Next Article in Journal
Monocyte and Macrophage in Neuroblastoma: Blocking Their Pro-Tumoral Functions and Strengthening Their Crosstalk with Natural Killer Cells
Next Article in Special Issue
Amazonia Phytotherapy Reduces Ischemia and Reperfusion Injury in the Kidneys
Previous Article in Journal
Environmental Enrichment Engages Vesicular Zinc Signaling to Enhance Hippocampal Neurogenesis
Previous Article in Special Issue
Immune Cells Are Differentially Modulated in the Heart and the Kidney during the Development of Cardiorenal Syndrome 3
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ischemic Tolerance—A Way to Reduce the Extent of Ischemia–Reperfusion Damage

by
Rastislav Burda
1,2,*,
Jozef Burda
3 and
Radoslav Morochovič
1,2
1
Department of Trauma Surgery, Faculty of Medicine, Pavol Jozef Šafárik University in Košice, Rastislavova 43, 040 01 Košice, Slovakia
2
Department of Trauma Surgery, Louis Pasteur University Hospital, Rastislavova 43, 040 01 Košice, Slovakia
3
Institute of Neurobiology, Slovak Academy of Sciences, 040 01 Košice, Slovakia
*
Author to whom correspondence should be addressed.
Cells 2023, 12(6), 884; https://doi.org/10.3390/cells12060884
Submission received: 21 February 2023 / Revised: 8 March 2023 / Accepted: 11 March 2023 / Published: 13 March 2023

Abstract

:
Individual tissues have significantly different resistance to ischemia–reperfusion damage. There is still no adequate treatment for the consequences of ischemia–reperfusion damage. By utilizing ischemic tolerance, it is possible to achieve a significant reduction in the extent of the cell damage due to ischemia–reperfusion injury. Since ischemia–reperfusion damage usually occurs unexpectedly, the use of preconditioning is extremely limited. In contrast, postconditioning has wider possibilities for use in practice. In both cases, the activation of ischemic tolerance can also be achieved by the application of sublethal stress on a remote organ. Despite very encouraging and successful results in animal experiments, the clinical results have been disappointing so far. To avoid the factors that prevent the activation of ischemic tolerance, the solution has been to use blood plasma containing tolerance effectors. This plasma is taken from healthy donors in which, after exposure to two sublethal stresses within 48 h, effectors of ischemic tolerance occur in the plasma. Application of this activated plasma to recipient animals after the end of lethal ischemia prevents cell death and significantly reduces the consequences of ischemia–reperfusion damage. Until there is a clear chemical identification of the end products of ischemic tolerance, the simplest way of enhancing ischemic tolerance will be the preparation of activated plasma from young healthy donors with the possibility of its immediate use in recipients during the initial treatment.

1. Ischemia–Reperfusion (IR) Injury

The term ischemia–reperfusion injury refers to a condition in which tissue ischemia occurs with insufficient oxygen supply for the tissue’s current metabolic needs, followed by the successful restoration of blood flow.
Ischemia with reperfusion initiates a wide complex of changes of an inflammatory nature that can worsen local conditions and also lead to dysfunction of a distant organ or the entire organism.
Ischemia occurs in acute arterial occlusions (an ischemic cerebrovascular accident, myocardial infarction, and ischemia of a limb), where reperfusion subsequently occurs as part of treatment (thrombolytic treatment, angioplasty, surgical revascularization, and operations with tourniquet application) [1]. Moreover, common surgical procedures (organ transplantation, transfer of free flaps, cardiopulmonary bypass, vascular operations, or polytrauma patients) lead to ischemia–reperfusion damage [2].
In transplant recipients, attention has been focused on post-transplant patient management in the past, with an obvious emphasis on immunosuppression, but “the biggest hit” to the donor organ is encountered during the process of donation and reperfusion at the time of transplantation. The problem is in the uncontrolled ROS (reactive oxygen species) formation during the reperfusion phase, which leads to mitochondrial disfunction (opening of mPTP/mitochondrial transition pores/and the release of DAMPs/damage-associated molecular patterns/in the intra- and extracellular space) [3,4,5,6].
In cardiomyocyte tissue cultures, 4 h of simulated ischemia resulted in the death of up to 17% of the cells, but the subsequent reperfusion resulted in the death of up to 73% of the ischemia-affected cells in the experiment [7]. More cells died because of the subsequent reperfusion than from the ischemia alone. Some of the cells of the ischemic tissue remain irreversibly damaged as a result of the action of ischemia. These cells inevitably die. The second group of cells, resulting from IR damage, is potentially vital but with a significantly increased risk of cell death. This second group represents cells that we can influence therapeutically (Figure 1).

2. Current Therapeutic Options for Influencing IR Damage

In the case of acute ischemic damage, it is important from a clinical point of view to shorten the duration of the ischemia as much as possible, correct the metabolic acidosis, and reduce the risk of acute renal failure and systemic inflammatory reaction (SIRS).
Currently, many experimental and clinical methodologies have been described to reduce or eliminate the extent of the IR damage. Some substances or procedures can only be used to prevent ischemia, while others can only be used to reduce the extent of the IR damage.
A list of options to reduce the range of IR damage in different ways follows:
  • A NO (Nitric Oxide) protective strategy [8];
  • Adenosine [9];
  • The influence of nuclear transcription factors [10];
  • The inhibition of apoptosis [11];
  • The inhibition of Ca 2+ excess in the cell [12];
  • Antioxidants [13,14];
  • Inhibitors of Na+ H+ channels [15];
  • Controlled reperfusion/reoxygenation [16];
  • Intermittent ischemia [17];
  • The depletion of neutrophil cells [18,19];
  • Aprotinin [20];
  • Poly (ADP-ribose) polymerase (PARP inhibitors) [21];
  • The blockade of the complement system [22];
  • The use of anesthetics [23,24];
  • Hypothermia [25,26];
  • Hyperbaric oxygen therapy [27];
  • MicroRNA as a therapeutic target [28,29];
  • Mesenchymal stromal cells therapy [30,31];
  • Ischemic tolerance (conditioning).
Most of the research in the field of IR damage is devoted to research on ischemia of the heart and brain. Despite the enormous progress in the understanding of IR damage, there has been no established clinically standard targeted therapy that leads in practice to an unequivocal suppression of IR damage. The problem is the lack of knowledge of the complex multifactorial damage in IR, while most treatment attempts are devoted to the blockade of only a certain mechanism of IR damage. The transfer of knowledge from experimental animals to humans is also problematic because it does not consider the complexity of the ischemic damage, the age of patients, or the comorbidity and co-medication of patients, which can significantly affect the treatment of ischemia.

3. Ischemic Tolerance

The use of ischemic tolerance seems to be a reproducible methodology that can adequately protect the organism from IR damage. Ischemic tolerance represents a robust internal defense mechanism and a state in which the cells are resistant to the devastating effect of ischemia, which would lead to cell death [32]. This is one of the forms of evolutionary adaptability [33]. Cells exposed to metabolic stress or sublethal ischemia (preconditioning) become temporarily resistant to the action of the subsequent lethal stress.
The power of protection obtained by ischemic tolerance is surprisingly large. Five minutes of global brain ischemia in rats killed nearly 40% of the most sensitive cells in the brain (CA1 of the hippocampus). A ten-minute global ischemia of the brain led to the death of up to 70% of these cells. The combination of two stresses, either as the action of smaller ischemia first (preconditioning), or, on the contrary, stronger ischemia first with the subsequent action of weaker ischemia (postconditioning), did not lead to a cumulative destructive effect and cell death but to the rescue of these cells [34].
The decisive factor for the activation of ischemic tolerance is the combination of two ischemic or metabolic stresses [34,35]. Activation of ischemic tolerance is a two-step process. The first stress is necessary for the initiation of ischemic tolerance, but for the activation of complete ischemic tolerance, the action of the second stress is absolutely necessary. The advantage is that these two stresses do not have to be of the same nature (cross tolerance), and it is also not important whether they affect the whole organism or only locally (distant tolerance).
According to the time in which the sublethal ischemia acts relative to the lethal ischemia, we can distinguish the following mechanisms that lead to the initiation of ischemic tolerance (Figure 2):
  • Ischemic preconditioning (IpreC) refers to the condition where the sublethal ischemia acts before the lethal ischemia itself;
  • Ischemic perconditioning refers to the condition where the sublethal ischemia and the lethal ischemia act simultaneously;
  • Ischemic postconditioning (IP) refers to the condition when the sublethal ischemia occurs after the lethal ischemia.
Figure 2. A schematic representation of the different ways of initiating ischemic tolerance. According to the time in which the sublethal ischemia acts relative to the lethal ischemia, we can distinguish the following mechanisms that lead to the initiation of ischemic tolerance. I—ischemic damage, IPreC—ischemic preconditioning, IP—ischemic postconditioning.
Figure 2. A schematic representation of the different ways of initiating ischemic tolerance. According to the time in which the sublethal ischemia acts relative to the lethal ischemia, we can distinguish the following mechanisms that lead to the initiation of ischemic tolerance. I—ischemic damage, IPreC—ischemic preconditioning, IP—ischemic postconditioning.
Cells 12 00884 g002

4. Cross Tolerance

Cross tolerance allows sublethal ischemia in any conditioning modality to be replaced by any one of many biological, physical, or chemical stressors.
Many stimuli or a combination of different stresses can lead to the activation of ischemic tolerance. Cross tolerance works both in the case of preconditioning and postconditioning.
Preconditioning can be induced by stimuli other than ischemia; for example, hypothermia, hyperthermia, hypoxia or hyperoxia, electroacupuncture, or exercise are referred to as cross preconditioning [33].

4.1. Hyperoxic and Hypoxic Preconditioning

Normobaric hyperoxia (95% O2) in rats reduced the extent of the cerebral infarct size [36]. The induction of tolerance in this case followed the path of genetic reprogramming [37].
Hyperbaric oxygenation (HBO) indicated tolerance by reducing cell apoptosis and suppressing COX-2 expression in global brain ischemia in rats [36].

4.2. Preconditioning Induced by Hyperthermia and Hypothermia

Hypothermia as preconditioning can activate rapidly (early tolerance); however, prolonging the duration of hypothermia does not significantly increase the neuroprotective effect. Instead, increasing the degree of hypothermia is more important [38]. Prolonged ischemic tolerance in this case probably acts through the proteosynthesis pathway.
Hyperthermia can also stimulate tolerance. Hyperthermia of young rats using a warm bath (41.5–42 °C) reduced neuronal damage after an ischemic insult 2 h after hyperthermia [39].
In human liver transplantation, livers were either conventionally cold stored or cold stored and subsequently treated by 1–2 h hypothermic oxygenated perfusion (HOPE) before implantation. Exploratory findings suggest that HOPE decreases the risk of severe liver graft-related events (fewer severe liver graft-related complications and a lower risk of liver-related graft loss) within a 12-month period after liver transplantation. The HOPE approach can be applied easily after organ transport during recipient hepatectomy [40]. Additionally, extracellular vesicles (EV) derived from mesenchymal stromal cells delivered during HOPE can be considered a new organ preservation strategy for increasing the donor pool and improving the transplant outcome because the gluconeogenesis system in HOPE+EV group was preserved [41].

4.3. Chemical/Pharmacological Preconditioning

Isoflurane, halothane, and other inhalation anesthetics lead to the activation of ischemic tolerance; they act as antagonists of the NMDA and AMPA receptors [42].
Sevoflurane reduced the extent of the infarct focus and improved the coordination of experimental animals after ischemia. Sevoflurane might exert beneficial effects on cerebral IR-induced neuronal injury through inhibiting the surface expression of the glutamate receptor 1 and blocking CP-AMPAR (calcium permeble-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor). Halothane is hepatotoxic; therefore, its clinical utility is limited, despite the eventual neuroprotective effect [43,44].
Low doses of liposaccharides can also activate ischemic tolerance in the rat brain [45]. Adenosine receptor agonists can also promote the release of adenosine and the activation of ATP-sensitive K+ channels, leading to the activation of ischemic tolerance [46]. A similar effect of pharmacological preconditioning can be achieved by 3-nitropropionic acid [34,47] but also by the application of bradykinin [48] and metformin [49].

