Next Article in Journal
Clinical Features of COVID-19 in Elderly Patients: Tools for Predicting Outcomes Are Needed
Next Article in Special Issue
Acute Heart Failure: Diagnostic–Therapeutic Pathways and Preventive Strategies—A Real-World Clinician’s Guide
Previous Article in Journal
The Effect of Maternal Coagulation Parameters on Fetal Acidemia in Placental Abruption
Previous Article in Special Issue
Effects of Cardiac Contractility Modulation Electrodes on Tricuspid Regurgitation in Patients with Heart Failure with Reduced Ejection Fraction: A Pilot Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Efficacy of the New Inotropic Agent Istaroxime in Acute Heart Failure

1
Division of Cardiology, Department of Advanced Biomedical Sciences, “Federico II” University, 80131 Naples, Italy
2
Division of Cardiology, Department of Medicine, Wilf Family Cardiovascular Research Institute, Einstein Institute for Aging Research, Albert Einstein College of Medicine, New York, NY 10461, USA
3
Casa di Cura “Montevergine”, Mercogliano, 83013 Avellino, Italy
4
Department of Mental and Physical Health and Preventive Medicine, University of Campania “Vanvitelli”, 81100 Caserta, Italy
5
Department of Molecular Pharmacology, Einstein-Mount Sinai Diabetes Research Center (ES-DRC), Einstein Institute for Neuroimmunology and Inflammation (INI), Fleischer Institute for Diabetes and Metabolism (FIDAM), Albert Einstein College of Medicine, New York, NY 10461, USA
*
Author to whom correspondence should be addressed.
J. Clin. Med. 2022, 11(24), 7503; https://0-doi-org.brum.beds.ac.uk/10.3390/jcm11247503
Submission received: 24 November 2022 / Revised: 14 December 2022 / Accepted: 15 December 2022 / Published: 18 December 2022

Abstract

:
Current therapeutic strategies for acute heart failure (AHF) are based on traditional inotropic agents that are often associated with untoward effects; therefore, finding new effective approaches with a safer profile is dramatically needed. Istaroxime is a novel compound, chemically unrelated to cardiac glycosides, that is currently being studied for the treatment of AHF. Its effects are essentially related to its inotropic and lusitropic positive properties exerted through a dual mechanism of action: activation of the sarcoplasmic reticulum Ca2+ ATPase isoform 2a (SERCA2a) and inhibition of the Na+/K+-ATPase (NKA) activity. The advantages of istaroxime over the available inotropic agents include its lower arrhythmogenic action combined with its capability of increasing systolic blood pressure without augmenting heart rate. However, it has a limited half-life (1 hour) and is associated with adverse effects including pain at the injection site and gastrointestinal issues. Herein, we describe the main mechanism of action of istaroxime and we present a systematic overview of both clinical and preclinical trials testing this drug, underlining the latest insights regarding its adoption in clinical practice for AHF.

1. Introduction

Heart failure (HF) is a clinical syndrome characterized by several symptoms, including dyspnea, ankle swelling, and fatigue, and signs, including peripheral edema, elevated jugular venous pressure, and pulmonary crackles. It is mainly due to structural and/or functional cardiac abnormalities causing a failure of the heart to pump enough to satisfy the metabolic requirements of the organism or generating an elevated intracardiac pressure to provide them [1].
With the aging of the population, HF incidence and prevalence are increasing worldwide [2,3]. HF incidence is about 6.0–7.9 per 1000 person-years in people >45 years old and about 21 per 1000 in people >65 years old [4,5]. The total percentage of people affected by HF is expected to rise from 2.4% in 2012 to 3.0% in 2030. HF prevalence is extremely variable with the lowest numbers in sub-Saharan Africa, but prevalence is projected to rise even in low- and middle-income countries as populations age and the burden of HF risk factors, such as elevated blood pressure, continues to increase [6,7]. HF has different classifications, based on the ejection fraction (EF), according to European Society of Cardiology (ESC) and the American Heart Association (AHA) guidelines. The ESC guidelines define HF as follows: HF with preserved ejection fraction (HFpEF) is characterized by an EF ≥ 50%, HF with mildly reduced ejection fraction (HFmrEF) is characterized by an 40% ≤EF < 49%, and HF with reduced ejection fraction (HFrEF) is characterized by an EF < 40%. The AHA defines: HFpEF with EF ≥ 50%; HFmrEF with 41% ≤ EF ≤ 49%; HF with improved EF (HFimpEF) for patients who have EF improved from a lower level to an EF >40% at follow-up; and HFrEF has EF ≤ 40% [1,8,9].
Acute HF (AHF) is universally considered as a gradual or rapid onset of signs and symptoms of HF resulting in a need for urgent therapy or hospitalization [8,10,11,12,13]. AHF can lead to cardiogenic shock, characterized by a life-threatening low-cardiac-output state causing end-organ hypoperfusion and hypoxia [8,14,15,16,17,18]. AHF in-hospital mortality ranges from 4 to 13% and the mortality rate at 1-year post-discharge is 25–30%, with more than 45% of readmissions [19,20,21,22,23].
Current therapy for AHF is based on inotropic agents that are often associated with serious adverse effects. Indeed, catecholamines, inhibitors of Na+/K+-ATPase (NKA), phosphodiesterase-3 inhibitors, and calcium (Ca2+) sensitizers, have been associated with tachycardia, ischemia, hypotension, and even with an excess of mortality, presumably related to arrhythmias in the short-term and the activation of signaling pathways that aggravate maladaptive remodeling of the failing heart in the long-term [24,25,26]. Therefore, there is an urgent need to identify new effective inotropic modulators with a safer profile.
Istaroxime is a relatively novel compound, a derivative of androstenedione, is chemically unrelated to cardiac glycosides, and possesses inotropic and lusitropic actions exerted through a dual mechanism: inhibition of the Na+/K+-ATPase (NKA) and activation of the sarcoplasmic reticulum calcium ATPase isoform 2a (SERCA2a) [27].

2. NKA: Na+/K+-ATPase Pump

NKA, first described in 1957 [28], is a ubiquitous enzyme that actively transports three Na+-ions through the cell membrane outside the cytoplasm in exchange for two K+ ions imported inside the cytoplasm [29,30]. NKA is a pump made up of two subunits: α and β. Subunit β does not seem to contain functional sites, but it is necessary to stabilize the α subunit and to guarantee the passage from the endoplasmic reticulum to the cell membrane [30]. The transport is accomplished by enzyme conformational changes between two states, E1 and E2. Using energy derived from ATP hydrolysis, NKA produces electrochemical gradients across the membrane necessary for electrical excitability as well as cellular uptake of ions, nutrients, and neurotransmitters. Electrochemical gradients across the membrane are necessary to regulate cell volume as well as intracellular pH [31].
NKA is a target of cardiotonic glycosides (CG), such as ouabain, digoxin, and digitoxin, which can bind NKA in the E2 state, inhibiting it [32,33,34]. This inhibition causes an increase in intracellular [Na+]. Consequently, the Na+/Ca2+ exchanger (NCX) ejects Na+ in exchange for Ca2+, thereby increasing the cytosolic concentration of Ca2+, activating excitation-contraction coupling, ultimately enhancing myocardial contractility [24,32,35,36].

3. SERCA2a: Sarcoendoplasmic Reticular Adenosine Triphosphate-Driven Ca2+ Pump

The family of the sarco/endoplasmic reticulum Ca2+ ATPase gene, also known as SERCA, counts different isoforms. The SERCA2 gene is known to codify for SERCA2a-d, whereas SERCA2a encodes for a 997-aminoacid protein, a specific isoform located in the cardiac muscle, slow-twitch skeletal muscle, and smooth muscle cells [37]. In cardiomyocytes, SERCA2a is able to generate an influx of Ca2+ from the cytosol into the sarcoplasmic reticulum (SR) against the gradient.
This process guarantees the relaxation of cardiac muscle fibers and a sufficient Ca2+ storage in the SR that can be utilized to start a new contractile activity for the ensuing contraction.
SERCA2a is inhibited by phospholamban (PLB) and this regulation depends on its state of phosphorylation mediated by protein kinase A (PKA): in its de-phosphorylated form, PLB inhibits SERCA2a, whereas in its phosphorylated form SERCA2a is released from such inhibition [38,39]. Stimulation of β-adrenergic receptors causes the phosphorylation of PLB by PKA, inactivating PLB and stimulating SERCA2a, improving myocardial contraction and relaxation (inotropic and lusitropic positive effects [40,41]). These processes explain why the dysregulation of SERCA2a and PLB interaction is intimately related to HF, along with the desensitization and downregulation of myocardial β-adrenergic receptors due to their chronic activation [42,43,44,45,46,47,48,49,50,51].

4. Ca2+ and SERCA2a Function in Cardiac Contractility

The contractile activity of the heart is due to the interaction of myosin and actin filaments, an interaction finely tuned by cytoplasmic Ca2+ levels [24]. Ca2+ fluxes in the myocardium are dysregulated in HF, producing a depression of myocardial contractility.
Ca2+ enters the cardiomyocyte during depolarization through L-type Ca2+ channels localized in the cell membrane in proximity of the SR. This Ca2+ influx induces a further release of Ca2+ from the SR in the cytoplasm, through intracellular Ca2+ release channels known as type 2 ryanodine receptors (RyR2) [52,53,54]. Thus, cytosolic [Ca2+] rises up to a critical concentration that activates the contractile system of the myocyte. To complete the contraction and start the relaxation phase, Ca2+ previously released by the SR needs to be re-uptaken. This step is possible thanks to SERCA2a, which brings intracytoplasmic Ca2+ into the SR against a concentration gradient [55,56,57].
Dysregulation of Ca2+ fluxes could be caused by alterations of RyR2 and/or loss of function of SERCA2a. Altered RyR2 generates an inappropriate diastolic release of Ca2+ from the SR, known as a Ca2+ leak; this diastolic leak reduces Ca2+ content in the SR and consequently, Ca2+ is available for the subsequent myocardial contraction [58,59,60,61].
Even the loss of function of SERCA2a reduces the availability of Ca2+ for the next contraction because of the inability of this pump to reuptake a sufficient amount of the ion. Moreover, SERCA2a loss of function compromises ventricular relaxation and causes diastolic dysfunction, reducing the quantity and speed of Ca2+ re-uptake from the cytosol [58,62].

