Next Article in Journal
A New Oxadiazole-Based Topsentin Derivative Modulates Cyclin-Dependent Kinase 1 Expression and Exerts Cytotoxic Effects on Pancreatic Cancer Cells
Next Article in Special Issue
Model Systems for Evidencing the Mediator Role of Riboflavin in the UVA Cross-Linking Treatment of Keratoconus
Previous Article in Journal
High Performance Liquid Chromatography versus Stacking-Micellar Electrokinetic Chromatography for the Determination of Potentially Toxic Alkenylbenzenes in Food Flavouring Ingredients
Previous Article in Special Issue
Direct and Indirect Chemiluminescence: Reactions, Mechanisms and Challenges
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Oxidative Crosslinking of Peptides and Proteins: Mechanisms of Formation, Detection, Characterization and Quantification

by
Eduardo Fuentes-Lemus
1,
Per Hägglund
1,
Camilo López-Alarcón
2 and
Michael J. Davies
1,*
1
Department of Biomedical Sciences, Panum Institute, University of Copenhagen, 2200 Copenhagen, Denmark
2
Departamento de Química Física, Facultad de Química y de Farmacia, Pontificia Universidad Catolica de Chile, Santiago 7820436, Chile
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(1), 15; https://doi.org/10.3390/molecules27010015
Submission received: 29 November 2021 / Revised: 17 December 2021 / Accepted: 18 December 2021 / Published: 21 December 2021
(This article belongs to the Special Issue Biomimetic Radical Chemistry and Applications 2021)

Abstract

:
Covalent crosslinks within or between proteins play a key role in determining the structure and function of proteins. Some of these are formed intentionally by either enzymatic or molecular reactions and are critical to normal physiological function. Others are generated as a consequence of exposure to oxidants (radicals, excited states or two-electron species) and other endogenous or external stimuli, or as a result of the actions of a number of enzymes (e.g., oxidases and peroxidases). Increasing evidence indicates that the accumulation of unwanted crosslinks, as is seen in ageing and multiple pathologies, has adverse effects on biological function. In this article, we review the spectrum of crosslinks, both reducible and non-reducible, currently known to be formed on proteins; the mechanisms of their formation; and experimental approaches to the detection, identification and characterization of these species.

1. Introduction

The formation of covalently linked peptides and proteins plays a key role in many biological processes, both physiologically and pathologically. These can be formed intentionally, such as in the oxidative folding of nascent proteins within mammalian cells in the endoplasmic reticulum or Golgi involving the generation of disulfide bonds from two cysteine (Cys) residues and in the assembly of insect exoskeletons via the crosslinking of two tyrosine (Tyr) residues, or as a result of accidental exposure to oxidizing species (low-molecular mass or enzymes) that chemically link two protein sites. These crosslinks can be formed between different sites within the same molecule (intramolecular or intrachain crosslinks), between two different chains in a single molecule (e.g., the interchain crosslinks in mammalian insulins), or between two separate species (intermolecular crosslinks). Some of these crosslinks play a key role in stabilizing or maintaining proteins structures and can be essential to functional activity [1], whereas others have negative effects of biological function (e.g., altered turnover, lifetime or activity) [2]. Whilst some crosslinks appear to be benign and devoid of adverse effects and end up as targets of catabolic processes (e.g., degradation by proteasomes, lysosomes, other proteases), others are strongly associated with adverse effects and are implicated (in some cases, causally) in the development of pathologies (e.g., [3,4]).
Whilst it is well established that crosslink formation can have major effects on biological systems—either positively or negatively—our knowledge of the full complement of crosslinks formed in biological systems (the ‘crosslink-ome’, ‘X-link-ome’) is far from complete [2]. The number of disulfide-containing proteins is large and relatively well defined, with these being particularly abundant in proteins found in biological fluids (e.g., human serum albumin in plasma contains 17 disulfides), in structural proteins such as receptors (e.g., the low-density lipoprotein receptor, LDLR, contains 30 unique intradomain disulfides [5]) and extracellular matrix proteins (e.g., laminin contains ~200 disulfides). Disulfides are also present at more modest levels in intracellular proteins, where they can function as structural elements, allosteric effectors, or be involved in catalytic cycles [1,6]. For proteins containing many disulfides, ensuring correct pairing of Cys residues into disulfides during synthesis and assembly is a complex problem [7].
Unlike the relatively well-characterized complement of disulfide-containing proteins, the number, occurrence and sites of other potential crosslinks is poorly understood, but the subject of active research. Recent developments in our understanding of the chemistry of crosslink formation, and the development of new and more sensitive methods for their detection is driving advances in this field. This article summarizes recent developments, with a focus on crosslinks formed by oxidation (either enzyme-mediated or via the reactions of low-molecular-mass oxidants). Crosslinks formed by deliberately added reagents (e.g., dicarbonyl and related crosslinking agents, such as glutaraldehyde) or those arising from glycation/glycoxidation reactions of sugars are not discussed for reasons of space. These have been discussed elsewhere [8,9,10].

2. Enzymatic Protein Crosslinking

Multiple enzymes can mediate the crosslinking of proteins, with a few key examples briefly summarized below. Enzyme-generated crosslinks are critical to the formation of many three-dimensional structures as these provide strength and rigidity, if biologically required. Examples include crosslinks formed within the extracellular matrix (ECM) of most, if not all, tissues, such as those formed between matrix proteins, and particularly collagens by the copper-containing lysyl oxidase (LOX) and LOX-like (LOXL) enzymes (reviewed in [11]). LOX oxidizes specific lysine (Lys) and hydroxylysine residues to carbonyls that undergo subsequent reactions to crosslink collagens (e.g., types I and III) and elastin [11,12,13,14]. In contrast, the LOXL family of enzymes acts on collagen type IV and drives the assembly of basement membranes [11,15]. Other enzymes also contribute to collagen crosslinking in the ECM with peroxidasin, a member of the heme peroxidase superfamily, mediating the formation of highly specific methionine (Met) to Lys crosslinks within the NC1 domains on collagen via generation of the oxidant hypobromous acid (HOBr). This species reacts rapidly with the Met residue to form an intermediate that then reacts with a suitably positioned Lys residue [16,17] (see also below). This type of crosslinking has been reported across many species [18]. Other members of the peroxidase superfamilies (e.g., horseradish peroxidase, myeloperoxidase, laccase) can also generate crosslinks via enzyme-mediated oxidation of substrates to radicals which then undergo radical–radical coupling. A classic example is oxidative coupling of Tyr and a wide range of other phenols via phenoxyl radical generation [19,20,21].
Isopeptide crosslinks involving reactions of the Lys side-chain amine with the carboxylic acid of aspartic acid (Asp), or the amides of asparagine (Asn) and glutamine (Gln) residues can be formed spontaneously or enzymatically [22,23,24]. These can be formed intracellularly, or within the ECM, with their function being associated in the latter case not only with enhanced ECM rigidity, but also in the attachment of bacterial pathogens to host tissue collagens and fibrinogens [25]. Enzymatic isopeptide bonds, formed by factor XIIIa (FXIIIa) between Lys and Gln residues, are critical to the formation of fibrin assemblies in blood clotting. Related crosslinks are generated by tissue transglutaminase enzymes [26]. These crosslinks are used to attach targeting moieties to specific proteins, including ubiquitin to proteins destined for proteosomal degradation [27], and small-ubiquitin-like modifier proteins (SUMOs, via the C-terminal glycine, Gly, residue) in the case of protein transport, targeting and regulation. These crosslinks involve the initial activation of the ubiquitin/SUMO, and subsequent transfer to the target, involving multiple activating (e.g., E1), conjugating (E2) and ligation (E3) enzymes [28]. Multiple glutamic, Glu and Gly residues have also been detected attached to Glu residues in the C-terminal regions of microtubulin, with this reported to be involved in microtubule function [29].
Covalent bonds are also widely and deliberately generated between proteins and co-factors to produce functional enzymes, with examples including the crosslinking of heme, flavin, pyridoxal, biotin, thiamine, molybdopterin and lipoic acid to proteins (see, for example, [30]). Whilst of considerable interest and importance, further discussion of these reactions lies outside the scope of this review, and subsequent sections are focused on oxidant-mediated reactions, though some of these enzyme-mediated crosslinks involve the formation and reactions of oxidants (e.g., by peroxidases and peroxidasin).
An overview of the crosslinks that are discussed further in this review is presented in Figure 1.

3. One-Electron (Radical–Radical) Reactions

Dimerization of two radicals to form a new covalent bond is typically a very fast process due to the low energy barriers for such reactions. Therefore, they are a major source of crosslinks in peptides and proteins when the radical flux is high and there are limited competing reactions. Most carbon-centered protein radicals (P) formed from aliphatic side-chains by hydrogen–atom abstraction reactions react rapidly with O2 at diffusion-controlled rates (k ~ 109 M−1 s−1) to give peptide or protein peroxyl radicals (P-OO) [31]. The rapidity of these reactions limits direct reactions of two P, except in circumstances where the O2 concentration is low. This is of biological relevance, as hypoxia is a common phenomenon, with endogenous levels of O2 being typically in the range 3–70 μM [32]. However, lower concentrations are present in situations where demand is great (e.g., high metabolic rates) or perfusion is poor (e.g., in the core of many solid tumors), thereby limiting P-OO formation and allowing (P-P) dimer formation [33]. For the limited number of P, where reaction with O2 is slow or modest, as is the case for Cys-derived thiyl radicals (RS, k < 107 M−1 s−1 [34]), tryptophan (Trp) indolyl radicals (Trp, k < 4 × 106 M−1 s−1 [35,36]) and Tyr phenoxyl radicals (Tyr, k < 103 M−1 s−1 [37]), formation of disulfides (cystine) from two RS, di-tyrosine from two Tyr, di-tryptophan from two Trp, and crossed dimers between these (e.g., Tyr–Trp) can be generated. The structure and mechanisms of the formation of these species are discussed below.
Light, particularly of wavelengths >~280 nm, which are not absorbed by the ozone layer, can penetrate significantly into biological structures and be absorbed either directly by protein residues, particularly Trp, Tyr and cystine [38], or by other species with high extinction coefficients in the long wavelength UV or visible regions. Energy absorption by non-protein species can give rise to indirect protein oxidation via the formation of excited states (e.g., singlet oxygen, 1O2 and reactive triplets) and/or radicals [38]. Direct UV absorption by proteins can form RS from homolysis of the –S–S– bond of cystine (with C-S cleavage being an alternative pathway), and Tyr and Trp radicals by photo-ionization of these side-chains. These species can then give rise to crosslinks.

4. Radical–Molecule Reactions

Radical–molecule reactions appear to be a limited pathway for the formation of protein crosslinks, due to the absence of double bonds to which radicals might add in proteins, and limited stability of adducts to aromatic rings. Notable exceptions are the rare amino acids dehydroalanine (DHA; 2-aminoacrylic acid) and dehydroaminobutyric acid (DHB; 2-aminocrotonic acid). These contain a double bond between the α- and β-carbons of the side-chain and are non-proteinogenic species [39], with these being generated via elimination reactions of serine residues (Ser), phospho-Ser and selenocysteine (Sec) residues (in the case of DHA) [39], and from threonine (Thr) and phospho-Thr (in the case of DHB) [40]. DHA can also be formed via cleavage of the carbon–sulfur bonds of the disulfide cystine, via mechanisms involving RS or nucleophilic elimination reactions [39].
Although radical addition to double bonds is typically rapid and energetically favorable due to low energy barriers, these reactions are rare as the concentrations of both DHA and DHB (with the former more abundant) and the radicals that might undergo addition with them are very low. Nevertheless, some examples are known for radicals that have relatively long lifetimes and modest rates of reaction with O2 (i.e., Cys thiyl, Tyr phenoxyl, Trp indolyl) [41].

5. Two-Electron (Molecule–Molecule) Reactions

Reactions between two molecules are typically much slower than between two radicals or radical–molecule reactions. However, the concentration of the reactants is often much higher than for reactive intermediates, and consequently, the overall rates of these reactions may be significant—and the yield of products greater—than for the processes outlined above. These reactions are therefore major sources of protein crosslinks. The rate constants for these reactions would be expected to vary enormously—though quantitative data is lacking for most systems—with some reactions involving unstable species (e.g., sulfenic acids (RSOH), S-nitrosothiols (RSNO), unsaturated aldehydes/ketones, quinones) being relatively rapid (i.e., occurring over seconds/minutes).

6. Types of Crosslinks Detected within and between Proteins and Peptides

The following sections and Table 1 summarize various types of crosslinks that have been detected within and between peptides and proteins, the nature of these species, their reversibility, mechanisms of formation and, subsequently, methods available to detect, identify, characterize and quantify these species.

6.2. Carbon–Carbon

6.2.1. Tyrosine–Tyrosine (Di-Tyrosine, Di-Tyr) Crosslinks

Di-Tyr crosslinks are formed deliberately by some enzymes, post-translationally, to give stability and elasticity to structural proteins (e.g., [125]). Nevertheless, protein exposure to oxidative environments, both biologically [126,127] and in food systems [128,129], can produce additional di-Tyr crosslinks. Thus, exposure of proteins to hydroxyl (HO) and peroxyl (ROO), peroxynitrous acid/peroxynitrite (ONOOH/ONOO), nitrogen dioxide (NO2), nitrosoperoxycarbonate (ONOOCO2), carbonate (CO3•−), lipid hydroperoxides (LOOH) and photoinduced reactions can yield di-Tyr crosslinks [130,131,132]. These can be formed inter- or intramolecularly between two proteins, from two free Tyr, or between free Tyr and proteins, by radical–radical reactions involving two Tyr (Figure 5) [130,133]. Electron delocalization over the benzene ring and the phenolic oxygen results in the formation of two regio-isomers, one involving a carbon–oxygen bond (C3–O, iso-dityrosine, vide infra) and another with a carbon–carbon (C3–C3, o,o’) bond. The latter is thermodynamically preferred (Figure 5) [19,134,135]. The production of di-Tyr crosslinks is limited by alternative reactions, such as that with O2 (though this is slow, k < 1 × 103 M−1 s−1, [130]), and also via fast reaction with other radicals such as O2•− (which gives short-lived peroxides [37,136]) and nitric oxide (NO) [137], and also with reducing agents such as ascorbate, thiols and polyphenols [138,139,140]. The extent of dimerization is also likely to be strongly modulated by steric and electronic interactions.
In pathophysiological contexts, di-Tyr crosslinking has been reported in low-density lipoproteins, arising from oxidative damage within human atherosclerotic lesions [61]; in erythrocytes exposed to a continuous flux of H2O2 [62]; in lens proteins, where it appears to originate from photo-oxidation of amino acids [63]; in tissues from a number of neurodegenerative conditions (e.g., amyloid-beta dimers in Alzheimer’s disease [64] and in α-synuclein [65,66], which may contribute to Parkinson’s disease); in lipofuscin pigments in the aged human brain [67]; in plasma of patients with chronic renal failure; and in urine of people with diabetes [68,69].
Di-Tyr crosslinks have also been detected on synthetic peptides exposed to peroxidase activity (e.g., hemoglobin and H2O2 [141]), and a large number of isolated proteins including (amongst others): lipoproteins [61,70,71], glucose 6-phosphate dehydrogenase (G6PDH) [53], ribonuclease A (RNAse A) [86], multiple extracellular matrix proteins [57,58,76], calmodulin [72], lysozyme [54,79], myoglobin [142,143], insulin [48,73,74] and human Δ25 centrin 2 [75].
The large body of evidence for di-Tyr crosslinks in oxidized proteins, fluids and tissues; the stability of this species (which is immune to reduction or repair); the association of this crosslink with the onset and development of human pathologies; and advances in analytical methods for the detection and quantification of di-Tyr (see below) have supported its use as a biomarker of protein oxidation in clinical studies [144,145].
Di-Tyr formation has also been reported in food systems, with both positive and negative effects. Protein crosslinking can give desirable textures and stability to processed foods, while negative effects have been reported with regard to poor protein digestibility and other undesirable properties [128,129]. Di-Tyr crosslinks have been detected in milk powders [146,147], during breadmaking [148], processing of myofibrillar proteins [149], in isolated caseins exposed to riboflavin-mediated photoreactions [51], and lactalbumin exposed to a laccase enzyme system [150] and UV-B light [50]. The presence of di-Tyr in foods, and consequent dietary intake, has resulted in studies on its toxicological properties. Intragastric administration of pure di-Tyr has been associated with metabolic alterations [151], including disrupted glucose metabolism [152], and renal alterations [153]. Interestingly, and of potential biological relevance, processed milks containing di-Tyr (and other Tyr oxidation products) can induce oxidative damage in plasma, liver and brain tissues, as well as spatial learning and memory impairments [154].
Di-Tyr crosslinks have also been detected in a wide variety of other organisms, including in insect cuticles, where di-Tyr crosslinks connect resilin proteins in a three-dimensional network [155]. Peroxidase-catalyzed di-Tyr crosslinks have also been detected in high concentrations in the oocyst walls of the apicomplexan parasite, Eimeria maxima, and in some bacterial spore proteins (e.g., those of Bacillus subtilis), with these providing a rigid framework that protects the oocyst or spores from harsh environmental conditions [60,77]. On the basis of these data, novel biomaterials have been developed that contain di-Tyr crosslinks that have excellent mechanical properties. Different strategies to induce efficient di-Tyr crosslink formation have been developed, including photoreactions mediated by ruthenium complexes and riboflavin [132].