4.4. Electroacupuncture

Continuous application of electroacupuncture can also act as adequate stress to induce preconditioning [50]. Physical activity, exercise, and transcranial light stimulation [51,52] can also act as adequate stress to induce preconditioning.

4.5. Multipotent Adult Progenitor Cells and Tissue Engineering

Normothermic machine perfusion (NMP) of donor kidneys prior to their transplantation provides a good platform for the direct delivery of cellular therapeutics to optimize the organ quality prior to transplantation.
Multipotent Adult Progenitor Cells (MAPC®) possess potent immunomodulatory properties that could minimize ischemia–reperfusion injury. Kidneys treated with MAPC cells demonstrated improvement in the clinically relevant parameters and injury biomarkers. This novel method of cell therapy delivery provides an exciting opportunity to recondition organs prior to transplantation [53].
Kidney conditioning with mesenchymal stromal cell (MSCs)/extracellular vesicle (EVs) delivery during hypothermic perfusions protected against ischemic damage by activating the CD73/ADO system [54].
The new strategy of including trimetazidine, tacrolimus, and carvedilol would allow us to shift from cold storage solutions to cold preservation solutions including multitarget pharmacological components, offering protection against IRI and thus protecting the more vulnerable organs [6,55,56].
Tissue engineering offers a promising toolset to tackle ischemia–reperfusion injuries. It devises tissue mimetics by using the unique therapeutic features of stem cells, functional biomaterials, and bioprocesses to design tools to emulate the macroscopic environment that interacts with tissues. This strategy allows the production of cell therapeutics capable of addressing ischemia–reperfusion injury [57].

5. Remote Conditioning

As in the case of preconditioning, postconditioning can also lead to the activation of the phenomenon of ischemic tolerance by an attack of sublethal ischemia in a distant organ. Kerendi described for the first time the possibility of postconditioning by clamping the renal artery after 30 min of cardiac artery occlusion. This mechanism of interorgan postconditioning is referred to as remote postconditioning (RP) [58]. It is the application of temporary and short-lasting ischemia on a remote organ (for example, a limb), which can activate resistance to the IR damage on the target organ or tissue.

6. Mechanism of the Operation of Local Ischemic Preconditioning

Ischemic preconditioning includes a complex cascade of mechanisms, while their induction also depends on the length of the acting ischemia. Several pathways by which preconditioning can be activated have been described.

6.1. Hypoxia-Inducible Factor (HIF-1α)

Hypoxia triggers the production of HIF-1α, which is an oxygen-dependent transcription factor. IPreC has a different effect on the expression of HIF-1α in different cells; its expression in neurons is fast, while in astrocytes it is slower but continuous [59,60].
HIF-1α is also formed in skeletal muscle after ischemia and during preconditioning through the transcription factor EGR1 (Early Growth Response 1) and the expression of cFOS genes, but it suppresses the expression of cJUN genes [61].
The formation of HIF-1α clearly promotes the expression of endothelial growth factor (VEGF) and the activation of nuclear factor kappa B (NF-κB), which can attenuate neuronal death induced by ischemic attack even in the hippocampus CA1 region, which is the brain region most sensitive to the influence of ischemia [62].

6.2. Glutamate Pathway

This pathway of the activation of ischemic tolerance is based on excitotoxicity glutamate. After the action of ischemia, there is a decrease in the intracellular ATP reserves, while at the same time there is an excessive accumulation of glutamate, which stimulates the N-methyl-D-aspartate (NMDA) receptor, which results in an excessive influx of calcium into the cell, damage to synaptic plasticity, and the accumulation of glutamate [63].
Without adequate activation of the NDMA receptors, the activation of ischemic tolerance will not occur; this activation mechanism includes the NF-κB and the tumor necrosis factor-α (TNF-α) pathway [64].
IPreC can block gap junctions (connections between astrocytes where connexin 43 is present), reduce extracellular glutamate, and reduce ROS in astrocytes, all of which lead to less damage to neurons [65].

6.3. NO Oxide Synthase (NOS)

NO is important in the mechanism of IPreC activation, but the exact mechanism of its action is still unknown. With the application of the NOS inhibitor (7-nitroindazole) after IPReC, the activation of ischemic tolerance was stopped [66].
eNO synthase is considered to protect against vascular spasm caused by a subarachnoid hemorrhage; that is, NOS is involved in the activation of ischemic tolerance by several mechanisms [66].

6.4. Pathway of the CD39-CD73-Adenosine Receptor

Adenosine is an antithrombotic agent, which protects against oxidative stress and suppresses the immune response. Evidently, therapies that promote adenosine generation or boost CD39 activity at the site of endothelial injury have promising benefits in the context of atherothrombotic stroke and can be extended to the current cerebral small-vessel disease’s known pathomechanisms [67], but its effect was also confirmed in the heart and liver, as mentioned below.
CD39 and CD73 (two of the ectonucleotidases) are involved in converting ATP into extracellular adenosine, which is the key metabolite that accumulates and inhibits the function of important immune cells, including T cells and NK cells, leading to an environment conducive to tumor growth. Consequently, multiple clinical strategies are being explored to target this pathway for the treatment not only of cancer, but also IR injury.
Tissue protection is ensured by CD73 enzyme-dependent adenosine generation, and it is signaled through the adenosine-mediated ADORA2B receptor, which triggers tissue protection. Moreover, the adenosine A2B receptor agonists may be used as a potential preventive therapy against IR injury in flap surgeries [68].
Hepatic ischemia should be also tempered with preconditioning, which is associated with the significant induction of the CD39 transcript, heightened protein expression, and improved outcomes after IR injury. Moreover, mimicking ischemic preconditioning with i.p. apyrase (a soluble ectonucleoside triphosphate diphosphohydrolase, NTPDase) in the absence of preconditioning attenuated hepatic injury after IR injury [69].
Ecto-5′-nucleotidase (CD73), the “pacemaker” enzyme of extracellular adenosine production, plays an important role in protection. Moreover, the use of soluble 5′-nucleotidase may be a potential therapeutic for hepatic ischemia [70]. The effectiveness of ecto-5′-nucleotidase (CD73) is ensured by the activation of cell-surface adenosine receptors (A1AR, A2AAR, A2BAR, A3AR); it is effective not only in the liver but also in cases of myocardial ischemia; so, 5′-nucleotidase or A2BAR agonists can be used as therapy for myocardial ischemia [71]. Ischemic preconditioning increases the level of CD39 in the heart and contributes to cardiac protection. Cardiac-specific expression of CD39 reduced myocardial dysfunction and infarct size following ischemia–reperfusion injury [72].
The production of anti-inflammatory adenosine reduces the donor kidney antigenic load, and these strategies may translate to improved transplant survival. CD39 in the vascular endothelium and in circulating cells, in particular, regulatory T cells (Treg), is upregulated in response to hypoxic stimuli and plays a critical role in regulating the immune response removing the proinflammatory ATP [73].
Administration of sesame could effectively protect kidney from IR injury by inhibiting inflammatory responses, which might be associated with promoting the adenosine–CD39–A2AR signaling pathway [74].

6.5. Immune Pathway

During ischemia, large numbers of inflammatory cells (such as microglia, lymphocytes, and neutrophils) migrate to the site of infarction (ischemia), and these cells mediate inflammatory brain damage [75].
The noncatalytic Toll-like (TLR) receptor pathway can induce the transcription factor NF-κB, which mediates the transcription of cytokines and chemokines. These are responsible for the initiation of the immune response and the formation of the inflammatory cascade.
IPreC itself provides neuroprotection by promoting an anti-inflammatory response by inhibiting the TLR4/NF-κB signaling system [76].

6.6. Enzymes and Receptors

The natural function of cyclooxygenase-2 (COX-2) is to assist in the oxidation of arachidonic acid to the prostaglandin, which has an important role in inflammatory damage after ischemia. IPreC led to the suppression of COX-2 formation in gerbils, thereby suppressing inflammatory manifestations in the ischemic brain lesion [77].
An increase in adenosine can be observed after brain ischemia; however, after IpreC, an increased expression of adenosine receptors has also been observed [78]. It is also assumed that adenosine receptor A1 is involved in the activation of ischemic tolerance. Adenosine kinase (ADK) inhibits the formation of adenosine, and its eventual suppression in the brain can lead to neuroprotection [79].
During reperfusion in the brain, unbound proteins accumulate in the endoplasmic reticulum. IPreC can inhibit stress-induced apoptosis, thereby contributing to the activation of tolerance [80].

6.7. Autophagy and Apoptosis

Autophagy is the process of internalizing and “digesting” one’s own damaged organelles. This process is involved in IPreC mechanisms. Ischemic injury activates autophagy [81]. Interestingly, significant autophagy activation appeared after three IPreC cycles of 5 min [82].
Conversely, apoptosis as a result of IR damage works mainly on a caspase-3-dependent pathway. Ischemia leads to overexpression of caspase-3, but IPreC can suppress this expression, resulting in the formation of ischemic tolerance [83]. Induction of autophagy appears to be a possible method of activating ischemic tolerance.

6.8. Energy Metabolism

During brain ischemia, there is a change in the energy metabolism from oxidative phosphorylation to excessive glycolysis, which allows the generation of ATP to cover the energy needs.
Excessive glycolysis increases the concentration of lactate and ROS, further promoting the expansion of ischemic changes. IPreC suppresses postischemic hyperglycolysis and promotes the utilization of β-hydroxybutyrate, thereby normalizing energy metabolism [84].

6.9. Permeability of the Blood–Brain Barrier

As a result of ischemia, there is a breakdown of the blood–brain barrier, which allows the infiltration of immune cells into the site of the ischemia [85]. IPreC can preserve the paracellular permeability of the membrane by supporting the formation of claudins and adherins, which are part of the tight and adherent connections of this membrane [86].

6.10. Transcriptional Regulation

Some works point to the importance of the transcription of survival genes after ischemia.
After hypoxic stress, HIF-1α and HIF-1β join together to form a HIF-1 heterodimer, which binds to the target genes, forming a transcriptional complex enabling the production of VEGF, erythropoetin, and glucose transporters [87].

6.11. Genetic Reprogramming

IPreC causes not only an increased expression of genes associated with cell protection but also a simultaneous suppression of the expression of the genes associated with cell degeneration after ischemic injury. Heat shock protein 72 (HSP-72) [88] and transforming growth factor (TGF-A) [89] are thought to be involved in this reprogramming.

6.12. Epigenetic Reprogramming

MicroRNA (miRNA) also has an important role in mediating the effect of IPreC. Lethal ischemia itself leads to suppression of miRNA transcription; in contrast, IPreC promotes its creation [90].

6.13. Activation of Ischemic Tolerance in the Brain

In an experimental model, the pre-exposure of animals to an enriched environment led to the activation of preconditioning. The major role was played by the activity-dependent transcription factor Npas4, whose expression is triggered by excessive Ca2+ influx. Furthermore, Npas4 regulates L-type voltage-gated Ca2+ channels through expression of the small Race-like GTPase Gem in ischemic neurons [91].
Preconditioning also initiates glial responses, especially in astrocytes, which transforms them into an ischemia-resistant phenotype. P2X7 receptors (P2X7Rs) in astrocytes play a main role in preconditioning; although P2X7Rs usually trigger inflammatory and toxic responses, preconditioning triggers P2X7Rs in astrocytes to function as a switch to protect the brain against ischemia [47,92].
However, because inhibitors of astrocytes activation abolish ischemic tolerance, it is interesting to consider the possibility that the neuron-dependent ischemic tolerance acquisition mechanism is not neuron-autonomous but rather neuron-non-autonomous as a result of communication with the glial cells. It should be noted that apart from microglia and astocytes, the vascular system and oligodendrocytes are involved in the activation of ischemic tolerance [93].

7. The Mechanism of Remote Ischemic Preconditioning

As of now, there has been no clearly confirmed method of propagating the initiation of ischemic tolerance by the body. Three theories of the spread of the triggers of ischemic tolerance are known: humoral, nervous, and immune. Following the transport of tolerance triggers to the target organ, the initiation of a common signaling pathway is predicted.

