Skip to main content

Emerging biomarkers for the detection of cardiovascular diseases

Abstract

Background

The prevalence of cardiovascular disease (CVD) has been continuously increasing, and this trend is projected to continue. CVD is rapidly becoming a significant public health issue. Every year there is a spike in hospital cases of CVD, a critical health concern in lower- and middle-income countries. Based on identification of novel biomarkers, it would be necessary to study and evaluate the diagnostic requirements or CVD to expedite early detection.

Main body

The literature review was written using a wide range of sources, such as well-known medical journals, electronic databases, manuscripts, texts, and other writings from the university library. After that, we analysed the specific markers of CVD and compiled a systematic review. A growing body of clinical research aims to identify people who are at risk for cardiovascular disease by looking for biomolecules. A small number of biomarkers have been shown to be useful and reliable in medicine. Biomarkers can be used for a variety of clinical applications, such as predicting heart disease risk, diagnosing disease, or predicting outcomes. As a result of the ability for a single molecule to act as a biomarker, its usefulness in medicine is expected to increase significantly.

Conclusions

Based on assessing the current trends in the application of CVD markers, we discussed and described the requirements for the application of CVD biomarkers in coronary heart disease, cerebrovascular disease, rheumatic heart disease, and other cardiovascular illnesses. Furthermore, the current review focuses on biomarkers for CVD and the procedures that should be considered to establish the comprehensive nature of the expression of biomarkers for cardiovascular illness.

Background

Coronary heart disease (CAD), cerebrovascular disease (CD), rheumatic heart disease (RHD), and other cardiovascular illnesses are non-communicable diseases and leading causes of global mortality [1]. CVD produces more fatalities yearly than any other disease, accounting for 17.9 million deaths per annum, which is about 31% of all worldwide deaths. Coronary heart disease claimed the lives of 7.3 million people, while stroke claimed the lives of 6.2 million [2]. By 2030, 23.6 million individuals will have died from CVD, principally heart disease and stroke. Biomarkers are used to diagnose and monitor cardiovascular disease to reduce the mortality rate [3].

Cardiac biomarkers are bio-molecular components obtained through venal punctures and secreted in the blood during an injured or strained heart [4]. Heart biomarker tests may assist in evaluating a person's risk of cardiac illnesses and monitoring and treating someone suspected of having an acute coronary syndrome (ACS) and cardiac ischemia. Biomarkers are a powerful tool for identifying high-risk people, promptly and reliably diagnosing illness conditions, and effectively diagnosing and treating patients [5, 6]. Most other biomarkers seem to be up-regulated or down-regulated in illnesses other than CVD. Current trends in identifying the number of biomarkers that may consistently aid prognosis in the early stages of the disease are limited. This review provides an overview of commonly used CVD biomarkers based on the carbon skeleton from which they are formed. Some of their merits and drawbacks, proposing a method for future, innovative biomarker research are shown in Fig. 1 and Table 1.

Fig. 1
figure 1

Schematic illustration of emerging biomarkers for the cardiovascular diseases respective to the biomolecular information currently available in the literature

Table 1 Notable emergent biomarkers in cardiovascular diseases

Main text

Carbohydrate-based biomarkers

Glycogen phosphorylase BB (GPBB)

The heart is highly susceptible to CO (carbon monoxide)-induced hypoxia due to its high oxygen demand. Due to a lack of apparent symptoms or specific electrocardiogram (ECG) changes, CO poisoning in cardiovascular patients may go undiagnosed and untreated. The importance of early and particular markers for myocardial hypoxia is growing. One candidate is the glycogen phosphorylase BB isoenzyme (GPBB) [7,8,9]. GPBB is a novel cardiac marker that might help identify myocardial ischemia early [10,11,12]. In hypoglycaemia or hypoxic settings, GPBB is an important enzyme that catalyses the first rate-limiting step in the degradation of glycogen to glucose 1-phosphate [13, 14].

In humans, three distinct isoforms have been identified: GPMM (present in muscles), GPLI (liver), and GPBB (brain). The diagnostic sensitivity of myoglobin and ischemia-modified albumin (IMA) is good. However, there is no heart muscle specificity [15]. GPBB is a more sensitive marker for diagnosing ACS within 4 h after the beginning of chest pain [16, 17], with a sensitivity of 81% and a specificity of 93%. This research revealed that GPBB is an excellent clinical marker of ischemia and may be used to diagnose ACS [15]. Although there is an expected increase in plasma GPBB concentration during pregnancy, which is also seen in the brain, there is no absolute specificity for identifying the cardiac injury. GPBB in plasma is a surrogate marker for acute coronary syndromes in many investigations [11, 16, 18]. As a result, GPBB may be employed as a secondary biomarker for detecting acute myocardial infarction [12].

Protein-based biomarkers

Aspartate aminotransferase (AST)

Aspartate aminotransferases, also called liver transaminases, are intracellular enzymes catalysing the transamination reaction between amino acids and ketonic acid [19]. Because it is present mainly in hepatocytes, AST is a commonly used serum marker of liver disease. Recent research has shown that liver transaminases can independently predict cardiac-related morbidity and mortality. The first cardiac biomarker, AST, was discovered in the liver, heart, skeletal muscle, brain, and kidney in 1954. It does not help diagnose AMI due to the lack of specificity of cardiac tissue [20, 21].

Total CK levels were tested for AMI in 1959 because it was an excellent indicator of skeletal muscle damage[22]. In 1960, LDH was used to diagnose AMI [23]. Finally, in 1979, the World Health Organization (WHO) approved a panel of CK, AST, and LDH to diagnose AMI [24].

Lactate dehydrogenase (LDH)

The intracellular enzyme lactate dehydrogenase (LDH), which is widely distributed, catalyses the conversion of pyruvic acid, the final by-product of the glycolytic process, to lactate. Since the 1960s, LDH has been employed to diagnose acute myocardial infarction as a marker of heart dysfunction. LDH is composed of two self-contained subunits, resulting in the formation of five isozymes. Each isozyme is expressed in a distinct organ: LDH1 is expressed in cardiomyocytes, LDH3 is found in lung tissue, and LDH5 is expressed in hepatocytes [25, 26]. Many organs, including skeletal muscle, kidney, liver, heart, lung, and erythrocytes, express LDH. LDH contains five isoenzymes, one of which is found in nature. Although it has been reported that an LDH1-to-LDH2 ratio higher than 1 is specific for AMI, it is presently not employed in diagnosing AMI [27]. LDH is now only used to differentiate acute from subacute myocardial infarction (MI) in patients with positive troponins whose CK and CK-MB readings have reverted to normal levels when they arrive at the hospital late in the disease [28].

Carbonic anhydrase III (CA-III)

Red blood cell primary protein constituents, carbonic anhydrases (CAs), are zinc-containing metalloproteinases that accelerate the hydration and dehydration of carbon dioxide [29]. CAs are found in various tissues and participate in multiple physiological functions, including gluconeogenesis, acid–base homeostasis, adipogenesis, and calcification [30]. CA-III protein is a unique member of the CAs family; it has a role in various disorders, including type-2 diabetes, MI, and skeletal muscle injury. CA-III protein is a cytosolic enzyme with lower activity than carbon dioxide hydratase [31]. It has been shown to scavenge oxygen radicals in vivo and increase survival from oxidative injury. The cytosolic isoform CA-III is abundant in skeletal muscles [32]. This enzyme possesses antioxidative properties and is critical for maintaining intracellular pH homeostasis. Given that high blood CA-III levels, often in conjunction with other muscle-specific proteins, are commonly used to diagnose overall muscle injury, it seemed worthwhile to review current CA-III investigations [33].

Brain natriuretic peptide (BNP and NT-pro BNP)

As a result of myocardial stretching, BNP and 32 amino acid (-a.a) N-terminal (NT)-proBNP are continually released from the ventricles into the bloodstream. The prohormone brain natriuretic peptide (proBNP), a 108-a.a peptide cleaved into the 76-a.a. BNP and 32-a.a. NT-proBNP, is produced as a result [34]. A sixfold increase in half-life compared to that of BNP is seen, and the peptide is eliminated from the body by many processes, including those involving the renal system and other systems [35]. BNP and NT-proBNP are high in individuals with heart failure, making them valuable diagnostic indicators for the disease. Brain natriuretic peptide (BNP) regulates several physiological and pathological processes [36]. BNP has been demonstrated to be involved in natriuresis, diuresis, vasodilatation, and inhibition of the sympathetic nervous system in addition to the renin–angiotensin–aldosterone system [37, 38]. Studies have shown that the function of BNP also influences diseases associated with the pathways mentioned above to a certain extent [39].

Mid-regional pro-a-type natriuretic peptide (MR-pro-ANP)

Natriuretic peptides are the gold standard in heart failure biomarkers and have been thoroughly researched in many clinical settings [40, 41]. In particular, B-type natriuretic peptide (BNP) is generated from scratch in response to ventricular stretch brought on by pressure and volume overload in both healthy adults and patients with left ventricular (LV) dysfunction [42]. N-terminal pro-B-type natriuretic peptide (NT-proBNP) is an essential cardiac biomarker that correlates well with left ventricular end-diastolic pressure and wall pressure. It has high sensitivity and specificity for the differential diagnosis of HF in patients with acute dyspnoea and has significant value in the diagnosis, evaluation, and prognosis of patients with either acute or chronic HF. Recently, a mid-regional sequence of pro-a-type natriuretic peptide (MR-pro-ANP), which is a more stable intermediate of the natriuretic peptides, was effectively employed in the clinical as a biomarker of prognosis and diagnosis of acute HF [18, 43].

Mid-regional pro-adrenomedullin (MR-proADM)

Adrenomedullin (ADM) is a vasoactive peptide in a human pheochromocytoma3 tumour. Even though a few organs and tissues produce ADM, it is principally synthesized by vascular endothelial cells and has a wide range of physiological functions [44, 45]. Patients with hypertension, congestive heart failure or myocardial infarction, renal disorders (including kidney failure and diabetes mellitus), acute phase of stroke (including septic shock), and arterial stiffness have been observed to have higher plasma ADM levels (including arterial stiffness). The magnitude of the increase is proportional to the severity of the vascular damage [46]. Sepsis, acute dyspnoea, acute and severe chronic HF with decreased ejection fraction (HFREF), and HF with preserved ejection fraction are all illnesses that are associated with higher levels of the neurohumoral marker of cardiac biochemical stress known as MR-pro-ADM (HFPEF) [47]. There is evidence that MR-proADM is an early predictor of in-hospital mortality owing to various causes, including respiratory infections, surgical procedures, and cardiovascular illness [48, 49].

Endothelin-1 (CT-proET-1)

Endothelin-1, a potent vasoconstrictor and potentiator of sympathetic neurohormones generated in blood vessel endothelial cells, is also a sympathetic activation marker. In heart failure, plasma levels of both endothelin-1 and large endothelin-1 are elevated and are linked to pulmonary artery pressure, disease severity, and death [50, 51]. ET-1 and adrenomedullin (ADM) levels in the blood are more significant in people with heart failure than in healthy individuals. It is believed that endothelin has a role in forming pulmonary vasoconstriction, smooth muscle cell proliferation, and remodelling the pulmonary vascular system [52]. In light of this, our team proposed that C-terminal pro-endothelin-1 (CT-proET-1) and mid-regional pro-adrenomedullin (MR-proADM) levels would be higher in heart failure patients compared to controls, and that the degree of this elevation would be related to cardiopulmonary reserve [53, 54]. C-terminal pro-endothelin-1, a newly discovered stable endothelin-1 surrogate marker, has also been proposed as a prognosticator. Increased levels of CT-proET-1, MR-proANP, and MR-proADM were independently linked with the risk of cardiac mortality or heart failure in a recent investigation on patients with stable coronary artery disease after adjusting for clinical cardiovascular risk factors and ejection fraction [55].