6.2.2. Tryptophan–Tryptophan (Di-Tryptophan, di-Trp) Crosslinks

Trp is a major target for many oxidants, and a large number of oxidation products have been characterized [156]. Many of these involve the initial formation of (nitrogen-centered) indolyl radicals (Trp) which, depending on the properties of the protein and environment, can self-react to generate di-Trp crosslinks (Figure 6), though less information is available on these species than di-Tyr. Di-Trp crosslinks have been reported on human superoxide dismutase 1 (hSOD1) upon exposure to H2O2 in the presence of carbonate ions [78], with the peroxidase activity of hSOD1 being responsible for the generation of CO3•− that induce one-electron oxidation of the single, solvent-exposed, Trp32 residue in this protein and consequent intermolecular di-Trp crosslinking. This results in irreversible dimerization and is associated with non-amyloid protein aggregation and the occurrence of amyotrophic lateral sclerosis (ALS) [157]. CO3•−-mediated di-Trp crosslinks have also been reported for both free Trp [158] and lysozyme [79]. In the latter case, oxidation mediated by hSOD1 resulted in the detection of both hetero-dimers of lysozyme with hSOD1 and hSOD1 dimers [79]. These reports indicate a high propensity of CO3•− to generate di-Trp crosslinks [79]. Di-Trp crosslinks have also been reported on lysozyme exposed to ONOOH/ONOO [159] and photo-oxidation mediated by kynurenic acid [160], as well as in solutions of free Trp and N-acetyl-Trp exposed to photo-reactions of riboflavin and kynurenic acid, respectively [161,162]. Di-Trp crosslinks have been also detected in lens proteins of patients with nuclear cataracts [80] and in α-crystallin proteins exposed to UV radiation and a photosensitizer [163].
Di-Trp crosslinks have also been detected in aggregates of α-lactalbumin (α-LA) exposed to UV-B light [50], and in FtsZ proteins of M. jannaschii, a thermophilic microorganism, exposed to peroxyl radicals [164]. The dimerization of Trp is fast (k ~ 2 − 6 × 108 M−1 s−1 [159]), but dependent on the residue environment in peptides and proteins. Due to electron delocalization across the pyrrole and benzene rings, and the presence of chiral carbons, a complex mixture of regio- and stereo-isomers can be formed. This heterogeneity is likely to contribute to the limited number of reports on this type of crosslink, due to the difficulty in identifying and quantifying all of these species. This is exacerbated by a current absence of antibodies against this linkage (unlike di-Tyr, see below). Nonetheless, di-Trp crosslinks appear to be generated in low yields, probably due to the large number of alternative fates of Trp (Figure 6) and the low abundance of Trp in proteins (~1–2%). Despite the modest rate constant for addition of O2 to Trp (k ≤ 4 × 106 M−1 s−1 [35,36]), oxygenated products and alternative crosslinks (e.g., Trp-Tyr, see below) are commonly detected, and at higher yields than di-Trp. Thus, only a small extent of the aggregation of α-LA induced by UV-B light was associated with di-Trp crosslinks, in contrast to high yields of disulfide bonds [50]. Nonetheless, under anaerobic conditions, or environments with low O2 levels (e.g., human lens [161]), the formation of di-Trp crosslinks appears to be favored.
The production of di-Trp crosslinks is also modulated by protein conformations, as reported for the extracellular matrix (ECM) protein fibronectin, when exposed to ONOO/ONOOH [57]. Fibronectin is polymerized into elastic fibrils (fibrillation) at the cell surface, with this then allowing binding of other ECM components. This process is controlled (amongst others) by alterations to the protein conformation, with the compact structure present in plasma being resistant to fibrillation. However, an extended conformation generated by cell-generated forces (as well as by altered pH and salt concentrations) is more prone to fibrillation. MS analysis of tryptic peptides from fibronectin exposed to ONOO/ONOOH showed that two of the four crosslinks detected correspond to di-Trp crosslinks [57]. One di-Trp crosslink (Trp445–Trp2264) was detected in both the compact and extended conformations, while another (Trp177–Trp2250) was only detected in the compact state. Interestingly, Trp2250 was oxidized to a high extent in the extended conformation, indicating that compact conformation of fibronectin favors radical–radical reactions (involving Trp177 and Trp2250), while the extended state favors other fates, including formation of 6-nitro-Trp [57].
Trp–Trp bonds have also been detected in the active sites of some enzymes, including amine dehydrogenases (e.g., methylamine dehydrogenase and aromatic amine dehydrogenases), where a tryptophan tryptophanylquinone crosslink is generated between C2 of an unmodified Trp residue, and C5 (on the benzene ring) of a Trp quinone (with the carbonyl groups present at C7 and C8) [165].

6.2.3. Tyrosine–Tryptophan (Tyr–Trp) Crosslinks

Cross-reaction between Tyr and Trp can lead to Tyr–Trp crosslinks. These species are less well characterized than those described above, and the precise structure of these species remains to be elucidated (i.e., the nature of the crosslink bond and sites on the two rings that are joined). MS analyses of solutions containing free Tyr and Trp incubated with CO3 have provided evidence for three different, isomeric, Tyr–Trp crosslinks, with these likely to involve carbon–carbon bonds, and probably between the ortho (C3) position on the Tyr ring, and C3 on the indole ring of Trp [158].
Tyr–Trp crosslinks have been detected in peptides and isolated proteins exposed to multiple oxidants, including a cytochrome c peroxidase mutant exposed to a heme iron-peroxide reaction [81]; in glucose-6-phosphate dehydrogenase (G6PDH) exposed to ROO [53]; model peptides exposed to pulsed UV light [166]; in lysozyme treated with ROO and exposed to photooxidation reactions mediated by riboflavin and rose Bengal [54,55,56]; in fibronectin exposed to ONOOH/ONOO [57]; in protein extracts of the Gram-positive bacterium Lactococcus lactis exposed to photooxidation [167]; and proteins extracted from human cataractous lenses [80].

6.2.4. Other Carbon–Carbon Crosslinks

As most protein carbon radicals, P, react rapidly with O2 (see above), termination reactions between two P are limited under most biological conditions. The formation of some P–P crosslinks has, however, been reported in model peptide systems exposed to radiation-induced radicals [33], and similar species may be formed in biological situations where the pO2 value is low (e.g., in solid tumors, tissues subject to hypoxia). This requires further study.

6.3. Carbon–Nitrogen (C–N) Crosslinks

A significant number of carbon–nitrogen crosslinks have been identified, with the majority of these arising from molecule–molecule reactions, with the nitrogen atom acting as a nucleophile, though a small number of radical–radical crosslinks have been characterized.
Delocalization of the unpaired electron over the aromatic rings of Tyr and Trp, followed by radical–radical reactions has been proposed, on the basis of MS data, to give di-Trp species with carbon–nitrogen (C3-N1) bonds (Figure 6) [158]. The exact structures of these species are not completely clarified, though it is clear that multiple isobaric species are formed [159,161]. Some of these crosslinks appear to be of pathological relevance, as they have been detected on lens crystallin proteins subjected to photo-oxidation [161] and in human cataractous lenses [80].
The much larger group of molecule–molecule reactions that generate C–N crosslinks primarily involve nucleophilic reactions of the nitrogen centers on Lys and hydroxy-Lys (Nε-amine), Arg (guanidine nitrogen) and His (imidazole nitrogen) with either the carbon atom of carbonyl functions (Schiff base reactions) or other electron-deficient carbon centers (via Michael addition). For the former type of reaction, the carbonyl can be formed enzymatically (e.g., by LOX and LOXL enzymes, see earlier) or as a result of oxidation of alcohols (Ser/Thr) or C–H bonds in the presence of O2 (e.g., via ROO and ROOH [31,168]).
In the case of the Cu2+-dependent LOX/LOXL reactions and collagen crosslinking, the initial Schiff base adducts can undergo multiple further condensation reactions that allow several chains to be linked together via a single site [11]. Thus, there is abundant evidence for lysyl pyrrole, hydroxylysyl pyrrole, lysyl pyridinoline and hydroxylysyl pyridinoline species involving three or four collagen chains [169]. The mechanism of formation of these species involves the initial formation of a two-chain crosslink and then further condensation with a Lys/hydroxy-Lys on third and fourth chains. These species are critical to the correct assembly of collagen-containing extracellular matrices in tissues [11].
Similar reactions occur with (Ca2+-dependent) transglutaminases that form isopeptide bonds between the Lys side-chain amine and amide (Gln, Asn) or carboxylate (Asp) residues, with the amide/carboxylate ‘activated’ via reaction with the Cys residue of a Cys–His–Asp catalytic triad on the enzyme. The reaction of the carbonyl function with the Cys residue generates a thioester that is then susceptible to reaction with the Lys amine [22,23,24]. Such isopeptide bonds are highly resistant to most proteases and are widely encountered in the formation of insoluble protein polymers/aggregates, such as those found in brain tissue lesions of people with Alzheimer’s disease [24,170], in some bacterial spore proteins [60] and in the food industry, where these reactions are used as glues to alter food texture and properties [171,172].
Reactions also occur with α,β-unsaturated aldehyde and ketones and related species via Michael addition reactions (Figure 7). Thus, intra- and interchain crosslinks can be generated by the reaction of Lys/Arg/His residues with an oxidized imidazole group of the amino acid His, which acts as a Michael acceptor. These Lys–His, Arg–His and His–His products have been detected on multiple proteins, and particularly those subjected to photo-oxidation, where His residues are a major initial target of damage (e.g., [82,84,85,86,173]). They have also been detected in antibodies exposed to (photo)oxidation [82,84]. His-containing crosslinks have also been reported in some collagens, with these probably formed by nucleophilic attack of an imidazole nitrogen on an α,β-unsaturated structure formed via other crosslinking reactions [174].
Related Michael reactions occur with Tyr-derived quinones formed on oxidation of Tyr residues, with nitrogen nucleophiles, such as the Nε-amine of Lys side-chains undergoing adduction reactions with the oxidized ring (Figure 7). These reactions occur with lower rate constants than the corresponding reaction with the RS from Cys [110,175] (i.e., C–S crosslinks predominate over C–N linkages, through this is dependent on the availability and abundance of Cys versus Lys residues, with the latter usually predominating numerically). These reactions are important in the coupling of proteins, including via N-terminal amine groups, by tyrosinase and polyphenol oxidases (e.g., [176]) and in the formation of insect exoskeletons/cuticles of many insects [177], as well as the ‘glues’ that attach mussels to rocks [178,179].

6.4. Carbon–Oxygen

As alcohols (ROH) are poorer nucleophiles when compared to RS and neutral amines (RNH2), there are limited numbers of known carbon–oxygen crosslinks formed via molecular processes such as Michael reactions. However, such coupling can occur via radical–radical reactions, with the most well-established example being the formation of iso-di-tyrosine (iso-di-Tyr), where two Tyr dimerize via the phenolate oxygen of one ring, and C3 on the second ring (Figure 5); this appears to be a minor reaction compared to carbon–carbon coupling (see above) [19]. Iso-di-Tyr crosslinks have been reported in the extensin proteins that contribute to the architecture of wall plant cells [180], and in systems exposed to the myeloperoxidase–H2O2 system of human phagocytes, suggesting that these species may be present in tissues subject to acute or chronic inflammation, such as atherosclerosis [19].

7. Secondary Reactions of Crosslinks

In most biological systems, protein crosslinks, and particularly the formation of irreversible covalent crosslinks such as di-Tyr, di-Trp and Tyr–Trp, are considered as ‘final’ oxidation products [181]. However, over-oxidation of these species is possible, particularly under conditions of extensive oxidative damage, or under environments with long-term protein exposure to oxidants, where secondary one-electron oxidation with formation of radicals such as di-Tyr, di-Trp, or Tyr–Trp may occur. Such radicals can mediate similar reactions to those described above for Tyr and Trp, including reaction with O2 to produce oxygenated products (e.g., alcohols and hydroperoxides) and self-reactions to generate trimers and oligomers. Thus, formation of tri-Tyr and pulcherosine crosslinks have been detected in human phagocytes [19], while di-, tri- and tetra-Tyr have been reported in structural proteins of plant parasitic nematodes [182]. In addition, oligomers of Tyr (n = 2–8) have been reported in α-lactalbumin exposed to a horseradish peroxidase–H2O2 system [183]. Tri-Trp has been reported in trimers of hSOD1 triggered by CO3 [78], while tri-Trp and a di-Trp hydroperoxide (di-Trp-OOH) were reported in solutions of free Trp and riboflavin illuminated with a high-intensity 365 nm light-emitting diode [162].
In contrast, photo-oxidation (at 320 nm) of di-Tyr, in the presence of O2, has been reported to occur via processes involving O2, singlet oxygen (1O2) and H2O2 [184]. These observations were ascribed to the action of di-Tyr crosslinks as photosensitizers that could induce photo-damage to other biomolecules [184]. These findings suggest that di-Tyr crosslinks may be able to extend oxidation processes, opening new pathways of reactions, though these are only likely to be of major impact in systems with very extensive extents of oxidation. The scope and role of these pathways is unexplored, as well as the ability of peroxides such as di-Trp-OOH to extend protein oxidation (in line with the capacity of other hydroperoxides [31]).

