Open access

Abnormal Folate Metabolism and Maternal Risk for Down Syndrome

Written By

Érika Cristina Pavarino, Bruna Lancia Zampieri, Joice Matos Biselli and Eny Maria Goloni Bertollo

Submitted: 03 November 2010 Published: 29 August 2011

DOI: 10.5772/19373

From the Edited Volume

Genetics and Etiology of Down Syndrome

Edited by Subrata Dey

Chapter metrics overview

3,880 Chapter Downloads

View Full Metrics

1. Introduction

Down syndrome (DS) or trisomy 21 (MIM 190685) is the most common genetic disorder with a prevalence of 1 in 660 live births (Jones, 2006). DS is the leading cause of genetically-defined intellectual disability (Contestabile et al., 2010) and its phenotype is complex and variable among individuals, who may present with a combination of dysmorphic features (Ahmed et al., 2005; Pavarino-Bertelli et al., 2009), congenital heart disease (Abbag, 2006), neurological abnormalities such as early manifestations of Alzheimer’s disease (Lott & Head, 2005), immunological impairments (Ram & Chinen, 2011), elevated risk of specific types of leukemia (Hasle et al., 2000), and other clinical complications (Venail et al., 2004).

Trisomy 21 can be caused by three types of chromosomal abnormalities: free trisomy, translocation, or mosaicism. Mosaicism accounts for the minority of DS cases (about 1%) and is characterized by some cells containing 46 chromosomes and others, 47 chromosomes. Translocations are attributed to 3-4% of the cases, with Robertsonian translocation involving chromosomes 14 and 21 being the most common type. Finally, free trisomy occurs in about 95% of cases (Ahmed et al., 2005; J.M. Biselli et al., 2008b) and is characterized by the presence of three complete copies of chromosome 21.

Free trisomy, the main chromosomal abnormality leading to DS, is caused by the failure of normal chromosome 21 segregation during meiosis (meiotic nondisjunction) (Hassold & Hunt, 2000). The parental origin of the extra chromosome 21 is maternal in about 80% of cases (Jyothy et al., 2001), and most (about 77%) occur during the first maternal meiotic division in the maturing oocyte, before conception (Antonarakis et al., 1992).

Advertisement

2. Meiosis and chromosomal segregation

Faithful transmission of a genome from one generation to another depends on the mechanism of cell division in which each pair of replicated chromosomes is separated and equally distributed to mother and daughter cells. Meiosis generates haploid gametes through a specialized cell division process that consists of one round of DNA replication followed by two cell divisions. The first division, meiosis I (MI), involves the segregation of homologous chromosomes from each other, whereas meiosis II (MII) involves the segregation of the sister chromatids (Hassold & Hunt, 2000).

Timing of chromosome attachment and loss of cohesion is essential to faithful chromosome segregation. During MI, the cohesion between sister chromatid arms assures physical attachment by the chiasmata of homologous chromosomes, ensuring their alignment on the meiosis-I spindle, and maintains them at the site of recombination. Chiasmata are resolved at anaphase I by the loss of cohesion between the arms of sister chromatids in the homologous chromosomes; the chromosomes then segregate to opposite poles of the cell. Cohesion, however, must be maintained at centromeres between sister chromatids beyond meiosis I to prevent premature chromatid separation (predivision) and ensure proper attachment of the sister chromatids to opposite spindle poles in meiosis II (Barbero, 2011; Sakuno & Watanabe, 2009; Vogt et al., 2008).

The centromeric cohesion during meiosis I results from the attachment of kinetochores of sister chromatids to only one spindle pole (Sakuno & Watanabe, 2009). Kinetochores are situated on opposite sides of the centromeric heterochromatin at the centromeres of each sister chromatid and they capture and stabilize microtubules for the formation of kinetochore fibers, only then they are capable of chromosome bi-orientation during the metaphase and chromosome segregation during the anaphase of meiosis (Vogt et al., 2008).

During cell division, several chromosomal mal-segregation mechanisms can occur. Classical nondisjunction is due to the failure to resolve chiasmata between homologous chromosomes, whereby both homologues segregate together. In addition, premature resolution of chiasmata or the failure to establish a chiasma between a pair of homologues results in the independent segregation of homologues at MI, which leads to an error if both segregate to the same pole of the MI spindle. A MI error can also involve the segregation of sister chromatids, rather than homologous chromosomes, whereby the premature separation of sister chromatids at MI can result in the segregation of a whole chromosome and a single chromatid to one of the poles. At MII, errors result from the failure of sister chromatid separation (Hassold & Hunt, 2000).

Advertisement

3. The origin of maternal chromosome 21 nondisjunction

The molecular mechanisms involved in meiotic nondisjunction leading to trisomy 21 are still poorly understood and the only well-established risk factor for DS is advanced maternal age at conception (35 years or older) (Allen et al., 2009; Jyothy et al., 2001; Lamb et al., 2005). Studies have suggested many explanations for the maternal age-associated increase in aneuploidy. One model attributes the effect of advanced maternal age to the uterine environment, indicating that there might be an age-related decline in the ability to recognize and then abort trisomic fetuses (Aymé & Lippman-Hand, 1982; Stein et al., 1986). However, the observation that the advanced maternal age effect is restricted to chromosome 21 nondisjunction of maternal origin, but not associated with cases resulting from sperm or post-zygotic mitotic errors, suggests that the uterus is the source of the age effect (Allen et al., 2009).

On the other hand, Zheng & Byers (1993) proposed that age-dependent trisomy 21 results primarily from a mechanism that favors maturation and utilization of euploid oocytes over the pre-existing aneuploid products of mitotic (premeiotic) nondisjunction at an early stage of the reproductive lifespan. In addition, decreased expression of checkpoint proteins in aging oocytes (Vogt et al., 2008) and failure to effectively replace cohesion proteins that are lost from chromosomes during aging (Chiang et al., 2010) also are pointed out as risk factors for predisposing oocytes to errors in chromosome segregation.

A link between altered recombination and maternal age-related nondisjunction has been described. It was observed that recombination is reduced among nondisjoined chromosomes 21 at MI, and this reduction seems to be age-related (Sherman et al., 1994). Lamb et al. (1996) proposed that at least two “hits” are required for chromosome 21 nondisjuntion: (1) the establishment in the fetal ovary of a susceptible pattern of meiotic recombination, and (2) the abnormal processing of susceptible chromosomes in the adult ovary. The second “hit” would involve degradation of a meiotic process (e.g., a spindle component, a sister chromatid cohesion protein, a meiotic motor protein, a checkpoint control protein) that increases the risk of improper segregation for these susceptible bivalents (Hassold & Sherman, 2000). Further studies have shown susceptible patterns of chromosome 21 meiotic recombination, including pericentromeric and telomeric exchanges, described as maternal risk factors for DS even in young DS mothers (Gosh et al., 2009; Lamb et al., 2005).

Besides advanced maternal age, the age of the maternal grandmother at the time of birth of the mother has also been pointed out as a risk factor for the occurrence of DS. At an advanced age, the grandmother's reproductive system may fail to make the essential proteins needed for proper meiotic segregation in the germ cells of her daughter, leading to nondisjunction of chromosome 21 during the embryogenesis of DS child’s mother when she was in the grandmother's womb (Malini & Ramachandra, 2006). However, more recent studies failed to support the suggestion that advanced age of the DS grandmother is responsible for meiotic disturbances in her daughter (Allen et al., 2009; Kovaleva et al., 2010).

Although the risk of bearing a child with DS increases substantially with increasing maternal age, many DS children are born to mothers aged less than 35 years-old, suggesting other risk factors influencing DS etiology. In 1999, James et al. produced the first evidence that the occurrence of DS independent of maternal age is associated with DNA hypomethylation due to impairments in folate metabolism.

Advertisement

4. Folate metabolism

Folate represents an essential nutrition component in the human diet, and is involved in many metabolic pathways, mainly the folate metabolism, i.e., a single-carbon transfer from one molecule to another through a series of interconnected biochemical reactions. Folate is a generic term for a family of compounds present in most foods, e.g., legumes, leafy greens, some fruits, vegetables (e.g., spinach, broccoli, asparagus, and lettuce), liver, milk, and dairy products (Lin & Young, 2000). Humans, as all mammals, are unable to synthesize folate, thus its ingestion, either from normal diet or nutritional supplements, is very important. After intestinal absorption, natural folate, known as polyglutamate, requires reduction into monoglutamate by conjugases in the small intestine before it can be absorbed. On the other hand, in its synthetic form, folic acid exists as monoglutamate and does not need to be reduced for release into the blood and cellular uptake (Bailey & Gregory, 1999; Hall & Solehdin, 1998). Another disadvantage of natural food folate is its poor stability especially under typical cooking conditions, which can substantially reduce the vitamin content before it is even ingested, a significant additional factor limiting the ability of natural food folates to enhance folate status (McNulty & Pentieva, 2004; McNulty & Scott, 2008).

Folate metabolism is a complex metabolic pathway that involves multiple enzymes and water-soluble B vitamins such as folate, vitamin B6 and vitamin B12, that play key roles as enzyme cofactors or substrates in this metabolism. It includes two main cycles: purine and pyrimidine synthesis, necessary for synthesis and repair of DNA, and DNA methylation, an epigenetic process that acts on the control associated with gene expression and genomic stability essential for normal cellular methylation reactions (Figure 1).

Figure 1.

Folate metabolism. BHMT = Betaine-homocysteine methyltransferase; B6 = vitamin B6; B12 = vitamin B12; CβS = Cystathionine β- synthase; CH3 = Methyl; dATP = Deoxyadenosine 5’-triphosphate; dGTP = Deoxyguanosine 5’-triphosphate; dTTP = Deoxythymidine 5’-triphosphate; DHF = Dihydrofolate; DHFR = Dihydrofolate reductase; Hcy = Homocysteine; MTHFD1 = Methylenetetrahydrofolate dehydrogenase 1; MTHFR = Methylenetetrahydrofolate reductase; MTR = Methionine synthase; MTRR = Methionine synthase reductase; RFC1 = Reduced folate carrier 1; SAH = S-adenosylhomocysteine; SAM = S- adenosylmethionine; cSHMT= Serine hydroxymethyltransferase; TC2 = Transcobalamin 2; THF = Tetrahydrofolate.

