Your activity: 22 p.v.
your limit has been reached. plz Donate us to allow your ip full access, Email: sshnevis@outlook.com

Vitamin K and the synthesis and function of gamma-carboxyglutamic acid

Vitamin K and the synthesis and function of gamma-carboxyglutamic acid
Authors:
Bruce Furie, MD
Beth A Bouchard, PhD
Section Editor:
Lawrence LK Leung, MD
Deputy Editor:
Jennifer S Tirnauer, MD
Literature review current through: Dec 2022. | This topic last updated: Jul 19, 2021.

INTRODUCTION — Vitamin K is a cofactor for the enzymatic conversion of glutamic acid (Glu) residues to gamma-carboxyglutamic acid (Gla) in vitamin K-dependent proteins, via the endoplasmic reticulum resident vitamin K-dependent gamma-glutamyl carboxylase. This carboxylase activity is found in essentially all mammalian tissues, and its reaction product, Gla, has been observed in both vertebrates and invertebrates; both play an important biological role in protein function [1].

The functions of Gla and the vitamin K-dependent biosynthesis of Gla will be discussed here. An overview of the blood coagulation cascade is presented separately. (See "Overview of hemostasis".)

BACKGROUND AND HISTORY — Vitamin K and its association with blood coagulation were initially described in the 1920s and 1930s, following investigation of a hemorrhagic disease of cattle caused by ingestion of spoiled sweet clover [2]. Since then, a number of observations, which will be discussed in detail below, have improved our understanding of the biological role of vitamin K [3]:

Discovery of vitamin K antagonists and their introduction as pharmacologic agents for anticoagulation (ie, the coumarins) [4].

Discovery of Gla in blood clotting proteins [5,6].

Identification of Gla as an amino acid that confers metal binding properties on proteins, a requirement for protein-membrane interaction [7,8].

Detection of an enzymatic activity (ie, the vitamin K-dependent gamma-glutamyl carboxylase) that catalyzes the incorporation of CO2 into glutamic acid to form Gla [7].

Identification of an intervening sequence (the propeptide) between the signal peptide and the mature vitamin K-dependent protein [9].

Discovery of the requirement for [10] and the sufficiency of [11] the gamma-carboxylation recognition site within the propeptide in directing synthesis of gamma-carboxyglutamic acid on the adjacent Gla domain of the precursor protein.

Additional advances include purification and cloning of the vitamin K-dependent carboxylase [12,13]; determination of the three-dimensional structure of the Gla domain of prothrombin and observation of an internal carboxylate-calcium network [14]; the proposal of a mechanism of vitamin K-mediated enhancement of carboxylase action [15]; and regulation of enzymatic vitamin K epoxidase activity by glutamate containing substrate [16].

BIOSYNTHESIS OF GAMMA-CARBOXYGLUTAMIC ACID — The requirement for vitamin K as an enzyme cofactor is unique to the vitamin K-dependent gamma-glutamyl carboxylase and the biosynthesis of gamma-carboxyglutamic acid (Gla). The mechanism by which vitamin K participates as a cofactor with the carboxylase is incompletely understood. The most attractive hypothesis is that an activated species of vitamin K abstracts a hydrogen from the gamma-carbon of Glu residues on specific proteins (factors VII, IX, X, and prothrombin; proteins S and C; and, in bone, osteocalcin [bone Gla protein] and matrix Gla protein), followed by transformation of the vitamin K intermediate to an epoxide. Carbon dioxide is subsequently added to the gamma-carbon of glutamic acid, utilizing the carboxylase activity of gamma-glutamyl carboxylase, to form Gla (figure 1) [17].

Based on a nonenzymatic model [15], a "base strength amplification mechanism" has been proposed to explain the conversion of the weak base form of vitamin K (vitamin KH2) into an oxygenated intermediate (the alkoxide) of sufficient basicity to abstract a hydrogen from the gamma-carbon of Glu (figure 1) [15]. It was initially proposed that a free cysteine in the gamma-glutamyl carboxylase was responsible for deprotonation of vitamin KH2, a hypothesis that is supported by numerous publications [18-28]. However, other studies suggest that the role of a free cysteine residue in carboxylation is uncertain, and implicate two lysine residues in the gamma-glutamyl carboxylase as important for catalysis [29,30].

Gamma-glutamyl carboxylase has two actions: it promotes the formation of Gla on the Glu residues; and it promotes the formation of the highly reactive alkoxide which then collapses into the epoxide form of vitamin K [16,18,31]. Thus, under normal conditions, for each molecule of Gla generated, one molecule of vitamin K epoxide is also formed [31,32].

The short-lived highly reactive alkoxide is potentially toxic, and it would be undesirable for it to be generated in the absence of Glu residues. Evidence indicates that no highly reactive vitamin K intermediate is generated by the carboxylase until a carboxylase substrate is bound to the enzyme and converts its vitamin K epoxidase function to an active state [16]. The epoxide is then recycled back to vitamin KH2 via a vitamin K reductase (see below).

The propeptide region of the vitamin K-dependent proteins functions as a recognition sequence, binding the carboxylase to its substrate on the adjacent glutamic acid rich (Gla) domain (see below) [10,11]. It also stimulates gamma-glutamyl carboxylation [33,34].

RECYCLING OF VITAMIN K — In its naturally occurring form, vitamin K is in the quinone oxidation state and must be reduced to the hydroquinone form (vitamin KH2), the active cofactor for the vitamin K-dependent carboxylase. The enzyme responsible for this conversion is known as vitamin K epoxide reductase (VKOR) because it also reduces the vitamin K epoxide formed during the carboxylation reaction [35]. Therapeutic doses of warfarin, which shares a common ring structure with vitamin K (figure 2), inhibit this reductase by binding to the same key residues and inducing the same conformational changes required for the VKOR catalytic cycle [36], resulting in insufficient generation of vitamin K hydroquinone to support full carboxylation and therefore full function of the vitamin K-dependent proteins of blood coagulation.

The gene encoding VKOR resides on chromosome 16 and encodes a protein of 163 amino acids [37,38]. VKOR is effective at low concentrations of vitamin K epoxide and vitamin K quinone and is the physiologically important enzyme for recycling vitamin K [39,40].

Homologs of VKOR have been identified in a variety of organisms [41-43]. Like VKOR, all the homologs contain an active site CXXC motif (C132-X-X-C135 in human VKOR) that forms the redox center and can directly interact with substrates, as well as an additional pair of cysteines that are conserved [37,38]. Molecular modeling and in vitro experiments with human VKOR suggest that residues F55, N80, and F83 act in a concerted manner to localize vitamin K epoxide close to the C135 catalytic residue [44]. Bacterial VKOR homologs catalyze disulfide bond formation in secreted proteins [43].

The reducing equivalents required by the bacterial VKOR are provided by periplasmic thioredoxin-like proteins or by a thioredoxin-like domain fused to the VKOR domain. Reducing equivalents required for conversion of vitamin K quinone or epoxide to the hydroquinone by mammalian VKOR are provided by endoplasmic reticulum anchored thioredoxin-like protein, TMX [45]. Determination of the crystal structure of a naturally occurring bacterial VKOR-thioredoxin domain fusion protein has suggested an electron transfer pathway that could apply to all VKOR homologs [46]. Newly synthesized proteins in the periplasm or endoplasmic reticulum reduce the CXXC motif of a thioredoxin-like protein, which in turn reduces a conserved disulfide bridge in a loop of VKOR. The loop cysteines reduce the CXXC motif of VKOR and finally the disulfide in this CXXC motif is reestablished by reduction of a quinone [46]. However, in mammalian cells, only a minor fraction (5.6 percent) of VKOR is in a fully reduced state, with the remainder split almost equally between partially oxidized and fully oxidized [47]. VKOR catalysis and electron transfer can initiate from either the fully reduced or partially oxidized state.

