27 Jul 2020

Synthetic vs. organic B12 metabolism—is cyanocobalamin inferior?

Cyanocobalamin is a common synthetic form of vitamin B12 used in supplements and fortified foods—how does it compare to natural forms?

Vitamin B12 (cobalamin, Cbl) has the most complex structure of all vitamins, which consists of a central cobalt atom bound to a corrin ring, a displaceable lower (a) ligand (5,6-dimethylbenzimidazole, DMBI) and a variable upper (b) ligand (e.g. cyano-, methyl-, 5’-deoxyadenosyl-, etc.) 1 (see).

Cbl was originally isolated as cyanocobalamin (CNCbl), which was later recognised as an artefact arising from extraction methods 2. Further advances led to identification of natural forms in microbes, animals and humans 2–5, where methylcobalamin (MeCbl) and adenosylcobalamin (AdoCbl) serve as vital coenzymes for methionine synthase (MS) and methylmalonyl-CoA mutase (MCM), respectively.

In the diet, animal foods mainly contain protein-bound Ado-, Me- and hydroxycobalamin (HOCbl) 6,7, while supplements can contain free-form CNCbl or natural forms. CNCbl is still widely used owing to its low cost and stability. Cbl stability in water is affected by various factors (e.g. light, heat, pH, ROS, other vitamins, polyphenols, etc.), where MeCbl is often least and CNCbl most stable 8–11. However, Me-, Ado- and CNCbl can all undergo photoaquation to HOCbl, which was the most stable in saline 11 and main form in natural seawater 12.

While CNCbl has favourable stability, as an unnatural form the burden of proof surely lies with bioactivity—does it work as well? The efficacy of different B12 forms has been the topic of recent reviews 13,14, which I will add to.

Bioavailability

On consumption, Cbl is initially bound by haptocorrin (HC) in the oral–gastric phase and then by intrinsic factor in the small intestine, which enables active (receptor-mediated) absorption in the ileum; while a minor fraction of free Cbl can also be absorbed passively 15. In blood, Cbl is transported by HC and transcobalamin (TC), the latter of which mediates transport to tissues and so better reflects ‘active B12’.

Early human studies with radiolabelled Cbl (synthetic and natural forms) and whole-body monitoring generally reported similar oral absorption (refs in 16), or lower for low-dose AdoCbl 17,18; while injection of high-dose HOCbl had greater retention and non-specific binding to blood plasma proteins 18,19. In recent trials, low-dose oral CNCbl elevated plasma B12 and/or holoTC more than HOCbl over 2 days 16 and 8 weeks 20,21. In 1 trial (2x 2.8ug CNCbl vs dietary B12) this was related to CNCbl accumulation on HC, although both forms similarly lowered functional biomarkers (i.e. homocysteine and MMA) 20. Whereas in another trial, with more equivalent 3ug supplements, HOCbl seemed to lower MMA slightly better at 8 weeks 21.

Recent animal studies also show similar absorption, but different metabolic efficacy 22,23. In particular, while CNCbl (human equiv. 60ug/day) increased B12 more in blood and kidney, HOCbl increased B12 more in liver, with a similar trend in other organs, and was better converted to other forms 23. Several earlier in vitro studies also suggested greater efficacy of natural Cbl in cells and mitochondria (refs in 13,19). For instance, in human cells, both HO- and MeCbl had greater cellular uptake (2x after 4hrs) and conversion to coenzymes (over 24hr), while CNCbl seemed to have more reversal of attachment and internalisation 19,24; although all forms supported similar methionine synthesis (after 1hr 25) and cell division (over days 24).

Cell processing

Cbl metabolism is highly redox-dependent, where the cobalt coordination (a/b ligands) and oxidation states (I–III) determine reactivity 1. Inside cells, the cytosolic chaperone MMACHC binds Cbl in the ‘base-off’ state (i.e. a-axial displacement of DMBI) and mediates reductive removal of upper ligands (e.g. decyanation, dealkylation or denitration), generating Cbl(I/II) products which are oxidised to Cbl(II) or aquacob(III)alamin (OH2Cbl) 26–30. Cbl processing and trafficking may be facilitated by multiprotein complexes which include MMACHC, MMADHC, MS and MSR 31. MMADHC acts as branch point for Cbl delivery to the enzymes MS (cytosol) and MCM (mitochondria), which mediate conversion to coenzymes 32. MSR stabilises MS, via reduction of OH2Cbl, and maintains MS activity, via reduction of Cbl(II) 33.

Regarding reaction rates, dealkylation proceeds quicker with MeCbl than AdoCbl 29. HO- and CNCbl go through similar redox-dependent processing 26,30,33, although the latter is slower (e.g. MMACHC 30, MS 25,34 and in vivo 21,23); perhaps due to Co-CN binding affinity and a lower redox potential 1,9,35,36.

Cbl processing is promoted by glutathione 5,34,37,38 and related Nrf2 activity 39,40. Accordingly, decyanation of CNCbl requires NADPH 26 and is supported by glutathione 25,37 (i.e. aerobic binding and anaerobic metabolism 30), while removal of natural ligands requires glutathione 27–30,41. Also, on cellular uptake of the TC-Cbl complex into lysosomes and degradation, some Cbl may be released in free form 42,43. Free HO/OH2Cbl can readily complex with glutathione to form glutathionylcobalamin (GSCbl) 25,30,36; a putative protector 44,45, antioxidant 46 and precursor to coenzymes 25,36–38. In particular, GSCbl had quicker coenzyme activity than HO-, Me- or CNCbl 25 (although see 34). Subsequently, bovine MMACHC was shown to mediate deglutathionylation of GSCbl, via a glutathione-dependent reaction analogous to dealkylation of coenzymes 41. The reaction with GSCbl is also rapid, being an order of magnitude higher than with coenzymes 41.

Host factors

Genetic, environmental and disease-related factors can affect B12 metabolism, where differences between forms may become more relevant. Foremost of which, rare inborn errors affecting MMACHC (i.e. CblC disorders), where HOCbl 14,26,35 or CNCbl 27 may be advantageous. More prevalently, various SNPs may slow Cbl uptake or processing (e.g. TCN2 and MTRR) 5,13,47.

Some other factors may specifically affect CNCbl. For instance, since the 1960s, elevated CNCbl has been reported in several populations, including smokers, Leber’s and kidney disease 2,48. Further, a meta-analysis of RCTs linked high-dose CNCbl to adverse CVD outcomes in people with impaired renal function, obscuring benefit in those with good function 49; potentially also of relevance in elderly 50. A human brain autopsy study also found very elevated CNCbl levels (15-fold; 1:1 ratio with MeCbl) in fetal samples, which may be a result of prenatal supplements and/or decreased metabolism 5. Note, in animal models, CNCbl also accumulated in kidneys and brain 23, suggesting organ-specific metabolism.

Host cyanide metabolism may be important. Intracellular processing of CNCbl releases cyanide 26, which is mainly metabolised via sulfur-dependent pathways 51,52; including MST (cytosol and mitochondria) and TST (aka. rhodanese; mitochondria) catalysis to thiocyanate (SCN), and non-enzymatic reactions 53,54. Even 1mg of CNCbl provides little cyanide compared to background dietary levels 55 (see). Conversely however, increased basal cyanide levels, due to exposure (e.g. smoking, diet or SNP) or impaired metabolism (e.g. Leber’s and kidney disease), may affect Cbl status 2,48. Indeed, cyanide has extremely high affinity for HOCbl (> glutathione), resulting in conversion to CNCbl 36, and also reacts with coenzymes 56,57. Note, treatment of neuro-symptoms in 5 HD patients with intravenous MeCbl was associated with higher CNCbl (4x) and slightly lower thiocyanate 48. Ultimately, saturation of cyanide detox pathways may favour auxiliary formation of CNCbl, sequestering cyanide and Cbl, at the expense of coenzymes 48. Importantly, cyanide metabolism is organ-dependent, being notably lower in brain—a major target of toxicity—and more so in early and later life 51,58. As noted above, even the normal brain can accumulate CNCbl, where concern was raised as to whether it could act as a competitive antagonist 5,23.

B12 metabolism may be even more generally impaired in ageing and chronic disease. For instance, early observations suggested MeCbl markedly declined during ageing in plasma, liver and brain 2. More recently, a marked decline in total, GS-, Ado- and MeCbl was reported in the ageing human brain (i.e. 61–80yrs), which occurred prematurely in autism and schizophrenia 5. In the periphery, functional B12 deficiency increases with age in relation to amnionless 59 and ‘oxidant risks’ 60, the presence of which impaired response to high-dose CNCbl 60. Note, in other trials on older people, very high doses of CNCbl were required to lower functional biomarkers 61,62, as compared to younger people (trials above) 20,21.

Host redox biology may be important. A recent systematic review found some evidence of a relation between B12 and redox status 63. Reciprocally, an oxidised redox status in ageing and disease may affect B12 status 5,39,40,60. In particular, depletion of glutathione may impair brain transport and conversion to coenzymes 5,34,37–40. In ageing neurons, alternate splicing of MS (via deletion of exons 19 and 20) may also increase Cbl(I) susceptibility to oxidation 38. Moreover, many diseases involve chronic inflammation, neutrophil activation and production of hypochlorous acid (HOCl) 64, a potent oxidant which can destroy Cbl to release toxic products, including redox-active cobalt, cyanogen chloride and cyanide 64,65.

Intriguingly, in healthy adults, a positive correlation was found between cerebrospinal fluid B12 and 8-OHdG—a marker of DNA oxidation 66. The authors speculated as to whether this may involve CNCbl supplements and reduced brain cyanide detoxification capacity 66. Note, other aspects of multivitamin-mineral supplements can also promote oxidation 67.

Gut microbiome?

Cbl is synthesised exclusively by microbes, yet potential effects on the gut microbiome seem rarely considered. Human faecal samples contain a variety of corrinoids, less than 2% of which is Cbl 68. Over 80% of sequenced human gut bacteria are predicted to use corrinoids, where they serve diverse functions as enzyme cofactors and genetic regulators (e.g. corrinoid riboswitches). However, since less than 25% possess synthetic capacity, they may be a precious resource for commensals 68 and pathogens 69. Further competition between host and microbes may occur under conditions of SIBO 68,70, and influence bacterial pathogenesis at various body sites 69,70.

B12 is often supplemented at high doses (e.g. >100x RDA), due to poor absorption (relative to body stores), which exceed active transport (via intrinsic factor) and so are absorbed by passive diffusion at a rate of ~1% 15. Consequently, most will end up in the gut 68. In mice, high-dose CNCbl (human equiv. 5mg for 16 days) greatly increased gut Cbl, but lowered microbial corrinoid analogues 71. This treatment had no effect on microbial diversity, SCFAs or IBD, but did selectively deplete Bacteroides 71. A couple of studies have compared CN- and MeCbl. In a human colonic simulation, they differently shifted the microbiome and metabolism, with MeCbl appearing more favourable 72. While in mice, only CNCbl (human equiv. 1.5mg for 3 days) aggravated IBD, in relation to increased Enterobacteriaceae (e.g. E. coli), modulation of bacterial enzymes and riboswitches 73.

In sum, while synthetic CNCbl has favourable stability, perhaps this comes at the cost of slower metabolism and some other potentially unfavourable characteristics. Could this hinder its efficacy in chronic disease and ageing? Many authors have long favoured or suggested natural B12 forms (e.g. HO- or MeCbl) to circumvent the conditions and issues discussed above 2,5,13,48,50,60,72,74. With natural form supplements the burden of proof perhaps lies more on stability.

References

1.           Dereven’kov, I. A., Salnikov, D. S., Silaghi-Dumitrescu, R., Makarov, S. V. & Koifman, O. I. Redox chemistry of cobalamin and its derivatives. Coord. Chem. Rev. 309, 68–83 (2016).

2.           Linnell, J. C. & Matthews, D. M. Cobalamin metabolism and its clinical aspects. Clin. Sci. (Lond). 66, 113–21 (1984).

3.           van Kapel, J., Spijkers, L. J., Lindemans, J. & Abels, J. Improved distribution analysis of cobalamins and cobalamin analogues in human plasma in which the use of thiol-blocking agents is a prerequisite. Clin. Chim. Acta. 131, 211–24 (1983).

4.           Djalali, M. et al. High-performance liquid chromatographic separation and dual competitive binding assay of corrinoids in biological material. J. Chromatogr. 529, 81–91 (1990).

5.           Zhang, Y. et al. Decreased Brain Levels of Vitamin B12 in Aging, Autism and Schizophrenia. PLoS One 11, e0146797 (2016).

6.           Farquharson, J. & Adams, J. F. The forms of vitamin B12 in foods. Br. J. Nutr. 36, 127–136 (1976).

7.           Czerwonka, M., Szterk, A. & Waszkiewicz-Robak, B. Vitamin B12 content in raw and cooked beef. Meat Sci. 96, 1371–1375 (2014).

8.           Johns, P. W. et al. Cocoa polyphenols accelerate vitamin B12 degradation in heated chocolate milk. Int. J. Food Sci. Technol. 50, 421–430 (2015).

9.           Hadinata Lie, A., V Chandra-Hioe, M. & Arcot, J. Sorbitol enhances the physicochemical stability of B12 vitamers. Int. J. Vitam. Nutr. Res. 1–9 (2019). doi:10.1024/0300-9831/a000578

10.        Schnellbaecher, A., Binder, D., Bellmaine, S. & Zimmer, A. Vitamins in cell culture media: Stability and stabilization strategies. Biotechnol. Bioeng. 116, 1537–1555 (2019).

11.        Juzeniene, A. & Nizauskaite, Z. Photodegradation of cobalamins in aqueous solutions and in human blood. J. Photochem. Photobiol. B. 122, 7–14 (2013).

12.        Heal, K. R. et al. Determination of four forms of vitamin B12 and other B vitamins in seawater by liquid chromatography/tandem mass spectrometry. Rapid Commun. Mass Spectrom. 28, 2398–2404 (2014).

13.        Paul, C. & Brady, D. M. Comparative Bioavailability and Utilization of Particular Forms of B12 Supplements With Potential to Mitigate B12-related Genetic Polymorphisms. Integr. Med. (Encinitas). 16, 42–49 (2017).

14.        Obeid, R., Fedosov, S. N. & Nexo, E. Cobalamin coenzyme forms are not likely to be superior to cyano- and hydroxyl-cobalamin in prevention or treatment of cobalamin deficiency. Mol. Nutr. Food Res. 59, 1364–72 (2015).

15.        Carmel, R. How I treat cobalamin (vitamin B12) deficiency. Blood 112, 2214–21 (2008).

16.        Greibe, E. et al. Increase in circulating holotranscobalamin after oral administration of cyanocobalamin or hydroxocobalamin in healthy adults with low and normal cobalamin status. Eur. J. Nutr. 57, 2847–2855 (2018).

17.        Adams, J. F., Ross, S. K., Mervyn, L., Boddy, K. & King, P. Absorption of cyanocobalamin, coenzyme B 12 , methylcobalamin, and hydroxocobalamin at different dose levels. Scand. J. Gastroenterol. 6, 249–52 (1971).

18.        HEINRICH, H. C. & GABBE, E. E. METABOLISM OF THE VITAMIN B12 COENZYME IN RATS AND MAN. Ann. N. Y. Acad. Sci. 112, 871–903 (1964).

19.        Hall, C., Begley, J. & Green-Colligan, P. The availability of therapeutic hydroxocobalamin to cells. Blood 63, 335–341 (1984).

20.        Naik, S. et al. Cyano-B12 or Whey Powder with Endogenous Hydroxo-B12 for Supplementation in B12 Deficient Lactovegetarians. Nutrients 11, (2019).

21.        Greibe, E. et al. Effect of 8-week oral supplementation with 3-µg cyano-B12 or hydroxo-B12 in a vitamin B12-deficient population. Eur. J. Nutr. 58, 261–270 (2019).

22.        Greibe, E., Nymark, O., Fedosov, S. N., Heegaard, C. W. & Nexo, E. Dietary Intake of Vitamin B12 is Better for Restoring a Low B12 Status Than a Daily High-Dose Vitamin Pill: An Experimental Study in Rats. Nutrients 10, 1096 (2018).

23.        Greibe, E. et al. The tissue profile of metabolically active coenzyme forms of vitamin B12 differs in vitamin B12-depleted rats treated with hydroxo-B12 or cyano-B12. Br. J. Nutr. 120, 49–56 (2018).

24.        Chu, R. C., Begley, J. A., Colligan, P. D. & Hall, C. A. The methylcobalamin metabolism of cultured human fibroblasts. Metabolism. 42, 315–9 (1993).

25.        Pezacka, E., Green, R. & Jacobsen, D. W. Glutathionylcobalamin as an intermediate in the formation of cobalamin coenzymes. Biochem. Biophys. Res. Commun. 169, 443–50 (1990).

26.        Kim, J., Gherasim, C. & Banerjee, R. Decyanation of vitamin B12 by a trafficking chaperone. Proc. Natl. Acad. Sci. U. S. A. 105, 14551–4 (2008).

27.        Gherasim, C., Ruetz, M., Li, Z., Hudolin, S. & Banerjee, R. Pathogenic mutations differentially affect the catalytic activities of the human B12-processing chaperone CblC and increase futile redox cycling. J. Biol. Chem. 290, 11393–402 (2015).

28.        Mascarenhas, R., Li, Z., Gherasim, C., Ruetz, M. & Banerjee, R. The human B12 trafficking protein CblC processes nitrocobalamin. J. Biol. Chem. 295, 9630–9640 (2020).

29.        Kim, J., Hannibal, L., Gherasim, C., Jacobsen, D. W. & Banerjee, R. A human vitamin B12 trafficking protein uses glutathione transferase activity for processing alkylcobalamins. J. Biol. Chem. 284, 33418–24 (2009).

30.        Li, Z., Gherasim, C., Lesniak, N. A. & Banerjee, R. Glutathione-dependent one-electron transfer reactions catalyzed by a B₁₂ trafficking protein. J. Biol. Chem. 289, 16487–97 (2014).

31.        Bassila, C. et al. Methionine synthase and methionine synthase reductase interact with MMACHC and with MMADHC. Biochim. Biophys. acta. Mol. basis Dis. 1863, 103–112 (2017).

32.        Mah, W. et al. Subcellular location of MMACHC and MMADHC, two human proteins central to intracellular vitamin B(12) metabolism. Mol. Genet. Metab. 108, 112–8 (2013).

33.        Yamada, K., Gravel, R. A., Toraya, T. & Matthews, R. G. Human methionine synthase reductase is a molecular chaperone for human methionine synthase. Proc. Natl. Acad. Sci. U. S. A. 103, 9476–81 (2006).

34.        Waly, M. I., Kharbanda, K. K. & Deth, R. C. Ethanol lowers glutathione in rat liver and brain and inhibits methionine synthase in a cobalamin-dependent manner. Alcohol. Clin. Exp. Res. 35, 277–83 (2011).

35.        Froese, D. S., Zhang, J., Healy, S. & Gravel, R. A. Mechanism of vitamin B12-responsiveness in cblC methylmalonic aciduria with homocystinuria. Mol. Genet. Metab. 98, 338–43 (2009).

36.        Xia, L., Cregan, A. G., Berben, L. A. & Brasch, N. E. Studies on the formation of glutathionylcobalamin: any free intracellular aquacobalamin is likely to be rapidly and irreversibly converted to glutathionylcobalamin. Inorg. Chem. 43, 6848–57 (2004).

37.        Zhang, Y., Hodgson, N., Trivedi, M. & Deth, R. Neuregulin 1 Promotes Glutathione-Dependent Neuronal Cobalamin Metabolism by Stimulating Cysteine Uptake. Oxid. Med. Cell. Longev. 2016, 3849087 (2016).

38.        Waly, M. et al. Alternatively Spliced Methionine Synthase in SH-SY5Y Neuroblastoma Cells: Cobalamin and GSH Dependence and Inhibitory Effects of Neurotoxic Metals and Thimerosal. Oxid. Med. Cell. Longev. 2016, 6143753 (2016).

39.        Schrier, M., Rose, N., Zhang, Y., Trivedi, M. & Deth, R. The role of Nrf2 in redox-dependent cobalamin processing. Free Radic. Biol. Med. 128, S92–S93 (2018).

40.        Schrier, M. S., Trivedi, M. S. & Deth, R. C. Redox-related epigenetic mechanisms in glioblastoma: Nuclear factor (erythroid-derived 2)-like 2, cobalamin, and dopamine receptor subtype 4. Front. Oncol. 7, 1–14 (2017).

41.        Jeong, J., Park, J., Park, J. & Kim, J. Processing of glutathionylcobalamin by a bovine B12 trafficking chaperone bCblC involved in intracellular B12 metabolism. Biochem. Biophys. Res. Commun. 443, 173–8 (2014).

42.        Quadros, E. V & Jacobsen, D. W. The dynamics of cobalamin utilization in L-1210 mouse leukemia cells: a model of cellular cobalamin metabolism. Biochim. Biophys. Acta 1244, 395–403 (1995).

43.        Hannibal, L. et al. Transcellular transport of cobalamin in aortic endothelial cells. FASEB J. 32, 5506–5519 (2018).

44.        Watson, W. P., Munter, T. & Golding, B. T. A new role for glutathione: protection of vitamin B12 from depletion by xenobiotics. Chem. Res. Toxicol. 17, 1562–7 (2004).

45.        Dereven’kov, I. A., Makarov, S. V, Shpagilev, N. I., Salnikov, D. S. & Koifman, O. I. Studies on reaction of glutathionylcobalamin with hypochlorite. Evidence of protective action of glutathionyl-ligand against corrin modification by hypochlorite. Biometals 30, 757–764 (2017).

46.        Birch, C. S., Brasch, N. E., McCaddon, A. & Williams, J. H. H. A novel role for vitamin B12: Cobalamins are intracellular antioxidants in vitro. Free Radic. Biol. Med. 47, 184–188 (2009).

47.        Gueant, J.-L. et al. Environmental influence on the worldwide prevalence of a 776C->G variant in the transcobalamin gene (TCN2). J. Med. Genet. 44, 363–367 (2007).

48.        Koyama, K. et al. Abnormal cyanide metabolism in uraemic patients. Nephrol. Dial. Transplant. 12, 1622–1628 (1997).

49.        Spence, J. D., Yi, Q. & Hankey, G. J. B vitamins in stroke prevention: time to reconsider. Lancet Neurol. 16, 750–760 (2017).

50.        Spence, J. D. Harm With High Levels of Serum B12 in Elderly Persons. J. Gerontol. A. Biol. Sci. Med. Sci. 74, 137 (2019).

51.        McMahon, T. F. & Birnbaum, L. S. Age-related changes in toxicity and biotransformation of potassium cyanide in male C57BL/6N mice. Toxicol. Appl. Pharmacol. 105, 305–14 (1990).

52.        Tor-Agbidye, J. et al. Bioactivation of cyanide to cyanate in sulfur amino acid deficiency: relevance to neurological disease in humans subsisting on cassava. Toxicol. Sci. 50, 228–35 (1999).

53.        Gyamfi, O. A. et al. Metabolism of Cyanide by Glutathione to Produce the Novel Cyanide Metabolite 2-Aminothiazoline-4-oxoaminoethanoic Acid. Chem. Res. Toxicol. 32, 718–726 (2019).

54.        Nagahara, N., Ito, T. & Minami, M. Mercaptopyruvate sulfurtransferase as a defense against cyanide toxication: molecular properties and mode of detoxification. Histol. Histopathol. 14, 1277–86 (1999).

55.        Schrenk, D. et al. Evaluation of the health risks related to the presence of cyanogenic glycosides in foods other than raw apricot kernels. EFSA J. 17, (2019).

56.        Brodie, S. J., Cregan, A. G., van Eldik, R. & Brasch, N. E. The reaction between methylcobalamin and cyanide revisited. Inorganica Chim. Acta 348, 221–224 (2003).

57.        Hamza, M. S. A., Zou, X., Brown, K. L. & Van Eldik, R. Equilibrium and kinetic studies on the reactions of alkylcobalamins with cyanide. Inorg. Chem. 40, 5440–5447 (2001).

58.        Sani, M. et al. Age-related changes in the activity of cerebral rhodanese in mice during the first four months of life. Brain Dev. 30, 279–86 (2008).

59.        Pannérec, A. et al. Vitamin B12 deficiency and impaired expression of amnionless during aging. J. Cachexia. Sarcopenia Muscle 9, 41–52 (2018).

60.        Solomon, L. R. Functional cobalamin (vitamin B12) deficiency: role of advanced age and disorders associated with increased oxidative stress. Eur. J. Clin. Nutr. 69, 687–92 (2015).

61.        Blacher, J. et al. Very low oral doses of vitamin B-12 increase serum concentrations in elderly subjects with food-bound vitamin B-12 malabsorption. J. Nutr. 137, 373–8 (2007).

62.        Eussen, S. J. P. M. et al. Oral cyanocobalamin supplementation in older people with vitamin B12 deficiency: a dose-finding trial. Arch. Intern. Med. 165, 1167–72 (2005).

63.        Lagemaat, E. E. van de, Groot, L. C. P. G. M. de & Heuvel, E. G. H. M. van den. Vitamin B12 in Relation to Oxidative Stress: A Systematic Review. Nutr. 2019, Vol. 11, Page 482 11, 482 (2019).

64.        Abu-Soud, H. M. et al. The reaction of HOCl and cyanocobalamin: corrin destruction and the liberation of cyanogen chloride. Free Radic. Biol. Med. 52, 616–25 (2012).

65.        Jeelani, R. et al. Melatonin prevents hypochlorous acid-mediated cyanocobalamin destruction and cyanogen chloride generation. J. Pineal Res. 64, (2018).

66.        Guest, J. et al. Novel relationships between B12, folate and markers of inflammation, oxidative stress and NAD(H) levels, systemically and in the CNS of a healthy human cohort. Nutr. Neurosci. 18, 355–64 (2015).

67.        Rabovsky, A. B., Buettner, G. R. & Fink, B. In vivo imaging of free radicals produced by multivitamin-mineral supplements. BMC Nutr. 1, 1–9 (2015).

68.        Degnan, P. H., Taga, M. E. & Goodman, A. L. Vitamin B12 as a modulator of gut microbial ecology. Cell Metab. 20, 769–78 (2014).

69.        Cordonnier, C. et al. Vitamin B12 Uptake by the Gut Commensal Bacteria Bacteroides thetaiotaomicron Limits the Production of Shiga Toxin by Enterohemorrhagic Escherichia coli. Toxins (Basel). 8, (2016).

70.        Rowley, C. A. & Kendall, M. M. To B12 or not to B12: Five questions on the role of cobalamin in host-microbial interactions. PLoS Pathog. 15, e1007479 (2019).

71.        Kelly, C. J. et al. Oral vitamin B12 supplement is delivered to the distal gut, altering the corrinoid profile and selectively depleting Bacteroides in C57BL/6 mice. Gut Microbes 10, 654–662 (2019).

72.        Xu, Y. et al. Cobalamin (Vitamin B12) Induced a Shift in Microbial Composition and Metabolic Activity in an in vitro Colon Simulation. Front. Microbiol. 9, 2780 (2018).

73.        Zhu, X. et al. Impact of Cyanocobalamin and Methylcobalamin on Inflammatory Bowel Disease and the Intestinal Microbiota Composition. J. Agric. Food Chem. 67, 916–926 (2019).

74.        Freeman, A. G. Cyanocobalamin - a case for withdrawal: discussion paper. J. R. Soc. Med. 85, 686–87 (1992).

 

No comments:

Post a Comment