HOMOCYSTEINE is a non–protein-forming, sulfur amino acid whose metabolism is at the intersection of two metabolic pathways1: remethylation and transsulfuration (Fig 1). In remethylation, homocysteine acquires a methyl group from N-5-methyltetrahydrofolate (MTHF ) or from betaine to form methionine. The reaction with MTHF occurs in all tissues and is vitamin B12–dependent, whereas the reaction with betaine is confined mainly to the liver and is vitamin B12–independent. A considerable proportion of methionine is then activated by adenosine triphosphate (ATP) to form S-adenosylmethionine (SAM). SAM serves primarily as a universal methyl donor to a variety of acceptors including guanidinoacetate, nucleic acids, neurotransmitters, phospholipids, and hormones. S-adenosylhomocysteine (SAH), the byproduct of these methylation reactions, is subsequently hydrolyzed, thus regenerating homocysteine, which then becomes available to start a new cycle of methyl-group transfer. In the transsulfuration pathway, homocysteine condenses with serine to form cystathionine in an irreversible reaction catalyzed by the pyridoxal-5′-phosphate (PLP)–containing enzyme, cystathionine β-synthase (CBS). Cystathionine is hydrolyzed by a second PLP-containing enzyme, gamma-cystathionase, to form cysteine and α-ketobutyrate. Excess cysteine is oxidized to taurine and inorganic sulfates or excreted in the urine. Thus, in addition to the synthesis of cysteine, this transsulfuration pathway effectively catabolizes excess homocysteine which is not required for methyltransfer and delivers sulfate for the synthesis of heparin, heparan sulfate, dermatan sulfate, and chondroitin sulfate. It is important to note that because homocysteine is not a normal dietary constituent, the sole source of homocysteine is methionine.

Fig. 1.

Homocysteine metabolic pathways. In the methylation pathway, homocysteine acquires a methyl group either from betaine, a reaction that occurs mainly in the liver or from 5-methyltetrahydrofolate, a reaction which occurs in all tissues and is vitamin B12–dependent. In the transsulfuration pathway homocysteine condenses with serine to form cystathionine in a reaction that is catalyzed by CBS and requires PLP.

Fig. 1.

Homocysteine metabolic pathways. In the methylation pathway, homocysteine acquires a methyl group either from betaine, a reaction that occurs mainly in the liver or from 5-methyltetrahydrofolate, a reaction which occurs in all tissues and is vitamin B12–dependent. In the transsulfuration pathway homocysteine condenses with serine to form cystathionine in a reaction that is catalyzed by CBS and requires PLP.

Close modal

Because of the existence of a cellular homocysteine export mechanism,2,3 plasma normally contains a small amount of homocysteine averaging 10 μmol/L. This export mechanism complements the catabolism of homocysteine through transsulfuration; together these mechanisms help maintain low intracellular concentrations of this potentially cytotoxic sulfur amino acid. In hyperhomocysteinemia, plasma homocysteine levels are elevated and, barring impaired renal function, the occurrence of hyperhomocysteinemia indicates that homocysteine metabolism has in some way been disrupted and that the export mechanism is disposing into the blood excess homocysteine. This export mechanism limits intracellular toxicity, but leaves vascular tissue exposed to the possibly deleterious effects of excess homocysteine.

The more severe cases of hyperhomocysteinemia are caused by homozygous defects in genes encoding for enzymes of homocysteine metabolism. In such cases, a defect in an enzyme involved in either homocysteine remethylation or transsulfuration leads to large elevations of homocysteine in the blood and urine. The classic form of such a disorder — congenital homocystinuria — is caused by homozygous defects in the gene encoding for CBS. In these individuals, fasting plasma homocysteine concentrations can be as high as 400 μmol/L.4 Two different forms of the disease can be distinguished on the basis of the responsiveness to treatment with large dosages of vitamin B6.5,6 Several CBS mutations are known7-10: the most frequent are 833TC and 919GA, located in exon 8; and 1224-2AC, which causes the entirety of exon 12 being deleted. The 833TC is spread in several ethnic groups; the 919GA mutation has been almost exclusively reported in patients of Celtic origin. In 20 homocystinuric patients from 16 unrelated Italian families, characterization of 24 of 30 independent alleles disclosed 13 mutations, including 11 novel mutations (146CT, 172CT, 262CT, 346GA, 374GA, 376AG, 452del27, 770CT, 844ins68, 869CT, 904GA). Two previously reported mutations (833TC and 341CT) were found in 26.6% and 16.6% of the alleles. Hence, most of the mutations are ‘private’ and clustered on exons 8, 3, and 1.9,10 

Homozygous defects of other genes that lead to similarly severe elevations in plasma homocysteine include those encoding for methylenetetrahydrfolate reductase (MTHFR) or for any of the enzymes which participate in the synthesis of methylated vitamin B12.11-17 Genetic impairments for the vitamin B12–dependent methyltetrahydrofolate:homocysteine methyltransferase have not been reported.4 MTHFR deficiency was first described by Mudd et al14 in two unrelated teenagers. In contrast to patients with CBS deficiency, these patients had slightly reduced methionine levels in their plasma and normal CBS activity in skin fibroblast extracts. Although the hyperhomocysteinemia associated with this defect is less severe than in homozygous CBS deficiency, the prognosis for these patients is generally worse than in CBS deficiency,18 related in part to the unresponsiveness to any form of treatment. Most patients with MTHFR deficiency have hypomethioninemia, but in contrast to patients who have hypomethioninemia due to inborn errors of vitamin B12 metabolism and who develop severe megaloblastic anemia,19 patients with MTHFR deficiency do not become anemic. However, expression of MTHFR deficiency is variable14 (see below).

Vitamin B12, an essential nutrient, is converted to methylcobalamin, which functions as a cofactor for methionine synthase, and to 5-deoxyadenosylcobalamin, the coenzyme for mitochondrial methylmalonyl-CoA mutase. Several distinct genetic defects of cobalamin transport and metabolism resulting in decreased methionine synthase activity and hyperhomocysteinemia have been described (transcobalamin II deficiency, Cbl C, D, E, and G mutations). These defects may selectively hamper methylation of vitamin B12. In addition, such defects may impair synthesis of 5-deoxyadenosylcobalamin, with resulting methylmalonic aciduria.19 

The more common causes of hyperhomocysteinemia are also moderate in character and may be caused by less severe defects in genes encoding for enzymes or from inadequate status of those vitamins that are involved in homocysteine metabolism. Plasma homocysteine concentrations in these instances may differ depending on which arm of the two metabolic pathways of homocysteine metabolism is defective.1 An impairment in the remethylation pathway, even if it is mild, will lead to a substantial increase in plasma homocysteine concentrations under fasting conditions. Such an impairment may be due to inadequate status of either folate or vitamin B12 or to defects in the gene encoding for MTHFR.1,17,20-33 MTHFR contains flavin adenine dinucleotide (FAD) as a prosthetic group, which raises the possibility that vitamin B2 status is also a determinant of fasting plasma homocysteine levels. In contrast, a mild impairment in the transsulfuration pathway will lead, at most, to a very slight increase in fasting plasma homocysteine levels. This mild impairment, which may be caused by heterozygous defects in the CBS gene or inadequate levels of vitamin B6,1,20,34-38 is normally identified by an abnormal increase in plasma homocysteine after a methionine loading test or following a meal.20,38-41 The different phenotype anticipated in remethylation and transsulfuration defects is supported by studies conducted in vitamin-deficient animal models. Thus, fasting plasma homocysteine concentration is 10-fold higher in folate deficient rats than in folate supplemented rats.42 This high concentration of homocysteine in plasma was in part due to a lack of MTHF for homocysteine transmethylation and to insufficient S-adenosylmethionine for the activation of the transsulfuration pathway.42 In both humans and rats, mild vitamin B6 deficiency was associated with normal fasting plasma homocysteine levels. Fasting hyperhomocysteinemia in vitamin B6 deficiency may occur only if the deficiency is severe and sustained over a long period of time.36 After methionine load, the homocysteine concentration increased 35-fold in rats that were vitamin B6–deficient compared with about fourfold in control rats and less than 35% in folate-deficient rats.42 Evidence of two distinct forms of hyperhomocysteinemia in humans is largely derived from preliminary data obtained from 274 consecutive participants in the Family Heart Study.43 Plasma total homocysteine were measured under fasting conditions and 4 hours after a methionine load for each participant. Using homocysteine values greater than the 90th percentile for the definition of hyperhomocysteinemia (both fasting and methionine load), it was shown that of 43 hyperhomocysteinemic individuals, 20 (46%) had fasting hyperhomocysteinemia only, 17 (34.5%) had postmethionine load hyperhomoccysteonemia only and just 7 (14%) had both types of hyperhomocysteinemia.

Existence of an interrelationship between vitamin status and plasma homocysteine was first reported by Kang et al,17 who showed an inverse relationship between homocysteine and plasma folate concentrations. Other studies have shown the existence of an inverse correlation between homocysteine and folate or vitamin B12 plasma concentrations and the efficacy of vitamin supplementation in the lowering of plasma homocysteine levels.22-24,26-31,33-35,44-51 The independent associations between individual nutrients and homocysteine concentrations were studied in an established cohort of Americans52 (The Framingham Heart Study). After controlling for age, sex, and other vitamins, nonfasting plasma homocysteine exhibited a strong, nonlinear inverse association with plasma folate. Minimum levels of homocysteine were observed around 10 nmol/L of folate and above. Nonfasting plasma homocysteine exhibited weaker inverse associations with plasma concentrations of vitamin B12 and pyridoxal-5′-phosphate (PLP, the active form of vitamin B6).