A second enzyme, DT-diaphorase, an NAD(P)H dehydrogenase, reduces the quinone form of vitamin K but not vitamin K epoxide [35]. This enzyme requires high concentrations of vitamin K and probably does not play a role in recycling at physiologic tissue concentrations of vitamin K [39]. It may, however, be important when vitamin K, in the quinone form, is used to overcome warfarin intoxication or poisoning with vitamin K antagonist rodenticides [48]. These "superwarfarins" are VKOR inhibitors that are two orders of magnitude more potent than warfarin and have half-lives measured in weeks. Treatment with high doses of vitamin K quinine for several months or years is often required to reverse the coagulopathy associated with superwarfarin poisoning. (See "Overview of rodenticide poisoning", section on 'Anticoagulants (superwarfarins and warfarins)' and "Management of warfarin-associated bleeding or supratherapeutic INR".)

Vitamin K epoxide is not detectable in the plasma of normal subjects, even after a pharmacologic dose of 10 mg of vitamin K [49]. However, it is measurable following the administration of warfarin [50,51]. The finding in warfarin-treated patients of a significant positive correlation between the INR and plasma vitamin K epoxide concentrations suggests that the latter reflects the pharmacodynamic activity of warfarin in anticoagulated patients [52].

Genetic variants of the VKOR complex — Pathogenic variants affecting the VKOR complex have been described in a number of families with deficiency of the vitamin K-dependent coagulation proteins [53-56]. Consanguinity was present in one family. The defect was corrected by supplementation with vitamin K in a number of these families and partially corrected in another [55]. (See "Rare inherited coagulation disorders", section on 'Multiple vitamin K-dependent factor deficiencies'.)

Hereditary warfarin resistance — Missense mutations within the gene for VKOR (VKORC1), have been proposed to be involved in resistance to vitamin K antagonist anticoagulants [37,57-60]. In addition, common variants within the VKORC1 gene appear to modulate the mean daily dose of warfarin required to acquire target anticoagulation intensity. Despite this observation, several studies have demonstrated that genotyping to determine the initial dosing of warfarin does not improve outcomes [61]. (See "Biology of warfarin and modulators of INR control", section on 'Genetic factors'.)

Other factors that affect warfarin dosing are discussed separately. (See "Warfarin and other VKAs: Dosing and adverse effects", section on 'Warfarin resistance'.)

Variants in VKORC1 may also be associated with increased risk for arterial vascular disease [62]. Missense mutation of VKORC1 coupled with limited vitamin K ingestion in rats leads to areas of massive vascular calcification associated with increased expression of uncarboxylated matrix Gla protein [63]; this is consistent with previous observations made in mice deficient in matrix Gla protein [64].

THE GAMMA-CARBOXYLATION RECOGNITION SEQUENCE — The propeptides of the vitamin K-dependent proteins contain a gamma-carboxylation recognition site [10,11]. In the vitamin K-dependent coagulation proteins (factors VII, IX, X, prothrombin) and regulatory proteins (proteins S and C), the recognition site is within the propeptide region; in contrast, the gamma-carboxylation recognition site resides within the mature protein sequence itself in matrix Gla protein [65].

The amino acids of this recognition site bind directly to the vitamin K-dependent carboxylase [66]. Although no consensus sequence prevails in the carboxylation recognition sites, these sites are best defined by a Z-F-Z-X-X-X-X-A motif, where Z is an aliphatic hydrophobic residue (isoleucine, valine, or leucine), F is phenylalanine, A is alanine, and X is any amino acid. Phenylalanine at residue 16 is preferred in carboxylase substrates, but leucine, valine, and lysine at this position also support carboxylation [67].

Disruption of the carboxylation recognition site in the propeptide region of factor IX yields a final protein that either lacks or is deficient in Gla [10]. Thus, this recognition site is required for gamma-carboxylation to occur. It is most likely that this region of the blood coagulation protein precursor docks with the membrane-bound carboxylase, bringing the active site of the carboxylase in close proximity to the substrate Glu residues on the precursor form of the vitamin K-dependent proteins. A mutation in factor VII, factor VII Tokushima (Cys22 —> Gly), blocks gamma carboxylation due to a disrupted gamma-carboxylase recognition site [68]. Glu residues contribute significantly to the recognition of the protein substrate by the carboxylase [11].

To test this hypothesis concerning the function of the recognition site, a prothrombin propeptide/thrombin chimera was constructed, which had the signal peptide and gamma-carboxylation recognition site-containing propeptide of prothrombin juxtaposed to a glutamate-rich C-terminal region of thrombin. Of the eight glutamic acids within the first 40 residues of the NH2-terminus of prothrombin adjacent to the propeptide, which are not normally gamma-carboxylated, at least seven underwent complete carboxylation when this chimeric protein was expressed in CHO cells.

These results indicate that the prothrombin gamma-carboxylation recognition site on the propeptide is sufficient to direct carboxylation of adjacent Glu residues in the propeptide/thrombin chimera by the carboxylase, without regard for the sequence context of the Glu substrate or structures defined by disulfide bonds. Thus, we hypothesize that any protein will undergo gamma-carboxylation if it meets the following criteria:

The protein includes a gamma-carboxylation recognition site that interacts with the gamma-glutamyl carboxylase

The protein is routed through the rough endoplasmic reticulum during biosynthesis

The cell has the carboxylase enzyme associated with the rough endoplasmic reticulum

There are Glu sites within 40 residues of the gamma-carboxylation recognition site

Intracellular vitamin K is present

Recognition site mutations — Additional factors may also be important for the carboxylation recognition site. As an example of the importance of alanine at position 10 as part of the carboxylation recognition site, an experimentally induced mutation at this site is associated with diminished gamma-carboxylation [10,69,70]. In addition, naturally occurring mutations of alanine 10 to threonine or valine in factor IX are associated with a marked increase in sensitivity to warfarin or phenprocoumon that is limited to the patient's factor IX activity [70-73].

The phenotype, factor IX levels, and factor IX carboxylation of these patients with mutations at alanine 10 were all normal; however, there was a more marked reduction in factor IX levels than in other vitamin K-dependent proteins after exposure to a vitamin K antagonist. It is possible that, with a reduced affinity for the carboxylase, the mutant profactor IX may be at a kinetic disadvantage when the concentration of the other carboxylase substrate, reduced vitamin K, is limited by coumarin-induced inhibition of the recycling enzymes. In contrast, the other unaffected vitamin K-dependent proteins can still compete effectively for enzyme and cofactor.

VITAMIN K-DEPENDENT CARBOXYLASE — The vitamin K-dependent carboxylase is an integral membrane protein, requiring carbon dioxide, molecular oxygen, and the hydroquinone form of vitamin K to convert Glu residues to Gla (figure 1) [13,74]. Molecular cloning of the human and bovine vitamin K-dependent carboxylase predicted a single chain protein of 758 amino acids with a molecular weight of about 94,000 [13,75]. The human carboxylase gene is localized on chromosome 2p12, and is encoded by 15 exons [76].

