Most proteins in nature are chemically modified after they are made to control how, when, and where they function. The 3 core features of proteins are posttranslationally modified: amino acid side chains can be modified, peptide bonds can be cleaved or isomerized, and disulfide bonds can be cleaved. Cleavage of peptide bonds is a major mechanism of protein control in the circulation, as exemplified by activation of the blood coagulation and complement zymogens. Cleavage of disulfide bonds is emerging as another important mechanism of protein control in the circulation. Recent advances in our understanding of control of soluble blood proteins and blood cell receptors by functional disulfide bonds is discussed as is how these bonds are being identified and studied.

There are 3 fundamental types of posttranslational modifications of proteins that appear to operate in all life forms.1-3  Amino acid side chains can be covalently modified (type 1), peptide bonds can be hydrolytically cleaved or isomerized (type 2), and disulfide bonds can be reductively cleaved (type 3) (Figure 1). These chemical changes are involved in all aspects of protein function and vastly increase the sophistication of the proteome.

Figure 1

The 3 fundamental types of posttranslational modifications. The ribbon structure of a model protein is shown. The amino acid side chains and the peptide and disulfide bonds that bind the polypeptide backbone can be posttranslationally modified. Type 1 modifications are covalent additions of a molecule to an amino acid side chain. The side chains of 15 of the 20 common amino acids of proteins can be covalently modified in reactions that usually involve an enzyme and a cosubstrate.1  The lysine and tyrosine side chains are shown. The most frequent Type 1 modification in humans is phosphorylation of tyrosine. Type 2 modifications are hydrolytic cleavage or isomerization of certain peptide bonds. Hydrolytic cleavage is catalyzed by proteases, which are tightly regulated in space and time because the cleavage is irreversible. Isomerization of the peptide bond on the C-terminal side of proline residues is catalyzed by peptidyl prolyl cis-trans isomerases. Type 3 modifications are reductive cleavage of certain disulfide bonds, known as allosteric disulfides. Allosteric disulfide bonds control the function of the mature protein in which they reside by mediating a change when they are cleaved by oxidoreductases or by thiol-disulfide exchange.

Figure 1

The 3 fundamental types of posttranslational modifications. The ribbon structure of a model protein is shown. The amino acid side chains and the peptide and disulfide bonds that bind the polypeptide backbone can be posttranslationally modified. Type 1 modifications are covalent additions of a molecule to an amino acid side chain. The side chains of 15 of the 20 common amino acids of proteins can be covalently modified in reactions that usually involve an enzyme and a cosubstrate.1  The lysine and tyrosine side chains are shown. The most frequent Type 1 modification in humans is phosphorylation of tyrosine. Type 2 modifications are hydrolytic cleavage or isomerization of certain peptide bonds. Hydrolytic cleavage is catalyzed by proteases, which are tightly regulated in space and time because the cleavage is irreversible. Isomerization of the peptide bond on the C-terminal side of proline residues is catalyzed by peptidyl prolyl cis-trans isomerases. Type 3 modifications are reductive cleavage of certain disulfide bonds, known as allosteric disulfides. Allosteric disulfide bonds control the function of the mature protein in which they reside by mediating a change when they are cleaved by oxidoreductases or by thiol-disulfide exchange.

Close modal

The type 1 modifications of side chains nearly always require an enzyme and a cosubstrate, so these events are usually restricted to specific intracellular environments where all 3 components are available.1  Type 2 modifications of peptide bonds and type 3 modifications of disulfide bonds are suited to the circulation. Cleavage of peptide or disulfide bonds usually does not require a cofactor, so the factors that mediate these events need only to find their substrate to function.

Protein disulfide bonds are the links between the sulfur atoms of 2 cysteine amino acids (the cystine residue) that form as proteins mature in the cell. These bonds have accrued during the evolution of eukaryotic proteins and, once acquired, have almost always been retained.4 

The tertiary structures or partial tertiary structures of 4104 human proteins are currently known or inferred from a homologous protein/domain; these contain 16 538 disulfide bonds (UniProt annotation). About half the disulfide bonds (7264) are in membrane proteins (1987) and most of the rest (8424) are in proteins containing a secretion signal sequence (1204). Interestingly, 587 proteins that reside in the cytoplasm or nucleoplasm contain 935 disulfide bonds, which are environments traditionally thought not to be conducive to disulfide bond formation. The mechanism of formation of disulfide bonds in cytoplasm and nucleoplasm proteins is largely unknown, although protein disulphide isomerase (PDI) has recently been found to interact with the actin cytoskeleton.5  These numbers are representative of the protein disulfide bonds in and on leukocytes.

Mass spectrometry analysis of human plasma has identified 1929 different proteins.6  The tertiary structures of 817 of these proteins are currently known or inferred from a homologous structure/domain; these contain 4594 disulfide bonds. Assuming this average protein to disulfide bond ratio of 1:5 holds for the remaining proteins with unknown structure, the 2000 or so plasma proteins will contain about 10 000 disulfide bonds.

Most of these disulfide bonds, as with most of the peptide bonds, perform a structural role. They stabilize the mature protein structure and remain unchanged for the life of the protein. However, some of the disulfide bonds—the allosteric disulfides—control the function of the mature protein in which they reside when they are cleaved.

Allosteric control is defined as a change in 1 site—the allosteric site—that influences another site by exploiting the protein's flexibility.7  Thus, cleavage of an allosteric disulfide bond results in a functional change at another site in the protein. Changes in ligand binding, substrate hydrolysis, proteolysis, or oligomer formation have been identified in blood proteins.2  The allosteric disulfide bonds are reduced by the catalytic disulfides of oxidoreductases (Figure 2A) or by thiol-disulfide exchange (Figure 2B).2,8 

Figure 2

Mechanisms of cleavage of allosteric disulfide bonds. Disulfide bond reduction occurs via a second-order nucleophilic substitution (SN2)-type reaction mechanism in which the 3 sulfur atoms involved must form an ∼180° angle. (A) For oxidoreductase cleavage, the active site sulfur ion nucleophile of the oxidoreductase (green) attacks 1 of the sulfur atoms of the allosteric disulfide bond (gray). The mixed disulfide that forms then spontaneously decomposes, releasing oxidized oxidoreductase and the substrate protein containing a reduced allosteric disulfide. (B) For thiol-disulfide exchange, the protein contains a sulfur ion nucleophile that is unreactive until a conformational change brings the sulfur ion in line with the allosteric bond, where it attacks 1 of the sulfur atoms of the disulfide, cleaving the bond. The conformational change can be mediated by ligand binding or by mechanical shear of the protein. Intramolecular cleavage is shown, but this can also occur intermolecularly.

Figure 2

Mechanisms of cleavage of allosteric disulfide bonds. Disulfide bond reduction occurs via a second-order nucleophilic substitution (SN2)-type reaction mechanism in which the 3 sulfur atoms involved must form an ∼180° angle. (A) For oxidoreductase cleavage, the active site sulfur ion nucleophile of the oxidoreductase (green) attacks 1 of the sulfur atoms of the allosteric disulfide bond (gray). The mixed disulfide that forms then spontaneously decomposes, releasing oxidized oxidoreductase and the substrate protein containing a reduced allosteric disulfide. (B) For thiol-disulfide exchange, the protein contains a sulfur ion nucleophile that is unreactive until a conformational change brings the sulfur ion in line with the allosteric bond, where it attacks 1 of the sulfur atoms of the disulfide, cleaving the bond. The conformational change can be mediated by ligand binding or by mechanical shear of the protein. Intramolecular cleavage is shown, but this can also occur intermolecularly.

