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 -- 1493 2022-09-13 19:58:19 |
2 format correct Meta information modification 1493 2022-09-14 02:52:10 |

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.
Carbone, F.;  Liberale, L.;  Preda, A.;  Schindler, T.H.;  Montecucco, F. Pathophysiology of Septic Cardiomyopathy. Encyclopedia. Available online: https://encyclopedia.pub/entry/27143 (accessed on 19 April 2024).
Carbone F,  Liberale L,  Preda A,  Schindler TH,  Montecucco F. Pathophysiology of Septic Cardiomyopathy. Encyclopedia. Available at: https://encyclopedia.pub/entry/27143. Accessed April 19, 2024.
Carbone, Federico, Luca Liberale, Alberto Preda, Thomas Hellmut Schindler, Fabrizio Montecucco. "Pathophysiology of Septic Cardiomyopathy" Encyclopedia, https://encyclopedia.pub/entry/27143 (accessed April 19, 2024).
Carbone, F.,  Liberale, L.,  Preda, A.,  Schindler, T.H., & Montecucco, F. (2022, September 13). Pathophysiology of Septic Cardiomyopathy. In Encyclopedia. https://encyclopedia.pub/entry/27143
Carbone, Federico, et al. "Pathophysiology of Septic Cardiomyopathy." Encyclopedia. Web. 13 September, 2022.
Pathophysiology of Septic Cardiomyopathy
Edit

Septic cardiomyopathy may be broadly defined as an acute cardiac dysfunction unrelated to ischemia that manifests in different ways: arrhythmias, left and/or right ventricular impairment during systole or diastole, with or without reduction in cardiac output. Endothelial, metabolic, and immune response abnormalities are generally involved in the pathogenesis of ventricular dysfunction and arrhythmias during sepsis, whereas the potential role of myocardial ischemia seems limited. Impaired blood flow autoregulation in coronary microcirculation and altered metabolism of lactate, free fatty acid, and glucose likely play a leading role. 

sepsis septic shock cardiomyopathy inflammation

1. Inflammatory Pathways and Cardiomyocyte Dysfunction

The hyperinflammatory—not counterbalanced—response is historically reported as a paradigm of sepsis. Early clinical trials then focused on blunting inflammatory response by common anti-inflammatory drugs such as glucocorticoids, non-steroidal anti-inflammatory drugs or target therapy, with very limited results [1]. The extreme variability in pro-/anti-inflammatory balance and the wide number of molecular pathways may explain such disheartening results [2][3]. Ancestral signals released by pathogens (pathogen-associated molecular patterns, PAMPs) or by damaged host tissues (damage-associated molecular patterns, DAMPs) and receptors such as Toll-like receptors (TLRs) trigger multiple intracellular pathways, including the activation of nuclear factor-kB (NF-kB) and mitogen-activated protein kinase (MAPK) [4][5][6]. The DAMPs with a recognized role in the pathophysiology of septic cardiomyopathy include: heparan sulphate, which increases intranuclear transcription of pro-inflammatory cytokines and vascular permeability [7]; high mobility group protein B1, which induces loss of calcium from sarcoplasmic reticulum [8]; and histones, which are able to interfere with the production of cellular ATP by reducing mitochondrial membrane potential and causing LV dysfunction and arrhythmias [9]. Among PAMPs, bacterial lipopolysaccharide (LPS) partially mimics hemodynamic effects of septic shock when administered to animal models or humans [10][11]. Detrimental effects of endotoxemia on cardiac contractility are largely due to mitochondrial dysfunction that leads to abnormal calcium handling, disruption of ATP synthesis, endothelium reticulum stress, and autophagy. Furthermore, LPS leads to electrophysiological dysfunctions through a both a direct and cytokine-mediated effect on sodium current kinetics and non-selective cation channel transient receptor potential vanilloid 1 (TRPV1) [12][13][14][15]. Furthermore, LPS and its downstream mediator TNF-α, IL-1β, IL-6, and C5a and ROS [16][17] critically alter calcium (Ca++) homeostasis by blunting the amplitude of intracellular currents and concentrations. Dysfunction of intracellular calcium transporters and decreased calcium sensitivity in cardiac myofilaments (due to phosphorylation of inhibitory troponin I) are cornerstones of the impaired excitation-contraction coupling. Ultimately, endotoxemia impairs sarcolemma diastolic Ca++ extrusion with consequent overload [18][19][20][21] that determines systolic and diastolic dysfunction during sepsis [22][23]. The characteristic delayed reversibility of septic cardiomyopathy might be explained by the synthesis of new myofilaments to replace the previously phosphorylated (and therefore inactive) ones.
