13

Homocysteine, vitamin B12, folates, vitamin B6, choline, betaine

13

Homocysteine, vitamin B12, folates, vitamin B6, choline, betaine

13

Homocysteine, vitamin B12, folates, vitamin B6, choline, betaine

13

Homocysteine, vitamin B12, folates, vitamin B6, choline, betaine

13.1 Hyperhomocysteinemia and degenerative diseases

Wolfgang Herrmann, Rima Obeid

13.1.1 Introduction

Homocysteine is a non proteinogenic, sulfur-containing amino acid of the methionine metabolism. Only 1–2% of homocysteine circulate freely in the blood in the reduced form, 70–90% of plasma homocysteine are protein-bound and the remaining proportions are disulfides, such as homocystine or the mixed disulfide homocysteine-cysteine (Fig. 13.1-1 – Homocysteine forms).

Homocystinuria is a congenital disorder of amino acid metabolism. This metabolic disorder is associated with a strong increase in plasma homocysteine concentration and markedly increased urinary excretion of the oxidation product homocystin /1/. If untreated, affected individuals develop already in their youth severe atherosclerotic vascular alterations, mental retardation and bone deformation besides other symptoms. According to the homocysteine theory of atherosclerosis, there is a causal association between significantly increased plasma homocysteine concentrations and atherosclerosis /1/. Numerous studies have provided evidence that even moderately elevated plasma homocysteine concentrations are positively associated with an increased risk of cardiovascular disease, venous thrombosis, peripheral arterial occlusion disease and stroke.

Hyperhomocysteinemia is considered as an independent risk factor for cardiovascular disease and thought to account for about 10% of the total risk /23/. Hyperhomocysteinemia is also a risk factor for neurodegenerative diseases such as vascular dementia, Alzheimer’s dementia or cognitive impairment /4/. Older adults, in particular, have a high prevalence of hyperhomocysteinemia due to common vitamin deficiencies /5/ that can easily be treated with adequate vitamin substitution.

13.1.2 Homocysteine metabolism

Homocysteine stands at the intersection of two metabolic pathways: catabolic transsulfuration and the remethylation cycle (Fig. 13.1-2 – Homocysteine metabolism/6/. Homocysteine is either transsulfurated to cystathionine or remethylated to methionine. The remethylation rate depends on the dietary methionine intake. During remethylation, homocysteine binds a methyl group of 5-methyltetrahydrofolate (5-methyl-THF) and forms methionine. The reaction occurs in all tissues and is catalyzed by the vitamin B12-dependent enzyme methionine synthase (EC 2.1.1.13).

Alternatively, the methyl group can also be provided by betaine. However, this reaction is present only in liver and kidney. Transfer of the methyl group from betaine to homocysteine is catalyzed by the enzyme betaine homocysteine methyltransferase (BHMT, EC 2.1.1.5) /7/.

The product of betaine de methylation is dimethylglycine (DMG), which can be further converted to sarcosine and later to glycine. Betaine is formed by enzymatic oxidation of choline. Consequently, choline deficiency can be a cause of hyperhomocysteinemia /8/.

Methionine is converted to S-adenosylmethionine (SAM), which functions as a universal donor of methyl groups.

The de methylation of SAM results in the formation of S-adenosylhomocysteine (SAH), from which homocysteine is released by hydrolytic cleavage. SAH is a potent competitor for binding sites of SAM on methyltransferases and, therefore, can inhibit transmethylation /9/.

In the catabolic transsulfuration pathway, homocysteine condenses with serine to form cystathionine. This reaction is catalyzed by cystathionine-β-synthase (CBS, EC 4.4.1.1), a vitamin B6-dependent enzyme.

In a second reaction step catalyzed by γ-cystathionase (EC 4.4.1.1), likewise a vitamin B6-dependent enzyme, cystathionine is hydrolyzed to form cysteine and α-keto butyrate. SAM plays an important role in the regulation between the metabolic pathways of remethylation and transsulfuration /10/.

SAM activates the CBS and inhibits the synthesis of 5-methyl-THF catalyzed by 5,10-methylenetetrahydrofolate reductase (MTHFR, EC 1.5.1.20).

The remethylation activity catalyzed by glycine N-methyl transferase (GNMT, EC 2.1.1.20) plays an important role in the regulation of SAM consumption, where the methyl group is transferred from SAM to glycine. The reaction is inhibited by 5-methyl-THF.

13.1.3 Hyperhomocysteinemia

As homocysteine is highly cytotoxic, its intracellular concentration is kept low by remethylation to form methionine and by a cellular homocysteine export mechanism to the plasma /11/. Most of the plasma homocysteine (about 70%) is metabolized renally (remethylated) and only a small proportion is excreted unchanged in urine.

Moderate hyperhomocysteinemia with plasma concentrations of 12–30 μmol/L can have various causes:

  • Deficiencies in folate, vitamin B12 or vitamin B6
  • Enzyme deficiency, the best known being MTHFR deficiency
  • Chronic kidney disease /12/.

Refer to Tab. 13.1-1 – Homocyteine level in fasting plasma and following oral methionine load.

Plasma homocysteine levels above 30 μmol/L are primarily caused by genetic enzyme deficiencies or renal dysfunction. Homocysteine metabolism is influenced by numerous substances, drugs, diseases and life-style factors, mostly as direct or indirect antagonists of co factors and enzyme activities, but also by disulfide exchange reactions, resorption disorders and enzyme induction (Tab. 13.1-2 – Causes of homocysteine changes and Tab. 13.1-3 – Mutations associated with hyperhomocysteinemia).

13.1.3.1 Genetic influences

Homocysteine levels between 30–100 μmol/L are primarily caused by heterozygous enzyme defects. Severe hyperhomocysteinemia with concentrations above 100 μmol/L is likely caused by an enzyme defect with a marked loss of enzyme activity (Tab. 13.1-3 – Mutations associated with hyperhomocysteinemia).

13.1.3.2 Associated diseases

Diseases associated with hyperhomocysteinemia are atherosclerosis, cardiovascular disease, congestive heart failure, thrombosis, stroke, neural tube defects, pregnancy complications, cognitive dysfunctions, depression, vascular dementia, Alzheimer’s dementia, hypertension and tumors.

Diseases, drugs, and conditions associated with hyperhomocysteinemia are described in Tab. 13.1-4 – Causes and diseases possibly associated with hyperhomocysteinemia.

13.1.3.3 Age and gender dependency

Homocysteine concentration generally increases with age; in younger age, men usually have higher levels than women. In about 40-year-olds, the gender-related difference is about 2 μmol/L and can be explained with the estrogen effect in women. The gender-related difference decreases rapidly during menopause.

Age-dependent increase in homocysteine is, at least partly, explained with the physiological decline in renal function /5/. The increase in homocysteine concentration is linear up to 60–65 years of age and then becomes much faster, increasing by about 10% or 1 μmol/L per decade on average /3/.

13.1.4 Homocysteine reduction through folic acid supplementation

The homocysteine-lowering effect of different folic acid doses was studied /44454647/. Participants with higher baseline homocysteine concentrations (16.6 μmol/L) had the greatest reductions in plasma homocysteine in response to daily folic acid doses of 0.2 mg (–20.6%), 0.4 mg (–20.7%), and 0.8 mg (–27.8%). In those with lower baseline homocysteine levels (10.1 μmol/L), the responses to the same doses were –8.2%, –8.9% and –8.3%, respectively /44/. Folic acid supplementation caused a linear, dose-dependent increase in folate concentration in blood. Due to the six-month supplementation of 200 μg/d, plasma folate was increased from 18 to 34 nmol/L and reached about 50% of the increase attained with 800 μg/d. Long-term supplementation with low folic acid doses (200 μg/d) are sufficient to efficiently improve homocysteine metabolism. Refer in addition to Fig. 13.1-5 – Median changes in homocysteine concentrations after 26 weeks of treatment with placebo or 0.2 mg, 0.4 mg or 0.8 mg folic acid per day).

13.1.5 Hyperhomocysteinemia as risk factor

Hyperhomocysteinemia is a general risk factor for many organ systems.

13.1.5.1 Vascular disease

Meta-analyses /48/ of retrospective and prospective studies emphasize the relation between hyperhomocysteinemia and degenerative vascular disease. Statistical analyses of this data regarding an increase in homocysteine of 5 μmol//L calculated an odds ratio of 1.32 for the incidence of ischemic heart disease, 1.60 for venous thrombosis and 1.59 for stroke. Moreover, these meta analyses show that lowering of the plasma homocysteine concentration by 3 μmol/L compared to the baseline level could be associated with a 16% lower risk of ischemic heart disease, 25% lower risk of thrombosis or 24% lower risk of stroke /2/. Furthermore, a prospective study over more than four years reported a significant decrease in the plaque volume of the carotid artery after supplementation with B vitamins /49/. In addition, a significant decrease of the carotid intima media thickness has been reported in patients at risk of cerebral ischemia after one-year supplementation with B vitamins /50/.

Secondary prevention studies

Despite the association between hyperhomocysteinemia and cardiovascular disease, the lowering of the disease rate by reduced homocysteine concentration following vitamin therapy as a key element of the causality chain has remained unsolved.

More than 50,000 individuals have been included in intervention studies worldwide to clarify the potential benefit of homocysteine lowering therapy with B vitamins (secondary prevention):

  • NORVIT Study /51/ (3,749 patients who had had an acute myocardial infarction were additionally treated with B vitamins for three years)
  • VISP Study /52/ (3,860 stroke patients were treated with low-dose or high-dose B vitamins for two years)
  • HOPE-2 Study /53/ (5,522 patients who had vascular disease or diabetes were treated with B vitamins or placebo for five years)
  • Search Study /54/ (12,064 survivors of myocardial infarction received 2 mg of folic acid plus 1 mg of vitamin B12 and/or placebo daily for 6.7 years).

In these studies, homocysteine serum level was lowered significantly by 17–28%, but no significant risk reduction regarding the end points (myocardial infarction, stroke) was observed. The clinical significance of secondary prevention studies is clearly limited, primarily because, due to the basic medication of participating patients, it is difficult to provide evidence of the additional effect of homocysteine reduction by B vitamins on the cardiovascular end points and, as a result, the diagnostic power is only considered adequate in large-scale meta analyses.

The situation is different in stroke prevention. Here, the expected risk reduction is about 25%; hence, a meta analysis requires considerably fewer patients in order to draw a well-founded conclusion.

Meta analyses

One meta analysis /55/ comprises data from 75 studies (22,068 patients and 23,618 controls) as well as 14 randomized trials on 39,597 participants with cardiovascular disease (CVD) events and homocysteine lowering. As a result, an independent significant association between plasma homocysteine and the risk of ischemic cardiovascular disease was found. The CVD risk was 16% higher in individuals who were carriers of genotype MTHFR-677 TT than in those with genotype CC (wild type); the difference in plasma homocysteine between the genotypes TT and CC was 1.9 μmol/L. The 14 randomized trials showed no significant reduction in CVD risk, despite a reduction in homocysteine by 3.3 μmol/L. However, in studies with low-dose aspirin use, homocysteine reduction was associated with a reduction in CVD risk by 7%. The CVD risk was not reduced in studies with high-dose aspirin intake. It is therefore concluded that folic acid would have a role in the primary prevention of CVD, when aspirin is not taken routinely, but not in secondary prevention /55/. For further meta analyses refer to References /565758/.

13.1.5.2 Neurotoxic effect

Homocysteine is able to compete with receptors for glutamate and aspartate in the central nervous system (CNS). As a result, hyperhomocysteinemia can have an excitotoxic effect on various N-methyl-D-aspartate (NMDA) glutamate receptor subtypes and, thus, contributes to an increased formation of hydroxyl radicals /59/ (Fig. 13.1-7 – Mechanisms of homocysteine neurotoxicity). The activation of NMDA receptors is thought to promote intracellular calcium accumulation and a resulting release of cellular proteases and, possibly, apoptosis. This excitotoxic mechanism has been postulated for a number of neurodegenerative and psychiatric diseases, from metabolic and toxic encephalopathy to schizophrenia. It has also been shown that both NMDA receptors and group I metabotropic glutamate receptors are involved in homocysteine-induced neurotoxicity.

13.1.5.3 Methylation capacity defect

In addition to its effect as a vessel damaging agent, hyperhomocysteinemia is also an indicator for defective methylation capacity (elevated SAH and/or low SAM). Intact interaction between folate and vitamin B12, as occurs with normal homocysteine levels, is very important for DNA synthesis and various methylation reactions (methylation of DNA, RNA, myelin, phospholipids, receptors and neurotransmitters). It has been reported that patients with neuropsychiatric disorders had reduced concentrations of folate /38/ and SAM as the active methyl group donor in cerebrospinal fluid /60/. The SAM levels post-mortem were very low in the brain and also in the cerebrospinal fluid of patients with Alzheimer’s dementia.

In Alzheimer’s disease, changes in the SAM status are considered to be important for the expression of presenilin, a key factor for β-amyloid formation, and for the formation of phosphorylated tau protein) /61/.

Refer to Fig. 13.1-8 – Correlation between SAH and the SAM/SAH ratio in cerebrospinal fluid (CSF) with β-amyloid 1–42 concentration in the CSF in dementia-free patients.

The concentrations of phosphorylated tau protein in cerebrospinal fluid are positively and age-independently associated with the SAH concentration in cerebrospinal fluid (Fig. 13.1-9 – Concentrations of SAH in cerebrospinal fluid (CSF) in relation to P-tau 181 concentrations in the CSF observed in three age groups/62/.

It is suggested that the over expression of the gene is linked to hypo methylation of the gene’s promotor region. The synthesis and catabolism of catecholamines and myelin formation can be modulated by reduced SAM concentration. Patients with depression and high plasma homocysteine had low concentrations of monoamine metabolites in the cerebrospinal fluid, which points to defective turnover of serotonin, dopamine and noradrenaline. Elevated homocysteine concentrations, but low methionine and tryptophan, the precursor of serotonin, have been reported in patients with Alzheimer’s disease.

Normal folate metabolism is also important for the synthesis of sufficient tetrahydrobiopterin, a cofactor determining the speed of serotonin and dopamine synthesis. Low vitamin B12 and folate lead to decreased formation of tetrahydrobiopterin in the CNS /63/. Elevated homocysteine levels and markedly low methionine and tryptophan concentrations were found in patients with early stage of Alzheimer’s disease.

13.1.5.4 Atherothrombotic effect

The homocysteine metabolism in cardiovascular cells is limited to folate- and vitamin B12-dependent remethylation. This conclusion is based on the fact that no transsulfuration has been proven in the endothelium cells of human vessels to date /64/. Because there is no irreversible degradation of homocysteine to cysteine, homocysteine synthesis can quickly exceed cellular export and cause specific cell injury up to cell death. More than any other organ, the cardiovascular system is especially sensitive to increases in homocysteine. Such increases can induce morphological changes of the vessel wall, stimulate inflammation, activate the endothelium and coagulation cascade and inhibit fibrinolysis. In general, hyperhomocysteinemia induces the impairment of antithrombotic endothelial function and creation of a pro coagulatory milieu /65/. Atherogenic mechanisms most frequently discussed in hyperhomocysteinemia include thickening of the vessel wall intima, increased platelet turnover, increased platelet activation with increased thromboxane synthesis, endothelial dysfunction, activated leukocytes, increased LDL oxidation, increased foam cell formation due to lipid deposits in the vessel wall and increased proliferation of smooth muscle cells.

The atherogenic process in hyperhomocysteinemia is mostly thought to be triggered by oxidative stress. The formation of homocysteine thiolactone, a highly reactive internal thioester, and inflammation are discussed as key mechanisms in this context.

References

1. McCully KS. Vascular pathology of homocysteinemia: implications for the pathogenesis of arteriosclerosis. Am J Pathol 1969; 56: 111–28.

2. Wald DS, Law M, Morris JK. Homocysteine and cardiovascular disease: evidence on causality from a meta-analysis. BMJ 2002; 325: 1202.

3. de Bree A, Verschuren WM, Kromhout D, Kluijtmans LA, Blom HJ. Homocysteine determinants and the evidence to what extent homocysteine determines the risk of coronary heart disease. Pharmacol Rev 2002; 54: 599–618.

4. Seshadri S, Wolf PA, Beiser AS et al. Association of plasma total homocysteine levels with subclinical brain injury: cerebral volumes, white matter hyperintensity, and silent brain infarcts at volumetric magnetic resonance imaging in the Framingham Offspring Study. Arch Neurol 2008; 65: 642–9.

5. Obeid R, Schorr H, Eckert R, Herrmann W. Vitamin B12 status in the elderly as judged by available biochemical markers. Clin Chem 2004; 50: 238–41.

6. Carmel R, Jacobsen DW. Homocysteine in health and disease. Cambridge: University Press, 2001.

7. Ueland PM, Holm PI, Hustad S. Betaine: a key modulator of one-carbon metabolism and homocysteine status. Clin Chem Lab Med 2005; 43: 1069–75.

8. Lever M, Slow S. The clinical significance of betaine, an osmolyte with a key role in methyl group metabolism. Clin Biochem 2010; 43: 732–44.

9. Selhub J. Homocysteine metabolism. Annu Rev Nutr 1999; 19: 217–46.

10. Finkelstein JD, Martin JJ. Methionine metabolism in mammals. Distribution of homocysteine between competing pathways. J Biol Chem 1984; 259: 9508–13.

11. 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 non-transformed, transformed, proliferating, and quiescent cells in culture. J Cell Physiol 1991; 146: 52–62.

12. Herrmann W, Schorr H, Obeid R, Makowski J, Fowler B, Kuhlmann MK. Disturbed homocysteine and methionine cycle intermediates s-adenosylhomocysteine and s-adenosylmethionine are related to degree of renal insufficiency in type 2 diabetes. Clin Chem 2005; 51: 891–7.

13. Mudd SH, Skovby F, Levy HL et al. The natural history of homocystinuria due to cystathionine beta-synthase deficiency. Am J Hum Genet 1985; 37: 1–31.

14. Geisel J, Zimbelmann I, Schorr H et al. Genetic defects as important factors for moderate hyperhomocysteinemia. Clin Chem Lab Med 2001; 39: 698–704.

15. Dettmers C, Schwarting A, Rao ML, Liappis N, Hartmann A. Homozygous and heterozygous homocystinuria. The risk factors of juvenile arteriosclerosis? Dtsch Med Wochenschr 1992; 117: 473–9.

16. Herrmann W, Obeid R, Schorr H, Zarzour W, Geisel J. Homocysteine, methylenetetrahydrofolate reductase C677T polymorphism and the B-vitamins: a facet of nature-nurture interplay. Clin Chem Lab Med 2003; 41: 547–53.

17. Dong X, Wu J, Liang P, Li J, Yuan L, Liu X. Methylenetetrahydrofolate reductase C677T and A1298C polymorphisms and gastric cancer: a metaanalysis. Arch Med Res 2010; 41: 125–33.

18. Liu H, Yang M, Li GM et al. The MTHFR C677T polymorphism contributes to an increased risk for vascular dementia: a metaanalysis. J Neurol Sci 2010; 294: 74–80.

19. Flynn A, Hirvonen T, Mensink GB et al. Intake of selected nutrients from foods, from fortification and from supplements in various European countries. Food Nutr Res 2009; 53.

20. Nationale Verzehrstudie II. www.bmelv.de/SharedDocs/Downloads/Ernaehrung/NVS_ErgebnisberichtTeil2.pdf?__blob=publicationFile.

21. Beitz R, Mensink GB, Fischer B, Thamm M. Vitamins – dietary intake and intake from dietary supplements in Germany. Eur J Clin Nutr 2002; 56: 539–45.

22. Herrmann W, Obeid R. Causes and early diagnosis of vitamin B12 deficiency. Dtsch Arztebl Int 2008; 105: 680–5.

23. Selhub J, Jacques PF, Wilson PW, Rush D, Rosenberg IH. Vitamin status and intake as primary determinants of homocysteinemia in an elderly population. JAMA 1993; 270: 2693–8.

24. Herrmann W, Schorr H, Geisel J, Riegel W. Homocysteine, cystathionine, methylmalonic acid and B-vitamins in patients with renal disease. Clin Chem Lab Med 2001; 39: 739–46.

25. Baigent C, Burbury K, Wheeler D. Premature cardiovascular disease in chronic renal failure. Lancet 2000; 356: 147–52.

26. Mallamaci F, Zoccali C, Tripepi G et al. Hyperhomocysteinemia predicts cardiovascular outcomes in hemodialysis patients. Kidney Int 2002; 61: 609–14.

27. Guttormsen AB, Ueland PM, Svarstad E, Refsum H. Kinetic basis of hyperhomocysteinemia in patients with chronic renal failure. Kidney Int 1997; 52: 495–502.

28. van Guldener C, Kulik W, Berger R et al. Homocysteine and methionine metabolism in ESRD: A stable isotope study. Kidney Int 1999; 56: 1064–71.

29. Obeid R, Kuhlmann MK, Kohler H, Herrmann W. Response of homocysteine, cystathionine, and methylmalonic acid to vitamin treatment in dialysis patients. Clin Chem 2005; 51: 196–201.

30. Kessler H, Bleich S, Falkai P, Supprian T. Homocysteine and dementia. Fortschr Neurol Psychiatr 2003; 71: 150–6.

31. Polyak Z, Stern F, Berner YN et al. Hyperhomocysteinemia and vitamin score: correlations with silent brain ischemic lesions and brain atrophy. Dement Geriatr Cogn Disord 2003; 16: 39–45.

32. Smith AD, Smith SM, de Jager CA et al. Homocysteine-lowering by B vitamins slows the rate of accelerated brain atrophy in mild cognitive impairment: a randomized controlled trial. PLoS ONE 2010; 5: e12244.

33. Vermeulen EG, Stehouwer CD, Valk J et al. Effect of homocysteine-lowering treatment with folic acid plus vitamin B on cerebrovascular atherosclerosis and white matter abnormalities as determined by MRA and MRI: a placebo-controlled, randomized trial. Eur J Clin Invest 2004; 34: 256–61.

34. Seshadri S, Beiser A, Selhub J et al. Plasma homocysteine as a risk factor for dementia and Alzheimer’s disease. N Engl J Med 2002; 346: 476–83.

35. Ravaglia G, Forti P, Maioli F et al. Homocysteine and folate as risk factors for dementia and Alzheimer disease. Am J Clin Nutr 2005; 82: 636–43.

36. Ravaglia G, Forti P, Maioli F et al. Homocysteine and cognitive function in healthy elderly community dwellers in Italy. Am J Clin Nutr 2003; 77: 668–73.

37. Bottiglieri T, Godfrey P, Flynn T, Carney MW, Toone BK, Reynolds EH. Cerebrospinal fluid S-adenosylmethionine in depression and dementia: effects of treatment with parenteral and oral S-adenosylmethionine. J Neurol Neurosurg Psychiatry 1990; 53: 1096–8.

38. Serot JM, Christmann D, Dubost T, Bene MC, Faure GC. CSF-folate levels are decreased in late-onset AD patients. J Neural Transm 2001; 108: 93–9.

39. Nurk E, Refsum H, Tell GS et al. Plasma total homocysteine and memory in the elderly: The Hordaland Homocysteine study. Ann Neurol 2005; 58: 847–57.

40. Duthie SJ, Whalley LJ, Collins AR, Leaper S, Berger K, Deary IJ. Homocysteine, B vitamin status, and cognitive function in the elderly. Am J Clin Nutr 2002; 75: 908–13.

41. Kado DM, Karlamangla AS, Huang MH et al. Homocysteine versus the vitamins folate, B6, and B12 as predictors of cognitive function and decline in older high-functioning adults: MacArthur Studies of Successful Aging. Am J Med 2005; 118: 161–7.

42. Wang HX, Wahlin A, Basun H, Fastbom J, Winblad B, Fratiglioni L. Vitamin B(12) and folate in relation to the development of Alzheimer’s disease. Neurology 2001; 56: 1188–94.

43. McCaddon A, Hudson P, Davies G, Hughes A, Williams JH, Wilkinson C. Homocysteine and cognitive decline in healthy elderly. Dement Geriatr Cogn Disord 2001; 12: 309–13.

44. Tighe P, Ward M, McNulty H et al. A dose-finding trial of the effect of long-term folic acid intervention: implications for food fortification policy. Am J Clin Nutr 2011; 93: 11–8.

45. Fohr IP, Prinz-Langenohl R, Bronstrup A et al. 5,10-Methylenetetrahydrofolate reductase genotype determines the plasma homocysteine-lowering effect of supplementation with 5-methyltetrahydrofolate or folic acid in healthy young women. Am J Clin Nutr 2002; 75: 275–82.

46. Wright AJ, Finglas PM, Dainty JR et al. Differential kinetic behavior and distribution for pteroylglutamic acid and reduced folates: a revised hypothesis of the primary site of PteGlu metabolism in humans. J Nutr 2005; 135: 619–23.

47. Venn BJ, Green TJ, Moser R, Mann JI. Comparison of the effect of low-dose supplementation with L-5-methyltetrahydrofolate or folic acid on plasma homocysteine: a randomized placebo-controlled study. Am J Clin Nutr 2003; 77: 658–62.

48. Homocysteine Studies Collaboration. Homocysteine and risk of ischemic heart disease and stroke: a meta-analysis. JAMA 2002; 288: 2015–22.

49. Peterson JC, Spence DJ. Vitamins and progression of atherosclerosis in hyperhomocysteinaemia. Lancet 1998; 351: 263.

50. Till U, Rohl P, Jentsch A et al. Decrease of carotid intima-media thickness in patients at risk to cerebral ischemia after supplementation with folic acid, Vitamins B6 and B12. Atherosclerosis 2005; 181: 131–5.

51. Bonaa KH, Njolstad I, Ueland PM et al. Homocysteine lowering and cardiovascular events after acute myocardial infarction. N Engl J Med 2006; 354: 1578–88.

52. Toole JF, Malinow MR, Chambless LE et al. Lowering homocysteine in patients with ischemic stroke to prevent recurrent stroke, myocardial infarction, and death: the Vitamin Intervention for Stroke Prevention (VISP) randomized controlled trial. JAMA 2004; 291: 565–75.

53. Lonn E, Yusuf S, Arnold MJ et al. Homocysteine lowering with folic acid and B vitamins in vascular disease. N Engl J Med 2006; 354: 1567–77.

54. Armitage JM, Bowman L, Clarke RJ et al. Effects of homocysteine-lowering with folic acid plus vitamin B12 vs. placebo on mortality and major morbidity in myocardial infarction survivors: a randomized trial. JAMA 2010; 303: 2486–94.

55. Wald DS, Morris JK, Wald NJ. Reconciling the evidence on serum homocysteine and ischaemic heart disease: a metaanalysis. PLoS ONE 2011; 6: e16473.

56. Wang X, Qin X, Demirtas H et al. Efficacy of folic acid supplementation in stroke prevention: a metaanalysis. Lancet 2007; 369: 1876–82.

57. Lee M, Hong KS, Chang SC, Saver JL. Efficacy of homocysteine-lowering therapy with folic acid in stroke prevention: a metaanalysis. Stroke 2010; 41: 1205–12.

58. Yang Q, Botto LD, Erickson JD et al. Improvement in stroke mortality in Canada and the United States, 1990 to 2002. Circulation 2006; 113: 1335–43.

59. Obeid R, Herrmann W. Mechanisms of homocysteine neurotoxicity in neurodegenerative diseases with special reference to dementia. FEBS Lett 2006; 580: 2994–3005.

60. Linnebank M, Popp J, Smulders Y et al. S-adenosylmethionine is decreased in the cerebrospinal fluid of patients with Alzheimer’s disease. Neurodegener Dis 2010; 7: 373–8.

61. Nicolia V, Fuso A, Cavallaro RA, Di LA, Scarpa S. B vitamin deficiency promotes tau phosphorylation through regulation of GSK3beta and PP2A. J Alzheimer Dis 2010; 19: 895–907.

62. Obeid R, Kasoha M, Knapp JP et al. Folate and methylation status in relation to phosphorylated tau protein(181P) and beta-amyloid(1–42) in cerebrospinal fluid. Clin Chem 2007; 53: 1129–36.

63. Bottiglieri T, Laundy M, Crellin R, Toone BK, Carney MW, Reynolds EH. Homocysteine, folate, methylation, and monoamine metabolism in depression. J Neurol Neurosurg Psychiatry 2000; 69: 228–32.

64. Chen P, Poddar R, Tipa EV et al. Homocysteine metabolism in cardiovascular cells and tissues: implications for hyperhomocysteinemia and cardiovascular disease. Adv Enzyme Regul 1999; 39: 93–109.

65. Stanger O, Weger M, Renner W, Konetschny R. Vascular dysfunction in hyperhomocyst(e)inemia. Implications for atherothrombotic disease. Clin Chem Lab Med 2001; 39: 725–33.

13.2 Homocysteine

Wolfgang Herrmann, Rima Obeid

13.2.1 Indication

Patients with atherosclerotic vascular and neurodegenerative disease.

Target groups for homocysteine determination:

  • Vascular disease: coronary heart disease, myocardial infarction, atherosclerosis of the carotid artery, peripheral arterial occlusion disease, atherosclerosis of the cerebral arteries, cerebral insult, venous thrombosis, pulmonary artery embolism
  • Risk groups for cardiovascular disease: positive family history, arterial hypertension, smoking, hyperlipidemia, renal insufficiency, diabetes, metabolic syndrome, adiposity
  • Risk groups for vitamin deficiency: elderly, vegetarians, inflammatory gastrointestinal disease (gastritis, malabsorption), alcohol abuse, one-sides eating habits, drugs
  • Other risk groups: dementia, cognitive impairment, depression, homo cystinuria, hypothyroidism, rheumatism, AIDS, pregnancy complications (preeclampsia, HELLP syndrome, neural tube defect, intrauterine growth retardation)
  • Prophylaxis: preconceptional prevention of neural tube defects, healthy individuals > 50 years.

13.2.2 Method of determination

High performance liquid chromatography (HPLC)

HPLC is the standard method for determining the homocysteine level in blood plasma or serum /1/. Because of their detection limit, simplicity and high throughput capacity, the HPLC method in combination with fluorescence detection (HPLC-FD) are widely used methods for aminothiol quantification /2/. There are considerable variations between the procedures for sample processing, chromatographic conditions and sample detection which limit the standardization of the method.

The applied methods can principally be divided into procedures using derivatization and those working without derivatization. The first procedure uses pre-column derivatization with fluorogenic reagents reacting with SH groups. Three derivatization reagents are commonly used: monobromobimane, 4-(aminosulfonyl)-7-fluoro-2,1,3-benzoxadiazole-4-sulfonate and ammonium 7-fluoro-2,1,3-benzoxadiazole-4-sulfonate, with 4-(aminosulfonyl)-7-fluoro-2,1,3-benzoxadiazole 4-sulfonate being the most suited. Electrochemical detection, which does not use sample derivatization, is attractive because of very easy sample processing, high methodical specificity, the possibility of auto injection with high sample throughput and the possibility of co-determining other thiols.

This procedure has the disadvantage that the flow cell and electrodes become easily contaminated. Amperometric (gold/mercury electrode) and coulometric (carbon electrode) methods are also used.

Gas-liquid chromatography-mass spectrometry (GC-MS)

Deuterated internal standard (D8-homocysteine) is added to the plasma samples, homocysteine is released due to reduction with 1,4-dithio-D,L-threitol and the sample is purified by solid phase extraction using column chromatography with an anion exchange resin. Homocysteine is then derivatized with N-methyl-N-(tert-butyldimethylsilyl)-trifluoroacetamide, and the butyldimethylsilyl derivatives are separated by means of capillary gas chromatography (using an automatic sample injection system) and identified and determined with a mass detector using the single ion monitoring (SIM) mode /3/.