Folate requires several transport systems to enter the cells and the one best characterized is the reduced folate carrier (RFC1), an enzyme located on intestinal cell membranes that carries out the transport of 5-methyltetrahydrofolate (5-methyl-THF) to the interior of a variety of cells, representing an important determinant of folate concentration in the interior of cells (Nguyen et al., 1997). In addition to the folate transport system, several genes and their respective enzymes play important roles in folate metabolism. The Dihydrofolate reductase (DHFR) gene encodes an enzyme that catalyzes the conversion of dihydrofolate (DHF) into tetrahydrofolate (THF) (Stanisiawska-Sachadyn et al., 2008), which is then converted into the corresponding 10-formyl, 5,10-methenyl, and 5,10-methylene derivatives by Methylenetetrahydrofolate dehydrogenase 1 (MTHFD1), a trifunctional nicotinamide adenine dinucleotide phosphate-dependent cytoplasmic enzyme. The donor cofactors for de novo purine and pyrimidine biosynthesis and, thus, the biosynthesis of DNA (Hum, 1988) are 10-formyl-THF and 5,10-methylene-THF. By an alternative route, THF is converted into 5,10-methylene-THF and glycine by the cytosolic form of the enzyme Serine hydroxymethyltransferase (cSHMT) (Steck et al., 2008).

Methylenetetrahydrofolate reductase (MTHFR) is responsible for the conversion of 5,10-methylene-THF to 5-methyl-THF, the main circulating form of folate that donates methyl groups for homocysteine (Hcy) remethylation into methionine. This latter reaction is catalyzed by the enzyme Methionine synthase (MTR), which requires vitamin B12 or cobalamin (Cbl) as a cofactor, and results in the formation of S-adenosylmethionine (SAM), the primary methyl (CH3) donor for DNA methylation reactions (Finkelstein & Martin, 2000). SAM is demethylated to form S-adenosylhomocysteine (SAH) and then hydrolyzed to form adenine and Hcy. The DNA methyltransferase (DNMTs) enzymes catalyze the transfer of the methyl group, obtained from conversion of SAM into SAH, to position 5’ of cytosine residues located mainly in dinucleotide cytosine-guanine (CpG) (Bestor, 2000; De Angelis et al., 2008).

Methionine synthase reductase (MTRR), an enzyme codified by the MTRR gene, is responsible for the maintenance of the active form of the enzyme MTR. During remethylation of Hcy to methionine, a reaction catalyzed by MTR, methylcob(III)alamine acts as a methyl donor. In this reaction, the transfer of a methyl group from methylcob(III)alamine results in the formation of highly reactive cob(I)alamine, which is oxidized into cob(II)alamine, resulting in MTR inactivation (Yamada et al., 2006). In this inactivation process, a complex is formed between the enzymes MTR and MTRR, and derivative electrons from the oxidation of nicotinamide adenine dinucleotide phosphate (NADPH), catalyzed by MTR, are transferred to the inactive form of MTR. This process favors the transfer of methyl from the SAM to the MTR enzyme, resulting in methylcob(III)alamine, thus reestablishing MTR activity (Leclerc et al., 1999; Olteanu et al., 2001, 2002).

Betaine-homocysteine methyltransferase (BHMT) catalyses the conversion of Hcy to methionine by an alternative pathway of remethylation using the amino acid bethaine as methyl donor. When the Hcy folate-dependent remethylation catalyzed by the MTR enzyme is impaired by genetics or environmental factors, the BHMT enzyme plays an important role maintaining the homeostasis of Hcy (Pajares & Pérez-Salab, 2006).

In the transsulfuration cycle, Hcy is converted into cystathionine by Cystathionine β-synthase (CβS), a vitamin B6-dependent enzyme, and then into cysteine (Kraus et al., 1998). Under normal physical conditions, all Hcy is remethylated into methionine or catalyzed into cystathionine. The increase of Hcy concentration represents impairment in folate metabolism and thus in methylation reactions (Fenech, 2002).

Besides the enzymes that act directly on folate metabolism, cobalamin-transporting proteins also play an important role in this metabolic pathway, since the MTR enzyme is cobalamin-dependent. The enzyme Transcobalamin 2 (TC2) is synthesized in the intestinal villi and binds itself to Cbl in the interstitial fluid. This formed complex goes into the intestinal villi microcirculation and then reaches the systemic circulation. This circulation distributes the vitamin to all tissues where specific receptors on cell membranes bind and internalize the TC2-Cbl complex by endocytosis (Quadros et al., 1999; Seetharam & Li, 2000).

Advertisement

5. Folate metabolism, genomic stability, and maternal risk for chromosome 21 nondisjunction

Based on evidence that stable centromeric DNA chromatin may depend on the epigenetic inheritance of specific centromeric methylation patterns and on the binding of specific methyl-sensitive proteins to maintain the higher order DNA architecture necessary for kinetochore assembly (Karpen & Allshire, 1997), James et al. (1999) hypothesized that pericentromeric hypomethylation, resulting from impaired folate metabolism secondary to polymorphism of the MTHFR gene, could impair chromosomal segregation and increase the risk for chromosome 21 nondisjunction in young mothers. They observed that the risk of having a child with DS was 2.6-fold higher in mothers with 677 C→T substitution in one or both alleles of the MTHFR gene than in mothers without the 677 C→T substitution. In addition, DS mothers displayed a significant increase in plasma Hcy concentrations and lymphocyte methotrexate cytotoxicity, consistent with abnormal folate and methyl metabolism.

As described above, the MTHFR enzyme plays an important role in regulating DNA methylation through the reduction of 5,10-methylene-THF to 5-methyl-THF (Figure 1). The 677 C→T polymorphism is known to decrease the affinity of the enzyme for the flavin-adenine-dinucleotide (FAD) cofactor, decreasing enzyme activity (Guenther et al., 1999; Yamada et al., 2001). The MTHFR 677 CT genotype seems to reduce enzyme activity by about 35% and the homozygous TT genotype by 70% (Frosst et al., 1995). Since the study by James et al. (1999), polymorphisms in the MTHFR gene are the most frequently investigated in attempt to clarify the role of folate and methyl metabolism in the maternal risk for DS (Martínez-Frías et al., 2008). Several studies have associated the MTHFR 677C→T polymorphism and the risk of bearing a child with DS (da Silva et al., 2005; Meguid et al., 2008; Sadiq et al., 2011; Wang et al., 2008) as well as with increasing plasma Hcy concentration (P.M. Biselli et al., 2007; da Silva et al., 2005; Narayanan et al., 2004; Ulvik et al., 2007).

Another common polymorphism in the MTHFR gene, the substitution of alanine for cytosine at the 1298 position, was already associated with DS risk and increased plasma Hcy concentration (Martínez-Frías et al., 2006; Meguid et al., 2008; Narayanan et al., 2004, Rai et al., 2006; Scala et al., 2006; Weisberg et al., 2001). This polymorphism proved to have an impact on enzyme activity resulting in an even more pronounced decrease in its activity in homozygous 1298 CC compared to the heterozygous individuals (van der Putt et al., 1998).

In addition to the MTHFR gene, other genetic polymorphisms involved in the folate pathway seem to modulate the maternal risk for bearing a child with DS (Bosco et al., 2003; J.M. Biselli et al., 2008a; Meguid et al., 2008; Pozzi et al., 2009; Sadiq et al., 2011; Scala et al., 2006; Wang et al., 2008) as well as the concentrations of metabolites involved in the folate pathway (Ananth et al. 2007; Barbosa et al., 2008; Cheng et al., 2010; Devos et al., 2008). The MTR 2756 A→G polymorphism has been associated with increased maternal risk for DS in the presence of AG or GG genotypes, as well as when combined with polymorphisms MTRR 66 A→G (MTR 2756AG/MTRR 66AG) (Bosco et al., 2003) and MTHFR 677 C→T (MTHFR 677TT/MTR 2756AA). In addition, the allele MTR 2756 G proved to be more frequent, both in homozygosis and heterozygosis, in DS mothers as compared to mothers of individuals without the syndrome (Pozzi et al., 2009). Concerning its influence on Hcy concentrations, studies have shown conflicting results, since some have associated the MTR 2756 A allele to increased Hcy concentration (Fredriksen et al., 2007; Harmon et al., 1999), while others found the same association, but with the polymorphic 2756 G alelle (Feix et al., 2001; Fillon-Emery et al., 2004).

As to the MTRR 66 A→G polymorphism, some studies have supported an independent role for this polymorphism in the maternal risk for DS in the presence of the homozygous MTRR 66 GG genotype (Hobbs et al., 2000; Pozzi et al., 2009; Wang et al., 2008). Most of the studies have associated this polymorphism with the risk of DS and increased Hcy concentration when combined to other polymorphisms, such as MTHFR 677 C→T (Hobbs et al., 2000; Martínez-Frías et al., 2006; O’Leary et al, 2002; Yang et al., 2008). Additionally, a steady state kinetic analysis showed a significantly decreased affinity of MTRR for MTR accompanying substitution 66 A→G, revealing a significant difference in the relative efficacies of the MTRR enzyme (Olteanu et al., 2002). However, several studies have failed to find association between DS risk and the MTRR 66 A→G polymorphism, whether alone or combined with other genetic variants ( Coppedè et al., 2009 ; Chango et al., 2005; Scala et al., 2006).