7.1. Humoral Pathway

The assumption of the existence of a humoral pathway is confirmed by two further assumptions. First, after preconditioning, a period of reperfusion is required, meaning that the tolerance triggers must be blood-borne before reaching the target organ. Secondly, in models of the cross-administration of activated blood or plasma, not only intergeneric but also interspecies transfer of tolerance triggers were proven. Shimizu described the possibility of transferring activated plasma (limb ischemia) to other individuals, and only by administering this plasma did the activation of ischemic tolerance and protection of the myocardium from subsequent ischemic damage occur. Based on this study, it can be assumed that the triggers of ischemic tolerance were <15 kDa molecules. The activated plasma effect can be blocked by naloxone’s blockers; so, it should be expected that opioid receptors will be involved in the activation of tolerance, and the functioning of the transfer of the tolerance triggers will be independent of the nervous system [94]. Based on other works, it can be assumed that interleukins, stromal factor+-α, TNF-α, bradykinin-2, adenosine, opioids, NO, miRNA, and possibly catecholamines can be possible triggers of tolerance [95]. The assumption of humoral transfer clearly supports the transfer of ischemic tolerance with blood plasma from a 2× conditioned donor [96,97].

7.2. Nerve Pathway

Administration of the ganglia blocker (hexamethonium) leads to the cessation of the activation of ischemic tolerance after the application of limb RIPreC; therefore, it can be concluded that the nerve pathway is also involved in the transmission of tolerance triggers [98].
The activation of dorsal neurons in the brainstem (vagal preganglionic neurons) produces a similar effect to the application of RIPreC. Moreover, by activating parasympathetic efferent nerves, neuroprotection will be activated, and the size of the infarcted focus will decrease [99,100].
It has been described in animal models that after transection of the femoral nerve, possibly the spinal cord, the RIPreC effect was abolished [101]. The neuroprotective effect of RIPreC on mice was only partially suppressed, when the femoral or sciatic nerve was damaged; so, it is necessary to think about the simultaneous connection of the humoral and nerve activation during the transfer of tolerance triggers [56,101].
Transection of the femoral nerve in an animal model stopped the activation of tolerance after limb RIPreC, but electrical stimulation of the damaged femoral nerve restored this activation of tolerance [102].
The release of endogenous humoral factors from the conditioned organ probably first activates the afferent nerve fibers in the conditioned organ and subsequently activates the efferent nerve fibers leading to the target organ.
Moreover, in patients with peripheral neuropathy, the RIPreC effect was weakened and even suppressed, which points to the involvement of the nerve pathway [99,103].

7.3. Inflammatory Pathway

Limb RIPreC inhibits inflammatory changes. It suppresses proinflammatory genes and stimulates anti-inflammatory genes. The influence of the genes is manifested by the reduction in the number of circulating neutrophils and the inhibition of the release of cytokines. This effect is already noticeable 24 h after the action of RIPreC [104,105].
The activation of immune cells and the regulation of inflammatory genes are associated with the release of endogenous opioids; therefore, the humoral and immunological pathways of activation will also be interconnected [106].

7.4. Common Final Mechanisms

Based on the previous results, it can be concluded that in the transmission of ischemic tolerance triggers, there is a mutual combination of their transmission via the humoral, nervous, and immunological pathways. After their transfer to the target organ, the common path of the activation of tolerance itself is activated. Upon activation of the tolerance in the target organ after the arrival of triggers, autophagy is activated, which promotes the degradation of the damaged cell organelles and accumulated proteins. Limb IRPreC stimulates autophagy and supports the emergence of neuroprotection [107]. Mitochondria also participate in this common mechanism of tolerance activation; the action of RIPreC supports the preservation of the structural and functional integrity of mitochondria, the reduction in mitochondrial degradation and, consequently, the reduction in apoptosis [108].

8. Cell Signaling Pathways in Ischemic Postconditioning

8.1. Akt Signaling Pathway Activation

The PI3K/Akt is a cellular pathway important in cell cycle regulation that regulates cell survival by inhibiting apoptosis and cell growth [109].
IP leads to an improvement and prolongation of Akt phosphorylation, which ultimately leads to a reduction in the extent of the necrosis at the ischemic site [110].
In the case of RIP, this pathway also applies, leading to a reduction in mitochondrial damage [110].

8.2. The mTOR Signaling Pathway of Activation (Mammalian Target of Rapamycin = Thr/Ser Protein Kinase)

The mTOR pathway plays a major role in cell metabolism, growth, differentiation, development, and cell survival [111]. However, its role in IR damage is controversial. mTOR is inhibited by hypoxia and the lack of ATP; in contrast, antioxidants, melatonin, and estradiol stimulate and affect this pathway [112].
In IP, blocking the mTOR pathway using its inhibitor rapamycin cancels the protective effect of IP [113,114].

8.3. The Mitogen-Activated Protein Kinase (MAPK) Pathway

MAPK is a major kinase pathway that transduces extracellular signals into an appropriate cellular response. As with the mTOR pathway, this pathway also mediates cell growth, development, division, and differentiation.
IP increases the expression of ERK1/2 kinase and activates the MAPK pathway; however, after the application of kinase blockers, the inhibition of the neuroprotective effect does not occur, as would be expected; therefore, the role of the MAPK pathway is rather unclear in cooperation with postconditioning [110].

8.4. The Protein Kinase C (PKC) Pathway

PKC is a serine/threonine protein kinase that has at least 12 isozymes. IP blocks cell death by promoting εPKC phosphorylation [115]. δPKC activity promotes cell death, but εPKC activity promotes neuronal survival [116].

8.5. The Toll-Like Receptor 4 (TLR4) Pathway

The TLR4 pathway is an important mediator of the immune response that leads to inflammation and is part of the activation of ischemic tolerance. Both IP and RIP lead to suppression of the TLR4/NF-κB pathway [117,118].
In the brain, the mechanism causing apoptosis of the CA1 neurons in the hippocampus also plays an important role. After ischemia, despite restoration of energy metabolism, there is no restoration of protein synthesis. That is caused by phosphorylation of the small α subunit of initiation factor 2 (eIF2α). The phosphorylation of eIF2α causes inhibition of the translation initiation with the subsequent disintegration of the polyribosomes into subunits [119].
If protein synthesis is not restored, the cells die by so called delayed death of neurons. Postconditioning allows the unblocking of proteosynthesis [88,120].

8.6. The Mitochondrial Role in Postconditioning

Mitochondria play a critical role in IR injury and the subsequent development of ventricular systolic dysfunction and possible compensatory heart hypertrophy [121].
The mitochondrial permeability transition pore (mPTP) has a pivotal influence on cells’ ability to survive or die. Postconditioning can block the opening of mPTP, which has detrimental effect on the activation of a robust anti-ischemic protection. Mitochondrial calcium is decreased during preconditioning; in contrast, it is increased significantly either in postconditioning or after the inhibition of mPTP [122,123]. Calcium changes in assembly explain why reperfusion injuries are diminished in postconditioned hearts. Postconditioning itself is attributed to the opening of the mitochondrial KATP channels (mKATP), and the inhibition of the mPTP opening leads to a reduced infarct size [124].
It is also expected that the activator of transcription 3 (STAT3) contributes to cardioprotection by the stimulation of respiration and the inhibition of the mPTP opening [125]. Moreover, mitochondrial ATP-sensitive K+ channels (mKATP) and connexin 43 have tasks in cardioprotection in postconditioning [126,127].
Mitochondria influence H2O2 production and redox stress during reperfusion through mKATP activation [128,129]. Moreover, phosphatidylinositol 3-kinase (PI3K) is involved in the regulation of mPTP [130].
Hydrogen sulfide (H2S), carbon monoxide (CO), and nitric oxide (NO) are recognized as three gaseous mediators for cardioprotection, which should also be produced within mitochondria [131,132]. Anesthetics (isoflurane, sevoflurane, and propofol) are also agents targeting mitochondria with prominent postconditioning effects [133,134,135].
Mitochondria are important players in many types of apoptotic and necrotic cell death [128]. Postconditioning signaling converges on the mitochondria, and it increases the levels of antiapoptotic markers, including the phospho-GSK-3β and Pim-1 kinases, while decreasing the proapoptotic markers, namely cytochrome c, thus preserving the mitochondrial morphology [136,137].

9. Pharmacological Postconditioning

The mechanism of ischemic postconditioning includes activation of the adenosine receptors, α-adrenergic receptors, opioid receptors, and bradykinin, with activation of pathways such as the PKC, TK, and MAP kinase, or the activation mechanism involves an endogenous trigger such as NO and Ca2+ ions. The administration of drugs whose effect mimics the mechanism of the action of ischemic postconditioning is called pharmacological postconditioning (FAP) [138,139].
FAP has been described in the heart, brain, and liver. Different mechanisms and receptors are involved in its functioning, while the protection obtained by FAP is based on inflammatory, antiapoptotic, and antioxidant processes. The advantage of FAP over IP is the possibility of its application even at a time when IP cannot be used, for example due to the impassability of a local artery in the basin of which ischemia has occurred.
The mechanism of FAP is direct interference with the course of the path of the lethal damage or the induction of a low-threshold stress, which can trigger the activation of tolerance against a stronger, i.e., lethal stress [140].
For example, the application of isoflurane after 60 min of global brain ischemia reduced the extent of brain damage in both in vivo and vitro models [141].
Moreover, the application of morphine, which leads to the stimulation of opioid receptors, can activate neuroprotection after the end of lethal ischemia, similar to ischemic postconditioning [142].