The known functional properties of this enzyme include:

A carboxylase active site

An epoxidase active site

A propeptide binding site that allows substrate to attach and shares sequence similarity with the propeptide of the carboxylase substrate [77]

A propeptide binding site that stimulates carboxylase and epoxidase activity

A glutamate binding site [78]

A vitamin K binding site

Mutations — Mutations of the carboxylase gene have been described in a number of families, leading to a congenital bleeding disorder with deficiency of all vitamin K-dependent coagulation factors [79-81]. A similar defect has been found in the Devon Rex cat [82], and mice carrying a null mutation of the carboxylase gene die at birth of massive intra-abdominal hemorrhage [83]. In most of the affected humans and in cats, the activity of the vitamin K-dependent coagulation factors can be normalized by vitamin K supplementation [53,82,84,85]. This normalization of activity is possible because the effect of most of the reported mutations is to weaken the binding of vitamin K to the enzyme. Thus, increasing the concentration of vitamin K raises the levels of the vitamin K-dependent coagulation factors. (See "Rare inherited coagulation disorders", section on 'Multiple vitamin K-dependent factor deficiencies'.)

INTRACELLULAR SITE OF GAMMA-CARBOXYLATION — Use of anticarboxylase antibodies has confirmed the intracellular localization of the carboxylase in both the endoplasmic reticulum and the Golgi apparatus. Studies in Chinese hamster ovary cells expressing prothrombin showed that uncarboxylated proprothrombin is completely gamma-carboxylated in the endoplasmic reticulum [86]. The carboxylated proprothrombin leaves the endoplasmic reticulum intact and is further processed in the Golgi apparatus to remove the propeptide.

Prothrombin and the other gamma-carboxylated extracellular vitamin K-dependent proteins bind to acidic membranes in the presence of calcium ions. The question therefore arises as to the mechanism which prevents the fully gamma-carboxylated intracellular precursor vitamin K-dependent proteins, such as proprothrombin, from binding to endoplasmic reticular membranes during transit through the biosynthetic pathway. The calcium concentration in the endoplasmic reticulum is high, in contrast to the cytoplasm, and is sufficient to support protein-membrane interaction.

Fully carboxylated profactor IX does not bind to membranes in the presence of calcium ions, whereas factor IX does [87]. Thus, the propeptide attached to factor IX appears to prevent proper folding of the Gla domain, expression of the phospholipid binding site, and the interaction of profactor IX with membranes.

FUNCTION OF GAMMA-CARBOXYGLUTAMIC ACID — Gamma-carboxyglutamic acid has been primarily studied in two protein families: the vitamin K-dependent blood coagulation (factors VII, IX, X, and prothrombin) and regulatory (proteins C and S) proteins; and proteins of mineralized tissue (bone Gla protein and matrix Gla protein). In addition, several proteins that contain Gla domains homologous to those of blood coagulation but of unknown function have been identified. These include Gas6 and four proteins that contain transmembrane regions as well as a Gla domain.

Gamma-carboxyglutamic acid distinguishes itself from aspartic acid and glutamic acid by containing two carboxyl groups in its side chain. The bivalent nature of Gla is similar to the immunoglobulins or fibrinogen. In these cases, a structural framework is formed via the linking of one ligand to another through a common bivalent, symmetrical molecule. As is described below, this also applies to the formation of the calcium-carboxylate network that stabilizes the Gla domains and, in the vitamin K-dependent proteins of blood coagulation, allows expression of the phospholipid binding site.

Plasma proteins of blood coagulation — The vitamin K-dependent blood coagulation and regulatory proteins contain 10 to 12 Gla residues in the Gla domain, which is located within the first 40 residues of the N-termini of the mature proteins. The Gla domain, in association with the adjacent aromatic amino acid stack domain, functions as a membrane binding component of these proteins.

Gla is an amino acid that confers metal binding properties on the vitamin K-dependent proteins. With the addition of calcium ions, these proteins undergo a structural transition that leads to exposure of a phospholipid binding site. In most cases, neither aspartic acid nor Glu will substitute for the function of Gla, emphasizing the importance of the presence of both carboxyl groups on a single amino acid.

The inability of Glu to substitute for Gla and the importance of vitamin K are best illustrated clinically by the usefulness of warfarin. Warfarin prevents the recycling of vitamin K from the oxidized (epoxide) to the reduced form (hydroquinone), leading to decreased gamma-glutamyl carboxylation because of insufficient active cofactor (the hydroquinone) (figure 1). The abnormal forms of the vitamin K-dependent proteins that result from the use of warfarin are undercarboxylated, having Glu residues at some or all of the positions that are usually carboxylated (ie, Gla). These proteins are for the most part biologically inert. Thus, in patients treated with warfarin, circulating coagulant activity correlates closely with the quantity of fully carboxylated proteins remaining in the circulation [88].

Elimination of even one Gla residue can significantly reduce the biological function of these proteins [89]. In addition, systematic mutation of individual carboxylatable Glu residues in recombinant prothrombin [89] or protein C [90-92] has yielded molecules with modest to severe functional defects, again indicating the importance of the bifunctional Gla side chain.

Prothrombin — The x-ray crystal structure of prothrombin fragment 1 indicates that the Gla domain is highly structured; many of the Gla side chains point inward to a linear array of internal calcium ions [14]. Several of these calcium ions are completely sequestered inside the core of the Gla domain.

Factor IX — The factor IX Gla domain is characterized by a fold of the polypeptide backbone, similar to that seen in prothrombin [93]. The location of the calcium binding sites in the internal core structure is nearly identical in prothrombin and factor IX, although there are some important distinguishing features [94,95].

Factor VII — The crystal structure of the factor VII-tissue factor complex indicates that the structure of the Gla domain of factor VII in the presence of calcium ions is nearly identical to its homolog in prothrombin [96].

Phospholipid membrane binding site — The phospholipid membrane binding site of the vitamin K-dependent proteins is expressed on the surface of the Gla domain. X-ray structure had shown unexpected exposure of three hydrophobic residues on the surface of these proteins, suggesting a potential role for these residues in membrane interaction [14]. Site-specific mutagenesis of homologous residues in protein C interfered with the membrane binding properties of this protein, but mutation of other hydrophobic residues did not perturb membrane binding significantly [97].

It has been concluded that calcium-induced exposure of hydrophobic amino acids in the Gla domain is critical for membrane binding [98]. Direct comparison of the structures of the calcium-stabilized form of factor IX, which binds to membranes, and the magnesium-stabilized form of factor IX, which does not, implicated residues 1 to 11 that form a loop in the Gla domain [99]. Correlation of membrane-binding properties of the vitamin K-dependent proteins with homology considerations suggests a possible alternative membrane contact site that involves residues 11, 33, and 34 [100].

The structure of the lysophosphatidylserine binding site in the bovine prothrombin Gla domain was determined by x-ray crystallography [101]. The serine head group binds Gla domain-bound calcium ions and Gla residues 17 and 21, fixed elements of the Gla domain fold, predicting the structural basis for phosphatidylserine specificity among Gla domains. A molecular dynamics simulation study of anionic phospholipid binding to the Gla domain of bovine fragment 1 in the presence of calcium ions has identified a second phosphatidylserine binding site with the head group bound by a Gla-bound calcium ion, Gla 30 and Lys 11 [102]. It can be concluded that Gla domains provide a unique mechanism for protein-phospholipid membrane interaction.