The RFC1 gene is polymorphic at nucleotide 80 (A→G), and investigation of the impact of this polymorphism on protein function have demonstrated a difference in its affinity for subtracts and/or efficiency in transport in comparison with the wild type enzyme (Whetstine et al., 2001). Few studies have evaluated the influence of the RFC1 80 A→G polymorphism on DS risk (J.M. Biselli, 2008a, 2008c; Chango et al., 2005; Coppedè et al., 2006). Some studies have found no association between this polymorphism and DS (Chango et al. 2005; Fintelman-Rodrigues et al., 2009); however, Coppedè et al. (2006) and J.M. Biselli et al. (2008a) suggest a role for this polymorphism when combined with other polymorphisms in genes involved in folate metabolism. Supporting this hypothesis, the combined RFC1 80 GG/MTHFR 677 TT genotype has been associated with increased Hcy concentration and the RFC1 80 AA/MTHFR 677 CT combined genotype with higher plasma folate concentration (Chango et al., 2000).

A common polymorphism in the CβS gene, 68-base pair (bp) insertion at nucleotide position 844 (844ins68), is also investigated in the risk for DS, but there is no evidence that this variant plays an independent role on this risk (da Silva et al., 2005; Chango et al., 2005; Scala et al., 2006). The CβS 844ins68 polymorphism has been associated with reduction of Hcy concentration in the presence of the insertion (Tsai et al., 1996; Tsai et al., 1999; Tsai et al., 2000), and it is believed that this insertion is related to increased enzyme activity (Tsai et al., 1996, Tsai et al., 1999). This variant is always found to be associated in cis with an additional polymorphism in the CβS gene, a thymine-to-cytosine transition at nucleotide position 833, which causes a threonine-to-isoleucine amino acid substitution, and is reported, together with CβS 844ins68, as a 833 T→C/844ins68 in cis double mutation (Pepe et al., 1999; Vyletal et al., 2007). Da Silva et al., (2005) observed that the 844ins68 polymorphism, in association with other polymorphisms of the folate pathway, is related to increased risk for DS. Concerning its influence on folate metabolite concentrations, such as folate, Hcy, and vitamin B12, the CβS 844ins68 polymorphism showed no significant association with any of the biochemical variables involved in folate metabolism (Bowron et al., 2005; Kumar et al., 2010; Summers et al., 2008).

The MTHFD1 gene presents a functional polymorphism, a guanine-to-adenine substitution at position 1958 (1958 G→A), that has been shown to reduce the activity and stability of the variant enzyme (Christensen et al., 2008). There are only two studies to date on the influence of this polymorphism on maternal risk for DS. Scala et al. (2006) showed an association of the MTHFD1 1958 AA genotype with DS risk, but only when combined with the RFC1 80 GG genotype; however, more recently, Neagos et al. (2010) failed to find association. Thus, further investigations are necessary to clarify the role of MTHFD1 1958 G→A in the chromosome 21 nondisjunction.

Johnson et al. (2004) described a 19-base pair (bp) deletion polymorphism in intron-1 of the DHFR gene and hypothesized that this polymorphism could be functional since the deletion removes a possible transcription factor binding site that affects gene regulation. A study with mothers of individuals with spina bifida showed that the expression of the messenger ribonucleic acid (mRNA) from the DHFR gene was 50% higher in the presence of del/del genotype than in the ins/ins genotype (Parle-McDermott et al., 2007). This polymorphism has been associated with the modulation of metabolites’ concentrations involved in the folate pathway. Gellekink et al. (2007) reported association between the del/del genotype and reduction of plasma Hcy concentration, but found no association between this genotype and concentrations of serum and erythrocyte folate. Another study found no effect on Hcy concentration, but found increased plasma and erythrocyte folate levels in del/del individuals (Stanislawska-Sachadyn et al., 2008). The results of the only study that investigated the 19-bp deletion polymorphism of DHFR gene in DS mothers did not support an association between this variant and the maternal risk for DS. In addition, the polymorphism was not associated with variations in serum folate and plasma Hcy and methylmalonic acid (MMA) concentrations in the study population (Mendes et al., 2010).

The TC2 gene, which codifies a transporting protein required for the cellular uptake of vitamin B12 (Seetharam &Li, 2000), is polymorphic at nucleotide position 776 (C→G). There is evidence that the presence of the TC2 776 CC genotype may be more efficient in delivering vitamin B12 to tissues, resulting in enhanced B12 functional status (Miller et al., 2002; Namour et al., 1998). In other studies, the presence of the TC2 776 GG genotype was shown to affect negatively the serum concentration of the TC2 protein-vitamin B12 complex (von Castel-Dunwoody et al., 2005) and was associated with low concentrations of SAM in childbearing-age women (Barbosa et al., 2008). Considering that SAM is the major methyl donor for DNA methylation reactions, it was hypothesized that the variant TC2 776 C→G could influence the maternal risk for DS by modifying the DNA methylation pattern. This polymorphism has only been investigated in DS risk by two groups to date (J.M. Biselli et al., 2008c; Fintelman-Rodrigues et al., 2009), but no association has been found.

The conflicting results shown by literature have raised the suggestion that the presence of individual polymorphisms in genes involved in folate metabolism might not increase the risk of having a child with DS, although the effect of combined risk genotypes might modify their individual effect and increase DS risk (J.M., Biselli et al., 2008a; Brandalize et al., 2010; Coppedè et al., 2006; Coppedè et al., 2009 ; da Silva et al., 2005; Martínez-Frías, et al., 2006; Scala et al., 2006; Wang et al., 2008). Moreover, there is evidence that the significance of genetic polymorphisms seems to depend on interactions with nutritional factors (Papoutsakis et al., 2010; Stover & Caudill, 2008).

Advertisement

6. Folate metabolism, genomic stability, and genetic polymorphisms

Both in vitro and in vivo studies have shown that DNA methylation is an important mechanism for the maintenance of genomic stability. Literature provides several examples that genome-wide DNA hypomethylation enhances the occurrence of aneuploidy and chromosomal rearrangements (Herrera et al., 2008), loss of heterozygosity (Matsuzaki et al., 2005), and chromosome malsegregation (Fenech et al., 2011). Folate and vitamin B12 are among the most important minerals and vitamins required for DNA maintenance and prevention of DNA damage that could be induced by inadequate intake of these antimutagenic vitamins (Fenech, 2002). In human cells, folate deficiency is associated with DNA hypomethylation (Chang et al., 2011; Linhart et al., 2009), DNA instability (strand breakage, uracil misincorporation) (Linhart et al., 2009; Williams & Jacobson, 2010), aneuploidy of chromosomes 17 and 21 (Beetstra et al., 2005; Wang et al., 2004), apoptosis (Li et al., 2003), and necrosis (Beetstra et al., 2005). Low vitamin B12 status is also associated with DNA hypomethylation (Brunaud et al., 2003) and genetic instability (Andreassi et al., 2003; Botto et al., 2003).

There is increasing evidence of association between polymorphisms in folate and Hcy metabolizing genes and levels of chromosome damage. The MTHFR 677 C→T polymorphism is associated with diminished levels of 5-methylcytosine and DNA hypomethylation (Chen et al., 2010; Friso et al., 2002; Paz et al., 2002), micronucleus formation (Andreassi et al., 2003; Botto et al., 2003), and microsatellite instability (Naghibalhossaini et al., 2010) in the presence of the variant T allele. The homozygous variant genotype of another polymorphism of the MTHFR gene, 1298 A→C, was more frequent in patients with Turner syndrome (de Oliveira et al., 2008), and a higher frequency of the C allele was observed in spontaneous abortions with fetal chromosomal aneuploidy as compared to those with normal fetal karyotypes (Kim et al., 2011), suggesting its involvement in the origin of chromosomal imbalances. The MTR 2756 A→G polymorphism was associated with reduced number of hypermethylated CpG islands of suppressor tumor genes and with higher micronucleus rates in the presence of the MTRR 66 GG variant genotype (Botto et al., 2003; Paz et al., 2002; Zijno et al., 2003).

The polymorphism RFC1 80 A→G has been associated with reduced percentage of 5-methylcytosine in the DNA of mothers of children with autism in the presence of homozygous and heterozygous genotypes for the G allele as compared to AA genotype (James et al., 2010); however, the presence of the A allele was recently associated with increased oxidative DNA damage, while the cSHMT 1420 C→T polymorphism was associated with reduced oxidative DNA damage (CC>CT>TT) (Mohammad et al., 2011).

Moreover, Piskac-Collier et al. (2011) recently demonstrated that lymphocytes from lung cancer patients showed a considerably increased frequency of cytogenetic damage in the presence of MTHFR 677 C→T, MTHFR 1298 A→C, and cSHMT 435 C→T allelic variants, suggesting that interactions between genetic polymorphisms may also have a significant impact on genetic instability.

Advertisement

7. Predisposition to chromosome malsegregation in young DS mothers and its association with folate-metabolizing gene polymorphisms

Studies with women who have a DS child at a young age have suggested that they present genetic predispositions to chromosome malsegregation in both somatic and germ line cells. Migliore et al. (2006) observed increased frequency of binucleated-micronucleated lymphocytes in women who had a DS child before 35 years of age, and fluorescence in situ hybridization analysis revealed that micronuclei were mainly originating from chromosomal malsegregation events, including chromosome 21 malsegregation. Further studies from their group confirmed increased chromosome damage in blood cells of young DS mothers and showed a significant correlation between micronucleated cells and both MTHFR 677C→T and 1298A→C polymorphisms. The mean frequency of binucleated-micronucleated cells increased significantly with the increasing number of MTHFR 677 T alleles, and MTHFR 1298 AA women have significantly higher binucleated-micronucleated cells frequency than do MTHFR 1298 AC + CC carriers (Coppedè et al., 2007; Coppedè, 2009 ). In addition, mothers who had a DS child at a young age showed increased frequency (of about 5-fold) of Alzheimer’s disease (AD) (Schupf, et al., 2001). A unifying hypothesis trying to relate DS, trisomy 21, and AD has proposed that trisomy 21 mosaicism at the germ cell level or in brain cells could account for the familial aggregation of AD and DS (Potter, 1991). Together, these results suggest that young DS mothers are more prone to chromosome malsegregation, which could be true both for somatic (peripheral blood lymphocytes, brain) and for germ cells and, importantly, folate-metabolizing gene polymorphisms seem to play an important role on this susceptibility to aneuploidy.