Creatine kinase-MB (CK-MB)

A blood test for creatine kinase has been used to identify myocardial infarction since the 1960s. Because of the discovery of CK-MB, one of three creatine kinase isoenzymes found mainly in the heart, researchers have increased the accuracy of myocardial damage diagnosis and have used it to evaluate acute myocardial infarction (AMI) since the 1970s. CK-MB is one of three isoforms of the creatine kinase enzyme, which is present mainly in the heart but also at minor levels in skeletal muscle [56]. Creatine kinase levels are often elevated in myocardial infarction. As a result, the CK-MB is used to determine whether the elevated creatinine level is due to skeletal or cardiac muscle damage [57].

Hydroxybutyrate dehydrogenase (HBDH)

The determination of serum hydroxybutyrate dehydrogenase (HBD) activity may be used to diagnose myocardial infarction without the necessity for isoenzyme separation [58, 59]. Raised serum HBDH activity is uncommon in liver disorders, as one would assume [60]. HBDH levels rise in the blood for 8 to 10 h after a heart attack and then rise again in 2–4 days. They may stay high for up to 18 days, which is a considerable period compared to other heart attack indications [60].

Heart-fatty acid binding protein (H-FABP)

Heart-fatty acid binding protein (H-FABP) is a low molecular weight cytosolic protein in cardiac tissues that transports fatty acids from the plasma membrane to oxidation sites in mitochondria and peroxisomes, as well as to the endoplasmic reticulum for lipid synthesis [61]. It is primarily found in the myocardium and the brain, kidneys, and skeletal muscle to a lesser level. H-FABP is released into the serum after myocyte rupture very early [62]. H-FABP levels rise as soon as 30 min after myocardial infarction, peak at 6–8 h, and then return to baseline after around 24 h [63]. Aside from that, H-FABP has the potential to be employed as a predictive biomarker of mortality [64].

Cardiac Troponin

Cardiac troponin is a calcium-dependent regulator of the heart muscle's contraction mechanism, frequently referred to as a "switch" between contraction and relaxation. For myocardial infarction, cardiac troponin is presently the gold standard biomarker test. Troponin T (TNT), troponin I (TNI), and troponin C (TNC) are the three components of the cardiac troponin complex. Because the amino acid sequences of skeletal and cardiac isoforms of TNT and TNI differ significantly, monoclonal antibody-based immunoassays have been designed to detect cardiac-specific TNT and cardiac-specific TNI.

ACS patients benefit from cardiac troponins because they are sensitive, specific, and offer symptomatic information. As a result, troponin is now the preferred cardiac marker in individuals with ACS. They are also more susceptible to minor degrees of myocardial injury than CK-MB tests. Even though it is the gold standard marker for myocardial infarction, it has minimal usefulness in the diagnosis of reinfarction. A second myocardial infarction that occurs 7 to 10 days after the first may be difficult to identify with cardiac troponin tests alone due to its extended half-life [65].

C reactive protein (CRP)

The liver secretes C reactive protein (CRP), one of the acute phase proteins. In reaction to numerous inflammatory cytokines, their content in the blood is quickly growing [66]. More evidence has been gained in recent years concerning the likely involvement of inflammation in developing significant health disorders such as diabetes or heart disease. According to many epidemiological studies, CRP is a powerful independent predictor of future cardiac events, including myocardial infarction, ischemic stroke, peripheral vascular disease without a known heart disease, and sudden cardiac death [67]. This protein is significant because it occurs early in the bloodstream during numerous systemic inflammatory disorders. Lower stability, a lengthy detection period, and cross-reactivity with other blood proteins are some drawbacks in diagnosing this protein [68].

Tumour necrosis factor (TNF-α)

Tumour necrosis factor (TNF-α) is an inflammatory cytokine that is elevated in chronic HF patients [69]. TNF-α has also been linked to aberrant left atrial dysfunction and advanced left ventricular diastolic and systolic dysfunction disease severity in patients with newly diagnosed HF [70], as well as the use of TNF-α as a predictor of death in advanced HF patients [71].

Interleukin-6 (IL-6)

Interleukin-6 (IL-6) is a cytokine that has a role in inflammation, but it also has cardiovascular effects via regulating cardiomyocyte hypertrophy and apoptosis [72]. Cardiac IL-6 expression has been shown to rise in advanced HF, indicating that it may have a role in prognosis [73]. Furthermore, elevated IL-6 levels have been linked to left ventricular dysfunction before HF diagnosis, suggesting that it might be used as a risk marker for the start and progression of HF [74]. This finding suggested that IL-6 may play a role in distinguishing between the decompensated and compensated phases [75].

Pentraxin 3 (PTX-3)

In response to an inflammatory stimulus, vascular endothelial cells, smooth muscle cells, macrophages, and neutrophils generate PTX-3, a unique marker of vascular inflammation [76]. In patients with unstable angina pectoris, myocardial infarction, and heart failure, the PTX-3 level has been recommended as a predictive biomarker of poor prognosis [77, 78]. PTX-3, on the other hand, predicts advanced atherosclerosis and is more specific for artery wall inflammation than CRP [79].

Pregnancy-associated plasma protein A (PAPPA)

A metalloproteinase called pregnancy-associated plasma protein A (PAPPA) is crucial for the rupture of atherosclerotic plaque [80]. The leading producers are the placental syncytiotrophoblast, fibroblasts, vascular endothelial cells, and vascular smooth muscle cells. It has been linked to plaque progression and instability in atherosclerosis [81].

A soluble cluster of differentiation 40 ligands (SCD40l)

The TNF-α family member SCD-40 l is increased on platelets in the intraluminal thrombus. The release of CD40L into the circulation occurs when the inflammatory and coagulant pathways are activated during thrombogenesis, signifying plaque rupture and eventually myocardial infarction [82].

Copeptin

The cleavage of pre-PR vasopressin yields copeptin, neurophysin ii, and vasopressin, the latter of which is sometimes referred to as antidiuretic hormone and is crucial for fluid balance and has been associated with the severity of HF [83]. Vasopressin is known to be unstable in circulation, making clinical measures difficult. On the other hand, copeptin is thought to have better stability and is produced in equimolar amounts to vasopressin, making it a more dependable and repeatable option for indirect vasopressin detection [84]. Copeptin may thus be utilized as a diagnostic and prognostic marker for myocardial infarction [85].

Growth differentiation factor-15 (GDF-15)

Growth differentiation factor-15 (GDF-15) is a member of the transforming growth factor (TGF) family of cytokines primarily produced by the placenta and implicated in tissue damage inflammatory and apoptotic pathways. GDF-15 levels over a certain threshold have been proven to indicate high-risk individuals with cardiovascular disease [51, 86,87,88]. Increased GDG-15 levels have been linked to the promotion of protective mechanisms for inhibiting apoptosis, hypertrophy, and adverse remodelling in patients with chronic HF, with increased expression of GDF-15 related to the advancement of protective means for inhibition of apoptosis, hypertrophy, and adverse remodelling [89].

F2 Isoprostanes

Arachidonic acid metabolism produces F2 isoprostanes. F2 isoprostanes are released by a few cells, including monocytes, during atherosclerosis. The amount of F2 isoprostane in the urine of individuals with unstable angina is higher in studies. It may also be utilized as a predictor of complications in non-fatal myocardial infarction, the progression of heart failure, and fatality [90].

Suppression of tumorigenicity 2 (ST2)

Suppression of tumorigenicity 2 (ST2) is an interleukin (IL)-1 receptor that has been discovered as a target for IL-33, a cytokine produced when the heart is under metabolic stress [90]. ST2 is a soluble protein produced by cardiomyocytes and detectable in the bloodstream. Its levels rise in response to mechanical stress on the heart [91]. Its utility in HF has been recognized as an independent predictor of mortality or the need for transplantation in patients with severe chronic HF [92], as providing prognostic information for patients with acute HF when combined with natriuretic peptides [93], implying that it could be used as a biomarker in conjunction with current clinical testing strategies like BNP and NTPROBNP.

Matrix metalloproteinases (MMP) and tissue inhibitors of metalloproteinases (TIMP)

Cardiac extracellular matrix remodelling is an essential mechanism in the development and progression of HF. This remodelling is partly mediated by collagenases, MMP, and TIMP, which degrade collagen and other matrix proteins [94]. Fibrosis and ventricular remodelling may result from an imbalance between MMPS and TIMPS. Excessive collagen crosslinking has been linked to HF hospitalization in hypertensive individuals, according to a new study [95]. George et al. revealed that MMP-2 (but not MMP-3, -9, or TIMP-1) concentrations were an independent predictor of death in patients with HF in the first MMPS investigations [96]. MMP-3 and MMP-9 were univariate indicators of all-cause mortality in patients with HFREF. Only MMP-9 emerged as a poor independent prognosis, according to Buralli et al. [97]. TIMP-1, but not MMP-9, was shown to be of separate and additional value in Frantz et al.'s prediction of all-cause mortality [98, 99].

Galectin-3

Galectin-3 is implicated in tissue healing, myofibroblast proliferation, and fibrogenesis in the inflammatory pathway after injury and ventricular remodelling. Galectin-3 instillation into the pericardium resulted in a considerable increase in collagen deposition [100]. Galectin-3 is elevated in patients with acute or chronic HF, and univariable analyses are frequently associated with risk, but with adjustment for renal function or other biomarkers, galectin-3 loses its prognostic meaning in many studies [101, 102]. Serial galectin-3 measurements in chronic HF patients may contribute to a single measure, but no known medications may change galectin-3 readings at this time [103].

Procalcitonin

Procalcitonin is a pro-peptide of calcitonin typically produced by parafollicular C cells in the thyroid gland [104]. Calcium homeostasis and immunity are regulated by pro-calcitonin/calcitonin axed [105]. There is insufficient proof that serum procalcitonin levels are a reliable marker for chronic HF with predictive value, although serial procalcitonin measurements are advised to differentiate between in-hospital mortality in a variety of diseases associated with pro-inflammatory activation (pneumonia, chronic obstructive pulmonary disease, acute respiratory tract infections, etc.) [106]. Extensive clinical studies are needed to fully clarify the hypothesis and gather further information about the predictive significance of pro-calcitonin in HF patients with worsened symptoms.

Comorbidities

Renal dysfunction, hematologic abnormalities, and liver dysfunction are all critical indicators of a poor prognosis in heart failure. While serum creatinine, estimated glomerular filtration rate, and blood urea nitrogen are essential indices of renal function, NT-PROBNP provides additional prognostic information [107]. Their predictive significance is lower at lower levels of impairment, and the commencement of these biomarkers' increase is often delayed following acute renal injury (AKI). Cystatin C and trace protein (BPT) outperformed standard renal indicators in predicting prognosis in HF, owing to its improved capacity to assess renal function at lower degrees of abnormalities [108].

While high levels of NAGL are linked to poor clinical outcomes, this link was lost after controlling for many factors, including NT-PROBNP [109]. Furthermore, it has a reasonable capacity to predict impending AKI (68%sensitivity and 70% specificity) [110]. KIM-1 in the urine and NAG in the blood showed potential, but further research is required. Hematologic abnormalities, including anaemia, iron deficiency, and increased red blood cell distribution width (RDW), are common in HF patients and play an essential role in predicting prognosis, independent of LVEF [111]. It is uncertain if correcting hematologic abnormalities can enhance clinical outcomes in HF patients.

Myeloperoxidase (MPO)

Myeloperoxidase (MPO) is an enzyme secreted into an extracellular fluid by polymorphonuclear neutrophils and macrophages' azurophilic granules during inflammation. A valuable risk marker and diagnostic tool in acute coronary syndrome and chest discomfort are myeloperoxidase, which is implicated in oxidative stress and inflammation [112]. High concentrations of MPO, irrespective of traditional CVD risk factors, have been demonstrated to enhance CVD risk in epidemiological investigations [113, 114]. MPO is implicated in all phases of endotheliosis, from endothelial dysfunction to plaque rupture, according to in vitro studies, and it is involved in the initiation of CVD and acute cardiovascular events [115]. Since both myoglobin and haemoglobin have peroxidase activity and the substrate for myeloperoxidase (MPO) and general peroxidase is the same, detecting them is problematic [116].