Close modal

Oxidoreductase cleavage

Oxidoreductase active sites contain a reactive dithiol/disulfide in a CXXC motif that can reduce or oxidize a substrate disulfide bond (Figure 2A). Seven members of the PDI family,9  2 members of the thioredoxin family, and glutaredoxin-110  have been identified in human plasma6  (Table 1). Three other members of the PDI family, transmembrane TMX3, ERp44, and ERp29, and the flavoenzyme, Ero1α, have been detected on the platelet surface.11,12  ERp44 and ERp29, however, do not contain a CXXC active site motif, so they may not participate in thiol-disulfide exchange reactions.9  The roles in the circulation of the PDI family members—PDI and ERp57—are best understood to date. There is also emerging evidence for a role for ERp5 in platelet function and thioredoxin in the inflammatory response.

Table 1

Disulfide bond oxidoreductases identified in plasma and/or on the surface of platelets or leukocytes

Class/familyMemberAlternative namesGene name
PDI PDI Cellular thyroid hormone–binding protein, prolyl 4-hydroxylase subunit β, p55 P4HB 
PDI A3 58-kDa glucose-regulated protein, p58, disulphide isomerase ER-60, ERp57, ERp60 PDIA3 
PDI A4 ERp70, ERp72 PDIA4 
PDI A6 ERp5, PDI P5, thioredoxin domain–containing protein 7 PDIA6 
Thioredoxin domain–containing protein 5 ERp46, thioredoxin-like protein p46 TXNDC5 
Thioredoxin domain–containing protein 15  TXNDC15 
Thioredoxin domain–containing protein 17 TRP14, protein 42-9-9, thioredoxin-like protein 5 TXNDC17 
PDI TMX3 Thioredoxin domain–containing protein 10, thioredoxin-related transmembrane protein 3 TMX3 
Endoplasmic reticulum resident protein 44 ERp44, thioredoxin domain–containing protein 4 ERP44 
Endoplasmic reticulum resident protein 29 ERp29, ERp28, ERp31 ERP29 
Thioredoxin Thioredoxin ADF, SASP TXN 
Thioredoxin-like protein 1 32 kDa thioredoxin-related protein TXNL1 
Glutaredoxin Glutaredoxin-1 Thioltransferase-1 GLRX 
Flavoenzyme ERO1-like protein α Ero1α, endoplasmic oxidoreductin-1–like protein, oxidoreductin-1-L-α ERO1L 
Class/familyMemberAlternative namesGene name
PDI PDI Cellular thyroid hormone–binding protein, prolyl 4-hydroxylase subunit β, p55 P4HB 
PDI A3 58-kDa glucose-regulated protein, p58, disulphide isomerase ER-60, ERp57, ERp60 PDIA3 
PDI A4 ERp70, ERp72 PDIA4 
PDI A6 ERp5, PDI P5, thioredoxin domain–containing protein 7 PDIA6 
Thioredoxin domain–containing protein 5 ERp46, thioredoxin-like protein p46 TXNDC5 
Thioredoxin domain–containing protein 15  TXNDC15 
Thioredoxin domain–containing protein 17 TRP14, protein 42-9-9, thioredoxin-like protein 5 TXNDC17 
PDI TMX3 Thioredoxin domain–containing protein 10, thioredoxin-related transmembrane protein 3 TMX3 
Endoplasmic reticulum resident protein 44 ERp44, thioredoxin domain–containing protein 4 ERP44 
Endoplasmic reticulum resident protein 29 ERp29, ERp28, ERp31 ERP29 
Thioredoxin Thioredoxin ADF, SASP TXN 
Thioredoxin-like protein 1 32 kDa thioredoxin-related protein TXNL1 
Glutaredoxin Glutaredoxin-1 Thioltransferase-1 GLRX 
Flavoenzyme ERO1-like protein α Ero1α, endoplasmic oxidoreductin-1–like protein, oxidoreductin-1-L-α ERO1L 

PDI is secreted by platelets,13  endothelial cells,14,15  and neutrophils,16  and mediates protein dithiol-disulfide exchange events in the extracellular space.17,18  This oxidoreductase binds to β3 integrins in the thrombus19  and αvβ3 integrin on endothelial cells.20  PDI inhibitors reduce platelet thrombus size and fibrin formation in both micro- and macrovascular murine models.14,21-23  Platelet PDI is involved in αIIbβ3 integrin–mediated platelet accumulation,24  whereas endothelial PDI is required for fibrin formation.14  Neutrophil PDI modulates ligand binding to αMβ2 integrin and neutrophil recruitment during venous inflammation.25 

The PDI family member ERp57 translocates to the platelet surface after activation, binds to β3 integrin, and is involved in platelet activation and aggregation in vitro and incorporation of platelets into a growing thrombus in murine models.26,27  PDI and ERp57 appear to have distinct roles in β3 integrin function and platelet aggregation as addition of function blocking anti-ERp57 antibodies to PDI-deficient platelets further diminishes integrin αIIbβ3 activation and platelet aggregation.24  Another PDI family member, ERp5, is also recruited to the platelet surface after activation in vitro, and function-blocking anti-ERp5 antibodies inhibit platelet aggregation, fibrinogen binding, and P-selectin exposure.28  A role for ERp5 in platelet function in vivo has not been reported to date.

Thioredoxin has been implicated in the inflammatory response in the circulation. Expression of this reductant is upregulated, and the protein is secreted by immune cells during inflammation, leading to a high local concentration and elevated blood levels.29  Serum levels of thioredoxin are increased in patients with asthma, rheumatoid arthritis, and heart failure30,31  and positively correlate with disease activity in rheumatoid arthritis.32,33 

PDI, ERp57, possibly ERp5, and thioredoxin are involved, therefore, in thrombosis and/or inflammation in mammals. The target disulfide bonds of these oxidoreductases are being defined. PDI is involved in reduction or formation of disulfide bonds in some platelet and leukocyte integrins and possibly leukocyte tissue factor (TF) (see the following section). Thioredoxin reduces the allosteric disulfides in β2-glycoprotein I34,35  and mast cell β tryptase36  in vitro; the role of this reductase in vivo is being studied (see the following section).

The general question of which oxidoreductase cleaves which allosteric disulfide will be determined by some of the same factors that govern which protease cleaves which peptide bond. Some proteases cleave many peptide bonds (such as thrombin), whereas others cleave only 1 or a very restricted number (such as factor VIIa), and the same scenario will likely apply to oxidoreductases and allosteric disulfide bonds. Oxidoreductase substrate selectivity will largely be driven by steric factors (accessibility of the disulfide bond in the substrate protein and conformation of the active site pocket of the enzyme) and by environmental context (being in the right place at the right time). The highly directional nature of disulfide bond cleavage will further restrict the substrate selection of oxidoreductases. Reduction of a disulfide bond proceeds via a second-order nucleophilic substitution (SN2)-type reaction mechanism in which the 3 sulfur atoms involved (the sulfur ion nucleophile and the 2 sulfur atoms of the disulfide bond) must form an ∼180° angle37,38  (Figure 2). For instance, stretching, twisting, or pulling of proteins makes their disulfide bonds easier or harder to cleave by changing the alignment of the 3 sulfur atoms.39,40  Another factor that influences the substrate selectivity of oxidoreductases is the redox potential of the oxidoreductase catalytic disulfide and the allosteric disulfide. Catalytic disulfides will only cleave allosteric disulfides with a bigger (less negative) redox potential.

An open question is whether the oxidoreductases act catalytically in the circulation to reduce several substrate disulfides or reduce only 1 disulfide bond. Intracellular oxidoreductases act as catalysts because factors regenerate the reduced enzyme. The thioredoxin active site disulfide, for example, is reduced by thioredoxin reductase and reduced NAD phosphate (NADPH) in the cytoplasm.10  NADPH provides the hydrogens and electrons used to reduce the oxidized thioredoxin. Thioredoxin reductase is found in plasma,6  but NADPH is not. It is possible, and seems likely, that a mechanism or mechanisms exist to cycle some oxidoreductases in blood. For instance, Ero1α has been implicated in oxidation of PDI on the platelet surface.12  On the other hand, restricting oxidoreductases to cleavage of a single substrate disulfide bond would be an efficient mechanism of regulation of their activity. Proteases, for instance, only require a source of water to cleave peptide bonds, so protease inhibitors have coevolved with proteases to control their activity.1  Controlling access to a source of hydrogens and electrons is a simple means of restricting oxidoreductase activity. Endogenous inhibitors of the oxidoreductases in the extracellular space have not been identified to date, which supports the idea that oxidoreductases may function as single-turnover enzymes.

