24 research outputs found

    Polyunsaturated fatty acids in atrial fibrillation: Looking for the proper candidates

    Get PDF
    This Document is Protected by copyright and was first published by Frontiers. All rights reserved. it is reproduced with permissionAtrial fibrillation (AF) is the most common sustained arrhythmia encountered in clinical practice with growing prevalence in developed countries. Several medical and interventional therapies, such as atrial specific drugs and pulmonary vein isolation, have demonstrated prevention of recurrences. However, their suboptimal long-term success and significant rate of secondary effects have led to intensive research in the last decade focused on novel alternative and supplemental therapies. One such candidate is polyunsaturated fatty acids (PUFAs). Because of their biological properties, safety, simplicity, and relatively cheap cost, there is a special clinical interest in omega-3 PUFAs as a possible antiarrhythmic agent. Obtained from diets rich in fish, they represent one of the current supplemental therapies. At the cellular level, an increasing body of evidence has shown that n-3 PUFAs exert a variety of effects on cardiac ion channels, membrane dynamic properties, inflammatory cascade, and other targets related to AF prevention. In this article, we review the current basic and clinical evidence pertinent to n-3 PUFAs in AF treatment and prevention.We also discuss controversial outcomes among clinical studies and propose specific subsets of AF patients who will benefit most from n-3 PUFAsNHLBI Grant K99-HL105574 to SFN and the Alfonso Martín Escudero Foundation Grant to DF

    A null mutation of the neuronal sodium channel NaV1.6 disrupts action potential propagation and excitation‐contraction coupling in the mouse heart

    Full text link
    Evidence supports the expression of brain‐type sodium channels in the heart. Their functional role, however, remains controversial. We used global NaV1.6‐null mice to test the hypothesis that NaV1.6 contributes to the maintenance of propagation in the myocardium and to excitation‐contraction (EC) coupling. We demonstrated expression of transcripts encoding full‐length NaV1.6 in isolated ventricular myocytes and confirmed the striated pattern of NaV1.6 fluorescence in myocytes. On the ECG, the PR and QRS intervals were prolonged in the null mice, and the Ca2+ transients were longer in the null cells. Under patch clamping, at holding potential (HP) = –120 mV, the peak INa was similar in both phenotypes. However, at HP = –70 mV, the peak INa was smaller in the nulls. In optical mapping, at 4 mM [K+]o, 17 null hearts showed slight (7%) reduction of ventricular conduction velocity (CV) compared to 16 wild‐type hearts. At 12 mM [K+]o, CV was 25% slower in a subset of 9 null vs. 9 wild‐type hearts. These results highlight the importance of neuronal sodium channels in the heart, whereby NaV1.6 participates in EC coupling, and represents an intrinsic depolarizing reserve that contributes to excitation.—Noujaim, S. F., Kaur, K., Milstein, M., Jones, J. M., Furspan, P., Jiang, D., Auerbach, D. S., Herron, T., Meisler, M. H., Jalife, J. A null mutation of the neuronal sodium channel NaV1.6 disrupts action potential propagation and excitation‐contraction coupling in the mouse heart. FASEB J. 26, 63–72 (2012). www.fasebj.orgPeer Reviewedhttps://deepblue.lib.umich.edu/bitstream/2027.42/154524/1/fsb2fj10179770.pd

    Human influenza A virus causes myocardial and cardiac-specific conduction system infections associated with early inflammation and premature death.