Liquid chromatography-tandem mass spectrometry (LC-MS/MS)

The LC-MS/MS method is a novel method of homocysteine determination characterized by high specificity and sensitivity and good analytical precision. A daily throughput of up to 200 plasma samples can be handled and sample processing is less time consuming (40 samples in less than 1 h) /4/.

Capillary electrophoresis

This determination method uses laser-induced fluorescence detection. For sample processing, plasma samples are mixed with internal standard, reduced with tri-N-butylphosphine and then deproteinized (trichloroacetic acid) and derivatized (5-bromomethyl-fluorescein). Electrophoretic separation takes approximately 7 min. /5/.

Immunological methods

Fluorescence polarization immunoassay (FPIA). This method is based on the principle of reductive release of homocysteine, the conversion of homocysteine to S-adenosylhomocysteine (SAH) catalyzed by SAH hydrolase and subsequent immunological detection with an SAH antibody. Besides serving as fluorescence polarization immunoassay, this method can also be used as latex-enhanced nephelometric assay. Compared to the standard method HPLC, immunological assays also show good correlations and assay characteristics within a concentration range of 7 to 40 μmol/L and CVs of 3–5% /6/.

Enzymatic method

The enzymatic homocysteine determination methods are based on the reactions of the homocysteine/methionine cycle described in Section 13.1-1 – Introduction. The detection limit and specificity of the assays are increased by enzymatic recycling. As in the other methods, oxidized homocysteine is first of all reduced to homocysteine by tris-(2-carboxyethyl)-phosphine (TCEP).

In the homocysteine assay catalyzed by cystathionine-β-synthase (Fig. 13.2-1 – Principle of the homocysteine assay catalyzed by cystathionine-β-synthase), homocysteine and serine are condensed to form cystathionine. Cystathionine is then converted by cystathionine-β-lyase to form homocysteine and pyruvate. The reduction of NADH absorption at 340 nm allows the calculation of the amount of pyruvate present and converted to lactate. In a different enzymatic homocysteine assay, homocysteine is not measured directly, but in a cyclical enzymatic assay method the co-substrate conversion product is measured instead of the co-substrate itself. In a reaction with S-adenosylmethionine (SAM) catalyzed by homocysteine methyl transferase (HMTase), homocysteine is converted to SAH. SAH is then hydrolytically cleaved by SAH-hydrolase to form homocysteine and adenosine. The adenosine formed in this cyclic reaction is measured by converting the adenosine by means of adenosine desaminase to inosine and ammonia. Glutamate is formed by ammonia and ketoglutarate in an NADH-dependent reaction catalyzed by glutamate dehydrogenase (GLD) and its reduction of absorption is measured at 340 nm based on NAD formation. The change in absorption at 340 nm is correlated to the homocysteine concentration in the sample.

13.2.3 Specimen

EDTA fasting plasma taken from EDTA blood after 12 h fasting. After sampling, cool the EDTA blood on ice, centrifuge within 30 min. and separate the plasma.

13.2.3.1 Oral methionine load

In this provocation assay, a repeat measurement of homocysteine is performed after 4 and 6 hours following oral methionine administration (Tab. 13.1-1 –Homocysteine level in fasting plasma following oral methionine load):

  • Collection of EDTA fasting blood for determining the homocysteine baseline concentration (store cool, separate plasma within 30 min.)
  • Oral administration of L-methionine (0.1 g/kg of body weight) in 200 mL of fruit juice and a small, methionine-poor snack
  • Collect EDTA blood again 4 and/or 6 hours following methionine administration to determine the post-methionine load concentration.

The homocysteine level primarily reflects the cystathionine-β-synthetase activity and/or availability of vitamin B6 /7/. By contrast, the fasting homocysteine levels are not sensitive for vitamin B6 deficiency. Within a fasting homocysteine range of 12–15 μmol/L, a high number of individuals yield pathological findings following methionine loading (above 38 μmol/L) /8/.

13.2.4 Reference interval

It does not make sense to specify a reference interval because a graded relation has been found between risk increase and/or the manifestation of cardiovascular and other degenerative disease and homocysteine concentrations already above 10 μmol/L (dose-response relation) /12/. An increase in homocysteine concentration by 1 μmol/L, the risk increases by 6–7% /13/. Risk intervals considering the differentiation by prophylactic and therapeutic aspects have been defined for practical use. Tab. 13.1-1 – Homocysteine level in fasting plasma following oral methionine load shows a concentration-dependent classification of hyperhomocysteinemias.

13.2.5 Clinical significance

If moderately elevated homocysteine concentration is detected (12–30 μmol/L), a second analysis is recommended after 4–6 weeks. The intraindividual variability of homocysteine is generally very low. Repeated measurements in healthy individuals after 6–18 months show very good conformity with the baseline values, with an individual, non-significant deviation of only 0.85–1.2 μmol/L /14/. Despite the low variability, repeated measurements in the decisive interval can improve diagnosis.

Moderate hyperhomocysteinemia (above 12 μmol/L) has been reported in about 10% of the general population, up to 40% of patients with vascular disease and about 50% of the elderly above 65 years of age. Synergistic interactions between homocysteine and additional risk factors (smoking, arterial hypertension, diabetes, hyperlipidemia) result in a proportional increase in total risk. Consequently, it is especially important to identify individuals at high risk of degenerative vascular disease.

13.2.5.1 Treatment goals and options

Prophylaxis

Supplementation remains a recommended option for prophylaxis. Doses within the sense of prophylactic measures are listed in Fig. 13.2-2 – Decision tree for the diagnosis and prophylaxis/therapy in hyperhomocysteinemia. Moreover, dietary measures such as conversion to a vitamin-rich diet are recommended.

Therapy

In patients with manifest vascular disease, stroke, patients with neurodegenerative disease or cognitive impairment, the goal should be to reach a plasma homocysteine level below 10 μmol/L /12/. Homocysteine levels above 12 μmol/L, vitamin deficiency as well as functional disorders of the kidneys and thyroid should be excluded as causes of elevated homocysteine. The findings should be interpreted due to consideration of the influencing factors. See

For example, a change of drugs, changed dose or treatment of hypothyroidism can result in marked lowering of homocysteine.

Recommendations for vitamin supplementation

The extent to which the homocysteine concentration is lowered by vitamin supplementation depends on the baseline concentration. Daily administration of 0.2–5 mg of folic acid can achieve a reduction by 16–39% /16/. Additional administration of vitamin B12 is to avoid relative folate deficiency (i.e., in support of folate utilization).

Therapy with folic acid alone should not be given (especially not in the long term), but only in combination with vitamin B12. Vitamin B6 has little influence on the fasting homocysteine concentration, but is an important cofactor in catabolic transsulfuration and should, therefore, also be taken into account in B vitamin supplementation.

If homocysteine concentration is moderately elevated, daily vitamin supplementation of 0.2–0.8 mg of folic acid, 3–30 μg of vitamin B12 (at least 100 μg in older patients due to malabsorption) and 2–20 mg of vitamin B6 should be started. If a reduction into the range below 10 μmol/L of homocysteine is achieved, the homocysteine concentration can be checked every 1–2 years.

Other causes of increased vitamin requirement

If the decrease in homocysteine level is not satisfactory, other causes must be searched for. For example, renal and thyroid dysfunctions can be present besides vitamin deficiency. It should be noted that a vitamin B12 concentration within the lower reference range does not exclude intracellular B12 deficiency /14/. Genetic mutations of enzymes involved in metabolism, for example the most widely known 5,10 methylene tetrahydrofolate reductase (MTHFR) 677 C>T polymorphism, can also cause higher vitamin requirement.

For other possible causes, see

Supplementation in renal dysfunction and enzyme deficiency

Very high vitamin doses (possibly also daily administration of 3 g of betaine) can be required in manifest renal dysfunction (insufficiency, dialysis), but in many cases it is not possible to reach normal concentrations. Homocysteine levels above 30 μmol/L detected in patients with no renal dysfunction can be caused by congenital enzyme deficiency, the prevalence of which has been underestimated to date. If a primary therapy attempt with pharmacological doses (e.g., 5–15 mg of folic acid, 1 mg of vitamin B12 and more than 20 mg of vitamin B6) fails, the patient should be referred to a specialist.

13.2.6 Comments and problems

Specimen

Plasma is to be preferred to serum. Sampling tubes other than those with added EDTA (e.g., 2-deaza-adenosine as stabilizer) are used in some cases to extend the period until centrifugation without significant change in homocysteine concentration. It is advisable for better comparability to pursue the principle of immediate separation by centrifugation or short-term cooled storage. On the other hand, 2-deaza-adenosine is not suitable as stabilizer for immunological assays.

There have been commercial efforts to use stabilizers for avoiding immediate centrifugation of the blood. Some manufacturers use SAH hydrolase inhibitors that inhibit the release of homocysteine from SAH. However, an increase in SAH yields incorrect homocysteine concentrations in immunological methods based on SAH determination /18/. In citrate monovette or primavette samples, the plasma homocysteine did not change significantly at 4 °C after 24 h and remained unchanged in the primavette at room temperature, but the SAH increased continuously /18/.

Sample processing after blood collection

After blood collection, homocysteine is generated in the blood cells and continuously released into the plasma, causing a gradual increase in plasma homocysteine concentration (5–15% per hour at room temperature) /18/. An increase in homocysteine can occur in serum samples due to cellular release during coagulation. The blood sample (typically EDTA blood) should always be stored in a cool place immediately after collection and, within 30 min, be centrifuged at 2,000 g for 10 min in a cooling centrifuge.

Method of determination

As a rule, total homocysteine is measured in plasma after a reduction step as the total of bound and free homocysteine. Therefore, the result does not take quantitative changes of the fractions into account.

Enzymatic assay: enzymatic homocysteine assays show good correlation with the HPLC method and the immunological assay over a large concentration range /6/. The coefficients of variation (day to day) reported for human sample specimens are below 7% for the low and high concentration ranges. Compared to the HPLC or immunological methods, the absolute values are slightly higher (up to 7% on average). The precision of the enzymatic assays verified against standard reference material has not always been satisfactory /6/. Moderately severe lipemia, icteric samples or hemolysis can affect the results /6/. Enzymatic homocysteine measurement in the plasma of patients with renal disease or patients with liver cirrhosis can be affected by interferences resulting in higher measurement results. In enzymatic assays based on the conversion of cystathionine to homocysteine and pyruvate, interference can be caused by elevated cystathionine levels in renal patients. In enzymatic homocysteine assays based on the conversion of ammonia to ketoglutarate in the indicator reaction, falsely high homocysteine concentrations can be measured in plasma with elevated ammonia concentrations (liver cirrhosis).

Influence factors

Moderately severe lipemia, hyperbilirubinemia or hemolysis can affect the assay quality /6/. Besides elevated homocysteine, elevated SAH can be measured in the plasma of patients with renal insufficiency. However, this error is not significant due to the markedly elevated homocysteine levels in these patients.

Other body fluids

Although no systematic analyses on the homocysteine concentration in other body fluids are available, smaller-scale studies provide information on homocysteine concentrations in other body fluids. Tab. 13.2-1 – Homocysteine concentrations in other body fluids shows selected information in this context. The homocysteine level in cerebrospinal fluid (CSF), determined using LC-MS/MS or HPLC, is about hundredfold lower than in plasma. CSF homocysteine is negatively correlated with serum and CSF folate, but positively correlated with CSF SAH and age. CSF and serum homocysteine are also correlated. CSF homocysteine has been reported to be higher in patients with Alzheimer’s dementia than in patients without neurological disease.

In a study, homocysteine levels in amniotic fluid showed comparable levels than in plasma (Tab. 13.2-1 – Homocysteine concentrations in other body fluids/10/. It is reported that homocysteine concentrations in amniotic fluid are higher in small for gestational age newborns than in mature newborns. The homocysteine concentration in amniotic fluid correlates with fetal gestational age, femur length and biparietal diameter.

Determinations of the homocysteine concentration in ejaculate within the scope of research on in vitro fertilization showed that the homocysteine levels in the ejaculate were similar in fertile and idiopathic sub fertile men, but lower in sub fertile men /11/.

Drugs

Drug-induced alterations of homocysteine levels are to be taken into account (Tab. 13.1-2 – Causes of homocysteine changes). For example, elevated homocysteine can be caused by therapy with drugs containing SAM. Moreover, drugs such as methotrexate, carbamazepine, phenytoin, NO, anticonvulsives or 6-azauridine triacetate can affect the patient’s homocysteine metabolism causing elevated homocysteine concentrations.

Stability

If not immediately centrifuged, the blood can be stored on ice or in the icebox for up to 1 h after collection. Without quick centrifugation and separation of the blood cells, the homocysteine increases rapidly as a function of temperature and time, resulting in falsely elevated measurement results. After centrifugation, homocysteine in plasma is stable: for 24 h at room temperature, for up to one week in the icebox (2–8 °C), for months at –20 °C, for up to 10 years at –70 °C /19/.

13.2.7 S-Adenosyl homocysteine (SAH), S-Adenosyl homocyteine (SAM)

SAH is formed via de methylation of SAM and is a potent inhibitor for many transmethylation reactions. SAM (activated methionine) plays an important role in intermediary metabolism as universal methyl group donor. It has been pointed out in recent publications that SAH might be a better biomarker for the development of Alzheimer’s dementia than homocysteine /20/. In a study on Alzheimer patients, SAH in CSF was reported to correlate significantly with phosphorylated tau protein as the key protein for the development of Alzheimer’s dementia, whereas no such association was found for homocysteine /21/. It was shown that biomarkers of neurodegeneration (amyloid precursor protein (APP) and α-synuclein) are correlated to biomarkers of methylation (SAH and SAM) in patients with Parkinson’s disease /22/. Better cognitive function was related to higher methylation potential (SAM/SAH ratio) /22/. It has been suggested that hypo methylation as an essential pathomechanism in hyperhomocysteinemia might be better reflected by SAH or the SAM/SAH ratio than by homocysteine.

The fast and reliable stable-isotope dilution ultra performance liquid chromatography-tandem mass spectrometry (UPLC-MS/MS) method has been developed for the simultaneous quantification of SAM and SAH in body fluids /23/. The method comprises a phenylboronic acid-containing solid-phase extraction procedure, serving for binding and clean-up of SAM and SAH. After extraction, SAM and SAH are separated using UPLC (Acquity UPLC BEH C18 column) and quantified using the tandem mass spectrometer.

The pre analytical procedures are to be followed precisely because SAM is unstable and degraded to form SAH. The EDTA blood collected for SAM and SAH determination should (as for homocysteine) be stored in a cool place until centrifugation, centrifuged within 30 min., and the plasma should be separated immediately and acetic acid be added for deproteination (1 mL of plasma + 0.1 mL of 1 N acetic acid). After thorough mixing, the mixture should be centrifuged again and the supernatant be stored at –70 °C until analysis. Plasma concentrations in normal individuals were 85.5 ± 11.1 nmol/L for SAM and 13.3 ± 5.0 nmol/L for SAH. The UPLC-MS/MS method is highly sensitive and specific and provides good precision and accuracy.

13.2.8 Pathophysiology

Based on different quantification approaches, up to 25% of cardiovascular events might be prevented by lowering elevated homocysteine concentrations /24/. Besides its significance as a risk factor with additional prognostic function, homocysteine is a sensitive diagnostic biomarker for folate, vitamin B12 and vitamin B6 deficiency /25/. Moreover, the determination of homocysteine is suited for documenting successful therapy following vitamin supplementation. In addition to their function as co factors for the enzymes involved in homocysteine metabolism, the vitamins B12, B6 and folic acid have important, independent properties. Folic acid and vitamin B6 deficiency are independent risk factors for cardiovascular disease.

Low 5-methyl-THF

Low folate concentration, a mutation in the MTHFR gene or low vitamin B12 concentration can cause impaired synthesis of 5-methyl-THF resulting in impaired remethylation of homocysteine to methionine /26/. It has been attempted for compensation to increasingly catabolize homocysteine using transsulfuration. However, the increased amount of formed homocysteine cannot be converted by transsulfuration because the SAM concentration is too low to activate the catabolic metabolic pathway. The increase in homocysteine is additionally enhanced by the low 5-methyl-THF concentration that does not inhibit the de methylation of SAM and thus contributes to further lowering of SAM.

Elevated 5-methyl-THF

Vitamin B12 or methionine synthase deficiency inhibit the remethylation of homocysteine to methionine /27/. The resulting elevated 5-methyl-THF inhibits the de methylation of SAM, and elevated SAM concentrations activate cystathionine-β-synthetase. Hyperhomocysteinemia in vitamin B12 deficiency is slightly less pronounced than in folate deficiency because the homocysteine catabolism in vitamin B12 deficiency is more active than in folate deficiency.

Chronic renal insufficiency

Premature atherosclerotic vascular disease in renal patients are partly explained with the high prevalence of hyperhomocysteinemia. The median homocysteine level in patients with hemodialysis and peritoneal dialysis is approximately 30 μmol/L /28/. In many cases, kidney transplantees also have moderately elevated homocysteine (median: 20 μmol/L). Impaired remethylation of homocysteine to methionine (only partly reversible by vitamin supplementation) has been discussed as the main cause of the high prevalence of hyperhomocysteinemia in renal patients (about 70% of transplantees and about 90% of patients with hemodialysis) /29/. Besides renal dysfunction, renal patients also have substantial B vitamin deficiency (B12, B6, folate) /28/.

Moderately impaired transsulfuration

Transsulfuration is only moderately impaired in vitamin B6 deficiency or in the presence of a heterozygous defect in the Cystathionine-β-synthetase gene /26/. In combination with non-impaired remethylation, the residual transsulfuration activity prevents hyperhomocysteinemia as long as the homocysteine load remains low. However, pronounced hyperhomocysteinemia develops following food intake or after a methionine load test with existing vitamin B6 deficiency. Therefore, the methionine loading test is primarily used for identifying individuals with impaired transsulfuration, who have a normal or borderline homocysteine level under fasting conditions.

References

1. McMenamin ME, Himmelfarb J, Nolin TD. Simultaneous analysis of multiple aminothiols in human plasma by high performance liquid chromatography with fluorescence detection. J Chromatogr B Analyt Technol Biomed Life Sci 2009; 877: 3274–81.

2. Ducros V, Demuth K, Sauvant MP et al. Methods for homocysteine analysis and biological relevance of the results. J Chromatogr B Analyt Technol Biomed Life Sci 2002; 781: 207–26.

3. 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 1988; 81: 466–74.

4. Arndt T, Guessregen B, Hohl A, Heicke B. Total plasma homocysteine measured by liquid chromatography-tandem mass spectrometry with use of 96-well plates. Clin Chem 2004; 50: 755–7.

5. Vecchione G, Margaglione M, Grandone E et al. Determining sulfur-containing amino acids by capillary electrophoresis: a fast novel method for total homocyst(e)ine human plasma. Electrophoresis 1999; 20: 569–74.

6. La’ulu SL, Rawlins ML, Pfeiffer CM, Zhang M, Roberts WL. Performance characteristics of six homocysteine assays. Am J Clin Pathol 2008; 130: 969–75.

7. Ubbink JB, van der Merwe A, Delport R et al. The effect of a subnormal vitamin B-6 status on homocysteine metabolism. J Clin Invest 1996; 98: 177–84.

8. Graham IM, Daly LE, Refsum HM et al. Plasma homocysteine as a risk factor for vascular disease. The European Concerted Action Project. JAMA 1997; 277: 1775–81.

9. Obeid R, Kostopoulos P, Knapp JP et al. Biomarkers of folate and vitamin B12 are related in blood and cerebrospinal fluid. Clin Chem 2007; 53: 326–33.

10. Grandone E, Colaizzo D, Vecchione G et al. Homocysteine levels in amniotic fluid. Relationship with birth-weight. Thromb Haemost 2006; 95: 625–8.

11. Ebisch IM, Peters WH, Thomas CM, Wetzels AM, Peer PG, Steegers-Theunissen RP. Homocysteine, glutathione and related thiols affect fertility parameters in the (sub)fertile couple. Hum Reprod 2006; 21: 1725–33.

12. Nygard O, Nordrehaug JE, Refsum H, Ueland PM, Farstad M, Vollset SE. Plasma homocysteine levels and mortality in patients with coronary artery disease. N Engl J Med 1997; 337: 230–6.

13. Bots ML, Launer LJ, Lindemans J et al. Homocysteine and short-term risk of myocardial infarction and stroke in the elderly: the Rotterdam Study. Arch Intern Med 1999; 159: 38–44.

14. Clarke R, Woodhouse P, Ulvik A et al. Variability and determinants of total homocysteine concentrations in plasma in an elderly population. Clin Chem 1998; 44: 102–7.

15. Herrmann W, Herrmann M, Obeid R. Hyperhomocysteinaemia: a critical review of old and new aspects. Curr Drug Metab 2007; 8: 17–31.

16. Homocysteine Lowering Trialists’ Collaboration. Lowering blood homocysteine with folic acid based supplements: meta-analysis of randomised trials. BMJ 1998; 316: 894–8.

17. Herrmann W, Schorr H, Bodis M et al. Role of homocysteine, cystathionine and methylmalonic acid measurement for diagnosis of vitamin deficiency in high-aged subjects. Eur J Clin Invest 2000; 30: 1083–9.

18. Hübner U, Schorr H, Eckert R, Geisel J, Herrmann W. Stability of plasma homocysteine, S-adenosylmethionine, and S-adenosylhomocysteine in EDTA, acidic citrate, and Primavette collection tubes. Clin Chem 2007; 53: 2217–8.

19. Israelson B, Brattstrom L, Refsum H. Homocystein in frozen plasma samples. A short cut to establish hyperhomocysteinemia as a risk factor for arteriosclerosis? Scand J Clin Lab Invest 1993; 53: 46–9.

20. Chang PY, Lu SC, Chen CH. S-adenosylhomocysteine: a better marker of the development of Alzheimer’s disease than homocysteine? J Alzheimers Dis 2010; 21: 65–6.

21. Popp J, Lewczuk P, Linnebank M et al. Homocysteine metabolism and cerebrospinal fluid markers for Alzheimer’s disease. J Alzheimers Dis 2009; 18: 819–28.

22. Obeid R, Schadt A, Dillmann U, Kostopoulos P, Fassbender K, Herrmann W. Methylation status and neurodegenerative markers in Parkinson disease. Clin Chem 2009; 55: 1852–60.

23. Kirsch SH, Knapp JP, Geisel J, Herrmann W, Obeid R. Simultaneous quantification of S-adenosyl methionine and S-adenosyl homocysteine in human plasma by stable-isotope dilution ultra performance liquid chromatography tandem mass spectrometry. J Chromatogr B Analyt Technol Biomed Life Sci 2009; 877: 3865–70.

24. Wald DS, Law M, Morris JK. Homocysteine and cardiovascular disease: evidence on causality from a meta-analysis. BMJ 2002; 325: 1202.

25. Selhub J, Jacques PF, Wilson PW, Rush D, Rosenberg IH. Vitamin status and intake as primary determinants of homocysteinemia in an elderly population. JAMA 1993; 270: 2693–8.

26. Kluijtmans LA, Young IS, Boreham CA et al. Genetic and nutritional factors contributing to hyperhomocysteinemia in young adults. Blood 2003; 101: 2483–8.

27. Selhub J. Homocysteine metabolism. Annu Rev Nutr 1999; 19: 217–46.

28. Herrmann W, Schorr H, Geisel J, Riegel W. Homocysteine, cystathionine, methylmalonic acid and B-vitamins in patients with renal disease. Clin Chem Lab Med 2001; 39: 739–46.

29. Moelby L, Rasmussen K, Ring T, Nielsen G. Relationship between methylmalonic acid and cobalamin in uremia. Kidney Int 2000; 57: 265–73.

13.3 Vitamin B12, Holotranscobalamin

Wolfgang Herrmann, Rima Obeid

Vitamin B12, also called cobalamin is a cofactor of two enzymes methionine synthase and L-methyl-malonyl-coenzyme A mutase. The basic structure of cobalamin forms is shown in Fig. 13.3-1 – Basic structure of cobalamin forms.

Methionine synthase forms methionine from homocysteine and 5-methyltetrahydrofolate.

L-methylmalonyl-CoA-mutase catalyzes the formation of methyl malonic acid from succinyl-CoA.

The main dietary sources of vitamin B12 are liver, meat, fish, milk, cheese and eggs.

Vitamin B12 is necessary /1/:

  • Together with folate, for the maturation of erythropoietic cells. Deficiency in either of the two vitamins causes dyssynchrony between maturation of cytoplasm and that of nuclei and leads to macrocytosis, immature nuclei, and hyper segmentation in granulocytes.
  • For the development and initial myelination of the central nervous system as well as for maintenance of its normal function. Vitamin B12 deficiency can lead to demyelination of the cranial and thoracic dorsal and lateral columns of the spinal cord, occasional demyelination of cranial and peripheral nerves, and demyelination of white matter in the brain.

About 80% of serum vitamin B12 are bound to the transport protein haptocorrin and are, in this form, metabolically inactive. The residual proportion is bound to trans cobalamin and biologically active (holotranscobalamin).

13.3.1 Indication

Populations at risk for vitamin B12 deficiency are listed in Tab. 13.3-1 – Populations at risk for vitamin B12 deficiency.

13.3.2 Method of determination

Besides total vitamin B12, holotranscobalamin (holoTC), the active form of vitamin B12 is clinically relevant for determination of the vitamin B12 status.

13.3.2.1 Total vitamin B12

Chemiluminescence immunoassay

Competitive immunoassay with direct chemiluminescence detection. Vitamin B12 is released from its endogenous protein binding site by reduction in an alkaline environment. Free vitamin B12 in the patient sample competes with a defined amount of labeled vitamin B12, (e.g., labeled with acridiniumester, for binding sites of a limited amount of pure intrinsic factor). The intrinsic factor binds to paramagnetic particles as the solid phase. After magnetic separation, the chemiluminescence reaction is triggered. There is an inversely proportional relationship between the amount of vitamin B12 in the patient sample and the measured relative light units.

Solid phase radioimmunoassay

Reaction tubes coated with antiserum against vitamin B12/bovine serum albumin are used. A radio iodinated tyrosine methyl ester of vitamin B12 (B12-TME-125I) is incorporated as radioactive tracer. In the competitive assay, endogenous vitamin B12 and B12-TME-125I compete for the antibody binding sites on the vessel wall. The radioactivity measured after rinsing of the reaction tubes is reversely proportional to the vitamin B12 concentration in the sample /2/.

Competitive radio ligand assay

Endogenous vitamin B12 is extracted from the binding proteins in the serum and a defined amount of isotope-labeled vitamin B12 and vitamin B12-binding protein ( e.g., IF is added). Then, the free and the protein-bound vitamin B12 fractions are separated. The radioactivity in the free fraction is reversely proportional to the vitamin B12 concentration in the sample.

13.3.2.2 Holotranscobalamin (holoTC)

Holotranscobalamin= total vitamin B12 = TVB12

Enzyme immunoassay I

This method is based on the micro particle enzyme immunoassay (MEIA) principle using an analyzer for measurement. The assay uses micro particles coated with monoclonal anti-holoTC antibodies /4/. HoloTC antigen in the sample binds to the coated micro particles forming an antigen-antibody complex on the micro particles. Alkaline phosphatase (ALP) conjugate is added and, after a washing step, the substrate 4-methylumbelliferyl-phosphate is added to the matrix cell. The ALP-labeled conjugate catalyzes the cleavage of a phosphate group from the substrate. The resulting fluorescent product, 4-methylumbelliferone, is measured by the optical assembly of the MEIA. The signal is directly proportional to the holoTC concentration in the sample /5/.

In a different method, magnetic beads coated with vitamin B12 precipitate apo-transcolobalamin in the plasma. The holoTC present in the supernatant is measured by ELISA. ELISA uses immobilized anti-human TC (from rabbit) to capture and detect holoTC in the plasma with a biotinylated antibody in a horseradish peroxidase avidin-mediate color reaction. The color reaction is directly proportional to the plasma holoTC concentration /6/.

Radioimmunoassay

HoloTC is bound by a monoclonal anti-TC II antibody immobilized on magnetic micro spheres and then magnetically separated from the non-bound fraction remaining in the liquid phase. Cobalamin is released by reduction and extraction from the magnetically separated holoTC, a defined amount of 57Co-tracer is added, and labeled cobalamin and unlabeled cobalamin compete for binding sites on immobilized intrinsic factor. After the non-bound fractions have been washed out, the radioactivity is measured. The radioactivity is inversely proportional to the holoTC concentration in the sample /7/.

13.3.2.3 Methyl malonic acid

Gas-liquid chromatography-mass spectrometry (GC-MS)

Deuterated internal standard is added to the samples, and the samples are purified by solid phase extraction using column chromatography with an anion exchange resin. MMA is then derivatized with N-methyl-N-(t-butyldimethylsilyl)-trifluoroacetamide and the derivatives are separated by means of capillary gas chromatography and identified and quantified with a mass detector using the single ion monitoring (SIM) mode /89/. Besides MMA, homocysteine is also used as a B12 marker. However, MMA is more specific because the homocysteine concentration increases in B12 and folate deficiency.

13.3.2.4 Vitamin B12 resorption test

Principle: the so-called Schilling test analyzes urinary excretion of orally administered, radioactively labeled vitamin B12 /10/.

Preparation: any vitamin B12 supplementation is discontinued at least 2 days before the test. If the patient received radioactive substances in the recent past before the test, 24-h urine is collected and checked for radioactivity before administration of the capsule with 57Co-vitamin B12.

Test procedure: the bladder is emptied and the fasted patient is given a 57Co- or 58Co-vitamin B12 capsule with an activity of about 20 kBq. This is followed 2 hours later by intramuscular injection of 1 mg of vitamin B12. Urine is collected for 24 h and the urine volume is measured. The total activity excreted in urine is correlated to the orally administered dose.

Clinical significance: the percentage excretion of labeled vitamin B12, related to the administered dose, is evaluated. The Schilling test is used to clarify the causes of vitamin B12 deficiency. In case of low radioactivity excretion, an extended test can be added for differentiation between malabsorption and intrinsic factor deficiency. The patient is administered the 57Co-vitamin B12 capsule and 35 mg of intrinsic factor concentrate at the same time.

13.3.2.5 Deoxyuridine suppression test

The deoxyuridine suppression test provides objective evidence of impaired DNA synthesis in patients with megaloblastic anemia and can be used for differentiation between vitamin B12 deficiency and folate deficiency. The test analyzes the capacity of leukocytes/bone marrow cells for incorporation of 3[H]-thymidine into their DNA, both before and after adding the corresponding cofactor (vitamin B12 or folate). The rate of incorporation is compared with that of control cells. Normoblastic bone marrow cells incorporate less than 10% of 3[H]-thymidine and megaloblastic bone marrow cells incorporate more than 10% of the controls /11/.

13.3.3 Specimen

Vitamin B12, holoTC, MMA, homocysteine:

  • For the separate determination of individual biomarkers: 1 mL of serum or plasma, respectively
  • For the combined determination of holoTC, MMA and homocysteine: 2 mL of serum or plasma.

Schilling test: refer entire 24-hour urine sample without supplement to the laboratory.

Deoxyuridin suppression test: leukocytes or bone marrow cells.