8. Detection of Crosslinks, including Advantages and Disadvantages of Different Methods

Despite considerable methodological advancements, analysis of intra- or intermolecular protein crosslinks remains a challenging task, and the mechanisms underlying the formation of some known crosslinks remain to be established. There are also, undoubtedly, more types of crosslinks that remain to be discovered. This lack of knowledge arises partly due to the complexity of some structures, and the relatively poor current ‘tool-box’ to detect and predict the chemistry of the species and the sites involved. The following section therefore provides an overview of current approaches to the identification and characterization of oxidation-induced crosslinks, with a focus on both mass spectrometry (MS) and techniques (e.g., UPLC, immunological methods) that provide complementary information. It is clearly advisable to use combinations of methods and to investigate both gross modifications (e.g., changes in the molecular mass) and structural/conformational changes in crosslinked vs. native proteins with methods that allow the chemical nature and the residues involved in the formation of these species to be determined (Figure 8). Each of these methods has individual advantages and disadvantages.

8.1. Analysis of Changes in Molecular Mass by Electrophoresis and Size Exclusion Chromatography (SEC)

Protein electrophoresis (e.g., SDS-PAGE) and SEC-derived methodologies are excellent tools to assess the presence of crosslinked proteins in samples. Separation by electrophoresis is typically achieved through the use of polyacrylamide gels with different pore sizes. Diverse strategies can be utilized with this approach, such as running gels under native, denaturing and/or reducing conditions. This allows the investigation and differentiation of the contributions of reducible intermolecular bonds (e.g., disulfides between two protein chains) from non-reducible bonds (e.g., carbon–carbon bonds). Similarly, SEC can separate and fractionate soluble proteins based on their hydrodynamic radius and molecular mass, and depending on the mobile phase used, it can also provide information about the nature of the crosslinks (i.e., reducible or non-reducible intermolecular bonds).
Whilst these techniques can provide information on the presence of intermolecular bonds, they rarely provide information about the presence of intramolecular species. Moreover, data analysis needs to be carried out with care, as multiple proteins may be present in each band/fraction from complex matrices. This complexity can be overcome by using two-dimensional gel electrophoresis, with proteins separated firstly on the basis of their isoelectric point, and subsequently by molecular mass. Both methodologies have been successfully employed in combination with MS-based strategies [54,56,80,84,93,94,141,185], with isolation and enrichment of fractions containing dimers (and species with higher degrees of oligomerization) from non-crosslinked monomers before MS analysis. This approach has been used successfully to characterize crosslinked forms of IgG, α-synuclein and lysozyme [56,84,141]. Two-dimensional gel electrophoresis can also be used for specific detection of proteins connected by intermolecular disulfides [185]. In this approach, the proteins are first separated by non-reducing SDS-PAGE in the first dimension, and then separated by SDS-PAGE under reducing conditions in the second dimension. Proteins lacking intermolecular disulfide bonds line up on the diagonal, whereas proteins connected by intermolecular disulfide bonds dissociate from each other in the second dimension and migrate below the diagonal.

8.2. Analysis of Protein Crosslinks by Western (Immuno-) Blotting and ELISA Assays

The use of antibodies to investigate the formation of oxidation products, including di-Tyr, is a widely used strategy. These are typically examined using immunoblotting or ELISA assays, with the former providing (limited) information on the nature and identity of the proteins on which the crosslinks are present, and whether these are intramolecular (in a monomer) or interchain species. However, there are few well-characterized antibodies against crosslinked species, and these vary significantly in their specificity and selectivity, with some having significant cross-reactivity with other materials. Furthermore, crosslinks buried within highly aggregated species may be poorly, or not, recognized by (large) antibodies. Thus, appropriate control experiments are critical, and both positive and negative data should be validated by alternative methods.
Despite these caveats, immunoblotting has been widely used to detect di-Tyr present in α-synuclein in Lewy Bodies in Parkinson’s disease [186], in atherosclerotic plaques [187], and on multiple isolated proteins, including tropoelastin, α-synuclein, caseins, glucose-6-phosphate dehydrogenase, laminins and fibronectin [51,53,58,65,76,188]. Relative yields (i.e., versus the parent) can be achieved by ELISA, and this can serve as a very useful screening tool.

8.3. Direct Detection by Spectrophotometric and Fluorometric Assays

Some crosslinked species, as well as heavily aggregated proteins, can be monitored by spectrophotometry and fluorescence spectroscopy. For example, turbidity changes in solutions can be used (in an approximate manner) to monitor the time-course of protein crosslinking and aggregation [189,190]. However, this approach does not provide information on the type and nature of the crosslinked proteins and is useful only when there is extensive formation of heavily aggregated species, as soluble dimers and oligomers do not contribute significantly to the turbidity of solutions.
In contrast, fluorescence experiments can provide relatively specific data on crosslinks such as di-Tyr (excitation and emission maximum at ~280 and ~410 nm, respectively) [191]. Such data need to be interpreted with care, particularly with intact proteins or complex systems, as other fluorescent or optically absorbing species (e.g., Tyr, Trp, Trp-derived products, co-factors) may be present that distort excitation or emission processes. Thus, depending on the sample complexity, di-Tyr detection by fluorescence can provide useful information, but requires further confirmation using other methodologies. In this context, coupling fluorescence detection methods with HPLC, UPLC or GC separation (see below) has proven to be a valuable strategy to detect and quantify di-Tyr in mixtures of amino acids and oxidation products arising from acid hydrolysis of complex protein and tissue samples [191,192].

8.4. HPLC/UPLC Methodologies

These techniques can provide important quantitative data on both the consumption of the parent amino acid residues, and product formation, including Trp- and Tyr-derived crosslinks [158,191,193,194]. Before analysis, proteins are often isolated by precipitation (e.g., from homogenates or cell lysates) using acids (e.g., trichloroacetic acid) or organic solvents (e.g., ice-cold acetone or ethanol), followed by sample clean-up (e.g., delipidation), and then hydrolysis (with acid, base or enzymes) to give free amino acids and products. The advantages and disadvantages of this approach have been recently reviewed in [192], and therefore, the following is focused solely on the detection and quantification of crosslinked species.
The free amino acids and crosslinked products obtained from hydrolysis are separated by HPLC/UPLC and can be quantified by MS [158,159,193], by fluorescence detection (directly for di-Tyr [51], or by pre-column tagging for parent amino acids using sensitive fluorescent tags such as o-phthaldialdehyde [191]), UV absorption, or, for some species, electrochemical oxidation [191,195]. For HPLC/UPLC methodologies coupled to MS detection, heavy atom labeling (usually 2H, 13C, or 15N) can be utilized for accurate quantification [193]. Unfortunately, these approaches do not usually provide information on the sites of modification within proteins, or information on which proteins they were located on, if multiple species were present. This drawback, along with the limited number of crosslinked species that can be detected and quantified using these methods (di-Tyr, di-Trp, Trp-Tyr and cystine), makes it sensible to combine this approach with other methods (e.g., peptide mass mapping or immunoblotting).

8.5. Detection and Characterization of Crosslinked Proteins Using Other Biophysical Approaches

Biophysical techniques including circular dichroism (CD), light scattering, small angle neutron scattering (SANS), small angle X-ray scattering (SAXS), X-ray crystallography, NMR spectroscopy, and electron microscopy can provide useful information on protein structure. These methods are sensitive to modified structures, supplying valuable information on changes on morphology (i.e., mass, size and shape), secondary structure and solubility [45,196,197]. Some of these can also yield data on increased electron density between residues, thus supporting the presence of both intra- and intermolecular crosslinked species. This approach has been used in X-ray crystallographic studies, to determine the exact sites of crosslinks in oxidized peroxiredoxin 5, thioredoxin 2 and γS-crystallin, and to elucidate a covalent crosslink between Cys and Lys containing an N–O–S bridge [45,197,198,199]. However, most of these methods (with the exception of X-ray crystallography, NMR spectroscopy and cryogenic electron microscopy) cannot provide a structure of sufficiently high-resolution to provide definitive identifications and they must therefore be combined with other methodologies. Moreover, these methods are currently limited to homogeneous (single protein) samples that are available in large quantities (mg amounts).

8.6. Mass Spectrometry (MS)-Based Detection and Structural Characterization of Crosslinked Proteins

MS is a highly versatile technique for analysis of protein crosslinks that can be applied to (i) detect crosslinks and quantify their abundance, (ii) localize the specific crosslinking sites within polypeptides and (iii) reveal the identity of the crosslinked proteins. All of these questions cannot, however, be readily answered in a single experiment, and careful consideration must be given to appropriate workflows for specific applications.

8.6.1. MS Analysis of Crosslinked Amino Acids

Crosslinked amino acids can be detected and quantified by GC-MS or LC-MS, with the latter being most commonly used as it limits the need for derivatization and high-temperature conditions. Free crosslinked amino acids (i.e., not protein-bound) such as di-Tyr can be detected in body fluids such as urine, plasma and cerebrospinal fluid [200,201]. As outlined above, crosslinked amino acids can also be released from proteins either through acidic or alkaline hydrolysis, or through the use of unspecific proteases (e.g., pronase that hydrolyses proteins to single amino acids) [61,66,186]. The choice of hydrolysis method depends on the chemical stability of the investigated crosslink. For example, di-Trp crosslinks are unstable to acid hydrolysis (as the indole ring undergoes acid-catalyzed cleavage) and therefore, alkaline hydrolysis or pronase digestion is preferred [159].

8.6.2. MS Analysis of Crosslinked Peptides

Degradation of proteins using specific proteases such as trypsin releases peptides that can be analyzed by MS. This is the principle behind protein identification in standard ‘bottom-up’ proteomics, where the experimental mass-to-charge (m/z) values for the experimentally generated peptides are compared to theoretical values based on the known specificity of the enzyme. This type of workflow can also be applied for the identification of crosslinked peptides, since the mass of the crosslinked peptides can be predicted (for example, di-Tyr yields a mass shift of −2 Da relative to the sum of the two crosslinked peptides). The analysis of crosslinked peptides is, however, often more challenging, since the mass spectra of crosslinked peptides are more complex than those from regular (non-crosslinked) peptides. This is mainly due to the fact that the MS/MS spectra of crosslinked peptides may contain fragments from both of the linked individual peptides. Furthermore, the number of potential combinations of crosslinked peptides to take into consideration may be very large due to the complex chemistry of oxidation-induced crosslinking. This is particularly problematic if complex (cell/tissue) samples are analyzed and if intermolecular crosslinks are considered in database searches against large proteins databases (e.g., the entire human proteome). Due to these issues, there is a significant risk of false positive identifications, and it is therefore important to carefully validate the data, for example, using an 18O-labeling approach. This strategy relies on the ability of trypsin (and other related proteases) to incorporate two 18O atoms from isotope-labeled water molecules into each C-termini of proteolytic peptides [202]. Since crosslinked peptides contain two C-termini, a total of four 18O atoms will be incorporated instead of two 18O for regular (non-crosslinked peptides). Thus, it is possible to distinguish crosslinked peptides from non-crosslinked peptides based on their isotope labeling pattern [167,203]. This strategy has been successfully applied to validate oxidation-induced crosslinks in a wide range of proteins, including glucose-6-phosphate dehydrogenase, lysozyme, RNase A, elastin, fibronectin, cyclic AMP receptor protein, C-reactive protein and β-2-microglobulin (B2M) [53,54,55,56,57,59,86,93,94,203].
Identification of crosslinked peptides can also be facilitated by MS-based fragmentation of the crosslink bond, yielding individual mass spectra of the two formerly linked peptides, which can be analyzed separately. For stable crosslinks such as di-Tyr, this has recently been accomplished by ultraviolet photodissociation (UVPD) fragmentation [141]. For more labile crosslinks, such as disulfides, in-source fragmentation can be utilized [204,205]. Disulfides can also be probed indirectly through MS analysis of alkylated Cys thiol groups released after in vitro disulfide reduction using chemical (e.g., dithiothreitol) or enzymatic (e.g., thioredoxin [206]) treatment.

9. Crosslink Quantification

At present, there are few methods that allow absolute quantification of crosslink concentrations, with a major limitation being the non-availability of pure standards, particularly from commercial sources. Thus, there is a pressing need for further pure crosslink standards for quantitative analyses. Disulfides are a major exception, together with di-Tyr (which is commercially available) and a few species generated via glycation reactions (e.g., pentosidine). Di-Tyr can therefore be quantified in absolute terms using some of the methods outlined above, using the purified material to construct standard curves (e.g., for MS or fluorescence detection, and UPLC/LC separation). Relative concentrations can also be obtained from ELISA assays using an anti-di-Tyr antibody, and approximate concentrations from MS analyses. In nearly all cases, even with isolated proteins where protein turnover is not a confounding factor, the absolute concentrations determined by these approaches give low values, suggesting that there are many other important crosslinks which are either undetected, or which we cannot quantify.

10. Future Perspectives

Irreversible protein crosslinks have been associated with the onset and development of pathological conditions and human diseases. Similarly, reversible crosslinks, such as disulfides, appear to play a key role in cell signaling events, primarily as a result of reversible thiol–disulfide switches or related species (e.g., in the case of protein tyrosine phosphatase-1B, PTP-1B). The growing realization of the importance of these events has increased interest in the detection and quantification of these species together with irreversible crosslinks and other protein oxidation products (such as 3-nitroTyr, 3-chloroTyr, etc.) as biomarkers of biological events. This includes both physiological signaling and the assessment of the extent of oxidative damage for clinical purposes, such as diagnosis, prognosis, prevention or decisions on therapies/treatments. Di-Tyr crosslinks are recognized as reliable biomarkers of several pathologies, particularly neurodegenerative disorders [144]. However, this species only represents a fraction of the total pool of covalent crosslinks generated when proteins are exposed to oxidative conditions. As a consequence, new investigations aimed at understanding the role that this and other crosslinks (e.g., di-Trp or Tyr–Trp) play in controlling biological activity and contributing to human disease, and their use as clinical biomarkers, are necessary. In particular, it will be of interest and importance to determine how individual crosslinks modulate protein lifetime and turnover (e.g., by proteasomes). Current data do not typically allow such an assessment, as the methods used to generate crosslinks also induce multiple other alterations on the proteins, prohibiting the identification of cause and effect relationships. Studies focused on the mechanisms underlying crosslink formation in biological systems as well as the analytical strategies for their detection and quantification should also be considered. Such investigations are also likely to be valuable in other contexts where oxidation is relevant, such as in the processes triggered by physical exercise [207]; in food processing, formulation and degradation; in determining the shelf-life of protein-based medicines and therapeutics; and the development of new biomaterials.

Author Contributions

All the authors contributed to the drafting, editing and final review of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This project received funding from the European Union’s Horizon 2020 Research and Innovation Programme under the Marie Skłodowska-Curie grant agreement no. 890681 (to E.F.L.), the Novo Nordisk Foundation (NNF13OC0004294, NNF19OC0058493, and NNF20SA0064214 to M.J.D.), the Carlsberg Foundation (CF19-0451 to P.H.) and Fondecyt (grant no. 1180642 to C.L.A.).