10. Ischemic Tolerance (Closing Remarks)—Application Possibilities

Even though the phenomenon of ischemic tolerance has been known for more than 30 years, we still only have partial information about its activation and functioning. So far, there is a lack of a comprehensive understanding not only of the mechanisms of its functioning but also of how to adequately transform the experimental results into real clinical use.
Individual tissues have significantly different resistance to IR damage. This resistance is partially dependent on the energy metabolism of the cells but is also not fully understood. In the brain, cells in the CA1 hippocampus are the most sensitive to ischemia, which begin to die after 5 min of global cerebral ischemia. Ten minutes of global ischemia of the rat brain led to the death of up to 70% of these cells [143]. Conversely, in the case of limbs, skeletal muscle is significantly more vulnerable than peripheral nervous tissue. In skeletal muscle, necrotic changes appeared after 2–3 h of ischemia, but in the peripheral nerve, signs of nerve cell degeneration did not appear until after 5 h of ischemia [144]. This different sensitivity of cells to IR damage must also be considered when investigating ischemic tolerance.
Until now, the clinical use of ischemic tolerance was only preconditioning. In this case, the application of sublethal stress activates temporary ischemic tolerance to stresses of different genesis, which allows survival under other circumstances of lethal stress.
This classic preconditioning is applicable only in the case of elective surgical interventions, during which ischemia of the organism or its part or organ is expected in advance, for example during the transfer of free lobes, elective replantations, revascularization, or operations such as aortocoronary bypass.
The preconditioning itself must be applied before the ischemia of the organ occurs, i.e., at the latest at the beginning of the revascularization operation, for example, before the ischemia occurring during the suture of the distal vascular anastomosis during the suture of the graft to the coronary vessel [145].
The disadvantage of local preconditioning is the one-time or repeated clamping of the vascular pedicle of the lobe, which can lead to intimal damage to the supply artery of the lobe. This problem also applies to local postconditioning.
The limited applicability of preconditioning lies in the fact that it is not always possible to predict the onset of ischemic damage. This fact significantly limits the applicability of preconditioning in clinical practice. Unfortunately, the exception is a war conflict, where it is possible to assume the occurrence of damage, the consequences of which could be significantly reduced using preconditioning in young, healthy, and unmedicated people.
From a clinical point of view, postconditioning has a much greater application due to the frequent unexpected occurrence of ischemia (for example, amputation of fingers, thrombosis of vascular pedicles of free lobes, damage to soft tissues in fractures of the proximal and distal tibia) and many other life, organ, or limb threatening conditions. The applicability of this methodology after exposure to lethal stress is an indisputable advantage.
The application of postconditioning is more useful in clinical practice, as it allows the living organism to activate the mechanism of ischemic tolerance, which can reverse the process of cell degeneration even within a few hours after the occurrence of lethal stress. It is necessary to emphasize that despite the fact that five or more minutes of brain ischemia in humans represents the beginning of irreversible changes, neurons undergo apoptosis similar to the so-called delayed death of neurons, and the therapeutic window of their rescue is significantly longer than in other types of cell death.
The use of short-term skeletal muscle ischemia (delayed postconditioning) can reverse the death of muscle cells after 3 h of ischemia and also prevent muscle edema [146]. These results correspond with those of Tsubota [147], who applied remote postconditioning to the contralateral limb in mice. In the case of remote postconditioning, the setting of an adequately large ischemia (its length and intensity) necessary to activate ischemic tolerance is still an open question. The application of ischemia to the upper limb has not been proven in humans, there is a preference for the application of ischemia to the lower limb. The reason is probably the larger amount of skeletal muscle in the lower limb, the ischemia of which leads to a significantly stronger stimulus for the activation of ischemic tolerance. Moreover, the determination of the length of the ischemia time interval has been extensively researched; we ourselves are inclined to 20 min of muscle ischemia, the most effective according to the results of Pignataro [148]. This period of ischemia is not dangerous for animals or humans, during which there is no damage to the vitality of the muscles.
If short-term ischemia of a relatively large skeletal muscle is used as postconditioning, it induces two synergistic beneficial effects [97]:
-
Induction of ischemic tolerance;
-
Temporary elevation of blood pressure and blood flow.
The effect of a temporary increase in perfusion caused by remote ischemic postconditioning has a positive effect in clinical practice, for example during the transfer of free, rotary flaps, or during finger replantation. Even today, free flaps are a reliable method of covering defects in trauma surgery, but their rejection (complete necrosis) occurs in up to 25% of cases, and partial necrosis is present in up to 36% of cases. Rejection is a consequence of the insufficient perfusion of the lobe. In addition, postconditioning-induced postischemic reactive hyperemia can act as a prevention of edema, the “no reflow” phenomenon, and an overall improvement in local microcirculation [149].
Although the use of remote ischemic local postconditioning is an effective method in properly selected cases, its universal clinical application is rather complex (tourniquet loading).
Despite the very encouraging and successful results in animal experiments, the clinical results in humans have been clearly disappointing so far. Most of the human studies concerned the use of the phenomenon of ischemic tolerance to reduce the extent of heart attack and improve the biochemical parameters [150,151,152]. Ultimately, RIPreC did not reduce the number of serious cardiac and cerebrovascular complications for cardiac surgery. Adequate protection of the myocardium has also not been demonstrated.
Due to the abovementioned facts, it will be clearly necessary to transfer the results from young healthy animals to the gerontic population, which suffers from numerous comorbidities (diabetes mellitus, hypertension, and hyperlipidemia) and co-medications (antioxidants, statins, β-blockers, ACE inhibitors, AT1 receptor antagonists, antagonists Ca channels, and nitrates) that block the emergence of ischemic tolerance [150,153,154]. Comorbidities clearly lead to the suppression of ischemic tolerance [155].
Since the effectiveness of remote ischemic postconditioning was also proven in the case of delayed neuronal death (“apoptosis”) induced by temporary global ischemia of the brain or kainate intoxication, it can be assumed that conditioning products are able to overcome the blood–brain barrier [156], i.e., they are probably evenly spread throughout the body.
Based on Shimizu’s results, it can be assumed that triggers of ischemic tolerance are contained in plasma, are low molecular hydrophobic substances, are independent of local neurogenic activity, and require the activation of local opioid receptors [94]. The idea of transferring these substances between individuals of the same species but also between species is revolutionary. However, even with the transfer of triggers from the donor, the activation of tolerance may be blocked (possible comorbidity and comedication).
The effect of preconditioned plasma was also demonstrated by Zhao [157]. Preconditioned plasma (taken from donor animals 48 h after limb ischemia) was able to reduce the extent of IR damage to the myocardium after transfer to recipients, which was clinically manifested by a reduction in the incidence and duration of ventricular tachycardia.
A similar result was also achieved by Weber [158]. By applying ischemic preconditioned plasma in volunteers, he achieved protection of human umbilical endothelial cells from damage indicated by hypoxia. In this case, the plasma was collected directly after the preconditioning of the volunteers.
In all cases mentioned, it was only a transfer of “partially” activated plasma. If the plasma was activated by only one stress, only triggers of ischemic tolerance were present in it. In our experiments, twice-activated plasma was used (exposure to 2 stresses in a period of 48 h, with plasma withdrawal 6 h after the second stress). Withdrawal of doubly activated plasma involves the transfer of effectors, not triggers, of ischemic tolerance. The use of a proteosynthesis inhibitor (cycloheximide), which blocked the emergence of tolerance, serves as proof of the necessity of synthesizing protein effectors [35].
Our results clearly confirmed the efficacy of doubly activated plasma when administered to recipient animals after lethal ischemia. The advantage of doubly activated plasma containing complete tolerance also lies in the fact that co-medication and co-morbidity in the recipient should not have an adverse effect on its immediate effectiveness, which will allow its easy application and immediate effectiveness.
Until there is a clear chemical identification of the end products (end effectors) of ischemic tolerance, the simplest way to activate ischemic tolerance will be the preparation of activated plasma from young healthy donors with the possibility of its immediate use in recipients during the initial treatment.
The next expected stage is the exact identification of the effector (product) of ischemic tolerance and its direct administration to the patient. Based on several works [36,97], it can be concluded that the end effector of ischemic tolerance passes through the blood–brain barrier, meaning that its application will not be limited to only one organ, and its action will be possible immediately in the whole organism. These assumptions were also confirmed by our results [96,97], where we clearly demonstrated the unequivocal effect of activated plasma on the survival of skeletal muscle cells as well as brain cells in the CA1 hippocampus.