13.3.4 Reference interval

Refer to

13.3.5 Clinical significance

The prevalence of vitamin B12 deficiency is high in at-risk populations with sometimes irreversible consequences. In patients with low holo TC concentration (< 400 pmol/L) physicians are not sure whether there is a deficiency of vitamin B12 and functional vitamin B12 (methylmalonic acid, MMA) is ordered. Commercially available assays evaluate functional vitamin B12 using the determination of MMA.

Laboratory evaluation on vitamin B12 deficiency:

  • The first step is to determine the vitamin B12 serum level. Falsely positive and falsely negative results (up to 50%) can be of disadvantage if the reference interval of a given test is used for assessment.
  • If vitamin B12 deficiency is suspected clinically at levels within the reference interval, the determination of holotranscobalamin or methylmalonic acid is recommended.

Refer to Fig. 13.3-2 – Algorithm for diagnosis of vitamin B12 deficiency.

13.3.5.1 Assessment of makers of vitamin B12 deficiency

Determination of total vitamin B12 is a first line marker, but provides limited diagnostic specificity and sensitivity, especially in individuals with vitamin B12 levels below 400 pmol/L /1213/. A total vitamin B12 level within the lower reference interval between 156–400 pmol/L suggests that vitamin B12 deficiency cannot be excluded /514/. It has been shown that individuals with vitamin B12 levels within the reference range (above 156 pmol/L) may have symptoms of vitamin B12 deficiency /1617/ and that individuals with normal vitamin B12 concentrations can have elevated MMA concentrations (above 300 nmol/L) and low holoTC concentrations indicating intracellular, metabolically manifest (functional) vitamin B12 deficiency /1318/. Conversely, normal MMA in low vitamin B12 concentration as falsely positive finding is possible.

Isolated low serum holoTC concentration is considered to be the earliest marker of vitamin B12 deficiency and indicates that the body has a low vitamin B12 supply and vitamin B12 stores are being depleted /13/. Consequently, low holoTC is indicative of a negative B12 balance. Clinical or hematological manifestations are not yet found in this stage. In metabolically manifested vitamin B12 deficiency [low holoTC, elevated MMA (and homocysteine)], there still may be no clinical symptoms. The diagnostic use of holoTC allows early treatment before irreversible neurological damage occurs.

HoloTC as metabolically active vitamin B12 fraction correlates significantly with MMA. Vitamin B12 , on the other hand, correlates significantly with holoTC within the reference interval, but less within the low concentration range /13/. Therefore, holoTC and MMA are better suited for the detection of vitamin B12 deficiency than total vitamin B12 /51419/.

In a second study /19/ patients with a Holo TC level < 300 pmol/L were investigated for MMA. A positive correlation was found between Holo TC level and MMA in 71% of cases. A disagreement was evaluated between Holo TC level und MMA in 18.8% of cases. The optimized cut-off value for MMA was 45.5 pmol/L. The diagnostic sensitivity for MMA was 66.7%, the diagnostic specificity 60%, the positive predivtive value 30.7% ,and the negative predictive value 88.9%.

13.3.5.2 Stages and early detection of vitamin B12 deficiency

The development of vitamin B12 deficiency passes through various stages /20/. The stage of store depletion in the plasma and cells (negative vitamin B12 balance) is followed by functional vitamin B12 deficiency, in which metabolic disorders occur indicated by elevated homocysteine and/or MMA.

Store depletion

Store depletion is solely reflected by the holoTC level, which is low in this stage.

Functional B12 deficiency

Because of depleted stores, the cell is no longer capable to have a balanced vitamin B12-dependent metabolism. Besides low holoTC and holohaptocorrin, the metabolic biomarkers homocysteine and MMA are elevated, where MMA is a sensitive, but not specific biomarker of functional vitamin B12 deficiency /12/. In addition, homocysteine is found to be elevated in folate or vitamin B6 deficiency.

Clinical B12 deficiency

The clinical manifestation stage is characterized by macrocytic anemia with elevated MCV, decreased MCHC and neutrophil hyper segmentation. Even persistent functional vitamin B12 deficiency does not necessarily result in clinically abnormal hematological manifestation. Vitamin B12 deficiency can cause neurological damage without the presence of classic hematological changes /20/. However, metabolic disorders primarily caused by vitamin B12 deficiency, such as hyperhomocysteinemia and hypo methylation, are considered to be important factors in the development of neurodegenerative diseases /21/. Early diagnosis of vitamin B12 deficiency is important not only because neurological damage is irreversible, but also because metabolic disorders can easily be corrected by adequate vitamin supplementation.

Early diagnosis is possible with modern biomarkers such as holoTC and MMA. However, a reliable diagnosis of vitamin B12 deficiency cannot be achieved based on individual biomarkers. The combined use of the biomarkers holoTC, MMA and homocysteine provides diagnostic validity.

In a study on adults with serum vitamin B12 ≤ 221 pmol/L, holoTC was determined in parallel /4/. 68% of the individuals in the lower normal range of vitamin B12 (139–221 pmol/L) showed holoTC ≤ 40 pmol/L. Only 40% of the group with pathologically low total vitamin B12 also had holoTC ≤ 40 pmol/L. In this study, a diagnostic sensitivity of 86% and a specificity of 66% were reported for holo TC (threshold 40 pmol/L).

False low holotranscobalamin levels in patients with normal vitamin B12 concentration support the presence of a homozygous single point mutation c.855T>A in exon 6 of the TCN2 gene, corresponding to a asparagine to lysine substitution in position 267 of the mature protein /24/.

13.3.5.3 Causes of vitamin B12 deficiency

As a coenzyme, vitamin B12 is involved in metabolic processes in hematopoiesis, the development of the nervous system and regeneration of mucus layers. It is directly interrelated with folate metabolism because methyl cobalamin catalyzes the transfer of the methyl group from 5-methyltetrahydrofolate to homocysteine for methionine synthesis. Except in vegetarians, who have a life-style related vitamin B12 deficiency, it is mostly caused by intestinal malabsorption or disturbed formation of intrinsic factor. The classic vitamin B12 deficiency is pernicious anemia initially manifested as fatigue, jaundice and palpitation.

Using more sensitive and more specific biomarkers such as MMA or holoTC, it has been shown that the prevalence of subclinical functional vitamin B12 deficiency in the population is higher than assumed to date /2223/. The prevalence of vitamin B12 deficiency in younger adults is reported 5 to 7% /23/. Functional B12 deficiency [elevated MMA and low holoTC (vitamin B12)] is common in older individuals and was diagnosed in elderly, healthy individuals above 65 years of age with a prevalence of 10–30% /222326/.

It has been reported that despite complying with the recommended dietary intake of vitamin B12 (> 2.4 μg/day) in most cases, elderly adults show a high prevalence of slightly abnormal vitamin B12 status (elevated MMA and homocysteine) attributed to non-dietary causes (such as malabsorption) /27/. It was shown in elderly patients from Strassburg with vitamin B12 deficiency that 53% had vitamin B12 malabsorption, 33% had pernicious anemia, only 2% had diet related B12 deficiency and 11% had B12 deficiency of unclarified etiology /28/. It should be noted that, because of the low recommended dietary allowance B12 dietary insufficiency in older people is underestimated.

Children of mothers with vitamin B12 deficiency can be born with a deficiency or, if fed exclusively with breast milk, can acquire the deficiency within the first 4–6 months. Typical symptoms of vitamin B12 deficiency are deficits in brain development, growth, general development, hypotension, lethargy, tremor, hyperexcitability and anemia. Evidence of reduced myelination and atrophy has been provided by brain imaging. Vitamin B12 substitution achieves rapid improvement /1/. However, permanent impairment is increasingly likely with longer lasting vitamin deficiency.

In vegetarians, the prevalence of abnormal vitamin B12, MMA or holoTC depends on the duration and strictness of the vegetarian diet /1332/. In a study on lacto vegetarians and lacto-ovo vegetarians, 63% of the individuals showed elevated MMA (above 271 nmol/L), 73% a holoTC concentration ≤ 35 pmol/L and 33% a homocysteine concentration above 12 μmol/L. In vegans, elevated MMA was detected in 86%, low holoTC in 90% and hyperhomocysteinemia in 55% of the individuals /13/.

13.3.5.4 Differential diagnosis of vitamin B12 deficiency

In order to clarify the causes of vitamin B12 deficiency confirmed by laboratory and clinical findings, further laboratory tests should be performed.

  • Antibodies against parietal cells: based on determination using ELISA in pernicious anemia, these antibodies have a diagnostic sensitivity of 80% at 50% specificity decreasing with age.
  • Antibodies against IF: in pernicious anemia, these antibodies have a diagnostic sensitivity of 40% at 90% specificity
  • Determination of gastrin: the determination of gastrin can help in many cases, if negative parietal and intrinsic factor antibody results are negative. High concentrations of gastrin indicate achlorhydria in pernicious anemia
  • Vitamin B12 resorption test (Schilling test): the test is helpful for the detection and differentiation of B12 malabsorption. Reduced urinary vitamin B12 excretion following oral administration of radioactive vitamin B12 is usually based on intrinsic factor deficiency caused by autoimmune destruction of the parietal cell function in pernicious anemia. Intrinsic factor deficiency is confirmed by normalization of the assay upon concurrent administration of intrinsic factor and radioactive vitamin B12. If the test still yields pathologic results, it can be repeated following treatment with antibiotics or vermicides to eliminate, for example, intestinal bacterial overgrowth or infestation with fish tapeworm. Continuously pathologic results are evidence of disease of terminal ileum (e.g., Crohn’s disease).

Diseases and defects associated with vitamin B12 deficiency are shown in Tab. 13.3-4 – Diseases and defects associated with vitamin B12 deficiency.

13.3.6 Comments and problems

Sample material

Serum can be stored in the dark for 1 day at 2–8 °C (longer storage at –20 °C). Hemolytic samples affect the test result and should not be used; lipemic samples are to be centrifuged at 13,000 × g prior to the test.

Reference interval

Because the reference intervals for the assays are different, the intervals specified for a given assay and manufacturer should be noted. The reference intervals for the assay published in the literature are only to a limited extent applicable to immunoassay findings. The reference intervals for adults shown in Tab. 13.3-2 and Tab. 13.3-3 refer to chemiluminescence immunoassay and those for children refer to radioimmunoassay.

The reference interval of the microbiological method differs from the information shown in Tab. 13.3-2 [adults 25–500 ng/L (18.4–500 pmol/L), immunoassay] /2/.

Diagnostic efficiency of vitamin B12, holoTC, MMA

The diagnostic sensitivity and specificity of vitamin B12 assays are not satisfactory. Values in the lower reference interval, in particular, are not conclusive /1335/. For example, about 25% of patients with vitamin B12 below 148 pmol/L (200 ng/L) show no clinical symptoms of vitamin B12 deficiency /36/. If MMA is used as biomarker of functional B12 deficiency, 90–95% of the patients with elevated MMA also have low vitamin B12 /34/.

In studies on vegetarians, the diagnostic efficiency of detecting vitamin B12 deficiency was about 40% higher when using MMA instead of vitamin B12 assay /13/. By contrast, diagnostic comparison yielded close agreement between MMA and holoTC.

In assays on serum samples with unconfirmed vitamin B12 deficiency, samples with normal creatinine tested for elevated MMA (≥ 300 nmol/L) showed a significantly larger area under the receiver operating characteristic (ROC) curve compared to total vitamin B12 (cutoff ≤ 156 pmol/L), 0.71 vs. 0.60 /14/. Accordingly, the diagnostic sensitivity and specificity were higher for holoTC than for total vitamin B12. In studies on younger, healthy individuals [median age 34 (21–68) years], low vitamin B12 levels were detected in 1% of the individuals, whereas holoTC was low in 11% and MMA was elevated in 5% /12/. Consequently, the prevalence of functional vitamin B12 deficiency was significantly higher than that of low vitamin B12 levels.

Vitamin B12 deficiency has been reported to increase significantly with age /22/. In elderly individuals [median age 82 (69–92) years], low vitamin B12 levels were found in 7%, but low holoTC concentrations in 21% and high MMA concentrations in 42% of the subjects /25/. Besides the elderly, vegetarians and patients with renal disease are particularly affected by a high prevalence of vitamin B12 deficiency /13/.

Individuals with increased vitamin requirement are another group at risk of B12 deficiency, for example pregnant and lactating women, patients with autoimmune disease or with HIV infection.

Moreover, B12 deficiency can also develop under intake of proton pump inhibitors /13/.

Examinations of gastrectomized patients yielded holoTC below 42 pmol/L in 25% compared to total vitamin B12 below 139 pmol/L in only 7.8% of the individuals /15/. Moreover, it was found that 44% of these patients with borderline total vitamin B12 (139–295 pmol/L) had low holoTC (below 42 pmol/L) /15/.

Limitations in renal patients

Discrepant findings between the metabolic biomarker MMA and holoTC and vitamin B12 have been reported in patients with chronic kidney disease. Markedly elevated MMA were found, while serum levels of holoTC or vitamin B12 were elevated /2530/. Refer to Fig. 13.3-3– Decrease in methylmalonic acid in patients with kidney disease after injection of vitamin B12.

The folate form distribution in women (pregnant women and control cohort), umbilical cord blood and older adults is shown in Tab. 13.4-5 – Folate forms in serum using UPLC-MS/MS.

Similar constellations have also been observed in older individuals with impaired renal function. In this case, increased concentrations of circulating vitamin B12 are necessary to meet the intracellular requirement of vitamin B12 /2530/. Consequently, the reference intervals for holoTC and vitamin B12 do not apply to renal patients. Moreover, the utilization of MMA as a metabolic biomarker of vitamin B12 in renal patients is only possible to a limited degree. In renal patients, the MMA concentration decreases as a function of increasing circulating vitamin B12 concentration until a vitamin B12 resistant level is reached. Beyond this level, further increase in vitamin B12 in plasma will not achieve a further reduction in MMA concentration /3038/. However, elevated MMA can also have other causes, such as intestinal bacterial overgrowth.

13.3.7 Pathophysiology

The term vitamin B12 (cobalamin, Cbl) summarizes porphyrin-like compounds consisting of four reduced pyrrole rings with a central cobalt atom (Fig. 13.3-1 – Basic structure of cobalamin forms). The corrin ring carries an α-axial and a β-axial ligand at the cobalt atom. Only cobalamines with an α-axial ligand consisting of 5,6-dimethylbenzimidazole have a vitamin function in the human body /37/. In medical literature, the term vitamin B12 summarizes all cobalamines with a biological effect in the human body.

Corrinoids are usually classified as primitive coenzymes. Substitution at the β ligand of the cobalt atom by cyanide creates cyanocobalamin (CNCbl), substitution by OH¯ creates hydroxo-Cbl, substitution by CH3 creates methyl-Cbl and substitution by 5-deoxyadenosyl creates 5’-deoxyadenosyl-Cbl (Fig. 13.3-1 – Basic structure of cobalamin forms). The final molecules of human biosynthesis are 5’-deoxyadenosyl-Cbl and methyl-Cbl12. In humans, only two enzymes are catalyzed by vitamin B12 -methionine synthase requiring methyl-Cbl as a cofactor, and L-methyl malonyl-CoA mutase, requiring adenosyl-Cbl as a cofactor. Methionine synthase catalyzes the re methylation of homocysteine to methionine, while L-methyl malonyl-CoA mutase catalyzes the isomerization of methyl malonyl-CoA to succinyl-CoA.

Various proteins are required for the transport and cellular uptake of vitamin B12: intrinsic factor, trans cobalamin (TC), haptocorrin (also known as R-binder protein) and the membrane-bound intrinsic factor and TC receptors (Fig. 13.3-4 – Transport and cellular uptake of vitamin B12).

In the stomach, vitamin B12 is released from food proteins by the action of gastric acid and mainly bound to haptocorrin. In the upper intestine, the haptocorrin-cobalamin complex is degraded by the action of pancreatic enzymes and alkaline pH environment and vitamin B12 is bound to a different protein, intrinsic factor (IF). In the lower ileum, the Cbl-IF complex is taken up into the cells via a special receptor on the surface of the enterocyte membrane. Vitamin B12 absorption via this receptor is limited to below 3 μg per meal. About 1% of the food vitamin B12 is absorbed independently of a receptor through passive diffusion.

Holotranscobalamine

In the enterocytes, the Cbl-IF complex is degraded and vitamin B12 is transferred to a third protein, TC. This complex, referred to as holoTC, enters the bloodstream via portal circulation. Cellular internalization is effected via the TC receptor present on all cells. After internalization, holoTC undergoes lysosomal hydrolyzation, releasing cobalamin, which in the form of methyl-Cbl or adenosyl-Cbl activates the corresponding enzymes intracellularly by catalyzation. Cbl is released to the cytosol as Cbl3+ and there reduced to Cbl2+. Cbl2+ binds to apomethionine synthase and catalyzes the activity of this enzyme.

HoloTC is metabolically active vitamin B12 with a relatively short biological half-life of about 1–2 hours. Only 10–30% of the vitamin B12 circulating in the plasma are bound to TC (Fig. 13.3-5 – Cellular uptake and biological activity of vitamin B12).

Approximately 80–90% of the plasma vitamin B12 are bound to haptocorrin, which is also known as TC I and has a biological half-life of 9–10 days. This fraction of little metabolic activity does not contribute to the supply of peripheral cells with vitamin B12 but is thought to be involved in transporting peripherally excessive vitamin B12 back to the liver. TC III, an R-binder protein of the granulocytes, is a small fraction with a similar metabolic function as TC I.

Moreover, MMA as a sensitive biomarker of intracellular, functional vitamin B12 deficiency plays a prominent role /3540/. In intracellular vitamin B12 deficiency, homocysteine increases because of the impaired methionine synthase reaction and MMA increases because of the impaired L-methyl malonyl-CoA mutase catalyzed reaction (Fig. 13.3-6 – Cobalamin-dependent biological reactions in the cytosol and mitochondria and in cobalamin deficiency). Homocysteine also increases in folate and vitamin B6 deficiency, while MMA is considered a specific and sensitive biomarker of functional vitamin B12 deficiency /40/.

Clinical effects of vitamin B12 deficiency

Because the human body has significant vitamin B12 stores, it takes years before under supply becomes clinically evident. Vitamin B12 deficiency generally passes through various stages (Fig. 13.3-7 – Cellular uptake, intracellular distribution and synthesis of cobalamin coenzymes). Besides its significance as a risk factor for cardiovascular and neurodegenerative disease, hyperhomocysteinemia in vitamin B12 deficiency indicates vitamin deficiency and hypo methylation, for example, of DNA, RNA, myelin, phospholipids or neurotransmitters. Hypo methylation is caused by reduced availability of S-adenosylmethionine (based on the vitamin B12 deficiency-induced inhibition of methionine synthase and reduced methionine) functioning as methyl donor for all methyltransferases. Hypo methylation of the basic myelin protein at position 107 (arginine), for example, can destabilize the protein and, thus, lead to the development of neuropathies. Funicular spinal disease is a common secondary neurological disease of Cbl deficiency. Psychiatric and neurological diseases, such as cognitive impairment, depression and dementia, can precede hematological anomalies by years or are even not detected.

Nevertheless, morphological changes in blood and bone marrow cells belong to the most important symptoms of vitamin B12 deficiency. Because of the high cell turnover rate, hematopoiesis quickly and sensitively responds to blocked nucleic acid metabolism. Megaloblastic anemia due to vitamin B12 deficiency develops based on defective DNA synthesis and resulting impaired nucleus maturing, while the cytoplasm develops normally. Cells with more or less strong morphological changes are detectable in the periphery (MCV above 110 fL, MCH above 40 pg). Studies have also shown that deficiency of vitamin B12 (low total vitamin B12 or holoTC) in the elderly is associated with significant atrophy of the brain /42/. Supplementation with B vitamins for two years in older adults with mild cognitive impairment clearly slowed the rate of brain atrophy. A higher rate of accelerated brain atrophy was associated with poorer cognitive performance /4344/.

References

1. Stabler, SB. Vitamin B12 deficiency. N Engl J Med 368; 2013: 149–60.

2. Endres DB, Painter K, Niswender GD. A solid-phase radioimmunoassay for vitamin B12 in serum, with use of radioiodinated tyrosine methyl ester of vitamin B12. Clin Chem 1978; 24: 460–5.

3. Hicks JM, Cook J, Godwin ID, Soldin SJ. Vitamin B12 and folate. Pediatric reference ranges. Arch Pathol Lab Med 1993; 17: 704–6.

4. Orning L, Rian A, Campbell A et al. Characterization of a monoclonal antibody with specificity for holo-transcobalamin. Nutr Metab (Lond) 2006; 3: 3.

5. Bamonti F, Moscato GA, Novembrino C et al. Determination of serum holotranscobalamin concentrations with the AxSYM active B(12) assay: cut-off point evaluation in the clinical laboratory. Clin Chem Lab Med 2010; 48: 249–53.

6. Nexo E, Christensen AL, Hvas AM, Petersen TE, Fedosov SN. Quantification of holo-transcobalamin, a marker of vitamin B12 deficiency. Clin Chem 2002; 48: 561–2.

7. Ulleland M, Eilertsen I, Quadros EV et al. Direct assay for cobalamin bound to transcobalamin (holo-transcobalamin) in serum. Clin Chem 2002; 48: 526–32.

8. Stabler SP, Lindenbaum J, Savage DG, Allen RH. Elevation of serum cystathionine levels in patients with cobalamin and folate deficiency. Blood 1993; 81: 3404–13.

9. Allen RH, Stabler SP, Savage DG, Lindenbaum J. Elevation of 2-methylcitric acid I and II levels in serum, urine, and cerebrospinal fluid of patients with cobalamin deficiency. Metabolism 1993; 42: 978–88.

10. Nickoloff E. Schilling test: physiologic basis for and use as a diagnostic test. Crit Rev Clin Lab Sci 1988; 26: 263–76.

11. Carmel R, Rasmussen K, Jacobsen DW, Green R. Comparison of the deoxyuridine suppression test with serum levels of methylmalonic acid and homocysteine in mild cobalamin deficiency. Br J Haematol 1996; 93: 311–8.

12. Loikas S, Lopponen M, Suominen P et al. RIA for serum holo-transcobalamin: method evaluation in the clinical laboratory and reference interval. Clin Chem 2003; 49: 455–62.

13. Herrmann W, Schorr H, Obeid R, Geisel J. Vitamin B-12 status, particularly holotranscobalamin II and methylmalonic acid concentrations, and hyperhomocysteinemia in vegetarians. Am J Clin Nutr 2003; 78: 131–6.

14. Obeid R, Herrmann W. Holotranscobalamin in laboratory diagnosis of cobalamin deficiency compared to total cobalamin and methylmalonic acid. Clin Chem Lab Med 2007; 45: 1746–50.

15. Lee YK, Kim HS, Kang HJ. Holotranscobalamin as an indicator of vitamin B12 deficiency in gastrectomized patients. Ann Clin Lab Sci 2009; 39: 361–6.

16. Lesho EP, Hyder A. Prevalence of subtle cobalamin deficiency. Arch Intern Med 1999; 159: 407.

17. Lindenbaum J, Healton EB, Savage DG et al. Neuropsychiatric disorders caused by cobalamin deficiency in the absence of anemia or macrocytosis. N Engl J Med 1988; 318: 1720–8.

18. Herrmann W, Obeid R. Cobalamin deficiency. In: Herrmann W, Obeid R, eds. Vitamins in the Prevention of Human Diseases. Berlin – New York: Walter de Gruyter 2011: 213–42.

19. Guillerne J, Feugray G, Muraine MQ. Preliminary evaluation of the diagnostic performance of Roche Elecsys active vitamin B12 versus total vitamin B12 deficiency screening. Sage Journals. doi: 10.1177/0004563223119457.

20. Herbert V. Staging vitamin B-12 (cobalamin) status in vegetarians. Am J Clin Nutr 1994; 59: 1213S–22S.

21. Obeid R, McCaddon A, Herrmann W. The role of hyperhomocysteinemia and B-vitamin deficiency in neurological and neuropsychiatric diseases. Clin Chem Lab Med 2007; 45: 1590–606.

22. Hughes CF, Ward M, Hoey L, McNulty H. Vitamin B12 and ageing: current issues and interaction with folate. Ann Clin Biochem 2013; 50: 315–29.

23. Herrmann W, Obeid R, Schorr H, Geisel J. The usefulness of holotranscobalamin in predicting vitamin B12 status in different clinical settings. Curr Drug Metab 2005; 6: 47–53.

24. Keller P, Rufener J, Schild C, Fedosov SN, Nissen PH, Nexo E. False low holotranscobalamin levels in a patient with novel TCN2 mutation. Clin Chem Lab Med 2016; 54: 1739–43.

25. Herrmann W, Obeid R, Schorr H, Geisel J. Functional vitamin B12 deficiency and determination of holotranscobalamin in populations at risk. Clin Chem Lab Med 2003; 41: 1478–88.

26. Herrmann W, Schorr H, Bodis M et al. Role of homocysteine, cystathionine and methylmalonic acid measurement for diagnosis of vitamin deficiency in high-aged subjects. Eur J Clin Invest 2000; 30: 1083–9.

27. Howard JM, Azen C, Jacobsen DW, Green R, Carmel R. Dietary intake of cobalamin in elderly people who have abnormal serum cobalamin, methylmalonic acid and homocysteine levels. Eur J Clin Nutr 1998; 52: 582–7.

28. Henoun LN, Noel E, Ben AM, Locatelli F, Blickle JF, Andres E. Cobalamin deficiency due to non-immune atrophic gastritis in elderly patients. A report of 25 cases. J Nutr Health Aging 2005; 9: 462.

29. Herrmann W, Schorr H, Geisel J, Riegel W. Homocysteine, cystathionine, methylmalonic acid and B-vitamins in patients with renal disease. Clin Chem Lab Med 2001; 39: 739–46.

30. Obeid R, Kuhlmann MK, Kohler H, Herrmann W. Response of homocysteine, cystathionine, and methylmalonic acid to vitamin treatment in dialysis patients. Clin Chem 2005; 51: 196–201.

31. Obeid R, Kuhlmann M, Kirsch CM, Herrmann W. Cellular uptake of vitamin B12 in patients with chronic renal failure. Nephron Clin Pract 2005; 99: c42–c48.

32. Obeid R, Geisel J, Schorr H, Hubner U, Herrmann W. The impact of vegetarianism on some haematological parameters. Eur J Haematol 2002; 69: 275–9.

33. Herbert V, Das KC. The role of vitamin B12 and folic acid in hemato- and other cell-poiesis. Vitam Horm 1976; 34: 1–30.

34. Rajan S, Wallace JI, Brodkin KI, Beresford SA, Allen RH, Stabler SP. Response of elevated methylmalonic acid to three dose levels of oral cobalamin in older adults. J Am Geriatr Soc 2002; 50: 1789–95.

35. Joosten E, van den Berg A, Riezler R et al. Metabolic evidence that deficiencies of vitamin B-12 (cobalamin), folate, and vitamin B-6 occur commonly in elderly people. Am J Clin Nutr 1993; 58: 468–76.

36. Stabler SP, Allen RH, Savage DG, Lindenbaum J. Clinical spectrum and diagnosis of cobalamin deficiency. Blood 1990; 76: 871–81.

37. Lindenbaum J, Rosenberg IH, Wilson PW, Stabler SP, Allen RH. Prevalence of cobalamin deficiency in the Framingham elderly population. Am J Clin Nutr 1994; 60: 2–11.

38. Moelby L, Rasmussen K, Ring T, Nielsen G. Relationship between methylmalonic acid and cobalamin in uremia. Kidney Int 2000; 57: 265–73.

39. Linnell JC, Bhatt HR. Inherited errors of cobalamin metabolism and their management. Baillieres Clin Haematol 1995; 8: 567–601.

40. Savage DG, Lindenbaum J, Stabler SP, Allen RH. Sensitivity of serum methylmalonic acid and total homocysteine determinations for diagnosing cobalamin and folate deficiencies. Am J Med 1994; 96: 239–46.

41. Rosenblatt DS, Fenton WA. Inherited disorders of folate and cobalamin transport and metabolism. In: Scriver CR, Beaudet AL, Sly WS, Valle D, eds. The metabolic and molecular bases of inherited disease. New York: Mcgraw-Hill Medical Publishing Division 2001: 3897–933.

42. Vogiatzoglou A, Refsum H, Johnston C et al. Vitamin B12 status and rate of brain volume loss in community-dwelling elderly. Neurology 2008; 71: 826–32.

43. Smith AD, Smith SM, de Jager CA et al. Homocysteine-lowering by B vitamins slows the rate of accelerated brain atrophy in mild cognitive impairment: a randomized controlled trial. PLoS ONE 2010; 5: e12244.

44. Andres E. Signs and symptoms of vitamin B12 (cobalamin) deficiency: a critical review of the literature. In: Herrmann W, Obeid R, eds. Vitamins in the Prevention of Human Diseases. Berlin – New York: Walter de Gruyter 2011: 242–53.

13.4 Folate (Vitamin B9)

Wolfgang Herrmann, Susanne H. Kirsch, Rima Obeid

Folate is the general term for the water-soluble vitamin B9 naturally found in foods such as leafy green vegetables, legumes, egg yolks, liver and some citrus fruits. This vitamin is essential for normal cell growth and replication /1/. The bio availability of naturally occurring folate is less than that of folic acid, a synthetic compound that is used as supplements and in fortified foods. The term folate, as used in this publication, encompasses folic acid, natural food folate and folate derivatives (Fig. 13.4-1 – Structure of folate derivatives).

13.4.1 Indication

Serum or erythrocyte folate should be determined in:

  • Anemia, especially megaloblastic anemia
  • Older adults
  • Folate malabsorption
  • Inflammatory bowel disease
  • Chronic alcoholism
  • Chronic liver disease
  • Complications during pregnancy
  • Previous complications during pregnancy (preeclampsia or HELLP syndrome)
  • Atherosclerotic vascular disease
  • Dementia, depression and cognitive impairment
  • Hyperhomocysteinemia

13.4.2 Method of determination

Commercially available ligand assays are used to detect metabolically active folates. Folate is determined in serum or – in specific problems – in the erythrocytes. Determination is predominantly based on ligand assays. Some assays allow the simultaneous determination of folate and vitamin B12. It is recommended to use liquid chromatography-tandem mass spectrometry (LC-MS/MS) methods for the quantification of individual folate forms, such as 5-methyl-THF in serum and erythrocytes.

The whole-blood folate concentration is determined by hemolyzing 1 part of EDTA-blood with 20 parts of 1% ascorbic acid solution and measuring the folate concentration in the hemolysate. The resulting value is to be multiplied with 21 and divided by the hematocrit value. For determining the erythrocyte folate concentration, the serum folate concentration is subtracted from the whole-blood folate and related to the hematocrit value of the sample:

Whole-blood folate – serum folate × [(100 – hematocrit)/hematocrit]

In simultaneous determination of folate and vitamin B12, the sample must be divided because ascorbic acid interferes with the determination of vitamin B12.

Competitive immunoassay

Principle: a release reagent (e.g., DTT) is added to the patient sample to separate the folates from their endogenous binding proteins. The free folates of the sample then compete with added folic acid labeled, for example, with biotin or acridinium ester, for a limited amount of folic acid-binding protein, labeled, for example, with ruthenium or biotin. The folic acid-binding protein loaded with folic acid/folate is separated and measured.