Conflicts of Interest

M.J.D. declares consultancy contracts with Novo Nordisk A/S. This funder had no role in the design of the study; in the collection, analyses or interpretation of data; in the writing of the manuscript, or in the decision to publish the results. The other authors declare no conflict of interest.

Abbreviations

ALSamyotrophic lateral sclerosis
α-LAα-lactalbumin
CDcircular dichroism
CTQcysteine tryptophanylquinone
DHAdehydroalanine
DHBdehydrobutyric acid
Di-Trpdi-tryptophan
Di-Trp-OOHdi-tryptophan hydroperoxides
Di-Tyrdi-tyrosine
DOPA3,4-hydroxyphenylalanine
ELISAenzyme-linked immunoassay;
ECMextracellular matrix
FtsZfilamenting temperature-sensitive mutant Z
GC-MSgas chromatography mass spectrometry
G6PDHglucose 6-phosphate dehydrogenase
GPx’sglutathione peroxidases
GSAglutathione sulfonamide
GSHglutathione
GSSGglutathione disulfide
H2O2hydrogen peroxide
HOBrthe physiological mixture of hypobromous acid and its anion, hypobromite
HOClthe physiological mixture of hypochlorous acid and its anion, hypochlorite
HPLChigh performance liquid chromatography
hSOD1human superoxide dismutase 1
Iso-di-Tyriso-di-tyrosine
LC-MSliquid chromatography mass spectrometry
LDLRlow-density-lipoprotein receptor
LOOHlipid hydroperoxide
LOXlysyl oxidase
LOXLlysyl oxidase-like enzyme
MSmass spectrometry
NMRnuclear magnetic resonance
1O2singlet oxygen
ONOO/ONOOHthe physiological mixture of peroxynitrite anion and peroxynitrous acid
Pprotein carbon-centered radical
POOprotein peroxyl radical
P-Pprotein dimers
PTP1Bprotein tyrosine phosphatase 1B
RSthiyl radical RS
RSthiolate anion
R-NHBrbromamines
R-NHClchloramines
ROHalcohol
ROOHalkyl hydroperoxide
RNAse Aribonuclease A
R-NH2neutral amine
RS-NHR′sulfenamide
RSNOS-nitroso thiol
RSOHsulfenic acid
RSO2Hsulfinic acid
RSO3Hsulfonic acid
RS-(O)-NHR′sulfinamide
RS-(O)2-NHR′sulfonamide
RS-Xsulfenyl halide
SANSsmall angle neutron scattering
SAXSsmall angle X-ray scattering
SDS-PAGEsodium dodecyl sulphate–polyacrylamide gel electrophoresis
SECsize exclusion chromatography
SUMOssmall-ubiquitin-like modifier proteins
Tri-Trptri-tryptophan
Trptryptophan indolyl radical
Trxthioredoxin
TrxRthioredoxin reductase
TTQtryptophan-tryptophanylquinone
Tyrtyrosine phenoxyl radical
UPLCultra-performance liquid chromatography
UVPDultraviolet photodissociation