Signals regulated by the TLRs expressed on myocyte surface (TLR2, TRL3, TLR4 and TLR9) lead to transcription of several pro-inflammatory cardiac depressive factors and cytokines such as interleukin (IL)-1, IL-6, tumor necrosis factor (TNF)-α and complement anaphylatoxin C5a [24][25]. LV ejection fraction (LVEF) was found higher in septic mouse with TRL3 and TLR9 gene deletion, suggesting their role in the mechanism of sepsis-induced cardiac dysfunction [26]. Likewise, TLR4 regulates oxidative stress in ryanodine receptor 2 involved in the storage of calcium within sarcoplasmic reticulum of cardiomyocytes [27], and its inhibition showed protective effects [28]. C5a is so far the only complement factor with reported direct myocardial depressive effect, despite several cardiac enzymes related to cardiac dysfunction during sepsis, such as serca2, NCX and Na+/K+-atpase, which are complement receptor dependent [29]. Finally, adhesion molecules take part in homing mechanisms and are upregulated during inflammatory response. In murine coronary endothelium and cardiomyocytes, specific adhesion molecules were upregulated after LPS and TNF-α infusion. Moreover, the use of antibody blocking these molecules such as anti-Intercellular adhesion molecule-1 (ICAM-1) and vascular adhesion molecule-1 (VCAM-1) showed to prevent myocardial neutrophil accumulation and cardiac dysfunction in animal models of sepsis [30]. On the contrary, neutrophils depletion did not protect against myocardial dysfunction during sepsis, suggesting their lesser cardiotoxic potential.
Nitric oxide (NO) is synthesized by nitric oxide synthase (NOS) in different cells of the CV system including cardiac myocytes, and it exerts important roles in maintaining tissue homeostasis by reducing oxidative stress. Indeed, via the cGMP pathway, NO regulates vascular tone, has an antioxidant effect, inhibits leukocyte and platelet adhesion to the endothelium, and increases myocardial contractility [31]. Inflammation and oxidative stress show deep interplay, fueling each other and facilitating the onset of deleterious vicious circles in different diseases including sepsis. Differently from its endothelial and neuronal isoforms, inducible NOS (iNOS) is not constitutively active and when highly expressed is a major responsible for vasodilatation and hypotension in shock [32]. Indeed, iNOS can produce large amount of NO when an inflammatory response occurs [33]. Sepsis leads to overexpression of iNOS not only in immune cells but also in the myocardium [34]. Such increased expression has detrimental effects on contractile function of cardiomyocytes, in part through the paradoxical induction of reactive oxygen species (ROS), i.e., peroxynitrite [35], in part through the down-regulation of adrenaline receptors and decreased sensitivity to calcium [7]. Confirming the important role played by iNOS in myocardial dysfunction during sepsis, several non-specific anti-NOS drugs such as melatonin and methylene blue showed beneficial effects in terms of cardiovascular function and prognosis in preclinical studies [35][36].

2. Adrenergic System

Increased activation of the sympathetic system is a major compensatory reaction to septic-related vasoplegia. During sepsis, β-adrenergic receptors (βARs) are downregulated, and the responsiveness to catecholamines is reduced in the whole cardiovascular system [37][38]. When persistent, this adaptive response become maladaptive [39]. In myocardial tissue, excessive stimulation of cardiac βARs suppresses their expression and leads to an inhibitory response with reverse of adrenergic G protein coupling [40], intracellular calcium overload [41], increased production of ROS [42], disruption of membrane potential via the inhibition of Na+/K+-ATPase pump [43], and induction of apoptosis [44]. A relevant role is played by Cardiac G Protein-Coupled Receptor Kinase 2 (GRK2). They are the major negative regulators of βAR pro-contractile signal in sepsis-induced myocardial dysfunction [45] and other different cardiac diseases including chronic heart failure and Takotsubo syndrome [46].