Advertisement

8. Folate supplementation and DS prevention

Two important emerging areas of nutrition science are nutrigenomics, which refers to the effect of diet on DNA stability, and nutrigenetics, which refers to the impact of genetic differences between individuals on their response to a specific dietary pattern, functional food, or supplement for a specific health outcome. On these terms, two premises are important: (a) inappropriate nutrient supply can cause considerable levels of genome mutation and alter the expression of genes required for genome maintenance, and (b) common genetic polymorphisms may alter the activity of genes that affect the bioavailability of micronutrients and/or the affinity for micronutrient cofactors in key enzymes involved in DNA metabolism or repair, resulting in a lower or higher reaction rate (Bull & Fenech, 2008; Fenech, 2005).

As mentioned before, the folate-dependent biosynthesis of nucleotide precursors for DNA synthesis and genome methylation is dependent on the availability of many vitamins, including B12, B6, niacin, riboflavin, and minerals (zinc, cobalt), and is subject to regulation by other nutrients, such as iron and vitamin A, not directly involved in DNA or SAM biosynthesis (Stover, & Caudill 2008). Therefore, impairments in one-carbon metabolism, and the SAM cycle in particular, induced by nutritional deficiencies and/or genetic polymorphisms that encode folate-dependent enzymes, alter genome methylation patterns and gene expression levels (Stover, 2004; Stover, & Caudill 2008).

Since 1992, supplementation with 0.4 mg/daily of folic acid is recommended for women of childbearing age for the prevention of neural tube defects (Centers for Disease Control, 1992). Barkai et al. (2003) observed that families at risk for neural tube defects present with a higher frequency of DS cases and vice-versa, suggesting that both disorders are influenced by the same folate-related risk factors. However, two issues ought to be considered in the prevention of DS by folic acid: the dose and the timing of folic acid intake (Scala et al., 2006). It has been proposed that genomic instability is reduced at plasma folate concentrations above 34 nmol/L and Hcy concentrations below 7.5 μmol/L; these concentrations can only be reached with the ingestion of more than 0.4 mg/day of folic acid (Fenech, 2002). A report of a decreased occurrence of DS offspring in mothers supplemented with high doses of folic acid (6 mg/day) (Czeizel & Puho, 2005) supports the hypothesis of an involvement of folate in the etiology of DS. Concerning the timing of folate intake, it should be remembered that maternal MI errors in the primary oocyte may occur in a process that begins during fetal life and ends at the time of ovulation, whereas MII errors occur at the time of fertilization (Yoon et al., 1996). Therefore, it is likely that only MII errors would be immediately affected by folic acid intake in adult women (Ray et al., 2003).

Advertisement

9. Conclusion

Currently available literature suggests that abnormal folate metabolism is associated with increased maternal risk for DS, with a complex interaction between genetic polymorphisms, environmental factors (i.e., nutritional factors), and epigenetic processes. However, given the complexity of the folate pathway, these complex interactions cannot be easily understood and none of the polymorphisms studied so far can be used in genetic counseling to predict the maternal risk for having a DS child ( Coppedè et al., 2009 ). However, nutrigenetics and nutrigenomics are promising areas for evaluating the possibility of DS prevention with folic acid supplementation associated with susceptible genotypes. Thus, further large-scale studies are necessary to better understand the complex association between chromosomal 21 nondisjunction and folate metabolism.

