’Connor, A.; Gurunathan, S.; Simpson, D.A.; McGeown, J.G.; Curtis, T.M. VEGF-induced retinal angiogenic signaling is critically dependent on Ca2+ signaling by Ca2+/calmodulin-dependent protein kinase II. Investig. Ophthalmol. Vis. Sci. 2011, 52, 3103–3111. [Google Scholar] [CrossRef]
  171. Song, H.B.; Jun, H.O.; Kim, J.H.; Fruttiger, M.; Kim, J.H. Suppression of transient receptor potential canonical channel 4 inhibits vascular endothelial growth factor-induced retinal neovascularization. Cell Calcium 2015, 57, 101–108. [Google Scholar] [CrossRef]
  172. Yi, F.X.; Boeldt, D.S.; Magness, R.R.; Bird, I.M. [Ca2+]i signaling vs. eNOS expression as determinants of NO output in uterine artery endothelium: Relative roles in pregnancy adaptation and reversal by VEGF165. Am. J. Physiol. Heart Circ. Physiol. 2011, 300, H1182–H1193. [Google Scholar] [CrossRef] [PubMed]
  173. Bird, I.M.; Sullivan, J.A.; Di, T.; Cale, J.M.; Zhang, L.; Zheng, J.; Magness, R.R. Pregnancy-dependent changes in cell signaling underlie changes in differential control of vasodilator production in uterine artery endothelial cells. Endocrinology 2000, 141, 1107–1117. [Google Scholar] [CrossRef] [PubMed]
  174. Anaya, H.A.; Yi, F.X.; Boeldt, D.S.; Krupp, J.; Grummer, M.A.; Shah, D.M.; Bird, I.M. Changes in Ca2+ Signaling and Nitric Oxide Output by Human Umbilical Vein Endothelium in Diabetic and Gestational Diabetic Pregnancies. Biol. Reprod. 2015, 93, 60. [Google Scholar] [CrossRef] [PubMed]
  175. Boeldt, D.S.; Bird, I.M. Vascular adaptation in pregnancy and endothelial dysfunction in preeclampsia. J. Endocrinol. 2017, 232, R27–R44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Holmes, K.; Chapman, E.; See, V.; Cross, M.J. VEGF stimulates RCAN1.4 expression in endothelial cells via a pathway requiring Ca2+/calcineurin and protein kinase C-delta. PLoS ONE 2010, 5, e11435. [Google Scholar] [CrossRef] [PubMed]
  177. Lee, M.; Spokes, K.C.; Aird, W.C.; Abid, M.R. Intracellular Ca2+ can compensate for the lack of NADPH oxidase-derived ROS in endothelial cells. FEBS Lett. 2010, 584, 3131–3136. [Google Scholar] [CrossRef]
  178. Mirzapoiazova, T.; Kolosova, I.; Usatyuk, P.V.; Natarajan, V.; Verin, A.D. Diverse effects of vascular endothelial growth factor on human pulmonary endothelial barrier and migration. Am. J. Physiol. Lung Cell Mol. Physiol. 2006, 291, L718–L724. [Google Scholar] [CrossRef] [Green Version]
  179. McLaughlin, A.P.; De Vries, G.W. Role of PLCgamma and Ca2+ in VEGF- and FGF-induced choroidal endothelial cell proliferation. Am. J. Physiol. Cell Physiol. 2001, 281, C1448–C1456. [Google Scholar] [CrossRef]
  180. Huang, K.; Andersson, C.; Roomans, G.M.; Ito, N.; Claesson-Welsh, L. Signaling properties of VEGF receptor-1 and -2 homo- and heterodimers. Int. J. Biochem. Cell Biol. 2001, 33, 315–324. [Google Scholar] [CrossRef]
  181. Wu, H.M.; Yuan, Y.; Zawieja, D.C.; Tinsley, J.; Granger, H.J. Role of phospholipase C, protein kinase C, and calcium in VEGF-induced venular hyperpermeability. Am. J. Physiol. 1999, 276, H535–H542. [Google Scholar] [CrossRef] [Green Version]
  182. Pal, S.; Wu, J.; Murray, J.K.; Gellman, S.H.; Wozniak, M.A.; Keely, P.J.; Boyer, M.E.; Gomez, T.M.; Hasso, S.M.; Fallon, J.F.; et al. An antiangiogenic neurokinin-B/thromboxane A2 regulatory axis. J. Cell Biol. 2006, 174, 1047–1058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Bazzazi, H.; Popel, A.S. Computational investigation of sphingosine kinase 1 (SphK1) and calcium dependent ERK1/2 activation downstream of VEGFR2 in endothelial cells. PLoS Comput. Biol. 2017, 13, e1005332. [Google Scholar] [CrossRef] [PubMed]
  184. Yokota, Y.; Nakajima, H.; Wakayama, Y.; Muto, A.; Kawakami, K.; Fukuhara, S.; Mochizuki, N. Endothelial Ca2+ oscillations reflect VEGFR signaling-regulated angiogenic capacity in vivo. eLife 2015, 4, e08817. [Google Scholar] [CrossRef] [PubMed]
  185. Szewczyk, M.M.; Pande, J.; Akolkar, G.; Grover, A.K. Caloxin 1b3: A novel plasma membrane Ca2+-pump isoform 1 selective inhibitor that increases cytosolic Ca2+ in endothelial cells. Cell Calcium 2010, 48, 352–357. [Google Scholar] [CrossRef] [PubMed]
  186. Paszty, K.; Caride, A.J.; Bajzer, Z.; Offord, C.P.; Padanyi, R.; Hegedus, L.; Varga, K.; Strehler, E.E.; Enyedi, A. Plasma membrane Ca2+-ATPases can shape the pattern of Ca2+ transients induced by store-operated Ca2+ entry. Sci. Signal. 2015, 8, 19. [Google Scholar] [CrossRef] [PubMed]
  187. Holton, M.L.; Wang, W.; Emerson, M.; Neyses, L.; Armesilla, A.L. Plasma membrane calcium ATPase proteins as novel regulators of signal transduction pathways. World J. Biol. Chem. 2010, 1, 201–208. [Google Scholar] [CrossRef] [PubMed]
  188. Grgic, I.; Eichler, I.; Heinau, P.; Si, H.; Brakemeier, S.; Hoyer, J.; Kohler, R. Selective blockade of the intermediate-conductance Ca2+-activated K+ channel suppresses proliferation of microvascular and macrovascular endothelial cells and angiogenesis in vivo. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 704–709. [Google Scholar] [CrossRef]
  189. Guerra, G.; Lucariello, A.; Perna, A.; Botta, L.; De Luca, A.; Moccia, F. The Role of Endothelial Ca2+ Signaling in Neurovascular Coupling: A View from the Lumen. Int. J. Mol. Sci. 2018, 19, 938. [Google Scholar] [CrossRef]
  190. Sheng, J.Z.; Braun, A.P. Small-and intermediate-conductance Ca2+-activated K+ channels directly control agonist-evoked nitric oxide synthesis in human vascular endothelial cells. Am. J. Physiol. Cell Physiol. 2007, 293, C458–C467. [Google Scholar] [CrossRef]
  191. Ye, X.; Beckett, T.; Bagher, P.; Garland, C.J.; Dora, K.A. VEGF-A inhibits agonist-mediated Ca2+ responses and activation of IKCa channels in mouse resistance artery endothelial cells. J. Physiol. 2018, 596, 3553–3566. [Google Scholar] [CrossRef]
  192. Jiang, L.; Jha, V.; Dhanabal, M.; Sukhatme, V.P.; Alper, S.L. Intracellular Ca2+ signaling in endothelial cells by the angiogenesis inhibitors endostatin and angiostatin. Am. J. Physiol. Cell Physiol. 2001, 280, C1140–C1150. [Google Scholar] [CrossRef] [PubMed]
  193. Berridge, M.J. The endoplasmic reticulum: A multifunctional signaling organelle. Cell Calcium 2002, 32, 235–249. [Google Scholar] [CrossRef]
  194. Taylor, C.W.; Machaca, K. IP3 receptors and store-operated Ca2+ entry: A license to fill. Curr. Opin. Cell Biol. 2018, 57, 1–7. [Google Scholar] [CrossRef] [PubMed]
  195. Petersen, O.H.; Verkhratsky, A. Endoplasmic reticulum calcium tunnels integrate signalling in polarised cells. Cell Calcium 2007, 42, 373–378. [Google Scholar] [CrossRef] [PubMed]