Liquid chromatography-tandem mass spectrometry (LC-MS/MS)

The LC-MS/MS methods allow the quantification of folic acid and reduced folates in the mono glutamate or poly glutamate form and folate degradation products, such as p-aminobenzoic acid, from small amounts of patient specimens /2/. An ultra performance liquid chromatography-tandem mass spectrometry (UPLC-MS/MS) method /3/ allows the simultaneous quantification of folic acid, 5-methyl-THF, 5-formyl-THF, 5,10-methenyl-THF and THF from serum or whole-blood hemolysates.

Principle: for determining the whole-blood folate concentration, 1 part of EDTA blood is hemolyzed with 10 parts of a 10 g/L ascorbic acid solution and 0.2% triton X-100 and incubated for conversion of the polyglutamates to monoglutamates by means of endogenous conjugases in blood. The erythrocyte folate is also calculated according to the formula above.

Serum and blood hemolysate samples are mixed with commercially available isotope-labeled (13C5) internal standards and processed by solid phase extraction (OASIS MAX). The sample is concentrated for better quantification of the minor folate forms. The concentration is determined in the acidic mobile phase using liquid chromatography, followed by detection in the mass spectrometer using positive electrospray ionization in the multiple reaction mode (MRM). Sample processing is fast and the duration of analysis per sample is 2.5 min.

13.4.3 Specimen

  • Serum: 1 mL
  • EDTA blood for folate determination in whole-blood hemolysate: 2 mL

For LC-MS/MS methods, 250 μL of serum or 200 μL of EDTA blood are required.

13.4.4 Reference interval

The serum folate concentration depends on the nutritional state. Consequently, reference intervals can only to a limited extent be applied to specific patient groups. Therefore, the values listed in Tab. 13.4-1 – Folate reference intervals for fasting serum are only guidance values, which are influenced by a given assay and by the folate status of the reference cohort.

Different reference ranges have been introduced for Europe and the United States because of the different eating habits (fortification of food with folic acid and widespread vitamin intake in the United States) (Tab. 13.4-2 – Serum and whole-blood folate reference intervals in the USA and in Europe).

13.4.5 Clinical significance

Folic acid fortification of stable food in the USA is seen to be the reason for different normal intervals of folate in US and Europe. Accordingly, folate deficiency cannot be excluded if folate concentrations are within the lower fraction of the reference interval.

13.4.5.1 Folate deficiency

Folate deficiency is present when the serum folate level is below 3.5 μg/L and below 250 μg/L in the erythrocytes /5/. However, the observed symptoms are mainly restricted to the stages of pre latent and latent folate deficiency, while the diagnosis of clinically manifest megaloblastic anemia is an exception (Tab. 13.4-3 – Stages of folate deficiency/5/.

The determination of erythrocyte folate provides information on the long-term folate status /6/. Because erythrocytic folate is less dependent on short-term nutritional effects, non-fasting samples can also be analyzed. However, the analysis of erythrocyte folate has a lower precision.

Folate deficiency at first leads to reduced urinary folate excretion, followed by a decrease in serum concentration and, after 3–4 weeks, by a decrease in erythrocyte folate concentration. Hyper segmentation of the neutrophil granulocytes is observed after 10–12 weeks, months later followed by thrombocytopenia, leukopenia and finally macrocytic anemia.

Megaloblastic anemia caused by folate deficiency cannot be distinguished from vitamin B12 deficiency in the blood count. Therefore, the presence of folate deficiency requires further diagnostic clarification as to whether vitamin B12 deficiency is also present. Other effects of folate deficiency: changes in the mucous layer of the oral cavity, diarrhea, impaired growth, reduced immune response, impaired fertility, congenital malformations, such as neural tube defect, and neurological and psychiatric disorders.

Because of its strong dependence on food quality, the folate status should be determined in two tests at intervals of several days. A level above 5.5 μg/L (12.5 nmol/L) most likely indicates adequate folate supply. Folate requirement is higher in individuals homozygous for 5,10-methylenetetrahydrofolate reductase (MTHFR) C677T polymorphism. Normal plasma homocysteine concentrations are found when the serum folate is above 15 nmol/L and the vitamin B12 status is adequate.

There are about 12–28 mg of folate in the human body, with approximately half being localized in the liver. About 10–90 μg of folate are excreted daily with the bile and practically completely reabsorbed. However, reabsorption is restricted in inflammatory bowel disease. The body’s reserves of folate are small (biological half-life ~100 days). In dietary folate withdrawal, the reserves are sufficient for about 3–4 weeks before the serum folate decreases and, after another 10–12 weeks, hyper segmentation of the neutrophil granulocytes is observed. Insufficient dietary intake and anti folate medication are the main causes of folate deficiency.

Folate deficiency is very common in Europe and promoted by lack of fresh fruit and vegetables. Good food sources for folate are, for example, green vegetables, cereal products, fruit, yeast and liver. Up to 90% of the folates can be lost in the processing of cereal products and other food /7/. Folates are also lost because folate is sensitive to heat, storage and light.

The recommendations of several specialist associations to eat five portions of fruit and vegetables a day (600–700 g) are hardly realizable /8/. The daily intake of folate from foods in Central Europe is currently between 230 and 280 μg (Fig. 13.4-2 – Folate intake in the population of various European countries/9/; accordingly, a substantial part of the normal population does not reach the required amount of folate through their natural diet alone /10/. The recommended daily allowance (RDA) of folate is 400 μg (600 μg for pregnant women) for dietary folate equivalents (DFE).

Because of the under supply, functional disorders are to be expected, such as hyperhomocysteinemia, morphological cell changes (hyper segmentation of neutrophil granulocytes), neural tube defects, neurological symptoms and anemia. A mean daily intake of about 400 μg of folate equivalents would optimize all folate-dependent metabolism and maintain a low homocysteine concentration.

The mean homocysteine concentration reaches a steady-state level of almost 11.0 μmol/L at a folate intake of 400 μg/d /11/ or a serum folate concentration of about 14 nmol/L /12/.

In illness, folate deficiency can develop from inadequate intake, losses, malabsorption or increased requirement. Risk groups and diseases associated with folic acid deficiency are shown in Tab. 13.4-4 – Risk groups and diseases associated with folate deficiency.

13.4.6 Comments and problems

Sample material

Serum is recommended for analysis. Because serum folate levels are strongly dependent on the last food intake, serum from fasting blood should be used for determination. The blood samples should be sufficiently coagulated before centrifugation. Concentrations measured in heparin anticoagulated blood are higher than in serum. EDTA plasma is not recommended because it turned out to be unstable. For determining folate forms in serum or EDTA whole blood, the samples should be stored at –80 °C as quickly as possible after collection. Repeated freezing and defrosting of the samples should be avoided. Folates are sensitive to temperature. Therefore, samples should not be handled and stored at room temperature for an extended period of time.

Reference interval

The reference intervals for erythrocyte folate are shown in Tab. 13.4-2. The folate form distribution in women (pregnant women and control cohort), umbilical cord blood and older adults is shown in Tab. 13.4-5 – Folate forms in serum using UPLC-MS/MS.

Evidence has been provided using the LC-MS/MS method that the distribution of folate forms in the erythrocytes differs in the different MTHFR genotypes. Compared to 677 CC and CT genotypes, carriers of 677 TT polymorphism accumulate up to 43% non methyl THF in the total folate concentration (Tab. 13.4-6 – Folate forms in serum using UPLC-MS/MS and immunological method by MTHFR C677T genotype/30/. Like erythrocyte folate, folate concentrations and folate form distribution in serum are dependent on the MTHFR genotype /3/, which may affect DNA methylation. The differences between serum and erythrocyte folates only become evident in a low folate status /30/.

Reference intervals for the determination of folate forms using LC-MS/MS have not been specified to date because of the differences among the measuring methods. Some guidance values for serum and erythrocyte folate are shown in Tab. 13.4-7 – Categories for 5-methyl -THF in fasting serum/plasma and erythrocytes using UPLC-MS/MS or LC-MS/MS. The folate in serum/plasma and the cerebrospinal fluid (CSF) is shown in Tab. 13.4-8 – Folate in serum/plasma and CSF using LC-MS/MS or immunological method.

Hemolysis

As hemolysis leads to falsely high values because of the high folate concentration in the erythrocytes, hemolytic serum is not suited for analysis.

Exposure to light

Folic acid and reduced folates are sensitive to light. 24 h exposure to light at room temperature decreases folate concentration by 12% if stored in a plastic tube and by 19% if stored in a separator tube /37/.

Stability

At 2–8 °C, the sample should not be stored for more than 2 days. Deep-frozen serum samples keep for 1 month at –20 °C and for longer at –80 °C. Folic acid is relatively stable, while reduced folates, especially THF and dihydrofolate, are unstable. They are immediately subject to inter conversion and degradation that are dependent on enzymes and/or influenced by oxygen presence, temperature, exposure to light and pH value. Care should always be taken, especially in the determination of folate form distribution in serum and erythrocytes, to be aware of changes in the folate profile resulting from sample processing and analysis. Therefore, the folate forms are in many cases subdivided into two groups, 5-methyl-THF and non-methyl-THF.

13.4.7 Pathophysiology

Folate

Folate plays an essential role in cell growth and proliferation. It provides methyl groups to homocysteine for conversion to methionine. It is also essential for the synthesis of purine and pyrimidine nucleotides. For this reason, folates play a key role in numerous cellular functions. Folate deficiency impairs DNA synthesis, causes hypo methylation and this results in the development of clinical conditions, such as megaloblastic anemia, neural tube defects and neurological damage.

Folic acid is the synthetic form of the vitamin. Chemically, it is pteroylmonoglutamic acid and contains pteroic acid as a substructure, which consists of pteridine and p-aminobenzoic acid. Depending on the number of glutamyl residues, pteroylmonoglutamate, -triglutamate, -heptaglutamate and/or -polyglutamate are distinguished. The pteridine ring can be present in the oxidized, di hydrated and tetra hydrated form.

Natural folates differ by:

  • The hydration state of the pteridine ring
  • Substitutions at N-5 and N-10 of the pteridine ring; various C1 units, such as methyl, formaldehyde or formate groups can be bound
  • The length of the glutamyl chain.

The parent substance of the folate coenzymes is 5,6,7,8-tetrahydrofolate (THF), which acts as universal C1 acceptor. 5-Methyl-THF, 5,10-methylene-THF, 5,10-methenyl-THF, 5-formyl-THF, 10-formyl-THF and THF are the most important folate forms in the body. Unmetabolized folic acid can be detected in the blood after intake of more than 200 μg/d of folic acid; however, the significance of the presence of unmetabolized folic acid in serum has been controversially discussed. 5-Methyl THF occurs in serum predominantly as mono glutamate and intracellularly as poly glutamate-THF with or without C1 units.

Uptake and metabolism

Folic acid is the most stable form of the vitamin with the highest oxidation stage and is quantitatively resorbed (more than 90%) in this form. Folates occur in all foods of plant and animal origin. For example, liver, vegetables (lettuce, spinach, asparagus, tomatoes, cucumbers) or cereals are rich in this vitamin. Food mainly contains polyglutamylfolates that are hydrolyzed to mono glutamate by a γ-glutamyl-carboxypeptidase (EC 3.4.19.9) in the brush border of the mucosa cells in the duodenum and upper jejunum for resorption. Resorption is optimal at pH 6. Transport across the mucosal membrane is mostly active; only 20–30% of the folates are taken up by passive diffusion. Via portal circulation, the folates pass to the liver where they are converted to the methylated form. The predominant form in the blood is 5-methyl-THF (as well as THF and 10-formyl-THF), which is transported bound to albumin, α2-macroglobulin and transferrin.

The uptake of 5-methyl-THF into the erythrocytes is based on the principles of saturation kinetics using a membrane-bound carrier. Passage across the blood-brain barrier is similar; the folate concentrations in the cerebrospinal fluid being approximately equal /38/ to 3-fold higher than in serum (Tab. 13.4-8 – Folate in serum/plasma and cerebrospinal fluid). In the erythrocytes, the folates are present as poly glutamate, mostly with 4–7 glutamate residues that have a high affinity to deoxyhemoglobin. Folate concentrations in erythrocytes are about 20–40-fold higher than in serum. For retention, intracellular 5-methyl-THF is de methylated in a vitamin B12-dependent reaction and then converted to poly glutamate.

Folates and C1 metabolism

Folic acid itself is not biologically active, but THF and its derivatives are. They are the decisive coenzymes acting as acceptors and carriers of hydroxy methyl groups (activated formaldehyde) and formyl groups (activated formic acid). The C1 residues coming from various metabolic processes are bound to THF and passed on to suitable acceptors for the synthesis of substances. THF-C1 compounds are present in various oxidation stages and can be inter converted as shown in Fig. 13.4-3 – Function of folate forms in human cells.

Methylation reactions play a prominent role in metabolism utilizing the following donors: 5,10-methylene-THF provides the methyl group for the formation of thymidylate from d-uridylate during DNA synthesis, 5-methyl-THF provides the methyl groups for the methylation of homocysteine to methionine. In the liver, homocysteine can also be methylated by betaine. The methyl groups of betaine are provided by S-adenosylmethionine (i.e., indirectly by 5-methyl-THF).

All other methylation reactions are based on S-adenosylmethionine, the universal methyl group donor for various biosyntheses, such as:

  • The formation of phosphatidylcholine, choline, carnitine, acetylcholine, adrenalin, creatine, thymine (methyluracil)
  • The methylation of myelin, DNA, receptors and neurotransmitters.

Folate deficiency is attributed to

  • Under supply due to inadequate intake
  • Reduced folate resorption in the intestinal tract
  • Increased consumption
  • In existing vitamin B12 deficiency (secondary folate deficiency).

Refer additionally to Tab. 13.4-9 – Causes of folate deficiency.

Cell dysfunction

Persistent folate deficit causes the depletion of the folate stores in serum and in the erythrocytes and leads to hyper segmentation of the neutrophil granulocytes, later followed by thrombocytopenia, leukopenia and finally, after months, macrocytic anemia. Impaired erythropoiesis is based on impaired nucleic acid synthesis, impaired maturation and, to a lesser extent, reduced hemoglobin synthesis.

During DNA synthesis, deoxyuridine mono phosphate (dUMP) is methylated to deoxythymidine mono phosphate (dTMP = thymidylate) by thymidylate synthase, with the methyl group provided by 5,10-methylene-THF. dTMP is successively converted to thymidine triphosphate (dTTP) and incorporated into the DNA. The folate deficiency-induced inhibition of thymidylate synthase causes reduced formation of dTMP. Increased incorporation of uracil into the DNA due to dTMP deficiency is corrected by repair mechanisms (splicing) (Fig. 13.4-4 – Effects of folate deficiency and/or vitamin B12 deficiency on erythropoiesis). This can lead to chromosomal breakage. Moreover, DNA hypo methylation can cause the activation of tumor suppressor genes /41/. Tissues with fast proliferating cells (lymphatic tissue, intestinal brush border membrane, hair follicles or tumor tissue) constantly require dTMP.

Like dUMP, the 5-fluorodeoxyuridylate (FdUMP) used in cancer therapy binds to thymidylate synthase, but – unlike dUMP – is not released and therefore causes irreversible inhibition of the enzyme. dTMP deficiency leads to the death of fast proliferating cells.

Hematopoietic cells with folate or cobalamin deficiency undergo changes in morphology, protein synthesis and cell kinetics. The changes are especially pronounced in the final stage of cell division and in non-divisible cells and mainly affect erythropoiesis and granulopoiesis. Ineffective erythropoiesis causes the death of polychromatic megaloblasts.

References

1. World health Organization, Food and Agriculture Organization. Vitamin and mineral requirements in human nutrition, 2nd ed. World Health Organization 2004; accessed 6 February 2014.

2. Hannisdal R, Ueland PM, Svardal A. Liquid chromatography-tandem mass spectrometry analysis of folate and folate catabolites in human serum. Clin Chem 2009; 55: 1147–54.

3. Kirsch SH, Knapp JP, Herrmann W, Obeid R. Quantification of key folate forms in serum using stable-isotope dilution ultra performance liquid chromatography-tandem mass spectrometry. J Chromatogr B Analyt Technol Biomed Life Sci 2010; 878: 68–75.

4. Hicks JM, Cook J, Godwin ID, Soldin SJ. Vitamin B12 and folate. Pediatric reference ranges. Arch Pathol Lab Med 1993; 117: 704–6.

5. Bäsler KH, Golly I, Loew D, Pietrzik K. Vitamin-Lexikon. München: Urban & Fisher, 2002.

6. Hoffbrand AV, Newcombe FA, Mollin DL. Method of assay of red cell folate activity and the value of the assay as a test for folate deficiency. J Clin Pathol 1966; 19: 17–28.

7. Schroeder HA. Losses of vitamins and trace minerals resulting from processing and preservation of foods. Am J Clin Nutr 1971; 24: 562–73.

8. Stanger O, Herrmann W, Pietrzik K et al. Konsensuspapier der D.A.CH.-Liga Homocystein über den rationellen klinischen Umgang mit Homocystein und B-Vitaminen bei kardiovaskulären und thrombotischen Erkrankungen – Richtlinien und Empfehlungen. J Kardiol 2003; 10: 190–9.

9. Flynn A, Hirvonen T, Mensink GB et al. Intake of selected nutrients from foods, from fortification and from supplements in various European countries. Food Nutr Res 2009; 53.

10. Beitz R, Mensink GB, Fischer B, Thamm M. Vitamins – dietary intake and intake from dietary supplements in Germany. Eur J Clin Nutr 2002; 56: 539–45.

11. Selhub J, Jacques PF, Wilson PW, Rush D, Rosenberg IH. Vitamin status and intake as primary determinants of homocysteinemia in an elderly population. JAMA 1993; 270: 2693–8.

12. Lewis CA, Pancharuniti N, Sauberlich HE. Plasma folate adequacy as determined by homocysteine level. Ann N Y Acad Sci 1992; 669: 360–2.

13. Herrmann W, Obeid R, Schorr H, Geisel J. Functional vitamin B12 deficiency and determination of holotranscobalamin in populations at risk. Clin Chem Lab Med 2003; 41: 1478–88.

14. World health Organization Guideline:. Optimal serum and red blood cell concentrations in women of reproductive age for prevention of neural tube defects. World Health Organization 2015; ISBN 9789241549042

15. Lumley J, Watson L, Watson M, Bower C. Periconceptional supplementation with folate and/or multivitamins for preventing neural tube defects. Cochrane Database Syst Rev 2001; CD001056.

16. Franke C, Verwied-Jorky S, Campoy C et al. Dietary intake of natural sources of docosahexaenoic acid and folate in pregnant women of three European cohorts. Ann Nutr Metab 2008; 53: 167–74.

17. Herrmann W, Obeid R. The mandatory fortification of staple foods with folic acid: a current controversy in Germany. Dtsch Arztebl Int 2011; 108: 249–54.

18. Willems FF, Aengevaeren WR, Boers GH, Blom HJ, Verheugt FW. Coronary endothelial function in hyperhomocysteinemia: improvement after treatment with folic acid and cobalamin in patients with coronary artery disease. J Am Coll Cardiol 2002; 40: 766–72.

19. Sanjoaquin MA, Allen N, Couto E, Roddam AW, Key TJ. Folate intake and colorectal cancer risk: a meta-analytical approach. Int J Cancer 2005; 113: 825–8.

20. Cole BF, Baron JA, Sandler RS et al. Folic acid for the prevention of colorectal adenomas: a randomized clinical trial. JAMA 2007; 297: 2351–9.

21. Cooper K, Squires H, Carroll C et al. Chemoprevention of colorectal cancer: systematic review and economic evaluation. Health Technol Assess 2010; 14: 1–206.

22. Ericson U, Sonestedt E, Gullberg B, Olsson H, Wirfalt E. High folate intake is associated with lower breast cancer incidence in postmenopausal women in the Malmo Diet and Cancer cohort. Am J Clin Nutr 2007; 86: 434–43.

23. Stolzenberg-Solomon RZ, Chang SC, Leitzmann MF et al. Folate intake, alcohol use, and postmenopausal breast cancer risk in the Prostate, Lung, Colorectal, and Ovarian Cancer Screening Trial. Am J Clin Nutr 2006; 83: 895–904.

24. Oaks BM, Dodd KW, Meinhold CL, Jiao L, Church TR, Stolzenberg-Solomon RZ. Folate intake, post-folic acid grain fortification, and pancreatic cancer risk in the Prostate, Lung, Colorectal, and Ovarian Cancer Screening Trial. Am J Clin Nutr 2010; 91: 449–55.

25. Larsson SC, Giovanucci E, Wolk A. Methionine and vitamin B6 intake and risk of pancreatic cancer: a prospective study of Swedish women and men. Gastroenterology 2007; 132: 113–8.

26. Tworoger SS, Hecht JL, Giovannucci E, Hankinson SE. Intake of folate and related nutrients in relation to risk of epithelial ovarian cancer. Am J Epidemiol 2006; 163: 1101–11.

27. Stevens VL, Rodriguez C, Pavluck AL, McCullough ML, Thun MJ, Calle EE. Folate nutrition and prostate cancer incidence in a large cohort of US men. Am J Epidemiol 2006; 163: 989–96.

28. Hubner RA, Houston RS. Folate and colorectal cancer prevention Br J Cancer 2009; 100: 233–9.

29. Obeid R, Kasoha M, Kirsch SH, Munz W, Herrmann W. Concentrations of unmetabolized folic acid and primary folate forms in pregnant women at delivery and in umbilical cord blood. Am J Clin Nutr 2010; 92: 1416–22.

30. Fazili Z, Pfeiffer CM, Zhang M, Jain RB, Koontz D. Influence of 5,10-methylenetetrahydrofolate reductase polymorphism on whole-blood folate concentrations measured by LC-MS/MS, microbiologic assay, and bio-rad radioassay. Clin Chem 2008; 54: 197–201.

31. Wang Y, Zhang HY, Liang QL, et al. Simultaneous quantification of 11 pivotal metabolites in neural tube defects by HPLC-electrospray tandem mass spectrometry. J Chromatogr B Analyt Technol Biomed Life Sci 2008; 863: 94–100.

32. Bodnar LM, Himes KP, Venkataramanan R et al. Maternal serum folate species in early pregnancy and risk of preterm birth. Am J Clin Nutr 2010; 92: 864–71.

33. Fazili Z, Pfeiffer CM, Zhang M. Comparison of serum folate species analyzed by LC-MS/MS with total folate measured by microbiologic assay and Bio-Rad radioassay. Clin Chem 2007; 53: 781–4.

34. Summers CM, Mitchell LE, Stanislawska-Sachadyn A et al. Genetic and lifestyle variables associated with homocysteine concentrations and the distribution of folate derivatives in healthy premenopausal women. Birth Defects Res A Clin Mol Teratol 2010; 88: 679–88.

35. Smulders YM, Smith DE, Kok RM et al. Red blood cell folate vitamer distribution in healthy subjects is determined by the methylenetetrahydrofolate reductase C677T polymorphism and by the total folate status. J Nutr Biochem 2007; 18: 693–9.

36. Huang Y, Khartulyari S, Morales ME et al. Quantification of key red blood cell folates from subjects with defined MTHFR 677C>T genotypes using stable isotope dilution liquid chromatography/mass spectrometry. Rapid Commun Mass Spectrom 2008; 22: 2403–12.

37. Mastropaolo W, Wilson MA. Effect of light on serum B12 and folate stability. Clin Chem 1993; 39: 913.

38. Obeid R, Kostopoulos P, Knapp JP et al. Biomarkers of folate and vitamin B12 are related in blood and cerebrospinal fluid. Clin Chem 2007; 53: 326–33.

39. Hagnelius NO, Wahlund LO, Nilsson TK. CSF/serum folate gradient: physiology and determinants with special reference to dementia. Dement Geriatr Cogn Disord 2008; 25: 516–23.

40. Popp J, Lewczuk P, Linnebank M et al. Homocysteine metabolism and cerebrospinal fluid markers for Alzheimer’s disease. J Alzheimers Dis 2009; 18: 819–28.

41. Woodson K, Mason J, Choi SW et al. Hypomethylation of p53 in peripheral blood DNA is associated with the development of lung cancer. Cancer Epidemiol Biomarkers Prev 2001;10: 69–74.

13.5 Vitamin B6

Wolfgang Herrmann, Rima Obeid

Vitamin B6 is the collective term for three similar, water-soluble compounds, pyridoxol/pyridoxine, pyridoxal (PL) and pyridoxamine (Fig. 13.5-1 – B6-vitamins and their corresponding phosphates). All three forms are precursors of pyridoxal-5’-phosphate (PLP), the most active form of vitamin B6 involved as cofactor in more than 180 enzymatic reactions in the human body. The human body itself cannot produce the cofactor pyridoxal phosphate and therefore depends on the dietary supply of its precursors.

Vitamin B6 is ubiquitous in nature. It can be synthesized by microorganisms and also by higher plants. Vitamin B6 is especially abundant in meat, liver, potatoes, cereals, vegetables, certain fish species (mackerel) and dairy products. About 30–40% of vitamin B6 in animal products are lost by frying and cooking; the loss of vitamin B6 in plant-based products by cooking is much smaller. The bio availability of vitamin B6 ranges between 70 and 80%.

13.5.1 Indication

  • Diagnosis of deficiency states
  • Part of the individual risk profile for patients with degenerative cardiovascular and neurodegenerative disease

Target groups are listed in Tab. 13.5-1 – Indications for vitamin B6 determination.

13.5.2 Method of determination

High performance liquid chromatography (HPLC)

Principle: the first step consists of sample processing and includes precipitation followed by sample derivatization. In the precipitation step, higher molecular substances are separated. The precipitate is sedimented by centrifugation and the supernatant is extracted. Vitamin B6 is converted to a fluorescent derivative by semi carbazide or cyanide derivatization. The derivatization reagent is added simultaneously with the precipitation reagent or later (depending on the method used) and incubated for 20 minutes. After centrifugation, the supernatant is injected into the HPLC. Fractionation by HPLC is effected on a reversed-phase column using isocratic methods at 25 °C. The chromatograms is recorded by a fluorescence detector.

Liquid chromatography-tandem mass spectrometry (LC-MS/MS)

The LC-MS/MS method allow the simultaneous, direct and fast determination of various vitamin B6 forms (pyridoxal 5’-phosphate, pyridoxal, 4-pyridoxic acid, pyridoxine and pyridoxamine) /1/.

Principle: plasma is deproteinized by means of trichloroacetic acid and mixed with internal standard containing Delta2-pyridoxal-5’-phosphate, Delta3-pyridoxal and Delta8-riboflavin. PL, PLP, pyridoxine (PN), pyridoxine 5’-phosphate, pyridoxamine (PM), pyridoxamine-5’-phosphate, 4-pyridoxic acid (PA), riboflavin, flavin mononucleotide (FMN) and FAD are fractionated by means of a C8 reversed-phase column applying an acetonitril gradient in an acetic acid/heptafluorobutyric acid buffer. The analytes are detected and quantified using LC-MS/MS in the positive ion mode.

This method does not necessitate analyte derivatization, only requires a small sample volume, has short fractionation times (8 min. per cycle) and provides acceptable sensitivity and specificity /1/.

Enzymatic assay

Enzymatic methods are based on pyridoxal-5’-phosphate (PLP)-dependent enzymes, such as apo-tyrosine-decarboxylase, erythrocytic aspartate aminotransferase or apo-homocysteine-lyase /8/. The enzymatic reaction products are detected radioactively or chromogenically. The results of the assays show the total PLP activity of the vitamers.

Principle: the PLP assay uses the apo form of the PLP-dependent, recombinant enzyme homocysteine-α,γ-lyase (rHCYase). For apo-rHCYase extraction, PLP is removed from the holoenzyme rHCYase by incubation with hydroxylamine. Enzymatic activity is restored by holoenzyme reconstitution which is linearly correlated with the enzyme-bound PLP. By way of conversion (reconstituted holo-rHCYase) from millimolar concentrations of homocysteine to H2S, the amplification principle of the assay allows to measure PLP concentrations in the nanomolar range. In the indicator reaction, N,N-dibutylphenylenediamine (is used as chromophor for the quantification of the formed H2S. The enzymatic method correlates well with the chromatographic methods of determination.

13.5.3 Specimen

Serum or EDTA plasma: 1 mL

13.5.4 Reference interval

Pyridoxal-5 phosphate (PLP): ≥ 4.9 μg/L (20 nmol/L)

Method-dependent reference values of the vitamin B6 status are shown in Tab. 13.5-2 – Methods for determining the vitamin B6 status and their reference values. Other authors specify PLP above 30 nmol/L as normal concentrations /3/.

13.5.5 Clinical significance

The turnover time of the vitamin B6 pool is not known in detail. It has been reported that there are two pools: a fast one with a turnover of about 0.5 days, and a slow one with a turnover of 25–33 days. The slow turnover compartment is classified as storage pool. Accordingly, vitamin B6 can only be stored for a few weeks. Deficiency symptoms following store depletion include muscular dystrophy, skin alterations, nervous system disorders or impaired protein biosynthesis.

The determination of serum vitamin B6 only allows limited conclusions regarding the functional, intracellular vitamin B6 status. The detection of vitamin B6 deficiency therefore utilizes various other available analysis methods and biomarkers that are more or less common due to their different diagnostic specificity, sensitivity and practicability.

The following analysis methods have been described for the diagnosis of vitamin B6 deficiency:

  • The most common method is the determination of pyridoxal-5’-phosphate (PLP) and pyridoxal (PL) in plasma and/or whole-blood
  • Another test is to measure the excretion of xanthurenic acid following tryptophan loading. The test is based on the different response times of vitamin B6-dependent enzymes to vitamin B6 deficiency. Thus, urinary excretion of xanthurenic acid can be utilized as a marker of vitamin B6 deficiency.
  • Oxidation of PL by aldehyde oxidase or aldehyde dehydrogenase results in 4-pyridoxic acid. Urinary excretion of 4-pyridoxic acid can be analyzed.
  • Determination of the activity of erythrocytic aspartate aminotransferase (AST) with and without PLP is also utilized as a marker. Activity is reduced in B6 deficiency and increases after incubation with PLP.
  • In homocysteine metabolism, vitamin B6 also acts as a cofactor of two enzymes of the transsulfuration pathway, cystathionine β-synthase and γ-cystathionase. Consequently, cystathionine and homocysteine are functional biomarkers of intracellular vitamin B6 deficiency, with cystathionine being more sensitive. However, these biomarkers are not specific of vitamin B6 deficiency but can also be elevated in folate or vitamin B12 deficiency or in renal dysfunction.

The oral methionine loading test is especially sensitive in the detection of vitamin B6 deficiency, even at normal fasting homocysteine levels. Because, in vitamin B6 deficiency, the transsulfuration pathway in homocysteine catabolism is only moderately impaired, the fasting homocysteine level in intact remethylation may still range within the reference interval. However, non-fasting conditions or methionine load in the setting of vitamin B6 deficiency results in marked changes in homocysteine concentration.

13.5.5.1 Assessment of vitamin B6-status

Metabolization analysis and pharmacokinetic studies trace the profile of all B6 forms present. Determination of the pyridoxal-5’-phosphate (PLP) level including the precursor molecule pyridoxal (PL) has become the established method for analyzing the vitamin B6 supply situation. PLP and PL in serum and/or plasma (and, possibly, PL in the erythrocytes) form the main B6 vitamers available for the tissues. Impairment of the phosphorylation-based equilibrium between PLP and PL causes a change in the matrix-specific correlations of the two forms.