Thiol-disulfide exchange cleavage

Cleavage of allosteric disulfides by thiol-disulfide exchange does not require additional hydrogens and electrons (Figure 2B), so this mechanism of cleavage is particularly suited to the circulation. All that is required is a conformational change in the substrate protein, which could be triggered by ligand binding and/or the mechanical shear forces of flowing blood. The substrate protein contains a sulfur ion that is unreactive until a change in structure brings it in line with the allosteric bond, where it attacks and cleaves the disulfide bond. Allosteric disulfide bonds in plasma plasminogen41  and von Willebrand factor (VWF)42  are cleaved by thiol-disulfide exchange (see the following section).

An important difference between peptide and disulfide bond cleavage is that cleavage of disulfide bonds is potentially reversible; this can provide for a fine level of protein regulation not possible with peptide bond cleavage. Some allosteric disulfides exist in equilibrium between reduced and oxidized isoforms in the population of protein molecules, and perturbation of the equilibrium by oxidoreductase or ligand binding, for instance, leads to a biological change. Cleavage of other allosteric disulfides, though, is irreversible as the downstream effects of the cleavage prevent the disulfide bond from reforming. Both of these situations in different blood proteins are discussed here.

We have chosen to highlight 4 soluble blood proteins and a lymphocyte receptor whose functions are controlled by cleavage of allosteric disulfide bonds. These examples are the best characterized at the molecular level so far. We also briefly discuss 7 other examples at a more preliminary stage of characterization.

Angiotensinogen

Plasma angiotensinogen is proteolyzed by renin and angiotensin-converting enzyme to produce the angiotensin peptides that control blood pressure and fluid homeostasis. The angiotensinogen Cys18–Cys138 disulfide bond is cleaved in a fraction of the plasma protein, which is significant because the oxidized protein is more efficiently activated by renin than the reduced protein.43  The ratio of oxidized to reduced angiotensinogen in blood is approximately 60:40 in healthy volunteers (independent of age or gender). This ratio increases in pregnant females with preeclampsia, which correlates with increased cleavage by renin and elevated blood pressure. It has been suggested that exposure of maternal angiotensinogen to reactive oxygen species present in the placenta increases the level of oxidized angiotensinogen, contributing to the development of preeclampsia.43 

β2-glycoprotein I

Plasma β2-glycoprotein I is an autoantigen in the antiphospholipid syndrome, which is associated with vascular thrombosis, accelerated atherosclerosis, and recurrent miscarriages.44  The β2-glycoprotein I–antibody complex appears to prime the arterial and venous vasculature for thrombosis through effects on blood cells and coagulation and fibrinolysis proteins. β2-glycoprotein I consists of 4 complement modules and a unique fifth domain that contains a disulfide bond linking Cys288 and the C-terminal residue, Cys326.

The Cys288–Cys326 disulfide bond is cleaved in a fraction of plasma β2-glycoprotein I, and the bond is reduced in the purified protein by PDI and thioredoxin.34,35  The ratio of oxidized to reduced β2-glycoprotein I is elevated in the blood of patients with antiphospholipid syndrome compared with healthy volunteers and with patients with vascular thrombosis and no antiphospholipid antibodies.45  There is also more oxidized β2-glycoprotein I in the blood of patients with antiphospholipid syndrome with both anti–β2-glycoprotein I antibodies and lupus anticoagulant than patients with only anti–β2-glycoprotein I antibodies.45 

The elevated oxidized β2-glycoprotein I in patients with antiphospholipid syndrome is associated with increased immunogenicity of the protein and increased thrombosis.45  Anti–β2-glycoprotein I antibodies from mice and rabbit plasma and autoantibodies from the plasma of patients with antiphospholipid syndrome bind more avidly to oxidized β2-glycoprotein I than reduced protein. The increased thrombosis may also relate to loss of anti-thrombotic properties of reduced β2-glycoprotein, such as protection of endothelial cells from oxidative stress.34 

IL receptor subunit γ

Interleukin (IL) receptor subunit γ or CD132 is the common γ subunit of the cytokine receptors for IL-2, -4, -7, -9, -15, and -21 on the surface of lymphocytes. The CD132 Cys183–Cys232 disulfide bond is cleaved on the surface of cultured T cells by protein reductants and on the surface of thymocytes in mice after an inflammatory challenge.46  The disulfide bond is located at the subunit surface close to the IL-2 binding site.47  Reduction of the disulfide46  or replacement of Cys183 and Cys232 with alanine or serine48  inhibits IL-2 binding to the receptor complex and therefore receptor signaling and cell proliferation. Notably, Cys183 and Cys232 mutants are among those identified in patients with X-linked severe combined immunodeficiency.49,50 

Plasminogen

Plasma plasminogen is the zymogen form of plasmin. The zymogen consists of an N-terminal Pan-apple domain followed by 5 kringle domains and a C-terminal serine protease domain. Plasminogen is converted to plasmin by urokinase or tissue plasminogen activator through cleavage of the Arg561–Val562 peptide bond in the serine protease domain and the Pan-apple domain is autoproteolytically released to produce mature plasmin. Plasmin activates other zymogens and cleaves the fibrin meshwork during dissolution of the thrombus and components of the extracellular matrix.

The Cys462–Cys541 disulfide bond in the kringle 5 domain is cleaved in a fraction of plasma plasminogen, and the level of reduced plasminogen varies in healthy individuals.41,51  Fragments of plasminogen containing the kringle domains are generated in plasma; these fragments function as endogenous inhibitors of tumor angiogenesis.52  The kringle fragments are produced from the reduced form of the zymogen.41  Generation of the kringle fragments from reduced plasminogen involves conversion to reduced plasmin, reduction of an additional kringle 5 disulfide bond, and subsequent proteolysis of up to 3 peptide bonds.41,53-55  Plasmin ligands such as tumor-derived phosphoglycerate kinase56  induce a conformational change in reduced kringle 5 that leads to attack by the Cys541 thiolate on the Cys536 sulfur atom of the Cys512–Cys536 disulfide bond, resulting in reduction of the bond by thiol-disulfide exchange (Figure 2B). Cleavage of the Cys512–Cys536 disulfide bond leads to a conformational change in plasmin and exposure of the peptide backbone to proteolysis on the C-terminal side of residues Lys486, Arg474, and Arg530. The proteolysis can be catalyzed by plasmin itself or other serine- or metalloproteases.

VWF

VWF is a plasma protein produced by vascular endothelial cells and megakaryocytes that chaperones blood coagulation cofactor factor VIII and tethers platelets to the injured blood vessel wall. It is a large glycoprotein that circulates as a series of multimers containing variable numbers of 500-kDa dimeric units. VWF structure is influenced by the mechanical shear forces in blood, where it transitions from a loosely coiled ball to elongated fibers that bind platelets.57-63 

Fiber formation involves covalent self-association of VWF mediated by thiol-disulfide exchange in the VWF C2 domains.64,65  Both the C2 Cys2431–Cys2453 and nearby Cys2451–Cys2468 disulfide bonds are involved in fiber formation. Reduction of the C2 domain Cys2431–Cys2453 disulfide bond in 1 molecule, presumably by an oxidoreductase in blood, creates a Cys2431 thiolate anion that attacks the Cys2431 sulfur (of the Cys2431–Cys2453 disulfide bond) in another VWF molecule, resulting in a disulfide-linked dimer.42  The Cys2451–Cys2468 disulfide-dithiol then mediates formation of trimers and higher order oligomers.

Other blood proteins

Three blood cell receptors and 4 other soluble plasma proteins have been found to contain putative allosteric disulfides. These examples lack either in vivo evidence for a role for the bond and/or there is uncertainty as to the identity of the allosteric disulfide or disulfides in the protein. Their classification as allosteric disulfides should be considered preliminary at this stage.

