ArticlePDF Available

Guidelines for Experimental Models of Myocardial Ischemia and Infarction

Authors:

Abstract and Figures

Myocardial infarction is a prevalent major cardiovascular event that arises from myocardial ischemia with or without reperfusion, and basic and translational research is needed to better understand its underlying mechanisms and consequences for cardiac structure and function. Ischemia underlies a broad range of clinical scenarios ranging from angina to hibernation to permanent occlusion, and while reperfusion is mandatory for salvage from ischemic injury, reperfusion also inflicts injury on its own. In this consensus statement, we present recommendations for animal models of myocardial ischemia and infarction. With increasing awareness of the need for rigor and reproducibility in designing and performing scientific research to ensure validation of results, the goal of this review is to provide best practice information regarding myocardial ischemia-reperfusion and infarction models. Listen to this article’s corresponding podcast at ajpheart.podbean.com/e/guidelines-for-experimental-models-of-myocardial-ischemia-and-infarction/.
Content may be subject to copyright.
REVIEW Guidelines in Cardiovascular Research
Guidelines for experimental models of myocardial ischemia and infarction
XMerry L. Lindsey,
1,2
* Roberto Bolli,
3
John M. Canty, Jr.,
4
Xiao-Jun Du,
5
Nikolaos G. Frangogiannis,
6
Stefan Frantz,
7
Robert G. Gourdie,
8
Jeffrey W. Holmes,
9
XSteven P. Jones,
10
Robert A. Kloner,
11,12
David J. Lefer,
13
Ronglih Liao,
14,15
Elizabeth Murphy,
16
Peipei Ping,
17
Karin Przyklenk,
18
Fabio A. Recchia,
19,20
Lisa Schwartz Longacre,
21
Crystal M. Ripplinger,
22
Jennifer E. Van Eyk,
23
and Gerd Heusch
24
*
1
Mississippi Center for Heart Research, Department of Physiology and Biophysics, University of Mississippi Medical Center,
Jackson, Mississippi;
2
Research Service, G. V. (Sonny) Montgomery Veterans Affairs Medical Center, Jackson, Mississippi;
3
Division of Cardiovascular Medicine and Institute of Molecular Cardiology, University of Louisville, Louisville, Kentucky;
4
Division of Cardiovascular Medicine, Departments of Biomedical Engineering and Physiology and Biophysics, The Veterans
Affairs Western New York Health Care System and Clinical and Translational Science Institute, Jacobs School of Medicine
and Biomedical Sciences, University at Buffalo, Buffalo, New York;
5
Baker Heart and Diabetes Institute, Melbourne, Victoria,
Australia;
6
The Wilf Family Cardiovascular Research Institute, Department of Medicine (Cardiology), Albert Einstein College
of Medicine, Bronx, New York;
7
Department of Internal Medicine I, University Hospital, Würzburg, Germany;
8
Center for
Heart and Regenerative Medicine Research, Virginia Tech Carilion Research Institute, Roanoke, Virginia;
9
Department of
Biomedical Engineering, University of Virginia Health System, Charlottesville, Virginia;
10
Department of Medicine, Institute
of Molecular Cardiology, Diabetes and Obesity Center, University of Louisville, Louisville, Kentucky;
11
HMRI
Cardiovascular Research Institute, Huntington Medical Research Institutes, Pasadena, California;
12
Division of
Cardiovascular Medicine, Keck School of Medicine, University of Southern California, Los Angeles, California;
13
Cardiovascular Center of Excellence, Louisiana State University Health Science Center, New Orleans, Louisiana;
14
Harvard Medical School, Boston, Massachusetts;
15
Division of Genetics and Division of Cardiovascular Medicine,
Department of Medicine, Brigham and Women’s Hospital, Boston, Massachusetts;
16
Systems Biology Center, National Heart,
Lung, and Blood Institute, National Institutes of Health, Bethesda, Maryland;
17
National Institutes of Health BD2KBig Data
to Knowledge (BD2K) Center of Excellence and Department of Physiology, Medicine and Bioinformatics, University of
California, Los Angeles, California;
18
Cardiovascular Research Institute and Departments of Physiology and Emergency
Medicine, Wayne State University School of Medicine, Detroit, Michigan;
19
Institute of Life Sciences, Scuola Superiore
Sant’Anna, Fondazione G. Monasterio, Pisa, Italy;
20
Cardiovascular Research Center, Lewis Katz School of Medicine,
Temple University, Philadelphia, Pennsylvania;
21
Heart Failure and Arrhythmias Branch, Division of Cardiovascular
Sciences, National Heart, Lung, and Blood Institute, National Institutes of Health, Bethesda, Maryland;
22
Department of
Pharmacology, School of Medicine, University of California, Davis, California;
23
The Smidt Heart Institute, Department of
Medicine, Cedars Sinai Medical Center, Los Angeles, California; and
24
Institute for Pathophysiology, West German Heart
and Vascular Center, University of Essen Medical School, Essen, Germany
Submitted 14 June 2017; accepted in final form 3 January 2018
Lindsey ML, Bolli R, Canty JM Jr, Du XJ, Frangogiannis NG, Frantz S,
Gourdie RG, Holmes JW, Jones SP, Kloner RA, Lefer DJ, Liao R, Murphy E,
Ping P, Przyklenk K, Recchia FA, Schwartz Longacre L, Ripplinger CM, Van
Eyk JE, Heusch G. Guidelines for experimental models of myocardial ischemia
and infarction. Am J Physiol Heart Circ Physiol 314: H812–H838, 2018. First
published January 12, 2018; doi:10.1152/ajpheart.00335.2017.—Myocardial in-
farction is a prevalent major cardiovascular event that arises from myocardial
ischemia with or without reperfusion, and basic and translational research is needed
to better understand its underlying mechanisms and consequences for cardiac
structure and function. Ischemia underlies a broad range of clinical scenarios
ranging from angina to hibernation to permanent occlusion, and while reperfusion
is mandatory for salvage from ischemic injury, reperfusion also inflicts injury on its
own. In this consensus statement, we present recommendations for animal models
of myocardial ischemia and infarction. With increasing awareness of the need for
rigor and reproducibility in designing and performing scientific research to ensure
validation of results, the goal of this review is to provide best practice information
regarding myocardial ischemia-reperfusion and infarction models.
Listen to this article’s corresponding podcast at ajpheart.podbean.com/e/guide-
lines-for-experimental-models-of-myocardial-ischemia-and-infarction/.
Am J Physiol Heart Circ Physiol 314: H812–H838, 2018.
First published January 12, 2018; doi:10.1152/ajpheart.00335.2017.
Licensed under Creative Commons Attribution CC-BY 4.0: ©the American Physiological Society. ISSN 0363-6135. http://www.ajpheart.orgH812
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
animal models; cardiac remodeling; heart failure; myocardial infarction; reperfu-
sion; rigor and reproducibility
INTRODUCTION
Ischemia occurs when blood flow to the myocardium is
reduced (129). Ischemia of prolonged duration induces myo-
cardial infarction (MI), and MI is a common cause of heart
failure (295). Ischemic cardiomyopathy is the most common
cause of heart failure and can arise from remodeling after an
acute ST segment elevation myocardial infarction (STEMI)
from multiple small nontransmural infarctions or from chronic
repetitive ischemia in the absence of infarction (15). Ischemia
can range in its extent from low flow to total coronary occlu-
sion, can be of short to long duration, can be successfully
reversed by reperfusion in a timely manner or not reperfused at
all, and can induce injury or provide cardioprotection. Like-
wise, there is a diverse variety of animal models to address
each type of ischemia within this spectrum. Figure 1 shows the
range of models that reflect the scale of ischemia and variety of
models available to better understand how the heart responds to
ischemia and the mechanisms whereby the heart can either
adapt to ischemia or progress to failure.
Experimental models of myocardial ischemia serve two
nearly opposing aims, both worthy of investigation. The first
aim is to provide better mechanistic insight that cannot be
obtained from a clinical situation. To achieve this aim, exper-
imental studies may be reductionist with low direct applicabil-
ity to the clinical situation (e.g., when using temporally in-
duced cell specific over- or underexpression of a gene). The
second aim is to provide mechanistic insight from an experi-
mental study for translation to the clinical situation, and for this
aim experimental models must replicate the clinical setting as
closely as possible (127).
For cardiovascular science to continue advancing, experi-
mental results should be reproducible and replicable, and
rigorous experimental design is a fundamental element of
reproducibility. Reproducibility refers to results that can be
repeated by multiple scientists and is a means of validation
across laboratories. Rigor refers to robust and unbiased exper-
imental design, methodology, analysis, interpretation, and re-
porting of results. With increasing awareness by journals and
granting agencies of the need for reproducibility and rigor in
designing and performing scientific research in preclinical
studies, the goal of this consensus article is to provide best
practice information regarding myocardial ischemia and infarc-
tion models. The strengths and limitations of the different
models are discussed, with a summary shown in Table 1. We
also address ways to incorporate Animals in Research: Report-
ing In Vivo Experiments (ARRIVE) guidelines and similar
standard operating procedures (168). The extensive reference
list provided also serves as a resource for researchers new to
the field.
IN VITRO AND EX VIVO MODELS
Myocyte Cell Culture
Model rationale. Isolated fresh or cultured cardiomyocytes
can be used as a powerful in vitro model of ischemia-reperfu-
sion (I/R), whereby ischemia is simulated with hypoxia and
reperfusion with reoxygenation (H/R). This system allows
precise control of the cellular and extracellular environment,
notably the specific impact of hypoxia and reoxygenation on
cardiomyocytes without confounding influences of other cell
types (e.g., fibroblasts, endothelial cells, inflammatory/immune
cells, and platelets) or circulating factors (e.g., hormones,
neurotransmitters, and cytokines).
Variables measured. The most common use of this model
system is for in vitro testing of specific factors proposed to be
involved in I/R injury or cardioprotection (31, 166, 227, 240,
244, 330). After H/R, cultured cardiomyocytes undergo apo-
ptosis, accompanied by cytochrome crelease and caspase
activation or necrosis (200, 296). Thus, assays of cell viability
are often performed to assess the role of a particular genetic or
pharmacological intervention in exacerbating or protecting the
cell from H/R-induced cell death. Viability may be measured
with a variety of assays, including lactate dehydrogenase
(LDH) release or propidium iodide exclusion as an indicator of
membrane integrity (24, 31, 58). Apoptosis is assessed with
TUNEL or annexin V staining (24, 58, 166). Mitochondrial
damage, including disruption of mitochondrial membrane po-
tential, is also a key component of cellular injury after H/R and
may be assessed using fluorescent dyes, such as tetramethyl-
rhodamine methyl ester (TMRM). The loss of mitochondrial
membrane potential causes TMRM to leak from the mitochon-
dria, decreasing fluorescence (24, 31). In addition, reactive
oxygen species (ROS) have been implicated in H/R injury
(275); thus, ROS production is another common assessment
(24, 58, 124, 125, 227).
More detailed analyses of cardiomyocyte responses to H/R
include assessments of morphology, contractile function (i.e.,
cell shortening), intracellular Ca
2
handling, and action poten-
tials (124, 166, 207). Contractile function is an important but
often overlooked variable in H/R assays, since contraction
requires ~70% of total energy utilization within a myocyte
(182). Many H/R studies use quiescent myocytes; however,
markedly impaired recovery of myocyte function and increased
cell death result when cells are stimulated to contract through-
out the H/R protocol (207). For all variables assessed, technical
replicates on the same sample should be performed to establish
the variability of the measurement technique. Biological rep-
licates often include measurements on plates or myocytes from
the same heart/harvest. If these are to be treated as independent
samples, nvalues for both the plate/myocyte number as well as
heart/harvest number should be fully reported.
There is currently no standardized protocol for H/R in
cultured cardiomyocytes, but cardiomyocyte source and H/R
conditions must be carefully considered. Theoretically, freshly
isolated adult cardiomyocytes are the ideal gold standard for
H/R experiments (166, 207, 244), although neonatal cardiomy-
* M. L. Lindsey and G. Heusch contributed equally to this work.
Address for reprint requests and other correspondence: M. L. Lindsey, Dept.
of Physiology and Biophysics, Univ. of Mississippi Medical Center, 2500 N.
State St., Rm. G351-04, Jackson, MS 39216-4505 (e-mail: mllindsey@umc.
edu).
H813MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
ocytes (58, 227, 296), cardiac progenitor cells (12), induced
pluripotent stem cell-derived cardiomyocytes (iPSC-CMs) (24,
31), and various cell lines such as H9c2 and HL-1 have been
used (330). Of these, neonatal cardiomyocytes are most com-
monly used, due to their relative ease of isolation and robust
viability for several days in culture; however, neonatal cardi-
omyocytes are more resistant to hypoxia than adult myocytes
(236, 255), and the mechanisms of this resistance remain
incompletely resolved, which may limit interpretation of such
results when designing more translational studies (235). Adult
myocytes are preferred, because ischemic heart disease almost
exclusively occurs in adults, and fresh isolation eliminates the
potential confounding factor of phenotypic transformation in
culture. The caveat is that adult cardiomyocytes are more
difficult to isolate and do not survive long in culture, impeding
longer H/R protocols or those requiring pre-H/R transfection.
Therefore, many investigators turn to neonatal cardiomyo-
cytes or cell lines (e.g., H9c2 or HL-1) if genetic modifica-
tions are necessary, in which case these genetically engi-
neered cells may be an important complement to in vivo or
adult cardiomyocyte studies. In addition, adult myocytes do
not form monolayers as seen with neonatal cells, limiting
their use for studies on gap junctions and electrical conduc-
tivity. Thus, authors should carefully consider experimental
end points in the context of overall study design, because
either neonatal or adult cell sources will be the best choice
depending on the question posed. For both neonatal and
adult primary cells, isolation protocols must assure for a
cardiomyocyte-enriched population (i.e., free of fibroblasts
and other cell types). For neonatal cells, the differential
attachment technique is often used, whereas gravity separa-
tion is typical for adult cardiomyocytes (161, 287). Regard-
less of the isolation and enrichment method, manual or
automated cell counting, expression of cardiomyocyte-spe-
cific markers, and visualization or quantification of T-tubu-
lar structure should be performed to verify purity and
phenotype. Use of iPSC-CMs may overcome some of these
primary cell limitations, but the response of iPSC-CMs to
H/R has not yet been fully characterized and may depend on
their maturation state (256).
The most common in vitro conditions used to simulate in
vivo ischemia are anoxia (1% O
2
,5%CO
2
,94%N
2
) and
complete substrate depletion (serum-free, glucose-free me-
dium). There are many variations to the protocol with addi-
tional conditions that more closely mimic the ischemic heart,
such as partial hypoxia, partial or no substrate depletion,
hyperkalemia, acidosis, and use of electrical stimulation (207).
The cell type needs to be carefully considered when deter-
mining optimal control and ischemic conditions. Neonatal
cardiomyocytes and cell lines (H9c2 and HL-1) favor glu-
Ischemia
Model
Spectrum
Ex vivo
isolated
perfused
heart
In vitro
cardio-
myocytes
Angina
Stunning,
hibernation,
ischemic
cardio-
myopathy
Permanent
Occlusion
MI
Reperfused
MI
Ablation
Cardioprotection
Mouse Permanent Occlusion MI
7 days
Matrix cross-linking
Fibroblast apoptosis
Vascular maturation
Chemokine suppression
Fibroblast tissue deposition
Angiogenesis
Chemokine induction
Leukocyte infiltration
Proliferative phase
2-7 days
Inflammatory phase
3h-72h
Maturation phase
7-21 days
Pig 60 min ischemia + 3 h reperfusion
LV Wound Healing
Phases
4 days 4 wks
Mouse 60 min ischemia + reperfusion
40 min 3 h >6 h
Area
at risk
Rat 45 min ischemia + 2 h reperfusion
Mouse 48 h
after
cyroinjury
Wavefront progression of MI
Area
at
risk
Cardiac
output
Afterload
Preload
Preload
Pump
J
G
K
D
I
H
D
EF
EF
B
A
Aortic
pressure
Electrical
pacing
Pressure
volume
Inflammation
Remodeling
Scar formation
Area at risk Infarction No-reflow
Infarct size
Interventions
Preload
Fig. 1. The spectrum of ischemia encapsulates in vitro, ex vivo, and in vivo models of ischemia that range from transient to prolonged in duration with acute
to chronic consequences. The pig section (bottom left) is modified from Heusch (126). MI, myocardial infarction; LV, left ventricular;
H814 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
cose metabolism and under control conditions are often
cultured in hyperglycemic medium, which is known to
induce ROS production and cell death in adult cardiomyo-
cytes. Furthermore, neonatal cardiomyocytes are insulin
resistant and supraphysiological concentrations of insulin
are required to increase glucose uptake in these cells com-
pared with adult cardiomyocytes (201). Thus, studies focus-
ing on metabolism with H/R and/or metabolic pathology
need to carefully consider the cellular environment in both
control and ischemic conditions. In addition to the cellular
environment, the duration of hypoxia and reoxygenation is
also an important consideration, as myocyte viability not
only depends on the duration of hypoxia but also the
duration of reoxygenation (166, 244).
To summarize, the major strengths and limitations of studies
using isolated cardiomyocytes are shown in Table 1. The major
strength of H/R experiments in cultured cardiomyocytes is the
ability to control precisely the cellular and extracellular envi-
ronment, each factor present in ischemic conditions (e.g.,
hypoxia, metabolic inhibition, or acidosis) can be tested alone
and in combination to determine individual contributions to
cellular injury. Even with the most carefully designed experi-
mental protocol, in vitro conditions can never fully recapitulate
the full spectrum of I/R injury in vivo. Thus, although in vitro
experiments can be mechanistically informative and identify
new targets for intervention, it is imperative that the results are
later validated in an appropriate intact animal model. Nonethe-
less, even if there is a discrepancy between in vivo I/R and in
vitro H/R experiments, important insight is to be gained from
parallel studies. As an example, angiopoietin-like protein 4
(ANGPTL4) reduces infarct size after I/R in vivo but does not
prevent cardiomyocyte death in vitro (90). These findings
indicate that other cell types (e.g., endothelial cells, fibroblasts,
or immune cells) are key to the cardioprotective effect of
ANGPTL4. Likewise, factors that prevent myocyte cell death
in culture may not reduce infarct size in vivo, suggesting that
the cardioprotective effect may be outweighed by noncar-
diomyocyte factors.
Isolated Perfused Hearts
Model rationale. The isolated perfused heart is a conve-
nient and reproducible model to test mechanisms of myo-
cardial injury and cardioprotection (14, 192). The heart is
Table 1. Comparison of different approaches with strengths and limitations for each method
Approach End-Point Measurements Strengths Limitations and Pitfalls
In vitro
cardiomyocytes
Cell viability (live-dead assay) High throughput Reductionist
Type of cell death (i.e., apoptosis) Isolate effects of hypoxia/reoxygenation on
cardiomyocytes without other cell types or
circulating factors
Cardiomyocyte viability may not predict
changes in infarct size in vivo
Adult cardiomyocyte culture technically
challenging
Isolated perfused
hearts
Infarct size per area at risk Relatively easy and reproducible Tissue edema
Left ventricular function Can study ischemia and reperfusion May not fully represent the in vivo response
Assessment of cardiac troponin I as a
secondary cardiac injury index
Accurate measure of infarct size Glucose as the sole substrate
Ample sample for biochemistry Limited stability
Compatible with NMR studies Excessive coronary flow
Capacity for high throughput Reductionist
Neurohormonal factor independent
Eliminates confounding effect of intervention on
systemic blood vessels or circulating factors
Angina Regional flow and function Close to clinical situation Technically complex; time and cost
intensiveMetabolism, morphology, molecular
biology, nerve activity, rhythm
Hibernation/
stunning
Regional flow and function Close to clinical situation Technically complex; time and cost
intensiveMetabolism, morphology, molecular
biology, rhythm
Permanent
occlusion MI
Inflammation, wound healing, scar
formation, remote region myocytes,
electrical activity
In the era of percutaneous coronary intervention,
~15–25% patients are not successfully
reperfused in a timely manner (53, 104)
Does not reflect the reperfused MI
patient response
Robust remodeling response; large effect size
Ischemia-
reperfusion MI
Inflammation, wound healing, scar
formation, myocyte viability, electrical
activity
Close to clinical scenario More technically challenging surgery
Reperfusion injury can expand area of
damage
Ablation Inflammation, wound healing, scar
formation, myocyte electrical activity
Geometrically defined lesion Nonischemic lethal injury
Infarct size/location independent of coronary
anatomy
Mechanisms of cell death different from
ischemia
Cardioprotection Infarct size per area at risk
Left ventricular geometry and function; no
reflow; circulating biomarkers such as
cardiac troponin I
Mouse, rat, rabbit, and pig: models of low
collateral flow (measurement of regional
myocardial blood flow not required)
Rat and rabbit: reliable infarct production;
relatively high survival rate
Pig: mimics humans with low collateral flow
Dog: large amount of historical data; shows
effect of intervention in the setting of variable
collateral flow; mimics humans with high
collateral flow
Mouse: small size; substantial variability
requiring large nvalues
Rat, rabbit, pig, and dog: not high
throughput; higher cost
Rabbit, pig, and dog: potential for lethal
arrhythmias
Dog: variability in collateral perfusion
(regional myocardial blood flow must
be measured)
MI, myocardial infarction.
H815MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
removed from the animal and perfused, typically with a
physiological saline solution such as Krebs-Henseleit buf-
fer. For screening drugs or interventions for protective
properties, this model is ideal, because the isolated perfused
heart is studied independently of circulating factors or
neuroendocrine inputs from other organs but retains the
function, composition, and architecture of the intact heart.
This approach is also easily amenable to biochemistry or
imaging studies in a nuclear magnetic resonance (NMR)
magnet, which can provide useful information to decipher
mechanisms of cardioprotection.
Perfused hearts can be studied in a working heart mode or in
a nonworking Langendorff mode. In the Langendorff mode,
the perfusate enters the coronary arteries to perfuse and oxy-
genate the heart, which continues to beat for several hours
(292). Heart rate and left ventricular (LV) developed pressure
are measured with a fluid-filled balloon placed in the cavity of
the LV and connected to a pressure transducer as indexes of
cardiac function (physiology). The heart can be perfused at con-
stant pressure, in which case the flow rate can vary, or flow can be
set with a pump, in which case the perfusion pressure can vary. In
the Langendorff mode, the heart does not pump against a gradient
and does not perform external work (226). In the working heart
mode, the perfusate enters the atrium at a filling pressure set by the
operator, and the heart pumps perfusate against a hydrostatic
pressure set to different levels (30). A heart model performing
external work is technically more challenging, particularly with
smaller hearts.
In a model of global ischemia, perfusate flow to the entire
heart is stopped, whereas in a regional ischemic model, a suture
is tied around a single coronary artery for occlusion. After the
ischemic period (typically 20 40 min for rodent models),
perfusion is restarted and the heart will usually beat and
develop a lower LV developed pressure than at baseline,
reflecting postischemic contractile dysfunction or stunning.
Contractile dysfunction is a measure of ischemic injury but is
not synonymous with cell death. Both contractile dysfunction
and cell death often result from the same mechanisms. It is
therefore important not to infer that protection against contrac-
tile dysfunction is the same as protection against infarct size.
Variables measured. To measure cell death or infarct size, it
is necessary to reperfuse the heart for a sufficient duration (at
least 60 –120 min) to wash out reductive equivalents (79, 271).
Triphenyltetrazolium chloride (TTC) is then added to the
perfusate or hearts are cut into transverse slices and incubated
in TTC solution. TTC is a dye that stains viable myocardium
red due to a formazan reaction with NADH and NADPH,
which are washed out from irreversibly injured myocardium
(81, 172), whereas necrotic tissue remains unstained and thus
appears white. Necrotic tissue area is normalized to the total
ventricular area (global ischemia) or the ischemic area at risk
for infarction (regional ischemia). For regional ischemia prep-
arations, the area at risk is measured after coronary reocclusion
and staining of the nonischemic myocardium with a dye such
as Evans blue.
The susceptibility of the heart to arrhythmias during I/R is
readily assessed through recording of an electrocardiogram
(313). Updated guidelines exist for the quantification of such
arrhythmias (56). The isolated heart is amenable to monitoring
of intracellular ions by optical methods using fluorescent indi-
cators (where the signal originates from a thin layer of epicar-
dial cells) or by NMR spectroscopy (where the signal is a
global average from the whole heart). Intracellular Na
and
Ca
2
concentrations have been monitored and intracellular H
concentration (i.e., intracellular pH) has been estimated in
isolated rodent hearts perfused and subjected to I/R within the
vertical bore of the NMR magnet (288). Intracellular high-
energy phosphate (ATP and creatine phosphate) have also been
monitored by this method, and recent developments using
hyperpolarized substrates now also allow real-time analysis of
metabolic flux through distinct pathways (167).
The rate of occurrence of cell death is determined by the
work that the heart performs at the time it becomes ischemic.
When global flow is stopped completely, the heart will con-
tinue to beat for a short period of time and continue to consume
ATP. Reducing work at the start of ischemia is cardioprotec-
tive, and this is the basis of cardioplegic solutions. Therefore,
it is important to assure that work is similar between control
and experimental hearts, which means that heart rate and
temperature must be controlled and held constant in the dif-
ferent treatment groups (281, 292). Due to the lack of neuro-
humoral factor influences on the perfused heart, heart rate is
typically lower than in an intact animal, and it can be con-
trolled by pacing. A slight (1°C) difference in temperature
can result in a large difference of infarct size. Temperature is
usually measured by a probe in the heart and controlled by
immersing the heart in a fluid bath.
In the absence of blood, the reduction in the oxygen-carrying
ability of the saline perfusate results in edema and an increase
in flow rate. Because the mouse has a high heart rate, it is likely
that oxygen delivery is on the edge of oxygen demand under
baseline perfusion conditions. Furthermore, Krebs-Henseleit
buffer typically contains only glucose as a substrate, whereas
the heart normally uses fatty acids as its prime substrate. Fatty
acids can be given as substrates but require the addition of a
vehicle such as BSA and also require specialized methods for
oxygenating the buffer. An advantage of the perfused heart
model is that it allows one to examine the impact of different
substrates either alone or in combinations. The effect of dif-
ferences associated with perfusion with long-chain versus
short-chain fatty acids also can be studied. Ex vivo hearts can
be readily perfused with radioactive- or stable isotope-labeled
substrates, allowing evaluation of substrate selection and me-
tabolism (167, 211, 245).
The no-reflow phenomenon (179) can also impact infarct
size and its measurement in isolated hearts. During ischemia
and early reflow, the heart goes into contracture, which restricts
flow to the subendocardium. The severity of no reflow depends
on the severity of ischemic injury and can vary between control
and protected hearts. In isolated hearts, no-reflow issues can be
reduced by deflating the balloon in the LV for a few minutes
right at the start of reperfusion (94).
To summarize, the major strengths and limitations of the
isolated perfused heart are shown in Table 1. In consideration
for all of the above factors, it is imperative that control and
treated hearts are studied under identical conditions. Although
ischemic pre- and postconditioning were originally described
in an in vivo dog model, much of the information about the
molecular signaling pathways responsible for protection was
established in perfused heart models (128).
H816 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
IN VIVO MODELS
Chronic Coronary Artery Disease
Coronary stenosis and stress-induced myocardial ischemia:
model rationale and variables measured. Reversible episodes
of ischemia can lead to contractile dysfunction in the absence
of significant myocyte necrosis (42). Because ischemic entities
are frequently encountered in clinical practice, the task of
understanding their pathophysiology and testing therapeutic
interventions has stimulated the development of animal models
in which acute and chronic adaptations to ischemia and the
ensuing functional recovery can be evaluated over time. Most
of these entities are consequences of either brief or chronic
episodes of ischemia, and the models developed to study them
will be discussed separately below.
Chronic stable angina is a clinical condition whereby a
patient has one or more coronary stenoses that have largely
compromised or even exhausted autoregulatory coronary re-
serve. Frequently, these limitations are partially compensated
by collateral blood flow from adjacent less-compromised cor-
onary arteries such that myocardial blood flow and contractile
function remain normal at rest. Stress situations such as exer-
cise, emotions, or pain, however, can precipitate acute myo-
cardial ischemia with or without typical chest pain. Chronic
stable angina in patients does not usually inflict global myo-
cardial ischemia but is a regional event. The acute precipitation
of myocardial ischemia requires an in vivo model where an
acute coronary stenosis can be produced to reduce coronary
blood flow. Alternatively, a stable stenosis must be created
where blood flow is maintained at baseline but acute ischemia
is elicited, e.g., by pacing or adrenergic activation in anesthe-
tized animals or by exercise in conscious animals (9, 93, 131).
