ArticlePDF AvailableLiterature Review

Adverse Cardiac Remodelling after Acute Myocardial Infarction: Old and New Biomarkers

Authors:

Abstract and Figures

The prevalence of heart failure (HF) due to cardiac remodelling after acute myocardial infarction (AMI) does not decrease regardless of implementation of new technologies supporting opening culprit coronary artery and solving of ischemia-relating stenosis with primary percutaneous coronary intervention (PCI). Numerous studies have examined the diagnostic and prognostic potencies of circulating cardiac biomarkers in acute coronary syndrome/AMI and heart failure after AMI, and even fewer have depicted the utility of biomarkers in AMI patients undergoing primary PCI. Although complete revascularization at early period of acute coronary syndrome/AMI is an established factor for improved short-term and long-term prognosis and lowered risk of cardiovascular (CV) complications, late adverse cardiac remodelling may be a major risk factor for one-year mortality and postponded heart failure manifestation after PCI with subsequent blood flow resolving in culprit coronary artery. The aim of the review was to focus an attention on circulating biomarker as a promising tool to stratify AMI patients at high risk of poor cardiac recovery and developing HF after successful PCI. The main consideration affects biomarkers of inflammation, biomechanical myocardial stress, cardiac injury and necrosis, fibrosis, endothelial dysfunction, and vascular reparation. Clinical utilities and predictive modalities of natriuretic peptides, cardiac troponins, galectin 3, soluble suppressor tumorogenicity-2, high-sensitive C-reactive protein, growth differential factor-15, midregional proadrenomedullin, noncoding RNAs, and other biomarkers for adverse cardiac remodelling are discussed in the review.
This content is subject to copyright. Terms and conditions apply.
Review Article
Adverse Cardiac Remodelling after Acute Myocardial Infarction:
Old and New Biomarkers
Alexander E. Berezin
1
and Alexander A. Berezin
2
1
Internal Medicine Department, State Medical University, Ministry of Health of Ukraine, Zaporozhye 69035, Ukraine
2
Internal Medicine Department, Medical Academy of Post-Graduate Education, Ministry of Health of Ukraine,
Zaporozhye 69096, Ukraine
Correspondence should be addressed to Alexander E. Berezin; aeberezin@gmail.com
Received 15 May 2019; Revised 6 January 2020; Accepted 22 May 2020; Published 12 June 2020
Academic Editor: Roberta Rizzo
Copyright © 2020 Alexander E. Berezin and Alexander A. Berezin. This is an open access article distributed under the Creative
Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the
original work is properly cited.
The prevalence of heart failure (HF) due to cardiac remodelling after acute myocardial infarction (AMI) does not decrease
regardless of implementation of new technologies supporting opening culprit coronary artery and solving of ischemia-relating
stenosis with primary percutaneous coronary intervention (PCI). Numerous studies have examined the diagnostic and
prognostic potencies of circulating cardiac biomarkers in acute coronary syndrome/AMI and heart failure after AMI, and even
fewer have depicted the utility of biomarkers in AMI patients undergoing primary PCI. Although complete revascularization at
early period of acute coronary syndrome/AMI is an established factor for improved short-term and long-term prognosis and
lowered risk of cardiovascular (CV) complications, late adverse cardiac remodelling may be a major risk factor for one-year
mortality and postponded heart failure manifestation after PCI with subsequent blood ow resolving in culprit coronary artery.
The aim of the review was to focus an attention on circulating biomarker as a promising tool to stratify AMI patients at high
risk of poor cardiac recovery and developing HF after successful PCI. The main consideration aects biomarkers of
inammation, biomechanical myocardial stress, cardiac injury and necrosis, brosis, endothelial dysfunction, and vascular
reparation. Clinical utilities and predictive modalities of natriuretic peptides, cardiac troponins, galectin 3, soluble suppressor
tumorogenicity-2, high-sensitive C-reactive protein, growth dierential factor-15, midregional proadrenomedullin, noncoding
RNAs, and other biomarkers for adverse cardiac remodelling are discussed in the review.
1. Introduction
Heart failure (HF) is a global health problem with serious
economic burden that has been considered as the dominant
cause of cardiovascular (CV) morbidity and mortality in
the developed and developing countries [1, 2]. HF aects
around 26 million people worldwide (including 5.7 million
and 3.4 million people in the US and in the EU, respectively),
and the estimated expenditures for HF care were around $31
billion [1]. It is expected that by 2030 more than 40 million
people will have this condition, and the HF diagnosis and
therapy will increase twice and even more [3]. The clinical
outcomes remain poor with a ve-year survival rate of
approximately 50% regardless of phenotype of HF that
completely correspond to the expected survival rate in non-
metastatic cancer [3, 4]. Despite sucient improvements in
diagnosis, prevention and treatment of HF new incidences
of HF with reduced ejection fraction (HFrEF) and midrange
ejection fraction (HFmrEF) in contrast to HF with preserved
ejection fraction (HFpEF) continue to occur as a need for
heart transplantation and mechanical support device use
[4]. Additionally, increased prevalence of HFpEF represents
the most frequent cause of CV and sudden death, primary
hospitalization, and readmission to the hospital due to acute
decompensation of HF [5].
The most common primary causes of HFrEF/HFmrEF
remain acute ST-segment elevation myocardial infarction
(STEMI) and hypertension, while incidences of HFpEF were
rather associated with hypertension, acute non-ST-segment
elevation myocardial infarction (non-STEMI), and
Hindawi
Disease Markers
Volume 2020, Article ID 1215802, 21 pages
https://doi.org/10.1155/2020/1215802
alternative causes (atrial brillation, cardiomyopathy, myo-
carditis, valvular heart disease, and diabetes mellitus) com-
pared with STEMI [68]. Contemporary strategy for the
prevention of HF after acute STEMI is based on early com-
plete cardiac revascularization and prevention of negative
impact of comorbidities, such as diabetes mellitus, abdomi-
nal obesity, hypertension, thyroid dysfunction, kidney fail-
ure, and conventional CV risk factors (smoking,
dyslipidaemia, insulin resistance, and hyperuricemia) [9,
10]. In fact, complete recovering of blood ow through cul-
prit coronary artery and other ischemia-related stenosis with
primary percutaneous coronary intervention (PCI) is not
warranted for full prevention of late adverse cardiac remod-
elling [11, 12]. Although improvement of prognosis, increase
in quality of life, and delay in progression or reversal of
ischemia-induced cardiac remodelling and chronic HF
remain prime targets for the treatment of AMI [13, 14], there
are no clear approaches for risk stratication in AMI patients
after successful PCI [15]. For instance, hyperemic microcir-
culatory resistance and no-reow phenomenon were found
as strong predictors for the extent of infarct size and early
cardiac remodelling [16]. Additionally, optic coherent
tomography or intravascular ultrasound performed over 3
months after initial major cardiac event frequently allows
identifying several factors contributing advance in late car-
diac remodelling, such as silent restenosis, progression of
old stenotic lesions, late stent thrombosis, and several post-
PCI technical problems with incomplete stent branches
expansion, stent malposition, and underpressed culprit pla-
que [17, 18]. Except for early revascularization, cardiac
remodelling could be prevented by pharmacotherapy includ-
ing complex neurohormonal blockade and device-based
therapies, which are addressed in the improvement of ven-
tricular dyssynchrony and prevention from fatal arrhythmias
[19]. In this context, new diagnostic and predictive options
are needed to prevent cardiac remodelling and HF. The aim
of the review was focused on the circulating biomarker as a
tool to stratify postmyocardial infarction patients at high risk
of poor cardiac recovery after reperfusion with primary PCI
and developing HF.
2. Adverse Cardiac Remodelling after Acute
Myocardial Infarction: Definitions and
Contributing Factors
2.1. Denition of Adverse Cardiac Remodelling. Adverse car-
diac remodelling after AMI is dened as complex interac-
tions between cellular and extracellular components of
myocardium, which are neurohumoral and epigenetic regu-
lations, leading to changes in the cardiac architectonics and
geometry frequently aecting both ventricles and atrials,
worsening diastolic lling and systolic function and associ-
ated with developing heart failure [17]. Additionally, there
is a large number of denitions of cardiac remodelling after
STEMI, which are based on multiple imaging modalities,
such as presentation of akinesia area, left ventricle enlarge-
ment, reduced LVEF, and early diastolic dysfunction (includ-
ing longitudinal strain increase, twist of LV apex, and
tethering eect). In fact, an impact of passive mechanical
constraint of surrounding myocardium on infarct zone
mediates infarct expansion and decline in both regional and
global systolic function [20].
Other criteria of cardiac remodelling aect shaping
stunned and hibernated myocardium after incomplete revas-
cularization or delay of PCI performing with inadequate per-
fusion recovery [21]. However, non-STEMI is also associated
with cardiac remodelling, rather mild-to-moderate than
severe, and frequently nondistinguished from STEMI-
induced cardiac disorders in prognostic aspects, but the
canonic model of pathogenesis of adverse cardiac remodel-
ling was based on STEMI impact on cardiac architectonic.
There is a sustainable option that STEMI-induced cardiac
remodelling frequently relates to HFrEF/HFmrEF, but non-
STEMI-induced cardiac remodelling is rather associated with
developing HFpEF than HFrEF.
2.2. Contributing Factors of Adverse Cardiac Remodelling. In
fact, there are at least two dierent variants of adverse car-
diac remodelling after acute MI, which distinguished each
other in pathogenesis, so called the early (at 2-3 weeks after
initial event) and late (at 3-6 months after AMI) remodel-
ling (see Figure 1).
Contemporary point of view is based on an idea that
early complete primary revascularization of culprit artery
and ischemia-induced stenosis/occlusions in other coronary
arteries at rst hours of STEMI is independent and the
most powerful factor preventing early LV cavity dilation,
declining LV pump function and the developing of HF. It
has been postulated that preserved systolic function and
LV dimensions at early stage of various revascularization
procedures can accompany with myocardial biomechanical
and energetics stress, mitochondrial dysfunction, and oxi-
dative stress that lead to potent fatal arrhythmias even prior
to diastolic dysfunction developing [22]. Over the next
three months after restoring TIMI III blood ow through
culprit artery with PCI, the primary causes inducing
adverse cardiac remodelling can be dierent from the afore-
mentioned. Indeed, other factors that may contribute to
cardiac remodelling after successful primary PCI are arte-
rial healing, vessel remodelling, stent restenosis, thrombosis,
and incomplete expansion of stent branches (known as mal-
position), and stent fracture, which require ischemia-driven
target vessel revascularization further [23]. Performing of
optical coherence tomography (OCT) in STEMI patients
presenting with late and very late stent thrombosis has
yielded that stent malposition was determined in 55% cases,
quarter of which had been found evidence of positive vessel
remodelling [24]. Additionally, neoatherosclerosis and
uncovered stent struts were reported as the primary cause
of late thrombosis in 35% cases and 10% cases, respectively
[24]. Although coronary stent fracture is an underrecognized
event, it has been reported frequently in the drug-eluting
stent era [25]. However, investigators have shown that tech-
nical problems with rst-generation eluting stent implanta-
tions in STEMI patients were associated with higher in-
hospital mortality and posthospital target vessel failure or
cardiac death [24].
2 Disease Markers
Endothelial shear stress, neointima formation, and late
thrombosis can appear beyond inadequate PCI and stent
positioning and are result of accelerating atherosclerosis
and inadequate drug support, i.e., nonoptimal care with
statins, refusal from dual antiplatelet therapy, eective antic-
oagulation if needed, and adenosine intracoronary for pre-
vention no-reow/slow-ow phenomena. Even a novel
device (known as bioabsorbable cardiac matrix) was not able
to attenuate adverse cardiac remodelling after AMI [26],
while there were strong positive expectations regarding these
devices [27]. Despite implantation of second-generation
everolimus-eluting stent in STEMI appears to be better to
rst-generation eluting stents, there is evidence that even a
small degree of chronic intrastent conditions may signi-
cantly inuence on healing persistence [28]. Frequencies of
uncovered and malapposed struts as well as percentage of
stents fully covered with neointima were 1.2%, 0.4%, and
60.9%, respectively, for over a one-year period after PCI with
second-generation everolimus-eluting stent implantation
[28]. In fact, they were not associated with the incidence of
clinical events and intrastent thrombus.
The next factor contributing to early and late cardiac
remodelling is the no-reowphenomenon. Indeed, the
no-reowphenomenon can be considered as a component
of early cardiac remodelling after STEMI that relates to
microvascular obstruction and dysfunction causing severe
disturbance in regional perfusion [29]. In fact, the no-
reowphenomenon is a result in poor healing of the culprit
artery and adverse cardiac remodelling, increasing the risk
for major adverse cardiac events, such as recurrent MI, newly
diagnosed HF, and sudden death, but the slow-owphe-
nomenon appears to be a serious factor contributing to both
types of adverse cardiac remodelling [30, 31].
Additional factor that is involved onto a development of
late adverse remodelling is epigenetically mediating distur-
bance of endogenous vascular repair system [32, 33]. It has
been found that altered vascular repair has maintained vaso-
constriction and vascular dysfunction that accelerated ath-
erosclerosis and supported hibernation in the grey zone
around myocardial infarction. Overall, the development of
adverse cardiac remodelling after AMI regardless of initial
cause (even in asymptomatic patients) was consistently asso-
ciated with poor clinical outcomes, and it could be predicted
and completely resolved [34, 35].
The factors preventing late adverse cardiac remodelling
after successful reperfusion with primary PCI in STEMI
patients are indicated in Figure 2. Recognition of the hetero-
geneous pathophysiology of adverse cardiac remodelling
after AMI can create a powerful risk stratication score based
on biomarkers reecting various stages of pathogenesis of the
condition [36].
3. Pathogenetic Mechanisms of Adverse Cardiac
Remodelling after Acute
Myocardial Infarction
Advances in our understanding of the molecular mecha-
nisms of regulation toward late adverse cardiac remodelling
were associated with the breakthrough in the recognition of
Delaying revascularization
Incomplete rocovering
blood ow though culprit
coronary artery
Severe ischemia-related
stenosis
Advance hyperemic
microcirculatory resistance
“No-reow” phenomenon
“No-reow” phenomenon
Inadequate healing
Hibernation and stunning
Incomplete recovering
blood ow though culprit
coronary artery
Hibernation
Microvascular obstruction
Restenosis
Progression of old stenotic
lesions
Late stent thrombosis
Stent malapposition
Comorbidities
Dysfunction of endogenous
vascular repair system
Tri g ger s
Acute myocardial
infarction
Early adverse cardiac
remodelling
Infarct expansion
Early LV dilatation
Myocardial thinning
Akinesis area shaping
Diastolic dysfunction
Reducing LVEF
Presenting HF
Late adverse cardiac
remodelling
Late LV dilatation
LV diastolic dysfunction
Interventricular dissynchrony
Regional contractility
dysfunction
HFpEF/HFmrEF/HFrEF
Figure 1: Adverse cardiac remodelling after AMI: the role of dierent triggers in development of cardiac architectonic disorders and heart
failure. LV: left ventricular; HF: heart failure; HFpEF: HF with preserved ejection fraction; HFmrEF: HF with midrange ejection fraction;
HFrEF: HF with reduced ejection fraction.
3Disease Markers
interplaying between various processes translating ische-
mia/reperfusion injury on myocardium, such as disrupting
nitric oxide (NO) and vascular endothelial growth factor
(VEGF) signalling systems, p38 MAPK pathway and redox
dysregulation, cytokine release, and activation of apoptotic
and necrotic death pathways with subsequent stimulation
of oxidative stress, mitochondrial dysfunction, altered
myocardial cell metabolism, excessive brosis, and cardiac
cell remodelling [37]. Therefore, preserved microvascular
inammation, small vessel obstruction, endothelial dys-
function, and atherosclerotic lesions mediate a remote
eect on advance LV remodelling [38]. Additionally, there
are new explanations regarding individual susceptibility to
ischemia/reperfusion injury including early and remote
ischemic preconditioning [39]. Figure 3 yields main path-
ogenetic mechanisms that are involved in the pathogenesis
of late adverse cardiac remodelling.
In fact, restoration of adequate blood perfusion after a
critical period of ischemia and prevention of reperfusion
damage appear to be not the only protector over cardiac
damage. Early irreversible cardiac myocyte injury leading to
necrosis in the ischemic myocardium and expanding infarc-
tion zone are an attribute of susceptibility of cardiac cells to
impaired metabolism, loss of structural integrity and selec-
tive permeability of the cell membranes, altered ultrastruc-
ture of cell organoids, such as sarcolemma disruption,
deterioration of nucleus, ribosomes, mitochondria, and sar-
coplasmic reticulum, the presentation of mitochondrial
amorphous densities, and chromatin fragmentation [40].
During this early stage of AMI development, the mitochon-
drial dysfunction plays a pivotal role in cardiac myocyte apo-
ptosis in the ischemic/reperfused heart, cardiac necrosis, and
ischemia-induced preconditioning phenomenon [41, 42].
Numerous studies have shown that proapoptotic stim-
uli through involving cytokines, which belong to the B-cell
lymphoma 2 (Bcl-2) super family, mediate the permeabil-
ity of the mitochondrial membranes and stimulate the
release of a wide spectrum of the active apoptogenic mol-
ecules (cytochrome c, Bax) into the cytoplasm. They cause
the apoptotic response, peroxidation of membrane, and
disruption of mitochondrial chromatin materials, including
small interfering ribonucleic acid (RNA) and mitochon-
drial deoxyribonucleic acid (DNA) [43, 44]. Cytochrome
C is able to bind to the adaptor protein apoptotic protease
activating factor 1 (Apaf-1) and act as a trigger of its olig-
omerization that activates caspase cascade through initiat-
ing procaspase-9 recruitment. Caspases including caspase-
6 and caspase-9 cleave cellular proteins and DNAs/RNAs
emerging apoptosis [45]. This process is under the close
epigenetic regulation of long noncoding RNAs (LncRNA)
and microRNAs (miRNA-29b-1-5p, miRNA-195), which
negatively regulate Bcl2l2 gene expression and participate
in cardiac myocyte apoptosis, oxidative stress through
inducing hydrogen peroxide (H
2
O
2
), and inammation
Preprocedural
Prevention of late adverse cardiac remodelling
Perprocedural Early postprocedural Late postprocedural
Optimal BP
Optimal
glycaemia
Statin initiation
Eective DAPT
Optimal time for
PCI
Radial vs femoral
access
TIMI III blood
ow recovery
Complete
revascularization
BES instead BMS
Optimal
anticoagulation
Nitrates
Anti-GPIIb/IIIa
inhibitors in
selected cases
romboectomy
in selected cases
Rotational
atheroectomy
protocol
Venous gra PCI
protocol
Reduce ischemic
time
IVUS-VH or OCT
enhancement
Hemodynamic
stabilization
Selective IC
calcium inhibitors
Adenosine
Anti-GPIIb/IIIa
inhibitors
Statins
Statins
Optimal DAPT
Optimal DAPT
IVUS-VH or OCT
control for
plaque necrotic
core component,
stent
implantation,
early stent
thrombosis/
restenosis
Beta-blockers,
ACE inhibitors or
ARBs in HF
Beta-blockers,
ACE inhibitors or
ARBs in HF
MCRA in
LVEF<35% MCRA in
LVEF<35%
Optimal
treatment
comorbidities
Preventing
transient
reversible
ischemia
Qualitative oine
analysis of IVUS-
VH or OCT
image for stent
implantation, late
stent thrombosis
/ restenosis,
neoatheroscleros
is, positive vessel
remodelling
Figure 2: The factors preventing late adverse cardiac remodelling in AMI patients after successful reperfusion with PCI. IVUS-VH:
intravascular ultrasound virtual-histology; BMS: bare metal stent; BES: biolimus eluting stent; OCT: optical coherence tomography;
DAPT: dual antiplatelet therapy; ACE: angiotensin-converting enzyme; ARBs: angiotensin-II receptor antagonists; MCRA:
mineralocorticoid receptor antagonists; IC: intracoronary.
4 Disease Markers
via triggering proinammatory cytokine release [44, 46].
Therefore, downregulated miRNA-98 and miRNA-124 may
attenuate cell survival through diminished levels of STAT3
and p-STAT3 in response to ischemia and over production
of H
2
O
2
[47, 48]. During ischemia/reperfusion episodes, oxi-
dative stress, mitochondrial Ca
2+
overload, proinammatory
cytokines (interleukin- (IL-) 2, IL-6, tumor necrosis factor-
alpha, and interferon-gamma) stimulate the activity of the
matrix metalloproteinases and suppress release of their tissue
inhibitors [49]. MMPs (MMP-2, MMP-6, and MMP-9)
directly contribute to global and local myocardial contractile
dysfunction and induce cell death [50]. Other matricellular
proteins, such as thrombospondin- (TSP-) 1 and TSP-2, as
well as bone-related proteins (osteopontin, osteonectin, and
osteoprotegerin), were found to regulate cardiac reparation
and remodelling via activation of VEGF and transforming
growth factor (TGF-β) by binding to the latency-associated
propeptide, inhibition of MMP activity, and exertion of potent
angiostatic actions of antigen-presenting cells and T-cells [51].
Moreover, they are triggers for accumulation, degradation and
remodelling of extracellular matrix (ECM), cleaving big
endothelin-1 and attenuating vasoconstriction, and modica-
tion of architectonics of myocardium leading to cardiac
remodelling and HF development [51, 52].
Interestingly, susceptibility of myocardium to ischemia
and reperfusion may relate to various inhered causes, such
as mutations in genes encoding for angiotensin II,
angiotensin-converting enzyme, osteopontin, osteoproteg-
erin, CC chemokine receptor 2, the members of the family
of multidomain extracellular protease enzymes ADAMTS
(A Disintegrin and Metalloproteinase with Thrombospon-
din motifs), predominantly ADAMTS-2, ADAMTS-4,
ADAMTS-10, and ADAMTS-13, promoter region of endo-
thelial NO synthase, apelin, TGF-β, VEGF, galectin-3, co-
lin-1, S100 calcium-binding protein A9, and mitochondrial
aldehyde dehydrogenase 2 (NDUFC2) [5356]. Finally, sus-
ceptibility of cardiac cells to ischemia and reperfusion dam-
age may relate to the capability of endogenous redox
systems to protect cell membranes and cellular structures
(mitochondria, cytoskeletal proteins, growth factor receptors,
and microtubule-associated proteins) from the impaired eect
of the deteriorating energetic metabolism and detergenting
impact of oxidized lipids and proteins sustaining an eective
work of transmembrane ionic pumps [57]. This phenomenon
was called ischemic preconditioning, and now it is also recog-
nized as an early (before AMI or during acute phase of AMI)
and remote (overreparative period of AMI) phenomenon
depending on a period of onset of ischemia-reperfusion epi-
sodes. However, previous studies have revealed a reduction
of infarct size and peripheral area with hibernating/stunning
myocardium with both types of preconditioning due to intra-
cardiac protection that prevents cytosolic and mitochondrial
Ca
2+
overload, accumulation of reactive oxygen species
(ROS), lysosomal/nonlysosomal enzyme releasing, and
inammatory reaction [5860].
Recurrent episodes of ischemia/reperfusion induce cardi-
oprotective mechanisms in failed heart named postcondi-
tioning and remote conditioning [61, 62], which are
supported by various comorbidities (diabetes mellitus, insu-
lin resistance, obesity, and inammatory conditions) [63,
64]. The cardiac protective mechanisms may include upregu-
lation of caveolin, resolvin D1/E1, ubiquinone, long pentra-
xin PTX3, apelin, glucocorticoids, and long noncoding
RNAs expression for IL-19, VEGF, eNO synthase, haem
Acute MI/ Acute
coronary syndrome
Early period aer
successful PCI
Remote period
aer PCI
First hours before
restoration of coronary
circulation
Over 3 months aer PCIWithin 3 months aer PCI
Myocardial necrosis Distal embolization Remote conditioning
Ischemia/reperfusion damage
Expanding necrosis
Preconditioning
Severity of coronary
atherosclerosis
Comorbidities
Comorbidities
Preexisting HF
Individual susceptibility
Stunning/hubernation
no-reow/slow-ow
Postconditioning
Transient ischemia/reperfusion
Post-PCI technical problems
Atherosclerosis progression
Atherosclerosis progression
Inammation and brosis
Inammation and brosis
Impaired cardiac and vascular
reparation
Hibernating myocardium
Remote ischemia/reperfusion
episodes
ECM and cell remodelling
Impaired cardia and vascular
reparation
Figure 3: The main pathogenetic mechanisms underlying the initiation and progression of late adverse cardiac remodelling in AMI patients
after successful reperfusion with PCI. HF: heart failure; ECM: extracellular matrix; PCI: percutaneous coronary intervention.
5Disease Markers
oxygenase-1 (HO-1), calcitonin gene-related peptide, and
peroxisome proliferator-activated receptor gamma, and
downregulation of β-adrenergic signalling, G protein-
coupled receptor kinase-2, and β-arrestin 1 and 2 in cardiac
myocytes, broblasts/myobroblasts, tissue residence cells,
and circulating progenitor cells as well as mononuclears
[6568]. These factors reduce inammatory inltrates, stabi-
lize cell membrane, support membrane ionic channels, and
suppress the formation of key proinammatory cytokines,
such as tumor necrosis factor-alpha (TNF-α), IL-1β, and
IL-6. Additionally, IL-19 suppresses the polarization of pro-
inammatory subtype M1 macrophages and triggers M2
macrophage polarization in infarct myocardium that leads
to inhibition of cardiac remodelling [68].
During AMI and recurrent episodes of micronecrosis in
myocardium after PCI due to remote ischemia/reperfusion
damage, the important role in the regulation of cardiac
remodelling belongs to alarmins, which are released by
necrotic myocardium and act as a powerful trigger of inam-
matory cytokine synthesis [69]. Damaged and necrotic car-
diac myocytes secrete wide-spectrum factors called DAMPs
(Damage-Associated Molecular Patterns), such as high-
mobility group 1B protein (HMGB1), RNA, nucleotides, heat
shock proteins (HSP), members of the S100 family, and IL-
1a, which potentiate the inammatory response, attenuate
oxidative stress, act as direct cytotoxic agents, and induce
thrombus formation and circulating blood cell aggregation
[70]. Numerous molecules, such as HMGB1, S100 family
members, are able to induce apoptosis of circulating endo-
thelial progenitor cells and tissue residence cells via a multi-
ligand receptor for advanced glycation end products-
(RAGE-) mediated activation of endoplasmic reticulum
stress pathway [71]. Therefore, the DAMPs and other che-
mokines, such as CXC and CC (predominantly CCL2,
CCR2, CCR5, and ELR+CXC chemokines), recruit various
subpopulations of peripheral blood cells including proin-
ammatory mononuclears, regulatory T-cells, nature killers,
and neutrophils in the infarcted myocardium and endothe-
lium supporting inammatory response [72].
Inammation is a crucial element for clearance of cellu-
lar and matrix debris, while suppression of proinamma-
tory signalling is necessary to transform the inammatory
phase to the proliferative phase [73]. Indeed, proinamma-
tory mediators include uncoupling protein 2, superoxide
dismutase- (SOD-) 1 and SOD-2, ROS, through the activa-
tion of mTOR, hypoxia-induced factor- (HIF-) 1, Toll-like
receptor (TLR)/IL-1, and RAGE-dependent pathways in
surviving border-zone broblasts, cardiac myocytes, endo-
thelial cells, smooth muscle cells, mononuclears, and several
residence and progenitor cells mediating reparative pro-
cesses [7476]. It relates to the modication of cardiac
broblasts into myobroblasts that are enriched in α-
smooth muscle actin, accumulation of extracellular matrix,
neovascularization, and angiogenesis. However, the proin-
ammatory cytokines may have a detrimental impact on
cardiac remodelling and function directly maintaining
repetitive ischemia/reperfusion episodes, suppressing repa-
ration, supporting endothelial dysfunction, coagulation,
and thrombosis [7779].
Over a 3-month period after PCI, the extracellular matrix
is continually being remodelled, and tissue broblasts, myo-
broblasts, and antigen-presenting cells become quiescent
and undergo apoptosis, and cell debris is cleared by macro-
phages [8082]. The regulation of proliferative response
and changing of cellular phases of tissue inammation are
mediated by the renin-angiotensin-aldosterone system
(RAAS) and simpatico adrenal system, which also are central
players in the endogenous repair system [83, 84]. Addition-
ally, the autonomic nervous system may play a crucial role
in the inammatory and apoptotic remodelling following
AMI [85]. Thus, late adverse cardiac remodelling is a sophis-
ticated structural and functional response of failing heart to
numerous triggers (inammation, brosis, cell survival sig-
nalling, and β-adrenergic signalling) and damaged factors
(ischemia, reperfusion, necrosis, and apoptosis) that appear
consequently and mutually activate each other.
4. Biomarkers of Adverse Cardiac Remodelling
Although there are well-developed current clinical recom-
mendations for HF provided by the experts of the European
Society of Cardiology (2016) [86] and American College of
Cardiology/American Heart Association (2017) [87], there
is a lack of statements for the use of biomarker strategies
for diagnosis, prediction, stratication, and prevention of
adverse cardiac remodelling. In fact, cardiac remodelling
after AMI regardless of PCI and other approaches for revas-
cularization is strongly associated with the development and
progress of HF. In this context, early biomarkers of myocar-
dial injury and necrosis, as well as biomarkers of biomechan-
ical stress, neurohumoral and inammatory activation, and
brosis, having predictive and diagnostic evidence for acute
and chronic HF, are extrapolated over strategy regarding
diagnosis, outcomes, and stratication of adverse cardiac
remodelling (Table 1).
There is no complete agreement between experts from
the European Society of Cardiology and American College
of Cardiology/American Heart Association regarding the
utility of biomarkers in HF [88]. Natriuretic peptides (NPs)
are recommended by both guidelines for acute and chronic
HF diagnosis, prediction of HF-relating outcomes, including
death, and a risk stratication. In contrast, the European
Society of Cardiology (2016) HF clinical recommendation
does not consist the supporting evidence regarding other bio-
markers for multitask strategy in HF, and HF-guided therapy
is not routinely recommended, while the HF biomarker guid-
ance was previously approved by the American College of
Cardiology/American Heart Association. Additionally, there
was poor discrimination when NPs were used in patients
with HF at hospital discharge, which was inferior to its per-
formance in patients with ambulatory HF regardless of sever-
ity cardiac dysfunction and phenotypes.
However, there is a large body of evidence that other bio-
markers (growth/dierential factor-15, MMP-2, MMP-6,
MMP-9, adipocytokines (apelin, chemerin, and visfatin),
circulating endothelial and mononuclear progenitor cells,
activated and apoptotic endothelial cell-derived microvesi-
cles, miRNAs, and bone-related proteins) reecting dierent
6 Disease Markers
stages of the pathogenesis of adverse cardiac remodelling
after PCI can be considered as promising tools for further
strategies to improve prediction of clinical outcomes, attenu-
ate CV risk stratication, and develop personifying strategy
for treatment [89, 90]. For instance, miRNAs are speculated
to have crucial roles in the nature evolution of adverse car-
diac remodelling after AMI, and identication of key genes
associated with damaged heart response could improve pre-
diction models for the patients [91]. Moreover, miRNA pro-
ling and gene cards with information about a signature of
mutations involved in the regulation of the transcription fac-
tors, which mediate cardiac remodelling, appear to be prom-
ising for further precise medicine after PCI [92].
5. Biomarkers of Cardiac Injury and Necrosis
Elevated levels of high-specic cardiac troponins T (hs-TnT)
and I (hs-TnI) in peripheral blood are served as diagnostic
and predictive biomarkers for acute coronary syndromes
and AMI [93], as well as an independent prognosticator of
CV risk in the general population [94]. Cardiac troponins
are structure proteins of actin-myosin complex, which are
released from the cells due to necrosis or leakage from cytosol
through the permeable cell membrane [95]. High-sensitivity
cardiac troponin assay allows diagnosing patients with minor
myocardial injury and suggesting a size of infarction [96].
Cell-free pool of cardiac troponins was reported having a ten-
dency to decrease after AMI, while peak concentrations of
both hs-TnT and hs-TnI have strongly predicted major car-
diovascular events including death, recurrent MI, need of
PCI, and subsequent HF hospitalization [96, 97]. Moreover,
elevated concentrations of circulating cardiac troponins
remain useful independent predictive biomarkers of newly
post-AMI HF [98, 99]. Interestingly, elevated levels of hs-
TnI were associated with CV death, whereas hs-TnT has
more strongly predicted the risk of non-CV death [100]. In
fact, cardiac and noncardiac surgeries mediate the elevation
of troponins in the peripheral blood postprocedurally. It
requires specic approach to assay an impact of transient ele-
vation of these ndings on a risk of poor prognosis. Obvi-
ously, the combined biomarkersmodels are necessary.
After a prolonged period of hopes regarding improve-
ment of diagnostic and risk stratication in STEMI patients
with subsequent PCI using the combined biomarker models
(cardiac troponins, NPs, copeptin, choline, soluble ST2,
GDF-15, high-sensitivity C-reactive protein, galectin-3, and
lipoprotein-associated phospholipase A2) [101], it has clearly
become what large clinical trials need to evaluate diagnostic
and predictive values of various combinations of biomarkers,
because the evidence of previous studies in AMI patient
treated with PCI appeared to be controversial [102, 103].
Copeptin did not add diagnostic information to peak con-
centration of high-sensitive troponin T in STEMI patients
with subsequent PCI [104, 105]. Yet, hs-TnT/hs-TnI and
NT-probrain NP (NT-proBNP) were recognized to have
similar predictive values for all-cause mortality and rst
readmission in HFpEF [106, 107], whereas NT-proBNP
was superior to cardiac troponins for the prognostication of
HFrEF clinical outcomes [108, 109]. It has been noted that
the predictive value of hs-TnI for HF-related clinical out-
come was strongest in men with HFpEF/HFrEF than in
women [108]. Other biomarkers, including soluble ST2,
Table 1: Clinical relevance of circulating biomarkers for late adverse cardiac remodelling: overlap with HF.
Biomarkers Heart failure Adverse cardiac remodelling
Diagnosis Outcomes Guided therapy Risk stratication Diagnosis Outcomes Risk stratication
Currently used or recommended biomarkers
hs-troponin T/I
҂
-++ ++++
NPs
#҂
++ +++ + ++ ++ +++ ++
MR-proADM + +++ - ++ + +++ ++
Galectin-3
҂
-+ - + -++ +
sST2
҂
++ + - - +++ +
Promising biomarkers
Copeptin + ++ - + + ++ +
GDF15 ++ - + - ++ ++
hs-CRP +- - ++
IL-1β+- + ++
IL-6 +- + ++
MMP-2 +––++ +
MMP-9 +––++ +
CTPpC-I ++-+++
APpC-III ++-+++
miRNAs ++ + -+ +
-
Mildly disagree;
moderately disagree;
+
mildly agree;
++
moderately agree;
+++
strongly agree;
#
approved by the European Society of Cardiology (2016);
҂
approved by the American College of Cardiology/American Heart Association (2017). hs: high sensitive; HF: heart failure; NPs: natriuretic peptides; sST2:
soluble suppression of tumorigenicity-2; MR-proADM: midregional proadrenomedullin; GDF: growth/dierential factor; CRP: C-reactive protein; miRNAs:
microribonucleic acids; MMP: matrix metalloproteinase; CTPpC-I: carboxytelopeptides of procollagen type I; APpC-III: aminopeptide of procollagen type III.
7Disease Markers
high-sensitivity C-reactive protein, galectin-3, midregional
proadrenomedullin, and GDF-15, in combination with
hs-TnI/hs-TnT did not represent superiority in compari-
son with the isolated use of hs-TnI/hs-TnT in HFpEF,
whereas in patients with HFmrEF/HFrEF, multimarkers
strategy was better in the prognostication of poor progno-
sis [110, 111].
Although previous clinical trials did not nd signicant
interactions between stable HFpEF and HFrEF when con-
sidering the prognostic value of the NT-proBNP, cystatin-
C, hs-TnT, and soluble ST2 [112114], it can be otherwise
for HF that is associated with adverse cardiac remodelling
after AMI with subsequent PCI. Thus, clinical prediction
models for HF-related outcomes based on various biomarkers
of biomechanical stress (NT-proBNP, copeptin, midregional
proadrenomedullin (MR-proADM), and growth/dierential
factor- (GDF-) 15), inammation (high-sensitivity C-reactive
protein), and brosis (galectin-3, soluble ST2) were only
improved marginally by the addition of hs-TnT/hs-TnI.
Moreover, hs-TnT or hs-TnI added to NT-proBNP and
sST2 appears to be emerging biomarkers in the prediction of
adverse outcome of HF after AMI in a short-term period
[115], but whether this combination is most suitable for
remote prognostication in patients with known late adverse
cardiac remodelling and dierent phenotypes of ischemia-
induced HF is not fully clear.
6. Inflammatory Biomarkers
6.1. Interleukins. IL-1β, IL-6, and angiopoietin-like protein 2
(Angptl 2) are inammatory cytokines that inuence delete-
rious eects on myocardium structure and function unleash-
ing to cardiac remodelling [116]. There is strong evidence
clarifying that the myocardial expression levels of IL-1β, IL-
6, and Angptl 2 were signicantly higher in the AMI patients
than in the healthy volunteers [117]. Moreover, the levels of
Angptl 2 and IL-6 rather correlated with the severity of
coronary atherosclerosis than the size of the infarct area
and HF presence. In contrast, IL-1βlevels were associated
with prior HF admissions, functional cardiac impairment,
and higher NT-proBNP, sST2, and hs-TnT concentrations
[115]. In fact, circulating IL-1βlevels had been clinically
meaningful in HF patients interfering with the predictive
ability of sST2. Indeed, regardless of LVEF, HF patients
with low sST2 (35.0 ng/ml) and also low IL-1β
(49.1 pg/ml) had signicantly lower risk of CV death,
HF-related outcomes including readmission, than among
patients with high sST2 (>35.0 ng/ml) and also high IL-
1β(<49.1 pg/ml) levels [115].
6.2. Soluble Suppression of Tumorigenicity-2. Serum levels of
IL-33 and soluble suppression of tumorigenicity-2 (sST2),
which is the soluble form of IL-1 receptor-like 1 (IL-33), were
signicantly higher in HF regardless of the presence HF phe-
notypes associated with HF symptom severity, LV hypertro-
phy, and the risk of CV death and hospitalization than in
healthy volunteers [118, 119]. It was found that IL-33
improved cell viability after ischemia injury through ST2 sig-
nalling and suppression nuclear factor kappa-B that
unleashed the upexpression of the antiapoptotic factors
(XIAP, cIAP1, surviving) and HIF-1, preventing apoptosis
[120]. In patients with AMI, serum levels of sST2 were found
to be increased, and after adjustment for comorbidities,
the Killip class and troponin T sST2 independently pre-
dicted the excess risk of death and HF [121]. Development
of adverse cardiac remodelling due to AMI was strongly
associated with the elevated levels of sST2 in the periph-
eral blood [122].
