Skoči na glavni sadržaj

Pregledni rad

https://doi.org/10.15836/ccar2024.251

Razumijevanje zatajivanja srca: evolucija shvaćanja i liječenja

Anton Šmalcelj orcid id orcid.org/0009-0007-6410-1387 ; School of Medicine, University of Zagreb, University Hospital Centre Zagreb, Zagreb, Croatia (retired)


Puni tekst: hrvatski pdf 539 Kb

str. 251-269

preuzimanja: 0

citiraj

Puni tekst: engleski pdf 539 Kb

str. 251-269

preuzimanja: 0

citiraj

Preuzmi JATS datoteku


Sažetak

SAŽETAK
Stariji se kardiolozi mogu prisjetiti evolucije poimanja patofiziologije zatajivanja srca (HF) i pristupa liječenju tijekom profesionalne karijere. Milton Packer je nedavno napredak koncepcije HF-a sažeo u četiri stadija: 1) kardiorenalni model (od 40-ih do kraja 60-ih godina prošloga stoljeća), 2) kardiocirkulacijski model ili hemodinamska hipoteza (70-e i 80-e godine 20. st.), 3) neurohormonalni model (od 90-ih godina 20. st. donedavno) i 4) najnoviji model staničnog opterećenja. Kardiorenalni se model oslanjao na liječenje digitalisom i diureticima. Vazodilatatori i pozitivni inotropi pobudili su nade u vrijeme dominacije hemodinamske hipoteze, ali nisu opravdali očekivanja. Tek je neurohormonalni pristup s inhibitornima renin-angiotenzin-aldosteronskog sustava i beta-blokatorima kao djelatnim lijekovima postigao trajno poboljšanje najvažnijih kliničkih ishoda, uključujući smrtnost. Takvo je liječenje, međutim, bilo neučinkovito u zatajivanju srca s očuvanom ejekcijskom frakcijom (HFpEF), tj. u polovice svih bolesnika s HF-om. Sretan splet okolnosti otkrio je da su inhibitori suprijenosnika natrija-glukoze 2 (SGLT2 inhibitori), prvotno uvedeni kao antidijabetički lijekovi, korisni i za kardiovaskularni sustav. Ta je spoznaja potaknula klinička istraživanja kojima su dokazani korisni učinci na ishode liječenja ne samo zatajivanja srca s reduciranom ejekcijskom frakcijom nego i HFpEF. Kada se misterij načina SGLT2 inhibitora počeo razotkrivati, najavljen je model staničnog opterećenja u HF-u. Pregled je usredotočen na povijesne i nove patofiziološke koncepcije HF-a, zajedno s odgovarajućim lijekovima, ali su spomenuti i nefarmakološki načini liječenja. Na kraju se raspravlja o izgledima za daljnji napredak razumijevanja i liječenja HF-a.

Ključne riječi

zatajivanje srca; neurohormonalna inhibicija; stanično opterećenje

Hrčak ID:

320160

URI

https://hrcak.srce.hr/320160

Datum izdavanja:

22.8.2024.

Podaci na drugim jezicima: engleski

Posjeta: 0 *




Historical background

William Harvey’s epochal discovery of blood circulation driven by a heart pump in 1628 provided a framework for the basic concept of heart failure (HF). Ancient Greek and Roman physicians attributed edema, anasarca, and dyspnea to a variety of causes other than heart disease; pleural effusions were thought to originate in the brain, and the heart was believed to heat and distribute the vital spirit. Linking cardiac pathology to hemodynamic and clinical disorders following Harvey’s discovery took time. G.M. Lancisi (1654-1720) observed that dilatation weakens the heart. J.N. Corvisart wrote in 1806 that marked cardiac dilatation in association with valvular regurgitation portends HF and a bad prognosis. In 1892, W. Osler viewed cardiac hypertrophy as a step towards HF (“broken compensation”). Those and many other physicians, whether famous or less known, elaborated the clinical presentation and course of HF (13).

Early pathophysiological concepts were simplistic, limited to cardiac weakness causing low cardiac output and extracellular fluid retention with pulmonary and systemic congestion. Treatment options were empirical and ineffective. Lifestyle changes provided some relief. Bloodletting and leeches were used for centuries. Southey tubes for edema drainage have been long forgotten. Among many herbal treatments that have been tried, only digitalis (foxglove) preparations proved to be a lucky hit. Since introduction to medical use by W. Withering in 1785, digitalis had been a pillar of HF treatment for about 200 years until recently (49).

Approaching modern times: the story of diuretics

The creation of effective diuretics in late the 1950s heralded the modern era of pharmacological HF management. Until 1957, the only diuretics used in HF were intravenous or intramuscular mercurial agents, which were difficult to use and fraught with toxicity. The serendipitous discovery of modern diuretics took place during the study of sulphonamide side-effects. In 1937/8, research on sulphonamides revealed their diuretic effects. In 1945, the development of the carbonic anhydrase inhibitor acetazolamide improved the understanding of diuretic mechanisms in renal tubules. Among many compounds synthesized in the search for potent carbonic anhydrase inhibitors, researchers stumbled on chlorothiazide, which inhibited not only carbonic anhydrase but also the sodium chloride cotransport system. Introduced in 1958, chlorothiazide was the first useful oral diuretic. It is still used widely, but much more as an antihypertensive agent than a HF drug (1012).

Thiazides and thiazide-like diuretics may be useful in the management of mild HF, but more potent diuretics are needed in severe forms. Furosemide, with a brand-name Lasix (“lasts six hours”), came to the rescue. The introduction of the loop diuretics furosemide and ethacrynic acid in the early 1960s dramatically improved medical practice. Until then, HF had been considered a terminal condition, but new diuretics allowed amazing relief of fluid retention. It was obvious that they worked, and little in the way of clinical trials was needed to accept them in clinical practice. Furosemide, as an archetype of potent loop diuretics, was followed by now forgotten ethacrynic acid and bumetanide, in addition to its current rival torasemide, and has remained a cornerstone in the treatment of congestive HF (1214).

Spironolactone, a nonselective steroid aldosterone receptor antagonist, appeared in 1957 (15). Its independent diuretic action is weak, but the synergism with loop diuretics is strong, in addition to potassium-sparing effects. The value of spironolactone became evident in the RALES trial (1999) which showed a reduction in morbidity and death among patients with severe HF (16). Those data suggested a paradigm-shift in the concept of HF, demonstrating the neurohormonal actions of spironolactone. The selective aldosterone receptor antagonist epleronone shares beneficial effects with spironolactone but is devoid of side-effects on sexual steroid hormones (17). Finally, the use of a novel nonsteroidal aldosterone receptor antagonist fineronone is advantageous in patients with chronic kidney disease and type 2 diabetes (18).

Diuretics were accepted as symptom-relieving and lifesaving drugs before the days of large clinical trials with mortality and survival endpoints. Even nowadays, the lack of alternative to loop diuretics makes it hard to envision how to design such trials (19,20).

Practice, theories, and pathophysiological concepts

With digoxin, furosemide, and spironolactone available, the stage was set for a dynamic era of HF management. Past stagnation and incapacity gave way to active pursuit of advances in HF treatment that dominated in the decades to follow. This pursuit rekindled research on HF pathophysiology (21). Research requires concepts that representing the basic models of HF. Those have changed significantly since the middle of the 20th century. Senior cardiologists may recall them from their own professional experience. The formulation of those models may vary somewhat, but some of them, e.g. the neurohormonal model, have been widely accepted. Milton Packer has proposed four concepts: 1) the cardiorenal model (1940s through the 1960s), 2) the cardiocirculatory model or hemodynamic hypothesis (1970s and 1980s), 3) the neurohormonal model (from 1990s up to recently), and 4) the recent cellular stress model (2224). Such a conceptual framework is interesting as it may explain many advancements but also some side-tracks. We will try to explicate the concepts, adding some remarks.

THE CARDIORENAL MODEL

The cardiorenal model regarded HF as an edematous disorder with salt and water retention. The pillars of treatment were digitalis and diuretics. Although the congestion seen to be due mostly to increased venous pressures (“backward HF”) dominated the concept, low cardiac output (“forward HF”) was duly appreciated. Moreover, renal hypoperfusion due to low cardiac output was deemed essential for extracellular fluid retention by activating renal (glomerular, tubular, and peritubular) mechanisms of volume conservation. These may be viewed as atavistic responses to hypovolemia due to fluid loss. In due course, the concept was expanded by adrenal and neurohormonal (e.g. antidiuretic hormone secretion) mechanisms of volume retention with dilutional hyponatremia as an epiphenomenon. Late additions were synthesis with the neurohormonal HF concept. The erstwhile cardiorenal concept views HF as a cardiac disorder with anticipatable renal response (without disease), at variance with the modern concept of cardiorenal syndrome where the disease of any of those two organs induces the disorder of the other (21,22).

The central tenet of the erstwhile cardiorenal concept assumes the main role of the heart disorder in the pathogenesis of HF, with the kidneys playing supporting roles. The widespread use of digitalis preparations (mostly digoxin) was intended to alleviate cardiac disorder as the origin of HF. The evaluation of efficacy and clinical utility was based on clinical observations, experience, and judgement; only later were some respectable randomized trials conducted, with neutral results on mortality (2529). The weak positive inotropy of digitalis, due to the inhibition of Na+/K+-ATPase in cardiac myocytes with a consequent increase in cytosolic Ca2+ content through Na+/Ca2+ exchanger, was overrated. Negative chronotropy, desirable in tachycardia due to atrial fibrillation, was limited because of narrow therapeutic width. Despite limitations, digitalis preparations had remained the mainstay of HF treatment for two centuries (6). Perhaps M. Packer, when giving the European Society of Cardiology (ESC) Rene Laennec Lecture on Clinical Cardiology at the ESC Congress 2023, dated the cardiorenal model to the 1748-1965 period correctly; digitalis use dates back that far (24).

Potent diuretics became available not earlier than in the 1960s, when this period was at an end. The diuretics revolutionized HF treatment and have remained its mainstay irrespectively of the overall concept. They are needed whenever an excess of extracellular fluid arises (12,14).

Finally, sticking to the cardiorenal model impeded advances in HF treatment. Lowering arterial pressure was deemed risky because of renal and myocardial hypoperfusion. The use of vasodilators was avoided except in severe hypertension. Similarly, the use of beta-blockers (BB) was restrained (propranolol appeared in the 1960s as an antianginal drug), fearing not only arterial hypotension but also negative inotropy (22,30,31).

THE CARDIOCIRCULATORY MODEL

The cardiocirculatory model or the hemodynamic hypothesis (1965-1992), as it was formulated by M. Packer, followed the cardiorenal model. The quest for new treatment strategies spurred by the unmet expectations of digoxin and diuretics led to reappraisal of pathophysiological concepts. The ensuing paradigm shift viewed HF principally as a hemodynamic disorder involving the whole cardiovascular system, not only the heart but also the vessels. The focus shifted from the kidneys to peripheral vessels and redistribution of intravascular volumes (22).

In HF, adrenergically mediated arterial and venous vasoconstriction devastatingly impairs hemodynamics. The resulting increase in afterload exhausts the myocardium and impedes cardiac output. Reduction of the venous reservoir increases cardiac preload and aggravates congestion. The rationale for the use of vasodilating drugs was to relieve the failing heart of preload and afterload burden, expecting a recovery in cardiac function (21,22). The whole array of vasodilators was tested intravenously or orally to relieve vasoconstriction and to achieve a breakthrough in HF management. Senior cardiologists may recall the fear of causing hypotension with those unorthodox innovations (30,31). The use of some initially promising vasodilators, e.g. the adrenergic alpha-receptor blockers phentolamine and prazosin, remained a dead-end attempt (3236). Hydralazine and isosorbide dinitrate held some promise, which was not fulfilled (3638). Calcium channel blockers were expected to improve hemodynamics, but controlled clinical trials showed worsening of HF (3943). At best, oral vasodilators provided some temporary relief without long-term benefits on outcomes. Intravenously administered vasoactive agents for acute HF fared better, especially if guided by hemodynamic monitoring. Intravenous nitroglycerine, acting mostly as a venous vasodilator, is still valuable in acute cardiogenic pulmonary edema, while combined arterial and venous vasodilator sodium nitroprusside is helpful in severe acute HF in a critical care setting (4447).

Disappointment with oral vasodilators shifted the focus to the heart itself. As the impairment of myocardial contractility was deemed to be the primary problem, the whole array of inotropic agents was created. The basic approach was to increase contractility, stimulating the influx of calcium ions or maintaining higher calcium levels in the cytosol of cardiac myocytes throughout the action potential. Dozens of such drugs progressed to phase 3 clinical trials. Among them, dobutamine, amrinone, milrinone, enoximone, levosimendan, pimobendane, and xamoterol were the most promising. Dobutamine was the first of them, created in 1975 by modifying the chemical structure of isoproterenol. The rationale was to benefit from adrenergic stimulation but to avoid detrimental vasoconstriction. A series of phosphodiesterase 3 (PDE-3) inhibitors followed, amrinone first, followed by milrinone and enoximone. Calcium sensitizers were developed in the 1980s, with levosimendan as a prototype. It increases contractile apparatus sensitivity to calcium ions during systole, not interfering with their diastolic release. Levosimendan also inhibits PDE-3 and activates ATP sensitive K+ channels, causing strong vasodilation. Pimobendan shares calcium sensitizing and PDE-3 inhibiting effects. Xamoterol is a beta1 selective partial adrenergic agonist. The ideal was to develop a positive inotropic agent with vasodilator properties, like levosimendan, or a vasodilator with positive inotropic properties, like flosequinan. This drug has both vasodilating and inotropic properties that are not entirely understood but are believed to be distinct from β-adrenergic receptor agonists and PDE inhibitors (48).