    Get PDF
    Human influenza A virus (hIAV) infection is associated with important cardiovascular complications, although cardiac infection pathophysiology is poorly understood. We aimed to study the ability of hIAV of different pathogenicity to infect the mouse heart, and establish the relationship between the infective capacity and the associated in vivo, cellular and molecular alterations. We evaluated lung and heart viral titres in mice infected with either one of several hIAV strains inoculated intranasally. 3D reconstructions of infected cardiac tissue were used to identify viral proteins inside mouse cardiomyocytes, Purkinje cells, and cardiac vessels. Viral replication was measured in mouse cultured cardiomyocytes. Human-induced pluripotent stem cell-derived cardiomyocytes (hiPSC-CMs) were used to confirm infection and study underlying molecular alterations associated with the in vivo electrophysiological phenotype. Pathogenic and attenuated hIAV strains infected and replicated in cardiomyocytes, Purkinje cells, and hiPSC-CMs. The infection was also present in cardiac endothelial cells. Remarkably, lung viral titres did not statistically correlate with viral titres in the mouse heart. The highly pathogenic human recombinant virus PAmut showed faster replication, higher level of inflammatory cytokines in cardiac tissue and higher viral titres in cardiac HL-1 mouse cells and hiPSC-CMs compared with PB2mut-attenuated virus. Correspondingly, cardiac conduction alterations were especially pronounced in PAmut-infected mice, associated with high mortality rates, compared with PB2mut-infected animals. Consistently, connexin43 and NaV1.5 expression decreased acutely in hiPSC-CMs infected with PAmut virus. YEM1L protease also decreased more rapidly and to lower levels in PAmut-infected hiPSC-CMs compared with PB2mut-infected cells, consistent with mitochondrial dysfunction. Human IAV infection did not increase myocardial fibrosis at 4-day post-infection, although PAmut-infected mice showed an early increase in mRNAs expression of lysyl oxidase. Human IAV can infect the heart and cardiac-specific conduction system, which may contribute to cardiac complications and premature death.JV is a PhD fellow of the La Caixa Foundation International Fellowship Programme (La Caixa/CNB). This work was supported by the European Molecular Biology Organizat ion (STF-7649 to AF), the Spanish Ministry of Science, Innovation and Universities (MCIU), (BFU2011-26175 and BFU2014-57797-R to AN), and the network Ciber de Enfermedades Respiratorias (CIBERES) including the Improvement and Mobilit y Programme. The CNIC is a Severo Ochoa Center of Excellence (SEV-2015-0505). CNIC is supported by MCIU and the Pro CNIC Foundation. This study was supported by grants from Fondo Europeo de Desarrollo Regional (CB16/11/00458), grants SAF2015-65607-R and SAF2016-80324-R from MCIU (A.H. and D.F-R.) and fellowship SVP-2014-068595 to J.A.N-A. This study was supported by Frankel Cardiovascular Centre, Michigan Medicine (Grant 332475). JJ is supported in part by the National Heart, Lung, and Blood Institute (R01 Grant HL122352). S.F.N is supported in part by the National Heart, Lung, and Blood Institute grants R21HL138064 and R01HL129136.S

    The small molecule GAT1508 activates brain-specific GIRK1/2 channel heteromers and facilitates conditioned fear extinction in rodents

    Get PDF
    G-protein-gated inwardly-rectifying K+ (GIRK) channels are targets of Gi/o-protein-signaling systems that inhibit cell excitability. GIRK channels exist as homotetramers (GIRK2 and GIRK4) or heterotetramers with nonfunctional homomeric subunits (GIRK1 and GIRK3). Although they have been implicated in multiple conditions, the lack of selective GIRK drugs that discriminate among the different GIRK channel subtypes has hampered investigations into their precise physiological relevance and therapeutic potential. Here, we report on a highly-specific, potent, and efficacious activator of brain GIRK1/2 channels. Using a chemical screen and electrophysiological assays, we found that this activator, the bromothiophene-substituted small molecule GAT1508, is specific for brain-expressed GIRK1/2 channels rather than for cardiac GIRK1/4 channels. Computational models predicted a GAT1508-binding site validated by experimental mutagenesis experiments, providing insights into how urea-based compounds engage distant GIRK1 residues required for channel activation. Furthermore, we provide computational and experimental evidence that GAT1508 is an allosteric modulator of channel-phosphatidylinositol 4,5-bisphosphate interactions. Through brain-slice electrophysiology, we show that subthreshold GAT1508 concentrations directly stimulate GIRK currents in the basolateral amygdala (BLA) and potentiate baclofen-induced currents. Of note, GAT1508 effectively extinguished conditioned fear in rodents and lacked cardiac and behavioral side effects, suggesting its potential for use in pharmacotherapy for post-traumatic stress disorder. In summary, our findings indicate that the small molecule GAT1508 has high specificity for brain GIRK1/2 channel subunits, directly or allosterically activates GIRK1/2 channels in the BLA, and facilitates fear extinction in a rodent model