  196. Zhang, A.Y.; Teggatz, E.G.; Zou, A.P.; Campbell, W.B.; Li, P.L. Endostatin uncouples NO and Ca2+ response to bradykinin through enhanced O2-production in the intact coronary endothelium. Am. J. Physiol. Heart Circ. Physiol. 2005, 288, H686–H694. [Google Scholar] [CrossRef] [PubMed]
  197. Bazzazi, H.; Isenberg, J.S.; Popel, A.S. Inhibition of VEGFR2 Activation and Its Downstream Signaling to ERK1/2 and Calcium by Thrombospondin-1 (TSP1): In silico Investigation. Front. Physiol. 2017, 8, 48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Ayada, T.; Taniguchi, K.; Okamoto, F.; Kato, R.; Komune, S.; Takaesu, G.; Yoshimura, A. Sprouty4 negatively regulates protein kinase C activation by inhibiting phosphatidylinositol 4,5-biphosphate hydrolysis. Oncogene 2009, 28, 1076–1088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Gifford, S.M.; Grummer, M.A.; Pierre, S.A.; Austin, J.L.; Zheng, J.; Bird, I.M. Functional characterization of HUVEC-CS: Ca2+ signaling, ERK 1/2 activation, mitogenesis and vasodilator production. J. Endocrinol. 2004, 182, 485–499. [Google Scholar] [CrossRef]
  200. Munaron, L.; Fiorio Pla, A. Calcium influx induced by activation of tyrosine kinase receptors in cultured bovine aortic endothelial cells. J. Cell. Physiol. 2000, 185, 454–463. [Google Scholar] [CrossRef]
  201. Maffucci, T.; Raimondi, C.; Abu-Hayyeh, S.; Dominguez, V.; Sala, G.; Zachary, I.; Falasca, M. A phosphoinositide 3-kinase/phospholipase Cgamma1 pathway regulates fibroblast growth factor-induced capillary tube formation. PLoS ONE 2009, 4, e8285. [Google Scholar] [CrossRef]
  202. Mergler, S.; Dannowski, H.; Bednarz, J.; Engelmann, K.; Hartmann, C.; Pleyer, U. Calcium influx induced by activation of receptor tyrosine kinases in SV40-transfected human corneal endothelial cells. Exp. Eye Res. 2003, 77, 485–495. [Google Scholar] [CrossRef]
  203. Antoniotti, S.; Fiorio Pla, A.; Barral, S.; Scalabrino, O.; Munaron, L.; Lovisolo, D. Interaction between TRPC channel subunits in endothelial cells. J. Recept. Signal Transduct. Res. 2006, 26, 225–240. [Google Scholar] [CrossRef] [PubMed]
  204. Antoniotti, S.; Fiorio Pla, A.; Pregnolato, S.; Mottola, A.; Lovisolo, D.; Munaron, L. Control of endothelial cell proliferation by calcium influx and arachidonic acid metabolism: A pharmacological approach. J. Cell. Physiol. 2003, 197, 370–378. [Google Scholar] [CrossRef] [PubMed]
  205. Antoniotti, S.; Lovisolo, D.; Fiorio Pla, A.; Munaron, L. Expression and functional role of bTRPC1 channels in native endothelial cells. FEBS Lett. 2002, 510, 189–195. [Google Scholar] [CrossRef]
  206. Wiecha, J.; Munz, B.; Wu, Y.; Noll, T.; Tillmanns, H.; Waldecker, B. Blockade of Ca2+-activated K+ channels inhibits proliferation of human endothelial cells induced by basic fibroblast growth factor. J. Vasc. Res. 1998, 35, 363–371. [Google Scholar] [CrossRef] [PubMed]
  207. Wiecha, J.; Reineker, K.; Reitmayer, M.; Voisard, R.; Hannekum, A.; Mattfeldt, T.; Waltenberger, J.; Hombach, V. Modulation of Ca2+-activated K+ channels in human vascular cells by insulin and basic fibroblast growth factor. Growth Horm. IGF Res. 1998, 8, 175–181. [Google Scholar] [CrossRef]
  208. Scharbrodt, W.; Kuhlmann, C.R.; Wu, Y.; Schaefer, C.A.; Most, A.K.; Backenkohler, U.; Neumann, T.; Tillmanns, H.; Waldecker, B.; Erdogan, A.; et al. Basic fibroblast growth factor-induced endothelial proliferation and NO synthesis involves inward rectifier K+ current. Arterioscler. Thromb. Vasc. Biol. 2004, 24, 1229–1233. [Google Scholar] [CrossRef] [PubMed]
  209. Moccia, F.; Berra-Romani, R.; Tritto, S.; Signorelli, S.; Taglietti, V.; Tanzi, F. Epidermal growth factor induces intracellular Ca2+ oscillations in microvascular endothelial cells. J. Cell. Physiol. 2003, 194, 139–150. [Google Scholar] [CrossRef]
  210. Valley, C.C.; Lidke, K.A.; Lidke, D.S. The dddspatiotemporal organization of ErbB receptors: Insights from microscopy. Cold Spring Harb. Perspect. Biol. 2014, 6, a020735. [Google Scholar] [CrossRef]
  211. Ridefelt, P.; Yokote, K.; Claesson-Welsh, L.; Siegbahn, A. PDGF-BB triggered cytoplasmic calcium responses in cells with endogenous or stably transfected PDGF beta-receptors. Growth Factors 1995, 12, 191–201. [Google Scholar] [CrossRef]
  212. Moccia, F.; Bonetti, E.; Dragoni, S.; Fontana, J.; Lodola, F.; Romani, R.B.; Laforenza, U.; Rosti, V.; Tanzi, F. Hematopoietic progenitor and stem cells circulate by surfing on intracellular Ca2+ waves: A novel target for cell-based therapy and anti-cancer treatment? Curr. Signal Transduct. Ther. 2012, 7, 161–176. [Google Scholar] [CrossRef]
  213. Noels, H.; Zhou, B.; Tilstam, P.V.; Theelen, W.; Li, X.; Pawig, L.; Schmitz, C.; Akhtar, S.; Simsekyilmaz, S.; Shagdarsuren, E.; et al. Deficiency of endothelial CXCR4 reduces reendothelialization and enhances neointimal hyperplasia after vascular injury in atherosclerosis-prone mice. Arterioscler. Thromb. Vasc. Biol. 2014, 34, 1209–1220. [Google Scholar] [CrossRef] [PubMed]
  214. Gupta, S.K.; Lysko, P.G.; Pillarisetti, K.; Ohlstein, E.; Stadel, J.M. Chemokine receptors in human endothelial cells. Functional expression of CXCR4 and its transcriptional regulation by inflammatory cytokines. J. Biol. Chem. 1998, 273, 4282–4287. [Google Scholar] [CrossRef] [PubMed]
  215. Sameermahmood, Z.; Balasubramanyam, M.; Saravanan, T.; Rema, M. Curcumin modulates SDF-1alpha/CXCR4-induced migration of human retinal endothelial cells (HRECs). Investig. Ophthalmol. Vis. Sci. 2008, 49, 3305–3311. [Google Scholar] [CrossRef] [PubMed]
  216. Augustin, H.G.; Koh, G.Y.; Thurston, G.; Alitalo, K. Control of vascular morphogenesis and homeostasis through the angiopoietin-Tie system. Nat. Rev. Mol. Cell Biol. 2009, 10, 165–177. [Google Scholar] [CrossRef] [PubMed]
  217. Pafumi, I.; Favia, A.; Gambara, G.; Papacci, F.; Ziparo, E.; Palombi, F.; Filippini, A. Regulation of Angiogenic Functions by Angiopoietins through Calcium-Dependent Signaling Pathways. BioMed Res. Int. 2015, 2015, 965271. [Google Scholar] [CrossRef] [PubMed]
  218. Yang, C.; Ohk, J.; Lee, J.Y.; Kim, E.J.; Kim, J.; Han, S.; Park, D.; Jung, H.; Kim, C. Calmodulin Mediates Ca2+-Dependent Inhibition of Tie2 Signaling and Acts as a Developmental Brake During Embryonic Angiogenesis. Arterioscler. Thromb. Vasc. Biol. 2016, 36, 1406–1416. [Google Scholar] [CrossRef] [PubMed]