Gla-containing proteins of mineralized tissue — Whereas the role of Gla is well defined for the plasma coagulation proteins, its function remains less clear in proteins/peptides outside of this family. There have been no successful studies to date on the structure of Gla-containing bone proteins, including osteocalcin (bone Gla protein), the most abundant osteoblast-specific noncollagenous protein, and matrix Gla protein. (See "Normal skeletal development and regulation of bone formation and resorption".)

There is also no evidence to support the use of vitamin K administration in patients with osteoporosis or, in a patient receiving a vitamin K antagonist, to support the use of additional interventions over and above those routinely used for osteoporosis prevention. These subjects are discussed in more detail separately. (See "Overview of the management of osteoporosis in postmenopausal women", section on 'Therapies not recommended' and "Prevention of osteoporosis" and "Drugs that affect bone metabolism", section on 'Anticoagulants'.)

Some insights into the function of osteocalcin and matrix Gla protein have come from experimental models lacking these proteins.

Osteocalcin deficient mice are characterized by increased bone formation, including higher bone mass and bones of improved functional quality; these observations suggest an important role for osteocalcin in the regulation of bone remodeling [103].

Matrix Gla protein deficient mice suffer from spontaneous and ultimately fatal calcification of arteries and cartilage [64], suggesting that one of the functions of this protein is to control and limit extraosseous calcification. The Gla residues in matrix Gla protein prevent osteogenic differentiation and calcification via binding and inhibition of bone morphogenetic protein-4 [104], as well as via binding to vascular smooth muscle cell-derived vesicles [105].

Other mammalian vitamin K-dependent proteins — Several other families of vitamin K-dependent proteins described in mammalian systems are Gas6 and the proline-rich Gla proteins.

Gas6 protein — Gas6 (growth-arrest-specific gene 6) has marked sequence homology in the Gla domain to the vitamin K-dependent blood coagulation and regulatory proteins, in particular protein S [106]. Gas6 is released from and potentiates the growth of vascular smooth muscle cells. When synthesized in the presence of warfarin and therefore lacking Gla, Gas6 demonstrates no thrombin-inducible growth potentiating activity or receptor binding ability [107]. In contrast, Gas6 lacking the entire Gla domain is a functional growth factor. This observation suggests that the Gla domain may be a negative regulator of the structure of a growth factor domain located elsewhere on the molecule.

Vitamin K-dependent single-pass integral membrane proteins — This family of four proteins, identified by searching a database with a consensus sequence derived from analysis of known Gla domains, can be divided into two subclasses based on their gene organization and protein sequence.

Proline rich Gla proteins, PRGP1 and PRGP2, are named for their proline-rich Gla protein characteristic [108].

Transmembrane Gla proteins, TMG3 and TMG4, were subsequently identified [109].

All of these proteins are expressed variably in fetal and adult tissue. While their functions are unknown, all of these proteins contain the PPXY motif involved in diverse cellular functions. The gene for TMG4 is one of those deleted in the 11p14-p12 chromosome region deletion associated with WAGR syndrome and falls into the linkage region of 11p13-p12 as an autism candidate gene [110].

Gamma-glutamyl carboxylase — The vitamin K-dependent gamma-glutamyl carboxylase, which converts Glu residues to Gla, has sequence homology with the region of matrix Gla protein containing the gamma-carboxylation recognition site [111]. Under some conditions, the carboxylase can become carboxylated, although the role of Gla in this enzyme remains to be determined [112].

Protein Z — Protein Z is a 62,000 molecular weight vitamin K-dependent plasma glycoprotein whose structure is similar to coagulation factors VII, IX, X, and proteins C and S [113]. Protein Z and a protein Z-dependent protease inhibitor (ZPI) appear to serve as cofactors for the inhibition of activated factor X and activated factor XI by a number of different mechanisms [114-116]. ZPI inhibition of factor Xa bound to phospholipid membranes, but not in solution, is dependent upon an interaction between the Gla domains of protein Z and factor Xa [117]. Deficiencies of protein Z or ZPI, as with deficiencies of proteins C and S, might therefore result in a prothrombotic state (eg, arterial thrombosis, pregnancy complications, venous thromboembolism) [118-123].

In support of this supposition are the following highly preliminary observations:

Protein Z deficiency dramatically increases the severity of the prothrombotic phenotype of factor V Leiden in a transgenic mouse model [113]. Similarly, protein Z deficiency or specific polymorphisms within the protein Z gene appear to influence clinical symptoms of thromboembolism in human subjects with factor V Leiden [124,125].

In a comparison of 200 women with one unexplained primary episode of early fetal loss (10th to 15th week of gestation) and 200 controls who had normal pregnancies, protein Z deficiency was significantly more common in the women with fetal loss (22 versus 4 percent) [126].

In two studies, nonsense mutations of the ZPI gene were found in a significantly higher percent of patients with thrombosis compared with controls, with odds ratios of 5.7 and 3.3, respectively [116,127].

The relationships between protein Z levels and haplotypes and ischemic stroke, Sneddon's syndrome, and atherosclerotic disease are unclear [128-134]. The significance of low levels of protein Z in patients with antiphospholipid antibodies is similarly unknown [135,136].

Gla-rich proteins — The Gla-rich protein (GRP), so-named because it contains the highest Gla content of any known protein (16 of 74 residues), was identified in the calcified cartilage of the Adriatic sturgeon [137]. The Gla containing region of GRP is not homologous to any other known Gla containing protein. GRP has orthologs in all taxonomic groups of vertebrates. GRP mRNA was found in virtually all rat and sturgeon tissue with the highest levels in cartilage and bone [137]. The protein is expressed in chondrocytes, chondroblasts, osteocytes and osteoblasts. The function of this protein is unknown.

Invertebrate Gla-containing proteins — Only one invertebrate, the cone snail, has been found to contain Gla [138]. This tropical snail stings its prey, whether fish, molluscs, or worms, via a harpoon that injects conotoxins. Many of these conotoxins are Gla-containing peptides that are channel blockers [139,140]. The presence of Gla in cysteine-free conotoxins known as conomarphins appear to preferentially stabilize certain conomarphin conformations, which may be associated with their target specificity and selectivity [141].

SUMMARY

Vitamin K dependent carboxylase and synthesis of Gla – The requirement for vitamin K as an enzyme cofactor is unique to the vitamin K-dependent gamma-glutamyl carboxylase and the biosynthesis of gamma-carboxyglutamic acid (Gla). The vitamin K-dependent carboxylase is an integral membrane protein, requiring carbon dioxide, molecular oxygen, and the hydroquinone form of vitamin K to convert glutamic acid (Glu) residues to Gla (figure 1). Mutations of the carboxylase gene have been described, leading to a congenital bleeding disorder with deficiency of all vitamin K-dependent coagulation factors. (See 'Biosynthesis of gamma-carboxyglutamic acid' above and 'Vitamin K-dependent carboxylase' above.)

Vitamin K recycling – For each molecule of Gla generated, one molecule of vitamin K epoxide is also formed. Vitamin K epoxide reductase (VKOR) is responsible for the conversion of vitamin K to the active cofactor for the vitamin K-dependent carboxylase; it also reduces the vitamin K epoxide formed during the carboxylation reaction. Therapeutic doses of warfarin inhibit this reductase, resulting in insufficient generation of vitamin K hydroquinone to support full carboxylation and therefore full function of the vitamin K-dependent proteins of blood coagulation. (See 'Recycling of vitamin K' above.)

Functions of Gla in clotting – Gla is a post-translationally modified glutamic acid that confers metal binding properties on the vitamin K-dependent proteins. With the addition of calcium ions, these proteins undergo a structural transition that leads to exposure of a phospholipid binding site. Factors that require Gla for their function include prothrombin (factor II) and factors VII, IX, and X, as well as proteins C and S. (See 'Function of gamma-carboxyglutamic acid' above.)