    Nerves projecting from the intrinsic cardiac ganglia of the pulmonary veins modulate sinoatrial node pacemaker function

    Full text link
    Rationale: Autonomic nerves from sinoatrial node (SAN) ganglia are known to regulate SAN function. However, it is unclear whether remote pulmonary vein ganglia (PVG) also modulate SAN pacemaker rhythm. Objective: To investigate whether in the mouse heart PVG modulate SAN function. Methods and Results: In hearts from 45 C57BL and 7 Connexin40+/GFP mice, we used tyrosine-hydroxylase (TH) and choline-acetyltransferase (ChAT) immunofluorescence labeling to characterize adrenergic and cholinergic elements, repectively, within the PVG and SAN. PVG project postganglionic nerves to the SAN. TH and ChAT stained nerves, enter the SAN as an extensive, dense mesh-like neural network. Neurons in PVG are biphenotypic, containing ChAT and TH positive neurons. In Langendorff-perfused hearts, we compared effects of electrical stimulation of PVG, posterior (PRCVG) and anterior right vena cava ganglia (ARCVG) using 200-2000 ms trains of pulses (300μs, 0.2-0.6mA, 200Hz). Sympathetic and/or parasympathetic blockade was achieved using 0.5μM propranolol and 1μM atropine, respectively. Epicardial optical mapping of SAN activation was performed before, during and after ganglion stimulation. PVG stimulation increased the P-P interval by 36±9%; PRCVG stimulation increased the P-P interval by 42±11%. ARCVG stimulation produced no change. Propranolol perfusion increased the PVG stimulation effect to 43±13%. Atropine caused a 5±6% decrease. In optical mapping experiments of whole hearts and isolated atrial preparations, PVG stimulation shifted the origin of SAN discharges to varying locations. Conclusions: PVG contain cholinergic, adrenergic and biphenotipic neurons whose axons project across the right atrium to richly innervate the SAN region and contribute significantly to regulation of SAN function.Zarzoso Muñoz, M.; Rysevaite, K.; Milstein, ML.; Calvo Saiz, CJ.; Kean, AC.; Atienza Fernández, F.; Pauza, DH.... (2013). Nerves projecting from the intrinsic cardiac ganglia of the pulmonary veins modulate sinoatrial node pacemaker function. Cardiovascular Research. 566-575. doi:10.1093/cvr/cvt081S566575Johnson, T. A., Gray, A. L., Lauenstein, J.-M., Newton, S. S., & Massari, V. J. (2004). Parasympathetic control of the heart. I. An interventriculo-septal ganglion is the major source of the vagal intracardiac innervation of the ventricles. Journal of Applied Physiology, 96(6), 2265-2272. doi:10.1152/japplphysiol.00620.2003Rysevaite, K., Saburkina, I., Pauziene, N., Noujaim, S. F., Jalife, J., & Pauza, D. H. (2011). Morphologic pattern of the intrinsic ganglionated nerve plexus in mouse heart. Heart Rhythm, 8(3), 448-454. doi:10.1016/j.hrthm.2010.11.019Yuan, B.