• Search Menu
  • Sign in through your institution
  • Author videos
  • ESC Content Collections
  • Supplements
  • Author Guidelines
  • Submission Site
  • Open Access Options
  • Self-Archiving Policy
  • About European Heart Journal Supplements
  • About the European Society of Cardiology
  • ESC Publications
  • Editorial Board
  • Advertising and Corporate Services
  • Journals Career Network
  • Terms and Conditions
  • Journals on Oxford Academic
  • Books on Oxford Academic

Issue Cover

Article Contents

Introduction, pathophysiology of acute heart failure, diagnosis of acute heart failure.

  • < Previous

Understanding acute heart failure: pathophysiology and diagnosis

  • Article contents
  • Figures & tables
  • Supplementary Data

Mattia Arrigo, John T. Parissis, Eiichi Akiyama, Alexandre Mebazaa, Understanding acute heart failure: pathophysiology and diagnosis, European Heart Journal Supplements , Volume 18, Issue suppl_G, December 2016, Pages G11–G18, https://doi.org/10.1093/eurheartj/suw044

  • Permissions Icon Permissions

Acute heart failure (AHF) is a relevant public health problem causing the majority of unplanned hospital admissions in patients aged of 65 years or more. AHF was historically described as a pump failure causing downstream hypoperfusion and upstream congestion. During the last decades a more complex network of interactions has been added to the simplistic haemodynamic model for explaining the pathophysiology of AHF. In addition, AHF is not a specific disease but the shared clinical presentation of different, heterogeneous cardiac abnormalities. Persistence of poor outcomes in AHF might be related to the paucity of improvements in the acute management of those patients. Indeed, acute treatment of AHF still mainly consists of intravenous diuretics and/or vasodilators, tailored according to the initial haemodynamic status with little regard to the underlying pathophysiological particularities. Therefore, there is an unmet need for increased individualization of AHF treatment according to the predominant underlying pathophysiological mechanisms to, hopefully, improve patient's outcome. In this article we review current knowledge on pathophysiology and initial diagnosis of AHF.

Acute heart failure (AHF) is a relevant public health problem causing the majority of unplanned hospital admissions in patients aged of 65 years or more. 1 Despite major achievements in the treatment of chronic heart failure (HF) over the last decades, which led to marked improvement in long-term survival, outcomes of AHF remain poor with 90-day rehospitalization and 1-year mortality rates reaching 10–30%. 2 Persistence of poor outcomes in AHF might be related to the paucity of improvements in the acute management of those patients. Despite lacking evidence of beneficial effects on outcome, acute treatment of AHF still mainly consists of non-invasive ventilation in case of pulmonary oedema, intravenous diuretics and/or vasodilators. These interventions are tailored according to the initial haemodynamic status with little regard to the underlying pathophysiological particularities. 3–5

Acute heart failure was historically described as a pump failure causing downstream hypoperfusion and upstream congestion. During the last decades a more complex network of interactions has been added to the simplistic haemodynamic model for explaining the pathophysiology of AHF. 6 In addition, AHF is not a specific disease but the shared clinical presentation of different, heterogeneous cardiac abnormalities. Therefore, there is an unmet need for increased individualization of AHF treatment according to the predominant underlying pathophysiological mechanisms to, hopefully, improve patient’s outcome.

Acute heart failure is defined as new-onset or worsening of symptoms and signs of HF, 5 often requiring rapid escalation of therapy and hospital admission. The clinical presentation of AHF typically includes symptoms or signs related to congestion and volume overload rather than to hypoperfusion. 7 Since congestion plays a central role for the vast majority of AHF cases, understanding of the underlying pathophysiological mechanisms related to congestion is essential for treating AHF patients. 8 More importantly, the level of congestion and the number of congested organs have prognostic relevance in HF patients. 8

Mechanisms of congestion: fluid accumulation and fluid redistribution

Congestion in heart failure.

Congestion in heart failure.

Tissue oedema occurs when the transudation from capillaries into the interstitium exceeds the maximal drainage of the lymphatic system. Transudation of plasma fluid into the interstitium results from the relation between hydrostatic and oncotic pressures in the capillaries and in the interstitium as well as interstitial compliance. Increased transcapillary hydrostatic pressure gradient, decreased transcapillary oncotic pressure gradient and increased interstitial compliance promote oedema formation.

In healthy individuals, increased total body sodium is usually not accompanied by oedema formation since a large quantity of sodium may be buffered by interstitial glycosaminoglycan networks without compensatory water retention. 14 Moreover, the interstitial glycosaminoglycan networks display low compliance which prevents fluid accumulation in the interstitium. 15

In HF, when sodium accumulation persists, the glycosaminoglycan networks may become dysfunctional resulting in reduced buffering capacity and increased compliance. In AHF the presence of pulmonary or peripheral oedema correlates poorly with left- and right-sided filling pressures, 16 , 17 but in patients with dysfunctional glycosaminoglycan networks even mildly elevated venous pressures might lead to pulmonary and peripheral oedemas. 9 In addition, since a large amount of sodium is stored in the interstitial glycosaminoglycan networks and does not reach the kidneys, it escapes renal clearance and is particularly difficult to remove from the body. 9

Moreover, persistent neuro-humoral activation induces maladaptive processes resulting in detrimental ventricular remodelling and organ dysfunction. Based on that, pharmacological therapies that inhibit the sympathetic and renin-angiotensin-aldosterone systems, including beta-blockers, angiotensin-converting enzyme inhibitors, angiotensin receptor blockers, aldosterone antagonists and more recently the angiotensin receptor neprilysin inhibitor LCZ696 have become the mainstays of chronic HF therapy. 18–33

Fluid accumulation alone cannot explain the whole pathophysiology of AHF. Indeed, the majority of AHF patients display only a minor increase in body weight (<1 kg) before hospital admission. 11–13

In those patients, congestion is precipitated by fluid redistribution, rather than accumulation. Sympathetic stimulation has been shown to induce a transient vasoconstriction leading to a sudden displacement of volume from the splanchnic and peripheral venous system to the pulmonary circulation, without exogenous fluid retention. 34 , 35 Nonetheless, the prerequisite for fluid redistribution is the presence of a certain amount of peripheral and splanchnic congestion.

In physiological states, capacitance veins contain one fourth of the total blood volume and stabilize cardiac preload, buffering volume overload. 36 , 37 In hypertensive AHF, the primary alteration is a mismatch in the ventricular-vascular coupling relationship with increased afterload and decreased venous capacitance (increased preload). 38

Fluid accumulation and fluid redistribution both produce an increase in cardiac load and congestion in AHF, but their relevance is likely to vary according to different clinical scenarios. While fluid accumulation might be more common in decompensations of congestive heart failure (CHF) with reduced ejection fraction, fluid redistribution might be the predominant pathophysiological mechanism in AHF with preserved ejection fraction. 39 Accordingly, the decongestive therapy should be tailored. While diuretics might be useful in presence of fluid accumulation, vasodilators might be more appropriate in presence of fluid redistribution to modulate ventricular-vascular coupling.

Furthermore, recent experimental data from human models suggest that venous congestion is not simply an epiphenomenon secondary to cardiac dysfunction but rather plays an active detrimental role in the pathophysiology of AHF inducing pro-oxidant, pro-inflammatory and haemodynamic stimuli that contribute to acute decompensation. 40 How these pathophysiological changes are induced remains incompletely understood but the biomechanical forces generated by congestion significantly contribute to endothelial and neuro-humoral activation. Indeed, endothelial stretch triggers an intracellular signalling cascade and causes endothelial cells to undergo a phenotypic switch to a pro-oxidant, pro-inflammatory vasoconstricted state. 41

Congestion-induced organ dysfunction

Overview of congestion-induced organ dysfunction and clinical manifestations

Congested organClinical manifestationReferences
HeartThird heart sound, jugular vein distension, positive hepato-jugular reflux
Functional mitral and tricuspid regurgitation
Elevated NPs: BNP >100 pg/mL, NT-proBNP >300 pg/mL, MR-proANP >120 pmol/L
LungDyspnoea, orthopnoea, bendopnoea, paroxysmal nocturnal dyspnoea ,
Auscultatory rales, crackles, wheezing; tachypnoea and hypoxia
Pathological chest radiography (interstitial/alveolar oedema, pleural effusion)
B-lines (‘comets’) on lung ultrasound
KidneyDecreased urine output ,
Elevated creatinine levels, hyponatraemia
LiverRight-sided upper abdominal discomfort, hepatomegaly, icterus ,
Elevated parameters of cholestasis
BowelNausea, vomiting, abdominal pain ,
Ascites, increased abdominal pressure
Cachexia
Congested organClinical manifestationReferences
HeartThird heart sound, jugular vein distension, positive hepato-jugular reflux
Functional mitral and tricuspid regurgitation
Elevated NPs: BNP >100 pg/mL, NT-proBNP >300 pg/mL, MR-proANP >120 pmol/L
LungDyspnoea, orthopnoea, bendopnoea, paroxysmal nocturnal dyspnoea ,
Auscultatory rales, crackles, wheezing; tachypnoea and hypoxia
Pathological chest radiography (interstitial/alveolar oedema, pleural effusion)
B-lines (‘comets’) on lung ultrasound
KidneyDecreased urine output ,
Elevated creatinine levels, hyponatraemia
LiverRight-sided upper abdominal discomfort, hepatomegaly, icterus ,
Elevated parameters of cholestasis
BowelNausea, vomiting, abdominal pain ,
Ascites, increased abdominal pressure
Cachexia

The close interaction between cardiac and renal dysfunction is known as the cardio-renal syndrome. 42 Historically, renal dysfunction in HF was described as consequence of reduced cardiac index and arterial underfilling both causing renal hypoperfusion. 43 More recent data showed that venous congestion (assessed as increased central venous pressure) was the strongest haemodynamic determinant for the development of renal dysfunction and low cardiac index alone in AHF has minor effects on renal function. 44 , 45 However, the combination of elevated central venous pressure and low cardiac index is particularly unfavourable for renal function.

Visceral congestion may increase intra-abdominal pressure in HF, which further negatively affects renal function in HF. Recent data showed that reducing central venous and intra-abdominal pressures by decongestive therapy may ameliorate serum creatinine, presumably by alleviating renal and abdominal congestion. 46

Cardiac dysfunction is frequently associated with liver abnormalities (cardio-hepatic syndrome) and negatively influences prognosis in AHF. 47 , 48 Cholestatic liver dysfunction is common in HF and is mainly related to right-sided congestion, while rapid and marked elevation in transaminases in AHF indicates hypoxic hepatitis related to hypoperfusion. 49 , 50 Finally, bowel congestion may contribute to development of cachexia in patients with advanced HF. 51

Assessment of congestion

Detection of congestion at an early (asymptomatic) stage is still an unmet need. Improved diagnostic methods would be highly valuable to enable early initiation of appropriate therapy following the ‘time to therapy’ approach recently introduced into HF guidelines. 5 The guidelines emphasize the potentially greater benefit of early treatment in the setting of AHF, as has long been the case for acute coronary syndromes. Indeed, the congestive cascade often begins several days or weeks before symptom onset and includes a sub-clinical increase of cardiac filling and venous pressures (‘haemodynamic congestion’) which may further lead to redistribution of fluids within the lungs and visceral organs (‘organ congestion’) and finally to overt signs and symptoms of volume overload (‘clinical congestion’). 12 , 52 Clinical congestion may be the ‘tip of the iceberg’ of the congestive cascade 8 . Although organ congestion is usually related to haemodynamic congestion, this might not be always true: indeed, several mechanisms might prevent oedema formation despite increased venous pressures and conversely, oedema might develop even in absence of increased hydrostatic pressure. 53

To achieve early detection of congestion, several strategies including cardiac biomarkers, intrathoracic impedance monitoring and implantable haemodynamic monitoring have been proposed. 54–59 However, the use of classical biomarkers, in particular natriuretic peptides (NPs), which are released by the failing heart, reflect the severity of myocardial dysfunction and only indirectly haemodynamic congestion. 60 , 61 Novel vascular biomarkers (e.g. soluble CD146, CA125) might better correlate with congestion than NPs. 62–65

Early management of AHF.

Early management of AHF.

Clinical evaluation

The initial clinical evaluation of dyspnoeic patients should help to (i) assess severity of AHF (ii) confirm the diagnosis of AHF and (iii) identify precipitating factors of AHF.

Since congestion is a typical feature of AHF, patient history and physical examination should primarily focus on the presence of congestion which would support the diagnosis of AHF. Left-sided congestion may cause dyspnoea, orthopnoea, bendopnoea, paroxysmal nocturnal dyspnoea, cough, tachypnoea, pathological lung auscultation (rales, crackles, wheezing) and hypoxia. 8 The absence of rales and a normal chest radiography do not exclude the presence of left-sided congestion. Indeed, 40–50% of patients with elevated pulmonary-artery wedge pressure may have a normal chest radiography. 66 Right-sided congestion may cause increased body weight, bilateral peripheral oedema, decreased urine output, abdominal pain, nausea and vomiting, jugular vein distension or positive hepato-jugular reflux, ascites, hepatomegaly, icterus. 8

Symptoms and signs of hypoperfusion indicate severity and may include hypotension, tachycardia, weak pulse, mental confusion, anxiety, fatigue, cold sweated extremities, decreased urine output and angina due to myocardial ischaemia. The presence of inappropriate stroke volume and clinical and biological signs of hypoperfusion in AHF defines cardiogenic shock, the most severe form of cardiac dysfunction. 67 Cardiogenic shock is most frequently related to acute myocardial infarction and accounts for less than 10% of AHF cases but is associated with in-hospital mortality rates of 40–50%. 39 , 68

However, given the limited sensitivity and specificity of symptoms and signs of AHF, the clinical evaluation should integrate information from additional tests. 69 , 70

According to the presence of clinical symptoms or signs of organ congestion (‘wet’ vs. ‘dry’) and/or peripheral hypoperfusion (‘cold’ vs. ‘warm’), patients may be classified in four groups. 67 , 71 About two of three AHF patients are classified ‘wet-warm’ (congested but well perfused), about one of four are ‘wet-cold’ (congested and hypoperfused) and only a minority are ‘dry-cold’ (not congested and hypoperfused). The fourth group ‘dry-warm’ represent the compensated (decongested, well-perfused) status. This classification may help to guide initial therapy (mostly vasodilators and/or diuretics) and carries prognostic information. 70 Patients with cardiopulmonary distress should be managed in intensive cardiac care units.

Notably, the use of inotropes should be restricted to patients with cardiogenic shock or AHF resulting in hypotension and hypoperfusion to maintain end-organ function, 5 since their often inappropriate use is associated with increased morbidity and mortality. 72

Acute heart failure usually consists of acute decompensation of chronic HF (ADHF) or, less frequently, may arise in patients without previous history of symptomatic HF ( de novo AHF). 68 The distinction of these two scenarios is important because the underlying mechanisms leading to AHF are significantly different. Indeed, while pre-existing pathophysiological derangements predispose CHF patients to ADHF, de novo AHF is typically induced by severe haemodynamic alterations secondary to the initial insult. Common causes of de novo AHF include acute myocardial infarction, severe myocarditis, acute valve regurgitation and pericardial tamponade. 68 On the other hand, ADHF may be precipitated by several clinical conditions, while in some patients, no precipitant can be identified. 2 , 73 , – 75

Rapid identification of precipitants of AHF is crucial to optimize patient management. The most common precipitants are myocardial ischaemia, arrhythmias (in particular paroxysmal atrial fibrillation), sepsis and/or pulmonary disease, uncontrolled hypertension, non-compliance with medical prescriptions, renal dysfunction and iatrogenic causes. The identification of precipitants of AHF aims at detecting reversible or treatable causes and at assisting prognostication. Indeed, the initial management should include, in addition to vasodilators and/or diuretics, also specific treatments directed towards the underlying causes of AHF. In particular, early coronary angiography with revascularization is recommended in AHF precipitated by acute coronary syndrome, antiarrhythmic treatment and/or electrical cardioversion are recommended in AHF precipitated by arrhythmia, rapid initiation of antimicrobial therapy is recommended for AHF precipitated by sepsis. 76–79 Furthermore, identification of precipitants of AHF may allow risk stratification of patients with AHF. Indeed, AHF precipitated by acute coronary syndrome or infection is associated with poorer outcomes whereas outcomes tend to be better in AHF precipitated by atrial fibrillation or uncontrolled hypertension. 73 , 74

Additional tests

Additional laboratory tests are helpful in the evaluation of patients with AHF. Natriuretic peptides, including B-type NP (BNP), amino-terminal pro-B-type NP (NT-proBNP) and mid-regional pro-atrial NPs (MR-proANP) show high accuracy and excellent negative predictive value in differentiating AHF from non-cardiac causes of acute dyspnoea. 80 Natriuretic peptide levels in HFpEF are lower than in HFrEF. Low circulating NPs (thresholds: BNP <100 pg/mL, NT-proBNP <300 pg/mL, MR-proANP <120 pmol/L) make the diagnosis of AHF unlikely. This is true for both HFrEF and HFpEF. A recent meta-analysis indicated that at these thresholds BNP and NT-proBNP have sensitivities of 0.95 and 0.99 and negative predictive values of 0.94 and 0.98, respectively, for a diagnosis of AHF. 80 MRproANP had a sensitivity ranging from 0.95 to 0.97 and a negative predictive value ranging from 0.90 to 0.97. 80

However, elevated levels of NPs do not automatically confirm the diagnosis of AHF, as they may also be associated with a wide variety of cardiac and non-cardiac causes. Among them, atrial fibrillation, age, and renal failure are the most important factors impeding the interpretation of NP measurements. On the other hand, NP levels may be disproportionally low in obese patients and in those with flash pulmonary oedema. Natriuretic peptides should be measured in all patients with suspected AHF upon presentation to the emergency department or intensive cardiac care units. 3–5

Cardiac troponin may be helpful to exclude myocardial ischaemia as precipitating factor of AHF. However, cardiac troponin, in particular when measured with high-sensitive assays, is frequently elevated in patients with AHF, often without obvious myocardial ischaemia or an acute coronary event. Indeed, AHF is characterized by accelerated myocardial necrosis and remodelling. Troponin measurement may be considered for prognostication as elevated levels are associated with poorer outcomes. 81 Numerous clinical variables and biomarkers are independent predictors of in-hospital complications and longer-term outcomes in AHF syndromes, but their impact on management has not been adequately established. The easy-to perform AHEAD score based on the analysis of co-morbidities has been shown to provide relevant information on short and long term prognosis of patients hospitalized for AHF. 82

An electrocardiography (ECG) may be helpful to identify potential precipitants of AHF (e.g. arrhythmia, ischaemia) and to exclude ST-elevation myocardial infarction requiring immediate revascularization. However, ECG is rarely normal in AHF patients. Current guidelines do not recommend immediate echocardiography in all patients presenting with AHF. 3–5 However, all patients presenting with cardiogenic shock or suspicion of acute life-threatening structural or functional cardiac abnormalities (mechanical complications, acute valve regurgitation, aortic dissection) should receive immediate echocardiography. Early echocardiography should be considered in all patients with de novo AHF and in those with unknown cardiac function, however, the optimal timing is unknown (preferably within 24–48 h from admission). 3–5

Thoracic ultrasound and chest X-ray may both be useful to assess the presence of interstitial pulmonary oedema. While chest X-ray may also be helpful to rule-out alternative causes of dyspnoea (e.g. pneumothorax, pneumonia), both techniques provide complementary information about the presence of pulmonary oedema or pleural effusion. Abdominal ultrasound may be useful to measure inferior vena cava diameter and collapsibility and exclude the presence of ascites. 3–5

Reassessment and allocation

Most of the patients presenting with AHF require hospital admission. The level of care (discharge, observation, ward, telemetry and intensive cardiac care unit) should be based on history (including symptom severity, precipitating factors), physical examination (haemodynamic and respiratory status, degree of congestion) and biomarkers (NPs, cardiac troponin, renal function, serum lactate). Forty to 50% of AHF patients require admission to intensive cardiac care units. 39 , 68 Low risk AHF patients (those with slightly elevated NP levels, normal blood pressure, stable renal failure, normal troponin) and with good response to initial therapy may be considered for early discharge. Follow-up plans must be in place prior to discharge and clearly communicated to the primary care team. 3–5

Pathophysiology-based management

According to current knowledge on the pathophysiology of AHF, the initial treatment of AHF patients should include decongestive therapy (e.g. vasodilators and/or diuretics) and specific therapy directed towards the underlying causes of AHF (e.g. revascularization, antiarrhythmic treatments, antimicrobial drugs). Moreover, early administration of oral disease-modifying HF therapy (beta-blockers, angiotensin-converting enzyme inhibitors or angiotensin receptor blockers and mineralocorticoid receptor antagonists), before hospital discharge is recommended in all patients with AHF.

Conflict of interest: MA is recipient of a fellowship of the Collège de Médecine des Hôpitaux de Paris. JP has received honoraria for lectures from Novartis, Orion Pharma and Roche Diagnostics. AM has received speaker honoraria from Abbott, Novartis, Orion, Roche and Servier and fee as member of advisory board and/or steering committee from Cardiorentis, Adrenomed, MyCartis, ZS Pharma and Critical Diagnostics. The other authors declare no conflict of interest.

Townsend N Nichols M Scarborough P Rayner M. Cardiovascular disease in Europe–epidemiological update 2015 . Eur Heart J 2015 ; 36 : 2696 – 2705 .

Google Scholar

Ambrosy AP Fonarow GC Butler J Chioncel O Greene SJ Vaduganathan M Nodari S Lam CSP Sato N Shah AN Gheorghiade M. The global health and economic burden of hospitalizations for heart failure: lessons learned from hospitalized heart failure registries . J Am Coll Cardiol 2014 ; 63 : 1123 – 1133 .

Mebazaa A Yilmaz MB Levy P Ponikowski P Peacock WF Laribi S Ristic AD Lambrinou E Masip J Riley JP McDonagh T Mueller C deFilippi C Harjola V-P Thiele H Piepoli MF Metra M Maggioni A McMurray JJV Dickstein K Damman K Seferovic PM Ruschitzka F Leite-Moreira AF Bellou A Anker SD Filippatos G. Recommendations on pre-hospital and early hospital management of acute heart failure: a consensus paper from the Heart Failure Association of the European Society of Cardiology, the European Society of Emergency Medicine and the Society of Academic Emergency Medicine - short version . Eur Heart J 2015 ; 36 : 1958 – 1966 .

Mebazaa A Tolppanen H Mueller C Lassus J diSomma S Baksyte G Cecconi M Choi DJ Cohen-Solal A Christ M Masip J Arrigo M Nouira S Ojji D Peacock F Richards M Sato N Sliwa K Spinar J Thiele H Yilmaz MB Januzzi J. Acute heart failure and cardiogenic shock: a multidisciplinary practical guidance . Intensive Care Med 2016 ; 42 : 147 – 163 .

Ponikowski P Voors AA Anker SD Bueno H Cleland JGF Coats AJS Falk V González-Juanatey JR Harjola V-P Jankowska EA Jessup M Linde C Nihoyannopoulos P Parissis JT Pieske B Riley JP Rosano GMC Ruilope LM Ruschitzka F Rutten FH van der Meer P , Authors/Task Force Members, Document Reviewers . 2016 ESC Guidelines for the diagnosis and treatment of acute and chronic heart failure: The Task Force for the diagnosis and treatment of acute and chronic heart failure of the European Society of Cardiology (ESC) Developed with the special contribution of the Heart Failure Association (HFA) of the ESC . Eur Heart J 2016 ; 37 : 2129 – 2200 .

Mentz RJ O'connor CM. Pathophysiology and clinical evaluation of acute heart failure . Nat Rev Cardiol 2016 ; 13 : 28 – 35 .

Costanzo MR Jessup M. Treatment of congestion in heart failure with diuretics and extracorporeal therapies: effects on symptoms, renal function, and prognosis . Heart Fail Rev 2012 ; 17 : 313 – 324 .

Gheorghiade M Follath F Ponikowski P Barsuk JH Blair JEA Cleland JG Dickstein K Drazner MH Fonarow GC Jaarsma T Jondeau G Sendon JL Mebazaa A Metra M Nieminen M Pang PS Seferovic P Stevenson LW Van Veldhuisen DJ Zannad F Anker SD Rhodes A McMurray JJV Filippatos G , European Society of Cardiology, European Society of Intensive Care Medicine . Assessing and grading congestion in acute heart failure: a scientific statement from the acute heart failure committee of the heart failure association of the European Society of Cardiology and endorsed by the European Society of Intensive Care Medicine . Eur J Heart Fail 2010 ; 12 : 423 – 433 .

Nijst P Verbrugge FH Grieten L Dupont M Steels P Tang WHW Mullens W. The pathophysiological role of interstitial sodium in heart failure . J Am Coll Cardiol 2015 ; 65 : 378 – 388 .

McKie PM Schirger JA Costello-Boerrigter LC Benike SL Harstad LK Bailey KR Hodge DO Redfield MM Simari RD Burnett JC Chen HH. Impaired natriuretic and renal endocrine response to acute volume expansion in pre-clinical systolic and diastolic dysfunction . J Am Coll Cardiol 2011 ; 58 : 2095 – 2103 .

Yu C-M Chan JY-S Zhang Q Omar R Yip GW-K Hussin A Fang F Lam KH Chan HC-K Fung JW-H. Biventricular pacing in patients with bradycardia and normal ejection fraction . N Engl J Med 2009 ; 361 : 2123 – 2134 .

Chaudhry SI Wang Y Concato J Gill TM Krumholz HM. Patterns of weight change preceding hospitalization for heart failure . Circulation 2007 ; 116 : 1549 – 1554 .

Zile MR Bennett TD St John Sutton M Cho YK Adamson PB Aaron MF Aranda JM Abraham WT Smart FW Stevenson LW Kueffer FJ Bourge RC. Transition from chronic compensated to acute decompensated heart failure: pathophysiological insights obtained from continuous monitoring of intracardiac pressures . Circulation 2008 ; 118 : 1433 – 1441 .

Titze J Shakibaei M Schafflhuber M Schulze-Tanzil G Porst M Schwind KH Dietsch P Hilgers KF. Glycosaminoglycan polymerization may enable osmotically inactive Na+ storage in the skin . Am J Physiol Heart Circ Physiol 2004 ; 287 : H203 – H208 .

Guyton AC. Interstitial fluid pressure. Pressure-volume curves of interstitial space . Circ Res 1965 ; 16 : 452 – 460 .

Breidthardt T Irfan A Klima T Drexler B Balmelli C Arenja N Socrates T Ringger R Heinisch C Ziller R Schifferli J Meune C Mueller C. Pathophysiology of lower extremity edema in acute heart failure revisited . Am J Med 2012 ; 125 : 1124.e1 – 1124.e8 .

Zile MR Adamson PB Cho YK Bennett TD Bourge RC Aaron MF Aranda JM Abraham WT Stevenson LW Kueffer FJ. Hemodynamic factors associated with acute decompensated heart failure: part 1–insights into pathophysiology . J Card Fail 2011 ; 17 : 282 – 291 .

Packer M Bristow MR Cohn JN Colucci WS Fowler MB Gilbert EM Shusterman NH. 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 ; 334 : 1349 – 1355 .

CIBIS-II Investigators and Committees . The Cardiac Insufficiency Bisoprolol Study II (CIBIS-II): a randomised trial . Lancet 1999 ; 353 : 9 – 13 .

MERIT-HF Study Group . Effect of metoprolol CR/XL in chronic heart failure: Metoprolol CR/XL Randomised Intervention Trial in Congestive Heart Failure (MERIT-HF) . Lancet 1999 ; 353 : 2001 – 2007 .

Flather MD Shibata MC Coats AJS Van Veldhuisen DJ Parkhomenko A Borbola J Cohen-Solal A Dumitrascu D Ferrari R Lechat P Soler-Soler J Tavazzi L Spinarova L Toman J Böhm M Anker SD Thompson SG Poole-Wilson PA , 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 ; 26 : 215 – 225 .

The 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 ; 316 : 1429 – 1435 .

The SOLVD Investigators . Effect of enalapril on survival in patients with reduced left ventricular ejection fractions and congestive heart failure. The SOLVD Investigators . N Engl J Med 1991 ; 325 : 293 – 302 .

The SOLVD Investigators . Effect of enalapril on mortality and the development of heart failure in asymptomatic patients with reduced left ventricular ejection fractions. The SOLVD Investigattors . N Engl J Med 1992 ; 327 : 685 – 691 .

Pitt B Poole-Wilson PA Segal R Martinez FA Dickstein K Camm AJ Konstam MA Riegger G Klinger GH Neaton J Sharma D Thiyagarajan B. 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 ; 355 : 1582 – 1587 .

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 ; 345 : 1667 – 1675 .

Maggioni AP Anand I Gottlieb SO Latini R Tognoni G Cohn JN , Val-HeFT Investigators (Valsartan Heart Failure Trial) . Effects of valsartan on morbidity and mortality in patients with heart failure not receiving angiotensin-converting enzyme inhibitors . J Am Coll Cardiol 2002 ; 40 : 1414 – 1421 .

Granger CB McMurray JJV Yusuf S Held P Michelson EL Olofsson B Ostergren J Pfeffer MA Swedberg K , 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 ; 362 : 772 – 776 .

Konstam MA Neaton JD Dickstein K Drexler H Komajda M Martinez FA Riegger GAJ Malbecq W Smith RD Guptha S Poole-Wilson PA , HEAAL Investigators . Effects of high-dose versus low-dose losartan on clinical outcomes in patients with heart failure (HEAAL study): a randomised, double-blind trial . Lancet 2009 ; 374 : 1840 – 1848 .

Pitt B Pfeffer MA Assmann SF Boineau R Anand IS Claggett B Clausell N Desai AS Diaz R Fleg JL Gordeev I Harty B Heitner JF Kenwood CT Lewis EF O'meara E Probstfield JL Shaburishvili T Shah SJ Solomon SD Sweitzer NK Yang S McKinlay SM , TOPCAT Investigators . Spironolactone for heart failure with preserved ejection fraction . N Engl J Med 2014 ; 370 : 1383 – 1392 .

Pitt B Remme W Zannad F Neaton J Martinez F Roniker B Bittman R Hurley S Kleiman J Gatlin M , 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 ; 348 : 1309 – 1321 .

Zannad F McMurray JJV Krum H Van Veldhuisen DJ Swedberg K Shi H Vincent J Pocock SJ Pitt B , EMPHASIS-HF Study Group . Eplerenone in patients with systolic heart failure and mild symptoms . N Engl J Med 2011 ; 364 : 11 – 21 .

McMurray JJV Packer M Desai AS Gong J Lefkowitz MP Rizkala AR Rouleau JL Shi VC Solomon SD Swedberg K Zile MR , the PARADIGM-HF Investigators and Committees . Angiotensin-neprilysin inhibition versus enalapril in heart failure . N Engl J Med 2014 ; 371 : 993 – 1004 .

Sánchez-Marteles M Rubio Gracia J Giménez López I. Pathophysiology of acute heart failure: a world to know . Rev Clin Esp 2016 ; 216 : 38 – 46 .

Cotter G Metra M Milo-Cotter O Dittrich HC Gheorghiade M. Fluid overload in acute heart failure–re-distribution and other mechanisms beyond fluid accumulation . Eur J Heart Fail 2008 ; 10 : 165 – 169 .

Greenway CV Lister GE. Capacitance effects and blood reservoir function in the splanchnic vascular bed during non-hypotensive haemorrhage and blood volume expansion in anaesthetized cats . J Physiol (Lond) 1974 ; 237 : 279 – 294 .

Greenway CV. Role of splanchnic venous system in overall cardiovascular homeostasis . Fed Proc 1983 ; 42 : 1678 – 1684 .

Viau DM Sala-Mercado JA Spranger MD O'leary DS Levy PD. The pathophysiology of hypertensive acute heart failure . Heart 2015 ; 101 : 1861 – 1867 .

Rudiger A Harjola V-P Müller A Mattila E Säila P Nieminen M Follath F. Acute heart failure: clinical presentation, one-year mortality and prognostic factors . Eur J Heart Fail 2005 ; 7 : 662 – 670 .

Colombo PC Onat D Harxhi A Demmer RT Hayashi Y Jelic S LeJemtel TH Bucciarelli L Kebschull M Papapanou P Uriel N Schmidt AM Sabbah HN Jorde UP. Peripheral venous congestion causes inflammation, neurohormonal, and endothelial cell activation . Eur Heart J 2014 ; 35 : 448 – 454 .

Colombo PC Doran AC Onat D Wong KY Ahmad M Sabbah HN Demmer RT. Venous congestion, endothelial and neurohormonal activation in acute decompensated heart failure: cause or effect? . Curr Heart Fail Rep 2015 ; 12 : 215 – 222 .

Ronco C Haapio M House AA Anavekar N Bellomo R. Cardiorenal syndrome . J Am Coll Cardiol 2008 ; 52 : 1527 – 1539 .

Ljungman S Laragh JH Cody RJ. Role of the kidney in congestive heart failure. Relationship of cardiac index to kidney function . Drugs 1990 ; 39 Suppl 4 : 10 – 21 .

Nohria A Hasselblad V Stebbins A Pauly DF Fonarow GC Shah M Yancy CW Califf RM Stevenson LW Hill JA. Cardiorenal interactions: insights from the ESCAPE trial . J Am Coll Cardiol 2008 ; 51 : 1268 – 1274 .

Mullens W Abrahams Z Francis GS Sokos G Taylor DO Starling RC Young JB Tang WHW. Importance of venous congestion for worsening of renal function in advanced decompensated heart failure . J Am Coll Cardiol 2009 ; 53 : 589 – 596 .

Verbrugge FH Dupont M Steels P Grieten L Malbrain M Tang WHW Mullens W. Abdominal contributions to cardiorenal dysfunction in congestive heart failure . J Am Coll Cardiol 2013 ; 62 : 485 – 495 .

Poelzl G Ess M Mussner-Seeber C Pachinger O Frick M Ulmer H. Liver dysfunction in chronic heart failure: prevalence, characteristics and prognostic significance . Eur J Clin Invest 2012 ; 42 : 153 – 163 .

Nikolaou M Parissis J Yilmaz MB Seronde M-F Kivikko M Laribi S Paugam-Burtz C Cai D Pohjanjousi P Laterre P-F Deye N Poder P Cohen-Solal A Mebazaa A. Liver function abnormalities, clinical profile, and outcome in acute decompensated heart failure . Eur Heart J 2013 ; 34 : 742 – 749 .

Ebert EC. Hypoxic liver injury . Mayo Clin Proc 2006 ; 81 : 1232 – 1236 .

Henrion J Schapira M Luwaert R Colin L Delannoy A Heller F. Hypoxic hepatitis: clinical and hemodynamic study in 142 consecutive cases . Medicine (Baltimore) 2003 ; 82 : 392 – 406 .

Valentova M Haehling von S Bauditz J Doehner W Ebner N Bekfani T Elsner S Sliziuk V Scherbakov N Murin J Anker SD Sandek A. Intestinal congestion and right ventricular dysfunction: a link with appetite loss, inflammation, and cachexia in chronic heart failure . Eur Heart J 2016 ; 37 : 1684 – 1691 .

Picano E Gargani L Gheorghiade M. Why, when, and how to assess pulmonary congestion in heart failure: pathophysiological, clinical, and methodological implications . Heart Fail Rev 2010 ; 15 : 63 – 72 .

Arrigo M Huber LC. Eponyms in cardiopulmonary reflexes . Am J Cardiol 2013 ; 112 : 449 – 453 .

Bourge RC Abraham WT Adamson PB Aaron MF Aranda JM Magalski A Zile MR Smith AL Smart FW O'shaughnessy MA Jessup ML Sparks B Naftel DL Stevenson LW , COMPASS-HF Study Group . Randomized controlled trial of an implantable continuous hemodynamic monitor in patients with advanced heart failure: the COMPASS-HF study . J Am Coll Cardiol 2008 ; 51 : 1073 – 1079 .

Chaudhry SI Mattera JA Curtis JP Spertus JA Herrin J Lin Z Phillips CO Hodshon BV Cooper LS Krumholz HM. Telemonitoring in patients with heart failure . N Engl J Med 2010 ; 363 : 2301 – 2309 .

Hindricks G Taborsky M Glikson M Heinrich U Schumacher B Katz A Brachmann J Lewalter T Goette A Block M Kautzner J Sack S Husser D Piorkowski C Sogaard P , IN-TIME study group . Implant-based multiparameter telemonitoring of patients with heart failure (IN-TIME): a randomised controlled trial . Lancet 2014 ; 384 : 583 – 590 .

Abraham WT Adamson PB Bourge RC Aaron MF Costanzo MR Stevenson LW Strickland W Neelagaru S Raval N Krueger S Weiner S Shavelle D Jeffries B Yadav JS , CHAMPION Trial Study Group . Wireless pulmonary artery haemodynamic monitoring in chronic heart failure: a randomised controlled trial . Lancet 2011 ; 377 : 658 – 666 .

McMurray JJV Adamopoulos S Anker SD Auricchio A Böhm M Dickstein K Falk V Filippatos G Fonseca C Gomez-Sanchez MA Jaarsma T Køber L Lip GYH Maggioni AP Parkhomenko A Pieske BM Popescu BA Rønnevik PK Rutten FH Schwitter J Seferovic P Stepinska J Trindade PT Voors AA Zannad F Zeiher A , ESC Committee for Practice Guidelines . ESC Guidelines for the diagnosis and treatment of acute and chronic heart failure 2012: The Task Force for the Diagnosis and Treatment of Acute and Chronic Heart Failure 2012 of the European Society of Cardiology. Developed in collaboration with the Heart Failure Association (HFA) of the ESC . Eur Heart J 2012 ; 33 : 1787 – 1847 .

Troughton R Michael Felker G Januzzi JL. Natriuretic peptide-guided heart failure management . Eur Heart J 2014 ; 35 : 16 – 24 .

Ambrosy AP Pang PS Khan S Konstam MA Fonarow GC Traver B Maggioni AP Cook T Swedberg K Burnett JC Grinfeld L Udelson JE Zannad F Gheorghiade M , EVEREST Trial Investigators . Clinical course and predictive value of congestion during hospitalization in patients admitted for worsening signs and symptoms of heart failure with reduced ejection fraction: findings from the EVEREST trial . Eur Heart J 2013 ; 34 : 835 – 843 .

Sabatine MS Morrow DA de Lemos JA Omland T Desai MY Tanasijevic M Hall C McCabe CH Braunwald E. Acute changes in circulating natriuretic peptide levels in relation to myocardial ischemia . J Am Coll Cardiol 2004 ; 44 : 1988 – 1995 .

Gayat E Caillard A Laribi S Mueller C Sadoune M Seronde M-F Maisel A Bartunek J Vanderheyden M Desutter J Dendale P Thomas G Tavares M Cohen-Solal A Samuel J-L Mebazaa A. Soluble CD146, a new endothelial biomarker of acutely decompensated heart failure . Int J Cardiol 2015 ; 199 : 241 – 247 .

Kubena P Arrigo M Parenica J Gayat E Sadoune M Ganovska E Pavlusova M Littnerova S Spinar J Mebazaa A , GREAT network . Plasma levels of soluble CD146 reflect the severity of pulmonary congestion better than brain natriuretic peptide in acute coronary syndrome . Ann Lab Med 2016 ; 36 : 300 – 305 .

Núñez J Llàcer P Núñez E Ventura S Bonanad C Bodí V Miñana G Santas E Mascarell B Fonarow GC Chorro FJ Sanchis J. Antigen carbohydrate 125 and creatinine on admission for prediction of renal function response following loop diuretic administration in acute heart failure . Int J Cardiol 2014 ; 174 : 516 – 523 .

Arrigo M Truong QA Onat D Szymonifka J Gayat E Tolppanen H Sadoune M Demmer RT Wong KY Launay JM Samuel JL Cohen-Solal A Januzzi JL Singh JP Colombo PC Mebazaa A . Soluble CD146 is a Novel Marker of Systemic Congestion with Prognostic Role in Severe Heart Failure Patients: An Experimental Mechanistic and Transcardiac Prognostic Study . Clin Chem 2016 ; in press.

Chakko S Woska D Martinez H de Marchena E Futterman L Kessler KM Myerberg RJ. Clinical, radiographic, and hemodynamic correlations in chronic congestive heart failure: conflicting results may lead to inappropriate care . Am J Med 1991 ; 90 : 353 – 359 .

Nohria A Tsang SW Fang JC Lewis EF Jarcho JA Mudge GH Stevenson LW. Clinical assessment identifies hemodynamic profiles that predict outcomes in patients admitted with heart failure . J Am Coll Cardiol 2003 ; 41 : 1797 – 1804 .

Nieminen MS Brutsaert D Dickstein K Drexler H Follath F Harjola V-P Hochadel M Komajda M Lassus J Lopez-Sendon JL Ponikowski P Tavazzi L. EuroHeart Survey Investigators, Heart Failure Association, European Society of Cardiology . EuroHeart Failure Survey II (EHFS II): a survey on hospitalized acute heart failure patients: description of population . Eur Heart J 2006 ; 27 : 2725 – 2736 .

Stevenson LW Perloff JK. The limited reliability of physical signs for estimating hemodynamics in chronic heart failure . JAMA 1989 ; 261 : 884 – 888 .

Drazner MH Hellkamp AS Leier CV Shah MR Miller LW Russell SD Young JB Califf RM Nohria A. Value of clinician assessment of hemodynamics in advanced heart failure: the ESCAPE trial . Circ Heart Fail 2008 ; 1 : 170 – 177 .

Nohria A Lewis E Stevenson LW. Medical management of advanced heart failure . JAMA 2002 ; 287 : 628 – 640 .

Arrigo M Mebazaa A. Understanding the differences among inotropes . Intensive Care Med 2015 ; 41 : 912 – 915 .

Fonarow GC Abraham WT Albert NM Stough WG Gheorghiade M Greenberg BH O'connor CM Pieper K Sun JL Yancy CW Young JB , OPTIMIZE-HF Investigators and Hospitals . Factors identified as precipitating hospital admissions for heart failure and clinical outcomes: findings from OPTIMIZE-HF . Arch Intern Med 2008 ; 168 : 847 – 854 .

Arrigo M Tolppanen H Sadoune M Feliot E Teixeira A Laribi S Plaisance P Nouira S Yilmaz MB Gayat E Mebazaa A , Network OBOTG . Effect of precipitating factors of acute heart failure on readmission and long-term mortality . ESC Heart Fail 2016 ; 3 : 115 – 121 .

Arrigo M Gayat E Parenica J Ishihara S Zhang J Choi DJ Park JJ AlHabib KF Sato N Miro O Maggioni A Zhang Y Spinar J Cohen-Solal A Iwashyna TJ Mebazaa A. Identification of precipitating factors to predict short-term outcome of acute heart failure: a report from the intercontinental GREAT registry . Eur J Heart Fail ; in press.

Hochman JS Sleeper LA Webb JG Sanborn TA White HD Talley JD Buller CE Jacobs AK Slater JN Col J McKinlay SM LeJemtel TH. Early revascularization in acute myocardial infarction complicated by cardiogenic shock. SHOCK investigators. Should we emergently revascularize occluded coronaries for cardiogenic shock . N Engl J Med 1999 ; 341 : 625 – 634 .

Santoro GM Carrabba N Migliorini A Parodi G Valenti R. Acute heart failure in patients with acute myocardial infarction treated with primary percutaneous coronary intervention . Eur J Heart Fail 2008 ; 10 : 780 – 785 .

Arrigo M Jaeger N Seifert B Spahn DR Bettex D Rudiger A. Disappointing success of electrical cardioversion for new-onset atrial fibrillation in cardiosurgical ICU patients . Crit Care Med 2015 ; 43 : 2354 – 2359 .

Singer M Deutschman CS Seymour CW Shankar-Hari M Annane D Bauer M Bellomo R Bernard GR Chiche J-D Coopersmith CM Hotchkiss RS Levy MM Marshall JC Martin GS Opal SM Rubenfeld GD van der Poll T Vincent J-L Angus DC. The third international consensus definitions for sepsis and septic shock (Sepsis-3) . JAMA 2016 ; 315 : 801 – 810 .

Peacock WF De Marco T Fonarow GC Diercks D Wynne J Apple FS Wu AHB , ADHERE Investigators . Cardiac troponin and outcome in acute heart failure . N Engl J Med 2008 ; 358 : 2117 – 2126 .

Roberts E Ludman AJ Dworzynski K Al-Mohammad A Cowie MR McMurray JJV Mant J , NICE Guideline Development Group for Acute Heart Failure . The diagnostic accuracy of the natriuretic peptides in heart failure: systematic review and diagnostic meta-analysis in the acute care setting . BMJ 2015 ; 350 : h910.

Spinar J Jarkovsky J Spinarova L Mebazaa A Gayat E Vitovec J Linhart A Widimsky P Miklik R Zeman K Belohlavek J Malek F Felsoci M Kettner J Ostadal P Cihalik C Vaclavik J Taborsky M Dusek L Littnerova S Parenica J. AHEAD score–Long-term risk classification in acute heart failure . Int J Cardiol 2016 ; 202 : 21 – 26 .

Maisel AS Krishnaswamy P Nowak RM McCord J Hollander JE Duc P Omland T Storrow AB Abraham WT Wu AHB Clopton P Steg PG Westheim A Knudsen CW Perez A Kazanegra R Herrmann HC McCullough PA , Breathing Not Properly Multinational Study Investigators . Rapid measurement of B-type natriuretic peptide in the emergency diagnosis of heart failure . N Engl J Med 2002 ; 347 : 161 – 167 .

Januzzi JL Camargo CA Anwaruddin S Baggish AL Chen AA Krauser DG Tung R Cameron R Nagurney JT Chae CU Lloyd-Jones DM Brown DF Foran-Melanson S Sluss PM Lee-Lewandrowski E Lewandrowski KB. The N-terminal Pro-BNP investigation of dyspnea in the emergency department (PRIDE) study . Am J Cardiol 2005 ; 95 : 948 – 954 .