References

  1. Hogg, P.J. Disulfide bonds as switches for protein function. Trends Biochem. Sci. 2003, 28, 210–214. [Google Scholar] [CrossRef]
  2. Hägglund, P.; Mariotti, M.; Davies, M.J. Identification and characterization of protein cross-links induced by oxidative reactions. Expert Rev. Proteom. 2018, 18, 665–681. [Google Scholar] [CrossRef] [PubMed]
  3. Adam, O.; Theobald, K.; Lavall, D.; Grube, M.; Kroemer, H.K.; Ameling, S.; Schafers, H.J.; Bohm, M.; Laufs, U. Increased lysyl oxidase expression and collagen cross-linking during atrial fibrillation. J. Mol. Cell Cardiol. 2011, 50, 678–685. [Google Scholar] [CrossRef] [PubMed]
  4. Andringa, G.; Lam, K.Y.; Chegary, M.; Wang, X.; Chase, T.N.; Bennett, M.C. Tissue transglutaminase catalyzes the formation of alpha-synuclein crosslinks in Parkinson’s disease. FASEB J. 2004, 18, 932–934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Martinez-Olivan, J.; Fraga, H.; Arias-Moreno, X.; Ventura, S.; Sancho, J. Intradomain Confinement of Disulfides in the Folding of Two Consecutive Modules of the LDL Receptor. PLoS ONE 2015, 10, e0132141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Azimi, I.; Wong, J.W.; Hogg, P.J. Control of mature protein function by allosteric disulfide bonds. Antioxid. Redox Signal. 2011, 14, 113–126. [Google Scholar] [CrossRef]
  7. Depuydt, M.; Messens, J.; Collet, J.F. How proteins form disulfide bonds. Antioxid. Redox Signal. 2011, 15, 49–66. [Google Scholar] [CrossRef]
  8. Sinz, A.; Arlt, C.; Chorev, D.; Sharon, M. Chemical cross-linking and native mass spectrometry: A fruitful combination for structural biology. Protein Sci. 2015, 24, 1193–1209. [Google Scholar] [CrossRef] [Green Version]
  9. Chellan, P.; Nagaraj, R.H. Protein crosslinking by the Maillard reaction: Dicarbonyl-derived imidazolium crosslinks in aging and diabetes. Arch. Biochem. Biophys. 1999, 368, 98–104. [Google Scholar] [CrossRef] [PubMed]
  10. Nandi, S.K.; Nahomi, R.B.; Rankenberg, J.; Glomb, M.A.; Nagaraj, R.H. Glycation-mediated inter-protein cross-linking is promoted by chaperone-client complexes of alpha-crystallin: Implications for lens aging and presbyopia. J. Biol. Chem. 2020, 295, 5701–5716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Pehrsson, M.; Mortensen, J.H.; Manon-Jensen, T.; Bay-Jensen, A.C.; Karsdal, M.A.; Davies, M.J. Enzymatic cross-linking of collagens in organ fibrosis—Resolution and assessment. Expert Rev. Mol. Diagn. 2021, 21, 1049–1064. [Google Scholar] [CrossRef] [PubMed]
  12. Kagan, H.M. Lysyl oxidase: Mechanism, regulation and relationship to liver fibrosis. Pathol. Res. Pract. 1994, 190, 910–919. [Google Scholar] [CrossRef]
  13. López, B.; González, A.; Hermida, N.; Valencia, F.; de Teresa, E.; Díez, J. Role of lysyl oxidase in myocardial fibrosis: From basic science to clinical aspects. Am. J. Physiol. Heart Circ. Physiol. 2010, 299, H1–H9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Trackman, P.C. Lysyl Oxidase isoforms and potential therapeutic opportunities for fibrosis and cancer. Expert Opin. Ther. Targets 2016, 20, 935–945. [Google Scholar] [CrossRef] [Green Version]
  15. Bignon, M.; Pichol-Thievend, C.; Hardouin, J.; Malbouyres, M.; Brechot, N.; Nasciutti, L.; Barret, A.; Teillon, J.; Guillon, E.; Etienne, E.; et al. Lysyl oxidase-like protein-2 regulates sprouting angiogenesis and type IV collagen assembly in the endothelial basement membrane. Blood 2011, 118, 3979–3989. [Google Scholar] [CrossRef]
  16. McCall, A.S.; Cummings, C.F.; Bhave, G.; Vanacore, R.; Page-McCaw, A.; Hudson, B.G. Bromine is an essential trace element for assembly of collagen IV scaffolds in tissue development and architecture. Cell 2014, 157, 1380–1392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Bhave, G.; Cummings, C.F.; Vanacore, R.M.; Kumagai-Cresse, C.; Ero-Tolliver, I.A.; Rafi, M.; Kang, J.-S.; Pedchenko, V.; Fessler, L.I.; Fessler, J.H.; et al. Peroxidasin forms sulfilimine chemical bonds using hypohalous acids in tissue genesis. Nat. Chem. Biol. 2012, 8, 784–790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Peterfi, Z.; Geiszt, M. Peroxidasins: Novel players in tissue genesis. Trends Biochem. Sci. 2014, 39, 305–307. [Google Scholar] [CrossRef] [PubMed]
  19. Jacob, J.S.; Cistola, D.P.; Hsu, F.F.; Muzaffar, S.; Mueller, D.M.; Hazen, S.L.; Heinecke, J.W. Human phagocytes employ the myeloperoxidase-hydrogen peroxide system to synthesize dityrosine, trityrosine, pulcherosine, and isodityrosine by a tyrosyl radical-dependent pathway. J. Biol. Chem. 1996, 271, 19950–19956. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Selinheimo, E.; Autio, K.; Kruus, K.; Buchert, J. Elucidating the mechanism of laccase and tyrosinase in wheat bread making. J. Agric. Food Chem. 2007, 55, 6357–6365. [Google Scholar] [CrossRef] [PubMed]
  21. Dunford, H.B. Peroxidases. In Advances in Inorganic Biochemistry; Elsevier: Amsterdam, The Netherlands, 1982; pp. 41–68. [Google Scholar]
  22. Lorand, L.; Graham, R.M. Transglutaminases: Crosslinking enzymes with pleiotropic functions. Nat. Rev. Mol. Cell Biol. 2003, 4, 140–156. [Google Scholar] [CrossRef] [PubMed]
  23. Heck, T.; Faccio, G.; Richter, M.; Thony-Meyer, L. Enzyme-catalyzed protein crosslinking. Appl. MicroBiol. Biotechnol. 2013, 97, 461–475. [Google Scholar] [CrossRef] [Green Version]
  24. Wang, D.-S.; Dickson, D.W.; Malter, J.S. Tissue transglutaminase, protein cross-linking and Alzheimer’s disease: Review and views. Int. J. Clin. Experim. Pathol. 2008, 1, 5–18. [Google Scholar]
  25. Sridharan, U.; Ponnuraj, K. Isopeptide bond in collagen- and fibrinogen-binding MSCRAMMs. Biophys. Rev. 2016, 8, 75–83. [Google Scholar] [CrossRef] [PubMed]
  26. Siebenlist, K.R.; Mosesson, M.W. Evidence of intramolecular cross-linked A alpha.gamma chain heterodimers in plasma fibrinogen. Biochemistry 1996, 35, 5817–5821. [Google Scholar] [CrossRef] [PubMed]
  27. Kerscher, O.; Felberbaum, R.; Hochstrasser, M. Modification of proteins by ubiquitin and ubiquitin-like proteins. Ann. Rev. Cell Dev. Biol. 2006, 22, 159–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Scheffner, M.; Nuber, U.; Huibregtse, J.M. Protein ubiquitination involving an E1-E2-E3 enzyme ubiquitin thioester cascade. Nature 1995, 373, 81–83. [Google Scholar] [CrossRef] [Green Version]
  29. Westermann, S.; Weber, K. Post-translational modifications regulate microtubule function. Nat. Rev. Mol. Cell Biol. 2003, 4, 938–947. [Google Scholar] [CrossRef] [Green Version]
  30. Cao, Y.; Li, H. Dynamics of protein folding and cofactor binding monitored by single-molecule force spectroscopy. Biophys. J. 2011, 101, 2009–2017. [Google Scholar] [CrossRef] [Green Version]
  31. Davies, M.J. Protein oxidation and peroxidation. Biochem. J. 2016, 473, 805–825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Carreau, A.; El Hafny-Rahbi, B.; Matejuk, A.; Grillon, C.; Kieda, C. Why is the partial oxygen pressure of human tissues a crucial parameter? Small molecules and hypoxia. J. Cell Mol. Med. 2011, 15, 1239–1253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Dizdaroglu, M.; Simic, M.G. Isolation and characterization of radiation-induced aliphatic peptide dimers. Int. J. Radiat. Biol. 1983, 44, 231–239. [Google Scholar] [CrossRef] [PubMed]
  34. Schoneich, C. Thiyl radicals and induction of protein degradation. Free Radic. Res. 2016, 50, 143–149. [Google Scholar] [CrossRef] [Green Version]
  35. Fang, X.; Jin, F.; Jin, H.; von Sonntag, C. Reaction of the superoxide radical with the N-centred radical derived from N-acetyltryptophan methyl ester. J. Chem. Soc. Perkin. Trans. 1998, 259–263. [Google Scholar] [CrossRef]
  36. Candeias, L.P.; Wardman, P.; Mason, R.P. The reaction of oxygen with radicals from oxidation of tryptophan and indole-3-acetic acid. Biophys. J. 1997, 67, 229–237. [Google Scholar] [CrossRef]
  37. Hunter, E.P.; Desrosiers, M.F.; Simic, M.G. The effect of oxygen, antioxidants, and superoxide radical on tyrosine phenoxyl radical dimerization. Free Radic. Biol. Med. 1989, 6, 581–585. [Google Scholar] [CrossRef]
  38. Pattison, D.I.; Rahmanto, A.S.; Davies, M.J. Photo-oxidation of proteins. PhotoChem. PhotoBiol. Sci. 2012, 11, 38–53. [Google Scholar] [CrossRef]
  39. Siodlak, D. α,β-Dehydroamino acids in naturally occurring peptides. Amino Acids 2015, 47, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Friedman, M. Chemistry, biochemistry, nutrition, and microbiology of lysinoalanine, lanthionine, and histidinoalanine in food and other proteins. J. Agric. Food Chem. 1999, 47, 1295–1319. [Google Scholar] [CrossRef] [PubMed]
  41. Steinmann, D.; Mozziconacci, O.; Bommana, R.; Stobaugh, J.F.; Wang, Y.J.; Schoneich, C. Photodegradation pathways of protein disulfides: Human growth hormone. Pharm. Res. 2017, 34, 2756–2778. [Google Scholar] [CrossRef] [PubMed]
  42. Nagy, P.; Lechte, T.P.; Das, A.B.; Winterbourn, C.C. Conjugation of glutathione to oxidized tyrosine residues in peptides and proteins. J. Biol. Chem. 2012, 287, 26068–26076. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Rokhsana, D.; Howells, A.E.; Dooley, D.M.; Szilagyi, R.K. Role of the Tyr-Cys cross-link to the active site properties of galactose oxidase. Inorg. Chem. 2012, 51, 3513–3524. [Google Scholar] [CrossRef]
  44. Davies, C.G.; Fellner, M.; Tchesnokov, E.P.; Wilbanks, S.M.; Jameson, G.N. The Cys-Tyr cross-link of cysteine dioxygenase changes the optimal pH of the reaction without a structural change. Biochemistry 2014, 53, 7961–7968. [Google Scholar] [CrossRef]
  45. Wensien, M.; von Pappenheim, F.R.; Funk, L.M.; Kloskowski, P.; Curth, U.; Diederichsen, U.; Uranga, J.; Ye, J.; Fang, P.; Pan, K.T.; et al. A lysine-cysteine redox switch with an NOS bridge regulates enzyme function. Nature 2021, 593, 460–464. [Google Scholar] [CrossRef] [PubMed]
  46. Van Montfort, R.L.; Congreve, M.; Tisi, D.; Carr, R.; Jhoti, H. Oxidation state of the active-site cysteine in protein tyrosine phosphatase 1B. Nature 2003, 423, 773–777. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, Z.; Lyons, B.; Truscott, R.J.; Schey, K.L. Human protein aging: Modification and crosslinking through dehydroalanine and dehydrobutyrine intermediates. Aging Cell 2014, 13, 226–234. [Google Scholar] [CrossRef] [PubMed]
  48. Torosantucci, R.; Mozziconacci, O.; Sharov, V.; Schöneich, C.; Jiskoot, W. Chemical modifications in aggregates of recombinant human insulin induced by metal-catalyzed oxidation: Covalent cross-linking via michael addition to tyrosine oxidation products. Pharm. Res. 2012, 29, 2276–2293. [Google Scholar] [CrossRef] [Green Version]
  49. 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, 6209–6212. [Google Scholar] [CrossRef] [Green Version]
  50. Zhao, Z.; Engholm-Keller, K.; Poojary, M.M.; Boelt, S.G.; Rogowska-Wrzesinska, A.; Skibsted, L.H.; Davies, M.J.; Lund, M.N. Generation of aggregates of alpha-lactalbumin by UV-B light exposure. J. Agric. Food Chem. 2020, 68, 6701–6714. [Google Scholar] [CrossRef] [PubMed]
  51. Fuentes-Lemus, E.; Silva, E.; Leinisch, F.; Dorta, E.; Lorentzen, L.G.; Davies, M.J.; López-Alarcón, C. alpha- and beta-casein aggregation induced by riboflavin-sensitized photo-oxidation occurs via di-tyrosine cross-links and is oxygen concentration dependent. Food Chem. 2018, 256, 119–128. [Google Scholar] [CrossRef] [PubMed]
  52. Colombo, G.; Clerici, M.; Altomare, A.; Rusconi, F.; Giustarini, D.; Portinaro, N.; Garavaglia, M.L.; Rossi, R.; Dalle-Donne, I.; Milzani, A. Thiol oxidation and di-tyrosine formation in human plasma proteins induced by inflammatory concentrations of hypochlorous acid. J. Proteom. 2017, 152, 22–32. [Google Scholar] [CrossRef]
  53. Leinisch, F.; Mariotti, M.; Rykaer, M.; López-Alarcón, C.; Hägglund, P.; Davies, M.J. Peroxyl radical- and photo-oxidation of glucose 6-phosphate dehydrogenase generates cross-links and functional changes via oxidation of tyrosine and tryptophan residues. Free Radic. Biol. Med. 2017, 112, 240–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Fuentes-Lemus, E.; Mariotti, M.; Hägglund, P.; Leinisch, F.; Fierro, A.; Silva, E.; Davies, M.J.; López-Alarcón, C. Oxidation of lysozyme induced by peroxyl radicals involves amino acid modifications, loss of activity, and formation of specific crosslinks. Free Radic. Biol. Med. 2021, 167, 258–270. [Google Scholar] [CrossRef] [PubMed]
  55. Fuentes-Lemus, E.; Mariotti, M.; Hägglund, P.; Leinisch, F.; Fierro, A.; Silva, E.; López-Alarcón, C.; Davies, M.J. Binding of rose bengal to lysozyme modulates photooxidation and cross-linking reactions involving tyrosine and tryptophan. Free Radic. Biol. Med. 2019, 143, 375–386. [Google Scholar] [CrossRef] [PubMed]
  56. Fuentes-Lemus, E.; Mariotti, M.; Reyes, J.; Leinisch, F.; Hägglund, P.; Silva, E.; Davies, M.J.; López-Alarcón, C. Photo-oxidation of lysozyme triggered by riboflavin is O2-dependent, occurs via mixed type 1 and type 2 pathways, and results in inactivation, site-specific damage and intra- and inter-molecular crosslinks. Free Radic. Biol. Med. 2020, 152, 61–73. [Google Scholar] [CrossRef] [PubMed]
  57. Mariotti, M.; Rogowska-Wrzesinska, A.; Hägglund, P.; Davies, M.J. Cross-linking and modification of fibronectin by peroxynitrous acid: Mapping and quantification of damage provides a new model for domain interactions. J. Biol. Chem. 2021, 296, 100360. [Google Scholar] [CrossRef]
  58. Degendorfer, G.; Chuang, C.Y.; Hammer, A.; Malle, E.; Davies, M.J. Peroxynitrous acid induces structural and functional modifications to basement membranes and its key component, laminin. Free Radic. Biol. Med. 2015, 89, 721–733. [Google Scholar] [CrossRef] [PubMed]
  59. Leinisch, F.; Mariotti, M.; Andersen, S.H.; Lindemose, S.; Hägglund, P.; Mollegaard, N.E.; Davies, M.J. UV oxidation of cyclic AMP receptor protein, a global bacterial gene regulator, decreases DNA binding and cleaves DNA at specific sites. Sci. Rep. 2020, 10, 3106. [Google Scholar] [CrossRef] [Green Version]
  60. Ursem, R.; Swarge, B.; Abhyankar, W.R.; Buncherd, H.; de Koning, L.J.; Setlow, P.; Brul, S.; Kramer, G. Identification of native cross-links in Bacillus subtilis spore coat proteins. J. Proteome Res. 2021, 20, 1809–1816. [Google Scholar] [CrossRef]
  61. Leeuwenburgh, C.; Rasmussen, J.E.; Hsu, F.F.; Mueller, D.M.; Pennathur, S.; Heinecke, J.W. Mass spectrometric quantification of markers for protein oxidation by tyrosyl radical, copper, and hydroxyl radical in low density lipoprotein isolated from human atherosclerotic plaques. J. Biol. Chem. 1997, 272, 3520–3526. [Google Scholar] [CrossRef] [Green Version]
  62. Giulivi, C.; Davies, K.J. Dityrosine and tyrosine oxidation products are endogenous markers for the selective proteolysis of oxidatively modified red blood cell hemoglobin by (the 19 S) proteasome. J. Biol. Chem. 1993, 268, 8752–8759. [Google Scholar] [CrossRef]
