Next Article in Journal
The EPH/Ephrin System in Pancreatic Ductal Adenocarcinoma (PDAC): From Pathogenesis to Treatment
Previous Article in Journal
Novel Benzo[a]phenoxazinium Chlorides Functionalized with Sulfonamide Groups as NIR Fluorescent Probes for Vacuole, Endoplasmic Reticulum, and Plasma Membrane Staining
Previous Article in Special Issue
On the Bioactivity of Echinacea purpurea Extracts to Modulate the Production of Inflammatory Mediators
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Host-Derived Cytotoxic Agents in Chronic Inflammation and Disease Progression

Medical Faculty, Institute of Medical Physics and Biophysics, Leipzig University, Härtelstr. 16-18, 04107 Leipzig, Germany
Int. J. Mol. Sci. 2023, 24(3), 3016; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24033016
Submission received: 12 December 2022 / Revised: 20 January 2023 / Accepted: 1 February 2023 / Published: 3 February 2023
(This article belongs to the Special Issue New Trends in Inflammation Management)

Abstract

:
At inflammatory sites, cytotoxic agents are released and generated from invading immune cells and damaged tissue cells. The further fate of the inflammation highly depends on the presence of antagonizing principles that are able to inactivate these host-derived cytotoxic agents. As long as the affected tissues are well equipped with ready-to-use protective mechanisms, no damage by cytotoxic agents occurs and resolution of inflammation is initiated. However, long-lasting and severe immune responses can be associated with the decline, exhaustion, or inactivation of selected antagonizing principles. Hence, cytotoxic agents are only partially inactivated and contribute to damage of yet-unperturbed cells. Consequently, a chronic inflammatory process results. In this vicious circle of permanent cell destruction, not only novel cytotoxic elements but also novel alarmins and antigens are liberated from affected cells. In severe cases, very low protection leads to organ failure, sepsis, and septic shock. In this review, the major classes of host-derived cytotoxic agents (reactive species, oxidized heme proteins and free heme, transition metal ions, serine proteases, matrix metalloproteases, and pro-inflammatory peptides), their corresponding protective principles, and resulting implications on the pathogenesis of diseases are highlighted.

1. Introduction

Persistent, long-lasting inflammation is a general problem in many diseases. As these chronic inflammatory states are only insufficiently or not terminated, novel immune cells are again and again recruited to and activated at the inflamed loci. Often, an ongoing immunocompromised condition exists in these patients, a condition that further disturbs the recovery to normal tissue homeostasis.
During inflammation, immune responses are initiated by the ligation of pathogen-associated molecular patterns (PAMPs) and/or damage-associated molecular patterns (DAMPs) to pattern-recognition receptors (PRRs) [1,2]. DAMPs are host-derived molecules that are also known as danger signals or alarmins. PRRs are distributed in both cell membranes and cytoplasm [3,4]. As a result of PAMP and DAMP ligation to PRRs, signaling events are induced leading to the release of cytokines and antimicrobial agents and the attraction of immune cells [5]. These activities are generally directed to combat invading pathogens and to decrease tissue damage. A second activity initiated by PRRs concerns the maturation of dendritic cells for the presentation of antigens to T lymphocytes [5].
This very attractive concept of immune activation by molecular patterns covers a broad range of initiating molecules including constituents of pathogens (viruses, bacteria, fungi, and others) and agents from damaged host cells. This concept is closely associated with immune responses initiated by external and host-derived antigens. Important biomarkers of an ongoing inflammation are pro-inflammatory cytokines such as interleukin-1 (IL-1), interleukin-6 (IL-6), and tumor necrosis factor α (TNF-α) and acute-phase proteins such as C-reactive protein (CRP) and serum amyloid A (SAA) [6,7,8,9,10,11]. The termination of inflammation is also highly regulated and characterized in contrast to the initiation and propagation phases by an own set of cytokines and changed energy metabolism of immune cells [10,12]. Importantly, anti-inflammatory cytokines contribute not only to the resolution of immune responses but also to the replacement of damaged biological material by novel cells and components of the extracellular matrix [13,14]. A transient immunosuppression is typical of this phase of inflammation [15].
During inflammation, unperturbed cells and tissues of the host can be damaged. Despite huge progress in understanding molecular and regulatory aspects of inflammation, no clear answers are given about the general interplay between inflammation and cell and tissue destruction, the severity of the resulting damage, and the fate of the affected organism during chronic inflammation.
Besides physical factors such as traumata, heat, or cold impacts, the action of numerous cytotoxic agents can seriously affect intact cells and tissues [16]. Like the initiating factors of inflammation, cytotoxic agents can result from both external sources and affected cells and tissues of the host. In the initiation and propagation phases of inflammation, cytotoxic agents from activated immune cells and defective tissue cells act predominantly destructively. During the resolution of inflammation, a shifted balance in the synthesis of extracellular matrix can also affect tissue homeostasis.
In this review, the role of host-derived cytotoxic agents will be evaluated in the development of cell and tissue damage during inflammation. In addition, deviations in the balance between cytotoxic agents and protective principles will be highlighted. On this basis, the role of insufficient protection against damage in the development of chronic inflammatory states will be addressed.

2. The Balance between the Action of Cytotoxic Agents and Protective Principles

2.1. Major Classes of Host-Derived Cytotoxic Agents

The contact of cytotoxic agents with living matter worsens cell functions and can induce irreversible changes in cells and tissues including cell death. According to the source of these agents, they can be roughly divided into external and host-derived cytotoxic components. The group of external cytotoxic agents comprises pathogen-derived toxins and manifold external poisons that act on the organism by inhalation, direct contact, or uptake with food. A third group represents environmental cytotoxic agents (for details see Section 2.4). An overview about external cytotoxic agents is given in Figure 1. Selected examples of these agents are included.
Of course, these external cytotoxic agents can cause sufficient threat to the affected tissues and finally the death of the organism. However, these agents are usually not involved in long-lasting, chronic inflammation.
Host-derived cytotoxic agents result from activated immune cells such as neutrophils, eosinophils, monocytes, macrophages, and T cells, but also from affected tissue cells of non-immunological origin, for instance, muscle cells and red blood cells. Immune cells contain an arsenal of potentially cytotoxic agents that are needed to inactivate and kill pathogens and to remove and digest affected cells and destroyed tissues [10]. Usually these agents act within small, bounded compartments, e.g., within the phagosomes of neutrophils and macrophages. However, a certain amount of these cytotoxic agents is released from activated immune cells into the surrounding milieu, where they become dangerous to unperturbed cells. The following classes of immune-cell-derived cytotoxic agents are known: small reactive species, heme peroxidases, free metal ions, serine proteases, matrix metalloproteases, and small pro-inflammatory peptides. These agents are either pre-assembled or generated during cellular immune activation.
Damage to tissue cells of non-immunological origin can result in the uncontrolled release of heme proteins such as hemoglobin and myoglobin and the subsequent formation of free heme. Cellular stress is also associated with enhanced formation of reactive species and deviations in free metal ion metabolism. As a result, the metabolism of mitochondria is disturbed and numerous oxidative processes in biological constituents take place.
According to their mode of action, host-derived cytotoxic agents can be divided into oxidant- and protease-based agents (Figure 2). Products of the first group promote oxidative alterations of biological constituents, whereas members of the second group cause proteolytic cleavage in cell and tissue components.
An overview of the major host-derived cytotoxic agents is given in Table 1. This overview also includes key information about the modes of action of these dangerous molecules and naturally occurring protective principles to avoid substantial damage. More details about host-derived cytotoxic agents are given in Section 3.

2.2. Control of Cytotoxic Agents by Protective Principles

The destructive action of host-derived cytotoxic agents depends not only on the mass of released cytotoxic agents, but also on the current status of host-own protective principles [10,17]. In order to curtail or avoid destruction by these agents, numerous ready-to-use mechanisms exist in cells and tissues to inactivate immediately hazardous components released from activated immune and affected tissue cells. Usually, unperturbed cells and tissues are well equipped with protective principles. In this way, any threat to unperturbed tissue components is minimized. The major antagonizing principles are listed in Table 1 in relation to their cytotoxic agents.
The balance between host-derived cytotoxic agents and protective principles functions well as long as the activation of immune cells is moderate enough and neighboring tissues are well-equipped with ready-to-use protective mechanisms (Figure 3). Problems can arise with severe and long-lasting immune responses and with the decline, exhaustion, or inactivation of selected antagonizing principles despite an up-regulation of many protective proteins under stress situations. In turn, long-lasting inflammatory processes can result from the permanent release of cytotoxic agents from damaged cells in combination with insufficient inactivation of these agents. In other words, low expression of a few protective principles favors the continuous action of destructive agents and affects still-unperturbed cells. In this vicious circle of permanent cell destruction, not only novel cytotoxic elements but also novel alarmins and antigens are liberated from affected cells. In addition, pro-inflammatory peptides such as angiotensin II and bradykinin are formed by insufficient inactivation of serine proteases. Hence, the inflammation cannot be terminated sufficiently and flares up again and again. In severe cases, a very low level of protection leads to organ failure, sepsis, and septic shock.
To overcome chronic inflammation, it is highly essential, besides inhibition of selected pathways in the inflammatory cascade, to improve poorly expressed protective systems to better detoxify the damaging agents.

2.3. Disturbed Balance between De Novo Synthesis and Damage of Tissue Components during Resolution of Inflammation

Termination of inflammation is characterized by the down-regulation of pro-inflammatory cells, cytokines, and signaling pathways as well as by the formation of anti-inflammatory mediators and induction of repair processes. During this phase of inflammation, cytokines of the transforming growth factor β (TGF-β) family, which are secreted from M2-type macrophages and some other cells, suppress together with interleukin 10 (IL-10)-activated immune cells [15,18]. These cytokines also promote tissue repair by stimulating fibroblasts to synthesize collagen and other components of the extracellular matrix (ECM) and by the release of tissue inhibitors of metalloproteases (TIMPs) [19,20]. The latter inhibitors down-regulate the activity of matrix metalloproteases (MMPs) and thus prevent degradation of ECM components.

2.4. Selected Environmental Cytotoxic Agents

Although not host-derived, we can also be exposed to external cytotoxic agents (see Figure 1). Of these agents, environmental cytotoxic agents act more or less intensely and permanently on our organism. As this exposure concerns nearly all persons, these agents are usually detoxified by antagonizing principles when the exposure is moderate and does not exceed a critical level. Examples of environmental cytotoxic agents, their mode of action, and antagonizing principles are given in Table 2.

3. Selected Cytotoxic Agents and Their Counter-Regulating Principles

3.1. Small Reactive Species and Metal Ions

3.1.1. Superoxide Anion Radicals

The stepwise reduction of dioxygen yields the species superoxide anion radical (O2•−) and hydrogen peroxide (H2O2) [36]. These species are less dangerous concerning their direct action on tissue components. However, they are involved in the formation of highly reactive and tissue-damaging agents by interaction with radicals, metal ions, and iron-containing proteins.
Activated leukocytes are able to generate large amounts of O2•− by reducing dioxygen. This reaction is catalyzed by NADPH oxidase, which is assembled from several membranous and cytoplasmic components during the activation of neutrophils, eosinophils, monocytes, and macrophages [37,38]. NADPH oxidases are also distributed in cells of the blood vessel wall, respiratory tract, gastrointestinal tract, and thyroid gland [39,40,41]. However, these enzymes are less efficient in reducing dioxygen than NADPH oxidase from immune cells. Other sources for superoxide anion radicals are reactions of xanthine oxidase [42,43], autoxidation of hemoglobin and myoglobin [44,45], cytochrome P450-driven redox recycling of some xenobiotica [46,47], and one-electron reduction of dioxygen by different mitochondrial enzymes [48,49].
Superoxide anion radicals are unstable. Two superoxide anion radicals dismutate spontaneously to hydrogen peroxide and dioxygen [50]. The rate of this dismutation highly depends on pH, with a maximal rate around pH 4.8, the pka value of O2•−, and decreasing rates with increasing pH [51]. With one unit pH increase, the dismutation rate of O2•− decreases by one order of magnitude. At pH 7.4 this rate is 2 × 105 M−1s−1 [51].
Superoxide anion radical reacts in a very rapid reaction with nitrogen monoxide, also a radical species, under the formation of the powerful oxidant peroxynitrite [52,53]. In mitochondria, superoxide anion radical is able to release Fe2+ from molecules containing [4Fe-4S]2+ clusters such as aconitase [54,55].
In humans, control over O2•− is realized with three isoforms of superoxide dismutase (SOD) and cytochrome c. SOD1 is distributed in the cytoplasm, intermembrane space of mitochondria, and nuclei [56,57]. In the mitochondrial matrix, SOD2 dominates [58]. SOD3 is mostly found in blood vessel walls and lungs [59]. These enzymes catalyze the dismutation of O2•− with a rate several orders higher than the spontaneous dismutation reaction of O2•−. In the intermembrane space of mitochondria, oxidized cytochrome c oxidizes O2•− to O2, thus contributing to the detoxification of O2•− [60,61].
Figure 4 depicts the major pathways for the formation of reactive species with a special focus on processes in activated neutrophils and stressed mitochondria. In both systems, the generation of small reactive species starts with the reduction of dioxygen to superoxide anion radicals.

3.1.2. Hydrogen Peroxide

Spontaneous and SOD-catalyzed dismutation of O2•− represent the main route of formation of H2O2. Thus, all processes generating O2•− also yield H2O2. Otherwise, different peroxisomal enzymes are able to reduce O2 directly to H2O2 [62].
Due to its electronic structure, reactions of H2O2 are restricted to transition metal ions, complexes of these ions, and some proteins with selenocysteine (or cysteine) residues at the active site [63,64]. Hydrogen peroxide is freely permeable through biological membranes, unlike O2•−. The interaction of transition metal ions such as Fe2+ and Cu+ with H2O2 yields very reactive hydroxyl radicals and metal-based reactive species that can cause manifold damaging reactions on biological material [65,66].
Heme peroxidases, different cytochromes, hemoglobin, and myoglobin are activated by H2O2 leading to reactive states of the heme in these proteins. During immune response, H2O2 activates the heme peroxidases myeloperoxidase (MPO), eosinophil peroxidase (EPO), and lactoperoxidase (LPO), which are involved in both pro- and anti-inflammatory activities [17,67,68,69,70].
Several enzymes are known to catalyze the reduction of H2O2 to H2O (Figure 4). Glutathione peroxidase (GPX) utilizes glutathione (GSH) to reduce H2O2. The highly distributed isoforms GPX1 and especially GPX4 also detoxify peroxynitrite, lipid hydroperoxides, and other organic hydroperoxides [71,72]. GSH is recovered from the resulting oxidized glutathione (GSSG) by glutathione reductase [73]. Peroxiredoxins, which are closely coupled to the thioredoxin system, also efficiently reduce H2O2 to H2O [74]. Catalase removes H2O2 by both reduction to H2O and oxidation to O2 [75].

3.1.3. Transition Metal Ions and Hydroxyl Radicals

In the reaction between H2O2 and Fe2+, which is known as the Fenton reaction, the highly reactive hydroxyl radical is formed. Alternatively, iron–oxygen complexes such as ferryl or perferryl compounds are discussed as products of this reaction [65,66]. Similarly, the reaction of H2O2 with Cu+ also yields hydroxyl radicals [76]. Organic hydroperoxides are also oxidized by Fe2+ and Cu+ under the formation of reactive radical species that are involved in subsequent destructive reactions. Beyond Fenton chemistry, further mechanisms apparently contribute to metal-ion-induced tissue damage such as the interaction of Fe2+ with biological buffer components or the formation of Fe2+–O2 and Fe2+–O2–Fe3+ complexes [77,78,79,80,81].
Hydroxyl radicals react in a nearly diffusion-controlled manner with many substrates by abstraction of an H-atom or by addition to an unsaturated system under formation of a hydroxylated product [82]. In both reaction types, substrate radicals are formed that can undergo manifold further reactions.
To avoid the disastrous formation of reactive species such as hydroxyl radicals and others, the main strategy of living matter is the tight control of transport, storage, and utilization of free metal ions (Figure 4) as both iron and copper ions are necessary constituents of many proteins [83,84]. Major components controlling iron metabolism are hepcidin (intestinal absorption), transferrin (blood transport), transferrin receptor (uptake by cells), and ferritin (intracellular storage) [85,86,87,88]. Similarly, different import and export transporters and chaperones are involved in copper metabolism [89]. Ceruloplasmin is able to oxidize both Fe2+ and Cu+ [90]. Lactoferrin released from activated neutrophils binds Fe3+ and promotes its transfer to transferrin [91].

3.1.4. Peroxynitrite

As already mentioned, peroxynitrite is formed in a very rapid reaction between O2•− and NO [52,53]. Peroxynitrite is involved in the formation of thiyl radicals and nitration of tyrosine residues, and is able to induce lipid-peroxidation processes [92,93,94,95]. In reaction with CO2, it yields nitrosoperoxycarbonate, which can decompose into radical species [96,97].
At inflammatory sites where heme peroxidases are present, peroxynitrite is decomposed in its reaction with resting MPO [98,99,100]. Other redox-active heme proteins scavenge peroxynitrite and inactivate this powerful oxidant [101,102,103].

3.1.5. Heme Peroxidases and Hypohalous Acids

At an inflammatory site, the heme protein MPO can be released from activated neutrophils (Figure 4) [67,104]. Eosinophils contain a similar peroxidase, the eosinophil peroxidase (EPO) [105]. A third immunologically relevant heme peroxidase is LPO, which is distributed in mucous surfaces [68]. All three heme peroxidases are able to oxidize SCN to OSCN. MPO and EPO also oxidize Br to HOBr, whereas only MPO is able to yield HOCl from Cl oxidation [70].
The MPO product HOCl reacts efficiently with methionine and cysteine residues of proteins. Further major protein targets for HOCl are residues of cystine, histidine, tryptophan, lysine, and α-amino groups [106,107]. HOBr , like HOCl, also oxidizes many residues in proteins, especially cysteine and methionine ones. HOBr induces ring halogenation in tyrosine residues more efficiently than HOCl [108].
Both HOCl and HOBr are inactivated at a high rate by thiocyanate (SCN) [109,110]. HOCl is additionally inactivated by Br. Further antagonizing principles against both hypohalous acids are ascorbate, GSH, taurine, and, additionally for HOBr, urate [111].
In blood, MPO and EPO are inactivated by ceruloplasmin through the formation of a tight inhibitory complex between heme peroxidase and ceruloplasmin [112,113,114,115].

3.2. Hemoglobin and Myoglobin Metabolites

There is always a release of intact hemoglobin from red blood cells and myoglobin from muscles at low levels. Intravascular hemolysis and rhabdomyolysis can be markedly enhanced under stress and disease situations (Figure 5). Once released from red blood cells, tetrameric ferric hemoglobin dissociates into dimers and is easily oxidized to methemoglobin. This oxidation is usually caused by nitric monoxide. Excessive intravascular hemolysis can affect the bioavailability of NO [116,117]. The serum protein haptoglobin is able to scavenge free methemoglobin. The resulting haptoglobin–methemoglobin complex is eliminated from circulating blood by spleen and liver macrophages in a CD163-dependent process [118,119]. In a similar way, haptoglobin also scavenges metmyoglobin formed after the release of myoglobin from muscle cells.
Although it is an acute-phase protein, the capacity of haptoglobin is limited when severe intravascular hemolysis or rhabdomyolysis occur. Both methemoglobin and metmyoglobin spontaneously liberate ferric protoporphyrin IX, briefly known as free heme, a very dangerous molecule [120]. Free heme easily intercalates into the lipid phases of membranes and lipoproteins and the hydrophobic areas of proteins. At these loci, it catalyzes oxidative processes [121,122]. In intact red blood cells, free heme induces hemolytic processes, thus enhancing existing intravascular hemolysis [123,124]. Free heme is highly cytotoxic to kidney and liver [125,126]. It is also a ligand to toll-like receptor 4 and thus contributes to the intensification of inflammatory processes [127,128]. In the nucleus, free heme interacts with parallel guanine-rich quadruplex DNA and RNA structural elements, known as G4 structures [129,130].
In order to avoid the disastrous activities of free heme, different serum proteins are able to complex and inactivate free heme. Hemopexin binds free heme with high affinity. This free-heme–hemopexin complex is liberated from circulating blood via CD91-mediated internalization by hepatocytes [131]. In humans, unlike mice, hemopexin is not an acute-phase protein [132].
Inside cells, free heme is detoxified by an interaction with heme oxygenase [133,134].