Those and other positive inotropic agents, irrespective of mode of action, provided a transient hemodynamic and clinical improvement but failed to reduce mortality and morbidity. Moreover, they were associated with increased mortality, except the prognostically neutral digoxin (25,48,49). Many inotropic drugs were created, but only those designed for short-term intravenous use (e.g. 48 h) in severe acute HF remained in use. Dobutamine, milrinone, and levosimendan are still indispensable in our cardiac care units (5053).

The failure of inotropic agents to convert short-term improvements into long-term benefits, with an excess of mortality, can be explained only speculatively. A mechanistic explanation suggests that overstrain of an injured organ shortens its lifespan. Overstimulating sick myocardium unduly depletes its meagre stores of energy. The persistent adrenergic overstimulation is detrimental, whereas BBs reduce mortality and morbidity. In addition, the chronic use of drugs acting via cAMP modulation, like PDE-3 inhibitors and adrenergic stimulants, disrupts calcium homeostasis with desensitization of the contractile apparatus to calcium through impairment of early diastolic relaxation and ventricular arrhythmias. Disruption of cardiac myocyte energetics and calcium ion homeostasis are detrimental for cardiac myocytes. Furthermore, HF is a heterogeneous syndrome comprising a diverse spectrum of diseases. The harmful effects of chronic treatment with milrinone were conspicuously more prevalent in ischemic than in non-ischemic HF. The doses of inotropic agents used were perhaps too high, adjusted to achieve maximal immediate inotropic effect, notwithstanding later consequences. The two-century-long history of digitalis parallels the positive inotropes of the 1980s in overuse and overdosage deviations. More judicious and selective use of those inotropes may have yielded different end-results (48,49,54).

Even if the attempts to treat HF with inotropes failed, the concept itself may not be doomed to failure. Strengthening the weak heart was the primordial aim which appeared self-evident for generations of physicians and was incorporated in wishful thinking on digitalis inotropy. Positive inotropes are still expected to have significant roles in chronic HF treatment, but as part of an elaborate framework of concepts and not as a blunt overstimulation of already exhausted cardiac myocytes. A hiatus in the development of positive inotropes was followed by new cellular targets in the 2000s. Pharmacological and gene therapy approaches were directed at a key enzyme responsible for myocardial calcium homeostasis that is downregulated in HF: sarcoplasmic reticulum Ca2+-ATPase (SERCA2a). Another concept is related to cardiac myosin activators, which are a new class of myotropes that improve myocardial function by directly augmenting cardiac sarcomere function. Omecamtiv mecarbil, the first of this class, augments cardiac contractility by selectively binding to cardiac myosin, thus increasing the number of myosin heads that can bind to the actin filament and initiate a power stroke at the start of systole (48,54). The GALACTIC-HF trial (2021) showed that omecamtiv mecarbil reduced the incidence of a composite of a heart failure events and death from cardiovascular causes among patients with HF with reduced ejection fraction (HFrEF) (55,56). However, FDA has declined to approve omecamtiv mecarbil, citing a lack of evidence on efficacy in 2023. No positive inotrope is currently approved for long-term use in HF.

M. Packer, who formulated the hemodynamic model of HF in his original article (1993), did not pay much attention to the recognition of diastolic left ventricular dysfunction as the cause of HF. Though the pressure-volume relations during the cardiac cycle had already been recognized in the early 20th century and the term left ventricular lusitropy, denoting the rate of early diastolic relaxation, predated clinical concepts, clinical research did not recognize the concept of left ventricular diastolic dysfunction until the 1970s and clinical practice did not use it until the 1980s (5766). This was a real paradigm shift, since until then only the impairment of left ventricular systolic function (i.e. contractility) was regarded as a cause of HF. It was only then that cardiologists realized that left ventricular systolic function was preserved in 40-50% of HF cases. It was assumed that diastolic dysfunction was the main culprit, accounting for at least 30% of all HF cases (62,6769).

This led to the confusing question of how to treat those patients. Use of positive inotropic agents seemed senseless and diuretics necessary; the excess of fluid should be removed in any case. It was advised to do this cautiously, since the Frank Starling curve was supposed to be steep and shifted to the right. Therefore, a sudden contraction of intravascular volume could precipitate a sudden drop in stroke volume and cardiac output. Many small clinical studies and later some landmark hypertension trials showed that renin-angiotensin-aldosterone system (RAAS) antagonists may improve left ventricular diastolic function in parallel with left ventricular hypertrophy regression better than BBs. The latter may yet provide beneficial effects on diastolic function by prolonging diastole. Calcium channel blockers proved to be controversial. The evidence on efficacy of those treatments in HF was only circumstantial, limited to surrogate hemodynamic data without the data on morbidity and mortality outcomes (66,70,71).

The uncertainties about the treatment of HF due to the left ventricular diastolic dysfunction have been never resolved by clinical trials since the hemodynamic hypothesis fell into disrepute. The assessment of left ventricular diastolic dysfunction in clinical practice is often indeterminate or elusive (72,73). The conceptual advancements made clear that HF in patients with preserved systolic function cannot be simplistically reduced to diastolic dysfunction. The role of left ventricular diastolic dysfunction in the pathophysiology of HF has not been ignored but has instead been incorporated into broader and more comprehensive concepts of HF with preserved systolic function, normal ejection fraction (HFNEF), and preserved ejection fraction (HFpEF) (7479). These concepts outgrew the hemodynamically frame (7781).

THE NEUROHORMONAL HYPOTHESIS

The neurohormonal hypothesis (1992-2019), as it was named, formulated, and dated by M. Packer, presented a radical paradigm-shift and a conceptual breakthrough which viewed HF as a systemic disorder involving a complex neurohormonal response with renin-angiotensin-aldosterone (RAAS) and adrenergic systems in protagonist roles (2123). Senior cardiologists may remember that this approach emerged gradually, not because of ingenious thinking which revised pathophysiological concepts but rather arising empirically through studies on vasodilators. The first angiotensin converting enzyme inhibitor (ACEi) captopril, first isolated from snake venom, introduced in 1981 and initially considered a plain vasodilator useful in arterial hypertension, proved to be superior to prazosin, hydralazine, and other pure vasodilators in the treatment of chronic HF. The advantage was explained by the blockade of detrimental neuroendocrine responses (3436,82). Following captopril, many other ACEis with specific qualities were developed. They provided a quantum leap in the treatments across the cardiovascular continuum. A series of landmark clinical trials evaluating the treatment of chronic HF with impaired left ventricular systolic function demonstrated that ACEis provided not only short-term hemodynamic improvements but also long-term benefits to morbidity and mortality. The most cited game-changing trials were CONSENSUS I (enalapril, 1987), VHeFT I (enalapril, 1991), SOLVD (enalapril, 1991), SAVE (captopril after myocardial infarction, 1992), AIRE (ramipril after myocardial infarction, 1993), and TRACE (trandolapril, after myocardial infarction, 1995) (8389).

An unprecedented breakthrough in HF trials with ACEs reinvigorated experimental scientific research on RAAS to explain those results and chart further advances. It became clear that besides of the endocrine component of RAAS, there was also a widespread tissue RAAS, with the heart, vessels, nervous system, and the kidneys as the main players. Endocrine (in blood), paracrine (in the tissues), and intracrine (in the cells), signaling was identified in addition to autocrine and juxtacrine RAAS. RAAS was recognized as a ubiquitous system for homeostasis and pathologies, biologically fundamental, with deep evolutionary roots and composed of ancient molecules. Its pivotal molecule, angiotensin II, an octapeptide with strong vasoconstrictive properties, arose in the early Cambrian ~500 million years ago, primarily as an epigenetic regulator of protein synthesis and growth-promoting factor. It is a key molecule in the signaling pathway of pathological myocardial hypertrophy and a potent promotor of atherogenesis (9094).

Until the 90s, positive inotropic agents (mainly digoxin) were deemed a mainstay of HF treatment. As the poor left ventricular systolic function was viewed as the main cause of HF, all drugs with negative inotropic effect were “absolutely contraindicated” according to the practice guidelines up to 1995 (95). The idea of using BBs as a primary therapy for congestive HF to improve symptoms and prognosis seemed paradoxical and dissenting. The cardiac community reacted with skepticism and disbelief when, in 1975, the pioneering report of Waagstein et al. gave an account on 7 cases of refractory HF in patients with dilated cardiomyopathy treated successfully by already forgotten BBs alprenolol and practolol (96,97). This unorthodox approach may be viewed as a bailout in a desperate situation. Waagstein violated a taboo and stirred controversy. A change of opinion took time. Many small studies with surrogate endpoints only added to confusion. Decades were needed for scientific research to recognize the detrimental effects of maladaptive adrenergic response to declining systolic function and vindicate large controlled clinical trials (98). The methodology was honed in HF trials with ACEs. Landmark clinical trials convincingly demonstrated the efficacy of four BBs in improving morbidity and reducing mortality in the patients with “systolic” HF. Pioneering trials were: CIBIS I (bisoprolol, 1994), CIBIS II (bisoprolol 1999), US Carvedilol HF Trials Program (1996, carvedilol), and MERIT-HF (metoprolol, 2000) (99103). Other important trials with bisoprolol, carvedilol, and metoprolol followed, while the SENIORS trial affirming nebivolol took place slightly later (2005) (104).

While the ACEIs were introduced as vasodilators in the treatment of HF to improve hemodynamics and were only later recognized as neurohormonal agents, BBs were introduced primarily as neurohormonal agents. When M. Packer proposed his neurohormonal model of HF pathophysiology in 1993, BBs were only envisaged, but neither evaluated in clinical trials nor approved for the treatment of HF (23). The trials opened the door to the extensive use of BBs in the treatment of “systolic” HF and firmly established the neurohormonal concept. ACEIs and BBs took pole position in the guidelines for HF treatment on both sides of the Atlantic, sharing the first two positions (46,105).

The neurohormonal concept led to reappraisal of spironolactone. After 40 years spent in a modest role as an adjunct potassium sparing diuretic, following the RAALES trial (1999) spironolactone was revisited as an essential neurohormonal drug which reduces morbidity and mortality in patients with HF (16). Considering the key role of aldosterone receptors in the pathogenesis of cardiovascular and renal pathologies, the results of RAALES trial might have been expected, but the clear reduction in all-cause mortality sent a clear message. The EPHESUS trial (2003) showed that the selective aldosterone receptor blocker eplerenone reduced mortality and morbidity in patients with acute myocardial infarction complicated by HF (17). This evidence positioned mineralocorticoid receptor antagonists (MRAs) firmly in third place on the list of preferred HF drugs, after ACEIs and BBs, while digoxin was relegated to the low fourth position (106,107).

In 1992, research on angiotensin II led to losartan, an angiotensin II type 1 receptor antagonist with antihypertensive properties (108). It was a forerunner of the whole class of angiotensin receptor blockers (ARBs) which have been used as an alternative to ACEs ever since. Candesartan, eprosartan, irbesartan, valsartan, telmisartan, and olmesartan followed. Affirmed as an antihypertensive drug with heart and kidney protecting properties (RENAAL 2001; LIFE 2002), losartan was compared to captopril in the treatment of systolic HF ((ELITE I 1997; ELITE II, 2000) and acute myocardial infarction (OPTIMAAL 2002), showing non-inferiority with better tolerability (70,109113). The trials with newer ARBs upheld the message: ARBs are a noninferior alternative to ACEs in left ventricular systolic failure. Only valsartan (2001), candesartan (CHARM-Alternative 2003, CHARM-Added, 2023), losartan again (HEAAL, 2012), and to a lesser extent telmisartan (1999, 2010) were evaluated in “systolic” HF trials, while candesartan (CHARM-Preserved, 2003) and irbesartan (2008) were appraised in HF with preserved systolic function trials (114121). European HF guidelines (2021) do not claim any preference for ACEs or ARBs, recommending them both as an alternative (IA). American guidelines (2022) prefer ACEs as the first choice in RAAS-naive patients (46,105).

The use of ARBs in the treatment of HF with reduced ejection fraction (HFrEF) has been upgraded recently by neprilysin inhibition which fits neatly with the neurohormonal concept of HF (122). In PARADIGM-HF Trial (2014) the dual-acting angiotensin receptor-neprilysin inhibitor (ARNI) sacubitril valsartan reduced the composite endpoint of HF hospitalization and death in comparison with standard enalapril treatment (123,124). ARNI therapy now has a class I indication for the treatment of patients with HFrEF (46,105,125).

The evolutionarily highly conserved family of natriuretic peptides comprises the atrial, brain, and C-type peptides (ANP, BNP, and CNP). ANP and BNP, secreted by the atria and ventricles, operate via the natriuretic peptide receptors type A (NPR-A) and type B (NPR-B), which are coupled to guanyl cyclase, mediating biological effects. Those include vasodilatation, natriuresis, and diuresis, inhibition of the RAAS, endothelin, and vasopressin, along with lipid mobilization. ANP is degraded rapidly by endopeptidase neprilysin. The inhibition of neprilysin increases the levels of ANP in circulation, with beneficial effects in HF (122,126,127).