  63. Truscott, R.J.W.; Friedrich, M.G. The etiology of human age-related cataract. Proteins don’t last forever. Biochim. Biophys. Acta 2016, 1860, 192–198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Serpell, L.C.; Williams, T.L.; Stewart-Parker, M.; Ford, L.; Skaria, E.; Cole, M.; Bucher, W.G.r.; Morris, K.L.; Sada, A.A.; Thorpe, J.R. A central role for dityrosine crosslinking of Amyloid-β in Alzheimer’s disease. Acta Neuropath Commun. 2013, 1, 83. [Google Scholar] [CrossRef] [Green Version]
  65. Tiwari, M.K.; Leinisch, F.; Sahin, C.; Møller, I.M.; Otzen, D.E.; Davies, M.J.; Bjerrum, M.J. Early events in copper-ion catalyzed oxidation of α-synuclein. Free Radic. Biol. Med. 2018, 121, 38–50. [Google Scholar] [CrossRef] [PubMed]
  66. Pennathur, S.; Jackson-Lewis, V.; Przedborski, S.; Heinecke, J.W. Mass spectrometric quantification of 3-nitrotyrosine, ortho-tyrosine, and o,o′-dityrosine in brain tissue of 1-methyl-4-phenyl-1,2,3,6- tetrahydropyridine-treated mice, a model of oxidative stress in Parkinson’s disease. J. Biol. Chem. 1999, 274, 34621–34628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Kato, Y.; Maruyama, W.; Naoi, M.; Hashizume, Y.; Osawa, T. Immunohistochemical detection of dityrosine in lipofuscin pigments in the aged human brain. FEBS Lett. 1998, 439, 231–234. [Google Scholar] [CrossRef] [Green Version]
  68. Kato, Y.; Dozaki, N.; Nakamura, T.; Kitamoto, N.; Yoshida, A.; Naito, M.; Kitamura, M.; Osawa, T. Quantification of modified tyrosines in healthy and diabetic human urine using liquid chromatography/tandem mass spectrometry. J. Clin. Biochem. Nutr. 2009, 44, 67–78. [Google Scholar] [CrossRef] [Green Version]
  69. Wu, G.R.; Cheserek, M.; Shi, Y.H.; Shen, L.Y.; Yu, J.; Le, G.W. Elevated plasma dityrosine in patients with hyperlipidemia compared to healthy individuals. Ann. Nutr. Metab. 2014, 66, 44–50. [Google Scholar] [CrossRef] [PubMed]
  70. Ziouzenkova, O.; Asatryan, L.; Akmal, M.; Tetta, C.; Wratten, M.L.; Loseto-Wich, G.; Jurgens, G.; Heinecke, J.; Sevanian, A. Oxidative cross-linking of ApoB100 and hemoglobin results in low density lipoprotein modification in blood. Relevance to atherogenesis caused by hemodialysis. J. Biol. Chem. 1999, 274, 18916–18924. [Google Scholar] [CrossRef] [Green Version]
  71. Francis, G.A.; Mendez, A.J.; Bierman, E.L.; Heinecke, J.W. Oxidative tyrosylation of high density lipoprotein by peroxidase enhances cholesterol removal from cultured fibroblasts and macrophage foam cells. Proc. Natl. Acad. Sci. USA 1993, 90, 6631–6635. [Google Scholar] [CrossRef] [Green Version]
  72. Malencik, D.A.; Anderson, S.R. Dityrosine formation in calmodulin: Cross-linking and polymerization catalyzed by Arthromyces peroxidase. Biochemistry 1996, 35, 4375–4386. [Google Scholar] [CrossRef] [PubMed]
  73. Aeschbach, R.; Amado, R.; Neukom, H. Formation of dityrosine cross-links in proteins by oxidation of tyrosine residues. Biochim. Biophys. Acta 1976, 439, 292–301. [Google Scholar] [CrossRef]
  74. Das, A.B.; Nauser, T.; Koppenol, W.H.; Kettle, A.J.; Winterbourn, C.C.; Nagy, P. Rapid reaction of superoxide with insulin-tyrosyl radicals to generate a hydroperoxide with subsequent glutathione addition. Free Radic. Biol. Med. 2014, 70, 86–95. [Google Scholar] [CrossRef]
  75. Gatin, A.; Billault, I.; Duchambon, P.; Van der Rest, G.; Sicard-Roselli, C. Oxidative radicals (HO. or N3.) induce several di-tyrosine bridge isomers at the protein scale. Free Radic. Biol. Med. 2021, 162, 461–470. [Google Scholar] [CrossRef]
  76. Degendorfer, G.; Chuang, C.Y.; Kawasaki, H.; Hammer, A.; Malle, E.; Yamakura, F.; Davies, M.J. Peroxynitrite-mediated oxidation of plasma fibronectin. Free Radic. Biol. Med. 2016, 97, 602–615. [Google Scholar] [CrossRef]
  77. Mai, K.; Smith, N.C.; Feng, Z.P.; Katrib, M.; Šlapeta, J.; Šlapetova, I.; Wallach, M.G.; Luxford, C.; Davies, M.J.; Zhang, X.; et al. Peroxidase catalysed cross-linking of an intrinsically unstructured protein via dityrosine bonds in the oocyst wall of the apicomplexan parasite, Eimeria maxima. Int. J. Parasitol. 2011, 41, 1157–1164. [Google Scholar] [CrossRef] [PubMed]
  78. Medinas, D.B.; Gozzo, F.C.; Santos, L.F.A.; Iglesias, A.H.; Augusto, O. A ditryptophan cross-link is responsible for the covalent dimerization of human superoxide dismutase 1 during its bicarbonate-dependent peroxidase activity. Free Radic. Biol. Med. 2010, 49, 1046–1053. [Google Scholar] [CrossRef]
  79. Paviani, V.; Queiroz, R.F.; Marques, E.F.; Di Mascio, P.; Augusto, O. Production of lysozyme and lysozyme-superoxide dismutase dimers bound by a ditryptophan cross-link in carbonate radical-treated lysozyme. Free Radic. Biol. Med. 2015, 89, 72–82. [Google Scholar] [CrossRef] [PubMed]
  80. Paviani, V.; Junqueira de Melo, P.; Avakin, A.; Di Mascio, P.; Ronsein, G.E.; Augusto, O. Human cataractous lenses contain cross-links produced by crystallin-derived tryptophanyl and tyrosyl radicals. Free Radic. Biol. Med. 2020, 160, 356–367. [Google Scholar] [CrossRef] [PubMed]
  81. Bhaskar, B.; Immoos, C.E.; Shimizu, H.; Sulc, F.; Farmer, P.J.; Poulos, T.L. A novel heme and peroxide-dependent tryptophan-tyrosine cross-link in a mutant of cytochrome c peroxidase. J. Mol. Biol. 2003, 328, 157–166. [Google Scholar] [CrossRef] [Green Version]
  82. Liu, M.; Zhang, Z.; Cheetham, J.; Ren, D.; Zhou, Z.S. Discovery and characterization of a photo-oxidative histidine-histidine cross-link in IgG1 antibody utilizing 18O-labeling and mass spectrometry. Anal. Chem. 2014, 86, 4940–4948. [Google Scholar] [CrossRef]
  83. Powell, T.; Knight, M.J.; O’Hara, J.; Burkitt, W. Discovery of a photoinduced histidine-histidine cross-link in an IgG4 antibody. J. Am. Soc. Mass Spectrom. 2020, 31, 1233–1240. [Google Scholar] [CrossRef] [PubMed]
  84. Xu, C.F.; Chen, Y.; Yi, L.; Brantley, T.; Stanley, B.; Sosic, Z.; Zang, L. Discovery and characterization of histidine oxidation initiated cross-links in an IgG1 monoclonal antibody. Anal. Chem. 2017, 89, 7915–7923. [Google Scholar] [CrossRef]
  85. Agon, V.V.; Bubb, W.A.; Wright, A.; Hawkins, C.L.; Davies, M.J. Sensitizer-mediated photooxidation of histidine residues: Evidence for the formation of reactive side-chain peroxides. Free Radic Biol. Med. 2006, 40, 698–710. [Google Scholar] [CrossRef] [PubMed]
  86. Leinisch, F.; Mariotti, M.; Hägglund, P.; Davies, M.J. Structural and functional changes in RNAse A originating from tyrosine and histidine cross-linking and oxidation. Free Radic. Biol. Med. 2018, 126, 73–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Torosantucci, R.; Sharov, V.S.; Van Beers, M.; Brinks, V.; Schöneich, C.; Jiskoot, W. Identification of oxidation sites and covalent cross-links in metal catalyzed oxidized interferon beta-1a: Potential implications for protein aggregation and immunogenicity. Mol. Pharm. 2013, 10, 2311–2322. [Google Scholar] [CrossRef] [Green Version]
  88. Oka, O.B.V.; Bulleid, N.J. Forming disulfides in the endoplasmic reticulum. Biochim Biophys. Acta Mol. Cell Res. 2013, 1833, 2425–2429. [Google Scholar] [CrossRef] [Green Version]
  89. Nagy, P. Kinetics and mechanisms of thiol–disulfide exchange covering direct substitution and thiol oxidation-mediated pathways. Antioxid. Redox Signal. 2013, 18, 1623–1641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Singh, R.; Whitesides, G.M. Degenerate intermolecular thiolate-disulfide Interchange involving cyclic five-membered disulfides is faster by ~103 than that involving six- or seven-membered disulfides. J. Am. Chem. Soc. 1990, 112, 6304–6309. [Google Scholar] [CrossRef]
  91. Berndt, C.; Lillig, C.H.; Holmgren, A. Thiol-based mechanisms of the thioredoxin and glutaredoxin systems: Implications for diseases in the cardiovascular system. Am. J. Physiol. Heart Circ. Physiol. 2007, 292, H1227–H1236. [Google Scholar] [CrossRef]
  92. Carroll, L.; Jiang, S.; Irnstorfer, J.; Beneyto, S.; Ignasiak, M.T.; Rasmussen, L.M.; Rogowska-Wrzesinska, A.; Davies, M.J. Oxidant-induced glutathionylation at protein disulfide bonds. Free Radic. Biol. Med. 2020, 160, 513–525. [Google Scholar] [CrossRef]
  93. Jiang, S.; Carroll, L.; Mariotti, M.; Hägglund, P.; Davies, M.J. Formation of protein cross-links by singlet oxygen-mediated disulfide oxidation. Redox Biol. 2021, 41, 101874. [Google Scholar] [CrossRef]
  94. Jiang, S.; Hägglund, P.; Carroll, L.; Rasmussen, L.M.; Davies, M.J. Crosslinking of human plasma C-reactive protein to human serum albumin via disulfide bond oxidation. Redox Biol. 2021, 41, 101925. [Google Scholar] [CrossRef]
  95. Yang, J.; Carroll, K.S.; Liebler, D.C. The expanding landscape of the thiol redox proteome. Mol. Cell Proteom. 2016, 15, 1–11. [Google Scholar] [CrossRef] [Green Version]
  96. Gupta, V.; Paritala, H.; Carroll, K.S. Reactivity, selectivity, and stability in sulfenic acid detection: A comparative study of nucleophilic and electrophilic probes. Bioconjug. Chem. 2016, 27, 1411–1418. [Google Scholar] [CrossRef] [Green Version]
  97. Ronsein, G.E.; Winterbourn, C.C.; Di Mascio, P.; Kettle, A.J. Cross-linking methionine and amine residues with reactive halogen species. Free Radic. Biol. Med. 2014, 70, 278–287. [Google Scholar] [CrossRef]
  98. Harwood, D.T.; Kettle, A.J.; Winterbourn, C.C. Production of glutathione sulfonamide and dehydroglutathione from GSH by myeloperoxidase-derived oxidants and detection using a novel LC-MS/MS method. Biochem. J. 2006, 399, 161–168. [Google Scholar] [CrossRef] [Green Version]
  99. Raftery, M.J.; Yang, Z.; Valenzuela, S.M.; Geczy, C.L. Novel intra- and inter-molecular sulfinamide bonds in S100A8 produced by hypochlorite oxidation. J. Biol. Chem. 2001, 276, 33393–33401. [Google Scholar] [CrossRef] [Green Version]
  100. Fu, X.; Mueller, D.M.; Heinecke, J.W. Generation of intramolecular and intermolecular sulfenamides, sulfinamides, and sulfonamides by hypochlorous acid: A potential pathway for oxidative cross-linking of low-density lipoprotein by myeloperoxidase. Biochemistry 2002, 41, 1293–1301. [Google Scholar] [CrossRef]
  101. Pullar, J.M.; Vissers, M.C.; Winterbourn, C.C. Glutathione oxidation by hypochlorous acid in endothelial cells produces glutathione sulfonamide as a major product but not glutathione disulfide. J. Biol. Chem. 2001, 276, 22120–22125. [Google Scholar] [CrossRef] [Green Version]
  102. Carr, A.C.; Winterbourn, C.C. Oxidation of neutrophil glutathione and protein thiols by myeloperoxidase-derived hypochlorous acid. Biochem. J. 1997, 327, 275–281. [Google Scholar] [CrossRef]
  103. Harwood, D.T.; Kettle, A.J.; Brennan, S.; Winterbourn, C.C. Simultaneous determination of reduced glutathione, glutathione disulphide and glutathione sulphonamide in cells and physiological fluids by isotope dilution liquid chromatography-tandem mass spectrometry. J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 2009, 877, 3393–3399. [Google Scholar] [CrossRef] [PubMed]
  104. Kettle, A.J.; Turner, R.; Gangell, C.L.; Harwood, D.T.; Khalilova, I.S.; Chapman, A.L.; Winterbourn, C.C.; Sly, P.D. Oxidation contributes to low glutathione in the airways of children with cystic fibrosis. Eur. Respir. J. 2014, 44, 122–129. [Google Scholar] [CrossRef] [Green Version]
  105. Winterbourn, C.C.; Brennan, S.O. Characterization of the oxidation products of the reaction between reduced glutathione and hypochlorous acid. Biochem. J. 1997, 326, 87–92. [Google Scholar] [CrossRef] [Green Version]
  106. Erve, J.C.; Svensson, M.A.; von Euler-Chelpin, H.; Klasson-Wehler, E. Characterization of glutathione conjugates of the remoxipride hydroquinone metabolite NCQ-344 formed in vitro and detection following oxidation by human neutrophils. Chem. Res. Toxicol. 2004, 17, 564–571. [Google Scholar] [CrossRef] [PubMed]
  107. Shetty, V.; Spellman, D.S.; Neubert, T.A. Characterization by tandem mass spectrometry of stable cysteine sulfenic acid in a cysteine switch peptide of matrix metalloproteinases. J. Am. Soc. Mass Spectrom 2007, 18, 1544–1551. [Google Scholar] [CrossRef] [Green Version]
  108. Winterbourn, C.C. Comparative reactivities of various biological compounds with myeloperoxidase-hydrogen peroxide-chloride, and similarity of the oxidant to hypochlorite. Biochim. Biophys. Acta 1985, 840, 204–210. [Google Scholar] [CrossRef]
  109. Mozziconacci, O.; Kerwin, B.A.; Schoneich, C. Reversible hydrogen transfer reactions of cysteine thiyl radicals in peptides: The conversion of cysteine into dehydroalanine and alanine, and of alanine into dehydroalanine. J. Phys. Chem. B 2011, 115, 12287–12305. [Google Scholar] [CrossRef] [Green Version]
  110. Shu, N.; Lorentzen, L.G.; Davies, M.J. Reaction of quinones with proteins: Kinetics of adduct formation, effects on enzymatic activity and protein structure, and potential reversibility of modifications. Free Radic. Biol. Med. 2019, 137, 169–180. [Google Scholar] [CrossRef] [PubMed]
  111. Sauerland, M.; Mertes, R.; Morozzi, C.; Eggler, A.L.; Gamon, L.F.; Davies, M.J. Kinetic assessment of Michael addition reactions of alpha, beta-unsaturated carbonyl compounds to amino acid and protein thiols. Free Radic. Biol. Med. 2021, 169, 1–11. [Google Scholar] [CrossRef] [PubMed]
  112. Li, W.W.; Heinze, J.; Haehnel, W. Site-specific binding of quinones to proteins through thiol addition and addition-elimination reactions. J. Am. Chem. Soc. 2005, 127, 6140–6141. [Google Scholar] [CrossRef]
  113. Miura, T.; Kakehashi, H.; Shinkai, Y.; Egara, Y.; Hirose, R.; Cho, A.K.; Kumagai, Y. GSH-mediated S-transarylation of a quinone glyceraldehyde-3-phosphate dehydrogenase conjugate. Chem. Res. Toxicol. 2011, 24, 1836–1844. [Google Scholar] [CrossRef]
  114. Truscott, R.J. Age-related nuclear cataract-oxidation is the key. Exp. Eye Res. 2005, 80, 709–725. [Google Scholar] [CrossRef] [PubMed]
  115. McIntire, W.S. Quinoproteins. FASEB J. 1994, 8, 513–521. [Google Scholar] [CrossRef]
  116. Nappi, A.J.; Vass, E. The effects of glutathione and ascorbic acid on the oxidations of 6-hydroxydopa and 6-hydroxydopamine. Biochim. Biophys. Acta 1994, 1201, 498–504. [Google Scholar] [CrossRef]
  117. Datta, S.; Mori, Y.; Takagi, K.; Kawaguchi, K.; Chen, Z.W.; Okajima, T.; Kuroda, S.; Ikeda, T.; Kano, K.; Tanizawa, K.; et al. Structure of a quinohemoprotein amine dehydrogenase with an uncommon redox cofactor and highly unusual crosslinking. Proc. Natl. Acad. Sci. USA 2001, 98, 14268–14273. [Google Scholar] [CrossRef] [Green Version]
  118. Davidson, V.L.; Wilmot, C.M. Posttranslational biosynthesis of the protein-derived cofactor tryptophan tryptophylquinone. Ann. Rev. Biochem. 2013, 82, 531–550. [Google Scholar] [CrossRef] [Green Version]
  119. Mozziconacci, O.; Williams, T.D.; Kerwin, B.A.; Schoneich, C. Reversible intramolecular hydrogen transfer between protein cysteine thiyl radicals and alpha C-H bonds in insulin: Control of selectivity by secondary structure. J. Phys. Chem. B 2008, 112, 15921–15932. [Google Scholar] [CrossRef] [PubMed]