3.3. Oxidation of Cell and Tissue Components

In addition to proteolytic cleavage, lipids, proteins, nucleic acids, and carbohydrates are subjected under stress conditions to numerous chemical processes, whereby oxidative modifications predominate [135]. The major initiating agents of these oxidative processes are highly reactive species, free transition metal ions, free heme, and aldehydes. Besides the open chain form of glucose [136], aldehydes result mostly from oxidative modifications of lipids [137,138].
Oxidative alterations of biological substrates are counterbalanced by lipid- and water-based antioxidant mechanisms. In lipid phases, major natural antioxidants are α-tocopherol, β-carotene, ubiquinol, and dehydrolipoic acid. They are mainly involved in the scavenging of lipid peroxyl radicals [139,140,141]. Inactivation of lipid hydroperoxides is a further strategy to prevent oxidative processes. This is achieved most of all by the action of glutathione peroxidase 4 (GPX4). A high intracellular level of GSH is essential for the proper action of GPX4 and other glutathione peroxidases [71,72]. In addition, perturbed acyl residues in phospholipids are cleaved by phospholipases [142]. A thorough control over transition free metal ions also contributes to the prevention of oxidative processes in membranes and lipoproteins.
Urate and ascorbate are the main water-soluble antioxidants in our organism [143,144]. Different polyphenols are important dietary antioxidants [145]. They exert their protective action by radical scavenging, sequestration of free metal ions, and interaction with activated complexes of heme proteins [146,147,148].

3.4. Serine Proteases

3.4.1. Release of Serine Proteases from Immune Cells

At inflammatory sites, activated neutrophils can release the serine proteases elastase, cathepsin G, proteinase 3, and neutrophil serine protease 4. These proteases are primarily involved in the deactivation, killing, and digestion of phagocytosed microorganisms in neutrophils. Their pH optimum is around 8–9, a condition that predominates in early phagosomes of neutrophils [149,150]. Elastase exhibits a killing activity against Gram-negative bacteria [151,152] and a variety of cancer cells [153]. In cancer cells, unlike non-cancer cells, elastase cleaves CD95 to liberate a death domain fragment that acts cytotoxically together with histone H1 [153].
Serine proteases participate in the recruitment of neutrophils to a destination site by digestion of the surrounding tissue components and the induction and regulation of immune signaling. Elastase and proteinase 3 are able to cleave a broad range of chemokines and cytokines [154]. The substrate specificity of cathepsin G is also relatively broad but more restricted than that observed for elastase and proteinase 3 [155]. In these experiments, only a few cytokines and chemokines, such as TNF-α, interleukin 5 (IL-5), interleukin 8 (IL-8), macrophage colony-stimulating factor (M-CSF), monocyte chemoattractant protein 1 (MCP-1), IL-1α, and Rantes, were resistant to neutrophil serine proteases.
Elastase and other serine proteases are attached together with other neutrophil proteins to a DNA network in neutrophil extracellular traps. These traps can kill external microbes independent of phagocytosis [156,157].
Activated mast cells release the serine proteases chymase, tryptase, and cathepsin G [158]. These proteases are involved in matrix destruction, tissue remodeling, and regulation of inflammation. Mast cell tryptase and chymase are more restrictive than neutrophil serine proteases in the cleavage of chemokines and cytokines [154,155].

3.4.2. Activities of Neutrophil Serine Proteases

Although all serine proteases contribute to damaging reactions, the focus is mostly directed on elastase. An overview about multiple activities of neutrophil elastase during immune response is given in Figure 6. Once released from activated neutrophils, elastase can affect healthy tissues. Elastase is involved in the destruction of extracellular matrix components such as elastin, collagens, proteoglycans, and laminin [159].
Like cathepsin G, proteinase 3, and cathepsin B, neutrophil elastase is able to convert angiotensinogen and angiotensin I into angiotensin II [160,161,162]. This pro-inflammatory peptide can further foment inflammatory processes.
At inflammatory sites, neutrophil elastase activates MMP2, MMP3, and MMP9 from inactive precursors by cleaving an inhibitory protein residue [163,164,165,166]. Cathepsin G is also able to activate MMP3 [163]. Cathepsin G and proteinase 3 are involved in MMP2 activation [164]. Elastase might additionally degrade TIMP-1 [165,167].

3.4.3. Mast Cell Serine Proteases

Human mast cells contain several types of tryptases and two members of chymase-like proteins, namely α-chymase and cathepsin G, which are secreted in response to allergens and pathogens [158]. Mast cell proteases are known to stimulate the production of pro-inflammatory mediators such as IL-6 and IL-8 from bronchial epithelial cells and promote procollagen cleavage. With these activities they contribute to the recruitment of neutrophils and eosinophils at inflamed epithelium [168,169,170,171].
Other inflammation-promoting activities of chymase are the cleavage of angiotensin I into angiotensin II, activation of MMPs, and release of selected extracellular matrix elements [172]. Tryptase is involved in the degradation of fibronectin and chemokines [172]. Both tryptases and chymases contribute to the activation of different MMPs [173,174,175]. MMPs are implicated in the pathogenesis of atherosclerosis and abdominal aortic aneurysms [176,177,178,179,180]. Mast cell proteases are implied in airway epithelial remodeling and alterations in epithelium functions [181]. They also contribute to angiogenesis induction during tumor growth [182]. Chymase promotes the formation of active TGF-β from its precursor [183].
Tetrameric tryptase is stabilized by heparin and some other glycosoaminoglycans [184]. In this complex, tryptase is not accessible to anti-proteases such as A1AT, SPLI, and α2-macroglobulin [185,186]. Lactoferrin, myeloperoxidase, and antithrombin III, which are known to have heparin-binding domains, can inhibit tryptase activity [187,188,189,190]. Spontaneous dissociation of the tryptase tetramer is a further mechanism to control tryptase activity [184,191].

3.4.4. Antiproteases

Several antagonizing proteins against elastase and other serine proteases exist in blood and tissues (Figure 7). The most abundant anti-protease is the serpin α1-antitrypsin (A1AT). This serum protein is synthesized in the liver and represents an acute-phase protein. A1AT inhibits elastase and cathepsin G but not in the presence of heparin [192,193]. The activity of proteinase 3 is affected by A1AT to a lesser degree. Heparin, however, enhanced the inactivation of proteinase 3 [194].
Several factors contribute to the failure of A1AT to inhibit elastase. The inactivation of elastase requires two unperturbed methionine residues (Met-351 and Met-358) at the active site of A1AT. By oxidation of these residues A1AT loses its ability to inhibit elastase [195]. Under stress conditions methionine oxidation in A1AT can be initiated by highly reactive species such as hydroxyl radicals, peroxynitrite, HOCl, HOBr, and others [196]. Tobacco smoke and activated phagocytes are under discussion to contribute to methionine oxidation in A1AT and thus cause an acquired A1AT deficiency [196]. Furthermore, neutrophil elastase can bind to negatively charged surfaces and polymers. Surface-bound elastase cannot be inhibited by endogenous antiproteases [197].
Serpin A3, also known as α1-antichymotrypsin, is, like A1AT, an acute-phase protein. This antiprotease efficiently inactivates cathepsin G and mast cell chymase [198,199,200].
Secretory leukocyte protease inhibitor (SLPI) is able to inactivate several serine proteases such as neutrophil elastase, cathepsin G, tryptase, and chymase [201]. SLPI is constitutively expressed in mucous secretions [202,203] and also secreted from activated immune cells. It is assumed that SLPI exhibits an anti-apoptotic effect on immune cells and thus contributes to a better removal of dying cells and microbes at inflammatory sites [204].
Elafin, which is also known as proteinase inhibitor 3, is able to inactivate neutrophil elastase and proteinase 3 [205,206]. It exerts anti-inflammatory, anti-microbial, and wound-healing effects [205,206]. Contradictory results were reported about the action of elafin on tumorigenesis. These results range from promotion of cell proliferation and induction of resistance against chemotherapy to tumor-suppressive effects [207,208]. In early-stage hepatocellular carcinoma, elafin promotes metastasis formation via activation of EGFR/AKT signaling [209].
The antiprotease serpin B1 efficiently inactivates elastase, cathepsin G, and proteinase 3 [210]. Under oxidative stress, the cysteine residue at the active site in serpin B1 is oxidized with the loss of the antiprotease activity.
In contrast to the aforementioned antiproteases, which directly interact with the active site of proteases, α2-macroglobulin forms a tetrameric cage around active proteases, thus inhibiting the direct contact between protease and substrate molecules. In this way, large substrate molecules such as collagen are excluded from direct contact, whereas small peptide substrates can be digested [211,212]. Although α2-macroglobulin inhibits the activities of elastase, cathepsin G, proteinase 3, and MMP9 released from neutrophils [213,214,215], the complex between elastase and α2-macroglobulin is still active against small substrates [214]. Moreover, neutrophil-derived reactive species such as HOCl can hinder α2-macroglobulin to form tetramers and promote stabilization of dimers with the loss of the antiprotease activity [216,217].
High-affinity complexes are also known between ceruloplasmin and serine proteases of neutrophils [91]. In this way, a destructive action of serine proteases on tissue components is minimized.

3.5. Small Pro-Inflammatory Peptides

3.5.1. Angiotensin II

The peptide hormone angiotensin II is an essential part of the renin–angiotensin–aldosteron system. It is involved in the regulation of blood pressure and water metabolism. During this activity, angiotensin II is formed from angiotensin I by the angiotensin-converting enzyme (ACE).
At inflammatory sites, angiotensin II can also be produced from cleavage of both angiotensinogen and angiotensin I by serine proteases released from immune cells such as elastase, cathepsin G, proteinase 3, and mast cell chymase (Figure 8) [160,161,162,218]. Increased angiotensin II contributes via docking to AT1 and AT2 receptors to proteolysis, actin cleavage, apoptosis induction, and activation of the ubiquitin-mediated protein degradation [219,220,221]. It also promotes superoxide anion radical production via activation of NADH/NADPH oxidases [222]. Generally, these pro-inflammatory activities of angiotensin II mediate the prolonged existence of inflammatory states [223].
Angiotensin II is under the control of ACE2, which converts this octapeptide to angiotensin 1–7 [224]. This limits the devastating activity of angiotensin II. Moreover, angiotensin 1–7 exerts an anti-inflammatory activity [225].

3.5.2. Bradykinin

As an essential member of the contact system, the nonapeptide bradykinin is responsible for increased vascular permeability, vasodilation, hypotension, and other effects via interaction with its constitutively expressed B2 receptor [226,227]. A further pro-inflammatory metabolite is des-Arg9-bradykinin, which is formed from bradykinin by carboxypeptidase N. At inflammatory sites, des-Arg9-bradykinin acts selectively via bradykinin B1 receptors, which are only expressed in inflamed and injured tissue [228,229,230].
Bradykinin is a short-lived mediator of inflammation. It is inactivated by aminopeptidase P and the angiotensin-converting enzyme (ACE). Inhibition of ACE enhances bradykinin’s effects [229].

3.6. Inhibition of Matrix Metalloproteases

In human tissues, 23 MMPs and four TIMPs are found. Most MMPs are normally not expressed in healthy tissue. The activity of MMPs is essential in tissue remodeling, such as angiogenesis, bone growth, wound healing, and repair processes during the resolution of inflammation [231,232].
MMPs are secreted as inactive enzymes bearing an inhibitory prodomain that must be cleaved. In addition to neutrophil serine proteases (see Section 3.4.2), plasmin, chymases, and other MMPs are involved in MMP activation [233]. At low concentrations, highly reactive species such as HOCl, OH, and ONOO can activate MMPs. However, higher concentrations of these species inactivate active MMPs [234].
During the exudation and infiltration phase of inflammation, MMP2 and MMP9 are mainly secreted from invading immune cells, smooth muscle cells, and fibroblasts [235,236,237]. These and other MMPs contribute to cleaving the matrix components collagen and elastin.
The activity of MMPs is tightly controlled by TIMPs and α2-macroglobulin. The latter inhibitor, which has a very broad activity range against proteases, acts in blood and other biological fluids [238]. Generally, TIMPs have a broad spectrum of inhibition of MMPs. The constitutively expressed TIMP-2, like TIMP-3 and TIMP-4, is able to inhibit nearly all MMPs. TIMP-1 has a low activity against membrane-bound MMPs [232]. TIMP-3 additionally inhibits members of disintegrin metalloproteinases. Moreover, it is the only TIMP that binds to the ECM [239].

4. Enhanced Cell and Tissue Damage during Chronic Inflammatory Diseases

4.1. Most Prominent Degradative Agents

Of note, most aforementioned host-derived cytotoxic agents execute a dual role in cells and tissues. They are involved in numerous beneficial functions during metabolism and immune response. Thus, these agents are mandatory to ensure tissue homeostasis and normal functioning of the organism. To control their bad side numerous protective mechanisms help to minimize the destruction of biological constituents.
Despite the long list of host-derived damaging agents and counter-regulating principles, only a few of these agents are responsible for initiating cell and tissue degradation under pathological conditions. The reason for this damage is mainly the weakness or exhaustion of the corresponding protective system. In turn, this favors prolonged activity of the damaging agents, induces the release of DAMPs and antigens from perturbed cells and tissues, and causes attraction of further immune cells.
Considering the aforementioned data, the most prominent candidates for this failure are the loss of control over the sequestration of transition metal ions, exhaustion of haptoglobin and hemopexin, enhanced activity of elastase, the disastrous action of angiotensin II, and the disturbed balance between MMPs and TIMPs.

4.2. Diminished Control over Transition Metal Ions

Ferrous and cupric ions are several-fold involved in damaging reactions of biological constituents. In stressed mitochondria, enhanced formation of O2•− favors the release of Fe2+ from proteins with [4Fe-4S]2+ clusters and thus contributes to mitochondrial dysfunction and apoptosis induction [240,241]. Stress situations are also responsible for the increase in iron ions in biological fluids and cytoplasm [242,243]. These free ions can result from damaged biological material, necrotic cells, heme destruction, release from ferritin, release from the labile iron pool, and overload of protective systems with transition metal ions.
In the reduced state, transition metal ions catalyze the oxidation of hydrogen peroxide and organic hydroperoxides. As a result of these reactions, highly reactive hydroxyl radicals and different substrate radicals are generated [65,66]. In oxidized lipids, alkoxyl radical species are formed from lipid hydroperoxides by Fe2+. The latter reaction promotes further destructive actions in lipid phases [26,244,245].
Ferroptosis is a special form of programmed cell death that is caused by enhanced values of free iron ions and lipid hydroperoxides [246,247]. In addition to an increased concentration of free iron ions, ferroptosis is promoted by a disturbance in glutathione supply and diminished activity of GPX4. Glutathione is the cofactor for GPX4, which is able to remove lipid hydroperoxides within biological membranes [248,249].
Excessive accumulation of copper ions takes place in patients with Morbus Wilson. This condition is associated with destructive reactions in liver, brain, and other organs initiated by the interaction of copper ions with hydroperoxides [250]. Enhanced values of free copper not bound to ceruloplasmin apparently contribute to the pathogenesis of Alzheimer’s disease [251,252].
Highly reactive hydroxyl radicals can also be generated as a result of water radiolysis induced by X-ray or radioactive irradiation [33,253].

4.3. Haptoglobin and Hemopexin Exhaustion

Exhaustion of haptoglobin and hemopexin promotes the disastrous action of free heme. Both severe intravascular hemolysis of red blood cells and rhabdomyolysis of muscle cells contribute to a decline in these protective proteins.
Enhanced intravascular hemolysis is reported for several diseases such as thalassemia, glucose-6-phosphate dehydrogenase deficiency, malaria, paroxysmal nocturnal hemoglobinuria, hereditary spherocytosis, and some others [117,254,255,256]. Increased hemoglobin release from red blood cells is also favored by osmotic stress, sheer stress, lytic poisons, secreted components from Gram-positive bacteria, chirurgical actions on the cardiovascular system, autoantibodies, oxidative processes in membranes of red blood cells, burn-associated necrosis, hemorrhagic conditions, and storage of blood for transfusion [40,117,119,126]. Increased rhabdomyolysis is observed after intensive muscle exercise, traumata, alcohol and drug abuses, the use of certain medications, electrical injury, heat stroke, prolonged immobilization, and as a result of some infections [257,258].
Decline of plasma haptoglobin is regarded as a marker of intravascular hemolysis [259,260]. In hemolytic diseases, a decrease in hemopexin levels follows haptoglobin depletion [261].
A massive release of hemoglobin from red blood cells, e. g. during malaria [262], or myoglobin from traumatic muscles [263,264] can induce acute kidney injury by several mechanisms. Although tubular heme oxygenase is able to detoxify some amount of free hemoglobin and free myoglobin, this enzyme exerts pro-oxidative and damage-promoting activities at a higher load of these heme proteins [265]. Further damage of tubular cells results from free heme by inducing proteasome inhibition, accumulation of misfolded proteins, and favoring the unfolded protein response [133,134,266,267,268].

4.4. Inactivation of Antiproteases

At inflammatory sites, elastase and other serine proteases are released from activated neutrophils. Although different antiproteases limit the activity of elastase by the formation of inactive complexes, elastase can promote long-lasting degradation of extracellular matrix components under conditions of oxidative stress. The latter situation favors the oxidation of critical residues in antiproteases with inactivation of these proteins.
The disastrous action of elastase and the failure of antagonizing principles are discussed in chronic obstructive pulmonary disease (COPD) and other lung diseases [269,270,271,272,273,274,275]. In different lung diseases, elastase affects mucus production and causes mucus hyperplasia [276]. Importantly, the activity of surface-bound neutrophil elastase correlates with parameters of diminished airflow and hyperinflation in lungs [197]. A diminished level of SLPI favors the development of emphysema and fibrosis in the lung [277]. Cathepsin G also plays a role in the pathogenesis of COPD and cystic fibrosis [278].
In hereditary A1AT deficiency, the circulating level of A1AT is decreased and represents a risk factor for the development of COPD and emphysema [279]. This hereditary deficiency of A1AT also promotes fibrosis in liver tissue and the formation of liver cirrhosis.

4.5. Disastrous Action of Angiotensin II

A shifted balance between ACE and ACE2 towards ACE promotes angiotensin II effects on the cardiovascular system such as vasoconstriction, hypertension, and cardiac hypertrophy [280]. Angiotensin II is involved in damage to the respiratory system and contributes to acute lung injury and acute respiratory distress syndrome [280].
Importantly, ACE2 is the receptor for severe acute respiratory syndrome (SARS) viruses [281]. Upon infection with SARS virus, the expression of ACE2 receptors is markedly downregulated in lungs [282]. The inhibition of ACE2 by SARS-COVID-19 also markedly prolongs the fatal action of angiotensin II on lung tissues [283,284]. In COVID-19 patients, antibodies against angiotensin II were found [285]. Downregulation of ACE in COVID-19 also affects bradykinin metabolism and elevates bradykinin level [286].