5. Istaroxime

Istaroxime -(E,Z)-3-[(2-aminoethoxy)imino] androstane-6,17-dione hydrochloride- is a compound derived from androstenedione, hence considered a cardiotonic steroid, developed for the treatment of AHF [63]. Its application is limited to acute intravenous therapy due to its short plasma half-life (less than 1 h) because of the extensive hepatic processing that generates a long-lasting metabolite named PST3093 (E,Z)-[(6-beta-hydroxy-17-oxoandrostan-3-ylidene)amino]oxyacetic acid [64]. Istaroxime is a first-in-class drug with both inotropic and lusitropic effects without vasodilator properties; on the contrary, one of its characteristics is to rise systolic blood pressure (SBP) [65]. As mentioned above, istaroxime properties are related to its dual mechanism of action: NKA inhibition and SERCA2a stimulation. Istaroxime inhibits NKA by binding its E2 state from the extracellular side [66] (Figure 1).
Its inhibition of NKA is similar to that of digoxin [67] and leads to an increased [Ca2+], obtaining an inotropic effect. At nanomolar concentrations, istaroxime stimulates SERCA2a activity and Ca2+ uptake through a direct interaction with the SERCA2a/PLB complex, independent of cAMP/PKA and PLB phosphorylation [68] (Figure 2).
Henceforth, an influx of Ca2+ into the SR is generated during the diastolic phase, promoting myocardial relaxation, with a lusitropic effect [48,69,70] without inducing spontaneous Ca2+ efflux from the SR [68,71]. The increase in Ca2+ reserves through SERCA2a into the SR contributes to augment Ca2+ ions available for the following cardiac cycle, contributing to the inotropic effect of the compound [58]. The peculiar combination of istaroxime targets (Figure 3) seems to confer a better safety profile compared to digoxin; for instance, treatment with istaroxime has been associated with a lower risk of arrhythmias [72].

Other Therapeutical Applications of Istaroxime

Istaroxime has also been shown to have anti-cancer actions due to its antiproliferative capacity exhibited in tumor cell lines [73]. Its use could be taken into account in prostate cancer [74]. Indeed, the use of androgen deprivation therapy in prostate cancer can be limited by the onset of cardiovascular events, including HF. In this sense, istaroxime could combine its antineoplastic and inotropic actions [75]; remarkably, the antiproliferative capacity of istaroxime is shared with other cardiotonic steroids, so these compounds are being proposed as new antitumoral drugs [34,76].

6. Preclinical Studies

In an animal model of diabetic cardiomyopathy, istaroxime was shown to improve diastolic dysfunction (DD) stimulating SERCA2a and reducing alterations in intracellular Ca2+ handling [77]; equally important, in animal models of acute decompensated HF it significantly improved hemodynamic and echocardiographic parameters [78]. Istaroxime was also compared to dobutamine in a preclinical study investigating chronic ischemic HF, showing to be an effective inotropic agent without positive chronotropic actions [79]. The chronic use of istaroxime was tested in a hamster model of progressive HF and was demonstrated to improve cardiac function and heart rate variability [80].
Comparing the electrophysiological effects of istaroxime and digoxin in guinea pig ventricular myocytes, istaroxime was shown to inhibit (−43%) the transient inward current (ITI) induced by the transient flux of Ca2+ in the presence of a complete block of the NKA pump; this effect was not evident with digoxin. Therefore, the therapeutic index of istaroxime may be accounted for by inhibition of ITI, a current directly involved in digitalis-induced arrhythmias [81,82,83,84]. Furthermore, the toxicity of istaroxime was compared to ouabain in murine cells, showing that istaroxime can reach a significant inotropic effect without activating Ca2+/calmodulin-dependent kinase II (CaMKII) whose over-stimulation could lead to cardiomyocyte death. Interestingly, at its inotropic concentration, istaroxime does not evoke a significant increase in diastolic local (Ca2+ sparks) or propagated (Ca2+ waves) SR Ca2+ release. In fact, istaroxime breaks arrhythmogenic Ca2+ waves into mini waves, which are less arrhythmogenic. This aspect corroborates the substantial safety compared to digitalis compounds. Remarkably, these results were evident even in PLB-knockout myocytes, suggesting that the safety of istaroxime is not merely related to its capacity to dissociate the SERCA/PLB interaction [85]. The main preclinical studies are summarized in Table 1.