Author Contributions

Conceptualization: R.B. and R.M.; writing—original draft preparation: R.B. and J.B.; writing—review and editing: R.B. and J.B.; supervision: J.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Krause, G.S.; Kumar, K.; White, B.C.; Aust, S.D.; Wiegenstein, J.G. Ischemia, resuscitation, and reperfusion: Mechanisms of tissue injury and prospects for protection. Am. Heart J. 1986, 111, 768–780. [Google Scholar] [CrossRef]
  2. Dorweiler, B.; Pruefer, D.; Andrasi, T.B.; Maksan, S.M.; Schmiedt, W.; Neufang, A.; Vahl, C.F. Ischemia-Reperfusion Injury: Pathophysiology and Clinical Implications. Eur. J. Trauma Emerg. Surg. 2007, 33, 600–612. [Google Scholar] [CrossRef] [PubMed]
  3. Salvadori, M.; Rosso, G.; Bertoni, E. Update on ischemia-reperfusion injury in kidney transplantation: Pathogenesis and treatment. World J. Transplant. 2015, 5, 52–67. [Google Scholar] [CrossRef] [PubMed]
  4. Saat, T.C.; van den Akker, E.K.; IJzermans, J.N.; Dor, F.J.; de Bruin, R.W. Improving the outcome of kidney transplantation by ameliorating renal ischemia reperfusion injury: Lost in translation? J. Transl. Med. 2016, 14, 20. [Google Scholar] [CrossRef] [Green Version]
  5. Nieuwenhuijs-Moeke, G.J.; Pischke, S.E.; Berger, S.P.; Sanders, J.S.F.; Pol, R.A.; Struys, M.; Ploeg, R.J.; Leuvenink, H.G.D. Ischemia and Reperfusion Injury in Kidney Transplantation: Relevant Mechanisms in Injury and Repair. J. Clin. Med. 2020, 9, 253. [Google Scholar] [CrossRef] [Green Version]
  6. Mico-Carnero, M.; Zaouali, M.A.; Rojano-Alfonso, C.; Maroto-Serrat, C.; Ben Abdennebi, H.; Peralta, C. A Potential Route to Reduce Ischemia/Reperfusion Injury in Organ Preservation. Cells 2022, 11, 2763. [Google Scholar] [CrossRef]
  7. Becker, L.B. New concepts in reactive oxygen species and cardiovascular reperfusion physiology. Cardiovasc. Res. 2004, 61, 461–470. [Google Scholar] [CrossRef] [Green Version]
  8. Ahluwalia, A.; Foster, P.; Scotland, R.S.; McLean, P.G.; Mathur, A.; Perretti, M.; Moncada, S.; Hobbs, A.J. Antiinflammatory activity of soluble guanylate cyclase: cGMP-dependent down-regulation of P-selectin expression and leukocyte recruitment. Proc. Natl. Acad. Sci. USA 2004, 101, 1386–1391. [Google Scholar] [CrossRef] [Green Version]
  9. De Marco, C.; Charron, T.; Rousseau, G. Adenosine in Acute Myocardial Infarction-Associated Reperfusion Injury: Does it Still Have a Role? Front. Pharmacol. 2022, 13, 856747. [Google Scholar] [CrossRef] [PubMed]
  10. Mata, A.; Cadenas, S. The Antioxidant Transcription Factor Nrf2 in Cardiac Ischemia-Reperfusion Injury. Int. J. Mol. Sci. 2021, 22, 11939. [Google Scholar] [CrossRef]
  11. Toldo, S.; Breckenridge, D.G.; Mezzaroma, E.; Van Tassell, B.W.; Shryock, J.; Kannan, H.; Phan, D.; Budas, G.; Farkas, D.; Lesnefsky, E.; et al. Inhibition of apoptosis signal-regulating kinase 1 reduces myocardial ischemia-reperfusion injury in the mouse. J. Am. Heart Assoc. 2012, 1, e002360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Wang, R.; Wang, M.; He, S.; Sun, G.; Sun, X. Targeting Calcium Homeostasis in Myocardial Ischemia/Reperfusion Injury: An Overview of Regulatory Mechanisms and Therapeutic Reagents. Front. Pharmacol. 2020, 11, 872. [Google Scholar] [CrossRef] [PubMed]
  13. Rodrigo, R.; Retamal, C.; Schupper, D.; Vergara-Hernandez, D.; Saha, S.; Profumo, E.; Buttari, B.; Saso, L. Antioxidant Cardioprotection against Reperfusion Injury: Potential Therapeutic Roles of Resveratrol and Quercetin. Molecules 2022, 27, 2564. [Google Scholar] [CrossRef]
  14. Yoshitomi, T.; Nagasaki, Y. Self-Assembling Antioxidants for Ischemia-Reperfusion Injuries. Antioxid. Redox Signal. 2022, 36, 70–80. [Google Scholar] [CrossRef] [PubMed]
  15. Miura, T.; Liu, Y.; Goto, M.; Tsuchida, A.; Miki, T.; Nakano, A.; Nishino, Y.; Ohnuma, Y.; Shimamoto, K. Mitochondrial ATP-sensitive K+ channels play a role in cardioprotection by Na+-H+ exchange inhibition against ischemia/reperfusion injury. J. Am. Coll. Cardiol. 2001, 37, 957–963. [Google Scholar] [CrossRef] [Green Version]
  16. Fischesser, D.M.; Bo, B.; Benton, R.P.; Su, H.; Jahanpanah, N.; Haworth, K.J. Controlling Reperfusion Injury with Controlled Reperfusion: Historical Perspectives and New Paradigms. J. Cardiovasc. Pharmacol. Ther. 2021, 26, 504–523. [Google Scholar] [CrossRef]
  17. Sadeghi, M.; Horer, T.M.; Forsman, D.; Dogan, E.M.; Jansson, K.; Kindler, C.; Skoog, P.; Nilsson, K.F. Blood pressure targeting by partial REBOA is possible in severe hemorrhagic shock in pigs and produces less circulatory, metabolic and inflammatory sequelae than total REBOA. Injury 2018, 49, 2132–2141. [Google Scholar] [CrossRef]
  18. Wu, X.; You, D.; Cui, J.; Yang, L.; Lin, L.; Chen, Y.; Xu, C.; Lian, G.; Wan, J. Reduced Neutrophil Extracellular Trap Formation During Ischemia Reperfusion Injury in C3 KO Mice: C3 Requirement for NETs Release. Front. Immunol. 2022, 13, 781273. [Google Scholar] [CrossRef]
  19. Oliveira, T.H.C.; Marques, P.E.; Proost, P.; Teixeira, M.M.M. Neutrophils: A cornerstone of liver ischemia and reperfusion injury. Lab. Investig. 2018, 98, 51–62. [Google Scholar] [CrossRef] [Green Version]
  20. Khan, T.A.; Bianchi, C.; Voisine, P.; Feng, J.; Baker, J.; Hart, M.; Takahashi, M.; Stahl, G.; Sellke, F.W. Reduction of myocardial reperfusion injury by aprotinin after regional ischemia and cardioplegic arrest. J. Thorac. Cardiovasc. Surg. 2004, 128, 602–608. [Google Scholar] [CrossRef] [Green Version]
  21. Liu, S.; Luo, W.; Wang, Y. Emerging role of PARP-1 and PARthanatos in ischemic stroke. J. Neurochem. 2022, 160, 74–87. [Google Scholar] [CrossRef]
  22. Vogel, C.W. The Role of Complement in Myocardial Infarction Reperfusion Injury: An Underappreciated Therapeutic Target. Front. Cell Dev. Biol. 2020, 8, 606407. [Google Scholar] [CrossRef] [PubMed]
  23. Ergun, Y.; Darendeli, S.; Imrek, S.; Kilinc, M.; Oksuz, H. The comparison of the effects of anesthetic doses of ketamine, propofol, and etomidate on ischemia-reperfusion injury in skeletal muscle. Fundam. Clin. Pharmacol. 2010, 24, 215–222. [Google Scholar] [CrossRef]
  24. Lee, J.M.; Suh, J.K.; Jeong, J.S.; Cho, S.Y.; Kim, D.W. Antioxidant effect of lidocaine and procaine on reactive oxygen species-induced endothelial dysfunction in the rabbit abdominal aorta. Korean J. Anesthesiol. 2010, 59, 104–110. [Google Scholar] [CrossRef] [PubMed]
  25. Hypothermia after Cardiac Arrest Study Group. Mild therapeutic hypothermia to improve the neurologic outcome after cardiac arrest. N. Engl. J. Med. 2002, 346, 549–556. [Google Scholar] [CrossRef]
  26. Lascarrou, J.B.; Merdji, H.; Le Gouge, A.; Colin, G.; Grillet, G.; Girardie, P.; Coupez, E.; Dequin, P.F.; Cariou, A.; Boulain, T.; et al. Targeted Temperature Management for Cardiac Arrest with Nonshockable Rhythm. N. Engl. J. Med. 2019, 381, 2327–2337. [Google Scholar] [CrossRef]
  27. Chen, W.; Lv, L.; Nong, Z.; Chen, X.; Pan, X.; Chen, C. Hyperbaric oxygen protects against myocardial ischemia-reperfusion injury through inhibiting mitochondria dysfunction and autophagy. Mol. Med. Rep. 2020, 22, 4254–4264. [Google Scholar] [CrossRef] [PubMed]
  28. Bonauer, A.; Carmona, G.; Iwasaki, M.; Mione, M.; Koyanagi, M.; Fischer, A.; Burchfield, J.; Fox, H.; Doebele, C.; Ohtani, K.; et al. MicroRNA-92a controls angiogenesis and functional recovery of ischemic tissues in mice. Science 2009, 324, 1710–1713. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, J.X.; Jiao, J.Q.; Li, Q.; Long, B.; Wang, K.; Liu, J.P.; Li, Y.R.; Li, P.F. miR-499 regulates mitochondrial dynamics by targeting calcineurin and dynamin-related protein-1. Nat. Med. 2011, 17, 71–78. [Google Scholar] [CrossRef]
  30. Rowart, P.; Erpicum, P.; Detry, O.; Weekers, L.; Gregoire, C.; Lechanteur, C.; Briquet, A.; Beguin, Y.; Krzesinski, J.M.; Jouret, F. Mesenchymal Stromal Cell Therapy in Ischemia/Reperfusion Injury. J. Immunol. Res. 2015, 2015, 602597. [Google Scholar] [CrossRef] [Green Version]
  31. Shi, Y.; Zhang, X.; Wan, Z.; Liu, X.; Chen, F.; Zhang, J.; Leng, Y. Mesenchymal stem cells against intestinal ischemia-reperfusion injury: A systematic review and meta-analysis of preclinical studies. Stem Cell Res. Ther. 2022, 13, 216. [Google Scholar] [CrossRef] [PubMed]
  32. Zemke, D.; Smith, J.L.; Reeves, M.J.; Majid, A. Ischemia and ischemic tolerance in the brain: An overview. Neurotoxicology 2004, 25, 895–904. [Google Scholar] [CrossRef]
  33. Gidday, J.M. Cerebral preconditioning and ischaemic tolerance. Nat. Rev. Neurosci. 2006, 7, 437–448. [Google Scholar] [CrossRef]
  34. Burda, J.; Matiasova, M.; Gottlieb, M.; Danielisova, V.; Nemethova, M.; Garcia, L.; Salinas, M.; Burda, R. Evidence for a role of second pathophysiological stress in prevention of delayed neuronal death in the hippocampal CA1 region. Neurochem. Res. 2005, 30, 1397–1405. [Google Scholar] [CrossRef]
  35. Burda, J.; Danielisova, V.; Nemethova, M.; Gottlieb, M.; Matiasova, M.; Domorakova, I.; Mechirova, E.; Ferikova, M.; Salinas, M.; Burda, R. Delayed postconditionig initiates additive mechanism necessary for survival of selectively vulnerable neurons after transient ischemia in rat brain. Cell. Mol. Neurobiol. 2006, 26, 1141–1151. [Google Scholar] [CrossRef] [PubMed]
  36. Cheng, O.; Ostrowski, R.P.; Wu, B.; Liu, W.; Chen, C.; Zhang, J.H. Cyclooxygenase-2 mediates hyperbaric oxygen preconditioning in the rat model of transient global cerebral ischemia. Stroke 2011, 42, 484–490. [Google Scholar] [CrossRef] [Green Version]
  37. Bigdeli, M.R. Neuroprotection caused by hyperoxia preconditioning in animal stroke models. Sci. World J. 2011, 11, 403–421. [Google Scholar] [CrossRef] [PubMed]
  38. Yunoki, M.; Nishio, S.; Ukita, N.; Anzivino, M.J.; Lee, K.S. Hypothermic preconditioning induces rapid tolerance to focal ischemic injury in the rat. Exp. Neurol. 2003, 181, 291–300. [Google Scholar] [CrossRef] [PubMed]
  39. Ikeda, T.; Xia, X.Y.; Xia, Y.X.; Ikenoue, T. Hyperthermic preconditioning prevents blood-brain barrier disruption produced by hypoxia-ischemia in newborn rat. Dev. Brain Res. 1999, 117, 53–58. [Google Scholar] [CrossRef]
  40. Schlegel, A.; Mueller, M.; Muller, X.; Eden, J.; Panconesi, R.; von Felten, S.; Steigmiller, K.; Da Silva, R.X.S.; de Rougemont, O.; Mabrut, J.Y.; et al. A multicenter randomized-controlled trial of hypothermic oxygenated perfusion (HOPE) for human liver grafts before transplantation. J. Hepatol. 2023. [Google Scholar] [CrossRef]
  41. Rampino, T.; Gregorini, M.; Germinario, G.; Pattonieri, E.F.; Erasmi, F.; Grignano, M.A.; Bruno, S.; Alomari, E.; Bettati, S.; Asti, A.; et al. Extracellular Vesicles Derived from Mesenchymal Stromal Cells Delivered during Hypothermic Oxygenated Machine Perfusion Repair Ischemic/Reperfusion Damage of Kidneys from Extended Criteria Donors. Biology 2022, 11, 350. [Google Scholar] [CrossRef]
  42. Wang, L.; Traystman, R.J.; Murphy, S.J. Inhalational anesthetics as preconditioning agents in ischemic brain. Curr. Opin. Pharmacol. 2008, 8, 104–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kitano, H.; Kirsch, J.R.; Hurn, P.D.; Murphy, S.J. Inhalational anesthetics as neuroprotectants or chemical preconditioning agents in ischemic brain. J. Cereb. Blood Flow Metab. 2007, 27, 1108–1128. [Google Scholar] [CrossRef] [Green Version]