  120. Schoneich, C. Thiyl radical reactions in the chemical degradation of pharmaceutical proteins. Molecules 2019, 24, 4357. [Google Scholar] [CrossRef] [Green Version]
  121. Mozziconacci, O.; Haywood, J.; Gorman, E.M.; Munson, E.; Schoneich, C. Photolysis of recombinant human insulin in the solid state: Formation of a dithiohemiacetal product at the C-terminal disulfide bond. Pharm. Res. 2012, 29, 121–133. [Google Scholar] [CrossRef] [PubMed]
  122. Orian, L.; Mauri, P.; Roveri, A.; Toppo, S.; Benazzi, L.; Bosello-Travain, V.; De Palma, A.; Maiorino, M.; Miotto, G.; Zaccarin, M.; et al. Selenocysteine oxidation in glutathione peroxidase catalysis: An MS-supported quantum mechanics study. Free Radic. Biol. Med. 2015, 87, 1–14. [Google Scholar] [CrossRef]
  123. Mauri, P.; Benazzi, L.; Flohe, L.; Maiorino, M.; Pietta, P.G.; Pilawa, S.; Roveri, A.; Ursini, F. Versatility of selenium catalysis in PHGPx unraveled by LC/ESI-MS/MS. Biol. Chem. 2003, 384, 575–588. [Google Scholar] [CrossRef] [PubMed]
  124. Zhong, L.; Arner, E.S.; Holmgren, A. Structure and mechanism of mammalian thioredoxin reductase: The active site is a redox-active selenolthiol/selenenylsulfide formed from the conserved cysteine-selenocysteine sequence. Proc. Natl. Acad. Sci. USA 2000, 97, 5854–5859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Heinecke, J.W.; Shapiro, B.M. Respiratory burst oxidase of fertilization. Proc. Natl. Acad. Sci. USA 1989, 86, 1259–1263. [Google Scholar] [CrossRef] [Green Version]
  126. Sies, H.; Berndt, C.; Jones, D.P. Oxidative Stress. Ann. Rev. Biochem. 2017, 86, 715–748. [Google Scholar] [CrossRef]
  127. Sies, H. Oxidative stress: Concept and some practical aspects. Antioxidants 2020, 9, 852. [Google Scholar] [CrossRef]
  128. Hellwig, M. The chemistry of protein oxidation in food. Angew. Chem. Int. Ed. Engl. 2019, 58, 16742–16763. [Google Scholar] [CrossRef]
  129. Xiong, Y.L.; Guo, A. Animal and plant protein oxidation: Chemical and functional property significance. Foods 2020, 10, 40. [Google Scholar] [CrossRef] [PubMed]
  130. Houée-Lévin, C.; Bobrowski, K.; Horakova, L.; Karademir, B.; Schöneich, C.; Davies, M.J.; Spickett, C.M. Exploring oxidative modifications of tyrosine: An update on mechanisms of formation, advances in analysis and biological consequences. Free Radic. Res. 2015, 49, 347–373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Gross, A.J.; Sizer, I.W. The oxidation of tyramine, tyrosine, and related compounds by peroxidase. J. Biol. Chem. 1959, 234, 1611–1614. [Google Scholar] [CrossRef]
  132. Liu, C.; Hua, J.; Ng, P.F.; Fei, B. Photochemistry of bioinspired dityrosine crosslinking. J. Mater. Sci. Technol. 2021, 63, 182–191. [Google Scholar] [CrossRef]
  133. Heinecke, J.W.; Li, W.; Francis, G.A.; Goldstein, J.A. Tyrosyl radical generated by myeloperoxidase catalyzes the oxidative cross-linking of proteins. J. Clin. Investig. 1993, 91, 2866–2872. [Google Scholar] [CrossRef]
  134. Giulivi, C.; Traaseth, N.J.; Davies, K.J. Tyrosine oxidation products: Analysis and biological relevance. Amino Acids 2003, 25, 227–232. [Google Scholar] [CrossRef] [PubMed]
  135. Karmakar, S.; Datta, A. Understanding the reactivity of CO3.- and NO2. radicals toward S-containing and aromatic amino acids. J. Phys. Chem. B 2017, 121, 7621–7632. [Google Scholar] [CrossRef] [PubMed]
  136. Winterbourn, C.C.; Parsons-Mair, H.N.; Gebicki, S.; Gebicki, J.M.; Davies, M.J. Requirements for superoxide-dependent tyrosine hydroperoxide formation in peptides. Biochem. J. 2004, 381, 241–248. [Google Scholar] [CrossRef] [Green Version]
  137. Folkes, L.K.; Bartesaghi, S.; Trujillo, M.; Wardman, P.; Radi, R. The effects of nitric oxide or oxygen on the stable products formed from the tyrosine phenoxyl radical. Free Radic. Res. 2021, 55, 141–153. [Google Scholar] [CrossRef]
  138. Gebicki, J.M.; Nauser, T.; Domazou, A.; Steinmann, D.; Bounds, P.L.; Koppenol, W.H. Reduction of protein radicals by GSH and ascorbate: Potential biological significance. Amino Acids 2010, 39, 1131–1137. [Google Scholar] [CrossRef] [PubMed]
  139. Nauser, T.; Koppenol, W.H.; Gebicki, J.M. The kinetics of oxidation of GSH by protein radicals. Biochem. J. 2005, 392, 693–701. [Google Scholar] [CrossRef] [PubMed]
  140. Nauser, T.; Gebicki, J.M. Antioxidants and radical damage in a hydrophilic environment: Chemical reactions and concepts. Essays Biochem. 2020, 64, 67–74. [Google Scholar] [CrossRef] [Green Version]
  141. Mukherjee, S.; Kapp, E.A.; Lothian, A.; Roberts, A.M.; Vasil’Ev, Y.V.; Boughton, B.A.; Barnham, K.J.; Kok, W.M.; Hutton, C.A.; Masters, C.L.; et al. Characterization and identification of dityrosine cross-linked peptides using tandem mass spectrometry. Anal. Chem. 2017, 89, 6136–6145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Tew, D.; Ortiz de Montellano, P.R. The myoglobin protein radical. Coupling of Tyr-103 to Tyr-151 in the H2O2-mediated cross-linking of sperm whale myoglobin. J. Biol. Chem. 1988, 263, 17880–17886. [Google Scholar] [CrossRef]
  143. Das, A.B.; Nagy, P.; Abbott, H.F.; Winterbourn, C.C.; Kettle, A.J. Reactions of superoxide with the myoglobin tyrosyl radical. Free. Radic. Biol. Med. 2010, 48, 1540–1547. [Google Scholar] [CrossRef]
  144. Kehm, R.; Baldensperger, T.; Raupbach, J.; Hohn, A. Protein oxidation—Formation mechanisms, detection and relevance as biomarkers in human diseases. Redox Biol. 2021, 42, 101901. [Google Scholar] [CrossRef]
  145. DiMarco, T.; Giulivi, C. Current analytical methods for the detection of dityrosine, a biomarker of oxidative stress, in biological samples. Mass Spectrom. Rev. 2007, 26, 108–120. [Google Scholar] [CrossRef] [PubMed]
  146. Chen, Z.; Leinisch, F.; Greco, I.; Zhang, W.; Shu, N.; Chuang, C.Y.; Lund, M.N.; Davies, M.J. Characterisation and quantification of protein oxidative modifications and amino acid racemisation in powdered infant milk formula. Free Radic. Res. 2019, 53, 68–81. [Google Scholar] [CrossRef] [PubMed]
  147. Fenaille, F.; Parisod, V.; Vuichoud, J.; Tabet, J.C.; Guy, P.A. Quantitative determination of dityrosine in milk powders by liquid chromatography coupled to tandem mass spectrometry using isotope dilution. J. Chromatogr. A 2004, 1052, 77–84. [Google Scholar] [CrossRef] [PubMed]
  148. Rodriguez-Mateos, A.; Millar, S.J.; Bhandari, D.G.; Frazier, R.A. Formation of dityrosine cross-links during breadmaking. J. Agric. Food Chem. 2006, 54, 2761–2766. [Google Scholar] [CrossRef] [PubMed]
  149. Ma, J.B.; Wang, X.Y.; Li, Q.; Zhang, L.; Wang, Z.; Han, L.; Yu, Q.L. Oxidation of myofibrillar protein and crosslinking behavior during processing of traditional air-dried yak (Bos grunniens) meat in relation to digestibility. LWT Food Sci. Technol. 2021, 142, 110984. [Google Scholar] [CrossRef]
  150. Ma, L.; Li, A.L.; Li, T.Q.; Li, M.; Wang, X.D.; Hussain, M.A.; Qayum, A.; Jiang, Z.M.; Hou, J.C. Structure and characterization of laccase-crosslinked alpha-lactalbumin: Impacts of high pressure homogenization pretreatment. LWT Food Sci. Technol. 2020, 118, 108843. [Google Scholar] [CrossRef]
  151. Yang, Y.; Zhang, H.; Yan, B.; Zhang, T.; Gao, Y.; Shi, Y.; Le, G. Health effects of dietary oxidized tyrosine and dityrosine administration in mice with nutrimetabolomic strategies. J. Agric. Food Chem. 2017, 65, 6957–6971. [Google Scholar] [CrossRef] [PubMed]
  152. Ding, Y.Y.; Tang, X.; Cheng, X.R.; Wang, F.F.; Li, Z.Q.; Wu, S.J.; Kou, X.R.; Shi, Y.H.; Le, G.W. Effects of dietary oxidized tyrosine products on insulin secretion via the thyroid hormone T3-regulated TR beta 1-Akt-mTOR pathway in the pancreas. RSC Adv. 2017, 7, 54610–54625. [Google Scholar] [CrossRef] [Green Version]
  153. Li, Z.L.; Shi, Y.; Ding, Y.; Ran, Y.; Le, G. Dietary oxidized tyrosine (O-Tyr) stimulates TGF-beta1-induced extracellular matrix production via the JNK/p38 signaling pathway in rat kidneys. Amino Acids 2017, 49, 241–260. [Google Scholar] [CrossRef] [PubMed]
  154. Li, B.W.; Mo, L.; Yang, Y.H.; Zhang, S.; Xu, J.B.; Ge, Y.T.; Xu, Y.C.; Shi, Y.H.; Le, G.W. Processing milk causes the formation of protein oxidation products which impair spatial learning and memory in rats. RSC Adv. 2019, 9, 22161–22175. [Google Scholar] [CrossRef] [Green Version]
  155. Schaefer, J.; Kramer, K.J.; Garbow, J.R.; Jacob, G.S.; Stejskal, E.O.; Hopkins, T.L.; Speirs, R.D. Aromatic cross-links in insect cuticle: Detection by solid-state 13C and 15N NMR. Science 1987, 235, 1200–1204. [Google Scholar] [CrossRef]
  156. Bellmaine, S.; Schnellbaecher, A.; Zimmer, A. Reactivity and degradation products of tryptophan in solution and proteins. Free Radic. Biol. Med. 2020, 160, 696–718. [Google Scholar] [CrossRef]
  157. Coelho, F.R.; Iqbal, A.; Linares, E.; Silva, D.F.; Lima, F.S.; Cuccovia, I.M.; Augusto, O. Oxidation of the tryptophan 32 residue of human superoxide dismutase 1 caused by its bicarbonate-dependent peroxidase activity triggers the non-amyloid aggregation of the enzyme. J. Biol. Chem. 2014, 289, 30690–30701. [Google Scholar] [CrossRef] [Green Version]
  158. Figueroa, J.D.; Zarate, A.M.; Fuentes-Lemus, E.; Davies, M.J.; López-Alarcón, C. Formation and characterization of crosslinks, including Tyr-Trp species, on one electron oxidation of free Tyr and Trp residues by carbonate radical anion. RSC Adv. 2020, 10, 25786–25800. [Google Scholar] [CrossRef]
  159. Carroll, L.; Pattison, D.I.; Davies, J.B.; Anderson, R.F.; López-Alarcón, C.; Davies, M.J. Formation and detection of oxidant-generated tryptophan dimers in peptides and proteins. Free Radic. Biol. Med. 2017, 113, 132–142. [Google Scholar] [CrossRef]
  160. Zhuravleva, Y.S.; Sherin, P.S. Influence of pH on radical reactions between kynurenic acid and amino acids tryptophan and tyrosine. Part II. Amino acids within the protein globule of lysozyme. Free Radic. Biol. Med. 2021, 174, 211–224. [Google Scholar] [CrossRef]
  161. Sormacheva, E.D.; Sherin, P.S.; Tsentalovich, Y.P. Dimerization and oxidation of tryptophan in UV-A photolysis sensitized by kynurenic acid. Free Radic. Biol. Med. 2017, 113, 372–384. [Google Scholar] [CrossRef]
  162. Silva, E.; Barrias, P.; Fuentes-Lemus, E.; Tirapegui, C.; Aspee, A.; Carroll, L.; Davies, M.J.; López-Alarcón, C. Riboflavin-induced Type 1 photo-oxidation of tryptophan using a high intensity 365nm light emitting diode. Free Radic. Biol. Med. 2019, 131, 133–143. [Google Scholar] [CrossRef]
  163. Sherin, P.S.; Zelentsova, E.A.; Sormacheva, E.D.; Yanshole, V.V.; Duzhak, T.G.; Tsentalovich, Y.P. Aggregation of α-crystallins in kynurenic acid-sensitized UVA photolysis under anaerobic conditions. Phys. Chem. Chem. Phys. 2016, 18, 8827–8839. [Google Scholar] [CrossRef] [PubMed]
  164. Reyes, J.S.; Fuentes-Lemus, E.; Aspee, A.; Davies, M.J.; Monasterio, O.; López-Alarcón, C.M. Jannaschii FtsZ, a key protein in bacterial cell division, is inactivated by peroxyl radical-mediated methionine oxidation. Free Radic. Biol. Med. 2021, 166, 53–66. [Google Scholar] [CrossRef]
  165. Barik, S. The uniqueness of tryptophan in biology: Properties, metabolism, interactions and localization in proteins. Int. J. Mol. Sci. 2020, 21, 8776. [Google Scholar] [CrossRef] [PubMed]
  166. Leo, G.; Altucci, C.; Bourgoin-Voillard, S.; Gravagnuolo, A.M.; Esposito, R.; Marino, G.; Costello, C.E.; Velotta, R.; Birolo, L. Ultraviolet laser-induced cross-linking in peptides. Rapid Commun. Mass Spectrom. 2013, 27, 1660–1668. [Google Scholar] [CrossRef] [Green Version]
  167. Mariotti, M.; Leinisch, F.; Leeming, D.J.; Svensson, B.; Davies, M.J.; Hägglund, P. Mass-spectrometry-based identification of cross-links in proteins exposed to photo-oxidation and peroxyl radicals using 18O labeling and optimized tandem mass spectrometry fragmentation. J. Proteome Res. 2018, 17, 2017–2027. [Google Scholar] [CrossRef]
  168. Gebicki, J.M. Protein hydroperoxides as new reactive oxygen species. Redox Rep. 1997, 3, 99–110. [Google Scholar] [CrossRef] [PubMed]
  169. Hanson, D.A.; Eyre, D.R. Molecular site specificity of pyridinoline and pyrrole cross-links in type I collagen of human bone. J. Biol. Chem. 1996, 271, 26508–26516. [Google Scholar] [CrossRef] [Green Version]
  170. Wilhelmus, M.M.M.; Grunberg, S.C.S.; Bol, J.G.J.M.; Van Dam, A.M.; Hoozemans, J.J.M.; Rozemuller, A.J.M.; Drukarch, B. Transglutaminases and transglutaminase-catalyzed cross-links colocalize with the pathological lesions in Alzheimer’s disease brain. Brain Pathol. 2009, 19, 612–622. [Google Scholar] [CrossRef]
  171. De Jong, G.A.H.; Koppelman, S.J. Transglutaminase catalyzed reactions: Impact on food applications. J. Food Sci. 2002, 67, 2798–2806. [Google Scholar] [CrossRef]
  172. Griffin, M.; Casadio, R.; Bergamini, C.M. Transglutaminases: Nature’s biological glues. Biochem. J. 2002, 368, 377–396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Shen, H.-R.; Spikes, J.D.; Smith, C.J.; Kopeček, J. Photodynamic cross-linking of proteins: IV. Nature of the His-His bond(s) formed in the rose bengal-photosensitized cross-linking of N-benzoyl-L-histidine. J. PhotoChem. PhotoBiol. A Chem. 2000, 130, 1–6. [Google Scholar] [CrossRef]
  174. Tanzer, M.L.; Housley, T.; Berube, L.; Fairweather, R.; Franzblau, C.; Gallop, P.M. Structure of two histidine-containing crosslinks from collagen. J. Biol. Chem. 1973, 248, 393–402. [Google Scholar] [CrossRef]
  175. Li, Y.; Jongberg, S.; Andersen, M.L.; Davies, M.J.; Lund, M.N. Quinone-induced protein modifications: Kinetic preference for reaction of 1,2-benzoquinones with thiol groups in proteins. Free Radic. Biol. Med. 2016, 97, 148–157. [Google Scholar] [CrossRef] [PubMed]
  176. Marmelstein, A.M.; Lobba, M.J.; Mogilevsky, C.S.; Maza, J.C.; Brauer, D.D.; Francis, M.B. Tyrosinase-mediated oxidative coupling of tyrosine tags on peptides and proteins. J. Am. Chem. Soc. 2020, 142, 5078–5086. [Google Scholar] [CrossRef]
  177. Rzepecki, L.M.; Nagafuchi, T.; Waite, J.H. α,β-Dehydro-3,4-dihydroxyphenylalanine derivatives: Potential schlerotization intermediates in natural composite materials. Arch. Biochem. Biophys. 1991, 285, 17–26. [Google Scholar] [CrossRef]
  178. Burzio, L.A.; Waite, J.H. Cross-linking in adhesive proteins: Studies with model decapeptides. Biochemistry 2000, 39, 11147–11153. [Google Scholar] [CrossRef]
  179. Rzepecki, L.M.; Waite, J.H. Wresting the muscle from mussel beards: Research and applications. Mol. Mar. Biol. Biotechnol. 1995, 4, 313–322. [Google Scholar]
  180. Mishler-Elmore, J.W.; Zhou, Y.; Sukul, A.; Oblak, M.; Tan, L.; Faik, A.; Held, M.A. Extensins: Self-assembly, crosslinking, and the role of peroxidases. Front. Plant. Sci. 2021, 12, 664738. [Google Scholar] [CrossRef]