Other functions of Gla – Other Gla-requiring proteins (eg, the bone and matrix Gla proteins, protein Z) are discussed above. (See 'Gla-containing proteins of mineralized tissue' above and 'Other mammalian vitamin K-dependent proteins' above.)

ACKNOWLEDGMENT — The UpToDate editorial staff acknowledges Barbara C Furie, PhD, who contributed to an earlier version of this topic review.

  1. Vermeer C, De Boer-Van den Berg MA. Vitamin K-dependent carboxylase. Haematologia (Budap) 1985; 18:71.
  2. Roderick LM. A problem in the coagulation of the blood "sweet clover" disease of cattle. Am J Physiol 1931; 96:413.
  3. Dam H. The antihemorrhagic vitamin of the chick. Occurrence and chemical nature. Nature 1935; 135:652.
  4. Campbell HA, Link KP. Studies on the hemorrhagic sweet clover disease IV. The isolation and crystallization of the hemorrhagic agent. J Biol Chem 1941; 138:21.
  5. Stenflo J, Fernlund P, Egan W, Roepstorff P. Vitamin K dependent modifications of glutamic acid residues in prothrombin. Proc Natl Acad Sci U S A 1974; 71:2730.
  6. Nelsestuen GL, Zytkovicz TH, Howard JB. The mode of action of vitamin K. Identification of gamma-carboxyglutamic acid as a component of prothrombin. J Biol Chem 1974; 249:6347.
  7. Esmon CT, Sadowski JA, Suttie JW. A new carboxylation reaction. The vitamin K-dependent incorporation of H-14-CO3- into prothrombin. J Biol Chem 1975; 250:4744.
  8. Sperling R, Furie BC, Blumenstein M, et al. Metal binding properties of gamma-carboxyglutamic acid. Implications for the vitamin K-dependent blood coagulation proteins. J Biol Chem 1978; 253:3898.
  9. Kurachi K, Davie EW. Isolation and characterization of a cDNA coding for human factor IX. Proc Natl Acad Sci U S A 1982; 79:6461.
  10. Jorgensen MJ, Cantor AB, Furie BC, et al. Recognition site directing vitamin K-dependent gamma-carboxylation resides on the propeptide of factor IX. Cell 1987; 48:185.
  11. Furie BC, Ratcliffe JV, Tward J, et al. The gamma-carboxylation recognition site is sufficient to direct vitamin K-dependent carboxylation on an adjacent glutamate-rich region of thrombin in a propeptide-thrombin chimera. J Biol Chem 1997; 272:28258.
  12. Wu SM, Morris DP, Stafford DW. Identification and purification to near homogeneity of the vitamin K-dependent carboxylase. Proc Natl Acad Sci U S A 1991; 88:2236.
  13. Wu SM, Cheung WF, Frazier D, Stafford DW. Cloning and expression of the cDNA for human gamma-glutamyl carboxylase. Science 1991; 254:1634.
  14. Soriano-Garcia M, Padmanabhan K, de Vos AM, Tulinsky A. The Ca2+ ion and membrane binding structure of the Gla domain of Ca-prothrombin fragment 1. Biochemistry 1992; 31:2554.
  15. Dowd P, Hershline R, Ham SW, Naganathan S. Vitamin K and energy transduction: a base strength amplification mechanism. Science 1995; 269:1684.
  16. Sugiura I, Furie B, Walsh CT, Furie BC. Propeptide and glutamate-containing substrates bound to the vitamin K-dependent carboxylase convert its vitamin K epoxidase function from an inactive to an active state. Proc Natl Acad Sci U S A 1997; 94:9069.
  17. Friedman PA, Shia MA, Gallop PM, Griep AE. Vitamin K-dependent gamma-carbon-hydrogen bond cleavage and nonmandatory concurrent carboxylation of peptide-bound glutamic acid residues. Proc Natl Acad Sci U S A 1979; 76:3126.
  18. Morris DP, Soute BA, Vermeer C, Stafford DW. Characterization of the purified vitamin K-dependent gamma-glutamyl carboxylase. J Biol Chem 1993; 268:8735.
  19. Mack DO, Suen ET, Girardot JM, et al. Soluble enzyme system for vitamin K-dependent carboxylation. J Biol Chem 1976; 251:3269.
  20. Friedman PA, Shia M. Some characteristics of a vitamin K-dependent carboxylating system from rat liver microsomes. Biochem Biophys Res Commun 1976; 70:647.
  21. Suttie JW, Lehrman SR, Geweke LO, et al. Vitamin K-dependent carboxylase: requirements for carboxylation of soluble peptide and substrate specificity. Biochem Biophys Res Commun 1979; 86:500.
  22. Canfield LM, Sinsky TA, Suttie JW. Vitamin K-dependent carboxylase: purification of the rat liver microsomal enzyme. Arch Biochem Biophys 1980; 202:515.
  23. Canfield LM. Vitamin K-dependent oxygenase/carboxylase; differential inactivation by sulfhydryl reagents. Biochem Biophys Res Commun 1987; 148:184.
  24. Bouchard BA, Furie B, Furie BC. Glutamyl substrate-induced exposure of a free cysteine residue in the vitamin K-dependent gamma-glutamyl carboxylase is critical for vitamin K epoxidation. Biochemistry 1999; 38:9517.
  25. Tie J, Wu SM, Jin D, et al. A topological study of the human gamma-glutamyl carboxylase. Blood 2000; 96:973.
  26. Pudota BN, Miyagi M, Hallgren KW, et al. Identification of the vitamin K-dependent carboxylase active site: Cys-99 and Cys-450 are required for both epoxidation and carboxylation. Proc Natl Acad Sci U S A 2000; 97:13033.
  27. Bandyopadhyay PK, Garrett JE, Shetty RP, et al. gamma -Glutamyl carboxylation: An extracellular posttranslational modification that antedates the divergence of molluscs, arthropods, and chordates. Proc Natl Acad Sci U S A 2002; 99:1264.
  28. Tie JK, Mutucumarana VP, Straight DL, et al. Determination of disulfide bond assignment of human vitamin K-dependent gamma-glutamyl carboxylase by matrix-assisted laser desorption/ionization time-of-flight mass spectrometry. J Biol Chem 2003; 278:45468.
  29. Rishavy MA, Pudota BN, Hallgren KW, et al. A new model for vitamin K-dependent carboxylation: the catalytic base that deprotonates vitamin K hydroquinone is not Cys but an activated amine. Proc Natl Acad Sci U S A 2004; 101:13732.
  30. Tie JK, Jin DY, Loiselle DR, et al. Chemical modification of cysteine residues is a misleading indicator of their status as active site residues in the vitamin K-dependent gamma-glutamyl carboxylation reaction. J Biol Chem 2004; 279:54079.
  31. Kuliopulos A, Hubbard BR, Lam Z, et al. Dioxygen transfer during vitamin K dependent carboxylase catalysis. Biochemistry 1992; 31:7722.
  32. Wood GM, Suttie JW. Vitamin K-dependent carboxylase. Stoichiometry of vitamin K epoxide formation, gamma-carboxyglutamyl formation, and gamma-glutamyl-3H cleavage. J Biol Chem 1988; 263:3234.
  33. Knobloch JE, Suttie JW. Vitamin K-dependent carboxylase. Control of enzyme activity by the "propeptide" region of factor X. J Biol Chem 1987; 262:15334.