Integrins are transmembrane, heterodimeric cell-adhesion receptors that mediate interactions between the cytoskeleton of cells and extracellular ligands and also play a role in signal transduction. Integrin activation has been associated with reduction of 1 or more disulfide bonds in the receptors.18,20,66-77  There are 18 α integrin subunits and 8 β subunits in mammals, but the majority of the redox studies have been conducted on the β3 subunit of the platelet αIIbβ3 (fibrinogen) receptor and the platelet and endothelial αvβ3 (vitronectin) receptor. There are 4 epidermal growth factor (EGF)-like domains in the extracellular region of β3.78,79  The disulfides in the EGF domains are generally important for maintaining the inactive state of β3 because mutation of a single cysteine for most of the disulfide bonds within the EGF domains results in a constitutively active β3.80-82  A recent study of the function of an atypical disulfide bond in β3 EGF domains is revealing.83  The disulfide bond in each of the 4 EGF domains in both αIIbβ3 and αvβ3 was investigated by mutating the cysteines to serines. An allosteric role for the EGF-4 Cys560–Cys583 disulfide in both αIIbβ3 and αvβ3 and for the EGF-3 Cys523–Cys544 bond only in αvβ3 was identified. For instance, activation of the C523S/C544S αvβ3 mutant by an antibody and a reducing agent is impaired. These disulfides are possible PDI19,20  and/or ERp5726,27  substrates.

TF is a transmembrane cofactor for factor VIIa. The TF/VIIa complex proteolytically activates factor X to initiate blood coagulation and provide the thrombin burst required for a stable thrombus. Intravascular TF is found on monocytes and neutrophils and exists mostly in a noncoagulant or cryptic form bound nonproductively to VIIa. Acute events lead to local decryption of TF, which appears to involve both formation of a disulfide bond between unpaired Cys186 and Cys209 in cryptic TF and exposure of phosphatidylserine on the cell surface (reviewed in Chen and Hogg84 ). Thioredoxin has been implicated in the maintenance of cryptic TF by cleaving the TF Cys186–Cys209 disulfide,84,85  whereas PDI has been implicated in TF decryption.86,87  The molecular mechanism of action of PDI in TF decryption is not known. Both the oxidoreductase and chaperone functions of PDI are possibly involved (reviewed in Langer and Ruf88 ).

Two circulating serine proteases appear to be regulated by allosteric disulfides. Proteolytic activation of coagulation factor XI contributes to the intrinsic/amplification phase of blood coagulation. One or more factor XI disulfide bonds are reduced in a fraction of the zymogen in blood, and the Cys362–Cys482 and Cys118–Cys147 disulfides are reduced by thioredoxin in the isolated enzyme.89  The reduced protein is more efficiently activated by thrombin, factor XIIa, or factor XIa than the oxidized protein, and blood levels of the reduced factor XI are elevated in patients with antiphospholipid syndrome,89  which may contribute to the thrombosis in this syndrome. Mast cell βII-tryptase is associated with pathological inflammation. The βII-tryptase Cys220–Cys248 disulfide bond exists in oxidized and reduced states in the secreted enzyme and the bond is reduced by thioredoxin in the isolated enzyme.36  The oxidized and reduced isoforms have different specificity and catalytic efficiency for hydrolysis of substrates.

The activity of 2 growth factors is also regulated by disulfide bond cleavage. Platelets contain abundant transforming growth factor-β1 (TGF-β1) that functions in a variety of physiologic and pathologic events. TGF-β1 is secreted as an inactive form in complex with latency-associated peptide and latent TGF-β–binding protein 1. Platelet TGF-β1 is activated by shear-induced disulfide bond cleavage in the growth factor,90  and platelet PDI has been implicated in this event.91  The identity of the functional disulfide bond or bonds in TGF-β1 is not known at present. The structures of the lymphangiogenic growth factors, vascular endothelial growth factor (VEGF)-C and VEGF-D, are also regulated by disulfide cleavage.92  The VEGFs function as disulfide-linked antiparallel homodimers, and dimer formation of VEGF-C and VEGF-D is controlled by a unique unpaired cysteine that exchanges with the interdimer disulfide bond.93 

Allosteric disulfides are being identified using bioinformatics and experimental screens, and both approaches are proving useful for identifying these bonds in circulating proteins.

Bioinformatic approach

The bioinformatic technique relies on high-resolution 3-dimensional protein structures and a dataset of allosteric disulfides from which common features can be derived. The configuration, surface exposure, and secondary structural motifs that the disulfide links can be useful measures for the identification of allosteric bonds. These and other disulfide bond parameters are easily extracted from any protein structure in Protein Data Bank format.94  The most informative disulfide bond measure at this time is the configuration of the cystine residue. The geometry of cystine is defined by 5 dihedral or χ angles, which are calculated by the rotation around the bonds linking the 6 atoms (Figure 3A).95  There are 20 possible different cystine configurations, and all the types have been identified in protein crystal structures.95,96  Of the 20 different configurations, the right-handed (RH) –RHStaple, left-handed (LH) –LHHook, and −/+RHHook bonds are emerging as allosteric configurations (Figure 3B).2,8  The configurations of allosteric disulfide bonds in blood proteins and their mechanism of cleavage, where known, are summarized in Table 2.

Figure 3

Configurations of allosteric disulfide bonds. (A) Classification of disulfide bonds based on their geometry.95,96  The geometry is defined by the 5 bond angles (χ angles) linking the 2 α-carbons of the cystine residue. Cα, main chain carbon atom; Cβ, side chain carbon atom of each cysteine residue. The χ angles are recorded as being either positive or negative. The 3 basic types of bond configurations (spirals, hooks, and staples) are based on the signs of the central 3 angles, and they can be either RH or LH depending on whether the sign of the χ3 angle is positive or negative, respectively. These 6 bond types expand to 20 when the χ1 and χ1′ angles are taken into account. (B) Examples of the structures of the emerging allosteric configurations: –RHStaple, –LHHook, and −/+RHHook.

Figure 3

Configurations of allosteric disulfide bonds. (A) Classification of disulfide bonds based on their geometry.95,96  The geometry is defined by the 5 bond angles (χ angles) linking the 2 α-carbons of the cystine residue. Cα, main chain carbon atom; Cβ, side chain carbon atom of each cysteine residue. The χ angles are recorded as being either positive or negative. The 3 basic types of bond configurations (spirals, hooks, and staples) are based on the signs of the central 3 angles, and they can be either RH or LH depending on whether the sign of the χ3 angle is positive or negative, respectively. These 6 bond types expand to 20 when the χ1 and χ1′ angles are taken into account. (B) Examples of the structures of the emerging allosteric configurations: –RHStaple, –LHHook, and −/+RHHook.