7. Clinical Investigations

The first evaluation of istaroxime in humans was a phase I-II dose escalating study evaluating the safety and tolerability of istaroxime and its specific effects on ECG and hemodynamic parameters in patients with chronic HF with reduced systolic function [86]. Three cohorts of six patients were exposed to four sequentially increasing (0.005–5.0 μg/kg/min) 1-h infusions. In addition to safety, hemodynamic parameters were evaluated by an impedance cardiography, a digital Holter recorder, and by electrocardiography. Enhanced contractility was demonstrated with evidence of improvement of the acceleration index. Istaroxime improved the left cardiac work index, cardiac index, and pulse pressure at doses of ≥1 μg/kg/min, with evidence of activity at doses of 0.5 μg/kg/min. Istaroxime also shortened the QTc. The compound was tolerated at doses of up to 3.33 μg/kg/min with evidence of gastrointestinal symptoms and injection site pain at higher doses [86].
The HORIZON-HF (Hemodynamic, Echocardiographic, and Neurohormonal Effects of Istaroxime, a Novel Intravenous Inotropic and Lusitropic Agent) trial is a randomized, double-blind, placebo-controlled, dose-escalation trial (NCT00616161) in which patients were randomized to istaroxime or placebo (ratio 3:1) [87]; patients treated with istaroxime were randomized to 0.5 µg/kg/min, 1.0 µg/kg/min, or 1.5 µg/kg/min doses [88]. The primary endpoint was the change in pulmonary capillary wedge pressure (PCWP) compared with placebo after a 6 h continuous infusion [87]. All the three doses of istaroxime lowered PCWP (mean ± SD: −3.2 ± 6.8 mm Hg, −3.3 ± 5.5 mm Hg, and −4.7 ± 5.9 mm Hg compared to 0.0 ± 3.6 mm Hg with placebo; p < 0.05 for all doses); furthermore, secondary end points (changes in cardiac index, right atrial pressure, SBP, diastolic blood pressure (DBP), heart rate (HR), and stroke work index) improved. In addition, changes in left ventricular (LV) end-diastolic and systolic volumes, LV ejection fraction (EF), diastolic function indexes, neurohormones, renal function, troponin, pharmacokinetics, and safety were evaluated. A significant decrease in HR (p = 0.008, 0.02, and 0.006, with 0.5, 1.0, and 1.5 μg/kg/min, respectively) and an increased SBP (p = 0.005 and p < 0.001, with 1 and 1.5 μg/kg/min, respectively) were observed in patients treated with istaroxime. The cardiac index increased during the 1.5 μg/kg/min infusion vs placebo (p = 0.04), but not at the end of the 6 h infusion. At the ultrasound examination (echocardiography), the LV end-systolic volume was reduced in the 1.0 μg/kg/min istaroxime group compared to placebo (−15.8 ± 22.7 mL vs. −2.1 ± 25.5 mL; p = 0.03), and LV end-diastolic volume was reduced in the 1.5 μg/kg/min group compared to placebo (−14.1 ± 26.3 mL vs. +3.9 ± 32.4 mL; p = 0.02). E-wave deceleration time increased in the 1.5 μg/kg/min group (+30 ± 51 ms vs. +3 ± 51 ms; p = 0.04). There were no changes in neurohormones, renal function, or troponin I. Adverse events were dose-related gastrointestinal symptoms and injection site pain. In particular, vomiting and nausea occurred and were dose-related. No deaths occurred during the treatment period [89].
A recent phase II, multicenter, randomized, double-blind, placebo-controlled, parallel group investigation trial (NCT02617446) revealed that a 24 h infusion of istaroxime at 0.5 or 1.0 μg/kg/min in patients with AHF and reduced LVEF markedly improves LV diastolic and systolic function without major cardiac adverse effects [90]. The primary endpoint was the E/e’ ratio change from baseline to 24 h that was significantly reduced by both doses of istaroxime compared to placebo (cohort 1: −4.55 ± 4.75 istaroxime 0.5 μg/kg/min vs. −1.55 ± 4.11 placebo, p = 0.029; cohort 2: −3.16 ± 2.59 istaroxime 1.0 μg/kg/min vs. −1.08 ± 2.72 placebo, p = 0.009). Moreover, other parameters including E/A ratio, left atrial dimensions, and inferior cava diameter improved. Among others, there was a decrease in HR (from baseline by about 3 bpm with istaroxime 0.5 μg/kg/min and by 8–9 bpm with istaroxime 1.0 μg/kg/min with significant changes vs. placebo at 3 to 24 h in the high-dose group; 3 h: −10.61 ± 10.04, p < 0.001; 6 h: −8.89 ± 9.83, p = 0.001; 12 h: −9.49 ± 11.96, p = 0.005; 24 h: −9.61 ± 12.10, p = 0.004) and a significant increase in SBP with the 1.0 μg/kg/min dose (from baseline by about 3 mmHg with istaroxime 0.5 μg/kg/min and by 6–8 mmHg with istaroxime 1.0 μg/kg/min reaching statistical significance compared to placebo at 3 to 12 h in istaroxime 1.0 μg/kg/min group; 3 h: 7.63 ± 9.22, p < 0.001; 6 h: 8.08 ± 11.06, p = 0.001; 12 h: 9.00 ± 11.75, p = 0.006). These results occurred early after starting the infusion, with a significant difference already evident at 3h, most likely due to the pharmacokinetic profile of istaroxime, characterized by a rapid onset of action and a rapid washout after infusion termination. An increased estimated glomerular filtration rate (eGFR) was observed in the group treated with the high dose of the drug compared to placebo. Self-reported dyspnea and N-terminal pro-brain natriuretic peptide improved in all groups without significant differences between istaroxime and placebo [90].
Also in this trial, the most common adverse events were injection site reactions and gastrointestinal events, the latter primarily with istaroxime 1.0 μg/kg/min. A higher rate of patients experiencing abdominal pain, nausea, and vomiting was observed with istaroxime 1.0 μg/kg/min. At this dose, a high rate of injection site reaction was observed in patients with short intravenous catheters, leading to a necessary change of the peripheral line or use of peripheral long line or a central line. One case of treatment discontinuation due to injection site problems occurred. Neither major cardiovascular events nor an increase in arrhythmias occurred. These observations confirm the findings of the HORIZON-HF trial and further extend the safety of istaroxime to 24 h compared to the HORIZON-HF trial, in which the compound was tested only for 6 h [88,90].
The clinical effects of istaroxime appear to be favorable for patients with AHF with low or borderline SBP, for which inotropes have been associated with major cardiovascular adverse events and possibly an increased risk of mortality [91,92]. The safety and efficacy of istaroxime in patients with AHF-related pre-cardiogenic shock (stage B of the Classification of the Society for Cardiovascular Angiography and Interventions, SCAI) was tested in the SEISMiC trial (NCT04325035), a phase II, multicenter, randomized, double-blind, placebo-controlled, parallel group study [93]. The SEISMiC study was designed to compare the safety and efficacy of istaroxime with placebo in patients hospitalized for AHF-related SCAI stage B pre-cardiogenic shock with persistent hypotension (75 < SBP < 90 mmHg) but no clinical signs of hypoperfusion (both clinically and as evidenced by venous lactate levels <2.0 mml/L). Patients were randomized to continuous 24 h infusion of istaroxime 1.0 µg/kg/min or 1.5 µg/kg/min or placebo. The primary endpoint was the adjusted area under the curve (AUC) representing the change in SBP from baseline, start of study drug infusion, through 6 h. Secondary endpoints included: SBP AUC through 24 h, changes from baseline in SBP, in DBP, and mean arterial pressure (MAP) at 6 and 24 h; changes from baseline in HR, treatment failure score (treatment failure defined as death or need for circulatory, respiratory, or renal mechanical support or need for intravenous inotrope or vasopressor treatment); increase from baseline in SBP ≥ 5% and/or ≥ 10 mmHg; changes in quality of life (measured by the EuroQol 5 Dimension 5 Level, EQ-5D-5L), change from baseline to 24 h in echocardiography parameters. Secondary endpoints were also considered: changes in troponin and N-terminal pro-B-type natriuretic peptide (NT-proBNP), hospital readmission for HF and for any cause by day 30, in-hospital worsening HF to day 5, and length of in-hospital stay. Endpoints of safety were: incidence of adverse events, changes in vital signs, change in 12-lead ECG, incidence of supraventricular and ventricular arrhythmias, changes in laboratory parameters, renal function, cardiac troponin (I and T), and mortality through day 30. In the SEISMiC trial, istaroxime increased SBP and improved echocardiographic measures, including an increase in the cardiac index and a reduction in the left atrial and left ventricular dimensions. The adjusted mean 6 h AUC was 53.1 (standard error [SE] 6.88) mmHg × hour in the istaroxime group vs. 30.9 (SE 6.76) mmHg × hour in the placebo group (p = 0.017); an increase of 72% was observed. The adjusted mean 24 h SBP AUC was 291.2 (SE 27.5) mmHg × hour in istaroxime arm vs 208.7 (SE 27.0) mmHg × hour in placebo arm (p = 0.025), with an increase of 40%. The adjusted SBP increase at 6 h was 12.3 (SE 1.71) mmHg in the istaroxime group vs. 7.5 (SE 1.64) mmHg in the placebo group (p = 0.045). The corresponding adjusted changes in SBP at 24 h were 17.1 (SE 2.36) mmHg and 15.1 (SE 2.25) mmHg in the istaroxime group vs. placebo (p = 0.543). Increases were noted in DBP and MAP, as well. Of note, the concomitant increase in both the cardiac index (at 24 h: +0.16 ± 0.1 vs. −0.06 ± 0.1 L/min/m2; p = 0.016) and SBP had not been observed with any previous intravenous drugs administered to patients with cardiogenic shock related to AHF. Additionally, other echocardiographic measures besides the cardiac index that demonstrated improvements at 24 h in the istaroxime group as compared to the placebo group were: left atrial area (−1.8 ± 0.5 vs. 0.0±0.5 cm2; p = 0.008), LV end-systolic volume (−8.7 ± 4.2 vs. 3.3 ± 4.2 mL; p = 0.034), and LV end-diastolic volume (−6.5 ± 4.9 vs. 5.6 ± 4.8 mL; p = 0.061). Laboratory parameters did not suggest an effect mediated by istaroxime on end-organ damage. Istaroxime treatment was associated with more adverse events (such as nausea, vomiting, and pain at infusion site) than placebo, but it was not associated with arrhythmias or worsening of renal function. Among gastrointestinal adverse events, nausea was the most frequent (28%), followed by vomiting (14%). Injection site pain occurred in 14% of the patients [93].
In order to reduce gastrointestinal adverse effects and injection site pain related to istaroxime, several attempts have been made, including the development of a liposomal formulation of the molecule, encapsulating istaroxime in a drug delivery system conveniently designed to be quickly destabilized in plasma in order to minimize alterations of the pharmacokinetic profile of istaroxime. Poly ethylene glycol 660-hydroxystearate (PEG-HS) was chosen as an excipient to modulate the bilayer fluidity and the release properties of the liposomes, obtaining an almost complete release in physiological conditions in less than 10 min [94]. It is important to emphasize that istaroxime use may also be considered in pediatric patients [95].
In summary, istaroxime was shown to be a potential new inotropic agent, safer than currently available treatments for AHF. Its ability to improve overall cardiac function in HF with reduced arrhythmogenic risk launched a new field of investigation in AHF treatment, which has most recently led to the development of other molecules with highly selective SERCA2a activation and longer half-time starting from istaroxime long-lasting metabolite PST3093 [64,96]. The main clinical trials investigating istaroxime are summarized in Table 2.
Intriguingly, therapies based on cardiac contractility modulation (CCM) take advantage of SERCA2a upregulation in patients with chronic HF. CCM therapy is based on an implantable device that delivers a non-excitatory high-voltage bipolar signal to the right ventricle (RV) synchronized on the absolute refractory period of the action potential [97,98]. It stimulates SERCA and RyR upregulation, PLB phosphorylation, and downregulation of NCX. Its modulation of Ca2+ flux and increase of re-uptake of the ion into SR results in the increase of myocardial contractility [97]. CCM has been approved by the FDA based on the results of several randomized clinical trials [99,100,101,102,103,104], which revealed that CCM improves the signs and symptoms of HF, particularly in patients with a LVEF between 25% and 45%, NYHA III symptoms despite guideline-directed medical therapy, and a sinus rhythm with normal QRS length. CCM therapy is associated with a reduction in hospitalization for HF compared with the rate of hospitalization the year before the device implantation [105,106]. Furthermore, a recent study revealed that CCM improves LVEF, global longitudinal strain, and myocardial mechano-energetic efficiency in patients with HFrEF [107]. These pieces of evidence suggest that the production of a metabolite derived from istaroxime, harboring a selective SERCA2a action and a longer half-time, could overcame the pharmacodynamic limitations of istaroxime itself, making it a new effective pharmacologic tool not only for AHF but for chronic HF as well [108].

8. Future Perspectives

HF is a clinical syndrome with considerable medical implications in our society, both in terms of morbidity and mortality [109,110,111]. Specifically, de novo AHF and acute decompensated HF (ADHF) represent life-threatening conditions [112,113,114,115]. According to recent evidence examining its effectiveness and safety, it is reasonable to consider istaroxime as the first example of a new useful and safe category of drugs against HF in contraposition with traditional inotropic agents whose uses are often limited by several adverse effects. The development of new molecules derived from istaroxime with longer lifetime and less adverse effects is ongoing and current results are quite promising. Nonetheless, more investigations are warranted.

9. Conclusions

Istaroxime is a cardiotonic steroid currently under investigation for AHF treatment. Compared to the inotropic agents presently used for AHF in clinical practice with which it shares inotropic actions, it also has a lusitropic positive effect and a better safety profile. Its properties are exerted through a dual mechanism of action: activation of SERCA2a and inhibition of NKA activity. Istaroxime is being tested exclusively for acute intravenous therapy due to its half-time of only 1 h because of its rapid hepatic metabolism. Available data indicate that istaroxime has an overall safe profile with a reduced arrhythmogenic risk. Adverse events include gastrointestinal discomfort, most likely attributable to the systemic inhibition of NKA, and pain at the injection site. Hitherto, only studies on animal models and phase I and II trials are available; therefore, albeit there are promising perspectives for istaroxime as a new inotropic agent for AHF treatment, it is necessary to further expand the investigation on this compound to assess its effectiveness and long-term safety.

Author Contributions

Conceptualization, I.F. and G.S.; methodology and formal analysis, I.F., P.M., G.M., U.K., L.S., A.D.L., T.T. and F.V.; writing—original draft preparation, I.F.; writing—review and editing, G.S. All authors have read and agreed to the published version of the manuscript.