Fibrinogen

In 1836, fibrinogen was identified and characterized as the first clotting agent [117]. Cells participating in the atherogenic process may produce cytokines, which trigger an acute phase response that raises plasma fibrinogen levels [118]. Fibrinogen is a significant acute phase protein that creates the substrate for thrombin and constitutes the last stage in the coagulation cascade [119]. According to research, fibrinogen is a significant and independent cardiovascular risk factor related to established risk factors and genetic variants. Despite unsolved difficulties requiring definitive solutions, fibrinogen has emerged as an essential new predictor of cardiac risk [120] since we may detect low fibrinogen levels in individuals with severe hepatic illness [121].

Trimethylamine n-oxide (TMAO)

Trimethylamine n-oxide (TMAO) is a dietary metabolite linked to red meat consumption. Liver enzymes produce TMAO from the gaseous precursor trimethylamine (TMA), which is created when gut bacteria break down these nutrients [122]. In all diets, red meat has the highest concentration of l-carnitine, increasing the quantity of l-carnitine accessible for TMAO production. While a recent study has proven that high levels of TMAO in the blood are linked to an elevated risk of heart disease and death, it shows that TMAO may be used as a biomarker for both heart and kidney illness [123].

Cystatin C

Most nucleated cells regularly produce and release cystatin C, a tiny protease inhibitor. It is a well-described blood marker of renal failure in clinical practice. In addition, because of decreased renal function and the relationship between cardiac risk and illnesses, cystatin C is being studied as a cardiac marker [124].

Myoglobin

Myoglobin is an oxygen-binding protein that forms a compound with iron molecules in skeletal and cardiac muscle tissues. One hour after myocardial infarction, it is released into circulation. Within 2 to 4 h after the commencement of myocardial damage, it steadily rises, peaks in 8–12 h, and recovers to normal in 24–36 h. Due to substantial levels of myoglobin in skeletal muscles, myoglobin is an early indication of acute myocardial infarction but lacks cardiac specificity. Myoglobin should not be used alone to confirm a diagnosis of acute myocardial infarction; instead, it should be combined with other heart-specific markers like TNI or TNT to boost diagnostic value [125].

Ischemia-modified albumin (IMA)

When serum albumin from ischemic cardiac tissues comes into touch with IMA, it becomes a new sign of ischemia. The albumin cobalt binding test, based on the inability of IMA to bind cobalt, has been developed and commercialized as the first commercially accessible FDA-approved diagnostic of myocardial ischemia with a 30-min laboratory turnaround time [126]. IMA may distinguish ischemic and non-ischemic patients. Because IMA is elevated in illnesses other than chest pain, its ineffectiveness calls into doubt its specificity. Even while IMA assessment remains the sole clinical biomarker for myocardial ischemia, the process of IMA generation and the precise unit being evaluated are unknown [127].

Endothelial dysfunction (noninvasive method)

Endothelial cells are a monolayer of active cells lining the inside blood and lymphatic arteries. It acts as an active paracrine, endocrine, and autocrine organ to maintain vascular homeostasis [128]. It controls oxidative stress, as well as both local and systemic inflammation. According to current research on atherosclerosis and the basic mechanisms involved, harmful changes in endothelial physiology and biology, also known as endothelial dysfunction, may be a critical early step in the development of atherosclerosis and may also signal lesion progression and the occurrence of complications [129, 130]. Endothelial dysfunction is an independent predictor of poor cardiovascular outcomes and may have more predictive power than established risk variables alone in predicting future cardiovascular events [131].

Apoptosis antigen-1 (APO1/FAS)

FAS, also known as apoptosis antigen-1 (APO1), is a TNF family member that regulates apoptosis. FAS is present in the thymus, liver, ovary, lung, and heart, among other organs [132, 133]. FAS is increased in cardiomyocytes due to apoptotic cell death caused by hypoxia and overstretching [134]. A soluble version of FAS (SFAS) missing the transmembrane domain was also discovered in human serum (133), and its levels are expected to be higher in individuals with cardiovascular disease. Individuals with severe CHF had considerably greater SFAS levels in their blood than patients with moderate CHF. Furthermore, stepwise multivariate analysis revealed that individuals with high SFAS and BNP and a poor ejection fraction had a substantially greater mortality rate than those with low SFAS levels, indicating that plasma SFAS is a helpful prognostic marker in CHF aetiology.

Neutrophil gelatinase-associated lipocalin (NGAL)

Neutrophil gelatinase-associated lipocalin (NGAL) is a lipocalin protein having a polypeptide chain of 178-a.a and a molecular mass of 25 KDa expressed by neutrophils and different epithelial cells express [135, 136]. NGAL expression is elevated under acute phase protein [137, 138]. NAGL seems to be one of the first kidney indicators of nephrotoxic and ischemia damage in animal models, and it is shown to be high in the urine and blood of people shortly after acute injury (AKI) [138]. Compared to control participants, chronic HF patients had more significant amounts of NGAL in both blood and urine [139]. NGAL seems to function as a biomarker in the case of acute HF, according to research by Damman et al. As a result, further study into the function of NGAL in acute and chronic heart failure might be beneficial [140].

Uric acid (UA)

In humans, uric acid (UA) is the final product of purine metabolism. The inactivation of uricase and the subsequent rise in UA levels are hypothesized to have offered evolutionary benefits by guarding against oxidative damage [141]. Increased serum UA, even below the clinical threshold for hyperuricemia, has been linked to the development of CVD through increasing oxidative stress, inducing endothelial dysfunction, and boosting inflammation [142]. Recent research has shown an independent link between UA and cardiovascular mortality [143, 144]. However, the evidence supporting the outcomes is still inconclusive. Numerous epidemiological investigations, including prospective, retrospective, cross-sectional, and meta-analysis studies, have failed to find an independent link between UA and CVD [145, 146].

Neuregulin-1 (NRG-1)

Endothelial cells emit neuregulin-1 (NRG-1), which binds to a family of ERBB receptors on neighbouring cardiac myocytes to enhance cell survival, growth, and maintenance [147]. The NRG-1 gene has been linked to more than 15 distinct protein products. The most prevalent NRG-1 protein in the cardiac system is NRG-1. The paracrine impact of the NRG-1 ligand is mediated by the ERBB (ERBB2, ERBB3, ERBB4) tyrosine kinase receptor family. NRG-1 is activated by various cardiovascular stressors, including oxidative stress, ischemia, and exercise [148]. As a result of NRG-1/ERBB4 paracrine signalling in the heart, this system may be engaged in cardiac adaptation to different types of physiologic stress. NRG-1 levels were associated with more advanced stages of HF and a poorer outcome in HF patients with CAD [149]. Increased blood levels of NRG, like NT-PROBNP, may be an insufficient physiologic response to cardiovascular injury, and exogenous treatment of NRG may enhance cardiovascular function. These data support the theory that NRG-1 in the myocardium is activated in response to ischemia. NRG-1's potential as a CVD biomarker needs further investigation.

Human serum albumin (HAS)

Albumin is a significant protein in human plasma made by liver cells. Albumin has a variety of activities, including controlling fluid filtration and adsorption across capillary walls and transporting various chemicals in the bloodstream. Albumin concentrations are often evaluated in blood or urine and may be used as indicators of dehydration, malnutrition, liver or kidney illness, and cardiovascular disease [150].

Serum amyloid A (SAA)

The serum amyloid A (SAA) protein belongs to a group of apolipoproteins that are usually linked to high-density lipoproteins (HDL). When interleukin-6 (IL-6) is stimulated in response to infection, inflammation, injury, or stress, the acute phase SAA proteins SAA1 and SAA2 are released into the blood. Binding phase proteins are being examined extensively as possible indicators for predicting the progression of cardiovascular illnesses, and SAA has shown to be a promising candidate for a cardiac biomarker [151, 152].

Retinol-binding protein 4 (RBP4)

Retinol-binding protein 4 (RBP4) is a member of the lipocalin protein family that serves as a carrier protein for vitamin A in serum. RBP4 has been shown to have a critical role in insulin resistance. Several studies have recently revealed that RBP4 levels in the blood may be linked to cardiovascular disease and metabolic syndrome [153, 154].

Soluble lectin-like oxidized LDL receptor (SLOX-1)

The extracellular domain of LOX-1 is proteolytically cleaved to form the soluble lectin-like oxidized LDL (SLOX-1) receptor. LOX-1 is a transmembrane protein present on the cell surface of endothelial cells and smooth muscle cells, among other places. SLOX-1 has been suggested as a biomarker for plaque rupture in a few studies, and it may have clinical usefulness in detecting atherosclerosis-related disorders [155, 156].

Adiponectin (ADN)

Adipokines are a kind of protein hormone abundant in the human body. It is a crucial lipid and glucose metabolism regulator predominantly expressed by adipocytes. According to research, ADN is an insulin sensitizing hormone with anti-diabetic, anti-inflammatory, and anti-atherogenic characteristics. ADN levels in the blood have been linked to various lifestyle disorders, including atherosclerosis and heart failure [157, 158].

S100 protein

S100 proteins have a molecular weight of 10 to 12 KDa and are acidic calcium-binding proteins. In humans, there are over 20 distinct members of this family. These proteins seem to be involved in various biological activities, including cell proliferation and differentiation, as well as the inflammatory response. The involvement of several family members (S100B and S100A) in heart illness has been investigated [159].

Lipid derived biomarkers

Triglyceride-to-high-density lipoprotein cholesterol ratio

Triglyceride (TG) to high-density lipoprotein cholesterol ratio (HDLC) and total cholesterol (TC)-to-HDLC ratio, as well as a low ankle branchial pressure index (ABPI), are vital biomarkers for CVD [160]. Although triglyceride levels may be seen in most routine lipid panels, they can also be utilized to predict CVD risk by combining them with other lipid profile indicators [160]. A triglyceride-to-HDL cholesterol ratio of more than 3.5 seems to be linked to an elevated risk of cardiovascular disease. Apolipoprotein B (APO-B), lipoprotein A (LPA), high-density lipoprotein cholesterol (HDL-C), and low-density lipoprotein cholesterol (LDL-C) elevations are reported during HF. Important biochemical markers include a considerable drop in apolipoprotein A1 (APOA1), APOA1/APOB ratio, and PON1 activity/HDLC ratio [161]. People with type 2 diabetes mellitus (T2DM) who have higher levels of LDL, triglycerides, and total cholesterol (hyperlipidemia) have a greater risk of coronary artery disease (168). Total cholesterol, LDL-C, APOB, LP(A), and the APOA1/APOB proportions are also excellent biomarkers for predicting CVDs in people [162].

Low-density lipoprotein cholesterol

A high plasma concentration of low-density lipoprotein cholesterol (LDLC) is a crucial risk factor for atherosclerotic cardiovascular disease (ACD). Furthermore, a rapid increase in LDLC plasma content suggests a significant risk of cardiovascular disease and necessitates faster therapeutic intervention [163]. Similarly, autosomal dominant hypercholesterolemia (ADH) is characterized by high LDL cholesterol, which causes early morbidity and death due to atherosclerotic cardiovascular disease (ASCVD). Similarly, familial combination hyperlipidemia (FCHL) is characterized by elevated total cholesterol, triglycerides, or LDL cholesterol and a significant risk of ASCVD [164]. LRP1 (low-density lipoprotein receptor-related protein 1) is a large endocytic and signalling receptor mainly found in the liver, where it regulates blood coagulation factor VIII plasma levels by controlling its absorption and subsequent degradation. Similarly, tiny dense LDL cholesterol is a new risk factor for cardiovascular disease [165].