References

  1. 1. Abbag F. I. 2006Congenital heart diseases and other major anomalies in patients with Down syndrome. Saudi Medical Journal, 27 2February 2006), 219 222 0379-5284
  2. 2. Ahmed I. Ghafoor T. Samore N. A. Chattha M. N. 2005Down syndrome: clinical and cytogenetic analysis. Journal of the College of Physicians and Surgeons- Pakistan: JCPSP, 15 7July 2005), 426 429 0102-2386X
  3. 3. Allen E. G. Freeman S. B. Druschel C. Hobbs C. A. O’Leary L. A. Romitti P. A. Royle M. H. Torfs C. P. Sherman S. L. 2009Maternal age and risk for trisomy 21 assessed by the origin of chromosome nondisjunction: a report from the Atlanta and National Down Syndrome Projects. Human Genetics, 125 1February 2009), 41 52 0340-6717
  4. 4. Ananth C. V. Elsasser D. A. Kinzler W. L. Peltier M. R. Getahun D. Leclerc D. Rozen R. R. New Jersey. Placental Abruption. Study Investigators. 2007Polymorphisms in methionine synthase reductase and betaine-homocysteine S-methyltransferase genes: Risk of placental abruption. Molecular Genetics and Metabolism, 91 1May 2007), 104 110 1096-7192
  5. 5. Andreassi M. G. Botto N. Cocci F. Battaglia D. Antonioli E. Masetti S. Manfredi S. Colombo M. G. Biagini A. Clerico A. 2003Methylenetetrahydrofolate reductase gene C677T polymorphism, homocysteine, vitamin B12, and DNA damage in coronary artery disease. Human Genetics, 112 2February 2003), 171 177 0340-6717
  6. 6. Antonarakis S. E. Petersen M. B. Mc Innis M. G. Adelsberger P. A. Schinzel A. A. Binkert F. Pangalos C. Raoul O. Slaugenhaupt S. A. Hafez M. Cohen M. M. Roulson D. Schwartz S. Mikkelsen M. Tranebjaerg L. Greenberg F. Hoar D. I. Rudd N. L. Warren A. C. Metaxotou C. Bartsocas C. Chakravarti A. 1992The meiotic stage of nondisjunction in trisomy 21: determination by using DNA polymorphisms. American Journal of Human Genetics, 50 3March 1992), 544 550 0340-6717
  7. 7. Aymé S. Lippman-Hand A. 1982Maternal-age effect in aneuploidy: does altered embryonic selection play a role? American Journal of Human Genetics, 34 4July 1982), 558 565 0002-9297
  8. 8. Bailey L. B. Gregory J. F. 1999Folate metabolism and requirements. The Journal of Nutrition, 129 4April 1999), 779 782 0022-3166
  9. 9. Barbero J. L. 2011Chromatid Cohesion Control and Aneuploidy. Cytogenetic and Genome Research, (January 2011), Epub ahead of print, 1424-8581 1424 8581
  10. 10. Barbosa P. R. Stabler S. P. Trentin R. Carvalho F. R. Luchessi A. D. Hirata R. D. Hirata M. H. Allen R. H. Guerra-Shinohara E. M. 2008Evaluation of nutritional and genetic determinants of total homocysteine, methylmalonic acid and S-adenosylmethionine/S-adenosylhomocysteine values in Brazilian childbearing-age women. Clinica Chimica Acta; International Journal of Clinical Chemistry, 388 1-2February 2008), 139 147 0009-8981
  11. 11. Barkai G. Arbuzova S. Berkenstadt M. Heifetz S. Cuckle H. 2003Frequency of Down’s syndrome and neural-tube defects in the same family. Lancet, 361 9366April 2003), 1331 1335 0140-6736
  12. 12. Beetstra S. Thomas P. Salisbury C. Turner J. Fenech M. 2005Folic acid deficiency increases chromosomal instability, chromosome 21 aneuploidy and sensitivity to radiation-induced micronuclei. Mutation Research, 578 1-2October 2005), 317 326 0027-5107
  13. 13. Bestor T. H. 2000The DNA methyltransferases of mammals. Human Molecular Genetics, 9 16October 2000), 2395 2402 0964-6906
  14. 14. Biselli P. M. Sanches de Alvarenga. M. P. Abbud-Filho M. Ferreira-Baptista M. A. Galbiatti A. L. Goto M. T. Cardoso M. A. Eberlin M. N. Haddad R. Goloni-Bertoll,o E. M. Pavarino-Bertelli E. C. 2007Effect of folate, vitamin B6, and vitamin B12 intake and MTHFR C677T polymorphism on homocysteine concentrations of renal transplant recipients. Transplantation proceedings, 39 10December 2007), 3163 3165 0041-1345
  15. 15. Biselli J. M. Goloni-Bertollo E. M. Zampieri B. L. Haddad R. Eberlin M. N. Pavarino-Bertelli E. C. 2008aGenetic polymorphisms involved in folate metabolism and elevated concentrations of plasma homocysteine: maternal risk factors for Down syndrome in Brazil. Genetics and Molecular Research, 7 1January 2008), 33 42 1676-5680
  16. 16. Biselli J. M. Goloni-Bertollo E. M. Ruiz M. T. Pavarino-Bertelli E. C. 2008bCytogenetic profile of Down syndrome cases seen by a general genetics outpatient service in Brazil. Down’s syndrome, research and practice, 12 3February 2008), 0968-7912
  17. 17. Biselli J. M. Brumati D. Frigeri V. F. Zampieri B. L. Goloni-Bertollo E. M. Pavarino-Bertelli E. C. 2008cA80G polymorphism of reduced folate carrier 1 (RFC1) and C776G polymorphism of transcobalamin 2 (TC2) genes in Down’s syndrome etiology. Sao Paulo Medical Journal, 126 6November 2008), 329 332 1516-3180
  18. 18. Bosco P. Guéant-Rodriguez R. M. Anello G. Barone C. Namour F. Caraci F. Romano A. Romano C. Guéant J. L. 2003Methionine synthase (MTR) 2756 (A→G) polymorphism, double heterozygosity Methionine synthase 2756AG / Methionine synthase reductase (MTRR 66AG) and elevated homocysteinemia are three risk factors for having a child with Down syndrome. American Journal of Medical Genetics Part A, 121 3September 2003), 219 224 1552-4825
  19. 19. Botto N. Andreassi M. G. Manfredi S. Masetti S. Cocci F. Colombo M. G. Storti S. Rizza A. Biagini A. 2003Genetic polymorphisms in folate and homocysteine metabolism as risk factors for DNA damage. European Journal of Human Genetics, 11 9September 2003), 671 678 1018-4813
  20. 20. Bowron A. Scott J. Stansbie D. 2005The influence of genetic and environmental factors on plasma homocysteine concentrations in a population at high risk for coronary artery disease. Annals of Clinical Biochemistry, 42 Pt6November 2005), 459 462 0004-5632
  21. 21. Brandalize A. P. Bandinelli E. dos Santos. P. A. Schüler-Faccini L. 2010Maternal gene polymorphisms involved in folate metabolism as risk factors for Down syndrome offspring in Southern Brazil. Disease Markers, 29 2 95 101 0278-0240
  22. 22. Brunaud L. Alberto J. M. Ayav A. Gérard P. Namour F. Antunes L. Braun M. Bronowicki J. P. Bresler L. Guéant J. L. 2003Vitamin B12 is a strong determinant of low methionine synthase activity and DNA hypomethylation in gastrectomized rats. Digestion, 68 2-3November 2003), 133 140 0012-2823
  23. 23. Bull C. Fenech M. 2008Genome-health nutrigenomics and nutrigenetics: nutritional requirements or ‘nutriomes’ for chromosomal stability and telomere maintenance at the individual level. The Proceedings of Nutrition Society, 67 2May 2008), 146 156 0029-6651
  24. 24. Centers for Disease Control. 1992Recommendations for the use of folic acid to reduce the number of cases of spina bifida and other neural tube defects. MMWR. Recommendations and reports: Morbidity and mortality weekly report. Recommendations and reports / Centers for Disease Control, 41 RR-14September 1992), 1 7 1057-5987
  25. 25. Chang H. Zhang T. Zhang Z. Bao R. Fu C. Wang Z. Bao Y. Li Y. Wu L. Zheng X. Wu J. 2011Tissue-specific distribution of aberrant DNA methylation associated with maternal low-folate status in human neural tube defects. The Journal of Nutritional Biochemistry, (February 2011), Epub ahead of print, 0955-2863 0955 2863
  26. 26. Chango A. Fillon-Emery N. de Courcy G. P. Lambert D. Pfister M. Rosenblatt D. S. Nicolas J. P. 2000A polymorphism (80G->A) in the Reduced folate carrier gene and its associations with folate status and homocysteinemia. Molecular Genetics and Metabolism, 70 4August 2000), 310 315 1096-7192
  27. 27. Chango A. Fillon-Emery N. Mircher C. Bléhaut H. Lambert D. Herbeth B. James S. J. Réthoré M. O. Nicolas J. P. 2005No association between common polymorphisms in genes of folate/homocysteine metabolism and the risk of Down syndrome among French mothers. The British Journal of Nutrition, 95 2August 2005), 166 169 0007-1145
  28. 28. Chen X. Guo J. Lei Y. Zou J. Lu X.. Bao Y. Wu L. Wu J. Zheng X. Shen Y. Wu B. L. Zhang T. 2010Global DNA hypomethylation is associated with NTD-affected pregnancy: A case-control study. Birth Defects Research. Part A, Clinical and Molecular Teratology, 88 7July 2010), 575 581 1542-0760
  29. 29. Cheng D. M. Jiang Y. G. Huang C. Y. Kong H. Y. Pang W. Yang H. P. 2010Polymorphism of MTHFR C677T, serum vitamin levels and cognition in subjects with hyperhomocysteinemia in China. Nutritional Neuroscience, 13 4August 2010), 175 182 0102-8415X
  30. 30. Chiang T. Duncan F. E. Schindler K. Schultz R. M. Lampson M. A. 2010Evidence that weakened centromere cohesion is a leading cause of age-related aneuploidy in oocytes. Current Biology: CB, 20 17September 2010), 1522 1528 0960-9822
  31. 31. Christensen K. E. Rohlicek C. V. Andelfinger G. U. Michaud J. Bigras J. L. Richter A. Mackenzie R. E. Rozen R. 2008The MTHFD1 p.Arg653Gln variant alters enzyme function and increases risk for congenital heart defects. Human Mutation, 30 2February 2008), 212 220 1098-1004
  32. 32. Contestabile A. Benfenati F. Gasparini L. 2010Communication breaks-Down: from neurodevelopment defects to cognitive disabilities in Down syndrome. Progress in Neurobiology, 91 1May 2010), 1 22 0301-0082
  33. 33. Coppedè F. Marini G. Bargagna S. Stuppia L. Minichilli F. Fontana I. Colognato R. Astrea G. Palka G. Migliore L. 2006Folate gene polymorphisms and the risk of Down syndrome pregnancies in young Italian women. American Journal of Medical Genetics Part A, 140 10May 2006), 1083 1091 1552-4825
  34. 34. Coppedè F. Colognato R. Bonelli A. Astrea G. Bargagna S. Siciliano G. Migliore L. 2007Polymorphisms in folate and homocysteine metabolizing genes and chromosome damage in mothers of Down syndrome children. American Journal of Medical Genetics. Part A, 143A 17September 2007), 2006 2015 1552-4825
  35. 35. Coppedè F. Migheli F. Bargagna S. Siciliano G. Antonucci I. Stuppia L. Palka G. Migliore L. 2009Association of maternal polymorphisms in folate metabolizing genes with chromosome damageand risk of Down syndrome offspring. Neuroscience Letters, 449 1January 2009), 15 19 0304-3940