Funding

The Santulli’s Lab is supported in part by the National Institutes of Health (NIH): National Heart, Lung, and Blood Institute (NHLBI: R01-HL159062, R01HL164772, R01-HL146691, T32-HL144456), National Institute of Diabetes and Digestive and Kidney Diseases (NIDDK: R01-DK123259, R01-DK033823), National Center for Advancing Translational Sciences (NCATS: UL1TR002556–06) to G.S., by the Diabetes Action Research and Education Foundation (to G.S.), and by the Monique Weill-Caulier and Irma T. Hirschl Trusts (to G.S). F.V. and U.K. are supported in part by postdoctoral fellowships of the American Heart Association (AHA-22POST995561 and AHA-23POST1026190, respectively).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Stanislovas S. Jankauskas for the helpful discussion.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Santulli, G.; Wang, X.; Mone, P. Updated ACC/AHA/HFSA 2022 guidelines on heart failure: What is new? From epidemiology to clinical management. Eur. Heart J. Cardiovasc. Pharmacother. 2022, 8, e23–e24. [Google Scholar] [CrossRef]
  2. Savarese, G.; Becher, P.M.; Lund, L.H.; Seferovic, P.; Rosano, G.M.C.; Coats, A. Global burden of heart failure: A com-prehensive and updated review of epidemiology. Cardiovasc. Res. 2022. Online ahead of print. [Google Scholar] [CrossRef] [PubMed]
  3. Jenca, D.; Melenovsky, V.; Stehlik, J.; Stanek, V.; Kettner, J.; Kautzner, J.; Adamkova, V.; Wohlfahrt, P. Heart failure after myocardial infarction: Incidence and predictors. ESC Heart Fail. 2021, 8, 222–237. [Google Scholar] [CrossRef]
  4. Huffman, M.D.; Berry, J.D.; Ning, H.; Dyer, A.R.; Garside, D.B.; Cai, X.; Daviglus, M.L.; Lloyd-Jones, D.M. Lifetime risk for heart failure among white and black Americans: Cardiovascular lifetime risk pooling project. J. Am. Coll. Cardiol. 2013, 61, 1510–1517. [Google Scholar] [CrossRef] [Green Version]
  5. Emmons-Bell, S.; Johnson, C.; Roth, G. Prevalence, incidence and survival of heart failure: A systematic review. Heart 2022, 108, 1351–1360. [Google Scholar] [CrossRef]
  6. Virani, S.S.; Alonso, A.; Aparicio, H.J.; Benjamin, E.J.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Cheng, S.; Delling, F.N.; et al. Heart Disease and Stroke Statistics-2021 Update: A Report From the American Heart Association. Circulation 2021, 143, e254–e743. [Google Scholar] [CrossRef]
  7. Heidenreich, P.A.; Albert, N.M.; Allen, L.A.; Bluemke, D.A.; Butler, J.; Fonarow, G.C.; Ikonomidis, J.S.; Khavjou, O.; Konstam, M.A.; Maddox, T.M.; et al. Forecasting the impact of heart failure in the United States: A policy statement from the American Heart Association. Circ. Heart Fail. 2013, 6, 606–619. [Google Scholar] [CrossRef] [Green Version]
  8. Masip, J.; Frank Peacok, W.; Arrigo, M.; Rossello, X.; Platz, E.; Cullen, L.; Mebazaa, A.; Price, S.; Bueno, H.; Di Somma, S.; et al. Acute Heart Failure in the 2021 ESC Heart Failure Guidelines: A scientific statement from the Association for Acute CardioVascular Care (ACVC) of the European Society of Cardiology. Eur. Heart J. Acute Cardiovasc. Care 2022, 11, 173–185. [Google Scholar] [CrossRef]
  9. Heidenreich, P.A.; Bozkurt, B.; Aguilar, D.; Allen, L.A.; Byun, J.J.; Colvin, M.M.; Deswal, A.; Drazner, M.H.; Dunlay, S.M.; Evers, L.R.; et al. 2022 AHA/ACC/HFSA Guideline for the Management of Heart Failure: A Report of the American College of Cardiology/American Heart Association Joint Committee on Clinical Practice Guidelines. Circulation 2022, 145, e895–e1032. [Google Scholar] [CrossRef]
  10. Kurmani, S.; Squire, I. Acute Heart Failure: Definition, Classification and Epidemiology. Curr. Heart Fail. Rep. 2017, 14, 385–392. [Google Scholar] [CrossRef]
  11. Nieminen, M.S.; Harjola, V.P. Definition and epidemiology of acute heart failure syndromes. Am. J. Cardiol. 2005, 96, 5G–10G. [Google Scholar] [CrossRef] [PubMed]
  12. Gheorghiade, M.; Zannad, F.; Sopko, G.; Klein, L.; Pina, I.L.; Konstam, M.A.; Massie, B.M.; Roland, E.; Targum, S.; Collins, S.P.; et al. Acute heart failure syndromes: Current state and framework for future research. Circulation 2005, 112, 3958–3968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. ter Maaten, J.M.; Valente, M.A.; Damman, K.; Hillege, H.L.; Navis, G.; Voors, A.A. Diuretic response in acute heart failure-pathophysiology, evaluation, and therapy. Nat. Rev. Cardiol. 2015, 12, 184–192. [Google Scholar] [CrossRef] [PubMed]
  14. Masip, J.; Peacock, W.F.; Price, S.; Cullen, L.; Martin-Sanchez, F.J.; Seferovic, P.; Maisel, A.S.; Miro, O.; Filippatos, G.; Vrints, C.; et al. Indications and practical approach to non-invasive ventilation in acute heart failure. Eur. Heart J. 2018, 39, 17–25. [Google Scholar] [CrossRef] [Green Version]
  15. Schumann, J.; Henrich, E.C.; Strobl, H.; Prondzinsky, R.; Weiche, S.; Thiele, H.; Werdan, K.; Frantz, S.; Unverzagt, S. Inotropic agents and vasodilator strategies for the treatment of cardiogenic shock or low cardiac output syndrome. Cochrane Database Syst. Rev. 2018, 1, CD009669. [Google Scholar] [CrossRef] [PubMed]
  16. Choi, M.S.; Shim, H.; Cho, Y.H. Mechanical Circulatory Support for Acute Heart Failure Complicated by Cardiogenic Shock. Int. J. Heart Fail. 2020, 2, 23–44. [Google Scholar] [CrossRef] [Green Version]
  17. Mebazaa, A.; Tolppanen, H.; Mueller, C.; Lassus, J.; DiSomma, S.; Baksyte, G.; Cecconi, M.; Choi, D.J.; Cohen Solal, A.; Christ, M.; et al. Acute heart failure and cardiogenic shock: A multidisciplinary practical guidance. Intensive Care Med. 2016, 42, 147–163. [Google Scholar] [CrossRef]
  18. Thiele, H.; Zeymer, U.; Neumann, F.J.; Ferenc, M.; Olbrich, H.G.; Hausleiter, J.; Richardt, G.; Hennersdorf, M.; Empen, K.; Fuernau, G.; et al. Intraaortic balloon support for myocardial infarction with cardiogenic shock. N. Engl. J. Med. 2012, 367, 1287–1296. [Google Scholar] [CrossRef] [Green Version]
  19. Abraham, W.T.; Adams, K.F.; Fonarow, G.C.; Costanzo, M.R.; Berkowitz, R.L.; LeJemtel, T.H.; Cheng, M.L.; Wynne, J.; Committee, A.S.A.; ADHERE Scientific Advisory Committee and Investigators; et al. In-hospital mortality in patients with acute decompensated heart failure requiring intravenous vasoactive medications: An analysis from the Acute Decompensated Heart Failure National Registry (ADHERE). J. Am. Coll. Cardiol. 2005, 46, 57–64. [Google Scholar] [CrossRef] [Green Version]
  20. Nieminen, M.S.; Brutsaert, D.; Dickstein, K.; Drexler, H.; Follath, F.; Harjola, V.P.; Hochadel, M.; Komajda, M.; Lassus, J.; Lopez-Sendon, J.L.; et al. EuroHeart Failure Survey II (EHFS II): A survey on hospitalized acute heart failure patients: Description of population. Eur. Heart J. 2006, 27, 2725–2736. [Google Scholar] [CrossRef]
  21. Abraham, W.T.; Fonarow, G.C.; Albert, N.M.; Stough, W.G.; Gheorghiade, M.; Greenberg, B.H.; O’Connor, C.M.; Sun, J.L.; Yancy, C.W.; Young, J.B.; et al. Predictors of in-hospital mortality in patients hospitalized for heart failure: Insights from the Organized Program to Initiate Lifesaving Treatment in Hospitalized Patients with Heart Failure (OPTIMIZE-HF). J. Am. Coll. Cardiol. 2008, 52, 347–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. De Matteis, G.; Covino, M.; Burzo, M.L.; Della Polla, D.A.; Franceschi, F.; Mebazaa, A.; Gambassi, G. Clinical Characteristics and Predictors of In-Hospital Mortality among Older Patients with Acute Heart Failure. J. Clin. Med. 2022, 11, 439. [Google Scholar] [CrossRef] [PubMed]
  23. Spinar, J.; Parenica, J.; Vitovec, J.; Widimsky, P.; Linhart, A.; Fedorco, M.; Malek, F.; Cihalik, C.; Spinarova, L.; Miklik, R.; et al. Baseline characteristics and hospital mortality in the Acute Heart Failure Database (AHEAD) Main registry. Crit. Care 2011, 15, R291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Kansakar, U.; Varzideh, F.; Jankauskas, S.S.; Gambardella, J.; Trimarco, B.; Santulli, G. Advances in the understanding of excitation-contraction coupling: The pulsing quest for drugs against heart failure and arrhythmias. Eur. Heart J. Cardiovasc. Pharmacother. 2021, 7, e91–e93. [Google Scholar] [CrossRef] [PubMed]
  25. Psotka, M.A.; Gottlieb, S.S.; Francis, G.S.; Allen, L.A.; Teerlink, J.R.; Adams, K.F., Jr.; Rosano, G.M.C.; Lancellotti, P. Cardiac Calcitropes, Myotropes, and Mitotropes: JACC Review Topic of the Week. J. Am. Coll. Cardiol. 2019, 73, 2345–2353. [Google Scholar] [CrossRef]