4.6. Disturbed Balance between MMPs and TIMPs

At inflammatory sites, MMPs promote the cleavage of collagen and elastin and can thus impair the stability of blood vessel walls. Hence, instable atherosclerotic plaques, thrombotic events, and aortic aneurysms can result [287,288]. In the formation of aneurysms, inflammatory cells infiltrate into an injured vessel wall. Different MMPs, especially MMP2 and MMP9, contribute together with reactive species, neutrophil elastase, and angiotensin II to the pathogenesis of aneurysms [289,290,291,292]. An increased ratio of MMP to TIMP expression was found in aneurysmal aortic specimens [293,294].
In repair processes during the termination of inflammation, the right balance between TIMPs and MMPs is highly important for recovery to normal tissue homeostasis [295,296,297]. Excess accumulation of ECM constituents leads to scar formation and the development of fibrosis in many organs. The major constituent of scars is collagen. Organ fibrosis is often followed by organ failure [298,299]. Generally, these deviations are caused by a limited activity of MMPs and prolonged activation by TGF-ß cytokines. Several pathophysiological factors are discussed to contribute to fibrosis, namely conserved PAMPs from pathogens [300,301], uncontrolled TGF-β signaling [302], T-cell derived cytokines [303], autoantibodies [304], and the action of angiotensin II [305,306].
The pathophysiological consequences of a disturbed equilibrium between MMPs and TIMPs are schematically presented in Figure 9.

4.7. Cytotoxic Agents in Tumorigenesis

In order to survive, many types of tumors manipulate their microenvironment in such a way that immune cells are unable to eliminate these degenerated cells. For example, tumor-associated macrophages are driven by lactate accumulated in cancer cells into an anti-inflammatory M2 subtype [306]. Lactate and extracellular acidosis suppress antitumor immunity and promote angiogenesis and tumor progression [307,308]. Lactate-induced release of hyaluronan by adjacent fibroblasts supports tumor growth, invasion, and metastasis [309]. Moreover, tumor cells can secrete high levels of TGF-β, which dampens the activity of many types of immune cells [310,311].
Observations of tumor patients and experimental animals confirm the enhanced formation and release of cytotoxic agents in affected tissue regions. Dysfunctional mitochondria are described in many types of cancers [312]. Oxidative stress can promote the release of iron from 4Fe-4S clusters of mitochondrial proteins such as aconitase and can produce agents damaging mitochondrial DNA [311]. In mice with Lewis lung carcinoma, functional degeneration of mitochondria already occurs at an early disease state [313,314].
In patients with pancreatic cancer and other cancer types, the circulating level of TIMP-1 is up-regulated and associated with a poor clinical outcome [315,316]. TIMP-1 activates via binding to CD63 hepatic stellate cells and thus creates a pre-metastatic niche in the liver [317].
Intratumoral hemorrhages and hemoglobin level in tumors are also associated with a poor clinical prognosis for affected individuals [318,319,320]. In these hemorrhages, hemoglobin can be released from red blood cells and the formation of free heme is likely. Indeed, free heme contributes to the progression of prostate cancer by controlling the expression of key target genes via docking to guanine-rich (G4) elements [321]. In the blood of prostate cancer patients, an inverse correlation between the levels of free heme and hemopexin was observed [321].
Elastase released from human neutrophils efficiently kills a wide range of cancer cells in contrast to elastase from murine neutrophils [153]. In neutrophils of mice, unlike human neutrophils, elastase is co-released with the antiprotease SLPI, which dampens elastase’s effects [153]. In the tumor microenvironment, the ability of elastase to kill cancer cells can be abrogated by the release of serine protease inhibitors and suppressing effects on immune cells [322,323].
Otherwise, neutrophil elastase is able to drive tumorigenesis as shown in preclinical studies using elastase knockout mice and pharmacological inhibition [324]. Intriguingly, the incubation time of cancer cells with elastase affects the functional results. While short incubation (1 h) promotes cellular proliferation, prolonged treatment (6–24 h) induces death of cancer cells [325,326].
In 4T1 and CT26 syngeneic mouse tumor models, angiotensin II promotes the formation of an immunosuppressive microenvironment [327]. Overexpression of the AT1 receptor in tumors is associated with more aggressive tumor features and a poor prognosis [328,329].

4.8. Diminished Protection during Sepsis

Sepsis and septic shock are very dangerous clinical conditions that can be associated with multiple organ failure and lethal outcome. Immunocompromised persons such as injured, diseased, and elderly people and newborns are most frequently prone to the development of sepsis. On the basis of impaired host immunity, opportunistic microbes, fungi, and latent viruses can be activated during sepsis [330,331,332,333,334,335,336]. These pathogens further deteriorate the health status of patients. Thus, in sepsis, host-derived cytotoxic agents can contribute together with pathogen-derived cytotoxic agents to cell and tissue damage.
Concerning tissue damage in septic patients, no consensus exists about the predominating damaging mechanisms as the actual status of individual protective systems can vary widely from patient to patient [337,338]. Numerous reports exist about the accumulation of immune cells, in particular neutrophils, in patients with sepsis and septic shock. Neutrophils from septic patients exhibit delayed apoptosis and diminished chemotactic mobility [339,340]. These cells can be activated far from infection sites [341,342]. In septic shock patients, the percentage of immature neutrophils increases. This parameter is associated with a higher risk of lethal outcome after septic shock [340].
In septic patients, enhanced values were reported for neutrophil products such as elastase [343], myeloperoxidase [344,345,346], and neutrophil extracellular traps [347,348,349]. Moreover, neutrophil-derived cytotoxic agents, enhanced formation of free heme [350,351,352], dysregulation in the sequestration of transition metal ions [353,354], and dysfunctional mitochondria [355,356,357] were also reported to contribute to damaging reactions in sepsis.
In sepsis, it is very challenging to predict which protective mechanism will be exhausted first. Several factors contribute to this uncertainty, such as the local protective status in the affected organs, the energy and immune status of patients, existing comorbidities, and genetic predisposition [358]. A personalized analysis of the status of protective systems is mandatory for individual therapeutic approaches.

5. Conclusions

A functioning immune system is mandatory for long-term surveillance of our organism by removing any harm that can disturb the integrity of cells and tissues. Otherwise, the activation of immune cells is associated with the release of aggressive, cell-damaging agents. These host-derived cytotoxic agents are an essential part of the immune response. The interplay between these agents and ready-to-use antagonizing principles determines the further fate of the inflammatory process.
Multiple mechanisms contribute to cell and tissue damage and thus to the development of chronic inflammatory disease states. The permanent damage of unperturbed cells by cytotoxic agents causes again and again the release of alarmins and antigens and the recruitment of novel immune cells. An important pathophysiological aspect in this vicious circle of initiation of immune responses and destruction of biological material is the decline, exhaustion, or inactivation of ready-to-use acting protective mechanisms in the affected tissues. Further complications result from dysregulated synthesis of ECM components, the existence of long-lasting immunocompromised conditions, and the colonization of commensal and mutualistic pathogens in inflamed tissue areas.
Although protective principles can act very efficiently against the corresponding damaging agent, their capacity is limited. As long as damaging agents are only weakly expressed, efficient protection is given, additional protective systems can be induced, and supplementation of consumed systems is working. Problems arise under conditions of severe and long-lasting action of damaging agents. Then, protection is diminished step by step and biological material is progressively damaged.
Both enhancement of oxidative modifications in biological material and increased proteolytic cleavage of substrates are central events in cell and tissue damage initiated by host-derived cytotoxic agents. In the pathogenesis of chronic inflammatory diseases, several types of host-derived cytotoxic agents often act in concert and promote each other. Enhanced formation of reactive species not only affects the release of transition metal ions and heme components, which have a high potential to further promote oxidative reactions, but also contributes to increased proteolytic damage of ECM and other constituents by inactivation of antiproteases, activation of pro-inflammatory peptides, and initiation of disturbances in the synthesis of novel matrix components. Although some links have been shown in activation and inactivation pathways between different cytotoxic agents and their antagonizing principles, we are far from a thorough understanding of the underlying molecular processes.
It also remains very challenging to determine which cytotoxic agents predominate in damaging reactions in a certain pathology. Moreover, individual expression of protective mechanisms against host-derived cytotoxic agents varies considerably from patient to patient. These uncertainties substantially impede the implementation of personalized therapies based on substitution of inefficient counter-regulating principles for patients with chronic inflammatory diseases. Their successful implementation requires not only the development of novel therapeutic approaches, but also progress in the diagnosis of the individual status of protective mechanisms.

Funding

This research received no external funding.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article.

Conflicts of Interest

The author declares no conflict of interests.