To reflect the regional character of chronic stable angina,
regional myocardial blood flow and regional contractile func-
tion must be measured. The standard approach to monitor
regional blood flow is to use microspheres (142), which have
traditionally been labeled with radioactive isotopes (66) and,
more recently, nonradioactive colored dyes or fluorescent dyes
(107, 184). Analysis of regional myocardial blood flow during
acute ischemia reveals an inability of perfusion to increase
distal to a stenosis compared with normal remote myocardium
(33, 39). As coronary vasodilator reserve is exhausted, there is
a major redistribution of blood flow away from the ischemic
region toward the nonischemic myocardium where metabolic
vasodilation prevails. In addition, there is a transmural blood
flow redistribution from the ischemic subendocardium toward
the subepicardium (93).
The gold standard for experimental regional contractile
function measurements is sonomicrometry (265). Simultane-
ous measurements of regional myocardial blood flow and
regional contractile function provide a means to determine the
quantitative relationship between regional blood flow (as a
surrogate for oxygen/energy supply) and regional contractile
function (as a surrogate for oxygen/energy demand). The
relation between regional contractile function and subendocar-
dial perfusion (flow-function relation) demonstrates close cou-
pling during steady-state ischemia at rest as well as over a wide
range of cardiac workloads (33, 35, 37, 91, 309). During
steady-state acute myocardial ischemia, there appears to be no
imbalance between supply (blood flow) and demand (func-
tion); rather, there is a matched reduction in both parameters
(129, 130). Such matching between regional blood flow and
contractile function also persists during major changes in heart
rate, when both blood flow and contractile function are nor-
malized for a single cardiac cycle (92). This matching can
persist for several hours and contribute to the maintenance of
myocardial viability and full recovery of contractile function
after eventual reperfusion (212). More specifically, the hall-
marks of short-term myocardial hibernation are a perfusion-
contraction match (259), together with metabolic signs of
adaptation to ischemia (210, 267) and the potential to recruit an
inotropic reserve in the dysfunctional myocardium (267). Also,
all pharmacological interventions to attenuate acute myocardial
ischemia (e.g., by nitrates, -blockers, Ca
2
antagonists, or
their combinations) operate along a fixed flow-function rela-
tionship (213–215). The two major mechanisms that precipi-
tate myocardial ischemia and must therefore be pharmacolog-
ically addressed are tachycardia (112, 113) and coronary va-
soconstriction (134, 273).
Studies evaluating brief total coronary occlusions can be
conducted in a variety of species. In contrast, to study coronary
artery stenosis, the animal under study must be large enough
that coronary artery instrumentation along with regional flow
and function measurements is feasible (i.e., in dogs and pigs
that can be instrumented with a hydraulic occluder on the
coronary artery). Acutely anesthetized animals have provided
insight into short-term coronary flow regulation over minutes
to hours but cannot evaluate the effects of chronic coronary
stenosis on long-term microvascular remodeling and collateral
growth over days to weeks (285). Acute experiments have the
advantage that sequential myocardial biopsies can be taken for
the analysis of metabolic and molecular analyses (101, 174,
210, 283). Microdialysis probes can be implanted to evaluate
interstitial mediators (209, 270), and the activity of the cardiac
innervation can be measured (132).
A significant experimental challenge is maintaining a fixed
degree of coronary artery narrowing throughout a study using
a hydraulic occluder. This limitation can by circumvented by
perfusing the coronary artery at constant pressure from a
reservoir, controlling flow with an extracorporeal pump or
perfusion of the region of interest with an extracorporeal
pressure control system (38). A major limitation of studies in
acutely anesthetized animals are the substantial confounding
effects of anesthesia and neurohormonal activation on hemo-
dynamics and flow, which alter coronary autoregulation and
produce varying degrees of coronary vasodilation and vaso-
constriction that modulate autoregulatory responses (33, 36,
39). The limitations of acute studies can be circumvented by
studying conscious, chronically instrumented animals.
While strict control of hydraulic occluders and coronary
collaterals stimulated by repetitive ischemia and chronic
stenosis at first seems a limitation, these factors can be
capitalized upon by provoking coronary collateral develop-
ment to the point where the artery can be totally occluded
without reducing resting myocardial perfusion. This is typ-
ically accomplished in dogs using ameroid occluders, which
are hygroscopic and gradually swell to produce a progres-
sive stenosis resulting in a total occlusion within 3– 4 wk.
Collateral blood vessel growth can also be stimulated in
dogs by repetitive brief coronary occlusions using a hydraulic
occluder (303, 320). Once collaterals are developed, variability
H817MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
in the hydraulic stenosis severity is no longer a problem, and
intervention effects on stress-induced ischemia can be studied
under multiple conditions. While pigs can also develop collat-
eral-dependent myocardium after ameroid occluder placement
(260), collaterals grow more slowly than in dogs, and pigs
frequently develop a subendocardial infarction (230). The
admixture of infarcted and normal myocardium greatly com-
plicates measurements of perfusion and function. Infarction
can largely be circumvented by employing a fixed diameter
stenosis on the coronary artery of farm-bred swine, resulting in
a much more severe limitation of subendocardial flow reserve
than in dogs, and there is usually contractile dysfunction at rest
(74, 77). While this is an extremely useful model to study
chronic vascular adaptations and interventions to promote
angiogenesis, alterations in myocardial physiology can com-
plicate the interpretation of flow changes. A major drawback of
chronic large animal models of coronary stenosis and collat-
eral-dependent myocardium is their expense and the labor-
intensive nature of the animal care and handling. The guinea
pig has a very well-developed collateral circulation that pre-
vents infarction from occurring following occlusion of a single
main coronary artery; to obtain infarction, multiple coronary
arteries need to be ligated (216). Thus, guinea pigs are not
suitable for in vivo ligation studies but can be used for heart
perfusion with global ischemia experiments. These issues have
motivated studies to assess flow regulation using repetitive
coronary occlusions in rats and mice, including genetically
altered animals (303).
Coronary microembolization: model rationale and variables
measured. Subclinical atherosclerotic plaque rupture or erosion
that does not result in complete thrombotic occlusion of the
coronary artery but leaves a residual blood flow into the distal
coronary microcirculation occurs spontaneously, with or with-
out clinical symptoms. Coronary microembolization is also
induced iatrogenically during percutaneous coronary interven-
tions. Atherosclerotic debris from the culprit lesion, together
with thrombotic material, soluble vasoconstrictors, as well as
thrombogenic and inflammatory substances, is washed into the
coronary microcirculation where it causes microvascular ob-
struction with resulting patchy microinfarcts and an inflamma-
tory reaction (135, 173). The inflammatory response includes
increased expression of tumor necrosis factor-associated
with profound contractile dysfunction and upregulation of
signal transduction pathways involving nitric oxide, sphin-
gosine, and ROS, which contribute to impaired excitation-
contraction coupling (32, 299). Repetitive coronary microem-
bolization can result in global LV dysfunction and, even in the
absence of overt infarction, in heart failure (261). Coronary
microembolization can be simulated experimentally by intra-
coronary infusion of inert particles of various diameter (67)
and also by intracoronary infusion of autologous microthrombi
(191). When the target under study is ischemic heart failure,
repeated coronary microembolization can be used in both small
and large animal models. When the target under study is a
spontaneous or periprocedural minor infarction, the animal
must be large enough such that regional myocardial measure-
ments of flow, contractile function, metabolism, and morphol-
ogy are possible (i.e., dog or pig models are preferable).
The major strengths and limitations of angina models are
shown in Table 1.
Stunning, Hibernation, and Ischemic Cardiomyopathy
Stunning: model rationale and variables measured. When
ischemia caused by a total coronary occlusion is brief (e.g., as
may be experienced from coronary vasospasm), regional con-
tractile dysfunction persists for hours after reperfusion but then
completely normalizes within 24 h. This phenomenon was first
demonstrated after a 15-min circumflex coronary artery occlu-
sion in chronically instrumented dogs, was subsequently called
stunned myocardium, and is common in patients with acute
coronary syndrome (17, 19, 143). Most investigators assume
that the complete normalization of function, lack of evidence
of infarction by TTC staining, and lack of sarcolemmal dis-
ruption on electron microscopy indicate that no cardiomyocyte
death is associated with stunning. While pathological evidence
of myocyte necrosis is indeed absent, TUNEL staining per-
formed 1 h after reperfusion demonstrates that programmed
cell death or myocyte apoptosis develops in rare isolated
cardiac myocytes and circulating cardiac troponin I is in-
creased (314). Thus, while there is no evidence of infarction in
stunned myocardium, regional myocyte loss can develop when
stunning becomes repetitive.
Because the essence of stunned myocardium consists of
relatively rapid (24 48 h) reversibility of contractile dysfunc-
tion in the absence of TTC or pathological evidence of infarc-
tion, most studies use chronic large animal models in which
serial measurements of function can be performed. In addition,
regional ischemia is the preferred model to allow assessment of
the remote nonischemic regions of the heart as an internal
control. While stunned myocardium occurs after demand-
induced ischemia distal to a coronary stenosis (144), most
studies have used transient total coronary occlusion. Animals
are instrumented with a hydraulic occluder to produce brief
ischemia 1–2 wk after recovery from surgical instrumentation.
To assess regional function, most studies have used sono-
micrometry for direct measurements of subendocardial seg-
ment length shortening or wall thickening. Recent studies have
used transient occlusion of the left anterior descending coro-
nary artery (LAD) using a balloon angioplasty catheter in
closed-chest sedated animals where regional function can be
assessed with imaging approaches such as echocardiography
(314). The latter approach circumvents the need for chronic
surgical instrumentation through a prior thoracotomy. Echo-
cardiography can also be employed to assess stunning in mice
chronically instrumented with an occluder to produce transient
ischemia (63).
The dog (20, 143), pig (291, 314), and rabbit (18) are the
most commonly used species to study myocardial hibernation.
Pigs and rabbits offer the advantage of having little or no
collateral circulation, so that the severity of the ischemic insult
and of the subsequent contractile dysfunction are more uni-
form. In contrast, dogs exhibit a highly variable degree of
collateral circulation resulting in widely different degrees of
myocardial stunning (20). There are also species differences in
the time course of recovery despite similar occlusion durations
(277). Many studies have also used open-chest animal models,
although the severity of myocardial stunning in these prepara-
tions is significantly exacerbated versus conscious animal mod-
els (21, 304). Most experimental approaches to assess stunning
are quite straightforward, although ventricular fibrillation can
develop. This is more common in swine as opposed to canine
H818 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
models, because pigs have little innate coronary collateral flow
(164). Because myocardial function assessed using segment
shortening and wall thickening is load dependent, it is impor-
tant to ensure that heart rate, systolic blood pressure, and LV
end-diastolic pressure remain reasonably constant over the
time frame of the measurements. An advantage of chronic
models using regional ischemia is that each animal can poten-
tially serve as its own control; hence, it is possible to use the
same animals to study the effects of pharmacological interven-
tions on physiological end points.
Short-term hibernation: model rationale and variables
measured. A prolonged episode of moderately severe ischemia
can be sustained for a period of hours in the absence of
pathological evidence of infarction. This phenomenon is
termed short-term hibernation (139, 212). An approximate
50% reduction in perfusion leads to reduced function and
perfusion-contraction matching, which largely prevents irre-
versible myocyte injury. If the heart is reperfused within a few
hours, contractile dysfunction persists in a fashion similar to
stunned myocardium but with a more protracted time course of
recovery (i.e., lasting in the timeframe of days rather than
hours) as is typically seen with stunning after a brief total
occlusion. This in part appears to relate to reversible myofi-
brillar disassembly and myolysis in the absence of sarcolem-
mal disruption (279). Unfortunately, when the initial adaptive
response to moderate ischemia in short-term hibernation is
present for longer than 12 h, progressive myocardial necrosis
begins to develop, resulting in some degree of myocardial
infarction usually confined to the subendocardium (48, 185,
269). While imposition of acute moderate ischemia was ini-
tially proposed as a mechanism of chronic hibernating myo-
cardium, the development of progressive infarction when flow
reductions last longer than 12 h leads to a pathological entity
with myofibrillar disassembly and cardiac biomarker release
that can no longer be defined as hibernation but, rather, is more
in line with subendocardial infarction (48, 279).
Studies of short-term hibernation usually require closed-
chest animal models, although considerable insight about ad-
justments between flow and function has been gleaned from
open-chest studies of myocardial metabolism (137, 210, 239).
The latter studies usually use a cannulated branch of the left
coronary artery perfused at constant pressure. Closed-chest
animal studies usually employ chronically instrumented dogs
or pigs. In these studies, a hydraulic occluder is placed around
a coronary to reduce flow or coronary pressure to a fixed level,
which is released after several hours.
Chronic hibernation and stunning: model rationale and
variables measured. While both stunning and short-term hi-
bernation are characterized by complete functional recovery,
chronic contractile dysfunction can develop when recurrent
ischemia develops before functional normalization (21).
Chronic contractile dysfunction from repetitive ischemia de-
velops in the absence of histological infarction, and both
hibernation and stunning involve the loss of myocytes via
apoptosis in a fashion similar to what happens after brief
episodes of ischemia (193, 314). Unlike stunning, which was
initially an experimental observation that later became associ-
ated with multiple clinical correlates, chronic hibernating myo-
cardium was first characterized in patients with chronic isch-
emic heart disease displaying regional contractile dysfunction
in the absence of manifest ischemia (23, 139). Only later were
animal models used to identify cellular and molecular mecha-
nisms responsible for the adaptive responses to chronic isch-
emia (77).
While it was originally controversial whether or not flow
was reduced or normal at rest (34), it is now clear that chronic
repetitive ischemia initially results in contractile dysfunction
with normal resting flow or chronic stunning (73, 278). When
this situation persists, the reduction in function leads to a
secondary reduction in regional energy utilization accompa-
nied by reduction in resting flow (76). Thus, the reduction in
resting flow characteristic of chronic hibernating myocardium
is a result, rather than cause, of regional dysfunction.
In contrast to models of short-term ischemia, animal models
of hibernating myocardium are based on chronic coronary
stenoses that frequently progress to total occlusion and collat-
eral-dependent myocardium. In studies using ameroid occlud-
ers that gradually swell to produce chronic stenosis, dogs
usually do not develop contractile dysfunction at rest but can
do so when preexisting epicardial collaterals are ligated at the
time of instrumentation (36). Swine ameroid occluder models
frequently have contractile dysfunction in collateral-dependent
myocardium, and this is usually associated with some degree
of subendocardial infarction (230).
A more consistent model of hibernating myocardium can be
produced by instrumenting juvenile swine with a fixed diam-
eter stenosis (1.5-mm diameter) on the proximal LAD (73, 77).
As the animals grow over the subsequent 3 mo, there is a
slowly progressive limitation in coronary flow reserve, because
the LAD stenosis limits maximal myocardial perfusion, while
the mass of myocardium distal to the stenosis increases in
parallel with cardiac growth. As a result, there is a more
prolonged and gradual stimulus for coronary collateral devel-
opment so that LAD occlusion almost always develops in the
absence of infarction.
After 3 mo, regional contractile dysfunction with mild re-
ductions of resting flow in the absence of infarction is consis-
tently manifest and is similar to the changes seen in humans
with hibernating myocardium caused by a chronic LAD occlu-
sion (308). Serial studies of this animal model have demon-
strated that the heart progressively adapts from a state of
contractile dysfunction with normal resting flow (chronic stun-
ning) to a state where resting flow decreases, consistent with
hibernating myocardium (41). Such chronic hibernation is
associated with a downregulation in mitochondrial metabolism
and regional myocyte hypertrophy that maintains myocardial
wall thickness constant in the setting of regional apoptosis-
induced myocyte loss.
Over longer periods of time (up to 6 mo) the adaptive
response of hibernating myocardium persists unchanged (75),
and the downregulation in metabolism and upregulation of
proteins involved in cellular survival and cytoprotection pre-
vent cell death and, hence, further myocyte loss (62). While
infarction does not develop in this model, revascularization
only partially reverses myocardial dysfunction and does so
over a much longer time frame than seen with either myocar-
dial stunning or short-term hibernation (237). Chronic contrac-
tile dysfunction in the absence of infarction can also be induced
using a hydraulic occluder to produce an acute stenosis in
chronically instrumented animals. Chronic stunning can de-
velop in swine subjected to daily episodes of short-term hiber-
nation (169). A more rapid transition from chronic stunning to
H819MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
hibernating myocardium than the one observed in the fixed
diameter stenosis model can be achieved by acutely imposing
a critical stenosis on the LAD (301). The latter model can
develop reductions in flow with regional contractile dysfunc-
tion after 2 wk of a stenosis sufficient to prevent reactive
hyperemia.
The fixed diameter chronic stenosis model is advantageous
in that hibernating myocardium develops reproducibly in a
predictable time frame. A limitation of the fixed stenosis
porcine model is that it requires cardiac growth to produce
a progressive physiological impairment in maximum myocar-
dial perfusion, and the 3- or 4-mo period required to develop
hibernating myocardium is viewed as cost prohibitive. This
model has so far only been studied in juvenile farm bred swine
and may produce variable results if cardiac growth is attenu-
ated by limiting feeding. It is not clear whether the model can
be effected in purpose-bred swine and, particularly, in mini-
swine, where growth rates are substantially attenuated. An
additional disadvantage is that the chronic stenosis model is
associated with a high rate of spontaneous ventricular fibrilla-
tion (40). This has provided insight into the mechanisms of
sudden cardiac arrest in chronic coronary disease but reduced
the success of studying chronic adaptations to ischemia in
survivors. Finally, because of the long duration of the studies
in the presence of animal growth, it is not feasible to chroni-
cally instrument animals. Nevertheless, it is feasible to use
telemetry to assess chronically LV pressure and arrhythmias in
untethered conscious animals (242).
Ischemic cardiomyopathy: model rationale and variables
measured. Ischemic cardiomyopathy is the underlying cause of
LV dysfunction in two out of every three patients with heart
failure (105). Ischemic cardiomyopathy in humans can arise
from LV remodeling after a large myocardial infarction but,
more commonly, is the result of extensive multivessel coronary
artery disease with modest amounts of diffuse fibrosis and
patchy infarction in multiple coronary artery distributions (15).
Along these lines, preclinical studies have established that
chronic coronary artery stenosis can induce significant myo-
cyte loss with modest global replacement fibrosis that leads to
global LV dysfunction and varying degrees of congestive heart
failure when the area at risk is large. Conceptually, the stenosis
does not limit blood flow at rest. Rather, by reducing maximal
perfusion in response to stress, it sets the stage for repetitive
episodes of subendocardial ischemia. A key feature of all
animal models of ischemic cardiomyopathy is that the myo-
cardium at risk of repetitive ischemia represents a large portion
of the LV (70% of LV mass). This has been achieved using
stenosis of the left main coronary artery in rodents or multi-
vessel coronary artery stenoses in large animals. As a result of
the large area at risk, myocyte cell death arises from both
ischemia and myocyte stretch and slippage from increased LV
end-diastolic pressure (possibly also reflecting transient isch-
emia).
In rats, ischemic cardiomyopathy can be induced by produc-
ing a fixed coronary stenosis of ~5060% diameter reduction
on the proximal left coronary artery, which causes variable
degrees of LV dysfunction (44, 45). While replacement fibrosis
occurs in these animals, it is patchy and modest, only increas-
ing twofold over control for an average of 10% of LV
cross-sectional area. Interestingly, the degree of LV dysfunc-
tion is primarily related to myocyte cell loss (necrosis and
apoptosis) and the elevation in LV end-diastolic pressure
related to fibrosis. A similar model of ischemic cardiomyopa-
thy has also been obtained in mice (189). While rodent models
afford the ability to perform higher throughput studies and use
transgenic animals to study molecular mechanisms, they have
relatively high surgical and postoperative mortality. In addi-
tion, there is considerable variability in physiological out-
comes, such that frequently animals are retrospectively cate-
gorized into mild, moderate, and severe heart failure groups.
Reproducibility of ischemic cardiomyopathy models, there-
fore, is indeed a concern.
While left main coronary stenosis is not feasible in large
animals, multivessel coronary stenoses can produce a large
ischemic risk area and recapitulate ischemic cardiomyopathy.
When fixed diameter occluders are placed on both the proximal
LAD and circumflex arteries in growing farm-bred swine, LV
ejection fraction declines with elevated resting LV end-dia-
stolic pressure (74), consistent with compensated LV dysfunc-
tion and no overt evidence of heart failure. These animals also
exhibit primary myocyte loss with only an approximately
twofold increase in extracellular matrix accumulation. A sim-
ilar condition has been produced using multivessel ameroid
occluders in dogs (80). Aside from requiring survival surgery,
the major disadvantage of these approaches arises from the
development of sudden cardiac arrest, which in swine is related
to both ventricular fibrillation and to a lesser extent bradyar-
rhythmias. In mice, a state of ischemic cardiomyopathy can be
induced using repetitive brief coronary occlusions, and this
model is associated with substantial but reversible fibrosis of
the myocardial region subjected to repetitive ischemia (63).
Noninvasive imaging. Noninvasive cardiac imaging technol-
ogies such as echocardiography, magnetic resonance imaging
(MRI), and computed tomography can measure regional and
global contractile function and are increasingly available for
preclinical studies, particularly in larger animals. NMR spec-
troscopy can provide information on cardiac energetics (114).
More sophisticated imaging technologies such as positron
emission tomography can measure regional myocardial perfu-
sion and regional myocardial metabolism and sympathetic
innervation and are increasingly used in preclinical studies (77,
187, 268). MRI can serially measure myocardial perfusion
(264) and can provide reliable measurements of infarct size and
microvascular obstruction. MRI-derived edema, however, is
time dependent and sensitive to cardioprotective interventions
(141, 148). Therefore, MRI-derived edema can be used to
stratify an ischemic/reperfused myocardial region for protocol
assignment but not for quantitative normalization of infarct
size to area at risk.
To summarize, the major strengths and limitations of stun-
ning, hibernation, and ischemic cardiomyopathy models are
shown in Table 1.
Myocardial Infarction Models: Permanent Coronary Artery
Occlusion with Nonreperfused and Reperfused Myocardial
Infarction
MI: general considerations. Coronary occlusion causes im-
mediate cessation of aerobic metabolism in the ischemic myo-
cardium, leading to rapid ATP depletion and metabolite accu-
mulation and resulting in severe systolic dysfunction within
seconds (86). If the duration of the ischemic insult is 15 min
H820 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
in larger mammals such as dog and pig, restoration of flow
reverses the early ischemic cardiomyocyte changes (transient
mitochondrial swelling or glycogen depletion) and all cardio-
myocytes in the ischemic area can survive (158). Longer
periods of ischemia cause death of an increasing number of
cardiomyocytes. A 20- to 30-min interval of severe ischemia is
sufficient to induce irreversible changes in some cardiomyo-
cytes of the subendocardial area, inducing sarcolemmal disrup-
tion and striking perturbations in mitochondrial architecture,
such as ultrastructural evidence of amorphous matrix densities
and severe mitochondrial swelling (156). These early ultra-
structural alterations mark cardiomyocytes that cannot be sal-
vaged and will ultimately die in the infarct environment (157).
Experimental studies in the canine model of MI demonstrate
a transmural heterogeneity in the myocardial response to isch-
emia, suggesting that the subendocardium, where myocardial
oxygen demand is greatest, is more susceptible to ischemic
injury than the midmyocardium and subepicardium (2). Thus,
the prevailing paradigm suggests a wavefront of cardiomyo-
cyte death that progresses from the more susceptible subendo-
cardium to the less vulnerable subepicardium as the duration of
the ischemic insult increases (159, 252). Experimental studies
in large animal models have demonstrated that ischemic myo-
cardium cannot be salvaged by reperfusion after6hofcoro-
nary occlusion (251). The increased vulnerability of subendo-
cardial regions to coronary occlusion may reflect a greater
reduction of the subendocardial blood flow due to transmural
differences in vascularization (2, 25) and extravascular com-
pression (68, 286). The wavefront concept of ischemia devel-
oping into infarction was derived from experimental studies in
dogs, where a substantial coronary collateral circulation influ-
ences the time course of cardiomyocyte necrosis (86).
The major species difference in the MI response lies in the
temporal and spatial kinetics of events and differences due to
myocardial size. In mice, durations of coronary occlusion
exceeding 60 –90 min are considered irreversible, and inflam-
mation and wound healing processes are accelerated (64, 88,
221, 222). In mouse and rat models, reperfused infarcts are
typically midmyocardial, and subepicardial and subendocardial
regions are relatively spared (50, 69, 325). Studies in a sheep
model of reperfused infarction also suggest that the midmyo-
cardium may be most vulnerable to ischemic injury; in con-
trast, the subendocardium is relatively resistant (263). The pig
model of coronary occlusion-reperfusion comes closest to
human STEMI in its temporal and spatial development, but
other models are nevertheless useful to study fundamental
mechanisms of MI (140).
MI: technical considerations. Extensive protocols providing
technical details for performing permanent occlusion MI and
reperfused MI in mice and rats are available (221, 222, 228,
317, 327). While MI is most commonly performed in rodent
models, protocols in other animal models are also available
(151, 183, 218, 331). For mice, the quality of open-chest
surgery to induce coronary occlusion directly impacts study
outcomes (152, 221, 222). Minimizing the size of the thora-
cotomy and limiting bleeding by entering the thorax through
intercostal muscles are recommended.
Biomarkers that have been used to evaluate the presence of
MI include cardiac troponins and creatine kinase, and plasma
proteins such as macrophage migration inhibitory factor can
also be measured as indices of injury (47, 55). Infarct size is
widely measured as a key variable for testing genetic or
therapeutic intervention efficacy, and infarct size measure-
ments taken serially at both early and late time points can
evaluate the extent of infarct expansion (22). Echocardiogra-
phy can also be used for infarct sizing, with the caveat that
echocardiography does not distinguish between reversible isch-
emic dysfunction (stunning) and irreversible loss of function
and therefore a secondary method is needed for confirmation of
infarct size at early time points. For more details on measuring
cardiac function in mice, the reader is advised to consult the
article Guidelines for measuring cardiac physiology in mice
(196).
Permanent occlusion MI: model rationale and variables
measured. Permanent coronary occlusion is a relevant animal
model of acute STEMI reflective of patients who, due to
contraindications or logistic issues, do not receive timely or
successful reperfusion (53, 104). Permanent coronary occlu-
sion yields acute ST segment elevation infarction with robust
myocardial inflammation and long-term remodeling, thus pro-
viding a large effect size that reduces the sample size needed to
detect differences between groups. Infarction assessed in the
first 1–14 days after coronary ligation is histologically charac-
terized by coagulation band necrosis with a fulminant inflam-
matory infiltrate in the infarct and border zone regions. Infarc-
tion is geometrically and physiologically characterized by wall
thinning, increases in LV dimensions and volumes, and de-
creases in fractional shortening and ejection fraction.
Changes that occur over the first week provide information
on myocyte cell death and infarct development, inflammation
and leukocyte physiology, extracellular matrix (ECM) turnover
and fibroblast activation, and the role of endothelial cells in
neovascularization (83, 154, 165, 194, 205). Chronic evalua-
tion at time points 4 8 wk post-MI provides information on
long-term remodeling. Whether the infarct region or remote
region is the focus of investigation depends on the question
asked. Examining the infarct region provides details on active
inflammation and scar formation, while examining the remote
region provides details on still-viable myocytes within the
myocardium and remote inflammatory and ECM processes.
Perioperative and postoperative survival should be assessed,
and the time point of delineation between these two phases
should be defined. For some laboratories, the perioperative
phase includes the time until the animal recovers and becomes
ambulatory (usually within 1–3 h for mice). For other labora-
tories, the perioperative phase includes the first 24 h after
surgery. Perioperative death within 24 h post-MI in mice is
usually due to surgical errors (or very large infarct sizes), and,
in established laboratories, the 24 h surgical mortality rate due
to technical issues is 10%. In the permanent occlusion MI
model in mice, postoperative death (deaths at 24 h time
point) typically occurs during days 3–7 post-MI and is due to
rupture, acute heart failure, or arrhythmias (59, 98, 233).
Autopsy is strongly recommended for all mice that die prema-
turely, to evaluate early deaths due to technical issues and later
deaths due to complications of MI. Seven-day postoperative
mortality rates are ~10 –25% (75–90% survival) for female
young mice and 50 –70% (30–50% survival) for male young
mice (47, 61, 89, 98, 152, 170, 195, 202, 206, 234, 310 –312,
319, 323). Immediate survival from the surgery can also be
affected by baseline characteristics such as obesity, diabetes, or
high levels of circulating inflammatory cells, which, in turn,
H821MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
determine the response to anesthesia and surgery (60, 123, 202,
203). While there is no difference in infarct tolerance between
young and middle-aged mice (323), older mice may survive
better than younger mice (202, 319).