  26. van Diepen, S.; Katz, J.N.; Albert, N.M.; Henry, T.D.; Jacobs, A.K.; Kapur, N.K.; Kilic, A.; Menon, V.; Ohman, E.M.; Sweitzer, N.K.; et al. Contemporary Management of Cardiogenic Shock: A Scientific Statement From the American Heart Association. Circulation 2017, 136, e232–e268. [Google Scholar] [CrossRef]
  27. Tavares, M.; Rezlan, E.; Vostroknoutova, I.; Khouadja, H.; Mebazaa, A. New pharmacologic therapies for acute heart failure. Crit. Care Med. 2008, 36, S112–S120. [Google Scholar] [CrossRef]
  28. Skou, J.C. The influence of some cations on an adenosine triphosphatase from peripheral nerves. Biochim. Biophys. Acta 1957, 23, 394–401. [Google Scholar] [CrossRef]
  29. Manoj, K.M.; Gideon, D.A.; Bazhin, N.M.; Tamagawa, H.; Nirusimhan, V.; Kavdia, M.; Jaeken, L. Na,K-ATPase: A murzyme facilitating thermodynamic equilibriums at the membrane-interface. J. Cell Physiol. 2022, in press. [Google Scholar] [CrossRef]
  30. Fedosova, N.U.; Habeck, M.; Nissen, P. Structure and Function of Na,K-ATPase-The Sodium-Potassium Pump. Compr. Physiol. 2021, 12, 2659–2679. [Google Scholar] [CrossRef]
  31. Morth, J.P.; Pedersen, B.P.; Toustrup-Jensen, M.S.; Sorensen, T.L.; Petersen, J.; Andersen, J.P.; Vilsen, B.; Nissen, P. Crystal structure of the sodium-potassium pump. Nature 2007, 450, 1043–1049. [Google Scholar] [CrossRef]
  32. Kanai, R.; Cornelius, F.; Ogawa, H.; Motoyama, K.; Vilsen, B.; Toyoshima, C. Binding of cardiotonic steroids to Na(+),K(+)-ATPase in the E2P state. Proc. Natl. Acad. Sci. USA 2021, 118, e2020438118. [Google Scholar] [CrossRef]
  33. Kravtsova, V.V.; Bouzinova, E.V.; Matchkov, V.V.; Krivoi, I.I. Skeletal Muscle Na,K-ATPase as a Target for Circulating Ouabain. Int. J. Mol. Sci. 2020, 21, 2875. [Google Scholar] [CrossRef] [Green Version]
  34. Triana-Martinez, F.; Picallos-Rabina, P.; Da Silva-Alvarez, S.; Pietrocola, F.; Llanos, S.; Rodilla, V.; Soprano, E.; Pedrosa, P.; Ferreiros, A.; Barradas, M.; et al. Identification and characterization of Cardiac Glycosides as senolytic compounds. Nat. Commun. 2019, 10, 4731. [Google Scholar] [CrossRef] [Green Version]
  35. Altamirano, J.; Li, Y.; DeSantiago, J.; Piacentino, V., 3rd; Houser, S.R.; Bers, D.M. The inotropic effect of cardioactive glycosides in ventricular myocytes requires Na+-Ca2+ exchanger function. J. Physiol. 2006, 575, 845–854. [Google Scholar] [CrossRef]
  36. Shattock, M.J.; Ottolia, M.; Bers, D.M.; Blaustein, M.P.; Boguslavskyi, A.; Bossuyt, J.; Bridge, J.H.; Chen-Izu, Y.; Clancy, C.E.; Edwards, A.; et al. Na+/Ca2+ exchange and Na+/K+-ATPase in the heart. J. Physiol. 2015, 593, 1361–1382. [Google Scholar] [CrossRef] [Green Version]
  37. Primeau, J.O.; Armanious, G.P.; Fisher, M.E.; Young, H.S. The SarcoEndoplasmic Reticulum Calcium ATPase. Subcell. Biochem. 2018, 87, 229–258. [Google Scholar] [CrossRef]
  38. Tada, M.; Kadoma, M.; Fujii, J.; Kimura, Y.; Kijima, Y. Molecular structure and function of phospholamban: The regulatory protein of calcium pump in cardiac sarcoplasmic reticulum. Adv. Exp. Med. Biol. 1989, 255, 79–89. [Google Scholar] [CrossRef]
  39. Gruber, S.J.; Haydon, S.; Thomas, D.D. Phospholamban mutants compete with wild type for SERCA binding in living cells. Biochem. Biophys. Res. Commun. 2012, 420, 236–240. [Google Scholar] [CrossRef] [Green Version]
  40. Arnold, M.E.; Dostmann, W.R.; Martin, J.; Previs, M.J.; Palmer, B.; LeWinter, M.; Meyer, M. SERCA2a-phospholamban interaction monitored by an interposed circularly permutated green fluorescent protein. Am. J. Physiol. Heart Circ. Physiol. 2021, 320, H2188–H2200. [Google Scholar] [CrossRef]
  41. Liu, Y.; Chen, J.; Fontes, S.K.; Bautista, E.N.; Cheng, Z. Physiological and pathological roles of protein kinase A in the heart. Cardiovasc. Res. 2022, 118, 386–398. [Google Scholar] [CrossRef] [PubMed]
  42. Izzo, R.; Cipolletta, E.; Ciccarelli, M.; Campanile, A.; Santulli, G.; Palumbo, G.; Vasta, A.; Formisano, S.; Trimarco, B.; Iaccarino, G. Enhanced GRK2 expression and desensitization of betaAR vasodilatation in hypertensive patients. Clin. Transl. Sci. 2008, 1, 215–220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Santulli, G.; Campanile, A.; Spinelli, L.; Assante di Panzillo, E.; Ciccarelli, M.; Trimarco, B.; Iaccarino, G. G protein-coupled receptor kinase 2 in patients with acute myocardial infarction. Am. J. Cardiol. 2011, 107, 1125–1130. [Google Scholar] [CrossRef]
  44. Sorriento, D.; Ciccarelli, M.; Santulli, G.; Illario, M.; Trimarco, B.; Iaccarino, G. Trafficking GRK2: Cellular and Metabolic consequences of GRK2 subcellular localization. Transl. Med. UniSa 2014, 10, 3–7. [Google Scholar]
  45. Sorriento, D.; Santulli, G.; Franco, A.; Cipolletta, E.; Napolitano, L.; Gambardella, J.; Gomez-Monterrey, I.; Campiglia, P.; Trimarco, B.; Iaccarino, G.; et al. Integrating GRK2 and NFkappaB in the Pathophysiology of Cardiac Hypertrophy. J. Cardiovasc. Transl. Res. 2015, 8, 493–502. [Google Scholar] [CrossRef]
  46. Selby, D.E.; Palmer, B.M.; LeWinter, M.M.; Meyer, M. Tachycardia-induced diastolic dysfunction and resting tone in myocardium from patients with a normal ejection fraction. J. Am. Coll. Cardiol. 2011, 58, 147–154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Dhalla, N.S.; Panagia, V.; Singal, P.K.; Makino, N.; Dixon, I.M.; Eyolfson, D.A. Alterations in heart membrane calcium transport during the development of ischemia-reperfusion injury. J. Mol. Cell. Cardiol. 1988, 20 (Suppl. S2), 3–13. [Google Scholar] [CrossRef] [PubMed]
  48. Gambardella, J.; Wang, X.J.; Ferrara, J.; Morelli, M.B.; Santulli, G. Cardiac BIN1 Replacement Therapy Ameliorates Inotropy and Lusitropy in Heart Failure by Regulating Calcium Handling. JACC Basic Transl. Sci. 2020, 5, 579–581. [Google Scholar] [CrossRef]
  49. Santulli, G.; Iaccarino, G. Adrenergic signaling in heart failure and cardiovascular aging. Maturitas 2016, 93, 65–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Ciccarelli, M.; Santulli, G.; Pascale, V.; Trimarco, B.; Iaccarino, G. Adrenergic receptors and metabolism: Role in development of cardiovascular disease. Front. Physiol. 2013, 4, 265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Santulli, G. Adrenal signaling in heart failure: Something more than a distant ship’s smoke on the horizon. Hypertension 2014, 63, 215–216. [Google Scholar] [CrossRef] [PubMed]
  52. Gambardella, J.; Jankauskas, S.S.; D’Ascia, S.L.; Sardu, C.; Matarese, A.; Minicucci, F.; Mone, P.; Santulli, G. Glycation of ryanodine receptor in circulating lymphocytes predicts the response to cardiac resynchronization therapy. J. Heart Lung Transplant. 2022, 41, 438–441. [Google Scholar] [CrossRef] [PubMed]
  53. Jin, X.; Amoni, M.; Gilbert, G.; Dries, E.; Donate Puertas, R.; Tomar, A.; Nagaraju, C.K.; Pradhan, A.; Yule, D.I.; Martens, T.; et al. InsP(3)R-RyR Ca(2+) channel crosstalk facilitates arrhythmias in the failing human ventricle. Basic Res. Cardiol. 2022, 117, 60. [Google Scholar] [CrossRef] [PubMed]
  54. 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] [Green Version]
  55. Braunwald, E. Heart failure. JACC Heart Fail. 2013, 1, 1–20. [Google Scholar] [CrossRef]
  56. Maning, J.; Desimine, V.L.; Pollard, C.M.; Ghandour, J.; Lymperopoulos, A. Carvedilol Selectively Stimulates betaArrestin2-Dependent SERCA2a Activity in Cardiomyocytes to Augment Contractility. Int. J. Mol. Sci. 2022, 23, 11315. [Google Scholar] [CrossRef]
  57. Chung, Y.J.; Luo, A.; Park, K.C.; Loonat, A.A.; Lakhal-Littleton, S.; Robbins, P.A.; Swietach, P. Iron-deficiency anemia reduces cardiac contraction by downregulating RyR2 channels and suppressing SERCA pump activity. JCI Insight 2019, 4, e125618. [Google Scholar] [CrossRef] [Green Version]
  58. Gambardella, J.; Trimarco, B.; Iaccarino, G.; Santulli, G. New Insights in Cardiac Calcium Handling and Excitation-Contraction Coupling. Adv. Exp. Med. Biol. 2018, 1067, 373–385. [Google Scholar] [CrossRef] [Green Version]
  59. Kushnir, A.; Santulli, G.; Reiken, S.R.; Coromilas, E.; Godfrey, S.J.; Brunjes, D.L.; Colombo, P.C.; Yuzefpolskaya, M.; Sokol, S.I.; Kitsis, R.N.; et al. Ryanodine Receptor Calcium Leak in Circulating B-Lymphocytes as a Biomarker in Heart Failure. Circulation 2018, 138, 1144–1154. [Google Scholar] [CrossRef]
  60. Santulli, G.; Nakashima, R.; Yuan, Q.; Marks, A.R. Intracellular calcium release channels: An update. J. Physiol. 2017, 595, 3041–3051. [Google Scholar] [CrossRef] [Green Version]