References

  1. Matzinger, P. Tolerance, danger, and the extended family. Annu. Rev. Immunol. 1994, 12, 991–1045. [Google Scholar] [CrossRef] [PubMed]
  2. Janeway, C.A.; Medzhitov, R. Innate immune recognition. Annu. Rev. Immunol. 2002, 20, 197–216. [Google Scholar] [CrossRef] [PubMed]
  3. Takeuchi, O.; Akira, S. Pattern recognition receptors and inflammation. Cell 2010, 140, 805–820. [Google Scholar] [CrossRef] [PubMed]
  4. Thompson, M.R.; Kaminski, J.J.; Kurt-Jones, E.A.; Fitzgerald, K.A. Pattern recognition receptors and the innate immune response to viral infection. Viruses 2011, 3, 920–940. [Google Scholar] [CrossRef]
  5. Suresh, R.; Moser, D.M. Pattern recognition in innate immunity, host defense, and immunopathology. Adv. Physiol. Educ. 2013, 37, 284–291. [Google Scholar] [CrossRef]
  6. Gewurz, H.; Mold, C.; Siegel, J.; Fiedel, B. C-Reactive protein and the acute phase response. Adv. Intern. Med. 1982, 27, 345–372. [Google Scholar] [CrossRef]
  7. Pepys, M.B.; Baltz, M.I. Acute phase proteins with special reference to C-reactive protein and related proteins (pentraxins) and serum amyloid A protein. Adv. Immunol. 1983, 34, 141–212. [Google Scholar] [CrossRef]
  8. Vandivier, R.W.; Henson, P.M.; Douglas, I.S. Burying the death: The impact of failed apoptotic cell removal (efferocytosis) on chronic inflammatory lung disease. Chest 2006, 129, 1673–1682. [Google Scholar] [CrossRef]
  9. Pober, J.S.; Sessa, W.C. Inflammation and the blood microvascular system. Cold Spring Harbor Perspect. Biol. 2015, 7, 016345. [Google Scholar] [CrossRef]
  10. Arnhold, J. Immune response and tissue damage. In Cell and Tissue Destruction. Mechanisms, Protection, Disorders; Academic Press: London, UK; San Diego, CA, USA; Cambridge, MA, USA; Oxford, UK, 2020; pp. 155–204. [Google Scholar] [CrossRef]
  11. Arnhold, J. Acute-phase proteins and additional protective systems. In Cell and Tissue Destruction. Mechanisms, Protection, Disorders; Academic Press: London, UK; San Diego, CA, USA; Cambridge, MA, USA; Oxford, UK, 2020; pp. 205–228. [Google Scholar] [CrossRef]
  12. Mosser, D.M.; Edwards, J.P. Exploring the full spectrum of macrophage activation. Nat. Rev. Immunol. 2008, 8, 958–969. [Google Scholar] [CrossRef]
  13. Yoshimura, A.; Wakabayashi, Y.; Mori, T. Cellular and molecular basis for the regulation of inflammation by TGF-β. J. Biochem. 2010, 147, 781–792. [Google Scholar] [CrossRef] [PubMed]
  14. Oishi, Y.; Manabe, I. Macrophages in inflammation, repair and regeneration. Int. Immunol. 2018, 30, 511–528. [Google Scholar] [CrossRef]
  15. Li, M.O.; Flavell, R.A. Contextual regulation of inflammation: A duet by transforming growth factor-beta and interleukin-10. Immunity 2008, 28, 468–476. [Google Scholar] [CrossRef] [PubMed]
  16. Arnhold, J. Cells and organisms as open systems. In Cell and Tissue Destruction. Mechanisms, Protection, Disorders; Academic Press: London, UK; San Diego, CA, USA; Cambridge, MA, USA; Oxford, UK, 2020; pp. 3–22. [Google Scholar] [CrossRef]
  17. Arnhold, J. The dual role of myeloperoxidase in immune response. Int. J. Mol. Sci. 2020, 21, 8057. [Google Scholar] [CrossRef]
  18. Li, M.O.; Wan, Y.Y.; Sanjabi, S.; Robertson, A.K.; Flavell, R.A. Transforming growth factor-β regulation of immune responses. Annu. Rev. Immunol. 2006, 24, 99–146. [Google Scholar] [CrossRef]
  19. Frangogiannis, N.G. Transforming growth factor-β in tissue fibrosis. J. Exp. Med. 2020, 217, e20190103. [Google Scholar] [CrossRef]
  20. Budi, E.H.; Schaub, J.R.; Decaris, M.; Turner, S.; Derynck, R. TGF-β as a driver of fibrosis: Physiological roles and therapeutic opportunities. J. Pathol. 2021, 254, 358–373. [Google Scholar] [CrossRef]
  21. Cadet, J.; Teoule, R. Comparative study of oxidation of nucleic acid components by hydroxyl radicals, singlet oxygen and superoxide anion radicals. Photochem. Photobiol. 1978, 28, 661–667. [Google Scholar] [CrossRef]
  22. Agnez-Lima, L.F.; Melo, J.T.A.; Silva, A.E.; Oliviera, A.H.S.; Timoteo, A.R.S.; Lima-Bessa, K.M.; Martinez, G.R.; Medeiros, M.H.G.; Di Mascio, P.; Galhardo, R.S.; et al. DNA damage by singlet oxygen and cellular protective mechanisms. Mutat. Res. 2012, 751, 15–28. [Google Scholar] [CrossRef] [PubMed]
  23. Di Mascio, P.; Devasagayam, T.P.A.; Kaiser, S.; Sies, H. Carotenoids, tocopherols and thiols as singlet oxygen quenchers. Biochem. Soc. Trans. 1990, 18, 1054–1056. [Google Scholar] [CrossRef] [PubMed]
  24. Conn, P.F.; Schalch, W.; Truscott, T.G. The singlet oxygen and carotenoid interaction. J. Photochem. Photobiol. B Biol. 1991, 11, 41–47. [Google Scholar] [CrossRef] [PubMed]
  25. Di Mascio, P.; Bechara, E.J.H.; Medeiros, H.G.; Briviba, K.; Sies, H. Singlet molecular oxygen production in the reaction of peroxynitrite with hydrogen peroxide. FEBS Lett. 1994, 355, 287–289. [Google Scholar] [CrossRef] [PubMed]
  26. Pryor, W.A. The formation of free radicals and the consequences of their reactions in vivo. Photochem. Photobiol. 1978, 28, 787–801. [Google Scholar] [CrossRef] [PubMed]
  27. Kermani, S.; Ben-Jebria, A.; Ultman, J.S. Kinetics of ozone reaction with uric acid, ascorbic acid, and glutathione at physiologically relevant conditions. Arch. Biochem. Biophys. 2006, 451, 8–16. [Google Scholar] [CrossRef] [PubMed]
  28. Behndig, A.F.; Blomberg, A.; Helleday, R.; Duggan, S.T.; Kelly, F.J.; Mudway, I.S. Antioxidant responses to acute ozone challenge in the healthy human airway. Inhal. Toxicol. 2009, 21, 933–942. [Google Scholar] [CrossRef]
  29. Devlin, R.B.; Folinsbee, L.J.; Biscardi, F.; Hatch, G.; Becker, S.; Madden, M.C.; Robbins, M.; Koren, H.S. Inflammation and cell damage induced by repeated exposure of humans to ozone. Inhal. Toxicol. 1997, 9, 211–234. [Google Scholar] [CrossRef]
  30. Frank, R.; Liu, M.C.; Spannhake, E.W.; Mlynarek, S.; Macri, K.; Weinmann, G.G. Repetitive ozone exposure of young adults. Evidence of persistent small airway dysfunction. Am. J. Respir. Crit. Care Med. 2001, 164, 1257–1260. [Google Scholar] [CrossRef] [Green Version]
  31. Tyrrell, R.M. UVA (320–380 nm) radiation as an oxidative stress. In Oxidative Stress: Oxidants and Antioxidants; Sies, H., Ed.; Academic Press: London, UK; San Diego, CA, USA; Cambridge, MA, USA; Oxford, UK, 1991; pp. 57–83. [Google Scholar]
  32. Korać, R.R.; Khambholja, K.M. Potential of herbs in skin protection from ultraviolet radiation. Pharmacogn. Rev. 2011, 5, 164–173. [Google Scholar] [CrossRef]
  33. Von Sonntag, C.; Schuchmann, H.-P. Pulse radiolysis. Meth. Enzymol. 1994, 233, 3–56. [Google Scholar] [CrossRef]
  34. Biakov, V.M.; Stepanov, S.V. Mechanism of primary radiobiologic action. Radiat. Biol. Radioecol. 1997, 37, 469–474. [Google Scholar]
  35. Nair, C.K.K.; Parida, D.K.; Nomura, T. Radioprotectors in radiotherapy. J. Radiat. Res. 2001, 42, 21–37. [Google Scholar] [CrossRef]
  36. Arnhold, J. Role of reactive species in destructions. In Cell and Tissue Destruction. Mechanisms, Protection, Disorders; Academic Press: London, UK; San Diego, CA, USA; Cambridge, MA, USA; Oxford, UK, 2020; pp. 23–54. [Google Scholar]
  37. Segal, A.W.; Jones, O.T.G. Novel cytochrome b system in phagocytic vacuoles from human granulocytes. Nature 1976, 276, 515–517. [Google Scholar] [CrossRef]
  38. Babior, B.M. NADPH oxidase. Curr. Opin. Immunol. 2004, 16, 42–47. [Google Scholar] [CrossRef]
  39. Cheng, G.; Cao, Z.; Xu, X.; van Meir, E.G.; Lambeth, J.D. Homologs of gp91phox: Cloning and tissue expression of Nox3, Nox4, and Nox5. Gene 2001, 269, 131–140. [Google Scholar] [CrossRef]
  40. Geiszt, M.; Witta, J.; Baffi, J.; Lekstrom, K.; Leto, T.L. Dual oxidases represent novel hydrogen peroxide sources supporting mucosal host defense. FASEB J. 2003, 17, 1502–1504. [Google Scholar] [CrossRef]
  41. Harper, R.W.; Xu, C.; Eiserich, J.P.; Chen, Y.; Kao, C.Y.; Thai, P.; Setiadi, H.; Wu, R. Differential regulation of dual NADPH/peroxidases. Duox1 and Duox2, by Th1 and Th2 cytokines in the respiratory tract epithelium. FEBS Lett. 2005, 579, 4911–4917. [Google Scholar] [CrossRef]
  42. Olson, J.S.; Ballou, D.P.; Palmer, G.; Massey, V. The reaction of xanthine oxidase with molecular oxygen. J. Biol. Chem. 1974, 249, 4350–4362. [Google Scholar] [CrossRef]
  43. Anderson, R.F.; Hille, R.; Massey, V. The radical chemistry of milk xanthine oxidase as studies by radiation chemistry technique. J. Biol. Chem. 1986, 261, 15870–15876. [Google Scholar] [CrossRef]
  44. Carrell, R.W.; Winterbourn, C.C.; Rachmilewitz, E.A. Activated oxygen and hemolysis. Br. J. Hematol. 1975, 30, 259–264. [Google Scholar] [CrossRef]
  45. Harel, S.; Kanner, J. Hemoglobin and myoglobin as inhibitors of hydroxyl radical generation in a model system of “iron redox” cycle. Free Radic. Res. Commun. 1989, 6, 1–10. [Google Scholar] [CrossRef]
  46. Land, E.J.; Mukherfee, T.; Swallow, A.J.; Bruce, J.M. One-electron reduction of adriamycin: Properties of the semiquinone. Arch. Biochem. Biophys. 1983, 225, 116–121. [Google Scholar] [CrossRef]
  47. Camhi, S.L.; Lee, P.; Choi, A.M.K. The oxidative stress response. New Horizons 1995, 3, 170–182. [Google Scholar] [PubMed]
  48. Goncalves, R.L.; Quinlan, C.L.; Perevoshchikova, I.V.; Hey-Mogensen, M.; Brand, M.D. Site of superoxide and hydrogen peroxide production by muscle mitochondria assessed ex vivo under conditions mimicking rest and exercise. J. Biol. Chem. 2015, 290, 209–227. [Google Scholar] [CrossRef]
  49. Brand, M.D. Mitochondrial generation of superoxide and hydrogen peroxide as the source of mitochondrial redox signaling. Free Radic. Biol. Med. 2016, 100, 14–31. [Google Scholar] [CrossRef]
  50. Koppenol, W.H. Thermodynamics of reactions involving oxyradicals and hydrogen peroxide. Bioelectrochem. Bioenerg. 1987, 18, 3–11. [Google Scholar] [CrossRef]
  51. Bielski, B.H.J.; Cabelli, D.E.; Arudi, R.L. Reactivity of HO2/O2 radicals in aqueous solution. J. Phys. Chem. Ref. Data 1985, 14, 1041–1100. [Google Scholar] [CrossRef]
  52. Huie, R.E.; Padmaja, S. Reactions of NO and O2. Free Radic. Res. Commun. 1993, 18, 195–199. [Google Scholar] [CrossRef]
  53. Kissner, R.; Nauser, T.; Bugnon, P.; Lye, P.G.; Koppenol, W.H. Formation and properties of peroxynitrite as studied by laser flash photolysis, high-pressure stopped-flow technique, and pulse radiolysis. Chem. Res. Toxicol. 1997, 10, 1285–1292. [Google Scholar] [CrossRef] [PubMed]
  54. Gardner, P.R. Superoxide-driven aconitase Fe-S cycling. Biosci. Rep. 1997, 17, 33–42. [Google Scholar] [CrossRef]
  55. Gardner, P.R. Aconitase: Sensitive target and measure of superoxide. Meth. Enzymol. 2002, 349, 9–23. [Google Scholar] [CrossRef]
  56. McCord, J.; Fridovich, I. Superoxide dismutase: An enzymic function for erythrocuprein (hemocuprein). J. Biol. Chem. 1960, 224, 6049–6055. [Google Scholar]
  57. Chang, L.Y.; Slot, J.W.; Geuze, H.J.; Crapo, J.D. Molecular immunocytochemistry of the CuZn superoxide dismutase in rat hepatocytes. J. Cell Biol. 1988, 107, 2169–2179. [Google Scholar] [CrossRef] [PubMed]
  58. Weisiger, R.A.; Fridovich, I. Mitochondrial superoxide dismutase. Site of synthesis and intramolecular localization. J. Biol. Chem. 1973, 248, 4793–4796. [Google Scholar] [CrossRef] [PubMed]
  59. Antonyuk, S.V.; Strange, R.W.; Marklund, S.L.; Hasnain, S.S. The structure of human extracellular copper-zinc superoxide dismutase at 1.7 Å resolution: Insights into heparin and collagen binding. J. Mol. Biol. 2009, 388, 310–326. [Google Scholar] [CrossRef]
  60. Starkov, A.A. The role of mitochondria in reactive oxygen species metabolism and signaling. Ann. N. Y. Acad. Sci. 2008, 1147, 37–52. [Google Scholar] [CrossRef]
  61. Pasdois, P.; Parker, J.E.; Griffiths, E.J.; Halestrap, A.P. The role of oxidized cytochrome c in regulating mitochondrial reactive species production and its perturbation in ischemia. Biochem. J. 2011, 436, 493–505. [Google Scholar] [CrossRef]
  62. Schrader, M.; Fahimi, H.D. Peroxisomes and oxidative stress. Biochim. Biophys. Acta 2006, 1763, 1755–1766. [Google Scholar] [CrossRef]
  63. Koppenol, W.H.; Butler, J. Mechanisms of reactions involving singlet oxygen and the superoxide anion. FEBS Lett. 1977, 83, 1–6. [Google Scholar] [CrossRef]
  64. Reich, H.J.; Hondal, R.J. Why nature chose selenium. ACS Chem. Biol. 2016, 11, 821–841. [Google Scholar] [CrossRef]
  65. Halliwell, B.; Gutteridge, J.M.C. Iron as a biological pro-oxidant. ISI Atlas Sci. Biochem. 1988, 1, 48–52. [Google Scholar]
  66. Koppenol, W.H. The centennial of the Fenton reaction. Free Radic. Biol. Chem. 1993, 15, 645–651. [Google Scholar] [CrossRef] [PubMed]
  67. Arnhold, J.; Flemmig, J. Human myeloperoxidase in innate and acquired immunity. Arch. Biochem. Biophys. 2010, 500, 92–106. [Google Scholar] [CrossRef]
  68. Flemmig, J.; Gau, J.; Schlorke, D.; Arnhold, J. Lactoperoxidase as potential drug target. Expert Opin. Ther. Targets 2016, 20, 447–461. [Google Scholar] [CrossRef]
  69. Arnhold, J. Heme peroxidases at unperturbed and inflamed mucous surfaces. Antioxidants 2021, 10, 1805. [Google Scholar] [CrossRef]
  70. Arnhold, J.; Malle, E. Halogenation activity of mammalian heme peroxidases. Antioxidants 2022, 11, 890. [Google Scholar] [CrossRef] [PubMed]
  71. Ursini, F.; Maiorino, M.; Roveri, A. Phospholipid hydroperoxide glutathione peroxidase (PHGPx): More than an antioxidant enzyme? Biomed. Environm. Sci. 1997, 10, 327–332. [Google Scholar]
  72. Lubos, E.; Loscalzo, J.; Handy, D.E. Glutathione peroxidase-1 in health and disease: From molecular mechanisms to therapeutic opportunities. Antioxid. Redox Signal. 2011, 15, 1957–1997. [Google Scholar] [CrossRef]
  73. Meister, A. Glutathione metabolism and its selective modification. J. Biol. Chem. 1988, 263, 17205–17208. [Google Scholar] [CrossRef]
  74. Low, F.M.; Hampton, M.P.; Winterbourn, C.C. Prx2 and peroxide metabolism in the erythrocyte. Antioxid. Redox Signal. 2008, 10, 1621–1630. [Google Scholar] [CrossRef]
  75. Goyal, M.M.; Basak, A. Human catalase: Looking for complete identity. Prot. Cell 2010, 1, 888–897. [Google Scholar] [CrossRef]
  76. Gunther, M.R.; Hanna, P.M.; Mason, R.P.; Cohen, M.S. Hydroxyl radical formation from cuprous ion and hydrogen peroxide: A spin-trapping study. Arch. Biochem. Biophys. 1995, 316, 515–522. [Google Scholar] [CrossRef]
  77. Ryan, T.P.; Aust, S.D. The role of iron in oxygen-mediated toxicities. Crit. Rev. Toxicol. 1992, 22, 119–141. [Google Scholar] [CrossRef]
  78. Reinke, L.A.; Rau, J.M.; McCay, P.B. Characteristics of an oxidant formed during iron(II) autoxidation. Free Radic. Biol. Med. 1994, 16, 485–492. [Google Scholar] [CrossRef]
  79. Qian, S.Y.; Buettner, G.R. Iron and dioxygen chemistry is an important route to initiation of biological free radical oxidations: An electron paramagnetic resonance spin trapping study. Free Radic. Biol. Med. 1999, 26, 1447–1456. [Google Scholar] [CrossRef]
  80. Urbanski, N.K.; Beresewicz, A. Generation of OH initiated by interaction of Fe2+ and Cu+ with dioxygen; comparison with the Fenton chemistry. Acta Biochim. Pol. 2000, 47, 951–962. [Google Scholar] [CrossRef] [PubMed]
  81. Flemmig, J.; Arnhold, J. Ferrous ion-induced strand breaks in the DNA plasmid pBR322 are not mediated by hydrogen peroxide. Eur. Biophys. J. 2007, 36, 377–384. [Google Scholar] [CrossRef] [PubMed]
  82. Bors, W.; Erben-Russ, M.; Saran, M. Fatty acid peroxyl radicals: Their generation and reactivities. Bioelectrochem. Bioenerg. 1987, 18, 37–49. [Google Scholar] [CrossRef]
  83. Crichton, R.R.; Ward, R.J. Iron homeostasis. Metal Ions Biol. Syst. 1998, 35, 633–665. [Google Scholar]
  84. Ponka, P. Cellular iron metabolism. Kidney Int. 1999, 55, S2–S11. [Google Scholar] [CrossRef]
  85. Zhao, N.; Zhang, A.-S.; Enns, C.A. Iron regulation by hepcidin. J. Clin. Investig. 2013, 123, 2337–2343. [Google Scholar] [CrossRef]
  86. Gkouvatsos, K.; Papanikolaou, G.; Pantopoulos, K. Regulation of iron transport and the role of transferrin. Biochim. Biophys. Acta 2011, 1820, 188–202. [Google Scholar] [CrossRef] [PubMed]
  87. Gamella, E.; Buratti, P.; Cairo, G.; Recalcati, S. The transferrin receptor: The cellular iron gate. Metallomics 2017, 9, 1367–1375. [Google Scholar] [CrossRef]
  88. Massover, W.H. Ultrastructure of ferritin and apoferritin: A review. Micron 1993, 24, 389–437. [Google Scholar] [CrossRef]
  89. Prohaska, J.R. Role of copper transporters in copper homeostasis. Am. J. Clin. Nutr. 2008, 88, 826S–829S. [Google Scholar] [CrossRef] [PubMed]
  90. Stoj, C.; Kosman, D.J. Cuprous oxidase activity of yeast Fet3p and human ceruloplasmin: Implication for function. FEBS Lett. 2003, 554, 422–426. [Google Scholar] [CrossRef]
  91. Sokolov, A.V.; Pulina, M.O.; Ageeva, K.V.; Runova, O.I.; Zakharova, E.T.; Vasilyev, V.B. Identification of leukocyte cationic proteins that interact with ceruloplasmin. Biochemistry 2007, 72, 872–877. [Google Scholar] [CrossRef]
  92. Radi, R.; Beckman, J.S.; Bush, K.M.; Freeman, B.A. Peroxynitrite-induced membrane lipid peroxidation: The cytotoxic potential of superoxide and nitric oxide. Arch. Biochem. Biophys. 1991, 288, 481–487. [Google Scholar] [CrossRef] [PubMed]
  93. Darley-Usmar, V.M.; Hogg, N.; O’Leary, V.J.; Wilson, M.T.; Moncada, S. The simultaneous generation of superoxide and nitric oxide can initiate lipid peroxidation in human low density lipoprotein. Free Radic. Res. Commun. 1992, 17, 9–20. [Google Scholar] [CrossRef]
  94. Goldstein, S.; Czapski, G. Mechanism of the nitrosation of thiols and amines by oxygenated NO solutions: The nature of the nitrosating intermediates. J. Am. Chem. Soc. 1996, 118, 3419–3425. [Google Scholar] [CrossRef]