  219. Inoue, K.; Xiong, Z.G. Silencing TRPM7 promotes growth/proliferation and nitric oxide production of vascular endothelial cells via the ERK pathway. Cardiovasc. Res. 2009, 83, 547–557. [Google Scholar] [CrossRef] [PubMed]
  220. Ashraf, S.; Bell, S.; O’Leary, C.; Canning, P.; Micu, I.; Fernandez, J.A.; O’Hare, M.; Barabas, P.; McCauley, H.; Brazil, D.P.; et al. CAMKII as a therapeutic target for growth factor-induced retinal and choroidal neovascularization. JCI Insight 2019, 4, 122442. [Google Scholar] [CrossRef]
  221. Chen, Z.; Li, B.; Dong, Q.; Qian, C.; Cheng, J.; Wang, Y. Repetitive Transient Ischemia-Induced Cardiac Angiogenesis is Mediated by Camkii Activation. Cell. Physiol. Biochem. 2018, 47, 914–924. [Google Scholar] [CrossRef]
  222. Zhang, Y.; Liu, N.M.; Wang, Y.; Youn, J.Y.; Cai, H. Endothelial cell calpain as a critical modulator of angiogenesis. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 1326–1335. [Google Scholar] [CrossRef] [PubMed]
  223. Tsai, F.C.; Seki, A.; Yang, H.W.; Hayer, A.; Carrasco, S.; Malmersjo, S.; Meyer, T. A polarized Ca2+, diacylglycerol and STIM1 signalling system regulates directed cell migration. Nat. Cell Biol. 2014, 16, 133–144. [Google Scholar] [CrossRef] [PubMed]
  224. Avraham, H.K.; Lee, T.H.; Koh, Y.; Kim, T.A.; Jiang, S.; Sussman, M.; Samarel, A.M.; Avraham, S. Vascular endothelial growth factor regulates focal adhesion assembly in human brain microvascular endothelial cells through activation of the focal adhesion kinase and related adhesion focal tyrosine kinase. J. Biol. Chem. 2003, 278, 36661–36668. [Google Scholar] [CrossRef] [PubMed]
  225. Brouet, A.; Sonveaux, P.; Dessy, C.; Balligand, J.L.; Feron, O. Hsp90 ensures the transition from the early Ca2+-dependent to the late phosphorylation-dependent activation of the endothelial nitric-oxide synthase in vascular endothelial growth factor-exposed endothelial cells. J. Biol. Chem. 2001, 276, 32663–32669. [Google Scholar] [CrossRef] [PubMed]
  226. Gelinas, D.S.; Bernatchez, P.N.; Rollin, S.; Bazan, N.G.; Sirois, M.G. Immediate and delayed VEGF-mediated NO synthesis in endothelial cells: Role of PI3K, PKC and PLC pathways. Br. J. Pharmacol. 2002, 137, 1021–1030. [Google Scholar] [CrossRef] [PubMed]
  227. Chen, F.; Zhu, L.; Cai, L.; Zhang, J.; Zeng, X.; Li, J.; Su, Y.; Hu, Q. A stromal interaction molecule 1 variant up-regulates matrix metalloproteinase-2 expression by strengthening nucleoplasmic Ca2+ signaling. Biochim. Biophys. Acta 2016, 1863, 617–629. [Google Scholar] [CrossRef] [PubMed]
  228. Kim, I.; Moon, S.O.; Kim, S.H.; Kim, H.J.; Koh, Y.S.; Koh, G.Y. Vascular endothelial growth factor expression of intercellular adhesion molecule 1 (ICAM-1), vascular cell adhesion molecule 1 (VCAM-1), and E-selectin through nuclear factor-kappa B activation in endothelial cells. J. Biol. Chem. 2001, 276, 7614–7620. [Google Scholar] [CrossRef] [PubMed]
  229. Zhu, L.; Song, S.; Pi, Y.; Yu, Y.; She, W.; Ye, H.; Su, Y.; Hu, Q. Cumulated Ca2+ spike duration underlies Ca2+ oscillation frequency-regulated NFkappaB transcriptional activity. J. Cell Sci. 2011, 124, 2591–2601. [Google Scholar] [CrossRef]
  230. Courtwright, A.; Siamakpour-Reihani, S.; Arbiser, J.L.; Banet, N.; Hilliard, E.; Fried, L.; Livasy, C.; Ketelsen, D.; Nepal, D.B.; Perou, C.M.; et al. Secreted frizzle-related protein 2 stimulates angiogenesis via a calcineurin/NFAT signaling pathway. Cancer Res. 2009, 69, 4621–4628. [Google Scholar] [CrossRef]
  231. Scholz, B.; Korn, C.; Wojtarowicz, J.; Mogler, C.; Augustin, I.; Boutros, M.; Niehrs, C.; Augustin, H.G. Endothelial RSPO3 Controls Vascular Stability and Pruning through Non-canonical WNT/Ca2+/NFAT Signaling. Dev. Cell 2016, 36, 79–93. [Google Scholar] [CrossRef]
  232. Takahashi, T.; Ueno, H.; Shibuya, M. VEGF activates protein kinase C-dependent, but Ras-independent Raf-MEK-MAP kinase pathway for DNA synthesis in primary endothelial cells. Oncogene 1999, 18, 2221–2230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Xia, P.; Aiello, L.P.; Ishii, H.; Jiang, Z.Y.; Park, D.J.; Robinson, G.S.; Takagi, H.; Newsome, W.P.; Jirousek, M.R.; King, G.L. Characterization of vascular endothelial growth factor’s effect on the activation of protein kinase C, its isoforms, and endothelial cell growth. J. Clin. Investig. 1996, 98, 2018–2026. [Google Scholar] [CrossRef] [PubMed]
  234. Parenti, A.; Morbidelli, L.; Cui, X.L.; Douglas, J.G.; Hood, J.D.; Granger, H.J.; Ledda, F.; Ziche, M. Nitric oxide is an upstream signal of vascular endothelial growth factor-induced extracellular signal-regulated kinase1/2 activation in postcapillary endothelium. J. Biol. Chem. 1998, 273, 4220–4226. [Google Scholar] [CrossRef] [PubMed]
  235. Lee, M.Y.; Luciano, A.K.; Ackah, E.; Rodriguez-Vita, J.; Bancroft, T.A.; Eichmann, A.; Simons, M.; Kyriakides, T.R.; Morales-Ruiz, M.; Sessa, W.C. Endothelial Akt1 mediates angiogenesis by phosphorylating multiple angiogenic substrates. Proc. Natl. Acad. Sci. USA 2014, 111, 12865–12870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Carmeliet, P.; Lampugnani, M.G.; Moons, L.; Breviario, F.; Compernolle, V.; Bono, F.; Balconi, G.; Spagnuolo, R.; Oosthuyse, B.; Dewerchin, M.; et al. Targeted deficiency or cytosolic truncation of the VE-cadherin gene in mice impairs VEGF-mediated endothelial survival and angiogenesis. Cell 1999, 98, 147–157. [Google Scholar] [CrossRef]
  237. Ruan, G.X.; Kazlauskas, A. Axl is essential for VEGF-A-dependent activation of PI3K/Akt. EMBO J. 2012, 31, 1692–1703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Cullen, P.J.; Lockyer, P.J. Integration of calcium and Ras signalling. Nat. Rev. Mol. Cell Biol. 2002, 3, 339–348. [Google Scholar] [CrossRef]
  239. Yoshioka, K.; Sugimoto, N.; Takuwa, N.; Takuwa, Y. Essential role for class II phosphoinositide 3-kinase alpha-isoform in Ca2+-induced, Rho- and Rho kinase-dependent regulation of myosin phosphatase and contraction in isolated vascular smooth muscle cells. Mol. Pharmacol. 2007, 71, 912–920. [Google Scholar] [CrossRef]
  240. Muller, M.R.; Rao, A. NFAT, immunity and cancer: A transcription factor comes of age. Nat. Rev. Immunol. 2010, 10, 645–656. [Google Scholar] [CrossRef]
  241. Graef, I.A.; Chen, F.; Chen, L.; Kuo, A.; Crabtree, G.R. Signals transduced by Ca2+/calcineurin and NFATc3/c4 pattern the developing vasculature. Cell 2001, 105, 863–875. [Google Scholar] [CrossRef]
  242. Kar, P.; Mirams, G.R.; Christian, H.C.; Parekh, A.B. Control of NFAT Isoform Activation and NFAT-Dependent Gene Expression through Two Coincident and Spatially Segregated Intracellular Ca2+ Signals. Mol. Cell 2016, 64, 746–759. [Google Scholar] [CrossRef] [PubMed]