  61. Sardu, C.; Santulli, G.; Guerra, G.; Trotta, M.C.; Santamaria, M.; Sacra, C.; Testa, N.; Ducceschi, V.; Gatta, G.; Amico, M.; et al. Modulation of SERCA in Patients with Persistent Atrial Fibrillation Treated by Epicardial Thoracoscopic Ablation: The CAMAF Study. J. Clin. Med. 2020, 9, 544. [Google Scholar] [CrossRef]
  62. Hoydal, M.A.; Kirkeby-Garstad, I.; Karevold, A.; Wiseth, R.; Haaverstad, R.; Wahba, A.; Stolen, T.L.; Contu, R.; Condorelli, G.; Ellingsen, O.; et al. Human cardiomyocyte calcium handling and transverse tubules in mid-stage of post-myocardial-infarction heart failure. ESC Heart Fail. 2018, 5, 332–342. [Google Scholar] [CrossRef] [Green Version]
  63. Rocchetti, M.; Alemanni, M.; Mostacciuolo, G.; Barassi, P.; Altomare, C.; Chisci, R.; Micheletti, R.; Ferrari, P.; Zaza, A. Modulation of sarcoplasmic reticulum function by PST2744 [istaroxime; (E,Z)-3-((2-aminoethoxy)imino) androstane-6,17-dione hydrochloride)] in a pressure-overload heart failure model. J. Pharmacol. Exp. Ther. 2008, 326, 957–965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Arici, M.; Ferrandi, M.; Barassi, P.; Hsu, S.C.; Torre, E.; Luraghi, A.; Ronchi, C.; Chang, G.J.; Peri, F.; Ferrari, P.; et al. Istaroxime metabolite PST3093 selectively stimulates SERCA2a and reverses disease-induced changes in cardiac function. J. Pharmacol. Exp. Ther. 2022, in press. [Google Scholar] [CrossRef] [PubMed]
  65. Khan, H.; Metra, M.; Blair, J.E.; Vogel, M.; Harinstein, M.E.; Filippatos, G.S.; Sabbah, H.N.; Porchet, H.; Valentini, G.; Gheorghiade, M. Istaroxime, a first in class new chemical entity exhibiting SERCA-2 activation and Na-K-ATPase inhibition: A new promising treatment for acute heart failure syndromes? Heart Fail. Rev. 2009, 14, 277–287. [Google Scholar] [CrossRef] [PubMed]
  66. Cornelius, F.; Mahmmoud, Y.A.; Toyoshima, C. Metal fluoride complexes of Na,K-ATPase: Characterization of fluoride-stabilized phosphoenzyme analogues and their interaction with cardiotonic steroids. J. Biol. Chem. 2011, 286, 29882–29892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Micheletti, R.; Mattera, G.G.; Rocchetti, M.; Schiavone, A.; Loi, M.F.; Zaza, A.; Gagnol, R.J.; De Munari, S.; Melloni, P.; Carminati, P.; et al. Pharmacological profile of the novel inotropic agent (E,Z)-3-((2-aminoethoxy)imino)androstane-6,17-dione hydrochloride (PST2744). J. Pharmacol. Exp. Ther. 2002, 303, 592–600. [Google Scholar] [CrossRef] [Green Version]
  68. Ferrandi, M.; Barassi, P.; Tadini-Buoninsegni, F.; Bartolommei, G.; Molinari, I.; Tripodi, M.G.; Reina, C.; Moncelli, M.R.; Bianchi, G.; Ferrari, P. Istaroxime stimulates SERCA2a and accelerates calcium cycling in heart failure by relieving phospholamban inhibition. Br. J. Pharmacol. 2013, 169, 1849–1861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Bossu, A.; Kostense, A.; Beekman, H.D.M.; Houtman, M.J.C.; van der Heyden, M.A.G.; Vos, M.A. Istaroxime, a positive inotropic agent devoid of proarrhythmic properties in sensitive chronic atrioventricular block dogs. Pharmacol. Res. 2018, 133, 132–140. [Google Scholar] [CrossRef] [PubMed]
  70. Schwinger, R.H.; Brixius, K.; Bavendiek, U.; Hoischen, S.; Muller-Ehmsen, J.; Bolck, B.; Erdmann, E. Effect of cyclopiazonic acid on the force-frequency relationship in human nonfailing myocardium. J. Pharmacol. Exp. Ther. 1997, 283, 286–292. [Google Scholar] [PubMed]
  71. Abrol, N.; de Tombe, P.P.; Robia, S.L. Acute inotropic and lusitropic effects of cardiomyopathic R9C mutation of phospholamban. J. Biol. Chem. 2015, 290, 7130–7140. [Google Scholar] [CrossRef] [PubMed]
  72. Alemanni, M.; Rocchetti, M.; Re, D.; Zaza, A. Role and mechanism of subcellular Ca2+ distribution in the action of two inotropic agents with different toxicity. J. Mol. Cell. Cardiol. 2011, 50, 910–918. [Google Scholar] [CrossRef] [PubMed]
  73. Alevizopoulos, K.; Dimas, K.; Papadopoulou, N.; Schmidt, E.M.; Tsapara, A.; Alkahtani, S.; Honisch, S.; Prousis, K.C.; Alarifi, S.; Calogeropoulou, T.; et al. Functional characterization and anti-cancer action of the clinical phase II cardiac Na+/K+ ATPase inhibitor istaroxime: In vitro and in vivo properties and cross talk with the membrane androgen receptor. Oncotarget 2016, 7, 24415–24428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Stagno, M.J.; Zacharopoulou, N.; Bochem, J.; Tsapara, A.; Pelzl, L.; Al-Maghout, T.; Kallergi, G.; Alkahtani, S.; Alevizopoulos, K.; Dimas, K.; et al. Istaroxime Inhibits Motility and Down-Regulates Orai1 Expression, SOCE and FAK Phosphorylation in Prostate Cancer Cells. Cell. Physiol. Biochem. 2017, 42, 1366–1376. [Google Scholar] [CrossRef] [Green Version]
  75. Wallner, M.; Khafaga, M.; Kolesnik, E.; Vafiadis, A.; Schwantzer, G.; Eaton, D.M.; Curcic, P.; Kostenberger, M.; Knez, I.; Rainer, P.P.; et al. Istaroxime, a potential anticancer drug in prostate cancer, exerts beneficial functional effects in healthy and diseased human myocardium. Oncotarget 2017, 8, 49264–49274. [Google Scholar] [CrossRef] [Green Version]
  76. Mijatovic, T.; Van Quaquebeke, E.; Delest, B.; Debeir, O.; Darro, F.; Kiss, R. Cardiotonic steroids on the road to anti-cancer therapy. Biochim. Biophys. Acta 2007, 1776, 32–57. [Google Scholar] [CrossRef]
  77. Torre, E.; Arici, M.; Lodrini, A.M.; Ferrandi, M.; Barassi, P.; Hsu, S.C.; Chang, G.J.; Boz, E.; Sala, E.; Vagni, S.; et al. SERCA2a stimulation by istaroxime improves intracellular Ca2+ handling and diastolic dysfunction in a model of diabetic cardiomyopathy. Cardiovasc. Res. 2022, 118, 1020–1032. [Google Scholar] [CrossRef]
  78. Sabbah, H.N.; Imai, M.; Cowart, D.; Amato, A.; Carminati, P.; Gheorghiade, M. Hemodynamic properties of a new-generation positive luso-inotropic agent for the acute treatment of advanced heart failure. Am. J. Cardiol. 2007, 99, 41A–46A. [Google Scholar] [CrossRef]
  79. Adamson, P.B.; Vanoli, E.; Mattera, G.G.; Germany, R.; Gagnol, J.P.; Carminati, P.; Schwartz, P.J. Hemodynamic effects of a new inotropic compound, PST-2744, in dogs with chronic ischemic heart failure. J. Cardiovasc. Pharmacol. 2003, 42, 169–173. [Google Scholar] [CrossRef]
  80. Lo Giudice, P.; Mattera, G.G.; Gagnol, J.P.; Borsini, F. Chronic istaroxime improves cardiac function and heart rate variability in cardiomyopathic hamsters. Cardiovasc. Drugs Ther. 2011, 25, 133–138. [Google Scholar] [CrossRef]
  81. Rocchetti, M.; Besana, A.; Mostacciuolo, G.; Ferrari, P.; Micheletti, R.; Zaza, A. Diverse toxicity associated with cardiac Na+/K+ pump inhibition: Evaluation of electrophysiological mechanisms. J. Pharmacol. Exp. Ther. 2003, 305, 765–771. [Google Scholar] [CrossRef] [PubMed]
  82. Tseng, G.N.; Wit, A.L. Characteristics of a transient inward current that causes delayed afterdepolarizations in atrial cells of the canine coronary sinus. J. Mol. Cell. Cardiol. 1987, 19, 1105–1119. [Google Scholar] [CrossRef] [PubMed]
  83. Hancox, J.C.; Levi, A.J. Actions of the digitalis analogue strophanthidin on action potentials and L-type calcium current in single cells isolated from the rabbit atrioventricular node. Br. J. Pharmacol. 1996, 118, 1447–1454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Holmes, M.; Hurley, M.E.; Sheard, T.M.D.; Benson, A.P.; Jayasinghe, I.; Colman, M.A. Increased SERCA2a sub-cellular heterogeneity in right-ventricular heart failure inhibits excitation-contraction coupling and modulates arrhythmogenic dynamics. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2022, 377, 20210317. [Google Scholar] [CrossRef]
  85. Racioppi, M.F.; Burgos, J.I.; Morell, M.; Gonano, L.A.; Vila Petroff, M. Cellular Mechanisms Underlying the Low Cardiotoxicity of Istaroxime. J. Am. Heart Assoc. 2021, 10, e018833. [Google Scholar] [CrossRef]
  86. Ghali, J.K.; Smith, W.B.; Torre-Amione, G.; Haynos, W.; Rayburn, B.K.; Amato, A.; Zhang, D.; Cowart, D.; Valentini, G.; Carminati, P.; et al. A phase 1-2 dose-escalating study evaluating the safety and tolerability of istaroxime and specific effects on electrocardiographic and hemodynamic parameters in patients with chronic heart failure with reduced systolic function. Am. J. Cardiol. 2007, 99, S47–S56. [Google Scholar] [CrossRef]
  87. Blair, J.E.; Macarie, C.; Ruzyllo, W.; Bacchieri, A.; Valentini, G.; Bianchetti, M.; Pang, P.S.; Harinstein, M.E.; Sabbah, H.N.; Filippatos, G.S.; et al. Rationale and design of the hemodynamic, echocardiographic and neurohormonal effects of istaroxime, a novel intravenous inotropic and lusitropic agent: A randomized controlled trial in patients hospitalized with heart failure (HORIZON-HF) trial. Am. J. Ther. 2008, 15, 231–240. [Google Scholar] [CrossRef]
  88. Shah, S.J.; Blair, J.E.; Filippatos, G.S.; Macarie, C.; Ruzyllo, W.; Korewicki, J.; Bubenek-Turconi, S.I.; Ceracchi, M.; Bianchetti, M.; Carminati, P.; et al. Effects of istaroxime on diastolic stiffness in acute heart failure syndromes: Results from the Hemodynamic, Echocardiographic, and Neurohormonal Effects of Istaroxime, a Novel Intravenous Inotropic and Lusitropic Agent: A Randomized Controlled Trial in Patients Hospitalized with Heart Failure (HORIZON-HF) trial. Am. Heart J. 2009, 157, 1035–1041. [Google Scholar] [CrossRef]
  89. Gheorghiade, M.; Blair, J.E.; Filippatos, G.S.; Macarie, C.; Ruzyllo, W.; Korewicki, J.; Bubenek-Turconi, S.I.; Ceracchi, M.; Bianchetti, M.; Carminati, P.; et al. Hemodynamic, echocardiographic, and neurohormonal effects of istaroxime, a novel intravenous inotropic and lusitropic agent: A randomized controlled trial in patients hospitalized with heart failure. J. Am. Coll. Cardiol. 2008, 51, 2276–2285. [Google Scholar] [CrossRef] [Green Version]
  90. Carubelli, V.; Zhang, Y.; Metra, M.; Lombardi, C.; Felker, G.M.; Filippatos, G.; O’Connor, C.M.; Teerlink, J.R.; Simmons, P.; Segal, R.; et al. Treatment with 24 hour istaroxime infusion in patients hospitalised for acute heart failure: A randomised, placebo-controlled trial. Eur. J. Heart Fail. 2020, 22, 1684–1693. [Google Scholar] [CrossRef]
  91. Thomas, S.S.; Nohria, A. Hemodynamic classifications of acute heart failure and their clinical application:—An update. Circ. J. 2012, 76, 278–286. [Google Scholar] [CrossRef] [PubMed]
  92. Weatherley, B.D.; Milo-Cotter, O.; Felker, G.M.; Uriel, N.; Kaluski, E.; Vered, Z.; O’Connor, C.M.; Adams, K.F.; Cotter, G. Early worsening heart failure in patients admitted with acute heart failure—A new outcome measure associated with long-term prognosis? Fundam. Clin. Pharmacol. 2009, 23, 633–639. [Google Scholar] [CrossRef]
  93. Metra, M.; Chioncel, O.; Cotter, G.; Davison, B.; Filippatos, G.; Mebazaa, A.; Novosadova, M.; Ponikowski, P.; Simmons, P.; Soffer, J.; et al. Safety and efficacy of istaroxime in patients with acute heart failure-related pre-cardiogenic shock—A multicentre, randomized, double-blind, placebo-controlled, parallel group study (SEISMiC). Eur. J. Heart Fail. 2022, 24, 1967–1977. [Google Scholar] [CrossRef] [PubMed]
  94. Luciani, P.; Fevre, M.; Leroux, J.C. Development and physico-chemical characterization of a liposomal formulation of istaroxime. Eur. J. Pharm. Biopharm. 2011, 79, 285–293. [Google Scholar] [CrossRef] [PubMed]
  95. Hoffman, T.M. Newer inotropes in pediatric heart failure. J. Cardiovasc. Pharmacol. 2011, 58, 121–125. [Google Scholar] [CrossRef] [PubMed]
  96. Luraghi, A.; Ferrandi, M.; Barassi, P.; Arici, M.; Hsu, S.C.; Torre, E.; Ronchi, C.; Romerio, A.; Chang, G.J.; Ferrari, P.; et al. Highly Selective SERCA2a Activators: Preclinical Development of a Congeneric Group of First-in-Class Drug Leads against Heart Failure. J. Med. Chem. 2022, 65, 7324–7333. [Google Scholar] [CrossRef]
  97. Lyon, A.R.; Samara, M.A.; Feldman, D.S. Cardiac contractility modulation therapy in advanced systolic heart failure. Nat. Rev. Cardiol. 2013, 10, 584–598. [Google Scholar] [CrossRef]
  98. Abi-Samra, F.; Gutterman, D. Cardiac contractility modulation: A novel approach for the treatment of heart failure. Heart Fail. Rev. 2016, 21, 645–660. [Google Scholar] [CrossRef] [Green Version]
  99. Neelagaru, S.B.; Sanchez, J.E.; Lau, S.K.; Greenberg, S.M.; Raval, N.Y.; Worley, S.; Kalman, J.; Merliss, A.D.; Krueger, S.; Wood, M.; et al. Nonexcitatory, cardiac contractility modulation electrical impulses: Feasibility study for advanced heart failure in patients with normal QRS duration. Heart Rhythm 2006, 3, 1140–1147. [Google Scholar] [CrossRef]
  100. Borggrefe, M.M.; Lawo, T.; Butter, C.; Schmidinger, H.; Lunati, M.; Pieske, B.; Misier, A.R.; Curnis, A.; Bocker, D.; Remppis, A.; et al. Randomized, double blind study of non-excitatory, cardiac contractility modulation electrical impulses for symptomatic heart failure. Eur. Heart J. 2008, 29, 1019–1028. [Google Scholar] [CrossRef] [PubMed]
  101. Kadish, A.; Nademanee, K.; Volosin, K.; Krueger, S.; Neelagaru, S.; Raval, N.; Obel, O.; Weiner, S.; Wish, M.; Carson, P.; et al. A randomized controlled trial evaluating the safety and efficacy of cardiac contractility modulation in advanced heart failure. Am. Heart J. 2011, 161, 329–337.e2. [Google Scholar] [CrossRef] [PubMed]
  102. Abraham, W.T.; Nademanee, K.; Volosin, K.; Krueger, S.; Neelagaru, S.; Raval, N.; Obel, O.; Weiner, S.; Wish, M.; Carson, P.; et al. Subgroup analysis of a randomized controlled trial evaluating the safety and efficacy of cardiac contractility modulation in advanced heart failure. J. Card. Fail. 2011, 17, 710–717. [Google Scholar] [CrossRef] [PubMed]
  103. Abraham, W.T.; Kuck, K.H.; Goldsmith, R.L.; Lindenfeld, J.; Reddy, V.Y.; Carson, P.E.; Mann, D.L.; Saville, B.; Parise, H.; Chan, R.; et al. A Randomized Controlled Trial to Evaluate the Safety and Efficacy of Cardiac Contractility Modulation. JACC Heart Fail. 2018, 6, 874–883. [Google Scholar] [CrossRef] [PubMed]
  104. Wiegn, P.; Chan, R.; Jost, C.; Saville, B.R.; Parise, H.; Prutchi, D.; Carson, P.E.; Stagg, A.; Goldsmith, R.L.; Burkhoff, D. Safety, Performance, and Efficacy of Cardiac Contractility Modulation Delivered by the 2-Lead Optimizer Smart System: The FIX-HF-5C2 Study. Circ. Heart Fail. 2020, 13, e006512. [Google Scholar] [CrossRef]
  105. Muller, D.; Remppis, A.; Schauerte, P.; Schmidt-Schweda, S.; Burkhoff, D.; Rousso, B.; Gutterman, D.; Senges, J.; Hindricks, G.; Kuck, K.H. Clinical effects of long-term cardiac contractility modulation (CCM) in subjects with heart failure caused by left ventricular systolic dysfunction. Clin. Res. Cardiol. 2017, 106, 893–904. [Google Scholar] [CrossRef] [Green Version]
  106. Liu, M.; Fang, F.; Luo, X.X.; Shlomo, B.H.; Burkhoff, D.; Chan, J.Y.; Chan, C.P.; Cheung, L.; Rousso, B.; Gutterman, D.; et al. Improvement of long-term survival by cardiac contractility modulation in heart failure patients: A case-control study. Int. J. Cardiol. 2016, 206, 122–126. [Google Scholar] [CrossRef] [PubMed]
  107. Masarone, D.; Kittleson, M.M.; De Vivo, S.; D’Onofrio, A.; Ammendola, E.; Nigro, G.; Contaldi, C.; Martucci, M.L.; Errigo, V.; Pacileo, G. The Effects of Device-Based Cardiac Contractility Modulation Therapy on Left Ventricle Global Longitudinal Strain and Myocardial Mechano-Energetic Efficiency in Patients with Heart Failure with Reduced Ejection Fraction. J. Clin. Med. 2022, 11, 5866. [Google Scholar] [CrossRef]
  108. Avvisato, R.; Jankauskas, S.S.; Santulli, G. Istaroxime and Beyond: New Therapeutic Strategies to Specifically Activate SERCA and Treat Heart Failure. J. Pharmacol. Exp. Ther. 2023, 384, 1–3. [Google Scholar] [CrossRef]
  109. Sze, S.; Pellicori, P.; Zhang, J.; Weston, J.; Clark, A.L. Which frailty tool best predicts morbidity and mortality in ambulatory patients with heart failure? A prospective study. Eur. Heart J. Qual. Care Clin. Outcomes 2022, in press. [Google Scholar] [CrossRef]
  110. Pocock, S.J.; Ferreira, J.P.; Gregson, J.; Anker, S.D.; Butler, J.; Filippatos, G.; Gollop, N.D.; Iwata, T.; Brueckmann, M.; Januzzi, J.L.; et al. Novel biomarker-driven prognostic models to predict morbidity and mortality in chronic heart failure: The EMPEROR-Reduced trial. Eur. Heart J. 2021, 42, 4455–4464. [Google Scholar] [CrossRef]
  111. Rich, J.D.; Burns, J.; Freed, B.H.; Maurer, M.S.; Burkhoff, D.; Shah, S.J. Meta-Analysis Global Group in Chronic (MAGGIC) Heart Failure Risk Score: Validation of a Simple Tool for the Prediction of Morbidity and Mortality in Heart Failure With Preserved Ejection Fraction. J. Am. Heart Assoc. 2018, 7, e009594. [Google Scholar] [CrossRef] [PubMed]
  112. Lee, W.C.; Wu, P.J.; Fang, H.Y.; Fang, Y.N.; Chen, H.C.; Tong, M.S.; Sung, P.H.; Lee, C.H.; Chung, W.J. Levosimendan Administration May Provide More Benefit for Survival in Patients with Non-Ischemic Cardiomyopathy Experiencing Acute Decompensated Heart Failure. J. Clin. Med. 2022, 11, 3997. [Google Scholar] [CrossRef] [PubMed]
  113. Dass, B.; Dimza, M.; Singhania, G.; Schwartz, C.; George, J.; Bhatt, A.; Radhakrishnan, N.; Bansari, A.; Bozorgmehri, S.; Mohandas, R. Renin-Angiotensin-Aldosterone System Optimization for Acute Decompensated Heart Failure Patients (ROAD-HF): Rationale and Design. Am. J. Cardiovasc. Drugs 2020, 20, 373–380. [Google Scholar] [CrossRef] [PubMed]