Since it is not always sufficient to use PLP alone as an indicator of the vitamin B6 status, the additional use of PL is recommended /4/. In a comparison of pregnant women with a nonpregnant control cohort, PLP was 50% lower in the pregnant women, but the total amount of PLP and PL levels did not differ significantly /5/. Relating the PLP and PL concentrations to albumin did not show any differences between the two groups.

Excretion of 4-pyridoxic acid is regarded as a short-term indicator of the vitamin B6 status. In B6 deficiency studies, the decrease in 4-pyridoxic acid in urine was similar to that of PLP in plasma /6/. Total urinary excretion of vitamin B6 (all forms) is not a sensitive indicator of the vitamin B6 status.

The aminotransferase activity in the erythrocytes is considered a long-term indicator of the vitamin B6 status /7/. The activity is analyzed with and without adding PLP. However, the reliability of this biomarker for the assessment of the vitamin B6 status is controversial /8/.

13.5.5.2 Interactions with drugs

Treatment with certain drugs can affect the vitamin B6 status and result in increased vitamin B6 requirement. Tab. 13.5-3 – Interactions between vitamin B6 and drugs shows the drugs and substances and their effects on the vitamin B6 status /9/. A common characteristic of drug interaction is the unfavorable effect on the function of the central nervous system (CNS). Moreover, many drugs react with pyridoxal-5’-phosphate (PLP) to form a Schiff base. This reaction can cause a decrease in the PLP tissue level, which may result, for example, in dysfunction in the brain. The drugs listed below may reduce the PLP level. Therefore, an adequate supply with vitamin B6 must be ensured upon intake of such drugs.

The following drugs cause interactions with vitamin B6:

  • Cycloserine (seromycine, tuberculosis treatment), isoniazid, (tuberculosis treatment), penicillamine (rheumatoid vasculitis, arthritis), theophylline (anti asthmatic), hydralazine (apresoline, antihypertensive). In most cases, vitamin B6 supplementation will eliminate undesired side effects and compensate for increased requirement /1011/.
  • Oral contraceptives do not react directly with pyridoxal-5’-phosphate (PLP), but induce the synthesis of enzymes, including some PLP-dependent ones. This results in increased PLP binding in the tissue and reduced plasma PLP concentration /12/.
  • Estrogens mainly influence enzymes of the tryptophan-niacin metabolism. In some cases, the vitamin B6 requirement can be higher in women under contraceptive treatment. Erythropoietin therapy reduces the vitamin B6 level in the erythrocytes so that supplementation may be required /13/.
  • Phenytoin (anticonvulsive) reduces the therapeutic effect of vitamin B6 supplementation. In concurrent antibiotics therapy and vitamin B6 supplementation, the two substances interfere with each other in intestinal absorption and should therefore not be taken at the same time.
  • Levodopa. Vitamin B6 reduces the effect of levodopa (anti-Parkinson drug), but also reduces the side effects of the treatment (vomiting, nausea). Therefore, care should be taken to ensure that the level of vitamin B6 supplementation attenuates the side effects, but does not reduce the effect of the drug.
  • Chemotherapeutics. In some chemotherapeutics like 5-fluorouracil, doxorubicin, vitamin B6 supplementation attenuates certain side effects without reducing the effect of the drug.
  • Tricyclic antidepressants. Their effect is also improved by vitamin B6 supplementation.

13.5.5.3 Vitamin B6-deficiency

Vitamin B6 deficiency can develop from vitamin B6-deficient diet, impaired vitamin B6 metabolism or dysfunctional vitamin B6-dependent enzymes, whose activity can partly be restored by high amount of vitamin B6. Pronounced vitamin B6 deficiency is rare and usually occurs as a result of other diseases. Because of the high number of vitamin B6-dependent reactions, deficiency can manifest in many ways. The significance of pyridoxal-5’-phosphate (PLP) for numerous cellular processes and functions is based on its participation as cofactor in various enzymatic reactions (Tab. 13.5-4 – Pyridoxal phosphate-catalyzed enzymatic reactions).

Processes and their functions influenced by PLP are shown in Tab. 13.5-5 – Vitamin B6-dependent cellular processes. Vitamin B6 deficiency has a negative effect on humoral and cellular immunity. Vitamin B6 treatment (50 mg/d) for 2 months improved lymphocyte responses in elderly women with impaired immunocompetence /14/.

13.5.5.4 Clinical symptoms and diseases

The most common symptoms in vitamin B6 deficiency are peripheral polyneuropathy, vomiting, depression, impaired cognitive competence and microcytic anemia. It cannot be clearly determined in most cases whether the vitamin B6 deficiency is based on increased catabolic rate, increased requirement, nutritional deficiency or impaired PLP synthesis. Diseases and conditions possibly associated with abnormal vitamin B6 metabolism are listed in Tab. 13.5-6 – Abnormal vitamin B6 metabolism, diseases and conditions.

Numerous diseases or pathological conditions are associated with abnormal vitamin B6 metabolism. The pyridoxal-5’-phosphate (PLP) concentration or tryptophan metabolism are regarded as primary indicators in this context. However, the use of PLP alone does not provide reliable information as to whether the metabolism has actually changed. Occurrences of pellagra and anemia have been reported historically in severe vitamin B6 deficiency. Mild vitamin B6 deficiency has been associated with premenstrual syndrome /15/, carpal tunnel syndrome /16/ and psychiatric disease /17/. High-dose vitamin B6 supplementation (100–200 mg/day) given to patients with carpal tunnel syndrome was found to reduce the risk of acute cardiac chest pain or myocardial infarction /18/. However, none of the above-mentioned associations are considered proven.

PLP is required for the conversion of tryptophan to niacin /3/. Diseases where tryptophan metabolism changes were determined and vitamin B6 supplemented include asthma, diabetes, certain tumors, pellagra and rheumatoid arthritis. Diseases or conditions where low PLP concentrations have been reported include asthma, diabetes, renal disease, alcoholism, cardiac disease, pregnancy and carcinomas (Tab. 13.5-7 – Diseases associated with vitamin B6 deficiency). Supplementation has been applied in numerous diseases or pathological conditions. High doses of pyridoxine (up to 750 mg/day) are given in enzyme dysfunction to restore normal functions. However, there is the risk that high doses of vitamin B6 can lead to peripheral sensory neuropathies and nerve degeneration /19/.

13.5.5.5 Recommended dietary allowance

The recommended dietary allowance (RDA) of vitamin B6 is 1.3 to 1.7 mg/day in the United States /20/ and 1.4 mg/day in Europe /21/. In 2003/4, the mean intake in the United States was 1.86 mg/day in nonusers and 1.92 mg/day in users (36% of the US population take vitamin supplements) /38/. In these cases, linear correlation between the plasma PLP level and vitamin B6 intake was observed. Mean plasma PLP level was 41 nmol/L in U.S. males not using vitamin supplementation and 29 nmol/L in females.

Vitamin B6 requirement also depends on protein turnover and increases as a function of protein intake because there is a positive correlation due to the involvement of pyridoxine in amino acid metabolism /39/. High protein intake induces the enzymes involved in amino acid metabolism to bind more PLP and, as a result, PLP is no longer available for other metabolic processes.

13.5.6 Comments and problems

Method of determination

The different procedures yield comparable results. However, the comparability of results obtained in different laboratories is not always satisfactory and leaves room for improving standardization /24/.

Procedures measuring one or more B6 vitamers or the metabolite 4-pyridoxic acid are regarded as direct indices of the vitamin B6 status. Indirect procedures determine metabolites or metabolic pathways, where pyridoxal-5’-phosphate (PLP) is essential as a cofactor and/or PLP-dependent enzyme activities are measured.

Reference interval

The reference interval for vitamin B6 strongly depends on the nutritional state and, consequently, the selection of the subject cohort. The reference range specified for vitamin B6 is only for orientation and should be confirmed by own data.

Stability

For determination, collect fasting blood in the morning before use of any medication.

Because vitamin B6 is highly sensitive to light and temperature, the sample should be protected from light, cooled and centrifuged immediately. The samples are stable for 1 week if stored at 2–8 °C in the dark and remain stable for months at –20 °C. Vitamin B6 can also be determined from whole blood. However, there are currently no generally valid reference values for this.

13.5.7 Pathophysiology

Vitamin B6 is the collective term for three similar, water-soluble compounds, pyridoxol/pyridoxine, pyridoxal (PL) and pyridoxamine (Fig. 13.5-1 – B6-vitamins and their corresponding phosphates). The corresponding 5’-phosphoric acid esters are the biologically active coenzymes. Chemically, they are derivatives of 4,5-bis(hydroxymethyl)-2-methylpyridine-3-ol and are distinguished by different substituents at position 4 that are involved in the coenzyme function. Having the same biological activity, the three species can be inter converted by metabolism. The following enzymes participate in the inter conversion of the different vitamers: pyridox(am)ine phosphate oxidase (PNPO, EC 1.4.3.5), pyridoxal kinase (PDXK, EC 2.7.1.35), pyridoxal phosphatase (PDXP, EC 3.1.3.74) and other phosphatases [alkaline phosphatase (ALP, EC 3.1.3.1), acidic phosphatase (ACP, EC 3.1.3.2)]. Pyridoxal is converted to 4-pyridoxate by means of aldehyde oxidase 1 (AOX1, EC 1.2.3.1). All three forms are precursors of pyridoxal-5’-phosphate (PLP), the most active form of vitamin B6 involved as cofactor in more than 180 enzymatic reactions in the human body. The human body itself cannot produce the cofactor pyridoxal phosphate and therefore depends on the dietary supply of its precursors.

Metabolism

Vitamin B6 is predominantly resorbed in the upper jejunum and ileum. Vitamin B6 generated in the colon is not available to the body. Before entering the intestinal mucosa cells, the vitamers must be de phosphorylated, a process catalyzed by the membrane-bound ALP /41/. The extents to which pyridoxal-5’-phosphate (PLP) and pyridoxal (PN) are resorbed in the human body are comparable. The resorption of phosphorylated compounds is slower. The resorption of B6 vitamers into the mucosa cells is a passive, non-saturable process. In the intestinal mucosa cell, the vitamers are re phosphorylated to PLP by pyridoxal kinase and then, before leaving the cell on the serosal side of the intestinal mucosa, de phosphorylated again. Following enteral resorption, phosphorylation to PLP and PNP takes place in the liver and other organs. Vitamin B6 in plasma occurs as PLP (60%), PN (15%) and PL (14%). Most derivatives are bound to albumin. The binding of PLP to albumin in the circulation protects against hydrolysis and enables PLP supply to the other tissues. The protein-bound PLP (via Schiff base) is thought to form the circulating depot form of vitamin B6 because it is membrane-impermeable and, therefore, not directly available to the cell. PLP is hydrolyzed to PL for tissue resorption by simple diffusion and, following resorption, PL is re phosphorylated again to the active coenzyme form. In the erythrocytes, PLP is mainly bound to hemoglobin, and its concentration is 4–5-fold higher than in plasma /3/. The vitamin B6 pool in adults is estimated at 40–50 mg. Most of it is bound to glycogen phosphorylase and localized in the muscles /42/.

Biochemistry

More than 180 enzymatic reactions are PLP-dependent /43/. In enzymatic reactions, PLP is covalently bound to the corresponding enzyme via a Schiff base (Fig. 13.5-2 – Key mechanism of the transamination reaction between enzyme-bound pyridoxal-5’-phosphate and an amino acid).

PLP is involved in the metabolism of amino acids, glucose and lipids, the synthesis of neurotransmitters, histamine and hemoglobin and the function and expression of genes. The liver is the main location of vitamin B6 metabolism. An overview of the types of PLP-catalyzed enzymatic reactions are provided in Tab. 13.5-4 – Pyridoxalphosphate-catalyzed enzymatic reactions. Reactions on the α, β and γ carbon atom of amino acids have been defined. PLP is also a cofactor in transamination, transsulfuration, deamination, decarboxylation, racemization and eliminations or replacements at the β or γ carbons /4243/. PLP is involved in the amino acid metabolism from synthesis to degradation: (a) Amino acid decarboxylation results in the formation of amines, which act as neurotransmitters or hormones, such as γ-aminobutyrate, histamine, noradrenaline (converted to adrenaline) or serotonin. (b) Amino acid transamination results in the formation of keto acids, which are then oxidized for metabolic energy. ( c ) A multitude of reactions pertain to the amino acid side chains, such as kynureninase (EC 3.7.1.3), cystathionine-β-synthase (EC 4.2.1.22) or γ-cystathionase (EC 4.4.1.1). (d) In phospholipid synthesis, phosphatidylserine decarboxylation results in the formation of phosphatidylethanolamine.

Besides folate and vitamin B12, vitamin B6 plays an important role as a cofactor in homocysteine catabolism. The cell toxic homocysteine stands at the intersection of two metabolic pathways: the re methylation to methionine and the transsulfuration to cysteine via cystathionine. The latter pathway is catalyzed by the enzymes cystathionine-β-synthase and γ-cystathionase, both of which need vitamin B6 as a cofactor. Impaired homocysteine degradation can lead to hyperhomocysteinemia, which is considered an independent risk factor for vasodegenerative and neurodegenerative disease. The correlation between vitamin B6 deficiency and postprandial increase in homocysteine has frequently been shown /44/. It has been shown in individuals with pathologic homocysteine concentration following a methionine loading test that vitamin B6 supplementation normalized the afterload homocysteine concentration in about 50% of the individuals /45/.

References

1. Gentili A, Caretti F, D’Ascenzo G et al. Simultaneous determination of water-soluble vitamins in selected food matrices by liquid chromatography/electrospray ionization tandem mass spectrometry. Rapid Commun Mass Spectrom 2008; 22: 2029–43.

2. Han Q, Xu M, Tang L et al. Homogeneous, nonradioactive, enzymatic assay for plasma pyridoxal 5-phosphate. Clin Chem 2002; 48: 1560–4.

3. Leklem JE. Vitamin B6. In: Rucker RB, Suttie JW, McCormick DB, Machlin LJ, eds. Handbook of Vitamins. New York: Marcel Dekker Inc 2001: 339–96.

4. Reynolds RD. Biochemical methods for status assessment. In: Raiten DJ, ed. Vitamin B6 Metabolism in Pregancy, Lactation and Infancy. Boca Raton: CRC Press 1995: 41–59.

5. Barnard HC, de Kock JJ, Vermaak WJ, Potgieter GM. A new perspective in the assessment of vitamin B-6 nutritional status during pregnancy in humans. J Nutr 1987; 117: 1303–6.

6. van den Berg H, Bogaards JJ. Vitamin B-6 metabolism in the pregnant rat: effect of progesterone on the (re)distribution in maternal vitamin B-6 stores. J Nutr 1987; 117: 1866–74.

7. Driskell JA, Clark AJ, Moak SW. Longitudinal assessment of vitamin B-6 status in southern adolescent girls. J Am Diet Assoc 1987; 87: 307–10.

8. Bender DA. Estrogens and vitamin B6-actions and interactions. World Rev Nutr Diet 1987; 51: 140–88.

9. Bhagavan HN. Interaction between vitamin B6 and drugs. In: Reynolds RD, Leklem JE, eds. Vitamin B6: Its role in health and disease. New York: Alan R. Liss 1985: 401–15.

10. Alvarez FG, Guntupalli KK. Isoniazid overdose: four case reports and review of the literature. Intensive Care Med 1995; 21: 641–4.

11. Chiang EP, Selhub J, Bagley PJ, Dallal G, Roubenoff R. Pyridoxine supplementation corrects vitamin B6 deficiency but does not improve inflammation in patients with rheumatoid arthritis. Arthritis Res Ther 2005; 7: R1404–R1411.

12. Leklem JE. Vitamin B-6 requirement and oral contraceptive use – a concern? J Nutr 1986; 116: 475–7.

13. Leklem JE, Brown RR, Rose DP, Linkswiler H, Arend RA. Metabolism of tryptophan and niacin in oral contraceptives users receiving controlled intakes of vitamin B6. Am J Clin Nutr 1975; 28: 146–56.

14. Talbott MC, Miller LT, Kerkvliet NI. Pyridoxine supplementation: effect on lymphocyte responses in elderly persons. Am J Clin Nutr 1987; 46: 659–64.

15. Wyatt K, Dimmock P, Jones P. Poor-quality studies suggest that vitamin B(6) use is beneficial in pre-menstrual syndrome. West J Med 2000; 172: 245.

16. Piazzini DB, Aprile I, Ferrara PE et al. A systematic review of conservative treatment of carpal tunnel syndrome. Clin Rehabil 2007; 21: 299–314.

17. Merete C, Falcon LM, Tucker KL. Vitamin B6 is associated with depressive symptomatology in Massachusetts elders. J Am Coll Nutr 2008; 27: 421–7.

18. Ellis JM, McCully KS. Prevention of myocardial infarction by vitamin B6. Res Commun Mol Pathol Pharmacol 1995; 89: 208–20.

19. Gdynia HJ, Muller T, Sperfeld AD et al. Severe sensorimotor neuropathy after intake of highest dosages of vitamin B6. Neuromuscul Disord 2008; 18: 156–8.

20. Institute of Medicine. Dietary Reference Intakes for Thiamin, Riboflavin, Niacin, Vitamin B6, Folate, Vitamin B12, Pantothenic Acid, Biotin, and Choline. USA: Washington, DC, National Academy Press 1998: 196–305.

21. Commission Directive 2008/100/EC of 28 October 2008 amending Council Directive 90/496/EEC on nutrition labelling for foodstuffs as regards recommended daily allowances, energy conversion factors and definitions. Official Journal of the European Union 2008; L 285–11.

22. Cotter PD, May A, Fitzsimons EJ et al. Late-onset X-linked sideroblastic anemia. Missense mutations in the erythroid delta-aminolevulinate synthase (ALAS2) gene in two pyridoxine-responsive patients initially diagnosed with acquired refractory anemia and ringed sideroblasts. J Clin Invest 1995; 96: 2090–6.

23. Danpure CJ, Jennings PR, Watts RW. Enzymological diagnosis of primary hyperoxaluria type 1 by measurement of hepatic alanine: glyoxylate aminotransferase activity. Lancet 1987; 1: 289–91.

24. Holmes RP. Pharmacological approaches in the treatment of primary hyperoxaluria. J Nephrol 1998;11 Suppl 1: 32–5.

25. Mikati MA, Trevathan E, Krishnamoorthy KS, Lombroso CT. Pyridoxine-dependent epilepsy: EEG investigations and long-term follow-up. Electroencephalogr Clin Neurophysiol 1991; 78: 215–21.

26. Plecko B, Stockler S. Vitamin B6 dependent seizures. Can J Neurol Sci 2009; 36 Suppl 2: S73–S77.

27. Larsson SC, Orsini N, Wolk A. Vitamin B6 and risk of colorectal cancer: a meta-analysis of prospective studies. JAMA 2010; 303: 1077–83.

28. Wu ET, Liang JT, Wu MS, Chang KC. Pyridoxamine prevents age-related aortic stiffening and vascular resistance in association with reduced collagen glycation. Exp Gerontol 2011; 46: 482–8.

29. Booth AA, Khalifah RG, Todd P, Hudson BG. In vitro kinetic studies of formation of antigenic advanced glycation end products (AGEs). Novel inhibition of post-Amadori glycation pathways. J Biol Chem 1997; 272: 5430–7.

30. Takatori A, Ishii Y, Itagaki S, Kyuwa S, Yoshikawa Y. Amelioration of the beta-cell dysfunction in diabetic APA hamsters by antioxidants and AGE inhibitor treatments. Diabetes Metab Res Rev 2004; 20: 211–8.

31. Voziyan PA, Metz TO, Baynes JW, Hudson BG. A post-Amadori inhibitor pyridoxamine also inhibits chemical modification of proteins by scavenging carbonyl intermediates of carbohydrate and lipid degradation. J Biol Chem 2002; 277: 3397–403.

32. den Heijer M. Vitamin B6 – Pyridoxine. In: Herrmann W, Obeid R, eds. Vitamins in the Prevention of Human Diseases. Berlin – New York: Walter de Gruyter 2011: 75–89.

33. Dierkes J, Hoffmann K, Klipstein-Grobusch K et al. Low plasma pyridoxal-5’phosphate and cardio-vascular disease risk in women: results from the Coronary Risk Factors for Atherosclerosis in Women Study. Am J Clin Nutr 2005; 81: 725–7.

34. Hron G, Lombardi R, Eichinger S, Lecchi A, Kyrle PA, Cattaneo M. Low vitamin B6 levels and the risk of recurrent venous thromboembolism. Haematologica 2007; 92: 1250–3.

35. Ray JG, Kearon C, Yi Q, Sheridan P, Lonn E. Homocysteine-lowering therapy and risk for venous thromboembolism: a randomized trial. Ann Intern Med 2007; 146: 761–7.

36. Selhub J, Troen A, Rosenberg IH. B vitamins and the aging brain. Nutr Rev 2010; 68 Suppl 2: S112–S118.

37. Smith AD, Smith SM, de Jager CA et al. Homocysteine-lowering by B vitamins slows the rate of accelerated brain atrophy in mild cognitive impairment: a randomized controlled trial. PLoS ONE 2010; 5: e12244.

38. Morris MS, Picciano MF, Jacques PF, Selhub J. Plasma pyridoxal 5’-phosphate in the US population: the National Health and Nutrition Examination Survey, 2003–2004. Am J Clin Nutr 2008; 87: 1446–54.

39. Hansen CM, Leklem JE, Miller LT. Vitamin B-6 status of women with a constant intake of vitamin B-6 changes with three levels of dietary protein. J Nutr 1996; 126: 1891–901.

40. Rybak ME, Jain RB, Pfeiffer CM. Clinical vitamin B6 analysis: an interlaboratory comparison of pyridoxal 5’-phosphate measurements in serum. Clin Chem 2005; 51: 1223–31.

41. Tarr JB, Tamura T, Stokstad EL. Availability of vitamin B6 and pantothenate in an average American diet in man. Am J Clin Nutr 1981; 34: 1328–37.

42. Combs GF. The vitamins: fundamental aspects in nutrition and health. San Diego: Elsevier 2008.

43. Percudani R, Peracchi A. The B6 database: a tool for the description and classification of vitamin B6-dependent enzymatic activities and of the corresponding protein families. BMC Bioinformatics 2009; 10: 273.

44. Brattstrom L, Israelsson B, Norrving B, et al. Impaired homocysteine metabolism in early-onset cerebral and peripheral occlusive arterial disease. Effects of pyridoxine and folic acid treatment. Atherosclerosis 1990; 81: 51–60.

45. Franken DG, Boers GH, Blom HJ, Trijbels FJ, Kloppenborg PW. Treatment of mild hyperhomocyste-inemia in vascular disease patients. Arterioscler Thromb 1994; 14: 465–70.

13.6 Betaine and choline

Wolfgang Herrmann, Susanne H. Kirsch, Rima Obeid

Betaine (N,N,N-Trimethylglycin)

The body’s betaine requirement is met through direct dietary intake and by oxidation of dietary choline. Betaine is abundant in cereals, especially wheat, spinach, chard and beets. Daily dietary intake of betaine is 100–300 mg. The intake differs between different populations and depends on the diet and the way the food is prepared (betaine is heat-stable). Physiologically, betaine has two functions: it acts as an osmolyte and is the source of methyl groups for numerous biochemical reactions. The differences in bio availability of betaine from different sources of food are small because betaine is strongly water-soluble and not bound to proteins. The mammalian intestine provides at least three transport systems for betaine /1/.

Choline

Choline is a water-soluble molecule often associated with the vitamin B complex. Choline is the main dietary source of methyl groups and is crucial for maintaining normal cell function and the structural integrity of cell membranes and transmembrane signaling. It has a direct influence on cholinergic neurotransmission and is required for normal muscle function and lipid transport from the liver /2/. In the prenatal period, choline is important for fetal brain development.

A choline-deficient diet causes liver injury, and choline deficiency increases the homocysteine concentration in blood, especially following methionine loading /3/. Choline deficiency can also cause developmental disorders, fetal brain injury, fatty liver or muscle injury. Moreover, deficiencies in choline, betaine or folate result in homocysteine accumulation. Hyperhomocysteinemia is an accepted risk factor for various diseases /3/. Furthermore, reduced plasma betaine concentrations have been correlated with lipid metabolism disorders, renal insufficiency and diabetes mellitus. The choline and folate metabolisms are closely interrelated. This explains the homocysteine-reducing effect of betaine and choline.

13.6.1 Indication

Homo cystinuria, pronounced hyperhomocysteinemia, alcoholic and non-alcoholic liver disease, diabetes mellitus.

13.6.2 Method of determination

Liquid chromatography-tandem mass spectrometry (LC-MS/MS)

The determination of betaine, choline [and dimethylglycine (DMG)] in blood plasma using liquid chromatography-tandem mass spectrometry (LC-MS/MS) provides maximum detection limit and test performance. Besides protein precipitation, no further derivatization steps are required for these methods.

The stable-isotope dilution ultra performance liquid chromatography-tandem mass spectrometry (UPLC-MS/MS) method enables concurrent measurement of betaine, choline, DMG and acetylcholine (ACh) based on simple sample processing /4/. EDTA plasma can be analyzed directly in this process.

High performance liquid chromatography (HPLC) and gas-liquid chromatography-mass spectrometry (GC-MS)

HPLC methods for betaine determination are widely used. However, these methods do not provide the sensitivity required for the quantification of serum and plasma concentrations. The first determination step using HPLC is the derivatization of betaine by alkylation of the carboxyl group to bind the functional group required for UV absorption or fluorescence. Common derivatization reagents are 2’-bromophenacyl bromide, 2’-bromophenacyl triflate and 2-naphthacyl triflate /9/.

Enzymatic methods

Enzymatic methods of betaine determination are based on reactions catalyzed by betaine-homocysteine methyl transferase (BHMT, EC 2.1.1.5) /10/. Choline determination is based on specific oxidation of choline by choline oxidase (EC 1.1.3.17) /11/. However, these methods provide little sensitivity and are time-consuming.

13.6.3 Specimen

EDTA plasma or urine: 1 mL

For (UP)LC-MS/MS methods, 100 μL of specimen are required. Betaine levels in plasma and serum are approximately identical. However, choline levels in serum are about twofold higher than in plasma (Fig. 13.6-1 – Betaine and choline levels in serum and EDTA plasma).

13.6.4 Reference interval

Tab. 13.6-1 – Fasting betaine and choline levels in EDTA plasma and urine shows guidance values for betaine, choline and DMG from the EDTA plasma and urine of several studies (LC-MS/MS method).

13.6.5 Clinical significance

Betaine and choline are important metabolites in the metabolism of lipids, neurotransmitters and amino acids. They are essential factors in homocysteine degradation and participate in the formation of S-adenosyl methionine (SAM), the universal methyl group donor. Insufficient supply can result in severe developmental disorders, fatty liver and muscle injury. Betaine and choline concentrations in plasma and urine are affected by diseases and environmental factors such as smoking and stress (Tab. 13.6-2 – Effect of diseases and lifestyle factors on plasma and urinary betaine and choline concentrations).

13.6.5.1 Clinical significance of betaine

Plasma betaine levels are remarkably stable for an extended period of time, even following significant changes in tissue concentrations. Plasma and urinary concentrations are characterized by great individuality. It has been shown that betaine concentrations are dependent on betaine intake (depending on the dose). In many cases, interpretation is exacerbated by minor changes observed under normal diet or major changes determined in supplementation studies. For this reason, plasma betaine determination is often not conclusive. Urinary betaine excretion expressed as ratio to creatinine is potentially more significant because it depends on the diet only to a small extent. Betaine levels in plasma and urine do not seem to be correlated. Urinary betaine concentrations are not correlated with age /8/. Excretion is very high in neonates, decreases within a short period of time and remains constant thereafter (betaine clearance approximately 1.5–3%). The concentration in adults seems to increase by about 1% a year. Betaine concentrations in men are about 15% higher than in women. During pregnancy, plasma betaine decreases by the same magnitude as homocysteine and related metabolites and reaches a plateau in gestation week 20 /12/.

Diseases associated with betaine deficiency are listed in Tab. 13.6-3 – Association between betaine deficiency and diseases.

13.6.5.2 Clinical significance of choline

The plasma choline status is influenced by the diet. It has been reported that choline levels were higher in individuals after a meal than in the fasting condition (Fig. 13.6-2 – Betaine and choline concentrations in EDTA plasma following overnight fasting). The marked increase in serum choline levels has been associated with a release of choline from phospholipids by serum cholinesterases (Fig. 13.6-2). These will be inhibited if EDTA, for example, is added. Choline levels are higher in men than in women and higher in older than in younger women /7/. Elevated choline concentrations (whole-blood and plasma) have been discussed as biomarkers for the risk of ischemic heart disease, particularly if no cardiac troponins are detectable on admission of the patient /13/. Choline concentrations decrease during physical activity. Several investigations have shown that plasma free choline decreases significantly in marathon runners /14/.

Choline deficiency causes the development of fatty liver and liver injury as it leads to phosphatidylcholine deficiency, which, in turn, impairs the removal of triglycerides (formed in the liver) via very-low-density lipoprotein (VLDL) /15/. Choline deficiency can also result in muscle injury with elevated creatine kinase activities /16/. Choline deficiency has also been linked to DNA damage and the death of peripheral lymphocytes /17/. Choline deficits cause hyperhomocysteinemia and increased cardiovascular risk. Moreover, choline deficiency during pregnancy results in increased risk of neural tube defects /12/. Choline deficiency during the fetal period is associated with later learning disorders or attention deficits /18/. An increased mortality risk and increased risk of breast cancer have been reported in women with low choline intake /19/. Diseases associated with choline deficiency are listed in Tab. 13.6-4 – Association between choline deficiency and diseases.

13.6.5.3 Clinical significance of N,N-dimethylglycine (DMG)

DMG is a potential biomarker for an abnormal C1 cycle (exclusively formed via betaine-homocysteine methyl transferase). Plasma and urinary DMG concentrations show smaller intraindividual differences than betaine. DMG concentrations are elevated in folate deficiency because of increased homocysteine methylation with betaine as the methyl donor. Betaine administration results in a mild, temporary increase in plasma and urinary DMG concentrations. Clearance occurs only partly by urinary excretion. DMG is primarily cleared by the mitochondrial enzyme dimethylglycine dehydrogenase (DMGDH, EC 1.5.99.2) The DMG/betaine ratio is considered to be more conclusive than the DMG concentration because the plasma DMG concentration is partly affected by betaine intake. There is no evidence that urinary DMG concentration provides additional clinical information.

13.6.6 Comments and problems

Naturally occurring cholinesterases must be inhibited immediately after sampling. Carbamates, EDTA (Ca++ chelating agents) or organophosphates are suited for this purpose. Choline concentration in serum is significantly higher than in plasma (Fig. 13.6-1 – Betaine and choline concentrations in serum and EDTA plasma), presumably because of the constant, serum cholinesterase-induced release of choline from phospholipids in serum. Citrate plasma has the lowest choline and betaine concentrations. EDTA plasma is recommended as specimen for betaine, choline and DMG determination.

The concentrations of betaine, choline and DMG in deep-frozen (–70 °C) plasma or urine remain stable for at least six months. Moreover, no changes in the plasma betaine, choline and DMG concentrations have been found even after several freezing-defrosting cycles.

Fasting plasma

Contrary to the choline concentration, the plasma betaine concentration is not influenced by fasting conditions. Plasma choline concentrations are higher under non-fasting than under fasting conditions. However, there is a strong correlation between fasting and non-fasting choline concentrations (Fig. 13.6-2 – Betaine and choline concentrations in EDTA plasma following overnight fasting and non-fasting). Therefore, 12 h fasting plasma is generally recommended as specimen.