Potential interactions between vitamin status and genetic abnormalities of the enzymes involved in methionine metabolism in the pathogenesis of hyperhomocysteinemia is illustrated by the recent data on the thermolabile MTHFR, a variant of the enzyme which is caused by a mutation in the structural gene at a polymorphic site (see below).

Mutations that result in severely reduced MTHFR activity and hyperhomocysteinemia are rare.53-55 However, in 1988 Kang et al17 reported that two unrelated patients with moderate hyperhomocysteinemia and low folate levels had a variant of MTHFR that was distinguished from the normal enzyme (as measured in lymphocyte extracts) by its lower specific activity (50%) and its thermolability. In subsequent studies, Kang et al56,57 showed that MTHFR thermolability is an inherited recessive trait, which is present in approximately 5% of the general population and 17% of patients with proven coronary artery disease, but is not associated with neurological complications. Impaired activity of MTHFR, due to the thermolabile form of the enzyme, has been observed in as many as 28% of hyperhomocysteinemic patients with premature vascular disease.58 The cDNA for human MTHFR has been recently isolated54 and it has been shown that MTHFR thermolability is caused by a point mutation (677C to T transition) at a polymorphic site, resulting in a valine substitution for an alanine in this enzyme.59 The mutation was found in 38% of unselected chromosomes from 57 French Canadian individuals; the homozygous state for the mutation was present in 12% of these subjects and correlated with significantly raised tHcy.59 Preliminary evidence indicates that the frequency of homozygotes for the 677CT mutation may vary significantly in populations from different geographic areas (from 1.4% to 15%).60 In the Dutch population, homozygosity for this mutation is 15% and 5% in 60 vascular patients and 111 controls, respectively.61 In an Italian patient population with arterial or venous occlusive disease, the prevalence of homozygotes was 19 of 64 individuals (29.7%).62 In a control group (n = 258), on the other hand, homozygosity for the 677CT mutation was only 15.1%. Interestingly, none of 7 patients with isolated methionine intolerance was homozygous for the mutation.63 

The impact of MTHFR thermolabile variant on plasma homocysteine levels is as of yet unclear. The hyperhomocysteinemia seen in the original patients of Kang et al17 was associated with low folate plasma levels, and folate supplementation reduced homocysteine to normal levels. In the above-mentioned Italian study,62 18 of the 19 patients who were homozygotes for the 677CT mutation were hyperhomocysteinemic with levels of homocysteine in plasma that were higher than the 95th percentile in the respective age-matched controls. In a recent study the occurrence of an interaction between MTHFR thermolability genotype and folate status was shown.64 When plasma folate concentrations were above the median (15.4 nmol/L), plasma homocysteine levels were low and unrelated to the MTHFR genotype. However, when plasma folate concentrations were below the median, plasma homocysteine levels were significantly higher in homozygotes for the 677CT mutation than in those with the normal genotype.64 These data imply that the phenotypic expression of the MTHFR genotypes is dependent on the availability of folate, suggesting that homozygotes for the thermolabile genotype might have a higher folate requirement than individuals with a normal genotype. Because FAD is an essential prosthetic group for MTHFR activity, it stands to reason that vitamin B2 status is also a determinant of plasma homocysteine levels. The latter relationship has yet to be demonstrated in population-based studies.

The relationship between severe hyperhomocysteinemia and arterial disease was first suggested by McCully,65 who observed autopsy evidence of precocious arterial thrombosis and atherosclerosis in a homocystinuric patient with impaired cobalamin metabolism that was identical to what had earlier been described in homocystinuric patients with CBS deficiency.66 This deficiency in CBS is characterized by arteriosclerosis, thromboembolic complications, skeletal abnormalities, ectopia lentis, and mental retardation. In 1976, Wilcken and Wilcken67 showed that the concentration of homocysteine-cysteine mixed disulfide after a methionine load was slightly higher in coronary heart disease patients than in controls, thus providing the first evidence of an association between mild hyperhomocysteinemia and vascular disease. Mildly increased homocysteine levels were later reported in coronary artery disease patients,68-73 in cerebrovascular disease patients,69,74-78 and in peripheral artery disease patients.69,75,79-81 Association of mild hyperhomocysteinemia with occlusive disease76,77,80,82-86 was independent of the presence of risk factors like smoking, hyperlipidemia, hypertension, and diabetes. This suggested that hyperhomocysteinemia is an independent risk factor for the abovementioned clinical conditions. Homocysteine concentrations in patients with symptomatic vascular disease are on average 31% greater than in normal controls,87 and prospective assessment of vascular disease risk indicates that plasma homocysteine levels are associated with increased risk of myocardial infarction.88,89 

Although venous thromboembolism accounts for 50% of the vascular complications of homocystinuria,90 the link between less severe hyperhomocysteinemia and venous thromboembolic disease was overlooked until recently. Recurrent episodes of thromboembolism, events that occur at an early age, thrombosis after trivial provocation, and thrombosis at unusual sites are all features which should heighten the suspicion that an inherited metabolic abnormality is playing an etiologic role.91 Studies on the relationship between thrombosis and mild hyperhomocysteinemia are summarized in Table 1. Brattstrom et al92 found no significant difference in plasma homocysteine concentrations between 42 patients with venous thromboembolism and healthy control subjects, although the male patients showed a tendency of higher plasma homocysteine than male control subjects. Similarly, Amundsen et al93 studied 35 young adult patients (age <56 years) with verified deep venous thrombosis and found no difference in fasting and postmethionine load plasma homocysteine levels with age/sex-matched healthy controls. However, other investigators have reported a positive association between hyperhomocysteinemia and venous thrombotic outcomes. In one study94 elevated fasting plasma homocysteine levels were reported in 25% of patients who developed venous thrombosis before 60 years of age. Fasting and postmethionine load homocysteine levels were measured by Falcon et al95 in a series of 80 patients who had at least one verified episode of venous thromboembolism before the age of 40 years and were free from hemostatic abnormalities known to be associated with increased risk of venous thromboembolism. Fasting hyperhomocysteinemia was observed in 8.8% of patients, but postmethionine load hyperhomocysteinemia was present in 17.7% of the patients. About half of the patients with hyperhomocysteinemia had a positive family history of thrombosis and familial hyperhomocysteinemia was confirmed in over 50% of the families studied. In a cross-sectional 2-year evaluation of 157 consecutive unrelated patients with a history of venous or arterial occlusive disease occurring before the age of 45 years or at unusual sites, moderate hyperhomocysteinemia was detected in 13.1% and 19.2% of patients with venous or arterial occlusive disease.96 The prevalence of hyperhomocysteinemia was almost twice as high when based on homocysteine measurements done after oral methionine load as when based on fasting levels. Deficiencies of protein C, protein S, plasminogen, and activated protein C resistance were detected only in patients with venous occlusive disease, with an overall prevalence of 18.7%. Familial hyperhomocysteinemia was shown in 8 of the 12 families investigated. Event-free survival analysis showed that the relative risk in patients with moderate hyperhomocysteinemia and the other defects was 1.7 times greater than in patients without defects and that the risk conferred by hyperhomocysteinemia was similar to that of defects affecting the protein C system. A higher rate of recurrent thrombosis was also observed in patients with hyperhomocysteinemia and with the other defects than in patients without defects.97 Homocysteine levels above the 90th percentile of the control distribution were observed in another study in 25% of 185 patients with recurrent venous thrombosis, with a relative risk of recurrence two times greater in patients with hyperhomocysteinemia than in those without hyperhomocysteinemia.97 In this study, the relative risk of patients with postmethionine load homocysteine concentrations exceeding the 90th percentile (2.6) was similar to that of patients with fasting hyperhomocysteinemia. Twenty-seven of the 46 patients with fasting hyperhomocysteinemia also had postload hyperhomocysteinemia, whereas 17 patients had isolated methionine intolerance. Hence, the overall prevalence of hyperhomocysteinemia in this patients' population was 34.1%. However, because absolute postmethionine load values (instead of the postmethionine load above baseline levels) were considered in this study, the relative contribution of remethylation or transsulfuration defects in the risk conferred by hyperhomocysteinemia could not be evaluated.

Table 1.

Relationship Between Homocysteine and Thrombosis

AuthorsThrombosis TypeAge (yr)Homocysteine MeasuredHomocysteine (μmol/L)Hyperhomocysteinemia Cases
Patients (n)Controls (n)Patients (%)Controls (%)OR (95% CI)
Brattstrom et al92  Venous <50 Fasting 13.4 (29) 10.6 (30) 
   PML (Δ)* 23.2 20.4 
   PML (42) (42) 14.5 3.1 (0.4, 27.7) 
Amundsen et al93  Deep venous <56 Fasting 10.3 ± 3.5 (35) 10.3 ± 2.6 (39) 5.7 2.6 2.3 (0.1, 68.5) 
   PML(T)* 39.5 ± 14.2 37.2 ± 14.2 5.7 2.6 2.3 (0.1, 68.5) 
Bienvenu et al94  Venous + arterial <58 Fasting 13.8 ± 6.0 (50) 8.1 ± 2.2 (49) 36 13.2 (2.82, 83.0) 
Falcon et al95  Juvenile venous <40 Fasting 7.4 ± 5.5 (80) 7.9 ± 2.1 (51) 8.8 ∞ (1.16, ∞) 
   PML(Δ)* 10.3 ± 5.1 (79) 8.3 ± 3.4 (40) 18.8 2.5 9.1 (1.38, 194) 
Fermo et al96  Venous + arterial <45 Fasting 14.7 ± 9.1 (157) 11.6 ± 3 (60) 8.9 1.9 (0.5, 6.6) 
   PML(T)* 25.6 ± 11.6 (87) 19.3 ± 7.1 (60) 21.8 5  5.3 (1.7, 17.1) 
den Heijer et al97  Deep vein 44 Fasting 12.9 (269) 12.3 (269) 10 2.5 (1.2, 5.2) 
Cattaneo et al98  Deep vein N/A Fasting + PML(T) N/A (89) N/A (89) 13.5 6.7 2.2 (0.8, 6.0) 
    
Simioni et al99  Deep vein 62 Baseline 15 (60) 12 (148) 25 11.5 2.6 (1.1-5.9) 
den Heijer et al100  Recurrent thrombosis 67P Fasting N/A (185) N/A (220) 25 10 2.0 (1.5, 2.7) 
  51C       
   PML(T)   24 10 2.6 (1.9, 3.6) 
Petri et al101ρ Atherothrombotic lupus 38 Fasting N/A (337)  15 10 3.49 (0.97, 12.54) 
AuthorsThrombosis TypeAge (yr)Homocysteine MeasuredHomocysteine (μmol/L)Hyperhomocysteinemia Cases
Patients (n)Controls (n)Patients (%)Controls (%)OR (95% CI)
Brattstrom et al92  Venous <50 Fasting 13.4 (29) 10.6 (30) 
   PML (Δ)* 23.2 20.4 
   PML (42) (42) 14.5 3.1 (0.4, 27.7) 
Amundsen et al93  Deep venous <56 Fasting 10.3 ± 3.5 (35) 10.3 ± 2.6 (39) 5.7 2.6 2.3 (0.1, 68.5) 
   PML(T)* 39.5 ± 14.2 37.2 ± 14.2 5.7 2.6 2.3 (0.1, 68.5) 
Bienvenu et al94  Venous + arterial <58 Fasting 13.8 ± 6.0 (50) 8.1 ± 2.2 (49) 36 13.2 (2.82, 83.0) 
Falcon et al95  Juvenile venous <40 Fasting 7.4 ± 5.5 (80) 7.9 ± 2.1 (51) 8.8 ∞ (1.16, ∞) 
   PML(Δ)* 10.3 ± 5.1 (79) 8.3 ± 3.4 (40) 18.8 2.5 9.1 (1.38, 194) 
Fermo et al96  Venous + arterial <45 Fasting 14.7 ± 9.1 (157) 11.6 ± 3 (60) 8.9 1.9 (0.5, 6.6) 
   PML(T)* 25.6 ± 11.6 (87) 19.3 ± 7.1 (60) 21.8 5  5.3 (1.7, 17.1) 
den Heijer et al97  Deep vein 44 Fasting 12.9 (269) 12.3 (269) 10 2.5 (1.2, 5.2) 
Cattaneo et al98  Deep vein N/A Fasting + PML(T) N/A (89) N/A (89) 13.5 6.7 2.2 (0.8, 6.0) 
    
Simioni et al99  Deep vein 62 Baseline 15 (60) 12 (148) 25 11.5 2.6 (1.1-5.9) 
den Heijer et al100  Recurrent thrombosis 67P Fasting N/A (185) N/A (220) 25 10 2.0 (1.5, 2.7) 
  51C       
   PML(T)   24 10 2.6 (1.9, 3.6) 
Petri et al101ρ Atherothrombotic lupus 38 Fasting N/A (337)  15 10 3.49 (0.97, 12.54) 

Abbreviation: N/A, data not available.

*

T, total plasma homocysteine level after methionine load (fasting homocysteine level was not subtracted). Δ, Net increase in homocysteine after methionine load; ie, total fasting levels.

Males only.

Males and females.

ρ Prospective study.

These data represent strong evidence supporting the role of moderate hyperhomocysteinemia in the development of premature and/or recurrent venous thromboembolic disease. High plasma homocysteine levels are also a risk factor for deep-vein thrombosis in the general population. Fasting homocysteine concentrations were measured in 269 patients below 70 years of age with a first episode of deep-vein thrombosis and matched control subjects participating in the Leiden Thrombophilia Study.100 Hyperhomocysteinemia exceeding the 95th percentile of the control group was found in 10% of the patients, with a matched odds ratio of 2.5. The effect of hyperhomocysteinemia was independent of other well-established risk factors for thrombosis, including protein C, protein S, and antithrombin III deficiencies and activated protein C resistance. An unexpected finding of this study was the observation that the association between elevated homocysteine levels and venous thrombosis was stronger among women than among men. Because nutritional parameters were not evaluated in this study, it cannot be ruled out that the stronger association observed in women may be caused by a different vitamin status. In addition, postmethionine load homocysteine measurements were not carried out, resulting in a potential underestimation of the risk conferred by hyperhomocysteinemia.

Most recently, Petri et al101 reported a prospective study on the association between homocysteine and risk of stroke and thrombotic events in 337 systemic lupus erythematosus (SLE) patients with follow-up of 1,619 person-years (mean 4.8 [SD 1.7] years). A fasting blood sample was obtained at the beginning of this study from each patient who also had four follow-up assessments per year to establish risk factors for thrombosis and coronary artery disease. During the follow-up there were 29 cases of stroke and 31 arterial thrombotic events. Hyperhomocysteinemia, defined as total plasma homocysteine >14.1 μmol/L), was found in 15% of these patients and was significantly associated with arterial thrombotic events (odds ratio 3.74 [95% confidence interval (CI) 1.96 to 7.13], P = .0001) and with stroke (odds ratio 2.24 [95% CI 1.22 to 4.13], P = .01). After adjustment for established risk factors, hyperhomocysteinemia remained an independent risk factor for thromboses (P = .05) and stroke (P = .04).

As in the case for deficiencies of the protein C anticoagulant system, not all patients with hyperhomocysteinemia develop thrombosis. The possibility that factors synergistic to hyperhomocysteinemia may be required for the development of thrombotic manifestations was explored in 45 members of seven unrelated consanguineous kindreds in which at least one member was homozygous for homocystinuria.101 Thrombosis occurred in 6 of 11 patients with homocystinuria before the age of 8 years; all 6 patients also had activated protein C resistance. Conversely, of four patients with homocystinuria who did not have activated protein C resistance, none had thrombosis occurring before the age of 17 years. The investigators concluded that the combination of homocystinuria and activated protein C resistance conveys a substantial risk for thrombosis.102 Such a conclusion may cast doubts about an independent pathogenic role of hyperhomocysteinemia in venous thromboembolism. Both activated protein C resistance and hyperhomocysteinemia are highly prevalent in patients with early onset vascular occlusive disease.96 If the association of activated protein C resistance and moderate hyperhomocysteinemia markedly increases the thrombotic risk, one would anticipate its prevalence to be significantly higher than expected based on the prevalence of the isolated defects. In a series of 307 patients with early onset venous or arterial disease or with thrombosis occurring at unusual sites, the prevalence of isolated activated protein C resistance and moderate hyperhomocysteinemia (fasting or postmethionine-load) were 10% and 27%, respectively. The combined defect was detected in 3.6% of the patients, a figure slightly, but not significantly, higher than the 2.7% prevalence expected, assuming no effect of the association on the risk of thrombosis.103 Although other possibilities cannot be ruled out, these data are consistent with the notion that moderate hyperhomocysteinemia, per se, is an independent risk factor for both venous and arterial thromboembolic disease.

Despite the large number of studies suggesting that mild elevations of homocysteine in plasma are associated with an increased risk for occlusive vascular disease, thrombosis, and stroke, the question of whether homocysteine, per se, is responsible for these associations remains uncertain. The survey of Mudd et al,104 which assessed the parents and grandparents of homocystinuric children, concluded that heterozygosity for CBS deficiency is not associated with increased risk in heart attacks and stroke. Swift and Morrell105 questioned the validity of some of the methods used by Mudd et al and argued that the data actually point to increased mortality rates in this heterozygote population. Two studies that used a noninvasive (doppler ultrasound) technique also provided conflicting results. Clarke106 found no evidence of increased frequency of endothelial plaques in the neck arteries of 25 Irish heterozygotes, compared with 21 control subjects. Rubba et al,84 on the other hand, indicated more frequent early vascular lesions in the iliac and internal carotid arteries in 14 heterozygotes than in 47 controls. At the molecular biology level, Kozich et al107 examined the CBS alleles in four patients with premature occlusive arterial disease who were (1) hyperhomocysteinemic based on postmethionine load results, and (2) had lower enzyme activity in their fibroblasts. None of the eight alleles contained any mutation which resulted in diminished enzyme activity. In a prior study this group of investigators showed that cultured fibroblasts are not always reliable for testing the phenotypic expression in homocystinuric patients.7 

Other studies61 found that prevalence of the more common CBS mutations is not higher in the patient populations. A knockout mouse with CBS deficiency reported by Watanabe et al108 was found to lack manifestations of thrombosis or/and cardiovascular complications and instead exhibited fatty livers. This is despite the fact that the levels of homocysteine in plasma were 40-fold higher in the homozygote mice than in the control mice. It might be that these mice did not live long enough for cardiovascular complications to develop.