  114. Scrutinio, D.; Ammirati, E.; Guida, P.; Passantino, A.; Raimondo, R.; Guida, V.; Sarzi Braga, S.; Canova, P.; Mastropasqua, F.; Frigerio, M.; et al. The ADHF/NT-proBNP risk score to predict 1-year mortality in hospitalized patients with advanced decompensated heart failure. J. Heart Lung Transplant. 2014, 33, 404–411. [Google Scholar] [CrossRef]
  115. Arrigo, M.; Jessup, M.; Mullens, W.; Reza, N.; Shah, A.M.; Sliwa, K.; Mebazaa, A. Acute heart failure. Nat. Rev. Dis. Primers 2020, 6, 16. [Google Scholar] [CrossRef]
Figure 1. Istaroxime inhibits the Na+/K+-ATPase (NAK) pump. (A) NAK actively transports three Na+-ions through the cell membrane outside the cytosol in exchange for two K+ ions inside the cytoplasm. (B) Istaroxime inhibits NAK by binding it from the extracellular side with a consequent raise of [Na+] in the intracellular side. (C) Increased [Na+] leads to the activation of the Na+/Ca2+ exchanger (NCX), which exchanges three Na+ ions for one Ca2+ ion, eventually increasing intracytoplasmatic [Ca2+].
Figure 1. Istaroxime inhibits the Na+/K+-ATPase (NAK) pump. (A) NAK actively transports three Na+-ions through the cell membrane outside the cytosol in exchange for two K+ ions inside the cytoplasm. (B) Istaroxime inhibits NAK by binding it from the extracellular side with a consequent raise of [Na+] in the intracellular side. (C) Increased [Na+] leads to the activation of the Na+/Ca2+ exchanger (NCX), which exchanges three Na+ ions for one Ca2+ ion, eventually increasing intracytoplasmatic [Ca2+].
Jcm 11 07503 g001
Figure 2. Istaroxime activates SERCA2a. (A) PLB, in its dephosphorylated form, inhibits SERCA2a. (B) PKA phosphorylates PLB releasing SERCA2a inhibition, thereby Ca2+ can be re-uptaken from the cytosol to the sarcoplasmic reticulum (SR). (C) Istaroxime activates SERCA2a with a direct interaction with the SERCA2a/PLB complex, independent of PLB phosphorylation. PKA: protein kinase A; PLB: phospholamban; SERCA2a: sarco/endoplasmic reticulum Ca2+ ATPase 2a.
Figure 2. Istaroxime activates SERCA2a. (A) PLB, in its dephosphorylated form, inhibits SERCA2a. (B) PKA phosphorylates PLB releasing SERCA2a inhibition, thereby Ca2+ can be re-uptaken from the cytosol to the sarcoplasmic reticulum (SR). (C) Istaroxime activates SERCA2a with a direct interaction with the SERCA2a/PLB complex, independent of PLB phosphorylation. PKA: protein kinase A; PLB: phospholamban; SERCA2a: sarco/endoplasmic reticulum Ca2+ ATPase 2a.
Jcm 11 07503 g002
Figure 3. Dual mechanism of action of istaroxime. Istaroxime (structure depicted on the left) inhibits NAK by binding it from the extracellular side. NAK inhibitions increase [Na+] inside the cytoplasm, activating NCX that expels three Na+ -ions in one Ca2+ ion exchange, raising [Ca2+] inside the cytoplasm, increasing myocardial contractility (inotropic effect); istaroxime activates SERCA with a direct interaction with the complex formed by SERCA2a and PLB, independent of PLB phosphorylation. SERCA2a stimulation triggers an influx of Ca2+ against the gradient from the cytosol to the SR, guaranteeing the relaxation of cardiac muscle fibers (lusitropic effect) and a sufficient amount of Ca2+ storage in the SR that can be utilized to start a new contractile activity for the ensuing contraction (empowerment of inotropic effect). NAK: Na+/K+-ATPase pump; NCX: Na+/Ca2+ exchanger; PKA: protein kinase A; PLB: phospholamban; SERCA: sarco/endoplasmic reticulum Ca2+ ATPase.
Figure 3. Dual mechanism of action of istaroxime. Istaroxime (structure depicted on the left) inhibits NAK by binding it from the extracellular side. NAK inhibitions increase [Na+] inside the cytoplasm, activating NCX that expels three Na+ -ions in one Ca2+ ion exchange, raising [Ca2+] inside the cytoplasm, increasing myocardial contractility (inotropic effect); istaroxime activates SERCA with a direct interaction with the complex formed by SERCA2a and PLB, independent of PLB phosphorylation. SERCA2a stimulation triggers an influx of Ca2+ against the gradient from the cytosol to the SR, guaranteeing the relaxation of cardiac muscle fibers (lusitropic effect) and a sufficient amount of Ca2+ storage in the SR that can be utilized to start a new contractile activity for the ensuing contraction (empowerment of inotropic effect). NAK: Na+/K+-ATPase pump; NCX: Na+/Ca2+ exchanger; PKA: protein kinase A; PLB: phospholamban; SERCA: sarco/endoplasmic reticulum Ca2+ ATPase.
Jcm 11 07503 g003
Table 1. Summary of designs and conclusions of preclinical studies investigating istaroxime.
Table 1. Summary of designs and conclusions of preclinical studies investigating istaroxime.
Clinical ConditionAnimal ModelEndpointConclusion
DCMDCM STZ induced in ratsDDIstaroxime improved DD stimulating SERCA2a and reducing alterations in intracellular Ca2+ handling.
ADHFCanine model of HF produced by multiple sequential intracoronary embolizations with microspheresLVEF (%);
LVEDV (mL);
LVESV (mL).
Istaroxime improved hemodynamic and echocardiographic parameters.
Chronic ischemic HF, comparing istaroxime to dobutamineCanine model of HF produced by ligation of the left anterior descending coronary artery and intracoronary embolizationsLV functionIstaroxime was shown to be an effective inotropic agent without positive chronotropic actions.
Progressive HFHamster model of progressive HFHeart/body weight ratio;
max dP/dT; min dP/dT;
LVSP; CFR
Istaroxime improved cardiac function and heart rate variability
Electrophysiological effects of istaroxime and digoxinGuinea pig isolated ventricular myocytesEffects on ITIIstaroxime inhibited ITI (effect not evident with digoxin)
Cardiotoxic effects of equi-inotropic concentrations of istaroxime and ouabainRat isolated ventricular myocytesCell viability; Apoptosis; CaMKII activation.Istaroxime had a significant inotropic effect, neither activating CaMKII nor promoting cardiomyocytes death (contrary to digoxin)
ADHF: acute decompensated heart failure; CaMKII: Ca2+/calmodulin-dependent kinase II; CFR: coronary flow rate; DCM: diabetic cardiomyopathy; DD: diastolic dysfunction; dP/dT: derivative of LV pressure; LVEDV: LV end-diastolic volume; LVESV: LV end-systolic volume; HF: heart failure; ITI: transient inward current of Ca2+; LV: left ventricle; LVEF: LV ejection fraction; LVSP: LV systolic pressure; MI: myocardial infarction; STZ: streptozotocin.
Table 2. Results of the main clinical trials on Istaroxime.
Table 2. Results of the main clinical trials on Istaroxime.
Clinical TrialPrimary Endpoint Main Results
HORIZON-HF (NCT00616161)Change in PCWP (mmHg)Istaroxime: −3.2 ± 6.8, −3.3 ± 5.5, and −4.7 ± 5.9 vs. placebo: 0.0 ± 3.6; p < 0.05 (for all doses)
The Clinical Study of the Safety and Efficacy of Istaroxime in Treatment of ADHF (NCT02617446)E/e’ ratio change
from baseline to 24 h
cohort 1: istaroxime 0.5 μg/kg/min: −4.55 ± 4.75 vs. placebo: −1.55 ± 4.11, p = 0.029;
cohort 2: 59 istaroxime 1.0 μg/kg/min: −3.16 ± 2. vs. placebo: −1.08 ± 2.72, p = 0.009
SEISMiC (NCT04325035)AUC (mmHg × hour; change in SBP from baseline through 6 h)Istaroxime: 53.1 ± 6.88
vs. placebo: 30.9 ± 6.76, p = 0.017
ADHF: acute decompensated heart failure; AUC: area under curve; PCWP: pulmonary capillary wedge pressure; SBP: systolic blood pressure.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Forzano, I.; Mone, P.; Mottola, G.; Kansakar, U.; Salemme, L.; De Luca, A.; Tesorio, T.; Varzideh, F.; Santulli, G. Efficacy of the New Inotropic Agent Istaroxime in Acute Heart Failure. J. Clin. Med. 2022, 11, 7503. https://0-doi-org.brum.beds.ac.uk/10.3390/jcm11247503

AMA Style

Forzano I, Mone P, Mottola G, Kansakar U, Salemme L, De Luca A, Tesorio T, Varzideh F, Santulli G. Efficacy of the New Inotropic Agent Istaroxime in Acute Heart Failure. Journal of Clinical Medicine. 2022; 11(24):7503. https://0-doi-org.brum.beds.ac.uk/10.3390/jcm11247503

Chicago/Turabian Style

Forzano, Imma, Pasquale Mone, Gaetano Mottola, Urna Kansakar, Luigi Salemme, Antonio De Luca, Tullio Tesorio, Fahimeh Varzideh, and Gaetano Santulli. 2022. "Efficacy of the New Inotropic Agent Istaroxime in Acute Heart Failure" Journal of Clinical Medicine 11, no. 24: 7503. https://0-doi-org.brum.beds.ac.uk/10.3390/jcm11247503

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