  243. Lin, Y.P.; Bakowski, D.; Mirams, G.R.; Parekh, A.B. Selective recruitment of different Ca2+-dependent transcription factors by STIM1-Orai1 channel clusters. Nat. Commun. 2019, 10, 2516. [Google Scholar] [CrossRef] [PubMed]
  244. Noy, P.J.; Gavin, R.L.; Colombo, D.; Haining, E.J.; Reyat, J.S.; Payne, H.; Thielmann, I.; Lokman, A.B.; Neag, G.; Yang, J.; et al. Tspan18 is a novel regulator of the Ca2+ channel Orai1 and von Willebrand factor release in endothelial cells. Haematologica 2018. [Google Scholar] [CrossRef] [PubMed]
  245. Schweighofer, B.; Testori, J.; Sturtzel, C.; Sattler, S.; Mayer, H.; Wagner, O.; Bilban, M.; Hofer, E. The VEGF-induced transcriptional response comprises gene clusters at the crossroad of angiogenesis and inflammation. Thromb. Haemost. 2009, 102, 544–554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Minami, T.; Horiuchi, K.; Miura, M.; Abid, M.R.; Takabe, W.; Noguchi, N.; Kohro, T.; Ge, X.; Aburatani, H.; Hamakubo, T.; et al. Vascular endothelial growth factor- and thrombin-induced termination factor, Down syndrome critical region-1, attenuates endothelial cell proliferation and angiogenesis. J. Biol. Chem. 2004, 279, 50537–50554. [Google Scholar] [CrossRef] [PubMed]
  247. Schabbauer, G.; Schweighofer, B.; Mechtcheriakova, D.; Lucerna, M.; Binder, B.R.; Hofer, E. Nuclear factor of activated T cells and early growth response-1 cooperate to mediate tissue factor gene induction by vascular endothelial growth factor in endothelial cells. Thromb. Haemost. 2007, 97, 988–997. [Google Scholar] [CrossRef] [PubMed]
  248. Liu, D.; Jia, H.; Holmes, D.I.; Stannard, A.; Zachary, I. Vascular endothelial growth factor-regulated gene expression in endothelial cells: KDR-mediated induction of Egr3 and the related nuclear receptors Nur77, Nurr1, and Nor1. Arterioscler. Thromb. Vasc. Biol. 2003, 23, 2002–2007. [Google Scholar] [CrossRef]
  249. Fuentes, J.J.; Genesca, L.; Kingsbury, T.J.; Cunningham, K.W.; Perez-Riba, M.; Estivill, X.; de la Luna, S. DSCR1, overexpressed in Down syndrome, is an inhibitor of calcineurin-mediated signaling pathways. Hum. Mol. Genet. 2000, 9, 1681–1690. [Google Scholar] [CrossRef] [Green Version]
  250. Iizuka, M.; Abe, M.; Shiiba, K.; Sasaki, I.; Sato, Y. Down syndrome candidate region 1,a downstream target of VEGF, participates in endothelial cell migration and angiogenesis. J. Vasc. Res. 2004, 41, 334–344. [Google Scholar] [CrossRef]
  251. Alghanem, A.F.; Wilkinson, E.L.; Emmett, M.S.; Aljasir, M.A.; Holmes, K.; Rothermel, B.A.; Simms, V.A.; Heath, V.L.; Cross, M.J. RCAN1.4 regulates VEGFR-2 internalisation, cell polarity and migration in human microvascular endothelial cells. Angiogenesis 2017, 20, 341–358. [Google Scholar] [CrossRef]
  252. Robbs, B.K.; Cruz, A.L.; Werneck, M.B.; Mognol, G.P.; Viola, J.P. Dual roles for NFAT transcription factor genes as oncogenes and tumor suppressors. Mol. Cell. Biol. 2008, 28, 7168–7181. [Google Scholar] [CrossRef] [PubMed]
  253. Suehiro, J.; Kanki, Y.; Makihara, C.; Schadler, K.; Miura, M.; Manabe, Y.; Aburatani, H.; Kodama, T.; Minami, T. Genome-wide approaches reveal functional vascular endothelial growth factor (VEGF)-inducible nuclear factor of activated T cells (NFAT) c1 binding to angiogenesis-related genes in the endothelium. J. Biol. Chem. 2014, 289, 29044–29059. [Google Scholar] [CrossRef] [PubMed]
  254. Lisman, J.; Yasuda, R.; Raghavachari, S. Mechanisms of CaMKII action in long-term potentiation. Nat. Rev. Neurosci. 2012, 13, 169–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Beckendorf, J.; van den Hoogenhof, M.M.G.; Backs, J. Physiological and unappreciated roles of CaMKII in the heart. Basic Res. Cardiol. 2018, 113, 29. [Google Scholar] [CrossRef] [PubMed]
  256. Toussaint, F.; Charbel, C.; Allen, B.G.; Ledoux, J. Vascular CaMKII: Heart and brain in your arteries. Am. J. Physiol. Cell Physiol. 2016, 311, C462–C478. [Google Scholar] [CrossRef] [PubMed]
  257. Toussaint, F.; Charbel, C.; Blanchette, A.; Ledoux, J. CaMKII regulates intracellular Ca2+ dynamics in native endothelial cells. Cell Calcium 2015, 58, 275–285. [Google Scholar] [CrossRef] [PubMed]
  258. Chakravarti, B.; Yang, J.; Ahlers-Dannen, K.E.; Luo, Z.; Flaherty, H.A.; Meyerholz, D.K.; Anderson, M.E.; Fisher, R.A. Essentiality of Regulator of G Protein Signaling 6 and Oxidized Ca2+/Calmodulin-Dependent Protein Kinase II in Notch Signaling and Cardiovascular Development. J. Am. Heart Assoc. 2017, 6, e007038. [Google Scholar] [CrossRef]
  259. Wei, C.; Wang, X.; Chen, M.; Ouyang, K.; Song, L.S.; Cheng, H. Calcium flickers steer cell migration. Nature 2009, 457, 901–905. [Google Scholar] [CrossRef] [Green Version]
  260. Tsai, F.C.; Meyer, T. Ca2+ pulses control local cycles of lamellipodia retraction and adhesion along the front of migrating cells. Curr. Biol. 2012, 22, 837–842. [Google Scholar] [CrossRef]
  261. Iamshanova, O.; Fiorio Pla, A.; Prevarskaya, N. Molecular mechanisms of tumour invasion: Regulation by calcium signals. J. Physiol. 2017, 595, 3063–3075. [Google Scholar] [CrossRef]
  262. Lamalice, L.; Le Boeuf, F.; Huot, J. Endothelial cell migration during angiogenesis. Circ. Res. 2007, 100, 782–794. [Google Scholar] [CrossRef]
  263. Ariyoshi, H.; Yoshikawa, N.; Aono, Y.; Tsuji, Y.; Ueda, A.; Tokunaga, M.; Sakon, M.; Monden, M. Localized activation of m-calpain in migrating human umbilical vein endothelial cells stimulated by shear stress. J. Cell. Biochem. 2001, 81, 184–192. [Google Scholar] [CrossRef]
  264. Yoshikawa, N.; Ariyoshi, H.; Aono, Y.; Sakon, M.; Kawasaki, T.; Monden, M. Gradients in cytoplasmic calcium concentration ([Ca2+]i) in migrating human umbilical vein endothelial cells (HUVECs) stimulated by shear-stress. Life Sci. 1999, 65, 2643–2651. [Google Scholar] [CrossRef]
  265. Miyazaki, T.; Honda, K.; Ohata, H. Requirement of Ca2+ influx-and phosphatidylinositol 3-kinase-mediated m-calpain activity for shear stress-induced endothelial cell polarity. Am. J. Physiol. Cell Physiol. 2007, 293, C1216–C1225. [Google Scholar] [CrossRef]
  266. Zeng, Z.; Inoue, K.; Sun, H.; Leng, T.; Feng, X.; Zhu, L.; Xiong, Z.G. TRPM7 regulates vascular endothelial cell adhesion and tube formation. Am. J. Physiol. Cell Physiol. 2015, 308, C308–C318. [Google Scholar] [CrossRef] [Green Version]
  267. Baldoli, E.; Maier, J.A. Silencing TRPM7 mimics the effects of magnesium deficiency in human microvascular endothelial cells. Angiogenesis 2012, 15, 47–57. [Google Scholar] [CrossRef]