Maisel A Mueller C Nowak R Peacock WF Landsberg JW Ponikowski P Möckel M Hogan C Wu AHB Richards M Clopton P Filippatos GS Di Somma S Anand I Ng L Daniels LB Neath S-X Christenson R Potocki M McCord J Terracciano G Kremastinos D Hartmann O Haehling von S Bergmann A Morgenthaler NG Anker SD. Mid-region pro-hormone markers for diagnosis and prognosis in acute dyspnea: results from the BACH (Biomarkers in Acute Heart Failure) trial . J Am Coll Cardiol 2010 ; 55 : 2062 – 2076 .

  • heart failure, acute
  • hemodynamics
  • congenital heart defects
  • vasodilators
  • heterogeneity
  • public health medicine
  • hypoperfusion
  • hospital admission
Month: Total Views:
December 2016 28
January 2017 25
February 2017 76
March 2017 184
April 2017 160
May 2017 85
June 2017 51
July 2017 87
August 2017 206
September 2017 199
October 2017 198
November 2017 319
December 2017 1,279
January 2018 1,816
February 2018 2,191
March 2018 3,341
April 2018 3,764
May 2018 2,961
June 2018 2,655
July 2018 2,006
August 2018 1,989
September 2018 2,362
October 2018 2,313
November 2018 2,356
December 2018 1,636
January 2019 1,729
February 2019 2,056
March 2019 3,650
April 2019 3,634
May 2019 4,130
June 2019 2,823
July 2019 2,940
August 2019 3,205
September 2019 2,921
October 2019 2,020
November 2019 1,754
December 2019 1,172
January 2020 1,513
February 2020 1,621
March 2020 1,821
April 2020 2,392
May 2020 1,523
June 2020 1,412
July 2020 1,355
August 2020 1,500
September 2020 1,852
October 2020 1,926
November 2020 1,334
December 2020 1,011
January 2021 1,191
February 2021 1,158
March 2021 1,512
April 2021 1,275
May 2021 1,044
June 2021 841
July 2021 702
August 2021 756
September 2021 1,260
October 2021 1,129
November 2021 1,053
December 2021 626
January 2022 701
February 2022 769
March 2022 1,147
April 2022 1,134
May 2022 931
June 2022 438
July 2022 451
August 2022 450
September 2022 696
October 2022 746
November 2022 575
December 2022 391
January 2023 390
February 2023 541
March 2023 646
April 2023 582
May 2023 553
June 2023 359
July 2023 368
August 2023 352
September 2023 517
October 2023 466
November 2023 435
December 2023 260
January 2024 357
February 2024 417
March 2024 439
April 2024 492
May 2024 437
June 2024 176

Email alerts

More on this topic, related articles in pubmed, citing articles via.

  • Recommend to Your Librarian

Affiliations

  • Online ISSN 1554-2815
  • Print ISSN 1520-765X
  • Copyright © 2024 European Society of Cardiology
  • About Oxford Academic
  • Publish journals with us
  • University press partners
  • What we publish
  • New features  
  • Open access
  • Institutional account management
  • Rights and permissions
  • Get help with access
  • Accessibility
  • Advertising
  • Media enquiries
  • Oxford University Press
  • Oxford Languages
  • University of Oxford

Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide

  • Copyright © 2024 Oxford University Press
  • Cookie settings
  • Cookie policy
  • Privacy policy
  • Legal notice

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

For full access to this pdf, sign in to an existing account, or purchase an annual subscription.

clinical presentation of acute heart failure

Heart Failure Clinical Presentation

  • Author: Ioana Dumitru, MD; Chief Editor: Gyanendra K Sharma, MD, FACC, FASE  more...
  • Sections Heart Failure
  • Practice Essentials
  • Pathophysiology
  • Epidemiology
  • Patient Education
  • Physical Examination
  • Predominant Right-Sided Heart Failure
  • Heart Failure in Children
  • Heart Failure Criteria, Classification, and Staging
  • ACC/AHA Staging
  • Approach Considerations
  • Routine Laboratory Tests
  • Natriuretic Peptides
  • Genetic Testing
  • Assessment of Hypoxemia
  • Electrocardiography
  • Chest Radiography
  • Echocardiography
  • CT Scanning and MRI
  • Nuclear Imaging
  • Catheterization and Angiography
  • Assessment of Functional Capacity
  • Nonpharmacologic Therapy
  • Pharmacologic Therapy
  • Acute Heart Failure Treatment
  • Treatment of Heart Failure with Preserved LVEF
  • Treatment of Right Ventricular Heart Failure
  • Electrophysiologic Intervention
  • Revascularization Procedures
  • Valvular Surgery
  • Ventricular Restoration
  • Extracorporeal Membrane Oxygenation
  • Ventricular Assist Devices
  • Heart Transplantation
  • Total Artificial Heart
  • Guidelines Summary
  • Screening and Genetic Testing
  • Diagnostic Procedures
  • Mechanical Circulatory Support Devices
  • Management of Acute Decompensated Heart Failure (ADHF)
  • 2024 ISHLT Guidelines on Heart Failure-Related Cardiogenic Shock
  • 2024 ESC Guidelines on Management of Dietary Sodium and Fluid Intake in Heart Failure
  • Medication Summary
  • Beta-Blockers, Alpha Activity
  • Beta-Blockers, Beta-1 Selective
  • ACE Inhibitors
  • Inotropic Agents
  • Vasodilators
  • B-type Natriuretic Peptides
  • I(f) Inhibitors
  • Angiotensin Receptor-Neprilysin Inhibitors (ARNi)
  • Diuretics, Loop
  • Diuretics, Thiazide
  • Diuretics, Other
  • Diuretics, Potassium-Sparing
  • Aldosterone Antagonists, Selective
  • SGLT2 Inhibitors
  • Dual SGLT1/2 Inhibitors
  • Soluble Guanylate Cyclase Stimulators
  • Alpha/Beta Adrenergic Agonists
  • Calcium Channel Blockers
  • Anticoagulants, Cardiovascular
  • Opioid Analgesics
  • Questions & Answers
  • Media Gallery

In evaluating patients with heart failure, the clinician should ask about the following comorbidities and/or risk factors:

  • Previous myocardial infarction
  • Valvular heart disease, familial heart disease
  • Alcohol use
  • Hypertension
  • Dyslipidemia
  • Coronary/peripheral vascular disease
  • Sleep-disordered breathing
  • Collagen vascular disease, rheumatic fever
  • Pheochromocytoma
  • Thyroid disease
  • Substance abuse (previous/current history)
  • History of chemotherapy/radiation to the chest

The New York Heart Association (NYHA) classification of heart failure is widely used in practice and in clinical studies to quantify clinical assessment of heart failure (see Heart Failure Criteria, Classification, and Staging ). Breathlessness, a cardinal symptom of left ventricular (LV) failure, may manifest with progressively increasing severity as the following:

Exertional dyspnea

Paroxysmal nocturnal dyspnea, dyspnea at rest.

  • Acute pulmonary edema

Other cardiac symptoms of heart failure include chest pain/pressure and palpitations. Common noncardiac signs and symptoms of heart failure include anorexia, nausea, weight loss, bloating, fatigue, weakness, oliguria, nocturia, and cerebral symptoms of varying severity, ranging from anxiety to memory impairment and confusion. Findings from the Framingham Heart Study suggested that subclinical cardiac dysfunction and noncardiac comorbidities are associated with increased incidence of heart failure, supporting the idea that heart failure is a progressive syndrome and that noncardiac factors are extremely important. [ 31 , 32 , 58 ]

Older patients with heart failure frequently have preserved ejection fraction and an atypical and/or delayed presentation. [ 59 ]

The principal difference between exertional dyspnea in patients who are healthy and exertional dyspnea in patients with heart failure is the degree of activity necessary to induce the symptom. As heart failure first develops, exertional dyspnea may simply appear to be an aggravation of the breathlessness that occurs in healthy persons during activity, but as LV failure advances, the intensity of exercise resulting in breathlessness progressively declines; however, subjective exercise capacity and objective measures of LV performance at rest in patients with heart failure are not closely correlated. Exertional dyspnea, in fact, may be absent in sedentary patients.

Orthopnea is an early symptom of heart failure and may be defined as dyspnea that develops in the recumbent position and is relieved with elevation of the head with pillows. As in the case of exertional dyspnea, the change in the number of pillows required is important. In the recumbent position, decreased pooling of blood in the lower extremities and abdomen occurs. Blood is displaced from the extrathoracic compartment to the thoracic compartment. The failing LV, operating on the flat portion of the Frank-Starling curve, cannot accept and pump out the extra volume of blood delivered to it without dilating. As a result, pulmonary venous and capillary pressures rise further, causing interstitial pulmonary edema, reduced pulmonary compliance, increased airway resistance, and dyspnea.

Orthopnea occurs rapidly, often within a minute or two of recumbency, and develops when the patient is awake. Orthopnea may occur in any condition in which the vital capacity is low. Marked ascites, regardless of its etiology, is an important cause of orthopnea. In advanced LV failure, orthopnea may be so severe that the patient cannot lie down and must sleep sitting up in a chair or slumped over a table.

Cough, particularly during recumbency, may be an "orthopnea equivalent." This nonproductive cough may be caused by pulmonary congestion and is relieved by the treatment of heart failure.

Paroxysmal nocturnal dyspnea usually occurs at night and is defined as the sudden awakening of the patient, after a couple of hours of sleep, with a feeling of severe anxiety, breathlessness, and suffocation. The patient may bolt upright in bed and gasp for breath. Bronchospasm increases ventilatory difficulty and the work of breathing and is a common complicating factor of paroxysmal nocturnal dyspnea. On chest auscultation, the bronchospasm associated with a heart failure exacerbation can be difficult to distinguish from an acute asthma exacerbation, although other clues from the cardiovascular examination should lead the examiner to the correct diagnosis. Both types of bronchospasm can be present in a single individual.

In contrast to orthopnea, which may be relieved by immediately sitting up in bed, paroxysmal nocturnal dyspnea may require 30 minutes or longer in this position for relief. Episodes may be so frightening that the patient may be afraid to resume sleeping, even after the symptoms have subsided.

Dyspnea at rest in heart failure is the result of the following mechanisms:

  • Decreased pulmonary function secondary to decreased compliance and increased airway resistance
  • Increased ventilatory drive secondary to hypoxemia due to increased pulmonary capillary wedge pressure (PCWP); ventilation/perfusion (V/Q) mismatching due to increased PCWP and low cardiac output; and increased carbon dioxide production
  • Respiratory muscle dysfunction, with decreased respiratory muscle strength, decreased endurance, and ischemia

Pulmonary edema

Acute pulmonary edema is defined as the sudden increase in PCWP (usually >25 mm Hg) as a result of acute and fulminant LV failure. It is a medical emergency and has a very dramatic clinical presentation. The patient appears extremely ill, poorly perfused, restless, sweaty, tachypneic, tachycardic, hypoxic, and coughing, with an increased work of breathing and using respiratory accessory muscles and with frothy sputum that on occasion is blood tinged.

Chest pain/pressure and palpitations

Chest pain/pressure may occur as a result of either primary myocardial ischemia from coronary disease or secondary myocardial ischemia from increased filling pressure, poor cardiac output (and, therefore, poor coronary diastolic filling), or hypotension and hypoxemia.

Palpitations are the sensation a patient has when the heart is racing. It can be secondary to sinus tachycardia due to decompensated heart failure, or more commonly, it is due to atrial or ventricular tachyarrhythmias.

Fatigue and weakness

Fatigue and weakness are often accompanied by a feeling of heaviness in the limbs and are generally related to poor perfusion of the skeletal muscles in patients with a lowered cardiac output. Although they are generally a constant feature of advanced heart failure, episodic fatigue and weakness are also common in earlier stages.

Nocturia and oliguria

Nocturia may occur relatively early in the course of heart failure. Recumbency reduces the deficit in cardiac output in relation to oxygen demand, renal vasoconstriction diminishes, and urine formation increases. Nocturia may be troublesome for patients with heart failure because it may prevent them from obtaining much-needed rest. Oliguria is a late finding in heart failure, and it is found in patients with markedly reduced cardiac output from severely reduced LV function.

Cerebral symptoms

The following may occur in elderly patients with advanced heart failure, particularly in those with cerebrovascular atherosclerosis:

  • Memory impairment
  • Bad dreams or nightmares
  • Rarely, psychosis with disorientation, delirium, or hallucinations

Patients with mild heart failure appear to be in no distress after a few minutes of rest, but they may be obviously dyspneic during and immediately after moderate activity. Patients with left ventricular (LV) failure may be dyspneic when lying flat without elevation of the head for more than a few minutes. Those with severe heart failure appear anxious and may exhibit signs of air hunger in this position.

Patients with a recent onset of heart failure are generally well nourished, but those with chronic severe heart failure are often malnourished and sometimes even cachectic. Chronic marked elevation of the systemic venous pressure may produce exophthalmos and severe tricuspid regurgitation, and it may lead to visible pulsation of the eyes and of the neck veins. Central cyanosis, icterus, and malar flush may be evident in patients with severe heart failure.

In mild or moderate heart failure, stroke volume is normal at rest; in severe heart failure, it is reduced, as reflected by a diminished pulse pressure and a dusky discoloration of the skin. With very severe heart failure, particularly if cardiac output has declined acutely, systolic arterial pressure may be reduced. The pulse may be weak, rapid, and thready; the proportional pulse pressure (pulse pressure/systolic pressure) may be markedly reduced. The proportional pulse pressure correlates reasonably well with cardiac output. In one study, when pulse pressure was less than 25%, it usually reflected a cardiac index of less than 2.2 L/min/m 2 . [ 60 ]

Ascites occurs in patients with increased pressure in the hepatic veins and in the veins draining into the peritoneum; it usually reflects long-standing systemic venous hypertension. Fever may be present in severe decompensated heart failure because of cutaneous vasoconstriction and impairment of heat loss.

Increased adrenergic activity is manifested by tachycardia, diaphoresis, pallor, peripheral cyanosis with pallor and coldness of the extremities, and obvious distention of the peripheral veins secondary to venoconstriction. Diastolic arterial pressure may be slightly elevated.

Rales heard over the lung bases are characteristic of heart failure that is of at least moderate severity. With acute pulmonary edema, rales are frequently accompanied by wheezing and expectoration of frothy, blood-tinged sputum. The absence of rales does not exclude elevation of pulmonary capillary pressure due to LV failure.

Systemic venous hypertension is manifested by jugular venous distention. Normally, jugular venous pressure declines with respiration; however, it increases in patients with heart failure, a finding known as the Kussmaul sign (also found in constrictive pericarditis ). This reflects an increase in right atrial pressure and, therefore, right-sided heart failure. In general, elevated jugular venous pressure is the most reliable indicator of fluid volume overload in older patients, and thorough evaluation is needed. [ 59 ]

The hepatojugular reflux is the distention of the jugular vein induced by applying manual pressure over the liver; the patient's torso should be positioned at a 45° angle. The hepatojugular reflux occurs in patients with elevated left-sided filling pressures and reflects elevated capillary wedge pressure and left-sided heart failure.

Although edema is a cardinal manifestation of heart failure, it does not correlate well with the level of systemic venous pressure. In patients with chronic LV failure and low cardiac output, extracellular fluid volume may be sufficiently expanded to cause edema in the presence of only slight elevations in systemic venous pressure. Usually, a substantial gain of extracellular fluid volume (ie, a minimum of 5 L in adults) must occur before peripheral edema develops. Edema in the absence of dyspnea or other signs of LV or right ventricular (RV) failure is not solely indicative of heart failure and can be observed in many other conditions, including chronic venous insufficiency, nephrotic syndrome, or other syndromes of hypoproteinemia or osmotic imbalance.

Hepatomegaly is prominent in patients with chronic right-sided heart failure, but it may occur rapidly in acute heart failure. When hepatomegaly occurs acutely, the liver is usually tender. In patients with considerable tricuspid regurgitation, a prominent systolic pulsation of the liver, attributable to an enlarged right atrial V wave, is often noted. A presystolic pulsation of the liver, attributable to an enlarged right atrial A wave, can occur in tricuspid stenosis, constrictive pericarditis, restrictive cardiomyopathy involving the right ventricle, and pulmonary hypertension (primary or secondary).

Hydrothorax is most commonly observed in patients with hypertension involving both the systemic and pulmonary circulation. It is usually bilateral, although when unilateral, it is usually confined to the right side of the chest. When hydrothorax develops, dyspnea usually intensifies because of further reductions in vital capacity.

Cardiac findings

Protodiastolic (S 3 ) gallop is the earliest cardiac physical finding in decompensated heart failure in the absence of severe mitral or tricuspid regurgitation or left-to-right shunts. The presence of an S 3 gallop in adults is important, pathologic, and often the most apparent finding on cardiac auscultation in patients with significant heart failure.

Cardiomegaly is a nonspecific finding that nonetheless occurs in most patients with chronic heart failure. Notable exceptions include heart failure from acute myocardial infarction, constrictive pericarditis, restrictive cardiomyopathy, valve or chordae tendineae rupture, or heart failure due to tachyarrhythmias or bradyarrhythmias.

Pulsus alternans (during pulse palpation, this is the alternation of one strong and one weak beat without a change in the cycle length) occurs most commonly in heart failure due to increased resistance to LV ejection, as occurs in hypertension, aortic stenosis, coronary atherosclerosis, and dilated cardiomyopathy. Pulsus alternans is usually associated with an S 3 gallop, signifies advanced myocardial disease, and often disappears with treatment of heart failure.

Accentuation of the P 2 heart sound is a cardinal sign of increased pulmonary artery pressure; it disappears or improves after treatment of heart failure. Mitral and tricuspid regurgitation murmurs are often present in patients with decompensated heart failure because of ventricular dilatatation. These murmurs often disappear or diminish when compensation is restored. Note that the correlation is poor between the intensity of the murmur of mitral regurgitation and its significance in patients with heart failure. Severe mitral regurgitation may be accompanied by an unimpressively soft murmur.

Cardiac cachexia is found in long-standing heart failure, particularly of the RV, because of anorexia from hepatic and intestinal congestion and sometimes because of digitalis toxicity. Occasionally, impaired intestinal absorption of fat occurs and, rarely, protein-losing enteropathy occurs. Patients with heart failure may also exhibit increased total metabolism secondary to augmentation of myocardial oxygen consumption, excessive work of breathing, low-grade fever, and elevated levels of circulating tumor necrosis factor (TNF).

Ascites, congestive hepatomegaly, and anasarca due to elevated right-sided heart pressures transmitted backward into the portal vein circulation may result in increased abdominal girth and epigastric and right upper quadrant (RUQ) abdominal pain. Other gastrointestinal symptoms, caused by congestion of the hepatic and gastrointestinal venous circulation, include anorexia, bloating, nausea, and constipation. In preterminal heart failure, inadequate bowel perfusion can cause abdominal pain, distention, and bloody stools. Distinguishing right-sided heart failure from hepatic failure is often clinically difficult.

Dyspnea, prominent in left ventricular failure, becomes less prominent in isolated right-sided heart failure because of the absence of pulmonary congestion. However, when cardiac output becomes markedly reduced in patients with terminal right-sided heart failure (as may occur in isolated right ventricular infarction and in the late stages of primary pulmonary hypertension and pulmonary thromboembolic disease), severe dyspnea may occur as a consequence of the reduced cardiac output, poor perfusion of respiratory muscles, hypoxemia, and metabolic acidosis.

In children, manifestations of heart failure vary with age. [ 61 ] Signs of pulmonary venous congestion in an infant generally include tachypnea, respiratory distress (retractions), grunting, and difficulty with feeding. Often, children with heart failure have diaphoresis during feedings, which is possibly related to a catecholamine surge that occurs when they are challenged with eating while in respiratory distress.

Right-sided venous congestion is characterized by hepatosplenomegaly and, less frequently, with edema or ascites . Jugular venous distention is not a reliable indicator of systemic venous congestion in infants, because the jugular veins are difficult to observe. Also, the distance from the right atrium to the angle of the jaw may be no more than 8-10 cm, even when the infant is sitting upright. Uncompensated heart failure in an infant primarily manifests as a failure to thrive. In severe cases, failure to thrive may be followed by signs of renal and hepatic failure.

In older children, left-sided venous congestion causes tachypnea, respiratory distress, and wheezing (cardiac asthma). Right-sided congestion may result in hepatosplenomegaly, jugular venous distention, edema, ascites, and/or pleural effusions. Uncompensated heart failure in older children may cause fatigue or lower-than-usual energy levels. Patients may complain of cool extremities, exercise intolerance, dizziness, or syncope.

For more information, see the Medscape Drugs & Diseases article Pediatric Congestive Heart Failure .

Framingham system for diagnosis of heart failure

In the Framingham system, the diagnosis of heart failure requires that either two major criteria or one major and two minor criteria be present concurrently, as shown in Table 1 below. [ 1 ] Minor criteria are accepted only if they cannot be attributed to another medical condition.

Table 1. Framingham Diagnostic Criteria for Heart Failure (Open Table in a new window)

Paroxysmal nocturnal dyspnea

Nocturnal cough

Weight loss of 4.5 kg in 5 days in response to treatment

Dyspnea on ordinary exertion

Neck vein distention

A decrease in vital capacity by one third the maximal value recorded

Rales

Pleural effusion

Acute pulmonary edema

Tachycardia (rate of 120 bpm)

Hepatojugular reflux

Hepatomegaly

S gallop

Bilateral ankle edema

Central venous pressure >16 cm water

 

Circulation time of ≥25 seconds

 

Radiographic cardiomegaly

 

Pulmonary edema, visceral congestion, or cardiomegaly at autopsy

 

Ho KK, Pinsky JL, Kannel WB, Levy D. The epidemiology of heart failure: the Framingham Study.

American College of Cardiology/American Heart Association (ACC/AHA) stage A patients are at high risk for heart failure; thus, this stage is now also known as "at risk for heart failure." [ 4 , 5 , 6 , 7 ]  Patients in this stage do not have structural heart disease or symptoms of heart failure. Thus, management in these cases focuses on prevention, through reduction of risk factors. Measures include the following [ 3 ] :

  • Treat hypertension (optimal blood pressure: < 130/80 mm Hg [ 62 ] )
  • Encourage smoking cessation
  • Treat lipid disorders
  • Encourage regular exercise
  • Discourage alcohol intake and illicit drug use

Patients who have a family history of dilated cardiomyopathy should be screened with a comprehensive history and physical examination together with echocardiography and transthoracic echocardiography every 2-5 years. [ 8 ]

ACC/AHA stage B patients are now also designated as "pre-heart failure.. [ 4 , 5 , 6 , 7 ]  These individuals are asymptomatic, with left ventricular (LV) dysfunction from previous myocardial infarction (MI), LV remodeling from LV hypertrophy (LVH), and asymptomatic valvular dysfunction, which includes patients with New York Heart Association (NYHA) class I heart failure (see Heart Failure Criteria, Classification, and Staging for a description of NYHA classes). [ 3 ] In addition to the heart failure education and aggressive risk factor modification used for stage A, treatment with an angiotensin-converting enzyme inhibitor/angiotensin-receptor blocker (ACEI/ARB) and/or beta-blockade is indicated.

Evaluation for coronary revascularization either percutaneously or surgically, as well as correction of valvular abnormalities, may be indicated. [ 3 ] Treatment with an implantable cardioverter-defibrillator (ICD) for primary prevention of sudden death in patients with an LV ejection fraction (LVEF) below 30% that is more than 40 days post-MI is reasonable if the expected survival is more than 1 year.

There is less evidence for implantation of an ICD in patients with nonischemic cardiomyopathy, an LVEF less than 30%, and no heart failure symptoms. There is no evidence for use of digoxin in these populations. [ 63 ] Aldosterone receptor blockade with eplerenone is indicated for post-MI LV dysfunction.

ACC/AHA stage C patients have structural heart disease and current or previous symptoms of heart failure; they are therefore designated as having "symptomatic heart failure." [ 4 , 5 , 6 , 7 ]  ACC/AHA stage C corresponds with NYHA class I-IV heart failure. The preventive measures used for stage A disease are indicated, as is dietary sodium restriction.

Drugs routinely used in these patients include ACEI/ARBs, beta-blockers, or angiotensin receptor–neprilysin inhibitors (ARNIs), in conjunction with evidence-based beta-blockers, and loop diuretics for fluid retention. [ 3 , 62 , 64 ] For selected patients, therapeutic measures include aldosterone receptor blockers, hydralazine and nitrates in combination, and cardiac resynchronization with or without an ICD (see Electrophysiologic Intervention ). [ 3 , 62 , 64 ]

A meta-analysis performed by Badve et al suggested that the survival benefit of treatment with beta-blockers extends to patients with chronic kidney disease and systolic heart failure (risk ratio 0.72). [ 65 ]

The 2016 and 2017 ACC/AHA focused updates to the 2013 guidelines added a class IIa recommendation for ivabradine, a sinoatrial node modulator, in patients with stage C heart failure. [ 62 , 64 ]  They indicate that ivabradine may reduce hospitalization for patients with symptomatic (NYHA class II-III) stable chronic heart failure with reduced ejection fraction (LVEF ≤35%) who are receiving recommended therapy, including a beta blocker at the maximum tolerated dose, and who are in sinus rhythm with a heart rate of 70 bpm or greater at rest. [ 62 , 64 ]

ACC/AHA stage D patients have refractory heart failure (NYHA class IV) that requires specialized interventions; they are in "advanced heart failure." [ 4 , 5 , 6 , 7 ] Therapy includes all the measures used in stages A, B, and C. Treatment considerations include heart transplantation or placement of an LV assist device in eligible patients; pulmonary catheterization; and options for end-of-life care. [ 3 ] For palliation of symptoms, continuous intravenous infusion of a positive inotrope may be considered.

Ho KK, Pinsky JL, Kannel WB, Levy D. The epidemiology of heart failure: the Framingham Study. J Am Coll Cardiol . 1993 Oct. 22(4 suppl A):6A-13A. [QxMD MEDLINE Link] . [Full Text] .

American Heart Association. Classes of heart failure. Available at http://www.heart.org/HEARTORG/Conditions/HeartFailure/AboutHeartFailure/Classes-of-Heart-Failure_UCM_306328_Article.jsp#.WUcGf-vyuHs . Updated: May 8, 2017; Accessed: June 18, 2017.

[Guideline] Yancy CW, Jessup M, Bozkurt B, et al, American College of Cardiology Foundation/American Heart Association Task Force on Practice Guidelines. 2013 ACCF/AHA guideline for the management of heart failure: a report of the American College of Cardiology Foundation/American Heart Association Task Force on practice guidelines. Circulation . 2013 Oct 15. 128(16):e240-327. [QxMD MEDLINE Link] . [Full Text] .

[Guideline] Heidenreich PA, Bozkurt B, Aguilar D, 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. J Am Coll Cardiol . 2022 May 3. 79 (17):e263-e421. [QxMD MEDLINE Link] . [Full Text] .

Heidenreich PA, Bozkurt B, Aguilar D, et al. 2022 AHA/ACC/HFSA Guideline for the management of heart failure: executive summary: s report of the American College of Cardiology/American Heart Association Joint Committee on Clinical Practice Guidelines. Circulation . 2022 May 3. 145 (18):e876-94. [QxMD MEDLINE Link] . [Full Text] .

Stiles S. New HF guidelines feature 'quad' therapy, tweaked terminology. Medscape Medical News. Available at https://www.medscape.com/viewarticle/971492 . April 3, 2022; Accessed: May 6, 2022.

Pina IL. Key points from the new heart failure guidelines [commentary]. TheHeart.org. Available at https://www.medscape.com/viewarticle/971949 . May 4, 2022; Accessed: May 6, 2022.

[Guideline] Ponikowski P, Voors AA, Anker SD, et al, for the Authors/Task Force Members. 2016 ESC guidelines for the diagnosis and treatment of acute and chronic heart failure: The task force for the diagnosis and treatment of acute and chronic heart failure of the European Society of Cardiology (ESC). Developed with the special contribution of the Heart Failure Association (HFA) of the ESC. Eur Heart J . 2016 Jul 14. 37(27):2129-200. [QxMD MEDLINE Link] . [Full Text] .

[Guideline] Lindenfeld J, Albert NM, Boehmer JP, et al, for the Heart Failure Society of America. HFSA 2010 comprehensive heart failure practice guideline. J Card Fail . 2010 Jun. 16(6):e1-194. [QxMD MEDLINE Link] . [Full Text] .

Stiles S. FDA approves heart sympathetic activity imaging agent for HF evaluation. Medscape News from WebMD. March 22, 2013. Available at http://www.medscape.com/viewarticle/781309 . Accessed: April 5, 2013.

Braunwald E. The pathogenesis of congestive heart failure: then and now. Medicine . 1991 Jan. 70(1):68-79. [Full Text] .

Braunwald E, Ross J Jr, Sonnenblick EH. Mechanisms of Contraction of the Normal and Failing Heart . 2nd ed. Boston: Little Brown & Co; 1976.

Greyson CR. Pathophysiology of right ventricular failure. Crit Care Med . 2008 Jan. 36(1 suppl):S57-65. [QxMD MEDLINE Link] .

Haddad F, Doyle R, Murphy DJ, Hunt SA. Right ventricular function in cardiovascular disease, part II: pathophysiology, clinical importance, and management of right ventricular failure. Circulation . 2008 Apr 1. 117(13):1717-31. [QxMD MEDLINE Link] .

Onwuanyi A, Taylor M. Acute decompensated heart failure: pathophysiology and treatment. Am J Cardiol . 2007 Mar 26. 99(6B):25D-30D. [QxMD MEDLINE Link] .

Ross J Jr, Braunwald E. Studies on Starling's law of the heart. IX. The effects of impeding venous return on performance of the normal and failing human left ventricle. Circulation . 1964 Nov. 30:719-27. [QxMD MEDLINE Link] . [Full Text] .

Gheorghiade M, Pang PS. Acute heart failure syndromes. J Am Coll Cardiol . 2009 Feb 17. 53(7):557-73. [QxMD MEDLINE Link] . [Full Text] .

Kajstura J, Leri A, Finato N, Di Loreto C, Beltrami CA, Anversa P. Myocyte proliferation in end-stage cardiac failure in humans. Proc Natl Acad Sci U S A . 1998 Jul 21. 95(15):8801-5. [QxMD MEDLINE Link] . [Full Text] .

Cohn JN. Structural basis for heart failure. Ventricular remodeling and its pharmacological inhibition. Circulation . 1995 May 15. 91(10):2504-7. [QxMD MEDLINE Link] . [Full Text] .

Cody RJ. Hormonal alterations in heart failure. In: Hosenpud JB, Greenberg BH, eds. Congestive Heart Failure: Pathophysiology, Diagnosis and Comprehensive Approach to Management . Philadelphia, PA: Lippincott Williams & Wilkins; 2000. 199-212.

Anversa P, Nadal-Ginard B. Myocyte renewal and ventricular remodelling. Nature . 2002 Jan 10. 415(6868):240-3. [QxMD MEDLINE Link] .

Leri A, Claudio PP, Li Q, et al. Stretch-mediated release of angiotensin II induces myocyte apoptosis by activating p53 that enhances the local renin-angiotensin system and decreases the Bcl-2-to-Bax protein ratio in the cell. J Clin Invest . 1998 Apr 1. 101(7):1326-42. [QxMD MEDLINE Link] . [Full Text] .

Kajstura J, Leri A, Castaldo C, Nadal-Ginard B, Anversa P. Myocyte growth in the failing heart. Surg Clin North Am . 2004 Feb. 84(1):161-77. [QxMD MEDLINE Link] .

Henes J, Rosenberger P. Systolic heart failure: diagnosis and therapy. Curr Opin Anaesthesiol . 2016 Feb. 29(1):55-60. [QxMD MEDLINE Link] .

Nicoara A, Jones-Haywood M. Diastolic heart failure: diagnosis and therapy. Curr Opin Anaesthesiol . 2016 Feb. 29(1):61-7. [QxMD MEDLINE Link] .

Feldman AM, Combes A, Wagner D, et al. The role of tumor necrosis factor in the pathophysiology of heart failure. J Am Coll Cardiol . 2000 Mar 1. 35(3):537-44. [QxMD MEDLINE Link] . [Full Text] .

Gary R, Davis L. Diastolic heart failure. Heart Lung . 2008 Nov-Dec. 37(6):405-16. [QxMD MEDLINE Link] .

Morris DA, Gailani M, Vaz Perez A, et al. Right ventricular myocardial systolic and diastolic dysfunction in heart failure with normal left ventricular ejection fraction. J Am Soc Echocardiogr . 2011 Aug. 24(8):886-97. [QxMD MEDLINE Link] .

Jousilahti P, Harald K, Jula A, et al, for the National Institute for Health and Welfare-THL - Helsinki - Finland. Salt intake and the risk of heart failure [abstract 1192]. Presented at: European Society of Cardiology 2017 Congress; August 27, 2017; Barcelona, Spain. Eur Heart J . Aug 2017. 38(suppl 1):240. [Full Text] .

Davenport L. High salt intake linked to increased heart-failure risk. Medscape News from WebMD. August 28, 2017. Available at http://www.medscape.com/viewarticle/884824 . Accessed: September 1, 2017.

Lam CS, Lyass A, Kraigher-Krainer E, et al. Cardiac dysfunction and noncardiac dysfunction as precursors of heart failure with reduced and preserved ejection fraction in the community. Circulation . 2011 Jul 5. 124(1):24-30. [QxMD MEDLINE Link] . [Full Text] .

Ho KK, Anderson KM, Kannel WB, Grossman W, Levy D. Survival after the onset of congestive heart failure in Framingham Heart Study subjects. Circulation . 1993 Jul. 88(1):107-15. [QxMD MEDLINE Link] . [Full Text] .

Halley CM, Houghtaling PL, Khalil MK, Thomas JD, Jaber WA. Mortality rate in patients with diastolic dysfunction and normal systolic function. Arch Intern Med . 2011 Jun 27. 171(12):1082-7. [QxMD MEDLINE Link] . [Full Text] .

Murphy RT, Starling RC. Genetics and cardiomyopathy: where are we now?. Cleve Clin J Med . 2005 Jun. 72(6):465-6, 469-70, 472-3 passim. [QxMD MEDLINE Link] .

Benjamin EJ, Blaha MJ, Chiuve SE, et al, for the American Heart Association Statistics Committee and Stroke Statistics Subcommittee. Heart disease and stroke statistics-2017 update: a report from the American Heart Association. Circulation . 2017 Mar 7. 135(10):e146-e603. [QxMD MEDLINE Link] . [Full Text] .

Ni H, Xu J. Recent trends in heart failure-related mortality: United States, 2000–2014. Centers for Disease Control and Prevention. December 31, 2015. Available at http://www.cdc.gov/nchs/data/databriefs/db231.htm . Accessed: January 5, 2016.

Brauser D. CDC: Heart-failure–related mortality rate climbs after decade-long decrease. Heartwire from Medscape. January 4, 2016. Available at http://www.medscape.com/viewarticle/856704 . Accessed: January 5, 2016.

Chen J, Normand SL, Wang Y, Krumholz HM. National and regional trends in heart failure hospitalization and mortality rates for Medicare beneficiaries, 1998-2008. JAMA . 2011 Oct 19. 306(15):1669-78. [QxMD MEDLINE Link] . [Full Text] .

Kolte D, Abbott JD, Aronow HD. Interventional therapies for heart failure in older adults. Heart Fail Clin . 2017 Jul. 13(3):535-70. [QxMD MEDLINE Link] .

Dharmarajan K, Rich MW. Epidemiology, pathophysiology, and prognosis of heart failure in older adults. Heart Fail Clin . 2017 Jul. 13(3):417-26. [QxMD MEDLINE Link] .

Jencks SF, Williams MV, Coleman EA. Rehospitalizations among patients in the Medicare fee-for-service program. N Engl J Med . 2009 Apr 2. 360(14):1418-28. [QxMD MEDLINE Link] . [Full Text] .

Stewart S, Wilkinson D, Hansen C, et al. Predominance of heart failure in the Heart of Soweto Study cohort: emerging challenges for urban African communities. Circulation . 2008 Dec 2. 118(23):2360-7. [QxMD MEDLINE Link] . [Full Text] .

Damasceno A, Cotter G, Dzudie A, Sliwa K, Mayosi BM. Heart failure in Sub-Saharan Africa: time for action. J Am Coll Cardiol . 2007 Oct 23. 50(17):1688-93. [QxMD MEDLINE Link] . [Full Text] .

Mbewu A, Mbanya JC. Cardiovascular diseases. In: Jamison DT, Feachem RG, Makgoba MW, et al, eds. Disease and mortality in Sub-Saharan Africa . 2nd ed. Washington, DC: World Bank Publications; 2006. 305-28.

Dries DL, Exner DV, Domanski MJ, Greenberg B, Stevenson LW. The prognostic implications of renal insufficiency in asymptomatic and symptomatic patients with left ventricular systolic dysfunction. J Am Coll Cardiol . 2000 Mar 1. 35(3):681-9. [QxMD MEDLINE Link] . [Full Text] .

Fonarow GC, Adams KF Jr, Abraham WT, et al, for the ADHERE Scientific Advisory Committee, Study Group and Investigators. Risk stratification for in-hospital mortality in acutely decompensated heart failure: classification and regression tree analysis. JAMA . 2005 Feb 2. 293(5):572-80. [QxMD MEDLINE Link] . [Full Text] .

Levy D, Kenchaiah S, Larson MG, et al. Long-term trends in the incidence of and survival with heart failure. N Engl J Med . 2002 Oct 31. 347 (18):1397-402. [QxMD MEDLINE Link] .

Lucas C, Johnson W, Hamilton MA, et al. Freedom from congestion predicts good survival despite previous class IV symptoms of heart failure. Am Heart J . 2000 Dec. 140(6):840-7. [QxMD MEDLINE Link] .

MacIntyre K, Capewell S, Stewart S, et al. Evidence of improving prognosis in heart failure: trends in case fatality in 66 547 patients hospitalized between 1986 and 1995. Circulation . 2000 Sep 5. 102(10):1126-31. [QxMD MEDLINE Link] . [Full Text] .

Ketchum ES, Levy WC. Establishing prognosis in heart failure: a multimarker approach. Prog Cardiovasc Dis . 2011 Sep-Oct. 54(2):86-96. [QxMD MEDLINE Link] .

van Diepen S, Bakal JA, McAlister FA, Ezekowitz JA. Mortality and readmission of patients with heart failure, atrial fibrillation, or coronary artery disease undergoing noncardiac surgery: an analysis of 38 047 patients. Circulation . 2011 Jul 19. 124(3):289-96. [QxMD MEDLINE Link] . [Full Text] .

Bursi F, McNallan SM, Redfield MM, et al. Pulmonary pressures and death in heart failure: a community study. J Am Coll Cardiol . 2012 Jan 17. 59(3):222-31. [QxMD MEDLINE Link] . [Full Text] .

Ho JE, Liu C, Lyass A, et al. Galectin-3, a marker of cardiac fibrosis, predicts incident heart failure in the community. J Am Coll Cardiol . 2012 Oct 2. 60(14):1249-56. [QxMD MEDLINE Link] .

Nayar P, Yu F, Chandak A, Kan GL, Lowes B, Apenteng BA. Risk factors for in-hospital mortality in heart failure patients: does rurality, payer or admission source matter?. J Rural Health . 2018 Dec. 34(1):103-8. [QxMD MEDLINE Link] .

Espersen C, Campbell RT, Claggett BL, et al. Predictors of heart failure readmission and all-cause mortality in patients with acute heart failure. Int J Cardiol . 2024 Jul 1. 406:132036. [QxMD MEDLINE Link] .

Dunlay SM, Eveleth JM, Shah ND, McNallan SM, Roger VL. Medication adherence among community-dwelling patients with heart failure. Mayo Clin Proc . 2011 Apr. 86(4):273-81. [QxMD MEDLINE Link] . [Full Text] .

DeWalt DA, Schillinger D, Ruo B, et al. Multisite randomized trial of a single-session versus multisession literacy-sensitive self-care intervention for patients with heart failure. Circulation . 2012 Jun 12. 125 (23):2854-62. [QxMD MEDLINE Link] .

Lainscak M, Cleland JG, Lenzen MJ, Follath F, Komajda M, Swedberg K. International variations in the treatment and co-morbidity of left ventricular systolic dysfunction: data from the EuroHeart Failure Survey. Eur J Heart Fail . 2007 Mar. 9(3):292-9. [QxMD MEDLINE Link] .

Panjrath G, Ahmed A. Diagnosis and management of heart failure in older adults. Heart Fail Clin . 2017 Jul. 13(3):427-44. [QxMD MEDLINE Link] .

Stevenson LW, Perloff JK. The limited reliability of physical signs for estimating hemodynamics in chronic heart failure. JAMA . 1989 Feb 10. 261(6):884-8. [QxMD MEDLINE Link] .

Steinhart B, Thorpe KE, Bayoumi AM, Moe G, Januzzi JL Jr, Mazer CD. Improving the diagnosis of acute heart failure using a validated prediction model. J Am Coll Cardiol . 2009 Oct 13. 54(16):1515-21. [QxMD MEDLINE Link] .

[Guideline] Yancy CW, Jessup M, Bozkurt B, et al. 2017 ACC/AHA/HFSA Focused Update of the 2013 ACCF/AHA Guideline for the Management of Heart Failure: A Report of the American College of Cardiology/American Heart Association Task Force on Clinical Practice Guidelines and the Heart Failure Society of America. Circulation . 2017 Aug 8. 136(6):e137-e161. [QxMD MEDLINE Link] . [Full Text] .

Rich MW, McSherry F, Williford WO, Yusuf S, for the Digitalis Investigation Group. Effect of age on mortality, hospitalizations and response to digoxin in patients with heart failure: the DIG study. J Am Coll Cardiol . 2001 Sep. 38(3):806-13. [QxMD MEDLINE Link] .

[Guideline] Yancy CW, Jessup M, Bozkurt B, et al, for the Writing Committee Members. 2016 ACC/AHA/HFSA focused update on new pharmacological therapy for heart failure: an update of the 2013 ACCF/AHA guideline for the management of heart failure: a report of the American College of Cardiology/American Heart Association Task Force on Clinical Practice Guidelines and the Heart Failure Society of America. Circulation . 2016 Sep 27. 134(13):e282-93. [QxMD MEDLINE Link] . [Full Text] .

Badve SV, Roberts MA, Hawley CM, et al. Effects of beta-adrenergic antagonists in patients with chronic kidney disease: a systematic review and meta-analysis. J Am Coll Cardiol . 2011 Sep 6. 58(11):1152-61. [QxMD MEDLINE Link] .

Maisel AS, Krishnaswamy P, Nowak RM, et al, for the Breathing Not Properly Multinational Study Investigators. Rapid measurement of B-type natriuretic peptide in the emergency diagnosis of heart failure. N Engl J Med . 2002 Jul 18. 347(3):161-7. [QxMD MEDLINE Link] .

Januzzi JL Jr, Camargo CA, Anwaruddin S, et al. The N-terminal pro-BNP investigation of dyspnea in the emergency department (PRIDE) study. Am J Cardiol . 2005 Apr 15. 95(8):948-54. [QxMD MEDLINE Link] .

Wang CS, FitzGerald JM, Schulzer M, Mak E, Ayas NT. Does this dyspneic patient in the emergency department have congestive heart failure?. JAMA . 2005 Oct 19. 294(15):1944-56. [QxMD MEDLINE Link] .

[Guideline] McDonagh TA, Metra M, Adamo M, et al, for the ESC Scientific Document Group. 2021 ESC Guidelines for the diagnosis and treatment of acute and chronic heart failure. Eur Heart J . 2021 Sep 21. 42(36):3599-726. [QxMD MEDLINE Link] . [Full Text] .

Groenveld HF, Januzzi JL, Damman K, et al. Anemia and mortality in heart failure patients a systematic review and meta-analysis. J Am Coll Cardiol . 2008 Sep 2. 52(10):818-27. [QxMD MEDLINE Link] .

Grote Beverborg N, van Veldhuisen DJ, van der Meer P. Anemia in heart failure: still relevant?. JACC Heart Fail . 2018 Mar. 6(3):201-8. [QxMD MEDLINE Link] . [Full Text] .

Beedkar A, Parikh R, Deshmukh P. Heart failure and the iron deficiency. J Assoc Physicians India . 2017 Nov. 65(11):79-80. [QxMD MEDLINE Link] .

Hoes MF, Grote Beverborg N, Kijlstra JD, et al. Iron deficiency impairs contractility of human cardiomyocytes through decreased mitochondrial function. Eur J Heart Fail . 2018 May. 20(5):910-9. [QxMD MEDLINE Link] . [Full Text] .

Rocha BML, Cunha GJL, Menezes Falcao LF. The burden of iron deficiency in heart failure: therapeutic approach. J Am Coll Cardiol . 2018 Feb 20. 71(7):782-93. [QxMD MEDLINE Link] .

Maisel AS, McCord J, Nowak RM, et al, for the Breathing Not Properly Multinational Study Investigators. Bedside B-Type natriuretic peptide in the emergency diagnosis of heart failure with reduced or preserved ejection fraction. Results from the Breathing Not Properly Multinational Study. J Am Coll Cardiol . 2003 Jun 4. 41(11):2010-7. [QxMD MEDLINE Link] . [Full Text] .

Januzzi JL, van Kimmenade R, Lainchbury J, et al. NT-proBNP testing for diagnosis and short-term prognosis in acute destabilized heart failure: an international pooled analysis of 1256 patients: the International Collaborative of NT-proBNP Study. Eur Heart J . 2006 Feb. 27(3):330-7. [QxMD MEDLINE Link] .