-X., Ardell, J. L., Hopkins, D. A., & Armour, J. A. (1993). Differential cardiac responses induced by nicotine sensitive canine atrial and ventricular neurones. Cardiovascular Research, 27(5), 760-769. doi:10.1093/cvr/27.5.760Rysevaite, K., Saburkina, I., Pauziene, N., Vaitkevicius, R., Noujaim, S. F., Jalife, J., & Pauza, D. H. (2011). Immunohistochemical characterization of the intrinsic cardiac neural plexus in whole-mount mouse heart preparations. Heart Rhythm, 8(5), 731-738. doi:10.1016/j.hrthm.2011.01.013Pauza, D. H., Pauziene, N., Pakeltyte, G., & Stropus, R. (2002). Comparative quantitative study of the intrinsic cardiac ganglia and neurons in the rat, guinea pig, dog and human as revealed by histochemical staining for acetylcholinesterase. Annals of Anatomy - Anatomischer Anzeiger, 184(2), 125-136. doi:10.1016/s0940-9602(02)80005-xPauza, D. H., Skripka, V., & Pauziene, N. (2002). Morphology of the Intrinsic Cardiac Nervous System in the Dog: A Whole-Mount Study Employing Histochemical Staining with Acetylcholinesterase. Cells Tissues Organs, 172(4), 297-320. doi:10.1159/000067198Arora, R. C., Waldmann, M., Hopkins, D. A., & Armour, J. A. (2003). Porcine intrinsic cardiac ganglia. The Anatomical Record, 271A(1), 249-258. doi:10.1002/ar.a.10030Gatti, P. J., Johnson, T. A., & John Massari, V. (1996). Can neurons in the nucleus ambiguus selectively regulate cardiac rate and atrio-ventricular conduction? Journal of the Autonomic Nervous System, 57(1-2), 123-127. doi:10.1016/0165-1838(95)00104-2Zhuang, S., Zhang, Y., Mowrey, K. A., Li, J., Tabata, T., Wallick, D. W., … Mazgalev, T. N. (2002). Ventricular Rate Control by Selective Vagal Stimulation Is Superior to Rhythm Regularization by Atrioventricular Nodal Ablation and Pacing During Atrial Fibrillation. Circulation, 106(14), 1853-1858. doi:10.1161/01.cir.0000031802.58532.04CHEN, J., WASMUND, S. L., & HAMDAN, M. H. (2006). Back to the Future: The Role of the Autonomic Nervous System in Atrial Fibrillation. Pacing and Clinical Electrophysiology, 29(4), 413-421. doi:10.1111/j.1540-8159.2006.00362.xArmour, J. A. (2008). Potential clinical relevance of the ‘little brain’ on the mammalian heart. Experimental Physiology, 93(2), 165-176. doi:10.1113/expphysiol.2007.041178LAZZARA, R., SCHERLAG, B. J., ROBINSON, M. J., & SAMET, P. (1973). Selective In Situ Parasympathetic Control of the Canine Sinoatrial and Atrioventricular Nodes. Circulation Research, 32(3), 393-401. doi:10.1161/01.res.32.3.393Gray, A. L., Johnson, T. A., Ardell, J. L., & Massari, V. J. (2004). Parasympathetic control of the heart. II. A novel interganglionic intrinsic cardiac circuit mediates neural control of heart rate. Journal of Applied Physiology, 96(6), 2273-2278. doi:10.1152/japplphysiol.00616.2003Pappone, C., Santinelli, V., Manguso, F., Vicedomini, G., Gugliotta, F., Augello, G., … Alfieri, O. (2004). Pulmonary Vein Denervation Enhances Long-Term Benefit After Circumferential Ablation for Paroxysmal Atrial Fibrillation. Circulation, 109(3), 327-334. doi:10.1161/01.cir.0000112641.16340.c7MIQUEROL, L., MEYSEN, S., MANGONI, M., BOIS, P., VANRIJEN, H., ABRAN, P., … GROS, D. (2004). Architectural and functional asymmetry of the His–Purkinje system of the murine heart. Cardiovascular Research, 63(1), 77-86. doi:10.1016/j.cardiores.2004.03.007Jalife, J., Slenter, V. A., Salata, J. J., & Michaels, D. C. (1983). Dynamic vagal control of pacemaker activity in the mammalian sinoatrial node. Circulation Research, 52(6), 642-656. doi:10.1161/01.res.52.6.642Fedorov, V. V., Hucker, W. J., Dobrzynski, H., Rosenshtraukh, L. V., & Efimov, I. R. (2006). Postganglionic