  44. Li, Y.; Liang, Z.; Lei, S.; Wu, X.; Yuan, T.; Ma, K.; Chi, K. Sevoflurane Preconditioning Downregulates GRIA1 Expression to Attenuate Cerebral Ischemia-Reperfusion-Induced Neuronal Injury. Neurotox. Res. 2023, 41, 29–40. [Google Scholar] [CrossRef]
  45. Vartanian, K.B.; Stevens, S.L.; Marsh, B.J.; Williams-Karnesky, R.; Lessov, N.S.; Stenzel-Poore, M.P. LPS preconditioning redirects TLR signaling following stroke: TRIF-IRF3 plays a seminal role in mediating tolerance to ischemic injury. J. Neuroinflamm. 2011, 8, 140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Heurteaux, C.; Lauritzen, I.; Widmann, C.; Lazdunski, M. Essential role of adenosine, adenosine A1 receptors, and ATP-sensitive K+ channels in cerebral ischemic preconditioning. Proc. Natl. Acad. Sci. USA 1995, 92, 4666–4670. [Google Scholar] [CrossRef] [Green Version]
  47. Zhu, H.; Sun, S.; Li, H.; Xu, Y. Cerebral ischemic tolerance induced by 3-nitropropionic acid is associated with increased expression of erythropoietin in rats. J. Huazhong Univ. Sci. Technol. Med. Sci. 2006, 26, 440–443. [Google Scholar] [CrossRef] [PubMed]
  48. Danielisova, V.; Gottlieb, M.; Nemethova, M.; Kravcukova, P.; Domorakova, I.; Mechirova, E.; Burda, J. Bradykinin postconditioning protects pyramidal CA1 neurons against delayed neuronal death in rat hippocampus. Cell. Mol. Neurobiol. 2009, 29, 871–878. [Google Scholar] [CrossRef]
  49. Jackson, C.W.; Escobar, I.; Xu, J.; Perez-Pinzon, M.A. Effects of ischemic preconditioning on mitochondrial and metabolic neruoprotection: 5’ adenosine monophosphate-activated protein kinase and sirtuins. Brain Circ. 2018, 4, 54–61. [Google Scholar] [CrossRef] [PubMed]
  50. Kim, J.H.; Choi, K.H.; Jang, Y.J.; Kim, H.N.; Bae, S.S.; Choi, B.T.; Shin, H.K. Electroacupuncture preconditioning reduces cerebral ischemic injury via BDNF and SDF-1alpha in mice. BMC Complement. Altern. Med. 2013, 13, 22. [Google Scholar] [CrossRef] [Green Version]
  51. Liebelt, B.; Papapetrou, P.; Ali, A.; Guo, M.; Ji, X.; Peng, C.; Rogers, R.; Curry, A.; Jimenez, D.; Ding, Y. Exercise preconditioning reduces neuronal apoptosis in stroke by up-regulating heat shock protein-70 (heat shock protein-72) and extracellular-signal-regulated-kinase 1/2. Neuroscience 2010, 166, 1091–1100. [Google Scholar] [CrossRef] [PubMed]
  52. Lee, H.I.; Park, J.H.; Park, M.Y.; Kim, N.G.; Park, K.J.; Choi, B.T.; Shin, Y.I.; Shin, H.K. Pre-conditioning with transcranial low-level light therapy reduces neuroinflammation and protects blood-brain barrier after focal cerebral ischemia in mice. Restor. Neurol. Neurosci. 2016, 34, 201–214. [Google Scholar] [CrossRef]
  53. Thompson, E.R.; Bates, L.; Ibrahim, I.K.; Sewpaul, A.; Stenberg, B.; McNeill, A.; Figueiredo, R.; Girdlestone, T.; Wilkins, G.C.; Wang, L.; et al. Novel delivery of cellular therapy to reduce ischemia reperfusion injury in kidney transplantation. Am. J. Transplant. 2021, 21, 1402–1414. [Google Scholar] [CrossRef]
  54. Grignano, M.A.; Bruno, S.; Viglio, S.; Avanzini, M.A.; Tapparo, M.; Ramus, M.; Croce, S.; Valsecchi, C.; Pattonieri, E.F.; Ceccarelli, G.; et al. CD73-Adenosinergic Axis Mediates the Protective Effect of Extracellular Vesicles Derived from Mesenchymal Stromal Cells on Ischemic Renal Damage in a Rat Model of Donation after Circulatory Death. Int. J. Mol. Sci. 2022, 23, 10681. [Google Scholar] [CrossRef] [PubMed]
  55. Kim, Y.H.; Her, A.Y.; Rha, S.W.; Choi, B.G.; Choi, S.Y.; Byun, J.K.; Mashaly, A.; Park, Y.; Jang, W.Y.; Kim, W.; et al. Impact of Trimetazidine Treatment on 5-year Clinical Outcomes in Patients with Significant Coronary Artery Spasm: A Propensity Score Matching Study. Am. J. Cardiovasc. Drugs 2018, 18, 117–127. [Google Scholar] [CrossRef] [PubMed]
  56. St Peter, S.D.; Post, D.J.; Rodriguez-Davalos, M.I.; Douglas, D.D.; Moss, A.A.; Mulligan, D.C. Tacrolimus as a liver flush solution to ameliorate the effects of ischemia/reperfusion injury following liver transplantation. Liver Transpl. 2003, 9, 144–149. [Google Scholar] [CrossRef] [Green Version]
  57. Zamorano, M.; Castillo, R.L.; Beltran, J.F.; Herrera, L.; Farias, J.A.; Antileo, C.; Aguilar-Gallardo, C.; Pessoa, A.; Calle, Y.; Farias, J.G. Tackling Ischemic Reperfusion Injury With the Aid of Stem Cells and Tissue Engineering. Front. Physiol. 2021, 12, 705256. [Google Scholar] [CrossRef] [PubMed]
  58. Kerendi, F.; Kin, H.; Halkos, M.E.; Jiang, R.; Zatta, A.J.; Zhao, Z.Q.; Guyton, R.A.; Vinten-Johansen, J. Remote postconditioning. Brief renal ischemia and reperfusion applied before coronary artery reperfusion reduces myocardial infarct size via endogenous activation of adenosine receptors. Basic Res. Cardiol. 2005, 100, 404–412. [Google Scholar] [CrossRef]
  59. Wang, C.; Weihrauch, D.; Schwabe, D.A.; Bienengraeber, M.; Warltier, D.C.; Kersten, J.R.; Pratt, P.F., Jr.; Pagel, P.S. Extracellular signal-regulated kinases trigger isoflurane preconditioning concomitant with upregulation of hypoxia-inducible factor-1alpha and vascular endothelial growth factor expression in rats. Anesth. Analg. 2006, 103, 281–288. [Google Scholar] [CrossRef]
  60. Hao, Y.; Xin, M.; Feng, L.; Wang, X.; Wang, X.; Ma, D.; Feng, J. Review Cerebral Ischemic Tolerance and Preconditioning: Methods, Mechanisms, Clinical Applications, and Challenges. Front. Neurol. 2020, 11, 812. [Google Scholar] [CrossRef]
  61. Fox, C.; Walsh, P.; Mulhall, K.J. Molecular Mechanism of Ischaemic Preconditioning of Skeletal Muscle In Vitro. Cureus 2018, 10, e3763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Lee, J.C.; Tae, H.J.; Kim, I.H.; Cho, J.H.; Lee, T.K.; Park, J.H.; Ahn, J.H.; Choi, S.Y.; Bai, H.C.; Shin, B.N.; et al. Roles of HIF-1alpha, VEGF, and NF-kappaB in Ischemic Preconditioning-Mediated Neuroprotection of Hippocampal CA1 Pyramidal Neurons Against a Subsequent Transient Cerebral Ischemia. Mol. Neurobiol. 2017, 54, 6984–6998. [Google Scholar] [CrossRef]
  63. Mark, L.P.; Prost, R.W.; Ulmer, J.L.; Smith, M.M.; Daniels, D.L.; Strottmann, J.M.; Brown, W.D.; Hacein-Bey, L. Pictorial review of glutamate excitotoxicity: Fundamental concepts for neuroimaging. AJNR Am. J. Neuroradiol. 2001, 22, 1813–1824. [Google Scholar] [PubMed]
  64. Watters, O.; O’Connor, J.J. A role for tumor necrosis factor-alpha in ischemia and ischemic preconditioning. J. Neuroinflammation 2011, 8, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ma, D.; Feng, L.; Cheng, Y.; Xin, M.; You, J.; Yin, X.; Hao, Y.; Cui, L.; Feng, J. Astrocytic gap junction inhibition by carbenoxolone enhances the protective effects of ischemic preconditioning following cerebral ischemia. J. Neuroinflammation 2018, 15, 198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Centeno, J.M.; Orti, M.; Salom, J.B.; Sick, T.J.; Perez-Pinzon, M.A. Nitric oxide is involved in anoxic preconditioning neuroprotection in rat hippocampal slices. Brain Res. 1999, 836, 62–69. [Google Scholar] [CrossRef]
  67. Lee, N.T.; Ong, L.K.; Gyawali, P.; Nassir, C.; Mustapha, M.; Nandurkar, H.H.; Sashindranath, M. Role of Purinergic Signalling in Endothelial Dysfunction and Thrombo-Inflammation in Ischaemic Stroke and Cerebral Small Vessel Disease. Biomolecules 2021, 11, 994. [Google Scholar] [CrossRef]
  68. Ulker, P.; Ozkan, O.; Amoroso, M.; Aslan, M.; Bassorgun, I.; Ubur, M.C.; Unal, K.; Ozcan, F.; Ozkan, O. Does ischemic preconditioning increase flap survival by ADORA2B receptor activation? Clin. Hemorheol. Microcirc. 2020, 75, 151–162. [Google Scholar] [CrossRef]
  69. Hart, M.L.; Gorzolla, I.C.; Schittenhelm, J.; Robson, S.C.; Eltzschig, H.K. SP1-dependent induction of CD39 facilitates hepatic ischemic preconditioning. J. Immunol. 2010, 184, 4017–4024. [Google Scholar] [CrossRef] [Green Version]
  70. Hart, M.L.; Much, C.; Gorzolla, I.C.; Schittenhelm, J.; Kloor, D.; Stahl, G.L.; Eltzschig, H.K. Extracellular adenosine production by ecto-5′-nucleotidase protects during murine hepatic ischemic preconditioning. Gastroenterology 2008, 135, 1739–1750.e3. [Google Scholar] [CrossRef]
  71. Eckle, T.; Krahn, T.; Grenz, A.; Kohler, D.; Mittelbronn, M.; Ledent, C.; Jacobson, M.A.; Osswald, H.; Thompson, L.F.; Unertl, K.; et al. Cardioprotection by ecto-5′-nucleotidase (CD73) and A2B adenosine receptors. Circulation 2007, 115, 1581–1590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Smith, S.B.; Xu, Z.; Novitskaya, T.; Zhang, B.; Chepurko, E.; Pu, X.A.; Wheeler, D.G.; Ziolo, M.; Gumina, R.J. Impact of cardiac-specific expression of CD39 on myocardial infarct size in mice. Life Sci. 2017, 179, 54–59. [Google Scholar] [CrossRef]
  73. Dwyer, K.M. Burnstock oration—purinergic signalling in kidney transplantation. Purinergic Signal. 2022, 18, 387–393. [Google Scholar] [CrossRef] [PubMed]
  74. Li, K.; Gong, X.; Kuang, G.; Jiang, R.; Wan, J.; Wang, B. Sesamin protects against renal ischemia reperfusion injury by promoting CD39-adenosine-A2AR signal pathway in mice. Am. J. Transl. Res. 2016, 8, 2245–2254. [Google Scholar] [PubMed]
  75. Wang, Q.; Tang, X.N.; Yenari, M.A. The inflammatory response in stroke. J. Neuroimmunol. 2007, 184, 53–68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Perez-Pinzon, M.A.; Vitro, T.M.; Dietrich, W.D.; Sick, T.J. The effect of rapid preconditioning on the microglial, astrocytic and neuronal consequences of global cerebral ischemia. Acta Neuropathol. 1999, 97, 495–501. [Google Scholar] [CrossRef] [PubMed]
  77. Colangelo, V.; Gordon, W.C.; Mukherjee, P.K.; Trivedi, P.; Ottino, P. Downregulation of COX-2 and JNK expression after induction of ischemic tolerance in the gerbil brain. Brain Res. 2004, 1016, 195–200. [Google Scholar] [CrossRef]
  78. Lee, J.C.; Tae, H.J.; Cho, G.S.; Kim, I.H.; Ahn, J.H.; Park, J.H.; Chen, B.H.; Cho, J.H.; Shin, B.N.; Cho, J.H.; et al. Ischemic preconditioning protects neurons from damage and maintains the immunoreactivity of kynurenic acid in the gerbil hippocampal CA1 region following transient cerebral ischemia. Int. J. Mol. Med. 2015, 35, 1537–1544. [Google Scholar] [CrossRef] [Green Version]
  79. Shen, H.Y.; Lusardi, T.A.; Williams-Karnesky, R.L.; Lan, J.Q.; Poulsen, D.J.; Boison, D. Adenosine kinase determines the degree of brain injury after ischemic stroke in mice. J. Cereb. Blood Flow Metab. 2011, 31, 1648–1659. [Google Scholar] [CrossRef] [Green Version]
  80. Hayashi, T.; Saito, A.; Okuno, S.; Ferrand-Drake, M.; Dodd, R.L.; Chan, P.H. Damage to the endoplasmic reticulum and activation of apoptotic machinery by oxidative stress in ischemic neurons. J. Cereb. Blood Flow Metab. 2005, 25, 41–53. [Google Scholar] [CrossRef] [Green Version]
  81. Mathew, R.; Khor, S.; Hackett, S.R.; Rabinowitz, J.D.; Perlman, D.H.; White, E. Functional role of autophagy-mediated proteome remodeling in cell survival signaling and innate immunity. Mol. Cell 2014, 55, 916–930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Xia, D.Y.; Li, W.; Qian, H.R.; Yao, S.; Liu, J.G.; Qi, X.K. Ischemia preconditioning is neuroprotective in a rat cerebral ischemic injury model through autophagy activation and apoptosis inhibition. Braz. J. Med. Biol. Res. 2013, 46, 580–588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Cho, B.B.; Toledo-Pereyra, L.H. Caspase-independent programmed cell death following ischemic stroke. J. Investig. Surg. 2008, 21, 141–147. [Google Scholar] [CrossRef]