  36. 36. Coppedè F. 2009The complex relationship between folate/homocysteine metabolism and risk of Down syndrome. Mutation Research, 682 1July-August 2009), 54 70 0027-5107
  37. 37. Czeizel A. E. Puhó E. 2005Maternal use of nutritional supplements during the first month of pregnancy and decreased risk of Down’s syndrome: case-control study. Nutrition, 21 6June 2005), 698 704 0899-9007
  38. 38. da Silva. L. R. Vergani N. Galdieri Lde. C. Ribeiro Porto. M. P. Longhitanom S. B. Brunoni D. D’Almeida V. Alvarez Perez. A. B. 2005Relationship between polymorphisms in genes involved in homocysteine metabolism and maternal risk for Down syndrome in Brazil. American Journal of Medical Genetics Part A, 135 3June 2005), 263 267 1552-4825
  39. 39. De Oliveira K. C. Bianco B. B. Verreschi I. T. Guedes A. D. Galera B. B. Galera M. F. Barbosa C. P. Lipay M. V. 2008Prevalence of the polymorphism MTHFR A1298C and not MTHFR C677T is related to chromosomal aneuploidy in Brazilian Turner Syndrome patients. Arquivos Brasileiros de Endocrinologia e Metabologia, 52 8November 2008), 1374 1381 0004-2730
  40. 40. De Angelis J. T. Farrington W. J. Tollefsbol T. O. 2008An overview of epigenetic assays. Molecular Biotechnology, 38 2February 2008), 179 183 1073-6085
  41. 41. De Vos L. Chanson A. Liu Z. Ciappio E. D. Parnell L. D. Mason J. B. Tucker K. L. Crott J. W. 2008Associations between single nucleotide polymorphisms in folate uptake and metabolizing genes with blood folate, homocysteine, and DNA uracil concentrations. American Journal of Clinical Nutrition, 88 4October 2008), 1149 1158 0002-9165
  42. 42. Feix A. Fritsche-Polanz R. Kletzmayr J. Vychytil A. Horl W. H. Sunder-Plassmann G. Födinger M. 2001Increased prevalence of combined MTR and MTHFR genotypes among individuals with severely elevated total homocysteine plasma levels. American journal of kidney diseases: the official journal of the National Kidney Foundation, 38 5November 2001), 956 964 0272-6386
  43. 43. Fenech M. 2002Micronutrients and genomic stability: a new paradigm for recommended dietary allowances (RDAs). Food and chemical toxicology: an international journal published for the British Industrial Biological Research Association, 40 8August 2002), 1113 1117 0278-6915
  44. 44. Fenech M. Baghurst P. Luderer W. Turner J. Record S. Ceppi M. Bonassi S. 2005Low intake of calcium, folate, nicotinic acid, vitamin E, retinol, b-carotene and high intake of pantothenic acid, biotin and riboflavin are significantly associated with increased genome instability- results from a dietary intake and micronucleus index survey in South Australia. Carcinogenesis, 26 5May 2005), 991 999 0143-3334
  45. 45. Fenech M. Kirsch-Volders M. Natarajan A. T. Surralles J. Crott J. W. Parry J. Norppa H. Eastmond D. A. Tucker J. D. Thomas P. 2011Molecular mechanisms of micronucleus, nucleoplasmic bridge and nuclear bud formation in mammalian and human cells. Mutagenesis, 26 1January 2011), 125 132 0267-8357
  46. 46. Fillon-Emery N. Chango A. Mircher C. Barbé F. Bléhaut H. Herbeth B. Rosenblatt D. S. Réthoré M. O. Lambert D. Nicolas J. P. 2004Homocysteine concentrations in adults with trisomy 21: effect of B vitamins and genetic polymorphisms. The American Journal of Clinical Nutrition,80 6December 2004), 1551 1557 0002-9165
  47. 47. Finkelstein J. D. Martin J. J. 2000Homocysteine. The International Journal of Biochemistry & Cell Biology, 32 4April 2000), 385 389 1357-2725
  48. 48. Fintelman-Rodrigues N. Corrêa J. C. Santos J. M. Pimentel M. M. Santos-Rebouças C. B. 2009Investigation of CBS, MTR, RFC-1 and TC polymorphisms as maternal risk factors for Down syndrome. Disease Markers,26 4 155 161 0278-0240
  49. 49. Fredriksen A. Meyer K. Ueland P. M. Vollset S. E. Grotmol T. Schneede J. 2007Large-scale population-based metabolic phenotyping of thirteen genetic polymorphisms related to one-carbon metabolism. Human Mutation,28 9September 2007), 856 865 1098-1004
  50. 50. Friso S. Choi S. W. Girelli D. Mason J. B. Dolnikowski G. G. Bagley P. J. Olivieri O. Jacques P. F. Rosenberg I. H. Corrocher R. Selhub J. 2002A common mutation in the 5,10-methylenetetrahydrofolate reductase gene affects genomic DNA methylation through an interaction with folate status. Proceedings of the National Academy of Sciences of the United States of America, 99 8April 2002), 5606 5611 0027-8424
  51. 51. Frosst P. Blom H. J. Milos R. Goyette P. Sheppard C. A. Matthews R. G. Boers G. J. H. den Heijer. M. Kluijtmans L. A. J. van den Heuve. L. P. Rozen R. 1995A candidate genetic risk factor for vascular disease: a common mutation in Methylenetetrahydrofolate reductase. Nature Genetics,1May 1995), 111 113 1061-4036
  52. 52. Gellekink H. Blom H. J. van der Linden I. J. den Heijer. M. 2007Molecular genetic analysis of the human dihydrofolate reductase gene: relation with plasma total homocysteine, serum and red blood cell folate levels. European Journal of Human Genetics: EJHG,15 1January 2007), 103 109 1018-4813
  53. 53. Ghosh S. Feingold E. Dey S. K. 2009Etiology of Down syndrome: Evidence for consistent association among altered meiotic recombination, nondisjunction, and maternal age across populations. American Journal of Medical Genetics Part A, 149A 7July 2009), 1415 1420 1552-4825
  54. 54. Guenther B. D. F. Sheppard C. A. Tran P. Rozen R. Matthews R. G. Ludwig M. L. 1999The structure and properties of methylenetetrahydrofolate reductase from Escherichia coli suggest how folate ameliorates human hyperhomocysteinemia. Nature Structural and Molecular Biology, 6 4April 1999), 359 365 1072-8368
  55. 55. Hall J. Solehdin F. 1998Folic acid for the prevention of congenital anomalies. European Journal of Pediatrics,157 6June 1998), 445 450 0340-6199
  56. 56. Harmon D. L. Shields D. C. Woodside J. V. Mc Master D. Yarnell J. W. G. Young I. S. Peng K. Shane B. Evans A. E. Whitehead A. S. 1999Methionine synthase D919G polymorphism is a significant but modest determinant of circulating homocysteine concentrations. Genetic Epidemiology, 17 4November 1999),298 309 0741-0395
  57. 57. Hasle H. Clemmensen I. H. Mikkelsen M. 2000Risks of leukaemia and solid tumours in individuals with Down’s syndrome. Lancet, 355 9199January 2000), 165 169 0140-6736
  58. 58. Hassold T. Sherman S. 2000Down syndrome: genetic recombination and the origin of the extra chromosome 21. Clinical Genetics, 57 2February 2000), 95 100 0009-9163
  59. 59. Hassold T. Hunt P. 2001To err (meiotically) is human: the genesis of human aneuploidy. Nature Reviews. Genetics, 2 4April 2001), 280 291 1471-0056
  60. 60. Herrera L. A. Prada D. Andonegui M. A. Dueñas-González A. 2008The epigenetic origin of aneuploidy. Current Genomics, 9 1March 2008), 43 50 1389-2029
  61. 61. Hobbs C. A. Cleves M. A. Lauer R. M. Burns T. L. James S. J. 2002Preferential transmission of the MTHFR 677 T allele to infants with Down syndrome: implications for a survival advantage. American Journal of Medical Genetics, 113 1November 2002), 9 14 1552-4825
  62. 62. Hum D. W. Bell A. W. Rozen R. Mac Kenzie. R. E. 1988Primary structure of a human trifunctional enzyme. Isolation of a cDNA encoding methylenetetrahydrofolate dehydrogenase-methenyltetrahydrofolate cyclohydrolase-formyltetrahydrofolate synthetase. The Journal of Biological Chemistry, 263 31November 1988), 15946 15950 0021-9258
  63. 63. James S. J. Melnyk S. Jernigan S. Pavliv O. Trusty T. Lehman S. Seidel L. Gaylor D. W. Cleves M. A. 2010A functional polymorphism in the reduced folate carrier gene and DNA hypomethylation in mothers of children with autism. American Journal of Medical Genetics. Part B, Neuropsychiatric genetics : the official publication of the International Society of Psychiatric Genetics, 153B 6September 2010), 1209 1220 1552-4841
  64. 64. Johnson W. G. Stenroos E. S. Spychala J. R. Chatkupt S. Ming S. X. Buyske S. 2004New 19 bp deletion polymorphism in intron-1 of dihydrofolate reductase (DHFR): a risk factor for spina bifida acting in mothers during pregnancy? American Journal of Medical Genetics Part A, 124 4February 2004), 339 345 1552-4825
  65. 65. Jones K. L. 2006Smith’s recognizable patterns of human malformation (6th edition), Elsevier Saunders, 0-72160-615-6
  66. 66. Jyothy A. Kumar K. S. Mallikarjuna G. N. Babu Rao. V. Uma Devi. B. Sujatha M. Reddy P. P. 2001Parental age and the origin of extra chromosome 21 in Down syndrome. Journal of Human Genetics, 46 6 347 350 1434-5161
  67. 67. Karpen G. H. Allshire R. C. 1997The case for epigenetic effects on centromere identity and function. Trends in Genetics: TIG, 13 12December 1997), 489 496 0168-9525
  68. 68. Kim S. Y. Park S. Y. Choi J. W. Kim D. J. Lee S. Y. Lim J. H. Han J. Y. Ryu H. M. Kim M. H. 2011Association between MTHFR 1298A>C polymorphism and spontaneous abortion with fetal chromosomal aneuploidy. American Journal of Reproductive Immunology (New York, N.Y.: 1989), (March 2011), Epub ahead of print, 0000-1046 1046 7408
  69. 69. Kovaleva N. V. Tahmasebi-Hesari M. Verlinskaia D. K. 2010Grandmaternal age in children with Down syndrome in St. Petersburg. Tsitologiia i Genetika, 44 5September-October 2010), 47 53 0564-3783
  70. 70. Kraus J. P. 1998Biochemistry and molecular genetics of cystathionine beta-synthase deficiency. European Journal of Pediatrics,157No.Suppl 2, (April 1998),S50 S53 0340-6199
  71. 71. Kumar J. Garg G. Karthikeyan G. Sengupta S. 2010Cystathionine beta-synthase 844Ins68 polymorphism is not associated with the levels of homocysteine and cysteine in an Indian population. Biomarkers: biochemical indicators of exposure, response, and susceptibility to chemicals, 15 3May 2010),283 287 0135-4750X
  72. 72. Lamb N. E. Freeman S. B. Savage-Austin A. Pettay D. Taft L. Hersey J. Gu Y. Shen J. Saker D. May K. M. Avramopoulos D. Petersen M. B. Hallberg A. Mikkelsen M. Hassold T. J. Sherman S. L. 1996Susceptible chiasmate configurations of chromosome 21 predispose to non-disjunction in both maternal meiosis I and meiosis II. Nature Genetics, 14 4December 1996), 400 405 1061-4036