Serum sST2 served as a predictive biomarker in
patients at risk of HF and in individuals with established
chronic HF [123], but the prognostic value of the bio-
marker was diminished after adjusting for the clinical sta-
tus including comorbidity presence (abdominal obesity,
diabetes mellitus, and obstructive pulmonary disease) and
NT-proBNP [124126]. Additionally, sST2 was able to be
helpful in short-term clinical outcome prognostication in
acute HF and actually decompensated HF patients regard-
less of worsening kidney function, whereas renal failure
was found to be a crucial factor for the NP predictive
value [127, 128]. In-patients survived after acute HF have
yielded the concentrations of sST2 at discharge which
were independently associated with sudden death, CV
death, HF-related death, and HF readmission during the
3-month period after discharge [127, 128]. Yet, sST2
yielded strong, independent predictive value for all-cause
and cardiovascular mortality, and HF hospitalization in
chronic HF, and deserves consideration to be part of a
multimarker panel together with NT-proBNP and hs-
TnT [129]. The PARADIGM-HF trial (Prospective Com-
parison of ARNI With ACEI to Determine Impact on
Global Mortality and Morbidity in Heart Failure) has
revealed the levels of sST2 increased at 1 month which
were associated with worse subsequent HF clinical out-
comes, and the decreased sST2 concentrations were related
to better prognosis particularly related to declined CV
death and HF admission [130].
6.3. C-Reactive Protein. High-sensitive C-reactive protein
(hs-CRP) has also markedly improved the risk stratication
of acute HF and acutely decompensated HF patients in mul-
tibiomarker models, which predominantly included MR-
proADM and NT-proBNP [131, 132]. However, circulating
levels of hs-CRP were associated with the New York Heart
Association functional class of HF, primary hospitalizations
and readmission predominantly in patients with HFrEF,
but not HFpEF [133]. Unfortunately, hs-CRP did not add
incremental value to NPs, sST2, and galectin-3 in patients
with HFpEF rather than HFpEF [133, 134]. The ASCEND-
HF trial has reported that the levels of hs-CRP at admission
in acute HF patients were not associated with acute dyspnea
improvement, in-hospital death, advancing HF, short-term
(30 days) and long-term (180 days) mortality, and HF read-
mission [135, 136]. On the contrary, at 30 days, elevated
levels of hs-CRP among survivors were associated with
higher 180-day mortality and readmission [135]. Although
hs-CRP is under ongoing investigations, potential treatment
options and goals of the therapy among HF individuals are
not fully determined.
8 Disease Markers
6.4. Growth Dierential Factor-15. Growth dierentiation
factor- (GDF-) 15 is determined as an inammation and oxi-
dative stress biomarker, which belongs to the TGF-βcyto-
kine superfamily and is highly expressed in myocardium
and endothelial cells in CV disease including HF [137]. Pre-
vious studies have shown that GDF-15 protected the myocar-
dium from ischemia and reperfusion injury [138, 139].
Higher serum levels of GDF-15 were associated with poor
prognosis in acute HF independent from concentrations of
NPs [140] and chronic HF irrespective of LVEF [141, 142].
Moreover, the Valsartan Heart Failure Trial has shown that
serial measurements of GDF-15 had increased the incre-
mental predictive power to the only measure at baseline
for the severity of HF and prognosis [143]. Additionally,
the elevated serum level of GDF-15 was the most prognos-
tic biomarker in comparison to NT-proBNP, hs-CRP, and
hs-TnT, in predicting long-term mortality in advanced HF
[144]. Overall, a multimarker model based on NT-proBNP,
hs-CRP, GDF-15, and hs-TnT had more predictable HFrEF
and HFpEF than the isolating biomarker [145, 146]. Prob-
ably, inammatory mediators, such as sST2 and GDF-15,
as it is expecting, can become molecular targets not only
for the diagnosis but also for the treatment of adverse car-
diac remodelling in the future.
7. Biomarkers of Cardiac Fibrosis
7.1. Galectin-3. Over the last decade, galectin-3 had been
widely investigated as a biomarker of brosis and inamma-
tion with a promising predictive value for HF development
and CV events [147]. Galectin-3 is multifunction β-galacto-
side-binding protein, which belongs to lectin family and is
expressed in several tissues and circulating cells, such as
mononuclears, macrophages, progenitor cells, mast cells,
and neutrophils [148]. Galectin-3 plays a pivotal role in
inammation, brosis, immunity, tissue repair, and cardiac
remodelling and acts as a mediator of the development and
progression of the diseases, for which these pathogenetic
stages are crucial [149, 150]. Indeed, galectin-3 is expressed
in myocardium releasing from activated macrophages and
contributes cardiac dysfunction through the remodelling of
ECM and accumulation of collagen [151]. Additionally,
galectin-3 is able to mediate cardiac and vascular brosis
induced by overexpressed aldosterone [152]. However, there
is evidence conrming the role of polymorphism of galectin-
3 gene in susceptibility to cardiac injury and brosis [153].
Being a mediator of both mutual relating processesinam-
mation and brosisgalectin-3 was approved by the Food
and Drugs Administration (USA) as a predictive biomarker
for HF development and progression [87, 154]. In fact, ele-
vated levels of galectin-3 were found in patients with adverse
cardiac remodelling regardless of HF phenotypes and it eth-
nologies [155, 156]. Therefore, galectin-3 having some
advantages to NPs (more stability and resistance against
hemodynamic overload and unloading state) predicted CV
mortality and rehospitalization in HFrEF and HFpEF [157,
158]. Moreover, the TRIUMPH (Translational Initiative on
Unique and Novel Strategies for Management of Patients
with Heart Failure) has shown that repeated measures of
serum levels of galectin-3 could be useful in routine clinical
practice for HF prognostication and treatment monitoring
[159]. However, head-to-head comparison of sST2 and
galectin3 has revealed the superiority of sST2 in long-term
risk stratication in an ambulatory stable HF [160]. For
future direction, these facts require to be investigated in detail
in large clinical trials with large sample size, because a meta-
analysis of a discriminative value of galectin-3 did not yield a
conrmation of previously received data [161].
7.2. Biomarkers of Collagen Turnover. It has been postulated
that biomarkers of collagen turnover, such as carboxy-
terminal telopeptide of collagen type I, amino-terminal pro-
peptide of type III procollagen, MMPs, and tissue inhibitors
of MMPs, may be useful for risk stratication of cardiac
remodelling associated with HFpEF and HFrEF [162, 163].
Indeed, myocardial brosis being a major cause of diastolic
dysfunction contributes predominantly to the HFpEF [164].
The ECM rearrangement corresponds to an intensity of the
inammation in myocardium, and serum levels of bio-
markers of collagen turnover are mediated by a balance
between degradation of ECM components and synthesis.
Proliferative phase complimented to myocardial brosis is
considered a typical response during late adverse cardiac
remodelling, whereas increased degradation of ECM is suit-
able for AMI and early cardiac dilatation [165]. In fact,
MMP-2, MMP-9, carboxytelopeptides of procollagen type I,
and aminopeptide of procollagen type III had a predictive
value for HFpEF that was equal NT-proBNP [163], while dis-
criminative ability of elevated serum levels of MMP-2 was
superior to NT-proBNP for early HFpEF [162, 163, 166].
Whether emerging biomarkers of ECM rearrangement and
collagen turnover is essential to identify asymptomatic
patients with HFpEF after AMI with subsequent PCI is not
fully clear, while a loss of myocardial collagen scaolding
plays a pivotal role in adverse cardiac remodelling with poor
prognosis. Interestingly, elevated levels of C-terminal telo-
peptide were associated with global LVEF, the risk of CV
death, and newly diagnosed or worsening HF due to various
causes [167169]. In this context, integrity of ECM bio-
markers into personifying predictive strategy in AMI patients
appears to be promised, because multiple biomarkers
approach with traditional biomarkers and indicators of
ECM turnover may have increased the sensitivity and speci-
city of clinical outcomes in patients with adverse cardiac
remodelling and isolated diastolic dysfunction.
8. Biomarkers of Biomechanical
Myocardial Stress
8.1. Natriuretic Peptides. The physiologically natriuretic pep-
tide (NP) system mediates water and sodium homeostasis
playing a pivotal role in blood pressure enhancement, uid
retention, vascular function, structure remodelling of the
heart, kidney, and vessels, and maintaining dierentiation
and repair tissue, and supports immunity, metabolic
response, and inammation [170]. There are at least four
members of NP system, such as atrial NP (ANP), brain NP
(BNP), C-type of NP, and D-type of NP [171]. Biological
9Disease Markers
eects of NPs are provided through interacting with appro-
priate receptors: NPR-A, NPR-B, and NPR-C. Kidney eects
of NPs are diuresis and wateresis due to the decreasing tubu-
lar reabsorption of sodium and water, increasing glomerular
ltration rate (GFR) in result of inducing aerent arteriole
vasodilation, and protection of the kidney from metabolic
and ischemia injury [172]. Vascular eects of NPs corre-
spond to vasodilation, support, capillary permeability and
vascular reparation, and antiproliferative and hypocoagula-
tive eects [173]. NPs mediate cardiac protection with
respect to decreasing preload and afterload, diminishing bio-
mechanical stress, and maintaining anti-ischemic, antiprolif-
erative, and antiapoptotic abilities. Therefore, NPs have
direct inotropic and antiarrhythmic eects [174]. Overall,
the NP system is a physiological antagonist of RAAS and
the sympathoadrenal system. The main triggers for synthesis
and release of NPs are myocardial stretching, uid retention,
increase of pre- and postload, BP elevation, decreasing GFR,
and ischemia of target organs (kidney, heart, and brain).
Therefore, adipocytes and glial cells can produce NPs as a
result of proinammatory stimulation [175].
Increased activity of a circulating and local NP system
was determined in patients with CV disease including LV
hypertrophy, AMI, stable coronary artery disease, hyperten-
sion, and HF [176]. However, there are large numbers of
causes distinguishing from CV and accompanying elevation
of circulating levels of NPs (see Table 2). There is a large body
of evidence showing that NP production occurs in close rela-
tion to the severity of LV systolic dysfunction, and the circu-
lating levels of BNP and ANP strongly correspond to the
New York Heart Association functional class of HF [88].
However, the production of NPs in advanced HF became
blunt and irrespective of how high concentration of NPs
in peripheral blood uid retention, vasoconstriction, and
cardiac dysfunction appears to progressed. In contrast,
adequate treatment of HF, which is associated with
improvement of clinical status and increase of tolerance
to physical exercise, corresponds to declining circulating
levels of BNP and ANP [177].
Therefore, patients with abdominal obesity frequently
present less levels of BNP that it is expected due to increased
circulating levels of neprilysin, which degradates BNP [178].
Although older age and female sex are the most common rea-
son association with increased levels of NPs in circulation
beyond relative causes, some structural abnormalities corre-
sponding to decreased mean e
velocity and increased mitral
early ow velocity/early diastolic tissue velocity ratio can be
found [179181].
Current clinical recommendations are considered NPs
predominantly BNP, NT-proBNP, and NT-proANP, as diag-
nostic and predictive biomarkers for HF regardless of LVEF,
as well as a tool for risk stratication in general population
[86, 87]. However, elevated levels of NPs (BNP 100 pg/ml
or NT proBNP 300 pg/ml;orBNP 300 pg/ml or NT
proBNP 900 pg/ml if in atrial brillation/utter) in patients
with suspected HFmrEF/HFpEF were found to conrm the
diagnosis [182]. NPs are also excellent prognostic biomarkers
of adverse cardiac remodelling after AMI, whereas the clini-
cal value of such discriminative ability is less clear than estab-
lished acute and chronic HF [183]. Therefore, decreased
levels of NT proBNP < 1000 pg/ml as a result of HF therapy
was associated with lower 180-day mortality and readmission
in comparison with NT proBNP 1000 pg/ml, whereas
NT-proBNP reduction of >30% from initial levels did not
improve 6-month outcomes and was not more eective than
a traditional treatment [184186]. Overall, elevated levels of
NPs including NT-proBNP and NT-proANP had higher
negative diagnostic value than the positive diagnostic value
for HF, while the positive predictive ability of NPs in elevat-
ing concentrations was superior to the negative predictive
value for asymptomatic cardiac remodelling, as well as HF
regardless of LVEF. In fact, high individual variability,
depending on the serum levels of NPs on comorbidities,
including GFR, abdominal obesity, and older age and
female sex, gives more opportunities to rule out major
structural cardiac abnormalities and HF, when NP levels
are normal or near normal. Conrmation of the HF and
cardiac remodelling with isolating diastolic dysfunction
requires more predictive information including clinical
conditions, diastolic characteristics, measure of LVEF, and
other biomarker assay.
8.2. Copeptin. Copeptin is a stable 39-aminoacid glycopep-
tide derived from C-terminal portion of the precursor of argi-
nine vasopressin, which is a key regulator of water
homeostasis and plasma osmolality [187]. Serum levels of
copeptin have exhibited close linear correlation with concen-
trations of arginine vasopressin and are use as surrogate bio-
marker of its secretion [188]. There is evidence that elevated
serum levels of copeptin are a diagnostic biomarker of
Table 2: CV and non-CV causes of elevating NPs in peripheral
blood.
CV causes Non-CV causes
Acute and chronic HF Sepsis/shock
LV hypertrophy Severe infections
Pulmonary hypertension Critical ill patients
ACS/AMI Acute and chronic kidney failure
Stable CAD Severe trauma/surgery
Multifocal atherosclerosis Chronic obstructive pulmonary
disease
Cardiomyopathies Severe bronchial asthma
Myocarditis Pneumonia
Atrial brillation and utter Large burns and frostbite
Hypertension Stroke
Congenital and acquired
valvular heart disease Kidney amyloidosis
Pericardial disease Diabetes mellitus
Cardiac toxicity due to
tumoricidal therapy Thyroid dysfunction
Electrical
cardioversion/ablation Anemia
Successful resuscitation Pleural disease
HF: heart failure; LV: left ventricular; ACS: acute coronary syndrome; AMI:
acute myocardial infarction; CAD: coronary artery disease.
10 Disease Markers
asymptomatic cardiac remodelling, HF, sepsis, acute kidney
injury, insulin resistance, and metabolic syndrome [189].
Several trials have yielded that increased levels of copeptin
were strong predictor of mortality in patients with acute
and chronic HF [189, 190], stroke [191], end stage of renal
disease [192], stable CAD [193], and diabetes mellitus
[194]. However, there is a large number of confounding fac-
tors (hydration status, gender, blood pressure, GFR, and
body mass), which make it dicult to interpret data of
copeptin levels in patients with known CV disease, as well
as in healthy individuals [195]. Additionally, copeptin was
not better than the NPs in the diagnosis and prognosis of
HF as well as in prognostication of adverse cardiac remodel-
ling after AMI [196].
8.3. Midregional Proadrenomedullin. Midregional proadre-
nomedullin (MR-proADM) is stable peptide fragment
that is precursor for adrenomedullin (ADM) and gener-
ated through posttranslational processing from pre-
proadrenomedullin [197]. ADM is expressed in several
tissues (adrenal medulla, brain, kidney, lung, spleen, liver,
and vasculature) and cells (endothelial cells, cardiac myo-
cytes, vascular smooth muscle cells, and epithelial cells)
and mediates natriuresis, diuresis, vasodilation, positive
inotropic eect, and hypotension [198].
Early clinical trials have shown that circulating levels of
MR-proADM were signicantly increased in patients with
acute HF and STEMI [199, 200], and a cut-ovalue of
0.79 nmol/l has been yielded to be asso ciated with adverse
outcomes including death [201, 202]. Additionally, serum
levels of MR-proADM >0.70 nmol/l were proposed to be
the rule-in criteria of AMI [203].
The MR-proADM has become a biomarker that was
specically investigated as a possible prognosticator of
acute HF and early outcomes in STEMI patients under-
going PCI. The BACH (Biomarkers in Acute Heart Fail-
ure) study revealed that increased serum levels of MR-
proANP were a useful diagnostic biomarker as BNP for
acute HF in patients with acute dyspnoe [204]. The
results of the DANAMI-3 (The Danish Study of Optimal
Acute Treatment of Patients with ST-segment-elevation
myocardial infarction) study have shown that elevated
levels of MR-proADM were strong predictor of short-
and long-term mortality and hospital admission for HF
after AMI [205]. Unfortunately, MR-proADM has dem-
onstrated predictive ability with high similarity to BNP,
MR-proANP, and copeptin for one-year all-cause mortal-
ity in acute HF [206]. However, the measure of MR-
proADM may give additional diagnostic and prognostic
information for incident CV events associated with
advanced atherosclerosis that is useful for risk stratica-
tion among patients with adverse cardiac remodelling
after AMI with subsequent PCI [207, 208]. Therefore,
MR-proADM was able to predict major adverse cardiac
events in patients suspecting AMI regardless of HF
[209]. Moreover, in contrast to NPs, MR-proADM did
not exhibit lowered concentration in obese patients with
known HF that may facilitate diagnosis and prognosis
of HF in this patient population [210].
9. Other Biomarkers of Cardiac Remodelling
9.1. Noncoding RNAs. Noncoding RNAs are powerful epige-
netic regulators of cardiac gene expression and mediators of
cardiac homeostasis and functions [211]. There are several
types of noncoding RNAs, such as microRNAs (miRNAs),
long noncoding RNAs, and circular RNAs, which play a cen-
tral role in the regulation of numerous pathogenetic mecha-
nisms and coordinate coupling of morbidity state with
susceptibility to inammatory and proliferative response
[212]. Among these types of noncoding RNAs, various miR-
NAs are widely investigated (see Figure 4). Although there is
a large body of evidence regarding up- and downregulation
of genes for potassium channels, SERCA, subunits of recep-
tors, signal molecules, proinammatory cytokines, apoptotic
mediators (Bax, caspase-9) in myocardium [213216], and
miRNAs are considered rather targets for personifying inter-
vention and translational therapy, as well as prognosticators
than diagnostic biomarkers for adverse cardiac remodelling
and HF [217]. However, having signatures of miRNAs,
which correspond to adverse cardiac remodelling, HF, sud-
den death, and cardiac abnormalities with established poor
prognosis, such as concentric LV hypertrophy, brosis, and
inammation, it has not completely understood whether
the miRNA cardpersonally created for each patient will
have clinical signicance in the prediction of HF [33, 218].
9.2. Circulating Mononuclear and Endothelial Progenitor
Cells. Mononuclears (MPCs) and endothelial progenitor cells
(EPCs) are essential components of endogenous vascular
repair system that is activated as a result of several triggers,
such as ischemia/hypoxia, inammation, shear stress, throm-
bosis, inltration of lipids, direct injury of vasculature, and
endothelium [32].
It has been hypothesized that mobilization of
MPCs/EPCs and increase in growth and dierentiation into
mature cells in vasculature accompany acute events, includ-
ing AMI and acute HF, and are associated with vascular rep-
aration [219, 220]. However, previous acute CV events and
chronic metabolic and CV diseases, such as diabetes mellitus,
abdominal obesity, and hypertension, were reported to be
causes of an exhausting pool of circulating angiopoetic
MPCs/EPCs with immune phenotypes CD45+CD34+,
CD45+CD34+CD133+, and CD45+CD34+CD133+CD184+
[221]. Consequently, advanced HF and progression of
AMI-induced adverse cardiac remodelling were related to
impaired vascular repair, vascularization, and angiogenesis
due to a declined number of circulating precursors and
lowered their function and survival [222]. This phenome-
non is known as progenitor cell dysfunction and considered
a promising predictive biomarker for CV mortality and HF
progression and admission [223, 224], as well as in patient
population with AMI submitted to PCI [225]. Probably,
coronary circulating proangiogenic MPCs/EPCs collected
from coronary sinus in AMI patients with subsequent PCI
can become a powerful biomarker with increased accuracy
in the prediction of adverse cardiac remodelling. However,
the number and functionality of proangiogenic circulating
precursors appear as promising biomarkers for the
11Disease Markers
prediction of cardiac remodelling and HF development.
Large clinical trials are required to clearly understand the
role of new biomarkers in the diagnostic and predictive
strategies among AMI patients with PCI.
9.3. Future Perspectives. There are numerous biomarkers,
which were investigated as candidates for risk stratication
and prognosis in AMI patients with PCI, such as activated
and apoptotic endothelial cell-derived microvesicles,
bone-related proteins (osteopontin, osteoprotegerin, and
osteonectin), adipokines, gastrointestinal hormones, apelin,
cardiotrophin-1, defensin-1 and defensin-2, macrophage
inhibitory cytokine-1, circular RNAs, and gene card.
Although the data received appear to be promising, there
is no clear understanding whether diagnostic and predic-
tive abilities of these biomarkers will be better than the
conventional biomarkers of biomechanical stress, inam-
mation, brosis, and cardiac injury.
10. Conclusions
Circulating biomarkers are a promising tool to stratify AMI
patients undergoing PCI at high risk of poor cardiac recovery
and HF development. NPs are traditionally recommended as
diagnostic and predictive biomarkers for acute HF and
chronic HF regardless of LVEF, whereas sST2, galectin-3,
and cardiac troponins can be used optionally. Previous clin-
ical studies have yielded that multimarker models, which
were based on the combination of biomarkers of several
pathological axes involved in the nature evolution of adverse
cardiac remodelling (biomechanical myocardial stress, necro-
sis and injury of cardiac myocytes, and inammation), have
provided incremental prognostic information for prediction
of CV death or HF in AMI patients with subsequent PCI.
Future clinical trials with larger sample sizes are required to
elucidate the role of personifying biomarker-based strategy
for diagnostic, prediction, and treatment among patients sus-
pecting adverse cardiac remodelling and HF.
Abbreviations
ACE: Angiotensin-converting enzyme
ADAMTS: A Disintegrin and Metalloproteinase with
Thrombospondin motifs
ADM: adrenomedullin
AMI: Acute myocardial infarction
Apaf-1: Adaptor protein apoptotic protease activating
factor 1
ARBs: Angiotensin-II receptor antagonists
Bcl-2: B-cell lymphoma 2
BES: Biolimus eluting stent
BMS: Bare metal stent
CRP: C-reactive protein
CV: Cardiovascular
DAMPs: Damage-Associated Molecular Patterns
DAPT: Dual antiplatelet therapy
DNA: Deoxyribonucleic acid
ECM: Extracellular matrix
EF: Ejection fraction
GDF15: Growth/dierential factor-15
Inammation
(regulation of
activity of
antigen-
presenting
cells,
macrophages
and 1-cells)
miRNA
146a
miRNA-155
miRNA-125a
Hypertension,
diabetes mellitus,
LVH, insulin
resistance,
metabolic
syndrome,
hypothyroid
dysfunction
Microvascular endothelial
dysfunction
miRNA-138 (eNOs)
miRNA-126 (VCAM-1)
miRNA-155 (antigen-
presenting cells)
miRNA-138 (eNOs)
miRNA-126 (VCAM-1)
miRNA-155 (cell-to-cell cooperation)
miRNA-26a (insulin resistance)
Late adverse cardiac
remodelling
Cell remodelling
miRNA-18
miRNA-19
miRNA-21
miRNA-22
miRNA-133
miRNA-30
miRNA-29
miRNA-195
miRNA-199
miRNA-210
miRNA-223
Fibrosis, extracellular
matrix accumulation
and angiogenesis
(regulation of cell
viability and apoptosis
of cardiac myocytes,
broblasts, smooth
muscle cells, regulation
of synthesis of MMPs,
growth factors {TGF-
beta, VEGF})
Figure 4: The role of miRNAs in the pathogenesis of late adverse cardiac remodelling in AMI. VEGF: vascular endothelial growth factor;
TGF: transforming growth factor; NO: nitric oxide; eNOs: endothelial NO synthase; MMP: matrix metalloproteinase; VCAM: vascular
adhesive molecule.
12 Disease Markers
H
2
O
2
: Hydrogen peroxide
HF: Heart failure
HFmrEF: Heart failure with midrange ejection fraction
HFpEF: Heart failure with preserved ejection fraction
HFrEF: Heart failure with reduced ejection fraction
HIF: Hypoxia-induce factor
HMGB1: High-mobility group 1B protein
HO-1: Haem oxygenase-1
HSP: Heat shock proteins
IC: Intracoronary
IL: Interleukin
IVUS-VH: Intravascular ultrasound virtual-histology
LncRNA: Long noncoding RNA
LV: Left ventricle
MAPK: Mitogen-activated protein kinase
MCRA: Mineralocorticoid receptor antagonists
miRNA: MicroRNA
MMP: Matrix metalloproteinase
NO: Nitric oxide
NPs: Natriuretic peptides
OCT: Optical coherence tomography
PCI: Percutaneous coronary intervention
RAAS: Renin-angiotensin-aldosterone system
RNA: Ribonucleic acid
ROS: Reactive oxygen species
SOD: Superoxide dismutase
sST2: Soluble suppression of tumorigenicity-2
STEMI: ST-segment elevation myocardial infarction
TIMI score: Thrombolysis in Myocardial Infarction score
TGF: Transforming growth factor
TLR: Toll-like receptor
TNF: Tumor necrosis factor
TSP: Thrombospondin
VEGF: Vascular endothelial growth factor.
Data Availability
This is a narrative review, so dataset was not created.
Conflicts of Interest
The authors declare that there is no conict of interest
regarding the publication of this paper.
References
[1] H. E. Carter, D. Schoeld, and R. Shrestha, Productivity
costs of cardiovascular disease mortality across disease types
and socioeconomic groups,Open Heart, vol. 6, no. 1, article
e000939, 2019.
[2] J. Yap, S. Y. Chia, F. Y. Lim et al., The Singapore Heart Fail-
ure Risk Score: prediction of survival in Southeast Asian
patients,Annals of the Academy of Medicine of Singapore,
vol. 48, no. 3, pp. 8694, 2019.
[3] X. Chen, G. Savarese, U. Dahlström, L. H. Lund, and M. Fu,
Age-dependent dierences in clinical phenotype and prog-
nosis in heart failure with mid-range ejection compared with
heart failure with reduced or preserved ejection fraction,
Clinical Research in Cardiology, vol. 108, no. 12, pp. 1394
1405, 2019, [Epub ahead of print].
[4] P. M. Seferović, M. Polovina, J. Bauersachs et al., Heart fail-
ure in cardiomyopathies: a position paper from the Heart
Failure Association of the European Society of Cardiology,
European Journal of Heart Failure, vol. 21, no. 5, pp. 553
576, 2019.
[5] K. Dharmarajan and M. W. Rich, Epidemiology, pathophys-
iology, and prognosis of heart failure in older adults,Heart
Failure Clinics, vol. 13, no. 3, pp. 417426, 2017.
[6] A. Slee, M. Saad, and S. Saksena, Heart failure progression
and mortality in atrial brillation patients with preserved or
reduced left ventricular ejection fraction,Journal of Inter-
ventional Cardiac Electrophysiology, vol. 55, no. 3, pp. 325
331, 2019, [Epub ahead of print].
[7] Y. Sato, A. Yoshihisa, M. Oikawa et al., Prognostic Impact of
Worsening Renal Function in Hospitalized Heart Failure
Patients With Preserved Ejection Fraction: A Report From
the JASPER Registry,Journal of Cardiac Failure, vol. 25,
no. 8, pp. 631642, 2019, [Epub ahead of print].
[8] M. Nakamura and J. Sadoshima, Cardiomyopathy in obe-
sity, insulin resistance and diabetes,The Journal of Physiol-
ogy, 2019, [Epub ahead of print].
[9] L. Shen, P. S. Jhund, K. F. Docherty et al., Prior pacemaker
implantation and clinical outcomes in patients with heart
failure and preserved ejection fraction,JACC: Heart Failure,
vol. 7, no. 5, pp. 418427, 2019, [Epub ahead of print].
[10] H. Seligman, M. J. Shun-Shin, A. Vasireddy et al., Fractional
ow reserve derived from microcatheters versus standard
pressure wires: a stenosis-level meta-analysis,Open Heart,
vol. 6, no. 1, article e000971, 2019.
[11] K. V. Patel, R. Mauricio, J. L. Grodin et al., Identifying a low-
ow phenotype in heart failure with preserved ejection frac-
tion: a secondary analysis of the RELAX trial,ESC Heart
Failure, vol. 6, no. 4, pp. 613620, 2019.
[12] G. V. Halade, V. Kain, B. Tourki, and J. K. Jadapalli, Lipox-
ygenase drives lipidomic and metabolic reprogramming in
ischemic heart failure,Metabolism, vol. 96, pp. 2232,
2019, [Epub ahead of print].
[13] J. P. Tsai, K. T. Sung, C. H. Su et al., Diagnostic accuracy of
left atrial remodelling and natriuretic peptide levels for pre-
clinical heart failure,ESC Heart Failure, vol. 6, no. 4,
pp. 723732, 2019, [Epub ahead of print].
[14] A. Burlacu, P. Simion, I. Nistor, A. Covic, and G. Tinica,
Novel percutaneous interventional therapies in heart failure
with preserved ejection fraction: an integrative review,Heart
Failure Reviews, vol. 24, no. 5, pp. 793803, 2019, [Epub
ahead of print].
[15] H. Zeng and J. X. Chen, Microvascular rarefaction and heart
failure with preserved ejection fraction,Frontiers in Cardio-
vascular Medicine, vol. 6, 2019.
[16] R. Scarsini, G. L. de Maria, A. Borlotti et al., Incremental value
of coronary microcirculation resistive reserve ratio in predict-
ing the extent of myocardial infarction in patients with STEMI.
Insights from the Oxford Acute Myocardial Infarction
(OxAMI) study,Cardiovascular Revascularization Medicine,
vol. 20, no. 12, pp. 11481155, 2019, [Epub ahead of print].
[17] B. Ky, B. French, A. May Khan et al., Ventricular-Arterial
Coupling, Remodeling, and Prognosis in Chronic Heart Fail-
ure,Journal of the American College of Cardiology, vol. 62,
no. 13, pp. 11651172, 2013.
[18] B. He, L. Gai, J. Gai et al., Correlation between major adverse
cardiac events and coronary plaque characteristics,
13Disease Markers
Experimental & Clinical Cardiology, vol. 18, no. 2, pp. e71
e76, 2013.
[19] A. S. Bhatt, A. P. Ambrosy, and E. J. Velazquez, Adverse
remodeling and reverse remodeling after myocardial infarc-
tion,Current Cardiology Reports, vol. 19, no. 8, 2017.
[20] A. S. Blom, J. J. Pilla, R. C. Gorman III et al., Infarct size
reduction and attenuation of global left ventricular remodel-
ing with the CorCap cardiac support device following acute
myocardial infarction in sheep,Heart Failure Reviews,
vol. 10, no. 2, pp. 125139, 2005.
[21] A. W. Schoenenberger, P. Jamshidi, R. Kobza et al., Progres-
sion of coronary artery disease during long-term follow-up of
the Swiss Interventional Study on Silent Ischemia Type II
(SWISSI II),Clinical Cardiology, vol. 33, no. 5, pp. 289
295, 2010.
[22] P. Erne, A. W. Schoenenberger, D. Burckhardt et al., Eects
of percutaneous coronary interventions in silent ischemia
after myocardial infarction: the SWISSI II randomized con-
trolled trial,JAMA, vol. 297, no. 18, pp. 19851991, 2007.
[23] G. A. Sgueglia, F. D'Errico, G. Giorè et al., Angiographic
and clinical performance of polymer-free biolimus-eluting
stent in patients with ST-segment elevation acute myocardial
infarction in a metropolitan public hospital: the BESAMI
MUCHO study,Catheterization and Cardiovascular Inter-
ventions, vol. 91, no. 5, pp. 851858, 2018.
[24] C. R. Jones, S. J. Khandhar, M. Ramratnam et al., Identica-
tion of intrastent pathology associated with late stent throm-
bosis using optical coherence tomography,Journal of
Interventional Cardiology, vol. 28, no. 5, pp. 439448, 2015.
[25] A. Al Mamary, G. Dariol, and M. Napodano, Late stent frac-
ture A potential role of left ventricular dilatation,Journal
of the Saudi Heart Association, vol. 26, no. 3, pp. 162165,
2014.
[26] S. V. Rao, U. Zeymer, P. S. Douglas et al., A randomized,
double-blind, placebo-controlled trial to evaluate the safety
and eectiveness of intracoronary application of a novel
bioabsorbable cardiac matrix for the prevention of ventricu-
lar remodeling after large ST- segment elevation myocardial
infarction: Rationale and design of the PRESERVATION I
trial,American Heart Journal, vol. 170, no. 5, pp. 929937,
2015, Epub 2015 Aug 24.
[27] S. V. Rao, U. Zeymer, P. S. Douglas et al., Bioabsorbable
intracoronary matrix for prevention of ventricular remodel-
ing after myocardial infarction,Journal of the American Col-
lege of Cardiology, vol. 68, no. 7, pp. 715723, 2016.
[28] T. Mizoguchi, T. Sawada, T. Shinke et al., Detailed compar-
ison of intra-stent conditions 12 months after implantation of
everolimus-eluting stents in patients with ST-segment eleva-
tion myocardial infarction or stable angina pectoris,Interna-
tional Journal of Cardiology, vol. 171, no. 2, pp. 224230,
2014, Epub 2013 Dec 21.
[29] S. H. Rezkalla, R. V. Stankowski, J. Hanna, and R. A. Kloner,
Management of no-reow phenomenon in the catheteriza-
tion laboratory,JACC: Cardiovascular Interventions,
vol. 10, no. 3, pp. 215223, 2017.
[30] J. Barrabes, Comments on the 2015 ESC Guidelines for the
Management of Acute Coronary Syndromes in Patients Pre-
senting Without Persistent ST-segment Elevation,Revista
Española de Cardiología (English Edition), vol. 68, no. 12,
pp. 10611067, 2015.
[31] T. K. Steigen, C. E. Buller, G. B. John Mancini et al., Myocar-
dial perfusion grade after late infarct artery recanalization is
associated with global and regional left ventricular function
at one year: analysis from the Total Occlusion Study of Can-
ada-2,Circulation: Cardiovascular Interventions, vol. 3,
no. 6, pp. 549555, 2010.
[32] A. E. Berezin, Endogenous vascular repair system in car-
diovascular disease: the role of endothelial progenitor
cells,Australasian Medical Journal, vol. 12, no. 2,
pp. 4248, 2019.
[33] A. Berezin, Epigenetics in heart failure phenotypes,BBA
Clinical, vol. 6, pp. 3137, 2016.
[34] A. Celik, N. Kalay, H. Korkmaz et al., Short-term left ven-
tricular remodeling after revascularization in subacute total
and subtotal occlusion with the infarct-related left anterior
descending artery,Cardiology Research, vol. 2, no. 5,
pp. 229235, 2011.
[35] M. E. Pfusterer, P. Buser, S. Osswald, P. Weiss, J. Bremerich,
and F. Burkart, Time dependence of left ventricular recovery
after delayed recanalization of an occluded infarct-related
coronary artery: ndings of a pilot study,Journal of the
American College of Cardiology, vol. 32, no. 1, pp. 97102,
1998.
[36] N. Reifart, Challenges in complicated coronary chronic total
occlusion recanalisation,Interventional Cardiology Review,
vol. 8, no. 2, pp. 107111, 2013.
[37] N. G. Frangogiannis, Pathophysiology of myocardial infarc-
tion,Comprehensive Physiology, vol. 20, pp. 18411875,
2015.
[38] M. Neri, I. Riezzo, N. Pascale, C. Pomara, and E. Turillazzi,
Ischemia/reperfusion injury following acute myocardial
infarction: a critical issue for clinicians and forensic patholo-
gists,Mediators Inamm, vol. 2017, article 7018393, 14
pages, 2017, Epub 2017 Feb 13.
[39] D. Y. Fuhrman and J. A. Kellum, Remote ischemic precon-
ditioning in the PICU: a simple concept with a complex past,
Pediatric Critical Care Medicine, vol. 17, no. 8, pp. e371e379,
2016.
[40] S. Hernandez-Resendiz, K. Chinda, S. B. Ong, H. Cabrera-
Fuentes, C. Zazueta, and D. Hausenloy, The role of redox
dysregulation in the inammatory response to acute myocar-
dial ischaemia-reperfusion injury - adding fuel to the re,
Current Medicinal Chemistry, vol. 25, no. 11, pp. 1275
1293, 2018.
[41] K. Przyklenk and P. Whittaker, Remote ischemic precondi-
tioning: current knowledge, unresolved questions, and future
priorities,Journal of Cardiovascular Pharmacology and
Therapeutics, vol. 16, no. 3-4, pp. 255259, 2016.
[42] D. M. Yellon and D. J. Hausenloy, Myocardial reperfusion
injury,The New England Journal of Medicine, vol. 357,
no. 11, pp. 11211135, 2007.
[43] P. Pagliaro, F. Moro, F. Tullio, M.-G. Perrelli, and C. Penna,
Cardioprotective pathways during reperfusion: focus on
redox signaling and other modalities of cell Signaling,Anti-
oxidants & Redox Signaling, vol. 14, no. 5, pp. 833850, 2011.
[44] B. Long, N. Li, X. X. Xu et al., Long noncoding RNA FTX
regulates cardiomyocyte apoptosis by targeting miR-29b-1-
5p and Bcl2l2,Biochemical and Biophysical Research Com-
munications, vol. 495, no. 1, pp. 312318, 2018.
[45] T. Kalogeris, Y. Bao, and R. J. Korthuis, Mitochondrial reac-
tive oxygen species: a double edged sword in ischemia/reper-
fusion vs preconditioning,Redox Biology, vol. 2, no. 1,
pp. 702714, 2014.
14 Disease Markers
[46] N. Zhang, X. Meng, L. Mei, J. Hu, C. Zhao, and W. Chen,
The long non-coding RNA SNHG1 attenuates cell apoptosis
by regulating miR-195 and BCL2-like protein 2 in human
cardiomyocytes,Cellular Physiology and Biochemistry,
vol. 50, no. 3, pp. 10291040, 2018.
[47] F. He, H. Liu, J. Guo et al., Inhibition of microRNA-124
reduces cardiomyocyte apoptosis following myocardial
infarction via targeting STAT3,Cellular Physiology and Bio-
chemistry, vol. 51, no. 1, pp. 186200, 2018.
[48] C. Sun, H. Liu, J. Guo et al., MicroRNA-98 negatively regu-
lates myocardial infarction-induced apoptosis by down-
regulating Fas and caspase-3,Scientic Reports, vol. 7,
no. 1, p. 7460, 2017.
[49] T. Okamoto, T. Akaike, T. Sawa, Y. Miyamoto, A. Van der
Vliet, and H. Maeda, Activation of matrix metalloprotein-
ases by peroxynitrite-induced protein S-glutathiolation via
disulde S-oxide formation,The Journal of Biological Chem-
istry, vol. 276, no. 31, pp. 2959629602, 2001.
[50] P.-Y. Cheung, G. Sawicki, M. Wozniak, W. Wang, M. W.
Radomski, and R. Schulz, Matrix metalloproteinase-2 con-
tributes to ischemia-reperfusion injury in the heart,Circula-
tion, vol. 101, no. 15, pp. 18331839, 2000.
[51] A. E. Berezin and T. A. Samura, Prognostic value of bio-
logical markers in myocardial infarction patients,Asian
Cardiovascular and Thoracic Annals, vol. 21, no. 2,
pp. 142150, 2013.
[52] C. Fernandez-Patron, M. W. Radomski, and S. T. Davidge,
Vascular matrix metalloproteinase-2 cleaves big
endothelin-1 yielding a novel vasoconstrictor,Circulation
Research, vol. 85, no. 10, pp. 906911, 1999.
[53] D. Edwards, M. Handsley, and C. Pennington, The ADAM
metalloproteinases,Molecular Aspects of Medicine, vol. 29,
no. 5, pp. 258289, 2008.
[54] O. V. Petyunina, M. P. Kopytsya, and A. E. Berezin, Bio-
marker-based prognostication of adverse cardiac remodeling
after STEMI: the role of single nucleotide polymorphism
T786C in endothelial NO-synthase gene,Journal of Cardiol-
ogy and Therapy, vol. 6, no. 1, pp. 768774, 2019.
[55] S. Cardin, M. P. Scott-Boyer, S. Praktiknjo et al., Dierences
in cell-type-specic responses to angiotensin II explain car-
diac remodeling dierences in C57BL/6 mouse substrains,
Hypertension, vol. 64, no. 5, pp. 10401046, 2014.
[56] Y. Li, X._. He, C. Li, L. Gong, and M. Liu, Identication of
candidate genes and microRNAs for acute myocardial infarc-
tion by weighted gene coexpression network analysis,
BioMed Research International, vol. 2019, Article ID
5742608, 11 pages, 2019.
[57] L. M. Buja, Myocardial ischemia and reperfusion injury,
Cardiovascular Pathology, vol. 14, no. 4, pp. 170175,
2005.
[58] C. P. Baines, How and when do myocytes die during ische-
mia and reperfusion: the late phase,Journal of Cardiovascu-
lar Pharmacology and Therapeutics, vol. 16, no. 3-4, pp. 239
243, 2016.
[59] Z. Zhao, Oxidative stress-elicited myocardial apoptosis dur-
ing reperfusion,Current Opinion in Pharmacology, vol. 4,
no. 2, pp. 159165, 2004.
[60] A. Prasad, G. W. Stone, D. R. Holmes, and B. Gersh, Reper-
fusion injury, microvascular dysfunction, and cardioprotec-
tion: the 'dark side' of reperfusion,Circulation, vol. 120,
no. 21, pp. 21052112, 2009.
[61] P. Ferdinandy, D. J. Hausenloy, G. Heusch, G. F. Baxter, and
R. Schulz, Interaction of risk factors, comorbidities, and
comedications with ischemia/reperfusion injury and cardio-
protection by preconditioning, postconditioning, and remote
conditioning,Pharmacological Reviews, vol. 66, no. 4,
pp. 11421174, 2014.
[62] R. S. Vander Heide and C. Steenbergen, Cardioprotection
and Myocardial Reperfusion,Circulation Research,
vol. 113, no. 4, pp. 464477, 2013.
[63] K. McCaerty, S. Forbes, C. Thiemermann, and M. M.
Yaqoob, The challenge of translating ischemic conditioning
from animal models to humans: the role of comorbidities,
Disease Models & Mechanisms, vol. 7, no. 12, pp. 1321
1333, 2014.
[64] A. Bonaventura, F. Montecucco, F. Dallegri et al., Novel
ndings in neutrophil biology and their impact on cardiovas-
cular disease,Cardiovascular Research, vol. 115, no. 8,
pp. 12661285, 2019, [Epub ahead of print].
[65] M. A. Razzaque, X. Xu, M. Han, A. Badami, and S. A. Akhter,
Inhibition of postinfarction ventricular remodeling by high
molecular weight polyethylene glycol,Journal of Surgical
Research, vol. 232, pp. 171178, 2018.
[66] W. An, Y. Yu, Y. Zhang, Z. Zhang, Y. Yu, and X. Zhao, Exog-
enous IL-19 attenuates acute ischaemic injury and improves
survival in male mice with myocardial infarction,British
Journal of Pharmacology, vol. 176, no. 5, pp. 699710, 2019.
[67] M. Arita, T. Ohira, Y.-P. Sun, S. Elangovan, N. Chiang, and
C. N. Serhan, Resolvin E1 selectively interacts with leukotri-
ene B4Receptor BLT1 and ChemR23 to regulate inamma-
tion,The Journal of Immunology, vol. 178, no. 6, pp. 3912
3917, 2007.
[68] V. Kain, K. A. Ingle, R. A. Colas et al., Resolvin D1 activates
the inammation resolving response at splenic and ventricu-
lar site following myocardial infarction leading to improved
ventricular function,Journal of Molecular and Cellular Car-
diology, vol. 84, pp. 2435, 2015.
[69] L. A. Grisanti, T. P. Thomas, R. L. Carter et al., Pepducin-
mediated cardioprotection via β-arrestin-biased β2-adrener-
gic receptor-specic signaling,Theranostics, vol. 8, no. 17,
pp. 46644678, 2018.
[70] Z. Guo, N. Liu, L. Chen, X. Zhao, and M. R. Li, Independent
roles of CGRP in cardioprotection and hemodynamic regula-
tion in ischemic postconditioning,European Journal of
Pharmacology, vol. 828, pp. 1825, 2018.
[71] Q. Huang, Z. Yang, J. P. Zhou, and Y. Luo, HMGB1 induces
endothelial progenitor cells apoptosis via RAGE-dependent
PERK/eIF2αpathway,Molecular and Cellular Biochemistry,
vol. 431, no. 1-2, pp. 6774, 2017, Epub 2017 Mar 1.
[72] S. Frankenreiter, P. Bednarczyk, A. Kniess et al., cGMP-ele-
vating compounds and ischemic conditioning provide cardi-
oprotection against ischemia and reperfusion injury via
cardiomyocyte-specic BK channels,Circulation,vol. 136,
no. 24, pp. 23372355, 2017.
[73] S. Raa, X. L. D. Chin, R. Stanzione et al., The reduction of
NDUFC2 expression is associated with mitochondrial
impairment in circulating mononuclear cells of patients with
acute coronary syndrome,International Journal of Cardiol-
ogy, vol. 286, pp. 127133, 2019.
[74] J. Wang, M. Liu, Q. Wu et al., Human embryonic stem cell-
derived cardiovascular progenitors repair infarcted hearts
through modulation of MacrophagesviaActivation of signal
transducer and activator of transcription 6,Antioxidants &
15Disease Markers
Redox Signaling, vol. 31, no. 5, pp. 369386, 2019, [Epub
ahead of print].
[75] Z. Zhong, J. Hou, Q. Zhang et al., Dierential expression of
circulating long non-coding RNAs in patients with acute
myocardial infarction,Medicine, vol. 97, no. 51, article
e13066, 2018.
[76] Z. A. Ibrahim, C. L. Armour, S. Phipps, and M. B. Sukkar,
RAGE and TLRs: relatives, friends or neighbours?,Molecu-
lar Immunology, vol. 56, no. 4, pp. 739744, 2013.
[77] J. J. De Haan, M. B. Smeets, G. Pasterkamp, and F. Arslan,
Danger signals in the initiation of the inammatory
response after myocardial infarction,Mediators of Inam-
mation, vol. 2013, Article ID 206039, 13 pages, 2013.
[78] A. Micera, B. O. Balzamino, A. D. Zazzo, F. Biamonte,
G. Sica, and S. Bonini, Toll-like receptors and tissue remod-
eling: the pro/cons recent ndings,Journal of Cellular Phys-
iology, vol. 231, no. 3, pp. 531544, 2016.
[79] K. Fujiu, J. Wang, and R. Nagai, Cardioprotective function
of cardiac macrophages,Cardiovascular Research, vol. 102,
no. 2, pp. 232239, 2014.
[80] B. Tourki and G. Halade, Leukocyte diversity in resolving
and nonresolving mechanisms of cardiac remodeling,The
FASEB Journal, vol. 31, no. 10, pp. 42264239, 2017.
[81] O. Dewald, P. Zymek, K. Winkelmann et al., CCL2/mono-
cyte chemoattractant protein-1 regulates inammatory
responses critical to healing myocardial infarcts,Circ. Res.,
vol. 96, no. 8, pp. 881889, 2005.
[82] P. M. Henson, D. L. Bratton, and V. A. Fadok, Apoptotic cell
removal,Current Biology, vol. 11, no. 19, pp. R795R805, 2001.
[83] S. Iravanian and S. C. Dudley Jr., The renin-angiotensin-
aldosterone system (RAAS) and cardiac arrhythmias,Heart
Rhythm, vol. 5, no. 6, pp. S12S17, 2008.
[84] M. Massa, V. Rosti, M. Ferrario et al., Increased circulating
hematopoietic and endothelial progenitor cells in the early
phase of acute myocardial infarction,Blood, vol. 105, no. 1,
pp. 199206, 2005.
[85] C. Gao, K. Howard-Quijano, C. Rau et al., Inammatory and
apoptotic remodeling in autonomic nervous system following
myocardial infarction,PLoS One, vol. 12, no. 5, article
e0177750, 2017.
[86] P. Ponikowski, A. A. Voors, S. D. Anker et al., 2016 ESC
guidelines for the diagnosis and treatment of acute and
chronic heart failure: the task force for the diagnosis and
treatment of acute and chronic heart failure of the European
Society of Cardiology (ESC). Developed with the special con-
tribution of the Heart Failure Association (HFA) of the ESC,
European Journal of Heart Failure, vol. 18, no. 8, pp. 891975,
2016.
[87] C. W. Yancy, M. Jessup, B. Bozkurt et al., 2017 ACC/A-
HA/HFSA focused update of the 2013 ACCF/AHA Guideline
for the management of heart failure: a report of the American
College of Cardiology/American Heart Association Task
Force on Clinical Practice Guidelines and the Heart Failure
Society of America,Circulation, vol. 136, no. 6, pp. e137
e161, 2017.
[88] B. Bozkurt, What is new in heart failure management in
2017? Update on ACC/AHA Heart Failure Guidelines,Cur-
rent Cardiology Reports, vol. 20, no. 6, 2018.
[89] A. E. Berezin, Circulating biomarkers in heart failure,
Advances in Experimental Medicine and Biology, vol. 1067,
pp. 89108, 2018.
[90] A. E. Berezin, Prognostication in dierent heart failure phe-
notypes: the role of circulating biomarkers,Journal of Circu-
lating Biomarkers, vol. 5, no. 6, p. 6, 2016.
[91] M. Cheng, S. An, and J. Li, Identifying key genes associated
with acute myocardial infarction,Medicine, vol. 96, no. 42,
article e7741, 2017.
[92] Y. Cao, R. Li, Y. Li et al., Identication of transcription
factor-gene regulatory network in acute myocardial infarc-
tion,Heart, Lung and Circulation, vol. 26, no. 4, pp. 343
353, 2017, Epub 2016 Jul 26.
[93] B. Ibanez, S. James, S. Agewall et al., 2017 ESC Guidelines for
the management of acute myocardial infarction in patients
presenting with ST-segment elevation: the Task Force for
the management of acute myocardial infarction in patients
presenting with ST-segment elevation of the European Soci-
ety of Cardiology (ESC),European Heart Journal, vol. 39,
no. 2, pp. 119177, 2018.
[94] M. N. Lyngbakken, H. Røsjø, O. L. Holmen, H. Dalen,
K. Hveem, and T. Omland, Temporal changes in cardiac
troponin I are associated with risk of cardiovascular events
in the general population: the Nord-Trøndelag Health
Study,Clinical Chemistry, vol. 65, no. 7, pp. 871881, 2019,
[Epub ahead of print].
[95] D. Soetkamp, K. Raedschelders, M. Mastali, K. Sobhani, C. N.
Bairey Merz, and J. Van Eyk, The continuing evolution of
cardiac troponin I biomarker analysis: from protein to pro-
teoform,Expert Review of Proteomics, vol. 14, no. 11,
pp. 973986, 2017.
[96] A. S. Shah, A. Anand, Y. Sandoval et al., High-sensitivity car-
diac troponin I at presentation in patients with suspected
acute coronary syndrome: a cohort study,Lancet, vol. 386,
no. 10012, pp. 24812488, 2015.
[97] D. Stelzle, A. S. V. Shah, A. Anand et al., High-sensitivity
cardiac troponin I and risk of heart failure in patients with
suspected acute coronary syndrome: a cohort study,Euro-
pean Heart Journal - Quality of Care and Clinical Outcomes,
vol. 4, no. 1, pp. 3642, 2018.
[98] A. E. Berezin, Circulating biomarkers in heart failure: diag-
nostic and prognostic importance,Journal of Laboratory
and Precision Medicine, vol. 3, article 36, 2018.
[99] X. Jia, W. Sun, R. C. Hoogeveen et al., High-Sensitivity Tro-
ponin I and Incident Coronary Events, Stroke, Heart Failure
Hospitalization, and Mortality in the ARIC Study,Circula-
tion, vol. 139, no. 23, pp. 26422653, 2019.
[100] P. Welsh, D. Preiss, C. Hayward et al., Cardiac troponin
T and troponin I in the general Population,Circulation,
vol. 139, no. 24, pp. 27542764, 2019, [Epub ahead of
print].
[101] T. Raskovalova, R. Twerenbold, P. O. Collinson et al.,
Diagnostic accuracy of combined cardiac troponin and
copeptin assessment for early rule-out of myocardial
infarction: a systematic review and meta-analysis,Euro-
pean Heart Journal: Acute Cardiovascular Care, vol. 3,
no. 1, pp. 1827, 2014.
[102] J. Searle, O. Danne, C. Müller, and M. Mockel, Biomarkers
in acute coronary syndrome and percutaneous coronary
intervention,Minerva Cardioangiologica, vol. 59, no. 3,
pp. 203223, 2011.
[103] C. Mueller, Biomarkers and acute coronary syndromes: an
update,European Heart Journal, vol. 35, no. 9, pp. 552
556, 2014, Epub 2013 Dec 18.
16 Disease Markers
[104] M. Möckel and J. Searle, Copeptin-marker of acute myocar-
dial infarction,Current Atherosclerosis Reports, vol. 16, no. 7,
p. 421, 2014.
[105] M. Karakas, J. L. Januzzi Jr., J. Meyer et al., Copeptin does
not add diagnostic information to high-sensitivity troponin
T in low- to intermediate-risk patients with acute chest pain:
results from the rule out myocardial infarction by computed
tomography (ROMICAT) study,Clinical Chemistry, vol. 57,
no. 8, pp. 11371145, 2011.
[106] S. Suzuki, H. Motoki, M. Minamisawa et al., Prognostic sig-
nicance of high-sensitivity cardiac troponin in patients with
heart failure with preserved ejection fraction,Heart Vessels,
vol. 34, no. 10, pp. 16501656, 2019, [Epub ahead of print].
[107] R. A. de Boer, M. Nayor, C. R. deFilippi et al., Association of
cardiovascular biomarkers with incident heart failure with
preserved and reduced ejection fraction,JAMA Cardiology,
vol. 3, no. 3, pp. 215224, 2018.
[108] A. Gohar, J. P. C. Chong, O. W. Liew et al., The prognostic
value of highly sensitive cardiac troponin assays for adverse
events in men and women with stable heart failure and a pre-
served vs. reduced ejection fraction,European Journal of
Heart Failure, vol. 19, no. 12, pp. 16381647, 2017, Epub
2017 Aug 28.
[109] R. Santhanakrishnan, J. P. C. Chong, T. P. Ng et al., Growth
dierentiation factor 15, ST2, high-sensitivity troponin T,
and N-terminal pro brain natriuretic peptide in heart failure
with preserved vs. reduced ejection fraction,European Jour-
nal of Heart Failure, vol. 14, no. 12, pp. 13381347, 2012.
[110] P. Moliner, J. Lupón, J. Barallat et al., Bio-proling and bio-
prognostication of chronic heart failure with mid-range ejec-
tion fraction,International Journal of Cardiology, vol. 257,
pp. 188192, 2018.
[111] P. Welsh, L. Kou, C. Yu et al., Prognostic importance of
emerging cardiac, inammatory, and renal biomarkers in
chronic heart failure patients with reduced ejection fraction
and anaemia: RED-HF study,European Journal of Heart
Failure, vol. 20, no. 2, pp. 268277, 2018.
[112] S. Sanders-van Wijk, V. van Empel, N. Davarzani et al., Cir-
culating biomarkers of distinct pathophysiological pathways
in heart failure with preserved vs. reduced left ventricular
ejection fraction,European Journal of Heart Failure,
vol. 17, no. 10, pp. 10061014, 2015.
[113] S. L. Seliger, J. de Lemos, I. J. Neeland et al., Older Adults,
MalignantLeft Ventricular Hypertrophy, and Associated
Cardiac-Specic Biomarker Phenotypes to Identify the Dif-
ferential Risk of New-Onset Reduced Versus Preserved Ejec-
tion Fraction Heart Failure: CHS (Cardiovascular Health
Study),JACC: Heart Failure, vol. 3, no. 6, pp. 445455,
2015, Epub 2015 May 14.
[114] C. Sinning, T. Kempf, M. Schwarzl et al., Biomarkers for
characterization of heart failure Distinction of heart failure
with preserved and reduced ejection fraction,International
Journal of Cardiology, vol. 227, pp. 272277, 2017.
[115] D. A. Pascual-Figal, S. Manzano-Fernández, M. Boronat
et al., Soluble ST2, high-sensitivity troponin T- and N-
terminal pro-B-type natriuretic peptide: complementary role
for risk stratication in acutely decompensated heart failure,
European Journal of Heart Failure, vol. 13, no. 7, pp. 718725,
2011.
[116] D. A. Pascual-Figal, A. Bayes-Genis, M. C. Asensio-Lopez
et al., The interleukin-1 axis and risk of death in patients
with acutely decompensated heart failure,Journal of the
American College of Cardiology, vol. 73, no. 9, pp. 1016
1025, 2019.
[117] Y. Cao, R. Li, F. Zhang, Z. Guo, S. Tuo, and Y. Li, Correlation
between angiopoietin-like proteins in inammatory media-
tors in peripheral blood and severity of coronary arterial
lesion in patients with acute myocardial infarction,Experi-
mental and Therapeutic Medicine, vol. 17, no. 5, pp. 3495
3500, 2019.
[118] E. O. Weinberg, M. Shimpo, S. Hurwitz, S. Tominaga, J. L.
Rouleau, and R. T. Lee, Identication of serum soluble ST2
receptor as a novel heart failure biomarker,Circulation,
vol. 107, no. 5, pp. 721726, 2003.
[119] J. L. Januzzi, W. F. Peacock, A. S. Maisel et al., Measurement
of the Interleukin Family Member ST2 in Patients With
Acute Dyspnea: Results From the PRIDE (Pro-Brain Natri-
uretic Peptide Investigation of Dyspnea in the Emergency
Department) Study,Journal of the American College of Car-
diology, vol. 50, no. 7, pp. 607613, 2007.
[120] K. Seki, S. Sanada, A. Y. Kudinova et al., Interleukin-33 pre-
vents apoptosis and improves survival after experimental
myocardial infarction through ST2 signaling,Circulation:
Heart Failure,vol. 2, no. 6, pp. 684691, 2009.
[121] W. S. Jenkins, V. L. Roger, A. S. Jae et al., Prognostic value
of soluble ST2 after myocardial infarction: a community per-
spective,The American Journal of Medicine, vol. 130, no. 9,
pp. 1112.e91112.e15, 2017, Epub 2017 Mar 23.
[122] R. A. P. Weir, A. M. Miller, G. E. J. Murphy et al., Serum sol-
uble ST2: a potential novel mediator in left ventricular and
infarct remodeling after acute myocardial infarction,Journal
of the American College of Cardiology, vol. 55, no. 3, pp. 243
250, 2010.
[123] W. H. W. Tang, Y. Wu, J. L. Grodin et al., Prognostic value
of baseline and changes in circulating soluble ST2 levels and
the eects of nesiritide in acute decompensated heart failure,
JACC: Heart Failure, vol. 4, no. 1, pp. 6877, 2016.
[124] S. Manzano-Fernández, J. L. Januzzi, F. J. Pastor-Pérez et al.,
Serial monitoring of soluble interleukin family member ST2
in patients with acutely decompensated heart failure,Cardi-
ology, vol. 122, no. 3, pp. 158166, 2012.
[125] T. Mueller, A. Gegenhuber, W. Poelz, and M. Haltmayer,
Diagnostic accuracy of B type natriuretic peptide and amino
terminal proBNP in the emergency diagnosis of heart fail-
ure,Heart, vol. 91, no. 5, pp. 606612, 2005.
[126] A. E. Berezin, Biomarkers for cardiovascular risk in patients
with diabetes,Heart, vol. 102, no. 24, pp. 19391941, 2016.
[127] M. S. Kim, T. D. Jeong, S. B. Han, W. K. Min, and J. J. Kim,
Role of soluble ST2 as a prognostic marker in patients with
acute heart failure and renal insuciency,Journal of Korean
Medical Science, vol. 30, no. 5, pp. 569575, 2015.
[128] D. A. Pascual-Figal, J. Ordoñez-Llanos, P. L. Tornel et al.,
Soluble ST2 for predicting sudden cardiac death in patients
with chronic heart failure and left ventricular systolic dys-
function,Journal of the American College of Cardiology,
vol. 54, no. 23, pp. 21742179, 2009.
[129] M. Emdin, A. Aimo, G. Vergaro et al., sST2 Predicts Out-
come in Chronic Heart Failure Beyond NTproBNP and
High- Sensitivity Troponin T,Journal of the American Col-
lege of Cardiology, vol. 72, no. 19, pp. 23092320, 2018.
[130] E. OMeara, M. F. Prescott, B. Claggett et al., Independent
prognostic value of serum soluble ST2 measurements in
patients with heart failure and a reduced ejection fraction in
17Disease Markers
the PARADIGM-HF trial (Prospective Comparison of ARNI
With ACEI to Determine Impact on Global Mortality and
Morbidity in Heart Failure),Circulation: Heart Failure,
vol. 11, no. 5, article e004446, 2018.
[131] J. Lassus, E. Gayat, C. Mueller et al., Incremental value of
biomarkers to clinical variables for mortality prediction in
acutely decompensated heart failure: the Multinational
Observational Cohort on Acute Heart Failure (MOCA)
study,International Journal of Cardiology, vol. 168, no. 3,
pp. 21862194, 2013.
[132] W. P. Huang, X. Zheng, L. He, X. Su, C. W. Liu, and M. X.
Wu, Role of soluble ST2 levels and beta-blockers dosage
on cardiovascular events of patients with unselected ST-
segment elevation myocardial infarction,Chinese Medical
Journal, vol. 131, no. 11, pp. 12821288, 2018.
[133] Y. Michowitz, Y. Arbel, D. Wexler et al., Predictive value of
high sensitivity CRP in patients with diastolic heart failure,
International Journal of Cardiology, vol. 125, no. 3, pp. 347
351, 2008.
[134] J. P. Araújo, P. Lourenço, A. Azevedo et al., Prognostic value
of high-sensitivity C-reactive protein in heart failure: a sys-
tematic review,Journal of Cardiac Failure, vol. 15, no. 3,
pp. 256266, 2009.
[135] A. P. Kalogeropoulos, W. H. W. Tang, A. Hsu et al., High-
sensitivity C-reactive protein in acute heart failure: insights
from the ASCEND-HF trial,Journal of Cardiac Failure,
vol. 20, no. 5, pp. 319326, 2014.
[136] K. Huynh, B. Van Tassell, and S. L. Chow, Predicting
therapeutic response in patients with heart failure: the
story of C-reactive protein,Expert Review of Cardiovascu-
lar Therapy, vol. 13, no. 2, pp. 153161, 2015, Epub 2015
Jan 12.
[137] K. C. Wollert, T. Kempf, and L. Wallentin, Growth dieren-
tiation factor 15 as a biomarker in cardiovascular disease,
Clinical Chemistry, vol. 63, no. 1, pp. 140151, 2017.
[138] T. Kempf, M. Eden, J. Strelau et al., The transforming
growth factor-beta superfamily member growth-
dierentiation factor-15 protects the heart from ischemia/re-
perfusion injury,Circulation Research, vol. 98, no. 3,
pp. 351360, 2006.
[139] A. Rohatgi, P. Patel, S. R. Das et al., Association of growth
dierentiation factor-15 with coronary atherosclerosis and
mortality in a young, multiethnic population: observations
from the Dallas Heart Study,Clinical Chemistry, vol. 58,
no. 1, pp. 172182, 2012.
[140] P. Bettencourt, J. Ferreira-Coimbra, P. Rodrigues et al.,
Towards a multi-marker prognostic strategy in acute heart
failure: a role for GDF-15,ESC Heart Failure, vol. 5, no. 6,
pp. 10171022, 2018.
[141] T. Kempf, S. von Haehling, T. Peter et al., Prognostic utility
of growth dierentiation factor-15 in patients with chronic
heart failure,Journal of the American College of Cardiology,
vol. 50, no. 11, pp. 10541060, 2007.
[142] M. M. Y. Chan, R. Santhanakrishnan, J. P. C. Chong et al.,
Growth dierentiation factor 15 in heart failure with pre-
served vs. reduced ejection fraction,European Journal of
Heart Failure, vol. 18, no. 1, pp. 8188, 2016.
[143] I. S. Anand, T. Kempf, T. S. Rector et al., Serial measurement
of growth-dierentiation factor-15 in heart failure: relation to
disease severity and prognosis in the Valsartan Heart Failure
Trial,Circulation, vol. 122, no. 14, pp. 13871395, 2010.
[144] D. J. Lok, I. J. T. Klip, S. I. Lok et al., Incremental Prognostic
Power of Novel Biomarkers (Growth-Dierentiation Factor-
15, High-Sensitivity C-Reactive Protein, Galectin-3, and
High- Sensitivity Troponin-T) in Patients With Advanced
Chronic Heart Failure,The American Journal of Cardiology,
vol. 112, no. 6, pp. 831837, 2013.
[145] B. G. Demissei, G. Cotter, M. F. Prescott et al., A multimar-
ker multi-time point-based risk stratication strategy in acute
heart failure: results from the RELAX-AHF trial,European
Journal of Heart Failure, vol. 19, no. 8, pp. 10011010, 2017.
[146] J. Li, Y. Cui, A. Huang et al., Additional diagnostic value of
growth dierentiation factor-15 (GDF-15) to N-terminal B-
type natriuretic peptide (NT-proBNP) in patients with dier-
ent stages of heart failure,Medical Science Monitor, vol. 24,
pp. 49924999, 2018.
[147] L. B. Daniels and Q. M. Bui, Should a high Gal-3 have us
scared sti?,Journal of the American College of Cardiology,
vol. 73, no. 18, pp. 22962298, 2019.
[148] J. Dumic, S. Dabelic, and M. Flogel, Galectin-3: an open-
ended story,Biochim Biophys Acta, vol. 1760, no. 4,
pp. 616635, 2006.
[149] H. Sano, D. K. Hsu, J. R. Apgar et al., Critical role of galectin-
3 in phagocytosis by macrophages,Journal of Clinical Inves-
tigation, vol. 112, no. 3, pp. 389397, 2003.
[150] A. Krzeslak and A. Lipinska, Galectin-3 as a multifunctional
protein,Cellular & Molecular Biology Letters, vol. 9, no. 2,
pp. 305328, 2004.
[151] U. C. Sharma, S. Pokharel, T. J. van Brakel et al., Galectin-3
marks activated macrophages in failure-prone hypertrophied
hearts and contributes to cardiac dysfunction,Circulation,
vol. 110, no. 19, pp. 31213128, 2004.
[152] L. Calvier, M. Miana, P. Reboul et al., Galectin-3 mediates
aldosterone-induced vascular brosis,Arteriosclerosis,
Thrombosis, and Vascular Biology, vol. 33, no. 1, pp. 6775,
2013.
[153] L. Yu, W. P. T. Ruifrok, M. Meissner et al., Genetic and
pharmacological inhibition of galectin-3 prevents cardiac
remodeling by interfering with myocardial brogenesis,Cir-
culation: Heart Failure, vol. 6, no. 1, pp. 107117, 2013.
[154] R. A. de Boer, A. A. Voors, P. Muntendam, W. H. van Gilst,
and D. J. van Veldhuisen, Galectin-3: a novel mediator of
heart failure development and progression,European Jour-
nal of Heart Failure, vol. 11, no. 9, pp. 811817, 2009.
[155] K. Karatolios, G. Chatzis, V. Holzendorf et al., Galectin-3 as
a Predictor of Left Ventricular Reverse Remodeling in
Recent- Onset Dilated Cardiomyopathy,Disease Markers,
vol. 2018, Article ID 2958219, 7 pages, 2018.
[156] R. A. de Boer, L. Yu, and D. J. van Veldhuisen, Galectin-3 in
cardiac remodeling and heart failure,Current Heart Failure
Reports, vol. 7, no. 1, pp. 18, 2010.
[157] R. A. de Boer, D. J. A. Lok, T. Jaarsma et al., Predictive value
of plasma galectin-3 levels in heart failure with reduced and
preserved ejection fraction,Annals of Medicine, vol. 43,
no. 1, pp. 6068, 2010.
[158] D. J. Lok, S. I. Lok, P. W. Bruggink-André de la Porte et al.,
Galectin-3 is an independent marker for ventricular remod-
eling and mortality in patients with chronic heart failure,
Clinical Research in Cardiology, vol. 102, no. 2, pp. 103110,
2013.
[159] L. C. van Vark, I. Lesman-Leegte, S. J. Baart et al., Prognostic
value of serial galectin-3 measurements in patients with acute
18 Disease Markers
heart failure,Journal of the American Heart Association,
vol. 6, no. 12, article e003700, 2017.
[160] A. Bayes-Genis, M. de Antonio, J. Vila et al., Head-to-head
comparison of 2 myocardial brosis biomarkers for long-
term heart failure risk stratication: ST2 versus galectin-3,
Journal of the American College of Cardiology, vol. 63, no. 2,
pp. 158166, 2014, Epub 2013 Sep 24.
[161] A. Chen, W. Hou, Y. Zhang, Y. Chen, and B. He, Prognostic
value of serum galectin-3 in patients with heart failure: A
meta- analysis,International Journal of Cardiology,
vol. 182, pp. 168170, 2015.
[162] J. Meluzin, J. Tomandl, H. Podrouzkova et al., Can markers
of collagen turnover or other biomarkers contribute to the
diagnostics of heart failure with normal left ventricular ejec-
tion fraction?,Biomedical Papers, vol. 157, no. 4, pp. 331
339, 2013, Epub 2013 Feb 14.
[163] R. Martos, J. Baugh, M. Ledwidge et al., Diagnosis of heart
failure with preserved ejection fraction: improved accuracy
with the use of markers of collagen turnover,European Jour-
nal of Heart Failure, vol. 11, no. 2, pp. 191197, 2009.
[164] F. H. Rutten, M. J. Cramer, and W. J. Paulus, Heart failure
with preserved ejection fraction: diastolic heart failure,
Nederlands Tijdschrift voor Geneeskunde, vol. 156, no. 45,
p. A5315, 2012.
[165] S. de Denus, J. Lavoie, A. Ducharme et al., Dierences in bio-
markers in patients with heart failure with a reduced vs a pre-
served left ventricular ejection fraction,Canadian Journal of
Cardiology, vol. 28, no. 1, pp. 6268, 2012.
[166] N. Maharaj, B. K. Khandheria, E. Libhaber et al., Relation-
ship between left ventricular twist and circulating biomarkers
of collagen turnover in hypertensive patients with heart fail-
ure,Journal of the American Society of Echocardiography,
vol. 27, no. 10, pp. 10641071, 2014.
[167] T. A. Zelniker, P. Jarolim, B. M. Scirica et al., Biomarker of
collagen turnover (C-terminal telopeptide) and prognosis in
patients with non-ST-elevation acute coronary syndromes,
Journal of the American Heart Association, vol. 8, no. 9, arti-
cle e011444, 2019.
[168] K. Nagao, T. Inada, A. Tamura et al., Circulating markers of
collagen types I, III, and IV in patients with dilated cardiomy-
opathy: relationships with myocardial collagen expression,
ESC Heart Failure, vol. 5, no. 6, pp. 10441051, 2018.
[169] A. M. Dupuy, N. Kuster, C. Curinier et al., Exploring colla-
gen remodeling and regulation as prognosis biomarkers in
stable heart failure,Clinica Chimica Acta, vol. 490,
pp. 167171, 2019.
[170] M. Volpe, M. Carnovali, and V. Mastromarino, The natri-
uretic peptides system in the pathophysiology of heart failure:
from molecular basis to treatment,Clinical Science, vol. 130,
no. 2, pp. 5777, 2016.
[171] M. Volpe, S. Rubattu, and J. Burnett, Natriuretic peptides in
cardiovascular diseases: current use and perspectives,Euro-
pean Heart Journal, vol. 35, no. 7, pp. 419425, 2014.
[172] R. Kerkelä, J. Ulvila, and J. Magga, Natriuretic peptides in
the regulation of cardiovascular physiology and metabolic
events,Journal of the American Heart Association, vol. 4,
no. 10, article e002423, 2015.
[173] C. Calvieri, S. Rubattu, and M. Volpe, Molecular mecha-
nisms underlying cardiac antihypertrophic and antibrotic
eects of natriuretic peptides,Journal of Molecular Medi-
cine, vol. 90, no. 1, pp. 513, 2012.
[174] D. K. Gupta and T. J. Wang, Natriuretic peptides and car-
diometabolic health,Circulation Journal, vol. 79, no. 8,
pp. 16471655, 2015.
[175] L. R. Potter, Natriuretic peptide metabolism, clearance and
degradation,FEBS Journal, vol. 278, no. 11, pp. 18081817,
2011.
[176] R. DAlessandro, D. Masarone, A. Buono et al., Natriuretic
peptides: molecular biology, pathophysiology and clinical
implications for the cardiologist,Future Cardiology, vol. 9,
no. 4, pp. 519534, 2013.
[177] Z. Cao, Y. Jia, and B. Zhu, BNP and NT-proBNP as diagnos-
tic biomarkers for cardiac dysfunction in both clinical and
forensic medicine,International Journal of Molecular Sci-
ences, vol. 20, no. 8, p. 1820, 2019.
[178] M. A. Alpert, C. J. Lavie, H. Agrawal, K. B. Aggarwal, and
S. A. Kumar, Obesity and heart failure: epidemiology,
pathophysiology, clinical manifestations, and manage-
ment,Translational Research, vol. 164, no. 4, pp. 345
356, 2014.
[179] E. H. Nah, S. Y. Kim, S. Cho, S. Kim, and H. I. Cho, Plasma
NT-proBNP levels associated with cardiac structural abnor-
malities in asymptomatic health examinees with preserved
ejection fraction: a retrospective cross-sectional study,BMJ
Open, vol. 9, no. 4, article e026030, 2019.
[180] U. L. Faxén, L. H. Lund, N. Orsini et al., N-terminal pro-B-
type natriuretic peptide in chronic heart failure: the impact of
sex across the ejection fraction spectrum,International Jour-
nal of Cardiology, vol. 287, pp. 6672, 2019, Epub 2019 Apr
11.
[181] F. S. Gaborit, C. Kistorp, T. Kümler et al., Prevalence of early
stages of heart failure in an elderly risk population: the
Copenhagen Heart Failure Risk Study,Open Heart, vol. 6,
no. 1, article e000840, 2019.
[182] Z. J. Han, X. D. Wu, J. J. Cheng et al., Diagnostic accuracy of
natriuretic peptides for heart failure in patients with pleural
eusion: a systematic review and updated meta-analysis,
PLoS One, vol. 10, no. 8, article e0134376, 2015.
[183] H.-P. B.-L. Rocca and S. S.-v. Wijk, Natriuretic peptides in
chronic heart failure,Cardiac Failure Review, vol. 5, no. 1,
pp. 4449, 2019.
[184] P. L. Myhre, M. Vaduganathan, B. Claggett et al., B-type
natriuretic peptide during treatment with sacubitril/valsar-
tan: the PARADIGM-HF Trial,Journal of the American
College of Cardiology, vol. 73, no. 11, pp. 12641272,
2019.
[185] S. Stienen, K. Salah, A. H. Moons et al., NT-proBNP (N-ter-
minal pro-B-type natriuretic peptide)-guided therapy in
acute decompensated heart failure: PRIMA II randomized
controlled trial (can NT-proBNP-guided therapy during hos-
pital admission for acute decompensated heart failure reduce
mortality and readmissions?),Circulation, vol. 137, no. 16,
pp. 16711683, 2018, Epub 2017 Dec 14.
[186] G. M. Felker, K. J. Anstrom, K. F. Adams et al., Eect of
natriuretic peptide-guided therapy on hospitalization or car-
diovascular mortality in high-risk patients with heart failure
and reduced ejection fraction: a randomized clinical trial,
JAMA, vol. 318, no. 8, pp. 713720, 2017.
[187] N. G. Morgenthaler, J. Struck, C. Alonso, and A. Bergmann,
Assay for the measurement of copeptin, a stable peptide
derived from the precursor of vasopressin,Clinical Chemis-
try, vol. 52, no. 1, pp. 112119, 2006.
19Disease Markers
[188] L. Dobša and K. C. Edozien, Copeptin and its potential role
in diagnosis and prognosis of various diseases,Biochemia
Medica, vol. 23, no. 2, pp. 172192, 2013.
[189] D. Bolignano, A. Cabassi, E. Fiaccadori et al., Copeptin
(CTproAVP), a new tool for understanding the role of vaso-
pressin in pathophysiology,Clinical Chemistry and Labora-
tory Medicine (CCLM), vol. 52, no. 10, pp. 14471456, 2014.
[190] N. G. Morgenthaler, Copeptin: a biomarker of cardiovascu-
lar and renal function,Congestive Heart Failure, vol. 16,
Suppl 1, pp. S37S44, 2010.
[191] W. J. Tu, G. Z. Ma, Y. Ni et al., Copeptin and NT-proBNP
for prediction of all-cause and cardiovascular death in ische-
mic stroke,Neurology, vol. 88, no. 20, pp. 18991905, 2017.
[192] W. Fenske, C. Wanner, B. Allolio et al., Copeptin levels asso-
ciate with cardiovascular events in patients with ESRD and
type 2 diabetes mellitus,Journal of the American Society of
Nephrology, vol. 22, no. 4, pp. 782790, 2011.
[193] I. Tasevska, S. Enhorning, M. Persson, P. M. Nilsson, and
O. Melander, Copeptin predicts coronary artery disease car-
diovascular and total mortality,Heart, vol. 102, no. 2,
pp. 127132, 2016.
[194] S. S. Bhandari, I. Loke, J. E. Davies, I. B. Squire, J. Struck, and
L. L. Ng, Gender and renal function inuence plasma levels
of copeptin in healthy individuals,Clinical Science, vol. 116,
no. 3, pp. 257263, 2009.
[195] G. Velho, S. Ragot, R. el Boustany et al., Plasma copeptin,
kidney disease, and risk for cardiovascular morbidity and
mortality in two cohorts of type 2 diabetes,Cardiovascular
Diabetology, vol. 17, no. 1, p. 110, 2018.
[196] A. E. Berezin, Up-to-date clinical approaches of biomarkers
use in heart failure,Biomedical Research and Therapy, vol. 4,
no. 6, pp. 13441370, 2017.
[197] K. Kitamura, K. Kangawa, and T. Eto, Adrenomedullin and
PAMP: discovery, structures, and cardiovascular functions,
Microscopy Research and Technique, vol. 57, no. 1, pp. 313,
2002.
[198] T. Nishikimi and Y. Nakagawa, Adrenomedullin as a bio-
marker of heart failure,Heart Failure Clinics, vol. 14, no. 1,
pp. 4955, 2018, Epub 2017 Oct 7.
[199] K. Kobayashi, K. Kitamura, N. Hirayama et al., Increased
plasma adrenomedullin in acute myocardial infarction,
American Heart Journal, vol. 131, no. 4, pp. 676680, 1996.
[200] Y. Miyao, T. Nishikimi, Y. Goto et al., Increased plasma
adrenomedullin levels in patients with acute myocardial
infarction in proportion to the clinical severity,Heart,
vol. 79, no. 1, pp. 3944, 1998.
[201] O. S. Dhillon, S. Q. Khan, H. K. Narayan et al., Prognostic
Value of Mid-Regional Pro-Adrenomedullin Levels Taken
on Admission and Discharge in NonST-Elevation Myocar-
dial Infarction: The LAMP (Leicester Acute Myocardial
Infarction Peptide) II Study,Journal of the American College
of Cardiology, vol. 56, no. 2, pp. 125133, 2010.
[202] I. T. Klip, A. A. Voors, S. D. Anker et al., Prognostic value of
mid-regional pro-adrenomedullin in patients with heart fail-
ure after an acute myocardial infarction,Heart, vol. 97,
no. 11, pp. 892898, 2011.
[203] S. Q. Khan, R. J. OBrien, J. Struck et al., Prognostic value of
midregional pro-adrenomedullin in patients with acute myo-
cardial infarction: the LAMP (Leicester Acute Myocardial
Infarction Peptide) study,Journal of the American College
of Cardiology, vol. 49, no. 14, pp. 15251532, 2007.
[204] A. Maisel, C. Mueller, R. Nowak et al., Mid-region pro-
hormone markers for diagnosis and prognosis in acute dys-
pnea: results from the BACH (Biomarkers in Acute Heart
Failure) trial,Journal of the American College of Cardiology,
vol. 55, no. 19, pp. 20622076, 2010.
[205] A. C. Falkentoft, R. Rørth, K. Iversen et al., MR-proADM as
a prognostic marker in patients with ST-segment-elevation
myocardial infarction-DANAMI-3 (a Danish Study of Opti-
mal Acute Treatment of Patients With STEMI) Substudy,
Journal of the American Heart Association, vol. 7, no. 11, arti-
cle e008123, 2018.
[206] A. Gegenhuber, J. Struck, B. Dieplinger et al., Comparative
evaluation of B-type natriuretic peptide, mid-regional pro-
A-type natriuretic peptide, mid-regional pro-adrenomedul-
lin, and copeptin to predict 1-year mortality in patients with
acute destabilized heart failure,Journal of Cardiac Failure,
vol. 13, no. 1, pp. 4249, 2007.
[207] O. Melander, C. Newton-Cheh, P. Almgren et al., Novel and
conventional biomarkers for prediction of incident cardio-
vascular events in the community,JAMA, vol. 302, no. 1,
pp. 4957, 2009.
[208] J. T. Neumann, S. Tzikas, A. Funke-Kaiser et al., Association
of MR-proadrenomedullin with cardiovascular risk factors
and subclinical cardiovascular disease,Atherosclerosis,
vol. 228, no. 2, pp. 451459, 2013.
[209] K. S. Shah, N. A. Marston, C. Mueller et al., Midregional
proadrenomedullin predicts mortality and major adverse car-
diac events in patients presenting with chest pain: results
from the CHOPIN trial,Academic Emergency Medicine,
vol. 22, no. 5, pp. 554563, 2015.
[210] C. Sinning, F. Ojeda, P. S. Wild et al., Midregional proadre-
nomedullin and growth dierentiation factor-15 are not
inuenced by obesity in heart failure patients,Clinical
Research in Cardiology, vol. 106, no. 6, pp. 401410, 2017.
[211] S. Dangwal, K. Schimmel, A. Foinquinos, K. Xiao, and
T. Thum, Noncoding RNAs in heart failure,in Heart Fail-
ure, vol. 243 of Handbook of Experimental Pharmacology,
pp. 423445, Springer, Cham, 2017.
[212] E. L. Vegter, P. van der Meer, L. J. de Windt, Y. M. Pinto, and
A. A. Voors, MicroRNAs in heart failure: from biomarker to
target for therapy,European Journal of Heart Failure,
vol. 18, no. 5, pp. 457468, 2016, Epub 2016 Feb 11.
[213] W. Li, M. Liu, C. Zhao et al., MiR-1/133 attenuates cardio-
myocyte apoptosis and electrical remodeling in mice with
viral myocarditis,Cardiology Journal, 2013, [Epub ahead of
print].
[214] A. Jodati, S. M. Pirouzpanah, N. Fathi Marouet al., Dier-
ent expression of Micro RNA-126, 133a and 145 in aorta and
saphenous vein samples of patients undergoing coronary
artery bypass graft surgery,Journal of Cardiovascular and
Thoracic Research, vol. 11, no. 1, pp. 4347, 2019.
[215] Y. Xiao, J. Zhao, J. P. Tuazon, C. V. Borlongan, and G. Yu,
MicroRNA-133a and myocardial infarction,Cell Trans-
plantation, vol. 28, no. 7, pp. 831838, 2019, [Epub ahead of
print].
[216] J. Song, Q. Xie, L. Wang et al., The TIR/BB-loop mimetic
AS-1 prevents Ang II-induced hypertensive cardiac hypertro-
phy via NF-κB dependent downregulation of miRNA-143,
Scientic Reports, vol. 9, no. 1, p. 6354, 2019.
[217] P. Shah, M. R. Bristow, and J. D. Port, MicroRNAs in heart
failure cardiac transplantation, and myocardial recovery:
20 Disease Markers
biomarkers with therapeutic potential,Current Heart Fail-
ure Reports, vol. 14, no. 6, pp. 454464, 2017.
[218] R. Verjans, W. J. A. Derks, K. Korn et al., Functional screen-
ing identies microRNAs as multi-cellular regulators of heart
failure,Scientic Reports, vol. 9, no. 1, p. 6055, 2019.
[219] A. Samman Tahhan, M. Hammadah, M. Raad et al., Progen-
itor cells and clinical outcomes in patients with acute coro-
nary syndromes,Circulation Research, vol. 122, no. 11,
pp. 15651575, 2018.
[220] E. Shantsila, B. J. Wrigley, A. Shantsila, L. D. Tapp, P. S. Gill,
and G. Y. H. Lip, Monocyte-derived and CD34+/KDR+
endothelial progenitor cells in heart failure,Journal of
Thrombosis and Haemostasis, vol. 10, no. 7, pp. 12521261,
2012.
[221] E. Nollet, V. Y. Hoymans, I. R. Rodrigus et al., Bone
marrow-derived progenitor cells are functionally impaired
in ischemic heart disease,Journal of Cardiovascular Transla-
tional Research, vol. 9, no. 4, pp. 266278, 2016.
[222] A. E. Berezin and A. A. Kremzer, Circulating endothelial
progenitor cells as markers for severity of ischemic chronic
heart failure,Journal of Cardiac Failure, vol. 20, no. 6,
pp. 438447, 2014.
[223] A. E. Berezin, A. A. Kremzer, Y. V. Martovitskaya et al., The
utility of biomarker risk prediction score in patients with
chronic heart failure,International Journal of Clinical and
Experimental Medicine, vol. 8, no. 10, pp. 1825518264, 2015.
[224] A. E. Berezin, A. A. Kremzer, T. A. Samura et al., Predictive
value of apoptotic microparticles to mononuclear progenitor
cells ratio in advanced chronic heart failure patients,Journal
of Cardiology, vol. 65, no. 5, pp. 403411, 2015.
[225] J. A. SuárezCuenca, R. RobledoNolasco, M. A. Alcántara
Meléndez et al., Coronary circulating mononuclear progen-
itor cells and soluble biomarkers in the cardiovascular prog-
nosis after coronary angioplasty,Journal of Cellular and
Molecular Medicine, vol. 23, no. 7, 2019.
21Disease Markers
... At present, copeptin was proven to show the same response as AVP to hypotension or hemodynamic stress. [60][61][62][63][64] In recent years, Copeptin has gained growing interest as part of a Dual Marker Strategy (DMS) in combination with Cardiac Troponin (cTn) in promptly ruling out Acute Myocardial Infarction (AMI) in patients presenting with symptoms suggestive of ACS. 63 Several studies, among which the most important is the Copeptin Helps in the Early Detection of Patients with Acute Myocardial Infarction (CHOPIN) trial, a multicenter international cohort study, support that copeptin and hs-cTnT in combination has the potential to allow a faster ruling of AMI when compared to the hs-cTnT-only-based algorithms and enables a useful reclassification of profile risk of patients. ...
... heart failure, metabolic syndrome, hypertension, acute kidney injury, pulmonary embolism, sepsis, acute pancreatitis,ischemic stroke) and that they are linked to the severity of these clinical situation. [62][63][64]66,[67][68][69][70] ...
... Based on current evidence, high levels of GDF-15 are associated with a poor prognosis for patients with acute heart failure. The elevated serum level of GDF-15 was the most prognostic biomarker in comparison to NT-proBNP, hs-CRP, and hs-TnT and an independent predictor of long-term mortality in advanced HF. 62,76 Recent studies propose an association between elevated GDF-15 levels and different clinical conditions (e.g., ineffective erythropoiesis inflammation, acute injury, cancer, and chronic kidney disease, ischemic stroke). Concerning ACS, it has been established that GDF-15 levels rise in just a few hours after MI. 64 However, in consideration of its important functions, a role in acute care must be further evaluated; for these reasons focused studies are necessary. ...
Article
Full-text available
Chest pain is one of the most prevalent causes of Emergency Department (ED) admission and could be a presenting symptom of Acute Coronary Syndrome (ACS). The aim of this review was to provide an overview of the research about troponin and its limitations and new biomarkers used in patients with cardiovascular diseases, with a special focus on soluble Suppression of Tumorigenicity 2 (sST2) and Soluble Urokinase Plasminogen Activator Receptor (suPAR). In January 2024, a PubMed and Reviews in Cardiovascular Medicine (RCM) search was carried out to identify all relevant papers in the past five years. 80 articles were included in the final review. ssT2 and suPAR are involved in both acute and chronic cardiovascular disease and can predict the risk of adverse events. sST2 and suPAR are promising biomarkers that, in combination with troponin, could help in the management of patients with chest pain in the ED. Further studies are needed to validate their role in management of ACS in this specific setting.
... Myocardial infarction (MI), which is predominantly caused by myocardial ischaemia, leads to large areas of myocardial cell necrosis and apoptosis (1). Cardiac remodelling after MI is characterized by inflammation, fibrosis and cardiac hypertrophy in the remaining myocardium (2). This adverse cardiac remodelling eventually leads to heart failure (3,4). ...
... The 30-day survival rate of acute ST segment elevation MI has increased to 95% (4). However, despite revascularization, patients with AMI have a 75% probability of developing heart failure within 5 years (2,3). An increasing number of patients that survive AMI develop heart failure. ...
Article
Full-text available
Meteorin-β (Metrnβ) is a protein that is secreted by skeletal muscle and adipose tissue, and participates in cardiovascular diseases. However, its role in myocardial infarction (MI) has not been fully elucidated to date. The aim of the present study was to investigate the role and underlying mechanism of Metrnβ in MI. In the present study, mice were subjected to left coronary ligation to induce a MI model before being injected with adeno-associated virus 9 (AAV9)-Metrnβ to overexpress Metrnβ. Mice were subjected to echocardiography and pressure-volume measurements 2 weeks after ligation. Cardiac injury was measured from the levels of cardiac troponin T and pro-inflammatory factors, which were detected using ELISA kits. Cardiac remodelling was determined from the cross-sectional areas detected using H&E and wheat germ agglutinin staining as well as from the transcriptional levels of hypertrophic and fibrosis markers detected using reverse transcription-quantitative PCR. Cardiac function was detected using echocardiography and pressure-volume measurements. In addition, H9c2 cardiomyocytes were transfected with Ad-Metrnβ to overexpress Metrnβ, before being exposed to hypoxia to induce ischaemic injury. Apoptosis was determined using TUNEL staining and caspase 3 activity. Cell inflammation was detected using ELISA assays for pro-inflammatory factors. Autophagy was detected using LC3 staining and assessing the protein level of LC3II using western blotting. H9c2 cells were also treated with rapamycin to induce autophagy. It was revealed that Metrnβ expression was reduced in both mouse serum and heart tissue 2 weeks post-MI. Metrnβ overexpression using AAV9-Metrnβ delivery reduced the mortality rate, decreased the infarction size and reduced the extent of myocardial injury 2 weeks post-MI. Furthermore, Metrnβ overexpression inhibited cardiac hypertrophy, fibrosis and inflammation post-MI. In ischaemic H9c2 cells, Metrnβ overexpression using adenovirus also reduced cell injury, cell death and inflammatory response. Metrnβ overexpression suppressed MI-induced autophagy in vitro. Following autophagy activation using rapamycin in vitro, the protective effects induced by Metrnβ were reversed. Taken together, these results indicated that Metrnβ could protect against cardiac dysfunction post-MI in mice by inhibiting autophagy.
... IL-1RA concentrations also peaked notably among diabetic individuals during G2 but not in G3 phases of LVR; these results are congruent with studies conducted by Scărlătescu et al., (2022). Significant variations were observed concerning Gal-3 levels across different LVR timeframes, emphasizing increased levels during extended remodeling periods as noted by Berezin and Berezin (2020). Contrarily, IMA presented no substantial fluctuation relative to remodeling times. ...
... Current interventions for advanced HF remain inadequate, with only a 50% survival rate observed 5 years after treatment across various HF phenotypes. 6,7 Heart transplantation is the only suitable therapeutic option but is constrained by limitations such as scarcity of heart donors, transplant rejection, and adverse reactions such as infections and malignant tumors induced by immunosuppressants. Hence, interventions for adverse developments after MI are pivotal in preventing the onset of HF. ...
Article
Myocardial infarction is a cardiovascular disease characterized by a high incidence rate and mortality. It leads to various cardiac pathophysiological changes, including ischemia/reperfusion injury, inflammation, fibrosis, and ventricular remodeling, which ultimately result in heart failure and pose a significant threat to global health. Although clinical reperfusion therapies and conventional pharmacological interventions improve emergency survival rates and short-term prognoses, they are still limited in providing long-lasting improvements in cardiac function or reversing pathological progression. Recently, cardiac patches have gained considerable attention as a promising therapy for myocardial infarction. These patches consist of scaffolds or loaded therapeutic agents that provide mechanical reinforcement, synchronous electrical conduction, and localized delivery within the infarct zone to promote cardiac restoration. This review elucidates the pathophysiological progression from myocardial infarction to heart failure, highlighting therapeutic targets and various cardiac patches. The review considers the primary scaffold materials, including synthetic, natural, and conductive materials, and the prevalent fabrication techniques and optimal properties of the patch, as well as advanced delivery strategies. Last, the current limitations and prospects of cardiac patch research are considered, with the goal of shedding light on innovative products poised for clinical application.
... These troponins are produced and released from cardiomyocytes in response to increased stretch or necrosis. Cardiac-specific troponins cTnT and cTnI serve as crucial predictive biomarkers for cardiac damage and acute myocardial infarction (AMI) (Berezin and Berezin, 2020). ...
Article
Full-text available
Background: Cardiovascular diseases are the leading cause of morbidity and mortality globally, with acute myocardial infarction (AMI) being a significant contributor. This study aimed to investigate the roles of cardiac biomarkers, including H-FABP, GPBB, and others, in detecting AMI. Method: Blood samples were collected from 80 individuals, including 50 with AMI and 30 healthy controls, admitted to the Coronary Care Unit (CCU) of an Educational Hospital in Diyala province between May and July 2022. Cardiac markers (hs-troponin-I, myoglobin, CK-MB, GPBB, and H-FABP) were measured using the Sandwich-ELISA technique. Results: There were no significant differences (p>0.05) in age or gender distribution between the study groups. Levels of cardiac markers were significantly higher in AMI patients compared to healthy controls (p<0.05). H-FABP demonstrated the highest sensitivity (100%), followed by GPBB (97%), hs-troponin-I (87%), CK-MB (85%), and myoglobin (78%), with significant differences (p<0.05) in detecting AMI. H-FABP and GPBB also exhibited the highest specificity (98% and 96%, respectively), while myoglobin and CK-MB had lower specificity (82% and 84%, respectively). Furthermore, positive correlations were observed between H-FABP, GPBB, and other markers (hs-troponin-I, myoglobin, CK-MB). Conclusion: H-FABP and GPBB show promise as predictive indicators for early diagnosis (within 1-4 hours of chest pain) of AMI, offering potential utility in clinical practice.
... Myocardial ischemia often causes myocardial hypertrophy, fibrosis, inflammation, and persistent stiffness. 45 Telomeres are located at the ends of eukaryotic chromosomes and help maintain their integrity. 46,47 Telomerase contains telomerase reverse transcriptase (TERT) and its RNA component (TERC). ...
Article
Full-text available
Background Cardiac repair remains a thorny issue for survivors of acute myocardial infarction (AMI), due to the regenerative inertia of myocardial cells. Cell-free therapies, such as exosome transplantation, have become a potential strategy for myocardial injury. The aim of this study was to investigate the role of engineered exosomes in overexpressing Growth Differentiation Factor-15 (GDF-15) (GDF15-EVs) after myocardial injury, and their molecular mechanisms in cardiac repair. Methods H9C2 cells were transfected with GDF-15 lentivirus or negative control. The exosomes secreted from H9C2 cells were collected and identified. The cellular apoptosis and autophagy of H2O2-injured H9C2 cells were assessed by Western blotting, TUNEL assay, electron microscopy, CCK-8 and caspase 3/7 assay. A rat model of AMI was constructed by ligating the left anterior descending artery. The anti-apoptotic, pro-angiogenic effects of GDF15-EVs treatment, as well as ensuing functional and histological recovery were evaluated. Then, mRNA sequencing was performed to identify the differentially expressed mRNAs after GDF15-EVs treatment. Results GDF15-EVs inhibited apoptosis and promoted autophagy in H2O2 injured H9C2 cells. GDF15-EVs effectively decreased the infarct area and enhanced the cardiac function in rats with AMI. Moreover, GDF15-EVs hindered inflammatory cell infiltration, inhibited cell apoptosis, and promoted cardiac angiogenesis in rats with AMI. RNA sequence showed that telomerase reverse transcriptase (TERT) mRNA was upregulated in GDF15-EVs-treated H9C2 cells. AMPK signaling was activated after GDF15-EVs. Silencing TERT impaired the protective effects of GDF15-EVs on H2O2-injured H9C2 cells. Conclusion GDF15-EVs could fulfil their protective effects against myocardial injury by upregulating the expression of TERT and activating the AMPK signaling pathway. GDF15-EVs might be exploited to design new therapies for AMI.
Article
In the present work, we carried out a comparative analysis of myocardial cytokine profile in patients with coronary heart disease (CHD) and in patients with ischemic cardiomyopathy (ICMP) associated with CHD. The concentrations of 41 cytokines secreted by 24-hour myocardial tissue culture intraoperatively sampled from the right atrial auricle (RAA, control) and peri-infarct left ventricular zone (PZ-LV) were determined by flow fluorimetry using a multiplex test system. The aim was to study in vitro cytokine profile of myocardial cells to search for possible predictors of adverse outcomes of surgical treatment of patients with CHD and ICMP. Myocardial secretion of proinflammatory molecules GM-CSF and IFN-γ increased significantly (up to 78-80 pg/g, p0.05) in patients with ICMP associated with CHD in contrast to zero values in CHD. At the same time, there was a three-fold decrease in the concentration of fractalkin 3 ligand (Flt-3L; FMS-like tyrosine kinase 3 ligand). A decrease in Flt-3L secretion was observed in the PZ-LV in comparison with the RAA. In addition, compared with RAA, concentrations of fibroblast growth factor-2 (FGF-2), platelet-derived growth factor-AB/BB (PDGFAB/BB), interleukins IL-15 and IL-4, and a regulated upon activation, normal T cell expressed and secreted (RANTES; CCL5) were strongly reduced in PZ-LV myocardial tissue culture. Differences in the course of CHD and ICMP are discussed, and possible predictors of surgical treatment risk in patients of the two groups are suggested using correlation and regression analyses. Proinflammatory cytokines (IL-5, IL-6) and chemokines (Flt-3L, IL-8), as well as angiogenesis factors (VEGF) and angiostasis (IP-10), are proposed to be considered as potential markers of adverse outcome of surgical treatment of cardiovascular disease.
Article
Aim. To investigate the polymorphic variants of IL10, FGF2, VEGFD, TRAIL, SELE, TNFA and TNFβ genes in patients with primary ST-segment elevation myocardial infarction (MI) (STEMI) and to evaluate their association with late post-infarction cardiac remodeling. Material and methods . The study includes 74 patients age 61±10,7 years with primary STEMI. Percutaneous coronary intervention with restoration of infarct-related artery patency was performed in all patients after 60 (40; 80) minutes since admission to the hospital. Serum levels of fibroblast growth factor (FGF), interleukin-10 (IL-10), tumor necrosis factor family cytokines (TNF-α, TNF-β and tumor necrosis factor-related apoptosis-inducing ligand (TRAIL)) were measured with the Multiplex Instrument FLEXMAP 3D system (Luminex Corporation) and the MILLIPLEX Human Cytokine/Chemokine Panel II on the 1st, 7th day of early post-MI period and after 6 and 12 months. The late adverse cardiac remodeling was determined after 12 months of long-term post-MI period according to 2D echo­cardiography. The increase in left ventricular end-diastolic volume by 15% or more by 12 months was considered late adverse cardiac remodeling. The patients were genotyped by 16 single-nucleotide polymorphisms (SNPs) in the TNFβ, TNF, Il10, TNFRSF1B, VEGFD, TRAIL, FGF2, SELE genes. Results. Adverse cardiac remodeling occurred in 19 patients (25,7%). The genetic association analysis revealed the significant association of rs1800629 TNFA (χ ² =4,748; p=0,029), rs5353 SELE (χ ² =10,85; p=0,004) and rs6632528 VEGFD (χ ² =8,127; p=0,017) with an increased risk of STEMI. Higher concentration of IL-10 was detected on the 7th day of MI (p=0,05) and 6 months later (p=0,028) in A/T rs3024492 genotype carriers, as well as FGF in T/T rs13122694 genotype carriers by the 6th month after the event (p=0,04). The dependence of the main LV indicators on the distribution of polymorphism genotypes rs3024492 IL10, rs13122694 FGF2 and rs4830939 VGEFD was discovered. In the first 24 hours of MI in rs3024492 IL10 heterozygotes, LV contractile function was worse in comparison with T/T genotype carriers. Also, carriers of the T/T rs13122694 FGF2 genotype were distinguished by higher LV ejection fraction, longitudinal global LV deformation and lower of LV end-systolic index in the early post-infarction period. In the long-term post-infarction period, T/T rs4830939 VEGFD carriers differed in a greater LV dilation than carriers of the C/C and C/T genotypes. Conclusion. The study showed the contribution of polymorphism of the inflammation system genes to a predisposition to STEMI — both at the levels of phenotype and individual signs.
Preprint
Full-text available
Introduction: Inflammation is a hallmark of post-ischemic myocardial injury. Expression of P-selectin by platelets and activated endothelial cells drives recruitment of immune cells to the deprived area and may serve as an early indicator of tissue injury. Due to its high soft tissue contrast, magnetic resonance imaging (MRI) is used for myocardial tissue characterization. Molecular imaging further allows for functional assessment using target-specific contrast agents. In this study, we assessed ischemic cardiac lesions non-invasively within the first hours after ischemia/reperfusion (I/R) in a porcine model using standard and advanced MRI techniques as well as molecular imaging targeting the cell adhesion molecule P-selectin. Methods: For molecular imaging, a monoclonal P-selectin antibody was functionalized with microparticles of iron oxide (MPIO). Specific binding to the target was confirmed by in vitro flow chamber using activated platelets as well as endothelial cells. In vivo, we used a closed-chest model of I/R of the circumflex artery in juvenile farm pigs by balloon-occlusion for 40 minutes, and real time MRI-guided coronary injection of MPIO-based contrast agents. 3T MRI was performed 2–4 hours after reperfusion, and lesions were characterized using injury (T1 mapping, LGE), edema (T2 mapping) and iron (T2* mapping) sensitive MRI. Results: Within the first hours after I/R, we detected increased inflammatory activity by means of higher numbers of innate immune cells in the blood. We found T1 mapping to be most sensitive for tissue injury, while no changes were detectable in edema-sensitive T2 mapping this early. Intriguingly, P-selectin MPIO contrast agent selectively enhanced the ischemic area in iron sensitive T2* mapping 4 hours after I/R which was confirmed histologically, while late gadolinium enhancement was always absent. Conclusion: By using real time MRI-guided coronary intervention, molecular MRI using P-selectin MPIO allows for sensitive detection of early myocardial inflammation after I/R beyond the capabilities of traditional edema sensitive imaging.
Article
Full-text available
Currently, there are no confident prognostic markers in patients with coronary artery disease (CAD) undergoing angioplasty. The present study aimed to explore whether basal coronary circulating Mononuclear Progenitor Cells (MPCs) and vascular injury biomarkers were related to development of major adverse cardiovascular events (MACEs) and may impact clinical prognosis. Methods The number of MPCs and soluble mediators such as IL‐1β, sICAM‐1, MMP‐9, malondialdehyde, superoxide dismutase and nitric oxide were determined in coronary and peripheral circulation. Prognostic ability for MACEs occurring at 6 months follow up was assessed by time‐to‐event and event free survival estimations. Results Lower coronary circulating MPCs subpopulations CD45⁺CD34⁺, CD45⁺CD34⁺CD133⁺CD184⁺, lower MMP‐9 and higher sICAM‐1 significantly associated with MACEs presentation and showed prognostic ability; while peripheral blood increase in malondialdehyde and decreased superoxide dismutase were observed in patients with MACEs. Conclusion Coronary concentration of biomarkers related with vascular repair, such as MPCs subpopulations and adhesion molecules, may predict MACEs and impact prognosis in patients with CAD undergoing angioplasty; whereas peripheral pro‐oxidative condition may be also associated.
Article
Full-text available
Background Small studies have suggested an association between markers of collagen turnover and adverse outcomes in heart failure ( HF ). We examined C‐terminal telopeptide (beta‐ CT x) and the risk of cardiovascular death or new or worsening HF in non– ST ‐elevation acute coronary syndrome. Methods and Results We measured baseline serum beta‐ CT x, NT ‐pro BNP (N‐terminal pro‐B‐type natriuretic peptide), hsTnT (high‐sensitivity cardiac troponin T) and hs CRP (high‐sensitivity C‐reactive protein) (Roche Diagnostics) in a nested biomarker analysis (n=4094) from a study of patients with non– ST ‐elevation acute coronary syndrome. The relationship between quartiles of beta‐ CT x and cardiovascular death or HF over a median follow‐up time of 12 months was analyzed using adjusted Cox models. Higher beta‐ CT x levels identified a significantly higher risk of cardiovascular death/ HF (Q4 10.9% versus Q1 3.8%, Logrank P <0.001). After multivariable adjustment, beta‐ CT x in the top quartile (Q4) was associated with cardiovascular death/ HF (Q4 versus Q1: adjusted hazard ratio 2.22 [1.50–3.27]) and its components (Q4 versus Q1: cardiovascular death: adjusted hazard ratio 2.48 [1.46–4.21]; HF : adjusted hazard ratio 2.04 [1.26–3.30]). In an adjusted multimarker model including NT ‐pro BNP , hsTnT, and hs CRP , beta‐ CT x remained independently associated with cardiovascular death/ HF (Q4 versus Q1: adjusted hazard ratio 1.98 [1.34–2.93]) and its components. Beta‐ CT x correlated weakly with NT ‐pro BNP ( r =0.17, P <0.001) and left ventricular ejection fraction ( r =−0.05, P =0.008) and did not correlate with hsTnT ( r =0.02, P =0.20), or hs CRP ( r =−0.03, P =0.09). Conclusions Levels of beta‐ CT x, a biomarker of collagen turnover, were associated with cardiovascular death and HF in patients with non– ST ‐elevation acute coronary syndrome. This biomarker, which correlated only weakly or not significantly with traditional biomarkers of cardiovascular death and HF , may provide complementary pathobiological insight and risk stratification in these patients.
Article
Full-text available
Background There is great interest in widening the use of high-sensitivity cardiac troponins for population cardiovascular disease (CVD) and heart failure screening. However, it is not clear whether cardiac troponin T (cTnT) and troponin I (cTnI) are equivalent measures of risk in this setting. We aimed to compare and contrast (1) the association of cTnT and cTnI with CVD and non-CVD outcomes, and (2) their determinants in a genome-wide association study. Methods High-sensitivity cTnT and cTnI were measured in serum from 19 501 individuals in Generation Scotland Scottish Family Health Study. Median follow-up was 7.8 years (quartile 1 to quartile 3, 7.1–9.2). Associations of each troponin with a composite CVD outcome (1177 events), CVD death (n=266), non-CVD death (n=374), and heart failure (n=216) were determined by using Cox models. A genome-wide association study was conducted using a standard approach developed for the cohort. Results Both cTnI and cTnT were strongly associated with CVD risk in unadjusted models. After adjusting for classical risk factors, the hazard ratio for a 1 SD increase in log transformed troponin was 1.24 (95% CI, 1.17–1.32) and 1.11 (1.04–1.19) for cTnI and cTnT, respectively; ratio of hazard ratios 1.12 (1.04–1.21). cTnI, but not cTnT, was associated with myocardial infarction and coronary heart disease. Both cTnI and cTnT had strong associations with CVD death and heart failure. By contrast, cTnT, but not cTnI, was associated with non-CVD death; ratio of hazard ratios 0.77 (0.67–0.88). We identified 5 loci (53 individual single-nucleotide polymorphisms) that had genome-wide significant associations with cTnI, and a different set of 4 loci (4 single-nucleotide polymorphisms) for cTnT. Conclusions The upstream genetic causes of low-grade elevations in cTnI and cTnT appear distinct, and their associations with outcomes also differ. Elevations in cTnI are more strongly associated with some CVD outcomes, whereas cTnT is more strongly associated with the risk of non-CVD death. These findings help inform the selection of an optimal troponin assay for future clinical care and research in this setting.
Article
Full-text available
Untreated pathological cardiac hypertrophy, which can be caused by sustained systemic hypertension, may lead to heart failure. In the present study, we investigated whether AS-1 had attenuating effects on hypertension-induced cardiac hypertrophy, and whether this process was mediated by the regulation of miRNA-143. To induce the hypertrophic response in vitro, cardiomyocytes were stimulated with Ang II for 24hs. AS-1 administration strongly attenuated Ang II-induced hypertrophic response of cardiomyocytes. Chronical infusion of Ang II via implanted osmotic mini-pump induced increased blood pressure and cardiac hypertrophy in vivo. AS-1 administration attenuated hypertension-induced cardiac hypertrophy by, at least in part, inhibin of MAPK signaling. We observed, for the first time, upregulated expression of miRNA-143 in Ang II-induced cardiomyocytes, and inhibition of miRNA-143 significantly reduced the Ang II-induced hypertrophic responses. Importantly, AS-1 administration diminished the Ang II-induced upregulation of miRNA-143. Overexpression of miRNA-143 abolished the attenuating effects of AS-1 on Ang II-induced hypertrophic response of cardiomyocytes. Additionally, AS-1 administration abrogates Ang II-induced nuclear translocation of p50 NF-κB subunit in hypertrophic cardiomyocytes. Application of NF-κB inhibitor significantly suppressed Ang II-induced upregulation of miRNA-143. Our data suggest a novel mechanism by which AS-1 attenuates Ang II-induced hypertrophic response through downregulation miRNA-143 expression in a NF-κB-dependent manner.
Article
Full-text available
Objectives Stage B heart failure (HF) is defined as an asymptomatic abnormality of the heart structure or function. The circulating level of N-terminal pro-B-type natriuretic peptide (NT-proBNP) is elevated in symptomatic patients with left ventricular (LV) dysfunction caused by a structural or functional abnormality. This study investigated the association of the NT-proBNP level with echocardiography-detected cardiac structural or diastolic abnormalities in asymptomatic subjects with preserved LV systolic function (ejection fraction >50%). Methods We retrospectively studied 652 health examinees who underwent echocardiography and an NT-proBNP test at a health-promotion centre in Seoul, between January 2016 and September 2018. The left ventricular mass index (LVMI) and the left atrial dimension (LAD) were used as markers for structural abnormalities, and the mean e’ velocity and mitral early flow velocity/early diastolic tissue velocity (E/e’) ratio were used as markers for diastolic dysfunction. The plasma NT-proBNP level was measured using electrochemiluminescence immunoassay (DPC Immulite 2000 XPi, Siemens Healthcare Diagnostics, Tarrytown, New York, USA). Results Subjects with preclinical structural abnormalities were older and had a higher body mass index (BMI), higher blood pressure, lower high-density lipoprotein cholesterol level, higher NT-proBNP level, and higher E/e’ (p<0.05). Multivariate regression analysis indicated that the factors associated with a higher NT-proBNP level were older age, female sex, lower BMI, higher creatinine level, higher LVMI and higher LAD (p<0.01). Conclusion Diastolic dysfunction is not associated with higher NT-proBNP levels, whereas preclinical cardiac structural abnormalities, as well as older age, female sex, lower BMI, and higher creatinine level, are associated with higher NT-proBNP levels.
Article
Neutrophils are the most abundant circulating leucocytes in healthy humans. These cells are central players during acute inflammatory responses, although a growing body of evidence supports a crucial role in chronic inflammation and chemokines and cytokines related to it as well. Thus, both humoral and cellular components are involved in the development of plaque formation and atherosclerosis. Accordingly, CANTOS trial using an interleukin-1 beta antibody confirmed that inflammatory cytokines contribute to the occurrence of myocardial infarction and cardiac death independent of changes in lipids. Recent data revealed that neutrophils are a heterogeneous population with different subsets and functional characteristics (i.e. CD177+ cells, OLFM4+ neutrophils, proangiogenic neutrophils, neutrophils undergoing reverse migration, and aged neutrophils). Importantly, neutrophils are able to synthesize de novo proteins. Neutrophil extracellular trap generation and NETosis have been considered as very important weapons in sterile inflammation. Neutrophil-derived microvesicles represent another mechanism by which neutrophils amplify inflammatory processes, being found at high levels both at the site of injury and in the bloodstream. Finally, neutrophil aging can influence their functions also in relation with host age. These recent acquisitions in the field of neutrophil biology might pave the way for new therapeutic targets to prevent or even treat patients experiencing cardiovascular (CV) diseases. Here, we discuss novel findings in neutrophil biology, their impact on CV and cerebrovascular diseases, and the potential implementation of these notions into daily clinical practice.
Article
Introduction: Numerous heart failure risk scores have been developed but there is none for Asians. We aimed to develop a risk calculator, the Singapore Heart Failure Risk Score, to predict 1- and 2-year survival in Southeast Asian patients hospitalised for heart failure. Materials and methods: Consecutive patients admitted for heart failure were identified from the Singapore Cardiac Databank Heart Failure registry. The follow-up was 2 to 4 years and mortality was obtained from national registries. Results: The derivation (2008-2009) and 2 validation cohorts (2008-2009, 2013) included 1392, 729 and 804 patients, respectively. Ten variables were ultimately included in the risk model: age, prior myocardial infarction, prior stroke, atrial fibrillation, peripheral vascular disease, systolic blood pressure, QRS duration, ejection fraction and creatinine and sodium levels. In the derivation cohort, predicted 1- and 2-year survival was 79.1% and 68.1% compared to actual 1- and 2-year survival of 78.2% and 67.9%. There was good agreement between the predicted and observed mortality rates (Hosmer-Lemeshow statistic = 14.36, P = 0.073). C-statistics for 2-year mortality in the derivation and validation cohorts were 0.73 (95% CI, 0.70-0.75) and 0.68 (95% CI, 0.64-0.72), respectively. Conclusion: We provided a risk score based on readily available clinical characteristics to predict 1- and 2-year survival in Southeast Asian patients hospitalised for heart failure via a simple online risk calculator, the Singapore Heart Failure Risk Score.
Article
Background: We assessed whether plasma troponin I measured by a high-sensitivity assay (hs-TnI) is associated with incident cardiovascular disease (CVD) and mortality in a community-based sample without prior CVD. Methods: ARIC study (Atherosclerosis Risk in Communities) participants aged 54 to 74 years without baseline CVD were included in this study (n=8121). Cox proportional hazards models were constructed to determine associations between hs-TnI and incident coronary heart disease (CHD; myocardial infarction and fatal CHD), ischemic stroke, atherosclerotic CVD (CHD and stroke), heart failure hospitalization, global CVD (atherosclerotic CVD and heart failure), and all-cause mortality. The comparative association of hs-TnI and high-sensitivity troponin T with incident CVD events was also evaluated. Risk prediction models were constructed to assess prediction improvement when hs-TnI was added to traditional risk factors used in the Pooled Cohort Equation. Results: The median follow-up period was ≈15 years. Detectable hs-TnI levels were observed in 85% of the study population. In adjusted models, in comparison to low hs-TnI (lowest quintile, hs-TnI ≤1.3 ng/L), elevated hs-TnI (highest quintile, hs-TnI ≥3.8 ng/L) was associated with greater incident CHD (hazard ratio [HR], 2.20; 95% CI, 1.64-2.95), ischemic stroke (HR, 2.99; 95% CI, 2.01-4.46), atherosclerotic CVD (HR, 2.36; 95% CI, 1.86-3.00), heart failure hospitalization (HR, 4.20; 95% CI, 3.28-5.37), global CVD (HR, 3.01; 95% CI, 2.50-3.63), and all-cause mortality (HR, 1.83; 95% CI, 1.56-2.14). hs-TnI was observed to have a stronger association with incident global CVD events in white than in black individuals and a stronger association with incident CHD in women than in men. hs-TnI and high-sensitivity troponin T were only modestly correlated ( r=0.47) and were complementary in prediction of incident CVD events, with elevation of both troponins conferring the highest risk in comparison with elevation in either one alone. The addition of hsTnI to the Pooled Cohort Equation model improved risk prediction for atherosclerotic CVD, heart failure, and global CVD. Conclusions: Elevated hs-TnI is strongly associated with increased global CVD incidence in the general population independent of traditional risk factors. hs-TnI and high-sensitivity troponin T provide complementary rather than redundant information.