Close modal
Table 2

Configurations of allosteric disulfide bonds in blood proteins and their mechanism of cleavage

ProteinDisulfide cysteinesDisulfide bond configuration (Protein Data Bank identifier)Cleavage mechanismReference
Angiotensinogen 18-138 −/+RHHook (2WXW) Oxidoreductase 43 
β2-glycoprotein I 288-326 −/+RHHook (1C1Z) Oxidoreductase 34, 35 
IL-2 receptor subunit γ (CD132) 183-232 −LHHook (2ERJ) Oxidoreductase 46 
Plasminogen 462-541 −/+RHHook (4DUR) Oxidoreductase? 41, 51 
512-536 −LHHook (4DUR) Thiol-disulfide exchange 
β3 integrin (αIIbβ3) 523-544 −LHHook (3FCS) Oxidoreductase? 83 
560-583 −LHHook (3FCS) 
β3 integrin (αvβ3) 523-544 −RHStaple (3IJE) Oxidoreductase? 83 
560-583 −LHHook (3IJE) 
Tissue factor 186-209 −RHStaple (1BOY) Oxidoreductase 98 
Factor XI 362-482 −/+RHHook (2F83) Oxidoreductase 89 
βII-tryptase 220-248 −RHHook and Oxidoreductase 36 
−RHSpiral (3FPZ) 
VEGF-C 156-165 −LHHook (2X1W) Thiol-disulfide exchange 93 
ProteinDisulfide cysteinesDisulfide bond configuration (Protein Data Bank identifier)Cleavage mechanismReference
Angiotensinogen 18-138 −/+RHHook (2WXW) Oxidoreductase 43 
β2-glycoprotein I 288-326 −/+RHHook (1C1Z) Oxidoreductase 34, 35 
IL-2 receptor subunit γ (CD132) 183-232 −LHHook (2ERJ) Oxidoreductase 46 
Plasminogen 462-541 −/+RHHook (4DUR) Oxidoreductase? 41, 51 
512-536 −LHHook (4DUR) Thiol-disulfide exchange 
β3 integrin (αIIbβ3) 523-544 −LHHook (3FCS) Oxidoreductase? 83 
560-583 −LHHook (3FCS) 
β3 integrin (αvβ3) 523-544 −RHStaple (3IJE) Oxidoreductase? 83 
560-583 −LHHook (3IJE) 
Tissue factor 186-209 −RHStaple (1BOY) Oxidoreductase 98 
Factor XI 362-482 −/+RHHook (2F83) Oxidoreductase 89 
βII-tryptase 220-248 −RHHook and Oxidoreductase 36 
−RHSpiral (3FPZ) 
VEGF-C 156-165 −LHHook (2X1W) Thiol-disulfide exchange 93 

It is important to keep in mind, though, that the protein backbone and the cystines that link it can change shape when in solution. That is, the crystal structure might represent only 1 of a number of different possible structures in solution. For instance, the same disulfide bond can exist in different configurations in different crystal structures.96  Dynamic isomerization of allosteric disulfide bonds is probably an important aspect of their function. Cleavage of the bond may only occur when the cystine adopts a particular configuration, which could be controlled by ligand binding or the mechanical shear in the circulation.

Experimental approach

A method based on differential cysteine labeling and mass spectrometry is currently the best experimental screen for allosteric disulfides.36,41,42,97  The approach is based on the principle that an allosteric disulfide will exist in equilibrium between reduced and oxidized states in a blood protein, such as observed in angiotensinogen, β2-glycoprotein I, and plasminogen. The technique measures the proportion of reduced and oxidized protein using different thiol alkylating agents. One alkylator is used to label the reduced portion of the allosteric disulfide bond in blood, or a fraction thereof, without or with previous treatment with a candidate oxidoreductase. The remaining disulfide bonds in the protein are then reduced using dithiothreitol and the new unpaired cysteines labeled with a second alkylator. Mass spectrometry is used to determine the identity of the protein and the ratio of differentially labeled cysteines. Typically, the cysteines showing significant increase in the ratio of the first to second alkylator after oxidoreductase treatment are allosteric disulfide candidates.

The functions of many blood proteins are controlled by cleavage of peptide bonds. For example, the blood coagulation and complement cascades are exquisitely regulated by discrete proteolysis. There are an increasing number of blood proteins found to be controlled by cleavage of the next most frequent covalent bond linking the polypeptide backbone of proteins: the disulfide bond. The allosteric and catalytic functional disulfide bonds contribute to the control of thrombosis and hemostasis, blood pressure, and inflammation in the circulation, and some bonds have been linked to disease. The pace of discovery of new allosteric disulfides in blood proteins is likely to increase during the coming years as new and better techniques for identifying and characterizing these bonds are developed. In terms of the oxidoreductases that cleave these bonds, the generation of specific inhibitors and cell-specific oxidoreductase-deficient mice has and will prove valuable for identifying the environments in which the enzymes work and their substrate disulfide bonds. Some allosteric and catalytic bonds will also likely be valid pharmaceutical targets and progress is being made on ways of targeting these bonds with small molecules or biologicals (reviewed in Hogg8 ).

This study was supported by grants from the National Health and Medical Research Council of Australia and the Cancer Council New South Wales.

Contribution: P.J.H. conceived the review and all authors contributed to the writing of the manuscript.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Philip Hogg, Level 2, Lowy Cancer Research Centre, University of New South Wales, Sydney 2052, Australia; e-mail: p.hogg@unsw.edu.au.

1
Walsh
 
CT
Posttranslational modification of proteins: expanding nature's inventory
2006
Greenwood Village, CO
Roberts & Company
2
Cook
 
KM
Hogg
 
PJ
Post-translational control of protein function by disulfide bond cleavage.
Antioxid Redox Signal
2013
, vol. 
18
 
15
(pg. 
1987
-
2015
)
3
Hogg
 
PJ
Disulfide bonds as switches for protein function.
Trends Biochem Sci
2003
, vol. 
28
 
4
(pg. 
210
-
214
)
4
Wong
 
JW
Ho
 
SY
Hogg
 
PJ
Disulfide bond acquisition through eukaryotic protein evolution.
Mol Biol Evol
2011
, vol. 
28
 
1
(pg. 
327
-
334
)
5
Sobierajska
 
K
Skurzynski
 
S
Stasiak
 
M
Kryczka
 
J
Cierniewski
 
CS
Swiatkowska
 
M
PDI directly interacts with β-actin Cys374 and regulates cytoskeleton reorganization [published online ahead of print January 10, 2014].
J Biol Chem
2014
6
Farrah
 
T
Deutsch
 
EW
Omenn
 
GS
, et al. 
A high-confidence human plasma proteome reference set with estimated concentrations in PeptideAtlas.
Mol Cell Proteomics
2011
, vol. 
10
 
9
pg. 
M110.006353
 
7
Monod
 
J
Wyman
 
J
Changeux
 
JP
On the nature of allosteric transitions: a plausible model.
J Mol Biol
1965
, vol. 
12
 (pg. 
88
-
118
)
8
Hogg
 
PJ
Targeting allosteric disulphide bonds in cancer.
Nat Rev Cancer
2013
, vol. 
13
 
6
(pg. 
425
-
431
)
9
Kozlov
 
G
Määttänen
 
P
Thomas
 
DY
Gehring
 
K
A structural overview of the PDI family of proteins.
FEBS J
2010
, vol. 
277
 
19
(pg. 
3924
-
3936
)
10
Holmgren
 
A
Thioredoxin and glutaredoxin systems.
J Biol Chem
1989
, vol. 
264
 
24
(pg. 
13963
-
13966
)
11
Holbrook
 
LM
Watkins
 
NA
Simmonds
 
AD
Jones
 
CI
Ouwehand
 
WH
Gibbins
 
JM
Platelets release novel thiol isomerase enzymes which are recruited to the cell surface following activation.
Br J Haematol
2010
, vol. 
148
 
4
(pg. 
627
-
637
)
12
Swiatkowska
 
M
Padula
 
G
Michalec
 
L
Stasiak
 
M
Skurzynski
 
S
Cierniewski
 
CS
Ero1alpha is expressed on blood platelets in association with protein-disulfide isomerase and contributes to redox-controlled remodeling of alphaIIbbeta3.
J Biol Chem
2010
, vol. 
285
 
39
(pg. 
29874
-
29883
)
13
Chen
 
K
Lin
 
Y
Detwiler
 
TC
Protein disulfide isomerase activity is released by activated platelets.
Blood
1992
, vol. 
79
 
9
(pg. 
2226
-
2228
)
14
Jasuja
 
R
Furie
 
B
Furie
 
BC
Endothelium-derived but not platelet-derived protein disulfide isomerase is required for thrombus formation in vivo.
Blood
2010
, vol. 
116
 
22
(pg. 
4665
-
4674
)
15
Ramachandran
 
N
Root
 
P
Jiang
 
XM
Hogg
 
PJ
Mutus
 
B
Mechanism of transfer of NO from extracellular S-nitrosothiols into the cytosol by cell-surface protein disulfide isomerase.
Proc Natl Acad Sci U S A
2001
, vol. 
98
 