  95. Schopfer, F.J.; Baker, P.R.S.; Freeman, B.A. NO-dependent protein nitration: A cell signaling event or an oxidative inflammatory response? Trends Biochem. Sci. 2003, 28, 646–654. [Google Scholar] [CrossRef]
  96. Augusto, O.; Bonini, M.G.; Amanso, A.M.; Linares, E.; Santos, C.C.; de Menezes, S.L. Nitrogen dioxide and carbonate radical anion: Two emerging radicals in biology. Free Radic. Biol. Med. 2002, 32, 841–859. [Google Scholar] [CrossRef] [PubMed]
  97. Ferrer-Sueta, G.; Radi, R. Chemical biology of ONOO-: Kinetics, diffusion, and radicals. ACS Chem. Biol. 2009, 4, 161–177. [Google Scholar] [CrossRef] [PubMed]
  98. Furtmüller, P.G.; Jantschko, W.; Zederbauer, M.; Schwanninger, M.; Jakoptisch, C.; Herold, S.; Koppenol, W.H.; Obinger, C. Peroxynitrite efficiently mediates the interconverion of redox intermediates of myeloperoxidase. Biochem. Biophys. Res. Commun. 2005, 337, 944–954. [Google Scholar] [CrossRef]
  99. Galijasevic, S.; Maitra, D.; Lu, T.; Sliskovic, I.; Abdulhamid, I.; Abu-Soud, H.M. Myeloperoxidase interaction with peroxynitrite: Chloride deficiency and heme depletion. Free Radic. Biol. Med. 2009, 47, 431–439. [Google Scholar] [CrossRef] [PubMed]
  100. Koyani, C.N.; Flemmig, J.; Malle, E.; Arnhold, J. Myeloperoxidase scavenges peroxynitrite: A novel anti-inflammatory action of the heme enzyme. Arch. Biochem. Biophys. 2015, 571, 1–19. [Google Scholar] [CrossRef] [PubMed]
  101. Su, J.; Groves, J.T. Mechanisms of peroxynitrite interactions with heme proteins. Inorg. Chem. 2010, 49, 6317–6329. [Google Scholar] [CrossRef]
  102. Keng, T.; Privalle, C.T.; Gilkeson, G.S.; Weinberg, J.B. Peroxynitrite formation and decreased catalase activity in autoimmune MRL-lpr/lpr mice. Mol. Med. 2000, 6, 779–792. [Google Scholar] [CrossRef]
  103. Gebicka, L.; Gebicki, J.L. Reaction of heme peroxidases with peroxynitrite. IUBMB Life 2000, 49, 11–15. [Google Scholar] [CrossRef]
  104. Klebanoff, S.J. Myeloperoxidase: Friend and foe. J. Leukoc. Biol. 2005, 77, 598–625. [Google Scholar] [CrossRef]
  105. Rothenberg, M.E.; Hogan, S.P. The eosinophil. Annu. Rev. Immunol. 2006, 24, 147–174. [Google Scholar] [CrossRef]
  106. Pattison, D.I.; Davies, M.J. Absolute rate constants for the reaction of hypochlorous acid with protein side chains and peptide bonds. Chem. Res. Toxicol. 2001, 14, 453–464. [Google Scholar] [CrossRef] [PubMed]
  107. Hawkins, C.L.; Pattison, D.I.; Davies, M.J. Hypochlorite-induced oxidation of amino acids, peptides, and proteins. Amino Acids 2003, 25, 259–274. [Google Scholar] [CrossRef] [PubMed]
  108. Pattison, D.I.; Davies, M.J. Kinetic analysis of the reaction of hypobromous acid with protein components: Implication for cellular damage and the use of 3-bromotyrosine as a marker of oxidative stress. Biochemistry 2004, 43, 4799–4809. [Google Scholar] [CrossRef] [PubMed]
  109. Ashby, M.T.; Carlson, A.C.; Scott, M.J. Redox buffering of hypochlorous acid by thiocyanate in physiologic fluids. J. Am. Chem. Soc. 2004, 126, 15976–15977. [Google Scholar] [CrossRef] [PubMed]
  110. Nagy, P.; Beal, J.L.; Ashby, M.T. Thiocyanate is an efficient endogenous scavenger of the phagocytic killing agent hypobromous acid. Chem. Res. Toxicol. 2006, 19, 587–593. [Google Scholar] [CrossRef]
  111. Davies, M.J.; Hawkins, C.L. The role of myeloperoxidase in biomolecule modification, chronic inflammation, and disease. Antioxid. Redox Signal. 2020, 32, 957–981. [Google Scholar] [CrossRef]
  112. Sokolov, A.V.; Ageeva, K.V.; Pulina, M.O.; Cherkalina, O.S.; Samygina, V.R.; Vlasova, I.I.; Panasenko, O.M.; Zakharova, E.T.; Vasilyev, V.B. Ceruloplasmin and myeloperoxidase in complex affect the enzymatic properties of each other. Free Radic. Res. 2008, 42, 989–998. [Google Scholar] [CrossRef]
  113. Chapman, A.L.P.; Mocatta, T.J.; Shiva, S.; Seidel, A.; Chen, B.; Khalilova, I.; Paumann-Page, M.E.; Jameson, G.N.L.; Winterbourn, C.C.; Kettle, A.J. Ceruloplasmin is an endogenous inhibitor of myeloperoxidase. J. Biol. Chem. 2013, 288, 6464–6477. [Google Scholar] [CrossRef]
  114. Samygina, V.R.; Sokolov, A.V.; Bourenkov, G.; Petoukhov, M.V.; Pulina, M.O.; Zakharova, E.T.; Vasilyev, V.B.; Bartunik, H.; Svergun, D.I. Ceruloplasmin: Macromolecular assemblies with iron-containing acute phase proteins. PLoS ONE 2013, 8, e67145. [Google Scholar] [CrossRef]
  115. Sokolov, A.V.; Kostevich, V.A.; Zakharova, E.T.; Samygina, V.R.; Panasenko, O.M.; Vasilyev, V.B. Interaction of ceruloplasmin with eosinophil peroxidase as compared to its interplay with myeloperoxidase: Reciprocal effect on enzymatic properties. Free Radic. Res. 2015, 49, 800–811. [Google Scholar] [CrossRef]
  116. Reiter, C.D.; Wang, X.; Tanus-Santos, J.E.; Hogg, N.; Cannon III, R.O.; Schechter, A.N.; Gladwin, M.T. Cell-free hemoglobin limits nitric oxide bioavailability in sickle-cell disease. Nat. Med. 2002, 8, 1383–1389. [Google Scholar] [CrossRef]
  117. Rother, R.P.; Bell, L.; Hillman, P.; Gladwin, M.T. The clinical sequelae of intravascular hemolysis and extracellular plasma hemoglobin. J. Am. Med. Assoc. 2005, 293, 1653–1662. [Google Scholar] [CrossRef]
  118. Kristiansen, M.; Graversen, J.H.; Jacobsen, C.; Sonne, O.; Hoffmann, H.J.; Law, S.K.; Moestrup, S.K. Identification of the haemoglobin scavenger receptor. Nature 2001, 409, 198–201. [Google Scholar] [CrossRef] [PubMed]
  119. Chiabrando, D.; Vinchi, F.; Fiorito, V.; Tolosano, E. Haptoglobin and hemopexin in heme detoxification and iron recycling. In Acute Phase Proteins- Regulation and Functions of Acute Phase Proteins; Veas, F., Ed.; Intech: Rijeka, Croatia, 2011; pp. 261–288. [Google Scholar] [CrossRef]
  120. Bunn, H.F.; Jandl, J.H. Exchange of heme among hemoglobins and hemoglobin and albumin. J. Biol. Chem. 1968, 243, 465–475. [Google Scholar] [CrossRef] [PubMed]
  121. Balla, G.; Jacob, H.S.; Eaton, J.W.; Belcher, J.D.; Vercelotti, G.M. Hemin: A possible physiological mediator of low density lipoprotein oxidation and endothelial injury. Arterioscler. Thromb. 1991, 11, 1700–1711. [Google Scholar] [CrossRef]
  122. Jeney, V.; Balla, J.; Yachie, A.; Varga, Z.; Vercelotti, G.M.; Eaton, J.W.; Balla, G. Pro-oxidant and cytotoxic effects of circulating heme. Blood 2002, 100, 879–887. [Google Scholar] [CrossRef] [PubMed]
  123. Kumar, S.; Bandyopadhyay, U. Free heme toxicity and its detoxification systems in human. Toxicol. Lett. 2005, 157, 175–188. [Google Scholar] [CrossRef]
  124. Flemmig, J.; Schlorke, D.; Kühne, F.-W.; Arnhold, J. Inhibition of the heme-induced hemolysis of red blood cells by the chlorite-based drug WF10. Free Radic. Res. 2016, 50, 1386–1395. [Google Scholar] [CrossRef]
  125. Lin, T.; Sammy, F.; Yang, H.; Thundivalappil, S.; Hellman, J.; Tracey, K.C.; Warren, H.S. Identification of hemopexin as an anti-inflammatory factor that inhibits synergy of hemoglobin with HMGB1 in sterile and infectious inflammation. J. Immunol. 2012, 189, 2017–2022. [Google Scholar] [CrossRef]
  126. Schaer, D.J.; Buehler, P.W.; Alayash, A.I.; Belcher, J.D.; Vercelotti, G.M. Hemolysis and free hemoglobin revisited: Exploring hemoglobin and hemin scavengers as a novel class of therapeutic proteins. Blood 2013, 121, 1276–1284. [Google Scholar] [CrossRef]
  127. Figueiredo, R.T.; Fernandez, P.L.; Mourao-Sa, D.S.; Porto, B.N.; Dutra, F.F.; Alves, L.S.; Oliviera, M.F.; Oliviera, P.L.; Graca-Souza, A.V.; Bozza, M.T. Characterization of heme as activator of toll-like receptor 4. J. Biol. Chem. 2007, 282, 20221–20229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Belcher, J.D.; Chen, C.; Nguyen, J.; Milbauer, L.; Abdulla, F.; Alayash, A.I.; Smith, A.; Nath, K.A.; Hebbel, R.P.; Vercelotti, G.M. Heme triggers TLR4 signaling leading to endothelial cell activation and vaso-occlusion in murine sickle cell disease. Blood 2014, 123, 377–390. [Google Scholar] [CrossRef]
  129. Poon, L.C.; Methot, S.P.; Morabi-Pazocki, W.; Pio, F.; Bennet, A.J.; Sen, D. Guanine-rich RNAs and DNAs that bind heme robustly catalyze oxygen transfer reactions. J. Am. Chem. Soc. 2011, 133, 1877–1884. [Google Scholar] [CrossRef] [PubMed]
  130. Gray, L.T.; Lombardi, E.P.; Verga, D.; Nicolas, A.; Teulade-Fichou, M.-P.; Londoño-Vallejo, A.; Maizels, N. G-Quadruplexes sequester free heme in living cells. Cell Chem. Biol. 2019, 26, 1681–1689. [Google Scholar] [CrossRef]
  131. Hvidberg, V.; Maniecki, M.B.; Jacobson, C.; Hojrup, P.; Moller, H.J.; Moestrup, S.K. Identification of the receptor scavenging hemopexin-heme complexes. Blood 2005, 106, 2572–2579. [Google Scholar] [CrossRef]
  132. Lin, T.; Maita, D.; Thundivalappil, S.R.; Riley, F.E.; Hambsch, J.; van Marter, L.J.; Christou, H.A.; Berra, L.; Fagan, S.; Christiani, D.C.; et al. Hemopexin in severe inflammation and infection: Mouse models and human diseases. Crit. Care 2015, 19, 166. [Google Scholar] [CrossRef] [PubMed]
  133. Santoro, A.M.; Lo Giudice, M.C.; D’Urso, A.; Lauceri, R.; Purello, R.; Milardi, D. Cationic porphyrins are reversible proteasome inhibitors. J. Am. Chem. Soc. 2012, 134, 10451–10457. [Google Scholar] [CrossRef]
  134. Vallelian, F.; Deuel, J.W.; Opitz, L.; Schaer, C.A.; Puglia, M.; Lönn, M.; Engelsberger, W.; Schauer, S.; Karnaukhova, E.; Spahn, D.R.; et al. Proteasome inhibition and oxidative reactions disrupt cellular homeostasis during heme stress. Cell Death Differ. 2015, 22, 597–611. [Google Scholar] [CrossRef]
  135. Arnhold, J. Oxidation and reduction of biological material. In Cell and Tissue Destruction. Mechanisms, Protection, Disorders; Academic Press: London, UK; San Diego, CA, USA; Cambridge, MA, USA; Oxford, UK, 2020; pp. 55–97. [Google Scholar]
  136. Hayward, L.D.; Angyal, S.J. A symmetry rule for the circular dichroism of reducing sugar, and the proportion of carbonyl forms in aqueous solutions thereof. Carbohydr. Res. 1977, 53, 13–20. [Google Scholar] [CrossRef]
  137. Pamplona, R. Advanced lipoxidation end-products. Chem. Biol. Interact. 2011, 192, 14–20. [Google Scholar] [CrossRef]
  138. Vistoli, G.; de Maddis, D.; Cipak, A.; Zarkovic, N.; Carini, M.; Aldini, G. Advanced glycoxidation and lipoxidation end products (AGEs amd ALEs): An overview of their mechanisms of formation. Free Radic. Res. 2013, 47 (Suppl. 1), 3–27. [Google Scholar] [CrossRef] [Green Version]
  139. Burton, G.W.; Joyce, A.; Ingold, K.U. Is vitamin E the only lipid-soluble, chain-breaking antioxidant in human blood plasma and erythrocyte membrane? Arch. Biochem. Biophys. 1983, 221, 281–290. [Google Scholar] [CrossRef]
  140. Buettner, G.R. The pecking order of free radicals and antioxidants: Lipid peroxidation, α-tocopherol, and ascorbate. Arch. Biochem. Biophys. 1993, 300, 535–543. [Google Scholar] [CrossRef]
  141. Niki, E. Antioxidants in relation to lipid peroxidation. Chem. Phys. Lipids 1987, 44, 227–253. [Google Scholar] [CrossRef]
  142. Van Kujik, F.J.G.M.; Sevanian, A.; Handelman, G.J.; Dratz, E.A. A new role for phospholipase A2: Protection of membranes from lipid peroxidation damage. Trends Biochem. Sci. 1987, 12, 31–34. [Google Scholar] [CrossRef]
  143. Hochstein, P.; Hatch, L.; Sevanian, A. Uric acid: Functions and determinations. Methods Enzymol. 1984, 105, 162–166. [Google Scholar]
  144. Nyyssönen, K.; Porkkala-Sarataho, E.; Kaikkonen, J.; Salonen, J.T. Ascorbate and urate are the strongest determinants of plasma antioxidative capacity and serum lipid resistance to oxidation in Finnish men. Atherosclerosis 1997, 130, 223–233. [Google Scholar] [CrossRef]
  145. Spencer, J.P.E.; Abd El Mohsen, M.M.; Minihane, A.-M.; Mathers, J.C. Biomarkers of the intake of dietary polyphenols: Strengths, limitations and application in nutrition research. Br. J. Nutr. 2008, 99, 12–22. [Google Scholar] [CrossRef]
  146. Steenken, S.; Neta, P. One-electron redox potentials of phenols. Hydroxy- and aminophenols and related compounds of biological interest. J. Phys. Chem. 1982, 86, 3661–3667. [Google Scholar] [CrossRef]
  147. Perron, N.R.; Brumaghim, J.L. A review of the antioxidant mechanisms of polyphenol compounds related to iron binding. Cell Biochem. Biophys. 2009, 53, 75–100. [Google Scholar] [CrossRef]
  148. Gau, J.; Furtmüller, P.G.; Obinger, C.; Prévost, M.; van Antwerpen, P.; Arnhold, J.; Flemmig, J. Flavonoids as promoters of the (pseudo-)halogenating activity of lactoperoxidase and myeloperoxidase. Free Radic. Biol. Med. 2016, 97, 307–319. [Google Scholar] [CrossRef]
  149. Wardlaw, A.C. The complement-dependent bacteriolytic activity of normal human serum. I. The effect of pH and ionic strength and the role of lysozyme. J. Exp. Med. 1962, 115, 1231–1249. [Google Scholar] [CrossRef]
  150. Korkmaz, B.; Horwitz, M.S.; Jenne, D.E.; Gauthier, F. Neutrophil elastase, proteinase 3, and cathepsin G as therapeutic targets in human diseases. Pharmacol. Rev. 2010, 62, 726–759. [Google Scholar] [CrossRef] [PubMed]
  151. Weinrauch, Y.; Drujan, D.; Shapiro, S.D.; Weiss, J.; Zychlinsky, A. Neutrophil elastase targets virulence factors of enterobacteria. Nature 2002, 417, 91–94. [Google Scholar] [CrossRef]
  152. Belaaouaj, A. Neutrophil elastase-mediated killing of bacteria: Lessons from targeted mutagenesis. Microb. Infect. 2002, 4, 1259–1264. [Google Scholar] [CrossRef] [PubMed]
  153. Cui, C.; Chakraborty, K.; Tang, X.A.; Zhou, G.; Schoenfelt, K.Q.; Becker, K.M.; Hoffmann, A.; Chang, Y.-F.; Blank, A.; Reardon, C.A.; et al. Neutrophil elastase selectively kills cancer cells and attenuates tumorigenesis. Cell 2021, 184, 3163–3177. [Google Scholar] [CrossRef] [PubMed]
  154. Fu, Z.; Akula, S.; Thorpe, M.; Hellman, L. Potent and broad but not unselectively cleavage of cytokines and chemokines by human neutrophil elastase and proteinase 3. Int. J. Mol. Sci. 2020, 21, 651. [Google Scholar] [CrossRef]
  155. Fu, Z.; Thorpe, M.; Alemayehu, R.; Roy, A.; Kervinen, J.; de Garavilla, L.; Abrink, M.; Hellman, L. Highly selective cleavage of cytokines and chemokines by the human mast cell chymase and neutrophil cathepsin G. J. Immunol. 2017, 198, 1474–1483. [Google Scholar] [CrossRef]
  156. Brinkmann, V.; Reichard, U.; Goosmann, C.; Fauler, B.; Uhlemann, Y.; Weiss, D.S.; Weinrauch, Y.; Zychlinsky, A. Neutrophil extracellular traps kill bacteria. Science 2004, 303, 1532–1535. [Google Scholar] [CrossRef]
  157. Papayannopoulos, V.; Zychlinsky, A. NETs: A new strategy for using old weapons. Trends Immunol. 2009, 30, 513–521. [Google Scholar] [CrossRef]
  158. Caughey, G.H. Mast cell tryptases and chymases in inflammation and host defense. Immunol. Rev. 2007, 217, 141–151. [Google Scholar] [CrossRef]
  159. Korkmaz, B.; Moreau, T.; Gauthier, F. Neutrophil elastase, proteinase 3 and cathepsin G: Physicochemical properties, activity and physiopathological functions. Biochimie 2008, 90, 227–242. [Google Scholar] [CrossRef]
  160. Ramaha, A.; Patston, P.A. Release and degradation of angiotensin I and II from angiotensinogen by neutrophil serine proteases. Arch. Biochem. Biophys. 2002, 397, 77–83. [Google Scholar] [CrossRef]
  161. Vidotti, D.B.; Casarini, D.E.; Christovam, P.C.; Leite, C.A.; Schor, N.; Boim, M.A. High glucose concentration stimulates intracellular renin activity and angiotensin II generation in rat mesangial cells. Am. J. Physiol. Renal Physiol. 2004, 286, F1039–F1045. [Google Scholar] [CrossRef]
  162. Penafuerte, C.A.; Gagnon, B.; Sirois, J.; Murphy, J.; MacDonald, N.; Tremblay, M.L. Identification of neutrophil-derived proteases and angiotensin II as biomarkers of cancer cachexia. Br. J. Cancer 2016, 114, 680–687. [Google Scholar] [CrossRef]
  163. Okada, Y.; Nakanishi, I. Activation of matrix metalloproteinase 3 (stromelysin) and matrix metalloproteinase 2 (‘gelatinase’) by human neutrophil elastase and cathepsin G. FEBS Lett. 1989, 249, 353–356. [Google Scholar] [CrossRef]
  164. Shamanian, P.; Schwartz, J.D.; Pocock, B.J.Z.; Monea, S.; Whiting, D.; Marcus, S.G.; Mignatti, P. Activation of progelatinase (MMP-2) by neutrophil elastase, cathepsin G, and proteinase-3: A role for inflammatory cells in tumor invasion and angiogenesis. J. Cell. Physiol. 2001, 189, 197–206. [Google Scholar] [CrossRef]
  165. Gaggar, A.; Li, Y.; Weathington, N.; Winkler, M.; Kong, M.; Jackson, P.; Blalock, J.E.; Clancy, J.P. Matrix metalloprotease-9 dysregulation in lower airway secretions of cystic fibrosis patients. Am. J. Physiol. Lung Cell Mol. Physiol. 2007, 293, L96–L104. [Google Scholar] [CrossRef]
  166. Garratt, L.W.; Sutanto, E.N.; Ling, K.-M.; Looi, K.; Iosifidis, T.; Martinovich, K.M.; Shaw, N.C.; Kicic-Starcevich, E.; Knight, D.A.; Ranganathan, S.; et al. on behalf of the Australian respiratory early surveillance team for cystic fibrosis. Matrix metalloproteinase activation by free neutrophil elastase contributes to bronchiectasis progression in early cystic fibrosis. Eur. Respir. J. 2015, 46, 384–394. [Google Scholar] [CrossRef]
  167. Jackson, P.I.; Xu, X.; Wilson, L.; Weathington, N.M.; Clancy, J.P.; Blalock, J.E.; Gaggar, A. Human neutrophil elastase-mediated cleavage sites of MMP-9 and TIMP-1: Implications to cystic fibrosis proteolytic dysfunction. Mol. Med. 2010, 16, 159–166. [Google Scholar] [CrossRef]
  168. Cairns, J.A.; Walls, A.F. Mast cell tryptase is a mitogen for epithelial cells. Stimulation of IL-8 production and intercellular adhesion molecule-1 expression. J. Immunol. 1996, 156, 275–283. [Google Scholar] [CrossRef]