Method of determination

The chemical determination of betaine is problematic because of the molecular structure. As it is strongly water-soluble, betaine is difficult to extract from aqueous solutions. An early method of determination utilized the low solubility of quaternary ammonium iodides in betaine /30/.

Betaine has a low-reactive amino group, and its likewise inactive carboxyl group is not as readily derivatized as other carboxyl compounds. Betaine derivatives present as quaternary ammonium cations can be separated using ion exchange chromatography /9/. Derivatization often also includes betaine metabolites, such as DMG. Derivatives may be formed that are alkylated both at the nitrogen and the carboxyl group, thus carrying two chromophores, are eluted too early by the silica gel column and escape reliable quantification /9/. It was not until modern chromatography/mass spectrometry procedures were introduced that significant methodological improvements were achieved.

Reference interval

No generally valid reference interval has been defined for betaine, choline and DMG in plasma or urine to date. Data from studies is either based on a small number of participants or a subpopulation. In general, however, the reference ranges of these studies coincide with each other.

Basically, the following applies to plasma and serum:

  • Betaine concentrations are tendentially higher in men than in women
  • DMG concentrations are below 10 μmol/L
  • Choline concentrations are tendentially higher under non-fasting conditions.

Determination in 24 h urine does not provide any further clinical information because excretion is stable across the day. In random urine specimens, the concentration is related to excretion per gram (g) of creatinine.

13.6.7 Pathophysiology

Betaine

Betaine has two important roles in human physiology. One is as an osmolyte to assist cellular volume regulation. The other one is as a methyl group donor for the betaine-homocysteine methyl transferase (BHMT)-catalyzed remethylation of homocysteine to methionine.

Rat tissue betaine concentrations are higher than blood plasma concentrations in almost all organs, especially the renal medulla /39/. There are various transporters that are osmoregulated, but none are specific for betaine, including the betaine GABA transporter (BGT-1) /40/ or the carnitine receptor OCTN2 /41/. As with most osmolytes, betaine is both a compensatory and counteracting solute /42/. It enhances protein stability and is particularly effective in counteracting the denaturing effect of urea /43/.

The two functions of betaine (osmolyte and methyl group donor) interact because BHMT is osmoregulated, with high tonicity reducing its expression so that betaine metabolism (and hence the mobilization of methyl groups) decreases when high osmolyte concentrations need to be maintained.

BHMT metabolism provides about 50% of the liver’s homocysteine methylation capacity and, therefore, is of extraordinary significance for the maintenance of methionine and SAM concentrations. In the human body, BHMT is expressed in the liver, cortex of the kidney, optic lens and other cells. Mammals express a different gene product referred to as BHMT2. This designation was chosen because a gene product was identified, which did not use betaine as a substrate and where the amino acid composition shared 73% identity with the protein BHMT. Instead of betaine it uses S-methyl methionine (SMM) as a methyl donor for the methylation of homocysteine /44/.

Plasma betaine and urinary betaine excretion are determinants of plasma homocysteine /12/. The role of betaine as a methyl group donor is especially important if folate supply is limited. BHMT gene expression and activity can be increased if betaine and choline are amply available or in dietary methionine deficiency /45/. DMG, the de methylation product of betaine, inhibits BHMT.

DMG is exclusively formed via the BHMT pathway and thus reflects BHMT-mediated homocysteine methylation. Via oxidative de methylation steps, the mitochondrial enzyme DMGDH transfers the methyl group from DMG as a one-carbon unit to tetrahydrofolate, resulting in the formation of 5,10-methylenetetrahydrofolate and sarcosine. Sarcosine, in turn, can be de methylated to form glycine. On the other hand, sarcosine can be formed from glycine via the enzyme glycine-N-methyl transferase (GNMT, EC 2.1.1.20) and SAM.

Choline

Eggs and liver are important dietary sources of choline. Moreover, betaine-containing foods can also have a choline-saving effect. The average daily dietary intake of choline is 8.4 mg/kg of body weight in men and 6.7 mg/kg of body weight in women /46/. Another important source of choline is endogenous synthesis in the liver via the enzyme phosphatidylethanolamine N-methyl transferase (PEMT, EC 2.1.1.17) (Fig. 13.1-2 – Homocysteine metabolism). PEMT needs 3 SAM molecules to convert phosphatidylethanolamine (PE) to phosphatidylcholine (PC). PC is either incorporated into the cell membrane or broken down for choline generation /2/.

Dietary choline (free or as ester) is released by pancreatic enzymes. Cholinesters include PC, phospho choline, glycerophosphocholine and sphingomyelin (SM) /2/. Choline is resorbed in the small intestine, enters portal circulation as free choline and is mainly taken up by the liver, while fat-soluble PC and SM are transported to the liver by the lymph and bypass. Therefore, the different forms of choline differ in bio availability. Choline is stored in all tissues, with uptake by the liver, kidneys, mammary glands, placenta and brain being of special significance. Choline stored in the kidney is primarily required for the formation of betaine and glycerophosphocholine. Both are organic osmolytes that promote water resorption from the renal tubule by the kidney /47/.

A small amount of choline is converted to acetylcholine (ACh) by means of choline acetyl transferase. High concentrations of choline acetyl transferase are found in the cholinergic nerve endings and in the placenta. In the brain, the availability of choline governs the rate of ACh synthesis /48/. The choline taken up by the brain may first enter a storage pool (perhaps the PC in membranes) before being converted to ACh. This pool is very important for the regeneration of cholinergic neurons and brain function. Brains of patients with Alzheimer’s disease were found to show abnormalities in phospholipid metabolism /49/.

One of the main functions of choline is the synthesis of membrane phospholipids. PC, from which SM can be formed, represents the key phospholipid class in membranes. Both PC and SM play a major role in signal transduction because they function as sources for “second messengers” /2/. Moreover, PC is also required for the formation of lipoproteins, which transport the triglycerides formed by the liver to other tissues. PC biosynthesis follows two metabolic pathways, the cytidine diphosphate-choline (CDP-choline) pathway (Kennedy pathway) and the PEMT pathway. Via the first pathway, choline is phosphorylated to form phospho choline and then converted further to CDP choline. CDP choline condenses with diacylglycerol to form PC. Alternatively, PC can be formed via the PEMT pathway using SAM as methyl donor. The PEMT pathway provides approximately 30% of the PC synthesized by the liver. Evidence of this pathway has also been provided in other tissues, including the brain and mammary glands. About 70% of the hepatic PC are formed via the CDP choline pathway. These two metabolic pathways produce different physical profiles of PC species, thus generating different signal molecules /50/.

In the inner mitochondrial membrane, choline can be oxidized to form betaine aldehyde, which is further oxidized to betaine via choline dehydrogenase (CHDH, ED 1.1.99.1) and betaine aldehyde dehydrogenase (BADH, EC 1.2.1.8). Via this pathway, choline supplies methyl groups for homocysteine re methylation. The liver and kidney are the major sites of choline oxidation. As betaine cannot be reduced to choline and choline is consumed for methylation reactions in the oxidative methylation pathway, the bio availability of choline for the alternative PC synthesis pathway is reduced.

References

1. Thwaites DT, Anderson CM. Deciphering the mechanisms of intestinal imino (and amino) acid transport: the redemption of SLC36A1. Biochim Biophys Acta 2007; 1768: 179–97.

2. Zeisel SH, Blusztajn JK. Choline and human nutrition. Annu Rev Nutr 1994; 14: 269–96.

3. da Costa KA, Gaffney CE, Fischer LM, Zeisel SH. Choline deficiency in mice and humans is associated with increased plasma homocysteine concentration after a methionine load. Am J Clin Nutr 2005; 81: 440–4.

4. Kirsch SH, Herrmann W, Rabagny Y, Obeid R. Quantification of acetylcholine, choline, betaine, and dimethylglycine in human plasma and urine using stable-isotope dilution ultra performance liquid chromatography-tandem mass spectrometry. J Chromatogr B Analyt Technol Biomed Life Sci 2010; 878: 3338–44.

5. Lever M, Slow S. The clinical significance of betaine, an osmolyte with a key role in methyl group metabolism. Clin Biochem 2010; 43: 732–44.

6. Holm PI, Ueland PM, Kvalheim G, Lien EA. Determination of choline, betaine, and dimethylglycine in plasma by a high-throughput method based on normal-phase chromatography-tandem mass spectrometry. Clin Chem 2003; 49: 286–94.

7. Konstantinova SV, Tell GS, Vollset SE, Nygard O, Bleie O, Ueland PM. Divergent associations of plasma choline and betaine with components of metabolic syndrome in middle age and elderly men and women. J Nutr 2008; 138: 914–20.

8. Lever M, Atkinson W, Slow S, Chambers ST, George PM. Plasma and urine betaine and dimethylglycine variation in healthy young male subjects. Clin Biochem 2009; 42: 706–12.

9. Storer MK, McEntyre CJ, Lever M. Separation of cationic aracyl derivatives of betaines and related compounds. J Chromatogr A 2006; 1104: 263–71.

10. Martin JJ, Finkelstein JD. Enzymatic determination of betaine in rat tissues. Anal Biochem 1981; 111: 72–6.

11. Hise MK, Mansbach CM. Determination of intracellular choline levels by an enzymatic assay. Anal Biochem 1983; 135: 78–82.

12. Velzing-Aarts FV, Holm PI, Fokkema MR, van der Dijs FP, Ueland PM, Muskiet FA. Plasma choline and betaine and their relation to plasma homocysteine in normal pregnancy. Am J Clin Nutr 2005; 81: 1383–9.

13. Apple FS, Wu AH, Mair J et al. Future biomarkers for detection of ischemia and risk stratification in acute coronary syndrome. Clin Chem 2005; 51: 810–24.

14. Buchman AL, Jenden D, Roch M. Plasma free, phospholipid-bound and urinary free choline all decrease during a marathon run and may be associated with impaired performance. J Am Coll Nutr 1999; 18: 598–601.

15. Zeisel SH. Choline: an essential nutrient for humans. Nutrition 2000; 16: 669–71.

16. da Costa KA, Badea M, Fischer LM, Zeisel SH. Elevated serum creatine phosphokinase in choline-deficient humans: mechanistic studies in C2C12 mouse myoblasts. Am J Clin Nutr 2004; 80: 163–70.

17. da Costa KA, Niculescu MD, Craciunescu CN, Fischer LM, Zeisel SH. Choline deficiency increases lymphocyte apoptosis and DNA damage in humans. Am J Clin Nutr 2006; 84: 88–94.

18. Meck WH, Smith RA, Williams CL. Organizational changes in cholinergic activity and enhanced visuospatial memory as a function of choline administered prenatally or postnatally or both. Behav Neurosci 1989; 103: 1234–41.

19. Xu X, Gammon MD, Zeisel SH et al. High intakes of choline and betaine reduce breast cancer mortality in a population-based study. FASEB J 2009; 23: 4022–8.

20. Schwab U, Torronen A, Toppinen L et al. Betaine supplementation decreases plasma homocysteine concentrations but does not affect body weight, body composition, or resting energy expenditure in human subjects. Am J Clin Nutr 2002; 76: 961–7.

21. Sparks JD, Collins HL, Chirieac DV et al. Hepatic very-low-density lipoprotein and apolipoprotein B production are increased following in vivo induction of betaine-homocysteine S-methyltransferase. Biochem J 2006; 395: 363–71.

22. Huang QC, Xu ZR, Han XY, LiWF. Effect of diatary betaine supplementation on lipogenic enzyme activities and fatty acid synthase mRNA expression in finishing pigs. Animal Feed Sci Technol 2008; i 140: 365–75.

23. Lever M, Atkinson W, George PM, Chambers ST. Sex differences in the control of plasma concentrations and urinary excretion of glycine betaine in patients attending a lipid disorders clinic. Clin Biochem 2007; 40: 1225–31.

24. Lever M, Sizeland PC, Bason LM, Hayman CM, Robson RA, Chambers ST. Abnormal glycine betaine content of the blood and urine of diabetic and renal patients. Clin Chim Acta 1994; 230: 69–79.

25. Lever M, George PM, Slow S, et al. Fibrate may cause an abnormal urinary betaine loss which is associated with elevations in plasma homocysteine. Cardiovasc Drug Ther 2009; 23: 395–401.

26. Mudd SH, Brosnan JT, Brosnan ME et al. Methyl balance and transmethylation fluxes in humans. Am J Clin Nutr 2007; 85: 19–25.

27. Detopoulou P, Panagiotakos DB, Antonopoulou S, Pitsavos C, Stefanadis C. Dietary choline and betaine intakes in relation to concentrations of inflammatory markers in healthy adults: the ATTICA study. Am J Clin Nutr 2008; 87: 424–30.

28. Dalmeijer GW, Olthof MR, Verhoef P, Bots ML, van der Schouw YT. Prospective study on dietary intakes of folate, betaine, and choline and cardiovascular disease risk in women. Eur J Clin Nutr 2008; 62: 386–94.

29. Graf D, Kurz AK, Reinehr R, Fischer R, Kircheis G, Haussinger D. Prevention of bile acid-induced apoptosis by betaine in rat liver. Hepatology 2002; 36: 829–39.

30. Barak AJ, Tuma DJ. Determination of choline, phosphorylcholine, and betaine. Methods Enzymol 1981; 72: 287–92.

31. Chiuve SE, Giovanni EL, Hankinson SE, et al. The association between betaine and choline intakes and the plasma concentrations of homocysteine in women. Am J Clin Nutr 2007; 86: 1073–81.

32. Shaw GM, Finnell RH, Blom HJ et al. Choline and risk of neural tube defects in a folate-fortified population. Epidemiology 2009; 20: 714–9.

33. Jones PA, Baylin SB. The epigenomics of cancer. Cell 2007; 128: 683–92.

34. Cho E, Willett WC, Colditz GA et al. Dietary choline and betaine and the risk of distal colorectal adenoma in women. J Natl Cancer Inst 2007; 99: 1224–31.

35. Garner SC, Mar MH, Zeisel SH. Choline distribution and metabolism in pregnant rats and fetuses are influenced by the choline content of the maternal diet. J Nutr 1995; 125: 2851–8.

36. Zeisel SH, Zola T, daCosta KA, Pomfret EA. Effect of choline deficiency on S-adenosylmethionine and methionine concentrations in rat liver. Biochem J 1989; 259: 725–9.

37. Niculescu MD, Craciunescu CN, Zeisel SH. Dietary choline deficiency alters global and gene-specific DNA methylation in the developing hippocampus of mouse fetal brains. FASEB J 2006; 20: 43–9.

38. Davison JM, Mellott TJ, Kovacheva VP, Blusztajn JK. Gestational choline supply regulates methylation of histone H3, expression of histone methyltransferases G9a (Kmt1c) and Suv39h1 (Kmt1a), and DNA methylation of their genes in rat fetal liver and brain. J Biol Chem 2009; 284: 1982–9.

39. Slow S, Lever M, Chambers ST, George PM. Plasma dependent and independent accumulation of betaine in male and female rat tissues. Physiol Res 2009; 58: 403–10.

40. Kempson SA, Montrose MH. Osmotic regulation of renal betaine transport: transcription and beyond. Pflugers Arch 2004; 449: 227–34.

41. Pochini L, Oppedisano F, Indiveri C, et al. Reconstitution into liposomes and functional characterization of the carnitine transportes from renal cell plasma membrane. Biochem Biophyrs Acta 2004; 1661: 78–86.

42. Gilles R. “Compensatory” organic osmolytes in high osmolarity and dehydration stresses: history and perspectives. Comp Biochem Physiol A Physiol 1997; 117: 279–90.

43. Burg MB, Kwon ED, Peters EM. Glycerophosphocholine and betaine counteract the effect of urea on pyruvate kinase. Kidney Int Suppl 1996; 57: S100–S104.

44. Szegedi SS, Castro CC, Koutmos M, Garrow TA. Betaine-homocysteine S-methyltransferase-2 is an S-methylmethionine-homocysteine methyltransferase. J Biol Chem 2008; 283: 8939–45.

45. Finkelstein JD, Martin JJ, Harris BJ, Kyle WE. Regulation of hepatic betaine-homocysteine methyltransferase by dietary betaine. J Nutr 1983; 113: 519–21.

46. Fischer LM, Scearce JA, Mar MH et al. Ad libitum choline intake in healthy individuals meets or exceeds the proposed adequate intake level. J Nutr 2005; 135: 826–9.

47. Burg MB, Ferraris JD. Intracellular organic osmolytes: function and regulation. J Biol Chem 2008; 283: 7309–13.

48. Cohen EL, Wurtman RJ. Brain acetylcholine: increase after systemic choline administration. Life Sci 1975; 16: 1095–102.

49. Nitsch RM, Blusztajn JK, Pittas AG, Slack BE, Growdon JH, Wurtman RJ. Evidence for a membrane defect in Alzheimer disease brain. Proc Natl Acad Sci U S A 1992; 89: 1671–5.

50. DeLong CJ, Shen YJ, Thomas MJ, Cui Z. Molecular distinction of phosphatidylcholine synthesis between the CDP-choline pathway and phosphatidylethanolamine methylation pathway. J Biol Chem 1999; 274: 29683–8.

Table 13.1-1 Homocysteine level in fasting plasma and following oral methionine load

Concentration*
(μmol/L)

Consequences
Assessment

Indicative of

< 10

No action required

Favorable

 

10–12

Action required in individuals with increased risk of degenerative disease

Tolerable

 

> 12–30

Action required, normalization of homocysteine level

Moderately elevated

B vitamin deficiency, MTHFR mutation, renal dysfunction

> 30–100

High health risk

Significantly elevated

Severe vitamin deficiencies, renal dysfunction, heterozygous mutations of enzymes (e.g., CBS)

> 100

High health risk

Severely elevated

Homozygous mutations of enzymes (e.g., CBS or methionine synthase)

Methionine load (oral ingestion of 0.1 g/kg of body weight in 200 mL fruit juice)

After 4–6 h

< 38

 

Pathologic in vitamin B6 deficiency or CBS deficiency

* fasting plasma; CBS, cystathionine-β-synthase; MS, methionine synthase; MTHFR, 5,10-methylenetetrahydrofolate reductase.

Table 13.1-2 Causes of changes in homocysteine concentration

Causes and/or factors

Mechanism

Old age > 60 yrs

Male gender

Postmenopause

Estrogen deficiency

Smoking

Interference with vitamin B6, B12 and folate, redox

Coffee

Vitamin B6 antagonist (caffeine), methyl group requirement

Alcohol

Interference with vitamin B6, B12 and folate, enzyme inhibition

Vegetarian diet

Reduced dietary B12 intake

Renal dysfunction

Impaired remethylation

B vitamin deficiency

Malabsorption of B vitamins, reduced intake, renal insufficiency

Thyroid dysfunction

 

Proliferative diseases

 

  • Psoriasis

Enzyme induction

  • ALL (acute lymphocytic leukemia)

Cell proliferation

  • RA (rheumatoid arthritis)

Cell proliferation

Theophylline

Vitamin B6 antagonist, inhibits pyridoxal kinase

Nitrous oxide (N2O)

Cobalt oxidation, cobalamin and MS inactivation

Fibrates

PPARa activation, renal function

Niacin

Vitamin B6 antagonist, inhibits pyridoxal kinase

Cholestipol/colestyramine

Folate and cobalamin resorption disorder

Methotrexate

Inhibits dihydrofolate reductase, folate antagonist

Trimethoprime

Inhibits dihydrofolate reductase

Postmenopausal hormone replacement

Estrogen effect

Anti-epileptic drugs

Folate antagonism, enzyme modulation

Metformin

Vitamin B12 homeostasis, Ca2+ binding

Omeprazole

Vitamin B12 malabsorption

L-DOPA

Increased methyl group requirement

D-penicillamine, N-acetylcysteine

Disulfide exchange

Table 13.1-3 Mutations associated with hyperhomocysteinemia

Affected enzyme

Name of mutation

5,10 methylenetetrahydrofolate

C677T or A1298C

Methionine synthase

A2756G (G919D)

Methionine synthase reductase

A66G

Cystathionine-β-synthase

68 bp D/I insertion

Transcobalamin

Pro259Arg

Thymidylate synthase

Tandem repeat polymorphism in the 5’ non-translated region

Table 13.1-4 Causes and diseases possibly associated with hyperhomocysteinemia

Clinical and laboratory findings

Homozygous homocystinuria

Classic homocystinuria is based on a homozygous defect in the cystathionine-β-synthase (CBS) gene with very strong loss of enzyme activity (prevalence in Europe 1 : 40,000 to 1  332,000) /13/. Homocysteine accumulates in the blood and cerebrospinal fluid and is excreted in urine as homocystin. Plasma homocysteine is continuously between 100 and 250 μmol/L. Besides somatic abnormalities (genu valgum, arachnodactyly, pes cavus, kyphoscoliosis), affected patients are mentally retarded. Recurrent thromboembolic events, lens dislocation and seizures occur. Premature atherosclerosis is inevitable. In many cases, patients suffer from myocardial infarction before 25 years of age and their life expectancy is below 50 years. In an analysis comprising 629 homocystinuria patients with homozygous CBS deficiency, thromboembolic events were detected in 25% of the patients /13/.

Treatment consists of dietary methionine restriction and the administration of high doses of vitamin B6 in order to enhance the still existing residual enzyme activity. Moreover, dietary cysteine fortification is recommended because this amino acid becomes essential due to reduced methionine degradation.

Besides CBS gene defects, homocystinuria has also been reported in severe deficits in MTHFR activity or in cobalamin C/D defects. Other mutations possibly affecting homocysteine metabolism are very rare and their clinical relevance has been little investigated to date /14/.

Heterozygous homocystinuria

The prevalence of heterozygous homocystinuria in the European population has been estimated at 1 : 70 to 1 : 290. Thus, it is among the common causes of premature atherosclerosis and one of the main causes of recurrent thromboses /15/. The methionine loading test is helpful for diagnosis in this context.

MTHFR C677T mutation

The enzyme MTHFR causes the irreversible reduction of 5,10-methylene-THF to 5-methyl-THF. Reduced MTHFR activity due to 677 C>T point mutation leads to moderately elevated homocysteine concentration in blood plasma, especially, if serum folate concentration is low or within the lower normal range /15/. The 677 C>T mutation results in an amino acid exchange from alanine to valine at position 222 of the MTHFR protein, which thus becomes thermally unstable and less active. The enzyme is stabilized by folate. Therefore, additional folate deficiency results in marked restriction of the enzyme’s function and moderate hyperhomocysteinemia can develop.

In Germany, Austria and Switzerland, 5–15% of the population are homozygous carriers of the nucleotide point mutation at position 677. Enzyme activity of the affected individuals is reduced by about 70%. Accordingly, homozygous carriers react to folate deficiency especially sensitive with an increase in homocysteine by about 25% (corresponding to approximately 2.6 μmol/L) /16/. Meta analyses comprising a sufficiently high number of cases find an association between homozygous genotype and risk increase by 16–23% for degenerative cardiovascular diseases that is explained with elevated homocysteine and/or folate deficiency /2/. Meta analyses have also provided evidence that the TT genotype of MTHFR 677 C>T polymorphism increases the risk of gastrointestinal tumors; however, no association with the CC genotype of MTHFR 1298 A>C polymorphism has been found /17/. Association between the MTHFR 677 TT genotype and the risk of neurodegenerative disease has also been reported /18/.

Vitamin deficiency

Vitamin under supply is by far the most common cause of hyperhomocysteinemia. Under supply can ensue from inadequate supply, reduced gastrointestinal absorption, increased consumption and interactions. Individuals with one-sided eating habits (vegetarians), older adults, pregnant women, patients with renal disease, malabsorption (inflammatory bowel disease) and tumor patients belong to the risk groups for vitamin deficits of clinical relevance. Moreover, alcohol abuse and the intake of certain drugs can lead to vitamin deficiency (Tab. 13.1-2 – Causes of homocysteine changes).

Folate deficiency, the most common vitamin deficiency in Europe, is promoted by lack of fresh fruit and vegetables. Up to 90% of the folates can be lost in the processing of cereal products and other food. Folates are also lost because folate is sensitive to heat, storage and light. The recommendations to eat five portions of fruit and vegetables a day (600–700 g) are hardly realizable. The required daily intake of about 400 μg folate equivalents is far from being reached in many cases. In most European countries, the average daily dietary folate intake is currently far below 300 μg /19/. It is specified as 230 to 280 μg/d on average. According to the National Consumption Study (NVS II) /20/, the median intake of folate equivalents in Germany is 283 μg/d for men and 252 μg/d for women /20/. Accordingly, a substantial part of the normal population does not reach the required amount of folate through their natural diet alone /21/.

Vitamin B12 is usually taken in sufficient amounts. Nevertheless, problems in risk groups may arise. In many cases, vitamin B12 deficiency in older people can reflect inadequate resorption due to age-related reduction in gastric acid secretion and/or mild pH increase or lack of intrinsic factor and can affect 30–40 % of older adults /22/. Vitamin B12 can only be synthesized by bacteria. Only animal source foods (fish, meat, eggs and dairy products) are good B12 sources. Contrary to folate, cobalamin is a relatively stable vitamin and is practically not destroyed by food preparation.

Vitamin B6 occurs primarily in meat, milk, cereals, potatoes, fruit and vegetables. The results from the Framingham Heart Study show a significant increase in homocysteine concentration at vitamin B6 intakes of less than approximately 1.4 mg/d /23/.

Chronic kidney disease

Homocysteine levels in renal patients between 20 and 80 μmol/L /24/ are linked with increased cardiovascular morbidity and mortality /25/. The risk of atherothrombotic events (fatal and non-fatal) in dialysis patients with marked hyperhomocysteinemia (> 38 μmol/L) was found to be 8-fold compared to those without hyperhomocysteinemia /26/.

Most of the re methylation of plasma homocysteine to methionine takes place in the kidney; hence, the kidney plays an important role in homocysteine clearance (only less than 1% are eliminated by urinary excretion). Impaired re methylation of homocysteine to methionine is considered to be one of the main causes of the high prevalence of hyperhomocysteinemia in renal patients /27/. It was shown by means of stable isotope technique that re methylation is significantly decreased in dialysis patients, whereas transsulfuration is not /28/.

Besides renal dysfunction causing hyperhomocysteinemia, renal patients also have B vitamin deficiency (B12, B6, folate), although the serum and/or plasma concentration of these vitamins is often found to be within the normal interval /24/. Homocysteine and methylmalonic acid (MMA) as metabolic biomarkers of folate and vitamin B12 deficiency are markedly elevated. The administration of folic acid and vitamin B12, in therapeutic doses significantly reduces the plasma homocysteine concentration /29/. Hence, the decrease in homocysteine in renal patients by B vitamins is, at least partly, explained with improved homocysteine re methylation. The reduction in MMA and homocysteine in renal patients following B vitamin substitution proves that this therapy improves the re methylation of homocysteine to methionine /29/.

Neurodegenerative disease

Dementia syndromes, organic depressive disorders, cerebral seizures as well as myelopathy and polyneuropathy have been reported as manifestations of disorders of the vitamins involved in methionine metabolism as co factors /30/. Silent stroke and cerebral atrophy are also correlated to hyperhomocysteinemia and B vitamin status /313233/.

Cognitive impairment

A positive correlation between the concentration of homocysteine or B vitamins (B12, B6, folate) and cognitive capacity was detected in healthy older adults (Fig. 13.1-3 – Dependency of cognitive performance in the elderly on homocysteine level/34/.

A negative correlation between the homocysteine level and cognitive function was found in both healthy older adults and depressive patients /35/. The fasting plasma homocysteine level has been classified as independent predictor for the decline in cognitive function in older adults (Fig. 13.1-4 – Cumulative incidence of dementia depending on basal homocysteine concentration/36/. Homocysteine as biomarker of vitamin B12 and folate status shows better correlation with cognitive impairment than MMA, which is considered as a sensitive and specific biomarker of vitamin B12 deficiency.

Low SAM and elevated SAH and/or homocysteine were found in the cerebrospinal fluid of patients suffering from dementia /37/. Moreover, low folate in cerebrospinal fluid has also been reported from patients with Alzheimer’s dementia /38/. Therefore, abnormal homocysteine metabolism in the brain can have negative effects on various metabolic pathways in the brain, increasing the risk of dementia. In the Hordaland Study, the baseline plasma homocysteine level was predictive of the decline in memory during the observation period of 6 years /39/.

Furthermore, it was found that homocysteine accounts for approximately 10% of the variance in cognitive tests or various cognitive functions (such as language skills, conceptional thinking, simple motoric and psychomotoric speed, executive functions, verbal memory and learning, visual memory and others) /40/. In a prospective study, low serum folate concentrations were also associated with lower cognitive performance /41/.

If homocysteine plays a causal role in the development of dementia, a reduction in its concentration should be associated with a delay in the development of dementia, reduced risk of stroke and improved morbidity of these diseases. There is a lack of sensitivity of the cognitive tests used in many studies. Tests quantifying cerebral damage are better suited than cognitive tests. Various, but not all, vitamin intervention studies have reported improvements of a number of cognitive parameters, reductions in risk of stroke or morbidity for other diseases. In a 24-month supplementation study with vitamin B12, vitamin B6 and folic acid on elderly individuals with mild cognitive impairment, markedly delayed brain atrophy was found compared to the placebo cohort /32/. The rate of brain atrophy in individuals with higher homocysteine levels at the start of the study was higher than in those with lower baseline concentration (Fig. 13.1-6 – Higher baseline homocysteine concentration are associated with a major reduction in brain atrophy/32/. In another randomized, placebo-controlled study, patients were treated with vitamin B6 and followed for 2 years. A tendential improvement of cerebral and cerebrovascular indices was found using magnetic resonance imaging and magnetic resonance angiography /33/.

Vascular and Alzheimer’s dementia

In patients with vascular dementia, hyperhomocysteinemia as a risk factor for subcortical vascular encephalopathy ranks before established risk factors (smoking, hyperlipidemia, hypertension). According to the data from the Framingham Study, a homocysteine level above 14 μmol/L doubles the risk of Alzheimer’s dementia /34/.

Low vitamin B12 level in serum and/or cerebrospinal fluid was found in patients suffering from dementia and is under discussion as an important cause of the development of neuropsychiatric diseases /42/. An early diagnosis of vitamin B12 deficiency could probably avoid irreversible neurological damage and/or delay its progression. Vitamin B12 deficiency-related elevated homocysteine concentrations in the central nervous system can result in damage of the vascular endothelium, which is also considered a cause of the development of dementia or can trigger a stroke. It was shown in a prospective study on elderly individuals that low folate (below 10 nmol/L) or vitamin B12 (below 150 pmol/L) develop Alzheimer’s dementia twice as often as individuals with higher B vitamin concentrations /43/.

Table 13.2-1 Homocysteine concentrations in other body fluids

Body fluid

Homocysteine concentration and patient cohort

Cerebrospinal fluid (CSF), nmol/L

Median: 90 (60–160)

Neurological patients without severe disease /9/

Mean ± SD: 84.9 ± 24.5

n = 15;

patients without neurological disease /9/

Mean ± SD: 110.6 ± 31.6

n = 18;

patients with Alzheimer’s dementia /9/

Amniotic fluid (17.1 ± 1.2 GW), μmol/L

Mean ± SD: 1.04 ± 0.72

(95% CI 0.43–2.41)

All newborns /10/

Mean: 1.01

(95% CI 0.94–1.08)

Mature newborns /10/

Mean: 1.29

(95% CI 1.05–1.51)

SGA /10/

Ejaculate, μmol/L

Mean ± SD: 5.9 ± 3.1

Fertile men /11/

Mean± SD: 5.8 ± 3.4

Idiopathic subfertile /11/

Mean ± SD: 4.2 ± 3.0

Subfertile /11/

MV, mean value; CI, confidence interval; SGA, small for gestational age; SD, standard deviation; GW, gestation week.