Lipoprotein-associated phospholipase A2 (LP-PLA2)

Lipoprotein-associated phospholipase A2 (LP-PLA2) is an enzyme that functions in blood vessel inflammation and is thought to contribute to atherosclerosis. This test determines how much LP-PLA2 is present in the blood. Recent research has shown that LP-PLA2 is a standalone risk factor for cardiovascular disease (CVD), including coronary heart disease (CHD) and ischemic stroke [166]. The sensitivity C-reactive protein (HS-CRP) test is linked to systemic inflammation, and higher levels are linked to an increased risk of cardiovascular disease. The LP-PLA2 test is linked to vascular inflammation and is considered to raise the risk of cardiac events, such as heart attack or stroke, at high levels. In persons with diabetes, a high level of LP-PLA2 is linked to vascular problems such as kidney damage [167]. Although it is a risk factor for heart disease, decreasing LP-PLA2 activity with medication is not advised for people with stable CHD [168].

Oxylipin

Oxylipins are bioactive lipid metabolites formed when polyunsaturated fatty acids are oxidized by the cyclooxygenase (COX), lipoxygenase, and cytochrome p450 pathways. They have a role in immunological, inflammatory, and vascular activities [169]. Abnormal oxylipin signalling has also been associated with CVD risk factors such as hypertension, hyperlipidemia, thrombosis, haemostasis, and diabetes. Oxylipins are a new age of risk indicators in numerous illnesses, including cardiovascular disease, and various oxylipins have a role in CVD development, albeit the mechanism is yet unknown [170].

Pre-disease biological marker

Hypertension (HTN)

Hypertension is a disease of the modern world. Although it seldom causes symptoms, it affects 16% to 37% of the world's population and is a significant risk factor for coronary artery disease, stroke, HF, and atrial fibrillation (AF) [171]. The available data support the claim that blood pressure (BP) monitoring is a substantial predictor of long-term risk and that it is better than clinic blood pressure measurement for this purpose. As a result, BP measurement combines higher precision with cheap cost and ease of application. According to American and European standards, all patients with known or suspected hypertension should have their BP checked and controlled at home [172, 173].

Nucleic acid-derived biomarkers

MicroRNAs (miRNAs)

MicroRNAs (miRNAs) are noncoding RNA molecules with a short length. They act via the seed region, a six- to an eight-nucleotide sequence that binds to messenger ribonucleic acid (mRNA) known as miRNA targets [174]. They primarily inhibit gene expression by downregulating translation at the post-expression stage, using two fundamental pathways: translational repression and mRNA degradation. The most reliable approach for quantitatively comparing miRNA expression levels is a real-time quantitative polymerase chain reaction, which has been the cornerstone of miRNA quantification. Lipid metabolism, glucose homeostasis, vascular integrity, and endothelial cell function are only a few pathways influenced by miRNA control [175, 176]. Earlier studies revealed that a miRNA panel's combined utility enhanced the predictive value of classic Framingham risk models, but no single miRNA imparted a clinically relevant reduction in acute MI risk [177].

Genome-based biomarkers

Soluble ST2

Over the last decade, more research has revealed that soluble ST2 is substantially more significant in people with heart failure and can aid with risk assessment for both acute and chronic heart failure. The US Food and Drug Administration recently approved a commercial-grade soluble ST2 assay (Presage, Critical Diagnostics, San Diego, CA) for use in chronic heart failure prognosis [49, 178, 179].

Blood gene expression in CAD

The Corus CAD (CardioDx, Palo Alto, CA) assay is now commercially available for this gene expression test. By standard protein biomarkers, the test is confined to patients with chest pain who do not have diabetes, chronic inflammatory illnesses, increased leukocyte levels, or acute coronary syndromes. The clinical efficacy of a transcriptome profile Corus CAD must be extensively validated against these current noninvasive standards and investigated in various groups [180, 181].

DEFA1/DEFA3

DEFA1/DEFA3 were the most strongly linked with CHD development among eight possible markers in a cross-sectional study of CHD prediction biomarkers in non-familial hyperlipidemia patients, and show promise for further application as inflammatory markers to synergistically predict the risk of CHD development in Thai hyperlipidemia patients [182, 183].

Conclusions

Over the last decade, various efforts have been undertaken to discover meaningful biomarkers for cardiovascular disease risk management. The search for the most beneficial and ideal cardiovascular indicators will likely go on forever. Many ancient and innovative substances will be studied as single markers and in multi-marker techniques in the future years. We explored the notion that an ideal cardiovascular biomarker may simultaneously provide clinicians with a range of information concerning predictive, diagnostic, and prognostic implications, in addition to satisfying the conventional biomarker designation requirements. Based on the gathered data, a few circulating molecules seem to be of interest in this regard since they can provide all the clinical data described above. However, further research is needed before an effective platform for biomarker analysis that may replace conventional CVD diagnosis in clinical practice can be provided. Innovative targets are expected to be discovered due to the expansion of public biomarker databases, advancements in the sensitivity and selectivity of analytical techniques, and the development and routine use of innovative platforms with proven potential.

Availability of data and materials

Not applicable.

References

  1. Al-Mawali A (2015) Non-communicable diseases: shining a light on cardiovascular disease, Oman’s biggest killer. Oman Med J 30(4):227

    Article  PubMed  PubMed Central  Google Scholar 

  2. Thomas H, Diamond J, Vieco A, Chaudhuri S, Shinnar E, Cromer S, Perel P, Mensah GA, Narula J, Johnson CO (2018) Global atlas of cardiovascular disease. Glob Heart 13:143–163

    Article  PubMed  Google Scholar 

  3. Sanchis-Gomar F, Perez-Quilis C, Leischik R, Lucia A (2016) Epidemiology of coronary heart disease and acute coronary syndrome. Ann Transl Med 4(13):256

    Article  PubMed  PubMed Central  Google Scholar 

  4. van Kimmenade RR, Januzzi JL Jr (2012) Emerging biomarkers in heart failure. Clin Chem 58(1):127–138

    Article  PubMed  Google Scholar 

  5. Haider A, Bengs S, Luu J, Osto E, Siller-Matula JM, Muka T, Gebhard C (2020) Sex and gender in cardiovascular medicine: presentation and outcomes of acute coronary syndrome. Eur Heart J 41(13):1328–1336

    Article  CAS  PubMed  Google Scholar 

  6. Schwartz GG, Steg PG, Szarek M, Bhatt DL, Bittner VA, Diaz R, Edelberg JM, Goodman SG, Hanotin C, Harrington RA (2018) Alirocumab and cardiovascular outcomes after acute coronary syndrome. N Engl J Med 379(22):2097–2107

    Article  CAS  PubMed  Google Scholar 

  7. AlGawwam G, Shake ME (2021) Evaluation of glycogen phosphorylase BB (GPBB) level in serum as marker in diagnosis of acute ischemic stroke. Indian J Forensic Med Toxicol 15(3):1603–1608

    CAS  Google Scholar 

  8. El-Nagdy SA, Elfakharany Y, Etewa R (2020) Role of glycogen phosphorylase in prediction of cardiotoxicity associated with acute carbon monoxide poisoning. Zagazig J Forensic Med 18(2):59–74

    Article  Google Scholar 

  9. Migocka-Patrzałek M, Elias M (2021) Muscle glycogen phosphorylase and its functional partners in health and disease. Cells 10(4):883

    Article  PubMed  PubMed Central  Google Scholar 

  10. Apple FS, Wu AH, Mair J, Ravkilde J, Panteghini M, Tate J, Pagani F, Christenson RH, Mockel M, Danne O (2005) Future biomarkers for detection of ischemia and risk stratification in acute coronary syndrome. Clin Chem 51(5):810–824

    Article  CAS  PubMed  Google Scholar 

  11. Krause E-G, Rabitzsch G, Noll F, Mair J, Puschendorf B (1996) Glycogen phosphorylase isoenzyme BB in diagnosis of myocardial ischaemic injury and infarction. Mol Cell Biochem 160(161):289–295

    Article  PubMed  Google Scholar 

  12. Singh N, Rathore V, Mahat RK, Rastogi P (2018) Glycogen phosphorylase BB: a more sensitive and specific marker than other cardiac markers for early diagnosis of acute myocardial infarction. Indian J Clin Biochem 33(3):356–360

    Article  CAS  PubMed  Google Scholar 

  13. Newgard CB, Hwang PK, Fletterick RJ (1989) The family of glycogen phosphorylases: structure and functio. Crit Rev Biochem Mol Biol 24(1):69–99

    Article  CAS  PubMed  Google Scholar 

  14. Park K-Y, Ay I, Avery R, Caceres JA, Siket MS, Pontes-Neto OM, Zheng H, Rost NS, Furie KL, Sorensen AG (2018) New biomarker for acute ischaemic stroke: plasma glycogen phosphorylase isoenzyme BB. J Neurol Neurosurg Psychiatry 89(4):404–409

    Article  PubMed  Google Scholar 

  15. Peetz D, Post F, Schinzel H, Schweigert R, Schollmayer C, Steinbach K, Dati F, Noll F, Lackner KJ (2005) Glycogen phosphorylase BB in acute coronary syndromes. Clin Chem Lab Med (CCLM) 43(12):1351–1358

    Article  CAS  Google Scholar 

  16. Rabitzsch G, Mair J, Lechleitner P, Noll F, Hofmann U, Krause E-G, Dienstl F, Puschendorf B (1995) Immunoenzymometric assay of human glycogen phosphorylase isoenzyme BB in diagnosis of ischemic myocardial injury. Clin Chem 41(7):966–978

    Article  CAS  PubMed  Google Scholar 

  17. Rabitzsch G, Mair J, Lechleitner P, Noll F, Hofmann U, Krause E-G, Dienstl F, Puschendorf B (1993) Isoenzyme BB of glycogen phosphorylase b and myocardial infarction. Lancet 341(8851):1032–1033

    Article  CAS  PubMed  Google Scholar 

  18. Wild PS, Schnabel RB, Lubos E, Zeller T, Sinning CR, Keller T, Tzikas S, Lackner KJ, Peetz D, Rupprecht HJ (2012) Midregional proadrenomedullin for prediction of cardiovascular events in coronary artery disease: results from the Athero Gene Study. Clin Chem 58(1):226–236

    Article  CAS  PubMed  Google Scholar 

  19. Parsanathan R, Jain SK (2020) Novel invasive and noninvasive cardiac-specific biomarkers in obesity and cardiovascular diseases. Metab Syndr Relat Disord 18(1):10–30

    Article  PubMed  PubMed Central  Google Scholar 

  20. Dolci A, Panteghini M (2006) The exciting story of cardiac biomarkers: from retrospective detection to gold diagnostic standard for acute myocardial infarction and more. Clin Chim Acta 369(2):179–187

    Article  CAS  PubMed  Google Scholar 

  21. Ladenson JH (2007) A personal history of markers of myocyte injury [myocardial infarction]. Clin Chim Acta 381(1):3–8

    Article  CAS  PubMed  Google Scholar 

  22. Ruzich RS (1992) Cardiac enzymes: how to use serial determinations to confirm acute myocardial infarction. Postgrad Med 92(7):85–92

    Article  CAS  PubMed  Google Scholar 

  23. Lee N, Barthel SR, Schatton T (2014) Melanoma stem cells and metastasis: mimicking hematopoietic cell trafficking? Lab Invest 94(1):13–30

    Article  CAS  PubMed  Google Scholar 

  24. Series CC, Challenge E, Mind OM, Doodle C (1979) Nomenclature and criteria for diagnosis of ischemic heart disease. Report of the Joint International Society and Federation of Cardiology/World Health Organization task force on standardization of clinical nomenclature. Circulation 59(3):607–609

    Article  Google Scholar 

  25. Yamaguchi S, Abe M, Arakaki T, Arasaki O, Shimabukuro M (2020) Prognostic value of lactate dehydrogenase for mid-term mortality in acute decompensated heart failure: a comparison to established biomarkers and brain natriuretic peptide. Heart Lung Circ 29(9):1318–1327

    Article  PubMed  Google Scholar 

  26. Aly M, Nayel M, Salama A, Ghazy E, Elshahawy I (2020) Cardiac troponin I as a cardiac biomarker has prognostic and predictive value for poor survival in Egyptian buffalo calves with foot-and-mouth disease. Vet World 13(5):890

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Schmiechen NJ, Han C, Milzman DP (1997) ED use of rapid lactate to evaluate patients with acute chest pain. Ann Emerg Med 30(5):571–577

    Article  CAS  PubMed  Google Scholar 

  28. Yun DD, Alpert JS (1997) Acute coronary syndromes. Cardiology 88(3):223–237

    Article  CAS  PubMed  Google Scholar 

  29. Silagi ES, Batista P, Shapiro IM, Risbud MV (2018) Expression of carbonic anhydrase III, a nucleus pulposus phenotypic marker, is hypoxia-responsive and confers protection from oxidative stress-induced cell death. Sci Rep 8(1):1–13

    Article  CAS  Google Scholar 

  30. Dowling P, Gargan S, Zweyer M, Sabir H, Swandulla D, Ohlendieck K (2021) Proteomic profiling of carbonic anhydrase CA3 in skeletal muscle. Expert Rev Proteomics 18:1–14

    Article  Google Scholar 

  31. Zamanova S, Shabana AM, Mondal UK, Ilies MA (2019) Carbonic anhydrases as disease markers. Expert Opin Ther Pat 29(7):509–533

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Feng H-Z, Jin J-P (2019) Transgenic expression of carbonic anhydrase III in cardiac muscle demonstrates a mechanism to tolerate acidosis. Am J Physiol Cell Physiol 317(5):C922–C931

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Zhang X-M, Tao Y-H, Zhou X-L, Shang X-L, Gong X-B, Liu Y-C, Huang Y-Y, Chen G, Yu Z-Y, Wang J-T (2021) The role of carbonic anhydrase III and autophagy in type 2 diabetes with cardio-cerebrovascular disease. Metab Brain Dis 36(8):2329–2341

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Kim YS, Karisa N, Jeon WY, Lee H, Kim Y-c, Ahn J (2019) High-level production of N-terminal pro-brain natriuretic peptide, as a calibrant of heart failure diagnosis, Escherichia coli. Appl Microbiol Biotechnol 103(12):4779–4788

    Article  CAS  PubMed  Google Scholar 

  35. York MK, Gupta DK, Reynolds CF, Farber-Eger E, Wells QS, Bachmann KN, Xu M, Harrell FE, Wang TJ (2018) B-type natriuretic peptide levels and mortality in patients with and without heart failure. J Am Coll Cardiol 71(19):2079–2088

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Almeida Junior GLGd, Braga F, Jorge JK, Nobre GF, Kalichsztein M, Faria PdMPd, Bussade B, Penna GL, Alves VO, Pereira MA (2020) Prognostic value of troponin-T and B-type natriuretic peptide in patients hospitalized for COVID-19. Arq Bras Cardiol 115:660–666

    Article  PubMed  Google Scholar 

  37. Ghomian N, Vakilian F, Shahri B, Rostaminejad V, Khadem-Rezaiyan M (2019) Can brain natriuretic peptide predict cardiovascular complications in severe preeclampsia? A case-control study. Int J Reprod BioMed 17(4):271

    CAS  PubMed Central  Google Scholar 

  38. Wong Y-K, Cheung CY, Tang CS, Hai JS, Lee C-H, Lau K-K, Au K-W, Cheung BM, Sham P-C, Xu A (2019) High-sensitivity Troponin I and B-type natriuretic peptide biomarkers for prediction of cardiovascular events in patients with coronary artery disease with and without diabetes mellitus. Cardiovasc Diabetol 18(1):1–12

    Article  Google Scholar 

  39. Li X, Liu C, Mao Z, Qi S, Song R, Zhou F (2020) Brain natriuretic peptide for predicting contrast-induced acute kidney injury in patients with acute coronary syndrome undergoing coronary angiography: a systematic review and Meta-analysis. J Int Cardiol 2020:1035089

    Article  Google Scholar 

  40. Cui K, Huang W, Fan J, Lei H (2018) Midregional pro-atrial natriuretic peptide is a superior biomarker to N-terminal pro-B-type natriuretic peptide in the diagnosis of heart failure patients with preserved ejection fraction. Medicine 97(36):e12277

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Pervez MO, Winther JA, Brynildsen J, Strand H, Christensen G, Høiseth AD, Myhre PL, Røysland R, Lyngbakken MN, Omland T (2018) Prognostic and diagnostic significance of mid-regional pro-atrial natriuretic peptide in acute exacerbation of chronic obstructive pulmonary disease and acute heart failure: data from the ACE 2 Study. Biomarkers 23(7):654–663

    Article  CAS  PubMed  Google Scholar 

  42. Gauffin E, Chisalita SI, Engvall J, Nyström FH, Östgren CJ (2021) Plasma mid-regional pro-atrial natriuretic peptide predicts cardiovascular events in patients with type 2 diabetes independently of subclinical organ damage. Diabetes Res Clin Pract 182:109095

    Article  CAS  PubMed  Google Scholar 

  43. Francis GS, Felker GM, Tang WW (2016) A test in context: critical evaluation of natriuretic peptide testing in heart failure. J Am Coll Cardiol 67(3):330–337

    Article  PubMed  PubMed Central  Google Scholar 

  44. Koyama T, Kuriyama N, Suzuki Y, Saito S, Tanaka R, Iwao M, Tanaka M, Maki T, Itoh H, Ihara M (2021) Mid-regional pro-adrenomedullin is a novel biomarker for arterial stiffness as the criterion for vascular failure in a cross-sectional study. Sci Rep 11(1):1–7

    Google Scholar 

  45. Czajkowska K, Zbroch E, Bielach-Bazyluk A, Mitrosz K, Bujno E, Kakareko K, Rydzewska-Rosolowska A, Hryszko T (2021) Mid-regional proadrenomedullin as a new biomarker of kidney and cardiovascular diseases—is it the future? J Clin Med 10(3):524

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. van Oers JA, Kluiters Y, Bons JA, de Jongh M, Pouwels S, Ramnarain D, de Lange DW, de Grooth H-J, Girbes AR (2021) Endothelium-associated biomarkers mid-regional proadrenomedullin and C-terminal proendothelin-1 have good ability to predict 28-day mortality in critically ill patients with SARS-CoV-2 pneumonia: a prospective cohort study. J Crit Care 66:173–180

    Article  PubMed  PubMed Central  Google Scholar 

  47. Bustamante A, García-Berrocoso T, Penalba A, Giralt D, Simats A, Muchada M, Zapata E, Rubiera M, Montaner J (2017) Sepsis biomarkers reprofiling to predict stroke-associated infections. J Neuroimmunol 312:19–23

    Article  CAS  PubMed  Google Scholar 

  48. Dres M, Hausfater P, Foissac F, Bernard M, Joly L-M, Sebbane M, Philippon A-L, Gil-Jardiné C, Schmidt J, Maignan M (2017) Mid-regional pro-adrenomedullin and copeptin to predict short-term prognosis of COPD exacerbations: a multicenter prospective blinded study. Int J Chron Obstr Pulm Dis 12:1047

    Article  CAS  Google Scholar 

  49. Lopes D, Falcao LM (2017) Mid-regional pro-adrenomedullin and ST2 in heart failure: Contributions to diagnosis and prognosis. Rev Port Cardiol (English Edition) 36(6):465–472

    Article  Google Scholar 

  50. Hülsmann M, Stanek B, Frey B, Sturm B, Putz D, Kos T, Berger R, Woloszczuk W, Maurer G, Pacher R (1998) Value of cardiopulmonary exercise testing and big endothelin plasma levels to predict short-term prognosis of patients with chronic heart failure. J Am Coll Cardiol 32(6):1695–1700

    Article  PubMed  Google Scholar 

  51. Kempf T, Wollert KC (2009) Growth differentiation factor-15: a new biomarker in cardiovascular disease. Herz 34(8):594–599

    Article  PubMed  Google Scholar 

  52. Jankowich M, Choudhary G (2020) Endothelin-1 levels and cardiovascular events. Trends Cardiovasc Med 30(1):1–8

    Article  CAS  PubMed  Google Scholar 

  53. Obokata M, Kane GC, Reddy YN, Melenovsky V, Olson TP, Jarolim P, Borlaug BA (2019) The neurohormonal basis of pulmonary hypertension in heart failure with preserved ejection fraction. Eur Heart J 40(45):3707–3717

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Then C, Thorand B, Then HL, Meisinger C, Heier M, Peters A, Koenig W, Rathmann W, Bidlingmaier M, Lechner A (2020) Serum uromodulin is inversely associated with arterial hypertension and the vasoconstrictive prohormone CT-proET-1 in the population-based KORA F4 study. PLoS ONE 15(8):e0237364

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Sabatine MS, Morrow DA, de Lemos JA, Omland T, Sloan S, Jarolim P, Solomon SD, Pfeffer MA, Braunwald E (2012) Evaluation of multiple biomarkers of cardiovascular stress for risk prediction and guiding medical therapy in patients with stable coronary disease. Circulation 125(2):233–240

    Article  CAS  PubMed  Google Scholar 

  56. Al-Hadi HA, Fox KA (2009) Cardiac markers in the early diagnosis and management of patients with acute coronary syndrome. Sultan Qaboos Univ Med J 9(3):231

    PubMed  PubMed Central  Google Scholar 

  57. Guzy PM (1977) Creatine phosphokinase-MB (CPK-MB) and the diagnosis of myocardial infarction. West J Med 127(6):455

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Elliott B, Jepson E, Wilkinson J (1962) Serum alpha-hydroxybutyrate dehydrogenase-a new test with improved specificity for myocardial lesions. Clin Sci 23(2):305

    CAS  Google Scholar 

  59. Rosalki S (1963) Serum α-hydroxybutyrate dehydrogenase: a new test for myocardial infarction. Br Heart J 25(6):795

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Elliott B, Wilkinson J (1963) Serum alpha-hydroxybutyrate dehydrogenase in diseases other than myocardial infarction. Clin Sci 24(3):343–350

    CAS  Google Scholar 

  61. Chan D, Ng LL (2010) Biomarkers in acute myocardial infarction. BMC Med 8(1):1–11

    Article  CAS  Google Scholar 

  62. Haltern G, Peiniger S, Bufe A, Reiss G, Gülker H, Scheffold T (2010) Comparison of usefulness of heart-type fatty acid binding protein versus cardiac troponin T for diagnosis of acute myocardial infarction. Am J Cardiol 105(1):1–9

    Article  CAS  PubMed  Google Scholar 

  63. Mad P, Domanovits H, Fazelnia C, Stiassny K, Russmüller G, Cseh A, Sodeck G, Binder T, Christ G, Szekeres T (2007) Human heart-type fatty-acid-binding protein as a point-of-care test in the early diagnosis of acute myocardial infarction. J Assoc Physicians 100(4):203–210

    CAS  Google Scholar 

  64. Ruzgar O, Bilge AK, Bugra Z, Umman S, Yilmaz E, Ozben B, Umman B, Meric M (2006) The use of human heart-type fatty acid-binding protein as an early diagnostic biochemical marker of myocardial necrosis in patients with acute coronary syndrome, and its comparison with troponin-T and creatine kinase–myocardial band. Heart Vessels 21(5):309–314

    Article  PubMed  Google Scholar 

  65. Brown CS, Bertolet BD (1997) Cardiac troponin: see ya later, CK! Chest 111(1):2–5

    Article  CAS  PubMed  Google Scholar 

  66. Du Clos TW (2000) Function of C-reactive protein. Ann Med 32(4):274–278

    Article  PubMed  Google Scholar 

  67. Clearfield MB (2005) C-reactive protein: a new risk assessment tool for cardiovascular disease. J Am Osteopath Assoc 105(9):409–416

    PubMed  Google Scholar 

  68. Chandra P, Suman P, Airon H, Mukherjee M, Kumar P (2014) Prospects and advancements in C-reactive protein detection. World J Methodol 4(1):1