  268. Fukumura, D.; Gohongi, T.; Kadambi, A.; Izumi, Y.; Ang, J.; Yun, C.O.; Buerk, D.G.; Huang, P.L.; Jain, R.K. Predominant role of endothelial nitric oxide synthase in vascular endothelial growth factor-induced angiogenesis and vascular permeability. Proc. Natl. Acad. Sci. USA 2001, 98, 2604–2609. [Google Scholar] [CrossRef] [Green Version]
  269. Papapetropoulos, A.; Garcia-Cardena, G.; Madri, J.A.; Sessa, W.C. Nitric oxide production contributes to the angiogenic properties of vascular endothelial growth factor in human endothelial cells. J. Clin. Investig. 1997, 100, 3131–3139. [Google Scholar] [CrossRef]
  270. Shu, X.; Keller, T.C.T.; Begandt, D.; Butcher, J.T.; Biwer, L.; Keller, A.S.; Columbus, L.; Isakson, B.E. Endothelial nitric oxide synthase in the microcirculation. Cell. Mol. Life Sci. 2015, 72, 4561–4575. [Google Scholar] [CrossRef]
  271. Reihill, J.A.; Ewart, M.A.; Hardie, D.G.; Salt, I.P. AMP-activated protein kinase mediates VEGF-stimulated endothelial NO production. Biochem. Biophys. Res. Commun. 2007, 354, 1084–1088. [Google Scholar] [CrossRef] [Green Version]
  272. Faehling, M.; Koch, E.D.; Raithel, J.; Trischler, G.; Waltenberger, J. Vascular endothelial growth factor-A activates Ca2+ -activated K+ channels in human endothelial cells in culture. Int. J. Biochem. Cell Biol. 2001, 33, 337–346. [Google Scholar] [CrossRef]
  273. De, A. Wnt/Ca2+ signaling pathway: A brief overview. Acta Biochim. Biophys. Sin. 2011, 43, 745–756. [Google Scholar] [CrossRef]
  274. Carvalho, J.R.; Fortunato, I.C.; Fonseca, C.G.; Pezzarossa, A.; Barbacena, P.; Dominguez-Cejudo, M.A.; Vasconcelos, F.F.; Santos, N.C.; Carvalho, F.A.; Franco, C.A. Non-canonical Wnt signaling regulates junctional mechanocoupling during angiogenic collective cell migration. eLife 2019, 8. [Google Scholar] [CrossRef]
  275. Shi, Y.N.; Zhu, N.; Liu, C.; Wu, H.T.; Gui, Y.; Liao, D.F.; Qin, L. Wnt5a and its signaling pathway in angiogenesis. Clin. Chim. Acta 2017, 471, 263–269. [Google Scholar] [CrossRef]
  276. Korn, C.; Scholz, B.; Hu, J.; Srivastava, K.; Wojtarowicz, J.; Arnsperger, T.; Adams, R.H.; Boutros, M.; Augustin, H.G.; Augustin, I. Endothelial cell-derived non-canonical Wnt ligands control vascular pruning in angiogenesis. Development 2014, 141, 1757–1766. [Google Scholar] [CrossRef] [Green Version]
  277. Arderiu, G.; Espinosa, S.; Pena, E.; Aledo, R.; Badimon, L. Monocyte-secreted Wnt5a interacts with FZD5 in microvascular endothelial cells and induces angiogenesis through tissue factor signaling. J. Mol. Cell. Biol. 2014, 6, 380–393. [Google Scholar] [CrossRef] [Green Version]
  278. Stefater, J.A., III; Rao, S.; Bezold, K.; Aplin, A.C.; Nicosia, R.F.; Pollard, J.W.; Ferrara, N.; Lang, R.A. Macrophage Wnt-Calcineurin-Flt1 signaling regulates mouse wound angiogenesis and repair. Blood 2013, 121, 2574–2578. [Google Scholar] [CrossRef] [Green Version]
  279. Cheng, C.W.; Yeh, J.C.; Fan, T.P.; Smith, S.K.; Charnock-Jones, D.S. Wnt5a-mediated non-canonical Wnt signalling regulates human endothelial cell proliferation and migration. Biochem. Biophys. Res. Commun. 2008, 365, 285–290. [Google Scholar] [CrossRef]
  280. Medina, R.J.; Barber, C.L.; Sabatier, F.; Dignat-George, F.; Melero-Martin, J.M.; Khosrotehrani, K.; Ohneda, O.; Randi, A.M.; Chan, J.K.Y.; Yamaguchi, T.; et al. Endothelial Progenitors: A Consensus Statement on Nomenclature. Stem Cells Transl. Med. 2017, 6, 1316–1320. [Google Scholar] [CrossRef]
  281. O’Neill, C.L.; McLoughlin, K.J.; Chambers, S.E.J.; Guduric-Fuchs, J.; Stitt, A.W.; Medina, R.J. The Vasoreparative Potential of Endothelial Colony Forming Cells: A Journey Through Pre-clinical Studies. Front. Med. 2018, 5, 273. [Google Scholar] [CrossRef] [Green Version]
  282. Moccia, F.; Ruffinatti, F.A.; Zuccolo, E. Intracellular Ca2+ Signals to Reconstruct A Broken Heart: Still A Theoretical Approach? Curr. Drug Targets 2015, 16, 793–815. [Google Scholar] [CrossRef]
  283. Poletto, V.; Rosti, V.; Biggiogera, M.; Guerra, G.; Moccia, F.; Porta, C. The role of endothelial colony forming cells in kidney cancer’s pathogenesis, and in resistance to anti-VEGFR agents and mTOR inhibitors: A speculative review. Crit. Rev. Oncol. Hematol. 2018, 132, 89–99. [Google Scholar] [CrossRef]
  284. Moccia, F.; Zuccolo, E.; Poletto, V.; Cinelli, M.; Bonetti, E.; Guerra, G.; Rosti, V. Endothelial progenitor cells support tumour growth and metastatisation: Implications for the resistance to anti-angiogenic therapy. Tumour Biol. 2015, 36, 6603–6614. [Google Scholar] [CrossRef]
  285. Poletto, V.; Dragoni, S.; Lim, D.; Biggiogera, M.; Aronica, A.; Cinelli, M.; De Luca, A.; Rosti, V.; Porta, C.; Guerra, G.; et al. Endoplasmic Reticulum Ca2+ Handling and Apoptotic Resistance in Tumor-Derived Endothelial Colony Forming Cells. J. Cell. Biochem. 2016, 117, 2260–2271. [Google Scholar] [CrossRef]
  286. Dragoni, S.; Laforenza, U.; Bonetti, E.; Lodola, F.; Bottino, C.; Berra-Romani, R.; Carlo Bongio, G.; Cinelli, M.P.; Guerra, G.; Pedrazzoli, P.; et al. Vascular endothelial growth factor stimulates endothelial colony forming cells proliferation and tubulogenesis by inducing oscillations in intracellular Ca2+ concentration. Stem Cells 2011, 29, 1898–1907. [Google Scholar] [CrossRef]
  287. Zuccolo, E.; Dragoni, S.; Poletto, V.; Catarsi, P.; Guido, D.; Rappa, A.; Reforgiato, M.; Lodola, F.; Lim, D.; Rosti, V.; et al. Arachidonic acid-evoked Ca2+ signals promote nitric oxide release and proliferation in human endothelial colony forming cells. Vascul. Pharmacol. 2016, 87, 159–171. [Google Scholar] [CrossRef]
  288. Sanchez-Hernandez, Y.; Laforenza, U.; Bonetti, E.; Fontana, J.; Dragoni, S.; Russo, M.; Avelino-Cruz, J.E.; Schinelli, S.; Testa, D.; Guerra, G.; et al. Store-operated Ca2+ entry is expressed in human endothelial progenitor cells. Stem Cells Dev. 2010, 19, 1967–1981. [Google Scholar] [CrossRef]
  289. Maeng, Y.S.; Choi, H.J.; Kwon, J.Y.; Park, Y.W.; Choi, K.S.; Min, J.K.; Kim, Y.H.; Suh, P.G.; Kang, K.S.; Won, M.H.; et al. Endothelial progenitor cell homing: Prominent role of the IGF2-IGF2R-PLCbeta2 axis. Blood 2009, 113, 233–243. [Google Scholar] [CrossRef]
  290. Lodola, F.; Laforenza, U.; Bonetti, E.; Lim, D.; Dragoni, S.; Bottino, C.; Ong, H.L.; Guerra, G.; Ganini, C.; Massa, M.; et al. Store-operated Ca2+ entry is remodelled and controls in vitro angiogenesis in endothelial progenitor cells isolated from tumoral patients. PLoS ONE 2012, 7, e42541. [Google Scholar] [CrossRef]