  84. Tohyama, Y.; Sako, K.; Yonemasu, Y. Hypothermia attenuates hyperglycolysis in the periphery of ischemic core in rat brain. Exp. Brain Res. 1998, 122, 333–338. [Google Scholar] [CrossRef]
  85. Wacker, B.K.; Freie, A.B.; Perfater, J.L.; Gidday, J.M. Junctional protein regulation by sphingosine kinase 2 contributes to blood-brain barrier protection in hypoxic preconditioning-induced cerebral ischemic tolerance. J. Cereb. Blood Flow Metab. 2012, 32, 1014–1023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Shi, Y.; Zhang, L.; Pu, H.; Mao, L.; Hu, X.; Jiang, X.; Xu, N.; Stetler, R.A.; Zhang, F.; Liu, X.; et al. Rapid endothelial cytoskeletal reorganization enables early blood-brain barrier disruption and long-term ischaemic reperfusion brain injury. Nat. Commun. 2016, 7, 10523. [Google Scholar] [CrossRef]
  87. Jones, N.M.; Bergeron, M. Hypoxic preconditioning induces changes in HIF-1 target genes in neonatal rat brain. J. Cereb. Blood Flow Metab. 2001, 21, 1105–1114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Burda, J.; Hrehorovska, M.; Bonilla, L.G.; Danielisova, V.; Cizkova, D.; Burda, R.; Nemethova, M.; Fando, J.L.; Salinas, M. Role of protein synthesis in the ischemic tolerance acquisition induced by transient forebrain ischemia in the rat. Neurochem. Res. 2003, 28, 1213–1219. [Google Scholar] [CrossRef]
  89. Dhodda, V.K.; Sailor, K.A.; Bowen, K.K.; Vemuganti, R. Putative endogenous mediators of preconditioning-induced ischemic tolerance in rat brain identified by genomic and proteomic analysis. J. Neurochem. 2004, 89, 73–89. [Google Scholar] [CrossRef]
  90. Atochin, D.N.; Clark, J.; Demchenko, I.T.; Moskowitz, M.A.; Huang, P.L. Rapid cerebral ischemic preconditioning in mice deficient in endothelial and neuronal nitric oxide synthases. Stroke 2003, 34, 1299–1303. [Google Scholar] [CrossRef] [Green Version]
  91. Takahashi, H.; Yamamoto, T.; Tsuboi, A. Molecular mechanisms underlying activity-dependent ischemic tolerance in the brain. Neurosci. Res. 2023, 186, 3–9. [Google Scholar] [CrossRef] [PubMed]
  92. Hirayama, Y.; Anzai, N.; Kinouchi, H.; Koizumi, S. P2X7 Receptors in Astrocytes: A Switch for Ischemic Tolerance. Molecules 2022, 27, 3655. [Google Scholar] [CrossRef] [PubMed]
  93. Koizumi, S.; Hirayama, Y. Ischemic Tolerance Induced by Glial Cells. Neurochem. Res. 2022, 47, 2522–2528. [Google Scholar] [CrossRef] [PubMed]
  94. Shimizu, M.; Tropak, M.; Diaz, R.J.; Suto, F.; Surendra, H.; Kuzmin, E.; Li, J.; Gross, G.; Wilson, G.J.; Callahan, J.; et al. Transient limb ischaemia remotely preconditions through a humoral mechanism acting directly on the myocardium: Evidence suggesting cross-species protection. Clin. Sci. 2009, 117, 191–200. [Google Scholar] [CrossRef] [Green Version]
  95. Kanoria, S.; Jalan, R.; Seifalian, A.M.; Williams, R.; Davidson, B.R. Protocols and mechanisms for remote ischemic preconditioning: A novel method for reducing ischemia reperfusion injury. Transplantation 2007, 84, 445–458. [Google Scholar] [CrossRef]
  96. Burda, R.; Danielisova, V.; Burda, J. The End Effector of Ischemic Tolerance Present in Blood Plasma from Double Conditioned Donors Ameliorates Trimethyltin Provoked Damage in Brain. OBM Neurobiol. 2019, 3, 1–12. [Google Scholar] [CrossRef] [Green Version]
  97. Burda, R.; Morochovič, R.; Némethová, M.; Burda, J. Remote ischemic postconditioning as well as blood plasma from double-conditioned donor ameliorate reperfusion syndrome in skeletal muscle. J. Plast. Surg. Hand Surg. 2020, 54, 59–65. [Google Scholar] [CrossRef]
  98. Wei, D.; Ren, C.; Chen, X.; Zhao, H. The chronic protective effects of limb remote preconditioning and the underlying mechanisms involved in inflammatory factors in rat stroke. PLoS ONE 2012, 7, e30892. [Google Scholar] [CrossRef]
  99. Mastitskaya, S.; Marina, N.; Gourine, A.; Gilbey, M.P.; Spyer, K.M.; Teschemacher, A.G.; Kasparov, S.; Trapp, S.; Ackland, G.L.; Gourine, A.V. Cardioprotection evoked by remote ischaemic preconditioning is critically dependent on the activity of vagal pre-ganglionic neurones. Cardiovasc. Res. 2012, 95, 487–494. [Google Scholar] [CrossRef]
  100. Sun, Z.; Baker, W.; Hiraki, T.; Greenberg, J.H. The effect of right vagus nerve stimulation on focal cerebral ischemia: An experimental study in the rat. Brain Stimul. 2012, 5, 1–10. [Google Scholar] [CrossRef] [Green Version]
  101. Donato, M.; Buchholz, B.; Rodriguez, M.; Perez, V.; Inserte, J.; Garcia-Dorado, D.; Gelpi, R.J. Role of the parasympathetic nervous system in cardioprotection by remote hindlimb ischaemic preconditioning. Exp. Physiol. 2013, 98, 425–434. [Google Scholar] [CrossRef] [PubMed]
  102. Steensrud, T.; Li, J.; Dai, X.; Manlhiot, C.; Kharbanda, R.K.; Tropak, M.; Redington, A. Pretreatment with the nitric oxide donor SNAP or nerve transection blocks humoral preconditioning by remote limb ischemia or intra-arterial adenosine. Am. J. Physiol. Heart Circ. Physiol. 2010, 299, H1598–H1603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Lim, S.Y.; Yellon, D.M.; Hausenloy, D.J. The neural and humoral pathways in remote limb ischemic preconditioning. Basic Res. Cardiol. 2010, 105, 651–655. [Google Scholar] [CrossRef]
  104. Konstantinov, I.E.; Arab, S.; Kharbanda, R.K.; Li, J.; Cheung, M.M.; Cherepanov, V.; Downey, G.P.; Liu, P.P.; Cukerman, E.; Coles, J.G.; et al. The remote ischemic preconditioning stimulus modifies inflammatory gene expression in humans. Physiol. Genom. 2004, 19, 143–150. [Google Scholar] [CrossRef] [Green Version]
  105. Shimizu, M.; Saxena, P.; Konstantinov, I.E.; Cherepanov, V.; Cheung, M.M.; Wearden, P.; Zhangdong, H.; Schmidt, M.; Downey, G.P.; Redington, A.N. Remote ischemic preconditioning decreases adhesion and selectively modifies functional responses of human neutrophils. J. Surg. Res. 2010, 158, 155–161. [Google Scholar] [CrossRef]
  106. Weber, C. Far from the heart: Receptor cross-talk in remote conditioning. Nat. Med. 2010, 16, 760–762. [Google Scholar] [CrossRef]
  107. Fan, J.; Zhang, Z.; Chao, X.; Gu, J.; Cai, W.; Zhou, W.; Yin, G.; Li, Q. Ischemic preconditioning enhances autophagy but suppresses autophagic cell death in rat spinal neurons following ischemia-reperfusion. Brain Res. 2014, 1562, 76–86. [Google Scholar] [CrossRef]
  108. Lv, S.; Ju, C.; Peng, J.; Liang, M.; Zhu, F.; Wang, C.; Huang, K.; Cheng, M.; Zhang, F. 25-Hydroxycholesterol protects against myocardial ischemia-reperfusion injury via inhibiting PARP activity. Int. J. Biol. Sci. 2020, 16, 298–308. [Google Scholar] [CrossRef] [Green Version]
  109. Wang, J.K.; Yu, L.N.; Zhang, F.J.; Yang, M.J.; Yu, J.; Yan, M.; Chen, G. Postconditioning with sevoflurane protects against focal cerebral ischemia and reperfusion injury via PI3K/Akt pathway. Brain Res. 2010, 1357, 142–151. [Google Scholar] [CrossRef] [PubMed]
  110. Pignataro, G.; Meller, R.; Inoue, K.; Ordonez, A.N.; Ashley, M.D.; Xiong, Z.; Gala, R.; Simon, R.P. In vivo and in vitro characterization of a novel neuroprotective strategy for stroke: Ischemic postconditioning. J. Cereb. Blood Flow Metab. 2008, 28, 232–241. [Google Scholar] [CrossRef] [Green Version]
  111. Jewell, J.L.; Guan, K.L. Nutrient signaling to mTOR and cell growth. Trends Biochem. Sci. 2013, 38, 233–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Wang, C.; Wang, Z.; Zhang, X.; Zhang, X.; Dong, L.; Xing, Y.; Li, Y.; Liu, Z.; Chen, L.; Qiao, H.; et al. Protection by silibinin against experimental ischemic stroke: Up-regulated pAkt, pmTOR, HIF-1alpha and Bcl-2, down-regulated Bax, NF-kappaB expression. Neurosci. Lett. 2012, 529, 45–50. [Google Scholar] [CrossRef] [PubMed]
  113. Xie, R.; Wang, P.; Ji, X.; Zhao, H. Ischemic post-conditioning facilitates brain recovery after stroke by promoting Akt/mTOR activity in nude rats. J. Neurochem. 2013, 127, 723–732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Chen, G.Z.; Shan, X.Y.; Li, X.S.; Tao, H.M. Remote ischemic postconditioning protects the brain from focal ischemia/reperfusion injury by inhibiting autophagy through the mTOR/p70S6K pathway. Neurol. Res. 2018, 40, 182–188. [Google Scholar] [CrossRef] [PubMed]
  115. Gao, X.; Zhang, H.; Takahashi, T.; Hsieh, J.; Liao, J.; Steinberg, G.K.; Zhao, H. The Akt signaling pathway contributes to postconditioning’s protection against stroke; the protection is associated with the MAPK and PKC pathways. J. Neurochem. 2008, 105, 943–955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Shimohata, T.; Zhao, H.; Steinberg, G.K. Epsilon PKC may contribute to the protective effect of hypothermia in a rat focal cerebral ischemia model. Stroke 2007, 38, 375–380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Qi, W.; Zhou, F.; Li, S.; Zong, Y.; Zhang, M.; Lin, Y.; Zhang, X.; Yang, H.; Zou, Y.; Qi, C.; et al. Remote ischemic postconditioning protects ischemic brain from injury in rats with focal cerebral ischemia/reperfusion associated with suppression of TLR4 and NF-small ka, CyrillicB expression. Neuroreport 2016, 27, 469–475. [Google Scholar] [CrossRef]
  118. Feng, R.; Li, S.; Li, F. Toll-like receptor 4 is involved in ischemic tolerance of postconditioning in hippocampus of tree shrews to thrombotic cerebral ischemia. Brain Res. 2011, 1384, 118–127. [Google Scholar] [CrossRef]
  119. Burda, J.; Martin, M.E.; Garcia, A.; Alcazar, A.; Fando, J.L.; Salinas, M. Phosphorylation of the alpha subunit of initiation factor 2 correlates with the inhibition of translation following transient cerebral ischaemia in the rat. Biochem. J. 1994, 302 Pt 2, 335–338. [Google Scholar] [CrossRef]
  120. Garcia, L.; Burda, J.; Hrehorovska, M.; Burda, R.; Martin, M.E.; Salinas, M. Ischaemic preconditioning in the rat brain: Effect on the activity of several initiation factors, Akt and extracellular signal-regulated protein kinase phosphorylation, and GRP78 and GADD34 expression. J. Neurochem. 2004, 88, 136–147. [Google Scholar] [CrossRef] [Green Version]
  121. Pagliaro, P.; Femmino, S.; Popara, J.; Penna, C. Mitochondria in Cardiac Postconditioning. Front. Physiol. 2018, 9, 287. [Google Scholar] [CrossRef] [Green Version]
  122. Argaud, L.; Gateau-Roesch, O.; Augeul, L.; Couture-Lepetit, E.; Loufouat, J.; Gomez, L.; Robert, D.; Ovize, M. Increased mitochondrial calcium coexists with decreased reperfusion injury in postconditioned (but not preconditioned) hearts. Am. J. Physiol. Heart Circ. Physiol. 2008, 294, H386–H391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Argaud, L.; Gateau-Roesch, O.; Raisky, O.; Loufouat, J.; Robert, D.; Ovize, M. Postconditioning inhibits mitochondrial permeability transition. Circulation 2005, 111, 194–197. [Google Scholar] [CrossRef] [PubMed]
  124. Dongworth, R.K.; Hall, A.R.; Burke, N.; Hausenloy, D.J. Targeting mitochondria for cardioprotection: Examining the benefit for patients. Future Cardiol. 2014, 10, 255–272. [Google Scholar] [CrossRef]
  125. Boengler, K.; Hilfiker-Kleiner, D.; Heusch, G.; Schulz, R. Inhibition of permeability transition pore opening by mitochondrial STAT3 and its role in myocardial ischemia/reperfusion. Basic Res. Cardiol. 2010, 105, 771–785. [Google Scholar] [CrossRef] [Green Version]