For permanent occlusion MI models, infarct size must be
measured in fresh LV slices at the time of necropsy by TTC
staining and typically ranges from 30% to 60% of the total LV
(47, 48, 61, 89, 98, 152, 195, 202, 206, 223, 310 –312, 319, 323).
The method for calculating infarct size varies across laboratories.
Some laboratories use area calculations, other laboratories mea-
sure length in the midmyocardium, and other laboratories measure
and average lengths in the subendocardium and subepicardium.
There is no need to use Evans blue for area at risk assessment in
permanent coronary occlusion models that pass the point from
ischemia to infarction, as the entire area at risk is infarcted. It
is important that MI surgical success is confirmed and that the
initial infarct injury is comparable across groups, to assess
remodeling differences at later stages. In mice, ligating the
coronary artery at the same anatomical location across groups
is important; 1 mm distal to the left atrium is the recommended
site to generate large infarcts (35– 60% of total LV). Failure to
induce MI can occur, usually due to missing the coronary
artery during the ligation step. Monitoring the electrocardio-
gram for ST segment elevation during the procedure reduces
this possibility. Echocardiography at 3 h after coronary occlu-
sion can be used to exclude animals with excessively small or
large MI before randomizing groups (153, 155, 195). Late
gadolinium-enhanced MRI is also useful for selecting animals
with consistent infarct sizes (262). When assessing effects of
treatments initiated post-MI, it is important to show that infarct
size is not different between groups before treatment. Plasma
sampling at 24 h post-MI can be used to assay cardiac bio-
markers, such as troponins and inflammatory cytokines, with
the caveat that these measurements can indicate presence or
absence of infarct and not extent of injury. After coronary
occlusion, care should be taken in performing these assess-
ments to minimize disturbing animals at times when cardiac
rupture may be triggered by stress, particularly at days 3–7
post-MI in untreated controls (96, 98). Small infarcts may
reflect technical issues in missing the coronary artery, resulting
in damage from the suture rather than reflecting the intended
myocardial ischemia and infarction. Infarct sizes 30% are
typically excluded. If included, small and large infarcts may
need to be grouped separately to reduce possible type II
statistical errors.
Cardiac wound healing and remodeling, typically assessed
days to weeks post-MI, can be examined using a wide variety
of approaches, including echocardiography, histology, bio-
chemistry, and cell biology (5, 217, 328). Serial measurements
of cardiac geometry and function by echocardiography are
useful for defining phenotypes. Cardiac dimensions vary de-
pending on heart rate and depth of anesthesia, and these
parameters must be carefully controlled and matched across
groups. Cardiac functional reserve can be assessed by measur-
ing the contractile response to inotropic drugs or volume
overload. Cardiac MRI and hemodynamic assessment by pres-
sure-volume catheterization are other ways to measure cardiac
physiology parameters. It is feasible to quantify infarct size
noninvasively and serially by using cardiac MRI (181). Hemo-
dynamic evaluation in mice is a terminal procedure, which
prevents its use in serial assessments.
Hematoxylin and eosin staining provides information on
areas of necrosis and inflammation, while picrosirius red stain-
ing provides information on total collagen accumulation both
in the scar and remote regions (316). Immunohistochemistry
for neutrophils, macrophages, lymphocytes, fibroblasts, and
endothelial cells provides information on the extent of inflam-
mation, scar formation, and neovascularization. Isolating indi-
vidual cell types and assessment of ex vivo phenotypes in
culture can further aid in understanding mechanisms. Studies
have revealed that inflammation evoked by acute myocardial
infarction also occurs systemically and that the spleen and liver
are important sources of cells and factors that influence LV
remodeling (71, 72, 95, 116, 198, 199, 293).
I/R MI: model rationale and variables measured. Implemen-
tation of myocardial reperfusion strategies has significantly
reduced mortality in acute STEMI. Reperfusion has contrib-
uted to the growing pool of patients who survive the acute
event and are at risk for adverse remodeling and subsequent
development of heart failure (133, 136). In addition to salvag-
ing cardiomyocytes, reperfusion has profound effects on cel-
lular events responsible for repair and remodeling.
Although timely reperfusion is essential to salvage viable
cardiomyocytes from ischemic death, extensive preclinical and
clinical evidence suggests that reperfusion itself causes injury
(119, 121, 147). Reperfusion-induced arrhythmias and myo-
cardial stunning are self-limited and reversible forms of rep-
erfusion injury, while microvascular obstruction and lethal
cardiomyocyte injury are irreversible and extend damage, thus
contributing to adverse outcomes following MI (13, 126, 177,
241, 318). In patients, no reflow during reperfusion may be
exacerbated due to the generation of microemboli composed of
atherosclerotic debris and thrombi during percutaneous coro-
nary interventions (135, 253).
MI both with or without reperfusion shares many of the
same technical guidelines, and this information is provided
above. The one technical difference is whether the ligation is
removed at 45– 60 min after the occlusion to reperfuse the
myocardium. Similar to permanent occlusion MI, studies in-
vestigating the inflammatory and reparative response following
MI with reperfusion need to take into account the dynamic
sequence of cellular events involved in repair. Common mea-
surements shared by the two MI models are shown in Table 2.
For studies aimed at investigating acute myocardial injury
using a reperfusion strategy, the duration of coronary occlusion
needs to be sufficient for the induction of significant MI but not
overly prolonged to cause irreversible injury in the entire area
at risk. From the cell physiology perspective, the reparative
response after MI can be divided into three distinct but over-
lapping phases: inflammation, proliferation, and maturation
(26, 65). In the infarcted myocardium, dying cardiomyocytes
release damage-associated molecular patterns and induce cy-
tokines and chemokines to recruit leukocytes into the infarcted
region, thus triggering an intense inflammatory reaction that
serves to clear the infarct from dead cells and ECM debris,
while initiating a reparative response (84). Early reperfusion
after irreversible cardiomyocyte injury accelerates and accen-
tuates the inflammatory reaction and has profound effects on
the pathological features of the infarct. Microvascular hyper-
permeability is evident in the myocardium with acute I/R (97).
Rapid extravasation of blood cells through the hyperpermeable
vessels may result in hemorrhagic changes (98, 178). Influx of
H822 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
phagocytotic macrophages is accelerated, resulting in more
rapid removal of dead cardiomyocytes compared with perma-
nent occlusion MI. In reperfused infarcts, dying cardiomyo-
cytes often exhibit large contraction bands, comprised of hy-
percontracted sarcomeres. Subsarcolemmal blebs and granular
mitochondrial densities, which are already present in irrevers-
ibly injured cardiomyocytes before restoration of blood flow,
become more prominent upon reperfusion.
Phagocytosis of dead cells by activated macrophages results
in the activation of endogenous anti-inflammatory pathways,
ultimately leading to resolution of the inflammatory infiltrate.
Suppression of inflammation is followed by recruitment of
activated myofibroblasts that deposit large amounts of ECM
proteins and by activation of angiogenesis (145). As the scar
matures, fibroblasts become quiescent and infarct neovessels
acquire a coat of mural cells (332). Compared with large
mammals, rodents exhibit an accelerated time course of infil-
tration with inflammatory and reparative cells (64).
Leukocyte infiltration during the inflammatory phase of
infarct healing and myofibroblast activation and accumulation
during the proliferative phase are predominantly localized in
the infarct region and border zone (87, 122, 280). During scar
maturation, the cellular content in the infarcted region is
reduced. At the same time, however, the number of activated
macrophages and fibroblasts in the remote remodeling myo-
cardium increases. Therefore, study of inflammatory and re-
parative cell infiltration and assessment of ECM protein depo-
sition should include systematic assessment of each end point
in the infarcted region, the peri-infarct area, and the remote
remodeling myocardium.
Sympathetic nerves are damaged by permanent coronary
occlusion but can regenerate after injury (220). In the setting of
chronic MI, regional hyperinnervation around the infarcted
region has been observed, and activation of cardiac sympa-
thetic nerves is important in triggering ventricular arrhythmias,
and such proarrhythmic action is dependent on the extent of
infarction (1, 70, 315, 326). In contrast, after I/R, chondroitin
sulfate proteoglycans prevent reinnervation (99, 100). Thus,
the model selected for sympathetic nerve evaluation should be
taken into consideration and depends on what question is being
asked.
MI: intervention considerations. The effects of interventions
on post-MI remodeling can be studied using both nonreper-
fused MI and reperfused MI/R models (11, 219, 232, 294,
322). Typically, nonreperfused MI yields accentuated dilative
remodeling and exacerbated dysfunction compared with a
reperfused infarct involving the same vascular territory, re-
flecting a combination of more extensive infarct and less
effective repair. In the reperfused MI/R model, the effects of
genetic or pharmacologic interventions implemented early af-
ter reperfusion may reflect differences in the extent of acute
cardiomyocyte injury rather than differences in wound healing
responses. With permanent occlusion MI (assuming a stan-
dardized area at risk) or very late reperfusion models, differ-
ences in geometry and function of the remodeling heart are
independent of acute cardiomyocyte injury and reflect effects
on inflammatory, reparative, or fibrotic cascades. In the pres-
ence of an occluded coronary artery, the delivery of systemi-
cally administered pharmacologic agents to the infarcted re-
gion of large animal models may be dependent on formation of
collaterals.
While the development of genetically targeted animals
(mice, rats, and rabbits) resulted in an explosion of studies
dissecting cell biological mechanisms and molecular pathways,
large animal models are considered closer to the clinical
situation for translational studies to test safety and effective-
Table 2. Common output measurements for in vivo MI and MI/reperfusion studies
Measurement Information Provided
Infarct size Infarct size (MI)
Infarct size per area at risk (MI/reperfusion)
Initial ischemic stimulus
Final area of damage
Effect of therapy or intervention
Plasma biomarkers Ischemia: creatine kinase, troponins
Inflammation: cytokines and chemokines
Scar formation: growth factors and the ECM
Neovascularization: angiogenic factors
Left ventricular physiology (echocardiography, MRI,
positron emission tomography imaging)
Geometry and function: dimensions, wall thickness, left ventricular dimensions and volumes,
fractional shortening, ejection fraction, remodeling index
Electrophysiological function: PR, QRS, and QT intervals/morphology; spontaneous and
inducible arrhythmias
Inflammation Immunohistochemistry and immunoblot analysis for cell numbers and inflammatory protein
expression
Flow cytometry analysis of the digested myocardium for individual cell phenotypes
Gene expression
Systemic and circulating inflammation
ECM scar Picrosirius red for collagen deposition
Immunohistochemistry and immunoblot analysis for ECM proteins and cross-linking enzymes
Gene expression
Scar strength assessment
Neovascularization Blood vessel numbers
Vessel type and quality
Microvascular damage Microvascular plugging
Hyperpermeability/edema
Hemorrhage
MI, myocardial infarction; ECM, extracellular matrix; MRI, magnetic resonance imaging.
H823MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
ness. Optimal study of molecular, cellular, and LV functional
end points and interpretation of the findings require under-
standing of the underlying pathophysiology. Assessment of
infarct size is typically the primary end point for investigations
examining the mechanisms of cardioprotection. Assessment of
chamber dimensions using echocardiography or MRI is crucial
to study the progression of adverse remodeling. Systolic and
diastolic cardiac geometry and function can be assessed non-
invasively using echocardiography (including Doppler ultra-
sound and speckle tracking), MRI, and hemodynamic assess-
ment. Mechanistic dissection of specific pathways may require
inclusion of additional cell physiology and molecular or pro-
teomic end points. In experimental models of MI, understand-
ing the time course of the cellular and molecular events is
critical for optimal study design. The effects of varying isch-
emic intervals on survival and activation of noncardiomyocyte
cellular and acellular (e.g., ECM) compartments are poorly
understood. Longer coronary occlusion times have distinct
effects on cardiac repair, by extending infarct size and by
influencing susceptible noncardiomyocyte populations, such as
endothelial cells, fibroblasts, pericytes, and immune cells (85).
Most studies characterizing responses to myocardial injury
have so far been performed in healthy young animals. Comor-
bid conditions, such as aging, diabetes, and metabolic dysfunc-
tion, affect the pattern of ischemic injury and modify the time
course and qualitative characteristics of the inflammatory and
reparative responses (27, 106, 202, 238, 298, 319, 323). These
comorbidities are relevant in the clinical context and must be
considered in translation of experimental findings to the clinic.
To summarize, the major strengths and limitations of the
nonreperfused and reperfused MI models are shown in Table 1.
Ablation
Model rationale and variables measured. The primary ad-
vantages of ablative injury techniques such as cryo-, thermal-,
and radio-frequency ablation are rigid and reproducible control
over the size, shape, and location of the region of damage.
With such methods, a wound can be stamped on the target
myocardial tissue with consistent dimensions, shape, and trans-
mural depth. Because the size of the damaged region is inde-
pendent of animal-to-animal variations in coronary anatomy
(223), the resultant ablation scar is also more reproducible than
ligation-induced injury (52, 160, 307), aiding studies of long-
term structural and functional remodeling and providing better
power to detect the effects of an experimental drug or cell
therapy. Infarct location can thus be controlled independently
from coronary anatomy and infarct transmurality can be con-
trolled (52, 160, 290, 305, 307). There are, however, important
differences in the modes of cell death in ablative vs. occlusion
injuries. For example, cryoinjury results in necrosis due to the
generation of ice crystals and disruption of the cell membrane
rather than direct ischemia. Furthermore, ablative injuries are
typically generated from the epicardial surface inward,
whereas ischemic infarcts tend to be propagated outward from
the inner myocardial layers (52, 160).
Unlike MI or MI with reperfusion, cryoinjury kills all (or
nearly all) cells within the core of the damaged region and
creates distinct wound margins. Thus, a number of studies have
used cryoinjury to avoid confounding effects of resident sur-
viving cells when testing stem cell and other related therapies
(6, 7, 258, 302). Ablation procedures typically apply a cooled/
heated probe to either the epicardial or endocardial surface of
the heart. The extent and depth of the lesion depend on both the
temperature of the probe and the time it remains in contact with
the tissue; damage can be extended by generating multiple
adjacent lesions or by repeat application at the same location.
Because these physical factors are central to injury formation,
investigators should report probe size and material, tempera-
ture, method and duration of preheating/cooling, precise ana-
tomic location and duration of probe application, and interle-
sion time and number of lesions (if applicable). Cryoablation
has been used to generate reproducible wounds and scar tissue
for the study of myocardial injury response in various species
including dogs (160, 171, 297), rabbits (6, 7, 302), rats (49, 82,
149, 190), and mice (109, 204, 257, 289, 305, 307). In mice,
survival rate after cryoinjury was nearly twice that of perma-
nent coronary ligation over an 8-wk period (307), whereas
dysfunction was similar. Lower mortality may be a conse-
quence of smaller infarct size. Ablative methodologies have
also been used in nonmammalian species such as zebrafish, to
probe the response of cardiac electrical properties to injury,
regeneration and scar formation (43, 46, 108). The ability to
destroy all cells within the cryoinfarct has provided interesting
clues regarding regeneration of fetal myocardium following
injury. In neonatal mice, mechanical or ischemic injuries to the
ventricular apex typically trigger regeneration, producing re-
covery of myocardial structure and function without scarring
(243, 276). Nontransmural cryoinfarcts in neonatal mice sim-
ilarly heal with minimal evidence of scarring and full func-
tional recovery with ongoing postnatal growth, while injuries
spanning the full thickness of the ventricular wall do not
regenerate muscle (57). Because neonatal mice can regenerate
myocardium during the first postnatal week, models of myo-
cardial ischemia in neonatal mice may be used to identify
pathways involved in cardiac regeneration (8, 208).
As with coronary ligation models, evaluation of LV geom-
etry and function with echocardiography and assessment of
electrophysiological remodeling and arrhythmia risk are rou-
tinely performed in cryoinjury models. Given the early time
course and different mechanisms of necrotic injury in cryoin-
jury versus ligation models, cryoinjury studies are typically
more focused on long-term myocardial regeneration or me-
chanical/electrophysiological remodeling rather than mecha-
nisms of acute postinjury cell death, inflammation, and scar
formation (52, 160). Ablation procedures are now used com-
monly in clinical electrophysiology, and it is therefore not
surprising that a number of experimental electrophysiology
studies have taken advantage of geometric control provided by
this model (54, 231). Cryoinjury and ischemic injury differ in
their transmural localization and the amount of surviving
myocardium (52, 109, 257, 258, 305). Finally, methods for
ablative targeting of cardiac neural tissues have proven useful
in studies of the role of autonomic inputs in normal and
arrhythmic hearts (254).
To summarize, the major strengths and limitations of the
cryoinjury model are shown in Table 1.
Cardioprotection
Model rationale and variables measured. An intervention
that is cardioprotective is broadly defined as serving to protect
H824 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
the heart (https://www.merriam-webster.com/dictionary/
cardioprotective), thereby in theory encompassing all of the
aforementioned aspects of cardiac damage and dysfunction. To
date, the only clinically established cardioprotective interven-
tion is early reperfusion (119, 133). The archetypal additive
cardioprotective intervention is, without question, ischemic
conditioning, encompassing the phenomena of ischemic pre-
conditioning, postconditioning, and remote conditioning (118,
175, 224, 247, 248, 329). Despite differences in the timing of
the protective stimulus (with preconditioning applied in a
prophylactic manner and postconditioning administered at the
time of reperfusion) and the site of the protective trigger (either
locally, or, for remote conditioning, in a tissue or organ distant
from the at-risk myocardium), all three forms of conditioning
share a common theme: there is overwhelming agreement that
ischemic preconditioning, postconditioning, and remote condi-
tioning render the heart resistant to lethal I/R-induced injury
(78, 118, 128, 163, 247).
This consensus with regard to ischemic conditioning and
cardioprotection is, however, a notable exception in the field.
Indeed, for the vast majority of the innumerable cardioprotec-
tive strategies that have been investigated, the current preclin-
ical literature on the topic of cardioprotection is fraught with
controversies and a lack of reproducibility among investigators
and laboratories (163, 188). The ensuing confusion in the field
may be attributed in part to two confounding factors: an overly
broad use of the term cardioprotection in some studies, to-
gether with false positive and false negative outcomes derived
from protocols executed in a suboptimal manner (162, 163,
188).
In some instances, a broad and suitably framed definition of
cardioprotection incorporating, for example, endothelial integ-
rity and vascular function, is appropriate (126, 225). A gener-
ally acknowledged and more narrowly focused hallmark of
cardioprotection is defined as an agent or intervention that,
when administered in the setting of ischemia/reperfusion, aug-
ments myocardial salvage and reduces myocardial infarct size
beyond that achieved by reperfusion alone (128). Accordingly,
for our purposes, we focus on cardiomyocyte viability and
define cardioprotection as a strategy that attenuates cardiomy-
ocyte death. Cardiomyocyte viability can be assessed in a full
spectrum of models, ranging from cardiomyocytes in culture to
isolated buffer-perfused hearts to in vivo studies in rodents
(mice and rats) or larger animals (including rabbits, dogs, pigs,
sheep, and, in a small number of studies, primates). There is no
ideal model that completely mirrors the clinical scenario to
fully ensure absolute translational relevance. Rather, each
model has merits and disadvantages (as shown in Table 3).
The overwhelming strength of the mouse species is the
availability of genetically modified strains to elucidate molec-
ular mechanisms once candidate cardioprotective strategies
have been identified, a benefit that is balanced by inherent
variability and resultant requirement for large sample sizes. An
additional problem of the mouse is the atypical geometry of
nontransmural infarcts, in which the subendocardium is spared
from death by diffusion of oxygen from the LV cavity and
occupies an inordinate proportion of the total LV wall thick-
ness. In all rodents, heart rate is much higher and therefore
infarct development is much faster than in larger mammals and
humans. Studies conducted in large animals (in particular, the
pig) are considered to have the greatest potential for preclinical
Table 3. Recommendations for cardioprotection studies
Model
Gold Standard
Primary End Point Required Covariates Potential Secondary End points Advantages Limitations
Cultured
cardiomyocytes
Cell viability (live-
dead assay)
None Types of cell death (i.e., apoptosis),
mitochondrial function; ROS
production
Capacity for high throughput Reductionist
Isolated buffer-perfused
hearts (mouse, rat,
and rabbit)
Infarct size (TTC) For models of regional
ischemia: area at
risk
Measures of LV function and coronary
flow; cTn as a secondary index of
cardiac injury
Throughput higher than in vivo; eliminates
confounding effect of intervention on
systemic blood vessels
Reductionist
Mouse Infarct size (TTC) Area at risk;
hemodynamics
Measures of LV function; measures of
no reflow; CK or cTn as secondary
indexes of cardiac injury
Availability of genetically modified
strains; model of low collateral flow
(measurement of RMBF not required)
Small size; differences between strains;
substantial variability requiring large
nvalues
Rat and rabbit Infarct size (TTC) Area at risk;
hemodynamics
Measures of LV function; measures of
no reflow; CK or cTn as secondary
index of cardiac injury
Reliable infarct production; relatively high
survival rate in experienced hands;
commercial availability of strains that
have comorbidities; model of low
collateral flow (measurement of RMBF
not required)
Not high throughput
Dog Infarct size (TTC) Collateral blood flow;
area at risk;
hemodynamics
Measures of LV function; measures of
no reflow; CK or cTn as secondary
indexes of cardiac injury
Large amount of historical data; shows
effect of intervention in the setting of
variable collateral flow; mimics humans
with high collateral flow
High cost; variability in collateral
perfusion (RMBF must be
measured); potential for lethal
arrhythmias; not high throughput
Pig Infarct size (TTC) Area at risk;
hemodynamics
Measures of LV function; measures of
no reflow; CK or cTn as secondary
indexes of cardiac injury
Model of low collateral flow
(measurement of RMBF not required);
mimics humans with low collateral flow
High cost; high incidence of lethal
arrhythmias; not high throughput
TTC, triphenyltetrazolium chloride; LV, left ventricular; CK, creatine kinase; cTn, cardiac troponin; RMBF, regional myocardial blood flow.
H825MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
relevance (103, 140). This advantage is accompanied by sub-
stantial costs incurred and, particularly in pigs, the well-
documented high incidence of lethal ventricular arrhythmias
(282).
In all studies focused on cardioprotection, the primary end
point must be a quantitative assessment of cardiomyocyte
viability. In cell culture models, these data may be obtained
using a live-dead assay such as trypan blue exclusion, pro-
pidium iodide exclusion, or other commercially available as-
says. In intact hearts, including isolated buffer-perfused hearts
and all in vivo models, the gold standard end point is myocar-
dial infarct size by TTC staining and, at later time points,
histopathologic analysis (79, 81). For models involving reper-
fusion, the area at risk must be quantified and infarct size must
be expressed as a proportion of the risk region. With global
rather than regional ischemia/reperfusion, the entire heart is
rendered at risk and thus infarct size is appropriately expressed
as a proportion of the total ventricular area. The duration of
ischemia must be sufficient to cause significant infarction but
not complete death of at-risk cardiomyocytes in the control
cohort. That is, if the duration of ischemia is selected such that
infarct size in controls is either inordinately small or exces-
sively large, the scope for salvage and probability of achieving
cardioprotection with any intervention is negligible. Irrespec-
tive of the model used, the protocol must involve I/R (or, in cell
culture models, H/R) rather than ischemia (or hypoxia) alone.
This requirement reflects the fact that even the most powerful
and well-established cardioprotective strategies such as isch-
emic preconditioning simply delay, rather than prevent, the
progression to cardiomyocyte death and infarction (224, 324).
The duration of reperfusion must be sufficient to allow for the
accurate and unambiguous delineation of necrotic and viable
myocardium. This is of particular importance when infarct size
is quantified using TTC staining: a minimum of 1–2 h of
reperfusion is considered mandatory in rodent hearts, while
longer periods of at least 3 h are standard in large animal
models (283).
Attention must be paid to essential covariates and possible
cofounders. Important considerations for all in vivo models
include body temperature, the choice of anesthetics and anal-
gesics, and changes in the determinants of myocardial oxygen
supply and demand, all of which are well recognized to have
profound effects on infarct size (111, 247). Particular care must
be taken to avoid the possibility of inadvertent preconditioning:
examples include triggering a protective phenotype by unin-
tentionally subjecting the myocardium to brief periods of
hypoxia or ischemia during surgical preparation, or intention-
ally imposing a period of ischemia in an effort to identify the
extent of the at-risk myocardium (180). Additional covariates
and confounders are model specific. In rodents, age, sex, and
strain of the animals have all been implicated or identified to
influence myocardial infarct size (10, 111, 306) Circadian
variation, the time of day at which experiments are performed,
may also be important (16, 28, 110, 266). The canine model is
known for its variability in the magnitude of collateral blood
flow. Accordingly, when using this model, measurement of
regional myocardial blood flow during coronary artery occlu-
sion and incorporation of collateral flow as a covariate in the
analysis of infarct size are mandatory (64, 284).
Finally, the overwhelming majority of studies conducted to
date have assessed the efficacy of candidate cardioprotective
strategies using healthy, juvenile, or adult animals. There is
evidence that the infarct-sparing effect of these purportedly
protective interventions may be lost or attenuated in the setting
of clinically relevant comorbidities, including aging, type 1 and
type 2 diabetes, hypercholesterolemia, and hypertension, and
may be influenced by diet or exercise (3, 4, 51, 78, 115, 117,
146, 150, 186, 197, 246, 249, 321). Accordingly, once proof of
principle is established, it is imperative that promising cardio-
protective therapies be reevaluated, adhering to the essential
elements of rigor described above, in comorbid models (127).
All protocols must include concurrent and appropriate control
cohorts. For example, when potential cardioprotective drugs
are evaluated, controls must receive matched volumes of ve-
hicle administered in an identical manner.
THE ISSUE OF TRANSLATION: TOWARD A RANDOMIZED
CONTROLLED STUDY OR TRIAL DESIGN
There is currently no established intervention, aside from
timely reperfusion, that limits damage to hearts of patients
experiencing myocardial ischemia to the extent that clinical
outcome is improved (133, 138). Discussions of prior failures
and hope for future successes have been reviewed elsewhere
(120, 127, 176). Several elements missing from preclinical
studies of infarct size reduction have been identified and
include absence of critical investigator blinding, statistical
weaknesses (underpowered studies), and insufficient method-
ological detail. These deficiencies explain in part the failure to
translate preclinical results into effective infarct-sparing treat-
ments in patients. Thus, many have questioned the reproduc-
ibility of interventions to protect from MI (i.e., reduce infarct
size).
Much of the lack of reproducibility has been ascribed to
limited or lacking scientific rigor (29, 250). In response to these
and other concerns, the United States National Institutes of
Health now includes explicit requirements for applicants to
show, and reviewers to evaluate, the level of scientific rigor in
grant applications. Whether suboptimal rigor fully explains and
underlies the reproducibility crisis is unclear (162). Nonethe-
less, advocating for reproducibility and scientific rigor is wel-
come.
Appropriate statistical issues should be considered before
initiating an infarct-sparing intervention. Investigators should
know the standard deviation of their primary, prespecified end
point, such as infarct size, chamber dimension, or ejection
fraction. A power analysis for the primary end point will
determine and justify the number of subjects to be enrolled in
each group. This may not be feasible when investigating an
entirely innovative strategy as there may be no basis for an
estimation of the expected effect size. The standard deviation
of the investigator’s most recent blinded study can be used to
determine group size (229). The choice of statistical analyses
must be appropriate for the study design (300). For two-group
studies, t-tests (or nonparametric equivalent) may be used; for
protocols involving multiple cohorts, ANOVA (or nonpara-
metric equivalent) is mandatory. Analysis of covariance, as-
sessing the effects including variations in risk region and
collateral blood flow, may also be applied.
Blinded assignment of animals and randomization to control
or treated groups is mandatory whenever possible. Incorpora-
tion of randomization with blinding is an easy and logical
H826 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
approach. For testing classical drug-based interventions or
when comparing mutant mice, individuals responsible for
blinding can use block randomization and label tubes or mice
with a simple unique alphanumeric code. The surgeon per-
forming the protocol should have no knowledge of the inter-
vention or genotype. We recommend that the same surgeon
perform all surgeries; if multiple surgeons are used, equal
numbers from all groups should be matched across the surgeon
pool. It is imperative that neither the individual analyzing
infarct size, nor the individual performing any other secondary
analyses, knows the intervention or genotype until all data are
compiled. If the potential for excluding animals exists, this
should be done based on previously declared exclusion and
inclusion criteria, and all decisions should be made before
disclosure of group assignment; details of such exclusions
should be made clear in any publications. Table 4 shows
ARRIVE guidelines for manuscript submission, modified to
focus on ischemia studies.