3. Microvascular Dysfunction and Vasoactive Peptides

Activation of endothelial cells in consequence of an infective stimulus leads to the synthesis of inflammatory cytokines, increased expression of cell adhesion molecules, loss of the barrier function by glycocalyx shedding, edema, apoptosis, hypercoagulative state and vasoplegia [47]. Several endogenous vasoactive peptides generate and maintain endothelial dysregulation. LPS and cytokines up-regulate cyclooxygenase (COX)-2 (inducible form) and then the production of prostanoids from arachidonic acid. This process mainly occurs in inflammatory cells, myofibroblasts, endothelium and even in cardiomyocytes (i.e., in myocardial infarction) [48][49]. Elevated levels of prostanoids such as thromboxane and prostacyclin are associated with coronary microvascular dysfunction [50]. However, clinical trials have failed so far in proving any beneficial effect of pharmacological COX inhibition on coronary microvascular homeostasis in sepsis. Less is known about the potential role of endothelin-1 (ET-1) in sepsis cardiomyopathy, whereas an increase has been reported in chronic cardiac disease (i.e., heart failure) [51]. ET-1 would have a major role in infectious disease, as a promoter of cytokines release, platelet aggregation and vasoreactivity [52]. In single mouse study, cardiac overexpression of ET-1 was associated with the onset of severe sepsis cardiomyopathy, characterized by interstitial infiltration of macrophages and T lymphocytes and increased levels of the pro-inflammatory cytokines TNF-α, INF-γ, IL-1, and IL-6 [53].

4. Energetic Dysmetabolism

Sepsis is characterized by altered myocardial lipoprotein metabolism and mobilization of triglycerides and free fatty acids to overwhelm the systemic suppression of energy production [54]. Physiologically, about 70% of cardiac ATP is produced via lipid oxidation, while the rest is produced via glucose oxidation. A minor part also derives from the catabolism of lactate and ketone bodies. Under pathological conditions, glucose oxidation becomes the prevalent energetic pathway [55]. During sepsis, a reduction in fatty acid oxidation is not compensated by the increase in glucose catabolism due to altered insulin action, exacerbation of the inhibitory effects of alternative substrates on glycolysis, and blunted glycogen synthesis [56]. Several enzymes involved in intracellular cardiac fatty acid mobilization and oxidation are also inhibited [57][58]. Cardiac transcriptional factors associated with fatty acid oxidation, such as peroxisome proliferator activated receptors (PPARs) are also suppressed in sepsis and their stimulation with PPARs agonists lead to improved survival in septic mice [59]. Mitochondrial dysfunction further contributes to the energetic failure in sepsis cardiomyopathy, thus being a non-negligible cause of reduced outcome [60][61][62]. Inflammation and oxidative stress alter mitochondrial structure determining swelling, cytoplasmic accumulation of denatured protein and lysosomal lesions [63][64]. Such a damage interferes with the respiratory chain [65] with falling in ATP synthesis, release of calcium and pro-apoptotic proteins [66]. The pro-oxidant environment induced by endotoxemia also trigger cardiac mitophagy, a defensive mechanism for the removal of damaged mitochondria [67]. The antioxidant N-acetylcysteine seems to have a protective roles against contractile dysfunction and mitophagy. Even levosimendam in both mice and human beings has beneficial effects due to calcium sensitization and antioxidant properties [68]. Lastly, when the degree of mitochondrial dysfunction is mild, myocardial hibernation may occur. This is an adaptive, self-protective mechanism during which all functions get reduced. Such a downregulation of mitochondrial gene transcription has been reported in sepsis cardiomyopathy [69] and represents a reversible condition that improves together with the resolution of sepsis [62].

References

  1. Cavaillon, J.M.; Singer, M.; Skirecki, T. Sepsis therapies: Learning from 30 years of failure of translational research to propose new leads. EMBO Mol. Med. 2020, 12, e10128.
  2. Venet, F.; Davin, F.; Guignant, C.; Larue, A.; Cazalis, M.A.; Darbon, R.; Allombert, C.; Mougin, B.; Malcus, C.; Poitevin-Later, F.; et al. Early assessment of leukocyte alterations at diagnosis of septic shock. Shock 2010, 34, 358–363.
  3. Davies, R.; O’Dea, K.; Gordon, A. Immune therapy in sepsis: Are we ready to try again? J. Intensive Care Soc. 2018, 19, 326–344.
  4. Conway-Morris, A.; Wilson, J.; Shankar-Hari, M. Immune activation in sepsis. Crit. Care Clin. 2018, 34, 29–42.
  5. Hawiger, J. Heartfelt sepsis: Microvascular injury due to genomic storm. Kardiol. Pol. 2018, 76, 1203–1216.
  6. Rosengart, M.R.; Nathens, A.B.; Arbabi, S.; Neff, M.J.; Garcia, I.; Martin, T.R.; Maier, R.V. Mitogen-activated protein kinases in the intensive care unit: Prognostic potential. Ann. Surg. 2003, 237, 94–100.
  7. Martin, L.; Schmitz, S.; De Santis, R.; Doemming, S.; Haase, H.; Hoeger, J.; Heinbockel, L.; Brandenburg, K.; Marx, G.; Schuerholz, T. Peptide 19-2.5 inhibits heparan sulfate-triggered inflammation in murine cardiomyocytes stimulated with human sepsis serum. PLoS ONE 2015, 10, e0127584.
  8. Wang, H.; Ward, M.F.; Sama, A.E. Novel hmgb1-inhibiting therapeutic agents for experimental sepsis. Shock 2009, 32, 348–357.
  9. Alhamdi, Y.; Abrams, S.T.; Cheng, Z.; Jing, S.; Su, D.; Liu, Z.; Lane, S.; Welters, I.; Wang, G.; Toh, C.H. Circulating histones are major mediators of cardiac injury in patients with sepsis. Crit. Care Med. 2015, 43, 2094–2103.
  10. Suffredini, A.F.; Fromm, R.E.; Parker, M.M.; Brenner, M.; Kovacs, J.A.; Wesley, R.A.; Parrillo, J.E. The cardiovascular response of normal humans to the administration of endotoxin. N. Engl. J. Med. 1989, 321, 280–287.
  11. Danner, R.L.; Elin, R.J.; Hosseini, J.M.; Wesley, R.A.; Reilly, J.M.; Parillo, J.E. Endotoxemia in human septic shock. Chest 1991, 99, 169–175.
  12. Yucel, G.; Zhao, Z.; El-Battrawy, I.; Lan, H.; Lang, S.; Li, X.; Buljubasic, F.; Zimmermann, W.H.; Cyganek, L.; Utikal, J.; et al. Lipopolysaccharides induced inflammatory responses and electrophysiological dysfunctions in human-induced pluripotent stem cell derived cardiomyocytes. Sci. Rep. 2017, 7, 2935.
  13. Sattler, K.; El-Battrawy, I.; Cyganek, L.; Lang, S.; Lan, H.; Li, X.; Zhao, Z.; Utikal, J.; Wieland, T.; Borggrefe, M.; et al. Trpv1 activation and internalization is part of the lps-induced inflammation in human ipsc-derived cardiomyocytes. Sci. Rep. 2021, 11, 14689.
  14. Liao, Z.; Li, Y.; Fan, X.; Yang, Z.; El-Battrawy, I.; Zhou, X.; Akin, I. Lipopolysaccharide modifies sodium current kinetics through ros and pkc signalling in induced pluripotent stem-derived cardiomyocytes from brugada syndrome patient. J. Cardiovasc. Dev. Dis. 2022, 9, 119.
  15. Fan, X.; Yang, G.; Kowitz, J.; Akin, I.; Zhou, X.; El-Battrawy, I. Takotsubo syndrome: Translational implications and pathomechanisms. Int. J. Mol. Sci. 2022, 23, 1951.
  16. Natanson, C.; Eichenholz, P.W.; Danner, R.L.; Eichacker, P.Q.; Hoffman, W.D.; Kuo, G.C.; Banks, S.M.; MacVittie, T.J.; Parrillo, J.E. Endotoxin and tumor necrosis factor challenges in dogs simulate the cardiovascular profile of human septic shock. J. Exp. Med. 1989, 169, 823–832.
  17. Kalbitz, M.; Fattahi, F.; Grailer, J.J.; Jajou, L.; Malan, E.A.; Zetoune, F.S.; Huber-Lang, M.; Russell, M.W.; Ward, P.A. Complement-induced activation of the cardiac nlrp3 inflammasome in sepsis. FASEB J. 2016, 30, 3997–4006.
  18. Wagner, S.; Schurmann, S.; Hein, S.; Schuttler, J.; Friedrich, O. Septic cardiomyopathy in rat lps-induced endotoxemia: Relative contribution of cellular diastolic ca(2+) removal pathways, myofibrillar biomechanics properties and action of the cardiotonic drug levosimendan. Basic Res. Cardiol. 2015, 110, 507.
  19. Koentges, C.; Cimolai, M.C.; Pfeil, K.; Wolf, D.; Marchini, T.; Tarkhnishvili, A.; Hoffmann, M.M.; Odening, K.E.; Diehl, P.; von Zur Muhlen, C.; et al. Impaired sirt3 activity mediates cardiac dysfunction in endotoxemia by calpain-dependent disruption of atp synthesis. J. Mol. Cell. Cardiol. 2019, 133, 138–147.
  20. Luo, Y.; Fan, C.; Yang, M.; Dong, M.; Bucala, R.; Pei, Z.; Zhang, Y.; Ren, J. Cd74 knockout protects against lps-induced myocardial contractile dysfunction through ampk-skp2-suv39h1-mediated demethylation of bclb. Br. J. Pharm. 2020, 177, 1881–1897.
  21. Joseph, L.C.; Reyes, M.V.; Lakkadi, K.R.; Gowen, B.H.; Hasko, G.; Drosatos, K.; Morrow, J.P. Pkcdelta causes sepsis-induced cardiomyopathy by inducing mitochondrial dysfunction. Am. J. Physiol. Heart Circ. Physiol. 2020, 318, H778–H786.
  22. Hobai, I.A.; Edgecomb, J.; LaBarge, K.; Colucci, W.S. Dysregulation of intracellular calcium transporters in animal models of sepsis-induced cardiomyopathy. Shock 2015, 43, 3–15.
  23. DeGrande, S.T.; Little, S.C.; Nixon, D.J.; Wright, P.; Snyder, J.; Dun, W.; Murphy, N.; Kilic, A.; Higgins, R.; Binkley, P.F.; et al. Molecular mechanisms underlying cardiac protein phosphatase 2a regulation in heart. J. Biol. Chem. 2013, 288, 1032–1046.
  24. Antonucci, E.; Fiaccadori, E.; Donadello, K.; Taccone, F.S.; Franchi, F.; Scolletta, S. Myocardial depression in sepsis: From pathogenesis to clinical manifestations and treatment. J. Crit. Care 2014, 29, 500–511.
  25. Zaky, A.; Deem, S.; Bendjelid, K.; Treggiari, M.M. Characterization of cardiac dysfunction in sepsis: An ongoing challenge. Shock 2014, 41, 12–24.
  26. Vallejo, J.G. Role of toll-like receptors in cardiovascular diseases. Clin. Sci. 2011, 121, 1–10.
  27. Yang, J.; Zhang, R.; Jiang, X.; Lv, J.; Li, Y.; Ye, H.; Liu, W.; Wang, G.; Zhang, C.; Zheng, N.; et al. Toll-like receptor 4-induced ryanodine receptor 2 oxidation and sarcoplasmic reticulum ca(2+) leakage promote cardiac contractile dysfunction in sepsis. J. Biol. Chem. 2018, 293, 794–807.
  28. Nath, K.A.; Belcher, J.D.; Nath, M.C.; Grande, J.P.; Croatt, A.J.; Ackerman, A.W.; Katusic, Z.S.; Vercellotti, G.M. Role of tlr4 signaling in the nephrotoxicity of heme and heme proteins. Am. J. Physiol. Ren. Physiol. 2018, 314, F906–F914.
  29. Fattahi, F.; Frydrych, L.M.; Bian, G.; Kalbitz, M.; Herron, T.J.; Malan, E.A.; Delano, M.J.; Ward, P.A. Role of complement c5a and histones in septic cardiomyopathy. Mol. Immunol. 2018, 102, 32–41.
  30. Raeburn, C.D.; Calkins, C.M.; Zimmerman, M.A.; Song, Y.; Ao, L.; Banerjee, A.; Harken, A.H.; Meng, X. Icam-1 and vcam-1 mediate endotoxemic myocardial dysfunction independent of neutrophil accumulation. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2002, 283, R477–R486.
  31. Rastaldo, R.; Pagliaro, P.; Cappello, S.; Penna, C.; Mancardi, D.; Westerhof, N.; Losano, G. Nitric oxide and cardiac function. Life Sci. 2007, 81, 779–793.
  32. Kakihana, Y.; Ito, T.; Nakahara, M.; Yamaguchi, K.; Yasuda, T. Sepsis-induced myocardial dysfunction: Pathophysiology and management. J. Intensive Care 2016, 4, 22.
  33. Spiller, F.; Oliveira Formiga, R.; Fernandes da Silva Coimbra, J.; Alves-Filho, J.C.; Cunha, T.M.; Cunha, F.Q. Targeting nitric oxide as a key modulator of sepsis, arthritis and pain. Nitric Oxide 2019, 89, 32–40.
  34. Khadour, F.H.; Panas, D.; Ferdinandy, P.; Schulze, C.; Csont, T.; Lalu, M.M.; Wildhirt, S.M.; Schulz, R. Enhanced no and superoxide generation in dysfunctional hearts from endotoxemic rats. Am. J. Physiol. Heart Circ. Physiol. 2002, 283, H1108–H1115.
  35. Pacher, P.; Beckman, J.S.; Liaudet, L. Nitric oxide and peroxynitrite in health and disease. Physiol. Rev. 2007, 87, 315–424.
  36. Kirov, M.Y.; Evgenov, O.V.; Evgenov, N.V.; Egorina, E.M.; Sovershaev, M.A.; Sveinbjørnsson, B.; Nedashkovsky, E.V.; Bjertnaes, L.J. Infusion of methylene blue in human septic shock: A pilot, randomized, controlled study. Crit. Care Med. 2001, 29, 1860–1867.
  37. Hollenberg, S.M.; Cunnion, R.E.; Zimmerberg, J. Nitric oxide synthase inhibition reverses arteriolar hyporesponsiveness to catecholamines in septic rats. Am. J. Physiol. 1993, 264, H660–H663.
  38. Rudiger, A.; Dyson, A.; Felsmann, K.; Carré, J.E.; Taylor, V.; Hughes, S.; Clatworthy, I.; Protti, A.; Pellerin, D.; Lemm, J.; et al. Early functional and transcriptomic changes in the myocardium predict outcome in a long-term rat model of sepsis. Clin. Sci. 2013, 124, 391–401.
  39. Liu, Y.C.; Yu, M.M.; Shou, S.T.; Chai, Y.F. Sepsis-induced cardiomyopathy: Mechanisms and treatments. Front. Immunol. 2017, 8, 1021.
  40. Rehman, A.; Baloch, N.U.; Morrow, J.P.; Pacher, P.; Haskó, G. Targeting of g-protein coupled receptors in sepsis. Pharmacol. Ther. 2020, 211, 107529.
  41. Celes, M.R.; Malvestio, L.M.; Suadicani, S.O.; Prado, C.M.; Figueiredo, M.J.; Campos, E.C.; Freitas, A.C.; Spray, D.C.; Tanowitz, H.B.; da Silva, J.S.; et al. Disruption of calcium homeostasis in cardiomyocytes underlies cardiac structural and functional changes in severe sepsis. PLoS ONE 2013, 8, e68809.
  42. Macarthur, H.; Westfall, T.C.; Riley, D.P.; Misko, T.P.; Salvemini, D. Inactivation of catecholamines by superoxide gives new insights on the pathogenesis of septic shock. Proc. Natl. Acad. Sci. USA 2000, 97, 9753–9758.
  43. Kalbitz, M.; Fattahi, F.; Herron, T.J.; Grailer, J.J.; Jajou, L.; Lu, H.; Huber-Lang, M.; Zetoune, F.S.; Sarma, J.V.; Day, S.M.; et al. Complement destabilizes cardiomyocyte function in vivo after polymicrobial sepsis and in vitro. J. Immunol. 2016, 197, 2353–2361.
  44. Nuding, S.; Schröder, J.; Presek, P.; Wienke, A.; Müller-Werdan, U.; Ebelt, H.; Werdan, K. Reducing elevated heart rates in patients with multiple organ dysfunction syndrome with the if (funny channel current) inhibitor ivabradine. Shock 2018, 49, 402–411.
  45. Dal-Secco, D.; DalBó, S.; Lautherbach, N.E.S.; Gava, F.N.; Celes, M.R.N.; Benedet, P.O.; Souza, A.H.; Akinaga, J.; Lima, V.; Silva, K.P.; et al. Cardiac hyporesponsiveness in severe sepsis is associated with nitric oxide-dependent activation of g protein receptor kinase. Am. J. Physiol. Heart Circ. Physiol. 2017, 313, H149–H163.
  46. Nakano, T.; Onoue, K.; Nakada, Y.; Nakagawa, H.; Kumazawa, T.; Ueda, T.; Nishida, T.; Soeda, T.; Okayama, S.; Watanabe, M.; et al. Alteration of β-adrenoceptor signaling in left ventricle of acute phase takotsubo syndrome: A human study. Sci. Rep. 2018, 8, 12731.
  47. Ince, C.; Mayeux, P.R.; Nguyen, T.; Gomez, H.; Kellum, J.A.; Ospina-Tascón, G.A.; Hernandez, G.; Murray, P.; De Backer, D. The endothelium in sepsis. Shock 2016, 45, 259–270.
  48. Liu, S.F.; Newton, R.; Evans, T.W.; Barnes, P.J. Differential regulation of cyclo-oxygenase-1 and cyclo-oxygenase-2 gene expression by lipopolysaccharide treatment in vivo in the rat. Clin. Sci. 1996, 90, 301–306.
  49. Zidar, N.; Dolenc-Strazar, Z.; Jeruc, J.; Jerse, M.; Balazic, J.; Gartner, U.; Jermol, U.; Zupanc, T.; Stajer, D. Expression of cyclooxygenase-1 and cyclooxygenase-2 in the normal human heart and in myocardial infarction. Cardiovasc. Pathol. 2007, 16, 300–304.
  50. Reines, H.D.; Halushka, P.V.; Cook, J.A.; Wise, W.C.; Rambo, W. Plasma thromboxane concentrations are raised in patients dying with septic shock. Lancet 1982, 2, 174–175.
  51. Stewart, D.J.; Cernacek, P.; Costello, K.B.; Rouleau, J.L. Elevated endothelin-1 in heart failure and loss of normal response to postural change. Circulation 1992, 85, 510–517.
  52. Freeman, B.D.; Machado, F.S.; Tanowitz, H.B.; Desruisseaux, M.S. Endothelin-1 and its role in the pathogenesis of infectious diseases. Life Sci. 2014, 118, 110–119.
  53. Yang, L.L.; Gros, R.; Kabir, M.G.; Sadi, A.; Gotlieb, A.I.; Husain, M.; Stewart, D.J. Conditional cardiac overexpression of endothelin-1 induces inflammation and dilated cardiomyopathy in mice. Circulation 2004, 109, 255–261.
  54. Carrara, M.; Ferrario, M.; Bollen Pinto, B.; Herpain, A. The autonomic nervous system in septic shock and its role as a future therapeutic target: A narrative review. Ann. Intensive Care 2021, 11, 80.
  55. Fillmore, N.; Mori, J.; Lopaschuk, G.D. Mitochondrial fatty acid oxidation alterations in heart failure, ischaemic heart disease and diabetic cardiomyopathy. Br. J. Pharm. 2014, 171, 2080–2090.
  56. Tessier, J.P.; Thurner, B.; Jüngling, E.; Lückhoff, A.; Fischer, Y. Impairment of glucose metabolism in hearts from rats treated with endotoxin. Cardiovasc. Res. 2003, 60, 119–130.
  57. Memon, R.A.; Fuller, J.; Moser, A.H.; Smith, P.J.; Feingold, K.R.; Grunfeld, C. In vivo regulation of acyl-coa synthetase mrna and activity by endotoxin and cytokines. Am. J. Physiol. 1998, 275, E64–E72.
  58. Memon, R.A.; Bass, N.M.; Moser, A.H.; Fuller, J.; Appel, R.; Grunfeld, C.; Feingold, K.R. Down-regulation of liver and heart specific fatty acid binding proteins by endotoxin and cytokines in vivo. Biochim. Biophys. Acta 1999, 1440, 118–126.
  59. Drosatos, K.; Khan, R.S.; Trent, C.M.; Jiang, H.; Son, N.H.; Blaner, W.S.; Homma, S.; Schulze, P.C.; Goldberg, I.J. Peroxisome proliferator-activated receptor-γ activation prevents sepsis-related cardiac dysfunction and mortality in mice. Circ. Heart Fail. 2013, 6, 550–562.
  60. Brealey, D.; Brand, M.; Hargreaves, I.; Heales, S.; Land, J.; Smolenski, R.; Davies, N.A.; Cooper, C.E.; Singer, M. Association between mitochondrial dysfunction and severity and outcome of septic shock. Lancet 2002, 360, 219–223.
  61. Stanzani, G.; Duchen, M.R.; Singer, M. The role of mitochondria in sepsis-induced cardiomyopathy. Biochim. Biophys. Acta Mol. Basis Dis. 2019, 1865, 759–773.
  62. Brealey, D.; Karyampudi, S.; Jacques, T.S.; Novelli, M.; Stidwill, R.; Taylor, V.; Smolenski, R.T.; Singer, M. Mitochondrial dysfunction in a long-term rodent model of sepsis and organ failure. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2004, 286, R491–R497.
  63. Hotchkiss, R.S.; Nicholson, D.W. Apoptosis and caspases regulate death and inflammation in sepsis. Nat. Rev. Immunol. 2006, 6, 813–822.
  64. Crouser, E.D. Mitochondrial dysfunction in septic shock and multiple organ dysfunction syndrome. Mitochondrion 2004, 4, 729–741.
  65. Dikalov, S. Cross talk between mitochondria and nadph oxidases. Free Radic. Biol. Med. 2011, 51, 1289–1301.
  66. Bernardi, P.; Di Lisa, F. The mitochondrial permeability transition pore: Molecular nature and role as a target in cardioprotection. J. Mol. Cell. Cardiol. 2015, 78, 100–106.
  67. Turdi, S.; Han, X.; Huff, A.F.; Roe, N.D.; Hu, N.; Gao, F.; Ren, J. Cardiac-specific overexpression of catalase attenuates lipopolysaccharide-induced myocardial contractile dysfunction: Role of autophagy. Free Radic. Biol. Med. 2012, 53, 1327–1338.
  68. Torraco, A.; Carrozzo, R.; Piemonte, F.; Pastore, A.; Tozzi, G.; Verrigni, D.; Assenza, M.; Orecchioni, A.; D’Egidio, A.; Marraffa, E.; et al. Effects of levosimendan on mitochondrial function in patients with septic shock: A randomized trial. Biochimie 2014, 102, 166–173.
  69. Matkovich, S.J.; Al Khiami, B.; Efimov, I.R.; Evans, S.; Vader, J.; Jain, A.; Brownstein, B.H.; Hotchkiss, R.S.; Mann, D.L. Widespread down-regulation of cardiac mitochondrial and sarcomeric genes in patients with sepsis. Crit. Care Med. 2017, 45, 407–414.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , ,
View Times: 396
Revisions: 2 times (View History)
Update Date: 14 Sep 2022
1000/1000