  73. 73. Lamb N. E. Yu K. Shaffer J. Feingold E. Sherman S. L. 2005Association between maternal age and meiotic recombination for trisomy 21. American Journal of Human Genetics, 76 1January 2005), 91 99 0002-9297
  74. 74. Leclerc D. Odièvre M. Wu Q. Wilson A. Huizenga J. J. Rozen R. Scherer S. W. Gravel R. A. 1999Molecular cloning, expression and physical mapping of the human methionine synthase reductase gene. Gene, 240 1November 1999),75 88 0378-1119
  75. 75. Li G. M. Presnell S. R. Gu L. 2003Folate deficiency, mismatch repair-dependent apoptosis, and human disease. The Journal of Nutritional Biochemistry, 14 10October 2003), 568 575 0955-2863
  76. 76. Lin M. Y. Young C. M. 2000Folate levels in cultures of lactic acid bacteria. International dairy journal / published in association with the International Dairy Federation., 10 409 414 0958-6946
  77. 77. Linhart H. G. Troen A. Bell G. W. Cantu E. Chao W. H. Moran E. Steine E. He T. Jaenisch R. 2009Folate deficiency induces genomic uracil misincorporation and hypomethylation but does not increase DNA point mutations. Gastroenterology, 136 1January 2009), 227 235e3, 0016-5085
  78. 78. Lott I. T. Head E. 2005Alzheimer disease and Down syndrome: factors in pathogenesis. Neurobiology of Aging, 26 3 383 389 0197-4580
  79. 79. Malini S. S. Ramachandra N. B. 2006Influence of advanced age of maternal grandmothers on Down syndrome. BMC Medical Genetics, 14January 2006), 7 4 1471-2350
  80. 80. Martínez-Frías M. L. Pérez B. Desviat L. R. Castro M. Leal F. Rodríguez L. Mansilla E. Martínez-Fernández M. L. Bermejo E. Rodríguez-Pinilla E. Prieto D. Ugarte M. Working E. C. E. M. C. Group 2006Maternal polymorphisms 677C-T and 1298A-C of MTHFR, and 66A-G MTRR genes: is there any relationship between polymorphisms of the folate pathway, maternal homocysteine levels, and the risk for having a child with Down syndrome? American Journal of Medical Genetics. Part A,140 9May 2006), 987 997 1552-4825
  81. 81. Martínez-Frías M. L. 2008The biochemical structure and function of methylenetetrahydrofolate reductase provide the rationale to interpret the epidemiological results on the risk for infants with Down syndrome. American Journal of Medical Genetics Part A, 146 11June 2008), 1477 1482 1552-4825
  82. 82. Matsuzaki K. Deng G. Tanaka H. Kakar S. Miura S. Kim Y. S. 2005The relationship between global methylation level, loss of heterozygosity, and microsatellite instability in sporadic colorectal cancer. Clinical Cancer Research: an official journal of the American Association for Cancer Research, 15 11December 2005), 8564 8569 1078-0432
  83. 83. Mc Nulty H. Pentieva K. 2004Folate bioavailability. The Proceedings of the Nutrional Society, 63 4November 2004), 529 536 1475-2719
  84. 84. Mc Nulty H. Scott J. M. 2008Intake and status of folate and related B-vitamins: considerations and challenges in achieving optimal status. The British Journal of Nutrition, 99No.Suppl 3, (June 2008), S48 S54 0007-1145
  85. 85. Meguid N. A. Dardir A. A. Khass M. Hossieny L. E. Ezzat A. El Awady M. K. 2008MTHFR genetic polymorphism as a risk factor in Egyptian mothers with Down syndrome children. Disease Markers, 24 1 19 26 0278-0240
  86. 86. Mendes C. C. Biselli J. M. Zampieri B. L. Goloni-Bertollo E. M. Eberlin M. N. Haddad R. Riccio M. F. Vannucchi H. Carvalho V. M. Pavarino-Bertelli E. C. 2010base pair deletion polymorphism of the dihydrofolate reductase (DHFR) gene: maternal risk of Down syndrome and folate metabolism. Sao Paulo Medical Journal,128 4July 2010), 215 218 1516-3180
  87. 87. Migliore L. Boni G. Bernardini R. Trippi F. Colognato R. Fontana I. Coppedè F. Sbrana I. 2006Susceptibility to chromosome malsegregation in lymphocytes of women who had a Down syndrome child in young age. Neurobiology of Aging, 27 5May 2006), 710 716 0197-4580
  88. 88. Miller J. W. Ramos M. I. Garrod M. G. Flynn M. A. Green R. 2002Transcobalamin II 775G>C polymorphism and indices of vitamin B12 status in healthy older adults. Blood, 100 2 718 720 0006-4971
  89. 89. Mohammad N. S. Yedluri R. Addepalli P. Gottumukkala S. R. Digumarti R. R. Kutala V. K. 2011Aberrations in one-carbon metabolism induce oxidative DNA damage in sporadic breast cancer. Molecular and Cellular Biochemistry, 349 1-2March 2011), 159 167 0300-8177
  90. 90. Naghibalhossaini F. Mokarram P. Khalili I. Vasei M. Hosseini S. V. Ashktorab H. Rasti M. Abdollahi K. 2010MTHFR C677T and A1298C variant genotypes and the risk of microsatellite instability among Iranian colorectal cancer patients. Cancer Genetics and Cytogenetics, 197 2March 2010), 142 151 0165-4608
  91. 91. Namour F. Guy M. Aimone-Gastin I. de Nonancourt M. Mrabet N. Guéant J. L. 1998Isoelectrofocusing phenotype and relative concentration of transcobalamin II isoproteins related to the codon 259 Arg/Pro polymorphism. Biochemical and Biophysical Research Communications, 251 3October 1999), 769 774 0000-6291X
  92. 92. Narayanan S. Mc Connell J. Little J. Sharp L. Piyathilake C. J. Powers H. Basten G. Duthie S. J. 2004Associations between two common variants C677T and A1298C in the methylenetetrahydrofolate reductase gene and measures of folate metabolism and DNA stability (strand breaks, misincorporated uracil, and DNA methylation status) in human lymphocytes in vivo. Cancer epidemiology, biomarkers & prevention : a publication of the American Association for Cancer Research, cosponsored by the American Society of Preventive Oncology, 13 9September 2004), 1436 1443 0055-9965
  93. 93. Nguyen T. T. Dyer D. L. Dunning D. D. Rubin S. A. Grant K. E. Said H. M. 1997Human intestinal folate transport: cloning, expression, and distribution of complementary RNA. Gastroenterology, 112 3March 1997), 783 791 0016-5085
  94. 94. O’Leary V. B. Parle Mc Dermott. A. Molloy A. M. Kirke P. N. Johnson Z. Conley M. Scott J. M. Mills J. L. 2002MTRR and MTHFR polymorphism: link to Down syndrome? American Journal of Medical Genetics, 107 2January 2002) 151 155 0148-7299
  95. 95. Olteanu H. Banerjee R. 2001Human methionine synthase reductase, a soluble P-450 reductase-like dual flavoprotein, is sufficient for NADPH-dependent methionine synthase activation. The Journal of Biological Chemistry, 276 38September 2001), 35558 35563 0021-9258
  96. 96. Olteanu H. Munson T. Banerjee R. 2002Differences in the efficiency of reductive activation of methionine synthase and exogenous electron acceptors between the common polymorphic variants of human methionine synthase reductase. Biochemistry, 41 45November 2002), 13378 13385 0006-2960
  97. 97. Pajares M. A. Pérez-Sala D. 2006Betaine homocysteine S-methyltransferase: just a regulator of homocysteine metabolism? Cellular and molecular life sciences : CMLS, 63 23December 2006), 2792 2803 0142-0682X
  98. 98. Papoutsakis C. Manios Y. Magkos F. Papaconstantinou E. Schulpis K. H. Zampelas A. Matalas A. L. Yiannakouris N. 2010Effect of the methylenetetrahydrofolate reductase (MTHFR 677C>T) polymorphism on plasma homocysteine concentrations in healthy children is influenced by consumption of folate-fortified foods. Nutrition (Burbank, Los Angeles County, Calif.), 26 10October 2010), 969 974 0899-9007
  99. 99. Parle Mc Dermott. A. Pangilinan F. Mills J. L. Kirke P. N. Gibney E. R. Troendle J. O’Leary V. B. Molloy A. M. Conley M. Scott J. M. Brody L. C. 2007The 19-bp deletion polymorphism in intron-1 of dihydrofolate reductase (DHFR) may decrease rather than increase risk for spina bifida in the Irish population. American Journal of Medical Genetics. Part A,143 11June 2007), 1174 1180 1552-4825
  100. 100. Pavarino-Bertelli E. C. Biselli J. M. Bonfim D. Goloni-Bertollo E. M. 2009Clinical profile of children with Down syndrome treated in a genetics outpatient service in the southeast of Brazil. Revista da Associação Médica Brasileira (1992), 55 5September-October 2009), 547 552 0104-4230
  101. 101. Paz M. F. Ávila S. Fraga M. F. Pollan M. Capella G. Peinado M. A. Sanchez-Cespedes M. Herman J. G. Esteller M. 2002Germ-line variants in methyl-group metabolism genes and susceptibility to DNA methylation in normal tissues and human primary tumors. Cancer Research, 62 15August 2002), 4519 4524 0008-5472
  102. 102. Pepe G. Vanegas O. C. Rickards O. Giusti B. Comeglio P. Brunelli T. Marcucci R. Prisco D. Gensini G. F. Abbate R. 1999World distribution of the T833C/844INS68 CBS in cis double mutation: a reliable anthropological marker. Human Genetics, 104 2February 1999), 126 129 0340-6717
  103. 103. Piskac-Collier A. L. Monroy C. Lopez M. S. Cortes A. Etzel C. J. Greisinger A. J. Spitz M. R. El -Zein R. A. 2011Variants in folate pathway genes as modulators of genetic instability and lung cancer risk. Genes, Chromosomes & Cancer, 50 1January 2011), 1 12 1045-2257
  104. 104. Potter H. 1991Review and hypothesis: Alzheimer disease and Down syndrome--chromosome 21 nondisjunction may underlie both disorders. American Journal of Human Genetics, 48 6June 1991), 1192 1200 0002-9297
  105. 105. Pozzi E. Vergani P. Dalprà L. Combi R. Silvestri D. Crosti F. Dell O. M. Valsecchi M. G. 2009Maternal polymorphisms for methyltetrahydrofolate reductase and methionine synthetase reductase and risk of children with Down syndrome. American Journal of Obstetrics and Gynecology, 200 6June 2009), 636e1-6, 0002-9378
  106. 106. Quadros E. V. Regec A. L. Khan K. M. Quadros E. Rothenberg S. P. 1999Transcobalamin II synthesized in the intestinal villi facilitates transfer of cobalamin to the portal blood. The American Journal of Physiology, 277 1Pt 1, (July 1999), G161 G166 0002-9513
  107. 107. Rai A. K. Singh S. Mehta S. Kumar A. Pandey L. K. Raman R. 2006MTHFR C677T and A1298C polymorphisms are risk factors for Down’s syndrome in Indian mothers. Journal of Human Genetics,51 4February 2006), 278 283 0004-5161
  108. 108. Ram G. Chinen J. 2011Infections and immunodeficiency in Down syndrome. Clinical and Experimental Immunology, 164 1April 2011), 9 16 0009-9104
  109. 109. Ray J. G. Meier C. Vermeulen M. J. Cole D. E. C. Wyatt P. R. 2003Prevalence of trissomy 21 following folic acid food fortification. American Journal of Medical Genetics. Part A, 120 3July 2003), 309 313 1552-4825
  110. 110. Sadiq M. F. Al-Refai E. A. Al-Nasser A. Khassawneh M. Al-Batayneh Q. 2011Methylenetetrahydrofolate Reductase Polymorphisms C677T and A1298C as Maternal Risk Factors for Down Syndrome in Jordan. Genetic Testing and Molecular Biomarkers,15 1-2January-February 2011), 51 57 1945-0265
  111. 111. Sakuno T. Watanabe Y. 2009Studies of meiosis disclose distinct roles of cohesion in the core centromere and pericentromeric regions. Chromosome Research: an international journal on the molecular, supramolecular and evolutionary aspects of chromosome biology, 17 2 239 249 0967-3849
  112. 112. Scala I. Granese B. Sellitto M. Salomè S. Sammartino A. Pepe A. Mastroiacovo P. Sebastio G. Andria G. 2006Analysis of seven maternal polymorphisms of genes involved in homocysteine/folate metabolism and risk of Down syndrome offspring. Genetics in medicine : official journal of the American College of Medical Genetics, 8 7July 2006), 409 416 1098-3600
  113. 113. Schupf N. Kapell D. Nightingale B. Lee J. H. Mohlenhoff J. Bewley S. Ottman R. Mayeux R. 2001Specificity of the fivefold increase in AD in mothers of adults with Down syndrome. Neurology, 57 6September 2001), 979 984 0028-3878
  114. 114. Seetharam B. Li N. 2000Transcobalamin II and its cell surface receptor. Vitamins and Hormones, 59 337 366 0083-6729
  115. 115. Sherman S. L. Petersen M. B. Freeman S. B. Hersey J. Pettay D. Taft L. Frantzen M. Mikkelsen M. Hassold T. J. 1994Non-disjunction of chromosome 21 in maternal meiosis I: evidence for a maternal age-dependent mechanism involving reduced recombination. Human Molecular Genetics, 3 9September 1994), 1529 1535 0964-6906
  116. 116. Stanisiawska-Sachadyn A. Brown K. S. Mitchell L. E. Woodside J. V. Young I. S. Scott J. M. Murray L. Boreham C. A. Mc Nulty H. Strain J. J. Whitehead A. S. 2008An insertion/deletion polymorphism of the Dihydrofolate reductase (DHFR) gene is associated with serum and red blood cell folate concentrations in women. Human Genetics, 123 3April 2008), 289 295 0340-6717
  117. 117. Steck S. E. Keku T. Butler L. M. Galanko J. Massa B. Millikan R. C. Sandler R. S. 2008Polymorphisms in Methionine synthase, Methionine synthase reductase and Serine hydroxymethyltransferase, folate and alcohol intake, and colon cancer risk. Journal of Nutrigenetics and Nutrigenomics, 2008,1 4June 2008), 196 204 1661-6499
  118. 118. Stein Z. Stein W. Susser M. 1986Attrition of trisomies as a maternal screening device. An explanation of the association of trisomy 21 with maternal age. Lancet, 1 8487April 1986), 944 947 0140-6736
  119. 119. Stover P. J. Caudill M. A. 2008Genetic and epigenetic contributions to human nutrition and health: managing genome-diet interactions. Journal of The American Dietetic Association, 108 9September 2008), 1480 1487 0002-8223
  120. 120. Stover P. J. 2004Physiology of folate and vitamin B12 in health and disease. Nutrition Reviews,62 6PtJune 2004), S3 S12discussion S13, 0029-6643
  121. 121. Summers C. M. Hammons A. L. Mitchell L. E. Woodside J. V. Yarnell J. W. G. Young I. S. Evans A. Whitehead A. S. 2008Influence of the cystathionine b-synthase 844ins68 and methylenetetrahydrofolate reductase 677C4T polymorphisms on folate and homocysteine concentrations. European Journal of Human Genetics: EJHG,16 8August 2008), 1010 1013 1018-4813
  122. 122. Tsai M. Y. Bignell M. Schwichtenberg K. Hanson N. Q. 1996High prevalence of a mutation in the cystathionine-b-synthase gene. American Journal of Human Genetics, 59 5October 1996), 1262 1267 0002-9297
  123. 123. Tsai M. Y. Welge B. C. Hanson N. Q. Bignell M. K. Vessey J. Schwichtenberg K. Yang F. Bullemer F. E. Rasmussen R. Graham K. J. 1999Genetic causes of mild hyperhomocysteinemia in patients with premature occlusive coronary artery diseases. Atherosclerosis, 143 1March 1999), 63 170 0021-9150
  124. 124. Tsai M. Y. Bignell M. Yang F. Welge B. G. Graham K. J. Hanson N. Q. 2000Polygenic influence on plasma homocysteine: association of two revalent mutations, the 844ins68 of cystathionine b-synthase and A2756G of methionine synthase, with lowered plasma homocysteine levels. Atherosclerosis, 149 1March 2000), 131 137 0021-9150
  125. 125. Ulvik A. Ueland P. M. Fredriksen A. Meyer K. Vollset S. E. Hoff G. Schneede J. 2007Functional inference of the Methylenetetrahydrofolate reductase 677 C > T and 1298A > C polymorphisms from a large-scale epidemiological study. Human Genetics, 121 1March 2007), 57 64 0340-6717
  126. 126. van der Put N. M. Gabreëls F. Stevens E. M. Smeitink J. A. Trijbels F. J. Eskes T. K. van den Heuvel. L. P. Blom H. J. 1998A second common mutation in the methylenetetrahydrofolate reductase gene: an additional risk factor for neural-tube defects? American Journal of Human Genetics, 62 5May 1998), 1044 1051 0002-9297
  127. 127. Venail F. Gardiner Q. Mondain M. 2004ENT and speech disorders in children with Down’s syndrome: an overview of pathophysiology, clinical features, treatments, and current management. Clinical Pediatrics (Phila), 43 9November-December 2004), 783 791 0009-9228
  128. 128. Vogt E. Kirsch-Volders M. Parry J. Eichenlaub-Ritter U. 2008Spindle formation, chromosome segregation and the spindle checkpoint in mammalian oocytes and susceptibility to meiotic error. Mutation Research, 651 1-2March 2008), 14 29 0027-5107
  129. 129. von-Dunwoody Castel. K. M. Kauwell G. P. Shelnutt K. P. Vaughn J. D. Griffin E. R. Maneval D. R. Theriaque D. W. Bailey L. B. 2005Transcobalamin 776C->G polymorphism negatively affects vitamin B-12 metabolism. American Journal of Clinical Nutrition,81 6June 2005), 1436 1441 0002-9165
  130. 130. Vyletal P. Sokolová J. Cooper D. N. Kraus J. P. Krawczak M. Pepe G. Rickards O. Koch H. G. Linnebank M. Kluijtmans L. A. Blom H. J. Boers G. H. Gaustadnes M. Skovby F. Wilcken B. Wilcken D. E. Andria G. Sebastio G. Naughten E. R. Yap S. Ohura T. Pronicka E. Laszlo A. Kozich V. 2007Diversity of cystathionine beta-synthase haplotypes bearing the most common homocystinuria mutation c.833T>C: a possible role for gene conversion. Human Mutation., 28 3March 2007), 255 264 1059-7794
  131. 131. Wang X. Thomas P. Xue J. Fenech M. 2004Folate deficiency induces aneuploidy in human lymphocytes in vitro-evidence using cytokinesis-blocked cells and probes specific for chromosomes 17 and 21. Mutation Research, 551 1-2July 2004), 167 180 0027-5107
  132. 132. Wang S. S. Qiao F. Y. Feng L. Lv J. J. 2008Polymorphisms in genes involved in folate metabolism as maternal risk factors for Down syndrome in China. Journal of Zhejiang University. Science. B, 9 2February 2008), 93 99 1673-1581
  133. 133. Weisberg I. S. Jacques P. F. Selhub J. Bostom A. G. Chen Z. Ellison C. Eckfeldt J. H. Rozen R. 2001The 1298A→C polymorphism in methylenetetrahydrofolate reductase (MTHFR): in vitro expression and association with Homocysteine. Atherosclerosis, 156 2June 2001), 409 415 0021-9150
  134. 134. Whetstine J. R. Gifford A. J. Witt T. Liu X. Y. Flatley R. M. Norris M. Haber M. Taub J. W. Ravindranath Y. Matherly L. H. 2001Single nucleotide polymorphisms in the human reduced folate carrier: characterization of a high-frequency G/A variant at position 80 and transport properties of the His(27) and Arg(27) carriers. Clinical cancer research : an official journal of the American Association for Cancer Research, 7 11November 2001), 3416 3422 1078-0432
  135. 135. Williams J. D. Jacobson M. K. 2010Photobiological implications of folate depletion and repletion in cultured human keratinocytes. Journal of photochemistry and photobiology. B, Biology, 99 1April 2010), 49 61 1011-1344
  136. 136. Yamada K. Chen Z. Rozen R. Matthews R. G. 2001Effects of common polymorphisms on the properties of recombinant human methylenetetrahydrofolate reductase. Proceedings of the National Academy of Sciences of the United States of America, 98 26December 2001), 14853 14858 0027-8424
  137. 137. Yamada K. Gravel R. A. Toraya T. Matthews R. G. 2006Human methionine synthase reductase is a molecular chaperone for human methionine synthase. Proceedings of the National Academy of Science of the United States of America, 103 25June 2006), 9476 9481 0027-8424
  138. 138. Yang Q. H. Botto L. D. Gallagher M. Friedman J. M. Sanders C. L. Koontz D. Nikolova S. Erickson J. D. Steinberg K. 2008Prevalence and effects of gene-gene and gene-nutrient interactions on serum folate and serum total homocysteine concentrations in the United States: findings from the third National Health and Nutrition Examination Survey DNA Bank. American Journal of Clinical Nutrition, 88 1July 2008), 232 246 0002-9165
  139. 139. Yoon P. W. Freeman S. B. Sherman S. L. Taft L. F. Gu Y. Pettay D. Flanders W. D. Khoury M. J. Hassold T. J. 1996Advanced maternal age and the risk of Down syndrome characterized by the meiotic stage of chromosomal error: A population-based study. American Journal of Human Genetics, 58 3March 19960, 628 633 0002-9297
  140. 140. Zheng C. J. Byers B. 1993Oocyte selection: a new model for the maternal-age dependence of Down syndrome. Human Genetics, 90 1-2September-October 1993), 1 6 0340-6717
  141. 141. Zijno A. Andreoli C. Leopardi P. Marcon F. Rossi S. Caiola S. Verdina A. Galati R. Cafolla A. Crebelli R. 2003Folate status, metabolic genotype, and biomarkers of genotoxicity in healthy subjects. Carcinogenesis, 24 6June 2003), 1097 1103 0143-3334

Written By

Érika Cristina Pavarino, Bruna Lancia Zampieri, Joice Matos Biselli and Eny Maria Goloni Bertollo

Submitted: 03 November 2010 Published: 29 August 2011