Table 13.3-1 Populations at risk for vitamin B12 deficiency /3/

Population-at-risk

Comment

Vegetarian, vegan and macrobiotic diets

Low dietary vitamin B12 intake

Hyperhomocysteinemia, homocystinuria

B vitamin deficiency, older adults, dialysis patients

Neonates and breastfed infants of vegetarian mothers

Low dietary vitamin B12 intake with breast milk

Older adults

Pernicious anemia, achlorhydria, gastrointestinal diseases causing cobalamin malabsorption (gastrointestinal surgery, gastritis, atrophy, intestinal bacterial overgrowth, drug-vitamin interactions, alcohol)

Neurodegenerative and neuropsychiatric diseases

Neuropathies, dementia, Alzheimer’s disease, cognitive disorders, schizophrenia

Chronic atrophic corpus gastritis

Malabsorption of vitamin B12

Diseases of terminal ileum

Crohn’s disease, ileal lymphomas, ileal resection, ileal bacterial overgrowth

Pancreatic insufficiency

 

Chronic hepatic and renal diseases

 

Macrocytic anemia

Low dietary vitamin B12 intake or pernicious anemia

Chronic alcoholism

Low dietary vitamin B12 intake, vitamin B12 malabsorption

Drugs

Proton pump inhibitors, H2 receptor antagonists, dinitrogen monoxide (laughing gas) inhalation

AIDS-associated myelopathy

Abnormal vitamin B12-dependent transmethylation

Table 13.3-2 Age- and gender-specific reference intervals for vitamin B12 /35/

Age
(years)

Female

Male

ng/L (pmol/L)

ng/L (pmol/L)

< 1

228–1515
(168–1115)

293–1210
(216–891)

2–3

416–1210
(307–892)

264–1215
(195–897)

4–6

313–1410
(231–1040)

245–1075
(181–795)

7–9

247–1175
(182–866)

271–1170
(200–863)

10–12

196–1020
(145–752)

183–1090
(135–803)

13–18

182–820
(134–605)

214–864
(158–638)

Adults

211–911
(156–672)

211–911
(156–672)

Vitamin B12 concentration in 45 patients with confirmed B12 deficiency diagnosis (95% interval): 32–246 ng/L (24–181 pmol/L).

Conversion factor: ng/L × 0.738 = pmol/L

Values for children and adolescents based on radioimmunoassay; values for adults based on chemiluminescence immunoassay.

Table 13.3-3 Reference intervals for biomarkers of the vitamin B12 status /51415/

Biomarker

Reference interval

Methylmalonic acid (MMA)

73–271 nmol/L

Holotranscobalamin (HoloTC)

35–171 pmol/L

Vitamin B12 resorption*

> 10% of administered dose

* Schilling test

Table 13.3-4 Disorders associated with vitamin B12 deficiency

Clinical and laboratory findings

Chronic kidney disease (CKD)

In many cases, patients with CKD have hyperhomocysteinemia and elevated MMA levels (100% of dialysis patients and approximately 60% of transplant recipients), although the serum vitamin B12 level is normal in most cases /2930/. Elevated homocysteine and MMA levels in patients with renal disease can be corrected by vitamin B12 substitution, which indicates the existence of vitamin B12 deficiency prior to treatment (Fig. 13.3-3 – Decrease in methyl malonic acid (MMA) in kidney patients after vitamin B12 supplementation/30/. This is presumably due to impaired cellular uptake of holoTC leading to intracellular functional cobalamin deficiency with metabolite increase. Investigations have shown that renal patients can have high holoTC concentrations which actually contradicts vitamin B12 deficiency /3031/. This can be explained by the role of the kidney in transcobalamin filtration and a resulting secondary accumulation of holoTC. Consequently, the holoTC level in the plasma of these patients does not reflect the functional vitamin B12 status. The reduction of the MMA concentration by more than 200 nmol/L by cobalamin injection is a reliable way of confirming vitamin B12 deficiency in patients with renal disease. Since renal patients may also have elevated MMA concentrations not induced by vitamin B12 deficiency, differentiation of B12 deficiency is only possible by therapeutic MMA lowering /30/. The elevated MMA concentration (vitamin B12 resistant MMA increase) remaining despite increased amounts of circulating vitamin B12 is explained with impaired renal function (Fig. 13.3-3).

Iron and vitamin B12 deficiency

Clinical diagnosis of vitamin B12 deficiency can be delayed by coexisting iron deficiency giving an equivocal hematological picture. The macrocytosis caused by vitamin B12 deficiency can be masked by concurrent iron deficiency /32/. Microcytosis becomes predominant over macrocytosis if iron deficiency is more severe than B12 deficiency. Vitamin B12 deficiency can cause an additional loss of iron by means of a secondary effect on the enterocytes /33/.

Malabsorption

Vitamin B12 deficiency does not occur until nutritional deficiency or malnutrition have persisted for an extended period of time or Cbl absorption has been persistently impaired and stores depleted due to gastrointestinal disease /18/. Vitamin B12 deficiency can have the following causes:

  • Long-term nutritional deficiency and malnutrition in vegetarians
  • Total or partial gastrectomy, chronic gastritis with resulting hypochlorhydria or achlorhydria, intrinsic factor deficiency
  • Exocrine pancreatic insufficiency, impaired cleavage of the B12 haptocorrin complex in the duodenum in trypsin deficiency
  • Bowel disease (e.g., tropical sprue, extensive intestinal involvement or preferential involvement of the terminal ileum in Crohn’s disease, blind loop syndrome)
  • Extensive parasite (e.g., fish tapeworm)
  • Selectively congenital disorder of B12 resorption (Imerslund-Grasbeck syndrome).

In many cases, individuals of advanced age develop malabsorption of food-bound vitamin B12. This is caused by atrophic gastritis, where inadequate amounts of HCl and pepsin, but adequate amounts of intrinsic factor are formed in the initial stage. In this stage, the uptake of food-bound B12 is no longer possible, but vitamin B12 administered as a supplement can still be absorbed. In advanced atrophy, adequate amounts of intrinsic factor can no longer be formed, causing a marked restriction in receptor-mediated uptake of vitamin B12 by the enterocytes. In this case, orally ingested vitamin B12 can only be absorbed by passive diffusion independently of receptors. Therefore, it is possible through administration of higher doses of vitamin B12 (up to 1 mg/day) to achieve a sufficient blood concentration of vitamin B12 (normalization of the B12 status) /34/. Sole substitution of vitamin B12 in older adults is not recommended, but rather always a combination of vitamin B12, B6 and folic acid (Fig. 13.3-3 – Decrease in methyl malonic acid (MMA) in kidney patients after vitamin B12 supplementation). If the passive resorption capacity is no longer sufficient, the vitamin B12 status can be normalized through bypassing of the gastrointestinal tract by injection (1 mg/month).

Chronic type A gastritis with destruction of the mucosa cells in the corpus and fundus of the stomach is based on an autoimmune process. Parietal cell antibodies are found in 80% of patients with pernicious anemia and intrinsic factor antibodies in up to 40%. It is very common in patients with pernicious anemia to concurrently have antibodies against thyroidal antigens, and, vice versa, one third of patients with thyroid antibodies also have parietal cell antibodies.

The effect of drugs (for example H2 blocker or proton pump inhibitors) can also impair the vitamin B12 resorption capacity.

Hereditary defects

Ten rare hereditary defects have been identified that impair the transport and metabolism of vitamin B12 in the human body (Fig. 13.3-7 – Cellular uptake, intracellular distribution and synthesis of cobalamin coenzymes). Three defects relate to absorption and transport, the other seven change cellular processing and coenzyme production. Defects relating to absorption and transport manifest already in infancy and early childhood as developmental delay with manifested megaloblastic anemia. The serum total vitamin B12 level can be low (in intrinsic factor or intrinsic factor receptor deficiency) or almost normal (TC II deficiency). Treatment of these diseases with Cobalamin (Cbl) injections has proven highly effective.

Clinical manifestation of defects in cellular Cbl processing and Cbl metabolism varies depending on whether only one or both coenzymes are affected. Two abnormalities in adenosyl-Cbl synthesis, referred to as CblA and CblB, lead to impaired methyl malonyl-CoA-mutase activity causing MMA acidemia. In patients with this defect, a marked reduction in MMA accumulation can in many cases be achieved by supplementation with cyano-Cbl or hydroxo-Cbl in addition to protein restriction. Oral antibiotics are additionally helpful in reducing propionate formation by intestinal bacteria. The exact localization of the CblA defect has not been determined, but is has been found that, in CblB defect, cob(I)alamin adenosyltransferase, the last step of adenosyl-Cbl biosynthesis, is impaired.

Two abnormalities in methyl-Cbl, referred to as CblE and CblG, lead to reduced methionine synthase activity with resulting homocystinuria, hyperhomocysteinemia and hypomethioninemia. Affected children have developmental disorders, avolition and megaloblastic anemia. Treatment with pharmacological vitamin B12 doses corrects most clinical abnormalities; in addition, administration of betaine is helpful to reduce hyperhomocysteinemia. The exact localization of the defects has not been determined, but the reducing system necessary to maintain the active form of the methyl-Cbl methyltransferase complex (CblE) or the enzyme methyltransferase itself (CblG) seem to be affected. Other mutations, referred to as CblC, CblD and CbF, lead to impaired synthesis of the two cobalamin, adenosly-Cbl and methyl-Cbl, and – accordingly – to a lack of enzyme activity of methylmalonyl-CoA-mutase and methionine synthase. Affected individuals have MMA aciduria and homocystinuria. The main clinical symptoms in patients are developmental retardation, avolition and hematological alterations, such as megaloblastic anemia and macrocytosis. Treatment is aimed at the relevant mutation and consists of a combination of protein restriction and pharmacological doses of hydroxo-Cbl, possibly in combination with betaine supplementation. The defects in CblC and CblD have not been clarified in detail, but affect the steps of intracellular Cbl metabolism and, possibly, cytosolic Cbl reduction. Defects in CblF seem to affect the lysosomal Cbl efflux. The above-mentioned diseases of Cbl metabolism are autosomal recessively inherited.

Table 13.4-1 Folate reference intervals for fasting serum

Adults

1.8–9.0 μg/L (4–20 nmol/L)

Children /4/

Female

Male

μg/L
(nmol/L)

μg/L
(nmol/L)

0–1 yrs

6.3–22.7
(14.3–51.5)

7.2–22.4
(16.3–50.8)

2–3 yrs

1.7–15.7
(3.9–35.6)

2.5–15.0
(5.7–34.0)

4–6 yrs

2.7–14.0
(6.1–31.9)

0.5–13.0
(1.1–29.4)

7–9 yrs

2.4–13.4
(5.4–30.4)

2.3–11.9
(5.2–27.0)

10–12 yrs

1.0–10.2
(2.3–23.1)

1.5–10.8
(3.4–24.5)

13–18 yrs

1.2–7.2
(2.7–16.3)

1.2–8.8
(2.7–19.9)

Conversion factor: μg/L × 2.27 = nmol/L. Values are the 2.5th and 97.5th percentiles.

Table 13.4-2 Serum and whole-blood folate reference intervals in the USA and in Europe

Serum folate

USA

Adults

μg/L

nmol/L

Deficiency

(clinically verified)

0.35–3.4

0.79–7.6

Not clearly assignable

3.4–5.4

7.6–12.2

Normally healthy

> 5.4

> 12.2

Europe

Adults

2.0–9.1

4.5–20.6

Whole-blood folate

USA

Adults

280–791

636–1796

Central Europe

Adults

150–450

341–1022

Values are the 2.5th and 97.5th percentiles.

Table 13.4-3 Stages of folate deficiency

Disorder

Prelatent

Latent

Manifested (morphologically and
functionally)

Stage

Normal

Early
depletion

Metabolic
disorder

Subclinical

Clinically
reversible

Clinically
irreversible

Folate in serum (μg/L)

> 5.38

5.38–3.38

< 3.38

< 3.38

< 3.38

< 3.38

Folate in erythrocytes (μg/L)

> 200

200–150

< 150

< 150

< 150

< 150

Homocysteine (μmol/L)

< 10

10–12

> 12

> 12

> 12

> 12

Hypersegmentation of neutrophils

no

no

no

yes

yes

yes

Macroovalocytosis/medullary megaloblastosis

no

no

no

no

yes

yes

Anemia (low hematocrit, Hb and red blood cell count)

no

no

no

no

yes?

yes

Clinically irreversible depression/dementia

no

no

no

no

possible

common

Table 13.4-4 Risk groups and diseases associated with folate deficiency

Clinical and laboratory findings

Risk groups

Individuals with one-sided eating habits, older adults, pregnant women, patients with renal disease or folate malabsorption (inflammatory bowel disease) and tumor patients are risk groups for folate deficits of clinical relevance. Moreover, alcohol abuse and the use of certain drugs lead to vitamin deficiency (Tab. 13.4-9 – Causes of folate deficiency).

Drugs

Various drugs lead to folate deficiency due to inhibition of the enzymes involved in folate metabolism. Oral contraceptives, for example, can inhibit intestinal folate conjugase, or methotrexate and trimethoprim inhibit dihydrofolate reductase. Serum of patients under methotrexate or leucovorin therapy is not suited for folate determination because of a cross-reactivity with folate binding proteins. The folate status in patients under methotrexate therapy can be analyzed using LC-MS/MS that allow a clear distinction between methotrexate and folates based on different mass transitions. An increase of the 5-formyl-THF peak is to be expected under leucovorin therapy; 5-methyl-THF remains unaffected.

Folate supply can also be affected by the folate-antagonistic effect of drugs. Drugs can lead to folate deficiency by restricting the bio availability of folate. Anti epileptic treatment (e.g., primidone, carbamazepine, valproic acid, phenobarbital) or the intake of folate antagonists, such as methotrexate, trimethoprim, triamterene, pyrimethamine, result in reduced folate resorption. Lower folate concentrations in women under contraceptive treatment have also been reported; reduced resorption and impaired cleavage of the folate polyglutamates have been discussed in this context.

Vitamin B12 deficiency

In vitamin B12 deficiency, methionine synthase requiring vitamin B12 as a cofactor is inhibited causing impaired transfer of the methyl group from 5-methyl-THF to homocysteine. This results in an increase in homocysteine and 5-methyl-THF and a loss of feedback control of 5-methyl-THF formation. THF regeneration is blocked due to the accumulation of 5-methyl-THF, thus causing functional folate deficiency. This folate deficiency is enhanced further by the fact that the retention of folate compounds is reduced in vitamin B12 deficiency due to decreased de methylation of 5-methyl-THF during cellular uptake leading to losses of 5-methyl-THF and increased renal excretion.

Inflammatory bowel disease

Folate resorption disorder develops in inflammatory bowel disease such as Crohn’s disease, ulcerative colitis or celiac disease.

Chronic kidney disease

Chronic kidney disease is characterized by a high prevalence of vitamin deficiencies, where a significant vitamin B12 and B6 deficiency is present besides folate deficiency. Folate deficiency in renal patients is enhanced further by the vitamin B12 deficiency-induced disorder of folate processing (folate accumulation) /13/.

Chronic hemodialysis

Hemodialysis patients ingest an inadequate amount of folate; in addition, hemodialysis itself causes folate losses.

Neural tube defects (NTD)

Periconceptional folic acid supplementation significantly reduces the incidence of NTDs by 20–60%. In a Cochrane Review comprising 6425 women from four trials, it was reported that periconceptional folate supplementation markedly reduced NTDs [relative risk (RR) 0.28; confidence interval 0.13 to 0.58] /15/. In 1992, the US Center for Disease Control and Prevention (CDC) determined folic acid supplementation of 400–800 μg/d for all women of reproductive age. A daily intake of 4 to 5 mg of folic acid is recommended for women with a positive NTD pregnancy history. The mandatory fortification of staple foods with folic acid for NTD prevention was introduced in the USA in 1998. Meanwhile, more than 50 countries have followed this procedure.

Mean daily folate intake in pregnant women in Europe is 327 μg/d /16/. The corresponding folate intake in Germany is 254–271 μg/d /16/. Only 6% of the individuals reached the folate intake of 600 μg/d recommended for pregnant women in Germany, Austria and Switzerland, and only 26% had a total dietary folate intake of 400 μg/d.

There are approximately 800 NTD pregnancies in Germany every year. Most of these pregnancies are terminated following positive prenatal screening. The incidence of NTD in Germany is very high compared to EUROCAT data (www.eurocat-network.eu): EUROCAT mean 7.88 per 10,000 births (between 2004 and 2008), in Germany about 12.36/10,000 births (mean of the registers in Mainz and Saxony-Anhalt). The high effectiveness and the favorable cost-benefit ratio of folic acid supplementation for NTD prevention speak in favor of the mandatory fortification of staple foods with folic acid /17/.

Thresholds associated with maximum reduction of neural tube defects are /14/:

  • Red blood cell folate: 400 μg/L (906 nmol/L)
  • Plasma folate 7 μg/L (15.9 nmol/L)

Vascular disease

Folate deficiency is the most common cause of hyperhomocysteinemia, an important cardiovascular and neurodegenerative risk factor /8/. Another approach for investigating the clinical effect of folate deficiency is the analysis of the endothelial function as the first identifiable correlate of vascular wall alteration. Significant improvement of the endothelial function (improved NO-mediated vasodilatation measured as increased tissue perfusion of the forearm) and significant improvement of the coronary endothelial function were observed after folic acid therapy in patients with cardiovascular disease /18/. The volumetric coronary blood flow increased by 96% compared to the initial values.

Neurological and psychiatric manifestations

Dementia syndromes, organic depressive disorders, cerebral seizures as well as myelopathy and polyneuropathy are neurological psychiatric manifestations of disorders of the vitamins involved in methionine metabolism as co factors. Silent stroke and cerebral atrophy also show a relation to hyperhomocysteinemia and B vitamin status. Moreover, low B vitamin concentration and hyperhomocysteinemia are associated with cognitive capacities and the volume of the cerebral gray matter. According to current study results, increased vitamin B intake can be CNS protective.

Refer additionally to Section 13.1 – Hyperhomocysteinemia and degenerative diseases

Malignant tumors – Generalized

High folate serum levels can have a protective effect on carcinogenesis.

– Colorectal carcinoma

According to a meta-analysis (five cross-sectional and seven case control studies), high dietary folate intake reduces the risk of colorectal tumors by approximately 25% /19/. However, it was found upon intake of 1 mg/d of folic acid for 6 to 8 years that the risk of advanced or multiple lesions tended to be increased, with no increased recurrence of colorectal adenomas /20/. In another meta-analysis on folic acid supplementation (0.5 to 5 mg/d for three years), no significant influence on the risk of colorectal carcinoma was found in patients with adenoma history /21/. Likewise, high-dosed folic acid supplementation (2.5 mg/d to 20 mg/d for two to five years and subsequent follow-up for five to seven years) in populations at no increased risk of developing colorectal cancer did not have a significant effect on the risk of colorectal cancer /21/.

– Breast Cancer

A protective effect of folate against breast cancer was found in a nine-year monitoring study comprising more than 11,000 postmenopausal women with a total intake of over 456 μg/d compared to 160 μg/d [hazard ratio (HR) 0.56] /22/. In a prospective study on more than 25,000 postmenopausal women, no increased risk of breast cancer was found upon folate intake of over 337 μg/d compared to ≤ 233 μg/d or supplementary folic acid intake below 400 μg/d /23/. In the same study, supplementary folic acid intake of ≥ 400 μg/d (corresponding to a total folate intake above 853 μg/d) was associated with a higher risk of breast cancer (HR 1.32). The currently available information does not suggest a conclusive association between risk of breast cancer and folate intake.

– Pancreatic cancer

According to a study on the risk of pancreatic cancer comprising more than 100,000 individuals, a daily folate intake of ≥ 253 μg/d, compared to ≤ 179 μg/d, is protective in women, but does not affect the risk of pancreatic cancer in men /24/. An 8.6-year Swedish prospective study of more than 81,000 men and women found that a folate intake above 350 μg/d, compared to ≤ 200 μg/d, is associated with lower risk of pancreatic cancer /25/.

– Ovarian cancer, Prostate carcinoma

Higher folate intake is not associated with increased risk of ovarian or prostate carcinoma. A 22-year prospective study did not show a significant relation between risk of ovarian carcinoma and folate intake /26/. In a nine-year monitoring study on 65,836 men (American Cancer Society Cancer Prevention Study II Nutrition Cohort), including 5158 men diagnosed with prostate cancer, there was no correlation between risk of prostate cancer and dietary or total folate intake /27/.

Adenoma history

A meta-analysis of human studies did not find a significant effect of high folic acid doses (0.5 to 20 mg/d for 3–7 years) on the recurrence or incidence of advanced adenomas in individuals with adenoma history /21/. A potentially harmful effect of unmetabolized, free folic acid in blood was not confirmed /117/. Consequently, a dual role might be proposed for folic acid in carcinogenesis (Fig. 13.4-5 – The dual role of folic acid regarding adenoma and cancer risk). Moderate intake of folic acid has a protective influence regarding the establishment of neoplastic foci, whereas excessive intake (much more than 1 mg/d) for an extended period of time can accelerate tumorigenesis /28/.

Table 13.4-5 Folate forms in serum using UPLC-MS/MS

Folate levels
(nmol/L)

Women, non-
supplemented
/29/

Pregnant women,
non-supplemented
/29/

Pregnant women,
supplemented
/29/a

Umbilical cord
blood
/35/b

Older adults, non-
supplemented
/1/

Older adults,
supplemented
/1/c

Total folate(d

18.1 (7.7–30.3)

13.7 (5.6–47.3)

26.6 (6.4–46.3)

39.8 (9.0–78.5)

11.0 (4.1–34.7)

102.9 (49.0–280.9)

5-methyl-THF

15.8 (6.4–27.7)

11.7 (3.9–44.5)

24.7 (4.8–39.6)

37.8 (7.7–76.5)

5.6 (2.4–19.1)

61.4 (33.2–139.0)

THF

2.1 (0.35–3.8)

1.6 (0.4–3.4)

2.2 (0.9–4.3)

2.2 (0.7–6.5)

3.4 (1.3–16.0)

14.5 (6.0–30.5)

Formyl-THF

0.16 (0.05–0.35)

0.13 (0.03–0.27)

0.25 (0.08–0.38)

0.19 (0.02–0.54)

 

 

5,10-methenyl-THF

0.02 (0.00–0.13)

0.17 (0.01–0.34)

0.20 (0.11–0.33)

0.16 (0.06–0.42)

 

 

Folic acid

0.10 (0.03–0.77)

0.13 (0.00–0.78)

0.19 (0.00–0.91)

0.21 (0.00–0.51)

0.08 (0.00–0.85)

15.3 (1.02–144.4)

Values are median and 10th and 90th percentiles.

a: 400 μg of folic acid/day

b: n = 10 pregnant women took 400 μg/day of folic acid starting in month 1 of pregnancy.

c: 5 mg of folic acid, 2 mg of cyanocobalamin and 40 mg of vitamin B6/day for 3 weeks.

d; total of folate forms. THF, tetrahydrofolate; UPLC-MS/MS, ultra performance liquid chromatography-tandem mass spectrometry.

Table 13.4-6 Folate forms in serum using UPLC-MS/MS and immunological method by MTHFR C677T genotype /3/

Folate form
(nmol/L)

MTHFR 677 CC
genotype

MTHFR 677 CT
and TT genotypes

Total folate,
immunological
method

28.59
(9.08–40.57)

17.50
(9.29–35.03)

Total folate,
UPLC-MS/MS

22.23
(8.00–30.55)

12.19
(5.97–29.90)

5-methyl-THF,
UPLC-MS/MS

18.48
(6.63–27.16)

10.72
(4.73–26.25)

THF, UPLC-MS/MS

2.77
(1.18–4.43)

1.09
(0.20–3.50)

Values expressed are median and 10th and 90th percentile).

a: Mann-Whitney-U assay, b: Total of folate forms

MTHFR, 5,10-methylenetetrahydrofolate reductase; THF, tetrahydrofolate; UPLC-MS/MS, ultra performance liquid chromatography-tandem mass spectrometry.

Table 13.4-7 Categories for 5-methyl-THF in fasting serum/plasma and erythrocytes using UPLC-MS/MS or LC-MS/MS

Supplemen-
tation

Age (years)

5-methyl-THF
(nmol/L)

32 men /3/

no

33 (20–51)

15.8 (5.6–26.7)

168(70% men) /2/

n/a

45 (21–69)

16.4 (5.8–71.3)

37 older adults /1/

no

81 (73–90)

6.5 (2.5–16.7)

61 Pregnant
women /29/

no

30 (22–37)

11.7 (3.9–44.5)

25 Pregnant
women /29/

400 μg folic
acid/day

24.7 (4.8–39.6)

50 Pregnant
women /31/

n. a.

n. a.

14.6 (9.1)

46 Pregnant
women
with NTD
fetus /31/

n. a.

n. a.

9.3 (4.5)

313 Pregnant women
< 16 GW /32/

n. a.

20–29a

15.8 (11.6–21.9)

100 Blood bank
USA /33/

n. a.

n. a.

23.7 (7.61–72.0)

23 African
women /34/

n = 15b

31.6 (6.0)

33.5 (17.2)

26 Caucasian
women /34/

n = 17b

33.3 (6.5)

48.4 (20.5)

75 Blood bank
Europe /30/

n. a.

n. a.

207 (30.2–462)

109 men /35/

no

36 (11.3)

427 (92.5–1089)

96 Blood bank
USA /30/

n. a.

n. a.

304 (94.7–703)

23 African
women /34/

n=15b

31.6 (6.0)

919 (334)

26 Caucasian
women /34/

n=17b

33.3 (6.5)

1040 (333)

30 Women
MTHFR 677 CC,
CT and TT, each /36/

n.a.

n.a.

930 (286), CC
1065 (360), CT
764 (292), TT

n.a., not achieved, n = 17b, 17 probands with multivitamins; n/a, no information; NTD, neurale tube defect

Table 13.4-8 Folate in serum/plasma andCSF using LC-MS/MS or immunological method

Folate (nmol/L)

Serum/plasma

CSF

Total folatea /38/

19.3 (11.3–42.4)

20.6 (14.0–27.6)

Total folatea /39/

13.4 (5.8)

30.4 (7.4)

5-methyl-THFb /40/

41.5 (13.7)

 

Values are median and 10th and 90th percentiles or mean and standard deviation.

a: immunological method; b: LC-MS/MS. CSF, cerebrospinal fluid;

THF, tetrahydrofolate;

LC-MS/MS, liquid chromatography-tandem mass spectrometry

Table 13.4-9 Causes of folate deficiency

Inadequate folate intake

Nutritional deficiency

Inadequate quantity of food

Malabsorption

Liver disease

Pancreatic insufficiency

Increased folate requirement

Pregnancy

Lactation period

Preterm neonate

Growth

Infection

Hemodialysis

Hemolytic anemia

Biological maturation

MTHFR genotype

Drug interferences

Folate analog

Methotrexate

Aminopterin

Trimethoprim

Pentamidine

Drug affecting resorption/utilization

Oral contraceptive

Acetylsalicyclic acid

Diphenylhydantoin

Barbiturate

Table 13.5-1 Indications for vitamin B6 determination

Nutritional
deficiency

Increased
requirement

Genetic
defect

Disease

Chronic
alcoholism

Pregnancy

Homo-
cystinuria

Premenstrual syndrome

Senile
cachexia

Lactation

Dialysis

Cystathio-
ninuria

Hyperoxal-
uria (type I)

Carpal tunnel syndrome
Hypertension
Hyperhomocysteinemia

Asthma

Diabetes mellitus

Long-term medication (see Tab. 13.5-3 – Interactions between vitamin B6 and drugs)

Table 13.5-2 Methods for determining the vitamin B6 status and their reference intervals /3/

Method

Reference interval

Direct, blood

Plasma pyridoxal phosphate

> 20 nmol/L

Plasma pyridoxal

Not available

Plasma total vitamin B6

> 40 nmol/L

Whole-blood pyridoxal phosphate

Not available

Direct, 24-hour urine

4-pyridoxine acid

> 3 μmol/24 h

Total vitamin B6

> 0.5 μmol/24 h

Indirect, blood

Homocysteine following oMLT

< 38 μmol/L

Erythrocyte ALT, ratio

> 1.24*

Erythrocyte AST, ratio

> 1.80*

Indirect, 24-hour urine

Xanthurenic acid following oTLT (2 g)

< 65 μmol/24 h

Cystathionine following oMLT (3 g)

< 351 μmol/24 h

Oxalate excretion

Not available

* Ratio, activity with PLP/activity without PLP; ALT, alanine aminotransferase; AST, aspartate aminotransferase; oMLT, oral methionine load test; oTLT, oral tryptophan load test; PLP, pyridoxal-5’-phosphate

Table 13.5-3 Interactions between vitamin B6 and drugs /3/

Ingredient
class

Example

Mechanism

Hydrazines

Iproniazid,
isoniazid,
hydralazine

Reacts with PL and PLP forming hydrazone

Methyl-
xanthines

Theophylline,
enprofylline

Inhibits pyridoxal kinase and specific aminotransferases

Antibiotics

Cycloserine

Reacts with PLP forming oxim

L-DOPA

L-3,4-dihydroxy-
phenylalanine

Reacts with PLP forming tetrahydrochinoline derivatives

Chelators

Penicillamine

Reacts with PLP forming thiazolidine

Oral contra- ceptives

Ethinylestradiol, mestranol

Elevated enzyme levels in the liver and other tissues, PLP retention

Alcohol

Ethanol

Increased PLP catabolism, low plasma concentration

PL, pyridoxal; PLP, pyridoxal-5’-phosphate.

Table 13.5-4 Pyridoxalphosphate-catalyzed enzymatic reactions /3/

Reaction type

Typical reaction or enzyme

Reactions at an α carbon

Transamination

Alanine Pyruvate + pyridoxamine phosphate

Racemate formation

D-amino acid L-amino acid

Decarboxylation

L-tryptophan tryptamine

Oxidative deamination

Histamine imidazol-4 acetaldehyde

Loss of side chain

THF + serine glycine + 5,10-methylene-THF

Reactions at a β carbon

Substitute (exchange)

Cysteine synthase

Elimination

Serine and threonine dehydratase

Reactions at a γ carbon

Substitute (exchange)

Cystathionine cysteine + α-ketobutyrate

Elimination

Homocysteine desulfhydrase

Cleavage

Kynurenine anthranilic acid

THF, tetrahydrofolate.

Table 13.5-5 Vitamin B6-dependent cellular processes /3/

Process

System/function

C1 metabolism,
hormone modulation

Immune function

Glycogen phosphorylase,
transamination

Gluconeogenesis

Tryptophan metabolism

Niacin formation

Heme synthesis,
transamination, O2 affinity

Metabolism and erythropoiesis

Neurotransmitter synthesis,
lipid metabolism

Nervous system

Hormone modulation,
binding of PLP and lysin or
hormone receptors

Hormone modulation

PLP, pyridoxal-5’-phosphate.