  34. Li S, Furie BC, Furie B, Walsh CT. The propeptide of the vitamin K-dependent carboxylase substrate accelerates formation of the gamma-glutamyl carbanion intermediate. Biochemistry 1997; 36:6384.
  35. Suttie JW. Vitamin K-dependent carboxylase. Annu Rev Biochem 1985; 54:459.
  36. Liu S, Li S, Shen G, et al. Structural basis of antagonizing the vitamin K catalytic cycle for anticoagulation. Science 2021; 371.
  37. Rost S, Fregin A, Ivaskevicius V, et al. Mutations in VKORC1 cause warfarin resistance and multiple coagulation factor deficiency type 2. Nature 2004; 427:537.
  38. Li T, Chang CY, Jin DY, et al. Identification of the gene for vitamin K epoxide reductase. Nature 2004; 427:541.
  39. Wallin R, Martin LF. Warfarin poisoning and vitamin K antagonism in rat and human liver. Design of a system in vitro that mimics the situation in vivo. Biochem J 1987; 241:389.
  40. Spohn G, Kleinridders A, Wunderlich FT, et al. VKORC1 deficiency in mice causes early postnatal lethality due to severe bleeding. Thromb Haemost 2009; 101:1044.
  41. Goodstadt L, Ponting CP. Vitamin K epoxide reductase: homology, active site and catalytic mechanism. Trends Biochem Sci 2004; 29:289.
  42. Singh AK, Bhattacharyya-Pakrasi M, Pakrasi HB. Identification of an atypical membrane protein involved in the formation of protein disulfide bonds in oxygenic photosynthetic organisms. J Biol Chem 2008; 283:15762.
  43. Dutton RJ, Boyd D, Berkmen M, Beckwith J. Bacterial species exhibit diversity in their mechanisms and capacity for protein disulfide bond formation. Proc Natl Acad Sci U S A 2008; 105:11933.
  44. Chatron N, Abi Khalil R, Benoit E, Lattard V. Structural Investigation of the Vitamin K Epoxide Reductase (VKORC1) Binding Site with Vitamin K. Biochemistry 2020; 59:1351.
  45. Schulman S, Wang B, Li W, Rapoport TA. Vitamin K epoxide reductase prefers ER membrane-anchored thioredoxin-like redox partners. Proc Natl Acad Sci U S A 2010; 107:15027.
  46. Li W, Schulman S, Dutton RJ, et al. Structure of a bacterial homologue of vitamin K epoxide reductase. Nature 2010; 463:507.
  47. Shen G, Cui W, Cao Q, et al. The catalytic mechanism of vitamin K epoxide reduction in a cellular environment. J Biol Chem 2021; 296:100145.
  48. Schulman S, Furie B. How I treat poisoning with vitamin K antagonists. Blood 2015; 125:438.
  49. Park BK, Choonara IA, Haynes BP, et al. Abnormal vitamin K metabolism in the presence of normal clotting factor activity in factory workers exposed to 4-hydroxycoumarins. Br J Clin Pharmacol 1986; 21:289.
  50. Choonara IA, Malia RG, Haynes BP, et al. The relationship between inhibition of vitamin K1 2,3-epoxide reductase and reduction of clotting factor activity with warfarin. Br J Clin Pharmacol 1988; 25:1.
  51. Cushman M, Booth SL, Possidente CJ, et al. The association of vitamin K status with warfarin sensitivity at the onset of treatment. Br J Haematol 2001; 112:572.
  52. Kamali F, Edwards C, Butler TJ, Wynne HA. The influence of (R)- and (S)-warfarin, vitamin K and vitamin K epoxide upon warfarin anticoagulation. Thromb Haemost 2000; 84:39.
  53. Brenner B. Hereditary deficiency of vitamin K-dependent coagulation factors. Thromb Haemost 2000; 84:935.
  54. Pauli RM, Lian JB, Mosher DF, Suttie JW. Association of congenital deficiency of multiple vitamin K-dependent coagulation factors and the phenotype of the warfarin embryopathy: clues to the mechanism of teratogenicity of coumarin derivatives. Am J Hum Genet 1987; 41:566.
  55. Oldenburg J, von Brederlow B, Fregin A, et al. Congenital deficiency of vitamin K dependent coagulation factors in two families presents as a genetic defect of the vitamin K-epoxide-reductase-complex. Thromb Haemost 2000; 84:937.
  56. Marchetti G, Caruso P, Lunghi B, et al. Vitamin K-induced modification of coagulation phenotype in VKORC1 homozygous deficiency. J Thromb Haemost 2008; 6:797.
  57. Bodin L, Horellou MH, Flaujac C, et al. A vitamin K epoxide reductase complex subunit-1 (VKORC1) mutation in a patient with vitamin K antagonist resistance. J Thromb Haemost 2005; 3:1533.
  58. Loebstein R, Dvoskin I, Halkin H, et al. A coding VKORC1 Asp36Tyr polymorphism predisposes to warfarin resistance. Blood 2007; 109:2477.
  59. Wilms EB, Touw DJ, Conemans JM, et al. A new VKORC1 allelic variant (p.Trp59Arg) in a patient with partial resistance to acenocoumarol and phenprocoumon. J Thromb Haemost 2008; 6:1224.
  60. Peoc'h K, Pruvot S, Gourmel C, et al. A new VKORC1 mutation leading to an isolated resistance to fluindione. Br J Haematol 2009; 145:841.
  61. Furie B. Do pharmacogenetics have a role in the dosing of vitamin K antagonists? N Engl J Med 2013; 369:2345.
  62. Wang Y, Zhang W, Zhang Y, et al. VKORC1 haplotypes are associated with arterial vascular diseases (stroke, coronary heart disease, and aortic dissection). Circulation 2006; 113:1615.
  63. Michaux A, Matagrin B, Debaux JV, et al. Missense mutation of VKORC1 leads to medial arterial calcification in rats. Sci Rep 2018; 8:13733.
  64. Luo G, Ducy P, McKee MD, et al. Spontaneous calcification of arteries and cartilage in mice lacking matrix GLA protein. Nature 1997; 386:78.
  65. Price PA, Fraser JD, Metz-Virca G. Molecular cloning of matrix Gla protein: implications for substrate recognition by the vitamin K-dependent gamma-carboxylase. Proc Natl Acad Sci U S A 1987; 84:8335.
  66. Hubbard BR, Ulrich MM, Jacobs M, et al. Vitamin K-dependent carboxylase: affinity purification from bovine liver by using a synthetic propeptide containing the gamma-carboxylation recognition site. Proc Natl Acad Sci U S A 1989; 86:6893.
  67. Tward JD, Furie BC, Bouchard BA, et al. Other amino acids than phenylalanine at position 16 in the carboxylation recognition site in the prothrombin propeptide can support gamma-carboxylation. Blood 1997; 90(Suppl 1):291a.
  68. Dissing C, Persson E. Factor VII Tokushima (Cys22→Gly) is not γ-carboxylated due to a disrupted γ-carboxylase recognition site. Thromb Res 2017; 158:108.
  69. Huber P, Schmitz T, Griffin J, et al. Identification of amino acids in the gamma-carboxylation recognition site on the propeptide of prothrombin. J Biol Chem 1990; 265:12467.
  70. Chu K, Wu SM, Stanley T, et al. A mutation in the propeptide of Factor IX leads to warfarin sensitivity by a novel mechanism. J Clin Invest 1996; 98:1619.
  71. Oldenburg J, Quenzel EM, Harbrecht U, et al. Missense mutations at ALA-10 in the factor IX propeptide: an insignificant variant in normal life but a decisive cause of bleeding during oral anticoagulant therapy. Br J Haematol 1997; 98:240.