nerve stimulation induces temporal inhibition of excitability in rabbit sinoatrial node. American Journal of Physiology-Heart and Circulatory Physiology, 291(2), H612-H623. doi:10.1152/ajpheart.00022.2006Saburkina, I., & Pauza, D. H. (2006). Location and variability of epicardiac ganglia in human fetuses. Anatomy and Embryology, 211(6), 585-594. doi:10.1007/s00429-006-0110-4Slavíková, J., Kuncová, J., Reischig, J., & Dvořáková, M. (2003). Neurochemical Research, 28(3/4), 593-598. doi:10.1023/a:1022837810357Tan, A. Y., Li, H., Wachsmann-Hogiu, S., Chen, L. S., Chen, P.-S., & Fishbein, M. C. (2006). Autonomic Innervation and Segmental Muscular Disconnections at the Human Pulmonary Vein-Atrial Junction. Journal of the American College of Cardiology, 48(1), 132-143. doi:10.1016/j.jacc.2006.02.054Vaitkevicius, R., Saburkina, I., Rysevaite, K., Vaitkeviciene, I., Pauziene, N., Zaliunas, R., … Pauza, D. H. (2009). Nerve Supply of the Human Pulmonary Veins: An Anatomical Study. Heart Rhythm, 6(2), 221-228. doi:10.1016/j.hrthm.2008.10.027Mabe, A. M., & Hoover, D. B. (2009). Structural and functional cardiac cholinergic deficits in adult neurturin knockout mice. Cardiovascular Research, 82(1), 93-99. doi:10.1093/cvr/cvp029Beau, S. L., Hand, D. E., Schuessler, R. B., Bromberg, B. I., Kwon, B., Boineau, J. P., & Saffitz, J. E. (1995). Relative Densities of Muscarinic Cholinergic and β-Adrenergic Receptors in the Canine Sinoatrial Node and Their Relation to Sites of Pacemaker Activity. Circulation Research, 77(5), 957-963. doi:10.1161/01.res.77.5.957Mangoni, M. E., & Nargeot, J. (2008). Genesis and Regulation of the Heart Automaticity. Physiological Reviews, 88(3), 919-982. doi:10.1152/physrev.00018.2007Brack, K. E., Coote, J. H., & Ng, G. A. (2003). Interaction between direct sympathetic and vagus nerve stimulation on heart rate in the isolated rabbit heart. Experimental Physiology, 89(1), 128-139. doi:10.1113/expphysiol.2003.002654Levy, M. N., & Zieske, H. (1969). Autonomic control of cardiac pacemaker activity and atrioventricular transmission. Journal of Applied Physiology, 27(4), 465-470. doi:10.1152/jappl.1969.27.4.465Hartzell, H. C. (1988). Regulation of cardiac ion channels by catecholamines, acetylcholine and second messenger systems. Progress in Biophysics and Molecular Biology, 52(3), 165-247. doi:10.1016/0079-6107(88)90014-4LEVY, M. N., YANG, T., & WALLICK, D. W. (1993). Assessment of Beat-by-Beat Control of Heart Rate by the Autonomic Nervous System: Molecular Biology Techniques Are Necessary, But Not Sufficient. Journal of Cardiovascular Electrophysiology, 4(2), 183-193. doi:10.1111/j.1540-8167.1993.tb01222.xLevy, M. N. (1971). Brief Reviews. Circulation Research, 29(5), 437-445. doi:10.1161/01.res.29.5.437Ng, G. A., Brack, K. E., & Coote, J. H. (2001). Effects of Direct Sympathetic and Vagus Nerve Stimulation on the Physiology of the Whole Heart - A Novel Model of Isolated Langendorff Perfused Rabbit Heart with Intact Dual Autonomic Innervation. Experimental Physiology, 86(3), 319-329. doi:10.1113/eph8602146Goldberg, J. (1975). Intra-SA-nodal pacemaker shifts induced by autonomic nerve stimulation in the dog. American Journal of Physiology-Legacy