Mechanistic reasoning may turn out to be overly simplistic when faced with the complexity of clinical medicine. Promising concepts may not work in clinical practice. Clinical trials may dash hopes placed in promising treatments. Nesiritide, a human recombinant B-type natriuretic peptide (BNP) with vasodilatory properties, binds to receptors in the vasculature, kidney, and other organs to mimic the actions of endogenous natriuretic peptides. It was approved by the FDA in 2001 for use in patients with acute HF based on studies showing hemodynamic and symptomatic improvements. A few years after its approval (2005), nesiritide fell out of use because small studies seemed to indicate an increased risk of kidney problems and an increased death rate. The ASCEND-HF study (2011) showed no impact of nesiritide on death or HF hospitalization (128132).

Elevated resting heart rate has been linked to poor outcomes in patients with chronic systolic HF. Ivabradine may be added to neurohormonal treatments as the adjunctive therapy for HF with reduced ejection fraction (HFrEF) to slow sinus rhythm. It inhibits “funny” pacemaker current (If) of the sinoatrial node, not affecting the AV node, inotropy, diastolic function, cardiac output, vascular resistance, or blood pressure (133,134). Ivabradine should be considered (IIa indication) in the patients with LVEF ≤35% in sinus rhythm and with a resting heart rate ≥70 bpm who remain symptomatic despite optimally up-titrated BBs, ACEi/ARNI, and MRA based treatment (or in BB intolerant patients) to reduce the risk of HF hospitalization and cardiovascular death (46). Ivabradine is not a substitute for BBs (133).

Neurohormonal activation is the crucial mechanism underlying the progression of HF, and therapeutic antagonism of neurohormonal systems has become the cornerstone of pharmacotherapy for HF. The perception of HF has changed: it is no longer regarded as a terminal syndrome with a dismal prognosis but as a treatable disorder (92).

However, it was not possible to ignore the fact that neurohormonal inhibition did not work in patients with HFpEF, representing at least a half of all patients with HF. Their share has risen to >50%, owing mostly to the aging of the population. The typical phenotype are obese elderly women with a small, well-contracting left ventricle, diabetes, hypertension, and atrial fibrillation (77,135). Diastolic left ventricular dysfunction is a risk factor but not the main cause of HF. Landmark HF studies with RAAS antagonists and BBs excluded patients with LV EF ≥40%. Trials with candesartan, irbesartan, and spironolactone, designed to explicitly address the efficacy of RAAS antagonists in patients with EF ≥40%, yielded disappointing results. The CHARM-Preserved trial with candesartan (2003) showed only a moderate reduction in hospital admissions among patients with HF with LVEF >40%. The I-PRESERVE trial (2008) with irbesartan was neutral about outcomes in patients with HF with LVEF ≥45%. The TOPCAT trial (2014) demonstrated that in patients with HF and LVEF ≥45%, the treatment with spironolactone did not significantly reduce the incidence of the primary composite outcome of cardiovascular death, aborted cardiac arrest, or hospitalisation (120,121,136).

Failure of neurohormonal inhibition to improve HFpEF took the cardiac community aback since the prognosis of HFpEF may be as grave as that of HFrEF (137). Cardiologists were at a loss since a huge population of patients with HFpEF was left without any treatment strategy. Judicious use of diuretics for decongestion was the only treatment remaining, along with the treatment of comorbidities (138,139). HFpEF is a phenotypically and pathophysiologically heterogeneous disorder in which therapy should target the underlying phenotypes, etiologies, and comorbidities, but such an approach was frustratingly complex (140). Relying on speculative treatment of LV diastolic dysfunction based on hemodynamics did not help. New treatment strategies were needed.

The cut-off point for HFpEF of 40-45% was set to distinguish HFpEF from HFrEF of earlier HF trials. However, trials like CHARM-Preserved and TOPCAT indicated the presence of a transitional EF range from 40% to 49% where the beneficial effects of candesartan and spironolactone were better than with EF ≥50%, albeit not as good as with <40%. The term HF mid-range EF (40-49%) was introduced (≈2014) and soon (2021) renamed to HF with mildly reduced EF (41-49%) or HFmrEF. It is an intermediate category between HFrEF and HFpEF with an estimated prevalence of 10% to 20% among all patients with HF. The 2021 ESC guidelines on HF stated that “although no strong recommendations for this HF phenotypes can be made ACEIs, ARBs, MRAs, BBs, and ARNI can be considered to reduce the risk of death and HF hospitalization in HFmEF (IIb)”. In other words, pharmacological neurohormonal inhibition may be beneficial in patients with HFmEF, but not as much as in patients with HFrEF, since the evidence is far weaker (46,139143). According to the 2023 ESC HF guidelines update, neurohormonal inhibitors comprising ACEI/ARNI/ARBs, BBs, and MRAs may be considered (class IIb recommendation) in patients with HFmrEF, while diuretics for fluid retention and dapagliflozin/empagliflozin are indicated with class I recommendation (125).

Two peculiar new non-neurohormonal drugs have raised hopes for the treatment of patients with HFrEF in addition to the standard neurohormonal inhibition. Vericiguat works via stimulation of soluble guanylate cyclase (sGC) in the NO-sGC-cGMP pathway, with resulting improvements in myocardial function and vasodilation (144). The cardiac myosin activator omecamtiv mecarbil, which potentiates the effects of myosin on actin, exerts positive inotropic effects. Both drugs reduced the composite endpoint of cardiovascular death and hospitalization in HFrEF trials, vericiguat in VICTORIA and omecamtiv mecarbil in GALACTIC-HF (55,56,145147).

THE CELLULAR STRESS HYPOTHESIS

The cellular stress hypothesis (2019-) is the final concept of HF pathophysiology proposed by M. Packer. It filled the substantial gaps in the understanding of HF left by the neurohormonal and the previous hypotheses. It has been inspired by the “game changer” role of SGLT2 inhibitors (SGLT2is). The essence is that cellular dysfunction perpetuates chronic HF. Once again, chance observations, clinical practice, and trials led the way.

Sodium-glucose cotransporters SGLT1 and SGLT2 are mediators of epithelial glucose transport. While SGLT1 accounts for most of the dietary glucose uptake in the intestine, SGLT2 is accountable for the majority of glucose reuptake in the tubular system of the kidney. The medications that inhibit SGLT2 suffix with flozins. The prototype phlorizin was identified in root bark from trees as early as in 1835. Although phlorizin did not show any obvious medicinal value, its blood glucose-lowering and glucosuric effects were described as early as 1886. Only recently (2012) was dapagliflozin introduced as antidiabetic drug. SGLT2is modulate the sodium-glucose cotransporter on the nephron, inhibiting glucose reuptake, inducing glucosuria, and lowering the serum glucose levels. In a way, SGLT2is act as diuretics stimulating osmotic diuresis along with glucosuria and natriuresis. Compared with placebo, SGLT2is reduce HbA1c levels by an average of 0.5-0.8% when used as monotherapy or add-on therapy (along with a modest weight loss). As antidiabetic agents, which was the primary indication, SGLT2is are modestly effective, but they proved to be exceptional in cardiac and renal protection (148151).

The stir caused by rosiglitazone, which was found to increase cardiovascular mortality in post-marketing surveillance, required that new antidiabetic drugs undergo cardiovascular outcome trials prior to FDA approval. This led to the EMPA-REG OUTCOME empagliflozin trial (2015) which unexpectedly revealed a significant reduction in primary composite outcomes of cardiovascular death, nonfatal myocardial infarction, and stroke in the treatment group (152). The serendipitous discovery led to several landmark confirmatory trials on cardiovascular outcomes including CANVAS (canagliflozin, 2017) and DECLARE-TIMI (dapagliflozin, 2019), with positive results, indicating that the favorable outcomes were attributable to a class effect of SGLT2 receptor inhibition. The risk of HF hospitalization was reduced in patients with and without HF history. As the benefits were observed already within 2 months, they were not attributed to the regulation of diabetes but to independent effects of SGLT2i (153,154). Moreover, marked beneficial effects on chronic kidney disease were also observed (155157).

Four landmark trials (two for HFrEF and two for HFpEF) set the standard for HF treatment, establishing the strongest (IA) recommendations for SGLT2is use across the whole EF range (HFrEF, HFmrEF, HFpEF) in HF guidelines (2022) (105,125). Those trials were: DAPA-HF (dapagliflozin 2019), EMPEROR REDUCED (empagliflozin 2020), EMPEROR PRESERVED (empagliflozin 2021), and DELIVER (dapagliflozin 2022). The results for dapagliflozin and empagliflozin were highly congruent: both reduced the composite endpoint of cardiovascular death and HF hospitalization by nearly 20% (158161).

The chance discovery which promoted the modest class of antidiabetic agents to a game changer in HF treatment stirred excitement but also took the cardiac community aback. The mechanisms behind the phenomenon were shrouded in mystery. The upheaval among clinical cardiologists stimulated a huge research effort to provide scientific explanations for the mechanisms of action. Experimental research proposed many explanations, mostly related to cytoprotection. SGLT2is exert cytoprotective effects on the failing heart via SGLT2-independent pathways to increase nutrient-deprivation signaling and autophagic flux, thus reducing cellular stress, improving mitochondrial vitality, and suppressing inflammatory signaling and apoptosis (162).

In the surge of data on the cellular effects of SGLT2, it is hardly possible to concisely outline a unifying concept. Perhaps closest to it is the hypothesis that SGLT2s provide cardiac and renal protection by inducing a state of fasting mimicry through the activation of low-energy sensors, which is not mediated through the SGLT2 protein. This state activates SIRT1/AMPK and suppresses Akt/mTOR signaling, which lead to a reduction in oxidative stress, normalized mitochondrial structure and function, suppression of inflammation, minimization of coronary microvascular injury, enhanced contractile performance, and myocardial protection. SGLT2is promote autophagy, independent of their effects on glucose. SGLT2is might enhance ATP and hemoglobin production by expanding the pool of reactive cytosolic Fe2+. This is due to SGLT2i-induced decline in hepcidin and ferritin levels. which alleviates functional iron deficit mediated by the HF inflammatory milieu (156,162,163).

In the failing heart, glucose transporter type 1 (GLUT1) levels are upregulated, along with excessive glycolysis and defective glucose oxidation, causing cytosolic accumulation of injurious glucose intermediates that activate mTOR and suppress nutrient-deprivation signaling. The uptake of long-chain fatty acids rises, but their oxidation is defective, impairing ATP production and causing cytosolic accumulation of toxic lipid intermediates, worsened by mitochondrial and nutrient-deprivation signaling dysfunction. The ensuing cytosolic accumulation of amino acids activates mTOR. SGLT2is cure the abnormalities in glucose, long-chain fatty acid, and amino acid metabolism by inhibiting GLUT1, stimulating nutrient-deprivation signaling, and restoring mitochondrial vitality. This improves nutrient oxidation and oxidative phosphorylation, preventing cytosolic accumulation of harmful glucose and lipid by-products (162,163).

With regard to HF, fluid accumulation pattern is critical. SGLT2is may differentially regulate the interstitial vs. intravascular compartment when compared with loop diuretics. In congestive HF, interstitial edema is the hallmark of disease. SGLT2 inhibitors may selectively reduce interstitial volume with minimal change in blood volume, whereas loop diuretics reduce both interstitial and intravascular volume. It has been assumed that this differential volume regulation by SGLT2is (interstitial > intravascular) may limit the aberrant reflex neurohormonal stimulation induced by intravascular volume depletion (153).

The crucial question arises how SGLT2is work in HFpEF. The answer is complex, reflecting the heterogeneity of HFpEF phenotypes and the intricacy of SGLT2i actions. Many pieces of this jigsaw puzzle have been put together, noting the considerable overlapping between HFpEF and HFrEF. SGLT2is have been shown to induce a nutrient-deprivation and hypoxic-like transcriptional paradigm, with increased ketosis, erythropoietin, and autophagic flux in addition to altering iron homeostasis, which may improve cardiac energetics and function. These agents also reduce epicardial adipose tissue and modify adipokine signaling attenuating inflammation and oxidative stress. SGLT2is may affect cardiomyocyte ionic homeostasis. Finally, they have been shown to reduce myofilament stiffness as well as extracellular matrix remodeling/fibrosis in the heart, improving diastolic function. The research on the salutary mechanisms of SGLT2is in HFpEF is expected to improve both the understanding of HFpEF and its treatment (164).

The key-message from SGLT2i trials is that cellular stress, unrecognized in previous concepts, is essential for HF pathophysiology. The quest for efficient treatments was the main driver behind this evolution of concepts. The new concept did not repeal but instead revised the previous ones, improving treatment strategies.

Modern HFrEF and partly HFmrEF treatments are based on four pillars (besides diuretics): 1) ACEIs or ARBs, preferably in ARNI combination, 2) BBs, 3) MRAs, and 4) SGLT2is (listed chronologically). With regard to HFpEF, besides omnipresent diuretics, there are SGLT2is which span the entire EF spectrum, rendering EF irrelevant. Quadruple therapy with ARNI, BB, MRAs, and SGLT2is has been established as first-line therapy for patients with HFrEF in current HF guidelines. There is increasing evidence that many patients with HF with an LVEF >40% may benefit from these medications. SGLT2is are beneficial regardless of ejection fraction. HFpEF treatment should be targeted according to the underlying phenotype, comorbidities ant etiology. The hypertensive phenotype often requires ACEIs, ARBs, BBs, and not rarely MRAs, irrespectively of preserved EF, blurring the distinction between HFrEF and HFpEF. Obese and diabetic phenotypes may benefit greatly from GLP1 receptor agonists, in addition to SGLT2i (140,165).

Some inconsistencies may be observed in clinical practice with regard to the use of MRAs in patients with HFpEF. The guidelines neither recommend the use of MRAs in those patients, nor oppose it. The question has been addressed in the 2023 update, since many patients with HFpEF who benefited from SGLT2i received MRAs and other neurohormonal inhibitors. The TOPCAT trial with spironolactone found generally neutral results but with ambiguity in interpretation due to marked regional differences in patient selection. A small trial (RAAM-PEF, 2011) demonstrated that the use of epleronone in patients with HFpEF was associated with significant reduction in markers of collagen turnover and improvement in diastolic function. Thus, it may make sense to prescribe a MRA drug along with a loop diuretic in patients with HFpEF (125,166,167).

A new entity called HF with recovered (or improved) ejection fraction (HFrecEF) has been recently recognized, along with uncertainties about prognosis and treatment (168171). The data from the DELIVER trial indicate that these patients benefit from SGLT2 inhibition (161).

The idea that the failing myocardium recapitulates the fetal signaling program has been generally accepted but has been recently revisited and reinvigorated. The reactivation of fetal beta-myosin heavy chain isoform replacing the mature α variant and the shift from fatty acids to glucose as the main “fuel” in the myocytes of the failing heart are well-known examples. Fetal reprogramming is adaptive in the short time but is deleterious if sustained for long periods (172).

The uridine diphosphate N-acetylglucosamine (UDP-GlcNAc) molecule has been identified recently as a hub of fetal reprogramming in the failing heart. Prolonged increases in glucose uptake in overburdened cardiac myocytes, especially when SIRT1 and AMPK signaling is suppressed, enhances the hexosamine biosynthesis pathway. Its final product, UDP-GlcNAc, acts as a critical nutrient surplus sensor. UDP-GlcNAc is the key step for O-GlcNAcylation, which in collusion with mTOR-mediated phosphorylation rapidly and reversibly modifies a multitude of intracellular proteins in the nucleus, cytoplasm, and mitochondria. The deleterious effects are impaired calcium kinetics with contractile dysfunction, arrhythmias related to activation of voltage-gated sodium channels and Ca2+/CaMKII, mitochondrial dysfunction, maladaptive hypertrophy, fibrosis, and HF. Damage can be prevented by muting of O-GlcNAcylation, which poses a challenge for innovations, including the use of m-RNA technologies (173).

It turns out that chronic HF is heterogeneous disorder defying unifying concepts and treatment approaches. Ejection fraction is only an imperfect tool to distinguish between the phenotypes with different treatment strategies (174). The concept of cellular stress identifies common denominators at a cellular level. This raises hopes that experimental research will create new drugs based on this concept.

Unfortunately, the experience of previous blind alleys in research paths teaches that disappointment is a companion of hope. Positive inotropic agents and nesiritide are the examples. Monoclonal antibodies directed against tumor necrosis factor α (TNFα) etanercept and infliximab raised high hopes in the treatment of HF, considering the role of detrimental inflammatory response in the process of adverse myocardial remodeling. However, the clinical trials RECOVER, RENNESAINCE, and ATTACH dashed these hopes in 2001 (175,176).

Concluding remarks

All that remains is to briefly comment on the evolving understanding of HF from simplistic clinical concepts to the thesis of a complex multicausal and multiorgan disorder. Viewpoints shifted from a mechanistic approach all the way down to the cellular level. Osler viewed HF as a terminal decompensated stage of many cardiac diseases. The diagnosis was based on clinical symptoms and signs alone. The traditional approach prevailed up to the 70s, when cardiac imaging provided hemodynamic quantitation. At the end of century, the neurohormonal concept stimulated the progress of laboratory diagnostics, leading to the launch of the NT-proBNP assay which enabled early recognition and monitoring of HF. Traditionally, the severity of symptoms was graded by NYHA functional classes from I to IV. The A-B-C-D HF staging was introduced in the USA in 2001. It describes the development of HF ranging from risk factors (A), structural heart disease without (B) and with prior or current failure symptoms (C), to refractory HF (D). European guidelines do not use such symbolic numeration but also describe the course of HF comprehensively, ranging from etiological factors to the advanced stage. Such a view contrasts with the historical notion of HF as a terminal edematous state with an ominous prognosis and detrimental clinical course (24,177,178).

This review focused on the pathophysiological concepts of HF in line with pharmacological management. A comprehensive approach includes other aspects, starting with preventive lifestyle changes. Devices, like cardiac resynchronization, physiological pacing, and implantable cardiac defibrillator (ICD), have improved clinical outcomes in addition to the drugs. Older outcome trials with HF drugs, conducted before the ICDs became a standard of care, are difficult to compare with the newer ones. Conversely, new HF treatments appear to reduce the benefit of ICD by decreasing the risk of sudden cardiac death (179,180). Surgical or alternatively percutaneous revascularization has showed hardly any benefits on outcomes in the treatment of HF (HFrEF) due to ischemic heart disease. It may (or should) yet be considered in patients carefully selected on an individual basis (181183). Epigenetic modulation, including mRNA technologies, is offering new prospects for treatment (184,185). Regenerative strategies, which caused enthusiasm in the cardiac community two decades ago but then stalled, have been recently revisited (186190). The modern armamentarium for HF treatment includes a gamut of approaches with cellular drugs and heart transplant at opposite sides of the spectrum. However, heart transplant is a bail-out action, while the drugs are the mainstay.

The diversity of treatment approaches reflects the complexity of HF pathophysiology, unmet needs (HF is still a detrimental disorder), and insufficient understanding of the underlying mechanisms. Reducing HF pathophysiology to the malfunction of the heart pump, with failure of the kidneys to excrete the excess of retained extracellular fluid, proved to be a simplistic concept insufficient for efficient treatment strategies. Adding the hemodynamics of circulation to the concept did not help much either; the resulting pharmacological interventions may have brought some temporary relief but without any impact on the final outcomes. Only a paradigm shift, discovering that the hub of HF pathophysiology is a detrimental systemic neurohormonal response, led to a breakthrough with reduction in adverse cardiovascular outcomes (including mortality) by pharmacological neurohormonal inhibition. HFpEF, where neurohormonal inhibition did not work, exposed the gaps in understanding. Chance discovery of SGLT2is, which are also beneficial for HFpEF and cardiorenal syndrome, highlighted the cellular aspects of HF. Revisiting cellular pathophysiology of the failing heart and the related organs, especially of the kidneys, holds promise to find new avenues of HF treatment. From the viewpoint of experimental research in cardiology, modulation of mitochondrial function may be one of those avenues (191194).

LITERATURE

1 

Katz AM. The “modern” view of heart failure: how did we get here? Circ Heart Fail. 2008 May;1(1):63–71. https://doi.org/10.1161/CIRCHEARTFAILURE.108.772756 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/19808272

2 

Katz AM. Evolving concepts of heart failure: cooling furnace, malfunctioning pump, enlarging muscle. Part II: Hypertrophy and dilatation of the failing heart. J Card Fail. 1998 March;4(1):67–81. https://doi.org/10.1016/S1071-9164(98)90509-7 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/9573505

3 

Osler W. Practice of medicine, designed for the use of practitioners and students of medicine. New York: A. Appleton and Company; 1892. p 634.

4 

Leech CB. An improvement of Southey’s tubes. J Am Med Assoc. 1936;106(22):1895–6. https://doi.org/10.1001/jama.1936.92770220001009a

5 

Jacobs MS. The history of digitalis therapy. Ann Med Hist. 1936;8(6):492–9. PubMed:https://pubmed.ncbi.nlm.nih.gov/33943518/

6 

Hauptman PJ, Kelly RA. Digitalis. Circulation. 1999 March 9;99(9):1265–70. https://doi.org/10.1161/01.CIR.99.9.1265 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/10069797

7 

Wade OL. Digoxin 1785-1985. I. Two hundred years of digitalis. J Clin Hosp Pharm. 1986 February;11(1):3–9. https://doi.org/10.1111/j.1365-2710.1986.tb00822.x PubMed: http://www.ncbi.nlm.nih.gov/pubmed/3514682

8 

Fisch C. William Withering: An account of the foxglove and some of its medical uses 1785-1985. J Am Coll Cardiol. 1985 May;5(5) Suppl A:1A–2A. https://doi.org/10.1016/S0735-1097(85)80456-3 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/3886745

9 

Davis RC, Hobbs FD, Lip GY. ABC of heart failure. History and epidemiology. BMJ. 2000 January 1;320(7226):39–42. https://doi.org/10.1136/bmj.320.7226.39 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/10617530

10 

Ray CT. Mercurial diuretics; their mechanism of action and application. AMA Arch Intern Med. 1958 December;102(6):1016–23. https://doi.org/10.1001/archinte.1958.00260230162020 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/13593909

11 

Moser M, Feig PU. Fifty years of thiazide diuretic therapy for hypertension. Arch Intern Med. 2009 November 9;169(20):1851–6. https://doi.org/10.1001/archinternmed.2009.342 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/19901136

12 

Felker GM, Ellison DH, Mullens W, Cox ZL, Testani JM. Diuretic Therapy for Patients With Heart Failure: JACC State-of-the-Art Review. J Am Coll Cardiol. 2020 March 17;75(10):1178–95. https://doi.org/10.1016/j.jacc.2019.12.059 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32164892

13 

Stokes W. A NEW EFFECTIVE DIURETIC--LASIX. BMJ. 1964 October 10;2(5414):910–4. https://doi.org/10.1136/bmj.2.5414.910 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/14185657

14 

Mullens W, Damman K, Harjola VP, Mebazaa A, Brunner-La Rocca HP, Martens P, et al. The use of diuretics in heart failure with congestion - a position statement from the Heart Failure Association of the European Society of Cardiology. Eur J Heart Fail. 2019 February;21(2):137–55. https://doi.org/10.1002/ejhf.1369 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30600580

15 

Sabbadin C, Calò LA, Armanini D. The story of spironolactones from 1957 to now: from sodium balance to inflammation. G Ital Nefrol. 2016 Feb;33 Suppl 66:33.S66.12. PubMed:https://pubmed.ncbi.nlm.nih.gov/26913880/

16 

Pitt B, Zannad F, Remme WJ, Cody R, Castaigne A, Perez A, et al. The effect of spironolactone on morbidity and mortality in patients with severe heart failure. Randomized Aldactone Evaluation Study Investigators. N Engl J Med. 1999 September 2;341(10):709–17. https://doi.org/10.1056/NEJM199909023411001 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/10471456

17 

Pitt B, Remme W, Zannad F, Neaton J, Martinez F, Roniker B, et al. Eplerenone Post-Acute Myocardial Infarction Heart Failure Efficacy and Survival Study Investigators. Eplerenone, a selective aldosterone blocker, in patients with left ventricular dysfunction after myocardial infarction. N Engl J Med. 2003 April 3;348(14):1309–21. https://doi.org/10.1056/NEJMoa030207 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/12668699

18 

Agarwal R, Filippatos G, Pitt B, Anker SD, Rossing P, Joseph A, et al. FIDELIO-DKD and FIGARO-DKD investigators. Cardiovascular and kidney outcomes with finerenone in patients with type 2 diabetes and chronic kidney disease: the FIDELITY pooled analysis. Eur Heart J. 2022 February 10;43(6):474–84. https://doi.org/10.1093/eurheartj/ehab777 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/35023547

19 

Hampton JR. Results of clinical trials with diuretics in heart failure. Br Heart J. 1994 August;72(2) Suppl:S68–72. https://doi.org/10.1136/hrt.72.2_Suppl.S68 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/7946764

20 

Cody RJ. Clinical trials of diuretic therapy in heart failure: research directions and clinical considerations. J Am Coll Cardiol. 1993 October;22(4) Suppl A:165A–71A. https://doi.org/10.1016/0735-1097(93)90484-I PubMed: http://www.ncbi.nlm.nih.gov/pubmed/8104203

21 

Packer M. Pathophysiology of chronic heart failure. Lancet. 1992 July 11;340(8811):88–92. https://doi.org/10.1016/0140-6736(92)90405-R PubMed: http://www.ncbi.nlm.nih.gov/pubmed/1352022

22 

Packer M. How should physicians view heart failure? The philosophical and physiological evolution of three conceptual models of the disease. Am J Cardiol. 1993 March 25;71(9):3C–11C. https://doi.org/10.1016/0002-9149(93)90081-M PubMed: http://www.ncbi.nlm.nih.gov/pubmed/8465799

23 

Packer M. The neurohormonal hypothesis: a theory to explain the mechanism of disease progression in heart failure. J Am Coll Cardiol. 1992 July;20(1):248–54. https://doi.org/10.1016/0735-1097(92)90167-L PubMed: http://www.ncbi.nlm.nih.gov/pubmed/1351488

24 

Packer M. ESC Rene Laennec Lecture on Clinical Cardiology. In: ESC Congress, 25 Aug 2023.

25 

Digitalis Investigation Group. The effect of digoxin on mortality and morbidity in patients with heart failure. N Engl J Med. 1997 February 20;336(8):525–33. https://doi.org/10.1056/NEJM199702203360801 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/9036306

26 

Digitalis Investigation Group; Ahmed A, Waagstein F, Pitt B, White M, Zannad F, Young JB, et al.. Effectiveness of digoxin in reducing one-year mortality in chronic heart failure in the Digitalis Investigation Group trial. Am J Cardiol. 2009 January 1;103(1):82–7. https://doi.org/10.1016/j.amjcard.2008.06.068 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/19101235

27 

Lopes RD, Rordorf R, De Ferrari GM, Leonardi S, Thomas L, Wojdyla DM, et al. ARISTOTLE Committees and Investigators. Digoxin and Mortality in Patients With Atrial Fibrillation. J Am Coll Cardiol. 2018 March 13;71(10):1063–74. https://doi.org/10.1016/j.jacc.2017.12.060 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/29519345

28 

Aguirre Dávila L, Weber K, Bavendiek U, Bauersachs J, Wittes J, Yusuf S, et al. Digoxin-mortality: randomized vs. observational comparison in the DIG trial. Eur Heart J. 2019 October 21;40(40):3336–41. https://doi.org/10.1093/eurheartj/ehz395 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/31211324

29 

Vamos M, Erath JW, Hohnloser SH. Digoxin-associated mortality: a systematic review and meta-analysis of the literature. Eur Heart J. 2015 July 21;36(28):1831–8. https://doi.org/10.1093/eurheartj/ehv143 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/25939649

30 

Packer M, Meller J. Oral vasodilator therapy for chronic heart failure: a plea for caution. Am J Cardiol. 1978 October;42(4):686–9. https://doi.org/10.1016/0002-9149(78)90642-2 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/358821

31 

Packer M. Vasodilator and inotropic therapy for severe chronic heart failure: passion and skepticism. J Am Coll Cardiol. 1983 November;2(5):841–52. https://doi.org/10.1016/S0735-1097(83)80230-7 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/6138375

32 

Perret Cl, Gardaz JP, Reynaert M, Grimbert F, Enrico JF. Phentolamine for vasodilator therapy in left ventricular failure complicating acute myocardial infarction. Haemodynamic study. Br Heart J. 1975 June;37(6):640–6. https://doi.org/10.1136/hrt.37.6.640 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/1148063

33 

Aronow WS, Lurie M, Turbow M, Whittaker K, Van Camp S, Hughes D. Effect of prazosin vs placebo on chronic left ventricular heart failure. Circulation. 1979 February;59(2):344–50. https://doi.org/10.1161/01.CIR.59.2.344 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/365391

34 

Mettauer B, Rouleau JL, Bichet D, Kortas C, Manzini C, Tremblay G, et al. Differential long-term intrarenal and neurohormonal effects of captopril and prazosin in patients with chronic congestive heart failure: importance of initial plasma renin activity. Circulation. 1986 March;73(3):492–502. https://doi.org/10.1161/01.CIR.73.3.492 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/3512121

35 

Bayliss J, Norell MS, Canepa-Anson R, Reid C, Poole-Wilson P, Sutton G. Clinical importance of the renin-angiotensin system in chronic heart failure: double blind comparison of captopril and prazosin. Br Med J (Clin Res Ed). 1985 June 22;290(6485):1861–5. https://doi.org/10.1136/bmj.290.6485.1861 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/3924285

36 

Rouleau JL, Chatterjee K, Benge W, Parmley WW, Hiramatsu B. Alterations in left ventricular function and coronary hemodynamics with captopril, hydralazine and prazosin in chronic ischemic heart failure: a comparative study. Circulation. 1982 April;65(4):671–8. https://doi.org/10.1161/01.CIR.65.4.671 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/7037220

37 

Loeb HS, Johnson G, Henrick A, Smith R, Wilson J, Cremo R, et al. Effect of enalapril, hydralazine plus isosorbide dinitrate, and prazosin on hospitalization in patients with chronic congestive heart failure. The V-HeFT VA Cooperative Studies Group. Circulation. 1993 Jun;87(6 Suppl):VI78-87. PubMed:https://pubmed.ncbi.nlm.nih.gov/8500244/

38 

Cole RT, Kalogeropoulos AP, Georgiopoulou VV, Gheorghiade M, Quyyumi A, Yancy C, et al. Hydralazine and isosorbide dinitrate in heart failure: historical perspective, mechanisms, and future directions. Circulation. 2011 May 31;123(21):2414–22. https://doi.org/10.1161/CIRCULATIONAHA.110.012781 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/21632515

39 

Elkayam U, Shotan A, Mehra A, Ostrzega E. Calcium channel blockers in heart failure. J Am Coll Cardiol. 1993 October;22(4) Suppl A:139A–44A. https://doi.org/10.1016/0735-1097(93)90478-J PubMed: http://www.ncbi.nlm.nih.gov/pubmed/8376684

40 

Elkayam U. Calcium channel blockers in heart failure. Cardiology. 1998;89 Suppl 1:38–46. https://doi.org/10.1159/000047278 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/9570428

41 

Konstam MA, Smith JJ, Patten R, Udelson JE. Calcium channel blockers in heart failure: help or hindrance? J Card Fail. 1996 December;2(4) Suppl:S251–7. https://doi.org/10.1016/S1071-9164(96)80085-6 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/8951587

42 

Packer M. Pathophysiological mechanisms underlying the adverse effects of calcium channel-blocking drugs in patients with chronic heart failure. Circulation. 1989 Dec;80(6 Suppl):IV59-67. PubMed:https://pubmed.ncbi.nlm.nih.gov/2688986/

43 

Packer M, Kessler PD, Lee WH. Calcium-channel blockade in the management of severe chronic congestive heart failure: a bridge too far. Circulation. 1987 Jun;75(6 Pt 2):V56-64. PubMed:https://pubmed.ncbi.nlm.nih.gov/3552317/

44 

Arrigo M, Jessup M, Mullens W, Reza N, Shah AM, Sliwa K, et al. Acute heart failure. Nat Rev Dis Primers. 2020 March 5;6(1):16. https://doi.org/10.1038/s41572-020-0151-7 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32139695

45 

Metra M, Teerlink JR, Voors AA, Felker GM, Milo-Cotter O, Weatherley B, et al. Vasodilators in the treatment of acute heart failure: what we know, what we don’t. Heart Fail Rev. 2009 December;14(4):299–307. https://doi.org/10.1007/s10741-008-9127-5 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/19096932

46 

McDonagh TA, Metra M, Adamo M, Gardner RS, Baumbach A, Böhm M, et al. ESC Scientific Document Group. 2021 ESC Guidelines for the diagnosis and treatment of acute and chronic heart failure. Eur Heart J. 2021 September 21;42(36):3599–726. https://doi.org/10.1093/eurheartj/ehab368 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/34447992

47 

Mullens W, Abrahams Z, Francis GS, Skouri HN, Starling RC, Young JB, et al. Sodium nitroprusside for advanced low-output heart failure. J Am Coll Cardiol. 2008 July 15;52(3):200–7. https://doi.org/10.1016/j.jacc.2008.02.083 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/18617068

48 

Ahmad T, Miller PE, McCullough M, Desai NR, Riello R, Psotka M, et al. Why has positive inotropy failed in chronic heart failure? Lessons from prior inotrope trials. Eur J Heart Fail. 2019 September;21(9):1064–78. https://doi.org/10.1002/ejhf.1557 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/31407860

49 

Packer M. The development of positive inotropic agents for chronic heart failure: how have we gone astray? J Am Coll Cardiol. 1993 October;22(4) Suppl A:119A–26A. https://doi.org/10.1016/0735-1097(93)90474-F PubMed: http://www.ncbi.nlm.nih.gov/pubmed/8397231

50 

Petersen JW, Felker GM. Inotropes in the management of acute heart failure. Crit Care Med. 2008 January;36(1) Suppl:S106–11. https://doi.org/10.1097/01.CCM.0000296273.72952.39 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/18158469

51 

Amin A, Maleki M. Positive inotropes in heart failure: a review article. Heart Asia. 2012 January 1;4(1):16–22. https://doi.org/10.1136/heartasia-2011-010068 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27326019

52 

Belletti A, Castro ML, Silvetti S, Greco T, Biondi-Zoccai G, Pasin L, et al. The Effect of inotropes and vasopressors on mortality: a meta-analysis of randomized clinical trials. Br J Anaesth. 2015 November;115(5):656–75. https://doi.org/10.1093/bja/aev284 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/26475799

53 

Bistola V, Arfaras-Melainis A, Polyzogopoulou E, Ikonomidis I, Parissis J. Inotropes in Acute Heart Failure: From Guidelines to Practical Use: Therapeutic Options and Clinical Practice. Card Fail Rev. 2019 November 4;5(3):133–9. https://doi.org/10.15420/cfr.2019.11.2 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/31768269

54 

Francis GS, Bartos JA, Adatya S. Inotropes. J Am Coll Cardiol. 2014 May 27;63(20):2069–78. https://doi.org/10.1016/j.jacc.2014.01.016 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/24530672

55 

Teerlink JR, Diaz R, Felker GM, McMurray JJV, Metra M, Solomon SD, et al. GALACTIC-HF Investigators. Cardiac Myosin Activation with Omecamtiv Mecarbil in Systolic Heart Failure. N Engl J Med. 2021 January 14;384(2):105–16. https://doi.org/10.1056/NEJMoa2025797 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/33185990

56 

Cleland JG, Teerlink JR, Senior R, Nifontov EM, Mc Murray JJ, Lang CC, et al. The effects of the cardiac myosin activator, omecamtiv mecarbil, on cardiac function in systolic heart failure: a double-blind, placebo-controlled, crossover, dose-ranging phase 2 trial. Lancet. 2011 August 20;378(9792):676–83. https://doi.org/10.1016/S0140-6736(11)61126-4 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/21856481

57 

Katz AM. Influence of altered inotropy and lusitropy on ventricular pressure-volume loops. J Am Coll Cardiol. 1988 February;11(2):438–45. https://doi.org/10.1016/0735-1097(88)90113-1 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/3276755

58 

Gibson DG, Brown D. Measurement of instantaneous left ventricular volumes and filling rate in man by echocardiography. Br Heart J. 1973 May;35(5):559. PubMed:https://pubmed.ncbi.nlm.nih.gov/4716043/

59 

Upton MT, Gibson DG, Brown DJ. Echocardiographic assessment of abnormal left ventricular relaxation in man. Br Heart J. 1976 October;38(10):1001–9. https://doi.org/10.1136/hrt.38.10.1001 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/973873

60 

Gaasch WH, Levine HJ, Quinones MA, Alexander JK. Left ventricular compliance: mechanisms and clinical implications. Am J Cardiol. 1976 November 4;38(5):645–53. https://doi.org/10.1016/S0002-9149(76)80015-X PubMed: http://www.ncbi.nlm.nih.gov/pubmed/136186

61 

Ng KS, Gibson DG. Impairment of diastolic function by shortened filling period in severe left ventricular disease. Br Heart J. 1989 October;62(4):246–52. https://doi.org/10.1136/hrt.62.4.246 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/2803869

62 

Vasan RS, Benjamin EJ, Levy D. Prevalence, clinical features and prognosis of diastolic heart failure: an epidemiologic perspective. J Am Coll Cardiol. 1995 December;26(7):1565–74. https://doi.org/10.1016/0735-1097(95)00381-9 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/7594087

63 

Yamamoto K, Redfield MM, Nishimura RA. Analysis of left ventricular diastolic function. Heart. 1996 June;75(6) Suppl 2:27–35. https://doi.org/10.1136/hrt.75.6_Suppl_2.27 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/8785701

64 

Vasan RS, Levy D. Defining diastolic heart failure: a call for standardized diagnostic criteria. Circulation. 2000 May 2;101(17):2118–21. https://doi.org/10.1161/01.CIR.101.17.2118 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/10790356

65 

van Kraaij DJ, van Pol PE, Ruiters AW, de Swart JB, Lips DJ, Lencer N, et al. Diagnosing diastolic heart failure. Eur J Heart Fail. 2002 August;4(4):419–30. https://doi.org/10.1016/S1388-9842(02)00020-X PubMed: http://www.ncbi.nlm.nih.gov/pubmed/12167379

66 

Jorge García M. Diagnóstico y guía terapéutica de la insuficiencia cardíaca diastólica [Diagnosis and therapeutic guidance of diastolic heart failure]. Rev Esp Cardiol. 2003 Apr;56(4):396-406. Spanish. https://doi.org/10.1016/S0300-8932(03)76884-5 https://doi.org/10.1016/S0300-8932(03)76884-5

67 

Zile MR, Baicu CF, Gaasch WH. Diastolic heart failure--abnormalities in active relaxation and passive stiffness of the left ventricle. N Engl J Med. 2004 May 6;350(19):1953–9. https://doi.org/10.1056/NEJMoa032566 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/15128895

68 

Bursi F, Weston SA, Redfield MM, Jacobsen SJ, Pakhomov S, Nkomo VT, et al. Systolic and diastolic heart failure in the community. JAMA. 2006 November 8;296(18):2209–16. https://doi.org/10.1001/jama.296.18.2209 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/17090767

69 

Kane GC, Karon BL, Mahoney DW, Redfield MM, Roger VL, Burnett JC Jr, et al. Progression of left ventricular diastolic dysfunction and risk of heart failure. JAMA. 2011 August 24;306(8):856–63. https://doi.org/10.1001/jama.2011.1201 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/21862747

70 

Dahlöf B, Devereux RB, Kjeldsen SE, Julius S, Beevers G, de Faire U, et al. LIFE Study Group. Cardiovascular morbidity and mortality in the Losartan Intervention For Endpoint reduction in hypertension study (LIFE): a randomised trial against atenolol. Lancet. 2002 March 23;359(9311):995–1003. https://doi.org/10.1016/S0140-6736(02)08089-3 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/11937178

71 

Greve AM, Olsen MH, Bella JN, Lønnebakken MT, Gerdts E, Okin PM, et al. Contrasting hemodynamic mechanisms of losartan- vs. atenolol-based antihypertensive treatment: a LIFE study. Am J Hypertens. 2012 September;25(9):1017–23. https://doi.org/10.1038/ajh.2012.81 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/22695506

72 

Oh JK, Miranda WR, Bird JG, Kane GC, Nagueh SF. The 2016 Diastolic Function Guideline: Is it Already Time to Revisit or Revise Them? JACC Cardiovasc Imaging. 2020 January;13(1 Pt 2):327–35. https://doi.org/10.1016/j.jcmg.2019.12.004 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/31918901

73 

Nagueh SF. Left Ventricular Diastolic Function: Understanding Pathophysiology, Diagnosis, and Prognosis With Echocardiography. JACC Cardiovasc Imaging. 2020 January;13(1 Pt 2):228–44. https://doi.org/10.1016/j.jcmg.2018.10.038 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30982669

74 

Senni M, Redfield MM. Heart failure with preserved systolic function. A different natural history? J Am Coll Cardiol. 2001 November 1;38(5):1277–82. https://doi.org/10.1016/S0735-1097(01)01567-4 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/11691495

75 

Burkhoff D, Maurer MS, Packer M. Heart failure with a normal ejection fraction: is it really a disorder of diastolic function? Circulation. 2003 February 11;107(5):656–8. https://doi.org/10.1161/01.CIR.0000053947.82595.03 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/12578861

76 

Maurer MS. Heart failure with a normal ejection fraction (HFNEF): embracing complexity. J Card Fail. 2009 September;15(7):561–4. https://doi.org/10.1016/j.cardfail.2009.04.004 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/19700131

77 

Lewis GA, Schelbert EB, Williams SG, Cunnington C, Ahmed F, McDonagh TA, et al. Biological Phenotypes of Heart Failure With Preserved Ejection Fraction. J Am Coll Cardiol. 2017 October 24;70(17):2186–200. https://doi.org/10.1016/j.jacc.2017.09.006 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/29050567

78 

Sanderson JE. HFNEF, HFpEF, HF-PEF, or DHF: what is in an acronym? JACC Heart Fail. 2014 February;2(1):93–4. https://doi.org/10.1016/j.jchf.2013.09.006 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/24622122

79 

Sanderson JE. Heart failure with a normal ejection fraction. Heart. 2007 February;93(2):155–8. https://doi.org/10.1136/hrt.2005.074187 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/16387829

80 

Parikh KS, Sharma K, Fiuzat M, Surks HK, George JT, Honarpour N, et al. Heart Failure With Preserved Ejection Fraction Expert Panel Report: Current Controversies and Implications for Clinical Trials. JACC Heart Fail. 2018 August;6(8):619–32. https://doi.org/10.1016/j.jchf.2018.06.008 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30071950

81 

Mishra S, Kass DA. Cellular and molecular pathobiology of heart failure with preserved ejection fraction. Nat Rev Cardiol. 2021 June;18(6):400–23. https://doi.org/10.1038/s41569-020-00480-6 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/33432192

82 

Cleland JG, Dargie HJ, Hodsman GP, Ball SG, Robertson JI, Morton JJ, et al. Captopril in heart failure. A double blind controlled trial. Br Heart J. 1984 November;52(5):530–5. https://doi.org/10.1136/hrt.52.5.530 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/6388612

83 

CONSENSUS Trial Study Group. Effects of enalapril on mortality in severe congestive heart failure. Results of the Cooperative North Scandinavian Enalapril Survival Study (CONSENSUS). N Engl J Med. 1987 June 4;316(23):1429–35. https://doi.org/10.1056/NEJM198706043162301 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/2883575

84 

Cohn JN, Johnson G, Ziesche S, Cobb F, Francis G, Tristani F, et al. A comparison of enalapril with hydralazine-isosorbide dinitrate in the treatment of chronic congestive heart failure. N Engl J Med. 1991 August 1;325(5):303–10. https://doi.org/10.1056/NEJM199108013250502 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/2057035

85 

Francis GS, Cohn JN, Johnson G, Rector TS, Goldman S, Simon A. Plasma norepinephrine, plasma renin activity, and congestive heart failure. Relations to survival and the effects of therapy in V-HeFT II. The V-HeFT VA Cooperative Studies Group. Circulation. 1993 Jun;87(6 Suppl):VI40-8. PubMed:https://pubmed.ncbi.nlm.nih.gov/8500238/

86 

SOLVD Investigators; Yusuf S, Pitt B, Davis CE, Hood WB, Cohn JN. Effect of enalapril on survival in patients with reduced left ventricular ejection fractions and congestive heart failure. N Engl J Med. 1991 August 1;325(5):293–302. https://doi.org/10.1056/NEJM199108013250501 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/2057034

87 

Pfeffer MA, Braunwald E, Moyé LA, Basta L, Brown EJ Jr, Cuddy TE, et al. Effect of captopril on mortality and morbidity in patients with left ventricular dysfunction after myocardial infarction. Results of the survival and ventricular enlargement trial. The SAVE Investigators. N Engl J Med. 1992 September 3;327(10):669–77. https://doi.org/10.1056/NEJM199209033271001 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/1386652

88 

Effect of ramipril on mortality and morbidity of survivors of acute myocardial infarction with clinical evidence of heart failure. The Acute Infarction Ramipril Efficacy (AIRE) Study Investigators. Lancet. 1993 Oct 2;342(8875):821-8. PubMed:https://pubmed.ncbi.nlm.nih.gov/8104270/

89 

Køber L, Torp-Pedersen C, Carlsen JE, Bagger H, Eliasen P, Lyngborg K, et al. A clinical trial of the angiotensin-converting-enzyme inhibitor trandolapril in patients with left ventricular dysfunction after myocardial infarction. Trandolapril Cardiac Evaluation (TRACE) Study Group. N Engl J Med. 1995 December 21;333(25):1670–6. https://doi.org/10.1056/NEJM199512213332503 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/7477219

90 

Patel S, Rauf A, Khan H, Abu-Izneid T. Renin-angiotensin-aldosterone (RAAS): The ubiquitous system for homeostasis and pathologies. Biomed Pharmacother. 2017 October;94:317–25. https://doi.org/10.1016/j.biopha.2017.07.091 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28772209

91 

Muñoz-Durango N, Fuentes CA, Castillo AE, González-Gómez LM, Vecchiola A, Fardella CE, et al. Role of the Renin-Angiotensin-Aldosterone System beyond Blood Pressure Regulation: Molecular and Cellular Mechanisms Involved in End-Organ Damage during Arterial Hypertension. Int J Mol Sci. 2016 June 23;17(7):797. https://doi.org/10.3390/ijms17070797 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27347925

92 

Hartupee J, Mann DL. Neurohormonal activation in heart failure with reduced ejection fraction. Nat Rev Cardiol. 2017 January;14(1):30–8. https://doi.org/10.1038/nrcardio.2016.163 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27708278

93 

Fournier D, Luft FC, Bader M, Ganten D, Andrade-Navarro MA. Emergence and evolution of the renin-angiotensin-aldosterone system. J Mol Med (Berl). 2012 May;90(5):495–508. https://doi.org/10.1007/s00109-012-0894-z PubMed: http://www.ncbi.nlm.nih.gov/pubmed/22527880

94 

Katz AM. Angiotensin II: hemodynamic regulator or growth factor? J Mol Cell Cardiol. 1990 July;22(7):739–47. https://doi.org/10.1016/0022-2828(90)90086-H PubMed: http://www.ncbi.nlm.nih.gov/pubmed/2231742

95 

Maggioni AP. Heart Failure: Treatment strategies for heart failure: beta blockers and antiarrhythmics. Heart. 2001 January;85(1):97–103. https://doi.org/10.1136/heart.85.1.97 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/11119476

96 

Waagstein F, Hjalmarson A, Varnauskas E, Wallentin I. Effect of chronic beta-adrenergic receptor blockade in congestive cardiomyopathy. Br Heart J. 1975 October;37(10):1022–36. https://doi.org/10.1136/hrt.37.10.1022 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/1191416

97 

Nicholls M. Finn Waagstein MD PhD: Finn Waagstein reflects on his work and research which saw beta-blockers revolutionise the treatment of heart failure patients. Mark Nicholls reports. Eur Heart J. 2017 November 21;38(44):3249–50. https://doi.org/10.1093/eurheartj/ehx618 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/29177413

98 

Waagstein F, Rutherford JD. The Evolution of the Use of β-Blockers to Treat Heart Failure: A Conversation With Finn Waagstein, MD. Circulation. 2017 September 5;136(10):889–93. https://doi.org/10.1161/CIRCULATIONAHA.117.029934 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28874420

99 

A randomized trial of beta-blockade in heart failure. The Cardiac Insufficiency Bisoprolol Study (CIBIS). CIBIS Investigators and Committees. Circulation. 1994 October;90(4):1765–73. https://doi.org/10.1161/01.CIR.90.4.1765 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/7923660

100 

The Cardiac Insufficiency Bisoprolol Study II (CIBIS-II). a randomised trial. Lancet. 1999 Jan 2;353(9146):9-13. PubMed:https://pubmed.ncbi.nlm.nih.gov/10023943/

101 

Erdmann E, Lechat P, Verkenne P, Wiemann H. Results from post-hoc analyses of the CIBIS II trial: effect of bisoprolol in high-risk patient groups with chronic heart failure. Eur J Heart Fail. 2001 August;3(4):469–79. https://doi.org/10.1016/S1388-9842(01)00174-X PubMed: http://www.ncbi.nlm.nih.gov/pubmed/11511434

102 

Packer M, Bristow MR, Cohn JN, Colucci WS, Fowler MB, Gilbert EM, et al. The effect of carvedilol on morbidity and mortality in patients with chronic heart failure. U.S. Carvedilol Heart Failure Study Group. N Engl J Med. 1996 May 23;334(21):1349–55. https://doi.org/10.1056/NEJM199605233342101 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/8614419

103 

Hjalmarson A, Goldstein S, Fagerberg B, Wedel H, Waagstein F, Kjekshus J, et al. Effects of controlled-release metoprolol on total mortality, hospitalizations, and well-being in patients with heart failure: the Metoprolol CR/XL Randomized Intervention Trial in congestive heart failure (MERIT-HF). MERIT-HF Study Group. JAMA. 2000 March 8;283(10):1295–302. https://doi.org/10.1001/jama.283.10.1295 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/10714728

104 

Flather MD, Shibata MC, Coats AJ, Van Veldhuisen DJ, Parkhomenko A, Borbola J, et al. SENIORS Investigators. Randomized trial to determine the effect of nebivolol on mortality and cardiovascular hospital admission in elderly patients with heart failure (SENIORS). Eur Heart J. 2005 February;26(3):215–25. https://doi.org/10.1093/eurheartj/ehi115 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/15642700

105 

Heidenreich PA, Bozkurt B, Aguilar D, Allen LA, Byun JJ, Colvin MM, 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 May 3;145(18):e895–1032. https://doi.org/10.1161/CIR.0000000000001063 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/35363499

106 

Struthers AD. Aldosterone blockade in heart failure. J Renin Angiotensin Aldosterone Syst. 2004 September;5 Suppl 1:S23–7. https://doi.org/10.3317/JRAAS.2004.021 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/15526239

107 

Struthers AD. Angiotensin blockade or aldosterone blockade as the third neuroendocrine-blocking drug in mild but symptomatic heart failure patients. Heart. 2006 December;92(12):1728–31. https://doi.org/10.1136/hrt.2005.068668 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/16339814

108 

Timmermans PB, Duncia JV, Carini DJ, Chiu AT, Wong PC, Wexler RR, et al. Discovery of losartan, the first angiotensin II receptor antagonist. J Hum Hypertens. 1995 Nov;9 Suppl 5:S3-18. PubMed:https://pubmed.ncbi.nlm.nih.gov/8583479/

109 

Brenner BM, Cooper ME, de Zeeuw D, Keane WF, Mitch WE, Parving HH, et al. RENAAL Study Investigators. Effects of losartan on renal and cardiovascular outcomes in patients with type 2 diabetes and nephropathy. N Engl J Med. 2001 September 20;345(12):861–9. https://doi.org/10.1056/NEJMoa011161 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/11565518

110 

Pitt B, Segal R, Martinez FA, Meurers G, Cowley AJ, Thomas I, et al. Randomised trial of losartan versus captopril in patients over 65 with heart failure (Evaluation of Losartan in the Elderly Study, ELITE). Lancet. 1997 March 15;349(9054):747–52. https://doi.org/10.1016/S0140-6736(97)01187-2 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/9074572

111 

Pitt B, Poole-Wilson PA, Segal R, Martinez FA, Dickstein K, Camm AJ, et al. Effect of losartan compared with captopril on mortality in patients with symptomatic heart failure: randomised trial--the Losartan Heart Failure Survival Study ELITE II. Lancet. 2000 May 6;355(9215):1582–7. https://doi.org/10.1016/S0140-6736(00)02213-3 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/10821361

112 

Pitt B. Evaluation of Losartan in the Elderly (ELITE) Trial: clinical implications. Eur Heart J. 1997 August;18(8):1197–9. https://doi.org/10.1093/oxfordjournals.eurheartj.a015425 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/9458406

113 

Dickstein K, Kjekshus J; OPTIMAAL Steering Committee of the OPTIMAAL Study Group. Effects of losartan and captopril on mortality and morbidity in high-risk patients after acute myocardial infarction: the OPTIMAAL randomised trial. Optimal Trial in Myocardial Infarction with Angiotensin II Antagonist Losartan. Lancet. 2002 September 7;360(9335):752–60. https://doi.org/10.1016/S0140-6736(02)09895-1 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/12241832

114 

Cohn JN, Tognoni G; Valsartan Heart Failure Trial Investigators. A randomized trial of the angiotensin-receptor blocker valsartan in chronic heart failure. N Engl J Med. 2001 December 6;345(23):1667–75. https://doi.org/10.1056/NEJMoa010713 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/11759645

115 

Granger CB, McMurray JJ, Yusuf S, Held P, Michelson EL, Olofsson B, et al. CHARM Investigators and Committees. Effects of candesartan in patients with chronic heart failure and reduced left-ventricular systolic function intolerant to angiotensin-converting-enzyme inhibitors: the CHARM-Alternative trial. Lancet. 2003 September 6;362(9386):772–6. https://doi.org/10.1016/S0140-6736(03)14284-5 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/13678870

116 

McMurray JJ, Ostergren J, Swedberg K, Granger CB, Held P, Michelson EL, et al. CHARM Investigators and Committees. Effects of candesartan in patients with chronic heart failure and reduced left-ventricular systolic function taking angiotensin-converting-enzyme inhibitors: the CHARM-Added trial. Lancet. 2003 September 6;362(9386):767–71. https://doi.org/10.1016/S0140-6736(03)14283-3 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/13678869

117 

Weir RA, McMurray JJ, Puu M, Solomon SD, Olofsson B, Granger CB, et al. CHARM Investigators. Efficacy and tolerability of adding an angiotensin receptor blocker in patients with heart failure already receiving an angiotensin-converting inhibitor plus aldosterone antagonist, with or without a beta blocker. Findings from the Candesartan in Heart failure: Assessment of Reduction in Mortality and morbidity (CHARM)-Added trial. Eur J Heart Fail. 2008 February;10(2):157–63. https://doi.org/10.1016/j.ejheart.2007.12.006 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/18242128

118 

Kiernan MS, Wentworth D, Francis G, Martinez FA, Dickstein K, Komajda M, et al. Predicting adverse events during angiotensin receptor blocker treatment in heart failure: results from the HEAAL trial. Eur J Heart Fail. 2012 December;14(12):1401–9. https://doi.org/10.1093/eurjhf/hfs145 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/23187648

119 

Parker AB, Azevedo ER, Baird MG, Smith SJ, Arnold JM, Humen DP, et al. ARCTIC: assessment of haemodynamic response in patients with congestive heart failure to telmisartan: a multicentre dose-ranging study in Canada. Am Heart J. 1999 November;138(5 Pt 1):843–8. https://doi.org/10.1016/S0002-8703(99)70008-6 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/10539814

120 

Yusuf S, Pfeffer MA, Swedberg K, Granger CB, Held P, McMurray JJ, et al. CHARM Investigators and Committees. Effects of candesartan in patients with chronic heart failure and preserved left-ventricular ejection fraction: the CHARM-Preserved Trial. Lancet. 2003 September 6;362(9386):777–81. https://doi.org/10.1016/S0140-6736(03)14285-7 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/13678871

121 

Massie BM, Carson PE, McMurray JJ, Komajda M, McKelvie R, Zile MR, et al. I-PRESERVE Investigators. Irbesartan in patients with heart failure and preserved ejection fraction. N Engl J Med. 2008 December 4;359(23):2456–67. https://doi.org/10.1056/NEJMoa0805450 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/19001508

122 

Hubers SA, Brown NJ. Combined Angiotensin Receptor Antagonism and Neprilysin Inhibition. Circulation. 2016 March 15;133(11):1115–24. https://doi.org/10.1161/CIRCULATIONAHA.115.018622 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/26976916

123 

McMurray JJ, Packer M, Desai AS, Gong J, Lefkowitz MP, Rizkala AR, et al. PARADIGM-HF Investigators and Committees. Angiotensin-neprilysin inhibition versus enalapril in heart failure. N Engl J Med. 2014 September 11;371(11):993–1004. https://doi.org/10.1056/NEJMoa1409077 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/25176015

124 

Docherty KF, Vaduganathan M, Solomon SD, McMurray JJV. Sacubitril/Valsartan: Neprilysin Inhibition 5 Years After PARADIGM-HF. JACC Heart Fail. 2020 October;8(10):800–10. https://doi.org/10.1016/j.jchf.2020.06.020 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/33004114

125 

McDonagh TA, Metra M, Adamo M, Gardner RS, Baumbach A, Böhm M, et al. ESC Scientific Document Group. 2023 Focused Update of the 2021 ESC Guidelines for the diagnosis and treatment of acute and chronic heart failure. Eur Heart J. 2023 October 1;44(37):3627–39. https://doi.org/10.1093/eurheartj/ehad195 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/37622666

126 

Inoue K, Naruse K, Yamagami S, Mitani H, Suzuki N, Takei Y. Four functionally distinct C-type natriuretic peptides found in fish reveal evolutionary history of the natriuretic peptide system. Proc Natl Acad Sci USA. 2003 August 19;100(17):10079–84. https://doi.org/10.1073/pnas.1632368100 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/12893874

127 

Fu S, Ping P, Wang F, Luo L. Synthesis, secretion, function, metabolism and application of natriuretic peptides in heart failure. J Biol Eng. 2018 January 12;12:2. https://doi.org/10.1186/s13036-017-0093-0 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/29344085

128 

Wylie JV, Tsao L. Nesiritide for the treatment of decompensated heart failure. Expert Rev Cardiovasc Ther. 2004 November;2(6):803–13. https://doi.org/10.1586/14779072.2.6.803 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/15500426

129 

Topol EJ. Nesiritide - not verified. N Engl J Med. 2005 July 14;353(2):113–6. https://doi.org/10.1056/NEJMp058139 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/16014879

130 

O’Connor CM, Starling RC, Hernandez AF, Armstrong PW, Dickstein K, Hasselblad V, et al. Effect of nesiritide in patients with acute decompensated heart failure. N Engl J Med. 2011 July 7;365(1):32–43. https://doi.org/10.1056/NEJMoa1100171 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/21732835

131 

Egom EE. BNP and Heart Failure: Preclinical and Clinical Trial Data. J Cardiovasc Transl Res. 2015 April;8(3):149–57. https://doi.org/10.1007/s12265-015-9619-3 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/25771949

132 

Kittleson MM. Nesiritide and Me. Circ Heart Fail. 2018 August;11(8):e005440. https://doi.org/10.1161/CIRCHEARTFAILURE.118.005440 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30354567

133 

Psotka MA, Teerlink JR. Ivabradine: Role in the Chronic Heart Failure Armamentarium. Circulation. 2016 May 24;133(21):2066–75. https://doi.org/10.1161/CIRCULATIONAHA.115.018094 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27217432

134 

Orasanu G, Al-Kindi SG, Oliveira GH. Ivabradine in Management of Heart Failure: a Critical Appraisal. Curr Heart Fail Rep. 2016 February;13(1):60–9. https://doi.org/10.1007/s11897-016-0276-x PubMed: http://www.ncbi.nlm.nih.gov/pubmed/26797824

135 

Lam CSP, Chandramouli C. Fat, Female, Fatigued: Features of the Obese HFpEF Phenotype. JACC Heart Fail. 2018 August;6(8):710–3. https://doi.org/10.1016/j.jchf.2018.06.006 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30078394

136 

Pitt B, Pfeffer MA, Assmann SF, Boineau R, Anand IS, Claggett B, et al. TOPCAT Investigators. Spironolactone for heart failure with preserved ejection fraction. N Engl J Med. 2014 April 10;370(15):1383–92. https://doi.org/10.1056/NEJMoa1313731 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/24716680

137 

Liang M, Bian B, Yang Q. Characteristics and long-term prognosis of patients with reduced, mid-range, and preserved ejection fraction: A systemic review and meta-analysis. Clin Cardiol. 2022 January;45(1):5–17. https://doi.org/10.1002/clc.23754 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/35043472

138 

Komajda M, Lam CS. Heart failure with preserved ejection fraction: a clinical dilemma. Eur Heart J. 2014 April;35(16):1022–32. https://doi.org/10.1093/eurheartj/ehu067 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/24618346

139 

Talha KM, Butler J. Breakthroughs in the treatment of heart failure with mildly reduced and preserved ejection fraction. Clin Cardiol. 2022 Jun;45 Suppl 1(Suppl 1):S31-S39. https://doi.org/10.1002/clc.23846 https://doi.org/10.1002/clc.23846

140 

Senni M, Paulus WJ, Gavazzi A, Fraser AG, Díez J, Solomon SD, et al. New strategies for heart failure with preserved ejection fraction: the importance of targeted therapies for heart failure phenotypes. Eur Heart J. 2014 October 21;35(40):2797–815. https://doi.org/10.1093/eurheartj/ehu204 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/25104786

141 

Lam CS, Teng TH. Understanding Heart Failure With Mid-Range Ejection Fraction. JACC Heart Fail. 2016 June;4(6):473–6. https://doi.org/10.1016/j.jchf.2016.03.025 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27256750

142 

Lam CS, Solomon SD. The middle child in heart failure: heart failure with mid-range ejection fraction (40-50%). Eur J Heart Fail. 2014 October;16(10):1049–55. https://doi.org/10.1002/ejhf.159 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/25210008

143 

Savarese G, Stolfo D, Sinagra G, Lund LH. Heart failure with mid-range or mildly reduced ejection fraction. Nat Rev Cardiol. 2022 February;19(2):100–16. https://doi.org/10.1038/s41569-021-00605-5 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/34489589

144 

Trujillo ME, Ayalasomayajula S, Blaustein RO, Gheyas F. Vericiguat, a novel sGC stimulator: Mechanism of action, clinical, and translational science. Clin Transl Sci. 2023 December;16(12):2458–66. https://doi.org/10.1111/cts.13677 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/37997225

145 

Armstrong PW, Pieske B, Anstrom KJ, Ezekowitz J, Hernandez AF, Butler J, et al. VICTORIA Study Group. Vericiguat in Patients with Heart Failure and Reduced Ejection Fraction. N Engl J Med. 2020 May 14;382(20):1883–93. https://doi.org/10.1056/NEJMoa1915928 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32222134

146 

Teerlink JR, Diaz R, Felker GM, McMurray JJV, Metra M, Solomon SD, et al. Omecamtiv Mecarbil in Chronic Heart Failure With Reduced Ejection Fraction: Rationale and Design of GALACTIC-HF. JACC Heart Fail. 2020 April;8(4):329–40. https://doi.org/10.1016/j.jchf.2019.12.001 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32035892

147 

Felker GM, Solomon SD, Claggett B, Diaz R, McMurray JJV, Metra M, et al. Assessment of Omecamtiv Mecarbil for the Treatment of Patients With Severe Heart Failure: A Post Hoc Analysis of Data From the GALACTIC-HF Randomized Clinical Trial. JAMA Cardiol. 2022 January 1;7(1):26–34. https://doi.org/10.1001/jamacardio.2021.4027 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/34643642

148 

Rieg T, Vallon V. Development of SGLT1 and SGLT2 inhibitors. Diabetologia. 2018 October;61(10):2079–86. https://doi.org/10.1007/s00125-018-4654-7 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30132033

149 

Wilcox T, De Block C, Schwartzbard AZ, Newman JD. Diabetic Agents, From Metformin to SGLT2 Inhibitors and GLP1 Receptor Agonists: JACC Focus Seminar. J Am Coll Cardiol. 2020 April 28;75(16):1956–74. https://doi.org/10.1016/j.jacc.2020.02.056 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32327107

150 

Helvacı Ö, Helvacı B. A Story of Serendipities: From Phlorizin to Gliflozins. Exp Clin Transplant. 2023 June;21 Suppl 2:105–8. https://doi.org/10.6002/ect.IAHNCongress.25 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/37496357

151 

Newman JD, Vani AK, Aleman JO, Weintraub HS, Berger JS, Schwartzbard AZ. The Changing Landscape of Diabetes Therapy for Cardiovascular Risk Reduction: JACC State-of-the-Art Review. J Am Coll Cardiol. 2018 October 9;72(15):1856–69. https://doi.org/10.1016/j.jacc.2018.07.071 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30286929

152 

Savage P, Dixon L, Grieve D, Watson C. SGLT2 Inhibition in Heart Failure: Clues to Cardiac Effects? Cardiol Rev. 2024 January 8. https://doi.org/10.1097/CRD.0000000000000637 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/38189526

153 

Verma S, McMurray JJV. SGLT2 inhibitors and mechanisms of cardiovascular benefit: a state-of-the-art review. Diabetologia. 2018 October;61(10):2108–17. https://doi.org/10.1007/s00125-018-4670-7 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30132036