  72. Oldenburg J, Kriz K, Wuillemin WA, et al. Genetic predisposition to bleeding during oral anticoagulant therapy: evidence for common founder mutations (FIXVal-10 and FIXThr-10) and an independent CpG hotspot mutation (FIXThr-10). Thromb Haemost 2001; 85:454.
  73. Kristensen SR. Warfarin treatment of a patient with coagulation factor IX propeptide mutation causing warfarin hypersensitivity. Blood 2002; 100:2676.
  74. Presnell SR, Stafford DW. The vitamin K-dependent carboxylase. Thromb Haemost 2002; 87:937.
  75. Rehemtulla A, Roth DA, Wasley LC, et al. In vitro and in vivo functional characterization of bovine vitamin K-dependent gamma-carboxylase expressed in Chinese hamster ovary cells. Proc Natl Acad Sci U S A 1993; 90:4611.
  76. Kuo WL, Stafford DW, Cruces J, et al. Chromosomal localization of the gamma-glutamyl carboxylase gene at 2p12. Genomics 1995; 25:746.
  77. Lin PJ, Jin DY, Tie JK, et al. The putative vitamin K-dependent gamma-glutamyl carboxylase internal propeptide appears to be the propeptide binding site. J Biol Chem 2002; 277:28584.
  78. Mutucumarana VP, Acher F, Straight DL, et al. A conserved region of human vitamin K-dependent carboxylase between residues 393 and 404 is important for its interaction with the glutamate substrate. J Biol Chem 2003; 278:46488.
  79. Brenner B, Sánchez-Vega B, Wu SM, et al. A missense mutation in gamma-glutamyl carboxylase gene causes combined deficiency of all vitamin K-dependent blood coagulation factors. Blood 1998; 92:4554.
  80. Spronk HM, Farah RA, Buchanan GR, et al. Novel mutation in the gamma-glutamyl carboxylase gene resulting in congenital combined deficiency of all vitamin K-dependent blood coagulation factors. Blood 2000; 96:3650.
  81. Rost S, Fregin A, Koch D, et al. Compound heterozygous mutations in the gamma-glutamyl carboxylase gene cause combined deficiency of all vitamin K-dependent blood coagulation factors. Br J Haematol 2004; 126:546.
  82. Soute BA, Ulrich MM, Watson AD, et al. Congenital deficiency of all vitamin K-dependent blood coagulation factors due to a defective vitamin K-dependent carboxylase in Devon Rex cats. Thromb Haemost 1992; 68:521.
  83. Zhu A, Sun H, Raymond RM Jr, et al. Fatal hemorrhage in mice lacking gamma-glutamyl carboxylase. Blood 2007; 109:5270.
  84. Mousallem M, Spronk HM, Sacy R, et al. Congenital combined deficiencies of all vitamin K-dependent coagulation factors. Thromb Haemost 2001; 86:1334.
  85. Soute BA, Jin DY, Spronk HM, et al. Characteristics of recombinant W501S mutated human gamma-glutamyl carboxylase. J Thromb Haemost 2004; 2:597.
  86. Bristol JA, Ratcliffe JV, Roth DA, et al. Biosynthesis of prothrombin: intracellular localization of the vitamin K-dependent carboxylase and the sites of gamma-carboxylation. Blood 1996; 88:2585.
  87. Wasley LC, Rehemtulla A, Bristol JA, Kaufman RJ. PACE/furin can process the vitamin K-dependent pro-factor IX precursor within the secretory pathway. J Biol Chem 1993; 268:8458.
  88. Blanchard RA, Furie BC, Kruger SF, et al. Immunoassays of human prothrombin species which correlate with functional coagulant activities. J Lab Clin Med 1983; 101:242.
  89. Ratcliffe JV, Furie B, Furie BC. The importance of specific gamma-carboxyglutamic acid residues in prothrombin. Evaluation by site-specific mutagenesis. J Biol Chem 1993; 268:24339.
  90. Zhang L, Jhingan A, Castellino FJ. Role of individual gamma-carboxyglutamic acid residues of activated human protein C in defining its in vitro anticoagulant activity. Blood 1992; 80:942.
  91. Zhang L, Castellino FJ. The contributions of individual gamma-carboxyglutamic acid residues in the calcium-dependent binding of recombinant human protein C to acidic phospholipid vesicles. J Biol Chem 1993; 268:12040.
  92. Jhingan A, Zhang L, Christiansen WT, Castellino FJ. The activities of recombinant gamma-carboxyglutamic-acid-deficient mutants of activated human protein C toward human coagulation factor Va and factor VIII in purified systems and in plasma. Biochemistry 1994; 33:1869.
  93. Freedman SJ, Furie BC, Furie B, Baleja JD. Structure of the calcium ion-bound gamma-carboxyglutamic acid-rich domain of factor IX. Biochemistry 1995; 34:12126.
  94. Huang M, Furie BC, Furie B. Crystal structure of the calcium-stabilized human factor IX Gla domain bound to a conformation-specific anti-factor IX antibody. J Biol Chem 2004; 279:14338.
  95. Shikamoto Y, Morita T, Fujimoto Z, Mizuno H. Crystal structure of Mg2+- and Ca2+-bound Gla domain of factor IX complexed with binding protein. J Biol Chem 2003; 278:24090.
  96. Banner DW, D'Arcy A, Chène C, et al. The crystal structure of the complex of blood coagulation factor VIIa with soluble tissue factor. Nature 1996; 380:41.
  97. Christiansen WT, Jalbert LR, Robertson RM, et al. Hydrophobic amino acid residues of human anticoagulation protein C that contribute to its functional binding to phospholipid Vesicles. Biochemistry 1995; 34:10376.
  98. Sunnerhagen M, Forsén S, Hoffrén AM, et al. Structure of the Ca(2+)-free Gla domain sheds light on membrane binding of blood coagulation proteins. Nat Struct Biol 1995; 2:504.
  99. Freedman SJ, Blostein MD, Baleja JD, et al. Identification of the phospholipid binding site in the vitamin K-dependent blood coagulation protein factor IX. J Biol Chem 1996; 271:16227.
  100. McDonald JF, Shah AM, Schwalbe RA, et al. Comparison of naturally occurring vitamin K-dependent proteins: correlation of amino acid sequences and membrane binding properties suggests a membrane contact site. Biochemistry 1997; 36:5120.
  101. Huang M, Rigby AC, Morelli X, et al. Structural basis of membrane binding by Gla domains of vitamin K-dependent proteins. Nat Struct Biol 2003; 10:751.
  102. Rodríguez Y, Mezei M, Osman R. The PT1-Ca2+ Gla domain binds to a membrane through two dipalmitoylphosphatidylserines. A computational study. Biochemistry 2008; 47:13267.
  103. Ducy P, Desbois C, Boyce B, et al. Increased bone formation in osteocalcin-deficient mice. Nature 1996; 382:448.
  104. Yao Y, Shahbazian A, Boström KI. Proline and gamma-carboxylated glutamate residues in matrix Gla protein are critical for binding of bone morphogenetic protein-4. Circ Res 2008; 102:1065.