  291. Dragoni, S.; Laforenza, U.; Bonetti, E.; Lodola, F.; Bottino, C.; Guerra, G.; Borghesi, A.; Stronati, M.; Rosti, V.; Tanzi, F.; et al. Canonical transient receptor potential 3 channel triggers vascular endothelial growth factor-induced intracellular Ca2+ oscillations in endothelial progenitor cells isolated from umbilical cord blood. Stem Cells Dev. 2013, 22, 2561–2580. [Google Scholar] [CrossRef]
  292. Dragoni, S.; Guerra, G.; Fiorio Pla, A.; Bertoni, G.; Rappa, A.; Poletto, V.; Bottino, C.; Aronica, A.; Lodola, F.; Cinelli, M.P.; et al. A functional Transient Receptor Potential Vanilloid 4 (TRPV4) channel is expressed in human endothelial progenitor cells. J. Cell. Physiol. 2015, 230, 95–104. [Google Scholar] [CrossRef] [PubMed]
  293. Tu, T.C.; Nagano, M.; Yamashita, T.; Hamada, H.; Ohneda, K.; Kimura, K.; Ohneda, O. A Chemokine Receptor, CXCR4, Which Is Regulated by Hypoxia-Inducible Factor 2alpha, Is Crucial for Functional Endothelial Progenitor Cells Migration to Ischemic Tissue and Wound Repair. Stem Cells Dev. 2016, 25, 266–276. [Google Scholar] [CrossRef] [PubMed]
  294. Zuccolo, E.; Di Buduo, C.; Lodola, F.; Orecchioni, S.; Scarpellino, G.; Kheder, D.A.; Poletto, V.; Guerra, G.; Bertolini, F.; Balduini, A.; et al. Stromal Cell-Derived Factor-1alpha Promotes Endothelial Colony-Forming Cell Migration Through the Ca2+-Dependent Activation of the Extracellular Signal-Regulated Kinase 1/2 and Phosphoinositide 3-Kinase/AKT Pathways. Stem Cells Dev. 2018, 27, 23–34. [Google Scholar] [CrossRef] [PubMed]
  295. Dragoni, S.; Turin, I.; Laforenza, U.; Potenza, D.M.; Bottino, C.; Glasnov, T.N.; Prestia, M.; Ferulli, F.; Saitta, A.; Mosca, A.; et al. Store-operated Ca2+ entry does not control proliferation in primary cultures of human metastatic renal cellular carcinoma. BioMed Res. Int. 2014, 2014, 739494. [Google Scholar] [CrossRef] [PubMed]
  296. Joo, H.J.; Song, S.; Seo, H.R.; Shin, J.H.; Choi, S.C.; Park, J.H.; Yu, C.W.; Hong, S.J.; Lim, D.S. Human endothelial colony forming cells from adult peripheral blood have enhanced sprouting angiogenic potential through up-regulating VEGFR2 signaling. Int. J. Cardiol. 2015, 197, 33–43. [Google Scholar] [CrossRef] [PubMed]
  297. Su, S.H.; Wu, C.H.; Chiu, Y.L.; Chang, S.J.; Lo, H.H.; Liao, K.H.; Tsai, C.F.; Tsai, T.N.; Lin, C.H.; Cheng, S.M.; et al. Dysregulation of Vascular Endothelial Growth Factor Receptor-2 by Multiple miRNAs in Endothelial Colony-Forming Cells of Coronary Artery Disease. J. Vasc. Res. 2017, 54, 22–32. [Google Scholar] [CrossRef] [PubMed]
  298. Ingram, D.A.; Mead, L.E.; Tanaka, H.; Meade, V.; Fenoglio, A.; Mortell, K.; Pollok, K.; Ferkowicz, M.J.; Gilley, D.; Yoder, M.C. Identification of a novel hierarchy of endothelial progenitor cells using human peripheral and umbilical cord blood. Blood 2004, 104, 2752–2760. [Google Scholar] [CrossRef]
  299. Wu, Y.; He, M.Y.; Ye, J.K.; Ma, S.Y.; Huang, W.; Wei, Y.Y.; Kong, H.; Wang, H.; Zeng, X.N.; Xie, W.P. Activation of ATP-sensitive potassium channels facilitates the function of human endothelial colony-forming cells via Ca2+/Akt/eNOS pathway. J. Cell. Mol. Med. 2017, 21, 609–620. [Google Scholar] [CrossRef]
  300. Aromolaran, A.S.; Zima, A.V.; Blatter, L.A. Role of glycolytically generated ATP for CaMKII-mediated regulation of intracellular Ca2+ signaling in bovine vascular endothelial cells. Am. J. Physiol. Cell Physiol. 2007, 293, C106–C118. [Google Scholar] [CrossRef]
  301. Regueiro, A.; Cuadrado-Godia, E.; Bueno-Beti, C.; Diaz-Ricart, M.; Oliveras, A.; Novella, S.; Gene, G.G.; Jung, C.; Subirana, I.; Ortiz-Perez, J.T.; et al. Mobilization of endothelial progenitor cells in acute cardiovascular events in the PROCELL study: Time-course after acute myocardial infarction and stroke. J. Mol. Cell. Cardiol. 2015, 80, 146–155. [Google Scholar] [CrossRef]
  302. Di Nezza, F.; Zuccolo, E.; Poletto, V.; Rosti, V.; De Luca, A.; Moccia, F.; Guerra, G.; Ambrosone, L. Liposomes as a Putative Tool to Investigate NAADP Signaling in Vasculogenesis. J. Cell. Biochem. 2017, 118, 3722–3729. [Google Scholar] [CrossRef] [PubMed]
  303. Zuccolo, E.; Lim, D.; Poletto, V.; Guerra, G.; Tanzi, F.; Rosti, V.; Moccia, F. Acidic Ca2+ stores interact with the endoplasmic reticulum to shape intracellular Ca2+ signals in human endothelial progenitor cells. Vascul. Pharmacol. 2015, 75, 70–71. [Google Scholar] [CrossRef]
  304. Moccia, F. Endothelial Ca2+ Signaling and the Resistance to Anticancer Treatments: Partners in Crime. Int. J. Mol. Sci. 2018, 19, 217. [Google Scholar] [CrossRef] [PubMed]
  305. Moccia, F. Remodelling of the Ca2+ Toolkit in Tumor Endothelium as a Crucial Responsible for the Resistance to Anticancer Therapies. Curr. Signal Transduct. Ther. 2017, 12, 3–18. [Google Scholar] [CrossRef]
  306. Thoppil, R.J.; Adapala, R.K.; Cappelli, H.C.; Kondeti, V.; Dudley, A.C.; Gary Meszaros, J.; Paruchuri, S.; Thodeti, C.K. TRPV4 channel activation selectively inhibits tumor endothelial cell proliferation. Sci. Rep. 2015, 5, 14257. [Google Scholar] [CrossRef] [PubMed]
  307. Bernardini, M.; Brossa, A.; Chinigo, G.; Grolez, G.P.; Trimaglio, G.; Allart, L.; Hulot, A.; Marot, G.; Genova, T.; Joshi, A.; et al. Transient Receptor Potential Channel Expression Signatures in Tumor-Derived Endothelial Cells: Functional Roles in Prostate Cancer Angiogenesis. Cancers 2019, 11, 956. [Google Scholar] [CrossRef]
  308. Dragoni, S.; Reforgiato, M.; Zuccolo, E.; Poletto, V.; Lodola, F.; Ruffinatti, F.A.; Bonetti, E.; Guerra, G.; Barosi, G.; Rosti, V.; et al. Dysregulation of VEGF-induced proangiogenic Ca2+ oscillations in primary myelofibrosis-derived endothelial colony-forming cells. Exp. Hematol. 2015, 43, 1019–1030. [Google Scholar] [CrossRef] [PubMed]
  309. Lodola, F.; Laforenza, U.; Cattaneo, F.; Ruffinatti, F.A.; Poletto, V.; Massa, M.; Tancredi, R.; Zuccolo, E.; Khdar, A.D.; Riccardi, A.; et al. VEGF-induced intracellular Ca2+ oscillations are down-regulated and do not stimulate angiogenesis in breast cancer-derived endothelial colony forming cells. Oncotarget 2017, 8, 95223–95246. [Google Scholar] [CrossRef]
  310. Moccia, F.; Fotia, V.; Tancredi, R.; Della Porta, M.G.; Rosti, V.; Bonetti, E.; Poletto, V.; Marchini, S.; Beltrame, L.; Gallizzi, G.; et al. Breast and renal cancer-Derived endothelial colony forming cells share a common gene signature. Eur. J. Cancer 2017, 77, 155–164. [Google Scholar] [CrossRef]