  169. Dougherty, R.H.; Sidhu, S.S.; Raman, K.; Solon, M.; Solberg, O.D.; Caughey, G.H.; Woodruff, P.G.; Fahy, J.V. Accumulation of intraepithelial mast cells with a unique protease phenotype in T(H)2-high asthma. J. Allergy Clin. Immunol. 2010, 125, 1046–1053. [Google Scholar] [CrossRef] [Green Version]
  170. Ramu, S.; Akbarshshi, H.; Mogren, S.; Berlin, F.; Cerps, S.; Menzel, M.; Hvidtfeldt, M.; Porsbjerg, C.; Uller, L.; Andersson, C.K. Direct effects of mast cell proteases, tryptase and chymase, on bronchial epithelial integrity proteins and anti-viral responses. BMC Immunol. 2021, 22, 35. [Google Scholar] [CrossRef]
  171. Karimi, N.; Morovati, S.; Chan, L.; Napoleoni, C.; Mehrani, Y.; Bridle, B.W.; Karimi, K. Mast cell tryptase and implications for SARS-CoV-2 pathogenesis. Bio. Med. 2021, 1, 136–149. [Google Scholar] [CrossRef]
  172. He, A.; Shi, G.-P. Mast cell chymase and tryptase as targets for cardiovascular and metabolic diseases. Curr. Pharm. Des. 2013, 19, 1114–1125. [Google Scholar] [CrossRef]
  173. Gruber, B.L.; Marchese, M.J.; Suzuki, K.; Schwartz, L.B.; Okada, Y.; Nagae, H.; Ramamurthy, N.S. Synovial procollagenase activation by human mast cell tryptase dependence upon matrix metalloproteinase 3 activation. J. Clin. Investig. 1989, 84, 1657–1662. [Google Scholar] [CrossRef]
  174. Lees, M.; Taylor, D.J.; Woolley, D.E. Mast cell proteinases activate precursor forms of collagenase and stromelysin, but not of gelatinases A and B. Eur. J. Biochem. 1994, 223, 171–177. [Google Scholar] [CrossRef]
  175. Yamamoto, K.; Kumagai, N.; Fukuda, K.; Fujitsu, Y.; Nishida, T. Activation of corneal fibroblast-derived matrix metalloproteinase-2 by tryptase. Curr. Eye Res. 2006, 31, 313–317. [Google Scholar] [CrossRef]
  176. Pyo, R.; Lee, J.K.; Shipley, J.M.; Curci, J.A.; Mao, D.; Ziporin, S.J.; Ennis, T.L.; Shapiro, S.D.; Senior, R.M.; Thompson, R.W. Targeted gene disruption of matrix metalloproteinase-9 (gelatinase B) suppresses development of experimental abdominal aortic aneurysms. J. Clin. Investig. 2000, 105, 1641–1649. [Google Scholar] [CrossRef]
  177. Galis, Z.S.M.; Johnson, C.; Godin, D.; Magid, R.; Shipley, J.M.; Senior, R.M.; Ivan, E. Targeted disruption of the matrix metalloproteinase-9 gene impairs smooth muscle cell migration and geometrical arterial remodeling. Circ. Res. 2002, 91, 852–859. [Google Scholar] [CrossRef]
  178. Longo, G.M.; Xiong, W.; Greiner, T.C.; Zhao, Y.; Fiotti, N.; Baxter, B.T. Matrix metalloproteinases 2 and 9 work in concert to produce aortic aneurysms. J. Clin. Investig. 2002, 110, 625–632. [Google Scholar] [CrossRef] [PubMed]
  179. Kuzuya, M.; Nakamura, K.; Sasaki, T.; Cheng, X.W.; Itohara, S.; Iguchi, A. Effect of MMP-2 deficiency on atherosclerotic lesion formation in apoE-deficient mice. Arterioscler. Thromb. Vasc. Biol. 2006, 26, 1120–1125. [Google Scholar] [CrossRef] [Green Version]
  180. Wågsäter, D.; Zhu, C.; Björkegren, J.; Skogsberg, J.; Eriksson, P. MMP-2 and MMP-9 are prominent matrix metalloproteinases during atherosclerosis development in the Ldlr(−/−)Apob(100/100) mouse. Int. J. Mol. Med. 2011, 28, 247–253. [Google Scholar] [CrossRef] [PubMed]
  181. Berlin, F.; Mogren, S.; Tutzauer, J.; Andersson, C.K. Mast cell proteases tryptase and chymase induce migratory and morphological alterations in bronchial epithelial cells. Int. J. Mol. Sci. 2021, 22, 5250. [Google Scholar] [CrossRef]
  182. De Souza Junior, D.A.; Santana, A.C.; da Silva, E.Z.M.; Oliver, C.; Jamur, M.C. The role of mast cell specific chymases and tryptases in tumor angiogenesis. BioMed Res. Int. 2015, 2015, 142359. [Google Scholar] [CrossRef] [PubMed]
  183. Arooj, M.; Kim, S.; Sakkiah, S.; Cao, G.P.; Lee, Y.; Lee, K.W. Molecular modeling study for inhibition mechanism of human chymase and its application in inhibitor design. PLoS ONE 2013, 8, e62740. [Google Scholar] [CrossRef] [PubMed]
  184. Schwartz, L.B.; Bradford, T.R. Regulation of tryptase from human lung mast cells by heparin. Stabilization of the active tetramer. J. Biol. Chem. 1986, 261, 7372–7379. [Google Scholar] [CrossRef] [PubMed]
  185. Schwartz, L.B.; Lewis, R.A.; Austen, K.F. Tryptase from human pulmonary mast cells. Purification and characterization. J. Biol. Chem. 1981, 256, 11939–11943. [Google Scholar] [CrossRef]
  186. Alter, S.C.; Yates, P.; Margolius, H.S.; Schwartz, L.B. Tryptase and kinin generation: Tryptase from human mast cells does not activate human urinary prokallikrein. Int. Arch. Allergy Appl. Immunol. 1987, 83, 321–324. [Google Scholar] [CrossRef]
  187. Cregar, L.; Elrod, K.C.; Putnam, D.; Moore, W.R. Neutrophil myeloperoxidase is a potent and selective inhibitor of mast cell tryptase. Arch. Biochem. Biophys. 1999, 366, 125–130. [Google Scholar] [CrossRef]
  188. Elrod, K.C.; Moore, W.R.; Abraham, W.M.; Tanaka, R.D. Lactoferrin, a potent tryptase inhibitor, abolishes late-phase airway responses in allergic sheep. Am. J. Respir. Crit. Care Med. 1997, 156, 375–381. [Google Scholar] [CrossRef] [PubMed]
  189. Alter, S.C.; Kramps, J.A.; Janoff, A.; Schwartz, L.B. Interactions of human mast cell tryptase with biological protease inhibitors. Arch. Biochem. Biophys. 1990, 276, 26–31. [Google Scholar] [CrossRef]
  190. Samoszuk, M.; Corwin, M.; Hazen, S.L. Effects of human mast cell tryptase and eosinophil granule proteins on the kinetics of blood clotting. Am. J. Hematol. 2003, 73, 18–25. [Google Scholar] [CrossRef]
  191. Schechter, N.M.; Eng, G.Y.; McCaslin, D.R. Human skin tryptase: Kinetic characterization of its spontaneous inactivation. Biochemistry 1993, 32, 2617–2625. [Google Scholar] [CrossRef] [PubMed]
  192. Frommherz, K.J.; Faller, B.; Bieth, J.G. Heparin strongly decreases the rate of inhibition of neutrophil elastase by α1-proteinase inhibitor. J. Biol. Chem. 1991, 266, 15356–15362. [Google Scholar] [CrossRef] [PubMed]
  193. Ermolieff, J.; Boudier, C.; Laine, A.; Meyer, B.; Bieth, J.G. Heparin protects cathepsin G against inhibition by protein proteinase inhibitors. J. Biol. Chem. 1994, 269, 29502–29508. [Google Scholar] [CrossRef]
  194. Fleddermann, J.; Pichert, A.; Arnhold, J. Interaction of serine proteases from polymorphonuclear leucocytes with the cell surface and heparin. Inflammation 2012, 35, 81–88. [Google Scholar] [CrossRef]
  195. Taggart, C.; Cervantes-Laurean, D.; Kim, G.; McElvaney, N.G.; Wehr, N.; Moss, J.; Levine, R.L. Oxidation of either methionine 351 or methionine 358 in alpha 1-antitrypsin causes loss of anti-neutrophil elastase activity. J. Biol. Chem. 2000, 275, 27258–27265. [Google Scholar] [CrossRef]
  196. Evans, M.D.; Pryor, W.A. Cigarette smoking. emphysema and damage to alpha 1-proteinase inhibitor. Am. J. Physiol. Lung Cell. Mol. Physiol. 1994, 266, L593–L611. [Google Scholar] [CrossRef]
  197. Dittrich, A.S.; Kühbandner, I.; Gehrig, S.; Rickert-Zacharias, V.; Twigg, M.; Wege, S.; Taggart, C.C.; Herth, F.; Schultz, C.; Mall, M.A. Elastase activity on sputum neutrophils correlates with severity of lung disease in cystic fibrosis. Eur. Respir. J. 2018, 51, 1701910. [Google Scholar] [CrossRef]
  198. Duranton, J.; Adam, C.; Blieth, J.G. Kinetic mechanism of the inhibition of cathepsin G by α1-antichymotrypsin and α1-proteinase inhibitor. Biochemistry 1997, 37, 11239–11245. [Google Scholar] [CrossRef]
  199. Travis, J.; Bowen, J.; Baugh, R. Human α1-antichymotrypsin: Interaction with chymotrypsin-like proteinases. Biochemistry 1978, 26, 5651–5656. [Google Scholar] [CrossRef] [PubMed]
  200. Kalsheker, N.A. α1-Antichymotrypsin. Int. J. Biochem. Cell Biol. 1996, 28, 961–964. [Google Scholar] [CrossRef] [PubMed]
  201. Thompson, R.C.; Ohlsson, K. Isolation, properties, and complete amino acid sequence of human secretory leukocyte protease inhibitor, a potent inhibitor of leukocyte elastase. Proc. Natl. Acad. Sci. USA 1986, 83, 6692–6696. [Google Scholar] [CrossRef]
  202. Franken, C.; Meijer, C.J.L.M.; Dijkman, J.H. Tissue distribution of antileukoprotease and lysozyme in humans. J. Histochem. Cytochem 1989, 37, 493–498. [Google Scholar] [CrossRef] [PubMed]
  203. Lee, C.H.; Igarashi, Y.; Hohman, R.J.; Kaulbach, H.; White, M.V.; Kaliner, M.A. Distribution of secretory leukoprotease inhibitor in the human nasal airway. Am. Rev. Respir Dis 1993, 147, 710–716. [Google Scholar] [CrossRef] [PubMed]
  204. McGarry, N.; Greene, C.M.; McElvaney, N.G.; Weldon, S.; Taggart, C.C. The ability of secretory leukocyte protease inhibitor to inhibit apoptosis in monocytes is independent of its antiprotease activity. J. Immunol. Res. 2015, 2015, 507315. [Google Scholar] [CrossRef]
  205. Williams, S.E.; Brown, T.I.; Roghanian, A.; Sallenave, J.M. SLPI and elafin: One glove, many fingers. Clin. Sci. 2006, 110, 21–35. [Google Scholar] [CrossRef]
  206. Verrier, T.; Solhonne, B.; Sallenave, J.M.; Garcia-Verdugo, I. The WAP protein Trappin-2/Elafin: A handyman in the regulation of inflammatory and immune responses. Int. J. Biochem. Cell Biol. 2012, 44, 1377–1380. [Google Scholar] [CrossRef]
  207. Labidi-Galy, S.I.; Clauss, A.; Ng, V.; Duraisamy, S.; Elias, K.M.; Piao, H.Y.; Bilal, E.; Davidowitz, R.A.; Lu, Y.; Badalian-Very, G.; et al. Elafin drives poor outcome in high-grade serous ovarian cancers and basal-like breast tumors. Oncogene 2015, 34, 373–383. [Google Scholar] [CrossRef]
  208. Hunt, K.K.; Wingate, H.; Yokota, T.; Liu, Y.; Mills, G.B.; Zhang, F.; Fang, B.; Su, C.H.; Zhang, M.; Yi, M.; et al. Elafin, an inhibitor of elastase, is a prognostic indicator in breast cancer. Breast Cancer Res. 2013, 15, R3. [Google Scholar] [CrossRef]
  209. Wang, C.; Liao, Y.; He, W.; Zhang, H.; Zuo, D.; Liu, W.; Yang, Z.; Qiu, J.; Yuan, Y.; Li, K.; et al. Elafin promotes tumour metastasis and attenuates the anti-metastatic effects of erlotinib via binding to EGFR in hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 2021, 40, 113. [Google Scholar] [CrossRef]
  210. Cooley, J.; Takayama, T.K.; Shapiro, S.D.; Schechter, N.M.; Remold-O’Donnell, E. The serpin MNEI inhibits elastase-like and chymotrypsin-like serine proteases through efficient reactions at two active sites. Biochemistry 2001, 40, 15762–15770. [Google Scholar] [CrossRef]
  211. Marrero, A.; Duquerroy, S.; Trapani, S.; Goulas, T.; Guevara, T.; Andersen, G.R.; Navaza, J.; Sottrup-Jensen, L.; Gomis-Rüth, F.X. The crystal structure of human alpha2-macroglobulin reveals a unique molecular cage. Angew. Chem. Int. Ed. Engl. 2012, 51, 3340–3344. [Google Scholar] [CrossRef]
  212. Vandooren, J.; Itoh, Y. Alpha-2-macroglobulin in inflammation, immunity and infections. Front. Immunol. 2021, 12, 803244. [Google Scholar] [CrossRef] [PubMed]
  213. Salvesen, G.; Virca, G.D.; Travis, J. Interaction of alpha 2-macroglobulin with neutrophil and plasma proteinases. Ann. N. Y. Acad. Sci. 1983, 421, 316–326. [Google Scholar] [CrossRef]
  214. Wewers, M.D.; Herzyk, D.J.; Gadek, J.E. Alveolar fluid neutrophil elastase activity in the adult respiratory distress syndrome is complexed to alpha-2-macroglobulin. J. Clin. Investig. 1988, 82, 1260–1267. [Google Scholar] [CrossRef] [PubMed]
  215. Rao, N.V.; Wehner, N.G.; Marshall, B.C.; Gray, W.R.; Gray, B.H.; Hoidal, J.R. Characterization of proteinase-3 (PR-3), a neutrophil serine proteinase. Structural and functional properties. J. Biol. Chem. 1991, 266, 9540–9548. [Google Scholar] [CrossRef] [PubMed]
  216. Siddiqui, T.; Zia, M.K.; Ali, S.S.; Ahsan, H.; Khan, F.H. Insight into the interactions of proteinase inhibitor- alpha-2-macroglobulin with hypochlorite. Int. J. Biol. Macromol. 2018, 117, 401–406. [Google Scholar] [CrossRef]
  217. Reddy, V.Y.; Desorchers, P.E.; Pizzo, S.V.; Gonias, S.L.; Sahakian, J.A.; Levine, R.L.; Weiss, S.J. Oxidative dissociation of human alpha 2-macroglobulin tetramers into dysfunctional dimers. J. Biol. Chem. 1994, 269, 4683–4691. [Google Scholar] [CrossRef]
  218. Ahmad, S.; Simmons, T.; Varagic, J.; Moniwa, N.; Chappell, M.C.; Ferrario, C.M. Chymase-dependent generation of angiotensin II from angiotensin-(1-12) in human atrial tissue. PLoS ONE 2011, 6, e28501. [Google Scholar] [CrossRef] [PubMed]
  219. Song, Y.H.; Li, Y.; Du, J.; Mitch, W.E.; Rosenthal, N.; Delafontaine, P. Muscle-specific expression of IGF-1 blocks angiotensin II-induced skeletal muscle wasting. J. Clin. Investig. 2005, 115, 451–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Sanders, P.M.; Russell, S.T.; Tisdale, M.J. Angiotensin II directly induces muscle protein catabolism through the ubiquitin-proteasome proteolytic pathway and may play a role in cancer cachexia. Br. J. Cancer 2005, 93, 425–434. [Google Scholar] [CrossRef]
  221. Trobec, K.; von Haehling, S.; Anker, S.D.; Lainsack, M. Growth hormone, insulin-like growth factor 1, and insulin signaling—A pharmacological target in body wasting and cachexia. J. Cachexia Sarcopenia Muscle 2011, 2, 191–200. [Google Scholar] [CrossRef]
  222. Semprun-Prieto, L.C.; Sukhanov, S.; Yoshida, T.; Rezk, B.M.; Gonzalez-Villalobos, R.A.; Vaughn, C.; Tabony, A.M.; Delafontaine, P. Angiotensin II induced catabolic effect and muscle atrophy are redox dependent. Biochem. Biophys. Res. Commun. 2011, 409, 217–221. [Google Scholar] [CrossRef]
  223. Benigni, A.; Cassis, P.; Remuzzi, G. Angiotensin II revisited: New roles in inflammation, immunology, and aging. EMBO Mol. Med. 2010, 2, 247–257. [Google Scholar] [CrossRef]
  224. Crackower, M.A.; Sarao, R.; Oudit, G.Y.; Yagil, C.; Kozieradzki, I.; Scanga, S.E.; Oliveira-dos-Santos, A.J.; da Costa, J.; Zhang, L.; Pey, Y.; et al. Angiotensin-converting enzyme 2 is an essential regulator of heart function. Nature 2002, 417, 822–828. [Google Scholar] [CrossRef] [PubMed]
  225. Ferrario, C.M.; Chappell, M.C. Novel angiotensin peptides. Cell. Mol. Life Sci. 2004, 61, 2720–2727. [Google Scholar] [CrossRef]
  226. Muller, F.; Renné, T. Novel roles for factor XII-driven plasma contact activation system. Curr. Opin. Hematol. 2008, 15, 516–521. [Google Scholar] [CrossRef] [PubMed]
  227. Oehmcke-Hecht, S.; Köhler, J. Interaction of the human contact system with pathogens—An update. Front. Immunol. 2018, 9, 312. [Google Scholar] [CrossRef]
  228. Marceau, F.; Hess, H.J.; Rachvarov, D.R. The B1 receptors for kinins. Pharmacol. Rev. 1998, 50, 357–386. [Google Scholar]
  229. Cyr, M.; Lepage, Y.; Blais Jr., C.; Gervais, N.; Cugno, M.; Rouleau, J.-L.; Adam, A. Bradykinin and des-Arg9-bradykinin metabolic pathways and kinetics of activation of human plasma. Am. J. Physiol. Heart Circ. Physiol. 2001, 281, H275–H283. [Google Scholar] [CrossRef]
  230. Hamza, M.; Wang, X.M.; Adam, A.; Brahim, J.S.; Rowan, J.S.; Carmona, G.N.; Dionne, R.A. Kinin B1 receptors contributes to acute pain following minor surgery in humans. Mol. Pain 2010, 6, 12. [Google Scholar] [CrossRef] [Green Version]
  231. Murphy, G.; Nagase, H. Progress in matrix metalloproteinase research. Mol. Aspects Med. 2008, 29, 290–308. [Google Scholar] [CrossRef]
  232. Löffek, S.; Schilling, O.; Franzke, C.W. Biological role of matrix metalloproteinases: A critical balance. Eur. Respir. J. 2011, 38, 191–208. [Google Scholar] [CrossRef]
  233. Ra, H.-J.; Parks, W.C. Control of matrix metalloproteinase catalytic activity. Matrix Biol. 2007, 26, 587–596. [Google Scholar] [CrossRef]
  234. Fu, X.; Kao, J.L.; Bergt, C.; Kassim, S.Y.; Huq, N.P.; d’Avignon, A.; Parks, W.C.; Mecham, R.P.; Heinecke, J.W. Oxidative cross-linking of tryptophan to glycine restrains matrix metalloproteinase activity: Specific structural motifs control protein oxidation. J. Biol. Chem. 2004, 279, 6209–6912. [Google Scholar] [CrossRef]
  235. Thompson, R.W.; Holmes, D.R.; Mertens, R.A.; Liao, S.; Botney, M.D.; Mecham, R.P.; Welgus, H.G.; Parks, W.C. Production and localization of 92-kilodalton gelatinase in abdominal aortic aneurysms. An elastolytic metalloproteinase expressed by aneurysm-infiltrating macrophages. J. Clin. Investig. 1995, 96, 318–326. [Google Scholar] [CrossRef]
  236. Davis, V.; Persidskaia, R.; Baca-Regen, L.; Itoh, Y.; Nagase, H.; Persidsky, Y.; Ghorpade, A.; Baxter, B.T. Matrix metalloproteinase-2 production and its binding to the matrix are increased in abdominal aortic aneurysms. Arterioscler. Thromb. Vasc. Biol. 1998, 18, 1625–1633. [Google Scholar] [CrossRef]
  237. Zhang, X.; Shen, Y.H.; LeMaire, S.A. Thoracic aortic dissection: Are matrix metalloproteinases involved? Vascular 2009, 17, 147–157. [Google Scholar] [CrossRef]
  238. Strickland, D.K.; Ashcom, J.D.; Williams, S.; Burgess, W.H.; Migliorini, M.; Argraves, W.S. Sequence identity between the alpha 2-macroglobulin receptor and low density lipoprotein receptor-related protein suggests that this molecule is a multifunctional receptor. J. Biol. Chem. 1990, 265, 17401–17404. [Google Scholar] [CrossRef] [PubMed]
  239. Lee, M.H.; Atkinson, S.; Murphy, G. Identification of the extracellular matrix (ECM) binding motifs of tissue inhibitor of metalloproteinases (TIMP)-3 and effective transfer to TIMP-1. J. Biol. Chem. 2007, 282, 6887–6898. [Google Scholar] [CrossRef] [PubMed]
  240. Wang, X. The expanding role of mitochondria in apoptosis. Genes Develop. 2001, 15, 2922–2933. [Google Scholar]
  241. Kagan, V.E.; Tyurin, V.A.; Jiang, J.; Tyurina, Y.Y.; Ritov, V.B.; Amoscato, A.A.; Osipov, A.N.; Belikova, N.A.; Kapralov, A.A.; Kini, V.; et al. Cytochrome c acts as a cardiolipin oxygenase required for the release of proapoptotic factors. Nat. Chem. Biol. 2005, 1, 223–232. [Google Scholar] [CrossRef]
  242. Kakhlon, O.; Cabantchik, Z.I. The labile iron pool: Characterization, measurement, and participation in cellular processes. Free Radic. Biol. Med. 2002, 33, 1037–1046. [Google Scholar] [CrossRef]