The Consortium for Preclinical Assessment of Cardiopro-
tective Therapies (CAESAR) endeavored to address the pri-
mary issues related to reproducibility in cardioprotection stud-
ies (163). Performing the same protocol at multiple sites and in
multiple species was extraordinarily challenging; the most
notable were challenges in identifying the underlying explana-
tions for differences in infarct size (or other variables) between
centers while they were implementing consonant protocols.
Interestingly, there were several instances of mice being or-
dered simultaneously from the same vendor, only to have
significantly different body weights at the time of study (de-
spite being fed the same chow). To be clear, the weight
differences were relatively small but on occasion were signif-
icantly different for reasons unknown. It is likely that small
differences, such as this, could theoretically affect the results
(or the perception of not being able to reproduce studies) of
published studies. During the CAESAR experience, numerous
seemingly extraneous factors were considered, such as differ-
ences in municipal water source, room temperature, relative
humidity, type of lighting, traffic in the room, and other
seemingly innocuous details that may or may not affect the
responsiveness of mice to an infarct-sparing regimen. All of
these variables reflect inherent challenges in performing in
vivo studies.
More important than slight differences in body weight or
other such ancillary factors were initial challenges in generat-
ing equivalent infarct sizes at different institutions, despite
using the same protocol. Such occasional variations emphasize
the unequivocal requirement in all protocols for concurrent and
appropriate controls (see Cardioprotection).
OVERALL DISCUSSION AND CONCLUSIONS
As highlighted throughout these guidelines, ischemia and
I/R have multiple consequences that show temporal variation
in terms of both incidence and influence on outcomes and may
be model dependent. Examples range from acute effects (in-
cluding biochemical perturbations in cardiomyocytes and other
cardiac cell types, disruption in cardiac conduction and devel-
opment of arrhythmias, contractile dysfunction, abnormalities
in endothelial, and vascular reactivity) to longer term outcomes
(such as cardiomyocyte death, microvascular obstruction and
no-reflow, scar healing, and LV remodeling) and, ultimately,
major adverse cardiac events, including heart failure and death
(136). Figure 1 shows the diversity in models available to
assess ischemia across its spectrum, and Tables 4 and 5 show
general recommendations for rigor and reproducibility in isch-
emia studies.
While writing these guidelines, the authors discussed
whether an algorithm to define the choice of model for a given
scientific question would be helpful. The consensus was that
the topic of myocardial ischemia and infarction and the many
clinical manifestations of coronary artery disease resulting in
and from myocardial ischemia or infarction are so broad and so
complex that we find ourselves unable to provide an algorithm
that truly covers all potential scientific approaches. Indeed,
such an algorithm may be used by regulatory and funding
agencies to actually limit research in the field, which would be
counterproductive to the goals of this document.
The approach used will vary depending on the questions
being addressed; as such, all of the approaches described above
may be considered good and a gold standard if appropriate to
address the target hypothesis. In vitro studies using isolated
cardiomyocytes or even isolated organelles (e.g., mitochon-
dria) are well suited to identify single molecular targets of
injury and protection (102). Isolated, buffer-perfused heart
Table 4. Checklist of considerations for rigor and reproducibility, modified from the ARRIVE Guidelines (168)
Item Details
Ethical statements Institutional Animal Care and Use Committee approval, Guide for the Care and Use of Laboratory Animals, welfare
assessments and interventions
Animal description, housing,
husbandry
Species, strain, source, age (mean and range), sex, genotypes, body weight (mean and range), health/immune status, housing
type (specific pathogen free), cage type, bedding material, number of cage companions, light-dark cycle, room
temperature and humidity, food type, food and water access
Study design Define model used; define groups; matching, randomization, and blinding protocols; order of treatment and assessment of
groups; method to confirm model success; inclusion/exclusion criteria (e.g., lower limit for infarct size); daily monitor to
record time and cause of death; define timing of perioperative and postoperative phases for survival analysis, flow chart
for complex designs
Experimental procedures Define area examined (e.g., infarct, remote, both regions); positive and negative controls; for drugs: formulation, dose, site,
and route of administration; anesthesia and analgesic use and pain monitoring; surgical procedure details and monitoring
records (e.g., electrocardiogram, heart rate, and anesthesia depth); method of euthanasia; time of day performed
Sample sizes Number of animals used per group for each experiment; sample size calculation; number of independent replicates for cell
culture studies
Variables measured Primary and secondary end points, list any animals or samples removed from analysis with reason
Statistics Methods used for each analysis; test for assumptions
ARRIVE, Animals in Research: Reporting In Vivo Experiments.
H827MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
models can be used to study the acute biochemical and func-
tional mechanisms of myocardial I/R injury and cardioprotec-
tion. Due to the need for stable stenosis and the required spatial
resolution of regional myocardial blood flow and contractile
function measurements, large animal preparations are recom-
mended as models for the clinical manifestations of chronic
stable angina and coronary microembolization. In vivo prepa-
rations are required to study more long-term effects of myo-
cardial I/R and respective therapeutic interventions. We distin-
guish between permanent occlusion MI and reperfused MI
models and also highlight measurements in common. Perma-
nent coronary occlusion MI and reperfused MI models are both
well suited to study repair and remodeling. I/R models are
mandatory to study cardioprotection, and the ischemia must be
of sufficient severity and duration to cause some infarction.
Cryo-/thermoinjury is not suited to study the pathophysiology
of MI but well suited to study repair and regeneration pro-
cesses.
Prospective planning of study design, i.e., randomization for
appropriate control versus treatment, blinding of investigators
(as much as possible for a given experimental protocol and the
subsequent data analysis), and adequate statistics, is mandatory
for reproducibility of all experimental models of myocardial
I/R and infarction. As larger data sets are being acquired,
consideration for how to harness big data should be given (272,
274). Compiling databases to incorporate results from across
studies and across laboratories will provide a means to use
epidemiological approaches or big data tools to validate pub-
lished findings, generate novel hypotheses, and assess individ-
ual variability in response to ischemia.
In conclusion, these guidelines provide recommendations to
help the investigator plan and execute a full range of studies
involving myocardial ischemia and infarction.
GRANTS
We acknowledge support from the following: National Institutes of
Health Grants HL-002066, HL-051971, HL-056728, HL-061610, HL-
075360, HL-078825, HL-088533, HL-092141, HL-093579, HL-107153,
HL-111600, HL-112730, HL-112831, HL-113452, HL-113530, HL-
116449, HL-128135, HL-129120, HL-129823, HL-130266, HL-131647,
HL-132075, HL-135772, GM-103492, HL-131647, HL-76246, HL-85440,
GM-104357, GM-114833, GM-115428, and UL1-TR-001412; American
Heart Association Grants 16GRNT30960054 and 16CSA28880004; Grand
Challenge Award; Department of Defense Grants 16W81XWH-16 –1-0592,
PR151051, PR151134, and PR151029; Biomedical Laboratory Research and
Development Service of the Veterans Affairs Office of Research and Devel-
opment Awards 1IO1BX002659 and I01BX000505; Australian National
Health and Medical Research Council Research Fellowship APP1043026;
Bundesministerium für Bildung und Forschung Grant BMBF01 EO1004; and
German Research Foundation Grants DFG He 1320/18-3 and SFB 1116 B8.
R. A. Kloner reports grant support from Stealth Biotherapeutics, Servier, Inc.,
and Faraday to test experimental compounds in experimental myocardial
infarction models.
DISCLAIMERS
The content is solely the responsibility of the authors and does not
necessarily represent the official views of the National Institutes of Health,
American Heart Association, United States Department of Defense, United
States Veterans Administration, National Health and Medical Research Coun-
cil, or German Research Foundation.
Table 5. Recommendations for ischemia studies
Common Experimental design should follow ARRIVE guidelines (see Table 4)
Cardiomyocytes When comparing groups, geometry and function end points by echocardiography assessment may not change. This
does not necessarily indicate no effect of the intervention.
Control groups can be shared across studies to reduce animal use, as long as the samples are collected within the
same timeframe, under identical conditions, and details are provided in the methods. Previously collected historical
controls should be avoided or clearly indicated.
For time-course studies, sham surgery can be replaced by day 0 negative controls if minimally invasive procedures
are used, which greatly reduces animal use.
Use to discern direct cardiomyocyte effects and responses
Isolated perfused hearts Use to discern cardiac effects and responses
Angina, stunning, hibernation, and
ischemic cardiomyopathy Use to reflect a particular clinical scenario
MI Use nonreperfused or reperfused MI to study repair and remodeling
Use nonreperfused MI to test interventions in a robust remodeling model and to test interventions in a model
clinically relevant to the nonreperfused patient
Use reperfused MI to test interventions in a model clinically relevant to the reperfused patient
Use reperfused MI to study cardioprotection
Essential to measure infarct size for nonreperfused MI and infarct size and area at risk for reperfused MI
Ablation Use to control size, shape, or location
Use to achieve maximal and uniform cell death in the target region
Use to investigate mechanisms of action of corresponding to clinical ablation technique
While not suited to study MI pathophysiology, is well suited to study repair and regeneration
Essential to quantify transmural extent of damage, to assess transmural variation
Essential to recognize that transmural extent of the lesion may evolve over time
Essential to standardize experimental protocol (e.g., probe temperature and contact time) to achieve consistent lesions
Cardioprotection Use to evaluate potentially protective strategies in the ischemia-reperfusion model
Essential to measure infarct size and area at risk
ARRIVE, Animals in Research: Reporting In Vivo Experiments; MI, myocardial infarction.
H828 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared by the authors.
AUTHOR CONTRIBUTIONS
M.L.L. and G.H. conceived and designed research; M.L.L., R.G.G., K.P.,
C.M.R., and G.H. prepared figures; M.L.L., R.B., J.M.C., X.-J.D., N.G.F., S.F.,
R.G.G., J.W.H., S.P.J., R.K., D.J.L., R.L., E.M., P.P., K.P., F.A.R., L.S.L.,
C.M.R., J.E.V.E., and G.H. drafted manuscript; M.L.L., R.B., J.M.C., X.-J.D.,
N.G.F., S.F., R.G.G., J.W.H., S.P.J., R.K., D.J.L., R.L., E.M., P.P., K.P.,
F.A.R., L.S.L., C.M.R., J.E.V.E., and G.H. edited and revised manuscript;
M.L.L., R.B., J.M.C., X.-J.D., N.G.F., S.F., R.G.G., J.W.H., S.P.J., R.K.,
D.J.L., R.L., E.M., P.P., K.P., F.A.R., L.S.L., C.M.R., J.E.V.E., and G.H.
approved final version of manuscript.
REFERENCES
1. Ajijola OA, Lux RL, Khahera A, Kwon O, Aliotta E, Ennis DB,
Fishbein MC, Ardell JL, Shivkumar K. Sympathetic modulation of
electrical activation in normal and infarcted myocardium: implications
for arrhythmogenesis. Am J Physiol Heart Circ Physiol 312: H608 –
H621, 2017. doi:10.1152/ajpheart.00575.2016.
2. Algranati D, Kassab GS, Lanir Y. Why is the subendocardium more
vulnerable to ischemia? A new paradigm. Am J Physiol Heart Circ
Physiol 300: H1090 –H1100, 2011. doi:10.1152/ajpheart.00473.2010.
3. Alleman RJ, Tsang AM, Ryan TE, Patteson DJ, McClung JM,
Spangenburg EE, Shaikh SR, Neufer PD, Brown DA. Exercise-
induced protection against reperfusion arrhythmia involves stabilization
of mitochondrial energetics. Am J Physiol Heart Circ Physiol 310:
H1360 –H1370, 2016. doi:10.1152/ajpheart.00858.2015.
4. Andres AM, Kooren JA, Parker SJ, Tucker KC, Ravindran N, Ito
BR, Huang C, Venkatraman V, Van Eyk JE, Gottlieb RA, Mentzer
RM Jr. Discordant signaling and autophagy response to fasting in hearts
of obese mice: implications for ischemia tolerance. Am J Physiol Heart
Circ Physiol 311: H219 –H228, 2016. doi:10.1152/ajpheart.00041.2016.
5. Aronsen JM, Espe EK, Skårdal K, Hasic A, Zhang L, Sjaastad I.
Noninvasive stratification of postinfarction rats based on the degree of
cardiac dysfunction using magnetic resonance imaging and echocardiog-
raphy. Am J Physiol Heart Circ Physiol 312: H932–H942, 2017. doi:10.
1152/ajpheart.00668.2016.
6. Atkins BZ, Hueman MT, Meuchel J, Hutcheson KA, Glower DD,
Taylor DA. Cellular cardiomyoplasty improves diastolic properties of
injured heart. J Surg Res 85: 234 –242, 1999. doi:10.1006/jsre.1999.
5681.
7. Atkins BZ, Hueman MT, Meuchel JM, Cottman MJ, Hutcheson KA,
Taylor DA. Myogenic cell transplantation improves in vivo regional
performance in infarcted rabbit myocardium. J Heart Lung Transplant
18: 1173–1180, 1999. doi:10.1016/S1053-2498(99)00096-0.
8. Aurora AB, Porrello ER, Tan W, Mahmoud AI, Hill JA, Bassel-
Duby R, Sadek HA, Olson EN. Macrophages are required for neonatal
heart regeneration. J Clin Invest 124: 1382–1392, 2014. doi:10.1172/
JCI72181.
9. Bache RJ, Arentzen CE, Simon AB, Vrobel TR. Abnormalities in
myocardial perfusion during tachycardia in dogs with left ventricular
hypertrophy: metabolic evidence for myocardial ischemia. Circulation
69: 409 –417, 1984. doi:10.1161/01.CIR.69.2.409.
10. Baker JE, Konorev EA, Gross GJ, Chilian WM, Jacob HJ. Resis-
tance to myocardial ischemia in five rat strains: is there a genetic
component of cardioprotection? Am J Physiol Heart Circ Physiol 278:
H1395–H1400, 2000. doi:10.1152/ajpheart.2000.278.4.H1395.
11. Barlow SC, Doviak H, Jacobs J, Freeburg LA, Perreault PE, Zellars
KN, Moreau K, Villacreses CF, Smith S, Khakoo AY, Lee T, Spinale
FG. Intracoronary delivery of recombinant TIMP-3 after myocardial
infarction: effects on myocardial remodeling and function. Am J Physiol
Heart Circ Physiol 313: H690 –H699, 2017. doi:10.1152/ajpheart.00114.
2017.
12. Bauer M, Kang L, Qiu Y, Wu J, Peng M, Chen HH, Camci-Unal G,
Bayomy AF, Sosnovik DE, Khademhosseini A, Liao R. Adult cardiac
progenitor cell aggregates exhibit survival benefit both in vitro and in
vivo. PLoS One 7: e50491, 2012. doi:10.1371/journal.pone.0050491.
13. Becerra R, Román B, Di Carlo MN, Mariangelo JI, Salas M, Sanchez
G, Donoso P, Schinella GR, Vittone L, Wehrens XH, Mundiña-
Weilenmann C, Said M. Reversible redox modifications of ryanodine
receptor ameliorate ventricular arrhythmias in the ischemic-reperfused
heart. Am J Physiol Heart Circ Physiol 311: H713–H724, 2016. doi:10.
1152/ajpheart.00142.2016.
14. Bell RM, Mocanu MM, Yellon DM. Retrograde heart perfusion: the
Langendorff technique of isolated heart perfusion. J Mol Cell Cardiol 50:
940 –950, 2011. doi:10.1016/j.yjmcc.2011.02.018.
15. Beltrami CA, Finato N, Rocco M, Feruglio GA, Puricelli C, Cigola E,
Quaini F, Sonnenblick EH, Olivetti G, Anversa P. Structural basis of
end-stage failure in ischemic cardiomyopathy in humans. Circulation 89:
151–163, 1994. doi:10.1161/01.CIR.89.1.151.
16. Bennardo M, Alibhai F, Tsimakouridze E, Chinnappareddy N,
Podobed P, Reitz C, Pyle WG, Simpson J, Martino TA. Day-night
dependence of gene expression and inflammatory responses in the re-
modeling murine heart post-myocardial infarction. Am J Physiol Regul
Integr Comp Physiol 311: R1243–R1254, 2016. doi:10.1152/ajpregu.
00200.2016.
17. Bolli R. Mechanism of myocardial “stunning”. Circulation 82: 723–738,
1990. doi:10.1161/01.CIR.82.3.723.
18. Bolli R, Bhatti ZA, Tang XL, Qiu Y, Zhang Q, Guo Y, Jadoon AK.
Evidence that late preconditioning against myocardial stunning in con-
scious rabbits is triggered by the generation of nitric oxide. Circ Res 81:
42–52, 1997. doi:10.1161/01.RES.81.1.42.
19. Bolli R, Marbán E. Molecular and cellular mechanisms of myocardial
stunning. Physiol Rev 79: 609 –634, 1999. doi:10.1152/physrev.1999.79.
2.609.
20. Bolli R, Zhu WX, Thornby JI, O’Neill PG, Roberts R. Time course
and determinants of recovery of function after reversible ischemia in
conscious dogs. Am J Physiol Heart Circ Physiol 254: H102–H114,
1988.
21. Bolli R, Zughaib M, Li XY, Tang XL, Sun JZ, Triana JF, McCay PB.
Recurrent ischemia in the canine heart causes recurrent bursts of free
radical production that have a cumulative effect on contractile function.
A pathophysiological basis for chronic myocardial “stunning”. J Clin
Invest 96: 1066 –1084, 1995. doi:10.1172/JCI118093.
22. Boyle MP, Weisman HF. Limitation of infarct expansion and ventric-
ular remodeling by late reperfusion. Study of time course and mechanism
in a rat model. Circulation 88: 2872–2883, 1993. doi:10.1161/01.CIR.
88.6.2872.
23. Braunwald E, Rutherford JD. Reversible ischemic left ventricular
dysfunction: evidence for the “hibernating myocardium”. J Am Coll
Cardiol 8: 1467–1470, 1986. doi:10.1016/S0735-1097(86)80325-4.
24. Brodarac A, Šaric´ T, Oberwallner B, Mahmoodzadeh S, Neef K,
Albrecht J, Burkert K, Oliverio M, Nguemo F, Choi YH, Neiss WF,
Morano I, Hescheler J, Stamm C. Susceptibility of murine induced
pluripotent stem cell-derived cardiomyocytes to hypoxia and nutrient
deprivation. Stem Cell Res Ther 6: 83, 2015. doi:10.1186/s13287-015-
0057-6.
25. Buckberg GD, Fixler DE, Archie JP, Hoffman JI. Experimental
subendocardial ischemia in dogs with normal coronary arteries. Circ Res
30: 67–81, 1972. doi:10.1161/01.RES.30.1.67.
26. Bujak M, Dobaczewski M, Chatila K, Mendoza LH, Li N, Reddy A,
Frangogiannis NG. Interleukin-1 receptor type I signaling critically
regulates infarct healing and cardiac remodeling. Am J Pathol 173:
57–67, 2008. doi:10.2353/ajpath.2008.070974.
27. Bujak M, Kweon HJ, Chatila K, Li N, Taffet G, Frangogiannis NG.
Aging-related defects are associated with adverse cardiac remodeling in
a mouse model of reperfused myocardial infarction. J Am Coll Cardiol
51: 1384 –1392, 2008. doi:10.1016/j.jacc.2008.01.011.
28. Bulluck H, Nicholas J, Crimi G, White SK, Ludman AJ, Pica S,
Raineri C, Cabrera-Fuentes HA, Yellon D, Rodriguez-Palomares J,
Garcia-Dorado D, Hausenloy DJ. Circadian variation in acute myocar-
dial infarct size assessed by cardiovascular magnetic resonance in rep-
erfused STEMI patients. Int J Cardiol 230: 149 –154, 2017. doi:10.1016/
j.ijcard.2016.12.030.
29. Bustin SA, Huggett JF. Reproducibility of biomedical research–the
importance of editorial vigilance. Biomol Detect Quantif 11: 1–3, 2017.
doi:10.1016/j.bdq.2017.01.002.
30. Bux AS, Lindsey ML, Vasquez HG, Taegtmeyer H, Harmancey R.
Glucose regulates the intrinsic inflammatory response of the heart to
surgically induced hypothermic ischemic arrest and reperfusion. Physiol
Genomics 49: 37–52, 2017. doi:10.1152/physiolgenomics.00102.2016.
31. Canfield SG, Sepac A, Sedlic F, Muravyeva MY, Bai X, Bosnjak ZJ.
Marked hyperglycemia attenuates anesthetic preconditioning in human-
induced pluripotent stem cell-derived cardiomyocytes. Anesthesiology
117: 735–744, 2012. doi:10.1097/ALN.0b013e3182655e96.
H829MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
32. Canton M, Skyschally A, Menabò R, Boengler K, Gres P, Schulz R,
Haude M, Erbel R, Di Lisa F, Heusch G. Oxidative modification of
tropomyosin and myocardial dysfunction following coronary microem-
bolization. Eur Heart J 27: 875–881, 2006. doi:10.1093/eurheartj/ehi751.
33. Canty JM Jr. Coronary pressure-function and steady-state pressure-flow
relations during autoregulation in the unanesthetized dog. Circ Res 63:
821–836, 1988. doi:10.1161/01.RES.63.4.821.
34. Canty JM Jr, Fallavollita JA. Resting myocardial flow in hibernating
myocardium: validating animal models of human pathophysiology. Am J
Physiol Heart Circ Physiol 277: H417–H422, 1999.
35. Canty JM Jr, Giglia J, Kandath D. Effect of tachycardia on regional
function and transmural myocardial perfusion during graded coronary
pressure reduction in conscious dogs. Circulation 82: 1815–1825, 1990.
doi:10.1161/01.CIR.82.5.1815.
36. Canty JM Jr, Klocke FJ. Reduced regional myocardial perfusion in the
presence of pharmacologic vasodilator reserve. Circulation 71: 370 –377,
1985. doi:10.1161/01.CIR.71.2.370.
37. Canty JM Jr, Klocke FJ. Reductions in regional myocardial function at
rest in conscious dogs with chronically reduced regional coronary artery
pressure. Circ Res 61: II107–II116, 1987.
38. Canty JM Jr, Mates RE. A programmable pressure control system for
coronary flow studies. Am J Physiol Heart Circ Physiol 243: H796–
H802, 1982.
39. Canty JM Jr, Smith TP Jr. Adenosine-recruitable flow reserve is absent
during myocardial ischemia in unanesthetized dogs studied in the basal
state. Circ Res 76: 1079 –1087, 1995. doi:10.1161/01.RES.76.6.1079.
40. Canty JM Jr, Suzuki G, Banas MD, Verheyen F, Borgers M,
Fallavollita JA. Hibernating myocardium: chronically adapted to isch-
emia but vulnerable to sudden death. Circ Res 94: 1142–1149, 2004.
doi:10.1161/01.RES.0000125628.57672.CF.
41. Canty JM Jr, Duncker DJ. Coronary blood flow and myocardial
ischemia. In: Braunwald’s Heart Disease, edited by Mann DL, Zipes DP,
Libby P, Bonow RO. Philadelphia, PA: Elsevier, 2014, p. 1029 –1056.
42. Canty JM Jr, Suzuki G. Myocardial perfusion and contraction in acute
ischemia and chronic ischemic heart disease. J Mol Cell Cardiol 52:
822–831, 2012. doi:10.1016/j.yjmcc.2011.08.019.
43. Cao H, Yu F, Zhao Y, Zhang X, Tai J, Lee J, Darehzereshki A,
Bersohn M, Lien CL, Chi NC, Tai YC, Hsiai TK. Wearable multi-
channel microelectrode membranes for elucidating electrophysiological
phenotypes of injured myocardium. Integr Biol 6: 789 –795, 2014.
doi:10.1039/C4IB00052H.
44. Capasso JM, Li P, Anversa P. Nonischemic myocardial damage in-
duced by nonocclusive constriction of coronary artery in rats. Am J
Physiol Heart Circ Physiol 260: H651–H661, 1991.
45. Capasso JM, Malhotra A, Li P, Zhang X, Scheuer J, Anversa P.
Chronic nonocclusive coronary artery constriction impairs ventricular
function, myocardial structure, and cardiac contractile protein enzyme
activity in rats. Circ Res 70: 148 –162, 1992. doi:10.1161/01.RES.70.1.
148.
46. Chablais F, Jaz´win´ ska A. Induction of myocardial infarction in adult
zebrafish using cryoinjury. J Vis Exp 2012: 3666, 2012. doi:10.3791/
3666.
47. Chan W, White DA, Wang XY, Bai RF, Liu Y, Yu HY, Zhang YY,
Fan F, Schneider HG, Duffy SJ, Taylor AJ, Du XJ, Gao W, Gao XM,
Dart AM. Macrophage migration inhibitory factor for the early predic-
tion of infarct size. J Am Heart Assoc 2: e000226, 2013. doi:10.1161/
JAHA.113.000226.
48. Chen C, Chen L, Fallon JT, Ma L, Li L, Bow L, Knibbs D, McKay
R, Gillam LD, Waters DD. Functional and structural alterations with
24-hour myocardial hibernation and recovery after reperfusion. A pig
model of myocardial hibernation. Circulation 94: 507–516, 1996. doi:
10.1161/01.CIR.94.3.507.
49. Cho SW, Gwak SJ, Kim IK, Cho MC, Hwang KK, Kwon JS, Choi
CY, Yoo KJ, Kim BS. Granulocyte colony-stimulating factor treatment
enhances the efficacy of cellular cardiomyoplasty with transplantation of
embryonic stem cell-derived cardiomyocytes in infarcted myocardium.
Biochem Biophys Res Commun 340: 573–582, 2006. doi:10.1016/j.bbrc.
2005.12.044.
50. Christia P, Bujak M, Gonzalez-Quesada C, Chen W, Dobaczewski
M, Reddy A, Frangogiannis NG. Systematic characterization of myo-
cardial inflammation, repair, and remodeling in a mouse model of
reperfused myocardial infarction. J Histochem Cytochem 61: 555–570,
2013. doi:10.1369/0022155413493912.
51. Chung Y. Myocardial PO
2
does not limit aerobic metabolism in the
postischemic heart. Am J Physiol Heart Circ Physiol 310: H226 –H238,
2016. doi:10.1152/ajpheart.00335.2015.
52. Ciulla MM, Paliotti R, Ferrero S, Braidotti P, Esposito A, Gianelli U,
Busca G, Cioffi U, Bulfamante G, Magrini F. Left ventricular remod-
eling after experimental myocardial cryoinjury in rats. J Surg Res 116:
91–97, 2004. doi:10.1016/j.jss.2003.08.238.
53. Cohen M, Boiangiu C, Abidi M. Therapy for ST-segment elevation
myocardial infarction patients who present late or are ineligible for
reperfusion therapy. J Am Coll Cardiol 55: 1895–1906, 2010. doi:10.
1016/j.jacc.2009.11.087.
54. Costa AR, Panda NC, Yong S, Mayorga ME, Pawlowski GP, Fan K,
Penn MS, Laurita KR. Optical mapping of cryoinjured rat myocardium
grafted with mesenchymal stem cells. Am J Physiol Heart Circ Physiol
302: H270 –H277, 2012. doi:10.1152/ajpheart.00019.2011.
55. Cullen LA, Mills NL, Mahler S, Body R. Early rule-out and rule-in
strategies for myocardial infarction. Clin Chem 63: 129 –139, 2017.
doi:10.1373/clinchem.2016.254730.
56. Curtis MJ, Hancox JC, Farkas A, Wainwright CL, Stables CL, Saint
DA, Clements-Jewery H, Lambiase PD, Billman GE, Janse MJ,
Pugsley MK, Ng GA, Roden DM, Camm AJ, Walker MJ. The
Lambeth Conventions (II): guidelines for the study of animal and human
ventricular and supraventricular arrhythmias. Pharmacol Ther 139: 213–
248, 2013. doi:10.1016/j.pharmthera.2013.04.008.
57. Darehzereshki A, Rubin N, Gamba L, Kim J, Fraser J, Huang Y,
Billings J, Mohammadzadeh R, Wood J, Warburton D, Kaartinen V,
Lien CL. Differential regenerative capacity of neonatal mouse hearts
after cryoinjury. Dev Biol 399: 91–99, 2015. doi:10.1016/j.ydbio.2014.
12.018.
58. Date T, Belanger AJ, Mochizuki S, Sullivan JA, Liu LX, Scaria A,
Cheng SH, Gregory RJ, Jiang C. Adenovirus-mediated expression of
p35 prevents hypoxia/reoxygenation injury by reducing reactive oxygen
species and caspase activity. Cardiovasc Res 55: 309 –319, 2002. doi:
10.1016/S0008-6363(02)00412-1.