  181. Breusing, N.; Grune, T. Biomarkers of protein oxidation from a chemical, biological and medical point of view. Exp. Gerontol. 2010, 45, 733–737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Lopezllorca, L.V.; Fry, S.C. Dityrosine, trityrosine and tetratyrosine, potential cross-links in structural proteins of plant-parasitic nematodes. Nematologica 1989, 35, 165–179. [Google Scholar] [CrossRef]
  183. Dhayal, S.K.; Sforza, S.; Wierenga, P.A.; Gruppen, H. Peroxidase induced oligo-tyrosine cross-links during polymerization of alpha-lactalbumin. Biochim Biophys. Acta Proteins Proteom. 2015, 1854, 1898–1905. [Google Scholar] [CrossRef] [PubMed]
  184. Reid, L.O.; Vignoni, M.; Martins-Froment, N.; Thomas, A.H.; Dantola, M.L. Photochemistry of tyrosine dimer: When an oxidative lesion of proteins is able to photoinduce further damage. PhotoChem. PhotoBiol. Sci. 2019, 18, 1732–1741. [Google Scholar] [CrossRef]
  185. Sommer, A.; Traut, R.R. Diagonal polyacrylamide-dodecyl sulfate gel electrophoresis for the identification of ribosomal proteins crosslinked with methyl-4-mercaptobutyrimidate. Proc. Natl. Acad. Sci. USA 1974, 71, 3946–3950. [Google Scholar] [CrossRef] [Green Version]
  186. Al-Hilaly, Y.K.; Biasetti, L.; Blakeman, B.J.F.; Pollack, S.J.; Zibaee, S.; Abdul-Sada, A.; Thorpe, J.R.; Xue, W.-F.; Serpell, L.C.; Goedert, M.; et al. The involvement of dityrosine crosslinking in α-synuclein assembly and deposition in Lewy bodies in Parkinson’s disease. Sci. Rep. 2016, 6, 39171. [Google Scholar] [CrossRef]
  187. Kato, Y.; Wu, X.; Naito, M.; Nomura, H.; Kitamoto, N.; Osawa, T. Immunochemical detection of protein dityrosine in atherosclerotic lesion of apo-E-deficient mice using a novel monoclonal antibody. Biochem. Biophys. Res. Commun. 2000, 275, 11–15. [Google Scholar] [CrossRef]
  188. Degendorfer, G.; Chuang, C.Y.; Mariotti, M.; Hammer, A.; Hoefler, G.; Hägglund, P.; Malle, E.; Wise, S.G.; Davies, M.J. Exposure of tropoelastin to peroxynitrous acid gives high yields of nitrated tyrosine residues, di-tyrosine cross-links and altered protein structure and function. Free Radic. Biol. Med. 2017, 115, 219–231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Huang, Y.R.; Hua, Y.F.; Qiu, A.Y. Soybean protein aggregation induced by lipoxygenase catalyzed linoleic acid oxidation. Food Res. Int. 2006, 39, 240–249. [Google Scholar] [CrossRef]
  190. Cui, X.H.; Xiong, Y.L.L.; Kong, B.H.; Zhao, X.H.; Liu, N. Hydroxyl radical-stressed whey protein isolate: Chemical and structural properties. Food Bioprocess. Technol. 2012, 5, 2454–2461. [Google Scholar] [CrossRef]
  191. Hawkins, C.L.; Morgan, P.E.; Davies, M.J. Quantification of protein modification by oxidants. Free Radic. Biol. Med. 2009, 46, 965–988. [Google Scholar] [CrossRef] [PubMed]
  192. Hawkins, C.L.; Davies, M.J. Detection, identification and quantification of oxidative protein modifications. J. Biol. Chem. 2019, 294, 19683–19708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Gamon, L.F.; Guo, C.; He, J.; Hägglund, P.; Hawkins, C.L.; Davies, M.J. Absolute quantitative analysis of intact and oxidized amino acids by LC-MS without prior derivatization. Redox Biol. 2020, 36, 101586. [Google Scholar] [CrossRef]
  194. Desmons, A.; Thioulouse, E.; Hautem, J.Y.; Saintier, A.; Baudin, B.; Lamaziere, A.; Netter, C.; Moussa, F. Direct liquid chromatography tandem mass spectrometry analysis of amino acids in human plasma. J. Chromatogr. A 2020, 1622, 461135. [Google Scholar] [CrossRef] [PubMed]
  195. Hensley, K.; Maidt, M.L.; Yu, Z.; Sang, H.; Markesbery, W.R.; Floyd, R.A. Electrochemical analysis of protein nitrotyrosine and dityrosine in the Alzheimer brain indicates region-specific accumulation. J. NeuroSci. 1998, 18, 8126–8132. [Google Scholar] [CrossRef]
  196. Verzini, S.; Shah, M.; Theillet, F.X.; Belsom, A.; Bieschke, J.; Wanker, E.E.; Rappsilber, J.; Binolfi, A.; Selenko, P. Megadalton-sized dityrosine aggregates of alpha-synuclein retain high degrees of structural disorder and internal dynamics. J. Mol. Biol. 2020, 432, 166689. [Google Scholar] [CrossRef]
  197. Thorn, D.C.; Grosas, A.B.; Mabbitt, P.D.; Ray, N.J.; Jackson, C.J.; Carver, J.A. The structure and stability of the disulfide-linked gammaS-crystallin dimer provide insight into oxidation products associated with lens cataract formation. J. Mol. Biol. 2019, 431, 483–497. [Google Scholar] [CrossRef]
  198. Evrard, C.; Capron, A.; Marchand, C.; Clippe, A.; Wattiez, R.; Soumillion, P.; Knoops, B.; Declercq, J.P. Crystal structure of a dimeric oxidized form of human peroxiredoxin 5. J. Mol. Biol. 2004, 337, 1079–1090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Smeets, A.; Evrard, C.; Landtmeters, M.; Marchand, C.; Knoops, B.; Declercq, J.P. Crystal structures of oxidized and reduced forms of human mitochondrial thioredoxin 2. Protein Sci. 2005, 14, 2610–2621. [Google Scholar] [CrossRef] [Green Version]
  200. Marvin, L.F.; Delatour, T.; Tavazzi, I.; Fay, L.B.; Cupp, C.; Guy, P.A. Quantification of o,o’-dityrosine, o-nitrotyrosine, and o-tyrosine in cat urine samples by LC/ electrospray ionization-MS/MS using isotope dilution. Anal. Chem. 2003, 75, 261–267. [Google Scholar] [CrossRef]
  201. Abdelrahim, M.; Morris, E.; Carver, J.; Facchina, S.; White, A.; Verma, A. Liquid chromatographic assay of dityrosine in human cerebrospinal fluid. J. Chromatogr. B Biomed. Sci. 1997, 696, 175–182. [Google Scholar] [CrossRef]
  202. Rose, K.; Savoy, L.A.; Simona, M.G.; Offord, R.E.; Wingfield, P. C-terminal peptide identification by fast atom bombardment mass spectrometry. Biochem. J. 1988, 250, 253–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Liu, M.; Zhang, Z.; Zang, T.; Spahr, C.; Cheetham, J.; Ren, D.; Zhou, Z.S. Discovery of undefined protein cross-linking chemistry: A comprehensive methodology utilizing 18O-labeling and mass spectrometry. Anal. Chem. 2013, 85, 5900–5908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Cramer, C.N.; Kelstrup, C.D.; Olsen, J.V.; Haselmann, K.F.; Nielsen, P.K. Generic workflow for mapping of complex disulfide bonds using in-source reduction and extracted ion chromatograms from data-dependent mass spectrometry. Anal. Chem. 2018, 90, 8202–8210. [Google Scholar] [CrossRef]
  205. Massonnet, P.; Haler, J.R.N.; Upert, G.; Smargiasso, N.; Mourier, G.; Gilles, N.; Quinton, L.; De Pauw, E. Disulfide connectivity analysis of peptides bearing two intramolecular disulfide bonds using MALDI in-source decay. J. Am. Soc. Mass Spectrom. 2018, 29, 1995–2002. [Google Scholar] [CrossRef] [Green Version]
  206. Hägglund, P.; Bunkenborg, J.; Maeda, K.; Svensson, B. Identification of thioredoxin disulfide targets using a quantitative proteomics approach based on isotope-coded affinity tags. J. Proteome Res. 2008, 7, 5270–5276. [Google Scholar] [CrossRef]
  207. Gomez-Cabrera, M.C.; Carretero, A.; Millan-Domingo, F.; Garcia-Dominguez, E.; Correas, A.G.; Olaso-Gonzalez, G.; Vina, J. Redox-related biomarkers in physical exercise. Redox. Biol. 2021, 42, 101956. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of crosslinks formed on proteins, their nature and mechanisms of formation.
Figure 1. Overview of crosslinks formed on proteins, their nature and mechanisms of formation.
Molecules 27 00015 g001
Figure 2. Mechanisms of traditional ‘thiol–disulfide exchange’ (top) and ‘oxidant-mediated thiol–disulfide exchange’ (bottom) reactions.
Figure 2. Mechanisms of traditional ‘thiol–disulfide exchange’ (top) and ‘oxidant-mediated thiol–disulfide exchange’ (bottom) reactions.
Molecules 27 00015 g002
Figure 3. Generation of crosslinks via oxidized thiol residues. Similar reactions of the ‘activated’ thiols (RS–OH, RS–Cl, RS–Br, RS–SCN, RS–NO) can occur with nitrogen nucleophiles (e.g., RNH2) to give new S–N bonded species (see text for further details).
Figure 3. Generation of crosslinks via oxidized thiol residues. Similar reactions of the ‘activated’ thiols (RS–OH, RS–Cl, RS–Br, RS–SCN, RS–NO) can occur with nitrogen nucleophiles (e.g., RNH2) to give new S–N bonded species (see text for further details).
Molecules 27 00015 g003
Figure 4. Michael addition reactions of nucleophiles to αβ-unsaturated carbonyl compounds.
Figure 4. Michael addition reactions of nucleophiles to αβ-unsaturated carbonyl compounds.
Molecules 27 00015 g004
Figure 5. Formation and reactions of Tyr phenoxyl radicals (Tyr). Tyr self-react to produce di-Tyr (o,o’-di-Tyr, red; iso-di-Tyr, black) or react with O2 to generate oxygenated products. Kinetic data from [130].
Figure 5. Formation and reactions of Tyr phenoxyl radicals (Tyr). Tyr self-react to produce di-Tyr (o,o’-di-Tyr, red; iso-di-Tyr, black) or react with O2 to generate oxygenated products. Kinetic data from [130].
Molecules 27 00015 g005
Figure 6. Formation and reactions of Trp indolyl radicals (Trp). Self-reactions of Trp produce carbon–carbon (C3–C3) and carbon–nitrogen (C3–N1) di-Trp crosslinks. It should be noted that multiple stereoisomers are potentially formed for both di-Trp dimers. Kinetic constants for self-reactions of Trp and their reaction with O2 are from [159] and [36], respectively.
Figure 6. Formation and reactions of Trp indolyl radicals (Trp). Self-reactions of Trp produce carbon–carbon (C3–C3) and carbon–nitrogen (C3–N1) di-Trp crosslinks. It should be noted that multiple stereoisomers are potentially formed for both di-Trp dimers. Kinetic constants for self-reactions of Trp and their reaction with O2 are from [159] and [36], respectively.
Molecules 27 00015 g006
Figure 7. Michael addition reactions of amino acid side-chains to oxidized His and Tyr residues.
Figure 7. Michael addition reactions of amino acid side-chains to oxidized His and Tyr residues.
Molecules 27 00015 g007
Figure 8. Overview of methods to detect and characterize crosslinked proteins and the sites/types of modifications. Abbreviations used: CD: circular dichroism, SANS: small angle neutron scattering, SAXS: small angle X-ray scattering, H–D: hydrogen–deuterium exchange mass spectrometry.
Figure 8. Overview of methods to detect and characterize crosslinked proteins and the sites/types of modifications. Abbreviations used: CD: circular dichroism, SANS: small angle neutron scattering, SAXS: small angle X-ray scattering, H–D: hydrogen–deuterium exchange mass spectrometry.
Molecules 27 00015 g008
Table 1. Examples of major non-disulfide protein crosslinks generated during non-enzymatic oxidative processes and methodologies employed to characterize them.
Table 1. Examples of major non-disulfide protein crosslinks generated during non-enzymatic oxidative processes and methodologies employed to characterize them.
Crosslinked ResiduesProtein(s)Chemical Nature and/or Mechanism of Formation of the CrosslinkMethod(s)Refs
Tyr-Cys
a)
Myoglobin
b)
Galactose oxidase
c)
Cysteine dioxygenase
1)
Michael addition from thiols (Cys) to oxidized Tyr species (a)
2)
Thioether bridge (C-S) (b and c)
Mass spectrometry (a)
X-ray crystallography (b, c)
[42,43,44]
Trp-CysHuman growth hormone (hGH)
1)
Michael addition from N (Trp indole) to DHA (formed from Cys)
2)
Thioether bridge (C-S)
Mass spectrometry[41]
Met-Hydroxy-lysineCollagen IVFormation of S=N bridge (sulfilimine bond) induced by peroxidasin/HOBrMass spectrometry[17]
Lys-CysTransaldolaseNitrogen–oxygen–sulfur (NOS) link/redox switchX-ray crystallography[45]
Cys-Ser
a)
Human growth hormone
b)
Tyrosine phosphatase 1B
1)
Formation of a vinyl ether between Ser and Cys that result in the elimination of the thiol group from Cys (a)
2)
Sulfenyl amide (S–N bridge) between Cys-OH and main-chain amide of Ser residue (b)
Mass spectrometry (a)
X-ray crystallography (b)
[41,46]
Cys-PhehGHCrosslink between thioaldehyde from Cys and dehydrophenylalanine generated from PheMass spectrometry[41]
Cys-DHA
Cys-DHB
Lens proteins (βB1, βB2, βA3, βA4 and γS crystallins)Nucleophilic addition from Cys (GSH) to DHA or DHBMass spectrometry[47]
Tyr-GlyInsulinMichael addition of primary amines (N-terminal Gly) to oxidized Tyr speciesMass spectrometry[48]
Trp-GlyMatrilysin (Matrix metalloproteinase 7)Crosslink between 3-chloroindolenine (3-Cl-Trp) and the main-chain amide adjacent to a Gly NMR spectroscopy[49]
Tyr-HisInsulinMichael addition from His to oxidized TyrMass spectrometry[48]
Tyr-Tyr
(selected data)
Isolated proteins including: α-lactalbumin, caseins, glucose 6-phosphate dehydrogenase, lysozyme, fibronectin, laminins, tropoelastin, cAMP receptor protein, α-synuclein, calmodulin, insulins, hemoglobin, human Δ25 centrin 2.
Human lipoproteins
Human plasma proteins, including those from people with chronic renal failure
Human atherosclerotic lesions
Erythrocytes exposed to H2O2
Brain proteins (amyloid-beta and α-synuclein) from Alzheimer’s subjects
Lipofuscin from aged human brain
Urine from people with diabetes
Human lens proteins
Bacterial spore coat proteins
Parasite oocysts
C–C and/or C–O crosslinks via radical–radical reactions Western blotting
UPLC/HPLC with various detection methods
Mass spectrometry
[48,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77]
Trp-Trp
a)
α-Lactalbumin
b)
Superoxide dismutase 1 (hSOD)
c)
Lysozyme-hSOD
d)
αB-Crystallin
e)
Fibronectin
C–C or C–N crosslinks via radical–radical reactionsMass spectrometry[50,57,78,79,80]
Tyr-Trp
a)
Cytochrome c peroxidase
b)
α-Lactalbumin
c)
Glucose 6-phosphate dehydrogenase
d)
Lysozyme
e)
β-Crystallin
f)
Human cataractous lenses
g)
Fibronectin
C–C (or C–O and C–N) crosslinks via radical–radical reactionsX-ray crystallography (a)
Mass spectrometry (b–g)
[50,53,56,57,80,81]
His-His
a)
Immunoglobulin G1
b)
Immunoglobulin G4
c)
N-Ac-His
Nucleophilic addition of His to oxidized HisMass spectrometry (a,b)
NMR (c)
[82,83,84,85]
His-ArgRibonuclease A (RNAse) Nucleophilic addition of Arg to oxidized HisMass spectrometry[86]
His-LysImmunoglobulin G1 Nucleophilic addition of Lys to oxidized HisMass spectrometry[82,84]
His-CysImmunoglobulin G1Nucleophilic addition of Cys to oxidized HisMass spectrometry[84]
Tyr-Lys
a)
RNAse
b)
Interferon beta-1a
c)
Insulin
Michael addition of Lys to oxidized TyrMass spectrometry[48,86,87]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fuentes-Lemus, E.; Hägglund, P.; López-Alarcón, C.; Davies, M.J. Oxidative Crosslinking of Peptides and Proteins: Mechanisms of Formation, Detection, Characterization and Quantification. Molecules 2022, 27, 15. https://doi.org/10.3390/molecules27010015

AMA Style

Fuentes-Lemus E, Hägglund P, López-Alarcón C, Davies MJ. Oxidative Crosslinking of Peptides and Proteins: Mechanisms of Formation, Detection, Characterization and Quantification. Molecules. 2022; 27(1):15. https://doi.org/10.3390/molecules27010015

Chicago/Turabian Style

Fuentes-Lemus, Eduardo, Per Hägglund, Camilo López-Alarcón, and Michael J. Davies. 2022. "Oxidative Crosslinking of Peptides and Proteins: Mechanisms of Formation, Detection, Characterization and Quantification" Molecules 27, no. 1: 15. https://doi.org/10.3390/molecules27010015

Article Metrics

Back to TopTop