    Article  PubMed  PubMed Central  Google Scholar 

  69. Chrysohoou C, Pitsavos C, Barbetseas J, Kotroyiannis I, Brili S, Vasiliadou K, Papadimitriou L, Stefanadis C (2009) Chronic systemic inflammation accompanies impaired ventricular diastolic function, detected by Doppler imaging, in patients with newly diagnosed systolic heart failure (Hellenic Heart Failure Study). Heart Vessels 24(1):22–26

    Article  PubMed  Google Scholar 

  70. Rivera M, Taléns-Visconti R, Sirera R, Bertomeu V, Salvador A, Cortés R, de Burgos FG, Climent V, Payá R, Martinez-Dolz L (2004) Soluble TNF-α and interleukin-6 receptors in the urine of heart failure patients. Their clinical value and relationship with plasma levels. Eur J Heart Fail 6(7):877–882

    Article  CAS  PubMed  Google Scholar 

  71. Wollert KC, Drexler H (2001) The role of interleukin-6 in the failing heart. Heart Fail Rev 6(2):95–103

    Article  CAS  PubMed  Google Scholar 

  72. Gabriele P, Zhi Fang S, Tonny DTT, Carsten K, Hideo AB, Michael E, Markus F, Thomas W, Hans HS, Mario CD (2001) Activation of the cardiac interleukin-6 system in advanced heart failure. Eur J Heart Fail 3(4):415–421

    Article  Google Scholar 

  73. Raymond RJ, Dehmer GJ, Theoharides TC, Deliargyris EN (2001) Elevated interleukin-6 levels in patients with asymptomatic left ventricular systolic dysfunction. Am Heart J 141(3):435–438

    Article  CAS  PubMed  Google Scholar 

  74. Pudil R, Tichý M, Andrýs C, Rehácek V, Bláha V, Vojácek J, Palicka V (2010) Plasma interleukin-6 level is associated with NT-proBNP level and predicts short-and long-term mortality in patients with acute heart failure. Acta Medica (Hradec Kralove) 53(4):225–228

    Article  CAS  Google Scholar 

  75. Inoue K, Sugiyama A, Reid PC, Ito Y, Miyauchi K, Mukai S, Sagara M, Miyamoto K, Satoh H, Kohno I (2007) Establishment of a high sensitivity plasma assay for human pentraxin3 as a marker for unstable angina pectoris. Arterioscler Thromb Vasc Biol 27(1):161–167

    Article  CAS  PubMed  Google Scholar 

  76. Latini R, Maggioni AP, Peri G, Gonzini L, Lucci D, Mocarelli P, Vago L, Pasqualini F, Signorini S, Soldateschi D (2004) Prognostic significance of the long pentraxin PTX3 in acute myocardial infarction. Circulation 110(16):2349–2354

    Article  CAS  PubMed  Google Scholar 

  77. Suzuki S, Takeishi Y, Niizeki T, Koyama Y, Kitahara T, Sasaki T, Sagara M, Kubota I (2008) Pentraxin 3, a new marker for vascular inflammation, predicts adverse clinical outcomes in patients with heart failure. Am Heart J 155(1):75–81

    Article  CAS  PubMed  Google Scholar 

  78. Knoflach M, Kiechl S, Mantovani A, Cuccovillo I, Bottazzi B, Xu Q, Xiao Q, Gasperi A, Mayr A, Kehrer M (2012) Pentraxin-3 as a marker of advanced atherosclerosis results from the Bruneck, ARMY and ARFY Studies. PLoS ONE 7(2):e31474

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Bayes-Genis A, Conover CA, Overgaard MT, Bailey KR, Christiansen M, Holmes DR Jr, Virmani R, Oxvig C, Schwartz RS (2001) Pregnancy-associated plasma protein A as a marker of acute coronary syndromes. N Engl J Med 345(14):1022–1029

    Article  CAS  PubMed  Google Scholar 

  80. Lund J, Qin QP, Ilva T, Nikus K, Eskola M, Porela P, Kokkala S, Pulkki K, Pettersson K, Voipio-Pulkki LM (2006) Pregnancy-associated plasma protein A: A biomarker in acute ST-elevation myocardial infarction (STEMI). Ann Med 38(3):221–228

    Article  CAS  PubMed  Google Scholar 

  81. Heeschen C, Dimmeler S, Hamm CW, van den Brand MJ, Boersma E, Zeiher AM, Simoons ML (2003) Soluble CD40 ligand in acute coronary syndromes. N Engl J Med 348(12):1104–1111

    Article  CAS  PubMed  Google Scholar 

  82. Neuhold S, Huelsmann M, Strunk G, Stoiser B, Struck J, Morgenthaler NG, Bergmann A, Moertl D, Berger R, Pacher R (2008) Comparison of copeptin, B-type natriuretic peptide, and amino-terminal pro-B-type natriuretic peptide in patients with chronic heart failure: prediction of death at different stages of the disease. J Am Coll Cardiol 52(4):266–272

    Article  CAS  PubMed  Google Scholar 

  83. Morgenthaler NG, Struck J, Alonso C, Bergmann A (2006) Assay for the measurement of copeptin, a stable peptide derived from the precursor of vasopressin. Clin Chem 52(1):112–119

    Article  CAS  PubMed  Google Scholar 

  84. Voors AA, Von Haehling S, Anker SD, Hillege HL, Struck J, Hartmann O, Bergmann A, Squire I, Van Veldhuisen DJ, Dickstein K (2009) C-terminal provasopressin (copeptin) is a strong prognostic marker in patients with heart failure after an acute myocardial infarction: results from the OPTIMAAL study. Eur Heart J 30(10):1187–1194

    Article  CAS  PubMed  Google Scholar 

  85. Reichlin T, Hochholzer W, Stelzig C, Laule K, Freidank H, Morgenthaler NG, Bergmann A, Potocki M, Noveanu M, Breidthardt T (2009) Incremental value of copeptin for rapid rule out of acute myocardial infarction. J Am Coll Cardiol 54(1):60–68

    Article  CAS  PubMed  Google Scholar 

  86. Wollert KC, Kempf T (2012) Growth differentiation factor 15 in heart failure: an update. Curr Heart Fail Rep 9(4):337–345

    Article  CAS  PubMed  Google Scholar 

  87. Xu J, Kimball TR, Lorenz JN, Brown DA, Bauskin AR, Klevitsky R, Hewett TE, Breit SN, Molkentin JD (2006) GDF15/MIC-1 functions as a protective and antihypertrophic factor released from the myocardium in association with SMAD protein activation. Circ Res 98(3):342–350

    Article  CAS  PubMed  Google Scholar 

  88. Kempf T, Björklund E, Olofsson S, Lindahl B, Allhoff T, Peter T, Tongers J, Wollert KC, Wallentin L (2007) Growth-differentiation factor-15 improves risk stratification in ST-segment elevation myocardial infarction. Eur Heart J 28(23):2858–2865

    Article  CAS  PubMed  Google Scholar 

  89. Kempf T, von Haehling S, Peter T, Allhoff T, Cicoira M, Doehner W, Ponikowski P, Filippatos GS, Rozentryt P, Drexler H (2007) Prognostic utility of growth differentiation factor-15 in patients with chronic heart failure. J Am Coll Cardiol 50(11):1054–1060

    Article  CAS  PubMed  Google Scholar 

  90. Sanada S, Hakuno D, Higgins LJ, Schreiter ER, McKenzie AN, Lee RT (2007) IL-33 and ST2 comprise a critical biomechanically induced and cardioprotective signaling system. J Clin Investig 117(6):1538–1549

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Weinberg EO, Shimpo M, Hurwitz S, Tominaga S-i, Rouleau J-L, Lee RT (2003) Identification of serum soluble ST2 receptor as a novel heart failure biomarker. Circulation 107(5):721–726

    Article  PubMed  Google Scholar 

  92. Rehman SU, Mueller T, Januzzi JL (2008) Characteristics of the novel interleukin family biomarker ST2 in patients with acute heart failure. J Am Coll Cardiol 52(18):1458–1465

    Article  CAS  PubMed  Google Scholar 

  93. Januzzi JL, Peacock WF, Maisel AS, Chae CU, Jesse RL, Baggish AL, O’Donoghue M, Sakhuja R, Chen AA, van Kimmenade RR (2007) Measurement of the interleukin family member ST2 in patients with acute dyspnea: results from the PRIDE (Pro-Brain Natriuretic Peptide Investigation of Dyspnea in the Emergency Department) study. J Am Coll Cardiol 50(7):607–613

    Article  CAS  PubMed  Google Scholar 

  94. López B, Ravassa S, González A, Zubillaga E, Bonavila C, Bergés M, Echegaray K, Beaumont J, Moreno MU, San José G (2016) Myocardial collagen crosslinking is associated with heart failure hospitalization in patients with hypertensive heart failure. J Am Coll Cardiol 67(3):251–260

    Article  PubMed  Google Scholar 

  95. George J, Patal S, Wexler D, Roth A, Sheps D, Keren G (2005) Circulating matrix metalloproteinase-2 but not matrix metalloproteinase-3, matrix metalloproteinase-9, or tissue inhibitor of metalloproteinase-1 predicts outcome in patients with congestive heart failure. Am Heart J 150:484–487

    Article  CAS  PubMed  Google Scholar 

  96. Buralli S, Dini FL, Ballo P, Conti U, Fontanive P, Duranti E, Metelli MR, Marzilli M, Taddei S (2010) Circulating matrix metalloproteinase-3 and metalloproteinase-9 and tissue Doppler measures of diastolic dysfunction to risk stratify patients with systolic heart failure. Am J Cardiol 105(6):853–856

    Article  CAS  PubMed  Google Scholar 

  97. Frantz S, Störk S, Michels K, Eigenthaler M, Ertl G, Bauersachs J, Angermann C (2008) Tissue inhibitor of metalloproteinases levels in patients with chronic heart failure: an independent predictor of mortality. Eur J Heart Fail 10(4):388–395

    Article  CAS  PubMed  Google Scholar 

  98. Ibrahim NE, Januzzi JL Jr (2017) Beyond natriuretic peptides for diagnosis and management of heart failure. Clin Chem 63(1):211–222

    Article  CAS  PubMed  Google Scholar 

  99. Tziakas DN, Chalikias GK, Stakos D, Chatzikyriakou SV, Papazoglou D, Mitrousi K, Lantzouraki A, Thomaidi A, Boudoulas H, Konstantinides S (2011) Independent and additive prognostic ability of serum carboxy-terminal telopeptide of collagen type-I in heart failure patients: a multi-marker approach with high-negative predictive value to rule out long-term adverse events. Eur J Prev Cardiol 19(1):62–71

    Article  PubMed  Google Scholar 

  100. Gullestad L, Ueland T, Kjekshus J, Nymo SH, Hulthe J, Muntendam P, Adourian A, Böhm M, van Veldhuisen DJ, Komajda M (2012) Galectin-3 predicts response to statin therapy in the Controlled Rosuvastatin Multinational Trial in Heart Failure (CORONA). Eur Heart J 33(18):2290–2296

    Article  CAS  PubMed  Google Scholar 

  101. van Kimmenade RR, Januzzi JL, Ellinor PT, Sharma UC, Bakker JA, Low AF, Martinez A, Crijns HJ, MacRae CA, Menheere PP (2006) Utility of amino-terminal pro-brain natriuretic peptide, galectin-3, and apelin for the evaluation of patients with acute heart failure. J Am Coll Cardiol 48(6):1217–1224

    Article  PubMed  Google Scholar 

  102. Motiwala SR, Szymonifka J, Belcher A, Weiner RB, Baggish AL, Sluss P, Gaggin HK, Bhardwaj A, Januzzi JL (2013) Serial measurement of galectin-3 in patients with chronic heart failure: results from the ProBNP Outpatient Tailored Chronic Heart Failure Therapy (PROTECT) study. Eur J Heart Fail 15(10):1157–1163

    Article  CAS  PubMed  Google Scholar 

  103. Ryu J-A, Yang JH, Lee D, Park C-M, Suh GY, Jeon K, Cho J, Baek SY, C. Carriere K, Chung CR, (2015) Clinical usefulness of procalcitonin and C-reactive protein as outcome predictors in critically ill patients with severe sepsis and septic shock. PLoS ONE 10(9):e0138150

    Article  PubMed  PubMed Central  Google Scholar 

  104. Reiner MM, Khoury WE, Canales MB, Chmielewski RA, Patel K, Razzante MC, Cloughtery CO, Rowland DY (2017) Procalcitonin as a biomarker for predicting amputation level in lower extremity infections. J Foot Ankle Surg 56(3):484–491

    Article  PubMed  Google Scholar 

  105. Hayashida K, Kondo Y, Hara Y, Aihara M, Yamakawa K (2017) Head-to-head comparison of procalcitonin and presepsin for the diagnosis of sepsis in critically ill adult patients: a protocol for a systematic review and meta-analysis. BMJ Open 7(3):e014305

    Article  PubMed  PubMed Central  Google Scholar 

  106. Peacock W, Hollander J, Diercks D, Lopatin M, Fonarow G, Emerman C (2008) Morphine and outcomes in acute decompensated heart failure: an ADHERE analysis. Emerg Med J 25(4):205–209

    Article  CAS  PubMed  Google Scholar 

  107. Manzano-Fernández S, Januzzi JL, Boronat-Garcia M, Bonaque-González JC, Truong QA, Pastor-Pérez FJ, Muñoz-Esparza C, Pastor P, Albaladejo-Otón MD, Casas T (2011) β-trace protein and cystatin C as predictors of long-term outcomes in patients with acute heart failure. J Am Coll Cardiol 57(7):849–858

    Article  PubMed  Google Scholar 

  108. Nymo SH, Ueland T, Askevold ET, Flo T, Kjekshus J, Hulthe J, Wikstrand J, McMurray J, Van Veldhuisen DJ, Gullestad L (2012) The association between neutrophil gelatinase-associated lipocalin and clinical outcome in chronic heart failure: results from CORONA. J Intern Med 271(5):436–443

    Article  CAS  PubMed  Google Scholar 

  109. Macdonald S, Arendts G, Nagree Y, Xu X-F (2012) Neutrophil Gelatinase-Associated Lipocalin (NGAL) predicts renal injury in acute decompensated cardiac failure: a prospective observational study. BMC Cardiovasc Disord 12(1):1–6

    Article  Google Scholar 

  110. Groenveld HF, Januzzi JL, Damman K, van Wijngaarden J, Hillege HL, van Veldhuisen DJ, van der Meer P (2008) Anemia and mortality in heart failure patients: a systematic review and meta-analysis. J Am Coll Cardiol 52(10):818–827

    Article  PubMed  Google Scholar 

  111. Loria V, Dato I, Graziani F, Biasucci LM (2008) Myeloperoxidase: a new biomarker of inflammation in ischemic heart disease and acute coronary syndromes. Mediat Inflamm 2008:1–4

    Article  Google Scholar 

  112. Baldus S, Heeschen C, Meinertz T, Zeiher AM, Eiserich JP, Münzel T, Simoons ML, Hamm CW (2003) Myeloperoxidase serum levels predict risk in patients with acute coronary syndromes. Circulation 108(12):1440–1445

    Article  CAS  PubMed  Google Scholar 

  113. Zhang R, Brennan M-L, Fu X, Aviles RJ, Pearce GL, Penn MS, Topol EJ, Sprecher DL, Hazen SL (2001) Association between myeloperoxidase levels and risk of coronary artery disease. JAMA 286(17):2136–2142

    Article  CAS  PubMed  Google Scholar 

  114. Schindhelm RK, van der Zwan LP, Teerlink T, Scheffer PG (2009) Myeloperoxidase: a useful biomarker for cardiovascular disease risk stratification? Clin Chem 55(8):1462–1470

    Article  CAS  PubMed  Google Scholar 

  115. Khan AA, Alsahli MA, Rahmani AH (2018) Myeloperoxidase as an active disease biomarker: recent biochemical and pathological perspectives. Med Sci 6(2):33

    Google Scholar 

  116. Klebanoff SJ (2005) Myeloperoxidase: friend and foe. J Leukoc Biol 77(5):598–625

    Article  CAS  PubMed  Google Scholar 

  117. Lowe GD (2010) Fibrinogen assays for cardiovascular risk assessment. Clin Chem 56(5):693

    Article  CAS  PubMed  Google Scholar 

  118. Welsh P, Woodward M, Rumley A, Lowe G (2008) Associations of plasma pro-inflammatory cytokines, Fibrinogen, viscosity and C-reactive protein with cardiovascular risk factors and social deprivation: the fourth Glasgow MONICA study. Br J Haematol 141(6):852–861

    Article  CAS  PubMed  Google Scholar 

  119. Budzynski AZ, Shainoff JR (1986) Fibrinogen and fibrin: biochemistry and pathophysiology. Crit Rev Oncol Hematol 6(2):97–146

    Article  CAS  PubMed  Google Scholar 

  120. Canseco-Ávila LM, Jerjes-Sánchez C, Ortiz-López R, Rojas-Martínez A, Guzmán-Ramírez D (2006) Fibrinogen. Cardiovascular risk factor or marker? Arch Cardiol Mexico 76(S4):158–172

    Google Scholar 

  121. Galanakis DK (1992) Fibrinogen anomalies and disease: a clinical update. Hematol Oncol Clin North Am 6(5):1171–1187

    Article  CAS  PubMed  Google Scholar 

  122. Ussher JR, Lopaschuk GD, Arduini A (2013) Gut microbiota metabolism of L-carnitine and cardiovascular risk. Atherosclerosis 231(2):456–461

    Article  CAS  PubMed  Google Scholar 

  123. Fogelman AM (2015) TMAO is both a biomarker and a renal toxin. Journal 116:396–397

    CAS  Google Scholar 

  124. Haidaris PJ, Francis CW, Sporn LA, Arvan DS, Collichio FA, Marder VJ (1989) Megakaryocyte and hepatocyte origins of human fibrinogen biosynthesis exhibit hepatocyte-specific expression of y chain-variant polypeptides. Blood 74(2):743–750

    Article  CAS  PubMed  Google Scholar 

  125. Singh T, Nigam A, Gupta A, Singh B (2011) Cardiac biomarkers: When to test?-physician perspective. JIACM 12(2):117–121

    Google Scholar 

  126. Eggers KM, Oldgren J, Nordenskjöld A, Lindahl B (2004) Diagnostic value of serial measurement of cardiac markers in patients with chest pain: limited value of adding myoglobin to Troponin I for exclusion of myocardial infarction. Am Heart J 148(4):574–581

    Article  CAS  PubMed  Google Scholar 

  127. Abadie JM, Blassingame CL, Bankson DD (2005) Albumin cobalt binding assay to rule out acute coronary syndrome. Ann Clin Lab Sci 35(1):66–72

    CAS  PubMed  Google Scholar 

  128. Gaze DC (2009) Ischemia modified albumin: a novel biomarker for the detection of cardiac ischemia. Drug Metab Pharmacokinet 24(4):333–341

    Article  CAS  PubMed  Google Scholar 

  129. El-Tamimi H, Mansour M, Wargovich T, Hill J, Kerensky R, Conti C, Pepine C (1994) Constrictor and dilator responses to intracoronary acetylcholine in adjacent segments of the same coronary artery in patients with coronary artery disease. Endothel Funct Revisit Circ 89(1):45–51

    CAS  Google Scholar 

  130. Lerman A, BurnettJr J (1992) Intact and altered endothelium in regulation of vasomotion. Circulation 86(6):III12-19

    CAS  PubMed  Google Scholar 

  131. Reriani MK, Flammer AJ, Jama A, Lerman LO, Lerman A (2012) Novel functional risk factors for the prediction of cardiovascular events in vulnerable patients following acute coronary syndrome. Circ J 2012:1203141649–1203141649

    Google Scholar 

  132. Jodo S, Kobayashi S, Nt K, Ogura N, Feng Y, Amasaki Y, Fujisaku A, Azuma M, Yagita H, Okumura K (1997) Serum levels of soluble Fas/APO-1 (CD95) and its molecular structure in patients with systemic lupus erythematosus (SLE) and other autoimmune diseases. Clin Exp Immunol 107(1):89–95

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Takahashi K, Ohyabu Y, Takahashi K, Solodushko V, Takatani T, Itoh T, Schaffer SW, Azuma J (2003) Taurine renders the cell resistant to ischemia-induced injury in cultured neonatal rat cardiomyocytes. J Cardiovasc Pharmacol 41(5):726–733

    Article  CAS  PubMed  Google Scholar 

  134. Cheng J, Zhou T, Liu C, Shapiro JP, Brauer MJ, Kiefer MC, Barr PJ, Mountz JD (1994) Protection from Fas-mediated apoptosis by a soluble form of the Fas molecule. Science 263(5154):1759–1762

    Article  CAS  PubMed  Google Scholar 

  135. Cowland JB, Borregaard N (1997) Molecular characterization and pattern of tissue expression of the gene for neutrophil gelatinase-associated lipocalin from humans. Genomics 45(1):17–23

    Article  CAS  PubMed  Google Scholar 

  136. Friedl A, Stoesz S, Buckley P, Gould M (1999) Neutrophil gelatinase-associated lipocalin in normal and neoplastic human tissues. Cell type-specific pattern of expression. Histochem J 31(7):433–441

    Article  CAS  PubMed  Google Scholar 

  137. Goetz DH, Holmes MA, Borregaard N, Bluhm ME, Raymond KN, Strong RK (2002) The neutrophil lipocalin NGAL is a bacteriostatic agent that interferes with siderophore-mediated iron acquisition. Mol Cell 10(5):1033–1043

    Article  CAS  PubMed  Google Scholar 

  138. Devarajan P (2008) NGAL in acute kidney injury: from serendipity to utility. Am J Kidney Dis 52(3):395–399

    Article  PubMed  Google Scholar 

  139. Yndestad A, Landrø L, Ueland T, Dahl CP, Flo TH, Vinge LE, Espevik T, Frøland SS, Husberg C, Christensen G (2009) Increased systemic and myocardial expression of neutrophil gelatinase-associated lipocalin in clinical and experimental heart failure. Eur Heart J 30(10):1229–1236

    Article  CAS  PubMed  Google Scholar 

  140. Damman K, Masson S, Hillege HL, Maggioni AP, Voors AA, Opasich C, Van Veldhuisen DJ, Montagna L, Cosmi F, Tognoni G (2011) Clinical outcome of renal tubular damage in chronic heart failure. Eur Heart J 32(21):2705–2712

    Article  CAS  PubMed  Google Scholar 

  141. Fabbrini E, Serafini M, Baric IC, Hazen SL, Klein S (2014) Effect of plasma uric acid on antioxidant capacity, oxidative stress, and insulin sensitivity in obese subjects. Diabetes 63(3):976–981

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Jin YL, Zhu T, Xu L, Zhang WS, Liu B, Jiang CQ, Yu H, Huang LM, Cheng KK, Thomas GN (2013) Uric acid levels, even in the normal range, are associated with increased cardiovascular risk: the Guangzhou Biobank Cohort Study. Int J Cardiol 168(3):2238–2241

    Article  PubMed  Google Scholar 

  143. Dutta A, Henley W, Pilling LC, Wallace RB, Melzer D (2013) Uric acid measurement improves prediction of cardiovascular mortality in later life. J Am Geriatr Soc 61(3):319–326

    Article  PubMed  Google Scholar 

  144. Zhao G, Huang L, Song M, Song Y (2013) Baseline serum uric acid level as a predictor of cardiovascular disease related mortality and all-cause mortality: a meta-analysis of prospective studies. Atherosclerosis 231(1):61–68

    Article  CAS  PubMed  Google Scholar 

  145. Fenech G, Rajzbaum G, Mazighi M, Blacher J (2014) Serum uric acid and cardiovascular risk: state of the art and perspectives. Joint Bone Spine 81(5):392–397

    Article  CAS  PubMed  Google Scholar 

  146. Jee SH, Lee SY, Kim MT (2004) Serum uric acid and risk of death from cancer, cardiovascular disease or all causes in men. Eur J Cardiovasc Prev Rehabil 11(3):185–191

    Article  PubMed  Google Scholar 

  147. Lee K-F, Simon H, Chen H, Bates B, Hung M-C, Hauser C (1995) Requirement for neuregulin receptor erbB2 in neural and cardiac development. Nature 378(6555):394–398

    Article  CAS  PubMed  Google Scholar 

  148. Kuramochi Y, Cote GM, Guo X, Lebrasseur NK, Cui L, Liao R, Sawyer DB (2004) Cardiac endothelial cells regulate reactive oxygen species-induced cardiomyocyte apoptosis through neuregulin-1β/erbB4 signaling. J Biol Chem 279(49):51141–51147

    Article  CAS  PubMed  Google Scholar 

  149. Lemmens K, Doggen K, De Keulenaer GW (2007) Role of neuregulin-1/ErbB signaling in cardiovascular physiology and disease: implications for therapy of heart failure. Circulation 116(8):954–960

    Article  CAS  PubMed  Google Scholar 

  150. Arques S (2018) Human serum albumin in cardiovascular diseases. Eur J Intern Med 52:8–12

    Article  CAS  PubMed  Google Scholar 

  151. Volpato S, Guralnik JM, Ferrucci L, Balfour J, Chaves P, Fried LP, Harris TB (2001) Cardiovascular disease, interleukin-6, and risk of mortality in older women: the women’s health and aging study. Circulation 103(7):947–953

    Article  CAS  PubMed  Google Scholar 

  152. Johnson BD, Kip KE, Marroquin OC, Ridker PM, Kelsey SF, Shaw LJ, Pepine CJ, Sharaf B, Bairey Merz CN, Sopko G (2004) Serum amyloid A as a predictor of coronary artery disease and cardiovascular outcome in women: the National Heart, Lung, and Blood Institute-Sponsored Women’s Ischemia Syndrome Evaluation (WISE). Circulation 109(6):726–732

    Article  CAS  PubMed  Google Scholar 

  153. Graham TE, Yang Q, Blüher M, Hammarstedt A, Ciaraldi TP, Henry RR, Wason CJ, Oberbach A, Jansson P-A, Smith U (2006) Retinol-binding protein 4 and insulin resistance in lean, obese, and diabetic subjects. N Engl J Med 354(24):2552–2563

    Article  CAS  PubMed  Google Scholar 

  154. Alkharfy KM, Al-Daghri NM, Vanhoutte PM, Krishnaswamy S, Xu A (2012) Serum retinol-binding protein 4 as a marker for cardiovascular disease in women. PLoS ONE 7(10):e48612

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Pirillo A, Norata GD, Catapano AL (2013) LOX-1, OxLDL, and atherosclerosis. Mediat Inflamm 48:1–12

    Article  Google Scholar 

  156. Kume N, Mitsuoka H, Hayashida K, Tanaka M, Kominami G, Kita T (2010) Soluble lectin-like oxidized LDL receptor-1 (sLOX-1) as a sensitive and specific biomarker for acute coronary syndrome—comparison with other biomarkers. J Cardiol 56(2):159–165

    Article  PubMed  Google Scholar 

  157. Heilbronn K, Smith R, Ravussin E (2003) The insulin-sensitizing role of the fat derived hormone adiponectin. Curr Pharm Des 9(17):1411–1418

    Article  CAS  PubMed  Google Scholar 

  158. Shimada K, Miyazaki T, Daida H (2004) Adiponectin and atherosclerotic disease. Clin Chim Acta 344(1–2):1–12

    Article  CAS  PubMed  Google Scholar 

  159. Wang X-j, Wang M (2010) The S100 protein family and its application in cardiac diseases. World J Emerg Med 1(3):165

    CAS  PubMed  PubMed Central  Google Scholar 

  160. Zhan Y, Yu J, Ding R, Sun Y, Hu D (2014) Triglyceride to high density lipoprotein cholesterol ratio, total cholesterol to high density lipoprotein cholesterol ratio and low ankle brachial index in an elderly population. Vasa 43(3):189–197

    Article  PubMed  Google Scholar 

  161. Fekih O, Triki S, Hellara I, Neffati F, Rejeb J, Ommezzine A, Chouchane S, Guediche MN, Bouslama A, Najjar MF (2014) Can paraoxonase 1 polymorphisms (L55 M and Q192 R) protect children with type 1 diabetes against lipid abnormalities? J Clin Lipidol 8(3):249–255

    Article  PubMed  Google Scholar 

  162. Shah SN, Arneja J (2013) Efficacy of rosuvastatin in achieving target HDL, LDL, triglycerides and total cholesterol levels in type 2 diabetes mellitus (T2DM) with newly diagnosed dyslipidaemia: an open label, nonrandomized, non-interventional and observational study in India. J Assoc Physicians India 61(10):721–726

    PubMed  Google Scholar 

  163. Merino J, Masana L, Guijarro C, Ascaso J, Lagares M, Civeira F (2014) Recomendations for clinical use of food enriched phytosterols/phytostanols handling hypercholesterolemia. ClínInvestig Arterioscler 26(3):147–158

    Google Scholar 

  164. Hoogeveen EK, Geleijnse JM, Kromhout D, van’t Sant P, Gemen EF, Kusters R, Giltay EJ (2015) No effect of n-3 fatty acids supplementation on NT-proBNP after myocardial infarction: the Alpha Omega Trial. Eur J Prev Cardiol 22(5):648–655

    Article  PubMed  Google Scholar 

  165. Hirayama S, Miida T (2012) Small dense LDL: an emerging risk factor for cardiovascular disease. Clin Chim Acta 414:215–224

    Article  CAS  PubMed  Google Scholar 

  166. Topbas C, Swick A, Razavi M, Anderson NL, Pearson TW, Bystrom C (2018) Measurement of lipoprotein-associated phospholipase A2 by use of 3 different methods: exploration of discordance between ELISA and activity assays. Clin Chem 64(4):697–704

    Article  CAS  PubMed  Google Scholar 

  167. Investigators S (2014) Darapladib for preventing ischemic events in stable coronary heart disease. N Engl J Med 370(18):1702–1711

    Article  Google Scholar 

  168. Wallentin L, Held C, Armstrong PW, Cannon CP, Davies RY, Granger CB, Hagström E, Harrington RA, Hochman JS, Koenig W (2016) Lipoprotein-associated phospholipase A2 activity is a marker of risk but not a useful target for treatment in patients with stable coronary heart disease. J Am Heart Assoc 5(6):e003407

    Article  PubMed  PubMed Central  Google Scholar 

  169. Tourdot BE, Ahmed I, Holinstat M (2014) The emerging role of oxylipins in thrombosis and diabetes. Front Pharmacol 4:176

    Article  PubMed  PubMed Central  Google Scholar 

  170. Nayeem MA (2018) Role of oxylipins in cardiovascular diseases. Acta Pharmacol Sin 39(7):1142–1154

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Idzikowska K, Zielińska M (2018) Midregional pro-atrial natriuretic peptide, an important member of the natriuretic peptide family: potential role in diagnosis and prognosis of cardiovascular disease. J Int Med Res 46(8):3017–3029

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Pickering TG, Miller NH, Ogedegbe G, Krakoff LR, Artinian NT, Goff D (2008) Call to action on use and reimbursement for home blood pressure monitoring: a joint scientific statement from the American Heart Association, American Society of Hypertension, and Preventive Cardiovascular Nurses Association. Hypertension 52(1):10–29

    Article  CAS  PubMed  Google Scholar 

  173. Parati G, Stergiou GS, Asmar R, Bilo G, De Leeuw P, Imai Y, Kario K, Lurbe E, Manolis A, Mengden T (2008) European Society of Hypertension guidelines for blood pressure monitoring at home: a summary report of the Second International Consensus Conference on Home Blood Pressure Monitoring. J Hypertens 26(8):1505–1526

    Article  CAS  PubMed  Google Scholar 

  174. Barwari T, Joshi A, Mayr M (2016) MicroRNAs in cardiovascular disease. J Am Coll Cardiol 68(23):2577–2584

    Article  CAS  PubMed  Google Scholar 

  175. Dimmeler S, Zeiher AM (2010) Circulating microRNAs: novel biomarkers for cardiovascular diseases? Journal 31:2705–2707

    Google Scholar 

  176. Karakas M, Schulte C, Appelbaum S, Ojeda F, Lackner KJ, Münzel T, Schnabel RB, Blankenberg S, Zeller T (2017) Circulating microRNAs strongly predict cardiovascular death in patients with coronary artery disease—results from the large AtheroGene study. Eur Heart J 38(7):516–523

    CAS  PubMed  Google Scholar 

  177. Zampetaki A, Willeit P, Tilling L, Drozdov I, Prokopi M, Renard J-M, Mayr A, Weger S, Schett G, Shah A (2012) Prospective study on circulating MicroRNAs and risk of myocardial infarction. J Am Coll Cardiol 60(4):290–299

    Article  CAS  PubMed  Google Scholar 

  178. Kim AJ, Ro H, Kim H, Chang JH, Lee HH, Chung W, Jung JY (2021) Soluble ST2 and galectin-3 as predictors of chronic kidney disease progression and outcomes. Am J Nephrol 52(2):119–130

    Article  CAS  PubMed  Google Scholar 

  179. Li C, Zhao M, Zhao M, Chen N, Guo Y, Du Y, Zhang Y, Cao B, Zhan B, Guo C (2022) IL-37 isoform D acts as an inhibitor of soluble ST2 to boost type 2 immune homeostasis in white adipose tissue. Cell Death Discov 8(1):1–12

    Google Scholar 

  180. Karthigan N, Lockwood S, White A, Yang J, Young MJ (2022) Mineralocorticoid receptor antagonists, heart failure and predictive biomarkers. J Endocrinol 253(3):R65–R70

    Article  PubMed  Google Scholar 

  181. Mincarone P, Bodini A, Tumolo MR, Vozzi F, Rocchiccioli S, Pelosi G, Caselli C, Sabina S, Leo CG (2021) Discrimination capability of pretest probability of stable coronary artery disease: a systematic review and meta-analysis suggesting how to improve validation procedures. BMJ Open 11(7):e047677

    Article  PubMed  PubMed Central  Google Scholar 

  182. Kanmura S, Morinaga Y, Tanaka A, Komaki Y, Iwaya H, Kumagai K, Mawatari S, Sasaki F, Tanoue S, Hashimoto S (2021) Increased gene copy number of DEFA1A3 is associated with the severity of ulcerative colitis. Clin Transl Gastroenterol 12(4):e00331

    Article  PubMed  PubMed Central  Google Scholar 

  183. Deng H, Li J, Shah AA, Lin G, Chen H, Ouyang W (2022) Commonly expressed key transcriptomic profiles of sepsis in the human circulation and brain via integrated analysis. Int Immunopharmacol 104:108518

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Authors would like to thank Osmania University and University of Hyderabad for providing the moral support for continuing our research.

Funding

None of the funding are available to complete this article.

Author information

Authors and Affiliations

Authors

Contributions

ST and RP provided the concept and overlook the preparation of this manuscript, other authors include SSRN; HR and SC contribute equally to complete this gathering relative information, proof reading and finalizing for the submission of the article. All authors have read and approved the manuscript.

Corresponding author

Correspondence to Raju Padiya.

Ethics declarations

Ethics approval and consent to participate

None of human and animals are used to complete this article, and this study does not require any ethical statement to present.

Consent for publication

Not applicable.

Competing interests

None of the other authors has any conflicts of interest, financial or otherwise, to disclose.

Additional information

Publisher's Note

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

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Thupakula, S., Nimmala, S.S.R., Ravula, H. et al. Emerging biomarkers for the detection of cardiovascular diseases. Egypt Heart J 74, 77 (2022). https://doi.org/10.1186/s43044-022-00317-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s43044-022-00317-2

Keywords