Maeda K, Tsutamoto T, Wada A, Hisanaga T, Kinoshita M. Plasma brain natriuretic peptide as a biochemical marker of high left ventricular end-diastolic pressure in patients with symptomatic left ventricular dysfunction. Am Heart J . 1998 May. 135(5 Pt 1):825-32. [QxMD MEDLINE Link] .

Fisher C, Berry C, Blue L, Morton JJ, McMurray J. N-terminal pro B type natriuretic peptide, but not the new putative cardiac hormone relaxin, predicts prognosis in patients with chronic heart failure. Heart . 2003 Aug. 89(8):879-81. [QxMD MEDLINE Link] . [Full Text] .

Hall C, Rouleau JL, Moye L, et al. N-terminal proatrial natriuretic factor. An independent predictor of long-term prognosis after myocardial infarction. Circulation . 1994 May. 89(5):1934-42. [QxMD MEDLINE Link] .

Andersson B, Hall C. N-terminal proatrial natriuretic peptide and prognosis in patients with heart failure and preserved systolic function. J Card Fail . 2000 Sep. 6(3):208-13. [QxMD MEDLINE Link] .

Chen HH, Burnett JC. Natriuretic peptides in the pathophysiology of congestive heart failure. Curr Cardiol Rep . 2000 May. 2(3):198-205. [QxMD MEDLINE Link] .

Cheng V, Kazanagra R, Garcia A, et al. A rapid bedside test for B-type peptide predicts treatment outcomes in patients admitted for decompensated heart failure: a pilot study. J Am Coll Cardiol . 2001 Feb. 37(2):386-91. [QxMD MEDLINE Link] .

Cowie MR, Struthers AD, Wood DA, et al. Value of natriuretic peptides in assessment of patients with possible new heart failure in primary care. Lancet . 1997 Nov 8. 350(9088):1349-53. [QxMD MEDLINE Link] .

Dao Q, Krishnaswamy P, Kazanegra R, et al. Utility of B-type natriuretic peptide in the diagnosis of congestive heart failure in an urgent-care setting. J Am Coll Cardiol . 2001 Feb. 37(2):379-85. [QxMD MEDLINE Link] .

Maeda K, Tsutamoto T, Wada A, Hisanaga T, Kinoshita M. Plasma brain natriuretic peptide as a biochemical marker of high left ventricular end-diastolic pressure in patients with symptomatic left ventricular dysfunction. Am Heart J . 1998 May. 135(5 pt 1):825-32. [QxMD MEDLINE Link] .

Maisel AS, Koon J, Hope J, et al. A rapid bedside test for brain natriuretic peptide accurately predicts cardiac function in patients referred for echocardiography. Am Heart J . 2001. 141:374-9.

Masson S, Vago T, Baldi G, et al. Comparative measurement of N-terminal pro-brain natriuretic peptide and brain natriuretic peptide in ambulatory patients with heart failure. Clin Chem Lab Med . 2002 Aug. 40(8):761-3. [QxMD MEDLINE Link] .

Song BG, Jeon ES, Kim YH, et al. Correlation between levels of N-terminal pro-B-type natriuretic peptide and degrees of heart failure. Korean J Intern Med . 2005 Mar. 20(1):26-32. [QxMD MEDLINE Link] .

Hobbs FD, Davis RC, Roalfe AK, Hare R, Davies MK, Kenkre JE. Reliability of N-terminal pro-brain natriuretic peptide assay in diagnosis of heart failure: cohort study in representative and high risk community populations. BMJ . 2002 Jun 22. 324(7352):1498. [QxMD MEDLINE Link] . [Full Text] .

Redfield MM, Rodeheffer RJ, Jacobsen SJ, Mahoney DW, Bailey KR, Burnett JC Jr. Plasma brain natriuretic peptide concentration: impact of age and gender. J Am Coll Cardiol . 2002 Sep 4. 40(5):976-82. [QxMD MEDLINE Link] .

St Peter JV, Hartley GG, Murakami MM, Apple FS. B-type natriuretic peptide (BNP) and N-terminal pro-BNP in obese patients without heart failure: relationship to body mass index and gastric bypass surgery. Clin Chem . 2006 Apr. 52 (4):680-5. [QxMD MEDLINE Link] .

Rivera M, Cortes R, Salvador A, et al. Obese subjects with heart failure have lower N-terminal pro-brain natriuretic peptide plasma levels irrespective of aetiology. Eur J Heart Fail . 2005 Dec. 7(7):1168-70. [QxMD MEDLINE Link] .

Hermann-Arnhof KM, Hanusch-Enserer U, Kaestenbauer T, et al. N-terminal pro-B-type natriuretic peptide as an indicator of possible cardiovascular disease in severely obese individuals: comparison with patients in different stages of heart failure. Clin Chem . 2005 Jan. 51(1):138-43. [QxMD MEDLINE Link] .

Seino Y, Ogawa A, Yamashita T, et al. Application of NT-proBNP and BNP measurements in cardiac care: a more discerning marker for the detection and evaluation of heart failure. Eur J Heart Fail . 2004 Mar 15. 6(3):295-300. [QxMD MEDLINE Link] .

Colucci WS, Elkayam U, Horton DP, et al. Intravenous nesiritide, a natriuretic peptide, in the treatment of decompensated congestive heart failure. Nesiritide Study Group. N Engl J Med . 2000 Jul 27. 343(4):246-53. [QxMD MEDLINE Link] .

Ezekowitz JA, Hernandez AF, Starling RC, et al. Standardizing care for acute decompensated heart failure in a large megatrial: the approach for the Acute Studies of Clinical Effectiveness of Nesiritide in Subjects with Decompensated Heart Failure (ASCEND-HF). Am Heart J . 2009 Feb. 157(2):219-28. [QxMD MEDLINE Link] .

Mills RM, LeJemtel TH, Horton DP, et al. Sustained hemodynamic effects of an infusion of nesiritide (human b-type natriuretic peptide) in heart failure: a randomized, double-blind, placebo-controlled clinical trial. Natrecor Study Group. J Am Coll Cardiol . 1999 Jul. 34(1):155-62. [QxMD MEDLINE Link] .

Silver MA, Horton DP, Ghali JK, Elkayam U. Effect of nesiritide versus dobutamine on short-term outcomes in the treatment of patients with acutely decompensated heart failure. J Am Coll Cardiol . 2002 Mar 6. 39(5):798-803. [QxMD MEDLINE Link] .

Miller WL, Hartman KA, Burritt MF, Borgeson DD, Burnett JC Jr, Jaffe AS. Biomarker responses during and after treatment with nesiritide infusion in patients with decompensated chronic heart failure. Clin Chem . 2005 Mar. 51(3):569-77. [QxMD MEDLINE Link] .

Fitzgerald RL, Cremo R, Gardetto N, et al. Effect of nesiritide in combination with standard therapy on serum concentrations of natriuretic peptides in patients admitted for decompensated congestive heart failure. Am Heart J . 2005 Sep. 150(3):471-7. [QxMD MEDLINE Link] .

Michels VV, Moll PP, Miller FA, et al. The frequency of familial dilated cardiomyopathy in a series of patients with idiopathic dilated cardiomyopathy. N Engl J Med . 1992 Jan 9. 326(2):77-82. [QxMD MEDLINE Link] .

Baig MK, Goldman JH, Caforio AL, Coonar AS, Keeling PJ, McKenna WJ. Familial dilated cardiomyopathy: cardiac abnormalities are common in asymptomatic relatives and may represent early disease. J Am Coll Cardiol . 1998 Jan. 31(1):195-201. [QxMD MEDLINE Link] .

Grunig E, Tasman JA, Kucherer H, Franz W, Kubler W, Katus HA. Frequency and phenotypes of familial dilated cardiomyopathy. J Am Coll Cardiol . 1998 Jan. 31 (1):186-94. [QxMD MEDLINE Link] .

McNally E, MacLeod H, Dellefave-Castillo L, et al. Arrhythmogenic right ventricular cardiomyopathy. GeneReviews . 1993. [QxMD MEDLINE Link] . [Full Text] .

[Guideline] Cheitlin MD, Alpert JS, Armstrong WF, et al. ACC/AHA guidelines for the clinical application of echocardiography. A report of the American College of Cardiology/American Heart Association Task Force on Practice Guidelines (Committee on Clinical Application of Echocardiography). Developed in collaboration with the American Society of Echocardiography. Circulation . 1997 Mar 18. 95(6):1686-744. [QxMD MEDLINE Link] . [Full Text] .

Patel AR, Alsheikh-Ali AA, Mukherjee J, et al. 3D echocardiography to evaluate right atrial pressure in acutely decompensated heart failure correlation with invasive hemodynamics. JACC Cardiovasc Imaging . 2011 Sep. 4 (9):938-45. [QxMD MEDLINE Link] .

Prior D, Coller J. Echocardiography in heart failure - a guide for general practice. Aust Fam Physician . 2010 Dec. 39(12):904-9. [QxMD MEDLINE Link] .

Kirkpatrick JN, Wiegers SE, Lang RM. Left ventricular assist devices and other devices for end-stage heart failure: utility of echocardiography. Curr Cardiol Rep . 2010 May. 12(3):257-64. [QxMD MEDLINE Link] .

Abraham J, Abraham TP. The role of echocardiography in hemodynamic assessment in heart failure. Heart Fail Clin . 2009 Apr. 5(2):191-208. [QxMD MEDLINE Link] .

Gupta VA, Nanda NC, Sorrell VL. Role of echocardiography in the diagnostic assessment and etiology of heart failure in older adults: opacify, quantify, and rectify. Heart Fail Clin . 2017 Jul. 13(3):445-66. [QxMD MEDLINE Link] .

Meersch M, Schmidt C, Zarbock A. Echophysiology: the transesophageal echo probe as a noninvasive Swan-Ganz catheter. Curr Opin Anaesthesiol . 2016 Feb. 29(1):36-45. [QxMD MEDLINE Link] .

Kim RJ, Wu E, Rafael A, et al. The use of contrast-enhanced magnetic resonance imaging to identify reversible myocardial dysfunction. N Engl J Med . 2000 Nov 16. 343(20):1445-53. [QxMD MEDLINE Link] .

[Guideline] Ritchie JL, Bateman TM, Bonow RO, et al. Guidelines for clinical use of cardiac radionuclide imaging. Report of the American College of Cardiology/American Heart Association Task Force on Assessment of Diagnostic and Therapeutic Cardiovascular Procedures (Committee on Radionuclide Imaging), developed in collaboration with the American Society of Nuclear Cardiology. J Am Coll Cardiol . 1995 Feb. 25(2):521-47. [QxMD MEDLINE Link] .

Taillefer R, DePuey EG, Udelson JE, Beller GA, Latour Y, Reeves F. Comparative diagnostic accuracy of Tl-201 and Tc-99m sestamibi SPECT imaging (perfusion and ECG-gated SPECT) in detecting coronary artery disease in women. J Am Coll Cardiol . 1997 Jan. 29(1):69-77. [QxMD MEDLINE Link] .

Bonow RO, Maurer G, Lee KL, et al, for the STICH Trial Investigators. Myocardial viability and survival in ischemic left ventricular dysfunction. N Engl J Med . 2011 Apr 28. 364(17):1617-25. [QxMD MEDLINE Link] .

Binanay C, Califf RM, Hasselblad V, et al, for the ESCAPE Investigators and ESCAPE Study Coordinators. Evaluation study of congestive heart failure and pulmonary artery catheterization effectiveness: the ESCAPE trial. JAMA . 2005 Oct 5. 294(13):1625-33. [QxMD MEDLINE Link] .

[Guideline] Dickstein K, Vardas PE, Auricchio A, et al al for the Task Force on Acute Heart Failure of the European Society of Cardiology. 2010 Focused Update of ESC Guidelines on device therapy in heart failure: an update of the 2008 ESC Guidelines for the diagnosis and treatment of acute and chronic heart failure and the 2007 ESC guidelines for cardiac and resynchronization therapy. Developed with the special contribution of the Heart Failure Association and the European Heart Rhythm Association. Eur Heart J . 2010 Nov. 31(21):2677-87. [QxMD MEDLINE Link] . [Full Text] .

Bitter T, Westerheide N, Prinz C, et. Cheyne-Stokes respiration and obstructive sleep apnoea are independent risk factors for malignant ventricular arrhythmias requiring appropriate cardioverter-defibrillator therapies in patients with congestive heart failure. Eur Heart J . 2011 Jan. 32(1):61-74. [QxMD MEDLINE Link] .

Ronco C, Haapio M, House AA, Anavekar N, Bellomo R. Cardiorenal syndrome. J Am Coll Cardiol . 2008 Nov 4. 52(19):1527-39. [QxMD MEDLINE Link] .

House AA, Haapio M, Lassus J, Bellomo R, Ronco C. Therapeutic strategies for heart failure in cardiorenal syndromes. Am J Kidney Dis . 2010 Oct. 56(4):759-73. [QxMD MEDLINE Link] .

Giamouzis G, Butler J, Starling RC, et al. Impact of dopamine infusion on renal function in hospitalized heart failure patients: results of the Dopamine in Acute Decompensated Heart Failure (DAD-HF) Trial. J Card Fail . 2010 Dec. 16(12):922-30. [QxMD MEDLINE Link] .

Pleister AP, Baliga RR, Haas GJ. Acute study of clinical effectiveness of nesiritide in decompensated heart failure: nesiritide redux. Curr Heart Fail Rep . 2011 Sep. 8(3):226-32. [QxMD MEDLINE Link] .

O'Connor CM, Starling RC, Hernandez AF, et al. Effect of nesiritide in patients with acute decompensated heart failure. N Engl J Med . 2011 Jul 7. 365(1):32-43. [QxMD MEDLINE Link] .

Konstam MA, Gheorghiade M, Burnett JC Jr, et al, for the Efficacy of Vasopressin Antagonism in Heart Failure Outcome Study With Tolvaptan (EVEREST) Investigators. Effects of oral tolvaptan in patients hospitalized for worsening heart failure: the EVEREST Outcome Trial. JAMA . 2007 Mar 28. 297(12):1319-31. [QxMD MEDLINE Link] .

Massie BM, O'Connor CM, Metra M, et al, for the PROTECT Investigators and Committees. Rolofylline, an adenosine A1-receptor antagonist, in acute heart failure. N Engl J Med . 2010 Oct 7. 363(15):1419-28. [QxMD MEDLINE Link] .

Roy D, Talajic M, Nattel S, et al, for the Atrial Fibrillation and Congestive Heart Failure Investigators. Rhythm control versus rate control for atrial fibrillation and heart failure. N Engl J Med . 2008 Jun 19. 358(25):2667-77. [QxMD MEDLINE Link] .

Wilton SB, Fundytus A, Ghali WA, et al. Meta-analysis of the effectiveness and safety of catheter ablation of atrial fibrillation in patients with versus without left ventricular systolic dysfunction. Am J Cardiol . 2010 Nov 1. 106(9):1284-91. [QxMD MEDLINE Link] .

MacDonald MR, Connelly DT, Hawkins NM, et al. Radiofrequency ablation for persistent atrial fibrillation in patients with advanced heart failure and severe left ventricular systolic dysfunction: a randomised controlled trial. Heart . 2011 May. 97(9):740-7. [QxMD MEDLINE Link] .

Chen YM, Li ZB, Zhu M, Cao YM. Effects of exercise training on left ventricular remodelling in heart failure patients: an updated meta-analysis of randomised controlled trials. Int J Clin Pract . 2012 Aug. 66(8):782-91. [QxMD MEDLINE Link] .

Mozaffarian D, Lemaitre RN, King IB, et al. Circulating long-chain omega-3 fatty acids and incidence of congestive heart failure in older adults: the cardiovascular health study: a cohort study. Ann Intern Med . 2011 Aug 2. 155(3):160-70. [QxMD MEDLINE Link] .

Marchioli R, Levantesi G, Silletta MG, et al, for the GISSI-HF Investigators. Effect of n-3 polyunsaturated fatty acids and rosuvastatin in patients with heart failure: results of the GISSI-HF trial. Expert Rev Cardiovasc Ther . 2009 Jul. 7 (7):735-48. [QxMD MEDLINE Link] .

Armstrong PW, Pieske B, Anstrom KJ, et al, for the VICTORIA Study Group. Vericiguat in patients with heart failure and reduced ejection fraction. N Engl J Med . 2020 May 14. 382 (20):1883-93. [QxMD MEDLINE Link] . [Full Text] .

Swedberg K, Komajda M, Böhm M, et al, for the SHIFT Investigators. Ivabradine and outcomes in chronic heart failure (SHIFT): a randomised placebo-controlled study. Lancet . 2010 Sep 11. 376 (9744):875-85. [QxMD MEDLINE Link] . [Full Text] .

Borer JS, Bohm M, Ford I, et al, for the SHIFT Investigators. Effect of ivabradine on recurrent hospitalization for worsening heart failure in patients with chronic systolic heart failure: the SHIFT Study. Eur Heart J . 2012 Nov. 33 (22):2813-20. [QxMD MEDLINE Link] . [Full Text] .

McMurray JJ, Packer M, Desai AS, et al, PARADIGM-HF Investigators and Committees. Angiotensin-neprilysin inhibition versus enalapril in heart failure. N Engl J Med . 2014 Sep 11. 371 (11):993-1004. [QxMD MEDLINE Link] . [Full Text] .

Solomon SD, McMurray JJV, Anand IS, et al, for the PARAGON-HF Investigators and Committees. Angiotensin-neprilysin inhibition in heart failure with preserved ejection fraction. N Engl J Med . 2019 Oct 24. 381 (17):1609-20. [QxMD MEDLINE Link] . [Full Text] .

Solomon SD, Zile M, Pieske B, et al, for the Prospective comparison of ARNI with ARB on Management Of heart failUre with preserved ejectioN fracTion (PARAMOUNT) Investigators. The angiotensin receptor neprilysin inhibitor LCZ696 in heart failure with preserved ejection fraction: a phase 2 double-blind randomised controlled trial. Lancet . 2012 Oct 20. 380 (9851):1387-95. [QxMD MEDLINE Link] .

Entresto (sacubitril/valsartan) [package insert]. East Hanover, NJ: Novartis Pharmaceuticals Corp. October 2019. Available at [Full Text] .

Packer M, Anker SD, Butler J, et al, for the EMPEROR-Reduced Trial Investigators. Cardiovascular and renal outcomes with empagliflozin in heart failure. N Engl J Med . 2020 Oct 8. 383 (15):1413-24. [QxMD MEDLINE Link] . [Full Text] .

McMurray JJV, Solomon SD, Inzucchi SE, et al, for the DAPA-HF Trial Committees and Investigators. Dapagliflozin in patients with heart failure and reduced ejection fraction. N Engl J Med . 2019 Nov 21. 381 (21):1995-2008. [QxMD MEDLINE Link] . [Full Text] .

Anker SD, Butler J, Filippatos G, et al, for the EMPEROR-Preserved Trial Investigators. Empagliflozin in heart failure with a preserved ejection fraction. N Engl J Med . 2021 Oct 14. 385 (16):1451-61. [QxMD MEDLINE Link] .

Solomon SD, McMurray JJV, Claggett B, et al, for the DELIVER Trial Committees and Investigators. Dapagliflozin in heart failure with mildly reduced or preserved ejection fraction. N Engl J Med . 2022 Sep 22. 387(12):1089-98. [QxMD MEDLINE Link] . [Full Text] .

Bhatt DL, Szarek M, Steg PG, et al, for the SOLOIST-WHF Trial Investigators. Sotagliflozin in patients with diabetes and recent worsening heart failure. N Engl J Med . 2021 Jan 14. 384(2):117-28. [QxMD MEDLINE Link] . [Full Text] .

Bhatt DL, Szarek M, Pitt B, et al, for the SCORED Investigators. Sotagliflozin in patients with diabetes and chronic kidney disease. N Engl J Med . 2021 Jan 14. 384(2):129-39. [QxMD MEDLINE Link] . [Full Text] .

Massie BM, Collins JF, Ammon SE, et al, for the WATCH Trial Investigators. Randomized trial of warfarin, aspirin, and clopidogrel in patients with chronic heart failure: the Warfarin and Antiplatelet Therapy in Chronic Heart Failure (WATCH) trial. Circulation . 2009 Mar 31. 119 (12):1616-24. [QxMD MEDLINE Link] .

Freeman JV, Yang J, Sung SH, Hlatky MA, Go AS. Effectiveness and safety of digoxin among contemporary adults with incident systolic heart failure. Circ Cardiovasc Qual Outcomes . 2013 Sep 1. 6 (5):525-33. [QxMD MEDLINE Link] .

Konstantinou DM, Karvounis H, Giannakoulas G. Digoxin in heart failure with a reduced ejection fraction: a risk factor or a risk marker. Cardiology . 2016. 134 (3):311-9. [QxMD MEDLINE Link] .

Owan TE, Hodge DO, Herges RM, Jacobsen SJ, Roger VL, Redfield MM. Trends in prevalence and outcome of heart failure with preserved ejection fraction. N Engl J Med . 2006 Jul 20. 355 (3):251-9. [QxMD MEDLINE Link] .

Hogg K, Swedberg K, McMurray J. Heart failure with preserved left ventricular systolic function; epidemiology, clinical characteristics, and prognosis. J Am Coll Cardiol . 2004 Feb 4. 43 (3):317-27. [QxMD MEDLINE Link] .

Gheorghiade M, Abraham WT, Albert NM, et al, for the OPTIMIZE-HF Investigators and Coordinators. Systolic blood pressure at admission, clinical characteristics, and outcomes in patients hospitalized with acute heart failure. JAMA . 2006 Nov 8. 296 (18):2217-26. [QxMD MEDLINE Link] .

Masip J, Roque M, Sanchez B, Fernandez R, Subirana M, Exposito JA. Noninvasive ventilation in acute cardiogenic pulmonary edema: systematic review and meta-analysis. JAMA . 2005 Dec 28. 294 (24):3124-30. [QxMD MEDLINE Link] .

Peter JV, Moran JL, Phillips-Hughes J, Graham P, Bersten AD. Effect of non-invasive positive pressure ventilation (NIPPV) on mortality in patients with acute cardiogenic pulmonary oedema: a meta-analysis. Lancet . 2006 Apr 8. 367 (9517):1155-63. [QxMD MEDLINE Link] .

Winck JC, Azevedo LF, Costa-Pereira A, Antonelli M, Wyatt JC. Efficacy and safety of non-invasive ventilation in the treatment of acute cardiogenic pulmonary edema--a systematic review and meta-analysis. Crit Care . 2006. 10 (2):R69. [QxMD MEDLINE Link] . [Full Text] .

Vital FM, Saconato H, Ladeira MT, et al. Non-invasive positive pressure ventilation (CPAP or bilevel NPPV) for cardiogenic pulmonary edema. Cochrane Database Syst Rev . 2008 Jul 16. CD005351. [QxMD MEDLINE Link] .

Maeder MT, Kaye DM. Heart failure with normal left ventricular ejection fraction. J Am Coll Cardiol . 2009 Mar 17. 53 (11):905-18. [QxMD MEDLINE Link] .

Gray A, Goodacre S, Newby DE, et al, for the 3CPO Trialists. Noninvasive ventilation in acute cardiogenic pulmonary edema. N Engl J Med . 2008 Jul 10. 359 (2):142-51. [QxMD MEDLINE Link] .

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 Jun 4. 316 (23):1429-35. [QxMD MEDLINE Link] .

Yusuf S, Pitt B, Davis CE, Hood WB, Cohn JN, for the SOLVD Investigators. Effect of enalapril on survival in patients with reduced left ventricular ejection fractions and congestive heart failure. N Engl J Med . 1991 Aug 1. 325 (5):293-302. [QxMD MEDLINE Link] .

Zannad F, Alla F, Dousset B, Perez A, Pitt B. Limitation of excessive extracellular matrix turnover may contribute to survival benefit of spironolactone therapy in patients with congestive heart failure: insights from the randomized aldactone evaluation study (RALES). Rales Investigators. Circulation . 2000 Nov 28. 102 (22):2700-6. [QxMD MEDLINE Link] .

Sodhi N, Lasala JM. Mechanical circulatory support in acute decompensated heart failure and shock. Interv Cardiol Clin . 2017 Jul. 6 (3):387-405. [QxMD MEDLINE Link] .

Velazquez EJ, Lee KL, Jones RH, et al, for the STICHES Investigators. Coronary-artery bypass surgery in patients with ischemic cardiomyopathy. N Engl J Med . 2016 Apr 21. 374(16):1511-20. [QxMD MEDLINE Link] .

Busko M. Clevidipine shows promise for acute HF with high BP. Medscape Medical News from WebMD. February 10, 2014. Available at http://www.medscape.com/viewarticle/820377 . Accessed: February 18, 2014.

Felker GM, Lee KL, Bull DA, et al, for the NHLBI Heart Failure Clinical Research Network. Diuretic strategies in patients with acute decompensated heart failure. N Engl J Med . 2011 Mar 3. 364 (9):797-805. [QxMD MEDLINE Link] .

Liu PP. Cardiorenal syndrome in heart failure: a cardiologist's perspective. Can J Cardiol . 2008 Jul. 24 suppl B:25B-9B. [QxMD MEDLINE Link] . [Full Text] .

Kramer BK, Schweda F, Riegger GA. Diuretic treatment and diuretic resistance in heart failure. Am J Med . 1999 Jan. 106 (1):90-6. [QxMD MEDLINE Link] .

Neuberg GW, Miller AB, O'Connor CM, et al, for the PRAISE Investigators. Prospective Randomized Amlodipine Survival Evaluation. Diuretic resistance predicts mortality in patients with advanced heart failure. Am Heart J . 2002 Jul. 144 (1):31-8. [QxMD MEDLINE Link] .

Costanzo MR, Saltzberg M, O'Sullivan J, Sobotka P. Early ultrafiltration in patients with decompensated heart failure and diuretic resistance. J Am Coll Cardiol . 2005 Dec 6. 46 (11):2047-51. [QxMD MEDLINE Link] .

Young JB, Abraham WT, Stevenson LW, et al. Results of the VMAC Trial: vasodilation in the management of acute congestive heart failure. Circ . 2000 Nov 28. 102(22):a2794. [Full Text] .

Publication Committee for the VMAC Investigators (Vasodilatation in the Management of Acute CHF). Intravenous nesiritide vs nitroglycerin for treatment of decompensated congestive heart failure: a randomized controlled trial. JAMA . 2002 Mar 27. 287 (12):1531-40. [QxMD MEDLINE Link] .

Costanzo MR, Jessup M. Treatment of congestion in heart failure with diuretics and extracorporeal therapies: effects on symptoms, renal function, and prognosis. Heart Fail Rev . 2012 Mar. 17 (2):313-24. [QxMD MEDLINE Link] .

Costanzo MR, Guglin ME, Saltzberg MT, et al, for the UNLOAD Trial Investigators. Ultrafiltration versus intravenous diuretics for patients hospitalized for acute decompensated heart failure. J Am Coll Cardiol . 2007 Feb 13. 49 (6):675-83. [QxMD MEDLINE Link] .

CIBIS Investigators and Committees. A randomized trial of beta-blockade in heart failure. The Cardiac Insufficiency Bisoprolol Study (CIBIS). CIBIS Investigators and Committees. Circulation . 1994 Oct. 90 (4):1765-73. [QxMD MEDLINE Link] .

Chaudhry SI, Mattera JA, Curtis JP, et al. Telemonitoring in patients with heart failure. N Engl J Med . 2010 Dec 9. 363 (24):2301-9. [QxMD MEDLINE Link] .

Koehler F, Winkler S, Schieber M, et al, Telemedical Interventional Monitoring in Heart Failure Investigators. Impact of remote telemedical management on mortality and hospitalizations in ambulatory patients with chronic heart failure: the telemedical interventional monitoring in heart failure study. Circulation . 2011 May 3. 123 (17):1873-80. [QxMD MEDLINE Link] .

P&T Community. FDA approves first implantable wireless device with remote monitoring to measure pulmonary artery pressure in certain heart failure patients (news release). Available at https://www.ptcommunity.com/news/20170422/fda-approves-first-implantable-device-remote-monitoring-measure-pa-pressure-heart . May 28, 2014; Accessed: June 2, 2014.

O'Riordan M. FDA approves first implantable device for remotely monitoring HF patients. Heartwire. Available at http://www.medscape.com/viewarticle/825805 . May 28, 2014; Accessed: June 2, 2014.

Abraham WT, Adamson PB, Bourge RC, et al, for the CHAMPION Trial Study Group. Wireless pulmonary artery haemodynamic monitoring in chronic heart failure: a randomised controlled trial. Lancet . 2011 Feb 19. 377 (9766):658-66. [QxMD MEDLINE Link] .

Eisen HJ, Kobashigawa J, Keogh A, et al, for the Mycophenolate Mofetil Cardiac Study Investigators. Three-year results of a randomized, double-blind, controlled trial of mycophenolate mofetil versus azathioprine in cardiac transplant recipients. J Heart Lung Transplant . 2005 May. 24 (5):517-25. [QxMD MEDLINE Link] .

Yancy CW, Lopatin M, Stevenson LW, De Marco T, Fonarow GC, for the ADHERE Scientific Advisory Committee and Investigators. Clinical presentation, management, and in-hospital outcomes of patients admitted with acute decompensated heart failure with preserved systolic function: a report from the Acute Decompensated Heart Failure National Registry (ADHERE) Database. J Am Coll Cardiol . 2006 Jan 3. 47 (1):76-84. [QxMD MEDLINE Link] .

Cardiac Rhythm News. FDA approves expanded indication for certain pacemakers and defibrillators used to treat heart failure. Available at https://cardiacrhythmnews.com/fda-approves-expanded-indication-for-certain-pacemakers-and-defibrillators-used-to-treat-heart-failure/ . April 15, 2014; Accessed: April 15, 2014.

Stiles S. FDA approves Medtronic CRT sevices for mild HF with AV block. Medscape Medical News. Available at http://www.medscape.com/viewarticle/823485 . April 10, 2014; Accessed: April 15, 2014.

Levy WC, Lee KL, Hellkamp AS, et al. Maximizing survival benefit with primary prevention implantable cardioverter-defibrillator therapy in a heart failure population. Circulation . 2009 Sep 8. 120 (10):835-42. [QxMD MEDLINE Link] .

Santangelo G, Bursi F, Negroni MS, et al. Arrhythmic event prediction in patients with heart failure and reduced ejection fraction. J Cardiovasc Med (Hagerstown) . 2021 Feb 1. 22 (2):110-7. [QxMD MEDLINE Link] .

Ojo A, Tariq S, Harikrishnan P, Iwai S, Jacobson JT. Cardiac resynchronization therapy for heart failure. Interv Cardiol Clin . 2017 Jul. 6 (3):417-26. [QxMD MEDLINE Link] .

Rosanio S, Schwarz ER, Ahmad M, et al. Benefits, unresolved questions, and technical issues of cardiac resynchronization therapy for heart failure. Am J Cardiol . 2005 Sep 1. 96 (5):710-7. [QxMD MEDLINE Link] .

Tang AS, Wells GA, Talajic M, et al, for the Resynchronization-Defibrillation for Ambulatory Heart Failure Trial Investigators. Cardiac-resynchronization therapy for mild-to-moderate heart failure. N Engl J Med . 2010 Dec 16. 363 (25):2385-95. [QxMD MEDLINE Link] .

Giraldi F, Cattadori G, Roberto M, et al. Long-term effectiveness of cardiac resynchronization therapy in heart failure patients with unfavorable cardiac veins anatomy comparison of surgical versus hemodynamic procedure. J Am Coll Cardiol . 2011 Jul 26. 58 (5):483-90. [QxMD MEDLINE Link] .

Abraham WT, Fisher WG, Smith AL, et al, for the MIRACLE Study Group. Multicenter InSync Randomized Clinical Evaluation. Cardiac resynchronization in chronic heart failure. N Engl J Med . 2002 Jun 13. 346 (24):1845-53. [QxMD MEDLINE Link] .

Moss AJ, Hall WJ, Cannom DS, et al, for the MADIT-CRT Trial Investigators. Cardiac-resynchronization therapy for the prevention of heart-failure events. N Engl J Med . 2009 Oct 1. 361 (14):1329-38. [QxMD MEDLINE Link] .

Arshad A, Moss AJ, Foster E, et al, for the MADIT-CRT Executive Committee. Cardiac resynchronization therapy is more effective in women than in men: the MADIT-CRT (Multicenter Automatic Defibrillator Implantation Trial with Cardiac Resynchronization Therapy) trial. J Am Coll Cardiol . 2011 Feb 15. 57 (7):813-20. [QxMD MEDLINE Link] .

Ruwald MH, Ruwald AC, Jons C, et al. Effect of metoprolol versus carvedilol on outcomes in MADIT-CRT (multicenter automatic defibrillator implantation trial with cardiac resynchronization therapy). J Am Coll Cardiol . 2013 Apr 9. 61 (14):1518-26. [QxMD MEDLINE Link] .

Cleland JG, Daubert JC, Erdmann E, et al, for the Cardiac Resynchronization-Heart Failure (CARE-HF) Study Investigators. The effect of cardiac resynchronization on morbidity and mortality in heart failure. N Engl J Med . 2005 Apr 14. 352 (15):1539-49. [QxMD MEDLINE Link] .

Bristow MR, Saxon LA, Boehmer J, et al, for the Comparison of Medical Therapy, Pacing, et al. Cardiac-resynchronization therapy with or without an implantable defibrillator in advanced chronic heart failure. N Engl J Med . 2004 May 20. 350 (21):2140-50. [QxMD MEDLINE Link] .

Curtis AB, Worley SJ, Adamson PB, et al, for the Biventricular versus Right Ventricular Pacing in Heart Failure Patients with Atrioventricular Block (BLOCK HF) Trial Investigators. Biventricular pacing for atrioventricular block and systolic dysfunction. N Engl J Med . 2013 Apr 25. 368 (17):1585-93. [QxMD MEDLINE Link] .

Veterans Administration Coronary Artery Bypass Surgery Cooperative Study Group. Eleven-year survival in the Veterans Administration randomized trial of coronary bypass surgery for stable angina. N Engl J Med . 1984 Nov 22. 311 (21):1333-9. [QxMD MEDLINE Link] .

VA Coronary Artery Bypass Surgery Cooperative Study Group. Eighteen-year follow-up in the Veterans Affairs Cooperative Study of Coronary Artery Bypass Surgery for stable angina. Circulation . 1992 Jul. 86 (1):121-30. [QxMD MEDLINE Link] .

Caracciolo EA, Davis KB, Sopko G, et al. Comparison of surgical and medical group survival in patients with left main equivalent coronary artery disease. Long-term CASS experience. Circulation . 1995 May 1. 91 (9):2335-44. [QxMD MEDLINE Link] .

Robinson TN, Morrell TD, Pomerantz BJ, Heimbach JK, Cairns CB, Harken AH. Therapeutically accessible clinical cardiac states. J Am Coll Surg . 2000 Oct. 191 (4):452-63. [QxMD MEDLINE Link] .

Senior R, Lahiri A, Kaul S. Effect of revascularization on left ventricular remodeling in patients with heart failure from severe chronic ischemic left ventricular dysfunction. Am J Cardiol . 2001 Sep 15. 88 (6):624-9. [QxMD MEDLINE Link] .

Velazquez EJ, Lee KL, Deja MA, et al, for the STICH Investigators. Coronary-artery bypass surgery in patients with left ventricular dysfunction. N Engl J Med . 2011 Apr 28. 364 (17):1607-16. [QxMD MEDLINE Link] . [Full Text] .

Elefteriades JA, Morales DL, Gradel C, Tollis G Jr, Levi E, Zaret BL. Results of coronary artery bypass grafting by a single surgeon in patients with left ventricular ejection fractions < or = 30%. Am J Cardiol . 1997 Jun 15. 79 (12):1573-8. [QxMD MEDLINE Link] .

Kron IL, Flanagan TL, Blackbourne LH, Schroeder RA, Nolan SP. Coronary revascularization rather than cardiac transplantation for chronic ischemic cardiomyopathy. Ann Surg . 1989 Sep. 210 (3):348-52; discussion 352-4. [QxMD MEDLINE Link] . [Full Text] .

Doenst T, Velazquez EJ, Beyersdorf F, et al, for the STICH investigators. To STICH or not to STICH: we know the answer, but do we understand the question?. J Thorac Cardiovasc Surg . 2005 Feb. 129 (2):246-9. [QxMD MEDLINE Link] .

Joyce D, Loebe M, Noon GP, et al. Revascularization and ventricular restoration in patients with ischemic heart failure: the STICH trial. Curr Opin Cardiol . 2003 Nov. 18 (6):454-7. [QxMD MEDLINE Link] .

Sharoni E, Song HK, Peterson RJ, Guyton RA, Puskas JD. Off pump coronary artery bypass surgery for significant left ventricular dysfunction: safety, feasibility, and trends in methodology over time--an early experience. Heart . 2006 Apr. 92 (4):499-502. [QxMD MEDLINE Link] . [Full Text] .

Calafiore AM, Di Giammarco G, Teodori G, et al. Late results of first myocardial revascularization in multiple vessel disease: single versus bilateral internal mammary artery with or without saphenous vein grafts. Eur J Cardiothorac Surg . 2004 Sep. 26 (3):542-8. [QxMD MEDLINE Link] .

Nishimura RA, Grantham JA, Connolly HM, Schaff HV, Higano ST, Holmes DR Jr. Low-output, low-gradient aortic stenosis in patients with depressed left ventricular systolic function: the clinical utility of the dobutamine challenge in the catheterization laboratory. Circulation . 2002 Aug 13. 106 (7):809-13. [QxMD MEDLINE Link] .

Carabello BA. Clinical practice. Aortic stenosis. N Engl J Med . 2002 Feb 28. 346 (9):677-82. [QxMD MEDLINE Link] .

Lindblom D, Lindblom U, Qvist J, Lundstrom H. Long-term relative survival rates after heart valve replacement. J Am Coll Cardiol . 1990 Mar 1. 15(3):566-73. [QxMD MEDLINE Link] .

Connolly HM, Oh JK, Schaff HV, et al. Severe aortic stenosis with low transvalvular gradient and severe left ventricular dysfunction:result of aortic valve replacement in 52 patients. Circulation . 2000 Apr 25. 101 (16):1940-6. [QxMD MEDLINE Link] .

deFilippi CR, Willett DL, Brickner ME, et al. Usefulness of dobutamine echocardiography in distinguishing severe from nonsevere valvular aortic stenosis in patients with depressed left ventricular function and low transvalvular gradients. Am J Cardiol . 1995 Jan 15. 75 (2):191-4. [QxMD MEDLINE Link] .

Vaquette B, Corbineau H, Laurent M, et al. Valve replacement in patients with critical aortic stenosis and depressed left ventricular function: predictors of operative risk, left ventricular function recovery, and long term outcome. Heart . 2005 Oct. 91 (10):1324-9. [QxMD MEDLINE Link] . [Full Text] .

Dujardin KS, Enriquez-Sarano M, Schaff HV, Bailey KR, Seward JB, Tajik AJ. Mortality and morbidity of aortic regurgitation in clinical practice. A long-term follow-up study. Circulation . 1999 Apr 13. 99 (14):1851-7. [QxMD MEDLINE Link] .

Chaliki HP, Mohty D, Avierinos JF, et al. Outcomes after aortic valve replacement in patients with severe aortic regurgitation and markedly reduced left ventricular function. Circulation . 2002 Nov 19. 106 (21):2687-93. [QxMD MEDLINE Link] .

Enriquez-Sarano M, Tajik AJ. Clinical practice. Aortic regurgitation. N Engl J Med . 2004 Oct 7. 351 (15):1539-46. [QxMD MEDLINE Link] .

Lancellotti P, Gerard PL, Pierard LA. Long-term outcome of patients with heart failure and dynamic functional mitral regurgitation. Eur Heart J . 2005 Aug. 26 (15):1528-32. [QxMD MEDLINE Link] .

Patel JB, Borgeson DD, Barnes ME, Rihal CS, Daly RC, Redfield MM. Mitral regurgitation in patients with advanced systolic heart failure. J Card Fail . 2004 Aug. 10 (4):285-91. [QxMD MEDLINE Link] .

Bolling SF, Pagani FD, Deeb GM, Bach DS. Intermediate-term outcome of mitral reconstruction in cardiomyopathy. J Thorac Cardiovasc Surg . 1998 Feb. 115 (2):381-6; discussion 387-8. [QxMD MEDLINE Link] .

Akasaka T, Yoshida K, Hozumi T, et al. Restricted coronary flow reserve in patients with mitral regurgitation improves after mitral reconstructive surgery. J Am Coll Cardiol . 1998 Dec. 32 (7):1923-30. [QxMD MEDLINE Link] .

Wu AH, Aaronson KD, Bolling SF, Pagani FD, Welch K, Koelling TM. Impact of mitral valve annuloplasty on mortality risk in patients with mitral regurgitation and left ventricular systolic dysfunction. J Am Coll Cardiol . 2005 Feb 1. 45 (3):381-7. [QxMD MEDLINE Link] .

McGee EC, Gillinov AM, Blackstone EH, et al. Recurrent mitral regurgitation after annuloplasty for functional ischemic mitral regurgitation. J Thorac Cardiovasc Surg . 2004 Dec. 128 (6):916-24. [QxMD MEDLINE Link] .

Srichai MB, Grimm RA, Stillman AE, et al. Ischemic mitral regurgitation: impact of the left ventricle and mitral valve in patients with left ventricular systolic dysfunction. Ann Thorac Surg . 2005 Jul. 80 (1):170-8. [QxMD MEDLINE Link] .

Morishita A, Shimakura T, Nonoyama M, Takasaki T. Mitral valve replacement in ischemic mitral regurgitation. Preservation of both anterior and posterior mitral leaflets. J Cardiovasc Surg (Torino) . 2002 Apr. 43 (2):147-52. [QxMD MEDLINE Link] .

Yun KL, Sintek CF, Miller DC, et al. Randomized trial comparing partial versus complete chordal-sparing mitral valve replacement: effects on left ventricular volume and function. J Thorac Cardiovasc Surg . 2002 Apr. 123 (4):707-14. [QxMD MEDLINE Link] .

Gillinov AM, Wierup PN, Blackstone EH, et al. Is repair preferable to replacement for ischemic mitral regurgitation?. J Thorac Cardiovasc Surg . 2001 Dec. 122 (6):1125-41. [QxMD MEDLINE Link] .

Miller DC. Ischemic mitral regurgitation redux--to repair or to replace?. J Thorac Cardiovasc Surg . 2001 Dec. 122 (6):1059-62. [QxMD MEDLINE Link] .

Orban M, Hausleiter J. Edge-to-edge mitral valve repair: solid data and a prosperous future. Heart . 2018 Feb. 104(4):280-1. [QxMD MEDLINE Link] .

Vahanian A, Urena M, Ince H, Nickenig G. Mitral valve: repair/clips/cinching/chordae. EuroIntervention . 2017 Sep 24. 13 (AA):AA22-30. [QxMD MEDLINE Link] .

Feldman T, Kar S, Rinaldi M, et al, for the EVEREST Investigators. Percutaneous mitral repair with the MitraClip system: safety and midterm durability in the initial EVEREST (Endovascular Valve Edge-to-Edge REpair Study) cohort. J Am Coll Cardiol . 2009 Aug 18. 54 (8):686-94. [QxMD MEDLINE Link] .

Chiarito M, Pagnesi M, Martino EA, et al. Outcome after percutaneous edge-to-edge mitral repair for functional and degenerative mitral regurgitation: a systematic review and meta-analysis. Heart . 2018 Feb. 104(4):306-12. [QxMD MEDLINE Link] .

Geis NA, Puls M, Lubos E, et al. Safety and efficacy of MitraClip therapy in patients with severely impaired left ventricular ejection fraction: results from the German transcatheter mitral valve interventions (TRAMI) registry. Eur J Heart Fail . 2018 Mar. 20(3):598-608. [QxMD MEDLINE Link] .

Buckberg GD. Ventricular restoration--a surgical approach to reverse ventricular remodeling. Heart Fail Rev . 2004 Oct. 9 (4):233-9; discussion 347-51. [QxMD MEDLINE Link] .

Yamaguchi A, Ino T, Adachi H, et al. Left ventricular volume predicts postoperative course in patients with ischemic cardiomyopathy. Ann Thorac Surg . 1998 Feb. 65 (2):434-8. [QxMD MEDLINE Link] .

Mickleborough LL, Carson S, Ivanov J. Repair of dyskinetic or akinetic left ventricular aneurysm: results obtained with a modified linear closure. J Thorac Cardiovasc Surg . 2001 Apr. 121 (4):675-82. [QxMD MEDLINE Link] .

Ott DA, Parravacini R, Cooley DA, et al. Improved cardiac function following left ventricular aneurysm resection: pre- and postoperative performance studies in 150 patients. Tex Heart Inst J . 1982 Sep. 9 (3):267-73. [QxMD MEDLINE Link] . [Full Text] .

Athanasuleas CL, Buckberg GD, Stanley AW, et al, for the RESTORE group. Surgical ventricular restoration in the treatment of congestive heart failure due to post-infarction ventricular dilation. J Am Coll Cardiol . 2004 Oct 6. 44 (7):1439-45. [QxMD MEDLINE Link] .

Jones RH, Velazquez EJ, Michler RE, et al, for the STICH Hypothesis 2 Investigators. Coronary bypass surgery with or without surgical ventricular reconstruction. N Engl J Med . 2009 Apr 23. 360 (17):1705-17. [QxMD MEDLINE Link] . [Full Text] .

Imamura T, Chung B, Nguyen A, Sayer G, Uriel N. Clinical implications of hemodynamic assessment during left ventricular assist device therapy. J Cardiol . 2018 Apr. 71 (4):352-8. [QxMD MEDLINE Link] .

Timms D. A review of clinical ventricular assist devices. Med Eng Phys . 2011 Nov. 33 (9):1041-7. [QxMD MEDLINE Link] .

Pamboukian SV, Tallaj JA, Brown RN, et al. Improvement in 2-year survival for ventricular assist device patients after implementation of an intensive surveillance protocol. J Heart Lung Transplant . 2011 Aug. 30 (8):879-87. [QxMD MEDLINE Link] .

Lietz K, Long JW, Kfoury AG, et al. Outcomes of left ventricular assist device implantation as destination therapy in the post-REMATCH era: implications for patient selection. Circulation . 2007 Jul 31. 116 (5):497-505. [QxMD MEDLINE Link] .

Daneshmand MA, Rajagopal K, Lima B, et al. Left ventricular assist device destination therapy versus extended criteria cardiac transplant. Ann Thorac Surg . 2010 Apr. 89 (4):1205-9; discussion 1210. [QxMD MEDLINE Link] .

US Food and Drug Administration. Recently-approved devices: HeartWare HVAD - P100047/S090. Available at https://www.fda.gov/MedicalDevices/ProductsandMedicalProcedures/DeviceApprovalsandClearances/Recently-ApprovedDevices/ucm581473.htm . September 27, 2017 [Updated October 31, 2017]; Accessed: October 31, 2017.

US National Library of Medicine. Evaluation of the Jarvik 2000 Left Ventricular Assist System with post-auricular connector--destination therapy study. ClinicalTrials.gov. Available at https://clinicaltrials.gov/ct2/show/NCT01627821 . Updated: October 16, 2017; Accessed: October 31, 2017.

Rose EA, Gelijns AC, Moskowitz AJ, et al, for theRandomized Evaluation of Mechanical Assistance for the Treatment of Congestive Heart Failure (REMATCH) Study Group. Long-term use of a left ventricular assist device for end-stage heart failure. N Engl J Med . 2001 Nov 15. 345 (20):1435-43. [QxMD MEDLINE Link] .

Park SJ, Tector A, Piccioni W, et al. Left ventricular assist devices as destination therapy: a new look at survival. J Thorac Cardiovasc Surg . 2005 Jan. 129 (1):9-17. [QxMD MEDLINE Link] .

Starling RC, Naka Y, Boyle AJ, et al. Results of the post-U.S. Food and Drug Administration-approval study with a continuous flow left ventricular assist device as a bridge to heart transplantation: a prospective study using the INTERMACS (Interagency Registry for Mechanically Assisted Circulatory Support). J Am Coll Cardiol . 2011 May 10. 57 (19):1890-8. [QxMD MEDLINE Link] .

Ventura PA, Alharethi R, Budge D, et al. Differential impact on post-transplant outcomes between pulsatile- and continuous-flow left ventricular assist devices. Clin Transplant . 2011 Jul-Aug. 25 (4):E390-5. [QxMD MEDLINE Link] .

Slaughter MS, Pagani FD, Rogers JG, et al, for the HeartMate II Clinical Investigators. Clinical management of continuous-flow left ventricular assist devices in advanced heart failure. J Heart Lung Transplant . 2010 Apr. 29 (4 suppl):S1-39. [QxMD MEDLINE Link] .

Kirklin JK, Naftel DC, Kormos RL, et al. Third INTERMACS Annual Report: the evolution of destination therapy in the United States. J Heart Lung Transplant . 2011 Feb. 30 (2):115-23. [QxMD MEDLINE Link] .

Taylor DO, Edwards LB, Boucek MM, et al. Registry of the International Society for Heart and Lung Transplantation: twenty-second official adult heart transplant report--2005. J Heart Lung Transplant . 2005 Aug. 24 (8):945-55. [QxMD MEDLINE Link] .

Boucek MM, Edwards LB, Keck BM, Trulock EP, Taylor DO, Hertz MI. Registry of the International Society for Heart and Lung Transplantation: eighth official pediatric report--2005. J Heart Lung Transplant . 2005 Aug. 24(8):968-82. [QxMD MEDLINE Link] .

United Network for Organ Sharing (US). National data. UNOS. Available at https://unos.org/data/ . October 24, 2017; Accessed: November 1, 2017.

Stevenson LW, Kormos RL, Bourge RC, et al. Mechanical cardiac support 2000: current applications and future trial design. June 15-16, 2000 Bethesda, Maryland. J Am Coll Cardiol . 2001 Jan. 37 (1):340-70. [QxMD MEDLINE Link] .

Gray NA Jr, Selzman CH. Current status of the total artificial heart. Am Heart J . 2006 Jul. 152 (1):4-10. [QxMD MEDLINE Link] .

Cooley DA, Liotta D, Hallman GL, Bloodwell RD, Leachman RD, Milam JD. Orthotopic cardiac prosthesis for two-staged cardiac replacement. Am J Cardiol . 1969 Nov. 24 (5):723-30. [QxMD MEDLINE Link] .

Platis A, Larson DF. CardioWest temporary total artificial heart. Perfusion . 2009 Sep. 24 (5):341-6. [QxMD MEDLINE Link] .

Roussel JC, Senage T, Baron O, et al. CardioWest (Jarvik) total artificial heart: a single-center experience with 42 patients. Ann Thorac Surg . 2009 Jan. 87 (1):124-9; discussion 130. [QxMD MEDLINE Link] .

Anderson E, Jaroszewski D, Pierce C, DeValeria P, Arabia F. Parallel application of extracorporeal membrane oxygenation and the CardioWest total artificial heart as a bridge to transplant. Ann Thorac Surg . 2009 Nov. 88 (5):1676-8. [QxMD MEDLINE Link] .

Morris RJ. Total artificial heart--concepts and clinical use. Semin Thorac Cardiovasc Surg . 2008 Fall. 20 (3):247-54. [QxMD MEDLINE Link] .

Meyer A, Slaughter M. The total artificial heart. Panminerva Med . 2011 Sep. 53 (3):141-54. [QxMD MEDLINE Link] .

US National Library of Medicine. European clinical evaluation of the Carmat total artificial heart (advance HF). ClinicalTrials.gov. Available at https://clinicaltrials.gov/ct2/show/NCT02962973 . Updated: November 25, 2016; Accessed: November 1, 2017.

[Guideline] Ackerman MJ, Priori SG, Willems S, et al, Heart Rhythm Society (HRS), European Heart Rhythm Association (EHRA). HRS/EHRA expert consensus statement on the state of genetic testing for the channelopathies and cardiomyopathies: this document was developed as a partnership between the Heart Rhythm Society (HRS) and the European Heart Rhythm Association (EHRA). Europace . 2011 Aug. 13 (8):1077-109. [QxMD MEDLINE Link] . [Full Text] .

Marcus FI, McKenna WJ, Sherrill D, et al. Diagnosis of arrhythmogenic right ventricular cardiomyopathy/dysplasia: proposed modification of the Task Force Criteria. Eur Heart J . 2010 Apr. 31 (7):806-14. [QxMD MEDLINE Link] . [Full Text] .

[Guideline] O'Meara E, McDonald M, Chan M, et al. CCS/CHFS heart failure guidelines: clinical trial update on functional mitral regurgitation, SGLT2 inhibitors, ARNI in HFpEF, and tafamidis in amyloidosis. Can J Cardiol . 2020 Feb. 36 (2):159-69. [QxMD MEDLINE Link] . [Full Text] .

[Guideline] Tracy CM, Epstein AE, Darbar D,et al. 2012 ACCF/AHA/HRS focused update of the 2008 guidelines for device-based therapy of cardiac rhythm abnormalities: a report of the American College of Cardiology Foundation/American Heart Association Task Force on Practice Guidelines and the Heart Rhythm Society. [corrected]. Circulation . 2012 Oct 2. 126 (14):1784-800. [QxMD MEDLINE Link] . [Full Text] .

[Guideline] Priori SG, Blomstrom-Lundqvist C, Mazzanti A, et al. 2015 ESC guidelines for the management of patients with ventricular arrhythmias and the prevention of sudden cardiac death: The Task Force for the Management of Patients with Ventricular Arrhythmias and the Prevention of Sudden Cardiac Death of the European Society of Cardiology (ESC). Endorsed by: Association for European Paediatric and Congenital Cardiology (AEPC). Eur Heart J . 2015 Nov 1. 36 (41):2793-867. [QxMD MEDLINE Link] . [Full Text] .

[Guideline] Rihal CS, Naidu SS, Givertz MM, et al, Society for Cardiovascular Angiography and Interventions (SCAI), Heart Failure Society of America (HFSA), et al. 2015 SCAI/ACC/HFSA/STS clinical expert consensus statement on the use of percutaneous mechanical circulatory support devices in cardiovascular care: endorsed by the American Heart Association, the Cardiological Society of India, & Sociedad Latino Americana de Cardiologia Intervencion; affirmation of value by the Canadian Association of Interventional Cardiology-Association Canadienne de Cardiologie d'intervention. J Am Coll Cardiol . 2015 May 19. 65 (19):e7-e26. [QxMD MEDLINE Link] . [Full Text] .

[Guideline] Feldman D, Pamboukian SV, Teuteberg JJ, et al., International Society for Heart and Lung Transplantation. The 2013 International Society for Heart and Lung Transplantation Guidelines for mechanical circulatory support: executive summary. J Heart Lung Transplant . 2013 Feb. 32 (2):157-87. [QxMD MEDLINE Link] . [Full Text] .

[Guideline] Peura JL, Colvin-Adams M, Francis GS, et al. Recommendations for the use of mechanical circulatory support: device strategies and patient selection: a scientific statement from the American Heart Association. Circulation . 2012 Nov 27. 126 (22):2648-67. [QxMD MEDLINE Link] . [Full Text] .

[Guideline] Baran DA, Billia F, Randhawa V, et al, for the Consensus Conference participants/collaborators. Consensus statements from the International Society for Heart and Lung Transplantation consensus conference: heart failure-related cardiogenic shock. J Heart Lung Transplant . 2024 Feb. 43 (2):204-16. [QxMD MEDLINE Link] . [Full Text] .

[Guideline] Mullens W, Damman K, Dhont S, et al. Dietary sodium and fluid intake in heart failure. A clinical consensus statement of the Heart Failure Association of the ESC. Eur J Heart Fail . 2024 Apr 12. [QxMD MEDLINE Link] . [Full Text] .

Masini M, Elia E, Vianello PF, et al. Frequency, predictors and prognostic impact of implantable cardioverter defibrillator shocks in a primary prevention population with heart failure and reduced ejection fraction. J Cardiovasc Med (Hagerstown) . 2021 Feb 1. 22 (2):118-25. [QxMD MEDLINE Link] .

[Guideline] Lloyd-Jones D, Adams RJ, Brown TM, et al, for the American Heart Association Statistics Committee and Stroke Statistics Subcommittee. Executive summary: heart disease and stroke statistics--2010 update: a report from the American Heart Association. Circulation . 2010 Feb 23. 121 (7):948-54. [QxMD MEDLINE Link] .

  • Heart Failure. This chest radiograph shows an enlarged cardiac silhouette and edema at the lung bases, signs of acute heart failure.
  • Heart Failure. Cardiac cirrhosis. Congestive hepatopathy with large renal vein.
  • Heart Failure. Cardiac cirrhosis. Congestive hepatopathy with large inferior vena cava.
  • Heart Failure. This electrocardiogram (ECG) is from a 32-year-old female with recent-onset congestive heart failure and syncope. The ECG demonstrates a tachycardia with a 1:1 atrial:ventricular relationship. It is not clear from this tracing whether the atria are driving the ventricles (sinus tachycardia) or the ventricles are driving the atria (ventricular tachycardia [VT]). At first glance, sinus tachycardia in this ECG might be considered with severe conduction disease manifesting as marked first-degree atrioventricular block with left bundle branch block. On closer examination, the ECG morphology gives clues to the actual diagnosis of VT. These clues include the absence of RS complexes in the precordial leads, a QS pattern in V6, and an R wave in aVR. The patient proved to have an incessant VT associated with dilated cardiomyopathy.
  • Heart Failure. This is a posteroanterior view of a right ventricular endocardial activation map during ventricular tachycardia in a patient with a previous septal myocardial infarction. The earliest activation is recorded in red; late activation displays as blue to magenta. Fragmented low-amplitude diastolic local electrocardiograms were recorded adjacent to the earliest (red) breakout area, and local ablation in this scarred zone (red dots) resulted in termination and noninducibility of this previously incessant arrhythmia.
  • Heart Failure. A 28-year-old woman presented with acute heart failure secondary to chronic hypertension. The enlarged cardiac silhouette on this anteroposterior (AP) radiograph is caused by acute heart failure due to the effects of chronic high blood pressure on the left ventricle. The heart then becomes enlarged, and fluid accumulates in the lungs (ie, pulmonary congestion).
  • Heart Failure. Epsilon wave on an electrocardiogram in a patient with arrhythmogenic right ventricular dysplasia (ARVD). ARVD is a congenital cardiomyopathy that is characterized by infiltration of adipose and fibrous tissue into the RV wall and loss of myocardial cells. Primary injuries usually are at the free wall of the RV and right atria, resulting in ventricular and supraventricular arrhythmias. The most significant of all rhythms associated with heart failure are the life-threatening ventricular arrhythmias.
  • Heart Failure. Electrocardiogram depicting ventricular fibrillation in a patient with a left ventricular assist device (LVAD). Ventricular fibrillation is often due to ischemic heart disease and can lead to myocardial infarction and/or sudden death.
  • Heart Failure. The rhythm on this electrocardiogram (ECG) is sinus with borderline PR prolongation. There is evidence of an acute/evolving anterior ischemia/myocardial infarction (MI) superimposed on the left bundle branch block (LBBB)–like pattern. Note the primary T-wave inversions in leads V2-V4, rather than the expected discordant (upright) T waves in the leads with a negative QRS. Although this finding is not particularly sensitive for ischemia/MI with LBBB, such primary T-wave changes are relatively specific. The prominent voltage with left atrial abnormality and leftward axis in concert with the left ventricular intraventricular conduction delay (IVCD) are consistent with underlying left ventricular hypertrophy. This ECG is an example of "bundle branch block plus." Image courtesy of http://ecg.bidmc.harvard.edu.
  • Heart Failure. This electrocardiogram (ECG) shows evidence of severe left ventricular hypertrophy (LVH) with prominent precordial voltage, left atrial abnormality, lateral ST-T abnormalities, and a somewhat leftward QRS axis (–15º). The patient had malignant hypertension with acute heart failure, accounting also for the sinus tachycardia (blood pressure initially 280/180 mmHg). The ST-T changes seen here are nonspecific and could be due to, for example, LVH alone or coronary artery disease. However, the ECG is not consistent with extensive inferolateral myocardial infarction. Image courtesy of http://ecg.bidmc.harvard.edu.
  • Heart Failure. The rhythm on this electrocardiogram is atrial tachycardia (rate, 154 beats/min) with a 2:1 atrioventricular (AV) block. Note the partially hidden, nonconducted P waves on the ST segments (eg, leads I and aVL). The QRS is very wide with an atypical intraventricular conduction defect (IVCD) pattern. The rSR' type complex in the lateral leads (I, aVL) is not due to a right bundle branch block (RBBB) but to an atypical left ventricular conduction defect. These unexpected rSR' complexes in the lateral leads (El-Sherif sign) correlate with underlying extensive myocardial infarction (MI) and, occasionally, ventricular aneurysm. (El-Sherif. Br Heart J. 1970;32:440-8.) The notching on the upstroke of the S waves in lead V4 with a left bundle branch block-type pattern also suggests underlying MI (Cabrera sign). This patient had severe cardiomyopathy secondary to coronary artery disease, with extensive left ventricular wall motion abnormalities. Image courtesy of http://ecg.bidmc.harvard.edu.
  • Heart Failure. On this electrocardiogram, baseline artifact is present, simulating atrial fibrillation. Such artifact may be caused by a variety of factors, including poor electrode contact, muscle tremor, and electrical interference. A single premature ventricular complex (PVC) is present with a compensatory pause such that the RR interval surrounding the PVC is twice as long as the preceding sinus RR interval. Evidence of a previous anterior myocardial infarction is present with pathologic Q waves in leads V1-V3. Borderline-low precordial voltage is a nonspecific finding. Cardiac catheterization showed a 90% stenosis in the patient's proximal portion the left anterior descending coronary artery, which was treated with angioplasty and stenting. Broad P waves in lead V1 with a prominent negative component is consistent with a left atrial abnormality. Image courtesy of http://ecg.bidmc.harvard.edu.
  • Heart Failure. This electrocardiogram (ECG) is from a patient who underwent urgent cardiac catheterization, which revealed diffuse severe coronary spasm (most marked in the left circumflex system) without any fixed obstructive lesions. Severe left ventricular wall motion abnormalities were present, involving the anterior and inferior segments. A question of so-called takotsubo cardiomyopathy (left ventricular apical ballooning syndrome) is also raised (see Bybee et al. Systematic review: transient left ventricular apical ballooning: a syndrome that mimics ST-segment elevation myocardial infarction. Ann Int Med 2004:141:858-65). The latter is most often reported in postmenopausal, middle-aged to elderly women in the context of acute emotional stress and may cause ST elevations acutely with subsequent T-wave inversions. A cocaine-induced cardiomyopathy (possibly related to coronary vasospasm) is a consideration but was excluded here. Myocarditis may also be associated with this type of ECG and the cardiomyopathic findings shown here. No fixed obstructive epicardial coronary lesions were detected by coronary arteriography. The findings in this ECG include low-amplitude QRS complexes in the limb leads (with an indeterminate QRS axis), loss of normal precordial R-wave progression (leads V1-V3), and prominent anterior/lateral T-wave inversions. Image courtesy of http://ecg.bidmc.harvard.edu.
  • Heart Failure. This electrocardiogram shows an extensive acute/evolving anterolateral myocardial infarction pattern, with ST-T changes most apparent in leads V2-V6, I, and aVL. Slow R-wave progression is also present in leads V1-V3. The rhythm is borderline sinus tachycardia with a single premature atrial complex (PAC) (fourth beat). Note also the low limb-lead voltage and probable left atrial abnormality. Left ventriculography showed diffuse hypokinesis as well as akinesis of the anterolateral and apical walls, with an ejection fraction estimated at 33%. Image courtesy of http://ecg.bidmc.harvard.edu.
  • Heart Failure. This electrocardiogram shows a patient is having an evolving anteroseptal myocardial infarction secondary to cocaine. There are Q waves in leads V2-V3 with ST-segment elevation in leads V2-V5 associated with T-wave inversion. Also noted are biphasic T waves in the inferior leads. These multiple abnormalities suggest occlusion of a large left anterior descending artery that wraps around the apex of the heart (or multivessel coronary artery disease). Image courtesy of http://ecg.bidmc.harvard.edu.
  • Heart Failure. A color-enhanced angiogram of the left heart shows a plaque-induced obstruction (top center) in a major artery, which can lead to myocardial infarction (MI). MIs can precipitate heart failure.
  • Heart Failure. Emphysema is included in the differential diagnosis of heart failure. In this radiograph, emphysema bubbles are noted in the left lung; these can severely impede breathing capacity.
  • Heart Failure. Cervicocephalic fibromuscular dysplasia (FMD) can lead to complications such as hypertension and chronic kidney failure, which can lead to heart failure. In this color Doppler and spectral Doppler ultrasonographic examination of the left internal carotid artery (ICA) in a patient with cervicocephalic FMD, stenoses of about 70% is seen in the ICA.
  • Heart Failure. Cervicocephalic fibromuscular dysplasia (FMD) can lead to complications such as hypertension and chronic kidney failure, which, in turn, can lead to heart failure. Nodularity in an artery is known as the "string-of-beads sign," and it can be seen this color Doppler ultrasonographic image from a 51-year-old patient with low-grade stenosing FMD of the internal carotid artery (ICA).
  • Heart Failure. Electrocardiogram from a 46-year-old man with long-standing hypertension. Note the left atrial abnormality and left ventricular hypertrophy with strain.
  • Heart Failure. Electrocardiogram from a 46-year-old man with long-standing hypertension. Left atrial abnormality and left ventricular hypertrophy with strain is revealed.
  • Heart Failure. Apical four-chamber echocardiogram in a 37-year-old man with arrhythmogenic right ventricular dysplasia (ARVD), a congenital cardiomyopathy. Note the prominent trabeculae and abnormal wall motion of the dilated RV. ARVD can result in ventricular and supraventricular arrhythmias. The most significant of all rhythms associated with heart failure are the life-threatening ventricular arrhythmias.
  • Heart Failure. Cardiac magnetic resonance image (CMRI), short-axis view. This image shows right ventricular (RV) dilatation, trabucular derangement, aneurysm formation, and dyskinetic free wall in a patient with arrhythmogenic RV dysplasia.
  • Heart Failure. This transthoracic echocardiogram demonstrates severe mitral regurgitation with a heavily calcified mitral valve and prolapse of the posterior leaflet into the left atrium.
  • Heart Failure. Echocardiogram of a patient with severe pulmonic stenosis. This image shows a parasternal short-axis view of a thickened pulmonary valve. Pulmonic stenosis can lead to pulmonary hypertension, which can result in hepatic congestion and in right-sided heart failure.
  • Heart Failure. Echocardiogram of a patient with severe pulmonic stenosis. This image shows a Doppler scan of the peak velocity (5.2 m/s) and gradients (peak 109 mmHg, mean 65 mmHg) across the valve.
  • Heart Failure. Echocardiogram of a patient with severe pulmonic stenosis. This image shows moderately severe pulmonary insufficiency (orange color flow) is also present.
  • Heart Failure. This video is an echocardiogram of a patient with severe pulmonic stenosis. The first segment shows the parasternal short-axis view of the thickened pulmonary valve. The second segment shows the presence of moderate pulmonary insufficiency (orange color flow). AV = aortic valve, PA = pulmonary artery, PI = pulmonary insufficiency, PV = pulmonary valve.
  • Heart Failure. Transesophageal echocardiogram with continuous wave Doppler interrogation across the mitral valve. An increased mean gradient of 16 mmHg is revealed, consistent with severe mitral stenosis.
  • Table 1. Framingham Diagnostic Criteria for Heart Failure
  • Table 2. Evidence-Based BNP and NT-proBNP Cutoff Values for Diagnosing HF
  • Table 3. 2013 American College of Cardiology Foundation/American Heart Association (ACCF/AHA) Heart Failure Staging System
  • Table 4. 2022 ACC/AHA/Heart Failure Society of America (HFSA) Heart Failure Staging System
  • Table 5. 2022 ACC/AHA/HFSA Classification of Heart Failure (HF) by Left Ventricular Ejection Fraction (LVEF)

Paroxysmal nocturnal dyspnea

Nocturnal cough

Weight loss of 4.5 kg in 5 days in response to treatment

Dyspnea on ordinary exertion

Neck vein distention

A decrease in vital capacity by one third the maximal value recorded

Rales

Pleural effusion

Acute pulmonary edema

Tachycardia (rate of 120 bpm)

Hepatojugular reflux

Hepatomegaly

S gallop

Bilateral ankle edema

Central venous pressure >16 cm water

 

Circulation time of ≥25 seconds

 

Radiographic cardiomegaly

 

Pulmonary edema, visceral congestion, or cardiomegaly at autopsy

 

Ho KK, Pinsky JL, Kannel WB, Levy D. The epidemiology of heart failure: the Framingham Study.

Age, y

>17

< 100 (0.13)

>500 (8.1)

-

-

>21

-

-

< 300 (0.02)

-

21-50

-

-

-

>450 (14)

50-75

-

-

-

>900 (5.0)

>75

-

-

-

>1800 (3.1)

Estimated GFR, < 60 mL/min

< 200 (0.13)

>500 (9.3)

-

-

BNP = B-type natriuretic peptide; GRF = glomerular filtration rate; HF = heart failure; LR = likelihood ratio; NPV = negative predictive value; NT-pro-BNP = N-terminal proBNP; PPV = positive predictive value; – = not specifically defined.

Derived from Breathing Not Properly data (1586 emergency department [ED] patients, prevalence of HF = 47%). ]

Derived from PRIDE data (1256 ED patients, prevalence of HF = 57%). , ]

Derived from subset of Breathing Not Properly data (452 ED patients, prevalence of HF = 49%). ]

At high risk for heart failure but without structural heart disease or symptoms of heart failure

Patients with coronary artery disease, hypertension, or diabetes mellitus without impaired left ventricular (LV) function, LV hypertrophy (LVH), or geometric chamber distortion

Patients with predisposing risk factors for developing heart failure

No corresponding New York Heart Association (NYHA) functional classification

Structural heart disease but without signs/symptoms of heart failure

Patients who are asymptomatic but who have LVH and/or impaired LV function

Corresponds with patients with NYHA class I

Structural heart disease with current or past symptoms of heart failure

Patients with known structural heart disease and shortness of breath and fatigue, as well as reduced exercise tolerance

The majority of patients with heart failure are in this stage

Corresponds with NYHA classes I, II, III and IV

Refractory heart failure requiring specialized interventions

Patients who have marked symptoms at rest despite maximal medical therapy

Patients in this stage may be eligible to receive mechanical circulatory support, receive continuous inotropic infusions, undergo procedures to facilitate fluid removal, or undergo heart transplantation or other procedures

Corresponds with patients with NYHA class IV

At risk of HF; asymptomatic, no structural heart disease nor cardiac biomarkers of stretch injury (eg, patients with hypertension, atherosclerotic cardiovascular disease, diabetes, metabolic syndrome and obesity, exposure to cardiotoxic agents, genetic variant for cardiomyopathy, or positive family history of cardiomyopathy)

No signs/symptoms of HF and evidence of one of the following:

Structural heart disease

Evidence for raised filling pressures by invasive hemodynamic measurements or by noninvasive imaging that suggests elevated filling pressures (eg, Doppler echocardiography)

Patients with risk factors and raised levels of B-type natriuretic peptides or persistently elevated cardiac troponin in the absence of competing diagnoses that result in such biomarker elevations (eg, acute coronary syndrome, chronic kidney disease, pulmonary embolus, or myopericarditis)

Structural heart disease with current or previous symptoms of HF

Marked HF symptoms that interfere with daily life and with repeated hospitalizations despite attempts to optimize guideline-directed medical therapy

HF = heart failure

LVEF ≤40%

Previous LVEF ≤40% and a followup LVEF >40%

LVEF of 41%-49%

Evidence of spontaneous/provokable increased LV filling pressures (eg, elevated natriuretic peptide, noninvasive and invasive hemodynamic measurement)

LVEF ≥50%

Evidence of spontaneous/provokable increased LV filling pressures (eg, elevated natriuretic peptide, noninvasive and invasive hemodynamic measurement)

Contributor Information and Disclosures

Ioana Dumitru, MD Associate Professor of Medicine, Division of Cardiology, Founder and Medical Director, Heart Failure and Cardiac Transplant Program, University of Nebraska Medical Center; Associate Professor of Medicine, Division of Cardiology, Veterans Affairs Medical Center Ioana Dumitru, MD is a member of the following medical societies: American College of Cardiology , Heart Failure Society of America , International Society for Heart and Lung Transplantation Disclosure: Nothing to disclose.

Mary L Windle, PharmD Adjunct Associate Professor, University of Nebraska Medical Center College of Pharmacy; Editor-in-Chief, Medscape Drug Reference Disclosure: Nothing to disclose.

Yasmine S Ali, MD, MSCI, FACC, FACP Assistant Clinical Professor of Medicine, Vanderbilt University School of Medicine; President, LastSky Writing, LLC Yasmine S Ali, MD, MSCI, FACC, FACP is a member of the following medical societies: American College of Cardiology , American College of Physicians , American Heart Association , American Medical Association , American Medical Writers Association , National Lipid Association , Tennessee Medical Association Disclosure: Serve(d) as a director, officer, partner, employee, advisor, consultant or trustee for: LastSky Writing;Philips Healthcare;M Health;Verywell Health; MedStudy;WellnessVerge;Prose Media; AKH, Inc.; LearnRoll, LLC; Eradigm; RxCe.com; M3 USA; M3 EU; Future US Inc.; Edanz Group; ChesterPA511<br/>Serve(d) as a speaker or a member of a speakers bureau for: RxCe.com.

Gyanendra K Sharma, MD, FACC, FASE Professor of Medicine and Radiology, Director, Adult Echocardiography Laboratory, Section of Cardiology, Medical College of Georgia at Augusta University Gyanendra K Sharma, MD, FACC, FASE is a member of the following medical societies: American Association of Cardiologists of Indian Origin, American Association of Physicians of Indian Origin , American College of Cardiology , American Society of Echocardiography , Society for Cardiovascular Magnetic Resonance , Society of Cardiovascular Computed Tomography Disclosure: Nothing to disclose.

Mathue M Baker, MD Cardiologist, BryanLGH Heart Institute and Saint Elizabeth Regional Medical Center Mathue M Baker, MD is a member of the following medical societies: American College of Cardiology Disclosure: Nothing to disclose.

Henry H Ooi, MD, MRCPI Director, Advanced Heart Failure and Cardiac Transplant Program, Nashville Veterans Affairs Medical Center; Assistant Professor of Medicine, Vanderbilt University School of Medicine Disclosure: Nothing to disclose.

Mariclaire Cloutier Freelance editor, Medscape Drugs & Diseases Disclosure: Nothing to disclose.

Barry E Brenner, MD, PhD, FACEP Professor of Emergency Medicine, Professor of Internal Medicine, Program Director, Emergency Medicine, Case Medical Center, University Hospitals, Case Western Reserve University School of Medicine

Barry E Brenner, MD, PhD, FACEP is a member of the following medical societies: Alpha Omega Alpha , American Academy of Emergency Medicine , American College of Chest Physicians , American College of Emergency Physicians , American College of Physicians , American Heart Association , American Thoracic Society , Arkansas Medical Society , New York Academy of Medicine , New York AcademyofSciences ,and Society for Academic Emergency Medicine

Disclosure: Nothing to disclose.

David FM Brown, MD Associate Professor, Division of Emergency Medicine, Harvard Medical School; Vice Chair, Department of Emergency Medicine, Massachusetts General Hospital

David FM Brown, MD is a member of the following medical societies: American College of Emergency Physicians and Society for Academic Emergency Medicine

William K Chiang, MD Associate Professor, Department of Emergency Medicine, New York University School of Medicine; Chief of Service, Department of Emergency Medicine, Bellevue Hospital Center

William K Chiang, MD is a member of the following medical societies: American Academy of Clinical Toxicology , American College of Medical Toxicology , and Society for Academic Emergency Medicine

Joseph Cornelius Cleveland Jr, MD Associate Professor, Division of Cardiothoracic Surgery, University of Colorado Health Sciences Center

Joseph Cornelius Cleveland Jr, MD is a member of the following medical societies: Alpha Omega Alpha , American Association for the Advancement of Science , American College of Cardiology , American College of Chest Physicians , American College of Surgeons , American Geriatrics Society , American Physiological Society , American Society of Transplant Surgeons , Association for Academic Surgery , Heart Failure Society of America , International Society for Heart and Lung Transplantation , Phi Beta Kappa , Society of Critical Care Medicine , Society of Thoracic Surgeons , and Western Thoracic Surgical Association

Disclosure: Thoratec Heartmate II Pivotal Tria; Grant/research funds Principal Investigator - Colorado; Abbott Vascular E-Valve E-clip Honoraria Consulting; Baxter Healthcare Corp Consulting fee Board membership; Heartware Advance BTT Trial Grant/research funds Principal Investigator- Colorado; Heartware Endurance DT trial Grant/research funds Principal Investigator-Colorado

Shamai Grossman, MD, MS Assistant Professor, Department of Emergency Medicine, Harvard Medical School; Director, The Clinical Decision Unit and Cardiac Emergency Center, Beth Israel Deaconess Medical Center

Shamai Grossman, MD, MS is a member of the following medical societies: American College of Emergency Physicians

John D Newell Jr, MD Professor of Radiology, Head, Division of Radiology, National Jewish Health; Professor, Department of Radiology, University of Colorado School of Medicine

John D Newell Jr, MD is a member of the following medical societies: American College of Chest Physicians , American College of Radiology , American Roentgen Ray Society , American Thoracic Society , Association of University Radiologists , Radiological Society of North America , and Society of Thoracic Radiology

Disclosure: Siemens Medical Grant/research funds Consulting; Vida Corporation Ownership interest Board membership; TeraRecon Grant/research funds Consulting; Medscape Reference Honoraria Consulting; Humana Press Honoraria Other

Craig H Selzman, MD, FACS Associate Professor of Surgery, Surgical Director, Cardiac Mechanical Support and Heart Transplant, Division of Cardiothoracic Surgery, University of Utah School of Medicine

Craig H Selzman, MD, FACS is a member of the following medical societies: Alpha Omega Alpha , American Association for Thoracic Surgery , American College of Surgeons , American Physiological Society , Association for Academic Surgery , International Society for Heart and Lung Transplantation , Society of Thoracic Surgeons , Southern Thoracic Surgical Association , and Western Thoracic Surgical Association

Gary Setnik, MD Chair, Department of Emergency Medicine, Mount Auburn Hospital; Assistant Professor, Division of Emergency Medicine, Harvard Medical School

Gary Setnik, MD is a member of the following medical societies: American College of Emergency Physicians , National Association of EMS Physicians , and Society for Academic Emergency Medicine

Disclosure: SironaHealth Salary Management position; South Middlesex EMS Consortium Salary Management position; ProceduresConsult.com Royalty Other

Brett C Sheridan, MD, FACS Associate Professor of Surgery, University of North Carolina at Chapel Hill School of Medicine

George A Stouffer III, MD Henry A Foscue Distinguished Professor of Medicine and Cardiology, Director of Interventional Cardiology, Cardiac Catheterization Laboratory, Chief of Clinical Cardiology, Division of Cardiology, University of North Carolina Medical Center

George A Stouffer III, MD is a member of the following medical societies: Alpha Omega Alpha , American College of Cardiology , American College of Physicians , American Heart Association , Phi Beta Kappa , and Society for Cardiac Angiography and Interventions

Francisco Talavera, PharmD, PhD Adjunct Assistant Professor, University of Nebraska Medical Center College of Pharmacy; Editor-in-Chief, Medscape Drug Reference

Disclosure: Medscape Salary Employment

What would you like to print?

  • Print this section
  • Print the entire contents of
  • Print the entire contents of article

Medscape Logo

  • Rapid Review Quiz: Decoding Heart Failure
  • Fast Five Quiz: Heart Failure With Reduced Ejection Fraction
  • Fast Five Quiz: Heart Failure Comorbidities
  • Trending Clinical Topic: Heart Failure Guidelines
  • Skill Checkup: A 54-Year-Old Black Man With Treated Heart Failure Has Worsening Symptoms
  • Skill Checkup: A Woman With Long‐standing Hypertension and Worsening Dyspnea on Exertion
  • Fast Five Quiz: Test Your Knowledge on Key Aspects of Heart Failure
  • SGLT2 Inhibitors Safe for HF in Congenital Heart Disease
  • Semaglutide Helps Heart Failure 'Regardless of Diuretics'
  • Yoga May Augment Medical Therapy in Heart Failure

A Global Health Tidal Wave? Cardiorenal Metabolic Syndrome

  • Drug Interaction Checker
  • Pill Identifier
  • Calculators

Cardiology Guidelines: 2017 Midyear Review

  • 2001/viewarticle/994736 Early or Delayed AF Ablation After Heart Failure Hospitalization?

Guiding Heart Failure: The Significance of Cardiac Biomarkers

  • 2001/s/viewarticle/998815 Cardiologists, Patients Can Talk Drug Costs
  • 2001/viewarticle/977754 Using the Power of Machine Learning to Hone in on HF Subtypes

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Published: 05 March 2020

Acute heart failure

  • Mattia Arrigo 1 ,
  • Mariell Jessup 2 ,
  • Wilfried Mullens 3 , 4 ,
  • Nosheen Reza 2 ,
  • Ajay M. Shah 5 ,
  • Karen Sliwa 6 &
  • Alexandre Mebazaa 7 , 8  

Nature Reviews Disease Primers volume  6 , Article number:  16 ( 2020 ) Cite this article

186k Accesses

240 Citations

214 Altmetric

Metrics details

  • Cardiovascular diseases
  • Heart failure

Acute heart failure (AHF) is a syndrome defined as the new onset (de novo heart failure (HF)) or worsening (acutely decompensated heart failure (ADHF)) of symptoms and signs of HF, mostly related to systemic congestion. In the presence of an underlying structural or functional cardiac dysfunction (whether chronic in ADHF or undiagnosed in de novo HF), one or more precipitating factors can induce AHF, although sometimes de novo HF can result directly from the onset of a new cardiac dysfunction, most frequently an acute coronary syndrome. Despite leading to similar clinical presentations, the underlying cardiac disease and precipitating factors may vary greatly and, therefore, the pathophysiology of AHF is highly heterogeneous. Left ventricular diastolic or systolic dysfunction results in increased preload and afterload, which in turn lead to pulmonary congestion. Fluid retention and redistribution result in systemic congestion, eventually causing organ dysfunction due to hypoperfusion. Current treatment of AHF is mostly symptomatic, centred on decongestive drugs, at best tailored according to the initial haemodynamic status with little regard to the underlying pathophysiological particularities. As a consequence, AHF is still associated with high mortality and hospital readmission rates. There is an unmet need for increased individualization of in-hospital management, including treatments targeting the causative factors, and continuation of treatment after hospital discharge to improve long-term outcomes.

Similar content being viewed by others

clinical presentation of acute heart failure

Acute kidney injury

clinical presentation of acute heart failure

Management of hypertensive crisis: British and Irish Hypertension Society Position document

clinical presentation of acute heart failure

Introduction

Heart failure (HF) is a chronic and progressive clinical syndrome induced by structural or functional cardiac abnormalities displaying either reduced (in HF with reduced ejection fraction (HFrEF)) or preserved (in HF with preserved ejection fraction (HFpEF)) left ventricular ejection fraction (LVEF) 1 . Cardiac dysfunction leads to elevated cardiac filling pressures at rest and during stress 1 . HF symptoms include dyspnoea (shortness of breath) and fatigue, often accompanied by typical physical signs, such as pulmonary rales (abnormal crackling sounds), peripheral oedema or distended jugular veins 1 . The substantial reduction in short-term mortality in patients with several cardiac conditions (particularly acute coronary syndromes and congenital heart disease) and the relevant improvement in long-term survival in patients with HFrEF (as a result of widespread use of effective disease-modifying oral therapies and devices), combined with several demographic changes, such as extended life expectancy, have sharply increased the number of patients living with HF 2 . In developed countries, HF has become a substantial public health problem, affecting 2% of the adult population, and the number of hospital admissions related to HF has tripled since the 1990s 2 .

Acute HF (AHF) is defined as new or worsening of symptoms and signs of HF and is the most frequent cause of unplanned hospital admission in patients of >65 years of age 3 . From a clinical perspective, we distinguish de novo HF — in which symptoms occur in patients without a previous history of HF — from acutely decompensated HF (ADHF) — in which symptoms increase in patients with previously diagnosed chronic HF. This classification provides little additional information in regard to the pathophysiology of AHF but has mainly clinical implications (de novo HF requires a more extensive diagnostic process to investigate the underlying cardiac pathology than ADHF). As HF is a chronic and progressive disease, the majority of hospitalizations are related to ADHF rather than de novo AHF 4 , 5 . The clinical presentation of AHF is characterized mostly by symptoms and signs related to systemic congestion (that is, extracellular fluid accumulation, initiated by increased biventricular cardiac filling pressures) 6 , 7 . Accordingly, the initial treatment in most patients with AHF consists of non-invasive ventilation and intravenous diuretics, which are administered alone or, especially in Europe and Asia, in combination with short-acting vasodilators 8 . Only a minority of patients with AHF present with cardiogenic shock, a critical condition characterized by the presence of clinical signs of peripheral tissue hypoperfusion; cardiogenic shock has a tenfold higher in-hospital mortality than AHF without shock and requires specific treatments 9 , 10 .

In contrast to the substantial improvements in the treatment of chronic HFrEF, AHF is still associated with poor outcomes, with 90-day readmission rates and 1-year mortality reaching 10–30% 11 , 12 . Although AHF is not a specific disease but the shared clinical presentation of different, heterogeneous cardiac abnormalities, most patients still receive decongestive drugs only, at best tailored according to the initial haemodynamic status with little regard to the underlying pathophysiological particularities. This approach might have contributed to the multitude of neutral or negative clinical trials assessing the effect of decongestive treatments on survival and to the persistence of poor outcomes in AHF. Thus, there is an unmet need for increased individualization and continuation of treatment after hospital discharge to improve long-term outcomes. This Primer reviews current concepts of epidemiology, pathophysiology, diagnosis and management of AHF to stimulate advances in research and clinical practice to improve patient outcomes. As cardiogenic shock is a separate entity with specific features, it is not discussed in this Primer.

Epidemiology

There are several reasons why global data on AHF are very limited. Differential coding of the syndrome, coupled with nuanced differences in case definitions, defies simple regional comparison. The International Classification of Disease (ICD) system classifies AHF and chronic HF as intermediate conditions and not underlying causes of death. The ICD also does not distinguish between de novo HF and ADHF as reasons for hospital admission. No global data on the proportion of HFrEF and HFpEF as underlying causes of AHF are available. The Global Burden of Disease (GBD) collaborators reported on global, regional and national age-specific and sex-specific mortality of 282 causes of death in 195 countries for the period 1980–2017, including cardiovascular diseases such as rheumatic heart disease, ischaemic heart disease and cardiomyopathy, but they did not list AHF 13 . The latest estimate by the GBD team in 2010 was 37.7 million cases of prevalent HF worldwide, leading to an average of 4.2 years lived with this disability for each patient, but data on the global incidence of AHF were not reported 14 . Data on annual hospitalizations for HF are only available for the USA and Europe and exceed 1 million in both regions 4 , 5 . Among these hospitalizations, >90% were due to symptoms and signs of fluid accumulation (indicating AHF). In addition, up to one in four patients (24%) are readmitted within 30 days, readmission rates in the first 3 months after hospitalization for AHF may reach 30% in the USA and in other countries 4 and one in two patients (50%) are readmitted within 6 months 4 , 5 . Recurrent fluid accumulation in patients with HF has uniformly been associated with worse outcomes independent of age and renal function 15 . In multiple studies of the 30-day to 90-day post-discharge period, ~25–30% of patients with AHF are readmitted during this time frame 16 , 17 , 18 , 19 , 20 . However, a substantial proportion of these patients are readmitted for a non-HF-related cause 21 , 22 . Medical comorbidities precipitate rehospitalization and, when poorly managed, contribute to worsening HF over time 22 . Psychosocial factors such as anxiety, depression, cognitive impairment and social isolation also confer increased risk of unplanned recurrent readmission or death of patients following hospitalization for AHF 23 .

There are no national data on the prevalence of AHF or chronic HF in low-income and middle-income countries. All registries of HF for these regions are based on hospital registries that included only patients admitted for AHF, without separating de novo HF from ADHF. Data from some of the key registries have recently been summarized 24 but focus on aetiology, risk factors, sociodemographic profile and mortality. The INTER-CHF study, one of the largest registries, reported on 5,823 patients with HF from 108 centres in six geographical regions 25 . The overall 1-year mortality was 16.5%, with the highest mortality in Africa (34%) and India (23%), about average mortality in southeast Asia (15%) and the lowest mortality in China (7%), South America (9%) and the Middle East (9%) 25 .

Risk factors

A systematic review of worldwide risk factors for HF found that ischaemic heart disease was the major underlying contributor to AHF admissions in >50% of patients in high-income regions, as well as eastern and central European regions 26 . In Asia Pacific high-income regions and Latin America, ischaemic heart disease contributed to 30–40% of admissions 26 , whereas in sub-Saharan Africa it contributed to <10% 27 . Hypertension was a consistent contributor to HF globally (17%) 26 . Of the other commonly reported risk factors, rheumatic heart disease was particularly prevalent in East Asia (34%) and sub-Saharan Africa (14%) 26 . The heterogeneous group of cardiomyopathies (which can include familial, peripartum, infective (for example, due to HIV infection), autoimmune, post-myocarditis and idiopathic cardiomyopathy, amongst others) were particularly prevalent in Africa (25.7%), with Chagas disease-associated cardiomyopathy being a specific cause in Latin America 26 . Chagas disease-associated acute myocarditis is commonly (>50% cases) associated with a substantial pericardial effusion, but it usually leads to AHF in only 1–5 of every 10,000 infected people 28 . However, Chagas disease remains common in Latin America and is the cause of HF in 10% of patients in the RAMADHF study and 28% in the GESICA study 29 , 30 .

In high-income regions with associated high scores in the human development index (a statistical tool that takes into account life expectancy, education and income), patients with AHF typically have a median age of >75 years at presentation, whereas in other areas, such as Latin America and sub-Saharan Africa, the median age of patients with AHF is up to two decades lower 25 . This difference could be due to poorly treated hypertension, ischaemic heart disease and late diagnosed rheumatic heart disease leading to HF presentation in younger age groups. In addition, there are differences between regions in the sex distribution; for example, rheumatic heart disease commonly affects women more than men 31 , 32 , and peripartum cardiomyopathy is particularly common in Africa 33 . As the obesity epidemic also affects women disproportionately, hypertensive heart disease leading to HF is commonly more prevalent in women than men 25 .

Morbidity and mortality

Globally, in-hospital AHF mortality hovers at ~4%, rises to ~10% within 60 to 90 days after discharge and increases further to 25–30% at 1 year 16 , 17 , 18 , 34 , 35 . The INTER-CHF prospective cohort study showed striking global variations in HF-associated mortality, with the highest 1-year overall and HF-related mortality in the countries with the youngest populations, such as India and African countries 25 . However, there was no analysis of HFpEF versus HFrEF as the underlying condition in the HF group.

Data from the THESUS-HF registry (a prospective study of AHF in nine sub-Saharan countries) were analysed to determine the predictors of readmission and outcome (including death) after an AHF event 35 . Similar to results in high-income countries, the predictors of 180-day mortality included malignancy, severe lung disease, smoking history, systolic blood pressure and heart rate either below or above their physiological ranges and symptoms and signs of congestion (orthopnoea (dyspnoea when lying flat), peripheral oedema and rales) at admission, kidney dysfunction, anaemia and HIV positivity. The risks predicted by calibration plots, comparing observed event rates with those predicted by the models, were generally low for all risk factors considered, suggesting that the main factors contributing to adverse outcomes in patients with AHF are still largely unknown 35 .

Mechanisms/pathophysiology

Pathophysiological mechanisms of ahf.

An underlying structural or functional cardiac condition is a prerequisite for AHF and includes a multitude of different acute (for example, myocardial infarction) or chronic (for example, dilated cardiomyopathy and ischaemic heart disease) cardiac pathologies. The underlying cardiac disease leads to the activation of several pathophysiological pathways (at first adaptive responses, which with time become maladaptive) that counter the negative effects of HF on oxygen delivery to the peripheral tissues, but such pathways can also eventually cause systemic congestion, ventricular remodelling and organ dysfunction 36 . Furthermore, some acute diseases can act as precipitating factors and trigger AHF either by directly impairing cardiac diastolic and/or systolic function or by further promoting systemic congestion 36 . Systemic congestion has a major effect on the clinical presentation in the majority of patients with AHF and is a relevant determinant of multi-organ dysfunction occurring in AHF (Fig.  1 ). The pathophysiology of AHF is heterogeneous, as it is greatly affected by the nature of the underlying cardiac disease. It is perhaps not surprising, therefore, that the responses to treatment may vary and that different patients may respond best to distinct treatment strategies that depend on the underlying pathophysiology.

figure 1

Acute heart failure (HF) results from the combination of an underlying but newly diagnosed cardiac dysfunction and precipitating factors or the onset of a new cardiac dysfunction (de novo HF) or the combination of an underlying chronic cardiac dysfunction and one or more precipitating factors (acutely decompensated HF (ADHF), that is, decompensation of chronic HF). Precipitating factors may directly affect left ventricular (LV) or right ventricular (RV) function (for example, myocardial ischaemia and arrhythmias) or may contribute to the development of congestion (for example, infection, hypertension and non-compliance with treatment recommendations). LV dysfunction (diastolic dysfunction in HF with preserved ejection fraction (HFpEF) or diastolic and systolic dysfunction in HF with reduced ejection fraction (HFrEF)) leads to pulmonary congestion, which in turn contributes to RV dysfunction and systemic congestion. Systemic congestion, neurohumoral activation and inflammation negatively affect ventricular function and further contribute to self-perpetuating congestion.

LV systolic and diastolic dysfunction

An acute change in cardiac function, mostly a worsening of left ventricular (LV) diastolic function, which in turn leads to an increase in LV filling pressures and pulmonary congestion, can result in AHF 37 ; an example of such sudden changes is acute myocardial ischaemia. Several pathophysiological mechanisms underlie the link between ischaemia, LV systolic and diastolic dysfunction and pulmonary congestion. LV contraction is highly dependent on oxidative energy generation and, therefore, ischaemia triggers systolic impairment, which leads to an increased residual LV end-diastolic volume and filling pressure. LV filling normally occurs in two phases, an early rapid phase that is highly dependent upon fast myocardial relaxation and a later phase that is dependent on left atrial contraction and the atrial-to-ventricular pressure gradient, which in turn is affected by the physical properties of the LV (for example, stiffness). Myocardial relaxation is also an active energy-requiring process that involves removing cytoplasmic calcium, mostly via re-uptake into the sarcoplasmic reticulum by the sarcoplasmic reticulum Ca 2+ ATPase (SERCA) pump and in part via extrusion across the cardiomyocyte plasma membrane. The end-diastolic properties of the LV are affected by the residual LV end-diastolic volume, structural changes (for example, fibrosis) and also by extremely delayed relaxation. The reduction in oxidative ATP generation in cardiomyocytes with the onset of severe acute ischaemia rapidly impairs myocardial relaxation, thereby affecting early LV filling and further increasing filling pressures. The presence of any coexisting conditions in which relaxation is already impaired or end-diastolic LV stiffness is increased will increase the likelihood of AHF. Conditions in which end-diastolic LV stiffness may be increased (and, therefore, also conditions with an increased risk of AHF precipitated by ischaemia) include chronic LV systolic dysfunction with raised LV end-diastolic volume and structural fibrosis and/or hypertrophy, both of which could result from diabetes mellitus, chronic hypertension, chronic kidney disease, chronic aortic stenosis and ageing 38 . LV filling may also be impaired by the sudden development of atrial fibrillation with the accompanying loss of atrial contraction, which may substantially increase filling pressures when there is already pre-existing diastolic dysfunction. For example, severe mitral stenosis (which is a common manifestation of rheumatic heart disease) is a type of diastolic dysfunction due to the valve abnormality rather than LV structural disease, and it can also induce atrial fibrillation, thereby increasing the risk of triggering AHF.

Fluid retention

In HF, an increase in the volume of extracellular fluid (referred to as fluid retention or fluid accumulation) and/or a change in the compliance of venous beds (which results in fluid redistribution without an increase in the overall volume) can lead to an increase in filling pressures. In fact, in the majority of patients, AHF occurs without acute changes in cardiac function but is induced by fluid accumulation and/or redistribution, which results in systemic congestion, especially in the presence of an underlying diastolic dysfunction 39 . The interactions between intravascular and interstitial fluid volumes are complex, and there is no linear correlation between central haemodynamics and volume changes 40 . Animal studies have shown that marked intravascular volume expansion does not lead to increased cardiac filling pressures if sympathetic activity is low 41 , 42 , and in patients with HF intravascular volume is only marginally reduced after diuretic therapy despite large reductions in body weight 40 . By contrast, only half of the patients exhibit a weight gain of >0.9 kg over the month preceding hospital presentation for ADHF, indicating that changes in the compliance state of the venous beds are also important drivers of congestion 43 . The majority of the retained sodium is stored in the extracellular compartment, which consists of both the intravascular compartment and the interstitium 44 . In healthy individuals, increased total body sodium is usually not accompanied by oedema formation, as a large quantity of sodium may be buffered by the interstitial glycosaminoglycan networks without compensatory water retention 45 . Moreover, the interstitial glycosaminoglycan networks display low compliance (limited elastic properties), which prevents fluid accumulation in the interstitium 46 . In patients with HF, when sodium accumulation persists, the glycosaminoglycan networks may become dysfunctional, resulting in reduced buffering capacity, increased interstitial compliance and oedema formation even in the presence of mildly elevated hydrostatic pressures 44 .

Fluid retention is frequently related to increased neurohumoral activation (that is, activation of the renin–angiotensin–aldosterone system and the vasopressin system) leading to renal salt and water retention, although it can also be iatrogenic (for example, caused by the administration of inappropriately large amounts of intravenous fluids). The neurohumoral pathway is already activated above the physiological baseline level early during disease progression in patients with chronic HF (even before the development of symptoms) or kidney disease, and, therefore, these patients are particularly prone to fluid accumulation. Mechanisms and consequences of neurohumoral activation have been extensively reviewed elsewhere 47 . Importantly, the resulting organ dysfunction contributes to self-perpetuation of congestion.

In HF, alterations in both proximal and distal nephron segments increase kidney sodium avidity 48 , which is already increased even before clinical symptoms of HF occur 49 , 50 . Furthermore, in several studies increased central venous pressure has been associated with worsening of renal function (WRF), often resulting in a further drop in natriuresis 51 , 52 , 53 . However, changes in renal function during AHF need to be interpreted within the specific clinical context, as this approach helps to correctly assess risk and determine further treatment strategies. In fact, it is possible that changes in renal function parameters occurring during AHF that would typically indicate WRF do not correspond to ‘true’ WRF, when accompanied by simultaneous favourable ongoing diuresis and improvement in HF status. Currently, misinterpretation of WRF in the AHF setting is a leading cause of decongestion not being achieved in AHF. To distinguish between ‘true’ WRF and ‘pseudo’ WRF during AHF, renal evaluation should include the assessment not only of changes in glomerular function (indicating the development of WRF), but also of the tubular response to diuretic therapy (diuretic response and/or efficiency), that is, the ability to eliminate residual congestion and the administered therapy.

Fluid redistribution

Sympathetic stimulation can induce a transient vasoconstriction leading to a sudden displacement of volume from the splanchnic and peripheral venous system to the pulmonary circulation, without exogenous fluid retention — that is, fluid redistribution 54 . Large veins physiologically contain one-quarter of the total blood volume and stabilize cardiac preload, buffering fluid retention 55 . Preload indicates the degree of stretch of cardiomyocytes at the end of diastole and correlates with the end-diastolic volume and pressure. By contrast, afterload indicates the pressure that the heart has to overcome to eject blood during ventricular contraction and correlates with systolic arterial pressure. A mismatch in the ventricular–vascular coupling relationship with increased afterload and decreased venous capacitance (leading to increased preload and end-diastolic volume) may excessively increase cardiac workload and exacerbate pulmonary and systemic congestion 56 . Finally, acute mechanical factors may also increase ventricular preload and cause AHF; for example, the sudden occurrence of mitral valve regurgitation due to ruptured papillary muscle chords or the sudden development of a ventricular septal defect.

Fluid accumulation and fluid redistribution both produce systemic congestion in AHF, but their relative contributions probably vary according to different clinical scenarios, and the decongestive therapy should be tailored accordingly (see Management) 36 .

Precipitating factors of AHF

The onset and increase in systemic congestion that precede AHF may develop over hours up to days, and can be triggered by several factors, either directly through stimulation of pathophysiological mechanisms leading to fluid accumulation or redistribution or indirectly through a worsening of cardiac diastolic or systolic function. The understanding of the pathophysiology involved in the development of AHF is important for providing the appropriate treatment. Although in many patients a progressive increase in body weight and pulmonary pressures may be observed as early as several days before hospital admission, in a relevant proportion of patients AHF is associated with only a minimal increase in body weight 39 , 43 . Several registries, including the North American OPTIMIZE-HF registry and the Euro-Asian registry of the GREAT network, have investigated the presence of precipitants in patients with AHF 57 , 58 . Acute coronary syndromes, arrhythmias (in particular atrial fibrillation), infections (in particular airway infections), uncontrolled hypertension and non-compliance with dietary recommendations and drug prescriptions are the most common identified precipitants 57 , 58 . Of note, in a relevant proportion of patients (~40–50%), no precipitants could be identified, whereas a combination of multiple factors were present in ~5–20% of patients 57 , 58 .

The identification of precipitants provides prognostic information, as highlighted by several studies showing an association between precipitating factors and both mortality and readmission rates 57 , 58 , 59 , 60 . AHF precipitated by acute coronary syndromes or infection is associated with higher short-term mortality than AHF precipitated by atrial fibrillation or uncontrolled hypertension 57 , 58 . Notably, although patients with AHF precipitated by acute coronary syndromes and those with AHF precipitated by infection have similar unfavourable prognoses, the risk of death changes with time differently in the two patient groups: it is the highest during the first days after admission in the first group and peaks ~3 weeks after admission in the second 58 , 61 . The explanation for this phenomenon is speculative; we might suggest a complex interaction between infection and a combination of endothelial dysfunction, atherosclerotic plaque instability, activated coagulation, fluid retention, inflammatory and ischaemic myocardial injury, arrhythmias and the risk of other precipitating non-cardiac illnesses that may lead to death 58 . Finally, and most importantly, the identification of precipitating factors enables the delivery of specific treatments directed towards the underlying causes of AHF, in addition to decongestive therapy.

Congestion and organ dysfunction

In the heart, elevated ventricular filling pressures lead to increased ventricular wall tension, myocardial stretch and remodelling, contributing to a progressive worsening in cardiac contractility, valvular regurgitation and systemic congestion 62 . In response to the increased wall tension, circulating natriuretic peptides (which stimulate diuresis and vasodilation) are physiologically released by atrial and ventricular cardiomyocytes as a compensatory mechanism, and often high-sensitivity cardiac troponins are detectable in a large proportion of patients with AHF, revealing nonischaemic myocyte injury or necrosis 63 . Increases in left atrial pressure and mitral valve regurgitation will increase the hydrostatic pressure in the pulmonary capillaries, thereby increasing fluid filtration rate from the capillaries to the pulmonary interstitium, causing lung stiffness and dyspnoea 64 . Notably, the relationship between hydrostatic pressure and interstitial fluid content is rather complex, as other mechanisms are involved in fluid homeostasis. For example, the lymphangiogenic factor VEGF-D has been found to regulate and mitigate pulmonary and systemic congestion in patients with HF or renal failure 65 , 66 , 67 . Indeed, in the early stage of lung congestion, the lymphatic system can cope with the large volume of interstitial fluid, but eventually, the drainage capacity is exceeded. Hence, fluid moves to pleural and intra-alveolar spaces causing pleural effusion and pulmonary oedema 68 .

Systemic congestion is a central feature in most patients with AHF 6 . In addition to poor cardiac function, numerous organs play a part in the development and propagation of congestion 69 . Congestion is the essential pathophysiological mechanism of impaired organ function in AHF, and hypoperfusion — if present — might cause further deterioration in organ function and is associated with increased mortality risk 6 . Improvement in organ function with decongestive therapies has been associated with a reduced risk of death, and, therefore, prevention and treatment of organ dysfunction is a key therapeutic target in patients with AHF.

AHF is associated with WRF. Elevated central venous pressure leads to renal venous hypertension, which in turn increases renal interstitial pressure. Ultimately, the hydrostatic pressure in the renal interstitium exceeds the intratubular hydrostatic pressure, resulting in the collapse of tubules and, therefore, reduced glomerular filtration rate 70 . In addition, renal venous hypertension induces a reduction in renal blood flow, renal hypoxia and ultimately interstitial fibrosis 51 , 52 , 71 . Other contributors to AHF-induced renal dysfunction include inflammatory processes, iatrogenic factors (for example, contrast media and nephrotoxic medications), impaired cardiac output and elevated intra-abdominal pressure 7 , 72 . Of note, an increase in plasma creatinine is often interpreted by clinicians as a sign of hypovolaemia, prompting a reduction in decongestive therapy, on the basis that excessive decongestion might result in renal tubular damage; however, this is not always the case, as discussed above (see Fluid retention) 73 , 74 . In patients with an increase in creatinine during decongestive therapy, it is recommended that decongestive therapy is pursued until euvolaemia is achieved 75 , as clinical outcomes are extremely poor if patients are discharged with ongoing congestion in the presence of WRF 76 . By contrast, relying exclusively on serial measurements of levels of biomarkers (such as circulating natriuretic peptides) to assess changes in volume might lead to inappropriate dose escalation of loop diuretics in patients without substantial residual congestion. This dose escalation may lead to adverse effects such as hypotension and/or further WRF. A multiparameter-based evaluation of congestion before discharge would be of benefit in patients with HF. In addition to biomarkers, clinical assessment at rest and during dynamic manoeuvres, supplemented with technical assessments (such as echocardiography or measurement of pulmonary pressures), is probably the best strategy, although it needs prospective evaluation 75 .

In patients with liver congestion, elevations in alkaline phosphatase, bilirubin and/or γ-glutamyl transferase (also known as glutathione hydrolase 1 proenzyme) are often observed 77 , 78 , 79 . Centrilobular necrosis and markedly elevated transaminases (alanine aminotransferase and aspartate aminotransferase) owing to hypoperfusion in the setting of hypoxic hepatitis are observed in severe hypoperfusion states such as cardiogenic shock 78 .

Splanchnic congestion results in increased intra-abdominal pressure and ischaemia of villi, which modify intestinal morphology, and alters intestinal permeability, nutrient absorption and the bacterial biolayer, possibly contributing to chronic inflammation and malnutrition 80 , 81 , 82 . Additionally, venous congestion and/or hypoperfusion impairs the splanchnic microcirculation and increases the risk of bowel ischaemia, enabling lipopolysaccharide or endotoxin produced by Gram-negative gut bacteria to enter the circulatory system and increase the pro-inflammatory environment of AHF 56 . Finally, congestion per se also results in endothelial activation, which further promotes a pro-inflammatory environment 83 , 84 .

Diagnosis, screening and prevention

The management of patients with HF is strikingly heterogeneous across the world according to sociocultural disparities and differences in health-care systems. Many cardiology societies have endeavoured to increase awareness of HF among the population in different countries and to educate health-care professionals to improve the management of patients with HF. The following sections about diagnosis and treatment of AHF reflect current recommendations in high-income countries and may be substantially different from management standards in low-income or developing countries depending on local availability of resources. The modern management of patients with AHF also includes an optimal interplay between accurate diagnosis, rapid implementation of disease-modifying drugs and devices, specific treatment of the underlying cardiac disease and frequent outpatient follow-up visits. Whereas loop diuretics to relieve congestion are inexpensive and widely available, disease-modifying drugs (particularly sacubitril (a neprilysin inhibitor)–valsartan (an angiotensin receptor blocker) 85 , which promotes vasodilation and natriuresis, and sodium-glucose cotransporter 2 inhibitors, which reduce blood glucose levels in patients with diabetes mellitus and have also been shown to have beneficial effects in patients with HF) 86 and cardiac devices are usually available only in high-income areas. Furthermore, accurate diagnosis of the underlying cardiac diseases and specific treatments often require multimodal imaging techniques, as well as interventional and surgical procedures, which are mostly available in high-volume centres in developed countries. Finally, frequent follow-up visits to reduce the need for hospital readmissions are only feasible in countries with an established network of health-care providers with sufficient expertise in the treatment of patients with HF.

Initial diagnosis

Clinical presentation.

Symptoms and signs related to systemic congestion characterize the clinical picture of patients presenting with AHF, to a similar extent regardless of LVEF 87 . The most common symptoms include dyspnoea during exercise or at rest, orthopnoea, fatigue and reduced exercise tolerance; symptoms are often accompanied by clinical signs such as peripheral oedema, jugular vein distension, the presence of a third heart sound (known as “S3 gallop”, an early diastolic low-frequency sound that may be present under different haemodynamic conditions and might represent termination of the rapid filling of the left ventricle), and pulmonary rales 88 . In patients presenting with chest discomfort, the differentiation between AHF and acute coronary syndrome may be challenging. Symptoms and signs related to peripheral hypoperfusion, such as cold and clammy skin, altered mental status and oliguria, characterize cardiogenic shock. Cardiogenic shock, as well as respiratory failure, myocardial infarction and arrhythmia, should be rapidly excluded during the initial triage of patients admitted for suspected AHF because these conditions require an appropriate level of monitoring and specific treatments 9 , 89 . Commonly accepted criteria for hospitalization in an intensive care unit or a cardiac care unit include haemodynamic instability (heart rate <40 beats per minute or >130 beats per minute, systolic blood pressure <90 mmHg or evidence of hypoperfusion) and respiratory distress (respiratory rate >25 breaths per minute, peripheral oxygen saturation <90% despite supplemental oxygen, use of accessory muscles for breathing or need for mechanical ventilatory support) 90 .

Several algorithms and scores, most of which include clinical variables and biomarkers, have been developed to predict in-hospital death, but most of these tools have not been adequately prospectively tested for triage or resources allocation purposes. The ADHERE risk tree is used to classify patients on the basis of whether three parameters collected at admission (that is, blood urea nitrogen, systolic blood pressure and serum creatinine) are above or below specific cut-off values; this tool enables patient stratification into five groups with substantially different in-hospital mortality ranging from 2% to 22% 91 . The GWTG-HF score is computed by adding the points derived from seven variables (age, systolic blood pressure, heart rate, blood urea nitrogen, plasma sodium, history of chronic obstructive pulmonary disease and black ethnicity) and enables stratification into nine categories with in-hospital risk of death ranging from <1% to >50% 92 . The MEESSI-AHF score includes 13 independent risk factors and may be used to estimate the 30-day mortality in patients with AHF 93 .

Diagnostic work-up

The clinical picture of AHF is neither sensitive nor specific enough for confirming or ruling out the diagnosis; thus, additional tests are required 94 . Cardiovascular biomarkers play a crucial part in the diagnostic process of AHF. Patients presenting with suspected AHF should undergo measurement of plasma natriuretic peptides (for example, brain natriuretic peptide (BNP), N-terminal pro-brain natriuretic peptide (NT-proBNP) or mid-regional pro-atrial natriuretic peptide (MR-proANP)). Although no diagnostic test can on its own reliably differentiate AHF from chronic HF, as all cardiovascular biomarkers are impaired in both patient groups, natriuretic peptides display high sensitivity for detecting underlying cardiac disease in patients presenting with acute dyspnoea. In patients with AHF, levels of circulating natriuretic peptides are elevated compared with levels in patients with shortness of breath of non-cardiac origin 95 , 96 , 97 ; thus the measurement of natriuretic peptides provides higher diagnostic accuracy than clinical evaluation alone 98 . By contrast, dyspnoea in patients with normal (or unchanged) circulating natriuretic peptides is very likely to be of non-cardiac origin. The measurement of natriuretic peptides is recommended in patients with suspected AHF upon admission 1 , 89 . In patients with chronically elevated natriuretic peptides owing to chronic HF, a relevant increase in circulating natriuretic peptides may indicate AHF. Additional tests, such as echocardiography or other imaging procedures, are required to confirm the diagnosis of AHF in patients with elevated natriuretic peptides. Several new biomarkers reflecting different pathophysiological aspects of AHF (for example, myocardial injury, systemic congestion, inflammation and fibrosis) may be useful for diagnostic or prognostic purposes, but their role in routine clinical practice is still not well established.

The initial diagnostic process should include a comprehensive evaluation not only of the clinical phenotype but also of the underlying cardiac disorders, precipitating factors and comorbidities. Our (M.A.) group has proposed a ‘7-P’ protocol for guiding evaluation and personalization of treatment. The seven elements are phenotype, pathophysiology, precipitants, pathology, polymorbidity, potential iatrogenic harms and patient preferences 99 (Box  1 ). The diagnosis of AHF is frequently made clinically based on history and clinical signs assisted by measuring circulating natriuretic peptides. The role of imaging for the initial assessment of AHF is limited to patients in whom the underlying cardiac condition is unknown (for example, patients with de novo HF, who require a more extensive diagnostic process than patients with ADHF) or the detection of congestion is uncertain. In these patients, echocardiography and lung ultrasonography may add valuable information. Transthoracic echocardiography should be performed in all patients with de novo HF or in patients with ADHF when a relevant change in cardiac pathology is suspected, to estimate LV and RV function and exclude severe valve disease or pericardial tamponade. Lung ultrasonography has emerged as a valuable modality to detect and monitor pulmonary congestion in patients with AHF. This bedside technique enables the detection of interstitial fluid in the pulmonary parenchyma in a rapid, inexpensive and reliable manner 100 , 101 . An ischaemic trigger of AHF, such as acute coronary syndromes, should be ruled out by electrocardiography and (serial) measurement of cardiac troponins; arrhythmias can be evaluated by electrocardiography, continuous electrocardiographic monitoring or interrogation of implantable cardioverter–defibrillator interrogation in selected patients; and infections by measurement of inflammatory markers (for example, C-reactive protein and procalcitonin) and additional investigations according to the clinical presentation (for example, analysis of microbiological specimens and imaging). Additional imaging modalities (for example, MRI) are rarely needed during the initial work-up but may be helpful during further investigations. The initial laboratory evaluation should also include a basic assessment of the function of other organ systems (for example, kidney, liver and blood).

Current recommendations on the management of AHF are mainly based on expert opinion rather than robust evidence, as randomized controlled trials are either lacking or their results are neutral or negative 1 , 3 , 9 . Recent data have shown that timely initiation of therapy may be a crucial factor in the treatment of AHF, with a positive association between short time from admission to diuretic administration and improved in-hospital survival. For this reason, the initial treatment should be delivered as soon as possible, ideally as early as during the diagnostic work-up 102 . However, because short-term intravenous therapy with diuretics or vasodilators is unlikely to change the mid-term and long-term clinical course in patients with AHF, the choice of initial treatment should take into account not only the clinical phenotype but also the underlying cardiac disorders, precipitating factors and comorbidities.

Box 1 The ‘7-P’ protocol

The assessment of the clinical phenotype based on peripheral perfusion (whereby normal perfusion is considered ‘warm’ and symptoms or signs of hypoperfusion are considered ‘cold’) and/or systemic congestion (whereby no congestion is considered ‘dry’ and the presence of congestion is considered ‘wet’) enables the classification of patients into one of four profiles. The vast majority of patients with AHF are well perfused but congested (‘warm–wet‘).

The initial treatment tackling haemodynamic disorders (for example, vasodilators and/or diuretics to reduce systemic congestion and positive inotropic drugs to improve peripheral perfusion) should be personalized according to the clinical phenotype and the leading pathophysiology (for example, fluid accumulation, fluid redistribution or peripheral hypoperfusion).

Identification of the precipitants of AHF is essential for providing optimal specific (medical and/or surgical) therapy and for estimating both prognosis and recovery potential.

Similarly, identification of the underlying cardiac pathology can contribute to tailoring the treatment.

The assessment of polymorbidity (for example, renal and hepatic dysfunction) or other relevant conditions (such as pregnancy, bleeding risk and allergies) should be integrated into the management plan.

Potential iatrogenic harms associated with diagnostic procedures and treatment should also be considered.

Patient preferences and ethical considerations should be integrated into the personalization of the treatment. Discussion with the patient or with relatives about resuscitation directives and treatment options are crucial and need to be evaluated early rather than late, particularly in patients with AHF who might show rapid deterioration. In the absence of long-term therapeutic options, palliation and supportive care should be offered to patients and relatives.

Screening and prevention

As mentioned above, AHF can arise de novo or in patients with previously diagnosed HF (ADHF). The prospective STOP-HF study investigated the efficacy of a natriuretic peptide-based screening programme and collaborative care in reducing newly diagnosed HF in an at-risk population 103 . However, although this study showed a significant reduction in the rate of emergency hospitalization for major cardiovascular events in the screening group, the reduction in the incidence of HF did not reach statistical significance. Thus, the role of screening in preventing HF — and more specifically AHF — has yet to be determined, and screening is not recommended by current guidelines 1 .

By contrast, prevention of decompensation in patients with previously diagnosed HF is of major importance. Hospital readmissions are frequent — in particular during the first months after hospital discharge for AHF — and are associated with adverse outcomes and relevant health-care expenditure 12 . The optimal strategy for reducing hospital readmission has not been prospectively validated in clinical trials. Residual congestion and lack of disease-modifying treatment implementation before hospital discharge have been associated with worse post-discharge outcomes 104 , 105 . Patient education and empowerment may play a crucial part: patients should understand the importance of compliance with medical treatment, be able to recognize symptoms or signs of worsening HF, have a plan about when and how to start or increase diuretic treatment, and know when to contact their cardiologist or the medical emergency system to avoid unnecessary delay. Furthermore, particular attention should be given to avoid self-medication or initiation of contraindicated drugs (for example, NSAIDs) by other physicians who are unaware of the HF diagnosis. Finally, a continuation of the chronic treatment of HF (diuretics and disease-modifying drugs) without interruption should be ensured, although this goal may be challenging, in particular in low-income countries and in the absence of insurance coverage for medical treatments.

Pre-hospital early management

There is a growing body of evidence that delayed treatment delivery is associated with poor outcomes in AHF 102 . For this reason, current guidelines advocate a ‘time-to-treatment’ concept, similar to those for acute myocardial infarctions or cerebrovascular accidents, and recommend early initiation of treatment in patients with AHF, optimally before hospital admission 1 , 9 , 89 . In the pre-hospital setting, patients with AHF should benefit from adequate non-invasive monitoring (that is, continuous electrocardiography and measurement of blood pressure and peripheral oxygen saturation (SpO 2 )), oxygen supplementation in case of hypoxia (SpO 2 <90%) or non-invasive ventilation in case of respiratory distress. Preclinical non-invasive ventilation treatment can reduce intubation rates and improve short-term outcome in patients with cardiogenic pulmonary oedema 106 . When the clinical diagnosis of AHF is straightforward, intravenous treatment (mostly vasodilators and/or diuretics) based on the clinical phenotype and involved pathophysiology should be delivered without waiting for additional testing. Diuretics are mainly used in the presence of fluid retention, whereas vasodilators are administered to reduce filling pressures and modulate ventricular–vascular coupling in the presence of fluid redistribution and preserved systolic blood pressure (>110 mmHg; caution should be used if the systolic blood pressure is 90–110 mmHg) 1 , 3 . The use of vasodilators is recommended by current guidelines 1 , 3 . However, in light of the new results of randomized clinical trials (such as RELAX-AHF-2, TRUE-AHF and GALACTIC) showing no prognostic benefit of vasodilatory agents in AHF, these recommendations may change. The use of inotropes should be restricted to patients in cardiogenic shock due to impaired myocardial contractility, as their inappropriate use is associated with arrhythmias, increased morbidity and mortality 107 . Notably, pre-hospital treatment should not delay rapid transfer to hospital, preferably to a site with a cardiology and cardiac care unit and/or an intensive care unit. Upon arrival at the hospital, patients should be triaged to exclude cardiopulmonary instability (that is, cardiogenic shock and respiratory failure) and undergo a detailed clinical evaluation.

In-hospital management

Individuals with AHF are at risk of death not only from cardiovascular failure but also from the consequences of organ dysfunction due to congestion and hypoperfusion; thus, it is imperative that the treatment strategy addresses both these issues. Despite the fact that there is little evidence from randomized controlled trials that tackling congestion improves survival, the effect of diuretics on symptoms and organ congestion are evident. Once oxygen saturation has been restored (with oxygen supplementation, non-invasive ventilation or mechanical ventilation), the initial treatment goals in patients presenting with AHF consist of achieving decongestion without residual fluid retention, optimizing perfusion pressures to preserve organ perfusion and maintaining or initiating disease-modifying oral therapies directed towards neurohumoral activation, as these medications also increase diuretic response and improve long-term survival 108 , 109 (Fig.  2 )

figure 2

Congestion is assessed on the basis of the presence of compatible clinical signs (for example, pulmonary rales, distended jugular veins and peripheral oedema), evidence of organ congestion on chest X-ray radiography or lung ultrasonography and elevated filling pressures on invasive monitoring. Abnormal peripheral perfusion is assessed on the basis of the presence of compatible clinical signs (for example, cold and clammy skin, oliguria and altered mental status) and other evidence of altered oxygen transport (for example, increased blood lactate and low central venous or mixed venous oxygen saturation). The response to fluid challenge (that is, change in cardiac output after administration of 250–500 ml of fluids), positive inotropic agents (that is, intravenous drugs that increase cardiac contractility) and vasopressors (that is, intravenous drugs that increase arterial blood pressure by causing peripheral vasoconstriction) should be closely assessed by measuring changes in stroke volume, either by echocardiography or by other invasive monitoring systems. HFpEF, heart failure with preserved ejection fraction; HFrEF, heart failure with reduced ejection fraction; MCS, mechanical circulatory support; SBP, systolic blood pressure.

Decongestive therapy

As patients with AHF present with a similar congestion profile irrespective of their LVEF 87 , the decongestive therapy is similar in patients with HFrEF or HFpEF 1 . The decongestive treatment should be tailored according to the haemodynamic phenotype and the underlying pathophysiology and administered (intravenously, to overcome reduced enteral absorption owing to gastrointestinal congestion) as soon as possible after presentation to increase its success. The practical approach to diuretic treatment has been extensively described in a consensus statement of the Heart Failure Association of the European Society of Cardiology 75 . Because loop diuretics are >90% protein-bound by albumin in the blood and need to be secreted into the proximal convoluted tubule through several organic anion transporters, when renal blood flow is reduced (such as in AHF), diuretic dosing needs to be adjusted to achieve a plasma concentration sufficient to obtain the desired effect. Furthermore, the peak effect of intravenous loop diuretics occurs within the first hours, with sodium excretion returning to baseline by 6–8 hours; however, to maintain the decongestive effect, the administration of diuretics should continue until euvolaemia is achieved, with three or four daily doses or continuous infusion.

The diuretic response may be evaluated by measuring the urinary volume output and spot urinary sodium content within the first hours after loop diuretic administration 75 . The measurement of spot urinary sodium content is particularly useful in patients with low to medium urine output. Whereas in patients producing high urinary volumes natriuresis is almost universally high, more-recent data indicate that in patients with a low to medium urine output, spot urinary sodium content offers independent prognostic information in addition to urinary volume output 110 . In patients with congestion, an hourly urine output of <100–150 ml during the first 6 hours and/or a spot urinary sodium content of <50–70 mmol 2 hours after loop diuretic administration generally indicates an inadequate response to diuretics 75 . Early evaluation of the diuretic response is recommended to identify patients with diuretic resistance, enabling rapid intensification (such as doubling) of the loop diuretic dose to attain the ceiling (maximum) dose quickly. As increasing the loop diuretic dose any further than the ceiling dose does not induce incremental diuresis and/or natriuresis, the addition of another diuretic agent with a different mode of action should be considered (sequential nephron blockade). In refractory forms, renal replacement therapy may be considered, although these technologies — despite being very effective in volume removal — have not been shown to improve outcomes 111 , 112 , 113 . Mechanisms and treatment approaches to diuretic resistance have been extensively reviewed elsewhere 114 .

Decongestive treatments should be continued until euvolaemia has been achieved and the medications are switched to an oral form. Loop diuretic therapy should then be reduced to the lowest dose that can maintain euvolaemia 1 , 115 . The quantification of fluid excess and the determination of euvolaemia may be challenging in clinical practice and may require a multimodal approach including symptoms, clinical signs, imaging (such as echocardiography, chest X-ray radiography and lung ultrasonography) and biomarkers 75 . Other techniques, such as data from implanted cardiac devices, pulmonary artery pressure sensors, bioelectrical impedance analysis and indicator dilution techniques, may provide additional valuable information, but their widespread use is limited by technical reasons and cost.

Comprehensive therapy

Specific treatments for the underlying cardiac disease and the precipitating factors should be implemented during hospitalization. For example, myocardial revascularization and optimal antimicrobial treatment should not be delayed when AHF is precipitated by myocardial ischaemia or infection, respectively. On the basis of the comorbidities identified during the initial evaluation and treatment, clinicians should be able to anticipate the need for particular drugs for some specific forms of HF (for example, HF associated with amyloidosis), surgical procedures (for example, for valvular heart disease), mechanical circulatory support (such as LV assist device) or cardiac transplantation. Finally, enrolment of patients in a comprehensive multidisciplinary HF care management programme, promoting medication adherence, up-titration of disease-modifying therapy, cardiac rehabilitation, treatment of underlying comorbidities and timely follow-up with the health-care team, is essential 1 .

Long-term management

Management goals and pre-discharge management.

Individuals who survive the first episode of AHF are at increased risk of experiencing another episode 12 . Thus, the management goals include improving survival and reducing the risk of hospital readmission due to subsequent episodes of AHF. Ensuring that the individual’s condition is sufficiently stabilized for a safe hospital discharge is the central element of pre-discharge management. Patients with AHF are considered ready for discharge after achieving adequate decongestion and stable renal function on guideline-directed oral therapy 1 . Congestion is the most common cause of AHF readmission, and persistent congestion and renal dysfunction are known markers of a poor post-discharge prognosis 69 . A variety of clinical markers (such as weight and fluid loss) and biochemical markers (such as natriuretic peptides) are used as proxies of congestion, but because HF decompensation can occur owing to both fluid accumulation and redistribution, these biomarkers cannot be applied uniformly across patients with AHF. Several studies have demonstrated the usefulness of natriuretic peptides and cardiac troponins in predicting the risk of death and readmission for HF 116 , 117 , 118 . Patients with AHF who have markedly elevated pre-discharge natriuretic peptide levels have worse clinical outcomes, including all-cause and cardiovascular mortality and morbidity, than patients with lower levels. However, the benefits of achieving specific natriuretic peptide target values prior to discharge have not been demonstrated. Abnormally elevated cardiac troponins are often detected in patients with AHF in the absence of overt myocardial ischaemia and are similarly associated with poor outcomes 116 , 117 . Another biomarker of myocardial fibrosis, soluble ST2 receptor (also known as IL-1 receptor-like 1, a protein involved in the process of myocardial fibrosis and hypertrophy) has been correlated with disease severity and a poor prognosis in patients with AHF 119 . ST2, along with other biomarkers of oxidative stress, inflammation and remodelling, requires further study and remains in preclinical exploration 120 . Overall, defining and achieving satisfactory decongestion remains the major hurdle in AHF management.

In addition to achieving adequate decongestion, implementation of the medical treatment of precipitating factors is recommended to improve post-discharge outcome. In patients with HFrEF, disease-modifying oral HF therapy according to HF guidelines (consisting of β-adrenergic receptor blockers, angiotensin-converting enzyme inhibitors or angiotensin receptor–neprilysin inhibitors, and mineralocorticoid receptor antagonists) should be continued or started during hospitalization and gradually titrated thereafter 1 , as it is associated with improved outcomes 105 . In patients with HFpEF, optimal control of comorbidities and precipitating factors is recommended 1 . Additional treatments, including appropriate drugs for some specific forms of HF or surgical procedures, should be evaluated during hospitalization.

Finally, pillars of pre-discharge management include ensuring a deliberate transition to outpatient care and creating a plan to assess and improve post-discharge prognosis. Care coordination for patients with HF is highly complex as clinicians, patients, care-givers and ancillary services must collaborate to titrate pharmacological therapy, monitor fluid volume status and electrolytes, treat comorbidities, initiate lifestyle changes and establish plans for adherence to treatment and emergency care 1 , 120 . Conversations regarding illness severity, barriers to self-care and advance care planning should be introduced before discharge.

Post-discharge management

In addition to continued supervised medical therapy, post-discharge management should incorporate efforts to improve symptoms and quality of life (QOL), delay disease progression and attempt to triage and prognosticate using a risk assessment framework to prevent hospital readmission and death. Generally, post-discharge prognostic tools are prediction models that take several patient clinical variables (for example, age, vital signs during hospitalization, laboratory data and comorbidities) into account and relate them to 30-day and 1-year mortality. Regardless of the time period considered, patients with AHF remain at persistently high risk of rehospitalization and death 121 . Thus, the American College of Cardiology Foundation–American Heart Association guideline for the management of HF recommends the first post-discharge telephone contact within 3 days and a follow-up visit 7–14 days after discharge, and the European Society of Cardiology guidelines recommend the first follow-up outpatient visit within 7 days of discharge 1 , 120 . Despite the complexity of factors associated with rehospitalization for HF, the readmission rate is a ubiquitous metric used to elucidate patient factors (as mentioned above) and health-care system factors that contribute to HF-related morbidity and mortality. Such health-care system factors include, for example, the quality of care provided, patient education, transitional support and medication reconciliation (that is, ensuring that the list of all medications a patient is taking is always as accurate and up-to-date as possible, to facilitate adjustments to the therapy whenever the patient is admitted to, or transferred or discharged from, a hospital). The public health and financial burdens of HF readmissions continue to grow, and evidence is surfacing that some national health policies, for example the Hospital Readmissions Reduction Program in the USA, which were intended to reduce these readmissions, may have had the unintended consequence of increasing post-discharge mortality 122 .

Clinicians should attempt to identify patients with AHF at high risk of readmission by incorporating clinical, laboratory, imaging and haemodynamic data into a comprehensive assessment. Concerning clinical characteristics in the post-discharge phase include multiple comorbidities (for example, chronic obstructive pulmonary disease, anaemia and chronic renal disease), low systolic blood pressure, high heart rate, progressive orthopnoea and jugular vein distension; laboratory parameters that should raise concerns include low serum sodium, elevated blood urea nitrogen and serum creatinine, low serum albumin and elevated natriuretic peptides 123 , 124 , 125 . In addition to traditional echocardiographic parameters used to evaluate biventricular filling pressures, other imaging techniques, such as lung ultrasonography and point-of-care ultrasonographic assessment of right internal jugular vein compliance, have shown promise in prediction of AHF rehospitalization in patients admitted with AHF 126 , 127 . Clinicians should prioritize a comprehensive clinical assessment of patients with AHF with close surveillance for these hallmarks of decompensation and perform targeted interventions focused on decongestion and patient education in the vulnerable early post-discharge phase 128 .

Implantable pulmonary artery pressure sensors to monitor the haemodynamic status and guide therapy can reduce the risk HF-related hospitalization in patients with HFrEF and HFpEF, but questions regarding true device efficacy remain, owing to concerns about potential bias and misconduct during trial execution 129 , 130 , 131 , 132 , 133 . Remote care using intrathoracic impedance monitoring has been associated with an increased risk of HF-related hospitalization 134 . Thus, the 2016 European Society of Cardiology guidelines provide a weak recommendation for the use of wireless implantable haemodynamic monitoring systems in patients with HF to reduce the risk of recurrent HF hospitalization 1 . Ultimately, prevention of readmission after an AHF hospitalization remains a challenge. Reliable identification of high-risk patients and of effective interventions to reduce the risk of rehospitalization has been elusive, as high-quality studies in representative patient cohorts are still needed.

Innovative care delivery models are being increasingly investigated as tools to improve post-discharge outcomes in patients with HF; however, results thus far have been disappointing. Telemonitoring alone did not reduce HF readmission in large multicentre and multinational trials 135 , 136 , 137 , 138 . Patient-centred transitional care approaches that include structured education, communication, clinical care and close surveillance did not improve outcomes compared with usual care models 139 . Questions remain regarding whether the use of these techniques alone can benefit certain subpopulations of patients and whether proving their efficacy will require a combination of patient-centred strategies.

Quality of life

Patients with AHF and chronic HF cope with numerous physical and psychological symptoms that adversely affect their QOL. Dyspnoea, fatigue, dry mouth, orthopnoea, sleep disturbance and difficulty concentrating are highly prevalent, distressing and burdensome and are predictive of reduced QOL in this population 140 (Fig.  3 ). Depression is more common among patients with HF than in the general population, with at least 20% of patients with HF meeting criteria for major depression 141 . Prevalence estimates of depression in the HF population vary widely, ranging from 9% to 60%, and such variation is thought to be largely due to differences in outcome ascertainment methods (that is, interviews versus self-reported questionnaires) and in HF severity at the time of assessment 141 , 142 . Patients with HF with more severe depression have increased health-care utilization, rehospitalization rates and mortality 141 , 143 , 144 , 145 . For clinicians, differentiating between symptoms due to HF and those due to depression can be challenging, highlighting a crucial need for a pragmatic and standardized approach to QOL assessment in routine clinical care.

figure 3

Physical and psychological symptoms that contribute to impaired quality of life in patients with acute heart failure (AHF).

In addition to the physiological alterations in patients with AHF, the stressors of the acute care environment can exacerbate physical and psychological impairments and lead to further declines in QOL 146 . Elderly hospitalized patients with AHF have a markedly higher symptom burden and worse QOL than age-matched cohorts with stable HFpEF and stable HFrEF 146 , 147 . For example, in a prospective, comprehensive, multicentre and multidimensional assessment of 27 patients of ≥60 years of age hospitalized with ADHF compared with three age-matched ambulatory cohorts with stable HF, 78% of the ADHF cohort had cognitive impairment and 30% had depressed mood, but only 11% had a previous diagnosis of depression, suggesting substantial under-recognition of depression in this population. In a sex-stratified analysis of several large international studies on chronic HF, disproportionately worse disease-specific and general QOL was observed in women than in men 148 . This sex-related difference was unexplained — possible hypotheses included differences in the perception of the effect of the disease between women and men and sex-related confounders that were not measured in this study (for example, access to health care, socioeconomic and educational factors, level of care-giver support, living alone or with other people and proactive help-seeking behaviour). In a global study of patients with LVEF <40% hospitalized with AHF, 13% of patients reported persistently unfavourable QOL, defined by Kansas City Cardiomyopathy Questionnaire (KCCQ) scores of <45, at 1 and 24 weeks after hospital discharge 149 . QOL issues also affect patient adherence to pharmacological and non-pharmacological treatment and place extraordinary stress on care-givers. Although many studies examining the QOL of patients with chronic stable HF have been published, there is a notable dearth of evidence regarding QOL in patients with AHF.

Similarly, interventions aimed at improving QOL in patients with AHF are not well studied. Guideline-directed medical therapy decreases symptoms and improves QOL in patients with HFrEF. Non-pharmacological and non-device-based or surgical strategies, such as multidisciplinary team management, exercise training, self-care education and lifestyle modifications, have been examined more rigorously in ambulatory patients with chronic HF, but have not been effective in improving QOL independently 1 . In a small single-centre study of hospitalized patients with HF, inpatient palliative care consultation was associated with improved symptom burden, depressive symptoms and QOL for up to 3 months after hospitalization 150 . Patient-centred outcomes such as QOL are increasingly incorporated into HF trials and recognized as predictors of clinical events. Further research into tools to assess and strategies to improve QOL in the AHF population should be prioritized as the global population of patients with HF continues to grow.

The development of new, effective interventions for the treatment of AHF has been unsuccessful since the 1990s. In contrast to substantial progress achieved in other fields of cardiology and oncology, for example, no new medication or device has been approved for AHF treatment. Many therapies have been tested in the setting of AHF, including inotropic agents (for example, levosimendan and omecamtiv mecarbil), vasodilators (for example, nesiritide, ularitide and serelaxin) and diuretics (for example, tolvaptan) 151 , 152 , 153 , 154 , but the results of these studies were neutral, and it is still unclear whether this neutrality was due to the inactivity of the tested drugs or inadequacy of the study designs. For instance, determining the best time to administer a tested drug is still a challenge. Few studies have assessed early end points and seem to indicate the use of effective agents as early as possible. On the one hand, if drugs that improve cardiac function are given as early as possible (for example, within 6 hours of presentation to the emergency department), they might prevent worsening of organ dysfunction and death. On the other hand, mortality in the first hours and days is related to severe and irreversible alteration in organ function, that is, excess congestion, hypoxia and/or hypoperfusion, and drugs that aim to improve heart function might not prevent death. Hence, studies have suggested that tested HF drugs should be administered within 48 hours of presentation. Furthermore, choosing the most appropriate primary end point also remains a challenge. For years, regulatory agencies sought ‘improvement in survival rate’ as the primary efficacy end point in both patients with AHF and patients with chronic HF, although intravenous drugs tested in patients with AHF were usually administered for 48 hours only, whereas oral therapy was given every day for years in patients with chronic HF. Because no drug has been shown to improve the survival rate in patients with AHF, experts and patient associations are asking to designate improvements in morbidity as the primary efficacy end point and mortality as a safety end point rather than a primary one.

Several new medications are being tested in AHF. These drugs act by modulating endothelial cell function via the adrenomedullin pathway (adrenomedullin is involved in the maintenance of the endothelial barrier function and in the regulation of the renin–angiotensin–aldosterone system and may have protective properties against fluid retention in AHF) 155 or improving cardiovascular function via the modulation of intracellular enzymes, such as dipeptidyl-peptidase 3 (a cytosolic enzyme involved in angiotensin II and enkephalin cleavage that has myocardial depressant properties and whose inhibition may improve haemodynamics) 156 , 157 , that are released into the circulation during cell necrosis. While these studies are ongoing, two challenges remain in the management of AHF. First, the implementation of disease-modifying oral HF therapy in patients with HFrEF is still a major challenge worldwide. Only a minority of patients receive the right classes and the right doses of angiotensin-converting enzyme inhibitors, β-adrenergic receptor blockers and mineralocorticoid receptor antagonists. Achieving this goal will certainly minimize episodes of AHF. The second challenge is the post-discharge medication for patients with AHF with HFpEF. Except for treating cardiovascular and metabolic comorbidities that are very frequent in these patients, no drug is recommended after discharge to prevent readmission for a new episode of acute dyspnoea.

Circulating biomarkers, such as natriuretic peptides, are increasingly used in the treatment of patients with AHF. However, during the acute episode, they indicate myocardial stretch but neither venous nor whole-body congestion. Furthermore, although observational studies have shown that a rapid decrease in natriuretic peptides levels is associated with improved outcomes, a recent trial showed no benefit from intensifying therapy to achieve low levels of natriuretic peptides 158 . Thus, a multimarker strategy based on serially evaluated biomarkers, such as natriuretic peptides, high-sensitivity cardiac troponins, soluble ST2, growth differentiation factor 15, cystatin-C, galectin-3 and high-sensitivity C-reactive protein, may provide increased prognostic accuracy and risk prediction but requires further investigation in different cohorts of patients with HF 159 . This multimarker strategy might identify high-risk patients who may benefit from novel therapies.

QOL is the main issue for individuals who survive an episode of AHF. Readmissions for dyspnoea are frequent in the months and years following an AHF episode, in particular if the patient does not have optimal doses of disease-modifying HF therapies and does not receive the appropriate devices when needed. Thus, patients seem to favour a rapid improvement in QOL, measured as the number of days out of hospital after discharge, rather than an improvement in survival rate with a bad QOL.

Basic and translational research is also needed to decipher mechanisms of decompensation in chronic HFrEF and HFpEF. AHF is associated with stimulation of the neuroendocrine system and worsening in congestion that harms many organs, including the lungs, kidney and liver. Studies need to elucidate the mechanisms that lead to organ dysfunctions in AHF to prevent worsening in organ function during AHF episodes.

In summary, AHF is a very frequent event that affects the QOL and survival in patients with chronic HFrEF or HFpEF. Signs and symptoms are often related to congestion and in a few patients to hypoperfusion. Mechanisms of decompensation are still unknown. The administration of symptomatic and causal treatments is recommended. Optimizing disease-modifying HF therapies as early as possible is probably the most effective way to prevent AHF episodes. Further research to decipher mechanisms of cardiac and neuroendocrine decompensation and to identify new treatments is needed.

Ponikowski, P. et al. 2016 ESC guidelines for the diagnosis and treatment of acute and chronic heart failure. Eur. Heart J. 37 , 2129–2200 (2016). The guidelines of the European Society of Cardiology provide evidence-based recommendations on diagnosis and treatment of chronic HF and AHF .

Article   PubMed   Google Scholar  

Braunwald, E. Heart failure. JACC Heart Fail. 1 , 1–20 (2013).

Mebazaa, A. et al. Recommendations on pre-hospital and early hospital management of acute heart failure: a consensus paper from the Heart Failure Association of the European Society of Cardiology, the European Society of Emergency Medicine and the Society of Academic Emergency Medicine–short version. Eur. Heart J. 36 , 1958–1966 (2015). This consensus document contains contemporary recommendations endorsed by different professional societies on the management of AHF .

Ambrosy, A. P. et al. The global health and economic burden of hospitalizations for heart failure: lessons learned from hospitalized heart failure registries. J. Am. Coll. Cardiol. 63 , 1123–1133 (2014).

Crespo-Leiro, M. G. et al. European Society of Cardiology Heart Failure long-term registry (ESC-HF-LT): 1-year follow-up outcomes and differences across regions. Eur. J. Heart Fail. 18 , 613–625 (2016).

Mentz, R. J. & O’Connor, C. M. Pathophysiology and clinical evaluation of acute heart failure. Nat. Rev. Cardiol. 13 , 28–35 (2015).

Article   PubMed   CAS   Google Scholar  

Ishihara, S. et al. Similar hemodynamic decongestion with vasodilators and inotropes: systematic review, meta-analysis, and meta-regression of 35 studies on acute heart failure. Clin. Res. Cardiol. 105 , 971–980 (2016).

Mebazaa, A. et al. Short-term survival by treatment among patients hospitalized with acute heart failure: the global ALARM-HF registry using propensity scoring methods. Intensive Care Med. 37 , 290–301 (2010).

Mebazaa, A. et al. Acute heart failure and cardiogenic shock: a multidisciplinary practical guidance. Intensive Care Med. 42 , 147–163 (2015).

Mebazaa, A. et al. Management of cardiogenic shock complicating myocardial infarction. Intensive Care Med. 44 , 760–773 (2018).

Follath, F. et al. Clinical presentation, management and outcomes in the acute heart failure global survey of standard treatment (ALARM-HF). Intensive Care Med. 37 , 619–626 (2011).

Article   CAS   PubMed   Google Scholar  

Greene, S. J. et al. The vulnerable phase after hospitalization for heart failure. Nat. Rev. Cardiol. 12 , 220–229 (2015). This paper reviews the topic of the vulnerable phase after an AHF episode, which is characterized by high mortality and readmission rates .

GBD 2017 Causes of Death Collaborators. Global, regional, and national age-sex-specific mortality for 282 causes of death in 195 countries and territories, 1980–2017: a systematic analysis for the Global Burden of Disease Study 2017. Lancet 392 , 1736–1788 (2018).

Article   Google Scholar  

Vos, T. et al. Years lived with disability (YLDs) for 1160 sequelae of 289 diseases and injuries 1990–2010: a systematic analysis for the Global Burden of Disease Study 2010. Lancet 380 , 2163–2196 (2012).

Article   PubMed   PubMed Central   Google Scholar  

Setoguchi, S., Stevenson, L. W. & Schneeweiss, S. Repeated hospitalizations predict mortality in the community population with heart failure. Am. Heart J. 154 , 260–266 (2007).

Gheorghiade, M. et al. Acute heart failure syndromes: current state and framework for future research. Circulation 112 , 3958–3968 (2005).

Hamo, C. E. et al. A critical appraisal of short-term endpoints in acute heart failure clinical trials. J. Card. Fail. 24 , 783–792 (2018).

Cleland, J. et al. The EuroHeart Failure survey programme—a survey on the quality of care among patients with heart failure in Europe: Part 1: patient characteristics and diagnosis. Eur. Heart J. 24 , 442–463 (2003).

Cuffe, M. S. et al. Short-term intravenous milrinone for acute exacerbation of chronic heart failure: a randomized controlled trial. JAMA 287 , 1541–1547 (2002).

Dharmarajan, K. et al. Diagnoses and timing of 30-day readmissions after hospitalization for heart failure, acute myocardial infarction, or pneumonia. JAMA 309 , 355–363 (2013).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Fudim, M. et al. Aetiology, timing and clinical predictors of early vs. late readmission following index hospitalization for acute heart failure: insights from ASCEND–HF. Eur. J. Heart Fail. 20 , 304–314 (2018).

Giamouzis, G. et al. Hospitalization epidemic in patients with heart failure: risk factors, risk prediction, knowledge gaps, and future directions. J. Card. Fail. 17 , 54–75 (2011).

Sokoreli, I. et al. Prognostic value of psychosocial factors for first and recurrent hospitalizations and mortality in heart failure patients: insights from the OPERA–HF study. Eur. J. Heart Fail. 20 , 689–696 (2018).

Ferreira, J. P. et al. World Heart Federation Roadmap for heart failure. Glob. Heart 14 , 197–214 (2019).

Dokainish, H. et al. Global mortality variations in patients with heart failure: results from the International Congestive Heart Failure (INTER-CHF) prospective cohort study. Lancet Glob. Health 5 , e665–e672 (2017). This study provides epidemiological data on mortality associated with HF across different continents .

Khatibzadeh, S., Farzadfar, F., Oliver, J., Ezzati, M. & Moran, A. Worldwide risk factors for heart failure: a systematic review and pooled analysis. Int. J. Cardiol. 168 , 1186–1194 (2013).

Damasceno, A. et al. The causes, treatment, and outcome of acute heart failure in 1006 Africans from 9 countries. Arch. Intern. Med. 172 , 1386–1394 (2012).

Parada, H., Carrasco, H. A., Añez, N., Fuenmayor, C. & Inglessis, I. Cardiac involvement is a constant finding in acute Chagas’ disease: a clinical, parasitological and histopathological study. Int. J. Cardiol. 60 , 49–54 (1997).

Bocchi, E. A. et al. Long-term prospective, randomized, controlled study using repetitive education at six-month intervals and monitoring for adherence in heart failure outpatients: the REMADHE trial. Circ. Heart Fail. 1 , 115–124 (2008).

Doval, H. C. et al. Randomised trial of low-dose amiodarone in severe congestive heart failure. Grupo de Estudio de la Sobrevida en la Insuficiencia Cardiaca en Argentina (GESICA). Lancet 344 , 493–498 (1994).

Zühlke, L. et al. Clinical outcomes in 3343 children and adults with rheumatic heart disease from 14 low- and middle-income countries: two-year follow-up of the global rheumatic heart disease registry (the REMEDY Study). Circulation 134 , 1456–1466 (2016).

Sliwa, K. et al. Incidence and characteristics of newly diagnosed rheumatic heart disease in urban African adults: insights from the Heart of Soweto study. Eur. Heart J. 31 , 719–727 (2010).

Bauersachs, J. et al. Pathophysiology, diagnosis and management of peripartum cardiomyopathy: a position statement from the Heart Failure Association of the European Society of Cardiology Study Group on peripartum cardiomyopathy. Eur. J. Heart Fail. 21 , 827–843 (2019).

Abraham, W. T. et al. Predictors of in-hospital mortality in patients hospitalized for heart failure: insights from the Organized Program to Initiate Lifesaving Treatment in Hospitalized Patients with Heart Failure (OPTIMIZE-HF). J. Am. Coll. Cardiol. 52 , 347–356 (2008).

Sliwa, K. et al. Readmission and death after an acute heart failure event: predictors and outcomes in sub-Saharan Africa: results from the THESUS-HF registry. Eur. Heart J. 34 , 3151–3159 (2013).

Arrigo, M., Parissis, J. T., Akiyama, E. & Mebazaa, A. Understanding acute heart failure: pathophysiology and diagnosis. Eur. Heart J. Suppl. 18 , G11–G18 (2016). This review summarizes the pathophysiology and the diagnostic process in AHF with special attention to treatment-relevant aspects .

Shah, A. M. Ventricular remodeling in heart failure with preserved ejection fraction. Curr. Heart Fail. Rep. 10 , 341–349 (2013).

Wang, M. & Shah, A. M. Age-associated pro-inflammatory remodeling and functional phenotype in the heart and large arteries. J. Mol. Cell. Cardiol. 83 , 101–111 (2015).

Zile, M. R. et al. Transition from chronic compensated to acute decompensated heart failure: pathophysiological insights obtained from continuous monitoring of intracardiac pressures. Circulation 118 , 1433–1441 (2008). This study provides unique insights into the pathophysiology of AHF using data obtained from continuous monitoring of intracardiac pressures .

Miller, W. L. & Mullan, B. P. Understanding the heterogeneity in volume overload and fluid distribution in decompensated heart failure is key to optimal volume management: role for blood volume quantitation. JACC Heart Fail. 2 , 298–305 (2014).

Kaye, D. M. et al. Neurochemical evidence of cardiac sympathetic activation and increased central nervous system norepinephrine turnover in severe congestive heart failure. J. Am. Coll. Cardiol. 23 , 570–578 (1994).

Fallick, C., Sobotka, P. A. & Dunlap, M. E. Sympathetically mediated changes in capacitance: redistribution of the venous reservoir as a cause of decompensation. Circ. Heart Fail. 4 , 669–675 (2011).

Chaudhry, S. I., Wang, Y., Concato, J., Gill, T. M. & Krumholz, H. M. Patterns of weight change preceding hospitalization for heart failure. Circulation 116 , 1549–1554 (2007).

Nijst, P. et al. The pathophysiological role of interstitial sodium in heart failure. J. Am. Coll. Cardiol. 65 , 378–388 (2015).

Titze, J. et al. Glycosaminoglycan polymerization may enable osmotically inactive Na+ storage in the skin. Am. J. Physiol. Heart Circ. Physiol. 287 , H203–H208 (2004).

Guyton, A. C. Interstitial fluid presure. II. Pressure-volume curves of interstitial space. Circ. Res. 16 , 452–460 (1965).

Hartupee, J. & Mann, D. L. Neurohormonal activation in heart failure with reduced ejection fraction. Nat. Rev. Cardiol. 14 , 30–38 (2017).

Mullens, W., Verbrugge, F. H., Nijst, P. & Tang, W. H. W. Renal sodium avidity in heart failure: from pathophysiology to treatment strategies. Eur. Heart J. 38 , 1872–1882 (2017).

McKie, P. M. et al. Impaired natriuretic and renal endocrine response to acute volume expansion in pre-clinical systolic and diastolic dysfunction. J. Am. Coll. Cardiol. 58 , 2095–2103 (2011).

Mullens, W. & Tang, W. H. W. The early intertwining of the heart and the kidney through an impaired natriuretic response to acute volume expansion. J. Am. Coll. Cardiol. 58 , 2104–2105 (2011).

Mullens, W. et al. Importance of venous congestion for worsening of renal function in advanced decompensated heart failure. J. Am. Coll. Cardiol. 53 , 589–596 (2009). This study highlights the impact of congestion (instead of reduced cardiac output) on renal function in patients with AHF .

Damman, K. et al. Increased central venous pressure is associated with impaired renal function and mortality in a broad spectrum of patients with cardiovascular disease. J. Am. Coll. Cardiol. 53 , 582–588 (2009).

Nijst, P., Martens, P., Dupont, M., Tang, W. H. W. & Mullens, W. Intrarenal flow alterations during transition from euvolemia to intravascular volume expansion in heart failure patients. JACC Heart Fail. 5 , 672–681 (2017).

Cotter, G., Metra, M., Milo-Cotter, O., Dittrich, H. C. & Gheorghiade, M. Fluid overload in acute heart failure – re-distribution and other mechanisms beyond fluid accumulation. Eur. J. Heart Fail. 10 , 165–169 (2014).

Greenway, C. V. & Lister, G. E. Capacitance effects and blood reservoir function in the splanchnic vascular bed during non-hypotensive haemorrhage and blood volume expansion in anaesthetized cats. J. Physiol. 237 , 279–294 (1974).

Verbrugge, F. H. et al. Abdominal contributions to cardiorenal dysfunction in congestive heart failure. J. Am. Coll. Cardiol. 62 , 485–495 (2013).

Fonarow, G. C. et al. Factors identified as precipitating hospital admissions for heart failure and clinical outcomes: findings from OPTIMIZE-HF. Arch. Intern. Med. 168 , 847–854 (2008).

Arrigo, M. et al. Precipitating factors and 90-day outcome of acute heart failure: a report from the intercontinental GREAT registry. Eur. J. Heart Fail. 19 , 201–208 (2017). This study shows the impact of precipitating factors on prognosis of patients with AHF in a large intercontinental registry .

Arrigo, M. et al. Effect of precipitating factors of acute heart failure on readmission and long-term mortality. ESC. Heart Fail. 3 , 115–121 (2016).

Platz, E. et al. Prevalence and prognostic importance of precipitating factors leading to heart failure hospitalization: recurrent hospitalizations and mortality. Eur. J. Heart Fail. 20 , 295–303 (2018).

Miró, Ò. et al. Time-pattern of adverse outcomes after an infection-triggered acute heart failure decompensation and the influence of early antibiotic administration and hospitalisation: results of the PAPRICA-3 study. Clin. Res. Cardiol. 109 , 34–45 (2020).

Parrinello, G. et al. Water and sodium in heart failure: a spotlight on congestion. Heart Fail. Rev. 20 , 13–24 (2015).

Volpe, M., Carnovali, M. & Mastromarino, V. The natriuretic peptides system in the pathophysiology of heart failure: from molecular basis to treatment. Clin. Sci. 130 , 57–77 (2016).

Article   CAS   Google Scholar  

MacIver, D. H., Adeniran, I., MacIver, I. R., Revell, A. & Zhang, H. Physiological mechanisms of pulmonary hypertension. Am. Heart J. 180 , 1–11 (2016).

Borné, Y. et al. Vascular endothelial growth factor D, pulmonary congestion, and incidence of heart failure. J. Am. Coll. Cardiol. 71 , 580–582 (2018).

Houston, B. A. et al. Relation of lymphangiogenic factor vascular endothelial growth factor-D to elevated pulmonary artery wedge pressure. Am. J. Cardiol. 124 , 756–762 (2019).

von Moos, S. et al. Vascular endothelial growth factor D is a biomarker of fluid overload in haemodialysis patients. Nephrol. Dial. Transplant. https://doi.org/10.1093/ndt/gfz281 (2020).

Ware, L. B. & Matthay, M. A. Clinical practice. Acute pulmonary edema. N. Engl. J. Med. 353 , 2788–2796 (2005).

Harjola, V.-P. et al. Organ dysfunction, injury and failure in acute heart failure: from pathophysiology to diagnosis and management. A review on behalf of the Acute Heart Failure Committee of the Heart Failure Association (HFA) of the European Society of Cardiology (ESC). Eur. J. Heart Fail. 19 , 821–836 (2017). This review summarizes pathophysiology, diagnosis and management of organ dysfunction occurring in patients with AHF .

Braam, B., Cupples, W. A., Joles, J. A. & Gaillard, C. Systemic arterial and venous determinants of renal hemodynamics in congestive heart failure. Heart Fail. Rev. 17 , 161–175 (2012).

Legrand, M., Mebazaa, A., Ronco, C. & Januzzi, J. L. When cardiac failure, kidney dysfunction, and kidney injury intersect in acute conditions: the case of cardiorenal syndrome. Crit. Care Med. 42 , 2109–2117 (2014).

Mullens, W. et al. Elevated intra-abdominal pressure in acute decompensated heart failure. J. Am. Coll. Cardiol. 51 , 300–306 (2008).

Ahmad, T. et al. Worsening renal function in patients with acute heart failure undergoing aggressive diuresis is not associated with tubular injury. Circulation 137 , 2016–2028 (2018).

Maisel, A. S. et al. Neutrophil gelatinase-associated lipocalin for acute kidney injury during acute heart failure hospitalizations: the AKINESIS study. J. Am. Coll. Cardiol. 68 , 1420–1431 (2016).

Mullens, W. 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. 21 , 137–155 (2019). This position statement from the Heart Failure Association of the European Society of Cardiology provides practical guidance on the use of diuretics to relieve congestion in patients with AHF .

Metra, M. et al. Is worsening renal function an ominous prognostic sign in patients with acute heart failure? The role of congestion and its interaction with renal function. Circ. Heart Fail. 5 , 54–62 (2012).

Auer, J. What does the liver tell us about the failing heart? Eur. Heart J. 34 , 711–714 (2013).

Møller, S. & Bernardi, M. Interactions of the heart and the liver. Eur. Heart J. 34 , 2804–2811 (2013).

Samsky, M. D. et al. Cardiohepatic interactions in heart failure: an overview and clinical implications. J. Am. Coll. Cardiol. 61 , 2397–2405 (2013).

Rogler, G. & Rosano, G. The heart and the gut. Eur. Heart J. 35 , 426–430 (2014).

Sandek, A. et al. Altered intestinal function in patients with chronic heart failure. J. Am. Coll. Cardiol. 50 , 1561–1569 (2007).

Valentova, M. et al. Intestinal congestion and right ventricular dysfunction: a link with appetite loss, inflammation, and cachexia in chronic heart failure. Eur. Heart J. 37 , 1684–1691 (2016).

Colombo, P. C. et al. Peripheral venous congestion causes inflammation, neurohormonal, and endothelial cell activation. Eur. Heart J. 35 , 448–454 (2014).

Colombo, P. C. et al. Venous congestion, endothelial and neurohormonal activation in acute decompensated heart failure: cause or effect? Curr. Heart Fail. Rep. 12 , 215–222 (2015).

McMurray, J. J. V. et al. Angiotensin–neprilysin inhibition versus enalapril in heart failure. N. Engl. J. Med. 371 , 993–1004 (2014).

McMurray, J. J. V. et al. Dapagliflozin in patients with heart failure and reduced ejection fraction. N. Engl. J. Med. 381 , 1995–2008 (2019).

Van Aelst, L. N. L. et al. Acutely decompensated heart failure with preserved and reduced ejection fraction present with comparable haemodynamic congestion. Eur. J. Heart Fail. 20 , 738–747 (2018).

Van de Werf, F. et al. Diastolic properties of the left ventricle in normal adults and in patients with third heart sounds. Circulation 69 , 1070–1078 (1984).

Mebazaa, A. et al. Recommendations on pre-hospital & early hospital management of acute heart failure: a consensus paper from the Heart Failure Association of the European Society of Cardiology, the European Society of Emergency Medicine and the Society of Academic Emergency Medicine. Eur. J. Heart Fail. 17 , 544–558 (2015).

Ponikowski, P. et al. 2016 ESC guidelines for the diagnosis and treatment of acute and chronic heart failure. Eur. J. Heart Fail. 18 , 891–975 (2016).

Fonarow, G. C. et al. Risk stratification for in-hospital mortality in acutely decompensated heart failure: classification and regression tree analysis. JAMA 293 , 572–580 (2005). This study provides one easy-to-use risk score (among many others available) to stratify patients admitted with AHF .

Peterson, P. N. et al. A validated risk score for in-hospital mortality in patients with heart failure from the American Heart Association Get With the Guidelines program. Circ. Cardiovasc. Qual. Outcomes 3 , 25–32 (2010).

Miró, Ò. et al. Predicting 30-day mortality for patients with acute heart failure in the emergency department: a cohort study. Ann. Intern. Med. 167 , 698–705 (2017).

Gheorghiade, M. et al. Assessing and grading congestion in acute heart failure: a scientific statement from the Acute Heart Failure Committee of the Heart Failure Association of the European Society of Cardiology and endorsed by the European Society of Intensive Care Medicine. Eur. J. Heart Fail. 12 , 423–433 (2010). This position statement from the Heart Failure Association of the European Society of Cardiology provides a statement on the assessment and quantification of congestion in AHF .

Maisel, A. S. et al. Rapid measurement of B-type natriuretic peptide in the emergency diagnosis of heart failure. N. Engl. J. Med. 347 , 161–167 (2002).

Januzzi, J. L. Jr et al. The N-terminal Pro-BNP Investigation of Dyspnea in the Emergency Department (PRIDE) study. Am. J. Cardiol. 95 , 948–954 (2005).

Maisel, A. et al. Mid-region pro-hormone markers for diagnosis and prognosis in acute dyspnea. J. Am. Coll. Cardiol. 55 , 2062–2076 (2010).

McCullough, P. A. et al. B-type natriuretic peptide and clinical judgment in emergency diagnosis of heart failure: analysis from Breathing Not Properly (BNP) multinational study. Circulation 106 , 416–422 (2002).

Arrigo, M., Nijst, P. & Rudiger, A. Optimising heart failure therapies in the acute setting. Card. Fail. Rev. 4 , 38–42 (2018).

Platz, E. et al. Detection and prognostic value of pulmonary congestion by lung ultrasound in ambulatory heart failure patients. Eur. Heart J. 37 , 1244–1251 (2016).

Aras, M. A. & Teerlink, J. R. Lung ultrasound: a ‘B-line’ to the prediction of decompensated heart failure. Eur. Heart J. 37 , 1252–1254 (2016).

Matsue, Y. et al. Time-to-furosemide treatment and mortality in patients hospitalized with acute heart failure. J. Am. Coll. Cardiol. 69 , 3042–3051 (2017). This study shows a beneficial association between early administration of decongestive treatment (loop diuretics) and mortality in AHF .

Ledwidge, M. et al. Natriuretic peptide-based screening and collaborative care for heart failure: the STOP-HF randomized trial. JAMA 310 , 66–74 (2013).

Rubio-Gracia, J. et al. Prevalence, predictors and clinical outcome of residual congestion in acute decompensated heart failure. Int. J. Cardiol. 258 , 185–191 (2018).

Gayat, E. et al. Heart failure oral therapies at discharge are associated with better outcome in acute heart failure: a propensity-score matched study. Eur. J. Heart Fail. 20 , 345–354 (2018). This study shows in a propensity-matched cohort a positive association between pre-discharge implementation of neuro-humoral blockers (β-adrenergic receptor blockers and renin–angiotensin inhibitors) and survival in patients with AHF .

Plaisance, P., Pirracchio, R., Berton, C., Vicaut, E. & Payen, D. A randomized study of out-of-hospital continuous positive airway pressure for acute cardiogenic pulmonary oedema: physiological and clinical effects. Eur. Heart J. 28 , 2895–2901 (2007).

Arrigo, M. & Mebazaa, A. Understanding the differences among inotropes. Intensive Care Med. 41 , 912–915 (2015).

Butler, J., Gheorghiade, M. & Metra, M. Moving away from symptoms-based heart failure treatment: misperceptions and real risks for patients with heart failure. Eur. J. Heart Fail. 18 , 350–352 (2016).

Kula, A. J. et al. Influence of titration of neurohormonal antagonists and blood pressure reduction on renal function and decongestion in decompensated heart failure. Circ. Heart Fail. 9 , e002333 (2016).

Brinkley, D. M. et al. Spot urine sodium as triage for effective diuretic infusion in an ambulatory heart failure unit. J. Card. Fail. 24 , 349–354 (2018).

Costanzo, M. R. et al. Extracorporeal ultrafiltration for fluid overload in heart failure. J. Am. Coll. Cardiol. 69 , 2428–2445 (2017).

Costanzo, M. R. et al. Ultrafiltration versus intravenous diuretics for patients hospitalized for acute decompensated heart failure. J. Am. Coll. Cardiol. 49 , 675–683 (2007).

Bart, B. A. et al. Ultrafiltration in decompensated heart failure with cardiorenal syndrome. N. Engl. J. Med. 367 , 2296–2304 (2012).

ter Maaten, J. M. et al. Diuretic response in acute heart failure–pathophysiology, evaluation and therapy. Nat. Rev. Cardiol. 12 , 184–192 (2015). This article reviews the mechanisms and the treatment of diuretic resistance occurring in AHF .

Martens, P., Nijst, P. & Mullens, W. Current approach to decongestive therapy in acute heart failure. Curr. Heart Fail. Rep. 12 , 367–378 (2015).

Yancy, C. W. et al. 2017 ACC/AHA/HFSA focused update of the 2013 ACCF/AHA guideline for the management of heart failure: a report of the American College of Cardiology/American Heart Association Task Force on Clinical Practice Guidelines and the Heart Failure Society of America. Circulation 136 , e137–e161 (2017).

Peacock, W. F. et al. Cardiac troponin and outcome in acute heart failure. N. Engl. J. Med. 358 , 2117–2126 (2008).

Logeart, D. et al. Predischarge B-type natriuretic peptide assay for identifying patients at high risk of re-admission after decompensated heart failure. J. Am. Coll. Cardiol. 43 , 635–641 (2004).

Rehman, S. U., Mueller, T. & Januzzi, J. L. Characteristics of the novel interleukin family biomarker ST2 in patients with acute heart failure. J. Am. Coll. Cardiol. 52 , 1458–1465 (2008).

Yancy, C. W. et al. 2013 ACCF/AHA guideline for the management of heart failure: executive summary: a report of the American College of Cardiology Foundation/American Heart Association Task Force on practice guidelines. J. Am. Coll. Cardiol. 62 , 1495–1539 (2013).

Passantino, A., Monitillo, F., Iacoviello, M. & Scrutinio, D. Predicting mortality in patients with acute heart failure: role of risk scores. World J. Cardiol. 7 , 902–911 (2015).

Wadhera, R. K. et al. Association of the Hospital Readmissions Reduction Program with mortality among Medicare beneficiaries hospitalized for heart failure, acute myocardial infarction, and pneumonia. JAMA 320 , 2542–2552 (2018).

Gheorghiade, M. et al. A comprehensive, longitudinal description of the in-hospital and post-discharge clinical, laboratory, and neurohormonal course of patients with heart failure who die or are re-hospitalized within 90 days: analysis from the EVEREST trial. Heart Fail. Rev. 17 , 485–509 (2012).

Grodin, J. L. et al. Prognostic implications of changes in amino-terminal Pro-B-type natriuretic peptide in acute decompensated heart failure: insights from ASCEND-HF. J. Card. Fail. 25 , 703–711 (2019).

Mueller, C. et al. Heart Failure Association of the European Society of Cardiology practical guidance on the use of natriuretic peptide concentrations. Eur. J. Heart Fail. 21 , 715–731 (2019).

Simon, M. A., Schnatz, R. G., Romeo, J. D. & Pacella, J. J. Bedside ultrasound assessment of jugular venous compliance as a potential point-of-care method to predict acute decompensated heart failure 30-day readmission. J. Am. Heart Assoc. 7 , e008184 (2018).

Platz, E. et al. Lung ultrasound in acute heart failure: prevalence of pulmonary congestion and short- and long-term outcomes. JACC Heart Fail. 7 , 849–858 (2019).

Palazzuoli, A. et al. Early readmission for heart failure: an avoidable or ineluctable debacle? Int. J. Cardiol. 277 , 186–195 (2019).

Abraham, W. T. et al. Wireless pulmonary artery haemodynamic monitoring in chronic heart failure: a randomised controlled trial. Lancet 377 , 658–666 (2011).

Adamson, P. B. et al. Wireless pulmonary artery pressure monitoring guides management to reduce decompensation in heart failure with preserved ejection fraction. Circ. Heart Fail. 7 , 935–944 (2014).

Givertz, M. M. et al. Pulmonary artery pressure-guided management of patients with heart failure and reduced ejection fraction. J. Am. Coll. Cardiol. 70 , 1875–1886 (2017).

Krahnke, J. S. et al. Heart failure and respiratory hospitalizations are reduced in patients with heart failure and chronic obstructive pulmonary disease with the use of an implantable pulmonary artery pressure monitoring device. J. Card. Fail. 21 , 240–249 (2015).

Loh, J. P., Barbash, I. M. & Waksman, R. Overview of the 2011 Food and Drug Administration Circulatory System Devices Panel of the Medical Devices Advisory Committee Meeting on the CardioMEMS Champion Heart Failure Monitoring System. J. Am. Coll. Cardiol. 61 , 1571–1576 (2013).

Van Veldhuisen, D. J. et al. Intrathoracic impedance monitoring, audible patient alerts, and outcome in patients with heart failure. Circulation 124 , 1719–1726 (2011).

Chaudhry, S. I. et al. Telemonitoring in patients with heart failure. N. Engl. J. Med. 363 , 2301–2309 (2010).

Jayaram, N. M. et al. Impact of telemonitoring on health status. Circ. Cardiovasc. Qual. Outcomes 10 , e004148 (2017).

Koehler, F. et al. Telemedical Interventional Monitoring in Heart Failure (TIM-HF), a randomized, controlled intervention trial investigating the impact of telemedicine on mortality in ambulatory patients with heart failure: study design. Eur. J. Heart Fail. 12 , 1354–1362 (2010).

Cleland, J. G. F. et al. Noninvasive home telemonitoring for patients with heart failure at high risk of recurrent admission and death: the Trans-European Network-Home-Care Management System (TEN-HMS) study. J. Am. Coll. Cardiol. 45 , 1654–1664 (2005).

Van Spall, H. G. C. et al. Effect of patient-centered transitional care services on clinical outcomes in patients hospitalized for heart failure: the PACT-HF randomized clinical trial. JAMA 321 , 753–761 (2019).

Zambroski, C. H., Moser, D. K., Bhat, G. & Ziegler, C. Impact of symptom prevalence and symptom burden on quality of life in patients with heart failure. Eur. J. Cardiovasc. Nurs. 4 , 198–206 (2005).

Rutledge, T., Reis, V. A., Linke, S. E., Greenberg, B. H. & Mills, P. J. Depression in heart failure: a meta-analytic review of prevalence, intervention effects, and associations with clinical outcomes. J. Am. Coll. Cardiol. 48 , 1527–1537 (2006).

Freedland, K. E. et al. Prevalence of depression in hospitalized patients with congestive heart failure. Psychosom. Med. 65 , 119–128 (2003).

Adelborg, K. et al. Mortality risk among heart failure patients with depression: a nationwide population-based cohort study. J. Am. Heart Assoc. 5 , e004137 (2016).

Sullivan, M., Simon, G., Spertus, J. & Russo, J. Depression-related costs in heart failure care. Arch. Intern. Med. 162 , 1860–1866 (2002).

O’Connor, C. M. & Joynt, K. E. Depression: are we ignoring an important comorbidity in heart failure? J. Am. Coll. Cardiol. 43 , 1550–1552 (2004).

Reeves, G. R. et al. Comparison of frequency of frailty and severely impaired physical function in patients ≥60 years hospitalized with acute decompensated heart failure versus chronic stable heart failure with reduced and preserved left ventricular ejection fraction. Am. J. Cardiol. 117 , 1953–1958 (2016).

Warraich, H. J. et al. Physical function, frailty, cognition, depression, and quality of life in hospitalized adults ≥60 years with acute decompensated heart failure with preserved versus reduced ejection fraction: insights from the REHAB-HF trial. Circ. Heart Fail. 11 , e005254 (2018).

PubMed   PubMed Central   Google Scholar  

Dewan, P. et al. Differential impact of heart failure with reduced ejection fraction on men and women. J. Am. Coll. Cardiol. 73 , 29–40 (2019).

Allen, L. A. et al. Identifying patients hospitalized with heart failure at risk for unfavorable future quality of life. Circ. Cardiovasc. Qual. Outcomes 4 , 389–398 (2011).

Sidebottom, A. C., Jorgenson, A., Richards, H., Kirven, J. & Sillah, A. Inpatient palliative care for patients with acute heart failure: outcomes from a randomized trial. J. Palliat. Med. 18 , 134–142 (2015).

Mebazaa, A. et al. Levosimendan vs dobutamine for patients with acute decompensated heart failure: the SURVIVE randomized trial. JAMA 297 , 1883–1891 (2007).

Teerlink, J. R. et al. Acute treatment with omecamtiv mecarbil to increase contractility in acute heart failure. J. Am. Coll. Cardiol. 67 , 1444–1455 (2016).

Packer, M. et al. Effect of ularitide on cardiovascular mortality in acute heart failure. N. Engl. J. Med. 376 , 1956–1964 (2017).

Konstam, M. A. et al. Effects of oral tolvaptan in patients hospitalized for worsening heart failure: the EVEREST outcome trial. JAMA 297 , 1319–1331 (2007).

Voors, A. A. et al. Adrenomedullin in heart failure: pathophysiology and therapeutic application. Eur. J. Heart Fail. 21 , 163–171 (2019).

Deniau, B. et al. Circulating dipeptidyl peptidase 3 is a myocardial depressant factor: dipeptidyl peptidase 3 inhibition rapidly and sustainably improves haemodynamics. Eur. J. Heart Fail. 14 , e0220866 (2019).

Google Scholar  

Takagi, K. et al. Circulating dipeptidyl peptidase 3 and alteration in haemodynamics in cardiogenic shock: results from the OptimaCC trial. Eur. J. Heart Fail. https://doi.org/10.1002/ejhf.1600 (2019).

Troughton, R., Michael Felker, G. & Januzzi, J. L. Natriuretic peptide-guided heart failure management. Eur. Heart J. 35 , 16–24 (2014).

Demissei, B. G. et al. A multimarker multi-time point-based risk stratification strategy in acute heart failure: results from the RELAX-AHF trial. Eur. J. Heart Fail. 19 , 1001–1010 (2017).

Download references

Acknowledgements

N.R. is supported by the National Institutes of Health, National Human Genome Research Institute, Ruth L. Kirschstein Institutional National Research Service T32 Award in Genomic Medicine (T32 HG009495).

Author information

Authors and affiliations.

Department of Cardiology, University Hospital Zurich, Zurich, Switzerland

Mattia Arrigo

Division of Cardiovascular Medicine, Department of Medicine, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA, USA

Mariell Jessup & Nosheen Reza

Ziekenhuis Oost Limburg, Genk, Belgium

Wilfried Mullens

University of Hasselt, Hasselt, Belgium

School of Cardiovascular Medicine & Sciences, King’s College London British Heart Foundation Centre, London, UK

Ajay M. Shah

Hatter Institute for Cardiovascular Research in Africa, Faculty of Health Sciences, Department of Medicine and Cardiology, University of Cape Town, Cape Town, South Africa

Karen Sliwa

Université de Paris, MASCOT, Inserm, Paris, France

Alexandre Mebazaa

Department of Anesthesia, Burn and Critical Care Medicine, AP-HP, Hôpital Lariboisière, Paris, France

You can also search for this author in PubMed   Google Scholar

Contributions

Introduction (A.M. and M.A.); Epidemiology (K.S.); Mechanisms/pathophysiology (M.A., W.M. and A.M.S.); Diagnosis, screening and prevention (A.M. and M.A.); Management (M.J., W.M. and N.R.); Quality of life (M.J. and N.R.); Outlook (A.M.); Overview of Primer (A.M.).

Corresponding author

Correspondence to Alexandre Mebazaa .

Ethics declarations

Competing interests.

All authors declare no competing interests.

Additional information

Peer review information.

Nature Reviews Disease Primers thanks A. Palazzuoli and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Reprints and permissions

About this article

Cite this article.

Arrigo, M., Jessup, M., Mullens, W. et al. Acute heart failure. Nat Rev Dis Primers 6 , 16 (2020). https://doi.org/10.1038/s41572-020-0151-7

Download citation

Accepted : 28 January 2020

Published : 05 March 2020

DOI : https://doi.org/10.1038/s41572-020-0151-7

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Establishment of in-hospital nutrition support program for middle-aged and elderly patients with acute decompendated heart failure.

  • Yongliang Li
  • Xiangdong Xu

BMC Cardiovascular Disorders (2024)

Risk factors and prognosis of perioperative acute heart failure in elderly patients with hip fracture: case-control studies and cohort study

  • Mingming Fu
  • Zhiyong Hou

BMC Musculoskeletal Disorders (2024)

Diagnostic and prognostic value of plasma miR-106a-5p levels in patients with acute heart failure

Journal of Cardiothoracic Surgery (2024)

Associations between static and dynamic changes of platelet counts and in-hospital mortality in critical patients with acute heart failure

Scientific Reports (2024)

The management of heart failure cardiogenic shock: an international RAND appropriateness panel

  • Stefan Williams
  • Antonis Kalakoutas
  • Alastair G. Proudfoot

Critical Care (2024)

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

clinical presentation of acute heart failure

brand logo

MICHAEL KING, MD, JOE KINGERY, DO, AND BARETTA CASEY, MD, MPH

Am Fam Physician. 2012;85(12):1161-1168

More recent articles on heart failure and cardiomyopathy are available.

Patient information: A handout on this topic is available at https://familydoctor.org/familydoctor/en/diseases-conditions/heart-failure.html .

Author disclosure: No relevant financial affiliations to disclose.

Heart failure is a common clinical syndrome characterized by dyspnea, fatigue, and signs of volume overload, which may include peripheral edema and pulmonary rales. Heart failure has high morbidity and mortality rates, especially in older persons. Many conditions, such as coronary artery disease, hypertension, valvular heart disease, and diabetes mellitus, can cause or lead to decompensation of chronic heart failure. Up to 40 to 50 percent of patients with heart failure have diastolic heart failure with preserved left ventricular function, and the overall mortality is similar to that of systolic heart failure. The initial evaluation includes a history and physical examination, chest radiography, electrocardiography, and laboratory assessment to identify causes or precipitating factors. A displaced cardiac apex, a third heart sound, and chest radiography findings of venous congestion or interstitial edema are useful in identifying heart failure. Systolic heart failure is unlikely when the Framingham criteria are not met or when B-type natriuretic peptide level is normal. Echocardiography is the diagnostic standard to confirm systolic or diastolic heart failure through assessment of left ventricular ejection fraction. Evaluation for ischemic heart disease is warranted in patients with heart failure, especially if angina is present, given that coronary artery disease is the most common cause of heart failure.

Heart failure is a common clinical syndrome characterized by dyspnea, fatigue, and signs of volume overload, which may include peripheral edema and pulmonary rales. There is no single diagnostic test for heart failure; therefore, it remains a clinical diagnosis requiring a history, physical examination, and laboratory testing. Symptoms of heart failure can be caused by systolic or diastolic dysfunction. Appropriate diagnosis and therapy for heart failure are important given the poor prognosis. Survival is 89.6 percent at one month from diagnosis, 78 percent at one year, and only 57.7 percent at five years. 1

Heart failure has an estimated overall prevalence of 2.6 percent. 2 It is becoming more common in adults older than 65 years because of increased survival after acute myocardial infarction and improved treatment of coronary artery disease (CAD), valvular disease, and hypertension.

The initial evaluation of patients with suspected heart failure should include a history and physical examination, laboratory assessment, chest radiography, and electrocardiography. Echocardiography can confirm the diagnosis.C
A displaced cardiac apex, a third heart sound, and chest radiography findings of pulmonary venous congestion or interstitial edema are good predictors to rule in the diagnosis of heart failure.C ,
Systolic heart failure can be effectively ruled out with a normal B-type natriuretic peptide or N-terminal pro–B-type natriuretic peptide level.C , , ,
Systolic heart failure can be effectively ruled out when the Framingham criteria are not met.C ,

Heart failure is defined by the American Heart Association and American College of Cardiology as “a complex clinical syndrome that can result from any structural or functional cardiac disorder that impairs the ability of the ventricle to fill with or eject blood.” 3 As cardiac output decreases because of stresses placed on the myocardium, activation of the sympathetic nervous and renin-angiotensin-aldosterone systems increases blood pressure (for tissue perfusion) and blood volume (enhancing preload, stroke volume, and cardiac output by the Frank-Starling mechanism). These compensatory mechanisms can also lead to further myocardial deterioration and worsening myocardial contractility. In systolic heart failure, cardiac output is decreased directly through reduced left ventricular function. In diastolic heart failure, cardiac output is compromised by poor ventricular compliance, impaired relaxation, and worsened end-diastolic pressure. 3 , 4

CAD is the underlying etiology in up to 60 to 70 percent of patients with systolic heart failure, 5 , 6 and a predictor for progression from asymptomatic to symptomatic left ventricular systolic dysfunction. Hypertension and valvular heart disease are significant risk factors for heart failure, with relative risks of 1.4 and 1.46, respectively. 6 Diabetes mellitus increases the risk of heart failure twofold by directly leading to cardiomyopathy and significantly contributing to CAD. Diabetes is one of the strongest risk factors for heart failure in women with CAD. 7 Smoking, physical inactivity, obesity, and lower socioeconomic status are often overlooked risk factors. 6 Numerous conditions can cause heart failure, either acutely without an underlying cardiac disorder or through decompensation of chronic heart failure ( Table 1 ) . 3 , 4 , 8 As a result, alternative causes should be promptly recognized, treated, and monitored to determine if the heart failure is reversible. 8

Common
Coronary artery disease
Hypertension
Idiopathic cardiomyopathy
Valvular heart disease
Less common
Arrhythmia (e.g., tachycardia, bradycardia, heart block)
Collagen vascular disease (e.g., systemic lupus erythematosus, scleroderma)
Endocrine/metabolic disorders (e.g., thyroid disease, diabetes mellitus, pheochromocytoma, other genetic disorders)
Hypertrophic cardiomyopathy
Myocarditis
Pericarditis
Postpartum cardiomyopathy
Restrictive cardiomyopathies (e.g., amyloidosis, hemochromatosis, sarcoidosis, other genetic disorders)
Toxic cardiomyopathy (e.g., alcohol, cocaine, radiation)
Anemia
Atrial fibrillation or other arrhythmias
Fluid overload (e.g., salt intake, water intake, medication compliance)
Fluid retention from drugs (e.g., chemotherapy, cyclooxygenase 1 and 2 inhibitors, excessive licorice, glitazones, glucocorticoids, androgens, estrogens)
Hyper- or hypothyroid disease
Pulmonary causes (e.g., cor pulmonale, pulmonary hypertension, pulmonary embolism)
Renal causes (e.g., renal failure, nephrotic syndrome, glomerulonephritis)
Sleep apnea
Systemic infection or septic shock

Classification

The most important consideration when categorizing heart failure is whether left ventricular ejection fraction (LVEF) is preserved or reduced (less than 50 percent). 3 , 8 A reduced LVEF in systolic heart failure is a powerful predictor of mortality. 9 As many as 40 to 50 percent of patients with heart failure have diastolic heart failure with preserved left ventricular function. 2 , 10 – 16 Overall, there is no difference in survival between diastolic and systolic heart failure that cannot be attributed to ejection fraction. 2 , 10 – 16 Patients with diastolic heart failure are more likely to be women, to be older, and to have hypertension, atrial fibrillation, and left ventricular hypertrophy, but no history of CAD. 11 – 14 , 17 , 18 Compared with systolic heart failure, which has well-validated therapies, diastolic heart failure lacks evidence-based treatment recommendations. 3 , 8 , 13

Heart failure symptoms can occur with preserved or reduced ejection fraction, (systolic or diastolic heart failure). The New York Heart Association classification system is the simplest and most widely used method to gauge symptom severity ( Table 2 ) . 19 The classification system is a well-established predictor of mortality and can be used at diagnosis and to monitor treatment response.

INo limitations of physical activity
No heart failure symptoms
IIMild limitation of physical activity
Heart failure symptoms with significant exertion; comfortable at rest or with mild activity
IIIMarked limitation of physical activity
Heart failure symptoms with mild exertion; only comfortable at rest
IVDiscomfort with any activity
Heart failure symptoms occur at rest

Initial Clinical Evaluation

Although no single item on clinical history, sign, or symptom has been proven to be diagnostic, many are helpful in assessing the probability of heart failure. The initial clinical evaluation, detailed in Tables 1 , 3 , 4 , 8 3 , 3 , 8 , 20 and 4 , 3 , 8 , 20 is directed at confirming heart failure, determining potential causes, and identifying comorbid illnesses. Table 5 lists findings for the initial evaluation of suspected heart failure, including history, physical examination, chest radiography, electrocardiography, and B-type natriuretic peptide (BNP) testing. 17 , 21 – 23 Evaluation for ischemic heart disease is warranted in patients with heart failure, especially if angina is present, given that CAD is the most common cause of heart failure.

Symptoms
Abdominal swelling
Dyspnea on exertion
Edema
Exercise intolerance
Fatigue
Orthopnea
Paroxysmal nocturnal dyspnea
Recent weight gain
Physical examination findings
Abdomen: hepatojugular reflux, ascites
Extremities: cool, dependent edema
Heart: bradycardia/tachycardia, laterally displaced point of maximal impulse, third heart sound (gallop or murmur)
Lungs: labored breathing, rales
Neck: elevated jugular venous pressure
Skin: cyanosis, pallor
Symptoms
Abdominal swelling (liver failure)
Anorexia, weight loss (sarcoidosis)
Chest pain (coronary artery disease)
Claudication (atherosclerotic disease)
Cough (pulmonary disease)
Diarrhea or skin lesions (amyloidosis)
Dyspnea on exertion (pulmonary disease, valvular disease)
Edema (liver or kidney failure)
Neurologic problems (sarcoidosis)
Palpitations (tachyarrhythmia)
Recent fevers, viral infection (endocarditis, myocarditis, infection)
Syncope (bradycardia, heart block)
Physical examination findings
Abdomen: distended, hepatosplenomegaly, tender, ascites (liver disease)
Extremities: joint inflammation/warmth (rheumatologic disease)
Heart: irregular rate or rhythm (arrhythmia)
Lungs: wheezing (pulmonary disease)
Neck: thyromegaly/nodule (thyroid disease)
Skin: cyanosis (anemia), jaundice (liver failure)
B-type natriuretic peptide level
Calcium and magnesium levels (diuretics, cause of arrhythmia)
Complete blood count (anemia)
Liver function (hepatic congestion, volume overload)
Renal function (renal causes)
Serum electrolyte level (electrolyte imbalance)
Thyroid-stimulating hormone level (thyroid disorders)
Urinalysis (renal causes)
Arterial blood gases (hypoxia, pulmonary disease)
Blood cultures (endocarditis, systemic infection)
Human immunodeficiency virus (cardiomyopathy)
Lyme serology (bradycardia/heart block)
Serum ferritin level, transferrin saturation (macrocytic anemia, hemochromatosis)
Thiamine level (deficiency, beriberi, alcoholism)
Troponin and creatine kinase-MB levels (myocardial infarction, myocardial injury)
A1C level (diabetes mellitus)
Lipid profile (hyperlipidemia)
Displaced cardiac apex 160.95
Third heart sound110.99
Chest radiography: interstitial edema120.97
Chest radiography: venous congestion120.96





History of heart failure5.80.90
Hepatojugular reflex6.40.96
Jugular venous distension5.10.92





Framingham criteria for systolic heart failure4.570.79
Framingham criteria for heart failure4.350.79
Framingham criteria for diastolic heart failure4.210.79
Initial clinical judgment4.40.86
History of myocardial infarction3.10.87
Rales (crackles)2.80.78
Murmur2.60.90
Paroxysmal nocturnal dyspnea2.60.84
Peripheral edema2.30.78
Orthopnea2.20.77
Elevated BNP level2.920.66
Elevated N-terminal pro-BNP level2.670.65
Chest radiography: cardiomegaly3.30.78
Chest radiography: pleural effusion3.20.92
ECG: atrial fibrillation3.80.93
ECG: new T-wave change3.00.92
ECG: any abnormality




2.20.78

History and Physical Examination

Patients with heart failure can have decreased exercise tolerance with dyspnea, fatigue, generalized weakness, and fluid retention, with peripheral or abdominal swelling and possibly orthopnea. 3 Patient history and physical examination are useful to evaluate for alternative or reversible causes ( Table 1 ) . 3 , 4 , 8 Nearly all patients with heart failure have dyspnea on exertion. However, heart failure accounts for only 30 percent of the causes of dyspnea in the primary care setting. 24 The absence of dyspnea on exertion only slightly decreases the probability of systolic heart failure, and the presence of orthopnea or paroxysmal nocturnal dyspnea has a small effect in increasing the probability of heart failure (positive likelihood ratio [LR+] = 2.2 and 2.6). 21 , 23

The presence of a third heart sound (ventricular filling gallop) is an indication of increased left ventricular end-diastolic pressure and a decreased LVEF. Despite being relatively uncommon findings, a third heart sound and displaced cardiac apex are good predictors of left ventricular dysfunction and effectively rule in the diagnosis of systolic heart failure (LR+ = 11 and 16). 21 , 23

The presence of jugular venous distention, hepatojugular reflux, pulmonary rales, and pitting peripheral edema is indicative of volume overload and enhances the probability of a heart failure diagnosis. Jugular venous distention and hepatojugular reflex have a moderate effect (LR+ = 5.1 and 6.4), whereas the others, along with cardiac murmurs, have only a small effect on the diagnostic probability (LR+ = 2.3 to 2.8). The absence of any of these findings is of little help in ruling out heart failure. 21

Laboratory Tests

Laboratory testing can help identify alternative and potentially reversible causes of heart failure. Table 4 lists laboratory tests appropriate for the initial evaluation of heart failure and other potential causes. 3 , 8 , 20 Other laboratory tests should be performed based on physician discretion to evaluate further causes or identify comorbid conditions that require enhanced control.

BNP and N-terminal pro-BNP (the cleaved inactive N-terminal fragment of the BNP precursor) levels can be used to evaluate patients with dyspnea for heart failure. BNP is secreted by the atria and ventricles in response to stretching or increased wall tension. 25 BNP levels increase with age, are higher in women and blacks, and can be elevated in patients with renal failure. 21 , 26 BNP appears to have better reliability than N-terminal pro-BNP, especially in older populations. 25 , 26 Multiple systematic reviews have concluded that BNP and N-terminal pro-BNP levels can effectively rule out a diagnosis of heart failure 22 , 25 , 27 , 28 because of their negative predictive value (negative likelihood ratio [LR–] = 0.1 and 0.14). 22 The average cutoff levels for heart failure were a BNP level of 95 pg per mL (95 ng per L) or a N-terminal pro-BNP level of 642 pg per mL (642 ng per L). 22

As BNP levels increase, the specificity increases and thus the likelihood of a heart failure diagnosis. 25 BNP levels are strong predictors of mortality at two to three months and cardiovascular events in acute heart failure, specifically when BNP level is greater than 200 pg per mL (200 ng per L) or N-terminal pro-BNP level is greater than 5,180 pg per mL (5,180 ng per L). 22 , 25 Limited evidence supports monitoring reduction of BNP levels in the acute and outpatient settings. A 30 to 50 percent reduction in BNP level at hospital discharge showed improved survival and reduced rehospitalization rates. Optimizing management for outpatient targets of a BNP level less than 100 pg per mL (100 ng per L) and an N-terminal pro-BNP level less than 1,700 pg per mL (1,700 ng per L) showed improvement in decompensations, hospitalizations, and mortality events. 22 , 25

Chest Radiography

Chest radiography should be performed initially to evaluate for heart failure because it can identify pulmonary causes of dyspnea (e.g., pneumonia, pneumothorax, mass). Pulmonary venous congestion and interstitial edema on chest radiography in a patient with dyspnea make the diagnosis of heart failure more likely (LR+ = 12). Other findings, such as pleural effusion or cardiomegaly, may slightly increase the likelihood of heart failure (LR+ = 3.2 and 3.3), but their absence is only slightly useful in decreasing the probability of heart failure (LR– = 0.33 to 0.48). 21

Electrocardiography

Electrocardiography (ECG) is useful for identifying other causes in patients with suspected heart failure. Changes such as left bundle branch block, left ventricular hypertrophy, acute or previous myocardial infarction, or atrial fibrillation can be identified and may warrant further investigation by echocardiography, stress testing, or cardiology consultation. Normal findings (or minor abnormalities) on ECG make systolic heart failure only slightly less likely (LR– = 0.27). 23 The presence of other findings such as atrial fibrillation, new T-wave changes, or any abnormality has a small effect on the diagnostic probability of heart failure (LR+ = 2.2 to 3.8). 21

Clinical Decision Making

The definition of heart failure continues to be debated, but it remains a clinical diagnosis. Several groups have published diagnostic criteria, but the Framingham criteria are widely accepted and include the components of the initial evaluation, which enhances their accuracy ( Table 6 ) . 17 A previous study validated the Framingham criteria for diagnosing systolic heart failure, 29 and a more recent study analyzed them for systolic and diastolic heart failure. 17 Both studies reported high sensitivity for systolic heart failure (97 percent compared with 89 percent for diastolic heart failure), which effectively rules out heart failure when the Framingham criteria are not met (LR– = 0.04). 17 , 29 The Framingham criteria only have a small effect on confirming a diagnosis of heart failure (LR+ = 4.21 to 4.57), but have a moderate effect on ruling out heart failure in general and diastolic heart failure (LR– = 0.1 and 0.13). 17

Acute pulmonary edema
Cardiomegaly
Hepatojugular reflex
Neck vein distension
Paroxysmal nocturnal dyspnea or orthopnea
Rales
Third heart sound gallop
Ankle edema
Dyspnea on exertion
Hepatomegaly
Nocturnal cough
Pleural effusion
Tachycardia (> 120 beats per minute)

Echocardiography is the most widely accepted and available method for identifying systolic dysfunction and should be performed after the initial evaluation to confirm the presence of heart failure. 3 Two-dimensional echocardiography with Doppler flow studies can assess LVEF, left ventricular size, wall thickness, valve function, and the pericardium. Echocardiography can assist in diagnosing diastolic heart failure if elevated left atrial pressure, impaired left ventricular relaxation, and decreased compliance are present. 2 , 3 Often, the diagnosis of diastolic heart failure is clinical without conclusive echocardiographic evidence. If echocardiography results are equivocal or inadequate, transesophageal echocardiography, radionuclide angiography, or cineangiography with contrast media (at catheterization) can be used to assess cardiac function. 30

If angina or chest pain is present with heart failure, the American Heart Association and the American College of Cardiology recommend that the patient undergo coronary angiography, unless there is a contraindication to potential revascularization. 3 Coronary angiography has been shown to improve symptoms and survival in patients with angina and reduced ejection fraction. 3 It is important to evaluate for CAD because it is the cause of heart failure and low ejection fraction in approximately two-thirds of patients. 4 , 5 Because wall motion abnormalities are common in nonischemic cardiomyopathy, noninvasive testing may not be adequate for assessing the presence of CAD, and cardiology consultation may be warranted.

Figure 1 is an algorithm for the evaluation and diagnosis of heart failure. When a patient presents with symptoms of heart failure, the initial evaluation is performed to identify alternative or reversible causes of heart failure and to confirm its presence. If the Framingham criteria are not met, or if the BNP level is normal, systolic heart failure is essentially ruled out. Echocardiography should be performed to assess LVEF when heart failure is suspected or if diastolic heart failure is still suspected when systolic heart failure is ruled out. Treatment options are guided by the final diagnosis and echocardiography results, with a consideration to evaluate for CAD.

Data Sources: A PubMed search was completed in Clinical Queries using the following key words in various combinations under the search by clinical study category: heart failure, symptoms, causes, diagnosis, diagnostic criteria, diastolic, systolic, brain natriuretic peptide. The categories searched included etiology, diagnosis, clinical prediction rules, and systematic reviews. The articles consisted of meta-analyses, systematic reviews, randomized controlled trials, and cohort studies. The related citations feature was used to locate similar research once appropriate articles had been discovered. We also searched the Agency for Healthcare Research and Quality Evidence Reports, Bandolier, the Cochrane Database of Systematic Reviews, the Database of Abstracts of Reviews of Effects, the Institute for Clinical Systems Improvement, and the National Guideline Clearinghouse database. Search dates: April 5 through 16, 2010; May 24 through 28, 2010; selected newer articles January 1 and April 20, 2011.

Loehr LR, Rosamond WD, Chang PP, Folsom AR, Chambless LE. Heart failure incidence and survival (from the Atherosclerosis Risk in Communities study). Am J Cardiol. 2008;101(7):1016-1022.

Redfield MM, Jacobsen SJ, Burnett JC, Mahoney DW, Bailey KR, Rodeheffer RJ. Burden of systolic and diastolic ventricular dysfunction in the community: appreciating the scope of the heart failure epidemic. JAMA. 2003;289(2):194-202.

Hunt SA, Abraham WT, Chin MH, et al. 2009 focused update incorporated into the ACC/AHA 2005 guidelines for the diagnosis and management of heart failure in adults: a report of the American College of Cardiology Foundation/American Heart Association Task Force on Practice Guidelines: developed in collaboration with the International Society for Heart and Lung Transplantation [published correction appears in Circulation . 2010;121(12):e258]. Circulation. 2009;119(14):e391-e479.

Dosh SA. Diagnosis of heart failure in adults. Am Fam Physician. 2004;70(11):2145-2152.

Gheorghiade M, Bonow RO. Chronic heart failure in the United States: a manifestation of coronary artery disease. Circulation. 1998;97(3):282-289.

He J, Ogden LG, Bazzano LA, Vupputuri S, Loria C, Whelton PK. Risk factors for congestive heart failure in US men and women: NHANES I epidemiologic follow-up study. Arch Intern Med. 2001;161(7):996-1002.

Bibbins-Domingo K, Lin F, Vittinghoff E, et al. Predictors of heart failure among women with coronary disease. Circulation. 2004;110(11):1424-1430.

Institute for Clinical Systems Improvement (ICSI). Heart failure in adults. Bloomington, Minn.: Institute for Clinical Systems Improvement (ICSI); 2009:95.

Solomon SD, Anavekar N, Skali H, et al.; Candesartan in Heart Failure Reduction in Mortality (CHARM) Investigators. Influence of ejection fraction on cardiovascular outcomes in a broad spectrum of heart failure patients. Circulation. 2005;112(24):3738-3744.

Aurigemma GP. Diastolic heart failure—a common and lethal condition by any name. N Engl J Med. 2006;355(3):308-310.

Lee DS, Gona P, Vasan RS, et al. Relation of disease pathogenesis and risk factors to heart failure with preserved or reduced ejection fraction: insights from the Framingham heart study of the National Heart, Lung, and Blood Institute. Circulation. 2009;119(24):3070-3077.

Bursi F, Weston SA, Redfield MM, et al. Systolic and diastolic heart failure in the community. JAMA. 2006;296(18):2209-2216.

Owan TE, Hodge DO, Herges RM, Jacobsen SJ, Roger VL, Redfield MM. Trends in prevalence and outcome of heart failure with preserved ejection fraction. N Engl J Med. 2006;355(3):251-259.

Bhatia RS, Tu JV, Lee DS, et al. Outcome of heart failure with preserved ejection fraction in a population-based study. N Engl J Med. 2006;355(3):260-269.

Vasan RS, Benjamin EJ, Levy D. Prevalence, clinical features and prognosis of diastolic heart failure: an epidemiologic perspective. J Am Coll Cardiol. 1995;26(7):1565-1574.

Persson H, Lonn E, Edner M, et al. Diastolic dysfunction in heart failure with preserved systolic function: need for objective evidence: results from the CHARM Echocardiographic Substudy-CHARMES. J Am Coll Cardiol. 2007;49(6):687-694.

Maestre A, Gil V, Gallego J, Aznar J, Mora A, Martín-Hidalgo A. Diagnostic accuracy of clinical criteria for identifying systolic and diastolic heart failure: cross-sectional study. J Eval Clin Pract. 2009;15(1):55-61.

Masoudi FA, Havranek EP, Smith G, et al. Gender, age, and heart failure with preserved left ventricular systolic function. J Am Coll Cardiol. 2003;41(2):217-223.

New York Heart Association Criteria Committee. Disease of the Heart and Blood Vessels: Nomenclature and Criteria for Diagnosis . 6th ed. Boston, Mass.: Little, Brown; 1964.

Remme WJ, Swedberg K Task Force for the Diagnosis and Treatment of Chronic Heart Failure, European Society of Cardiology. Guidelines for the diagnosis and treatment of chronic heart failure [published correction appears in Eur Heart J . 2001;22(23):2217–2218]. Eur Heart J. 2001;22(17):1527-1560.

Wang CS, FitzGerald JM, Schulzer M, Mak E, Ayas NT. Does this dyspneic patient in the emergency department have congestive heart failure?. JAMA. 2005;294(15):1944-1956.

Balion C, Santaguida PL, Hill S, et al. Testing for BNP and NT-proBNP in the diagnosis and prognosis of heart failure. Evid Rep Technol Assess (Full Rep). 2006;142:1-147.

Madhok V, Falk G, Rogers A, Struthers AD, Sullivan FM, Fahey T. The accuracy of symptoms, signs and diagnostic tests in the diagnosis of left ventricular dysfunction in primary care: a diagnostic accuracy systematic review. BMC Fam Pract. 2008;9:56.

Mulrow CD, Lucey CR, Farnett LE. Discriminating causes of dyspnea through clinical examination. J Gen Intern Med. 1993;8(7):383-392.

Chen WC, Tran KD, Maisel AS. Biomarkers in heart failure. Heart. 2010;96(4):314-320.

Ewald B, Ewald D, Thakkinstian A, Attia J. Meta-analysis of B type natriuretic peptide and N-terminal pro B natriuretic peptide in the diagnosis of clinical heart failure and population screening for left ventricular systolic dysfunction. Intern Med J. 2008;38(2):101-113.

Battaglia M, Pewsner D, Jüni P, Egger M, Bucher HC, Bachmann LM. Accuracy of B-type natriuretic peptide tests to exclude congestive heart failure: systematic review of test accuracy studies. Arch Intern Med. 2006;166(10):1073-1080.

Latour-Pérez J, Coves-Orts FJ, Abad-Terrado C, Abraira V, Zamora J. Accuracy of B-type natriuretic peptide levels in the diagnosis of left ventricular dysfunction and heart failure: a systematic review. Eur J Heart Fail. 2006;8(4):390-399.

Jimeno Sainz A, Gil V, Merino J, García M, Jordán A, Guerrero L. Validity of Framingham criteria as a clinical test for systolic heart failure [in Spanish]. Rev Clin Esp. 2006;206(10):495-498.

Naik MM, Diamond GA, Pai T, Soffer A, Siegel RJ. Correspondence of left ventricular ejection fraction determinations from two-dimensional echocardiography, radionuclide angiography and contrast cineangiography. J Am Coll Cardiol. 1995;25(4):937-942.

Continue Reading

clinical presentation of acute heart failure

More in AFP

More in pubmed.

Copyright © 2012 by the American Academy of Family Physicians.

This content is owned by the AAFP. A person viewing it online may make one printout of the material and may use that printout only for his or her personal, non-commercial reference. This material may not otherwise be downloaded, copied, printed, stored, transmitted or reproduced in any medium, whether now known or later invented, except as authorized in writing by the AAFP.  See permissions  for copyright questions and/or permission requests.

Copyright © 2024 American Academy of Family Physicians. All Rights Reserved.

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • v.320(7229); 2000 Jan 22

ABC of heart failure

Clinical features and complications, clinical features.

Patients with heart failure present with a variety of symptoms, most of which are non-specific. The common symptoms of congestive heart failure include fatigue, dyspnoea, swollen ankles, and exercise intolerance, or symptoms that relate to the underlying cause. The accuracy of diagnosis by presenting clinical features alone, however, is often inadequate, particularly in women and elderly or obese patients.

Symptoms and signs in heart failure

  • Paroxysmal nocturnal dyspnoea
  • Reduced exercise tolerance, lethargy, fatigue
  • Nocturnal cough
  • Ankle swelling
  • Cachexia and muscle wasting
  • Tachycardia
  • Pulsus alternans
  • Elevated jugular venous pressure
  • Displaced apex beat
  • Right ventricular heave
  • Crepitations or wheeze
  • Third heart sound
  • Hepatomegaly (tender)

Exertional breathlessness is a frequent presenting symptom in heart failure, although it is a common symptom in the general population, particularly in patients with pulmonary disease. Dyspnoea is therefore moderately sensitive, but poorly specific, for the presence of heart failure. Orthopnoea is a more specific symptom, although it has a low sensitivity and therefore has little predictive value. Paroxysmal nocturnal dyspnoea results from increased left ventricular filling pressures (due to nocturnal fluid redistribution and enhanced renal reabsorption) and therefore has a greater sensitivity and predictive value. Nocturnal ischaemic chest pain may also be a manifestation of heart failure, so left ventricular systolic dysfunction should be excluded in patients with recurrent nocturnal angina.

Common causes of lower limb oedema

  • Gravitational disorder—for example, immobility
  • Congestive heart failure
  • Venous thrombosis or obstruction, varicose veins
  • Hypoproteinaemia—for example, nephrotic syndrome, liver disease
  • Lymphatic obstruction

Fatigue and lethargy

Fatigue and lethargy in chronic heart failure are, in part, related to abnormalities in skeletal muscle, with premature muscle lactate release, impaired muscle blood flow, deficient endothelial function, and abnormalities in skeletal muscle structure and function. Reduced cerebral blood flow, when accompanied by abnormal sleep patterns, may occasionally lead to somnolence and confusion in severe chronic heart failure.

Sensitivity, specificity, and predictive value of symptoms, signs, and chest x ray findings for presence of heart failure (ejection fraction <40%) in 1306 patients with coronary artery disease undergoing cardiac catheterisation

History:
 Shortness of breath665223
 Orthopnoea2181 2
 Paroxysmal nocturnal  dyspnoea337626
 History of oedema238022
Examination:
 Tachycardia (>100  beats/min) 799 6
 Crepitations139127
 Oedema (on  examination)1093 3
 Gallop (S3)319561
 Neck vein distension1097 2
Chest ray examination:
 Cardiomegaly626732

Swelling of ankles and feet is another common presenting feature, although there are numerous non-cardiac causes of this symptom. Right heart failure may manifest as oedema, right hypochondrial pain (liver distension), abdominal swelling (ascites), loss of appetite, and, rarely, malabsorption (bowel oedema). An increase in weight may be associated with fluid retention, although cardiac cachexia and weight loss are important markers of disease severity in some patients.

Physical signs

Physical examination has serious limitations as many patients, particularly those with less severe heart failure, have few abnormal signs. In addition, some physical signs are difficult to interpret and, if present, may occasionally be related to causes other than heart failure.

Oedema and a tachycardia, for example, are too insensitive to have any useful predictive value, and although pulmonary crepitations may have a high diagnostic specificity they have a low sensitivity and predictive value. Indeed, the commonest cause of lower limb oedema in elderly people is immobility, and pulmonary crepitations may reflect poor ventilation with infection, or pulmonary fibrosis, rather than heart failure. Jugular venous distension has a high specificity in diagnosing heart failure in patients who are known to have cardiac disease, although some patients, even with documented heart failure, do not have an elevated venous pressure. The presence of a displaced apex beat in a patient with a history of myocardial infarction has a high positive predictive value. A third heart sound has a relatively high specificity, although its universal value is limited by a high interobserver variability, with interobserver agreement of less than 50% in non-specialists.

In patients with pre-existing chronic heart failure, other clinical features may be evident that point towards precipitating causes of acute heart failure or deteriorating heart failure. Common factors that may be obvious on clinical assessment and are associated with relapses in congestive heart failure include infections, arrhythmias, continued or recurrent myocardial ischaemia, and anaemia.

Clinical diagnosis and clinical scoring systems

Several epidemiological studies, including the Framingham heart study, have used clinical scoring systems to define heart failure, although the use of these systems is not recommended for routine clinical practice.

Precipitating causes of heart failure

  • Arrhythmias, especially atrial fibrillation
  • Infections (especially pneumonia)
  • Acute myocardial infarction
  • Angina pectoris or recurrent myocardial ischaemia
  • Alcohol excess
  • Iatrogenic cause—for example, postoperative fluid replacement or administration of steroids or non-steroidal anti-inflammatory drugs
  • Poor drug compliance, especially in antihypertensive treatment
  • Thyroid disorders—for example, thyrotoxicosis
  • Pulmonary embolism

In a patient with appropriate symptoms and a number of physical signs, including a displaced apex beat, elevated venous pressure, oedema, and a third heart sound, the clinical diagnosis of heart failure may be made with some confidence. However, the clinical suspicion of heart failure should also be confirmed with objective investigations and the demonstration of cardiac dysfunction at rest. It is important to note that, in some patients, exercise-induced myocardial ischaemia may lead to a rise in ventricular filling pressures and a fall in cardiac output, leading to symptoms of heart failure during exertion.

Classification

Symptoms and exercise capacity are used to classify the severity of heart failure and monitor the response to treatment. The classification of the New York Heart Association (NYHA) is used widely, although outcome in heart failure is best determined not only by symptoms (NYHA class) but also by echocardiographic criteria. As the disease is progressive, the importance of early treatment, in an attempt to prevent progression to more severe disease, cannot be overemphasised.

European Society of Cardiology's guidelines for diagnosis of heart failure

Essential features.

  • Symptoms of heart failure (for example, breathlessness, fatigue, ankle swelling)
  • Objective evidence of cardiac dysfunction (at rest)

Non-essential features

  • Response to treatment directed towards heart failure (in cases where the diagnosis is in doubt)

Complications

Arrhythmias, atrial fibrillation.

Atrial fibrillation is present in about a third (range 10-50%) of patients with chronic heart failure and may represent either a cause or a consequence of heart failure. The onset of atrial fibrillation with a rapid ventricular response may precipitate overt heart failure, particularly in patients with pre-existing ventricular dysfunction. Predisposing causes should be considered, including mitral valve disease, thyrotoxicosis, and sinus node disease. Importantly, sinus node disease may be associated with bradycardias, which might be exacerbated by antiarrhythmic treatment.

NYHA classification of heart failure

Class i: asymptomatic.

No limitation in physical activity despite presence of heart disease. This can be suspected only if there is a history of heart disease which is confirmed by investigations—for example, echocardiography

Class II: mild

Slight limitation in physical activity. More strenuous activity causes shortness of breath—for example, walking on steep inclines and several flights of steps. Patients in this group can continue to have an almost normal lifestyle and employment

Class III: moderate

More marked limitation of activity which interferes with work. Walking on the flat produces symptoms

Class IV: severe

Unable to carry out any physical activity without symptoms. Patients are breathless at rest and mostly housebound

Atrial fibrillation that occurs with severe left ventricular dysfunction following myocardial infarction is associated with a poor prognosis. In addition, patients with heart failure and atrial fibrillation are at particularly high risk of stroke and other thromboembolic complications.

Ventricular arrhythmias

Malignant ventricular arrhythmias are common in end stage heart failure. For example, sustained monomorphic ventricular tachycardia occurs in up to 10% of patients with advanced heart failure who are referred for cardiac transplantation. In patients with ischaemic heart disease these arrhythmias often have re-entrant mechanisms in scarred myocardial tissue. An episode of sustained ventricular tachycardia indicates a high risk for recurrent ventricular arrhythmias and sudden cardiac death.

Predisposing factors for ventricular arrhythmias

  • Recurrent or continued coronary ischaemia
  • Recurrent myocardial infarction
  • Hypokalaemia and hyperkalaemia
  • Hypomagnesaemia
  • Psychotropic drugs—for example, tricyclic antidepressants
  • Digoxin (leading to toxicity)
  • Antiarrhythmic drugs that may be cardiodepressant (negative inotropism) and proarrhythmic

Sustained polymorphic ventricular tachycardia and torsades de pointes are more likely to occur in the presence of precipitating or aggravating factors, including electrolyte disturbance (for example, hypokalaemia or hyperkalaemia, hypomagnesaemia), prolonged QT interval, digoxin toxicity, drugs causing electrical instability (for example, antiarrhythmic drugs, antidepressants), and continued or recurrent myocardial ischaemia. β Blockers are useful for treating arrhythmias, and these agents (for example, bisoprolol, metoprolol, carvedilol) are likely to be increasingly used as a treatment option in patients with heart failure.

Stroke and thromboembolism

Congestive heart failure predisposes to stroke and thromboembolism, with an overall estimated annual incidence of approximately 2%. Factors contributing to the increased thromboembolic risk in patients with heart failure include low cardiac output (with relative stasis of blood in dilated cardiac chambers), regional wall motion abnormalities (including formation of a left ventricular aneurysm), and associated atrial fibrillation. Although the prevalence of atrial fibrillation in some of the earlier observational studies was between 12% and 36%—which may have accounted for some of the thromboembolic events—patients with chronic heart failure who remain in sinus rhythm are also at an increased risk of stroke and venous thromboembolism. Patients with heart failure and chronic venous insufficiency may also be immobile, and this contributes to their increased risk of thrombosis, including deep venous thrombosis and pulmonary embolism.

Complications of heart failure

  • Arrhythmias —Atrial fibrillation; ventricular arrhythmias (ventricular  tachycardia, ventricular fibrillation); bradyarrhythmias
  • Thromboembolism —Stroke; peripheral embolism; deep venous  thrombosis; pulmonary embolism
  • Gastrointestinal —Hepatic congestion and hepatic dysfunction;  malabsorption
  • Musculoskeletal —Muscle wasting
  • Respiratory —Pulmonary congestion; respiratory muscle weakness;  pulmonary hypertension (rare)

Recent observational data from the studies of left ventricular dysfunction (SOLVD) and vasodilator heart failure trials (V-HeFT) indicate that mild to moderate heart failure is associated with an annual risk of stroke of about 1.5% (compared with a risk of less than 0.5% in those without heart failure), rising to 4% in patients with severe heart failure. In addition, the survival and ventricular enlargement (SAVE) study recently reported an inverse relation between risk of stroke and left ventricular ejection fraction, with an 18% increase in risk for every 5% reduction in left ventricular ejection fraction; this clearly relates thromboembolism to severe cardiac impairment and the severity of heart failure. As thromboembolic risk seems to be related to left atrial and left ventricular dilatation, echocardiography may have some role in the risk stratification of thromboembolism in patients with chronic heart failure.

Most long term (more than 10 years of follow up) longitudinal studies of heart failure, including the Framingham heart study (1971), were performed before the widespread use of angiotensin converting enzyme inhibitors. In the Framingham study the overall survival at eight years for all NYHA classes was 30%, compared with a one year mortality in classes III and IV of 34% and a one year mortality in class IV of over 60%. The prognosis in patients whose left ventricular dysfunction is asymptomatic is better than that in those whose left ventricular dysfunction is symptomatic. The prognosis in patients with congestive heart failure is dependent on severity, age, and sex, with a poorer prognosis in male patients. In addition, numerous prognostic indices are associated with an adverse prognosis, including NYHA class, left ventricular ejection fraction, and neurohormonal status.

Morbidity and mortality for all grades of symptomatic chronic heart failure are high, with a 20-30% one year mortality in mild to moderate heart failure and a greater than 50% one year mortality in severe heart failure. These prognostic data refer to patients with systolic heart failure, as the natural course of diastolic dysfunction is less well defined

Some predictors of poor outcome in chronic heart failure

  • High NYHA functional class
  • Reduced left ventricular ejection fraction
  • Low peak oxygen consumption with maximal exercise (% predicted value)
  • Increased pulmonary artery capillary wedge pressure
  • Reduced cardiac index
  • Diabetes mellitus
  • Reduced sodium concentration
  • Raised plasma catecholamine and natriuretic peptide concentrations

Survival can be prolonged in chronic heart failure that results from systolic dysfunction if angiotensin converting enzyme inhibitors are given. Longitudinal data from the Framingham study and the Mayo Clinic suggest, however, that there is still only a limited improvement in the one year survival rate of patients with newly diagnosed symptomatic chronic heart failure, which remains at 60-70%. In these studies only a minority of patients with congestive heart failure were appropriately treated, with less than 25% of them receiving angiotensin converting enzyme inhibitors, and even among treated patients the dose used was much lower than doses used in the clinical trials.  

Cardiac mortality in placebo controlled heart failure trials

CONSENSUSNYHA IV (cardiomegaly)73Enalapril38541
SOLVD-PAsymptomatic (EF <35%)83Enalapril13144
SOLVD-TSymptomatic (EF <35%)71Enalapril31364
SAVEPostmyocardial infarction (EF <40%)100Captopril17214
V-HeFT INYHA II-III (EF <45%)44H-ISDN37415
V-HeFT IINYHA II-III (EF <45%)52Enalapril28 34*5
PRAISENYHA III-IV (EF <30%)63Amlodipine28331.2

EF ejection fraction. SOLVD-P, SOLVD-T=studies of left ventricular dysfunction prevention arm (P) and treatment arm (T).

H-ISDN=hydralazine and isosorbide dinitrate.

*Treatment with H-ISDN.

Treatment with angiotensin converting enzyme inhibitors prevents or delays the onset of symptomatic heart failure in patients with asymptomatic, or minimally symptomatic, left ventricular systolic dysfunction. The increase in mortality with the development of symptoms suggests that the optimal time for intervention with these agents is well before the onset of substantial left ventricular dysfunction, even in the absence of overt clinical symptoms of heart failure. This benefit has been confirmed in several large, well conducted, postmyocardial infarction studies.

Key references

  • Doval HC, Nul DR, Grancelli HO, Perrone SV, Bortman GR, Curiel R, et al. Randomised trial of low-dose amiodarone in severe congestive heart failure. Lancet 1994;334:493-8.
  • Gradman A, Deedwania P, Cody R, Massie B, Packer M, Pitt B, et al. Predictors of total mortality and sudden death in mild to moderate heart failure. J Am Coll Cardiol 1989;14:564-70.
  • Guidelines for the diagnosis of heart failure. The Task Force on Heart Failure of the European Society of Cardiology. Eur Heart J 1995;16:741-51.
  • Rodeheffer RJ, Jacobsen SJ, Gersh BJ, Kottke TE, McCann HA, Bailey KR, et al. The incidence and prevalence of congestive heart failure in Rochester, Minnesota. Mayo Clin Proc 1993;68:1143-50.
  • The SOLVD Investigators. Effect of enalapril on mortality and the development of heart failure in asymptomatic patients with reduced left ventricular ejection fractions. N Engl J Med 1992;327:685-91.
  • The 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;316:1429-35.

Sudden death

The mode of death in heart failure has been extensively investigated, and progressive heart failure and sudden death seem to occur with equal frequency. Some outstanding questions still remain, however. Although arrhythmias are common in patients with heart failure and are indicators of disease severity, they are not powerful independent predictors of prognosis. Sudden death may be related to ventricular arrhythmias, although asystole is a common terminal event in severe heart failure. It has not been firmly established whether these arrhythmias are primary arrhythmias or whether some are secondary to acute coronary ischaemia or indicate in situ coronary thrombosis. The cause of death is often uncertain, especially as the patient may die of a cardiac arrest outside hospital or while asleep.

An external file that holds a picture, illustration, etc.
Object name is heart04.f1.jpg

Gross oedema of ankles, including bullae with serous exudate

An external file that holds a picture, illustration, etc.
Object name is heart04.f2.jpg

24 Hour Holter tracing showing frequent ventricular extrasystoles

Acknowledgments

The table on the sensitivity, specificity, and predictive value of symptoms, signs, and chest x ray findings is adapted with permission from Harlan et al ( Ann Intern Med 1977;86:133-8).

R D S Watson is consultant cardiologist in the university department of medicine and the department of cardiology, City Hospital, Birmingham.

The ABC of heart failure is edited by C R Gibbs, M K Davies, and G Y H Lip. CRG is research fellow and GYHL is consultant cardiologist and reader in medicine in the university department of medicine and the department of cardiology, City Hospital, Birmingham; MKD is consultant cardiologist in the department of cardiology, Selly Oak Hospital, Birmingham. The series will be published as a book in the spring.

  • Introduction
  • Conclusions
  • Article Information

a Information was obtained via screening logs, reporting the single criterion that did not meet eligibility.

ICU indicates intensive care unit; SGLT-2, sodium-glucose cotransporter 2.

Distribution of wins, ties, and losses for the dapagliflozin group among the 64 232 paired comparisons, stratified by each level of the hierarchical primary composite outcome. Every possible pair of participants between groups was compared in a hierarchical fashion with a win, a loss, or a tie determined by the outcome evaluated at each level of the hierarchy. Early ties were determined when both participants in the pair died during hospitalization. Percentages are calculated for each level of the hierarchy. The win ratio equals the total wins for the dapagliflozin group divided by the total losses for the dapagliflozin group: 27 143/26 929 = 1.01 (95% CI, 0.90-1.13; P  = .89). ICU indicates intensive care unit.

a Determined by physician’s assessment.

b Baseline serum creatinine levels. To convert creatinine from mg/dL to μmol/L, multiply by 88.4.

c According to the Simplified Acute Physiology Score 3 subgroup.

Shown is the win ratio for the composite hierarchical primary outcome of hospital mortality, initiation of kidney replacement therapy, and intensive care unit (ICU) length of stay stratified for prespecified subgroups. A win ratio greater than 1.0 indicates a favorable effect for the dapagliflozin group. The width of point estimates are scaled according to the number of participants in each subgroup. The 95% CIs were not adjusted for multiple comparisons and should not be used to infer treatment effects.

Trial Protocol

Statistical Analysis Plan

eTable 1. Inclusion criteria

eTable 2. Exclusion criteria

eTable 3. Key study dates and milestones

eTable 4. Protocol deviations

eTable 5. Results for the secondary outcomes from the frequentist analysis

eTable 6. Results for the exploratory post-hoc outcome

eTable 7. Hazard ratios from subdistribution and cause-specific hazard models for initiation of kidney replacement therapy

eTable 8. All investigator-reported serious adverse events

eFigure 1. Intersection between the organ dysfunction eligibility criteria for randomized participants

eFigure 2. Utilization of dapagliflozin during study follow-up

eFigure 3. Posterior predictive distributions by study group and posterior distributions of average marginal effects on hospital mortality

eFigure 4. Posterior predictive distributions by study group and posterior distributions of average marginal effects on initiation of kidney replacement therapy

eFigure 5. Posterior predictive distributions by study group and posterior distributions of average marginal effects on ICU-free days

eFigure 6. Posterior predictive distributions by study group and posterior distributions of average marginal effects on hospital-free days

eFigure 7. Posterior predictive distributions by study group and posterior distributions of average marginal effects on mechanical ventilation-free days

eFigure 8. Posterior predictive distributions by study group and posterior distributions of average marginal effects on kidney replacement therapy-free days

eFigure 9. Posterior predictive distributions by study group and posterior distributions of average marginal effects on vasopressor-free days

eFigure 10. Distribution of ICU-, hospital-, KRT-, mechanical ventilation-, and vasopressor-free days

eFigure 11. Boxplot of serum creatinine and pH Levels

eFigure 12. Posterior predictive distributions by study group and posterior distributions of average marginal effects on modified major adverse kidney events

eFigure 13. Cumulative incidence functions for hospital mortality and use of kidney replacement therapy

Nonauthor Collaborators. The DEFENDER study Investigators

Data Sharing Statement

  • Fluid Therapy for Critically Ill Adults With Sepsis: A Review JAMA Review June 13, 2023 This review summarizes the 4 phases of fluid therapy used for critically ill patients with sepsis: resuscitation, optimization, stabilization, and evacuation. Fernando G. Zampieri, MD, PhD; Sean M. Bagshaw, MD, MSc; Matthew W. Semler, MD, MSc
  • SGL-2 Therapy for Acute Organ Dysfunction JAMA Editorial June 14, 2024 Hernando Gómez, MD, MPH; Lennie P. G. Derde, MD, PhD

See More About

Select your interests.

Customize your JAMA Network experience by selecting one or more topics from the list below.

  • Academic Medicine
  • Acid Base, Electrolytes, Fluids
  • Allergy and Clinical Immunology
  • American Indian or Alaska Natives
  • Anesthesiology
  • Anticoagulation
  • Art and Images in Psychiatry
  • Artificial Intelligence
  • Assisted Reproduction
  • Bleeding and Transfusion
  • Caring for the Critically Ill Patient
  • Challenges in Clinical Electrocardiography
  • Climate and Health
  • Climate Change
  • Clinical Challenge
  • Clinical Decision Support
  • Clinical Implications of Basic Neuroscience
  • Clinical Pharmacy and Pharmacology
  • Complementary and Alternative Medicine
  • Consensus Statements
  • Coronavirus (COVID-19)
  • Critical Care Medicine
  • Cultural Competency
  • Dental Medicine
  • Dermatology
  • Diabetes and Endocrinology
  • Diagnostic Test Interpretation
  • Drug Development
  • Electronic Health Records
  • Emergency Medicine
  • End of Life, Hospice, Palliative Care
  • Environmental Health
  • Equity, Diversity, and Inclusion
  • Facial Plastic Surgery
  • Gastroenterology and Hepatology
  • Genetics and Genomics
  • Genomics and Precision Health
  • Global Health
  • Guide to Statistics and Methods
  • Hair Disorders
  • Health Care Delivery Models
  • Health Care Economics, Insurance, Payment
  • Health Care Quality
  • Health Care Reform
  • Health Care Safety
  • Health Care Workforce
  • Health Disparities
  • Health Inequities
  • Health Policy
  • Health Systems Science
  • History of Medicine
  • Hypertension
  • Images in Neurology
  • Implementation Science
  • Infectious Diseases
  • Innovations in Health Care Delivery
  • JAMA Infographic
  • Law and Medicine
  • Leading Change
  • Less is More
  • LGBTQIA Medicine
  • Lifestyle Behaviors
  • Medical Coding
  • Medical Devices and Equipment
  • Medical Education
  • Medical Education and Training
  • Medical Journals and Publishing
  • Mobile Health and Telemedicine
  • Narrative Medicine
  • Neuroscience and Psychiatry
  • Notable Notes
  • Nutrition, Obesity, Exercise
  • Obstetrics and Gynecology
  • Occupational Health
  • Ophthalmology
  • Orthopedics
  • Otolaryngology
  • Pain Medicine
  • Palliative Care
  • Pathology and Laboratory Medicine
  • Patient Care
  • Patient Information
  • Performance Improvement
  • Performance Measures
  • Perioperative Care and Consultation
  • Pharmacoeconomics
  • Pharmacoepidemiology
  • Pharmacogenetics
  • Pharmacy and Clinical Pharmacology
  • Physical Medicine and Rehabilitation
  • Physical Therapy
  • Physician Leadership
  • Population Health
  • Primary Care
  • Professional Well-being
  • Professionalism
  • Psychiatry and Behavioral Health
  • Public Health
  • Pulmonary Medicine
  • Regulatory Agencies
  • Reproductive Health
  • Research, Methods, Statistics
  • Resuscitation
  • Rheumatology
  • Risk Management
  • Scientific Discovery and the Future of Medicine
  • Shared Decision Making and Communication
  • Sleep Medicine
  • Sports Medicine
  • Stem Cell Transplantation
  • Substance Use and Addiction Medicine
  • Surgical Innovation
  • Surgical Pearls
  • Teachable Moment
  • Technology and Finance
  • The Art of JAMA
  • The Arts and Medicine
  • The Rational Clinical Examination
  • Tobacco and e-Cigarettes
  • Translational Medicine
  • Trauma and Injury
  • Treatment Adherence
  • Ultrasonography
  • Users' Guide to the Medical Literature
  • Vaccination
  • Venous Thromboembolism
  • Veterans Health
  • Women's Health
  • Workflow and Process
  • Wound Care, Infection, Healing

Others Also Liked

  • Download PDF
  • X Facebook More LinkedIn

Tavares CAM , Azevedo LCP , Rea-Neto Á, et al. Dapagliflozin for Critically Ill Patients With Acute Organ Dysfunction : The DEFENDER Randomized Clinical Trial . JAMA. Published online June 14, 2024. doi:10.1001/jama.2024.10510

Manage citations:

© 2024

  • Permissions

Dapagliflozin for Critically Ill Patients With Acute Organ Dysfunction : The DEFENDER Randomized Clinical Trial

  • 1 Hospital Israelita Albert Einstein, São Paulo, São Paulo, Brazil
  • 2 Geriatric Cardiology Unit, Instituto do Coração do Hospital das Clínicas da Faculdade de Medicina da Universidade de São Paulo, São Paulo, Brazil
  • 3 Center for Studies and Research in Intensive Care Medicine, Curitiba, Brazil
  • 4 Internal Medicine Department, Hospital de Clínicas, Federal University of Paraná, Curitiba, Brazil
  • 5 Hospital Santa Casa Curitiba, Curitiba, Brazil
  • 6 Hospital M´Boi Mirim, São Paulo, Brazil
  • 7 Hospital de Câncer de Barretos, Barretos, Brazil
  • 8 Intensive Care Division, Hospital de Base, Faculdade de Medicina de São José do Rio Preto, São José do Rio Preto, Brazil
  • 9 Hospital Universitário São Francisco de Assis na Providência de Deus, Bragança Paulista, Brazil
  • 10 Hospital de Emergência Dr Daniel Houly, Arapiraca, Brazil
  • 11 Hospital Santa Lucia, Poços de Caldas, Brazil
  • 12 Santa Casa de Misericórdia de Barretos, Barretos, Brazil
  • 13 Hospital Maternidade São José, Colatina, Brazil
  • 14 Hospital Municipal de Aparecida de Goiânia, Aparecida de Goiânia, Brazil
  • 15 Hospital Nossa Senhora de Oliveira, Vacaria, Brazil
  • 16 Centro Hospitalar Unimed, Joinville, Brazil
  • 17 Department of Cardiovascular Disease, Saint Luke’s Mid America Heart Institute, University of Missouri–Kansas City School of Medicine, Kansas City
  • 18 Australian and New Zealand Intensive Care Research Centre, School of Public Health and Preventive Medicine, Monash University, Melbourne, Australia
  • 19 Department of Intensive Care, Austin Hospital, Melbourne, Australia
  • 20 George Institute for Global Health, London, United Kingdom
  • 21 Imperial College London, London, United Kingdom
  • 22 Department of Critical Care Medicine, Faculty of Medicine and Dentistry, University of Alberta and Alberta Health Services, Edmonton, Canada
  • Editorial SGL-2 Therapy for Acute Organ Dysfunction Hernando Gómez, MD, MPH; Lennie P. G. Derde, MD, PhD JAMA
  • Review Fluid Therapy for Critically Ill Adults With Sepsis: A Review Fernando G. Zampieri, MD, PhD; Sean M. Bagshaw, MD, MSc; Matthew W. Semler, MD, MSc JAMA

Question   Does the addition of dapagliflozin to standard of care improve the hierarchical outcome of hospital mortality, initiation of kidney replacement therapy, and the length of stay in the intensive care unit (ICU) among critically ill patients with acute organ dysfunction?