  105. Schurgers LJ, Spronk HM, Skepper JN, et al. Post-translational modifications regulate matrix Gla protein function: importance for inhibition of vascular smooth muscle cell calcification. J Thromb Haemost 2007; 5:2503.
  106. Manfioletti G, Brancolini C, Avanzi G, Schneider C. The protein encoded by a growth arrest-specific gene (gas6) is a new member of the vitamin K-dependent proteins related to protein S, a negative coregulator in the blood coagulation cascade. Mol Cell Biol 1993; 13:4976.
  107. Nakano T, Kawamoto K, Kishino J, et al. Requirement of gamma-carboxyglutamic acid residues for the biological activity of Gas6: contribution of endogenous Gas6 to the proliferation of vascular smooth muscle cells. Biochem J 1997; 323 ( Pt 2):387.
  108. Kulman JD, Harris JE, Haldeman BA, Davie EW. Primary structure and tissue distribution of two novel proline-rich gamma-carboxyglutamic acid proteins. Proc Natl Acad Sci U S A 1997; 94:9058.
  109. Kulman JD, Harris JE, Xie L, Davie EW. Identification of two novel transmembrane gamma-carboxyglutamic acid proteins expressed broadly in fetal and adult tissues. Proc Natl Acad Sci U S A 2001; 98:1370.
  110. Xu S, Han JC, Morales A, et al. Characterization of 11p14-p12 deletion in WAGR syndrome by array CGH for identifying genes contributing to mental retardation and autism. Cytogenet Genome Res 2008; 122:181.
  111. Price PA, Williamson MK. Substrate recognition by the vitamin K-dependent gamma-glutamyl carboxylase: identification of a sequence homology between the carboxylase and the carboxylase recognition site in the substrate. Protein Sci 1993; 2:1987.
  112. Berkner KL, Pudota BN. Vitamin K-dependent carboxylation of the carboxylase. Proc Natl Acad Sci U S A 1998; 95:466.
  113. Yin ZF, Huang ZF, Cui J, et al. Prothrombotic phenotype of protein Z deficiency. Proc Natl Acad Sci U S A 2000; 97:6734.
  114. Han X, Fiehler R, Broze GJ Jr. Isolation of a protein Z-dependent plasma protease inhibitor. Proc Natl Acad Sci U S A 1998; 95:9250.
  115. Han X, Fiehler R, Broze GJ Jr. Characterization of the protein Z-dependent protease inhibitor. Blood 2000; 96:3049.
  116. Van de Water N, Tan T, Ashton F, et al. Mutations within the protein Z-dependent protease inhibitor gene are associated with venous thromboembolic disease: a new form of thrombophilia. Br J Haematol 2004; 127:190.
  117. Rezaie AR, Bae JS, Manithody C, et al. Protein Z-dependent protease inhibitor binds to the C-terminal domain of protein Z. J Biol Chem 2008; 283:19922.
  118. Broze GJ Jr. Protein-Z and thrombosis. Lancet 2001; 357:900.
  119. Broze GJ Jr. Protein Z-dependent regulation of coagulation. Thromb Haemost 2001; 86:8.
  120. Gris JC, Amadio C, Mercier E, et al. Anti-protein Z antibodies in women with pathologic pregnancies. Blood 2003; 101:4850.
  121. Corral J, González-Conejero R, Hernández-Espinosa D, Vicente V. Protein Z/Z-dependent protease inhibitor (PZ/ZPI) anticoagulant system and thrombosis. Br J Haematol 2007; 137:99.
  122. Zhang J, Tu Y, Lu L, et al. Protein Z-dependent protease inhibitor deficiency produces a more severe murine phenotype than protein Z deficiency. Blood 2008; 111:4973.
  123. Sofi F, Cesari F, Abbate R, et al. A meta-analysis of potential risks of low levels of protein Z for diseases related to vascular thrombosis. Thromb Haemost 2010; 103:749.
  124. Kemkes-Matthes B, Nees M, Kühnel G, et al. Protein Z influences the prothrombotic phenotype in Factor V Leiden patients. Thromb Res 2002; 106:183.
  125. Kemkes-Matthes B, Matthes KJ, Souri M, et al. R255h amino acid substitution of protein Z identified in patients with factor V Leiden mutation. Br J Haematol 2005; 128:248.
  126. Gris JC, Quéré I, Dechaud H, et al. High frequency of protein Z deficiency in patients with unexplained early fetal loss. Blood 2002; 99:2606.
  127. Corral J, González-Conejero R, Soria JM, et al. A nonsense polymorphism in the protein Z-dependent protease inhibitor increases the risk for venous thrombosis. Blood 2006; 108:177.
  128. Vasse M, Guegan-Massardier E, Borg JY, et al. Frequency of protein Z deficiency in patients with ischaemic stroke. Lancet 2001; 357:933.
  129. Kobelt K, Biasiutti FD, Mattle HP, et al. Protein Z in ischaemic stroke. Br J Haematol 2001; 114:169.
  130. Mushunje A, Zhou A, Huntington JA, et al. Antithrombin 'DREUX' (Lys 114Glu): a variant with complete loss of heparin affinity. Thromb Haemost 2002; 88:436.
  131. Ayoub N, Esposito G, Barete S, et al. Protein Z deficiency in antiphospholipid-negative Sneddon's syndrome. Stroke 2004; 35:1329.
  132. Sofi F, Cesari F, Pratesi G, et al. Low protein Z levels in patients with peripheral arterial disease. Thromb Haemost 2007; 98:1114.
  133. Nowak-Göttl U, Fröhlich B, Thedieck S, et al. Association of the protein Z ATG haplotype with symptomatic nonvascular stroke or thromboembolism in white children: a family-based cohort study. Blood 2009; 113:2336.
  134. Sofi F, Cesari F, Tu Y, et al. Protein Z-dependent protease inhibitor and protein Z in peripheral arterial disease patients. J Thromb Haemost 2009; 7:731.
  135. Steffano B, Forastiero R, Martinuzzo M, Kordich L. Low plasma protein Z levels in patients with antiphospholipid antibodies. Blood Coagul Fibrinolysis 2001; 12:411.
  136. McColl MD, Deans A, Maclean P, et al. Plasma protein Z deficiency is common in women with antiphospholipid antibodies. Br J Haematol 2003; 120:913.
  137. Viegas CS, Simes DC, Laizé V, et al. Gla-rich protein (GRP), a new vitamin K-dependent protein identified from sturgeon cartilage and highly conserved in vertebrates. J Biol Chem 2008; 283:36655.
  138. McIntosh JM, Olivera BM, Cruz LJ, Gray WR. Gamma-carboxyglutamate in a neuroactive toxin. J Biol Chem 1984; 259:14343.
  139. Rigby AC, Lucas-Meunier E, Kalume DE, et al. A conotoxin from Conus textile with unusual posttranslational modifications reduces presynaptic Ca2+ influx. Proc Natl Acad Sci U S A 1999; 96:5758.
  140. Yuan Y, Balsara RD, Zajicek J, et al. Discerning the Role of the Hydroxyproline Residue in the Structure of Conantokin Rl-B and Its Role in GluN2B Subunit-Selective Antagonistic Activity toward N-Methyl-d-Aspartate Receptors. Biochemistry 2016; 55:7112.
  141. Itang CEMM, Gaza JT, Masacupan DJM, et al. Identification of Conomarphin Variants in the Conus eburneus Venom and the Effect of Sequence and PTM Variations on Conomarphin Conformations. Mar Drugs 2020; 18.
Topic 1318 Version 20.0

References