  243. Kruszewski, M. Labile iron pool: The main determinant of cellular response to oxidative stress. Mutat. Res. 2003, 53, 81–92. [Google Scholar] [CrossRef]
  244. Frankel, E.N. Secondary products of lipid oxidation. Chem. Phys. Lipids 1987, 44, 73–85. [Google Scholar] [CrossRef]
  245. Stables, M.J.; Gillroy, D.W. Old and new generation of lipid mediators in acute inflammation and resolution. Prog. Lipid Res. 2011, 50, 35–51. [Google Scholar] [CrossRef]
  246. Dixon, S.J.; Lemberg, K.M.; Lamprecht, M.R.; Skouta, R.; Zaitsev, E.M.; Gleason, C.E.; Patel, D.N.; Cantley, A.M.; Yang, W.S.; Morrison, B., 3rd; et al. Ferroptosis: An iron-dependent form of nonapoptotic cell death. Cell 2012, 149, 1060–1072. [Google Scholar] [CrossRef]
  247. Chen, X.; Li, J.; Kang, R.; Klionsky, D.J.; Tang, D. Ferroptosis: Machinery and regulation. Autophagy 2021, 17, 2054–2081. [Google Scholar] [CrossRef]
  248. Yang, W.S.; Stockwell, B.R. Ferroptosis: Death by lipid peroxidation. Trends Cell. Biol. 2016, 26, 165–176. [Google Scholar] [CrossRef]
  249. Xie, Y.; Hou, W.; Song, X.; Yu, Y.; Huang, J.; Sun, X.; Kang, R.; Tang, D. Ferroptosis: Process and function. Cell Death Differ. 2016, 23, 369–379. [Google Scholar] [CrossRef] [PubMed]
  250. De Bie, P.; Muller, P.; Wijmenga, C.; Klomp, L.W.J. Molecular pathogenesis of Wilson and Menkes disease: Correlation of mutations with molecular defects and disease phenotypes. J. Med. Genet. 2007, 44, 673–688. [Google Scholar] [CrossRef] [PubMed]
  251. Squitti, R.; Pasqualetti, P.; Dal Forno, G.; Moffa, F.; Cassetta, E.; Lupoi, D.; Vernieri, F.; Rossi, L.; Baldassini, M.; Rossini, P.M. Excess of serum copper not related to ceruloplasmin in Alzheimer disease. Neurology 2005, 64, 1040–1046. [Google Scholar] [CrossRef] [PubMed]
  252. Squitti, R.; Bressi, F.; Pasqualetti, P.; Bonomini, C.; Ghidoni, R.; Binetti, G.; Cassetta, E.; Moffa, F.; Ventriglia, M.; Vernieri, F.; et al. Longitudinal prognostic value of serum ‘‘free” copper in patients with Alzheimer disease. Neurology 2009, 72, 50–55. [Google Scholar] [CrossRef] [PubMed]
  253. Wardman, P. Application of pulse radiolysis methods to study the reactions and structure of biomolecules. Rep. Prog. Phys. 1978, 41, 259–302. [Google Scholar] [CrossRef]
  254. Sarkar, S.; Prakash, D.; Marwaha, R.K.; Garewal, G.; Kumar, L.; Singhi, S.; Walia, B.N. Acute intravascular haemolysis in glucose-6-phosphate dehydrogenase deficiency. Ann. Trop. Paediatr. 1993, 13, 391–394. [Google Scholar] [CrossRef]
  255. Parker, C. Paroxysmal nocturnal hemoglobinuria. Curr. Opin. Hematol. 2012, 19, 141–148. [Google Scholar] [CrossRef]
  256. Kato, G.J.; Steinberg, M.H.; Gladwin, M.T. Intravascular hemolysis and the pathophysiology of sickle cell disease. J. Clin. Investig. 2017, 127, 750–760. [Google Scholar] [CrossRef]
  257. Sauret, J.M.; Marinides, G.; Wang, G.K. Rhabdomyolysis. Am. Fam. Physican 2002, 65, 907–912. [Google Scholar]
  258. Hunter, J.D.; Gregg, K.; Damani, Z. Rhabdomyolysis. Cont. Educ. Anaesth. Crit. Care Pain 2006, 6, 141–143. [Google Scholar] [CrossRef]
  259. Körmöczi, G.F.; Säemann, M.D.; Buchta, C.; Peck-Radosavljevic, M.; Mayr, W.R.; Schwartz, D.W.; Dunkler, D.; Spitzauer, S.; Panzer, S. Influence of clinical factors on the haemolysis marker haptoglobin. Eur. J. Clin. Invest. 2006, 36, 202–209. [Google Scholar] [CrossRef] [PubMed]
  260. Schaer, D.J.; Vinchi, F.; Ingoglia, G.; Tolosano, E.; Buehler, P.W. Haptoglobin, hemopexin, and related defense pathways—Basic science, clinical perspectives, and drug development. Front. Physiol. 2014, 5, 415. [Google Scholar] [CrossRef]
  261. Muller-Eberhard, U.; Javid, J.; Liem, H.H.; Hanstein, A.; Hanna, M. Plasma concentrations of hemopexin, haptoglobin and heme in patients with various hemolytic diseases. Blood 1968, 32, 811–815. [Google Scholar] [CrossRef]
  262. Tombe, M. Images in clinical medicine. Hemoglobinuria with malaria. N. Engl. J. Med. 2008, 358, 1837. [Google Scholar] [CrossRef] [PubMed]
  263. Vanholder, R.; Sever, M.S.; Erek, E.; Lameire, N. Acute renal failure related to the crush syndrome: Towards an era of seismo-nephrology? Nephrol. Dial. Transplant. 2000, 15, 1517–1521. [Google Scholar] [CrossRef]
  264. Genthon, A.; Wilcox, S.R. Crush syndrome: A case report and review of the literature. J. Emerg. Med. 2014, 46, 313–319. [Google Scholar] [CrossRef]
  265. Deuel, J.W.; Schaer, C.A.; Boretti, F.S.; Opitz, L.; Garcia-Rubio, I.; Baek, J.H.; Spahn, D.R.; Buehler, P.W.; Schaer, D.J. Hemoglobinuria-related acute kidney injury is driven by intrarenal oxidative reactions triggering a heme toxicity response. Cell Death Dis. 2016, 7, e2064. [Google Scholar] [CrossRef]
  266. Tracz, M.J.; Alam, J.; Nath, K.A. Physiology and pathophysiology of heme: Implications for kidney disease. J. Am. Soc. Nephrol. 2007, 18, 414–420. [Google Scholar] [CrossRef]
  267. Tabibzadeh, N.; Estournet, C.; Placier, S.; Perez, J.; Bilbault, H.; Girshovich, A.; Vandermeersch, A.; Jouanneau, C.; Letavenier, E.; Hammoudi, N.; et al. Plasma-heme induced renal toxicity is related to capillary rarefaction. Sci. Rep. 2017, 7, 40156. [Google Scholar] [CrossRef]
  268. Schröder, M.; Kaufman, R.J. The mammalian unfolded protein response. Annu. Rev. Biochem. 2005, 74, 730–789. [Google Scholar] [CrossRef] [PubMed]
  269. Wynn, T.A.; Ramalingam, T.R. Mechanisms of fibrosis: Therapeutic translation for fibrotic disease. Nat. Med. 2012, 18, 1028–1040. [Google Scholar] [CrossRef] [PubMed]
  270. Hogg, J.C. A brief review of chronic obstructive pulmonary disease. Can. Respir. J. 2012, 19, 381–384. [Google Scholar] [CrossRef] [PubMed]
  271. McDonough, J.E.; Yuan, R.; Suzuki, M.; Seyednejad, N.; Elliott, W.M.; Sanchez, P.G.; Wright, A.C.; Gefter, W.B.; Litzky, L.; Coxson, H.O.; et al. Small-airway obstruction and emphysema in chronic obstructive pulmonary disease. N. Engl. J. Med. 2011, 365, 1567–1575. [Google Scholar] [CrossRef]
  272. Taggart, C.C.; Greene, C.M.; Carroll, T.P.; O’Neill, S.J.; McElvaney, N.G. Elastolytic proteases. Inflammation resolution and dysregulation in chronic infective lung disease. Am. J. Respir. Crit. Care Med. 2005, 171, 1070–1076. [Google Scholar] [CrossRef]
  273. Brode, S.K.; Ling, S.C.; Chapman, K.R. Alpha-1 antitrypsin deficiency: A commonly overlooked cause of lung disease. Can. Med. Ass. J. 2012, 184, 1365–1371. [Google Scholar] [CrossRef]
  274. Gompertz, S.; O’Brien, C.; Bayley, D.L.; Hill, S.L.; Stockley, R.A. Changes in bronchial inflammation during acute exacerbations of chronic bronchitis. Eur. Respir. J. 2001, 17, 1112–1119. [Google Scholar] [CrossRef]
  275. Paone, G.; Conti, V.; Vestry, A.; Leone, A.; Puglisi, G.; Benassi, F.; Brunetti, G.; Schmid, G.; Cammarella, I.; Terzano, C. Analysis of sputum markers in the evaluation of lung inflammation and functional impairment in symptomatic smokers and COPD patients. Dis. Markers 2011, 31, 91–100. [Google Scholar] [CrossRef]
  276. Gramegna, A.; Amati, F.; Terranova, L.; Sotgui, G.; Tarsia, P.; Miglietta, D.; Calderazzo, M.A.; Aliberti, S.; Blasi, F. Neutrophil elastase in bronchiectasis. Respir. Res. 2017, 18, 211. [Google Scholar] [CrossRef]
  277. Weldon, S.; McNally, P.; McElvaney, N.G.; Elborn, J.S.; McAuley, D.F.; Wartelle, J.; Belaaouaj, A.; Levine, R.L.; Taggart, C.C. Decreased levels of secretory leucoprotease inhibitor in the Pseudomonas-infected cystic fibrosis lung are due to neutrophil elastase degradation. J. Immunol. 2009, 183, 8148–8156. [Google Scholar] [CrossRef]
  278. Guerra, M.; Frey, D.; Hagner, M.; Dittrich, S.; Paulsen, M.; Mall, M.A.; Schultz, C. Cathepsin G activity as a new marker for detecting airway inflammation by microscopy and flow cytometry. ACS Cent. Sci. 2019, 5, 539–548. [Google Scholar] [CrossRef] [PubMed]
  279. Janus, E.D.; Philipps, N.D.; Carrell, R.W. Smoking, lung function and alpha 1-antitrypsin deficiency. Lancet 1985, 8421, 152–154. [Google Scholar] [CrossRef] [PubMed]
  280. Imai, Y.; Kuba, K.; Ohto-Nakanishi, T.; Penninger, J.M. Angiotensin-converting enzyme 3 (ACE2) in disease pathogenesis. Circ. J. 2010, 74, 405–410. [Google Scholar] [CrossRef]
  281. Li, W.; Moore, M.J.; Vasilieva, N.; Sui, J.; Wong, S.K.; Berne, M.A.; Somasundaran, M.; Sullivan, J.L.; Luzuriaga, K.; Greenough, T.C.; et al. Angiotensin-converting enzyme 2 is a functional receptor for the SARS coronavirus. Nature 2003, 426, 450–454. [Google Scholar] [CrossRef] [PubMed]
  282. Kuba, K.; Imai, Y.; Rao, S.; Gao, H.; Guo, F.; Guan, B.; Huan, Y.; Yang, P.; Zhang, Y.; Deng, W.; et al. A crucial role of angiotensin converting enzyme 2 (ACE2) in SARS coronavirus-induced lung injury. Nat. Med. 2005, 11, 875–879. [Google Scholar] [CrossRef]
  283. Lumbers, E.R.; Delforce, S.J.; Pringle, K.G.; Smith, G.R. The lung, the heart, the novel coronavirus, and the renin-angiotensin system; the need for clinical trials. Front. Med. 2020, 7, 248. [Google Scholar] [CrossRef]
  284. Lamas-Barreiro, J.M.; Alonso-Suárez, M.; Fernández-Martín, J.J.; Saavedra-Alonso, J.A. Angiotensin II suppression in SARS-CoV-2 infection; a therapeutic approach. Nefrologia 2020, 40, 213–216. [Google Scholar] [CrossRef] [PubMed]
  285. Briquez, P.S.; Rouhani, S.J.; Yu, J.; Pyzer, A.R.; Trujillo, J.; Dugan, H.L.; Stamper, C.T.; Changrob, S.; Sperling, A.I.; Wilson, P.C.; et al. Severe COVID-19 induces autoantibodies against angiotensin II that correlate with blood pressure dysregulation and disease severity. Sci. Adv. 2022, 8, eabn3777. [Google Scholar] [CrossRef]
  286. Garvin, M.R.; Alvarez, C.; Miller, J.I.; Prates, E.T.; Walker, A.M.; Amos, B.K.; Mast, A.E.; Justice, A.; Aronow, B.; Jacobson, D. A mechanistic model and therapeutic interventions for COVID-19 involving a RAS-mediated bradykinin storm. eLife 2020, 9, e59177. [Google Scholar] [CrossRef]
  287. Gough, P.J.; Gomez, I.G.; Wille, P.T.; Raines, E.W. Macrophage expression of active MMP-9 induces acute plaque disruption in apoE-deficient mice. J. Clin. Investig. 2006, 116, 59–69. [Google Scholar] [CrossRef]
  288. Kadoglou, N.P.; Liapis, C.D. Matrix metalloproteinases: Contribution to pathogenesis, diagnosis, surveillance and treatment of abdominal aortic aneurysms. Curr. Med. Res. Opin. 2004, 20, 419–432. [Google Scholar] [CrossRef]
  289. Cohen, J.R.; Parikh, S.; Grella, L.; Sarfati, I.; Corbie, G.; Danna, D.; Wise, L. Role of the neutrophil in abdominal aortic aneurysm development. Cardiovasc. Surg. 1993, 1, 373–376. [Google Scholar] [CrossRef] [PubMed]
  290. Lysgaard Poulsen, J.; Stubbe, J.; Lindholt, J.S. Animal Models Used to Explore Abdominal Aortic Aneurysms: A Systematic Review. Eur. J. Vasc. Endovasc. Surg. 2016, 52, 487–499. [Google Scholar] [CrossRef]
  291. Maguire, E.M.; Pearce, S.W.A.; Xiao, R.; Oo, A.Y.; Xiao, Q. Matrix metalloproteinase in abdominal aortic aneurysm and aortic dissection. Pharmaceuticals 2019, 12, 118. [Google Scholar] [CrossRef] [PubMed]
  292. Zhang, X.; Ares, W.J.; Taussky, P.; Ducruet, A.; Grandhi, R. Role of matrix metalloproteinases in the pathogenesis of intracranial aneurysms. Neurosurg. Focus 2019, 47, E4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Freestone, T.; Turner, R.J.; Coady, A.; Higman, D.J.; Greenhalgh, R.M.; Powell, J.T. Inflammation and matrix metalloproteinases in the enlarging abdominal aortic aneurysm. Arterioscler. Thromb. Vasc. Biol. 1995, 15, 1145–1151. [Google Scholar] [CrossRef]
  294. Tamarina, N.A.; McMillan, W.D.; Shively, V.P.; Pearce, W.H. Expression of matrix metalloproteinases and their inhibitors in aneurysms and normal aorta. Surgery 1997, 122, 264–271. [Google Scholar] [CrossRef]
  295. Lauer-Fields, J.L.; Juska, D.; Fields, G.B. Matrix metalloproteinases and collagen catabolism. Biopolymers 2002, 66, 19–32. [Google Scholar] [CrossRef]
  296. Iredale, J.P.; Thompson, A.; Henderson, N.C. Extracellular matrix degradation in liver fibrosis: Biochemistry and regulation. Biochim. Biophys. Acta 2013, 1832, 876–883. [Google Scholar] [CrossRef]
  297. Lech, M.; Anders, H.-J. Macrophages and fibrosis: How resident and infiltrating mononuclear phagocytes orchestrate all phases of tissue injury and repair. Biochim. Biophys. Acta 2013, 1832, 989–997. [Google Scholar] [CrossRef]
  298. Gurtner, G.C.; Werner, S.; Barrandon, Y.; Longaker, M.T. Wound repair and regeneration. Nature 2008, 453, 314–321. [Google Scholar] [CrossRef]
  299. Horowitz, J.C.; Thannickal, V.J. Mechanisms for the resolution of organ fibrosis. Physiology 2019, 34, 43–55. [Google Scholar] [CrossRef] [PubMed]
  300. Otte, J.M.; Rosenberg, I.M.; Podolsky, D.K. Intestinal myofibroblasts in innate immune responses of the intestine. Gastroenterology 2003, 124, 1866–1878. [Google Scholar] [CrossRef] [PubMed]
  301. Meneghin, M.D.; Hogaboam, C. Infectious disease, the innate immune response, and fibrosis. J. Clin. Investig. 2007, 117, 530–538. [Google Scholar] [CrossRef]
  302. Varga, J.; Abraham, D. Systemic sclerosis: A prototypic multisystem fibrotic disorder. J. Clin. Investig. 2007, 117, 557–567. [Google Scholar] [CrossRef]
  303. Lee, C.G.; Homer, R.J.; Zhu, Z.; Lanone, S.; Wang, X.; Koteliansky, V.; Shipley, J.M.; Gotwals, P.; Noble, P.; Chen, Q.; et al. Interleukin-13 induces tissue fibrosis by selectively stimulating and activating transforming growth factor-β1. J. Exp. Med. 2001, 194, 809–821. [Google Scholar] [CrossRef]
  304. Hasegawa, M.; Fujimoto, M.; Takehara, K.; Sato, S. Pathogenesis of systemic sclerosis: Altered B cell function is the key linking systemic autoimmunity and tissue fibrosis. J. Dermatol. Sci. 2005, 389, 1–17. [Google Scholar] [CrossRef]
  305. Mezzano, S.A.; Ruiz-Ortega, M.; Egido, J. Angiotensin II and renal fibrosis. Hypertension 2001, 38, 635–638. [Google Scholar] [CrossRef] [PubMed]
  306. Watanabe, T.; Barker, T.A.; Berk, B.C. Angiotensin II and the endothelium: Diverse signals and effects. Hypertension 2005, 45, 163–169. [Google Scholar] [CrossRef]
  307. Mu, X.; Shi, W.; Xu, Y.; Xu, C.; Zhao, T.; Geng, B.; Yang, J.; Pan, J.; Hu, S.; Zhang, C.; et al. Tumor-derived lactate induces M2 macrophage polarization via the activation of the ERK/STAT3 signaling pathway in breast cancer. Cell Cycle 2018, 17, 428–438. [Google Scholar] [CrossRef]
  308. Végran, F.; Boidot, R.; Michiels, C.; Sonveaux, P.; Feron, O. Lactate influx through the endothelial cell monocarboxylate transporter MCT1 supports an NFκB/IL-8 pathway that drives angiogenesis. Cancer Res. 2011, 71, 2550–2560. [Google Scholar] [CrossRef] [PubMed]
  309. Chamnee, T.; Ontong, P.; Itano, N. Hyaluronan: A modulator of the tumor microenvironment. Cancer Lett. 2016, 375, 20–30. [Google Scholar] [CrossRef] [PubMed]
  310. Teicher, B.A. Malignant cells, directors of the malignant process: Role of transforming growth factor-β. Cancer Metastasis Rev. 2001, 20, 133–143. [Google Scholar] [CrossRef] [PubMed]
  311. Dong, M.; Blobe, G.C. Role of transforming growth factor-β in hematologic malignancies. Blood 2006, 107, 4589–4596. [Google Scholar] [CrossRef] [PubMed]
  312. Luo, Y.; Ma, J.; Lu, W. The significance of mitochondrial dysfunction in cancer. Int. J. Mol. Sci. 2020, 21, 5598. [Google Scholar] [CrossRef]
  313. Brown, J.L.; Rosa-Caldwell, M.E.; Lee, D.E.; Blackwell, T.A.; Brown, L.A.; Perry, R.A.; Haynie, W.S.; Hardee, J.P.; Carson, J.A.; Wiggs, M.P.; et al. Mitochondrial degeneration precedes the development of muscle atrophy in progression cancer cachexia in tumor-bearing mice. J. Cachexia Sarcopenia Muscle 2017, 8, 926–938. [Google Scholar] [CrossRef]
  314. Blackwell, T.A.; Cervenka, I.; Khatri, B.; Brown, J.L.; Rosa-Caldwell, M.E.; Lee, D.E.; Perry, R.A., Jr.; Brown, L.A.; Haynie, W.S.; Wiggs, M.P.; et al. Transcriptomic analysis of the development of skeletal muscle atrophy in cancer-cachexia in tumor-bearing mice. Physiol. Genomics 2018, 50, 1071–1082. [Google Scholar] [CrossRef]
  315. Prokopchuk, O.; Grünwald, B.; Nitsche, U.; Jäger, C.; Prokopchuk, O.L.; Schubert, E.C.; Friess, H.; Martignoni, M.E.; Krüger, A. Elevated systemic levels of the matrix metalloproteinase inhibitor TIMP-1 correlate with clinical markers of cachexia in patients with chronic pancreatitis and pancreatic cancer. BMC Cancer 2018, 18, 128. [Google Scholar] [CrossRef]
  316. Cao, Z.; Zhao, K.; Jose, I.; Hoogenraad, N.J.; Osellame, L.D. Biomarkers for cancer cachexia: A mini review. Int. J. Mol. Sci. 2021, 22, 4501. [Google Scholar] [CrossRef]
  317. Grünwald, B.; Harant, V.; Schaten, S.; Frühschütz, M.; Spallek, R.; Höchst, B.; Stutzer, K.; Berchtold, S.; Erkan, M.; Prokopchuk, O.; et al. Pancreatic pre-malignant lesions secrete TIMP-1, which activates hepatic stellate cells via CD63 signaling to create a pre-metastatic niche in the liver. Gastroenterology 2016, 151, 1011–1024. [Google Scholar] [CrossRef]
  318. Tarnawski, R.; Skladowski, K.; Maciejewski, B. Prognostic value of hemoglobin concentration in radiotherapy for cancer of supraglottic larynx. Int. J. Radiat. Oncol. Biol. Phys. 1997, 38, 1007–1011. [Google Scholar] [CrossRef]
  319. Grogan, M.; Thomas, G.M.; Malemed, J.; Wong, F.L.; Pearcey, R.G.; Joseph, P.K.; Portelance, L.; Crook, J.; Jones, K.D. The importance of hemoglobin levels during radiotherapy for carcinoma of the cervix. Cancer 1999, 86, 1528–1536. [Google Scholar] [CrossRef]
  320. Yin, T.; He, S.; Liu, X.; Jiang, W.; Ye, T.; Lin, Z.; Sang, Y.; Su, C.; Wan, Y.; Shen, G.; et al. Extravascular red blood cells and hemoglobin promote tumor growth and therapeutic resistance as endogenous danger signals. J. Immunol. 2015, 194, 429–437. [Google Scholar] [CrossRef] [PubMed]