Findings   In this multicenter, open-label, randomized clinical trial that included 507 participants with at least 1 acute organ dysfunction (hypotension, kidney injury, or respiratory), the use of 10 mg of dapagliflozin for up to 14 days did not significantly reduce the combined outcome of hospital mortality, initiation of kidney replacement therapy, and ICU length of stay, assessed by the win ratio method (win ratio, 1.01, not significant) through 28 days after randomization.

Meaning   The addition of dapagliflozin to standard care for individuals with critical illness and acute organ dysfunction did not improve clinical outcomes.

Importance   Sodium-glucose cotransporter 2 (SGLT-2) inhibitors improve outcomes in patients with type 2 diabetes, heart failure, and chronic kidney disease, but their effect on outcomes of critically ill patients with organ failure is unknown.

Objective   To determine whether the addition of dapagliflozin, an SGLT-2 inhibitor, to standard intensive care unit (ICU) care improves outcomes in a critically ill population with acute organ dysfunction.

Design, Setting, and Participants   Multicenter, randomized, open-label, clinical trial conducted at 22 ICUs in Brazil. Participants with unplanned ICU admission and presenting with at least 1 organ dysfunction (respiratory, cardiovascular, or kidney) were enrolled between November 22, 2022, and August 30, 2023, with follow-up through September 27, 2023.

Intervention   Participants were randomized to 10 mg of dapagliflozin (intervention, n = 248) plus standard care or to standard care alone (control, n = 259) for up to 14 days or until ICU discharge, whichever occurred first.

Main Outcomes and Measures   The primary outcome was a hierarchical composite of hospital mortality, initiation of kidney replacement therapy, and ICU length of stay through 28 days, analyzed using the win ratio method. Secondary outcomes included the individual components of the hierarchical outcome, duration of organ support–free days, ICU, and hospital stay, assessed using bayesian regression models.

Results   Among 507 randomized participants (mean age, 63.9 [SD, 15] years; 46.9%, women), 39.6% had an ICU admission due to suspected infection. The median time from ICU admission to randomization was 1 day (IQR, 0-1). The win ratio for dapagliflozin for the primary outcome was 1.01 (95% CI, 0.90 to 1.13; P  = .89). Among all secondary outcomes, the highest probability of benefit found was 0.90 for dapagliflozin regarding use of kidney replacement therapy among 27 patients (10.9%) in the dapagliflozin group vs 39 (15.1%) in the control group.

Conclusion and Relevance   The addition of dapagliflozin to standard care for critically ill patients and acute organ dysfunction did not improve clinical outcomes; however, confidence intervals were wide and could not exclude relevant benefits or harms for dapagliflozin.

Trial Registration   ClinicalTrials.gov Identifier: NCT05558098

Sodium-glucose cotransporter 2 (SGLT-2) inhibitors are effective at improving clinical outcomes in several randomized clinical trials across the spectrum of cardiovascular, metabolic, and kidney diseases. 1 - 3 Their use in acute illness, including patients with COVID-19 4 , 5 or acute heart failure 6 , 7 and immediately after experiencing myocardial infarction, 8 , 9 have been recently tested with promising but nondefinitive results. Although the exact mechanism underlying their benefits is debated, 10 various potential beneficial mechanisms are proposed, several of which could be useful for patients with critical illness. These include improvements in endothelial dysfunction, adrenergic tone modulation, oxidative stress, and cardiorenal effects, lending biological plausibility to their use in treating acute organ dysfunction. Experimental models that simulate acute intensive care unit (ICU) conditions reveal that SGLT-2 inhibitors attenuate inflammation and provide protection against organ injury. 11 , 12 In particular, nephroprotective effects of SGLT-2 inhibitors may be of interest to those treating critically ill populations, given the high incidence of acute kidney injury in this population. 13

There is no trial that assessed safety and effectiveness of SLGT-2 inhibitors in a broad population of critically ill patients with organ failure. Therefore, we conducted a randomized clinical trial to assess the effects of dapagliflozin when added to standard care of critically ill patients with acute organ dysfunction. We hypothesized that dapagliflozin could reduce the composite outcome of hospital mortality, initiation of kidney replacement therapy (KRT), and the duration of ICU stay.

The trial protocol (available in Supplement 1 ) was approved by the institutional review board from each site, and all patients or legal representatives provided written informed consent. The trial design and statistical analysis plan ( Supplement 2 ) were previously published. 14 This was an investigator-initiated, multicenter, open-label, randomized clinical trial conducted across 22 ICUs in Brazil. The trial operations were coordinated by the Academic Research Organization of the Hospital Israelita Albert Einstein. An independent data and safety monitoring board (DSMB) reviewed unblinded study data for safety. The trial was conducted in accordance with the Good Clinical Practice guidelines and is reported following the Consolidated Standards of Reporting Trials (CONSORT) 2010 reporting guideline statement for parallel-group randomized trials. 15

Eligible participants were aged 18 years or older, admitted to the ICU with an expected length of stay of 48 hours or longer, with at least 1 organ dysfunction criterion, (1) hypotension (mean arterial pressure <65 mm Hg, systolic blood pressure <90 mm Hg, or use of vasopressors), (2) signs of acute kidney injury (increase in 0.3 mg/dL [22.88 μmol/L] in serum creatinine or decrease in urine output <0.5 mL/kg/h for ≥6 hours), or (3) need of new use of high-flow nasal catheter, noninvasive, or invasive ventilation ( Table 1 ). Key exclusion criteria were the presence of organ dysfunction criteria for more than 24 hours, end-stage kidney disease undergoing maintenance dialysis, prior use of dapagliflozin or other SGLT-2 inhibitor, known type 1 diabetes, a history of diabetic ketoacidosis, and planned ICU admission following elective surgery ( Figure 1 ). Further details are found in eTables 1 and 2 in Supplement 3 .

Eligible patients were randomized in a 1:1 ratio to receive either 10 mg of open-label dapagliflozin in addition to standard care (dapagliflozin intervention group) or standard care alone (control group). Randomization was performed by a central, concealed, web-based automated system (Research Electronic Data Capture [REDCap]), stratified by study site with variable block sizes of 4, 8, and 12. There was no blinding.

Dapagliflozin, 10 mg/d, was given orally within 24 hours of randomization, preferably in the morning without fasting, for a duration of 14 days or until ICU discharge, whichever occurred first. For participants who were unable to swallow pills, dapagliflozin was administered enterally after macerating the medication and diluting it in water before administration. 4 , 5 Study protocol mandated that dapagliflozin administration was discontinued in the following situations: (1) absolute fasting or the inability to access the enteral route for drug administration, (2) occurrence of euglycemic diabetic ketoacidosis (blood glucose ≤250 mg/dL [13.88 mmol/L], metabolic acidosis, and moderate ketonuria [≥2 on urine stick] or ketonemia [blood ketones ≥1.5 mmol/L]), (3) more than 1 episode of severe hypoglycemia (blood glucose ≤50 mg/dL [2.77 mmol/L]), (4) withdrawal of consent, (5) suspected allergic reaction, and (6) initiation of KRT. Adherence was assessed daily for 14 days. Each study site was expected to provide standard of care treatment for critical illness for all trial participants, which was determined solely by the local site health care team and aligned with institutional protocols and international guidelines. This included various aspects of care, such as ventilation strategies, management of sepsis, delirium prevention and management, prophylaxis for deep venous thrombosis, sedation practices, pain management, and other relevant components of critical care. A minimum daily carbohydrate intake of 100 g of glucose was suggested for all study participants.

Baseline demographic information, comorbidities, concomitant medications, reasons for ICU admission, and illness severity were collected at enrollment. From days 1 to 5, monitoring included laboratory parameters such as blood gas analysis and serum creatinine levels. Participants were followed up for 28 days or until discharged home from the hospital. Adverse events were observed until trial follow-up was complete. Hospital outcomes were documented either at the time of hospital discharge or after 28 days of follow-up, whichever occurred earlier. All data collection was performed by trained site personnel using a dedicated electronic data capture system, and comprehensive data monitoring was conducted across all sites, either through remote means or on-site evaluations. Records of screening failures were documented in the form of weekly screening logs for each enrolling and active site. To ensure trial representativeness and diversity, self-reported race and ethnicity information was collected by site personnel, using available data from electronic medical records or directly from participants when feasible.

The primary outcome was a hierarchical composite of hospital mortality, initiation of KRT, and ICU length of stay through 28 days after randomization. For ICU length of stay, the cumulative number of calendar days (without fractions) spent in the ICU was calculated from randomization until hospital discharge.

The 7 prespecified secondary outcomes included hospital mortality, KRT use, ICU-free days, hospital-free days, vasopressor-free days, mechanical ventilation–free days, and KRT-free days. All secondary outcomes were evaluated within 28 days after randomization. To be considered free of vasopressor and mechanical ventilation, a cutoff of 6 hours or less within a calendar day was used. The ICU-free days, hospital-free days, and KRT-free days were defined as the count of full calendar days (without fractions) in which participants were alive and free from each respective component. These outcomes were measured on an ordinal scale ranging from 0 to 29, with higher values signifying more favorable outcomes. Participants who did not survive until hospital discharge were assigned a value of 0. For those discharged to home before day 28, it was assumed that they remained alive and free from the specified outcome beyond their discharge date.

Adverse events of special interest were collected during the trial: (1) elevation of elevated serum liver transaminases (exceeding 3 times the reference range), (2) skin lesions, (3) hypoglycemia (blood glucose ≤50 mg/dL), (4) urinary tract infections, (5) bloodstream infections, and (6) occurrence of diabetic ketoacidosis (metabolic acidosis and moderate ketonuria [≥2 on urine stick] or ketonemia [blood ketones ≥1.5 mmol/L]). These events were reported without regard to their severity or causality assessment. All serious adverse events occurring during study follow-up were recorded, regardless of presumed causality.

The sample size was calculated under the hypothesis that dapagliflozin would lead to reductions in all the individual components of the hierarchical composite primary outcome, anticipating a 2% absolute reduction in hospital mortality (from 30% to 28%), a 3% absolute reduction in the initiation KRT (from 10% to 7%), and a mean reduction in ICU length of stay by 0.5 days (with an assumed variance of 1.1 days). In simulations, enrolling 500 participants would provide the study with at least 85% statistical power to detect an intervention effect, with a 95% CI for the win ratio exceeding 1.0 and a resulting median simulated value for the win ratio of 1.40. Ten thousand simulations with samples of 500 participants each were conducted, with 95% CIs calculated using bootstrapping. Additional information is shown in Supplements 1 and 2 .

The hierarchical composite primary outcome was analyzed using the generalized pairwise comparison method 16 and the treatment effect quantified using the win ratio method. This approach involved comparing each participant in the dapagliflozin group with every participant in the control group, generating all conceivable participant pairs across trial groups. In each pairwise comparison, a win, loss, or tie was defined based on the comparative assessment of participant outcomes in hierarchical fashion. The primary composite outcome hierarchy consisted of 3 hierarchical levels, (1) hospital mortality, (2) initiation of KRT, and (3) ICU length of stay. For the first level of comparison, if both participants in a pair died before discharge, it was classified as an early tie . This signifies that the pair is not subjected to further comparison for the second or third hierarchical levels, thus emphasizing the higher importance of hospital mortality. 17 If both participants survived, the pair was subsequently evaluated for the initiation of KRT. In the event of a tie, the participants were then compared with respect to ICU length of stay. The win ratio was calculated by dividing the total number of wins in the dapagliflozin group by the total number of losses. A detailed win ratio hierarchy flowchart is shown in the eMethods section in Supplement 3 , and review of the generalized pairwise method is found elsewhere. 18

The secondary binary outcomes were assessed with a bayesian hierarchical logistic regression (multilevel) model, adjusted for study site (random intercept), participant age, clinical suspicion of sepsis, and the use of vasopressors and mechanical ventilation at randomization using a normally distributed neutral prior, centered at an odds ratio (OR) of 1.0 (corresponding to a 95% credible interval [CrI] ranging between 0.5 and 2.0). 19 Days-free secondary outcomes were analyzed with a hierarchical ordinal bayesian model adjusted for the same covariates. Treatment effects were quantified using adjusted OR, bayesian 95% CrIs, and probability of benefit for the dapagliflozin group. Further details regarding secondary models’ assumptions are provided in Supplement 2 and in the eMethods section in Supplement 3 .

To complement the bayesian analysis, additional prespecified frequentist analyses were conducted for the secondary outcomes. For hospital mortality and initiation of KRT, a logistic regression model was used, adjusting for the same covariates used in the bayesian models. For the ordinal secondary outcomes, differences between the groups were computed using the Hodges-Lehmann estimator, with results presented as differences in days between groups along with their corresponding 95% CIs. Comparisons of trends in serum creatinine and pH levels between study groups were conducted from days 1 to 5 using a linear mixed-effects model for repeated measures.

The efficacy and safety analyses included all the participants who underwent randomization (intention-to-treat principle). An additional sensitivity analysis for safety outcomes was conducted in the safety analysis population, comprising participants who received at least 1 dose of dapagliflozin.

Prespecified subgroup analyses for the primary outcome were conducted using the stratified win ratio method 20 for the following subgroups, (1) presence of clinical suspicion of sepsis at randomization, (2) prior diabetes, (3) serum creatinine levels at enrollment (<1.5 mg/dL, 1.5-3.0 mg/dL, and >3.0 mg/dL [to convert creatinine from mg/dL to μmol/L, multiply by 88.4]), (4) reason for ICU admission due to cardiovascular causes (from the table of reasons for ICU admission of Simplified Acute Physiology Score 3 [SAPS 3]), 21 and (5) age (<65 years and ≥65 years).

The DSMB led all planned safety analyses after the enrollment of 100, 250, and 375 participants. These analyses included the absolute and relative frequencies of all serious adverse events, adverse events of special interest, hospital mortality, and initiation of KRT, according to study groups. An interim analysis was performed when half of the intended trial population (250 participants) was enrolled. At this analysis, the DSMB would recommend halting the trial for safety reasons if the posterior probability of harm associated with dapagliflozin for the composite outcome of hospital mortality or KRT exceeded 80%. No interim analyses were conducted for efficacy or futility (see Supplement 3 ).

Post hoc exploratory analyses were conducted to assess the effect of dapagliflozin on (1) modified major adverse kidney events (MAKEs), defined as the composite outcome of death, initiation of KRT, or doubling the serum creatinine level during the first 5 days after enrollment, and (2) the use of KRT while accounting for the competing risk of death (see the eMethods section in Supplement 3 ).

For the primary outcome, a 2-sided P value of less than .05 was considered to indicate statistical significance and 95% CIs were calculated using the bootstrap method. 16  P values are presented exclusively for the primary outcome and subgroup analyses. The analyses of secondary outcomes were not adjusted for multiple comparisons. All analyses were conducted using R software version 4.2.1 or higher (R Foundation for Statistical Computing). 22

From November 22, 2022, to August 30, 2023, 4434 participants were screened, and 507 participants from 22 sites in Brazil were randomized: 248 to receive dapagliflozin plus standard care and 259 to receive standard care ( Figure 1 ; eTable 3 in Supplement 3 ). All 507 participants (mean age, 63.9 (SD, 15) years; 46.9% women) were included in the analysis, with no loss to follow-up. The database lock was performed on October 20, 2023. Two hundred four (39.6%) had ICU admission due to suspected infection, and the median time from ICU admission to randomization was 1 day (IQR, 0-1 day, Table 1 ). At randomization, 235 participants (46.4%) required respiratory support with mechanical ventilation and 253 (49.9%) received norepinephrine. Furthermore, and the number of trial participants who met the organ dysfunction eligibility criteria were 249 (49.5%) for respiratory, 227 (44.2%) for hypotension, and 212 (42.2%) for kidney injury. The most common inclusion criterion was kidney injury in isolation (140 patients [27.6%]), followed by respiratory dysfunction in isolation (120 patients [23.6%]), and hypotension in isolation (96 patients [18.9%]); remaining possible combinations and their frequencies are shown in eFigure 1 in Supplement 3 .

All 248 participants randomized to receive dapagliflozin received at least 1 dose of the study medication. None of the control group participants received dapagliflozin or any other SGLT-2 inhibitor during study follow-up (eFigure 2 and eTable 4 in Supplement 3 ).

Dapagliflozin treatment did not result in a higher number of wins than the standard care alone group for the primary hierarchical composite outcome. The total number of wins was 27 143 (42.3%) in the dapagliflozin group and 26 929 (41.9%) in the standard care alone group, a win ratio of 1.01 (95% CI, 0.90 to 1.13; P  = .89; Table 2 ). Among all pairwise comparisons, there were 10 160 ties (15.8%), with 7832 (12.2%) occurring in the hospital mortality comparison and classified as early ties ( Figure 2 ).

Within 28 days, hospital mortality occurred in 88 of 248 participants (35.5%) in the dapagliflozin group compared with 89 of 259 participants (34.4%) in the standard care alone group. The adjusted OR for the bayesian model, accounting for study site, age, sepsis at randomization, use of vasopressors at randomization, and use of mechanical ventilation, was 1.06 (95% CrI, 0.76-1.52; Table 2 ). Initiation of KRT occurred in 27 participants (10.9%) in the dapagliflozin group compared with 39 participants (15.1%) in the standard care alone group (adjusted OR, 0.76; 95% CrI, 0.50-1.18). The posterior probabilities indicating that the use of dapagliflozin reduced the risk of hospital mortality and initiation of KRT compared with standard of care alone, were .36 and .90, respectively ( Table 2 and eFigures 3 and 4 in Supplement 3 ).

For the secondary ordinal outcomes of ICU-free days, hospital-free days, mechanical ventilation-free days, KRT-free days, and vasopressor-free days, the results from the bayesian hierarchical logistic regression models were inconclusive about treatment effect on these outcomes, yielding posterior probabilities of benefit for dapagliflozin between 0.39 to 0.63 ( Table 2 and eFigures 5-10 in Supplement 3 ). The complementary frequentist analyses of the secondary outcomes yielded results that were consistent with bayesian analysis, with an adjusted OR of 1.08 (95% CI, 0.73-1.60) for hospital mortality and 0.67 (95%CI, 0.39-1.13) for use of KRT (eTable 5 in Supplement 3 ). No significant differences between study groups were observed for serum creatinine and pH levels during the initial 5 days of trial follow-up (eFigure 11 in Supplement 3 ).

No evidence of heterogeneity of treatment effect was detected in predefined subgroups ( Figure 3 ), as assessed in a one-at-a-time subgroup analysis. Among patients receiving dapagliflozin compared with standard care, there was a posterior probability of benefit of .60 for modified MAKEs (eTable 6 and eFigure 12 in Supplement 3 ). When considering the competing risk of death, the cause-specific adjusted hazard ratio for the dapagliflozin group for the use of KRT was 0.72 (95% CI, 0.44-1.18), a similar result was obtained from the Fine and Gray model (adjusted HR, 0.71; 95% CI, 0.45-1.13; eTable 7 and eFigure 13 in Supplement 3 ).

Investigator-reported serious adverse events were documented in 115 participants (46.4%) in the dapagliflozin group and in 123 participants (47.5%) in the control group ( Table 3 and eTable 8 in Supplement 3 ). Adverse events of special interest, including urinary tract infections (4 [1.6%] vs 3 [1.2%]), hypoglycemia (2 [0.8%] vs 0), and bloodstream infections (1 [0.4%] vs 4 [1.5%]) were reported in the dapagliflozin group vs the control group, respectively. There were no reported cases of ketoacidosis.

In this randomized, open-label, controlled clinical trial involving 507 participants, the addition of dapagliflozin to standard care was not associated with an increase in the win ratio for a hierarchical end point of hospital mortality, use of KRT, and ICU length of stay. Of the 7 secondary end points, a suggestion of benefit was found for only 1 (the use of KRT, 0.90 probability of benefit). As expected in critically ill patients, a substantial number of serious adverse events were reported in both trial groups. However, dapagliflozin use was well tolerated, with numerically fewer serious adverse events reported in this group than the standard care alone group.

There is increasing interest in SGLT-2 inhibitors for treating acutely ill patients. There is high-quality evidence to support their use in outpatients with diabetes, 1 heart failure, 2 and chronic kidney disease, 3 and there is some potential benefit for patients with myocardial infarction. 8 , 9 The benefits of SGLT-2 inhibitors may derive from their nephroprotective effects. 3 , 23 Experimental evidence suggests that this may be evident in models of sepsis, 12 and the biological rational may also involve different pathways (including inflammation, energy metabolism, and endothelial function), 24 - 26 which are mediators for organ dysfunction among acutely illness patients. 27 - 29 This trial was designed to extend the prior evidence and assess the effects of dapagliflozin in an unselected population of critically ill patients in a randomized trial.

These results have several implications. First, despite a neutral result for the primary end point, dapagliflozin use appeared safe in a population of critically ill patients with a hospital mortality rate of 35%. More specifically, adverse events of interest that have been suggested to occur with dapagliflozin use, including bloodstream or urinary infections, were uncommon and occurred at similar rates in both groups, and no ketoacidosis event was reported during the trial. Second, although the results were also inconclusive for all secondary end points, they do not exclude the potential benefits or harms from this therapy. Third, the probability of benefit for the prespecified secondary outcome of reducing KRT use was 0.90. This was not confirmed in a post hoc analysis that considered composite kidney end points or competing risks. Although the finding may be due to chance, it is aligned with several trials in the outpatient setting that suggested a nephroprotective effect of SGLT-2 inhibitors. For example, the DARE-19 trial 4 found that kidney events were numerically lower in patients with COVID-19 who were treated with dapagliflozin. Taken together, these trial results suggest that further study of SGLT-2 inhibitors on critically ill patients should continue and that renal outcomes could be favored as a potential target.

This trial has several limitations. First, the unblinded nature of the trial may introduce bias. Second, the trial enrolled an unselected and heterogeneous population of critically ill patients across various stages of acute illness. It is conceivable, for example, that participants with specific features (diabetes, chronic kidney disease, etc) may have differential treatment effects, but these were not observed. As a first trial of its kind, broad inclusion criteria were used to assess safety and the drug effects on clinical outcomes. 30 Third, no data were available on the biological response to dapagliflozin, and it is possible that inadequate absorption of the oral drug may have influenced the findings. Fourth, the analysis of secondary outcomes used models adjusted for clinical suspicion of sepsis based on physicians’ assessment rather than confirmed through objective criteria.

The addition of dapagliflozin to standard care for critically ill patients and acute organ dysfunction did not improve clinical outcomes; however, confidence intervals were wide and could not exclude relevant benefits or harms for dapagliflozin.

Accepted for Publication: May 17, 2024.

Published Online: June 14, 2024. doi:10.1001/jama.2024.10510

Corresponding Author: Fernando G. Zampieri, MD, PhD, Rua Comendador Elias Jafet, 755, São Paulo, SP, Brazil, 05653-000 ( [email protected] ).

Author Contributions: Drs Tavares, Schuler, Monfardini, Nieri, Damiani, and Zampieri had full access to all of the data in the study and take responsibility for the integrity of the data and the accuracy of the data analysis.

Concept and design: Tavares, Almeida, Figueiredo, Monfardini, Silva, Kosiborod, Pereira, Damiani, Corrêa, Serpa-Neto, Berwanger, Zampieri.

Acquisition, analysis, or interpretation of data: Tavares, Azevedo, Rea-Neto, Campos, Amendola, Kozesinski-Nakatani, David-João, Lobo, Filiponi, Almeida, Bergo, Guimarães, Castro, Schuler, Westphal, Carioca, Monfardini, Nieri, Neves, Paulo, Albuquerque, Silva, Pereira, Corrêa, Serpa-Neto, Berwanger.

Drafting of the manuscript: Tavares, Azevedo, Filiponi, Almeida, Carioca, Monfardini, Corrêa, Serpa-Neto, Zampieri.

Critical review of the manuscript for important intellectual content: Tavares, Rea-Neto, Campos, Amendola, Kozesinski-Nakatani, David-João, Lobo, Almeida, Bergo, Guimarães, Figueiredo, Castro, Schuler, Westphal, Nieri, Neves, Paulo, Albuquerque, Silva, Kosiborod, Pereira, Damiani, Corrêa, Serpa-Neto, Berwanger.

Statistical analysis: Tavares, Schuler, Monfardini, Nieri, Damiani, Zampieri.

Obtained funding: Berwanger, Zampieri.

Administrative, technical, or material support: Tavares, Azevedo, Campos, Filiponi, Almeida, Monfardini, Neves, Paulo, Albuquerque, Silva, Pereira, Serpa-Neto.

Supervision: Tavares, Azevedo, David-João, Figueiredo, Kosiborod, Pereira, Corrêa, Berwanger.

Other: Guimarães.

Other - Database Quality Review: Carioca.

Other - Patient recruitment: Westphal.

Conflict of Interest Disclosures: Dr Tavares reported receiving grants from Novo Nordisk outside the submitted work. Dr Azevedo reported receiving lecture fees from Baxter, MSD, Biolab, and Nestle; nonfinancial support from MSD; and a grant for congress participations outside the submitted work. Dr Lobo reported receiving personal fees from Edwards, Pfizer, and Roche outside the submitted work. Dr Kosiborod reported receiving to his institution personal fees from 35Pharma, Alnylam, Amgen, Applied Therapeutics, Arrowhead Pharmaceuticals, Bayer, Boehringer Ingelheim, Cytokinetics, Dexom, Eli Lilly, Esperion Therapeutics, Imbria Pharmaceuticals, Janssen, Lexicon Pharmaceutcials, Merck, NovoNordisk, Pfizer, Pharmacosmos, Regeneron, Sanofi, scPharmaceutical, Structure Therapeutics, Vifor Pharma, and Youngene Therapeutics; grants to his institution from AstraZeneca and Boehringer Ingelheim; and having stock options from Artera Health and Saghmos Therapeutics. Dr Pereira reported receiving grants from the Brazilian Ministry of Health during the conduct of the study and outside the submitted work. Dr Serpa-Neto reported receiving personal fees from Drager outside the submitted work. Dr Berwanger reported receiving grants to his previous institution from Amgen, AstraZeneca, Bayer, Novartis, Servier, and Pfizer outside the submitted work. Dr Zampieri reported receiving consulting fees from Baxter International and Bactiguard and receiving grants to his institution from Ionis Pharmaceuticals outside the submitted work. No other disclosures were reported.

Funding/Support: This trial is funded by Brazilian Ministry of Health through the Programa de Apoio ao Desenvolvimento Institucional do Sistema Único de Saúde—PROADI-SUS.

Role of the Funder/Sponsor: The Brazilian Ministry of Health approved the study but had no role in the design and conduct of the study; collection, management, analysis, interpretation of the data; manuscript preparation, review, or approval; nor in the decision to submit the manuscript for publication.

Group Information: A complete list of the DEFENDER study investigators appears in Supplement 4 .

Meeting Presentation: This paper was presented at the Critical Care Reviews Meeting; June 14, 2024; Belfast, United Kingdom.

Data Sharing Statement: See Supplement 5 .

  • Register for email alerts with links to free full-text articles
  • Access PDFs of free articles
  • Manage your interests
  • Save searches and receive search alerts

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

Acute heart failure: clinical presentation, one-year mortality and prognostic factors

Affiliation.

  • 1 Department of Medicine, University Hospital Zürich, Switzerland. [email protected]
  • PMID: 15921809
  • DOI: 10.1016/j.ejheart.2005.01.014

Aims: Acute heart failure (HF) is a common but ill-defined clinical entity. We describe patients hospitalised with acute HF in regard of clinical presentation, mortality, and risk factors for an unfavourable outcome.

Methods and results: We conducted a prospective study including 312 consecutive patients from two European centers hospitalised with acute HF, defined as new onset or worsening of symptoms and signs of HF within 7 days. The mean age was 73 years and 56% were men. Twenty-eight percent had de-novo acute HF and 72% a decompensation of chronic HF. Coronary heart disease (CHD) was the most frequent underlying heart disease, elevated blood pressure >150 mmHg and acute ischemia being the most important triggers. Four percent of the patients had cardiogenic shock, 13% presented with pulmonary edema. LV-EF was <35%, 35-50% and >50% in 35%, 32% and 33% of the patients, respectively. ICU-treatment was necessary in 39% of the patients. Thirty-day mortality (11%) was increased in the presence of shock or elevated troponin T levels. Twelve-month all-cause mortality (29%) increased in the presence of shock, left ventricular dysfunction, renal insufficiency, CHD, and age.

Conclusions: This prospective study shows that despite modern treatment, morbidity and mortality of patients hospitalised with acute HF remain high.

PubMed Disclaimer

Similar articles

  • Clinical Assessment in Acute Heart Failure. Kakouros NS, Kakouros SN. Kakouros NS, et al. Hellenic J Cardiol. 2015 Jul-Aug;56(4):285-301. Hellenic J Cardiol. 2015. PMID: 26233768 Review. No abstract available.
  • Acute pulmonary oedema: clinical characteristics, prognostic factors, and in-hospital management. Parissis JT, Nikolaou M, Mebazaa A, Ikonomidis I, Delgado J, Vilas-Boas F, Paraskevaidis I, Mc Lean A, Kremastinos D, Follath F. Parissis JT, et al. Eur J Heart Fail. 2010 Nov;12(11):1193-202. doi: 10.1093/eurjhf/hfq138. Epub 2010 Sep 13. Eur J Heart Fail. 2010. PMID: 20837636
  • [50-year trends in the incidence and prognosis of heart failure: the Framingham study]. Imai Y. Imai Y. Nihon Rinsho. 2007 Apr 28;65 Suppl 4:61-4. Nihon Rinsho. 2007. PMID: 17508541 Review. Japanese. No abstract available.
  • Heart failure with preserved left ventricular systolic function among patients with non-ST-segment elevation acute coronary syndromes. Bennett KM, Hernandez AF, Chen AY, Mulgund J, Newby LK, Rumsfeld JS, Hochman JS, Hoekstra JW, Ohman EM, Gibler WB, Roe MT, Peterson ED. Bennett KM, et al. Am J Cardiol. 2007 May 15;99(10):1351-6. doi: 10.1016/j.amjcard.2006.12.057. Epub 2007 Apr 9. Am J Cardiol. 2007. PMID: 17493458
  • Prevalence, clinical characteristics, quality of life, and prognosis of patients with congestive heart failure and isolated left ventricular diastolic dysfunction. Badano LP, Albanese MC, De Biaggio P, Rozbowsky P, Miani D, Fresco C, Fioretti PM. Badano LP, et al. J Am Soc Echocardiogr. 2004 Mar;17(3):253-61. doi: 10.1016/j.echo.2003.11.002. J Am Soc Echocardiogr. 2004. PMID: 14981424
  • Nitric Oxide in Cardiac Surgery: A Review Article. Kamenshchikov NO, Duong N, Berra L. Kamenshchikov NO, et al. Biomedicines. 2023 Apr 3;11(4):1085. doi: 10.3390/biomedicines11041085. Biomedicines. 2023. PMID: 37189703 Free PMC article. Review.
  • Outcomes and predictors of one-year mortality in patients hospitalized with Acute Heart Failure. Lorlowhakarn K, Arayakarnkul S, Trongtorsak A, Leesutipornchai T, Kewcharoen J, Sinphurmsukskul S, Siwamogsatham S, Puwanant S, Ariyachaipanich A. Lorlowhakarn K, et al. Int J Cardiol Heart Vasc. 2022 Nov 30;43:101159. doi: 10.1016/j.ijcha.2022.101159. eCollection 2022 Dec. Int J Cardiol Heart Vasc. 2022. PMID: 36467463 Free PMC article.
  • Mortality and Readmission Rates After Heart Failure: A Systematic Review and Meta-Analysis. Lan T, Liao YH, Zhang J, Yang ZP, Xu GS, Zhu L, Fan DM. Lan T, et al. Ther Clin Risk Manag. 2021 Dec 7;17:1307-1320. doi: 10.2147/TCRM.S340587. eCollection 2021. Ther Clin Risk Manag. 2021. PMID: 34908840 Free PMC article. Review.
  • Long-term prognostic value of functional status and delirium in emergency patients with decompensated heart failure. Alberto RM, Domingo R, Aitor A, Sergio HM, Pascual P, Mireia P, Salvador B, Herminia TO. Alberto RM, et al. Eur Geriatr Med. 2018 Aug;9(4):515-522. doi: 10.1007/s41999-018-0072-0. Epub 2018 Jun 14. Eur Geriatr Med. 2018. PMID: 34674495
  • Right atrial volume index to left atrial volume index ratio is associated with adverse clinical outcomes in cardiogenic shock. Patel PR, Stafford PL, Bilchick KC, Walker MR, Ibrahim S, Martin D, Betz Y, Patel TR, Kwon Y, Mehta N, Sodhi N, Mwansa H, Breathett K, Mazimba S. Patel PR, et al. J Echocardiogr. 2022 Mar;20(1):42-50. doi: 10.1007/s12574-021-00552-7. Epub 2021 Oct 8. J Echocardiogr. 2022. PMID: 34623621 Free PMC article.

Publication types

  • Search in MeSH

Related information

Linkout - more resources, full text sources.

  • MedlinePlus Health Information

Research Materials

  • NCI CPTC Antibody Characterization Program

Miscellaneous

  • NCI CPTAC Assay Portal

full text provider logo

  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

IMAGES

  1. PPT

    clinical presentation of acute heart failure

  2. Heart failure 1: pathogenesis, presentation and diagnosis

    clinical presentation of acute heart failure

  3. Conceptual frameworks of acute heart failure presentations. Patients

    clinical presentation of acute heart failure

  4. Heart Failure

    clinical presentation of acute heart failure

  5. Clinical presentation of acute heart failure according to the SBP on

    clinical presentation of acute heart failure

  6. Stages of Heart Failure

    clinical presentation of acute heart failure

VIDEO

  1. Acute Viral Hepatitis #shorts #hepatitis #gastroenterologist

  2. Heart failure Diagnosis

  3. Heart failure? How to manage unexpected challenges

  4. Complications due to uncontrolled hypertension #hypertension #highbloodpressure #heartdisease

  5. Heart failure (CHF) pathophysiology شرح بالعربى باثوفسيولوجى فشل القلب

  6. Teaser

COMMENTS

  1. Heart failure: Clinical manifestations and diagnosis in adults

    HF is a complex clinical syndrome identified by presence of current or prior characteristic symptoms, such as dyspnea and fatigue, and evidence of cardiac dysfunction as a cause of these symptoms (eg, abnormal left ventricular [LV] and/or right ventricular [RV] filling and elevated filling pressures) [ 1-5 ]. From a hemodynamic perspective, HF ...

  2. Acute Heart Failure: Definition, Classification and Epidemiology

    Acute heart failure is a clinical syndrome characterised by signs and symptoms of fluid overload which require hospitalisation. Patients may present with AHF as the first presentation of heart disease but more commonly as decompensation of a pre-existing cardiomyopathy. ... Pathogenesis and clinical presentation of acute heart failure. Rev Esp ...

  3. Acute Heart Failure: Types, Symptoms, Causes and Treatment

    Acute heart failure is a sudden, life-threatening condition that occurs when your heart can no longer do its job. ADHF occurs in people with a history of heart disease. De novo heart failure is due to other medical conditions affecting the heart. You should seek emergency medical treatment if you experience heart failure symptoms.

  4. Acute Heart Failure: Diagnostic-Therapeutic Pathways and Preventive

    Acute heart failure (AHF) is the most frequent cause of unplanned hospital admission in patients of >65 years of age and it is associated with significantly increased morbidity, mortality, and healthcare costs. ... management of AHF differs according to clinical presentation [3,17]. Figure 2. Open in a separate window.

  5. Understanding acute heart failure: pathophysiology and diagnosis

    Acute heart failure (AHF) is a relevant public health problem causing the majority of unplanned hospital admissions in patients aged of 65 years or more. 1 Despite major achievements in the treatment of chronic heart failure ... Acute heart failure: clinical presentation, one-year mortality and prognostic factors ...

  6. Acute heart failure

    The clinical presentation of AHF is characterized mostly by symptoms and signs related to systemic congestion (that is, extracellular fluid accumulation, initiated by increased biventricular cardiac filling pressures) 6,7. ... Acute heart failure (HF) results from the combination of an underlying but newly diagnosed cardiac dysfunction and ...

  7. Heart Failure Clinical Presentation

    Clinical presentation, management, and in-hospital outcomes of patients admitted with acute decompensated heart failure with preserved systolic function: a report from the Acute Decompensated Heart Failure National Registry (ADHERE) Database. ... and a somewhat leftward QRS axis (-15º). The patient had malignant hypertension with acute heart ...

  8. Pathogenesis and Clinical Presentation of Acute Heart Failure

    Acute heart failure (AHF) can be defined as a heterogeneous syndrome of signs and symptoms of new-onset or gradual/rapidly worsening heart failure (HF), requiring urgent therapy. 1,2 AHF constitutes a clinical syndrome with a complex and, more importantly, not completely understood pathophysiology. 3-5 Given the diversity of clinical ...

  9. Pathogenesis and clinical presentation of acute heart failure

    Abstract. Acute heart failure constitutes a heterogeneous clinical syndrome, whose pathophysiology is complex and not completely understood. Given the diversity of clinical presentations, several different pathophysiological mechanisms along with factors triggering circulatory decompensation are involved. This article discusses the available ...

  10. Acute heart failure

    Follath, F. et al. Clinical presentation, management and outcomes in the acute heart failure global survey of standard treatment (ALARM-HF). Intensive Care Med. 37 , 619-626 (2011).

  11. Pathophysiology and Therapeutic Approaches to Acute Decompensated Heart

    Definition and Clinical Presentation. The clinical presentation of symptoms and signs of congestion and poor organ perfusion due to HF requiring urgent, usually intravenous, therapy has been variously called AHF, ADHF, AHF syndrome, and hospitalized HF, as well as other terms. ... Diuretic response in acute heart failure: clinical ...

  12. Diagnosis and Evaluation of Heart Failure

    Appropriate diagnosis and therapy for heart failure are important given the poor prognosis. Survival is 89.6 percent at one month from diagnosis, 78 percent at one year, and only 57.7 percent at ...

  13. Pathogenesis and Clinical Presentation of Acute Heart Failure

    and Clinical Presentation of Acute Heart Failure Piotr Ponikowskia,b,* and Ewa A. Jankowskaa,b aDepartment b of Heart Diseases, Wroclaw Medical University, Wroclaw, Poland Centre for Heart Diseases, Military Hospital, Wroclaw, Poland ACUTE HEARTFAILURE:ACOMPLEXCLINICALSYNDROMEWITH DISTINCT PATHOPHYSIOLOGIES Acute heart failure (AHF) can be ...

  14. Approach to diagnosis and evaluation of acute decompensated heart

    INTRODUCTION. Acute decompensated heart failure (ADHF) is a clinical syndrome of new or worsening signs and symptoms of HF that often lead to hospitalization or an emergency department visit. For the purposes of this discussion, we use the term ADHF for all the acute HF syndromes. The clinical manifestations and diagnosis of ADHF will be ...

  15. PDF Clinical Assessment in Acute Heart Failure

    15451 Athens, Greece. [email protected]. eart failure (HF) is defined as "a H complex clinical syn drome that can result from any structural or functional cardiac disorder that impairs the ability of the ventricle to fill with, or eject blood.". HF has an estimated overall prev alence of 2.6%.

  16. Clinical presentation and outcome by age categories in acute heart

    Clinical presentation and outcome by age categories in acute heart failure: results from an international observational cohort ... The GREAT registry consisted of patients identified as presenting with acute heart failure at the emergency department. Four groups of patients were defined according to age: the young patient group (<65 years ...

  17. Clinical presentation and outcomes of acute heart failure in the

    Introduction. Heart failure is a clinical syndrome characterized by a set of clinical signs (elevated jugular venous pressure, pulmonary congestion) and non-specific symptoms (dyspnea, orthopnoea, lower limb swelling) caused by a structural and/or functional cardiac abnormality, leading to reduced cardiac output and/or elevated intracardiac pressures at rest or during stress. 1 The number of ...

  18. Pathogenesis and Clinical Presentation of Acute Heart Failure

    Acute heart failure (AHF) can be defined as a heterogeneous syndrome of signs and symptoms of new-onset or gradual/rapidly worsening heart failure (HF), requiring urgent therapy. 1, 2 AHF constitutes a clinical syndrome with a complex and, more importantly, not completely understood pathophysiology. 3, 4, 5 Given the diversity of clinical ...

  19. Clinical presentation, management and outcomes in the Acute Heart

    Purpose: We performed a survey on acute heart failure (AHF) in nine countries in four continents. We aimed to describe characteristics and management of AHF among various countries, to compare patients with de novo AHF versus patients with a pre-existing episode of AHF, and to describe subpopulations hospitalized in intensive care unit (ICU) versus cardiac care unit (CCU) versus ward.

  20. ABC of heart failure: Clinical features and complications

    Clinical features. Patients with heart failure present with a variety of symptoms, most of which are non-specific. The common symptoms of congestive heart failure include fatigue, dyspnoea, swollen ankles, and exercise intolerance, or symptoms that relate to the underlying cause. The accuracy of diagnosis by presenting clinical features alone ...

  21. Clinical presentation, management, and in-hospital outcomes of patients

    Objectives: The aims of this analysis were to describe the clinical characteristics, management, and outcomes of patients hospitalized for acute decompensated heart failure (HF) with preserved systolic function (PSF). Background: Clinically meaningful characteristics of these patients have not been fully studied in a large database

  22. Heart Failure Stages

    So this is really the bulk of our heart failure clinical work. And then of course, Stage D describes advanced heart failure. So these are individuals who have ongoing heart failure symptoms that interfere with daily life and usually recurrent hospitalizations despite all the attempts to optimize their stability with guideline-directed medical ...

  23. Dapagliflozin for Critically Ill Patients With Acute Organ Dysfunction

    Sodium-glucose cotransporter 2 (SGLT-2) inhibitors are effective at improving clinical outcomes in several randomized clinical trials across the spectrum of cardiovascular, metabolic, and kidney diseases. 1-3 Their use in acute illness, including patients with COVID-19 4,5 or acute heart failure 6,7 and immediately after experiencing myocardial ...

  24. Clinical presentation and outcomes of acute heart failure in the

    Clinical presentation and outcomes of acute heart failure in the critically ill patient: A prospective, observational, multicentre study Med Intensiva (Engl Ed). ... Aims: To assess the clinical profile and factors associated with 30-day mortality in patients with acute heart failure (AHF) admitted to the intensive care unit (ICU).

  25. Clinical importance of distinguishing true anemia from dilutional

    1 INTRODUCTION. A major concomitant of chronic heart failure (CHF) is volume overload, affecting the interstitial space as well as the blood volume (BV), leading to both peripheral edema and intravascular congestion with concomitant organ damage. 1, 2 The increase in BV is usually due to the expansion of plasma volume (PV), whereas red cell volume (RCV) may be decreased, constant, or increased ...

  26. Acute heart failure: clinical presentation, one-year mortality and

    Aims: Acute heart failure (HF) is a common but ill-defined clinical entity. We describe patients hospitalised with acute HF in regard of clinical presentation, mortality, and risk factors for an unfavourable outcome. Methods and results: We conducted a prospective study including 312 consecutive patients from two European centers hospitalised ...