Other inconsistencies relate to the question of whether MTHFR thermolability confers increased risk for the various diseases. The early study by Kang et al,57 who relied on enzyme activity, has shown a higher prevalence (17%) of this variant in a North American, coronary artery disease (CAD) population than in normal healthy controls (5%). Similarly, the abovementioned studies, which relied on molecular biology techniques, demonstrated higher prevalence of the 677CT homozygotes in vascular disease patients (15%) than in controls (5%) in a Dutch population; and to 29.7% in an Italian patient population, compared with only 15.1% in healthy controls.61,62 

However, other studies have been contradictory. No difference in the prevalence of the homozygosity for the 677CT mutation was found between coronary patients and controls in an Australian population,109 a second Dutch population,110 a French Canadian population,111 a British population,112 and in the Physician Health Study.113 

These apparent inconsistencies can be explained on the basis of the differences in the study population, both with respect to genetic background and dietary habits, differences in the pathology among species (eg, human beings v mice), and others. It is for these reasons that understanding the mechanism that underlies this relationship between homocysteine and disease is of primary importance. What follows is a brief summary of this effort.

Early animal studies suggested a toxic effect of hyperhomocysteinemia on endothelial cells, resulting in shortened platelet survival,114 but these data have not been confirmed.115,116 

In vitro studies of cultured endothelial cells also showed a toxic effect of homocysteine on cell viability and function, but these studies were conducted using extremely high homocysteine concentrations (1 to 10 mmol), exceeding the levels encountered even under the most severe pathologic conditions.114,117-119 Furthermore, at no time did these studies show that the observed effect was specific for homocysteine and cannot be observed with other sulfur-containing compounds, particularly cysteine, whose concentration in plasma is 20- to 25-fold higher than that of homocysteine. Nonspecific inhibition of prostacyclin synthesis120 and activation of factor V121 by high concentrations of homocysteine on cultured endothelial cells has been reported. Inhibition of protein C activation122 and downregulation of thrombomodulin expression123 at homocysteine concentrations greater than 5 mmol/L have been also observed. One to 5 mmol homocysteine specifically blocks t-PA, but not plasminogen binding to endothelial cells.124 The toxic effect of high homocysteine concentrations on endothelial cells125 also results in increased platelet adhesion,120 impaired regulation of endothelium-derived relaxing factor and related nitrogen oxides,126 induction of tissue factor,127 suppression of heparan sulfate expression,128 and stimulation of smooth muscle cell proliferation.129 Hyperhomocysteinemia may induce oxidation of low-density lipoprotein in vitro.130 Because homocysteine can participate in disulfide bond exchange reactions, it is possible that excessive homocysteine entering the circulation can alter plasma proteins by this process. It has been reported that homocysteine concentrations as low as 8 μmol/L markedly increases the affinity of Lp(a) for plasmin modified fibrin surfaces, thus inhibiting plasminogen activation.131 

It is generally held that different mechanisms are responsible for arterial and venous thromboembolic diseases, involving platelet function abnormalities in arterial thrombosis and abnormalities of coagulation and/or fibrinolysis in venous thromboembolism. Ex vivo studies looking for such abnormalities in patients with hyperhomocysteinemia have given inconclusive results.115,116,132-137 In subjects with severe hyperhomocysteinemia caused by homozygous CBS deficiency, an abnormally high in vivo biosynthesis of thromboxane A2 — as reflected by the urinary excretion of its major metabolite 11-dehydro-thromboxane B2 — has been observed.138 Administration of aspirin inhibits thromboxane production; urinary appearance returns to baseline levels over a time course consistent with platelet survival, suggesting platelets are the major source of increased thromboxane urinary excretion.138 Because thrombin is a potent inducer of platelet activation, the presence of hypercoagulable state was investigated in homocystinuric patients.139 Increased levels of prothrombin fragment 1.2, thrombin-antithrombin complex, and activated protein C were all observed in homocystinuric patients on vitamin treatment who were free of vascular disease. However, these abnormalities did not correlate with urinary thromboxane excretion. Interestingly, protein C levels, but not factor VII and factor II levels, were significantly reduced in homocystinuric patients and correlated with the degree of hyperhomocysteinemia.139 Diet-responsive deficiency of factor VII was previously reported in CBS-deficient patients.134,135,140 Reduced protein C levels may at least partly contribute to venous thrombotic manifestations of patients with homozygous CBS deficiency. The observation that the increased urinary thromboxane excretion was independent of homocysteine levels and was present both in vitamin B6–responsive and nonresponsive patients may have an impact on treatment of hyperhomocysteinemia.139 It is noteworthy that although the effectiveness of vitamin B6 in preventing thromboembolism in pyridoxine-responsive patients was shown to be statistically highly significant, the occurrence of thromboembolism was not abolished by vitamin supplementation.19 

A recent study by Lentz et al141 used a cymonologus monkey model to investigate possible mechanisms of action of mild hyperhomocysteinemia. Mild hyperhomocysteinemia was induced by a diet high in methionine, depleted of folate, and free of choline. Total homocysteine concentrations were 10.6 μmol/L in the experimental monkeys and 4.0 μmol/L in the control group. In response to activation of platelets by infusion of collagen, blood flow to the leg decreased by 42% compared with 14% in controls. The response of resistance vessels to the endothelium-dependent vasodilators, adenosine diphosphate (ADP) and acetylcholine, were markedly impaired in hyperhomocysteinemic monkeys, which indicates that increased vasoconstriction in response to collagen may be caused by decreased vasodilator responsiveness to platelet-generated ADP. Furthermore, thrombomodulin anticoagulant activity in aorta decreased by 34% in the hyperhomocysteinemic monkeys.

Another recent study with promising results showed that incubation of aortic endothelial cells with 50 μmol/L homocysteine for 4 hours led to 64% reduction in the level of glutathione peroxidase.142 Glutathione peroxidase is a key enzyme in the oxidative defense mechanism which was shown to potentiate the action of nitric oxide. These observations indicate that homocysteine may cause endothelial cell injury both by promoting the formation of peroxides as well as by impairing their inactivation. Potential antagonism between nitric oxide (NO)-related functions and homocysteine were first reported by Stamler et al126 and recently discussed in an editorial by Loscalzo,143 who regards this antagonism as stemming from a deleterious effect of chronic exposure to hyperhomocysteinemia on the production of NO, which ultimately leads to unopposed homocysteine-mediated oxidative injury to the endothelium. Existence of antagonism between homocysteine and NO was also implied in a more recent study by de Groote et al,144 which showed that the virulence of Salmonella typhimurium in mice was attenuated in bacterial strains that were incapable of synthesizing homocysteine. Homocysteine may act as NO+ receptor from S-nitrosothiol NO-donor compounds, thereby antagonizing the effect of NO in diverse processes including infection, atherosclerosis, and neurologic disease. Although further experiments are required, these data point to the possibility that elevated homocysteine exerts its effect on more than one system.

Elevations in plasma homocysteine are common in the general population, particularly in the elderly. Vitamin status is a primary determinant of mild to moderate hyperhomocysteinemia accounting for approximately two thirds of all such cases.50 Vitamin supplementation results in near normalization of plasma homocysteine in most cases.145,146 A recent meta-analysis of 38 studies evaluated the risk of hyperhomocysteinemia for arteriosclerotic vascular disease, estimated the reduction of homocysteine levels by folic acid administration, and calculated the potential reduction of coronary artery disease mortality by increasing folic acid intake.147 These analyses proposed that elevations of total homocysteine were an independent graded risk factor for arteriosclerotic vascular disease and calculated that folic acid fortification of food would reduce the annual mortality by 50,000. Vitamin supplementation may also reduce recurrence of venous thromboembolic disease in patients with hyperhomocysteinemia. However, at present the clinical efficacy of this approach has not been tested. In addition, the bulk of evidence indicates that fasting total homocysteine determinations may identify up to 50% of the total population of hyperhomocysteinemic subjects. Patients with isolated methionine intolerance may benefit from vitamin B6 supplementation.

The time is ripe for a placebo/controlled multicenter trial for determining the efficacy of vitamin supplementation in the reduction of morbidity and mortality among patients with occlusive vascular disease, stroke, and thrombosis. Because vitamins are relatively inexpensive, there is little incentive from the part of drug companies to support such a trial and it is up to government agencies to assume this task. For these reasons it is important that the design of such a trial take into account all the information available. Homocysteine metabolism requires the participation of folate as well as vitamin B12 and vitamin B6 coenzymes. Reduction of homocysteine levels in plasma requires that all three of these vitamins are supplemented.

Supported in part with Federal funds from the US Department of Agriculture (USDA), Agricultural Research Service (contract 53-3K06-5-10); the financial support of Telethon-Italy (Grant No. E.C472) is gratefully acknowledged.

Address reprint requests to Jacob Selhub, PhD, Jean Mayer USDA Human Nutrition Research Center on Aging at Tufts University, 711 Washington St, Boston, MA 02111.

1
Selhub
J
Miller
JW
The pathogenesis of homocysteinemia: Interruption of the coordinate regulation by S-adenosylmethionine of the remethylation and transsulfuration of homocysteine.
Am J Clin Nutr
55
1991
131
2
Ueland
PM
Refsum
H
Male
R
Lillehaug
JR
Disposition of endogenous homocysteine by mouse fibroblast C3H/10T1/2 CI 8 and the chemically transformed C3H/10T1/2 MCA CI 16 cells following methotrexate exposure.
J Natl Cancer Inst
77
1986
283
3
Christensen
B
Refsum
H
Vintermyr
O
Ueland
PM
Homocysteine export from cells cultured in the presence of physiological or superfluous levels of methionine: methionine loading of nontransformed, transformed, proliferating, and quiescent cells in culture.
J Cell Biol
146
1991
52
4
Mudd SH, Levy HL, Skovby F: Disorders of transsulfuration, in Scriver CR, Beaudet AL, Sly WS, Valle D (eds): The Metabolic and Molecular Basis of Inherited Disease. New York, NY, McGraw-Hill, 1995, p 1279
5
Fowler
B
Kraus
J
Packman
S
Rosenberg
LE
Homocystinuria: Evidence for three distinct classes of cystathionine-β-synthase mutants in cultured fibroblasts.
J Clin Invest
61
1978
645
6
Barber
GW
Spaeth
GL
The successful treatment of homocystinuria with pyridoxine.
J Pediatr
75
1969
463
7
Kozich
V
Kraus
JP
Screening for mutations by expressing cDNA segments in E. Coli: homocystinuria due to cystathionine β-synthase deficiency.
Hum Mutation
1
1992
113
8
Hu
FL
Gu
Z
Kozich
V
Kraus
JP
Ramesh
V
Shih
VE
Molecular basis of cystathionine β-synthase deficiency in pyridoxine responsive and nonresponsive homocystinuria.
Hum Mol Gen
2
1993
1857
9
Sebastio
G
Sperandeo
MP
Panico
M
de Franchis
R
Kraus
JP
Andria
G
The molecular basis of homocystinuria due to cystathionine-β-synthase deficiency in italian families and report of four novel mutations.
Am J Hum Genet
56
1995
1324
10
Sperandeo
MP
Panico
M
Pepe
A
de Franchis
R
Kraus
JP
Andria
G
Sebastio
G
Molecular analysis of patients affected by homocystinuria due to cystathionine-β-synthase deficiency: Report of a new mutation in exon 8 and a deletion in intron 11.
J Inher Met Dis
18
1995
211
11
Levy
HL
Mudd
SH
Schulman
JD
Dreyfus
PM
Abeles
RH
A derangement in B12 metabolism associated with homocystinemia, cystathioninemia, hypomethioninemia and methylmalonic aciduria.
Am J Med
48
1970
390
12
Goodman
SI
Moe
PG
Hammond
KB
Mudd
SH
Uhlendorf
BW
Homocystinuria with methylmalonic aciduria: Two cases in a sibship.
Biochem Med
4
1970
500
13
Mudd
SH
Levy
HL
Morrow
G
Deranged B12 metabolism: Effects on sulfur aminoacid metabolism.
Biochem Med
4
1970
193
14
Mudd
SH
Uhlendorf
BW
Freeman
JM
Finkelstein
JD
Shih
VE
Homocystinuria associated with decreased methylenetetrahydrofolate reductase activity.
Biochem Biophys Res Commun
46
1972
905
15
Shih
VE
Salam
MZ
Mudd
SH
Uhlendorf
BW
Adams
RD
A new form of homocystinuria due to 5,10-methylenetetrahydrofolate reductase deficiency.
Pediatr Res
6
1972
395
16
Kanwar
JS
Manaligod
JR
Wong
PWK
Morphologic studies in a patient with homocystinuria due to 5,10-methylenetetrahydrofolate reductase deficiency.
Pediatr Res
10
1976
598
17
Kang
SS
Zhou
J
Wong
PWK
Kowalisyn
J
Strokosch
G
Intermediate homocysteinemia: A thermolabile variant of methylenetetrahydrofolate reductase.
Am J Hum Genet
43
1988
414
18
Erbe RW: Inborn errors of folate metabolism, in Blakley RL, Whitehead VM (eds): Folate and pterins: Nutritional, pharmacological and physiological aspects. New York, NY, Wiley 1986, p 413
19
Fenton WA, Rosenberg LE: Inherited disorders of cobalamin transport and metabolism, in Scriver CR, Beaudet AL, Sly WS, Valle D (eds): The Metabolic Basis of Inherited Disease. New York, NY, McGraw-Hill 1989, p 2065
20
Dillon
MJ
England
JM
Gompertz
D
Goodey
PA
Grant
DB
Hussein
HA
Linnell
JC
Matthews
DM
Mudd
SH
News
GH
Seakins
JW
Uhlendorf
BW
Wise
IJ
Mental retardation, megaloblastic anaemia, methylmalonic aciduria and abnormal homocysteine metabolism due to an error in vitamin B12 metabolism.
Clin Sci Molecular Med
47
1974
43
21
Brattstrom
L
Israelsson
B
Norrving
B
Bergqvist
D
Throne
J
Hultberg
B
Hamfelt
A
Impaired homocysteine metabolism in early onset cerebral and peripheral occlusive arterial disease — Effects of pyridoxine and folic acid treatment.
Atherosclerosis
81
1990
51
22
Dudman
NPB
Wilcken
DEL
Wang
J
Lynch
JF
Macey
D
Lundberg
P
Disordered methionine/homocysteine metabolism in premature vascular disease: Its occurrence, cofactor therapy, and enzymology.
Arterioscl Thromb
13
1993
1253
23
Brattstrom
LE
Israelsson
B
Jeppsson
J-O
Hultberg
BL
Folic acid — An innocuous means to reduce plasma homocysteine.
Scand J Clin Lab Invest
48
1988
215
24
Wilcken
DEL
Dudman
NPB
Tyrrell
PA
Robertson
MR
Folic acid lowers elevated plasma homocysteine in chronic renal insufficiency: Possible implications for prevention of vascular disease.
Metabolism
37
1988
697
25
Kang
SS
Wong
PWK
Norusis
M
Homocysteinemia due to folate deficiency.
Metabolism
36
1987
458
26
Lindenbaum
J
Healton
EB
Savage
DG
Brust
JC
Garrett
TJ
Podell
ER
Marcell
PD
Stabler
SP
Allen
RH
Neuropsychiatric disorders caused by cobalamin deficiency in the absence of anemia or macrocytosis.
N Engl J Med
318
1988
1720
27
Kang
SS
Wong
PWK
Zhou
JM
Sora
J
Lessick
M
Ruggie
N
Grcevich
G
Thermolabile methylenetetrahydrofolate reductase in patients with coronary artery disease.
Metabolism
37
1988
611
28
Stabler
SP
Marcell
PD
Podell
ER
Allen
RH
Savage
DG
Lindenbaum
J
Elevation of total homocysteine in the serum of patients with cobalamin or folate deficiency detected by capillary gas-chromatography-mass spectrometry.
J Clin Invest
81
1988
466
29
Levy
HL
Cardinale
GJ
Sulfur aminoacid abnormalities in experimental vitamin B12 deficiency.
Fed Proc
29
1970
634
30
Hollowell
JG
Hall
WK
Cryell
ME
McPherson
J
Hahn
DA
Homocystinuria and organic aciduria in a patient with vitamin B12 deficiency.
Lancet
2
1969
1428
31
Lin
JY
Kang
SS
Zhou
J
Wong
PWK
Homocysteinemia in rats induced by folic acid deficiency.
Life Sciences
44
1989
319
32
Applegarth
DA
Hardwick
DF
Ingram
F
Auckland
NL
Bozoian
G
Excretion of S-adenosylmethionine and S-adenosylhomocysteine in homocystinuria.
N Engl J Med
285
1971
1265
33
Wilcken
DEL
Gupta
VJ
Betts
AK
Homocysteine in the plasma of renal transplant recipients: Effects of cofactors for methionine metabolism.
Clin Sci
61
1981
743
34
Swift
ME
Schultz
TD
Relationship of vitamins B6 and B12 to homocysteine levels: Risk for coronary heart disease.
Nutr Reports Internat
34
1986
1
35
Brattstrom
L
Israelsson
B
Lindgarde
F
Hultberg
B
Higher total plasma homocysteine in vitamin B12 deficiency than in heterozygosity for homocystinuria due to cystathionine β-synthase deficiency.
Metabolism
37
1988