59. De Jesus NM, Wang L, Lai J, Rigor RR, Francis Stuart SD, Bers
DM, Lindsey ML, Ripplinger CM. Anti-arrhythmic effects of interleu-
kin-1 inhibition following myocardial infarction. Heart Rhythm 14:
727–736, 2017. doi:10.1016/j.hrthm.2017.01.027.
60. DeLeon-Pennell KY, de Castro Brás LE, Iyer RP, Bratton DR, Jin
YF, Ripplinger CM, Lindsey ML. P. gingivalis lipopolysaccharide
intensifies inflammation post-myocardial infarction through matrix met-
alloproteinase-9. J Mol Cell Cardiol 76: 218 –226, 2014. doi:10.1016/j.
yjmcc.2014.09.007.
61. DeLeon-Pennell KY, Tian Y, Zhang B, Cates CA, Iyer RP, Cannon
P, Shah P, Aiyetan P, Halade GV, Ma Y, Flynn E, Zhang Z, Jin YF,
Zhang H, Lindsey ML. CD36 is a matrix metalloproteinase-9 sub-
strate that stimulates neutrophil apoptosis and removal during cardiac
remodeling. Circ Cardiovasc Genet 9: 14 –25, 2016. doi:10.1161/
CIRCGENETICS.115.001249.
62. Depre C, Kim SJ, John AS, Huang Y, Rimoldi OE, Pepper JR,
Dreyfus GD, Gaussin V, Pennell DJ, Vatner DE, Camici PG, Vatner
SF. Program of cell survival underlying human and experimental hiber-
nating myocardium. Circ Res 95: 433–440, 2004. doi:10.1161/01.RES.
0000138301.42713.18.
63. Dewald O, Frangogiannis NG, Zoerlein M, Duerr GD, Klemm C,
Knuefermann P, Taffet G, Michael LH, Crapo JD, Welz A, Entman
ML. Development of murine ischemic cardiomyopathy is associated
with a transient inflammatory reaction and depends on reactive oxygen
species. Proc Natl Acad Sci USA 100: 2700 –2705, 2003. doi:10.1073/
pnas.0438035100.
64. Dewald O, Ren G, Duerr GD, Zoerlein M, Klemm C, Gersch C,
Tincey S, Michael LH, Entman ML, Frangogiannis NG. Of mice and
dogs: species-specific differences in the inflammatory response following
myocardial infarction. Am J Pathol 164: 665–677, 2004. doi:10.1016/
S0002-9440(10)63154-9.
65. Dewald O, Zymek P, Winkelmann K, Koerting A, Ren G, Abou-
Khamis T, Michael LH, Rollins BJ, Entman ML, Frangogiannis NG.
CCL2/monocyte chemoattractant protein-1 regulates inflammatory re-
sponses critical to healing myocardial infarcts. Circ Res 96: 881–889,
2005. doi:10.1161/01.RES.0000163017.13772.3a.
66. Domenech RJ, Hoffman JI, Noble MI, Saunders KB, Henson JR,
Subijanto S. Total and regional coronary blood flow measured by
radioactive microspheres in conscious and anesthetized dogs. Circ Res
25: 581–596, 1969. doi:10.1161/01.RES.25.5.581.
H830 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
67. Dörge H, Neumann T, Behrends M, Skyschally A, Schulz R, Kasper
C, Erbel R, Heusch G. Perfusion-contraction mismatch with coronary
microvascular obstruction: role of inflammation. Am J Physiol Heart
Circ Physiol 279: H2587–H2592, 2000. doi:10.1152/ajpheart.2000.279.
6.H2587.
68. Downey JM, Kirk ES. Inhibition of coronary blood flow by a vascular
waterfall mechanism. Circ Res 36: 753–760, 1975. doi:10.1161/01.RES.
36.6.753.
69. Du CK, Zhan DY, Akiyama T, Inagaki T, Shishido T, Shirai M,
Pearson JT. Myocardial interstitial levels of serotonin and its major
metabolite 5-hydroxyindole acetic acid during ischemia-reperfusion. Am
J Physiol Heart Circ Physiol 312: H60 –H67, 2017. doi:10.1152/ajpheart.
00471.2016.
70. Du XJ, Cox HS, Dart AM, Esler MD. Sympathetic activation
triggers ventricular arrhythmias in rat heart with chronic infarction
and failure. Cardiovasc Res 43: 919 –929, 1999. doi:10.1016/S0008-
6363(99)00139-X.
71. Dutta P, Hoyer FF, Grigoryeva LS, Sager HB, Leuschner F, Courties
G, Borodovsky A, Novobrantseva T, Ruda VM, Fitzgerald K,
Iwamoto Y, Wojtkiewicz G, Sun Y, Da Silva N, Libby P, Anderson
DG, Swirski FK, Weissleder R, Nahrendorf M. Macrophages retain
hematopoietic stem cells in the spleen via VCAM-1. J Exp Med 212:
497–512, 2015. doi:10.1084/jem.20141642.
72. Dutta P, Hoyer FF, Sun Y, Iwamoto Y, Tricot B, Weissleder R,
Magnani JL, Swirski FK, Nahrendorf M. E-selectin inhibition miti-
gates splenic HSC activation and myelopoiesis in hypercholesterolemic
mice with myocardial infarction. Arterioscler Thromb Vasc Biol 36:
1802–1808, 2016. doi:10.1161/ATVBAHA.116.307519.
73. Fallavollita JA, Canty JM Jr. Differential
18
F-2-deoxyglucose uptake
in viable dysfunctional myocardium with normal resting perfusion:
evidence for chronic stunning in pigs. Circulation 99: 2798 –2805, 1999.
doi:10.1161/01.CIR.99.21.2798.
74. Fallavollita JA, Canty JM Jr. Ischemic cardiomyopathy in pigs with
two-vessel occlusion and viable, chronically dysfunctional myocardium.
Am J Physiol Heart Circ Physiol 282: H1370 –H1379, 2002. doi:10.
1152/ajpheart.00138.2001.
75. Fallavollita JA, Logue M, Canty JM Jr. Stability of hibernating
myocardium in pigs with a chronic left anterior descending coronary
artery stenosis: absence of progressive fibrosis in the setting of stable
reductions in flow, function and coronary flow reserve. J Am Coll Cardiol
37: 1989 –1995, 2001. doi:10.1016/S0735-1097(01)01250-5.
76. Fallavollita JA, Malm BJ, Canty JM Jr. Hibernating myocardium
retains metabolic and contractile reserve despite regional reductions in
flow, function, and oxygen consumption at rest. Circ Res 92: 48–55,
2003. doi:10.1161/01.RES.0000049104.57549.03.
77. Fallavollita JA, Perry BJ, Canty JM Jr.
18
F-2-deoxyglucose deposi-
tion and regional flow in pigs with chronically dysfunctional myocar-
dium. Evidence for transmural variations in chronic hibernating myocar-
dium. Circulation 95: 1900 –1909, 1997. doi:10.1161/01.CIR.95.7.1900.
78. Ferdinandy P, Hausenloy DJ, Heusch G, Baxter GF, Schulz R.
Interaction of risk factors, comorbidities, and comedications with isch-
emia/reperfusion injury and cardioprotection by preconditioning, post-
conditioning, and remote conditioning. Pharmacol Rev 66: 1142–1174,
2014. doi:10.1124/pr.113.008300.
79. Ferrera R, Benhabbouche S, Bopassa JC, Li B, Ovize M. One hour
reperfusion is enough to assess function and infarct size with TTC
staining in Langendorff rat model. Cardiovasc Drugs Ther 23: 327–331,
2009. doi:10.1007/s10557-009-6176-5.
80. Firoozan S, Wei K, Linka A, Skyba D, Goodman NC, Kaul S. A
canine model of chronic ischemic cardiomyopathy: characterization of
regional flow-function relations. Am J Physiol Heart Circ Physiol 276:
H446 –H455, 1999.
81. Fishbein MC, Meerbaum S, Rit J, Lando U, Kanmatsuse K, Mercier
JC, Corday E, Ganz W. Early phase acute myocardial infarct size
quantification: validation of the triphenyl tetrazolium chloride tissue
enzyme staining technique. Am Heart J 101: 593–600, 1981. doi:10.
1016/0002-8703(81)90226-X.
82. Fomovsky GM, Rouillard AD, Holmes JW. Regional mechanics de-
termine collagen fiber structure in healing myocardial infarcts. J Mol Cell
Cardiol 52: 1083–1090, 2012. doi:10.1016/j.yjmcc.2012.02.012.
83. Francis Stuart SD, De Jesus NM, Lindsey ML, Ripplinger CM. The
crossroads of inflammation, fibrosis, and arrhythmia following myocar-
dial infarction. J Mol Cell Cardiol 91: 114 –122, 2016. doi:10.1016/j.
yjmcc.2015.12.024.
84. Frangogiannis NG. The immune system and the remodeling infarcted
heart: cell biological insights and therapeutic opportunities. J Cardiovasc
Pharmacol 63: 185–195, 2014. doi:10.1097/FJC.0000000000000003.
85. Frangogiannis NG. Inflammation in cardiac injury, repair and regener-
ation. Curr Opin Cardiol 30: 240 –245, 2015. doi:10.1097/HCO.
0000000000000158.
86. Frangogiannis NG. Pathophysiology of myocardial infarction. Compr
Physiol 5: 1841–1875, 2015. doi:10.1002/cphy.c150006.
87. Frangogiannis NG, Michael LH, Entman ML. Myofibroblasts in
reperfused myocardial infarcts express the embryonic form of smooth
muscle myosin heavy chain (SMemb). Cardiovasc Res 48: 89 –100,
2000. doi:10.1016/S0008-6363(00)00158-9.
88. Frantz S, Bauersachs J, Ertl G. Post-infarct remodelling: contribution
of wound healing and inflammation. Cardiovasc Res 81: 474 –481, 2009.
doi:10.1093/cvr/cvn292.
89. Frey A, Saxon VM, Popp S, Lehmann M, Mathes D, Pachel C,
Hofmann U, Ertl G, Lesch KP, Frantz S. Early citalopram treatment
increases mortality due to left ventricular rupture in mice after myocar-
dial infarction. J Mol Cell Cardiol 98: 28 –36, 2016. doi:10.1016/j.yjmcc.
2016.07.002.
90. Galaup A, Gomez E, Souktani R, Durand M, Cazes A, Monnot C,
Teillon J, Le Jan S, Bouleti C, Briois G, Philippe J, Pons S, Martin
V, Assaly R, Bonnin P, Ratajczak P, Janin A, Thurston G, Valen-
zuela DM, Murphy AJ, Yancopoulos GD, Tissier R, Berdeaux A,
Ghaleh B, Germain S. Protection against myocardial infarction and
no-reflow through preservation of vascular integrity by angiopoietin-like
4. Circulation 125: 140 –149, 2012. doi:10.1161/CIRCULATIONAHA.
111.049072.
91. Gallagher KP, Matsuzaki M, Koziol JA, Kemper WS, Ross J Jr.
Regional myocardial perfusion and wall thickening during ischemia in
conscious dogs. Am J Physiol Heart Circ Physiol 247: H727–H738,
1984.
92. Gallagher KP, Matsuzaki M, Osakada G, Kemper WS, Ross J Jr.
Effect of exercise on the relationship between myocardial blood flow and
systolic wall thickening in dogs with acute coronary stenosis. Circ Res
52: 716 –729, 1983. doi:10.1161/01.RES.52.6.716.
93. Gallagher KP, Osakada G, Matsuzaki M, Kemper WS, Ross J Jr.
Myocardial blood flow and function with critical coronary stenosis in
exercising dogs. Am J Physiol Heart Circ Physiol 243: H698 –H707,
1982.
94. Ganote CE, Humphrey SM. Effects of anoxic or oxygenated reperfu-
sion in globally ischemic, isovolumic, perfused rat hearts. Am J Pathol
120: 129 –145, 1985.
95. Gao XM, Moore XL, Liu Y, Wang XY, Han LP, Su Y, Tsai A, Xu Q,
Zhang M, Lambert GW, Kiriazis H, Gao W, Dart AM, Du XJ.
Splenic release of platelets contributes to increased circulating platelet
size and inflammation after myocardial infarction. Clin Sci (Lond) 130:
1089 –1104, 2016. doi:10.1042/CS20160234.
96. Gao XM, White DA, Dart AM, Du XJ. Post-infarct cardiac rupture:
recent insights on pathogenesis and therapeutic interventions. Pharmacol
Ther 134: 156 –179, 2012. doi:10.1016/j.pharmthera.2011.12.010.
97. Gao XM, Wu QZ, Kiriazis H, Su Y, Han LP, Pearson JT, Taylor AJ,
Du XJ. Microvascular leakage in acute myocardial infarction: charac-
terization by histology, biochemistry, and magnetic resonance imaging.
Am J Physiol Heart Circ Physiol 312: H1068 –H1075, 2017. doi:10.
1152/ajpheart.00073.2017.
98. Gao XM, Xu Q, Kiriazis H, Dart AM, Du XJ. Mouse model of
post-infarct ventricular rupture: time course, strain- and gender-depen-
dency, tensile strength, and histopathology. Cardiovasc Res 65: 469–
477, 2005. doi:10.1016/j.cardiores.2004.10.014.
99. Gardner RT, Habecker BA. Infarct-derived chondroitin sulfate pro-
teoglycans prevent sympathetic reinnervation after cardiac ischemia-
reperfusion injury. J Neurosci 33: 7175–7183, 2013. doi:10.1523/
JNEUROSCI.5866-12.2013.
100. Gardner RT, Wang L, Lang BT, Cregg JM, Dunbar CL, Woodward
WR, Silver J, Ripplinger CM, Habecker BA. Targeting protein ty-
rosine phosphatase after myocardial infarction restores cardiac sym-
pathetic innervation and prevents arrhythmias. Nat Commun 6: 6235,
2015. doi:10.1038/ncomms7235.
101. Gedik N, Krüger M, Thielmann M, Kottenberg E, Skyschally A,
Frey UH, Cario E, Peters J, Jakob H, Heusch G, Kleinbongard P.
Proteomics/phosphoproteomics of left ventricular biopsies from patients
with surgical coronary revascularization and pigs with coronary occlu-
H831MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
sion/reperfusion: remote ischemic preconditioning. Sci Rep 7: 7629,
2017. doi:10.1038/s41598-017-07883-5.
102. Gedik N, Maciel L, Schulte C, Skyschally A, Heusch G, Kleinbon-
gard P. Cardiomyocyte mitochondria as targets of humoral factors
released by remote ischemic preconditioning. Arch Med Sci 13: 448–
458, 2017. doi:10.5114/aoms.2016.61789.
103. Gent S, Skyschally A, Kleinbongard P, Heusch G. lschemic precon-
ditioning in pigs: a causal role for signal transducer and activator of
transcription 3. Am J Physiol Heart Circ Physiol 312: H478 –H484, 2017.
doi:10.1152/ajpheart.00749.2016.
104. Gharacholou SM, Alexander KP, Chen AY, Wang TY, Melloni C,
Gibler WB, Pollack CV Jr, Ohman EM, Peterson ED, Roe MT.
Implications and reasons for the lack of use of reperfusion therapy in
patients with ST-segment elevation myocardial infarction: findings from
the CRUSADE initiative. Am Heart J 159: 757–763, 2010. doi:10.1016/
j.ahj.2010.02.009.
105. Gheorghiade M, Bonow RO. Chronic heart failure in the United States:
a manifestation of coronary artery disease. Circulation 97: 282–289,
1998. doi:10.1161/01.CIR.97.3.282.
106. Girod WG, Jones SP, Sieber N, Aw TY, Lefer DJ. Effects of hyper-
cholesterolemia on myocardial ischemia-reperfusion injury in LDL re-
ceptor-deficient mice. Arterioscler Thromb Vasc Biol 19: 2776–2781,
1999. doi:10.1161/01.ATV.19.11.2776.
107. Glenny RW, Bernard S, Brinkley M. Validation of fluorescent-labeled
microspheres for measurement of regional organ perfusion. J Appl
Physiol 74: 2585–2597, 1993. doi:10.1152/jappl.1993.74.5.2585.
108. González-Rosa JM, Martín V, Peralta M, Torres M, Mercader N.
Extensive scar formation and regression during heart regeneration after
cryoinjury in zebrafish. Development 138: 1663–1674, 2011. doi:10.
1242/dev.060897.
109. Grisel P, Meinhardt A, Lehr HA, Kappenberger L, Barrandon Y,
Vassalli G. The MRL mouse repairs both cryogenic and ischemic
myocardial infarcts with scar. Cardiovasc Pathol 17: 14 –22, 2008.
doi:10.1016/j.carpath.2007.01.007.
110. Gumz ML. Taking into account circadian rhythm when conducting
experiments on animals. Am J Physiol Renal Physiol 310: F454 –F455,
2016. doi:10.1152/ajprenal.00549.2015.
111. Guo Y, Flaherty MP, Wu WJ, Tan W, Zhu X, Li Q, Bolli R. Genetic
background, gender, age, body temperature, and arterial blood pH have
a major impact on myocardial infarct size in the mouse and need to be
carefully measured and/or taken into account: results of a comprehensive
analysis of determinants of infarct size in 1,074 mice. Basic Res Cardiol
107: 288, 2012. doi:10.1007/s00395-012-0288-y.
112. Guth BD, Heusch G, Seitelberger R, Ross J Jr. Elimination of
exercise-induced regional myocardial dysfunction by a bradycardiac
agent in dogs with chronic coronary stenosis. Circulation 75: 661–669,
1987. doi:10.1161/01.CIR.75.3.661.
113. Guth BD, Indolfi C, Heusch G, Seitelberger R, Ross J Jr. Mechanisms
of benefit in the ischemic myocardium due to heart rate reduction. Basic
Res Cardiol 85, Suppl 1: 157–166, 1990.
114. Guth BD, Martin JF, Heusch G, Ross J Jr. Regional myocardial blood
flow, function and metabolism using phosphorus-31 nuclear magnetic
resonance spectroscopy during ischemia and reperfusion in dogs. JAm
Coll Cardiol 10: 673–681, 1987. doi:10.1016/S0735-1097(87)80212-7.
115. Halade GV, Kain V, Black LM, Prabhu SD, Ingle KA. Aging
dysregulates D- and E-series resolvins to modulate cardiosplenic and
cardiorenal network following myocardial infarction. Aging (Albany NY)
8: 2611–2634, 2016. doi:10.18632/aging.101077.
116. Halade GV, Kain V, Ingle KA. Heart functional and structural com-
pendium of cardiosplenic and cardiorenal networks in acute and chronic
heart failure pathology. Am J Physiol Heart Circ Physiol 314:
H255H267, 2018. doi:10.1152/ajpheart.00528.2017.
117. Halade GV, Kain V, Ingle KA, Prabhu SD. Interaction of 12/15-
lipoxygenase with fatty acids alters the leukocyte kinetics leading to
improved postmyocardial infarction healing. Am J Physiol Heart Circ
Physiol 313: H89H102, 2017. doi:10.1152/ajpheart.00040.2017.
118. Hausenloy DJ, Barrabes JA, Bøtker HE, Davidson SM, Di Lisa F,
Downey J, Engstrom T, Ferdinandy P, Carbrera-Fuentes HA, Heu-
sch G, Ibanez B, Iliodromitis EK, Inserte J, Jennings R, Kalia N,
Kharbanda R, Lecour S, Marber M, Miura T, Ovize M, Perez-
Pinzon MA, Piper HM, Przyklenk K, Schmidt MR, Redington A,
Ruiz-Meana M, Vilahur G, Vinten-Johansen J, Yellon DM, Garcia-
Dorado D. Ischaemic conditioning and targeting reperfusion injury: a
30 year voyage of discovery. Basic Res Cardiol 111: 70, 2016. doi:10.
1007/s00395-016-0588-8.
119. Hausenloy DJ, Bøtker HE, Engstrom T, Erlinge D, Heusch G, Ibanez
B, Kloner RA, Ovize M, Yellon DM, Garcia-Dorado D. Targeting
reperfusion injury in patients with ST-segment elevation myocardial
infarction: trials and tribulations. Eur Heart J 38: 935–941, 2017.
doi:10.1093/eurheartj/ehw145.
120. Hausenloy DJ, Garcia-Dorado D, Bøtker HE, Davidson SM, Downey
J, Engel FB, Jennings R, Lecour S, Leor J, Madonna R, Ovize M,
Perrino C, Prunier F, Schulz R, Sluijter JPG, Van Laake LW,
Vinten-Johansen J, Yellon DM, Ytrehus K, Heusch G, Ferdinandy P.
Novel targets and future strategies for acute cardioprotection: Position
Paper of the European Society of Cardiology Working Group on Cellular
Biology of the Heart. Cardiovasc Res 113: 564 –585, 2017. doi:10.1093/
cvr/cvx049.
121. Hausenloy DJ, Yellon DM. Myocardial ischemia-reperfusion injury: a
neglected therapeutic target. J Clin Invest 123: 92–100, 2013. doi:10.
1172/JCI62874.
122. Hawkins HK, Entman ML, Zhu JY, Youker KA, Berens K, Doré M,
Smith CW. Acute inflammatory reaction after myocardial ischemic
injury and reperfusion. Development and use of a neutrophil-specific
antibody. Am J Pathol 148: 1957–1969, 1996.
123. Heaberlin JR, Ma Y, Zhang J, Ahuja SS, Lindsey ML, Halade GV.
Obese and diabetic KKAy mice show increased mortality but improved
cardiac function following myocardial infarction. Cardiovasc Pathol 22:
481–487, 2013. doi:10.1016/j.carpath.2013.06.002.
124. Heinzel FR, Luo Y, Dodoni G, Boengler K, Petrat F, Di Lisa F, de
Groot H, Schulz R, Heusch G. Formation of reactive oxygen species at
increased contraction frequency in rat cardiomyocytes. Cardiovasc Res
71: 374 –382, 2006. doi:10.1016/j.cardiores.2006.05.014.
125. Heinzel FR, Luo Y, Li X, Boengler K, Buechert A, García-Dorado D,
Di Lisa F, Schulz R, Heusch G. Impairment of diazoxide-induced
formation of reactive oxygen species and loss of cardioprotection in
connexin 43 deficient mice. Circ Res 97: 583–586, 2005. doi:10.1161/
01.RES.0000181171.65293.65.
126. Heusch G. The coronary circulation as a target of cardioprotection. Circ
Res 118: 1643–1658, 2016. doi:10.1161/CIRCRESAHA.116.308640.
127. Heusch G. Critical issues for the translation of cardioprotection. Circ Res
120: 1477–1486, 2017. doi:10.1161/CIRCRESAHA.117.310820.
128. Heusch G. Molecular basis of cardioprotection: signal transduction in
ischemic pre-, post-, and remote conditioning. Circ Res 116: 674 –699,
2015. doi:10.1161/CIRCRESAHA.116.305348.
129. Heusch G. Myocardial ischemia: lack of coronary blood flow or myo-
cardial oxygen supply/demand imbalance? Circ Res 119: 194 –196, 2016.
doi:10.1161/CIRCRESAHA.116.308925.
130. Heusch G. The regional myocardial flow-function relationship: a frame-
work for an understanding of acute ischemia, hibernation, stunning and
coronary microembolization. 1980. Circ Res 112: 1535–1537, 2013.
doi:10.1161/CIRCRESAHA.113.301446.
131. Heusch G, Deussen A. The effects of cardiac sympathetic nerve stim-
ulation on perfusion of stenotic coronary arteries in the dog. Circ Res 53:
8 –15, 1983. doi:10.1161/01.RES.53.1.8.
132. Heusch G, Deussen A, Thämer V. Cardiac sympathetic nerve activity
and progressive vasoconstriction distal to coronary stenoses: feed-back
aggravation of myocardial ischemia. J Auton Nerv Syst 13: 311–326,
1985. doi:10.1016/0165-1838(85)90020-7.
133. Heusch G, Gersh BJ. The pathophysiology of acute myocardial infarc-
tion and strategies of protection beyond reperfusion: a continual chal-
lenge. Eur Heart J 38: 774 –784, 2017.
134. Heusch G, Guth BD, Seitelberger R, Ross J Jr. Attenuation of
exercise-induced myocardial ischemia in dogs with recruitment of cor-
onary vasodilator reserve by nifedipine. Circulation 75: 482–490, 1987.
doi:10.1161/01.CIR.75.2.482.
135. Heusch G, Kleinbongard P, Böse D, Levkau B, Haude M, Schulz
R, Erbel R. Coronary microembolization: from bedside to bench and
back to bedside. Circulation 120: 1822–1836, 2009. doi:10.1161/
CIRCULATIONAHA.109.888784.
136. Heusch G, Libby P, Gersh B, Yellon D, Böhm M, Lopaschuk G, Opie
L. Cardiovascular remodelling in coronary artery disease and heart
failure. Lancet 383: 1933–1943, 2014. doi:10.1016/S0140-6736(14)
60107-0.
137. Heusch G, Post H, Michel MC, Kelm M, Schulz R. Endogenous nitric
oxide and myocardial adaptation to ischemia. Circ Res 87: 146–152,
2000. doi:10.1161/01.RES.87.2.146.
H832 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
138. Heusch G, Rassaf T. Time to give up on cardioprotection? A critical
appraisal of clinical studies on ischemic pre-, post-, and remote condi-
tioning. Circ Res 119: 676 –695, 2016. doi:10.1161/CIRCRESAHA.116.
308736.
139. Heusch G, Schulz R, Rahimtoola SH. Myocardial hibernation: a
delicate balance. Am J Physiol Heart Circ Physiol 288: H984–H999,
2005. doi:10.1152/ajpheart.01109.2004.
140. Heusch G, Skyschally A, Schulz R. The in-situ pig heart with regional
ischemia/reperfusion–ready for translation. J Mol Cell Cardiol 50: 951–
963, 2011. doi:10.1016/j.yjmcc.2011.02.016.
141. Heusch P, Nensa F, Heusch G. Is MRI really the gold standard for the
quantification of salvage from myocardial infarction? Circ Res 117:
222–224, 2015. doi:10.1161/CIRCRESAHA.117.306929.
142. Heymann MA, Payne BD, Hoffman JIE, Rudolph AM. Blood flow
measurements with radionuclide-labeled particles. Prog Cardiovasc Dis
20: 55–79, 1977. doi:10.1016/S0033-0620(77)80005-4.
143. Heyndrickx GR, Baig H, Nellens P, Leusen I, Fishbein MC, Vatner
SF. Depression of regional blood flow and wall thickening after brief
coronary occlusions. Am J Physiol Heart Circ Physiol 234: H653–H659,
1978.
144. Homans DC, Sublett E, Dai XZ, Bache RJ. Persistence of regional left
ventricular dysfunction after exercise-induced myocardial ischemia. J
Clin Invest 77: 66 –73, 1986. doi:10.1172/JCI112303.
145. House SL, Castro AM, Lupu TS, Weinheimer C, Smith C, Kovacs A,
Ornitz DM. Endothelial fibroblast growth factor receptor signaling is
required for vascular remodeling following cardiac ischemia-reperfusion
injury. Am J Physiol Heart Circ Physiol 310: H559 –H571, 2016.
doi:10.1152/ajpheart.00758.2015.
146. Hunter I, Soler A, Joseph G, Hutcheson B, Bradford C, Zhang FF,
Potter B, Proctor S, Rocic P. Cardiovascular function in male and
female JCR:LA-cp rats: effect of high-fat/high-sucrose diet. Am J Physiol
Heart Circ Physiol 312: H742–H751, 2017. doi:10.1152/ajpheart.00535.
2016.
147. Ibáñez B, Heusch G, Ovize M, Van de Werf F. Evolving therapies for
myocardial ischemia/reperfusion injury. J Am Coll Cardiol 65: 1454 –
1471, 2015. doi:10.1016/j.jacc.2015.02.032.
148. Ibanez B, Prat-González S, Speidl WS, Vilahur G, Pinero A, Cim-
mino G, García MJ, Fuster V, Sanz J, Badimon JJ. Early metoprolol
administration before coronary reperfusion results in increased myocar-
dial salvage: analysis of ischemic myocardium at risk using cardiac
magnetic resonance. Circulation 115: 2909 –2916, 2007. doi:10.1161/
CIRCULATIONAHA.106.679639.
149. Iop L, Chiavegato A, Callegari A, Bollini S, Piccoli M, Pozzobon M,
Rossi CA, Calamelli S, Chiavegato D, Gerosa G, De Coppi P, Sartore
S. Different cardiovascular potential of adult- and fetal-type mesenchy-
mal stem cells in a rat model of heart cryoinjury. Cell Transplant 17:
679 –694, 2008. doi:10.3727/096368908786092739.