17
(pg. 
9539
-
9544
)
16
Bennett
 
TA
Edwards
 
BS
Sklar
 
LA
Rogelj
 
S
Sulfhydryl regulation of L-selectin shedding: phenylarsine oxide promotes activation-independent L-selectin shedding from leukocytes.
J Immunol
2000
, vol. 
164
 
8
(pg. 
4120
-
4129
)
17
Jiang
 
XM
Fitzgerald
 
M
Grant
 
CM
Hogg
 
PJ
Redox control of exofacial protein thiols/disulfides by protein disulfide isomerase.
J Biol Chem
1999
, vol. 
274
 
4
(pg. 
2416
-
2423
)
18
Essex
 
DW
Li
 
M
Miller
 
A
Feinman
 
RD
Protein disulfide isomerase and sulfhydryl-dependent pathways in platelet activation.
Biochemistry
2001
, vol. 
40
 
20
(pg. 
6070
-
6075
)
19
Cho
 
J
Kennedy
 
DR
Lin
 
L
, et al. 
Protein disulfide isomerase capture during thrombus formation in vivo depends on the presence of β3 integrins.
Blood
2012
, vol. 
120
 
3
(pg. 
647
-
655
)
20
Swiatkowska
 
M
Szymański
 
J
Padula
 
G
Cierniewski
 
CS
Interaction and functional association of protein disulfide isomerase with alphaVbeta3 integrin on endothelial cells.
FEBS J
2008
, vol. 
275
 
8
(pg. 
1813
-
1823
)
21
Cho
 
J
Furie
 
BC
Coughlin
 
SR
Furie
 
B
A critical role for extracellular protein disulfide isomerase during thrombus formation in mice.
J Clin Invest
2008
, vol. 
118
 
3
(pg. 
1123
-
1131
)
22
Reinhardt
 
C
von Brühl
 
ML
Manukyan
 
D
, et al. 
Protein disulfide isomerase acts as an injury response signal that enhances fibrin generation via tissue factor activation.
J Clin Invest
2008
, vol. 
118
 
3
(pg. 
1110
-
1122
)
23
Jasuja
 
R
Passam
 
FH
Kennedy
 
DR
, et al. 
Protein disulfide isomerase inhibitors constitute a new class of antithrombotic agents.
J Clin Invest
2012
, vol. 
122
 
6
(pg. 
2104
-
2113
)
24
Kim
 
K
Hahm
 
E
Li
 
J
, et al. 
Platelet protein disulfide isomerase is required for thrombus formation but not for hemostasis in mice.
Blood
2013
, vol. 
122
 
6
(pg. 
1052
-
1061
)
25
Hahm
 
E
Li
 
J
Kim
 
K
Huh
 
S
Rogelj
 
S
Cho
 
J
Extracellular protein disulfide isomerase regulates ligand-binding activity of alphaMbeta2 integrin and neutrophil recruitment during vascular inflammation.
Blood
2013
, vol. 
121
 
19
 
3789-3800, S1-S15
26
Holbrook
 
LM
Sasikumar
 
P
Stanley
 
RG
Simmonds
 
AD
Bicknell
 
AB
Gibbins
 
JM
The platelet-surface thiol isomerase enzyme ERp57 modulates platelet function.
J Thromb Haemost
2012
, vol. 
10
 
2
(pg. 
278
-
288
)
27
Wang
 
L
Wu
 
Y
Zhou
 
J
, et al. 
Platelet-derived ERp57 mediates platelet incorporation into a growing thrombus by regulation of the αIIbβ3 integrin.
Blood
2013
, vol. 
122
 
22
(pg. 
3642
-
3650
)
28
Jordan
 
PA
Stevens
 
JM
Hubbard
 
GP
, et al. 
A role for the thiol isomerase protein ERP5 in platelet function.
Blood
2005
, vol. 
105
 
4
(pg. 
1500
-
1507
)
29
Lillig
 
CH
Holmgren
 
A
Thioredoxin and related molecules—from biology to health and disease.
Antioxid Redox Signal
2007
, vol. 
9
 
1
(pg. 
25
-
47
)
30
Yamada
 
Y
Nakamura
 
H
Adachi
 
T
, et al. 
Elevated serum levels of thioredoxin in patients with acute exacerbation of asthma.
Immunol Lett
2003
, vol. 
86
 
2
(pg. 
199
-
205
)
31
Burke-Gaffney
 
A
Callister
 
ME
Nakamura
 
H
Thioredoxin: friend or foe in human disease?
Trends Pharmacol Sci
2005
, vol. 
26
 
8
(pg. 
398
-
404
)
32
Jikimoto
 
T
Nishikubo
 
Y
Koshiba
 
M
, et al. 
Thioredoxin as a biomarker for oxidative stress in patients with rheumatoid arthritis.
Mol Immunol
2002
, vol. 
38
 
10
(pg. 
765
-
772
)
33
Lemarechal
 
H
Allanore
 
Y
Chenevier-Gobeaux
 
C
Ekindjian
 
OG
Kahan
 
A
Borderie
 
D
High redox thioredoxin but low thioredoxin reductase activities in the serum of patients with rheumatoid arthritis.
Clin Chim Acta
2006
, vol. 
367
 
1-2
(pg. 
156
-
161
)
34
Ioannou
 
Y
Zhang
 
JY
Passam
 
FH
, et al. 
Naturally occurring free thiols within beta 2-glycoprotein I in vivo: nitrosylation, redox modification by endothelial cells, and regulation of oxidative stress-induced cell injury.
Blood
2010
, vol. 
116
 
11
(pg. 
1961
-
1970
)
35
Passam
 
FH
Rahgozar
 
S
Qi
 
M
, et al. 
Beta 2 glycoprotein I is a substrate of thiol oxidoreductases.
Blood
2010
, vol. 
116
 
11
(pg. 
1995
-
1997
)
36
Cook
 
KM
McNeil
 
HP
Hogg
 
PJ
Allosteric control of βII-tryptase by a redox active disulfide bond.
J Biol Chem
2013
, vol. 
288
 
48
(pg. 
34920
-
34929
)
37
Fernandes
 
PA
Ramos
 
MJ
Theoretical insights into the mechanism for thiol/disulfide exchange.
Chemistry
2004
, vol. 
10
 
1
(pg. 
257
-
266
)
38
Wiita
 
AP
Ainavarapu
 
SR
Huang
 
HH
Fernandez
 
JM
Force-dependent chemical kinetics of disulfide bond reduction observed with single-molecule techniques.
Proc Natl Acad Sci U S A
2006
, vol. 
103
 
19
(pg. 
7222
-
7227
)
39
Baldus
 
IB
Gräter
 
F
Mechanical force can fine-tune redox potentials of disulfide bonds.
Biophys J
2012
, vol. 
102
 
3
(pg. 
622
-
629
)
40
Wiita
 
AP
Perez-Jimenez
 
R
Walther
 
KA
, et al. 
Probing the chemistry of thioredoxin catalysis with force.
Nature
2007
, vol. 
450
 
7166
(pg. 
124
-
127
)
41
Butera
 
D
Wind
 
T
Lay
 
A
Beck
 
J
Castellino
 
F
Hogg
 
P
Characterization of a reduced form of plasma plasminogen as the precursor for angiostatin formation.
J Biol Chem
2014
, vol. 
289
 
5
(pg. 
2992
-
3000
)
42
Ganderton
 
T
Wong
 
JW
Schroeder
 
C
Hogg
 
PJ
Lateral self-association of VWF involves the Cys2431-Cys2453 disulfide/dithiol in the C2 domain.
Blood
2011
, vol. 
118
 
19
(pg. 
5312
-
5318
)
43
Zhou
 
A
Carrell
 
RW
Murphy
 
MP
, et al. 
A redox switch in angiotensinogen modulates angiotensin release.
Nature
2010
, vol. 
468
 
7320
(pg. 
108
-
111
)
44
Giannakopoulos
 
B
Krilis
 
SA
The pathogenesis of the antiphospholipid syndrome.
N Engl J Med
2013
, vol. 
368
 
11
(pg. 
1033
-
1044
)
45
Ioannou
 
Y
Zhang
 
JY
Qi
 
M
, et al. 
Novel assays of thrombogenic pathogenicity in the antiphospholipid syndrome based on the detection of molecular oxidative modification of the major autoantigen β2-glycoprotein I.
Arthritis Rheum
2011
, vol. 
63
 
9
(pg. 
2774
-
2782
)
46
Metcalfe
 
C
Cresswell
 
P
Barclay
 
AN
Interleukin-2 signalling is modulated by a labile disulfide bond in the CD132 chain of its receptor.
Open Biol
2012
, vol. 
2
 
1
pg. 
110036
 
47
Wang
 
X
Rickert
 
M
Garcia
 
KC
Structure of the quaternary complex of interleukin-2 with its alpha, beta, and gammac receptors.
Science
2005
, vol. 
310
 
5751
(pg. 
1159
-
1163
)
48
Olosz
 
F
Malek
 
TR
Three loops of the common gamma chain ectodomain required for the binding of interleukin-2 and interleukin-7.
J Biol Chem
2000
, vol. 
275
 
39
(pg. 
30100
-
30105
)
49
Puck
 
JM
IL2RGbase: a database of γ c-chain defects causing human X-SCID.
Immunol Today
1996
, vol. 
17
 
11
(pg. 
507
-
511
)
50
Niemela
 
JE
Puck
 
JM
Fischer
 
RE
Fleisher
 
TA
Hsu
 
AP
Efficient detection of thirty-seven new IL2RG mutations in human X-linked severe combined immunodeficiency.
Clin Immunol
2000
, vol. 
95
 
1 Pt 1
(pg. 
33
-
38
)
51
Law
 
RH
Caradoc-Davies
 
T
Cowieson
 
N
, et al. 
The X-ray crystal structure of full-length human plasminogen.
Cell Rep
2012
, vol. 
1
 
3
(pg. 
185
-
190
)
52
Wang
 
J
Wang
 
J
Dai
 
J
, et al. 
A glycolytic mechanism regulating an angiogenic switch in prostate cancer.
Cancer Res
2007
, vol. 
67
 
1
(pg. 
149
-
159
)
53
Lay
 
AJ
Jiang
 
XM
Daly
 
E
Sun
 
L
Hogg
 
PJ
Plasmin reduction by phosphoglycerate kinase is a thiol-independent process.
J Biol Chem
2002
, vol. 
277
 
11
(pg. 
9062
-
9068
)
54
Stathakis
 
P
Fitzgerald
 
M
Matthias
 
LJ
Chesterman
 
CN
Hogg
 
PJ
Generation of angiostatin by reduction and proteolysis of plasmin. Catalysis by a plasmin reductase secreted by cultured cells.
J Biol Chem
1997
, vol. 
272
 
33
(pg. 
20641
-
20645
)
55
Stathakis
 
P
Lay
 
AJ
Fitzgerald
 
M
Schlieker
 
C
Matthias
 
LJ
Hogg
 
PJ
Angiostatin formation involves disulfide bond reduction and proteolysis in kringle 5 of plasmin.
J Biol Chem
1999
, vol. 
274
 
13
(pg. 
8910
-
8916
)
56
Lay
 
AJ
Jiang
 
XM
Kisker
 
O
, et al. 
Phosphoglycerate kinase acts in tumour angiogenesis as a disulphide reductase.
Nature
2000
, vol. 
408
 
6814
(pg. 
869
-
873
)
57
Barg
 
A
Ossig
 
R
Goerge
 
T
, et al. 
Soluble plasma-derived von Willebrand factor assembles to a haemostatically active filamentous network.
Thromb Haemost
2007
, vol. 
97
 
4
(pg. 
514
-
526
)
58
Bernardo
 
A
Ball
 
C
Nolasco
 
L
Choi
 
H
Moake
 
JL
Dong
 
JF
Platelets adhered to endothelial cell-bound ultra-large von Willebrand factor strings support leukocyte tethering and rolling under high shear stress.
J Thromb Haemost
2005
, vol. 
3
 
3
(pg. 
562
-
570
)
59
Savage
 
B
Sixma
 
JJ
Ruggeri
 
ZM
Functional self-association of von Willebrand factor during platelet adhesion under flow.
Proc Natl Acad Sci U S A
2002
, vol. 
99
 
1
(pg. 
425
-
430
)
60
Schneider
 
SW
Nuschele
 
S
Wixforth
 
A
, et al. 
Shear-induced unfolding triggers adhesion of von Willebrand factor fibers.
Proc Natl Acad Sci U S A
2007
, vol. 
104
 
19
(pg. 
7899
-
7903
)
61
Siedlecki
 
CA
Lestini
 
BJ
Kottke-Marchant
 
KK
Eppell
 
SJ
Wilson
 
DL
Marchant
 
RE
Shear-dependent changes in the three-dimensional structure of human von Willebrand factor.
Blood
1996
, vol. 
88
 
8
(pg. 
2939
-
2950
)
62
Steppich
 
DM
Angerer
 
JI
Opfer
 
J
, et al. 
Relaxation of ultralarge VWF bundles in a microfluidic-AFM hybrid reactor [published correction appears in Biochem Biophys Res Commun. 2013;441(1):272].
Biochem Biophys Res Commun
2008
, vol. 
369
 
2
(pg. 
507
-
512
)
63
Ulrichts
 
H
Vanhoorelbeke
 
K
Girma
 
JP
Lenting
 
PJ
Vauterin
 
S
Deckmyn
 
H
The von Willebrand factor self-association is modulated by a multiple domain interaction.
J Thromb Haemost
2005
, vol. 
3
 
3
(pg. 
552
-
561
)
64
Choi
 
H
Aboulfatova
 
K
Pownall
 
HJ
Cook
 
R
Dong
 
JF
Shear-induced disulfide bond formation regulates adhesion activity of von Willebrand factor.
J Biol Chem
2007
, vol. 
282
 
49
(pg. 
35604
-
35611
)
65
Li
 
Y
Choi
 
H
Zhou
 
Z
, et al. 
Covalent regulation of ULVWF string formation and elongation on endothelial cells under flow conditions.
J Thromb Haemost
2008
, vol. 
6
 
7
(pg. 
1135
-
1143
)
66
Essex
 
DW
Redox control of platelet function.
Antioxid Redox Signal
2009
, vol. 
11
 
5
(pg. 
1191
-
1225
)
67
Essex
 
DW
Li
 
M
Redox control of platelet aggregation.
Biochemistry
2003
, vol. 
42
 
1
(pg. 
129
-
136
)
68
Lahav
 
J
Gofer-Dadosh
 
N
Luboshitz
 
J
Hess
 
O
Shaklai
 
M
Protein disulfide isomerase mediates integrin-dependent adhesion.
FEBS Lett
2000
, vol. 
475
 
2
(pg. 
89
-
92
)
69
Lahav
 
J
Jurk
 
K
Hess
 
O
, et al. 
Sustained integrin ligation involves extracellular free sulfhydryls and enzymatically catalyzed disulfide exchange.
Blood
2002
, vol. 
100
 
7
(pg. 
2472
-
2478
)
70
Lahav
 
J
Wijnen
 
EM
Hess
 
O
, et al. 
Enzymatically catalyzed disulfide exchange is required for platelet adhesion to collagen via integrin alpha2beta1.
Blood
2003
, vol. 
102
 
6
(pg. 
2085
-
2092
)
71
Laragione
 
T
Bonetto
 
V
Casoni
 
F
, et al. 
Redox regulation of surface protein thiols: identification of integrin alpha-4 as a molecular target by using redox proteomics.
Proc Natl Acad Sci U S A
2003
, vol. 
100
 
25
(pg. 
14737
-
14741
)
72
Ni
 
H
Li
 
A
Simonsen
 
N
Wilkins
 
JA
Integrin activation by dithiothreitol or Mn2+ induces a ligand-occupied conformation and exposure of a novel NH2-terminal regulatory site on the beta1 integrin chain.
J Biol Chem
1998
, vol. 
273
 
14
(pg. 
7981
-
7987
)
73
Nolan
 
SM
Mathew
 
EC
Scarth
 
SL
Al-Shamkhani
 
A
Law
 
SK
The effects of cysteine to alanine mutations of CD18 on the expression and adhesion of the CD11/CD18 integrins.
FEBS Lett
2000
, vol. 
486
 
2
(pg. 
89
-
92
)
74
O’Neill
 
S
Robinson
 
A
Deering
 
A
Ryan
 
M
Fitzgerald
 
DJ
Moran
 
N
The platelet integrin alpha IIbbeta 3 has an endogenous thiol isomerase activity.
J Biol Chem
2000
, vol. 
275
 
47
(pg. 
36984
-
36990
)
75
Walsh
 
GM
Sheehan
 
D
Kinsella
 
A
Moran
 
N
O’Neill
 
S
Redox modulation of integrin [correction of integin] alpha IIb beta 3 involves a novel allosteric regulation of its thiol isomerase activity.
Biochemistry
2004
, vol. 
43
 
2
(pg. 
473
-
480
)
76
Yan
 
B
Smith
 
JW
A redox site involved in integrin activation.
J Biol Chem
2000
, vol. 
275
 
51
(pg. 
39964
-
39972
)
77
Yan
 
B
Smith
 
JW
Mechanism of integrin activation by disulfide bond reduction.
Biochemistry
2001
, vol. 
40
 
30
(pg. 
8861
-
8867
)
78
Xiong
 
JP
Mahalingham
 
B
Alonso
 
JL
, et al. 
Crystal structure of the complete integrin alphaVbeta3 ectodomain plus an alpha/beta transmembrane fragment.
J Cell Biol
2009
, vol. 
186
 
4
(pg. 
589
-
600
)
79
Zhu
 
J
Luo
 
BH
Xiao
 
T
Zhang
 
C
Nishida
 
N
Springer
 
TA
Structure of a complete integrin ectodomain in a physiologic resting state and activation and deactivation by applied forces.
Mol Cell
2008
, vol. 
32
 
6
(pg. 
849
-
861
)
80
Chen
 
P
Melchior
 
C
Brons
 
NH
Schlegel
 
N
Caen
 
J
Kieffer
 
N
Probing conformational changes in the I-like domain and the cysteine-rich repeat of human beta 3 integrins following disulfide bond disruption by cysteine mutations: identification of cysteine 598 involved in alphaIIbbeta3 activation.
J Biol Chem
2001
, vol. 
276
 
42
(pg. 
38628
-
38635
)
81
Kamata
 
T
Ambo
 
H
Puzon-McLaughlin
 
W
, et al. 
Critical cysteine residues for regulation of integrin alphaIIbbeta3 are clustered in the epidermal growth factor domains of the beta3 subunit.
Biochem J
2004
, vol. 
378
 
Pt 3
(pg. 
1079
-
1082
)
82
Mor-Cohen
 
R
Rosenberg
 
N
Landau
 
M
Lahav
 
J
Seligsohn
 
U
Specific cysteines in beta3 are involved in disulfide bond exchange-dependent and -independent activation of alphaIIbbeta3.
J Biol Chem
2008
, vol. 
283
 
28
(pg. 
19235
-
19244
)
83
Mor-Cohen
 
R
Rosenberg
 
N
Einav
 
Y
, et al. 
Unique disulfide bonds in epidermal growth factor (EGF) domains of β3 affect structure and function of αIIbβ3 and αvβ3 integrins in different manner.
J Biol Chem
2012
, vol. 
287
 
12
(pg. 
8879
-
8891
)
84
Chen
 
VM
Hogg
 
PJ
Encryption and decryption of tissue factor.
J Thromb Haemost
2013
, vol. 
11
 
Suppl 1
(pg. 
277
-
284
)
85
Wang
 
P
Wu
 
Y
Li
 
X
Ma
 
X
Zhong
 
L
Thioredoxin and thioredoxin reductase control tissue factor activity by thiol redox-dependent mechanism.
J Biol Chem
2013
, vol. 
288
 
5
(pg. 
3346
-
3358
)
86
Ahamed
 
J
Versteeg
 
HH
Kerver
 
M
, et al. 
Disulfide isomerization switches tissue factor from coagulation to cell signaling.
Proc Natl Acad Sci U S A
2006
, vol. 
103
 
38
(pg. 
13932
-
13937
)
87
Langer
 
F
Spath
 
B
Fischer
 
C
, et al. 
Rapid activation of monocyte tissue factor by antithymocyte globulin is dependent on complement and protein disulfide isomerase.
Blood
2013
, vol. 
121
 
12
(pg. 
2324
-
2335
)
88
Langer
 
F
Ruf
 
W
Synergies of phosphatidylserine and protein disulfide isomerase in tissue factor activation [published online ahead of print January 23, 2014].
Thromb Haemost
2014
, vol. 
111
 
3
89
Giannakopoulos
 
B
Gao
 
L
Qi
 
M
, et al. 
Factor XI is a substrate for oxidoreductases: enhanced activation of reduced FXI and its role in antiphospholipid syndrome thrombosis.
J Autoimmun
2012
, vol. 
39
 
3
(pg. 
121
-
129
)
90
Ahamed
 
J
Burg
 
N
Yoshinaga
 
K
Janczak
 
CA
Rifkin
 
DB
Coller
 
BS
In vitro and in vivo evidence for shear-induced activation of latent transforming growth factor-beta1.
Blood
2008
, vol. 
112
 
9
(pg. 
3650
-
3660
)
91
Brophy
 
TM
Coller
 
BS
Ahamed
 
J
Identification of the thiol isomerase-binding peptide, mastoparan, as a novel inhibitor of shear-induced transforming growth factor β1 (TGF-β1) activation.
J Biol Chem
2013
, vol. 
288
 
15
(pg. 
10628
-
10639
)
92
Leppänen
 
VM
Jeltsch
 
M
Anisimov
 
A
, et al. 
Structural determinants of vascular endothelial growth factor-D receptor binding and specificity.
Blood
2011
, vol. 
117
 
5
(pg. 
1507
-
1515
)
93
Chiu
 
J
Wong
 
JW
Gerometta
 
M
Hogg
 
PJ
Mechanism of dimerization of a recombinant mature vascular endothelial growth factor C.
Biochemistry
2014
, vol. 
53
 
1
(pg. 
7
-
9
)
94
Wong
 
JW
Hogg
 
PJ
Analysis of disulfide bonds in protein structures.
J Thromb Haemost
2010
, vol. 
8
 pg. 
2345
 
95
Schmidt
 
B
Ho
 
L
Hogg
 
PJ
Allosteric disulfide bonds.
Biochemistry
2006
, vol. 
45
 
24
(pg. 
7429
-
7433
)
96
Schmidt
 
B
Hogg
 
PJ
Search for allosteric disulfide bonds in NMR structures.
BMC Struct Biol
2007
, vol. 
7
 pg. 
49
 
97
Metcalfe
 
C
Cresswell
 
P
Ciaccia
 
L
Thomas
 
B
Barclay
 
AN
Labile disulfide bonds are common at the leucocyte cell surface.
Open Biol
2011
, vol. 
1
 
3
pg. 
110010
 
98
Chen
 
VM
Ahamed
 
J
Versteeg
 
HH
Berndt
 
MC
Ruf
 
W
Hogg
 
PJ
Evidence for activation of tissue factor by an allosteric disulfide bond.
Biochemistry
2006
, vol. 
45
 
39
(pg. 
12020
-
12028
)

Author notes

D.B. and K.M.C. contributed equally to this study.

Sign in via your Institution