Content, 229(4), 1116-1123. doi:10.1152/ajplegacy.1975.229.4.1116Shibata, N., Inada, S., Mitsui, K., Honjo, H., Yamamoto, M., Niwa, R., … Kodama, I. (2001). Pacemaker Shift in the Rabbit Sinoatrial Node in Response to Vagal Nerve Stimulation. Experimental Physiology, 86(2), 177-184. doi:10.1113/eph8602100Glukhov, A. V., Fedorov, V. V., Anderson, M. E., Mohler, P. J., & Efimov, I. R. (2010). Functional anatomy of the murine sinus node: high-resolution optical mapping of ankyrin-B heterozygous mice. American Journal of Physiology-Heart and Circulatory Physiology, 299(2), H482-H491. doi:10.1152/ajpheart.00756.2009Michaels, D. C., Matyas, E. P., & Jalife, J. (1987). Mechanisms of sinoatrial pacemaker synchronization: a new hypothesis. Circulation Research, 61(5), 704-714. doi:10.1161/01.res.61.5.704Boyett, M. (2000). The sinoatrial node, a heterogeneous pacemaker structure. Cardiovascular Research, 47(4), 658-687. doi:10.1016/s0008-6363(00)00135-8Lemery, R., Birnie, D., Tang, A. S. L., Green, M., & Gollob, M. (2006). Feasibility study of endocardial mapping of ganglionated plexuses during catheter ablation of atrial fibrillation. Heart Rhythm, 3(4), 387-396. doi:10.1016/j.hrthm.2006.01.009Pokushalov, E., Romanov, A., Shugayev, P., Artyomenko, S., Shirokova, N., Turov, A., & Katritsis, D. G. (2009). Selective ganglionated plexi ablation for paroxysmal atrial fibrillation. Heart Rhythm, 6(9), 1257-1264. doi:10.1016/j.hrthm.2009.05.018Scherlag, B. J., Nakagawa, H., Jackman, W. M., Yamanashi, W. S., Patterson, E., Po, S., & Lazzara, R. (2005). Electrical Stimulation to Identify Neural Elements on the Heart: Their Role in Atrial Fibrillation. Journal of Interventional Cardiac Electrophysiology, 13(S1), 37-42. doi:10.1007/s10840-005-2492-2Puodziukynas, A., Kazakevicius, T., Vaitkevicius, R., Rysevaite, K., Jokubauskas, M., Saburkina, I., … Pauza, D. H. (2012). Radiofrequency catheter ablation of pulmonary vein roots results in axonal degeneration of distal epicardial nerves. Autonomic Neuroscience, 167(1-2), 61-65. doi:10.1016/j.autneu.2012.01.001Bauer, A., Deisenhofer, I., Schneider, R., Zrenner, B., Barthel, P., Karch, M., … Schmidt, G. (2006). Effects of circumferential or segmental pulmonary vein ablation for paroxysmal atrial fibrillation on cardiac autonomic function. Heart Rhythm, 3(12), 1428-1435. doi:10.1016/j.hrthm.2006.08.025Armour, J. A. (2010). Functional anatomy of intrathoracic neurons innervating the atria and ventricles. Heart Rhythm, 7(7), 994-996. doi:10.1016/j.hrthm.2010.02.01

    Ventricular Fibrillation

    No full text

    The nervous heart

    No full text
    Many cardiac electrophysiological abnormalities are accompanied by autonomic nervous system dysfunction. Here, we review mechanisms by which the cardiac nervous system controls normal and abnormal excitability and may contribute to atrial and ventricular tachyarrhythmias. Moreover, we explore the potential antiarrhythmic and/or arrhythmogenic effects of modulating the autonomic nervous system by several strategies, including ganglionated plexi ablation, vagal and spinal cord stimulations, and renal sympathetic denervation as therapies for atrial and ventricular arrhythmias
    corecore