  311. Zuccolo, E.; Bottino, C.; Diofano, F.; Poletto, V.; Codazzi, A.C.; Mannarino, S.; Campanelli, R.; Fois, G.; Marseglia, G.L.; Guerra, G.; et al. Constitutive Store-Operated Ca2+ Entry Leads to Enhanced Nitric Oxide Production and Proliferation in Infantile Hemangioma-Derived Endothelial Colony-Forming Cells. Stem Cells Dev. 2016, 25, 301–319. [Google Scholar] [CrossRef]
  312. Moccia, F.; Dragoni, S.; Poletto, V.; Rosti, V.; Tanzi, F.; Ganini, C.; Porta, C. Orai1 and Transient Receptor Potential Channels as novel molecular targets to impair tumor neovascularisation in renal cell carcinoma and other malignancies. Anticancer Agents Med. Chem. 2014, 14, 296–312. [Google Scholar] [CrossRef] [PubMed]
  313. Kuang, C.Y.; Yu, Y.; Guo, R.W.; Qian, D.H.; Wang, K.; Den, M.Y.; Shi, Y.K.; Huang, L. Silencing stromal interaction molecule 1 by RNA interference inhibits the proliferation and migration of endothelial progenitor cells. Biochem. Biophys. Res. Commun. 2010, 398, 315–320. [Google Scholar] [CrossRef] [PubMed]
  314. Kuang, C.Y.; Yu, Y.; Wang, K.; Qian, D.H.; Den, M.Y.; Huang, L. Knockdown of transient receptor potential canonical-1 reduces the proliferation and migration of endothelial progenitor cells. Stem Cells Dev. 2012, 21, 487–496. [Google Scholar] [CrossRef] [PubMed]
  315. Shi, Y.; Song, M.; Guo, R.; Wang, H.; Gao, P.; Shi, W.; Huang, L. Knockdown of stromal interaction molecule 1 attenuates hepatocyte growth factor-induced endothelial progenitor cell proliferation. Exp. Biol. Med 2010, 235, 317–325. [Google Scholar] [CrossRef] [PubMed]
  316. Wang, L.Y.; Zhang, J.H.; Yu, J.; Yang, J.; Deng, M.Y.; Kang, H.L.; Huang, L. Reduction of Store-Operated Ca2+ Entry Correlates with Endothelial Progenitor Cell Dysfunction in Atherosclerotic Mice. Stem Cells Dev. 2015, 24, 1582–1590. [Google Scholar] [CrossRef] [PubMed]
  317. Du, L.L.; Shen, Z.; Li, Z.; Ye, X.; Wu, M.; Hong, L.; Zhao, Y. TRPC1 Deficiency Impairs the Endothelial Progenitor Cell Function via Inhibition of Calmodulin/eNOS Pathway. J. Cardiovasc. Transl. Res. 2018, 11, 339–345. [Google Scholar] [CrossRef] [PubMed]
  318. Fantozzi, I.; Zhang, S.; Platoshyn, O.; Remillard, C.V.; Cowling, R.T.; Yuan, J.X. Hypoxia increases AP-1 binding activity by enhancing capacitative Ca2+ entry in human pulmonary artery endothelial cells. Am. J. Physiol. Lung Cell Mol. Physiol. 2003, 285, L1233–1245. [Google Scholar] [CrossRef] [PubMed]
  319. Moccia, F.; Dragoni, S.; Cinelli, M.; Montagnani, S.; Amato, B.; Rosti, V.; Guerra, G.; Tanzi, F. How to utilize Ca2+ signals to rejuvenate the repairative phenotype of senescent endothelial progenitor cells in elderly patients affected by cardiovascular diseases: A useful therapeutic support of surgical approach? BMC Surg. 2013, 13, 46. [Google Scholar] [CrossRef]
  320. Balbi, C.; Lodder, K.; Costa, A.; Moimas, S.; Moccia, F.; van Herwaarden, T.; Rosti, V.; Campagnoli, F.; Palmeri, A.; De Biasio, P.; et al. Reactivating endogenous mechanisms of cardiac regeneration via paracrine boosting using the human amniotic fluid stem cell secretome. Int. J. Cardiol. 2019, 287, 87–95. [Google Scholar] [CrossRef]
  321. Favia, A.; Pafumi, I.; Desideri, M.; Padula, F.; Montesano, C.; Passeri, D.; Nicoletti, C.; Orlandi, A.; Del Bufalo, D.; Sergi, M.; et al. NAADP-Dependent Ca2+ Signaling Controls Melanoma Progression, Metastatic Dissemination and Neoangiogenesis. Sci. Rep. 2016, 6, 18925. [Google Scholar] [CrossRef]
Figure 1. The pro-angiogenic Ca2+ toolkit in vascular endothelial cells. Pro-angiogenic cues, such as growth factors and chemokines, bind to specific receptor tyrosine kinases (RTK) and Gq/11-protein Coupled Receptors (Gq/11PCR) thereby activating multiple phospholipase C (PLC) isoforms, which in turn cleave phosphatidylinositol 4,5-bisphosphate (PIP2) into inositol-1,4,5-trisphosphate (InsP3) and diacylglycerol (DAG). InsP3 triggers Ca2+ release from the endoplasmic reticulum (ER) through InsP3 receptors (InsP3R), while DAG stimulates extracellular Ca2+ entry through TRPC3 and TRPC6. However, the major Ca2+-entry pathway in vascular endothelial cells is provided by store-operated Ca2+ entry (SOCE), which is mainly mediated by the physical interaction between STIM1 and Orai1. In addition, SOCE may be sustained by the interplay among STIM1, Transient Receptor Potential (TRP) Canonical 1 (TRPC1) and TRPC4, with [35] or without the involvement of Orai1. Endogenous Ca2+ release may also be sustained by ryanodine receptors (RyR, not shown) and by endolysosomal two-pore channel 1-2 (TPC1-2), which are gated by nicotinic acid adenine dinucleotide phosphate (NAADP). Multiple Ca2+-transporting systems maintain resting Ca2+ concentration and clear cytosolic Ca2+ after the pro-angiogenic signal. These include Sarco-Endoplasmic Reticulum Ca2+-ATPase 2a (SERCA2a), Plasma Membrane Ca2+-ATPase 1 (PMCA1) and PMCA4, and Na+/Ca2+ exchanger (NCX). Please, see the text for a more detailed description of how the endothelial Ca2+ toolkit is recruited by pro-angiogenic cues.
Figure 1. The pro-angiogenic Ca2+ toolkit in vascular endothelial cells. Pro-angiogenic cues, such as growth factors and chemokines, bind to specific receptor tyrosine kinases (RTK) and Gq/11-protein Coupled Receptors (Gq/11PCR) thereby activating multiple phospholipase C (PLC) isoforms, which in turn cleave phosphatidylinositol 4,5-bisphosphate (PIP2) into inositol-1,4,5-trisphosphate (InsP3) and diacylglycerol (DAG). InsP3 triggers Ca2+ release from the endoplasmic reticulum (ER) through InsP3 receptors (InsP3R), while DAG stimulates extracellular Ca2+ entry through TRPC3 and TRPC6. However, the major Ca2+-entry pathway in vascular endothelial cells is provided by store-operated Ca2+ entry (SOCE), which is mainly mediated by the physical interaction between STIM1 and Orai1. In addition, SOCE may be sustained by the interplay among STIM1, Transient Receptor Potential (TRP) Canonical 1 (TRPC1) and TRPC4, with [35] or without the involvement of Orai1. Endogenous Ca2+ release may also be sustained by ryanodine receptors (RyR, not shown) and by endolysosomal two-pore channel 1-2 (TPC1-2), which are gated by nicotinic acid adenine dinucleotide phosphate (NAADP). Multiple Ca2+-transporting systems maintain resting Ca2+ concentration and clear cytosolic Ca2+ after the pro-angiogenic signal. These include Sarco-Endoplasmic Reticulum Ca2+-ATPase 2a (SERCA2a), Plasma Membrane Ca2+-ATPase 1 (PMCA1) and PMCA4, and Na+/Ca2+ exchanger (NCX). Please, see the text for a more detailed description of how the endothelial Ca2+ toolkit is recruited by pro-angiogenic cues.
Ijms 20 03962 g001
Figure 2. Ca2+-dependent activation of the extracellular signal-regulated kinase (ERK) pathway and of the Nuclear Factor of Activated T-cells (NFAT). VEGF binding to VEGFR2 triggers an increase in intracellular Ca2+ concentration that may stimulate the ERK1/2 phosphorylation cascade or NFAT nuclear translocation. VEGF-induced endothelial Ca2+ signals may be mediated by multiple Ca2+-entry/release pathways, depending on species and vascular bed. The Ca2+ response to VEGF may recruit the Ca2+-dependent PKCβ2 (cPKCβ2), which engages the downstream RAF1–MEK–ERK1/2 cascade to induce gene expression. An increase in [Ca2+]i is required to promote cPKCβ2 translocation to the plasma membrane, where it is activated by DAG. Moreover, VEGF-induced endothelial Ca2+ signals may be sensed by calmodulin (CaM), which in turn activates calcineurin to dephosphorylate NFAT, thereby inducing its nuclear translocation. See Figure 1, Section 4.1 (ERK) and Section 4.3 (calcineurin and NFAT) for further details.
Figure 2. Ca2+-dependent activation of the extracellular signal-regulated kinase (ERK) pathway and of the Nuclear Factor of Activated T-cells (NFAT). VEGF binding to VEGFR2 triggers an increase in intracellular Ca2+ concentration that may stimulate the ERK1/2 phosphorylation cascade or NFAT nuclear translocation. VEGF-induced endothelial Ca2+ signals may be mediated by multiple Ca2+-entry/release pathways, depending on species and vascular bed. The Ca2+ response to VEGF may recruit the Ca2+-dependent PKCβ2 (cPKCβ2), which engages the downstream RAF1–MEK–ERK1/2 cascade to induce gene expression. An increase in [Ca2+]i is required to promote cPKCβ2 translocation to the plasma membrane, where it is activated by DAG. Moreover, VEGF-induced endothelial Ca2+ signals may be sensed by calmodulin (CaM), which in turn activates calcineurin to dephosphorylate NFAT, thereby inducing its nuclear translocation. See Figure 1, Section 4.1 (ERK) and Section 4.3 (calcineurin and NFAT) for further details.
Ijms 20 03962 g002
Figure 3. The Ca2+-dependent activation of Ca2+/Calmodulin (CaM)-dependent protein kinase 2 (CaMKII). Endothelial Ca2+ oscillations recruit CaMKII, which, in turns, stimulate angiogenesis by phosphorylating multiple targets, as widely illustrated in Section 4.4. VEGF-induced endothelial Ca2+ oscillations are sensed by CaM, which in turn stimulates CaMKII to phosphorylate multiple targets, including FAK to promote endothelial cell migration and Akt, JNK and Src to induce gene expression. The Ca2+ entry/release pathways that are recruited by VEGF to engage endothelial CaMKII are yet to be fully elucidated.
Figure 3. The Ca2+-dependent activation of Ca2+/Calmodulin (CaM)-dependent protein kinase 2 (CaMKII). Endothelial Ca2+ oscillations recruit CaMKII, which, in turns, stimulate angiogenesis by phosphorylating multiple targets, as widely illustrated in Section 4.4. VEGF-induced endothelial Ca2+ oscillations are sensed by CaM, which in turn stimulates CaMKII to phosphorylate multiple targets, including FAK to promote endothelial cell migration and Akt, JNK and Src to induce gene expression. The Ca2+ entry/release pathways that are recruited by VEGF to engage endothelial CaMKII are yet to be fully elucidated.
Ijms 20 03962 g003
Figure 4. The Ca2+-dependent activation of the endothelial nitric oxide (NO) synthase (eNOS). VEGF binding to VEGFR2 causes an increase in intracellular Ca2+ concentration that displaces caveolin 1 (CaV1) from eNOS, thereby removing the tonic inhibition and inducing NO release. VEGF may impinge on several Ca2+ entry/release pathways to engage eNOS, including InsP3R, TPC2 and Orai1. See Figure 1 and Section 4.6. for further details.
Figure 4. The Ca2+-dependent activation of the endothelial nitric oxide (NO) synthase (eNOS). VEGF binding to VEGFR2 causes an increase in intracellular Ca2+ concentration that displaces caveolin 1 (CaV1) from eNOS, thereby removing the tonic inhibition and inducing NO release. VEGF may impinge on several Ca2+ entry/release pathways to engage eNOS, including InsP3R, TPC2 and Orai1. See Figure 1 and Section 4.6. for further details.
Ijms 20 03962 g004
Figure 5. The Ca2+ toolkit in endothelial colony forming cells (ECFC). Growth factors, such as VEGF, and chemokines, such as stromal derived factor-1α (SDF-1α), bind to specific RTK and Gq/11PCR, thereby activating multiple PLC isoforms, which in turn cleave PIP2 into InsP3 and DAG. InsP3 triggers ER-dependent Ca2+ release through InsP3R, while DAG gates TRPC3 exclusively in umbilical cord blood-derived ECFCs. SOCE is the major Ca2+ entry pathway also in ECFC, in which it is mediated by the dynamic interplay among STIM1, Orai1 and TRPC1. Endogenous Ca2+ release is further supported by NAADP, which evokes EL Ca2+ release through TPC1. SERCA and PMCA contribute to maintain resting Ca2+ levels and clear cytosolic Ca2+ after a pro-angiogenic Ca2+ signal. Pro-angiogenic Ca2+ signals may also be delivered by TRP Vanilloid 1 (TRPV1) and TRPV4. See Section 5.2. (SDF-1α) and Section 5.3. (VEGF and NAADP) for further details.
Figure 5. The Ca2+ toolkit in endothelial colony forming cells (ECFC). Growth factors, such as VEGF, and chemokines, such as stromal derived factor-1α (SDF-1α), bind to specific RTK and Gq/11PCR, thereby activating multiple PLC isoforms, which in turn cleave PIP2 into InsP3 and DAG. InsP3 triggers ER-dependent Ca2+ release through InsP3R, while DAG gates TRPC3 exclusively in umbilical cord blood-derived ECFCs. SOCE is the major Ca2+ entry pathway also in ECFC, in which it is mediated by the dynamic interplay among STIM1, Orai1 and TRPC1. Endogenous Ca2+ release is further supported by NAADP, which evokes EL Ca2+ release through TPC1. SERCA and PMCA contribute to maintain resting Ca2+ levels and clear cytosolic Ca2+ after a pro-angiogenic Ca2+ signal. Pro-angiogenic Ca2+ signals may also be delivered by TRP Vanilloid 1 (TRPV1) and TRPV4. See Section 5.2. (SDF-1α) and Section 5.3. (VEGF and NAADP) for further details.
Ijms 20 03962 g005

Share and Cite

MDPI and ACS Style

Moccia, F.; Negri, S.; Shekha, M.; Faris, P.; Guerra, G. Endothelial Ca2+ Signaling, Angiogenesis and Vasculogenesis: Just What It Takes to Make a Blood Vessel. Int. J. Mol. Sci. 2019, 20, 3962. https://doi.org/10.3390/ijms20163962

AMA Style

Moccia F, Negri S, Shekha M, Faris P, Guerra G. Endothelial Ca2+ Signaling, Angiogenesis and Vasculogenesis: Just What It Takes to Make a Blood Vessel. International Journal of Molecular Sciences. 2019; 20(16):3962. https://doi.org/10.3390/ijms20163962

Chicago/Turabian Style

Moccia, Francesco, Sharon Negri, Mudhir Shekha, Pawan Faris, and Germano Guerra. 2019. "Endothelial Ca2+ Signaling, Angiogenesis and Vasculogenesis: Just What It Takes to Make a Blood Vessel" International Journal of Molecular Sciences 20, no. 16: 3962. https://doi.org/10.3390/ijms20163962

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

Article Metrics

Back to TopTop