  321. Canesin, G.; Di Ruscio, A.; Li, M.; Ummarino, S.; Hedblom, A.; Choudhury, R.; Krzyzanowska, A.; Csizmadia, E.; Palominos, M.; Stiehm, A.; et al. Scavenging of labile heme by hemopexin is a key checkpoint in cancer growth and metastases. Cell Rep. 2020, 32, 108181. [Google Scholar] [CrossRef]
  322. Chang, H.Y.; Lee, S.H.; Liao, I.C.; Huang, S.H.; Cheng, H.C.; Liao, P.C. Secretomic analysis identifies alpha-1 antitrypsin (A1AT) as a required protein in cancer cell migration, invasion, and pericellular fibronectin assembly for facilitating lung colonization of lung adenocarcinoma cells. Mol. Cell. Proteomics 2012, 11, 1320–1339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Dudani, J.S.; Warren, A.D.; Bhatia, S.N. Harnessing protease activity to improve cancer care. Annu. Rev. Cancer Biol. 2018, 2, 353–376. [Google Scholar] [CrossRef]
  324. Lerman, I.; Hammes, S.R. Neutrophil elastase in the tumor microenvironment. Steroids 2018, 133, 96–101. [Google Scholar] [CrossRef] [PubMed]
  325. Houghton, A.M.; Rzymkiewicz, D.M.; Ji, H.; Gregory, A.D.; Egea, E.E.; Metz, H.E.; Stolz, D.B.; Land, S.R.; Marconcini, L.A.; Kliment, C.R.; et al. Neutrophil elastase-mediated degradation of IRS-1 accelerates lung tumor growth. Nat. Med. 2010, 16, 219–223. [Google Scholar] [CrossRef]
  326. Peng, B.; Hu, J.; Fu, X. ELANE: An emerging lane to selective anticancer therapy. Signal Transduction Targeted Therapy 2021, 6, 358. [Google Scholar] [CrossRef]
  327. Xie, G.; Cheng, T.; Lin, J.; Zhang, L.; Zhneg, J.; Liu, Y.; Xie, G.; Wang, B.; Yuan, Y. Local angiotensin II contributes to tumor resistance to checkpoint immunotherapy. J. Immunother. Cancer 2018, 6, 88. [Google Scholar] [CrossRef]
  328. Ino, K.; Shibata, K.; Kajiyama, H.; Yamamoto, E.; Nagasaka, T.; Nawa, A.; Nomura, S.; Kikkawa, F. Angiotensin II type 1 receptor expression in ovarian cancer and its correlation with tumour angiogenesis and patient survival. Br. J. Cancer 2006, 94, 552–560. [Google Scholar] [CrossRef] [PubMed]
  329. Pinter, M.; Jain, R.K. Targeting the renin-angiotensin system to improve cancer treatment: Implications for immunotherapy. Sci. Transl. Med. 2017, 9, eaan5616. [Google Scholar] [CrossRef] [PubMed]
  330. Pittet, D.; Wenzel, R.P. Nosocomial bloodstream infections. Secular trends in rates, mortality, and contribution to total hospital deaths. Arch. Intern. Med. 1995, 155, 1177–1184. [Google Scholar] [CrossRef] [PubMed]
  331. Llewelyn, M.J.; Cohen, J. Tracking the microbes in sepsis: Advancements in treatment bring challenges for microbial epidemiology. Clin. Infect. Dis. 2007, 44, 1343–1348. [Google Scholar] [CrossRef]
  332. Kollef, K.E.; Schramm, G.E.; Wills, A.R.; Reichley, R.M.; Mirek, S.T.; Kollef, M.H. Predictors of 30-day mortality and hospital costs in patients with ventilator-associated pneumonia attributed to potentially antibiotic-resistant gram-negative bacteria. Chest 2008, 134, 281–287. [Google Scholar] [CrossRef]
  333. Babu, M.; Menon, V.P.; Devi, P.U. Prevalence of antimicrobial resistant pathogens in severe sepsis and septic shock patients. J. Young Pharm. 2018, 10, 358–361. [Google Scholar] [CrossRef]
  334. Luyt, C.E.; Combes, A.; Deback, C.; Aubriot-Lorton, N.H.; Nieszkowska, A.; Trouillet, J.L.; Capron, F.; Agut, H.; Gibert, C.; Chastre, J. Herpes simplex virus lung infection in patients undergoing prolonged mechanical ventilation. Am. J. Respir. Crit. Care Med. 2007, 175, 935–942. [Google Scholar] [CrossRef]
  335. Limaye, A.P.; Kirby, K.A.; Rubenfeld, G.D.; Leisenring, W.M.; Bulger, E.M.; Neff, M.J.; Gibran, N.S.; Huang, M.-L.; Santo, T.K.; Corey, L.; et al. Cytomegalovirus reactivation in critically ill immunocompetent patients. J. Am. Med. Assoc. 2008, 300, 413–422. [Google Scholar] [CrossRef]
  336. Lin, G.-L.; McGinley, J.P.; Drysdale, S.B.; Pollard, A.J. Epidemiology and immune pathogenesis of viral sepsis. Front. Immunol. 2018, 9, 2147. [Google Scholar] [CrossRef]
  337. Crouser, E.; Exline, M.; Wewers, M.D. Sepsis: Links between pathogen sensing and organ damage. Curr. Pharm. Des. 2008, 14, 1840–1852. [Google Scholar] [CrossRef]
  338. Rittirsch, D.; Flierl, M.A.; Ward, P.A. Harmful molecular mechanisms in sepsis. Nat. Rev. Immunol. 2008, 8, 776–787. [Google Scholar] [CrossRef]
  339. Jiminez, M.F.; Watson, R.W.; Parodo, J.; Evans, D.; Foster, D.; Steinberg, M.; Rotstein, O.D.; Marshall, J.C. Dysregulated expression of neutrophil apoptosis in the systemic inflammatory response syndrome. Arch. Surg. 1997, 132, 1263–1270. [Google Scholar] [CrossRef] [PubMed]
  340. Demaret, J.; Venet, F.; Friggeri, A.; Cauzalis, M.-A.; Plassais, J.; Jallades, L.; Malcus, C.; Poitevin-Later, F.; Textoris, J.; Lepape, A.; et al. Marked alterations of neutrophil functions during sepsis-induced immunosuppression. J. Leukoc. Biol. 2015, 98, 1081–1090. [Google Scholar] [CrossRef] [PubMed]
  341. Brown, K.A.; Brain, S.D.; Pearson, J.D.; Edgeworth, J.D.; Lewis, S.M.; Treacher, D.F. Neutrophils in development of multiple organ failure in sepsis. Lancet 2006, 368, 157–169. [Google Scholar] [CrossRef] [PubMed]
  342. Kovach, M.A.; Standiford, T.J. The function of neutrophils in sepsis. Curr. Opin. Infect. Dis. 2012, 25, 321–327. [Google Scholar] [CrossRef] [PubMed]
  343. Tsujimoto, H.; Ono, S.; Majima, T.; Kawarabayashi, N.; Takayama, E.; Kinoshita, M.; Seki, S.; Hiraide, H.; Moldawer, L.L.; Mochizuki, H. Neutrophil elastase, MIP-2, and TLR-4 expression during human and experimental sepsis. Shock 2005, 23, 39–44. [Google Scholar] [CrossRef] [PubMed]
  344. Schrijver, I.T.; Kempermann, H.; Roest, M.; Kesecioglu, J.; de Lange, D.W. Myeloperoxidase can differentiate between sepsis and non-infectious SIRS and predicts mortality in intensive care patients with SIRS. Intens. Care Med. Exp. 2017, 5, 43. [Google Scholar] [CrossRef]
  345. Kothari, N.; Keshari, R.S.; Bogra, J.; Kohli, M.; Abbas, H.; Malik, A.; Dikshit, M.; Barthwal, M.K. Increased myeloperoxidase enzyme activity in plasma is an indicator of inflammation and onset of sepsis. J. Crit. Care 2011, 26, 435.e1–453.e7. [Google Scholar] [CrossRef]
  346. Cha, Y.S.; Yoon, J.M.; Jung, W.J.; Kim, Y.W.; Kim, T.H.; Kim, O.H.; Cha, K.C.; Kim, H.; Hwang, S.O.; Lee, K.H. Evaluation of usefulness of myeloperoxidase index (MPXI) for differential diagnosis of systemic inflammatory response syndrome (SIRS) in the emergency department. Emerg. Med. J. 2015, 32, 304–307. [Google Scholar] [CrossRef]
  347. Maruchi, Y.; Tsuda, M.; Mori, H.; Takenaka, N.; Gocho, T.; Huq, M.A.; Takeyama, N. Plasma myeloperoxidase-conjugated DNA level predicts outcomes and organ dysfunction in patients with septic shock. Crit. Care 2018, 22, 176. [Google Scholar] [CrossRef]
  348. Czaikoski, P.G.; Mota, J.M.S.C.; Nascimento, D.C.; Sȏnego, F.; Castanheira, F.V.S.; Melo, P.H.; Scortegagna, G.T.; Silva, R.L.; Borrosa-Sousa, R.; Souto, F.O.; et al. Neutrophil extracellular traps induce organ damage during experimental and clinical sepsis. PLoS ONE 2016, 11, e0148142. [Google Scholar] [CrossRef]
  349. Denning, N.-L.; Aziz, M.; Gurien, S.D.; Wang, P. DAMPs and NETs in sepsis. Front. Immunol. 2019, 10, 2536. [Google Scholar] [CrossRef] [PubMed]
  350. Larsen, R.; Gozzelino, R.; Jeney, V.; Tokaji, L.; Bozza, F.A.; Japiassú, A.M.; Bonaparte, D.; Cavalcante, M.M.; Chora, A.; Ferreira, A.; et al. A central role for free heme in the pathogenesis of severe sepsis. Sci. Transl. Med. 2010, 2, 51ra71-1. [Google Scholar] [CrossRef]
  351. Othmann, A.; Filep, J.G. Enemies at the gate: How cell-free hemoglobin and bacterial infection can cooperate to drive acute lung injury during sepsis. Am. J. Physiol. Heart Circ. Physiol. 2021, 321, H131–H134. [Google Scholar] [CrossRef] [PubMed]
  352. Verheij, M.W.; Bulder, I.; Wuillemin, W.A.; Voermans, C.; Zeerleder, S.S. Scavengers of hemoproteins as potential biomarkers for severe sepsis and septic shock. Transl. Med. Commun. 2021, 6, 8. [Google Scholar] [CrossRef]
  353. Lan, P.; Pan, K.-H.; Wang, S.-J.; Shi, Q.-C.; Yu, Y.-X.; Fu, Y.; Chen, Y.; Jiang, Y.; Hua, X.-T.; Zhou, J.-C.; et al. High serum iron level is associated with increased mortality in patients with sepsis. Sci. Rep. 2018, 8, 11072. [Google Scholar] [CrossRef] [PubMed]
  354. Brandtner, A.; Tymoszuk, P.; Nairz, M.; Lehner, G.F.; Fritsche, G.; Vales, A.; Falkner, A.; Schennach, H.; Theurl, I.; Joannidis, M.; et al. Linkage of alterations in systemic iron homeostasis to patients’ outcome in sepsis: A prospective study. J. Intens. Care 2020, 8, 76. [Google Scholar] [CrossRef]
  355. Galley, H.F. Oxidative stress and mitochondrial dysfunction in sepsis. Br. J. Anaesth. 2011, 107, 57–64. [Google Scholar] [CrossRef]
  356. Singer, M. The role of mitochondrial dysfunction in sepsis-induced multi-organ failure. Virulence 2014, 51, 66–72. [Google Scholar] [CrossRef]
  357. Rahmel, T.; Marko, B.; Nowak, H.; Bergmann, L.; Thon, P.; Rump, K.; Kreimendahl, S.; Rassow, J.; Peters, J.; Singer, M.; et al. Mitochondrial dysfunction in sepsis is associated with diminished intramitochondrial TFAM despite its increased cellular expression. Sci. Rep. 2020, 10, 21209. [Google Scholar] [CrossRef]
  358. Arnhold, J. Organ damage and failure. In Cell and Tissue Destruction. Mechanisms, Protection, Disorders; Academic Press: London, UK; San Diego, CA, USA; Cambridge, MA, USA; Oxford, UK, 2020; pp. 289–307. [Google Scholar] [CrossRef]
Figure 1. Classification of external cytotoxic agents. Three selected examples are indicated for each group.
Figure 1. Classification of external cytotoxic agents. Three selected examples are indicated for each group.
Ijms 24 03016 g001
Figure 2. Major classes of host-derived cytotoxic agents.
Figure 2. Major classes of host-derived cytotoxic agents.
Ijms 24 03016 g002
Figure 3. The interplay between host-derived cytotoxic agents and antagonizing principles. (Patho)physiological consequences of the release of cytotoxic agents at inflammatory sites highly depend on the status of protective mechanisms.
Figure 3. The interplay between host-derived cytotoxic agents and antagonizing principles. (Patho)physiological consequences of the release of cytotoxic agents at inflammatory sites highly depend on the status of protective mechanisms.
Ijms 24 03016 g003
Figure 4. Major pathways in formation of reactive species in activated neutrophils (upper panel) and stressed mitochondria (lower panel). Antagonizing principles against these species are displayed on grey backgrounds. In deactivation of transition metal ions, the term chelators stands for numerous proteins that scavenge, transport, and store iron and copper ions. Further explanations are given in the text. Abbreviations: MPO—myeloperoxidase, SOD—superoxide dismutase.
Figure 4. Major pathways in formation of reactive species in activated neutrophils (upper panel) and stressed mitochondria (lower panel). Antagonizing principles against these species are displayed on grey backgrounds. In deactivation of transition metal ions, the term chelators stands for numerous proteins that scavenge, transport, and store iron and copper ions. Further explanations are given in the text. Abbreviations: MPO—myeloperoxidase, SOD—superoxide dismutase.
Ijms 24 03016 g004
Figure 5. Formation of methemoglobin, metmyoglobin, and free heme as a result of excessive intravascular hemolysis and rhabdomyolysis. Protective mechanism are presented on grey backgrounds. Further explanations are given in the text.
Figure 5. Formation of methemoglobin, metmyoglobin, and free heme as a result of excessive intravascular hemolysis and rhabdomyolysis. Protective mechanism are presented on grey backgrounds. Further explanations are given in the text.
Ijms 24 03016 g005
Figure 6. Activities of neutrophil elastase at inflammatory sites.
Figure 6. Activities of neutrophil elastase at inflammatory sites.
Ijms 24 03016 g006
Figure 7. The interplay between neutrophil-derived serine proteases and antiproteases.
Figure 7. The interplay between neutrophil-derived serine proteases and antiproteases.
Ijms 24 03016 g007
Figure 8. Effects of serine proteases on the renin–angiotensin–aldosteron system. Further explanations are given in the text.
Figure 8. Effects of serine proteases on the renin–angiotensin–aldosteron system. Further explanations are given in the text.
Ijms 24 03016 g008
Figure 9. Disturbances of the equilibrium between MMPs and TIMPs and examples of resulting disorders.
Figure 9. Disturbances of the equilibrium between MMPs and TIMPs and examples of resulting disorders.
Ijms 24 03016 g009
Table 1. Major host-derived cytotoxic agents and their antagonizing principles.
Table 1. Major host-derived cytotoxic agents and their antagonizing principles.
Cytotoxic AgentMode of Cytotoxic ActionAntagonizing PrinciplesRemarks
Superoxide anion radicalRelease of Fe2+ from [4Fe-4S]2+ clusters, formation of peroxynitrite Superoxide dismutases, cytochrome c
Hydrogen peroxideFormation of hydroxyl radicals in reaction with Fe2+ or Cu+Catalase, peroxiredoxins, glutathione peroxidases
Hydroxyl radicalsDiffusion-controlled oxidation of many substratesNo antagonizing principles;
only limited protection by carbohydrates
Prevention of their formation is the main strategy
Very dangerous
PeroxynitriteFormation of substrate radicals, nitration of tyrosine residues, initiation of lipid peroxidationMyeloperoxidase, heme proteins
Hypochlorous acid, hypobromous acidPreferred oxidation of cysteine, methionine residues
Interaction with aromatic amino acid residues and amino groups
SCN, taurine, glutathione (GSH), ascorbate
Myeloperoxidase (MPO)Formation of HOCl, HOBr, substrate radicalsCeruloplasmin
Free transition metal ionsDangerous radical species in reaction with H2O2 and organic hydroperoxidesProper control over all aspects of iron and copper ion metabolismEnhanced yield of free transition metal ions is dangerous
Free methemoglobinFormation of free hemeHaptoglobin
Free metmyoglobinFormation of free hemeHaptoglobin
Free hemeOxidation at hydrophobic loci, hemolysis of red blood cells, cytotoxic to kidney and liver, interaction with G4 structures in nucleic acids, can act as DAMPHemopexin
Heme oxygenase
Very dangerous
Oxidative products in lipid phases such as lipid peroxyl radicals and lipid hydroperoxidesInduction of further oxidative modifications of yet-unperturbed moleculesLipid antioxidants such as α-tocopherol, carotinoids, ubiquinol, dehydrolipoic acid
Glutathione peroxidase 4 (GPX4), and GSHProper control over transition free metal ions
Oxidative products in water-exposed moleculesInduction of further oxidative modifications of yet-unperturbed moleculesUrate, ascorbate, polyphenols
Proper control over transition free metal ions
Neutrophil elastaseCleavage of many extracellular matrix components, formation of angiotensin IIα1-antitrypsin (A1AT), secretory leukocyte protease inhibitor (SLPI), elafin, serpin B1, α2-macroglobulinFailure of anti-proteases to inhibit elastase at severe oxidative stress
Very dangerous
Cathepsin GCleavage of extracellular matrix components, receptor shedding, formation of angiotensin IIA1AT, α1-antichymotrypsin, SLPI
Proteinase 3Cleavage of extracellular matrix components, in particular elastinA1AT, elafin
Mast cell tryptasesCleavage of extracellular matrix componentsHeparin-binding proteins such as lactoferrin, MPO, antithrombin IIIProtected by heparin against the action of anti-proteases
Mast cell chymaseCleavage of extracellular matrix components, chemokines, and cytokines, formation of angiotensin IIα1-antichymotrypsin
Angiotensin IIReceptor-mediated pro-inflammatory effectsAngiotensin converting enzyme 2 (ACE2)Very dangerous
BradykininReceptor-mediated pro-inflammatory effectsAminopeptidase P, angiotensin converting enzyme (ACE)
Matrix metalloproteases (MMPs)Cleavage of extracellular matrix componentsTissue inhibitors of metalloproteases (TIMPs)Problems at shifted balance between MMPs and TIMPs
Table 2. Environmental cytotoxic agents and their antagonizing principles.
Table 2. Environmental cytotoxic agents and their antagonizing principles.
Cytotoxic AgentMode of Cytotoxic ActionAntagonizing PrinciplesRemarks
Singlet oxygen (1O2)DNA damage, especially guanine [21,22]Carotenoids [23,24,25]Skin and eye exposure
OzoneFormation of ozonides and cytotoxic aldehydes [26]Ascorbate, GSH, urate [27,28]Exposure to respiratory system [29,30]
SunlightInduction of photooxidative processes, formation of 1O2 [31]Melanin, polyphenols, [32]Skin exposure
Ionizing irradiationWater radiolysis, formation of solvated electrons, O2•−, H2O2, OH, and substrate radicals [33,34,35]See remarks in Table 1Always present at very low level
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

Arnhold, J. Host-Derived Cytotoxic Agents in Chronic Inflammation and Disease Progression. Int. J. Mol. Sci. 2023, 24, 3016. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24033016

AMA Style

Arnhold J. Host-Derived Cytotoxic Agents in Chronic Inflammation and Disease Progression. International Journal of Molecular Sciences. 2023; 24(3):3016. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24033016

Chicago/Turabian Style

Arnhold, Jürgen. 2023. "Host-Derived Cytotoxic Agents in Chronic Inflammation and Disease Progression" International Journal of Molecular Sciences 24, no. 3: 3016. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24033016

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