150. Ip WT, McAlindon A, Miller SE, Bell JR, Curl CL, Huggins CE,
Mellor KM, Raaijmakers AJ, Bienvenu LA, McLennan PL, Pepe S,
Delbridge LM. Dietary omega-6 fatty acid replacement selectively
impairs cardiac functional recovery after ischemia in female (but not
male) rats. Am J Physiol Heart Circ Physiol 311: H768 –H780, 2016.
doi:10.1152/ajpheart.00690.2015.
151. Isorni MA, Casanova A, Piquet J, Bellamy V, Pignon C, Puymirat E,
Menasche P. Comparative analysis of methods to induce myocardial
infarction in a closed-chest rabbit model. BioMed Res Int 2015: 893051,
2015. doi:10.1155/2015/893051.
152. Iyer RP, de Castro Brás LE, Cannon PL, Ma Y, DeLeon-Pennell KY,
Jung M, Flynn ER, Henry JB, Bratton DR, White JA, Fulton LK,
Grady AW, Lindsey ML. Defining the sham environment for post-
myocardial infarction studies in mice. Am J Physiol Heart Circ Physiol
311: H822–H836, 2016. doi:10.1152/ajpheart.00067.2016.
153. Iyer RP, de Castro Brás LE, Patterson NL, Bhowmick M, Flynn ER,
Asher M, Cannon PL, Deleon-Pennell KY, Fields GB, Lindsey ML.
Early matrix metalloproteinase-9 inhibition post-myocardial infarction
worsens cardiac dysfunction by delaying inflammation resolution. J Mol
Cell Cardiol 100: 109 –117, 2016. doi:10.1016/j.yjmcc.2016.10.005.
154. Iyer RP, Jung M, Lindsey ML. MMP-9 signaling in the left ventricle
following myocardial infarction. Am J Physiol Heart Circ Physiol 311:
H190 –H198, 2016. doi:10.1152/ajpheart.00243.2016.
155. Iyer RP, Patterson NL, Zouein FA, Ma Y, Dive V, de Castro Brás
LE, Lindsey ML. Early matrix metalloproteinase-12 inhibition worsens
post-myocardial infarction cardiac dysfunction by delaying inflammation
resolution. Int J Cardiol 185: 198 –208, 2015. doi:10.1016/j.ijcard.2015.
03.054.
156. Jennings RB, Murry CE, Steenbergen C Jr, Reimer KA. Develop-
ment of cell injury in sustained acute ischemia. Circulation 82, Suppl:
II2–II12, 1990.
157. Jennings RB, Reimer KA. Lethal myocardial ischemic injury. Am J
Pathol 102: 241–255, 1981.
158. Jennings RB, Schaper J, Hill ML, Steenbergen C Jr, Reimer KA.
Effect of reperfusion late in the phase of reversible ischemic injury.
Changes in cell volume, electrolytes, metabolites, and ultrastructure. Circ
Res 56: 262–278, 1985. doi:10.1161/01.RES.56.2.262.
159. Jennings RB, Sommers HM, Smyth GA, Flack HA, Linn H. Myo-
cardial necrosis induced by temporary occlusion of a coronary artery in
the dog. Arch Pathol 70: 68 –78, 1960.
160. Jensen JA, Kosek JC, Hunt TK, Goodson WH 3rd, Miller DC.
Cardiac cryolesions as an experimental model of myocardial wound healing.
Ann Surg 206: 798 –803, 1987. doi:10.1097/00000658-198712000-00019.
161. Jian Z, Chen YJ, Shimkunas R, Jian Y, Jaradeh M, Chavez K,
Chiamvimonvat N, Tardiff JC, Izu LT, Ross RS, Chen-Izu Y. In vivo
cannulation methods for cardiomyocytes isolation from heart disease
models. PLoS One 11: e0160605, 2016. doi:10.1371/journal.pone.
0160605.
162. Jones SP. I’ll have the rigor, but hold the mortis. Circ Res 120:
1852–1854, 2017. doi:10.1161/CIRCRESAHA.117.311114.
163. Jones SP, Tang XL, Guo Y, Steenbergen C, Lefer DJ, Kukreja RC,
Kong M, Li Q, Bhushan S, Zhu X, Du J, Nong Y, Stowers HL, Kondo
K, Hunt GN, Goodchild TT, Orr A, Chang CC, Ockaili R, Salloum
FN, Bolli R. The NHLBI-sponsored Consortium for preclinicAl assESs-
ment of cARdioprotective therapies (CAESAR): a new paradigm for
rigorous, accurate, and reproducible evaluation of putative infarct-spar-
ing interventions in mice, rabbits, and pigs. Circ Res 116: 572–586, 2015.
doi:10.1161/CIRCRESAHA.116.305462.
164. Jorge E, Amorós-Figueras G, García-Sánchez T, Bragós R, Rosell-
Ferrer J, Cinca J. Early detection of acute transmural myocardial
ischemia by the phasic systolic-diastolic changes of local tissue electrical
impedance. Am J Physiol Heart Circ Physiol 310: H436 –H443, 2016.
doi:10.1152/ajpheart.00754.2015.
165. Jung M, Ma Y, Iyer RP, DeLeon-Pennell KY, Yabluchanskiy A,
Garrett MR, Lindsey ML. IL-10 improves cardiac remodeling after
myocardial infarction by stimulating M2 macrophage polarization and
fibroblast activation. Basic Res Cardiol 112: 33, 2017. doi:10.1007/
s00395-017-0622-5.
166. Kang PM, Haunstetter A, Aoki H, Usheva A, Izumo S. Morphological
and molecular characterization of adult cardiomyocyte apoptosis during
hypoxia and reoxygenation. Circ Res 87: 118 –125, 2000. doi:10.1161/
01.RES.87.2.118.
167. Khemtong C, Carpenter NR, Lumata LL, Merritt ME, Moreno KX,
Kovacs Z, Malloy CR, Sherry AD. Hyperpolarized 13C NMR detects
rapid drug-induced changes in cardiac metabolism. Magn Reson Med 74:
312–319, 2015. doi:10.1002/mrm.25419.
168. Kilkenny C, Browne WJ, Cuthill IC, Emerson M, Altman DG.
Improving bioscience research reporting: the ARRIVE guidelines for
reporting animal research. PLoS Biol 8: e1000412, 2010. doi:10.1371/
journal.pbio.1000412.
169. Kim SJ, Peppas A, Hong SK, Yang G, Huang Y, Diaz G, Sa-
doshima J, Vatner DE, Vatner SF. Persistent stunning induces
myocardial hibernation and protection: flow/function and metabolic
mechanisms. Circ Res 92: 1233–1239, 2003. doi:10.1161/01.RES.
0000076892.18394.B6.
170. Kingery JR, Hamid T, Lewis RK, Ismahil MA, Bansal SS, Rokosh G,
Townes TM, Ildstad ST, Jones SP, Prabhu SD. Leukocyte iNOS is
required for inflammation and pathological remodeling in ischemic heart
failure. Basic Res Cardiol 112: 19, 2017. doi:10.1007/s00395-017-
0609-2.
171. Klein GJ, Harrison L, Ideker RF, Smith WM, Kasell J, Wallace AG,
Gallagher JJ. Reaction of the myocardium to cryosurgery: electrophys-
iology and arrhythmogenic potential. Circulation 59: 364 –372, 1979.
doi:10.1161/01.CIR.59.2.364.
172. Klein HH, Schaper J, Puschmann S, Nienaber C, Kreuzer H,
Schaper W. Loss of canine myocardial nicotinamide adenine dinucle-
otides determines the transition from reversible to irreversible ischemic
damage of myocardial cells. Basic Res Cardiol 76: 612–621, 1981.
doi:10.1007/BF01908051.
H833MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
173. Kleinbongard P, Böse D, Baars T, Möhlenkamp S, Konorza T,
Schöner S, Elter-Schulz M, Eggebrecht H, Degen H, Haude M,
Levkau B, Schulz R, Erbel R, Heusch G. Vasoconstrictor potential of
coronary aspirate from patients undergoing stenting of saphenous vein
aortocoronary bypass grafts and its pharmacological attenuation. Circ
Res 108: 344 –352, 2011. doi:10.1161/CIRCRESAHA.110.235713.
174. Kleinbongard P, Skyschally A, Gent S, Pesch M, Heusch G. STAT3
as a common signal of ischemic conditioning: a lesson on “rigor and
reproducibility” in preclinical studies on cardioprotection. Basic Res
Cardiol 113: 3, 2018. doi:10.1007/s00395-017-0660-z.
175. Kleinbongard P, Skyschally A, Heusch G. Cardioprotection by remote
ischemic conditioning and its signal transduction. Pflugers Arch 469:
159 –181, 2017. doi:10.1007/s00424-016-1922-6.
176. Kloner RA. Current state of clinical translation of cardioprotective
agents for acute myocardial infarction. Circ Res 113: 451–463, 2013.
doi:10.1161/CIRCRESAHA.112.300627.
177. Kloner RA. No-reflow phenomenon: maintaining vascular integr-
ity. J Cardiovasc Pharmacol Ther 16: 244 –250, 2011. doi:10.1177/
1074248411405990.
178. Kloner RA, Ellis SG, Lange R, Braunwald E. Studies of experimental
coronary artery reperfusion. Effects on infarct size, myocardial function,
biochemistry, ultrastructure and microvascular damage. Circulation 68:
I8 –I15, 1983.
179. Kloner RA, Ganote CE, Jennings RB. The “no-reflow” phenomenon
after temporary coronary occlusion in the dog. J Clin Invest 54: 1496 –
1508, 1974. doi:10.1172/JCI107898.
180. Kloner RA, Przyklenk K, Whittaker P, Hale S. Preconditioning
stimuli and inadvertent preconditioning. J Mol Cell Cardiol 27: 743–747,
1995. doi:10.1016/0022-2828(95)90079-9.
181. Klotz L, Norman S, Vieira JM, Masters M, Rohling M, Dubé KN,
Bollini S, Matsuzaki F, Carr CA, Riley PR. Cardiac lymphatics are
heterogeneous in origin and respond to injury. Nature 522: 62–67, 2015.
doi:10.1038/nature14483.
182. Kolwicz SC Jr, Purohit S, Tian R. Cardiac metabolism and its inter-
actions with contraction, growth, and survival of cardiomyocytes. Circ
Res 113: 603–616, 2013. doi:10.1161/CIRCRESAHA.113.302095.
183. Koudstaal S, Jansen of Lorkeers S, Gho JM, van Hout GP, Jansen
MS, Grundeman PF, Pasterkamp G, Doevendans PA, Hoefer IE,
Chamuleau SA. Myocardial infarction and functional outcome assess-
ment in pigs. J Vis Exp 86: e51269, 2014. doi:10.3791/51269.
184. Kowallik P, Schulz R, Guth BD, Schade A, Paffhausen W, Gross R,
Heusch G. Measurement of regional myocardial blood flow with multi-
ple colored microspheres. Circulation 83: 974 –982, 1991. doi:10.1161/
01.CIR.83.3.974.
185. Kudej RK, Ghaleh B, Sato N, Shen YT, Bishop SP, Vatner SF.
Ineffective perfusion-contraction matching in conscious, chronically in-
strumented pigs with an extended period of coronary stenosis. Circ Res
82: 1199 –1205, 1998. doi:10.1161/01.RES.82.11.1199.
186. Lal N, Chiu AP, Wang F, Zhang D, Jia J, Wan A, Vlodavsky I,
Hussein B, Rodrigues B. Loss of VEGFB and its signaling in the
diabetic heart is associated with increased cell death signaling. Am J
Physiol Heart Circ Physiol 312: H1163–H1175, 2017. doi:10.1152/
ajpheart.00659.2016.
187. Lautamäki R, Schuleri KH, Sasano T, Javadi MS, Youssef A, Merrill
J, Nekolla SG, Abraham MR, Lardo AC, Bengel FM. Integration of
infarct size, tissue perfusion, and metabolism by hybrid cardiac positron
emission tomography/computed tomography: evaluation in a porcine
model of myocardial infarction. Circ Cardiovasc Imaging 2: 299 –305,
2009. doi:10.1161/CIRCIMAGING.108.846253.
188. Lefer DJ, Bolli R. Development of an NIH consortium for preclinicAl
AssESsment of CARdioprotective therapies (CAESAR): a paradigm
shift in studies of infarct size limitation. J Cardiovasc Pharmacol Ther
16: 332–339, 2011. doi:10.1177/1074248411414155.
189. Li B, Li Q, Wang X, Jana KP, Redaelli G, Kajstura J, Anversa P.
Coronary constriction impairs cardiac function and induces myocardial
damage and ventricular remodeling in mice. Am J Physiol Heart Circ
Physiol 273: H2508 –H2519, 1997.
190. Li RK, Jia ZQ, Weisel RD, Mickle DA, Zhang J, Mohabeer MK, Rao
V, Ivanov J. Cardiomyocyte transplantation improves heart function.
Ann Thorac Surg 62: 654 –660, discussion 660– 651, 1996.
191. Li S, Zhong S, Zeng K, Luo Y, Zhang F, Sun X, Chen L. Blockade of
NF-kappaB by pyrrolidine dithiocarbamate attenuates myocardial in-
flammatory response and ventricular dysfunction following coronary
microembolization induced by homologous microthrombi in rats. Basic
Res Cardiol 105: 139 –150, 2010. doi:10.1007/s00395-009-0067-6.
192. Liao R, Podesser BK, Lim CC. The continuing evolution of the
Langendorff and ejecting murine heart: new advances in cardiac pheno-
typing. Am J Physiol Heart Circ Physiol 303: H156 –H167, 2012.
doi:10.1152/ajpheart.00333.2012.
193. Lim H, Fallavollita JA, Hard R, Kerr CW, Canty JM Jr. Profound
apoptosis-mediated regional myocyte loss and compensatory hypertro-
phy in pigs with hibernating myocardium. Circulation 100: 2380 –2386,
1999. doi:10.1161/01.CIR.100.23.2380.
194. Lindsey ML, Iyer RP, Jung M, DeLeon-Pennell KY, Ma Y. Matrix
metalloproteinases as input and output signals for post-myocardial in-
farction remodeling. J Mol Cell Cardiol 91: 134 –140, 2016. doi:10.1016/
j.yjmcc.2015.12.018.
195. Lindsey ML, Iyer RP, Zamilpa R, Yabluchanskiy A, DeLeon-Pennell
KY, Hall ME, Kaplan A, Zouein FA, Bratton D, Flynn ER, Cannon
PL, Tian Y, Jin YF, Lange RA, Tokmina-Roszyk D, Fields GB, de
Castro Brás LE. A novel collagen matricryptin reduces left ventricular
dilation post-myocardial infarction by promoting scar formation and
angiogenesis. J Am Coll Cardiol 66: 1364 –1374, 2015. doi:10.1016/j.
jacc.2015.07.035.
196. Lindsey ML, Kassiri Z, Virag JA, de Castro Brás LE, Scherrer-
Crosbie M. Guidelines for measuring cardiac physiology in mice. Am J
Physiol Heart Circ Physiol. doi:10.1152/ajpheart.00339.2017.
197. Liu J, Wang P, Douglas SL, Tate JM, Sham S, Lloyd SG. Impact of
high-fat, low-carbohydrate diet on myocardial substrate oxidation, insu-
lin sensitivity, and cardiac function after ischemia-reperfusion. Am J
Physiol Heart Circ Physiol 311: H1–H10, 2016. doi:10.1152/ajpheart.
00809.2015.
198. Liu SQ, Ma XL, Qin G, Liu Q, Li YC, Wu YH. Trans-system
mechanisms against ischemic myocardial injury. Compr Physiol 5: 167–
192, 2015. doi:10.1002/cphy.c140026.
199. Liu SQ, Tefft BJ, Roberts DT, Zhang LQ, Ren Y, Li YC, Huang Y,
Zhang D, Phillips HR, Wu YH. Cardioprotective proteins upregulated
in the liver in response to experimental myocardial ischemia. Am J
Physiol Heart Circ Physiol 303: H1446 –H1458, 2012. doi:10.1152/
ajpheart.00362.2012.
200. Long X, Boluyt MO, Hipolito ML, Lundberg MS, Zheng JS, O’Neill
L, Cirielli C, Lakatta EG, Crow MT. p53 and the hypoxia-induced
apoptosis of cultured neonatal rat cardiac myocytes. J Clin Invest 99:
2635–2643, 1997. doi:10.1172/JCI119452.
201. Lopaschuk GD, Jaswal JS. Energy metabolic phenotype of the
cardiomyocyte during development, differentiation, and postnatal
maturation. J Cardiovasc Pharmacol 56: 130 –140, 2010. doi:10.
1097/FJC.0b013e3181e74a14.
202. Lopez EF, Kabarowski JH, Ingle KA, Kain V, Barnes S, Crossman
DK, Lindsey ML, Halade GV. Obesity superimposed on aging magni-
fies inflammation and delays the resolving response after myocardial
infarction. Am J Physiol Heart Circ Physiol 308: H269 –H280, 2015.
doi:10.1152/ajpheart.00604.2014.
203. Luo M, Guan X, Luczak ED, Lang D, Kutschke W, Gao Z, Yang J,
Glynn P, Sossalla S, Swaminathan PD, Weiss RM, Yang B, Rokita
AG, Maier LS, Efimov IR, Hund TJ, Anderson ME. Diabetes in-
creases mortality after myocardial infarction by oxidizing CaMKII. J
Clin Invest 123: 1262–1274, 2013. doi:10.1172/JCI65268.
204. Ma N, Ladilov Y, Moebius JM, Ong L, Piechaczek C, Dávid A,
Kaminski A, Choi YH, Li W, Egger D, Stamm C, Steinhoff G.
Intramyocardial delivery of human CD133cells in a SCID mouse
cryoinjury model: Bone marrow vs. cord blood-derived cells. Cardiovasc
Res 71: 158 –169, 2006. doi:10.1016/j.cardiores.2006.03.020.
205. Ma Y, Iyer RP, Jung M, Czubryt MP, Lindsey ML. Cardiac fibroblast
activation post-myocardial infarction: current knowledge gaps. Trends
Pharmacol Sci 38: 448 –458, 2017. doi:10.1016/j.tips.2017.03.001.
206. Ma Y, Yabluchanskiy A, Iyer RP, Cannon PL, Flynn ER, Jung M,
Henry J, Cates CA, Deleon-Pennell KY, Lindsey ML. Temporal
neutrophil polarization following myocardial infarction. Cardiovasc Res
110: 51–61, 2016. doi:10.1093/cvr/cvw024.
207. Maddaford TG, Hurtado C, Sobrattee S, Czubryt MP, Pierce GN. A
model of low-flow ischemia and reperfusion in single, beating adult
cardiomyocytes. Am J Physiol 277: H788 –H798, 1999.
208. Mahmoud AI, Porrello ER, Kimura W, Olson EN, Sadek HA.
Surgical models for cardiac regeneration in neonatal mice. Nat Protoc 9:
305–311, 2014. doi:10.1038/nprot.2014.021.
H834 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
209. Martin C, Schulz R, Post H, Boengler K, Kelm M, Kleinbongard P,
Gres P, Skyschally A, Konietzka I, Heusch G. Microdialysis-based
analysis of interstitial NO in situ: NO synthase-independent NO forma-
tion during myocardial ischemia. Cardiovasc Res 74: 46 –55, 2007.
doi:10.1016/j.cardiores.2006.12.020.
210. Martin C, Schulz R, Rose J, Heusch G. Inorganic phosphate content
and free energy change of ATP hydrolysis in regional short-term hiber-
nating myocardium. Cardiovasc Res 39: 318 –326, 1998. doi:10.1016/
S0008-6363(98)00086-8.
211. Masoud WG, Ussher JR, Wang W, Jaswal JS, Wagg CS, Dyck JR,
Lygate CA, Neubauer S, Clanachan AS, Lopaschuk GD. Failing
mouse hearts utilize energy inefficiently and benefit from improved
coupling of glycolysis and glucose oxidation. Cardiovasc Res 101:
30 –38, 2014. doi:10.1093/cvr/cvt216.
212. Matsuzaki M, Gallagher KP, Kemper WS, White F, Ross J Jr.
Sustained regional dysfunction produced by prolonged coronary stenosis:
gradual recovery after reperfusion. Circulation 68: 170 –182, 1983.
doi:10.1161/01.CIR.68.1.170.
213. Matsuzaki M, Gallagher KP, Patritti J, Tajimi T, Kemper WS,
White FC, Ross J Jr. Effects of a calcium-entry blocker (diltiazem) on
regional myocardial flow and function during exercise in conscious dogs.
Circulation 69: 801–814, 1984. doi:10.1161/01.CIR.69.4.801.
214. Matsuzaki M, Guth B, Tajimi T, Kemper WS, Ross J Jr. Effect of the
combination of diltiazem and atenolol on exercise-induced regional
myocardial ischemia in conscious dogs. Circulation 72: 233–243, 1985.
doi:10.1161/01.CIR.72.1.233.
215. Matsuzaki M, Patritti J, Tajimi T, Miller M, Kemper WS, Ross J Jr.
Effects of -blockade on regional myocardial flow and function during
exercise. Am J Physiol Heart Circ Physiol 247: H52–H60, 1984.
216. Maxwell MP, Hearse DJ, Yellon DM. Species variation in the coronary
collateral circulation during regional myocardial ischaemia: a critical
determinant of the rate of evolution and extent of myocardial infarction.
Cardiovasc Res 21: 737–746, 1987. doi:10.1093/cvr/21.10.737.
217. Mayorga M, Kiedrowski M, Shamhart P, Forudi F, Weber K,
Chilian WM, Penn MS, Dong F. Early upregulation of myocardial
CXCR4 expression is critical for dimethyloxalylglycine-induced cardiac
improvement in acute myocardial infarction. Am J Physiol Heart Circ
Physiol 310: H20 –H28, 2016. doi:10.1152/ajpheart.00449.2015.
218. McCall FC, Telukuntla KS, Karantalis V, Suncion VY, Heldman
AW, Mushtaq M, Williams AR, Hare JM. Myocardial infarction and
intramyocardial injection models in swine. Nat Protoc 7: 1479–1496,
2012. doi:10.1038/nprot.2012.075.
219. Meschiari CA, Jung M, Iyer RP, Yabluchanskiy A, Toba H, Garrett
MR, Lindsey ML. Macrophage overexpression of matrix metalloprotei-
nase-9 in aged mice improves diastolic physiology and cardiac wound
healing after myocardial infarction. Am J Physiol Heart Circ Physiol
314: H224H235, 2018. doi:10.1152/ajpheart.00453.2017.
220. Meyer C, Scherschel K. Ventricular tachycardia in ischemic heart
disease: the sympathetic heart and its scars. Am J Physiol Heart Circ
Physiol 312: H549 –H551, 2017. doi:10.1152/ajpheart.00061.2017.
221. Michael LH, Ballantyne CM, Zachariah JP, Gould KE, Pocius JS,
Taffet GE, Hartley CJ, Pham TT, Daniel SL, Funk E, Entman ML.
Myocardial infarction and remodeling in mice: effect of reperfusion. Am
J Physiol Heart Circ Physiol 277: H660 –H668, 1999.
222. Michael LH, Entman ML, Hartley CJ, Youker KA, Zhu J, Hall SR,
Hawkins HK, Berens K, Ballantyne CM. Myocardial ischemia and
reperfusion: a murine model. Am J Physiol Heart Circ Physiol 269:
H2147–H2154, 1995.
223. Morrissey PJ, Murphy KR, Daley JM, Schofield L, Turan NN,
Arunachalam K, Abbott JD, Koren G. A novel method of standardized
myocardial infarction in aged rabbits. Am J Physiol Heart Circ Physiol
312: H959 –H967, 2017. doi:10.1152/ajpheart.00582.2016.
224. Murry CE, Jennings RB, Reimer KA. Preconditioning with ischemia:
a delay of lethal cell injury in ischemic myocardium. Circulation 74:
1124 –1136, 1986. doi:10.1161/01.CIR.74.5.1124.
225. Nakanishi K, Vinten-Johansen J, Lefer DJ, Zhao Z, Fowler WC 3rd,
McGee DS, Johnston WE. Intracoronary L-arginine during reperfusion
improves endothelial function and reduces infarct size. Am J Physiol
Heart Circ Physiol 263: H1650 –H1658, 1992.
226. Neely JR, Liebermeister H, Battersby EJ, Morgan HE. Effect of
pressure development on oxygen consumption by isolated rat heart. Am
J Physiol Heart Circ Physiol 212: 804 –814, 1967.
227. Negoro S, Kunisada K, Fujio Y, Funamoto M, Darville MI, Eizirik
DL, Osugi T, Izumi M, Oshima Y, Nakaoka Y, Hirota H, Kishimoto
T, Yamauchi-Takihara K. Activation of signal transducer and activator
of transcription 3 protects cardiomyocytes from hypoxia/reoxygenation-
induced oxidative stress through the upregulation of manganese super-
oxide dismutase. Circulation 104: 979 –981, 2001. doi:10.1161/hc3401.
095947.
228. Nossuli TO, Lakshminarayanan V, Baumgarten G, Taffet GE, Bal-
lantyne CM, Michael LH, Entman ML. A chronic mouse model of
myocardial ischemia-reperfusion: essential in cytokine studies. Am J
Physiol Heart Circ Physiol 278: H1049 –H1055, 2000. doi:10.1152/
ajpheart.2000.278.4.H1049.
229. Nuzzo R. Scientific method: statistical errors. Nature 506: 150 –152,
2014. doi:10.1038/506150a.
230. O’Konski MS, White FC, Longhurst J, Roth D, Bloor CM. Ameroid
constriction of the proximal left circumflex coronary artery in swine. A
model of limited coronary collateral circulation. Am J Cardiovasc Pathol
1: 69 –77, 1987.
231. O’Quinn MP, Palatinus JA, Harris BS, Hewett KW, Gourdie RG. A
peptide mimetic of the connexin43 carboxyl terminus reduces gap junc-
tion remodeling and induced arrhythmia following ventricular injury.
Circ Res 108: 704 –715, 2011. doi:10.1161/CIRCRESAHA.110.235747.
232. Oduk Y, Zhu W, Kannappan R, Zhao M, Borovjagin AV, Oparil S,
Zhang J. VEGF nanoparticles repair the heart after myocardial infarc-
tion. Am J Physiol Heart Circ Physiol 314: H278H284, 2018. doi:10.
1152/ajpheart.00471.2017.
233. Omiya S, Omori Y, Taneike M, Protti A, Yamaguchi O, Akira S,
Shah AM, Nishida K, Otsu K. Toll-like receptor 9 prevents cardiac
rupture after myocardial infarction in mice independently of inflamma-
tion. Am J Physiol Heart Circ Physiol 311: H1485–H1497, 2016.
doi:10.1152/ajpheart.00481.2016.
234. Ongstad EL, O’Quinn MP, Ghatnekar GS, Yost MJ, Gourdie RG. A
connexin43 mimetic peptide promotes regenerative healing and improves
mechanical properties in skin and heart. Adv Wound Care (New Rochelle)
2: 55–62, 2013. doi:10.1089/wound.2011.0341.
235. Ostádal B, Ostádalová I, Kolár F, Charvátová Z, Netuka I. Ontoge-
netic development of cardiac tolerance to oxygen deprivation–possible
mechanisms. Physiol Res 58, Suppl 2: S1–S12, 2009.
236. Ostadalova I, Ostadal B, Kolár F, Parratt JR, Wilson S. Tolerance to
ischaemia and ischaemic preconditioning in neonatal rat heart. J Mol Cell
Cardiol 30: 857–865, 1998. doi:10.1006/jmcc.1998.0653.
237. Page BJ, Banas MD, Suzuki G, Weil BR, Young RF, Fallavollita JA,
Palka BA, Canty JM Jr. Revascularization of chronic hibernating
myocardium stimulates myocyte proliferation and partially reverses
chronic adaptations to ischemia. J Am Coll Cardiol 65: 684 –697, 2015.
doi:10.1016/j.jacc.2014.11.040.
238. Panizzi P, Swirski FK, Figueiredo JL, Waterman P, Sosnovik DE,
Aikawa E, Libby P, Pittet M, Weissleder R, Nahrendorf M. Impaired
infarct healing in atherosclerotic mice with Ly-6C(hi) monocytosis. JAm
Coll Cardiol 55: 1629 –1638, 2010. doi:10.1016/j.jacc.2009.08.089.
239. Pantely GA, Malone SA, Rhen WS, Anselone CG, Arai A, Bristow J,
Bristow JD. Regeneration of myocardial phosphocreatine in pigs despite
continued moderate ischemia. Circ Res 67: 1481–1493, 1990. doi:10.
1161/01.RES.67.6.1481.
240. Park KM, Teoh JP, Wang Y, Broskova Z, Bayoumi AS, Tang Y, Su
H, Weintraub NL, Kim IM. Carvedilol-responsive microRNAs, miR-
199a-3p and -214 protect cardiomyocytes from simulated ischemia-
reperfusion injury. Am J Physiol Heart Circ Physiol 311: H371–H383,
2016. doi:10.1152/ajpheart.00807.2015.
241. Parrish DC, Francis Stuart SD, Olivas A, Wang L, Nykjaer A,
Ripplinger CM, Habecker BA. Transient denervation of viable myo-
cardium after myocardial infarction does not alter arrhythmia suscepti-
bility. Am J Physiol Heart Circ Physiol 314: H415H423, 2018. doi:
10.1152/ajpheart.00300.2017.
242. Pizzuto MF, Suzuki G, Banas MD, Heavey B, Fallavollita JA, Canty
JM Jr. Dissociation of hemodynamic and electrocardiographic indexes
of myocardial ischemia in pigs with hibernating myocardium and sudden
cardiac death. Am J Physiol Heart Circ Physiol 304: H1697–H1707,
2013. doi:10.1152/ajpheart.00166.2013.
243. Porrello ER, Mahmoud AI, Simpson E, Hill JA, Richardson JA,
Olson EN, Sadek HA. Transient regenerative potential of the neo-
natal mouse heart. Science 331: 1078 –1080, 2011. doi:10.1126/
science.1200708.
244. Portal L, Martin V, Assaly R, d’Anglemont de Tassigny A,
Michineau S, Berdeaux A, Ghaleh B, Pons S. A model of hypoxia-
reoxygenation on isolated adult mouse cardiomyocytes: characterization,
H835MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
comparison with ischemia-reperfusion, and application to the cardiopro-
tective effect of regular treadmill exercise. J Cardiovasc Pharmacol Ther
18: 367–375, 2013. doi:10.1177/1074248412475158.
245. Pound KM, Sorokina N, Ballal K, Berkich DA, Fasano M, Lanoue
KF, Taegtmeyer H, O’Donnell JM, Lewandowski ED. Substrate-
enzyme competition attenuates upregulated anaplerotic flux through
malic enzyme in hypertrophied rat heart and restores triacylglyceride
content: attenuating upregulated anaplerosis in hypertrophy. Circ Res
104: 805–812, 2009. doi:10.1161/CIRCRESAHA.108.189951.
246. Przyklenk K. Ischaemic conditioning: pitfalls on the path to clinical
translation. Br J Pharmacol 172: 1961–1973, 2015. doi:10.1111/bph.
13064.
247. Przyklenk K. Reduction of myocardial infarct size with ischemic “con-
ditioning”: physiologic and technical considerations. Anesth Analg 117:
891–901, 2013. doi:10.1213/ANE.0b013e318294fc63.
248. Przyklenk K, Bauer B, Ovize M, Kloner RA, Whittaker P. Regional
ischemic “preconditioning” protects remote virgin myocardium from
subsequent sustained coronary occlusion. Circulation 87: 893–899, 1993.
doi:10.1161/01.CIR.87.3.893.
249. Przyklenk K, Maynard M, Greiner DL, Whittaker P. Cardioprotec-
tion with postconditioning: loss of efficacy in murine models of type-2
and type-1 diabetes. Antioxid Redox Signal 14: 781–790, 2011. doi:10.
1089/ars.2010.3343.
250. Ramirez FD, Motazedian P, Jung RG, Di Santo P, MacDonald ZD,
Moreland R, Simard T, Clancy AA, Russo JJ, Welch VA, Wells GA,
Hibbert B. Methodological rigor in preclinical cardiovascular studies:
targets to enhance reproducibility and promote research translation. Circ
Res 120: 1916 –1926, 2017. doi:10.1161/CIRCRESAHA.117.310628.
251. Reimer KA, Jennings RB, Tatum AH. Pathobiology of acute myocar-
dial ischemia: metabolic, functional and ultrastructural studies. Am J
Cardiol 52: 72A–81A, 1983. doi:10.1016/0002-9149(83)90180-7.
252. Reimer KA, Lowe JE, Rasmussen MM, Jennings RB. The wavefront
phenomenon of ischemic cell death. 1. Myocardial infarct size vs dura-
tion of coronary occlusion in dogs. Circulation 56: 786 –794, 1977.
doi:10.1161/01.CIR.56.5.786.
253. Rezkalla SH, Kloner RA. Coronary no-reflow phenomenon: from the
experimental laboratory to the cardiac catheterization laboratory. Cath-
eter Cardiovasc Interv 72: 950 –957, 2008. doi:10.1002/ccd.21715.
254. Ripplinger CM, Noujaim SF, Linz D. The nervous heart. Prog Biophys
Mol Biol 120: 199 –209, 2016. doi:10.1016/j.pbiomolbio.2015.12.015.
255. Riva E, Hearse DJ. Age-dependent changes in myocardial susceptibility
to ischemic injury. Cardioscience 4: 85–92, 1993.
256. Robertson C, Tran DD, George SC. Concise review: maturation phases
of human pluripotent stem cell-derived cardiomyocytes. Stem Cells 31:
829 –837, 2013. doi:10.1002/stem.1331.
257. Robey TE, Murry CE. Absence of regeneration in the MRL/MpJ mouse
heart following infarction or cryoinjury. Cardiovasc Pathol 17: 6–13,
2008. doi:10.1016/j.carpath.2007.01.005.
258. Roell W, Fan Y, Xia Y, Stoecker E, Sasse P, Kolossov E, Bloch W,
Metzner H, Schmitz C, Addicks K, Hescheler J, Welz A, Fleisch-
mann BK. Cellular cardiomyoplasty in a transgenic mouse model.
Transplantation 73: 462–465, 2002. doi:10.1097/00007890-200202150-
00022.
259. Ross J Jr. Myocardial perfusion-contraction matching. Implications for
coronary heart disease and hibernation. Circulation 83: 1076 –1083,
1991. doi:10.1161/01.CIR.83.3.1076.
260. Roth DM, Maruoka Y, Rogers J, White FC, Longhurst JC, Bloor
CM. Development of coronary collateral circulation in left circumflex
ameroid-occluded swine myocardium. Am J Physiol 253: H1279 –
H1288, 1987.
261. Sabbah HN, Stein PD, Kono T, Gheorghiade M, Levine TB, Jafri S,
Hawkins ET, Goldstein S. A canine model of chronic heart failure
produced by multiple sequential coronary microembolizations. Am J
Physiol Heart Circ Physiol 260: H1379 –H1384, 1991.
262. Saeed M, Watzinger N, Krombach GA, Lund GK, Wendland MF,
Chujo M, Higgins CB. Left ventricular remodeling after infarction:
sequential MR imaging with oral nicorandil therapy in rat model. Radi-
ology 224: 830 –837, 2002. doi:10.1148/radiol.2243011372.
263. Sakamoto H, Parish LM, Hamamoto H, Ryan LP, Eperjesi TJ,
Plappert TJ, Jackson BM, St John-Sutton MG, Gorman JH 3rd,
Gorman RC. Effect of reperfusion on left ventricular regional remod-
eling strains after myocardial infarction. Ann Thorac Surg 84: 1528–
1536, 2007. doi:10.1016/j.athoracsur.2007.05.060.
264. Sánchez-González J, Fernandez-Jiménez R, Nothnagel ND, López-
Martín G, Fuster V, Ibañez B. Optimization of dual-saturation single
bolus acquisition for quantitative cardiac perfusion and myocardial blood
flow maps. J Cardiovasc Magn Reson 17: 21, 2015. doi:10.1186/s12968-
015-0116-2.
265. Sasayama S, Franklin D, Ross J Jr, Kemper WS, McKown D.
Dynamic changes in left ventricular wall thickness and their use in
analyzing cardiac function in the conscious dog. Am J Cardiol 38:
870 –879, 1976. doi:10.1016/0002-9149(76)90800-6.
266. Schloss MJ, Horckmans M, Nitz K, Duchene J, Drechsler M, Bid-
zhekov K, Scheiermann C, Weber C, Soehnlein O, Steffens S. The
time-of-day of myocardial infarction onset affects healing through oscil-
lations in cardiac neutrophil recruitment. EMBO Mol Med 8: 937–948,
2016. doi:10.15252/emmm.201506083.
267. Schulz R, Guth BD, Pieper K, Martin C, Heusch G. Recruitment of an
inotropic reserve in moderately ischemic myocardium at the expense of
metabolic recovery. A model of short-term hibernation. Circ Res 70:
1282–1295, 1992. doi:10.1161/01.RES.70.6.1282.
268. Schulz R, Kappeler C, Coenen H, Bockisch A, Heusch G. Positron
emission tomography analysis of [1-
11
C]acetate kinetics in short-term
hibernating myocardium. Circulation 97: 1009 –1016, 1998. doi:10.
1161/01.CIR.97.10.1009.
269. Schulz R, Post H, Neumann T, Gres P, Lüss H, Heusch G. Progressive
loss of perfusion-contraction matching during sustained moderate isch-
emia in pigs. Am J Physiol Heart Circ Physiol 280: H1945–H1953, 2001.
doi:10.1152/ajpheart.2001.280.5.H1945.
270. Schulz R, Rose J, Post H, Heusch G. Involvement of endogenous
adenosine in ischaemic preconditioning in swine. Pflugers Arch 430:
273–282, 1995. doi:10.1007/BF00374659.
271. Schwarz ER, Somoano Y, Hale SL, Kloner RA. What is the required
reperfusion period for assessment of myocardial infarct size using tri-
phenyltetrazolium chloride staining in the rat? J Thromb Thrombolysis
10: 181–187, 2000. doi:10.1023/A:1018770711705.
272. Scruggs SB, Watson K, Su AI, Hermjakob H, Yates JR III, Lindsey
ML, Ping P. Harnessing the heart of big data. Circ Res 116: 1115–1119,
2015. doi:10.1161/CIRCRESAHA.115.306013.
273. Seitelberger R, Guth BD, Heusch G, Lee JD, Katayama K, Ross J Jr.
Intracoronary alpha 2-adrenergic receptor blockade attenuates ischemia
in conscious dogs during exercise. Circ Res 62: 436 –442, 1988. doi:10.
1161/01.RES.62.3.436.
274. Seldin MM, Kim ED, Romay MC, Li S, Rau CD, Wang JJ, Krishnan
KC, Wang Y, Deb A, Lusis AJ. A systems genetics approach
identifiesTrp53inp2as a link between cardiomyocyte glucose utilization
and hypertrophic response. Am J Physiol Heart Circ Physiol 312:
H728 –H741, 2017. doi:10.1152/ajpheart.00068.2016.
275. Semenza GL. Cellular and molecular dissection of reperfusion injury:
ROS within and without. Circ Res 86: 117–118, 2000. doi:10.1161/01.
RES.86.2.117.
276. Sen S, Sadek HA. Neonatal heart regeneration: mounting support and
need for technical standards. J Am Heart Assoc 4: e001727, 2015.
doi:10.1161/JAHA.114.001727.
277. Shen YT, Vatner SF. Differences in myocardial stunning following
coronary artery occlusion in conscious dogs, pigs, and baboons. Am J
Physiol Heart Circ Physiol 270: H1312–H1322, 1996.
278. Shen YT, Vatner SF. Mechanism of impaired myocardial function
during progressive coronary stenosis in conscious pigs. Hibernation
versus stunning? Circ Res 76: 479 –488, 1995. doi:10.1161/01.RES.76.
3.479.
279. Sherman AJ, Klocke FJ, Decker RS, Decker ML, Kozlowski KA,
Harris KR, Hedjbeli S, Yaroshenko Y, Nakamura S, Parker MA,
Checchia PA, Evans DB. Myofibrillar disruption in hypocontractile
myocardium showing perfusion-contraction matches and mismatches.
Am J Physiol Heart Circ Physiol 278: H1320 –H1334, 2000. doi:10.
1152/ajpheart.2000.278.4.H1320.
280. Shinde AV, Humeres C, Frangogiannis NG. The role of -smooth
muscle actin in fibroblast-mediated matrix contraction and remodeling.
Biochim Biophys Acta 1863: 298 –309, 2017. doi:10.1016/j.bbadis.2016.
11.006.
281. Skrzypiec-Spring M, Grotthus B, Szelag A, Schulz R. Isolated heart
perfusion according to Langendorff–still viable in the new millennium. J
Pharmacol Toxicol Methods 55: 113–126, 2007. doi:10.1016/j.vascn.
2006.05.006.
282. Skyschally A, Amanakis G, Neuhäuser M, Kleinbongard P, Heusch
G. Impact of electrical defibrillation on infarct size and no-reflow in pigs
H836 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
subjected to myocardial ischemia-reperfusion without and with ischemic
conditioning. Am J Physiol Heart Circ Physiol 313: H871–H878, 2017.
doi:10.1152/ajpheart.00293.2017.
283. Skyschally A, Gent S, Amanakis G, Schulte C, Kleinbongard P,
Heusch G. Across-species transfer of protection by remote ischemic
preconditioning with species-specific myocardial signal transduction
by reperfusion injury salvage kinase and survival activating factor
enhancement pathways. Circ Res 117: 279 –288, 2015. doi:10.1161/
CIRCRESAHA.117.306878.
284. Skyschally A, Schulz R, Heusch G. Pathophysiology of myocardial
infarction: protection by ischemic pre- and postconditioning. Herz 33:
88 –100, 2008. doi:10.1007/s00059-008-3101-9.
285. Sorop O, Merkus D, de Beer VJ, Houweling B, Pistea A, McFalls EO,
Boomsma F, van Beusekom HM, van der Giessen WJ, VanBavel E,
Duncker DJ. Functional and structural adaptations of coronary mi-
crovessels distal to a chronic coronary artery stenosis. Circ Res 102:
795–803, 2008. doi:10.1161/CIRCRESAHA.108.172528.
286. Spaan JA, Breuls NP, Laird JD. Diastolic-systolic coronary flow
differences are caused by intramyocardial pump action in the anesthe-
tized dog. Circ Res 49: 584 –593, 1981. doi:10.1161/01.RES.49.3.584.
287. Sreejit P, Kumar S, Verma RS. An improved protocol for primary
culture of cardiomyocyte from neonatal mice. In Vitro Cell Dev Biol
Anim 44: 45–50, 2008. doi:10.1007/s11626-007-9079-4.
288. Steenbergen C, Perlman ME, London RE, Murphy E. Mechanism of
preconditioning. Ionic alterations. Circ Res 72: 112–125, 1993. doi:10.
1161/01.RES.72.1.112.
289. Strem BM, Zhu M, Alfonso Z, Daniels EJ, Schreiber R, Beygui R,
MacLellan WR, Hedrick MH, Fraser JK. Expression of cardiomyo-
cytic markers on adipose tissue-derived cells in a murine model of
acute myocardial injury. Cytotherapy 7: 282–291, 2005. doi:10.1080/
14653240510027226.
290. Strungs EG, Ongstad EL, O’Quinn MP, Palatinus JA, Jourdan LJ,
Gourdie RG. Cryoinjury models of the adult and neonatal mouse heart
for studies of scarring and regeneration. Methods Mol Biol 1037: 343–
353, 2013. doi:10.1007/978-1-62703-505-7_20.
291. Sun JZ, Tang XL, Knowlton AA, Park SW, Qiu Y, Bolli R. Late
preconditioning against myocardial stunning. An endogenous protective
mechanism that confers resistance to postischemic dysfunction 24 h after
brief ischemia in conscious pigs. J Clin Invest 95: 388 –403, 1995.
doi:10.1172/JCI117667.
292. Sutherland FJ, Hearse DJ. The isolated blood and perfusion fluid
perfused heart. Pharmacol Res 41: 613–627, 2000. doi:10.1006/phrs.
1999.0653.
293. Swirski FK, Nahrendorf M, Etzrodt M, Wildgruber M, Cortez-
Retamozo V, Panizzi P, Figueiredo JL, Kohler RH, Chudnovskiy A,
Waterman P, Aikawa E, Mempel TR, Libby P, Weissleder R, Pittet
MJ. Identification of splenic reservoir monocytes and their deployment
to inflammatory sites. Science 325: 612–616, 2009. doi:10.1126/science.
1175202.
294. Takawale A, Zhang P, Azad A, Wang W, Wang X, Murray AG,
Kassiri Z. Myocardial overexpression of TIMP3 after myocardial infarc-
tion exerts beneficial effects by promoting angiogenesis and suppressing
early proteolysis. Am J Physiol Heart Circ Physiol 313: H224 –H236,
2017. doi:10.1152/ajpheart.00108.2017.
295. Tanai E, Frantz S. Pathophysiology of heart failure. Compr Physiol 6:
187–214, 2015. doi:10.1002/cphy.c140055.
296. Tanaka M, Ito H, Adachi S, Akimoto H, Nishikawa T, Kasajima T,
Marumo F, Hiroe M. Hypoxia induces apoptosis with enhanced expres-
sion of Fas antigen messenger RNA in cultured neonatal rat cardiomy-
ocytes. Circ Res 75: 426 –433, 1994. doi:10.1161/01.RES.75.3.426.
297. Taylor CB, Davis CB Jr, Vawter GF, Hass GM. Controlled myocar-
dial injury produced by a hypothermal method. Circulation 3: 239 –253,
1951. doi:10.1161/01.CIR.3.2.239.
298. Thakker GD, Frangogiannis NG, Bujak M, Zymek P, Gaubatz JW,
Reddy AK, Taffet G, Michael LH, Entman ML, Ballantyne CM.
Effects of diet-induced obesity on inflammation and remodeling after
myocardial infarction. Am J Physiol Heart Circ Physiol 291: H2504 –
H2514, 2006. doi:10.1152/ajpheart.00322.2006.
299. Thielmann M, Dörge H, Martin C, Belosjorow S, Schwanke U, van
De Sand A, Konietzka I, Büchert A, Krüger A, Schulz R, Heusch G.
Myocardial dysfunction with coronary microembolization: signal trans-
duction through a sequence of nitric oxide, tumor necrosis factor-alpha,
and sphingosine. Circ Res 90: 807–813, 2002. doi:10.1161/01.RES.
0000014451.75415.36.
300. Thiese MS, Arnold ZC, Walker SD. The misuse and abuse of statistics
in biomedical research. Biochem Med (Zagreb) 25: 5–11, 2015. doi:10.
11613/BM.2015.001.
301. Thomas SA, Fallavollita JA, Suzuki G, Borgers M, Canty JM Jr.
Dissociation of regional adaptations to ischemia and global myolysis in
an accelerated Swine model of chronic hibernating myocardium. Circ
Res 91: 970 –977, 2002. doi:10.1161/01.RES.0000040396.79379.77.
302. Thompson RB, Emani SM, Davis BH, van den Bos EJ, Morimoto
Y, Craig D, Glower D, Taylor DA. Comparison of intracardiac cell
transplantation: autologous skeletal myoblasts versus bone marrow
cells. Circulation 108, Suppl 1: II264 –II271, 2003. doi:10.1161/01.
cir.0000087657.29184.9b.
303. Toyota E, Warltier DC, Brock T, Ritman E, Kolz C, O’Malley P,
Rocic P, Focardi M, Chilian WM. Vascular endothelial growth factor
is required for coronary collateral growth in the rat. Circulation 112:
2108 –2113, 2005. doi:10.1161/CIRCULATIONAHA.104.526954.
304. Triana JF, Li XY, Jamaluddin U, Thornby JI, Bolli R. Postischemic
myocardial “stunning”. Identification of major differences between the
open-chest and the conscious dog and evaluation of the oxygen radical
hypothesis in the conscious dog. Circ Res 69: 731–747, 1991. doi:10.
1161/01.RES.69.3.731.
305. van Amerongen MJ, Harmsen MC, Petersen AH, Popa ER, van
Luyn MJ. Cryoinjury: a model of myocardial regeneration. Cardiovasc
Pathol 17: 23–31, 2008. doi:10.1016/j.carpath.2007.03.002.
306. van den Borne SW, van de Schans VA, Strzelecka AE, Vervoort-
Peters HT, Lijnen PM, Cleutjens JP, Smits JF, Daemen MJ, Janssen
BJ, Blankesteijn WM. Mouse strain determines the outcome of wound
healing after myocardial infarction. Cardiovasc Res 84: 273–282, 2009.
doi:10.1093/cvr/cvp207.
307. van den Bos EJ, Mees BM, de Waard MC, de Crom R, Duncker DJ.
A novel model of cryoinjury-induced myocardial infarction in the mouse:
a comparison with coronary artery ligation. Am J Physiol Heart Circ
Physiol 289: H1291–H1300, 2005. doi:10.1152/ajpheart.00111.2005.
308. Vanoverschelde JL, Wijns W, Depré C, Essamri B, Heyndrickx GR,
Borgers M, Bol A, Melin JA. Mechanisms of chronic regional post-
ischemic dysfunction in humans. New insights from the study of nonin-
farcted collateral-dependent myocardium. Circulation 87: 1513–1523,
1993. doi:10.1161/01.CIR.87.5.1513.
309. Vatner SF. Correlation between acute reductions in myocardial blood
flow and function in conscious dogs. Circ Res 47: 201–207, 1980.
doi:10.1161/01.RES.47.2.201.
310. Vogel B, Shinagawa H, Hofmann U, Ertl G, Frantz S. Acute DNase1
treatment improves left ventricular remodeling after myocardial infarc-
tion by disruption of free chromatin. Basic Res Cardiol 110: 15, 2015.
doi:10.1007/s00395-015-0472-y.
311. Voorhees AP, DeLeon-Pennell KY, Ma Y, Halade GV, Yabluchan-
skiy A, Iyer RP, Flynn E, Cates CA, Lindsey ML, Han HC. Building
a better infarct: Modulation of collagen cross-linking to increase infarct
stiffness and reduce left ventricular dilation post-myocardial infarction. J
Mol Cell Cardiol 85: 229 –239, 2015. doi:10.1016/j.yjmcc.2015.06.006.
312. Wan F, Letavernier E, Le Saux CJ, Houssaini A, Abid S, Czibik G,
Sawaki D, Marcos E, Dubois-Rande JL, Baud L, Adnot S, De-
rumeaux G, Gellen B. Calpastatin overexpression impairs postinfarct
scar healing in mice by compromising reparative immune cell recruit-
ment and activation. Am J Physiol Heart Circ Physiol 309: H1883–
H1893, 2015. doi:10.1152/ajpheart.00594.2015.
313. Warren M, Sciuto KJ, Taylor TG, Garg V, Torres NS, Shibayama J,
Spitzer KW, Zaitsev AV. Blockade of CaMKII depresses conduction
preferentially in the right ventricular outflow tract and promotes ischemic
ventricular fibrillation in the rabbit heart. Am J Physiol Heart Circ
Physiol 312: H752–H767, 2017. doi:10.1152/ajpheart.00347.2016.
314. Weil BR, Young RF, Shen X, Suzuki G, Qu J, Malhotra S, Canty JM
Jr. Brief Myocardial ischemia produces cardiac troponin i release and
focal myocyte apoptosis in the absence of pathological infarction in
swine. JACC Basic Transl Sci 2: 105–114, 2017. doi:10.1016/j.jacbts.
2017.01.006.
315. White IA, Gordon J, Balkan W, Hare JM. Sympathetic reinnervation
is required for mammalian cardiac regeneration. Circ Res 117: 990 –994,
2015. doi:10.1161/CIRCRESAHA.115.307465.
316. Whittaker P, Kloner RA, Boughner DR, Pickering JG. Quantitative
assessment of myocardial collagen with picrosirius red staining and
circularly polarized light. Basic Res Cardiol 89: 397–410, 1994. doi:10.
1007/BF00788278.
H837MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
317. Wu Y, Yin X, Wijaya C, Huang MH, McConnell BK. Acute myocar-
dial infarction in rats. J Vis Exp 2011: 2464, 2011. doi:10.3791/2464.
318. Xu Y, Huo Y, Toufektsian MC, Ramos SI, Ma Y, Tejani AD, French
BA, Yang Z. Activated platelets contribute importantly to myocardial
reperfusion injury. Am J Physiol Heart Circ Physiol 290: H692–H699,
2006. doi:10.1152/ajpheart.00634.2005.
319. Yabluchanskiy A, Ma Y, DeLeon-Pennell KY, Altara R, Halade GV,
Voorhees AP, Nguyen NT, Jin YF, Winniford MD, Hall ME, Han
HC, Lindsey ML. Myocardial infarction superimposed on aging:
MMP-9 deletion promotes M2 macrophage polarization. J Gerontol A
Biol Sci Med Sci 71: 475–483, 2016. doi:10.1093/gerona/glv034.
320. Yamamoto H, Tomoike H, Shimokawa H, Nabeyama S, Nakamura
M. Development of collateral function with repetitive coronary occlusion
in a canine model reduces myocardial reactive hyperemia in the absence
of significant coronary stenosis. Circ Res 55: 623–632, 1984. doi:10.
1161/01.RES.55.5.623.
321. Yamamoto T, Tamaki K, Shirakawa K, Ito K, Yan X, Katsumata Y,
Anzai A, Matsuhashi T, Endo J, Inaba T, Tsubota K, Sano M,
Fukuda K, Shinmura K. Cardiac Sirt1 mediates the cardioprotective
effect of caloric restriction by suppressing local complement system
activation after ischemia-reperfusion. Am J Physiol Heart Circ Physiol
310: H1003–H1014, 2016. doi:10.1152/ajpheart.00676.2015.
322. Yang J, Brown ME, Zhang H, Martinez M, Zhao Z, Bhutani S, Yin
S, Trac D, Xi JJ, Davis ME. High-throughput screening identifies
microRNAs that target Nox2 and improve function after acute myocar-
dial infarction. Am J Physiol Heart Circ Physiol 312: H1002–H1012,
2017. doi:10.1152/ajpheart.00685.2016.
323. Yang Y, Ma Y, Han W, Li J, Xiang Y, Liu F, Ma X, Zhang J, Fu Z,
Su YD, Du XJ, Gao XM. Age-related differences in postinfarct left
ventricular rupture and remodeling. Am J Physiol Heart Circ Physiol
294: H1815–H1822, 2008. doi:10.1152/ajpheart.00831.2007.
324. Yellon DM, Hausenloy DJ. Myocardial reperfusion injury. N Engl J
Med 357: 1121–1135, 2007. doi:10.1056/NEJMra071667.
325. Yoshihara HA, Bastiaansen JA, Berthonneche C, Comment A,
Schwitter J. An intact small animal model of myocardial ischemia-
reperfusion: characterization of metabolic changes by hyperpolarized
13
C
MR spectroscopy. Am J Physiol Heart Circ Physiol 309: H2058 –H2066,
2015. doi:10.1152/ajpheart.00376.2015.
326. Yu Y, Wei SG, Zhang ZH, Weiss RM, Felder RB. ERK1/2 MAPK
signaling in hypothalamic paraventricular nucleus contributes to sympa-
thetic excitation in rats with heart failure after myocardial infarction. Am
J Physiol Heart Circ Physiol 310: H732–H739, 2016. doi:10.1152/
ajpheart.00703.2015.
327. Zamilpa R, Zhang J, Chiao YA, de Castro Brás LE, Halade GV, Ma
Y, Hacker SO, Lindsey ML. Cardiac wound healing post-myocardial
infarction: a novel method to target extracellular matrix remodeling in the
left ventricle. Methods Mol Biol 1037: 313–324, 2013. doi:10.1007/978-
1-62703-505-7_18.
328. Zhang F, Xia Y, Yan W, Zhang H, Zhou F, Zhao S, Wang W, Zhu
D, Xin C, Lee Y, Zhang L, He Y, Gao E, Tao L. Sphingosine
1-phosphate signaling contributes to cardiac inflammation, dysfunction,
and remodeling following myocardial infarction. Am J Physiol Heart
Circ Physiol 310: H250 –H261, 2016. doi:10.1152/ajpheart.00372.2015.
329. Zhao ZQ, Corvera JS, Halkos ME, Kerendi F, Wang NP, Guyton
RA, Vinten-Johansen J. Inhibition of myocardial injury by ischemic
postconditioning during reperfusion: comparison with ischemic precon-
ditioning. Am J Physiol Heart Circ Physiol 285: H579 –H588, 2003.
doi:10.1152/ajpheart.01064.2002.
330. Zhu HL, Wei X, Qu SL, Zhang C, Zuo XX, Feng YS, Luo Q, Chen
GW, Liu MD, Jiang L, Xiao XZ, Wang KK. Ischemic postconditioning
protects cardiomyocytes against ischemia/reperfusion injury by inducing
MIP2. Exp Mol Med 43: 437–445, 2011. doi:10.3858/emm.2011.43.8.
049.
331. Ziv O, Schofield L, Lau E, Chaves L, Patel D, Jeng P, Peng X, Choi
BR, Koren G. A novel, minimally invasive, segmental myocardial
infarction with a clear healed infarct borderzone in rabbits. Am J Physiol
Heart Circ Physiol 302: H2321–H2330, 2012. doi:10.1152/ajpheart.
00031.2012.
332. Zymek P, Bujak M, Chatila K, Cieslak A, Thakker G, Entman ML,
Frangogiannis NG. The role of platelet-derived growth factor signaling
in healing myocardial infarcts. J Am Coll Cardiol 48: 2315–2323, 2006.
doi:10.1016/j.jacc.2006.07.060.
H838 MYOCARDIAL ISCHEMIA GUIDELINES
AJP-Heart Circ Physiol doi:10.1152/ajpheart.00335.2017 www.ajpheart.org
Downloaded from www.physiology.org/journal/ajpheart by ${individualUser.givenNames} ${individualUser.surname} (158.046.158.224) on July 25, 2018.
Copyright © 2018 American Physiological Society. All rights reserved.
... To induce ischemia-like conditions, we replaced the normal buffer with a buffer with no energy sources and saturated it with nitrogen, as previously reported. 15,16 During reperfusion, this buffer was replaced with the original buffer and oxygenation was resumed (Figure 3a). ...
... To confirm whether this novel ex vivo system could recapitulate intracellular ATP dynamics in kidney diseases, we first induced culture conditions that mimicked IR injury 15,16 Electron microscopy analysis of these slices as early as 1 h after reperfusion revealed mitochondrial damage in PTs (Figure 4a) and a lower mitochondrial length-to-width ratio (L/W ratio) after longer ischemia ( Figure 4b). Additionally, the slices 6 h after reperfusion showed loss of brush borders in PTs and tubular degeneration after longer ischemia (Figure 4, c and d). ...
Preprint
ATP depletion plays a central role in the pathogenesis of kidney diseases. Recently, we reported spatiotemporal intracellular ATP dynamics during ischemia reperfusion (IR) using GO-ATeam2 mice systemically expressing an ATP biosensor. However, observation from the kidney surface did not allow visualization of deeper nephrons or accurate evaluation of ATP synthesis pathways. Here we established a novel ATP imaging system using slice culture of GO-ATeam2 mouse kidneys, evaluated ATP synthesis pathway, and analyzed intracellular ATP dynamics using ex vivo IR and cisplatin nephropathy models. We demonstrated that proximal tubules (PTs) were strongly dependent on oxidative phosphorylation (OXPHOS), whereas podocytes relied on both OXPHOS and glycolysis using ATP synthesis inhibitors. We also confirmed that the ex vivo IR model could recapitulate ATP dynamics in vivo ; ATP recovery in PTs after reperfusion varied depending on ischemia length, whereas ATP in distal tubules (DTs) recovered well even after long ischemia. After cisplatin administration, ATP levels in PTs decreased first, followed by a decrease in DTs. An organic cation transporter 2 inhibitor suppressed cisplatin uptake in kidney slices, leading to better ATP recovery in PTs, but not in DTs. Finally, we confirmed that a mitochondria protection reagent delayed cisplatin-induced ATP decrease in PTs. This novel system may provide new insights into the energy dynamics and pathogenesis of kidney disease.
... Считаем, что при рассмотрении сценариев с сильным ветвлением коронарных сосудов, задействованных в ишемическом процессе, общий подход к моделированию может оставаться тем, который был разработан для случая крупноочагового инфаркта с неосложненной морфологией [21][22][23]. Предполагаем, в частности, что а) общее расположение кардиомиоцитов может быть охарактеризовано наличием некоторой единой оси миокардиальных клеток, которая обнаруживается in vivo на срезах в среднем слое миокарда и вдоль которой происходит более интенсивный, чем в других направлениях, перенос веществ, б) хотя любая локальная область миокарда имеет уникальный морфологический рисунок, все же в среднем его внутренняя структура достаточно однородна, в) архитектура сердца по своей природе трехмерна, но этим фактом можно пренебречь, т. к. большинство измеряемых характеристик (от структуры миокарда до характеристик происходящих там процессов) получены на двумерных гистологических срезах или диссоциированных клетках (см., например, [24][25][26][27]), г) повреждение, захватывающее всю толщу стенки левого желудочка, и происходящие там процессы можно считать однородными по данному направлению. ...
Article
The study of the inflammatory phase of acute myocardial infarction in multivessel coronary lesion was performed using the methodology of mathematical modeling. The minimal reaction-diffusion mathematical model is focused on the description of the functional M1/M2 polarization of macrophages and the influence of factors of aseptic inflammation on the process of cardiomyocyte death. The initial conditions and dynamics of the process in the infarction nucleus are assumed to be consistent with laboratory measurement data. The nature of the spatiotemporal distribution of substances (cell populations and inflammatory mediators) and the features of the formation of nonlinear dynamic structures of demarcation inflammation are studied using model examples. The patterns of functioning of the basic mechanisms of the inflammatory response are analyzed, and the role of the main inflammatory mediators is evaluated. The previously obtained estimates of the effectiveness of anti-inflammatory therapeutic strategies based on cytokine management and macrophage polarization in complex heart attack scenarios with multivessel coronary lesion have been confirmed. The research results allow us to consider the accepted reaction-diffusion model with constant diffusion coefficients as an example of a formal mathematical description of an active environment in which dissipative (diffusion) and local biochemical processes compete with each other, as well as the pro-inflammatory link of innate immunity opposes the anti-inflammatory one. The ability of macrophages to functionally M1/M2 polarization and reprogramming plays a crucial role in this competition.The adequacy of the research results is confirmed by quantitative and qualitative agreement with experimental data.
... Initially, this work consisted in studying the effect of different concentrations of MGO on cardiac function in the Langendorff model. This ex vivo technique allowed us to assess the direct impacts of MGO without the complications associated with in vivo models and by avoiding the confounding effect of the neuronal and hormonal environment of the living animal [34,35]. In this first exploratory experiment, the toxicity of MGO described a concentrationdependent effect relationship. ...
Preprint
Methylglyoxal (MGO) is an endogenous, highly reactive dicarbonyl metabolite generated under hyperglycaemic conditions. MGO plays a role in developing pathophysiological conditions, including diabetic cardiomyopathy. However, the mechanisms involved and the molecular targets of MGO in the heart have not been elucidated. In this work, we studied the exposure-related effects of MGO on cardiac function in an isolated perfused rat heart ex vivo model. The effect of MGO on calcium homeostasis in cardiomyocytes was studied in vitro by the fluorescence indicator of intracellular calcium Fluo-4. We demonstrated that MGO induced cardiac dysfunction, both in contractility and diastolic function. In rat heart, the effects of MGO treatment were significantly limited by aminoguanidine, a scavenger of MGO, ruthenium red, a general cation channel blocker, and verapamil, an L-type voltage-dependent calcium channel blocker, demonstrating that this dysfunction involved alteration of calcium regulation. MGO induced a significant concentration-dependent increase of intracellular calcium in neonatal rat cardiomyocytes, which was limited by aminoguanidine and verapamil. These results suggest that the functionality of various calcium channels is altered by MGO, particularly the L-type calcium channel, thus explaining its cardiac toxicity. Therefore, MGO could participate in the development of diabetic cardiomyopathy through its impact on calcium homeostasis in cardiac cells.
... By utilizing various approaches to inhibit RUNX1 in mice following MI, including an adenoviral vector expressing Runx1-shRNA, a cardiotropic adenoassociated virus serotype 9 (AAV9) expressing a shRNA targeting Runx1, and a small molecule inhibitor (Ro5-3335), this study showed that both gene therapeutic and pharmacological approaches to antagonize RUNX1 preserve cardiac function following MI (Martin et al. 2023). However, although this study included infarct size measurement at 7-days post-MI performed by Sirius red staining on fixed histological sections (Martin et al. 2023), the acute ischemic injury and the tissue viability following MI, which can be assessed by triphenyltetrazolium chloride (TTC) staining for fresh tissue sample (Lindsey et al. 2018), has not been examined. Since Runx1 expression is increased as early as 1-day post-MI, its impact on infarct size at the early stage of tissue injury post-MI needs to be evaluated. ...
Article
Full-text available
Myocardial infarction (MI) results in prolonged ischemia and the subsequent cell death leads to heart failure which is linked to increased deaths or hospitalizations. New therapeutic targets are urgently needed to prevent cell death and reduce infarct size among patients with MI. Runt-related transcription factor-1 (RUNX1) is a master-regulator transcription factor intensively studied in the hematopoietic field. Recent evidence showed that RUNX1 has a critical role in cardiomyocytes post-MI. The increased RUNX1 expression in the border zone of the infarct heart contributes to decreased cardiac contractile function and can be therapeutically targeted to protect against adverse cardiac remodelling. This study sought to investigate whether pharmacological inhibition of RUNX1 function has an impact on infarct size following MI. In this work we demonstrate that inhibiting RUNX1 with a small molecule inhibitor (Ro5-3335) reduces infarct size in an in vivo rat model of acute MI. Proteomics study using data-independent acquisition method identified increased cathepsin levels in the border zone myocardium following MI, whereas heart samples treated by RUNX1 inhibitor present decreased cathepsin levels. Cathepsins are lysosomal proteases which have been shown to orchestrate multiple cell death pathways. Our data illustrate that inhibition of RUNX1 leads to reduced infarct size which is associated with the suppression of cathepsin expression. This study demonstrates that pharmacologically antagonizing RUNX1 reduces infarct size in a rat model of acute MI and unveils a link between RUNX1 and cathepsin-mediated cell death, suggesting that RUNX1 is a novel therapeutic target that could be exploited clinically to limit infarct size after an acute MI.
... Afterward blood was collected for detection of troponin I levels and hearts were removed for subsequent analyses. The extent of infarct size was determined according to the state of the art as defined by current guidelines by calculating the percentage of infarction compared with the area at risk (AAR) [13,[59][60][61]. For this purpose, a double staining technique using triphenyl tetrazolium chloride (TTC) to mark vital and necrotic tissue and Evans Blue staining to negatively mark the AAR was used [27]. ...
Article
Full-text available
Neutrophils are not only involved in immune defense against infection but also contribute to the exacerbation of tissue damage after ischemia and reperfusion. We have previously shown that genetic ablation of regulatory Gαi proteins in mice has both protective and deleterious effects on myocardial ischemia reperfusion injury (mIRI), depending on which isoform is deleted. To deepen and analyze these findings in more detail the contribution of Gαi2 proteins in resident cardiac vs circulating blood cells for mIRI was first studied in bone marrow chimeras. In fact, the absence of Gαi2 in all blood cells reduced the extent of mIRI (22,9% infarct size of area at risk (AAR) Gnai2−/− → wt vs 44.0% wt → wt; p < 0.001) whereas the absence of Gαi2 in non-hematopoietic cells increased the infarct damage (66.5% wt → Gnai2−/−vs 44.0% wt → wt; p < 0.001). Previously we have reported the impact of platelet Gαi2 for mIRI. Here, we show that infarct size was substantially reduced when Gαi2 signaling was either genetically ablated in neutrophils/macrophages using LysM-driven Cre recombinase (AAR: 17.9% Gnai2fl/fl LysM-Cre+/tg vs 42.0% Gnai2fl/fl; p < 0.01) or selectively blocked with specific antibodies directed against Gαi2 (AAR: 19.0% (anti-Gαi2) vs 49.0% (IgG); p < 0.001). In addition, the number of platelet-neutrophil complexes (PNCs) in the infarcted area were reduced in both, genetically modified (PNCs: 18 (Gnai2fl/fl; LysM-Cre+/tg) vs 31 (Gnai2fl/fl); p < 0.001) and in anti-Gαi2 antibody-treated (PNCs: 9 (anti-Gαi2) vs 33 (IgG); p < 0.001) mice. Of note, significant infarct-limiting effects were achieved with a single anti-Gαi2 antibody challenge immediately prior to vessel reperfusion without affecting bleeding time, heart rate or cellular distribution of neutrophils. Finally, anti-Gαi2 antibody treatment also inhibited transendothelial migration of human neutrophils (25,885 (IgG) vs 13,225 (anti-Gαi2) neutrophils; p < 0.001), collectively suggesting that a therapeutic concept of functional Gαi2 inhibition during thrombolysis and reperfusion in patients with myocardial infarction should be further considered.
Article
Full-text available
Since the invention of cardiopulmonary bypass, cardioprotective strategies have been investigated to mitigate ischemic injury to the heart during aortic cross-clamping and reperfusion injury with cross-clamp release. With advances in cardiac surgical and percutaneous techniques and post-operative management strategies including mechanical circulatory support, cardiac surgeons are able to operate on more complex patients. Therefore, there is a growing need for improved cardioprotective strategies to optimize outcomes in these patients. This review provides an overview of the basic principles of cardioprotection in the setting of cardiac surgery, including mechanisms of cardiac injury in the context of cardiopulmonary bypass, followed by a discussion of the specific approaches to optimizing cardioprotection in cardiac surgery, including refinements in cardiopulmonary bypass and cardioplegia, ischemic conditioning, use of specific anesthetic and pharmaceutical agents, and novel mechanical circulatory support technologies. Finally, translational strategies that investigate cardioprotection in the setting of cardiac surgery will be reviewed, with a focus on promising research in the areas of cell-based and gene therapy. Advances in this area will help cardiologists and cardiac surgeons mitigate myocardial ischemic injury, improve functional post-operative recovery, and optimize clinical outcomes in patients undergoing cardiac surgery.
Article
Maternal mortality rates are at an all-time high across the world and are set to increase in subsequent years. Cardiovascular disease is the leading cause of death during pregnancy and postpartum, especially in the US. Therefore, understanding the physiological changes in the cardiovascular system during normal pregnancy is necessary to understand disease-related pathology. Significant systemic and cardiovascular physiological changes occur during pregnancy that are essential for supporting the maternal-fetal dyad. The physiological impact of pregnancy on the cardiovascular system has been examined in both experimental animal models and in humans. However, there is a continued need in this field of study to provide increased rigor and reproducibility. Therefore, these guidelines aim to provide information regarding best practices and recommendations to accurately and rigorously measure cardiovascular physiology during normal and cardiovascular-disease-complicated pregnancies in human and animal models.
Article
The intracardiac nervous system is considered to be involved in ischemic conditioning’s cardioprotection through the release of acetylcholine (ACh). However, we demonstrate that hypoxic preconditioning (HPC) protects from hypoxia/reoxygenation injury and increases intra- and extracellular ACh during hypoxia in isolated adult ventricular rat cardiomyocytes. HPC’s protection involves cardiomyocyte muscarinic and nicotinic ACh receptor activation. Thus, besides the intracardiac nervous system, a nonneuronal cholinergic cardiac system may also be causally involved in cardiomyocyte protection by ischemic conditioning.
Article
Full-text available
Cardiovascular disease is a leading cause of death, and translational research is needed to understand better mechanisms whereby the left ventricle responds to injury. Mouse models of heart disease have provided valuable insights into mechanisms that occur during cardiac aging and in response to a variety of pathologies. The assessment of cardiovascular physiological responses to injury or insult is an important and necessary component of this research. With increasing consideration for rigor and reproducibility, the goal of this guidelines review is to provide best-practice information regarding how to measure accurately cardiac physiology in animal models. In this article, we define guidelines for the measurement of cardiac physiology in mice, as the most commonly used animal model in cardiovascular research. Listen to this article’s corresponding podcast at http://ajpheart.podbean.com/e/guidelines-for-measuring-cardiac-physiology-in-mice/.
Article
Full-text available
The controversy regarding the mechanism(s) of left ventricular (LV) dysfunction in chronic coronary artery disease is, in part, related to the lack of an appropriate animal model for this condition. We have developed such a model by placing Ameroid constrictors on proximal portions of coronary arteries in dogs who were euthanized (mean of 6 wk) after the development of severe global LV dysfunction noted on two-dimensional echocardiography. The LV end-systolic size nearly doubled ( P < 0.001) over the observation period, and the percent change in LV size from end diastole to end systole decreased by >50% ( P< 0.001). Regional dysfunction was noted in 23 of 24 myocardial beds analyzed within regions showing no gross evidence of infarction. In 10 of these beds, severe dysfunction was noted without a decrease in radiolabeled microsphere-derived myocardial blood flow (MBF). In 13 myocardial beds, decrease in function was associated with a decrease in MBF ( P < 0.001), with close coupling noted between percent wall thickening and MBF. In the beds that exhibited an ultimate decrease in MBF, the decrease in function preceded the decrease in MBF. In conclusion, we describe chronic LV dysfunction in a canine model of multivessel stenosis that closely mimics chronic ischemic LV dysfunction in humans. Whereas regional function is severely reduced in this model, MBF is varied in different segments and at different times during the observation period. These results provide new insights regarding flow-function relations in chronic ischemic LV dysfunction.
Article
Full-text available
Ischemic conditioning before (ischemic preconditioning, IPC) or after (ischemic postconditioning, POCO) sustained myocardial ischemia/reperfusion (I/R), induced locally or remotely from the heart (remote IPC, RIPC), reduces infarct size. However, none of the identified signaling steps of ischemic conditioning was robust across models and species to be successfully translated to humans. In prior separate studies in pigs, activation of signal transducer and activator of transcription 3 (STAT3) was causal for infarct size reduction by IPC, POCO, and RIPC but it remains unclear whether or not STAT3 is truly a common denominator of cardioprotective signaling. We therefore, now analyzed the phosphorylation of STAT3 and other signaling proteins in left ventricular biopsies from our prior studies on IPC, POCO and RIPC in one approach. We developed a strategy for the quantification of protein phosphorylation in multiple samples from many experiments on different gels/membranes by Western blot. Along with reduced infarct size, the ratio of STAT3tyr705 phosphorylation/total STAT3 protein at early reperfusion was significantly increased by IPC (IPC 2.0 ± 0.3 vs. I/R 1.2 ± 0.2 arbitrary units), but only trendwise by POCO and RIPC (1.3 ± 0.2; 1.4 ± 0.2 arbitrary units); storage time for IPC samples was shorter than for POCO and RIPC samples. No other signaling protein phosphorylation was associated with reduced infarct size. We confirmed STAT3 phosphorylation with IPC. For POCO and RIPC we could not reproduce the findings from our earlier more focused studies. At this point, we can not distinguish between lack of robustness of the biological signal and methodological issues of our retrospective approach.
Article
Full-text available
Heart failure (HF) secondary to myocardial infarction (MI) is linked to kidney complications that comprise cellular, structural, functional, and survival indicators. However, HF research is focused on left ventricular (LV) pathology. Here, we determined comprehensive functional analysis of the LV using echocardiography in transition from acute heart failure (AHF) to progressive chronic heart failure (CHF) pathology and developed a histological compendium of the cardiosplenic and cardiorenal networks in pathological remodeling. In surgically induced MI using permanent coronary ligation, the LV dysfunction is pronounced, with myocardium necrosis, wall thinning, and 20–30% LV rupture events that indicated AHF and CHF pathological remodeling in C57BL/6 male mice (2–4 mo old, n = 50). Temporal LV function analysis indicated that fractional shortening and strain are reduced from day 1 to day 5 in AHF and sustained to advance to CHF from day 28 to day 56 compared with naïve control mice ( n = 6). During the transition of AHF ( day 1 to day 5) to advanced CHF ( day 28 to day 56), histological and cellular changes in the spleen were definite, with bimodal inflammatory responses in kidney inflammatory biomarkers. Likewise, there was a unidirectional, progressive, and irreversible deposition of compact collagen in the LV along with dynamic changes in the cardiosplenic and cardiorenal networks post-MI. The renal histology and injury markers suggested that cardiac injury triggers irreversible dysregulation that actively alters the cardiosplenic and cardiorenal networks. In summary, the novel strategies or pathways that modulate comprehensive cardiosplenic and cardiorenal networks in AHF and CHF would be effective approaches to study either cardiac repair or cardiac pathology. NEW & NOTEWORTHY The present compendium shows irreversible ventricular dysfunction as assessed by temporal echocardiography while histological and structural measurements of the spleen and kidney added a novel direction to study cardiosplenic and cardiorenal networks in heart failure pathology. Therefore, the consideration of systems biology and integrative approach is essential to develop novel treatments. Listen to this article's corresponding podcast at http://ajpheart.podbean.com/e/temporal-dynamics-of-acute-and-chronic-heart-failure/ .
Article
Full-text available
Vascular endothelial growth factor (VEGF) is a well-characterized proangiogenic cytokine that has been shown to promote neovascularization in hearts of patients with ischemic heart disease but can also lead to adverse effects depending on the dose and mode of delivery. We investigated whether prolonged exposure to a low dose of VEGF could be achieved by encapsulating VEGF in polylactic coglycolic acid nanoparticles and whether treatment with VEGF-containing nanoparticles improved cardiac function and protected against left ventricular remodeling in the hearts of mice with experimentally induced myocardial infarction. Polylactic coglycolic acid nanoparticles with a mean diameter of ~113 nm were generated via double emulsion and loaded with VEGF; the encapsulation efficiency was 53.5 ± 1.7% (107.1 ± 3.3 ng VEGF/mg nanoparticles). In culture, VEGF nanoparticles released VEGF continuously for at least 31 days, and in a murine myocardial infarction model, VEGF nanoparticle administration was associated with significantly greater vascular density in the peri-infarct region, reductions in infarct size, and improvements in left ventricular contractile function 4 wk after treatment. Thus, our study provides proof of principle that nanoparticle-mediated delivery increases the angiogenic and therapeutic potency of VEGF for the treatment of ischemic heart disease. NEW & NOTEWORTHY Vascular endothelial growth factor (VEGF) is a well-characterized proangiogenic cytokine but has a short half-life and a rapid clearance rate. When encapsulated in nanoparticles, VEGF was released for 31 days and improved left ventricular function in infarcted mouse hearts. These observations indicate that our new platform increases the therapeutic potency of VEGF.
Article
Full-text available
Remote ischemic preconditioning (RIPC) by repeated brief cycles of limb ischemia/reperfusion reduces myocardial ischemia/reperfusion injury. In left ventricular (LV) biopsies from patients undergoing coronary artery bypass grafting (CABG), only the activation of signal transducer and activator of transcription 5 was associated with RIPC’s cardioprotection. We have now used an unbiased, non-hypothesis-driven proteomics and phosphoproteomics approach to analyze LV biopsies from patients undergoing CABG and from pigs undergoing coronary occlusion/reperfusion without (sham) and with RIPC. False discovery rate-based statistics identified a higher prostaglandin reductase 2 expression at early reperfusion with RIPC than with sham in patients. In pigs, the phosphorylation of 116 proteins was different between baseline and early reperfusion with RIPC and/or with sham. The identified proteins were not identical for patients and pigs, but in-silico pathway analysis of proteins with ≥2-fold higher expression/phosphorylation at early reperfusion with RIPC in comparison to sham revealed a relation to mitochondria and cytoskeleton in both species. Apart from limitations of the proteomics analysis per se, the small cohorts, the sampling/sample processing and the number of uncharacterized/unverifiable porcine proteins may have contributed to this largely unsatisfactory result.
Article
The past two decades have witnessed an explosive growth of knowledge regarding postischemic myocardial dysfunction or myocardial “stunning.” The purpose of this review is to summarize current information regarding the pathophysiology and pathogenesis of this phenomenon. Myocardial stunning should not be regarded as a single entity but rather as a “syndrome” that has been observed in a wide variety of experimental settings, which include the following: 1) stunning after a single, completely reversible episode of regional ischemia in vivo; 2) stunning after multiple, completely reversible episodes of regional ischemia in vivo; 3) stunning after a partly reversible episode of regional ischemia in vivo (subendocardial infarction); 4) stunning after global ischemia in vitro; 5) stunning after global ischemia in vivo; and 6) stunning after exercise-induced ischemia (high-flow ischemia). Whether these settings share a common mechanism is unknown. Although the pathogenesis of myocardial stunning has not been definitively established, the two major hypotheses are that it is caused by the generation of oxygen-derived free radicals (oxyradical hypothesis) and by a transient calcium overload (calcium hypothesis) on reperfusion. The final lesion responsible for the contractile depression appears to be a decreased responsiveness of contractile filaments to calcium. Recent evidence suggests that calcium overload may activate calpains, resulting in selective proteolysis of myofibrils; the time required for resynthesis of damaged proteins would explain in part the delayed recovery of function in stunned myocardium. The oxyradical and calcium hypotheses are not mutually exclusive and are likely to represent different facets of the same pathophysiological cascade. For example, increased free radical formation could cause cellular calcium overload, which would damage the contractile apparatus of the myocytes. Free radical generation could also directly alter contractile filaments in a manner that renders them less responsive to calcium (e.g., oxidation of critical thiol groups). However, it remains unknown whether oxyradicals play a role in all forms of stunning and whether the calcium hypothesis is applicable to stunning in vivo. Nevertheless, it is clear that the lesion responsible for myocardial stunning occurs, at least in part, after reperfusion so that this contractile dysfunction can be viewed, in part, as a form of “reperfusion injury.” An important implication of the phenomenon of myocardial stunning is that so-called chronic hibernation may in fact be the result of repetitive episodes of stunning, which have a cumulative effect and cause protracted postischemic dysfunction. A better understanding of myocardial stunning will expand our knowledge of the pathophysiology of myocardial ischemia and provide a rationale for developing new therapeutic strategies designed to prevent postischemic dysfunction in patients.
Article
Cardiac sympathetic nerves stimulate heart rate and force of contraction. Myocardial infarction (MI) leads to the loss of sympathetic nerves within the heart, and clinical studies indicate that sympathetic denervation is a risk factor for arrhythmias and cardiac arrest. Two distinct types of denervation have been identified in the mouse heart after MI caused by ischemia-reperfusion: transient denervation of peri-infarct myocardium, and sustained denervation of the infarct. Sustained denervation is linked to increased arrhythmia risk, but it is not known if acute nerve loss in peri-infarct myocardium also contributes to arrhythmia risk. Peri-infarct sympathetic denervation requires the p75 neurotrophin receptor (p75NTR), but removal of p75NTR alters the pattern of sympathetic innervation in the heart and increases spontaneous arrhythmias. Therefore, we targeted the p75NTR co-receptor sortilin, and the p75NTR-induced protease TACE/ADAM17 (Tumor necrosis factor-α Converting Enzyme/A disintegrin and metalloproteinase domain 17) in order to selectively block peri-infarct denervation. Sympathetic nerve density was quantified using immunohistochemistry for tyrosine hydroxylase. Genetic deletion of sortilin had no effect on the timing or extent of axon degeneration, but inhibiting TACE/ADAM17 with the protease inhibitor marimastat prevented the loss of axons from viable myocardium. We then asked if retention of nerves in peri-infarct myocardium had an impact on cardiac electrophysiology three days after MI, using ex vivo optical mapping of transmembrane potential and intracellular calcium. Preventing acute denervation of viable myocardium after MI did not significantly alter cardiac electrophysiology or calcium handling, suggesting that transient denervation at this early timepoint has minimal impact on arrhythmia risk.
Article
Matrix metalloproteinase (MMP)-9 increases in the myocardium with advanced age and after MI. Because young transgenic (TG) mice overexpressing human MMP-9 only in macrophages mice show better outcomes post-MI, while aged TG mice show a worse aging phenotype, we wanted to evaluate the combination effect to see if the detrimental effect of aging counteracted the benefits of macrophage MMP-9 overexpression. We used 17-28 month old male and female C57BL/6J wild type (WT) mice and TG mice (n = 10-21/group) to evaluate the effects of aging superimposed on MI. Despite similar infarct areas and mortality rates at day (D) 7 post-MI, TG aging mice showed improved diastolic properties and remodeling index compared to WT (both P<0.05). Macrophage numbers were higher in TG compared to WT at both D0 and D7 post-MI, and the post-MI increase was due to elevated CD18 protein levels (all P<0.05). RNA-seq analysis of cardiac macrophages isolated from D7 post-MI infarcts identified 1,276 statistically different genes (994 increased and 282 decreased in TG, all P<0.05). Reduced vascular endothelial growth factor-A (VEGFA), platelet-derived growth factor subunit A (Pdgfa), and transforming growth factor beta 3 (Tgfb3) along with elevated tissue inhibitor of MMP-4 (Timp4) macrophage expression reveal mechanisms of indirect downstream effects on fibroblasts and neovascularization. While collagen accumulation was enhanced in TG compared to WT at both D0 and D7 post-MI (P<0.05 for both), the post-MI collagen cross-linking ratio was higher in the WT (P<0.05), consistent with increased diastolic volumes. Vessel numbers (by GSL-I lectin staining) were decreased in the TG compared to WT at both D0 and D7 post-MI (P<0.05 for both). In conclusion, macrophage-derived MMP-9 improved post-MI cardiac wound healing through direct and indirect mechanisms to improve diastolic physiology and remodeling.