175
36
Miller
JW
Ribaya-Mercado
JD
Russell
RM
Shepard
DC
Morrow
FD
Cochary
EF
Sadowsky
JA
Gershoff
SN
Selhub
J
Effect of vitamin B6 deficiency on plasma homocysteine concentrations.
Am J Clin Nutr
55
1992
1154
37
Boers
GHJ
Fowler
B
Smals
AGH
Trijbels
JMF
Leermakers
AL
Kleijer
WJ
Kloppenborg
PWC
Improved identification of heterozygotes for homocystinuria due to cysthationine synthase activity by the combination of methionine loading and enzyme determination in cultured fibroblasts.
Hum Genet
69
1985
164
38
Sardharwalla
IB
Fowler
B
Robins
AJ
Komrower
GM
Detection of heterozygotes for homocystinuria.
Arch Dis Child
49
1974
553
39
Smolin
LA
Bebvenga
NJ
Factors affecting the accumulation of homocyst(e)ine in rats deficient in vitamin B-6.
J Nutr
114
1984
103
40
Park YK, Linkswiler H. Effect of vitamin B6 depletion in adult man on the excretion of cystathionine and other methionine metabolites. J Nutr 100:110, 1970
41
Miller
JW
Nadeau
MR
Smith
D
Selhub
J
Vitamin B6 deficiency vs folate deficiency: Comparison of responses to methionine loading in rats.
Am J Clin Nutr
59
1994
1033
42
Miller
JW
Nadeau
MR
Smith
J
Smith
D
Selhub
J
Folate deficiency-induced homocysteinemia in rats: Disruption of S-adenosylmethionine's coordinate regulation of homocysteine metabolism.
Biochem J
298
1994
415
43
Bostom
AG
Jacques
PF
Nadeau
MR
Williams
RR
Ellison
RC
Selhub
J
Post-methionine load hyperhomocysteinemia in persons with normal fasting total plasma homocysteine — Initial results from The NHLBI Family Heart Study.
Atherosclerosis
116
1995
147
44
Lindgren
F
Israelsson
B
Lindgren
A
Hultberg
B
Andersson
A
Brattstrom
L
Plasma homocysteine in acute myocardial infarction: Homocysteine-lowering effect of folic acid.
J Intern Med
237
1995
381
45
Ubbink
JB
Vermaak
WJH
van der Merwe
A
Becker
PJ
Vitamin B12, vitamin B6, and folate nutritional status in men with hyperhomocysteinemia.
Am J Clin Nutr
57
1993
47
46
Ubbink
JB
Vermaak
WJH
van der Merwe
A
Becker
PJ
Delport
R
Potgieter
HC
Vitamin requirements for the treatment of hyperhomocysteinemia in humans.
J Nutr
124
1994
1927
47
Franken
DG
Boers
GHJ
Blom
HJ
Trijbels
JMF
Effect of various regimens of vitamin B-6 and folic acid on mild hyperhomocysteinemia in vascular patients.
J Inher Metab Dis
17
1994
159
48
van den Berg
N
Franken
DG
Boers
GHJ
Blom
HJ
Jakobs
C
Stehouwer
CDA
Rauwerda
JA
Combined vitamin B-6 plus folic acid therapy in young patients with arteriosclerosis and hyperhomocysteinemia.
J Vasc Surg
20
1994
933
49
Pancharunity
N
Lewis
CA
Sauberlich
HE
Perkins
LL
Go
RC
Alvarez
JO
Macaluso
M
Acton
RT
Copeland
RB
Cousins
AL
Gore
TB
Cornwell
PE
Roseman
JM
Plasma homocysteine, folate and B6 concentrations and the risk for early-onset coronary artery disease.
Am J Clin Nutr
59
1994
940
50
Naurath
HJ
Joosten
E
Riezler
R
Stabler
SP
Allen
RH
Lindenbaum
J
Effects of vitamin B12, folate, and vitamin B6 supplements in elederly people with normal serum vitamin concentrations.
Lancet
346
1995
85
51
Dalery
K
Lussier-Cacan
S
Selhub
J
Davignon
J
Latour
Y
Genest
J
Homocysteine and coronary artery disease in french canadian subjects: Relationship with vitamin B12, B6, pyridoxal phosphate and folate.
Am J Cardiol
75
1995
1107
52
Selhub
J
Jacques
PF
Wilson
PWF
Rush
D
Rosenberg
IH
Vitamin status and intake as primary determinants of homocysteinemia in the elderly.
JAMA
270
1993
2693
53
Rosenblatt DS: Inherited disorders of folate transport and metabolism, in Scriver CR, Beaudet AL, Sly WS, Valle D (eds): The Metabolic Basis of Inherited Disease. New York, NY, McGraw-Hill, 1989, p 2049
54
Goyette
P
Sumner
JS
Milos
R
Duncan
MV
Rosenblatt
DS
Matthews
RG
Rozen
R
Human methylenetetrahydrofolate reductase: Isolation of cDNA, mapping and mutation identification.
Nat Gene
7
1994
195
55
Goyette
P
Frosst
P
Rosenblatt
DS
Rozen
R
Seven novel mutations in the methylenetetrahydrofolate reductase gene and genotype/phenotype correlations in severe methylenetetrahydrofolate reductase deficiency.
Am J Hum Genet
56
1995
1052
56
Kang
SS
Wong
PWK
Susmano
A
Sora
J
Norusis
M
Ruggie
N
Thermolabile methylenetetrahydrofolate reductase: An inherited risk factor for coronary artery disease.
Am J Hum Genet
48
1991
536
57
Kang
SS
Passen
EL
Ruggie
N
Wong
PWK
Sora
H
Thermolabile defect of methylenetetrahydrofolate reductase in coronary artery disease.
Circulation
88
1993
1463
58
Engbersen
AMT
Franken
DG
Boers
GHJ
Stevens
EMB
Trijbels
FJM
Blom
HJ
Thermolabile 5,10-methylenetetrahydrofolate reductase as a cause of mild hyperhomocysteinemia.
Am J Hum Genet
56
1995
142
59
Frosst
P
Blom
HJ
Milos
R
Goyette
P
Sheppard
CA
Matthews
RG
Boers
GJH
den Heijer
M
Kluijtmans
LAJ
van den Heuvel
Rozen R
A candidate genetic risk factor for vascular disease: A common mutation in methylenetetrahydrofolate reductase.
Nat Genet
10
1995
111
60
Motulsky
A
Nuritional ecogenetics: Homocysteine-related arteriosclerotic vascular disease, neural tube defects, and folic acid.
Am J Hum Genet
58
1996
17
61
Kluijtmans
L
van den Heuvel
L
Boers
G
Frosst
P
Stevens
E
van Oost
B
den Heijer
M
Trijbels
FJ
Rozen
R
Blom
HJ
Molecular genetic analysis in mild hyperhomocysteinemia: A common mutation in the 5,10-methylenetetrahydrofolate reductase gene is a genetic factor for cardiovascular disease.
Am J Hum Genet
58
1996
35
62
de Franchis
R
Mancini
FP
D'Angelo
A
Sebastio
G
Fermo
I
De Stefano
V
Margaglione
M
Mazzola
G
Di Minno
G
Andria
G
Elevated total plasma homocysteine (tHcy) and 677CT mutation of the 5,10-methylenetetrahydrofolate reductase gene in thrombotic vascular disease.
Am J Hum Genet
59
1996
262
63
D'Angelo
A
Fermo
I
Vigano'D'Angelo
S
Methionine loading, B-6 status, and premature thromboembolic disease: Implications for the pathogenesis and treatment of hyperhomocysteinemia — Reply.
Ann Intern Med
125
1996
419
64
Jacques
PF
Bostom
AG
Williams
RR
Ellison
RC
Eckfeldt
JH
Rosenberg
IH
Selhub
J
Rozen
R
Relation between folate status, a common mutation in methylenetetrahydrofolate reductase, and plasma homocysteine concentrations.
Circulation
93
1996
7
65
McCully
KS
Vascular pathology of homocysteinemia: Implications for the pathogenesis of arteriosclerosis.
Am J Pathol
56
1969
111
66
Mudd
SH
Finkelstein
JD
Irreverre
F
Laster
L
Homocystinuria: An enzymatic defect.
Science
143
1964
1443
67
Wilcken
DEL
Wilcken
B
The pathogenesis of coronary artery disease. A possible role for methionine metabolism.
J Clin Invest
57
1976
1079
68
Murphy-Cuthorian
DR
Wexman
MP
Grieco
AJ
Heiringer
JE
Glassman
E
Guall
GE
Ng
SK
Feit
F
Wexman
K
Fox
AC
Methionine intolerance: A possible risk factor for coronary artery disease.
J Am Coll Cardiol
6
1985
725
69
Clarke
R
Daly
L
Robinson
K
Naughten
E
Cahalane
S
Fowler
B
Graham
I
Hyperhomocysteinemia: An independent risk factor for vascular disease.
N Engl J Med
324
1991
1149
70
Kang
SS
Wong
PWK
Cook
HY
Norusis
M
Messer
JV
Protein-bound homocyst(e)ine. A possible risk factor for coronary artery disease.
J Clin Invest
77
1986
1482
71
Israelsson
B
Brattstrom
LE
Hultberg
BL
Homocysteinemia and myocardial infarction.
Atherosclerosis