  126. Garlid, K.D.; Halestrap, A.P. The mitochondrial K(ATP) channel–fact or fiction? J. Mol. Cell Cardiol. 2012, 52, 578–583. [Google Scholar] [CrossRef] [Green Version]
  127. Di Lisa, F.; Canton, M.; Carpi, A.; Kaludercic, N.; Menabo, R.; Menazza, S.; Semenzato, M. Mitochondrial injury and protection in ischemic pre- and postconditioning. Antioxid. Redox Signal. 2011, 14, 881–891. [Google Scholar] [CrossRef]
  128. Penna, C.; Rastaldo, R.; Mancardi, D.; Raimondo, S.; Cappello, S.; Gattullo, D.; Losano, G.; Pagliaro, P. Post-conditioning induced cardioprotection requires signaling through a redox-sensitive mechanism, mitochondrial ATP-sensitive K+ channel and protein kinase C activation. Basic Res. Cardiol. 2006, 101, 180–189. [Google Scholar] [CrossRef]
  129. Serviddio, G.; Di Venosa, N.; Federici, A.; D’Agostino, D.; Rollo, T.; Prigigallo, F.; Altomare, E.; Fiore, T.; Vendemiale, G. Brief hypoxia before normoxic reperfusion (postconditioning) protects the heart against ischemia-reperfusion injury by preventing mitochondria peroxyde production and glutathione depletion. FASEB J. 2005, 19, 354–361. [Google Scholar] [CrossRef] [PubMed]
  130. Bopassa, J.C.; Ferrera, R.; Gateau-Roesch, O.; Couture-Lepetit, E.; Ovize, M. PI 3-kinase regulates the mitochondrial transition pore in controlled reperfusion and postconditioning. Cardiovasc. Res. 2006, 69, 178–185. [Google Scholar] [CrossRef] [Green Version]
  131. Penna, C.; Perrelli, M.G.; Pagliaro, P. Mitochondrial pathways, permeability transition pore, and redox signaling in cardioprotection: Therapeutic implications. Antioxid. Redox Signal. 2013, 18, 556–599. [Google Scholar] [CrossRef] [PubMed]
  132. Penna, C.; Granata, R.; Tocchetti, C.G.; Gallo, M.P.; Alloatti, G.; Pagliaro, P. Endogenous Cardioprotective Agents: Role in Pre and Postconditioning. Curr. Drug Targets 2015, 16, 843–867. [Google Scholar] [CrossRef] [Green Version]
  133. Obal, D.; Dettwiler, S.; Favoccia, C.; Scharbatke, H.; Preckel, B.; Schlack, W. The influence of mitochondrial KATP-channels in the cardioprotection of preconditioning and postconditioning by sevoflurane in the rat in vivo. Anesth. Analg. 2005, 101, 1252–1260. [Google Scholar] [CrossRef]
  134. Pravdic, D.; Mio, Y.; Sedlic, F.; Pratt, P.F.; Warltier, D.C.; Bosnjak, Z.J.; Bienengraeber, M. Isoflurane protects cardiomyocytes and mitochondria by immediate and cytosol-independent action at reperfusion. Br. J. Pharmacol. 2010, 160, 220–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. He, W.; Zhang, F.J.; Wang, S.P.; Chen, G.; Chen, C.C.; Yan, M. Postconditioning of sevoflurane and propofol is associated with mitochondrial permeability transition pore. J. Zhejiang Univ. Sci. B 2008, 9, 100–108. [Google Scholar] [CrossRef] [Green Version]
  136. Murphy, E.; Steenbergen, C. What makes the mitochondria a killer? Can we condition them to be less destructive? Biochim. Biophys. Acta 2011, 1813, 1302–1308. [Google Scholar] [CrossRef] [Green Version]
  137. Penna, C.; Perrelli, M.G.; Raimondo, S.; Tullio, F.; Merlino, A.; Moro, F.; Geuna, S.; Mancardi, D.; Pagliaro, P. Postconditioning induces an anti-apoptotic effect and preserves mitochondrial integrity in isolated rat hearts. Biochim. Biophys. Acta 2009, 1787, 794–801. [Google Scholar] [CrossRef] [Green Version]
  138. Seewald, M.; Coles, J.A., Jr.; Sigg, D.C.; Iaizzo, P.A. Featured Article: Pharmacological postconditioning with delta opioid attenuates myocardial reperfusion injury in isolated porcine hearts. Exp. Biol. Med. 2017, 242, 986–995. [Google Scholar] [CrossRef]
  139. Khan, H.; Kashyap, A.; Kaur, A.; Singh, T.G. Pharmacological postconditioning: A molecular aspect in ischemic injury. J. Pharm. Pharmacol. 2020, 72, 1513–1527. [Google Scholar] [CrossRef]
  140. Bahde, R.; Spiegel, H.U. Hepatic ischaemia-reperfusion injury from bench to bedside. Br. J. Surg. 2010, 97, 1461–1475. [Google Scholar] [CrossRef] [PubMed]
  141. Meybohm, P.; Gruenewald, M.; Zacharowski, K.D.; Albrecht, M.; Lucius, R.; Fosel, N.; Hensler, J.; Zitta, K.; Bein, B. Mild hypothermia alone or in combination with anesthetic post-conditioning reduces expression of inflammatory cytokines in the cerebral cortex of pigs after cardiopulmonary resuscitation. Crit. Care 2010, 14, R21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Pateliya, B.B.; Singh, N.; Jaggi, A.S. Possible role of opioids and KATP channels in neuroprotective effect of postconditioning in mice. Biol. Pharm. Bull. 2008, 31, 1755–1760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Burda, J.; Danielisova, V.; Nemethova, M.; Gottlieb, M.; Kravcukova, P.; Domorakova, I.; Mechirova, E.; Burda, R. Postconditioning and anticonditioning: Possibilities to interfere to evoked apoptosis. Cell. Mol. Neurobiol. 2009, 29, 821–825. [Google Scholar] [CrossRef]
  144. Korthals, J.K.; Maki, T.; Gieron, M.A. Nerve and muscle vulnerability to ischemia. J. Neurol. Sci. 1985, 71, 283–290. [Google Scholar] [CrossRef]
  145. Jenkins, D.P.; Pugsley, W.B.; Alkhulaifi, A.M.; Kemp, M.; Hooper, J.; Yellon, D.M. Ischaemic preconditioning reduces troponin T release in patients undergoing coronary artery bypass surgery. Heart 1997, 77, 314–318. [Google Scholar] [CrossRef] [Green Version]
  146. Burda, R. How to Influence the Extent of Reperfusion Syndromme in Replantations and Revascularisations, Faculty of Medicine, Pavol Jozef Šafárik University, Košice, Slovakia. 2014. Available online: https://opac.crzp.sk/?fn=detailBiblioForm&sid=CB068CE69013CFB6D36EE4794CD3 (accessed on 20 February 2023).
  147. Tsubota, H.; Marui, A.; Esaki, J.; Bir, S.C.; Ikeda, T.; Sakata, R. Remote postconditioning may attenuate ischaemia-reperfusion injury in the murine hindlimb through adenosine receptor activation. Eur. J. Vasc. Endovasc. Surg. 2010, 40, 804–809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Pignataro, G.; Esposito, E.; Sirabella, R.; Vinciguerra, A.; Cuomo, O.; Di Renzo, G.; Annunziato, L. nNOS and p-ERK involvement in the neuroprotection exerted by remote postconditioning in rats subjected to transient middle cerebral artery occlusion. Neurobiol. Dis. 2013, 54, 105–114. [Google Scholar] [CrossRef]
  149. Kolbenschlag, J.; Sogorski, A.; Kapalschinski, N.; Harati, K.; Lehnhardt, M.; Daigeler, A.; Hirsch, T.; Goertz, O. Remote Ischemic Conditioning Improves Blood Flow and Oxygen Saturation in Pedicled and Free Surgical Flaps. Plast. Reconstr. Surg. 2016, 138, 1089–1097. [Google Scholar] [CrossRef]
  150. Heusch, G.; Gersh, B.J. The pathophysiology of acute myocardial infarction and strategies of protection beyond reperfusion: A continual challenge. Eur. Heart J. 2017, 38, 774–784. [Google Scholar] [CrossRef] [Green Version]
  151. Hausenloy, D.J.; Candilio, L.; Laing, C.; Kunst, G.; Pepper, J.; Kolvekar, S.; Evans, R.; Robertson, S.; Knight, R.; Ariti, C.; et al. Effect of remote ischemic preconditioning on clinical outcomes in patients undergoing coronary artery bypass graft surgery (ERICCA): Rationale and study design of a multi-centre randomized double-blinded controlled clinical trial. Clin. Res. Cardiol. 2012, 101, 339–348. [Google Scholar] [CrossRef]
  152. Meybohm, P.; Bein, B.; Brosteanu, O.; Cremer, J.; Gruenewald, M.; Stoppe, C.; Coburn, M.; Schaelte, G.; Boning, A.; Niemann, B.; et al. A Multicenter Trial of Remote Ischemic Preconditioning for Heart Surgery. N. Engl. J. Med. 2015, 373, 1397–1407. [Google Scholar] [CrossRef] [PubMed]
  153. Kleinbongard, P.; Heusch, G. Extracellular signalling molecules in the ischaemic/reperfused heart—druggable and translatable for cardioprotection? Br. J. Pharmacol. 2015, 172, 2010–2025. [Google Scholar] [CrossRef] [Green Version]
  154. Tyagi, S.; Singh, N.; Virdi, J.K.; Jaggi, A.S. Diabetes abolish cardioprotective effects of remote ischemic conditioning: Evidences and possible mechanisms. J. Physiol. Biochem. 2019, 75, 19–28. [Google Scholar] [CrossRef] [PubMed]
  155. Schwartz Longacre, L.; Kloner, R.A.; Arai, A.E.; Baines, C.P.; Bolli, R.; Braunwald, E.; Downey, J.; Gibbons, R.J.; Gottlieb, R.A.; Heusch, G.; et al. New horizons in cardioprotection: Recommendations from the 2010 National Heart, Lung, and Blood Institute Workshop. Circulation 2011, 124, 1172–1179. [Google Scholar] [CrossRef]
  156. Burda, R.; Danielisova, V.; Gottlieb, M.; Nemethova, M.; Bonova, P.; Matiasova, M.; Morochovic, R.; Burda, J. Delayed remote ischemic postconditioning protects against transient cerebral ischemia/reperfusion as well as kainate-induced injury in rats. Acta Histochem. 2014, 116, 1062–1067. [Google Scholar] [CrossRef] [PubMed]
  157. Zhao, Y.; Zheng, Z.N.; Liu, X.; Dai, G.; Jin, S.Q. Effects of preconditioned plasma collected during the late phase of remote ischaemic preconditioning on ventricular arrhythmias caused by myocardial ischaemia reperfusion in rats. J. Int. Med. Res. 2018, 46, 1370–1379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Weber, N.C.; Riedemann, I.; Smit, K.F.; Zitta, K.; van de Vondervoort, D.; Zuurbier, C.J.; Hollmann, M.W.; Preckel, B.; Albrecht, M. Plasma from human volunteers subjected to remote ischemic preconditioning protects human endothelial cells from hypoxia-induced cell damage. Basic Res. Cardiol. 2015, 110, 17. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic illustration of the effect of IR damage. As a result of ischemia, some of the cells are primarily irreversibly damaged; they can no longer be influenced therapeutically. Reversibly damaged cells can be influenced. The next approach can lead to their recovery or to irreversible necrosis. Explanations: the yellow circle with the black one inside represents all cells that experienced ischemia, the yellow circle represents the cells with reversible damage, the black circle represents the dead cells (irreversible damage), and the green circle represents the surviving viable cells.
Figure 1. Schematic illustration of the effect of IR damage. As a result of ischemia, some of the cells are primarily irreversibly damaged; they can no longer be influenced therapeutically. Reversibly damaged cells can be influenced. The next approach can lead to their recovery or to irreversible necrosis. Explanations: the yellow circle with the black one inside represents all cells that experienced ischemia, the yellow circle represents the cells with reversible damage, the black circle represents the dead cells (irreversible damage), and the green circle represents the surviving viable cells.
Cells 12 00884 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Burda, R.; Burda, J.; Morochovič, R. Ischemic Tolerance—A Way to Reduce the Extent of Ischemia–Reperfusion Damage. Cells 2023, 12, 884. https://doi.org/10.3390/cells12060884

AMA Style

Burda R, Burda J, Morochovič R. Ischemic Tolerance—A Way to Reduce the Extent of Ischemia–Reperfusion Damage. Cells. 2023; 12(6):884. https://doi.org/10.3390/cells12060884

Chicago/Turabian Style

Burda, Rastislav, Jozef Burda, and Radoslav Morochovič. 2023. "Ischemic Tolerance—A Way to Reduce the Extent of Ischemia–Reperfusion Damage" Cells 12, no. 6: 884. https://doi.org/10.3390/cells12060884

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop