Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2102 word(s) 2102 2021-12-30 02:23:28 |
2 format correct Meta information modification 2102 2022-01-04 02:58:04 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Cairrao, E. Cyclic Nucleotides Regulate Vascular Tone. Encyclopedia. Available online: https://encyclopedia.pub/entry/17711 (accessed on 12 May 2024).
Cairrao E. Cyclic Nucleotides Regulate Vascular Tone. Encyclopedia. Available at: https://encyclopedia.pub/entry/17711. Accessed May 12, 2024.
Cairrao, Elisa. "Cyclic Nucleotides Regulate Vascular Tone" Encyclopedia, https://encyclopedia.pub/entry/17711 (accessed May 12, 2024).
Cairrao, E. (2022, January 03). Cyclic Nucleotides Regulate Vascular Tone. In Encyclopedia. https://encyclopedia.pub/entry/17711
Cairrao, Elisa. "Cyclic Nucleotides Regulate Vascular Tone." Encyclopedia. Web. 03 January, 2022.
Cyclic Nucleotides Regulate Vascular Tone
Edit

Several mediators and drugs regulate blood flow and blood pressure. The cyclic nucleotides cAMP and cGMP are the key messengers mediating vasodilation under physiological conditions and are therefore involved in the physiological regulation of vascular tone. Their function is due to the existence of several mechanisms that include the reduction of [Ca2+]i and a decrease in the sensitivity of the contractile machinery; both mechanisms may occur together and decrease MLC phosphorylation. On the other hand, these mechanisms are regulated by several cellular effectors, including ion channels that regulate the membrane potential and Ca2+ influx. All these mechanisms promote vessel relaxation by reducing vascular contractility and tone. 

compartmentation cAMP cGMP vascular smooth muscle cells cardiovascular diseases

1. cAMP Effector Proteins

cAMP is generated in response to the activation of a wide range of membrane receptors belonging to the GPCR superfamily. After binding a ligand, a G-protein is activated and promotes the activation of the AC that generates cAMP from ATP. This second intracellular messenger, in turn, regulates an infinity of physiological and pathological processes in different organs, including the CV system, and this is achieved by activating the so-called cAMP effectors. These effectors include cAMP-dependent protein kinase (protein kinase A, PKA), exchange protein directly activated by cAMP (EPAC), cyclic nucleotide-gated (CNG) ion channels, and the recently discovered Popeye domain-containing proteins (POPDC) [1].
The major effects of cAMP in eukaryotic cells result from the activation of PKA [2]. PKA was first discovered by Walsh et al. in 1968 [3]. Its catalytic subunit (C) was the first protein kinase to be crystallized. Since then, many studies have focused on it, and is currently one of the most studied protein kinases. However, and despite intense research, some important aspects such as its activation and inactivation in the complex cellular environment still need to be investigated [4].
Structurally, the PKA holoenzyme is a tetramer formed by two catalytic C subunits, and a dimer regulatory (R) subunit, where cAMP binds [5]. The three isoforms of the C subunit (α, β, γ) have virtually identical kinetic and physicochemical properties, while the four regulatory subunits (RIα, RIIα, RIβ, and RIIβ) exhibit distinct binding affinities for cAMP and are differentially located within cells [5][6][7][8][9]. PKA holoenzymes containing the RI subunits are PKA type I (RIα and RIβ) and are predominantly in the cytoplasm. In contrast, PKA holoenzymes containing type II subunits are called PKA type II (RIIα and RIIβ) and appear to be associated with cellular structures and organelles [5][9]. Initially it was thought that this could be due to the anchoring of A-kinase anchoring proteins (AKAP) with greater affinity for the RII subunits [5]. Still, other studies have demonstrated that there are AKAP with dual specificity for the RI and RII subunits and AKAP that bind only to the RI subunits. The studies show both PKA type I and II may be anchored in subcellular compartments within the cells [9][10][11][12] and, therefore, the specificity of PKA signaling is achieved by binding to these anchoring proteins [13][14][15]. Overall, AKAP direct holoenzymes PKA to different subcellular sites near neighbouring proteins, optimizing signal transduction and allowing events responsive to local cAMP to occur within specific compartments of the cell (i.e., resulting in compartmentalized cAMP signaling, explained in more detail below) [15][16]. Thus, an imbalance in the expression or activity of some of the PKA subunits [17][18] or anchoring by AKAP [19] may lead to the development of cardiovascular diseases (e.g., myocardial infarction).
In the absence of cAMP, PKA is in an inactive state, with two regulatory subunits, R, and two catalytic subunits, C, forming a complex called a tetrameric holoenzyme (R2C2) [20]. In each regulatory subunit R, there are two binding sites (A and B) to which cAMP cooperatively binds during activation [9]. The inactive PKA holoenzyme exposes the available B site for cAMP binding. When occupied, the binding of cAMP to site A is increased, which leads to an intramolecular conformational change: the regulatory subunits dissociate from the R2C2 complex and generate two active C subunits, which phosphorylate their substrates in the cytosol and the nucleus [9][15][21][22][23][24]. Importantly, only the C subunit is in the nucleus [25]. In other words, the binding of the C subunit to the inhibitory sites of the respective R subunit renders the kinase-inactive, while the cAMP allosteric binding to two C-terminal tandem cAMP binding domains (CNB-A and CNB-B) of the R subunits triggers the catalytic activity of the holoenzyme [26][27]. However, it has been suggested that cAMP can also activate PKA without releasing the catalytic subunits C [4][28][29][30] and there are intact and active holoenzymes within the cytoplasm in the presence of cAMP. Other studies suggest that a more significant number of R subunits than C are essential for reducing the release of catalytic C subunits and increasing their recapture [25].
PKA may also play a critical role in the phenotypic modulation of the vascular SMCs, that mainly depend on intracellular ATP concentrations. Hogarth at al demonstrated that low micromolar ATP concentrations induce a negligible effect on DNA synthesis but induce higher serum response factor activity, and showed a higher gene expression of the markers of differentiated smooth muscle cells, such as SM-α-actin and SM22, indicating a contractile SMCs phenotype. On the contrary, high micromolar ATP concentration inhibits serum response factor activity and the markers genes expression and promotes cultured cell growth in a manner dependent on PKA activation. The change into the contractile phenotype by high ATP intracellular can be prevented, and even reversed, by inhibition of PKA activity [31].
Although PKA is the primary protein kinase activated by cAMP, several studies suggest that some of the biological effects of cAMP may, at least in part, be mediated by activation of PKG [32][33][34] which is classically activated by cGMP. The cyclic nucleotide-binding domain of PKG binds cGMP with greater affinity than cAMP. Moreover, intracellular concentrations of cAMP are generally higher than cGMP, which show that cAMP is also an agonist of PKG [35]. In this sense, Eckly-Michel et al. (1997) showed clearly that cAMP mediates PKG activation in vascular smooth muscle [36].
In addition to PKA, another important downstream effector of cAMP [37] is the EPAC family [3][38][39]. The EPAC proteins (1 and 2) act as guanine nucleotide exchange factors (GEFs) for small Ras-like GTPases (Rap1 and Rap2), and for this reason are also designated as cAMP-GEF proteins. The EPAC1 encoded by the RAPGEF3 gene is present in the human vasculature [37]. Although EPAC and PKA can act independently, they often act together in the same biological process in which they mediate synergistic or opposite effects. Discovery of EPAC discovery is relatively recent, but it modulates several different systems, including the CV system, which plays a critical role [37]. Normally, in vascular SMCs, activation of RhoA/Rho kinase (ROCK) signalling phosphorylates MLCP, inhibiting its phosphatase activity. Thus, phosphorylation of MLC by MLCK increases, activated by the Ca2+-CaM complex, and contraction occurs. EPAC induces a vasorelaxation of vascular SMCs through inhibition of RhoA/ROCK signalling. The cAMP-mediated activation of EPAC/Rap1 releases the inhibitory effect of RhoA/ROCK on the MLCP. In turn, this leads to dephosphorylation of MLC and subsequent vasorelaxation of SMCs [37]. However, the role of EPAC in the vasculature is controversial: these proteins may promote or inhibit SMCs proliferation and migration. Moreover, if EPACs induce vasorelaxation of vascular SMCs extracted from large vessels, on the other hand they induce vasoconstriction of micro vascular SMCs. In other words, EPAC proteins may act as mediator of vascular SMCs phenotypic switching or, on the other hand, have a protective role [37]. This disagreement of the effects of EPACs seems to be based on the use of different cellular model systems, as it is known that the susceptibility of vascular SMCs to phenotypic switching is variable according to their origin, the vascular bed from which they were isolated, as well as differences in species, strain, age, or sex of the animal model. In this sense, there is no doubt that their role in the vasculature is crucial, and that anomalous EPAC signaling may be frequently implicated in pathological CV conditions [40]. Therefore, Rap1-independent EPAC signalling may be a useful target for drugs designed to treat atherosclerosis and hypertension.
Additionally, other effectors of cAMP are CNGs, expressed in vascular SMCs [41][42] and EC [42][43][44]. CNGs are nonselective cationic channels that open in response to direct binding of cAMP and cGMP [45], modulating the vascular tone.
Moreover, POPDC also play the role of cAMP effectors. The POPDC family of proteins (1-3) contain an extracellular N-terminal domain, three transmembrane domains, and a cytosolic Popeye domain [46]. This latter domain serves as the binding site for high-affinity cAMP. POPDC is expressed in vascular SMCs colocalized with α-smooth muscle actin and appears to be involved in the mechanisms of vasculogenesis [46][47]. However, its role as a cAMP effector in vascular SMCs remains unclear.

2. cGMP Effector Proteins

cGMP concentration is critical for the maintenance of CV homeostasis in several cell types. Moreover, the classical and significant physiological effect of an increase in the intracellular concentration of cGMP in vascular cells is vasodilation [2]. The cGMP pathway begins with the activation by one ligand (NO, ANP, BNP, and CNP) to guanylyl cyclase, either soluble or particulate, that induces the increase of the intracellular concentration of cGMP, which may activate three classes of cGMP effector proteins: (1) PKG, (2) CNG channels and (3) PDE, whose activity can be regulated by cGMP as previously described [48][49][50].
This protein kinase (also known as cGKs) alters the activity of target proteins, phosphorylating specific radicals of serine and threonine [51][52]. Moreover, PKG is considered the most important target of cGMP in the CV system. In mammals, two PKG genes have been identified, namely PRKG1 AND PRKG2, that encode PKG1 and PKG2, respectively [53][54][55]. The membrane PKG 2 is not expressed in the CV system. On the other hand, cytosolic PKG 1 has two different splice variants (PKG 1α and PKG 1β) that differ only in the first~100 amino acids. However, the one most expressed in vascular SMCs is PKG 1β [53][56][57][58]. In these cells, both PKG 1α and PKG 1β are expressed, while EC expresses only PKG 1β [48]. However, the expression of PKG is dependent of the cell density in culture, low density, or reduced expression of PKG [59]. Moreover, the increase in the expression of PKG contributes to change for the contractile phenotypic in cultured vascular SMCs, and the suppression of PKG expression during cultured growth in vitro may facilitate the modulation to a more synthetic, dedifferentiated phenotype [60]. The crystal structure of PKG 1 was elucidated only in 2016 [61]. PKG1 consists of a homodimer divided into a regulatory (R) and a catalytic (C) region. The regulatory region is composed of four functional domains, leucine zipper, auto-inhibitory (AI) (which engages the catalytic region of the PKG in the committed/inactive state), and two tandemly arranged low and high-affinity cGMP binding domains (CNB-A and CNB-B) [61][62]. Selective interaction of cGMP with the cyclic nucleotide-binding domain binding pockets leads to a 10-fold higher sensitivity for PKG 1α compared to PKG 1β [63]. Two domains constitute the C-terminal catalytic region, the kinase domain (where the Mg2+/ATP binds) and the AGC-kinase C-terminal domain (where substrate/downstream proteins bind and phosphorylation occurs) [62]. This protein can bind or activate several signaling pathways, but the exact mechanisms underlying their interactions with PKG have not yet been elucidated. Briefly, PKG operates in parallel through several mechanisms to reduce vascular contractility: (1) PKG inhibits IP3 stimulated SR Ca2+ release by the IP3R by phosphorylating the IP3 receptor-associated PKG-substrate (IRAG) [64][65]; (2) PKG can inhibit by direct phosphorylation the IP3R [65]; (3) PKG phosphorylates phospholamban which removes the brake on SERCA activity, increasing Ca2+ import into the SR; (4) IP3 concentration due to phospholipase C activation is suppressed by PKG phosphorylation of the regulator of G-protein signalling 2 (RGS2) [66]; (5) PKG prevents the inactivation of MLCP by ROCK [67][68]; (6) PKG phosphorylate Ser-695 on the myosin phosphatase target subunit 1 (MYPT1) of the MLCP, which prevents ROCK from phosphorylating the adjacent Thr-696 that usually leads to enzyme inactivation [69][70]; (7) PKG phosphorylates the small heat shock protein HSP20 which induces vascular relaxation (by arresting the reorganization of the actin cytoskeleton required for vasoconstriction) and inhibits of agonist-induced constriction [71][72]; (8) PKG activates large conductance Ca2+-activated K+ channel (BKCa) [73]; (9) PKG activate Kv [73], and (10) PKG inhibits the voltage-dependent Ca2+ channels (see review [74]).
Concerning the CNG channels, these channels are nonselective cation channels activated by the binding of cGMP or cAMP [45], as previously mentioned for cAMP effector proteins. The CNG channels reveal a higher sensitivity for cGMP than for cAMP. In physiological conditions, CNG channels carry inward Na+ and Ca2+ currents [75]. Even if the divalent cations can permeate the channel, higher concentrations induce these cations to bind to specific sites within the channel pore and block further ion flow. Knowledge of CNG channels is greater regarding the visual and olfactory systems while in other areas is less. Despite that, the CNG channels are used with the patch-clamp recordings to observe cyclic nucleotide changes [76].

References

  1. Amunjela, J.N.; Swan, A.H.; Brand, T. The Role of the Popeye Domain Containing Gene Family in Organ Homeostasis. Cells 2019, 8, 1594.
  2. Morgado, M.; Cairrão, E.; Santos-Silva, A.J.; Verde, I. Cyclic nucleotide-dependent relaxation pathways in vascular smooth muscle. Experientia 2011, 69, 247–266.
  3. Walsh, D.A.; Perkins, J.P.; Krebs, E.G. An adenosine 3’,5’-monophosphate-dependant protein kinase from rabbit skeletal muscle. J. Biol. Chem. 1968, 243, 3763–3765.
  4. Smith, F.D.; Esseltine, J.L.; Nygren, P.J.; Veesler, D.; Byrne, D.P.; Vonderach, M.; Strashnov, I.; Eyers, C.E.; Eyers, P.A.; Langeberg, L.K.; et al. Local protein kinase A action proceeds through intact holoenzymes. Science 2017, 356, 1288–1293.
  5. Colledge, M.; Scott, J.D. AKAPs: From structure to function. Trends Cell Biol. 1999, 9, 216–221.
  6. Dostmann, W.R.; Taylor, S.S.; Genieser, H.G.; Jastorff, B.; Døskeland, S.O.; Ogreid, D. Probing the cyclic nucleotide binding sites of cAMP-dependent protein kinases I and II with analogs of adenosine 3′,5′-cyclic phosphorothioates. J. Biol. Chem. 1990, 265, 10484–10491.
  7. Cummings, D.E.; Brandon, E.P.; Planas, J.V.; Motamed, K.; Idzerda, R.L.; McKnight, G.S. Genetically lean mice result from targeted disruption of the RIIβ subunit of protein kinase A. Nature 1996, 382, 622–626.
  8. Taskén, K.; Sklhegg, B.S.; Austlid, T.K.; Solberg, R.; Knutsen, H.K.; Levy, F.O.; Sandberg, M.; Ørstavik, S.; Larsen, T.; Johansen, A.K.; et al. Structure, function, and regulation of human cAMP-dependent protein kinases. Adv. Second. Messenger Phosphoprot. Res. 1997, 31, 191–204.
  9. Søberg, K.; Jahnsen, T.; Rognes, T.; Skålhegg, B.S.; Laerdahl, J.K. Evolutionary Paths of the cAMP-Dependent Protein Kinase (PKA) Catalytic Subunits. PLoS ONE 2013, 8, e60935.
  10. Wang, L.; Sunahara, R.K.; Krumins, A.; Perkins, G.; Crochiere, M.L.; Mackey, M.; Bell, S.; Ellisman, M.H.; Taylor, S.S. Cloning and mitochondrial localization of full-length D-AKAP2, a protein kinase A anchoring protein. Proc. Natl. Acad. Sci. USA 2001, 98, 3220–3225.
  11. Taskén, K.; Aandahl, E.M. Localized Effects of cAMP Mediated by Distinct Routes of Protein Kinase A. Physiol. Rev. 2004, 84, 137–167.
  12. Jarnæss, E.; Tasken, K. Spatiotemporal control of cAMP signalling processes by anchored signalling complexes. Biochem. Soc. Trans. 2007, 35, 931–937.
  13. Carr, D.W.; Stofko-Hahn, R.E.; Fraser, I.D.; Bishop, S.M.; Acott, T.S.; Brennan, R.G.; Scott, J.D. Interaction of the regulatory subunit (RII) of cAMP-dependent protein kinase with RII-anchoring proteins occurs through an amphipathic helix binding motif. J. Biol. Chem. 1991, 266, 14188–14192.
  14. Scott, J.D.; Stofko, R.E.; McDonald, J.R.; Comer, J.D.; Vitalis, E.A.; Mangili, J.A. Type II regulatory subunit dimerization determines the subcellular localization of the cAMP-dependent protein kinase. J. Biol. Chem. 1990, 265, 21561–21566.
  15. Turnham, R.E.; Scott, J.D. Protein kinase A catalytic subunit isoform PRKACA.; History, function and physiology. Gene 2015, 577, 101–108.
  16. Bauman, A.L.; Soughayer, J.; Nguyen, B.T.; Willoughby, D.; Carnegie, G.K.; Wong, W.; Hoshi, N.; Langeberg, L.K.; Cooper, D.M.; Dessauer, C.; et al. Dynamic Regulation of cAMP Synthesis through Anchored PKA-Adenylyl Cyclase V/VI Complexes. Mol. Cell 2006, 23, 925–931.
  17. Kirschner, L.S.; Carney, J.A.; Pack, S.; Taymans, S.E.; Giatzakis, C.; Cho, Y.S.; Cho-Chung, Y.S.; Stratakis, C.A. Mutations of the gene encoding the protein kinase A type I-α regulatory subunit in patients with the Carney complex. Nat. Genet. 2000, 26, 89–92.
  18. Beuschlein, F.; Fassnacht, M.; Assié, G.; Calebiro, D.; Stratakis, C.A.; Osswald, A.; Ronchi, C.; Wieland, T.; Sbiera, S.; Faucz, F.R.; et al. Constitutive Activation of PKA Catalytic Subunit in Adrenal Cushing’s Syndrome. N. Engl. J. Med. 2014, 370, 1019–1028.
  19. Gold, M.G.; Gonen, T.; Scott, J.D. Local cAMP signaling in disease at a glance. J. Cell Sci. 2013, 126, 4537–4543.
  20. Krebs, E.G.; Beavo, J.A. Phosphorylation-Dephosphorylation of Enzymes. Annu. Rev. Biochem. 1979, 48, 923–959.
  21. Bjorn, S.S.; Skalhegg, B.S.; Tasken, K. Specificity in the cAMP/PKA signaling pathway. differential expression, regulation, and subcellular localization of subunits of PKA. Front. Biosci. 2000, 5, d678–d693.
  22. Kim, C.; Xuong, N.-H.; Taylor, S.S. Crystal Structure of a Complex Between the Catalytic and Regulatory (RIα) Subunits of PKA. Science 2005, 307, 690–696.
  23. Sjoberg, T.J.; Kornev, A.P.; Taylor, S.S. Dissecting the cAMP-inducible allosteric switch in protein kinase A RIα. Protein Sci. 2010, 19, 1213–1221.
  24. Anand, G.S.; Krishnamurthy, S.; Bishnoi, T.; Kornev, A.; Taylor, S.S.; Johnson, D.A. Cyclic AMP- and (Rp)-cAMPS-induced Conformational Changes in a Complex of the Catalytic and Regulatory (RIα) Subunits of Cyclic AMP-dependent Protein Kinase. Mol. Cell. Proteom. 2010, 9, 2225–2237.
  25. Wong, W.; Scott, J.D. AKAP signalling complexes: Focal points in space and time. Nat. Rev. Mol. Cell Biol. 2004, 5, 959–970.
  26. Taylor, S.S.; Buechler, J.A.; Yonemoto, W. cAMP-DEPENDENT PROTEIN KINASE: FRAMEWORK FOR A DIVERSE FAMILY OF REGULATORY ENZYMES. Annu. Rev. Biochem. 1990, 59, 971–1005.
  27. Herberg, F.W.; Taylor, S.S.; Dostmann, W.R. Active Site Mutations Define the Pathway for the Cooperative Activation of cAMP-Dependent Protein Kinase. Biochemistry 1996, 35, 2934–2942.
  28. Yang, S.; Fletcher, W.H.; Johnson, D.A. Regulation of cAMP-dependent protein kinase: Enzyme activation without dissociation. Biochemistry 1995, 34, 6267–6271.
  29. Bittinger, M.A.; McWhinnie, E.; Meltzer, J.; Iourgenko, V.; Latario, B.; Liu, X.; Chen, C.H.; Song, C.; Garza, D.; Labow, M. Activation of cAMP Response Element-Mediated Gene Expression by Regulated Nuclear Transport of TORC Proteins. Curr. Biol. 2004, 14, 2156–2161.
  30. Smith, F.D.; Reichow, S.; Esseltine, J.L.; Shi, D.; Langeberg, L.K.; Scott, J.D.; Gonen, T. Intrinsic disorder within an AKAP-protein kinase A complex guides local substrate phosphorylation. eLife 2013, 2, e01319.
  31. Hogarth, D.K.; Sandbo, N.; Taurin, S.; Kolenko, V.; Miano, J.M.; Dulin, N.O. Dual role of PKA in phenotypic modulation of vascular smooth muscle cells by extracellular ATP. Am. J. Physiol. Physiol. 2004, 287, C449–C456.
  32. Lincoln, T.M.; Cornwell, T.L.; Taylor, A.E. cGMP-dependent protein kinase mediates the reduction of Ca2+ by cAMP in vascular smooth muscle cells. Am. J. Physiol. Physiol. 1990, 258, C399–C407.
  33. White, R.E.; Kryman, J.P.; El-Mowafy, A.M.; Han, G.; Carrier, G.O. cAMP-Dependent Vasodilators Cross-Activate the cGMP-Dependent Protein Kinase to Stimulate BK Ca Channel Activity in Coronary Artery Smooth Muscle Cells. Circ. Res. 2000, 86, 897–905.
  34. Wu, Y.-J.; Bond, M.; Sala-Newby, G.B.; Newby, A.C. Altered S-Phase Kinase-Associated Protein-2 Levels Are a Major Mediator of Cyclic Nucleotide–Induced Inhibition of Vascular Smooth Muscle Cell Proliferation. Circ. Res. 2006, 98, 1141–1150.
  35. VanSchouwen, B.; Selvaratnam, R.; Giri, R.; Lorenz, R.; Herberg, F.W.; Kim, C.; Melacini, G. Mechanism of cAMP Partial Agonism in Protein Kinase G (PKG). J. Biol. Chem. 2015, 290, 28631–28641.
  36. Eckly-Michel, A.; Martin, V.; Lugnier, C. Involvement of cyclic nucleotide-dependent protein kinases in cyclic AMP-mediated vasorelaxation. Br. J. Pharmacol. 1997, 122, 158–164.
  37. Wehbe, N.; Nasser, S.A.; Al-Dhaheri, Y.; Iratni, R.; Bitto, A.; El-Yazbi, A.F.; Badran, A.; Kobeissy, F.; Baydoun, E.; Eid, A.H. EPAC in Vascular Smooth Muscle Cells. Int. J. Mol. Sci. 2020, 21, 5160.
  38. Kawasaki, H.; Springett, G.M.; Mochizuki, N.; Toki, S.; Nakaya, M.; Matsuda, M.; Housman, D.E.; Graybiel, A.M. A Family of cAMP-Binding Proteins That Directly Activate Rap1. Science 1998, 282, 2275–2279.
  39. De Rooij, J.; Zwartkruis, F.J.T.; Verheijen, M.H.G.; Cool, R.H.; Nijman, S.M.B.; Wittinghofer, A.; Bos, J.L. Epac is a Rap1 guanine-nucleotide-exchange factor directly activated by cyclic AMP. Nature 1998, 396, 474–477.
  40. Lezoualc’H, F.; Fazal, L.; Laudette, M.; Conte, C. Cyclic AMP Sensor EPAC Proteins and Their Role in Cardiovascular Function and Disease. Circ. Res. 2016, 118, 881–897.
  41. Leung, Y.K.; Du, J.; Huang, Y.; Yao, X. Cyclic Nucleotide-Gated Channels Contribute to Thromboxane A2-Induced Contraction of Rat Small Mesenteric Arteries. PLoS ONE 2010, 5, e11098.
  42. Yao, X.; Leung, P.-S.; Kwan, H.-Y.; Wong, T.-P.; Fong, M.-W. Rod-type cyclic nucleotide-gated cation channel is expressed in vascular endothelium and vascular smooth muscle cells. Cardiovasc. Res. 1999, 41, 282–290.
  43. Cheng, K.-T.; Chan, F.L.; Huang, Y.; Chan, W.-Y.; Yao, X. Expression of olfactory-type cyclic nucleotide-gated channel (CNGA2) in vascular tissues. Histochem. Cell Biol. 2003, 120, 475–481.
  44. Shen, B.; Cheng, K.-T.; Leung, Y.-K.; Kwok, Y.-C.; Kwan, H.-Y.; Wong, C.-O.; Chen, Z.-Y.; Huang, Y.; Yao, X. Epinephrine-induced Ca2+ influx in vascular endothelial cells is mediated by CNGA2 channels. J. Mol. Cell. Cardiol. 2008, 45, 437–445.
  45. Biel, M.; Michalakis, S. Cyclic nucleotide-gated channels. In Handbook of Experimental Pharmacology; Springer: Berlin/Heidelberg, Germany, 2009; pp. 111–136.
  46. E.Reeseabc, D.; Zavaljevskib, M.; Streiff, N.L.; Baderabc, D. bves:A Novel Gene Expressed during Coronary Blood Vessel Development. Dev. Biol. 1999, 209, 159–171.
  47. Wada, A.M.; Reese, D.E.; Bader, D.M. Bves: Prototype of a new class of cell adhesion molecules expressed during coronary artery development. Development 2001, 128, 2085–2093.
  48. Bork, N.I.; Nikolaev, V.O. cGMP Signaling in the Cardiovascular System—The Role of Compartmentation and Its Live Cell Imaging. Int. J. Mol. Sci. 2018, 19, 801.
  49. Lugnier, C. Cyclic nucleotide phosphodiesterase (PDE) superfamily: A new target for the development of specific therapeutic agents. Pharmacol. Ther. 2006, 109, 366–398.
  50. Ahmed, W.S.; Geethakumari, A.M.; Biswas, K.H. Phosphodiesterase 5 (PDE5): Structure-function regulation and therapeutic applications of inhibitors. Biomed. Pharmacother. 2020, 134, 111128.
  51. Walsh, D.A.; Van Patten, S.M. Multiple pathway signal transduction by the cAMP-dependent protein kinase. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 1994, 8, 1227–1236.
  52. Taylor, S.S.; Kim, C.; Vigil, D.; Haste, N.M.; Yang, J.; Wu, J.; Anand, G.S. Dynamics of signaling by PKA. Biochim. Biophys. Acta (BBA)-Proteins Proteom. 2005, 1754, 25–37.
  53. Hofmann, F.; Feil, R.; Kleppisch, T.; Schlossmann, J. Function of cGMP-dependent protein kinases as revealed by gene deletion. Physiol. Rev. 2006, 86, 1–23.
  54. Lincoln, T.M.; Dey, N.; Sellak, H. Invited review: cGMP-dependent protein kinase signaling mechanisms in smooth muscle: From the regulation of tone to gene expression. J. Appl. Physiol. 2001, 91, 1421–1430.
  55. Lucas, K.A.; Pitari, G.M.; Kazerounian, S.; Ruiz-Stewart, I.; Park, J.; Schulz, S.; Chepenik, K.P.; Waldman, S.A. Guanylyl cyclases and signaling by cyclic GMP. Pharmacol. Rev. 2000, 52, 375–414.
  56. Geiselhöringer, A.; Werner, M.; Sigl, K.; Smital, P.; Wörner, R.; Acheo, L.; Stieber, J.; Weinmeister, P.; Feil, R.; Feil, S.; et al. IRAG is essential for relaxation of receptor-triggered smooth muscle contraction by cGMP kinase. EMBO J. 2004, 23, 4222–4231.
  57. Hofmann, F. The Biology of Cyclic GMP-dependent Protein Kinases. J. Biol. Chem. 2005, 280, 1–4.
  58. Feil, R.; Hofmann, F. Alternative splicing of cGMP-dependent protein kinase I and nitrate tolerance. Circ. Res. 2003, 93, e143.
  59. Cornwell, T.L.; Soff, G.A.; Traynor, A.E.; Lincoln, T.M. Regulation of the Expression of Cyclic GMP-Dependent Protein Kinase by Cell Density in Vascular Smooth Muscle Cells. J. Vasc. Res. 1994, 31, 330–337.
  60. Boerth, N.J.; Dey, N.B.; Cornwell, T.L.; Lincoln, T.M. Cyclic GMP-Dependent Protein Kinase Regulates Vascular Smooth Muscle Cell Phenotype. J. Vasc. Res. 1997, 34, 245–259.
  61. Kim, J.J.; Lorenz, R.; Arold, S.T.; Reger, A.S.; Sankaran, B.; Casteel, D.E.; Herberg, F.W.; Kim, C. Crystal Structure of PKG I:cGMP Complex Reveals a cGMP-Mediated Dimeric Interface that Facilitates cGMP-Induced Activation. Structure 2016, 24, 710–720.
  62. Maryam, A.; Khalid, R.R.; Vedithi, S.C.; Ece, A.; Çınaroğlu, S.S.; Siddiqi, A.R.; Blundell, T.L. Exploring the structural basis of conformational heterogeneity and autoinhibition of human cGMP-specific protein kinase Iα through computational modelling and molecular dynamics simulations. Comput. Struct. Biotechnol. J. 2020, 18, 1625–1638.
  63. Huang, G.Y.; Kim, J.J.; Reger, A.S.; Lorenz, R.; Moon, E.-W.; Zhao, C.; Casteel, D.E.; Bertinetti, D.; VanSchouwen, B.; Selvaratnam, R.; et al. Structural Basis for Cyclic-Nucleotide Selectivity and cGMP-Selective Activation of PKG I. Structure 2013, 22, 116–124.
  64. Schlossmann, J.; Ammendola, A.; Ashman, K.; Zong, X.; Huber, A.; Neubauer, G.; Wang, G.-X.; Allescher, H.-D.; Korth, M.; Wilm, M.; et al. Regulation of intracellular calcium by a signalling complex of IRAG, IP3 receptor and cGMP kinase Iβ. Nature 2000, 404, 197–201.
  65. Komalavilas, P.; Lincoln, T.M. Phosphorylation of the Inositol 1,4,5-Trisphosphate Receptor. J. Biol. Chem. 1996, 271, 21933–21938.
  66. Osei-Owusu, P.; Sun, X.; Drenan, R.M.; Steinberg, T.H.; Blumer, K.J. Regulation of RGS2 and Second Messenger Signaling in Vascular Smooth Muscle Cells by cGMP-dependent Protein Kinase. J. Biol. Chem. 2007, 282, 31656–31665.
  67. Sawada, N.; Itoh, H.; Yamashita, J.; Doi, K.; Inoue, M.; Masatsugu, K.; Fukunaga, Y.; Sakaguchi, S.; Sone, M.; Yamahara, K.-I.; et al. cGMP-Dependent Protein Kinase Phosphorylates and Inactivates RhoA. Biochem. Biophys. Res. Commun. 2001, 280, 798–805.
  68. Sauzeau, V.; Le Jeune, H.; Cario-Toumaniantz, C.; Smolenski, A.; Lohmann, S.M.; Bertoglio, J.; Chardin, P.; Pacaud, P.; Loirand, G. Cyclic GMP-dependent Protein Kinase Signaling Pathway Inhibits RhoA-induced Ca2+ Sensitization of Contraction in Vascular Smooth Muscle. J. Biol. Chem. 2000, 275, 21722–21729.
  69. Somlyo, A.V. Cyclic GMP Regulation of Myosin Phosphatase. Circ. Res. 2007, 101, 645–647.
  70. Wooldridge, A.A.; MacDonald, J.A.; Erdodi, F.; Ma, C.; Borman, M.A.; Hartshorne, D.J.; Haystead, T.A.J. Smooth Muscle Phosphatase Is Regulated in Vivo by Exclusion of Phosphorylation of Threonine 696 of MYPT1 by Phosphorylation of Serine 695 in Response to Cyclic Nucleotides. J. Biol. Chem. 2004, 279, 34496–34504.
  71. Beall, A.; Bagwell, D.; Woodrum, D.; Stoming, T.A.; Kato, K.; Suzuki, A.; Rasmussen, H.; Brophy, C.M. The Small Heat Shock-related Protein, HSP20, Is Phosphorylated on Serine 16 during Cyclic Nucleotide-dependent Relaxation. J. Biol. Chem. 1999, 274, 11344–11351.
  72. Beall, A.C.; Kato, K.; Goldenring, J.R.; Rasmussen, H.; Brophy, C.M. Cyclic Nucleotide-dependent Vasorelaxation Is Associated with the Phosphorylation of a Small Heat Shock-related Protein. J. Biol. Chem. 1997, 272, 11283–11287.
  73. Cairrão, E.; Santos-Silva, A.J.; Verde, I. PKG is involved in testosterone-induced vasorelaxation of human umbilical artery. Eur. J. Pharmacol. 2010, 640, 94–101.
  74. Manoury, B.; Idres, S.; Leblais, V.; Fischmeister, R. Ion channels as effectors of cyclic nucleotide pathways: Functional relevance for arterial tone regulation. Pharmacol. Ther. 2020, 209, 107499.
  75. Podda, M.V.; Grassi, C. New perspectives in cyclic nucleotide-mediated functions in the CNS: The emerging role of cyclic nucleotide-gated (CNG) channels. Pflügers Arch.-Eur. J. Physiol. 2013, 466, 1241–1257.
  76. Feil, R.; Lehners, M.; Stehle, D.; Feil, S. Visualising and understanding cGMP signals in the cardiovascular system. Br. J. Pharmacol. 2021, 1–19.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 442
Revisions: 2 times (View History)
Update Date: 04 Jan 2022
1000/1000