71
1988
227
72
Olszewski
AJ
Szostak
WB
Homocysteine content of plasma proteins in ischemic heart disease.
Atherosclerosis
69
1988
109
73
Ubbink
JG
Vermack
WJH
Bennett
JM
Becker
PG
VanStaden
DA
Bissbort
S
The prevalence of homocysteinemia and hypercholesterolemia in angiographically defined coronary artery disease.
Klin Wochenschr
69
1991
527
74
Boers
GHJ
Smals
AGH
Trijbels
FJM
Fowler
B
Bakkeren
JAJM
Schoonderwaldt
HC
Klijer
WJ
Kloppenborg
PWC
Heterozygosity for homocystinuria in premature peripheral and cerebral occlusive arterial disease.
N Engl J Med
313
1985
709
75
Araki
A
Sako
Y
Fukushima
Y
Matsumoto
M
Asada
T
Kita
T
Plasma sulphydryl-containing amino acids in patients with cerebral infarction and in hypertensive subjects.
Atherosclerosis
79
1989
139
76
Coull
BM
Malinow
MR
Beamer
N
Sexton
G
Nordt
F
deGarmo
P
Elevated plasma homocysteine in acute stroke and TIA: A possible independent risk factor for stroke.
Stroke
21
1990
572
77
Brattstrom
L
Lindgren
A
Israelsson
B
Malinow
MR
Norrving
B
Upson
B
Hyperhomocysteinaemia in stroke — Prevalence, cause, and relationships to type of stroke and stroke risk factors.
Eur J Clin Invest
22
1992
214
78
Brattstrom
L
Hardebo
JE
Hultberg
B
Moderate homocysteinemia. A possible risk factor for arteriosclerotic cerebrovascular disease.
Stroke
15
1984
1012
79
Malinow
MR
Kang
SS
Taylor
LM
Wong
PWK
Coull
B
Inahara
T
Mukerjee
D
Sexton
G
Upson
B
Prevalence of hyperhomocysteinemia in patients with peripheral arterial occlusive disease.
Circulation
79
1989
1180
80
Molgaard
J
Malinow
MR
Lassvik
C
Holm
AC
Upson
B
Olsson
AG
Hyperhomocyst(e)inaemia: An independent risk factor for intermittent claudication.
J Intern Med
231
1992
273
81
Taylor
LM
De Frang
RD
Harris
EJ
Jr
Poryter
JM
The association of elevated plasma homocyst(e)ine with progression of symptomatic peripheral arterial disease.
J Vasc Surg
13
1991
128
82
Glueck
CJ
Shaw
P
Lang
JE
Tracy
T
Sieve-Smith
L
Wang
Y
Evidence that homocysteine is an independent risk factor for atherosclerosis in hyperlipidemic patients.
Am J Cardiol
72
1995
132
83
Genest
JJ
McNamara
JR
Salem
DN
Wilson
PWF
Shaefer
EJ
Malinow
MR
Plasma homocyst(e)ine levels in men with premature coronary artery disease.
J Am Coll Cardiol
16
1990
1114
84
Rubba
P
Faccenda
F
Pauciullo
P
Carbone
L
Mancini
M
Strissciuglio
P
Carrozzo
R
Sartorio
R
Del Giudice
E
Andria
G
Early signs of vascular disease in homocystinuria: A non-invasive study by ultrasound methods in eight families with cystathionine-β-synthase deficiency.
Metabolism
39
1990
1191
85
Malinow
MR
Nieto
FJ
Sklo
M
Chambless
LE
Bond
G
Carotid artery intimal-medial wall thickening and plasma homocyst(e)ine in asymptomatic adults. The Atherosclerosis in Communities (ARIC) Study.
Circulation
87
1993
1107
86
Selhub
J
Jacques
PF
Bostom
AG
D'Agostino
RB
Wilson
PWF
Belanger
AJ
O'Leary
DH
Wolf
PA
Schaefer
EJ
Rosenberg
IH
Association between plasma homocysteine concentrations and extracranial carotid-artery stenosis.
N Engl J Med
332
1995
286
87
Ueland PM, Refsum H, Brattstrom L: Plasma homocysteine and cardiovascular disease, in Francis RBJ (ed): Atherosclerotic Cardiovascular Disease Hemostasis, and Endothelial Function. New York, NY, Dekker, 1992, p 183
88
Stampfer
MJ
Malinow
MR
Willet
WC
Newcomer
LM
Upson
B
Ullmann
D
Tishler
PV
Hennekens
CH
A prospective study of plasma homocyst(e)ine and risk of myocardial infarction in US physicians.
JAMA
268
1992
877
89
Arnesen
E
Refsum
H
Bonaa
KH
Ueland
PM
Forde
OH
Nordrehaug
JE
Serum total homocysteine and coronary heart disease.
Int J Epidemiol
24
1995
704
90
Mudd
SH
Levy
HL
Skovby
F
The natural history of homocystinuria due to cystathionine β-synthase deficiency.
Am J Hum Genet
37
1985
1
91
Hirsh J, Prins MH, Samama M: Approach to the thrombophilic patient, in Colman RW, Hirsh J, Marder VJ, Salzman E (eds): Hemostasis and Thrombosis: Basic and Clinical Practice. Philadelphia, PA, Lippincott, 1994, p 1543
92
Brattstrom
L
Tengborn
L
Lagerstedt
C
Israelsson
B
Hultberg
B
Plasma homocysteine in venous thromboembolism.
Haemostasis
21
1991
51
93
Amundsen
T
Ueland
PM
Waage
A
Plasma homocysteine levels in patients with deep venous thrombosis.
Arterioscler Thromb Vasc Biol
15
1995
1321
94
Bienvenu
T
Ankri
A
Chadefauz
B
Montalescot
G
Kamoun
P
Elevated total plasma homocysteine, a risk factor for thrombosis; relation to coagulation and fibrinolytic parameters.
Thromb Res
70
1993
123
95
Falcon
CR
Cattaneo
M
Panzeri
D
Martinelli
I
Mannucci
PM
High prevalence of hyperhomocysteinemia in patients with juvenile venous thrombosis.
Arterioscler Thromb
14
1994
1080
96
Fermo
I
Vigano'D'Angelo
S
Paroni
R
Mazzola
G
Calori
G
D'Angelo
A
Prevalence of moderate hyperhomocysteinemia in patients with early-onset venous and arterial occlusive disease.
Ann Intern Med
123
1995
747
97
den Heijer
M
Blom
HJ
Gerrits
WBJ
Rosendaal
FR
Haak
HL
Wijermans
PW
Bos
GMJ
Is hyperhomocysteinemia a risk factor for recurrent thrombosis?
Lancet
345
1995
882
98
Cattaneo
M
Martinelli
I
Mannucci
PM
Hyperhomocysteinemia as a risk factor for deep-vein thrombosis.
N Engl J Med
335
1996
974
99
Simioni
P
Prandoni
P
Burlina
A
Tormene
D
Sardella
C
Ferrari
V
Benedetti
L
Girolami
A
Hyperhomocysteinemia and deep-vein thrombosis. A case control study.
Thromb Haemost
76
1996
883
100
den Heijer
M
Koster
T
Blom
HJ
Bos
GMJ
Briet
E
Reitsma
PH
Vandenbroucke
JP
Rosendaal
F
Hyperhomocysteinemia as a risk factor for deep vein thrombosis.
N Engl J Med
334
1996
759
101
Petri
M
Roubenoff
R
Dallal
GE
Nadeau
MR
Selhub
J
Rosenberg
IH
Plasma homocysteoine as a risk factor for atherothrombotic events in systemic lupus erythmatosus.
Lancet
348
1996
1120
102
Mandel H, Brenner B, Berant M, Rosenberg N, Lanir N, Jakobs C, Fowler B, Seligsohn U. Coexistence of hereditary homocystinuria and factor V Leiden — Effect on thrombosis. N Engl J Med 334:763, 1996
103
D'Angelo
A
Fermo
I
Vigano'D'Angelo
S
Coexistence of hereditary homocystinuria and factor V Leiden — Effect on thrombosis.
N Engl J Med
335
1996
1289
104
Mudd
SH
Havlik
R
Levy
HL
McKusick
VA
Feinleib
M
A study of cardiovascular risk in heterozygotes for homocystinuria.
Am J Hum Genet
33
1981
883
105
Swift
M
Morrell
D
Cardiovascular risk in homocystinuria family members.
Am J Hum Genet
34
1982
1016
106
Clarke R: The Irish experience, in Robinson K (ed): Homocysteinaemia and Vascular Disease. Luxembourg, Commission of the Europen Communities, 1990, p 41
107
Kozich
V
Kraus
E
de Franchis
R
Fowler
B
Boers
GHJ
Graham
I
Kraus
JP
Hyperhomocysteinemia in premature arterial disease: Examination of cystathionine β-synthase alleles at the molecular level.
Hum Mol Genet
4
1995
623
108
Watanabe
M
Osada
J
Aratani
Y
Kluckman
K
Reddick
R
Malinow
MR
Maeda
N
Mice deficient in cystathionine β-synthase: Animal models for mild and severe homocyst(e)inemia.
Proc Natl Acad Sci USA
92
1995
1585
109
Wilcken
DE
Wang
XL
Sim
AS
McCredie
RM
Distribution inhealthy and coronary populations of the methylenetetrahydrofolate reductase (MTHFR) C677T mutation.
Arter Thromb Vasc Biol
16
1996
878
110
Schmitz