154 

Garg V, Verma S, Connelly K. Mechanistic insights regarding the role of SGLT2 inhibitors and GLP1 agonist drugs on cardiovascular disease in diabetes. Prog Cardiovasc Dis. 2019 July-August;62(4):349–57. https://doi.org/10.1016/j.pcad.2019.07.005 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/31381891

155 

Vasquez-Rios G, Nadkarni GN. SGLT2 Inhibitors: Emerging Roles in the Protection Against Cardiovascular and Kidney Disease Among Diabetic Patients. Int J Nephrol Renovasc Dis. 2020 October 28;13:281–96. https://doi.org/10.2147/IJNRD.S268811 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/33149657

156 

Packer M. SGLT2 Inhibitors Produce Cardiorenal Benefits by Promoting Adaptive Cellular Reprogramming to Induce a State of Fasting Mimicry: A Paradigm Shift in Understanding Their Mechanism of Action. Diabetes Care. 2020 March;43(3):508–11. https://doi.org/10.2337/dci19-0074 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32079684

157 

Chen K, Nie Z, Shi R, Yu D, Wang Q, Shao F, et al. Time to Benefit of Sodium-Glucose Cotransporter-2 Inhibitors Among Patients With Heart Failure. JAMA Netw Open. 2023 August 1;6(8):e2330754. https://doi.org/10.1001/jamanetworkopen.2023.30754 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/37615988

158 

McMurray JJV, Solomon SD, Inzucchi SE, Køber L, Kosiborod MN, Martinez FA, et al. DAPA-HF Trial Committees and Investigators. Dapagliflozin in Patients with Heart Failure and Reduced Ejection Fraction. N Engl J Med. 2019 November 21;381(21):1995–2008. https://doi.org/10.1056/NEJMoa1911303 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/31535829

159 

Packer M, Anker SD, Butler J, Filippatos G, Pocock SJ, Carson P, et al. EMPEROR-Reduced Trial Investigators. Cardiovascular and Renal Outcomes with Empagliflozin in Heart Failure. N Engl J Med. 2020 October 8;383(15):1413–24. https://doi.org/10.1056/NEJMoa2022190 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32865377

160 

Anker SD, Butler J, Filippatos G, Ferreira JP, Bocchi E, Böhm M, et al. EMPEROR-Preserved Trial Investigators. Empagliflozin in Heart Failure with a Preserved Ejection Fraction. N Engl J Med. 2021 October 14;385(16):1451–61. https://doi.org/10.1056/NEJMoa2107038 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/34449189

161 

Solomon SD, Vaduganathan M, Claggett BL, de Boer RA, DeMets D, Hernandez AF, et al. Baseline Characteristics of Patients With HF With Mildly Reduced and Preserved Ejection Fraction: DELIVER Trial. JACC Heart Fail. 2022 March;10(3):184–97. https://doi.org/10.1016/j.jchf.2021.11.006 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/35241246

162 

Packer M. SGLT2 inhibitors: role in protective reprogramming of cardiac nutrient transport and metabolism. Nat Rev Cardiol. 2023 July;20(7):443–62. https://doi.org/10.1038/s41569-022-00824-4 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/36609604

163 

Packer M. Critical Reanalysis of the Mechanisms Underlying the Cardiorenal Benefits of SGLT2 Inhibitors and Reaffirmation of the Nutrient Deprivation Signaling/Autophagy Hypothesis. Circulation. 2022 November;146(18):1383–405. https://doi.org/10.1161/CIRCULATIONAHA.122.061732 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/36315602

164 

Pandey AK, Bhatt DL, Pandey A, Marx N, Cosentino F, Pandey A, et al. Mechanisms of benefits of sodium-glucose cotransporter 2 inhibitors in heart failure with preserved ejection fraction. Eur Heart J. 2023 October 1;44(37):3640–51. https://doi.org/10.1093/eurheartj/ehad389 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/37674356

165 

Docherty KF, Bayes-Genis A, Butler J, Coats AJS, Drazner MH, Joyce E, et al. The four pillars of HFrEF therapy: is it time to treat heart failure regardless of ejection fraction? Eur Heart J Suppl. 2022 December 19;24 Suppl L:L10–9. https://doi.org/10.1093/eurheartjsupp/suac113 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/36545228

166 

Pfeffer MA, Braunwald E. Treatment of Heart Failure With Preserved Ejection Fraction: Reflections on Its Treatment With an Aldosterone Antagonist. JAMA Cardiol. 2016 April 1;1(1):7–8. https://doi.org/10.1001/jamacardio.2015.0356 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27437645

167 

Deswal A, Richardson P, Bozkurt B, Mann DL. Results of the Randomized Aldosterone Antagonism in Heart Failure with Preserved Ejection Fraction trial (RAAM-PEF). J Card Fail. 2011 August;17(8):634–42. https://doi.org/10.1016/j.cardfail.2011.04.007 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/21807324

168 

Punnoose LR, Givertz MM, Lewis EF, Pratibhu P, Stevenson LW, Desai AS. Heart failure with recovered ejection fraction: a distinct clinical entity. J Card Fail. 2011 July;17(7):527–32. https://doi.org/10.1016/j.cardfail.2011.03.005 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/21703523

169 

Basuray A, French B, Ky B, Vorovich E, Olt C, Sweitzer NK, et al. Heart failure with recovered ejection fraction: clinical description, biomarkers, and outcomes. Circulation. 2014 June 10;129(23):2380–7. https://doi.org/10.1161/CIRCULATIONAHA.113.006855 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/24799515

170 

Kalogeropoulos AP, Fonarow GC, Georgiopoulou V, Burkman G, Siwamogsatham S, Patel A, et al. Characteristics and Outcomes of Adult Outpatients With Heart Failure and Improved or Recovered Ejection Fraction. JAMA Cardiol. 2016 August 1;1(5):510–8. https://doi.org/10.1001/jamacardio.2016.1325 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27434402

171 

Wilcox JE, Fang JC, Margulies KB, Mann DL. Heart Failure With Recovered Left Ventricular Ejection Fraction: JACC Scientific Expert Panel. J Am Coll Cardiol. 2020 August 11;76(6):719–34. https://doi.org/10.1016/j.jacc.2020.05.075 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32762907

172 

Barth AS, Kumordzie A, Frangakis C, Margulies KB, Cappola TP, Tomaselli GF. Reciprocal transcriptional regulation of metabolic and signaling pathways correlates with disease severity in heart failure. Circ Cardiovasc Genet. 2011 October;4(5):475–83. https://doi.org/10.1161/CIRCGENETICS.110.957571 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/21828333

173 

Packer M. Foetal recapitulation of nutrient surplus signalling by O-GlcNAcylation and the failing heart. Eur J Heart Fail. 2023 August;25(8):1199–212. https://doi.org/10.1002/ejhf.2972 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/37434410

174 

Cikes M, Solomon SD. Beyond ejection fraction: an integrative approach for assessment of cardiac structure and function in heart failure. Eur Heart J. 2016 June 1;37(21):1642–50. https://doi.org/10.1093/eurheartj/ehv510 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/26417058

175 

Louis A, Cleland JG, Crabbe S, Ford S, Thackray S, Houghton T, et al. Clinical Trials Update: CAPRICORN, COPERNICUS, MIRACLE, STAF, RITZ-2, RECOVER and RENAISSANCE and cachexia and cholesterol in heart failure. Highlights of the Scientific Sessions of the American College of Cardiology, 2001. Eur J Heart Fail. 2001 June;3(3):381–7. https://doi.org/10.1016/S1388-9842(01)00149-0 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/11378012

176 

Coletta AP, Clark AL, Banarjee P, Cleland JG. Clinical trials update: RENEWAL (RENAISSANCE and RECOVER) and ATTACH. Eur J Heart Fail. 2002 August;4(4):559–61. https://doi.org/10.1016/S1388-9842(02)00121-6 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/12167397

177 

Hunt SA, Baker DW, Chin MH, Cinquegrani MP, Feldman AM, Francis GS, et al. American College of Cardiology/American Heart Association. ACC/AHA guidelines for the evaluation and management of chronic heart failure in the adult: executive summary. A report of the American College of Cardiology/American Heart Association Task Force on Practice Guidelines (Committee to revise the 1995 Guidelines for the Evaluation and Management of Heart Failure). J Am Coll Cardiol. 2001 December;38(7):2101–13. https://doi.org/10.1016/S0735-1097(01)01683-7 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/11738322

178 

Stevenson LW, Naftilan AJ. Changing the Stage Directions for Heart Failure? J Am Coll Cardiol. 2020 March 31;75(12):1439–42. https://doi.org/10.1016/j.jacc.2020.02.027 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32216912

179 

Disertori M, Masè M, Rigoni M, Nollo G, Ravelli F. Declining clinical benefit of ICD in heart failure patients: Temporal trend of mortality outcomes from randomized controlled trials. J Cardiol. 2020 February;75(2):148–54. https://doi.org/10.1016/j.jjcc.2019.06.001 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/31526627

180 

Seferovic PM, Ponikowski P, Anker SD, Bauersachs J, Chioncel O, Cleland JGF, et al. Clinical practice update on heart failure 2019: pharmacotherapy, procedures, devices and patient management. An expert consensus meeting report of the Heart Failure Association of the European Society of Cardiology. Eur J Heart Fail. 2019 October;21(10):1169–86. https://doi.org/10.1002/ejhf.1531 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/31129923

181 

Li Kam Wa ME, Assar SZ, Kirtane AJ, Perera D. Revascularisation for Ischaemic Cardiomyopathy. Interv Cardiol. 2023 August 1;18:e24. https://doi.org/10.15420/icr.2023.06 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/37655258

182 

Liga R, Colli A, Taggart DP, Boden WE, De Caterina R. Myocardial Revascularization in Patients With Ischemic Cardiomyopathy: For Whom and How. J Am Heart Assoc. 2023 March 21;12(6):e026943. https://doi.org/10.1161/JAHA.122.026943 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/36892041

183 

Ryan M, Morgan H, Petrie MC, Perera D. Coronary revascularisation in patients with ischaemic cardiomyopathy. Heart. 2021 April;107(8):612–8. https://doi.org/10.1136/heartjnl-2020-316856 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/33436491

184 

Gorica E, Mohammed SA, Ambrosini S, Calderone V, Costantino S, Paneni F. Epi-Drugs in Heart Failure. Front Cardiovasc Med. 2022 July 13;9:923014. https://doi.org/10.3389/fcvm.2022.923014 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/35911511

185 

Ambrosini S, Gorica E, Mohammed SA, Costantino S, Ruschitzka F, Paneni F. Epigenetic remodeling in heart failure with preserved ejection fraction. Curr Opin Cardiol. 2022 May 1;37(3):219–26. https://doi.org/10.1097/HCO.0000000000000961 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/35275888

186 

Ptaszek LM, Mansour M, Ruskin JN, Chien KR. Towards regenerative therapy for cardiac disease. Lancet. 2012 March 10;379(9819):933–42. https://doi.org/10.1016/S0140-6736(12)60075-0 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/22405796

187 

Doppler SA, Deutsch MA, Lange R, Krane M. Cardiac regeneration: current therapies-future concepts. J Thorac Dis. 2013 October;5(5):683–97. https://doi.org/10.3978/j.issn.2072-1439.2013.08.71 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/24255783

188 

Bartunek J, Behfar A, Dolatabadi D, Vanderheyden M, Ostojic M, Dens J, et al. Cardiopoietic stem cell therapy in heart failure: the C-CURE (Cardiopoietic stem Cell therapy in heart failURE) multicenter randomized trial with lineage-specified biologics. J Am Coll Cardiol. 2013 June 11;61(23):2329–38. https://doi.org/10.1016/j.jacc.2013.02.071 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/23583246

189 

Silvestre JS, Menasché P. The Evolution of the Stem Cell Theory for Heart Failure. EBioMedicine. 2015 November 5;2(12):1871–9. https://doi.org/10.1016/j.ebiom.2015.11.010 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/26844266

190 

Tzahor E, Poss KD. Cardiac regeneration strategies: Staying young at heart. Science. 2017 June 9;356(6342):1035–9. https://doi.org/10.1126/science.aam5894 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28596337

191 

Rosca MG, Hoppel CL. Mitochondria in heart failure. Cardiovasc Res. 2010 October 1;88(1):40–50. https://doi.org/10.1093/cvr/cvq240 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/20668004

192 

Brown DA, Perry JB, Allen ME, Sabbah HN, Stauffer BL, Shaikh SR, et al. Expert consensus document: Mitochondrial function as a therapeutic target in heart failure. Nat Rev Cardiol. 2017 April;14(4):238–50. https://doi.org/10.1038/nrcardio.2016.203 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28004807

193 

Zhou B, Tian R. Mitochondrial dysfunction in pathophysiology of heart failure. J Clin Invest. 2018 August 31;128(9):3716–26. https://doi.org/10.1172/JCI120849 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30124471

194 

Kumar AA, Kelly DP, Chirinos JA. Mitochondrial Dysfunction in Heart Failure With Preserved Ejection Fraction. Circulation. 2019 March 12;139(11):1435–50. https://doi.org/10.1161/CIRCULATIONAHA.118.036259 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30856000


This display is generated from NISO JATS XML with jats-html.xsl. The XSLT engine is libxslt.