Table 13.5-6 Abnormal vitamin B6 metabolism, diseases and conditions /3/

Disease or condition

  • Old age
  • Nutritional deficiency or malnutrition
  • Atherosclerotic vascular diseases
  • Thrombotic vascular diseases
  • Cerebrovascular diseases
  • Neurodegenerative diseases (e.g., Alzheimer’s disease)
  • Impaired renal function, dialysis
  • Inflammatory bowel disease, celiac disease
  • Hyperhomocysteinemia/homocystinuria
  • Alcoholism
  • Pregnancy complications (e.g., preeclampsia)
  • Spasms in neonates and infants
  • Rheumatoid arthritis
  • Tumor diseases
  • Sickle cell anemia
  • Premenstrual syndrome
  • Asthma
  • Hodgkin’s disease
  • Pellagra (niacin deficiency)
  • Drug interactions

Table 13.5-7 Diseases associated with vitamin B6 deficiency

Clinical and laboratory findings

Homocystinuria

In the rare genetic homozygous homocystinuria with plasma homocysteine levels above 100 μmol/L associated with severe juvenile atherosclerosis, the focus is on treatment with vitamin B6. Homocystinuria is caused by severe cystathionine-β-synthase deficiency. Because saturation of the apoenzyme with high doses of the cofactor vitamin B6 increases the stability of the enzyme, leading to considerable amounts of the active enzyme.

Anemia /22/

Vitamin B6 plays an important role in the function and metabolism of red cells.Pyridoxal-5’-phosphate (PLP) is a cofactor of transaminases. Moreover, pyridoxal (PL) binds to the hemoglobin α-chain and increases the hemoglobin’s oxygen binding affinity, while PLP bound to the hemoglobin S β-chain reduces the oxygen binding affinity. The effect of PLP and PL on the oxygen binding affinity is important in sickle cell anemia. Treatment of low values with high doses of vitamin B6 (100 mg/day) improves the symptoms of the disease.

PLP is also a coenzyme of α-aminolevulinic acid synthase (ALS), a key enzyme of heme synthesis. In the setting of vitamin B6 deficiency or a genetic defect of ALS, hypochromic, iron-refractory, microcytic anemia can develop. ALS is reduced in such patients and can be normalized by high-dose treatment with vitamin B6 (200–600 mg/day).

Nephrolithiasis

Vitamin B6 deficiency can also result in increased oxalic acid excretion and development of nephrolithiasis. Primary oxalosis type I is caused by a defect of peroxisomal alanine glyoxylate aminotransferase, as a result of which glyoxylic acid degradation is inhibited /23/. High doses of vitamin B6 trigger the induction of glyoxylate transaminase and thus induce an alternative pathway for degradation /24/.

Neurological disorders in neonates

PLP is a cofactor of enzymes involved in the synthesis of various neurotransmitters, such as 5-hydroxytryptamine, dopamine, noradrenaline, histamine or γ-aminobutyric acid. The role of PLP as cofactor in the formation of neurotransmitters and the observation of neurological abnormalities in neonates with vitamin B6 deficiency underscores the importance of the vitamin for CNS function. Treatment of these neonates with high doses of vitamin B6 resulted in normalization of the electroencephalogram.

Neonatal seizures /26/

Neonatal seizures caused by pyridox(am)ine phosphate oxidase (PNPO) deficiency are an important disease with disorder of the vitamin B6 metabolism. It has been known for a long time that some neonatal seizures respond to high doses of folate and pyridoxine or PLP.

Malignant tumor /27/

There is a strong correlation between vitamin B6 intake or plasma PLP level and risk of cancer, especially regarding colorectal cancer as well as gastrointestinal tumors. In contrast, no significant relation with the vitamin B6 status has been reported for breast cancer and prostate carcinoma.

Vascular disease

Because homocysteine is a separate risk factor for degenerative cardiovascular and neurodegenerative diseases, vitamin B6, as a cofactor involved in homocysteine metabolism, is of special significance in the diagnosis and therapy of such diseases.

It was recently reported that pyridoxamine could prevent vascular aging induced by cross-linking of glycated collagen /28/. Pyridoxamine plays an important role as inhibitor of the formation of advanced glycation end products (AGE) by non-enzymatic protein glycation from sugar-protein adducts /29/. Pyridoxamine also inhibits the oxidative conversion of intermediate products of the Amadori reaction /30/. Moreover, pyridoxamine eliminates numerous toxic carbonyl species from sugars or lipids and thus prevents oxidative modification of proteins /31/. Because of the significance of glycation for the pathogenesis of complications in diabetes mellitus, pyridoxamine has the potential of preventing clinical consequences of the disease.

Association between low plasma PLP and cardiovascular disease has been confirmed in numerous retrospective and prospective studies /32/. The fact that low vitamin B6 can lead to elevated homocysteine concentrations might also explain the increased cardiovascular risk in PLP deficiency. Moreover, plasma PLP is also decreased by factors such as smoking or inflammation, which – in turn – are directly associated with cardiovascular risk. The current studies do not allow a clear conclusion as to whether low PLP is a causal and independent risk factor for cardiovascular disease or only a surrogate marker /33/.

Venous thrombosis

Low plasma PLP is also associated with an increased risk of getting venous thrombosis. This association has been reported from retrospective and prospective studies /34/. However, secondary prevention studies with B vitamins (B6, B12 and folic acid) did not show a lower risk of thrombosis /35/. It remains open whether B vitamins have a protective effect in primary thrombosis prevention.

Neurological disorders

Deficiencies in vitamin B6, B12 and folic acid are associated with an increased risk of neurological and psychiatric dysfunctions and cognitive deficit. In the elderly, the high incidence of cognitive impairment and dementia may be related to the high prevalence of inadequate B vitamin status and hyperhomocysteinemia in this population /36/. Accelerated brain atrophy in the elderly with mild cognitive impairment (MCI) could be slowed down significantly by treatment with homocysteine-lowering B vitamins (B6, B12 and folic acid) /37/. Accelerated brain atrophy is characteristic of individuals with MCI developing Alzheimer’s dementia. Clinical studies performed to date underscore the significance of vitamin B6 for the development and prevention of neurodegenerative disorders.

Table 13.6-1 Fasting betaine and choline levels in EDTA plasma and urine using LC-MS/MS. Modified according to Ref. /5/

Study
group

Plasma (μmol/L)

Urine (mmol/mol creatinine)

Betaine

Choline

DMG

ACh*

Betaine

Choline

DMG

Men and
women /6/

27–
41a

7.0–
9.3a

1.3–
2.0a

Men and
women,
non-fasting /6/

36–
47a

9.0–
12.3a

1.6–
2.5a

Women,
47–49 yrs /7/

21.7–
46.3

6.8–
11.5

Men,
47–49 yrs /7/

31.0–
58.2

7.4–
12.6

Women,
71–74 yrs /7/

24.9–
50.3

7.2–
12.8

Men,
71–74 yrs /7/

31.6–
61.1

8.1–
14.1

Women,
50–73 yrss /4/b

18.7–
37.5

7.0–
11.7

1.5–
3.7

2.6–
5.8

3.3–
11.4

1.1–
4.0

1.3–
4.7

Men,
57–75 yrs /4/b

25.2–
46.8

7.4–
12.4

2.0–
4.3

1.8–
6.6

2.4–
25.3

1.0–
4.7

0.6–
7.0

Women /58/

17–60c

Men /58/

21–78c

Women /58/

 

0.4–
12.3c

Men /58/

 

0.8–
9.6c

Men and
women /58/

 

1.8–
35.9c

Men and
women /58/

 

0.4–3
0.5c

Values are the 10th and 90th percentiles; • μmol/mol creatinine; a Values are the 25th and 75th percentiles; bUPLC-MS/MS method; c central 95% range.

ACh, acetylcholine; DMG, dimethylglycine; LC-MS/MS, liquid chromatography-tandem mass spectrometry.

Table 13.6-2 Effect of diseases and lifestyle factors on plasma and urinary betaine and choline concentrations /10/

Disease/lifestyle factor

Betaine

Choline

Plasma

  • Metabolic syndrome

  • Folate deficiency

  • Homocysteine

  • Stress

  • Smoking

  • Exercise

  • Body mass index

  • HDL cholesterol

  • Triglycerides

Urinary excretion

  • Metabolic syndrome

↑↑

  • Diabetes mellitus

↑↑

  • Chronic renal failure (without diabetes mellitus)

↑↑

  • Fibrate therapy

  • Homocysteine

Table 13.6-3 Association between betaine deficiency and diseases

Clinical and laboratory findings

Adiposity

The role of betaine in adiposity might be based on correlation with lipid metabolism. Various cross-sectional studies have shown /7/ that plasma betaine is inversely correlated with lipid parameters (triglycerides, apolipoprotein B, LDL-cholesterol) and, therefore, low plasma betaine is associated with higher vascular risk. Moreover, betaine supplementation in healthy individuals caused an increase in plasma cholesterol /20/ and a mild increase in plasma triglycerides and LDL cholesterol. In addition, animal assays have shown that betaine supplementation leads to increased hepatic apolipoprotein B synthesis /21/ and increased plasma LDL and triglycerides. These changes are associated with a marked decrease in tissue lipids. Betaine also reduced fatty acid synthesis /22/.

In cross-sectional studies, low plasma betaine is associated with elevated lipids and other parameters of metabolic syndrome /10/. This conforms to findings that patients with dyslipidemia had lower plasma betaine /34/. Therefore, numerous studies suggest an association between betaine deficiency and lipid abnormalities.

Diabetes mellitus/metabolic syndrome

More than 20% of patients with diabetes mellitus have elevated urinary betaine excretion /24/. Abnormal betaine excretion is also common in patients with chronic renal disease /24/, in patients with lipid metabolism disorders /23/ and in patients under fibrate therapy /25/. As there is a positive correlation in diabetic patients between urinary betaine excretion and sorbitol, another renal osmolyte, sorbitol overproduction has been discussed in diabetes mellitus.

Patients with metabolic syndrome also show a tendency toward lower plasma betaine /7/. The missing association between the substrate choline and its product betaine (formed by oxidation of choline by the mitochondrial enzyme choline dehydrogenase) reflects the disorder of this metabolic pathway as part of the mitochondrial dysfunction in metabolic syndrome. High urinary betaine excretion is common in metabolic syndrome /8/. This is in line with the finding that, in various disorders of lipid metabolism, large amounts of betaine are also excreted in urine /23/, which might result in problems in the betaine supply to the tissues of such patients. These patients can presumably benefit from betaine supplementation.

Vascular disease

High doses of betaine (≥ 6 g/d), alone or in combination with B vitamins, have been applied in the treatment of genetic homocystinuria for many years. Betaine is the key determinant of plasma homocysteine under non-fasting conditions. Various studies have investigated the effect of moderate betaine supplementation on plasma homocysteine concentrations and found a dose-dependent reduction in fasting homocysteine by up to 20% and in homocysteine following methionine load by 29–40% for the interaction of time and treatment /20/. Methionine, formed from homocysteine by methylation, can only be metabolized to homocysteine through conversion; in this cyclic process, the steady-state concentrations of plasma homocysteine only change within a small margin compared to the high betaine turnover /26/.

Betaine can be effective in critical supply with methyl groups because methyl groups from dietary betaine intake rapidly appear in tissue methionine. Betaine deficiency can manifest as hyperhomocysteinemia. The relation between betaine supply and vascular disease has been analyzed in studies. Betaine supplementation did not improve blood flow-mediated vasodilatation indicating endothelial function, despite significantly reduced homocysteine /26/.

In a study /37/ on healthy individuals with high choline and betaine intake, low plasma concentrations of inflammation biomarkers were measured. Because inflammation plays an important role in atherogenesis, high choline and betaine intake is considered protective against cardiovascular disease. However, in two prospective studies [of the Dutch PROSPECT-EPIC cohort and the Atherosclerosis Risk in Communities (ARIC) Study], no association between choline and/or betaine intake and cardiovascular risk was found /28/.

Liver disease

Betaine has been known to attenuate the toxic effect of alcohol on the liver and especially to counteract the development of fatty liver. This effect is presumably based on maintenance of the SAM level. Alcohol inhibits methionine synthase and thus increases betaine requirement for maintenance of the methylation capacity. The methylation capacity is especially important in the liver for the de novo synthesis of phosphatidylcholine (PC) essential for VLDL formation and secretion.

Betaine also improves the liver function in non-alcoholic fatty liver and in liver disease caused by xenobiotics or bile acids /29/. This positive effect is achieved by providing methyl groups for phosphatidylcholine synthesis or the osmolytic effect of betaine /29/.

Table 13.6-4 Association between choline deficiency and diseases

Clinical and laboratory findings

Cardiovascular disease

Choline and betaine are donators of methyl groups that lower the plasma homocysteine level. Clinical studies have shown that homocysteine is inversely correlated with choline intake /28/. In other studies, however, choline intake was not, or only tendentially, associated with the incidence of cardiovascular disease /31/.

Neural tube defect (NTD)

Choline is required as methyl group donor for neural tube closure. Pregnant women with low choline intake have an increased risk of giving birth to an NTD-affected child /32/.

Malignant tumor

Choline deficiency is associated with higher incidence of liver carcinoma. Various pathomechanisms have been discussed. For example, epigenetic alterations can trigger mechanisms involved in carcinogenesis /33/. Choline deficiency triggers protein kinase C (PKC) signaling that may cause altered cell proliferation, cell apoptosis and possibly carcinogenesis.

Intakes of choline and betaine are associated with risk of cancer. High choline intake reduced the risk of breast cancer and high choline and betaine consumption were also associated with lower breast cancer mortality /19/. However, positive association was observed between choline intake and risk of colorectal carcinomas indicating different etiologies for breast cancer and colorectal tumors /34/. Individuals with single nucleotide polymorphism (SNP) in the choline metabolism gene PEMT rs12325817 had a higher risk of breast cancer, while women with SNP in the gene BHMT rs3733890 had lower breast cancer mortality.

Epigenetic mechanisms of choline

The mechanisms of how choline influences fetal memory have remained unknown. It is assumed that, in adults, increased dietary choline intake leads to increased formation and release of ACh. In the fetus, maternal choline supplementation rather leads to greater concentrations of phosphoryl choline and betaine in the fetal brain than of choline and ACh /35/. Epigenetic mechanisms convey signals by DNA methylation and histone modification that lead to altered gene expression. SAM is utilized as methyl donor for the methylation of DNA and histones. The availability of SAM is directly affected by dietary choline intake. Choline deficiency or choline pathway disorders cause depleted SAM and increased S-adenosyl homocysteine (SAH) concentrations /36/. Choline deficiency causes reduced methylation of various genes and increased apoptosis in the fetal hippocampus /37/. The availability of choline during gestation also influences histone methylation in the developing fetus resulting in altered gene expression, especially in genes regulating methylation and neuronal cell differentiation /38/.

Figure 13.1-1 Homocysteine forms.

Homocysteine (1–2%) Homocysteine (5–10%) Cysteine-homocysteine (5–10%) HS-CH 2 -CH 2 -CH-COOH NH 2 - S-S-CH 2 -CH 2 -CH-COOH NH 2 Protein Protein Protein-bound homocysteine (70–90%) CH 2 -CH 2 -CH-COOH S S CH 2 -CH 2 -CH-COOH NH 2 NH 2 CH 2 -CH-COOH S S CH 2 -CH 2 -COOH NH 2 NH 2 S Protein

Figure 13.1-2 Homocysteine metabolism. ADP, adenosine diphosphate; ATP, adenosine triphosphate; BHMT, betaine homocysteine methyltransferase; CBS, cystathionine-β-synthase; CDP, cytidine diphosphate; CMP, cytidine monophosphate; CTP, cytidine triphosphate; DAG, diacylglycerol; GCT, γ-cystathionase; Hcy, homocysteine; MS, methionine synthase; 5-MTHF, 5-methyltetrahydrofolate; PC, phosphatidylcholine; PE, phosphatidylethanolamine; PEMT, phosphatidylethanolamine N-methyltransferase; PPi, pyrophosphate; SAM, S-adenosylmethionine; SAH, S-adenosylhomocysteine; THF, tetrahydrofolate; X-CH3, methylation product.

Cholin PC PE Phosphorylcholine CDP-Choline Acetylcholine Trimethylamine Choline-Oxidase Betaine aldehydedehydrogenase Choline-Kinase ATP ADP SAM Hcy Methionine SAH Betaine Dimethylglycine PEMT BHMT 5-MTHF THF B 12 - Cysteine Glutamate Glutathione Glycine SAH-Hydrolase CDP-Choline:DAGcholinphosphotransferase CTP PPi CTP:Phosphocholine cytidylyltransferase DAG CMP Membrane phospholipids Choline Phospholipase D Sphingomyelin Sphingomyelin-Synthase Diet CBS GCT X CH 3 -X 3 SAM 3 SAH γ-Glutamylcysteine

Figure 13.1-3 Dependency of cognitive performance (MMSE ,mini-mental state examination Score) in the elderly on homocysteine concentration. Modified according to Ref. /49/.

Homocysteine (μmol/l) 161514131211 n = 650 Healthy SeniorsAverage age 73 ± 6 Years * = 24–25 26–28 > 28 *

Figure 13.1-4 Cumulative incidence of dementia over a median of 8 years depending on basal homocysteine concentration. Modified according to Ref. /47/.

Cumulative incidence for dementia (%)201510 5 0 Individuals in the highesthomocysteine quartile All other individuals Years 0 2 4 6 8 10 12

Figure 13.1-5 Median changes in homocysteine concentrations after 26 weeks of treatment with placebo or 0.2 mg, 0.4 mg or 0.8 mg folic acid per day. Modified according to Ref. /44/.

11.5 12.6 12.6 13.1 11.7 10.6 10.6 10.3 1.4 –14.3 –14.2 –18.3 0–5–10–15–20 Placebo 0.2 mg Before 0.4 mg 0.8 mg After 26Weeks % Difference Homocysteine (μmol/L) % Differencecompared with baseline 20151050

Figure 13.1-6 Higher baseline homocysteine concentration was associated with a major reduction in brain atrophy following 24-month treatment with daily intake of folic acid (0.8 mg), vitamin B12 (0.5 mg) and B6 (20 mg). Modified according to Ref. /32/.

11.2% 43% 0 10 20 30 40 50 Hcy ≤ Median Hcy > Median

Figure 13.1-7 Mechanisms of homocysteine neurotoxicity. APP, amyloid precursor protein; MAP, mitogen-activated protein; PC, phosphatidylcholine; PE, phosphatidylethanolamine; PKC, protein kinase C; PS1, presenilin; SAM, S-adenosylmethionine; SAH, S-adenosylhomocysteine; sAPP, soluble amyloid precursor protein; tHcy, total homocysteine; UPS, ubiquitin proteasome system.

Inflammation Oxidative stress ↓↓ SAM/SAH ↑ tHcy UPS Unfoldedprotein cannot beremoved ↑ β-Amyloid ↓↓ Folate, B 12 , B 6 ↑ PS1 ↑ α-Synuclein ↑ α-Synuclein-aggregates ↑ Phosphatidylethanolamine ↑ PE/PC ? ? ↓ sAPPα non-amyloidogener Pathway + ↑ PS1 APP production Changeof themembrane fluidity ↓ Phosphatidic ↓ Diacyl glycine ↓ PKC L-Isoaspartyl D-Aspartyl L-Isoaspartyl D-Aspartyl Proteine L-Isoaspartyl,D-Aspartyl-O-Methyltransferase“repair protein“ α-Synuclein ↑ SAH SAM ↓↓ sAPP α ↑↑ sAPP β ↓↓ B 6 MAP-KinaseCell SignallingPathway ↓ PEMT-Activity

Figure 13.1-8 Correlation between SAH and the SAM/SAH ratio in cerebrospinal fluid (CSF) with β-amyloid 1–42 concentration in the CSF in dementia-free patients. SAM, S-adenosylmethionine; SAH, S-adenosylhomocysteine. Modified according to Ref. /62/.

2,000 1,000 600 400 200 100 CSF β-amyloid (1–42), ng/l CSF SAH, nmol/l 6050403020108765 40302010864 r = –0.31; p = 0.001 CSF SAM/SAH ratio r = 0.34; p < 0.001,after correction for age

Figure 13.1-9 Concentrations of SAH in cerebrospinal fluid (CSF) in relation to P-tau 181 concentrations in the CSF observed in three age groups. Age (in years):  ≤ 40,  41–60,  61–86. SAH, S-adenosylhomocysteine. Modified according to Ref. /62/.

< 38.5 ng/L ≥ 38.5 ng/L 0 10 20 30 CSF SAH (nmol/L) P-tau (181p) 9.7 11.6 12.6 12.3 15.2 19.0 p = 0.031 p = 0.038 n = 27 23 22 12 21 36 p = 0.001

Figure 13.2-1 Principle of the homocysteine assay catalyzed by cystathionine-β-synthase. Pyruvate detection is at 340 nm by means of NAD/LD. CBS, cystathionine-β-synthase; CBL, cystathionine-β-lyase.

HS OH NH2 l-Homocysteine OH NH S O HO NH 2 Cystathionine CBS CBL l-Serine Pyruvate + AmmoniaPyruvate will be detected by 340 nm via NADH/LDH O O

Figure 13.2-2 Decision tree for the diagnosis and prophylaxis/therapy in hyperhomocysteinemia (not applicable to kidney patients). B12, vitamin B12.

Risk groupsfor a stroke,cardiovascular disease Healthy > 50 Jahre,risk groups for vitamin deficiency(Table 13.1-2) Homocysteine ​​measurement > 12.0 μmol/L 10.1–12.0 μmol/L Folate: 0.2–0.5 mg, B 12 : 10–30 μg (> 60 y.: 100 μg), Vitamine B 6 : 2–10 mg ≤ 10.0 μmol/L Patients with a mild cognitivedysfunction, risk groups forneurodegenerative diseases No increased risk,control after twoyears Tolerable,yearly control Maintain supplementation, regularly control Control after four to six weeks Further diagnostics (renal function, malabsorption, MTHFR) Optimize dosage (e.g. increase B 12 to 1 mg/dL in patients with malobsorption ir impaired kidney function) ≤ 12.0 μmol/L > 12.0 μmol/L

Figure 13.3-1 Basic structure of cobalamin forms.

O O O O O O O O O O O P H 2 N H 2 N H 2 N NH 2 NH 2 NH 2 H 3 C CH 3 CH 3 CH 3 CH 3 CH 3 CH 3 CH 3 H 3 C H 3 C H 3 C HN N N N Co N N N R O OH HO
–R Name –CN Cyanocobalamin –OH Hydroxocobalamin H 2 O Aquacobalamin NO 2 Nitritocobalamin –CH 3 Methylcobalamin 5’-deoxyadenosyl 5’-Adenosylcobalamin

Figure 13.3-2 Algorithm for diagnosis of vitamin B12 deficiency. HoloTC, holotranscobalamin; MMA, methylmalonic acid.

HoloTC < 35 pmol/L B 12 deficiency Suspicion of vitamin B 12 deficiency HoloTC ≥ 35 pmol/L B 12 deficiency unlikely*

* MMA measurement recommended for concentration range 35–70 pmol/L in impaired renal function or diabetes mellitus. High MMA despite normal plasma and/or serum holoTC concentration suggests intracellular cobalamin deficiency that can be verified by reducing MMA following B12 supplementation.

Figure 13.3-3 Decrease in methylmalonic acid (MMA) in kidney patients after i.v. injections of 0.7 mg vitamin B12 3 ×/week for 4 weeks. The MMA lowering effect persisted in the following months. The MMA concentration was not normalized by the vitamin B12 treatment. SD, standard deviation.

4 2 Base-line 1,8001,6001,4001,2001,000800600 B 12 0.7 mg i.v./3 × per week MMA (average value ± 2 SD, nmol/L) 20 weeks 8 6 p < 0.001 compared with before treatment Depletion

Figure 13.3-4 Transport and cellular uptake of vitamin B12. Hcy, homocysteine; B12, vitamin B12; TC, transcobalamin; MS, methionine synthase; Cbl, cobalamin; Ado, adenosyl.

Diet Protein bound B 12 Haptocorrin (R-binder) Stomach Intestinallumen(Ileum) Intestinalcell Blood Intrinsic factor (IF) IF-receptor TC-receptor Lysosome Cells MethylmalonylCoA mutase Methionine Haptocorrin-B 12 IF-B 12 TC-B 12 Haptocorrin-B 12 (80%) TC-B 12 (10–30%) Succinyl CoA MethylmalonylCoA Hcy MS Methyl-B 12 OH-B 12 Ado-B 12

Figure 13.3-5 Cellular uptake and biological activity of vitamin B12. Hcy, homocysteine; TC, transcobalamin; MMA, methylmalonic acid.

Hcy MMA Only 10–30% serum total cobalamine is active 70–90% Haptocorrin-B 12 Holohaptocorrin Transcobalamin-B 12 HoloTC “Biological inert“ 10–30% “Biological active“ TC-receptor Hcy MMA Blood Cell TC–

Figure 13.3-6 Cobalamin-dependent biological reactions in the cytosol and mitochondria and in cobalamin deficiency. B12, vitamin B12; Hcy, homocysteine; Meth, methionine; MMA, methylmalonic acid; MM-CoA, methylmalonate-CoA; 5-MTHF, 5-methyltetrahydrofolate; Succ-CoA, succinyl-CoA; THF, tetrahydrofolate.

Meth+THF ↑ Hcy +5-MTHF Methionine synthase Cytosol ↑ MM-CoA Succ-CoA MM-CoA-mutase Mitochondrion ↓ CH 3 -B 12 OH-B 12 ↓ Adenosyl-B 12 Cobalamin-dependent biological Reaction ↑ MMA

Figure 13.3-7 Cellular uptake, intracellular distribution and synthesis of cobalamin coenzymes. Inherited genetic defects of cobalamin metabolism are categorized into classes A–G. The relevant defects have the following locus: 1) Lysosomal Cbl efflux, 2) Methionine synthase reductase, 3) Methionine synthase, 4) Cytosolic Cbl reductase/β ligand transferase, 5) Mitochondrial Cbl reductase and 6) CblI-adenosyl transferase. Ado, adenosyl; Cbl, cobalamin; Hcy, homocysteine; MS, methionine synthase; MCM, L-methylmalonyl-CoA-mutase; SAM, S-adenosyl methionine; SAH, S-adenosyl homocysteine; TC, transcobalamin. Modified according to Ref. /41/.

L-methylmalonyl CoA Succinyl CoA Cbl III Cbl II Cbl I MCM/ AdoCbl OH-Cbl Methyl-Cbl Lysosomal degradation Cbl C/D 4 Cbl F 1 Cbl A 5 Cbl B 6 TCII-Cbl Hcy Meth SAM SAH MS Cbl G 3 TC Cbl Cbl E 2 TC receptor

Figure 13.4-1 Structure of folate derivatives.

5 10 R2 R1 H 2 N HN HN N HN O COOH COOH O N N
Derivatives of pteroylmonoglutamic acid R1 R2 5-methyl 10-formyl –OH 5-formyl 5,10-methenyl Glutamic acid 5-formimino 5,10-methylene (–glutamyl) n

Figure 13.4-2 Folate intake in the population of various European countries. Modified according to Ref. /9/.

6004002000 Absorption (μg) Folate (Normal diet, women) Folate (Normal diet, men) DK FI DE IE IT PL ES NL UK DK FI DE IE IT PL ES NL UK

Figure 13.4-3 Function of folate forms in human cells. B12, vitamin B12; DHF, dihydrofolate; dTMP, deoxythymidine monophosphate; dUMP, deoxyuridine monophosphate; Hcy, homocysteine; Meth, methionine; MTHFR, 5,10-methylenetetrahydrofolate reductase; SAH, S-adenosyl homocysteine; SAM, S-adenosyl methionine; THF, tetrahydrofolate.

Active folate Folate DHF 10-Formyl THF 5-Formyl THF 5-Formimino THF 5,10-Methenyl THF 5,10-methylene THF 5-Methyl THF Purine dTMP DNA MTHFR Vitamin supplements Meth Hcy SAM SAH Methylation reactions B 12 DHF dUMP Thymidylate- synthase Diet THF polyglutamate THF

Figure 13.4-4 Effects of folate deficiency and/or vitamin B12 deficiency on erythropoiesis. DHF, dihydrofolate; dTMP, deoxythymidine monophosphate; dTTP, deoxythymidine triphosphate; dU, deoxyuridine; dUDP, deoxyuridine diphosphate; dUMP, deoxyuridine monophosphate; dUTP, deoxyuridine triphosphate; Hcy, homocysteine; 5-MTHF, 5-methyltetrahydrofolate; 5,10-MTHF, 5,10-methylenetetrahydrofolate; THF, tetrahydrofolate.

HCY Methionine THF DHF Folate 5,10-MTHF 5-MTHF dTMP dTTP DNA-synthesis dU dUMP dUTP Normoblast Megaloblast Hypersegmented neutrophil Normalneutrophil B 12 Normalerythropoiesis Disturbederythropoiesis Thymidylatesynthase

Figure 13.4-5 The dual role of folic acid regarding adenoma and cancer risk. Folate undersupply is considered an increased risk for various tumors (incorporation of uracil instead of thymidine into the DNA); oversupply with folic acid supplements has also been discussed to cause an increased risk of cancer (genetic dysregulation). There is no upper limit for natural folates. Modified according to Ref. /17/.

Uracil misincorporation U A Epigeneticregulation G C – CH 3 Adenoma- and cancer risk Low risk Total folate incorporation/d Folateoverdose≥ 1 mg/d Fortificationarea~200–400 g/d Lack of Folate≤ 400 μg/d Highrisk High risk

Figure 13.5-1 B6-vitamins (top row) and their corresponding phosphates (bottom row). A) Phosphatase; B) Pyridoxine kinase; C) Pyridoxine phosphate oxidase; D) Various transaminases.

CH 2 -OH CH 2 -OH HO HO Pyridoxine phosphate Pyridoxine CH 2 -OH CH 2 - HO HO P B A CHO CH 2 -OH HO HO Pyridoxal phosphate Pyridoxal CHO CH 2 - HO HO P B A CH 2 -OH HO HO Pyridoxamine phosphate Pyridoxamine CH 2 -NH CH 2 -NH CH 2 - HO HO P B A C D C N N N N N N

Figure 13.5-2 Key mechanism of the transamination reaction between enzyme-bound pyridoxal-5’-phosphate and an amino acid.

CH 2 HO HO P CH 2 HO HO R-CH NH 2 COOH H-C N Enzymelysine side chain H-C N R-CH COOH N N P

Figure 13.6-1 Betaine and choline concentrations in serum and EDTA plasma in 74 females (median age 35 years). Betaine concentrations in plasma and serum are almost identical, while choline concentrations in serum are twofold higher than in plasma. Modified according to Ref. /4/.

Betainen = 74r = 0.792p < 0.001 252015105 403020100 5 10 15 20 25 Plasma, μmol/L Serum, μmol/L Cholinen = 74r = 0.319p = 0.006 4 6 8 10

Figure 13.6-2 Betaine and choline concentrations in EDTA plasma in 28 individuals (10 males, median age: 52 years) following overnight fasting and following a small meal (non-fasting). Non-fasting individuals show moderately decreased betaine levels but higher choline concentrations. Modified according to Ref. /4/.

Choliner = 0.779p <0.001 Fasting plasma concentrations, μmol/L Non-fasting plasma concentrations, μmol/L 120100806040200 201816141210 Betainer = 0.536p = 0.003 4 6 8 10 12 14 20 40 60 80 100 120 Male (n = 10)Female (n = 18) Male (n = 10)Female (n = 18)
Goto top <