C
Lindpaintner
K
Verhoef
P
Gazaino
JM
Buring
J
Genetic polymorphism of methylenetetrahydrofolate reductase and myocardial infraction: A case-control study.
Circulation
94
1996
1812
111
Christensen
B
Frosst
P
Lussier-Cacan
S
Selhub
J
Goyette
P
Rosenblatt
D
Genest
J
Rozen
R
Correlation of a common mutation in the methylenetetrahydrofolate reductase (MTHFR) gene with plasma homocysteine in patients with premature coronary artery disease.
Arterioscler Thromb Vasc Biol
17
1997
569
112
Adams
M
Smith
PD
Martin
D
Thompson
JR
Lodwick
D
Samani
NJ
Genetic analysis of thermolabile methylenetetrahydrofolate reductase as a risk factor for myocardial infarction.
Q J Med
89
1996
437
113
Ma
J
Stampfer
MJ
Hennekens
CH
Forst
P
Selhub
J
Horsford
J
Malinow
MR
Willett
W
Roze
R
Methylenetetrahydrofolate reductase polymorphism and risk of myocardial infarction in the U.S. physicians.
Circulation
94
1996
2410
114
Harker
LA
Harlan
JM
Ross
R
Effect of sulfinpyrazone on homocysteine-induced endothelial injury and arteriosclerosis in baboons.
Circ Res
53
1983
731
115
Uhlemann ER, TenPas JH, Lucky AW, Schulman JD, Mudd SH, Shulamn NR: Platelet survival and morphology in homocystinuria due to cysthationine synthase deficiency 295:1283, 1976
116
Hill-Zobel
R
Pyeritz
RE
Scheffel
U
Malpica
O
Engin
S
Camargo
EE
Abbott
M
Guilarte
TR
Hill
J
McIntyre
PA
Murphy
EA
Min-Fu
T
Kinetics and distribution of 111-Indium-labeled platelets in patients with homocystinuria.
N Engl J Med
307
1982
781
117
Wall
RT
Harlan
JM
Harker
LA
Striker
GE
Homocysteine-induced endothelial cell injury in vitro: A model for the study of vascular injury.
Thromb Res
18
1980
113
118
De Groot
PG
Willems
C
Boers
GHJ
Gonsalves
MD
van Aken
WG
van Mourik
JA
Endothelial cell dysfunction in homocystinuria.
Eur J Clin Invest
13
1983
405
119
Starkebaum
G
Harlan
JM
Endothelial cell injury die to copper-catalyzed hydrogen peroxide generation from homocysteine.
J Clin Invest
77
1986
1370
120
Wang
J
Dudman
NPB
Wilcken
DE
Effects of homocysteine and related compounds on prostacyclin production by cultured human vascular endothelial cells.
Thromb Haemost
6
1993
1047
121
Rodgers
GM
Kane
WH
Activation of endogenous factor V by a homocysteine-induced vascular endothelial cell activator.
J Clin Invest
77
1986
1909
122
Rodgers
GM
Conn
MT
Homocysteine, an atherogenic stimulus, reduces protein C activation by arterial and venous endothelial cells.
Blood
75
1990
895
123
Lentz
SR
Sadler
JE
Inhibition of thrombomodulin surface expression and protein C activation by the thrombogenic agent homocysteine.
J Clin Invest
88
1991
1906
124
Hajjar
KA
Homocysteine-induced modulation of tissue plasminogen activator binding to its endothelial cell membrane receptor.
J Clin Invest
91
1993
2873
125
Blann
AD
Endothelial cell damage and homocysteine.
Atherosclerosis
94
1992
89
126
Stamler
JS
Osborne
JA
Jaraki
M
Rabbini
LE
Mullins
M
Singel
D
Loscalzo
J
Adverse vascular effects of homocysteine are modulated by endothelium-derived relaxing factor and related oxides of nitrogen.
J Clin Invest
91
1993
308
127
Fryer
RH
Wilson
BD
Gubler
DB
Fitzgerald
LA
Rodgers
GM
Homocysteine, a risk factor for premature vascular disease and thrombosis, induces tissue factor activity in endothelial cells.
Arterioscler Thromb
13
1993
1327
128
Nishinaga
M
Ozawa
T
Shimada
K
Homocysteine, a thrombogenic agent, suppresses anticoagulant heparan sulfate expression in cultured porcine aortic endothelial cells.
J Clin Invest
92
1993
1381
129
Tsai
J-C
Perrella
MA
Yoshizumi
M
Hsieh
CM
Haber
E
Schlegel
R
Lee
M
Promotion of vascular smooth muscle cell growth by homocysteine: a link to athersoclerosis.
Proc Natl Acad Sci USA
91
1994
6369
130
Pathasarathy
S
Oxidation of low-density lipoprotein by thiol compounds leads to its recognition by the acetyl LDL receptor.
Biochim Biophys Acta
917
1987
337
131
Harpel
PC
Chang
VT
Borth
W
Homocysteine and other sulfhydryl compounds enhance the binding of lipoprotein(a) to fibrin: A potential biochemical link between thrombosis, atherogenesis and sulfhydryl compounds metabolism.
Proc Natl Acad Sci USA
89
1992
10193
132
Harker
LA
Slichter
SJ
Scott
CR
Ross
R
Homocysteinemia. Vascular injury and arterial thrombosis.
N Engl J Med
291
1974
537
133
Giannini
MJ
Coleman
M
Innerfield
I
Antithrombin activity on homocystinuria.
Lancet
1
1974
1094
134
Maruyama
I
Fukuda
R
Kaszmama
M
Abe
T
Yoshida
Y
Igata
A
A case of homocystinuria with low antithrombin activity.
Acta Haematol Jpn
40
1977
267
135
Palareti
G
Salardi
S
Pizzi
S
Legnani
C
Poggi
M
Grauso
F
Caniato
A
Coccheri
S
Cacciari
E
Blood coagulation changes in homocystinuria: Effects of pyridoxine and other specific therapy.
J Pediatr
109
1986
1001
136
Merckz
J
Kuntz
F
Deficit en facteur VII et homocystinurie: Association fortuite ou syndrome?
Nouv Presse Med
10
1981
3769
137
Munnich
A
Saudbray
JM
Dautzenberg
MD
Parvy
P
Ogier
H
Girot
R
Manigne
P
Frezal
J
Diet-responsive proconvertin (factor VII) deficiency in homocystinuria.
J Pediatr
102
1983
730
138
Di Minno
G
Davi'G
Margaglione M
Cirillo
F
Grandone
E
Ciabattoni
E
Catalano
I
Strisciuglio
P
Andria
G
Patrono
C
Mancini
M
Abnormally high thromboxane biosynthesis in homozygous homocystinuria. Evidence for platelet involvement and Probucol-sensitive mechanism.
J Clin Invest
92
1993
1400
139
Di Minno G, Coppola A, D'Angelo A, Davi' G, Cerbone AM, D'Angelo SA, della Valle P, Strisciuglio P, Rubba P, Mancini FP, Andria G, Patrono C: Does raised in vivo thromoboxane A2 biosynthesis reflect hypercoagulable state in homocystinuria due to homozygous cystathionine β-synthase deficiency. Arteriosc Thromb Vasc Biol 1996 (in press)
140
Brattstrom
L
Israelsson
B
Tengborn
L
Hultberg
B
Homocysteine, factor VII and antithrombin III in subjects with different gene dosage for cystathionine β-synthase.
J Inher Metab Dis
12
1989
475
141
Lentz
RS
Sobey
CG
Pigeors
DJ
Bhopatkar
MY
Faraci
FM
Faraci
FM
Malinow
MR
Heistad
DD
Vascular dysfunction in monkeys with diet-induced hyperhomocyst(e)inemia.
J Clin Invest
98
1996
24
142
Upchurch GR Jr, Welch GN, Freedman JE, Loscalzo J: Homocysteine attenuates endothelial glutathione peroxidase and thereby potentiates peroxide-mediated injury. Circulation 92:I-228, 1995
143
Loscalzo
J
The oxidant stress of homocyst(e)inemia.
J Clin Invest
98
1996
5
144
de Groote
MA
Testerman
T
Xu
Y
Stauffer
G
Fang
FC
Homocysteine antagonism of nitric oxide-related cytostasis in Salmonella typhimurium.
Science
272
1996
414
145
Franken
DG
Boers
GHJ
Blom
HJ
Trijbels
FJM
Kloppenborg
PW
Treatment of mild hyperhomocysteinemia in vascular disease patients.
Arterioscl Thromb
14
1994
465
146
Brattstrom
L
Lindgren
A
Israelsson
B
Andersson
A
Hultberg
B
Homocysteine and cysteine: Determinants of plasma levels in middle-aged and elderly subjects.
J Intern Med
236
1994
631
147
Boushey
CJ
Beresford
SAA
Omenn
GS
Motulsky
AG
A quantitative assessment of plasma homocysteine as a risk factor for vascular disease. Probable benefits of increasing folic acid intakes.
JAMA
274
1995
1049
Sign in via your Institution