Hostname: page-component-8448b6f56d-mp689 Total loading time: 0 Render date: 2024-04-16T19:14:45.116Z Has data issue: false hasContentIssue false

Intestinal metabolism of sulfur amino acids

Published online by Cambridge University Press:  19 October 2009

Caroline Bauchart-Thevret
Affiliation:
Department of Pediatrics, Baylor College of Medicine, USDA/ARS Children's Nutrition Research Center, Houston, Texas77030, USA
Barbara Stoll
Affiliation:
Department of Pediatrics, Baylor College of Medicine, USDA/ARS Children's Nutrition Research Center, Houston, Texas77030, USA
Douglas G. Burrin*
Affiliation:
Department of Pediatrics, Baylor College of Medicine, USDA/ARS Children's Nutrition Research Center, Houston, Texas77030, USA
*
*Corresponding author: Dr Douglas G. Burrin, fax +1 713 798 7057, email dburrin@bcm.tmc.edu
Rights & Permissions [Opens in a new window]

Abstract

The gastrointestinal tract (GIT) is a metabolically significant site of sulfur amino acid (SAA) metabolism in the body and metabolises about 20 % of the dietary methionine intake which is mainly transmethylated to homocysteine and trans-sulfurated to cysteine. The GIT accounts for about 25 % of the whole-body transmethylation and trans-sulfuration. In addition, in vivo studies in young pigs indicate that the GIT is a site of net homocysteine release and thus may contribute to the homocysteinaemia. The gut also utilises 25 % of the dietary cysteine intake and the cysteine uptake by the gut represents about 65 % of the splanchnic first-pass uptake. Moreover, we recently showed that SAA deficiency significantly suppresses intestinal mucosal growth and reduces intestinal epithelial cell proliferation, and increases intestinal oxidant stress in piglets. These recent findings indicate that intestinal metabolism of dietary methionine and cysteine is nutritionally important for intestinal mucosal growth. Besides their role in protein synthesis, methionine and cysteine are precursors of important molecules. S-adenosylmethionine, a metabolite of methionine, is the principal biological methyl donor in mammalian cells and a precursor for polyamine synthesis. Cysteine is the rate-limiting amino acid for glutathione synthesis, the major cellular antioxidant in mammals. Further studies are warranted to establish how SAA metabolism regulates gut growth and intestinal function, and contributes to the development of gastrointestinal diseases. The present review discusses the evidence of SAA metabolism in the GIT and its functional and nutritional importance in gut function and diseases.

Type
Review Article
Copyright
Copyright © The Authors 2009

Introduction

Increasing evidence indicates that sulfur amino acids, methionine and cysteine, are implicated in numerous biological functions and diseases, aside from their role in protein synthesis(Reference Stipanuk1Reference Brosnan and Brosnan3). Methionine is an indispensable amino acid and is transmethylated intracellularly to homocysteine via S-adenosylmethionine (SAM), the principal biological methyl donor in mammalian cells and a precursor for polyamine synthesis(Reference Mato, Corrales and Lu4) (Fig. 1). Reduced SAM concentrations, as a consequence of low methionine intake or folate deficiency, mainly lead to a deregulation in DNA methylation and are implicated in various cancers, including colorectal cancer(Reference Kim5, Reference Giovannucci, Rimm and Ascherio6). Homocysteine is a sulfur-containing amino acid present in the blood and tissues but not incorporated into protein(Reference MacCoss, Fukagawa and Matthews7). Numerous clinical studies have shown that elevated plasma homocysteine levels are strongly associated with increased risk for several diseases, including CVD, stroke, Alzheimer's disease and osteoporosis(Reference Zhou and Austin8Reference Herrmann, Peter and Umanskaya10). Homocysteine can be converted into cysteine via cystathionine through the trans-sulfuration pathway, an irreversible process(Reference Stipanuk1) (Fig. 1). Homocysteine can also be methylated back to methionine via the remethylation pathway (Fig. 1). The combination of transmethylation and remethylation pathways comprises the methionine cycle which occurs in most cells. However, the trans-sulfuration pathway has a limited tissue distribution and is restricted to the liver, kidney, intestine, pancreas and adrenals(Reference Brosnan and Brosnan3, Reference Zlotkin and Anderson11). Cysteine is considered a semi-indispensable amino acid whose availability is dependent upon methionine intake(Reference Stipanuk1). However, dietary cysteine can satisfy a proportion of the sulfur amino acid requirement, the so-called cysteine-sparing effect on dietary methionine requirement(Reference Ball, Courtney-Martin and Pencharz12). Cysteine is a constituent amino acid of the tripeptide glutathione (γ-Glu-Cys-Gly), the major cellular antioxidant in mammals, and serves also as a precursor for the synthesis of taurine, pyruvate, sulfate and hydrogen sulfide (H2S) (Fig. 1).

Fig. 1 Sulfur amino acid metabolism. AMT, aminotransferase; BHMT, betaine-homocysteine methyltransferase; CBS, cystathionine β-synthase; CDO, cysteine dioxygenase; CGL, cystathionine γ-lyase; CSD, cysteine sulfinate decarboxylase; DHFR, dihydrofolate reductase; DMG, dimethyl-glycine; dTMP, thymidylate; dUMP, deoxyuridylate; MAT, methionine adenosyltransferase; MS, methionine synthase; MT, methyl transferases; MTHFR, methylenetetrahydrofolate reductase, SAH, S-adenosylhomocysteine; SAHH, S-adenosylhomocysteine hydrolase; SAM, S-adenosylmethionine; SHMT, serine hydroxymethyltransferase; TS, thymidylate synthetase.

Although the liver is usually considered as one of the major organs in the body that metabolises sulfur amino acids, the gastrointestinal tract (GIT), and especially the intestine, appears to be also a significant site of dietary sulfur amino acid metabolism. Several reports indicate that intestinal sulfur amino acid metabolism is nutritionally important for normal gut function and intestinal mucosal growth. Dysregulation in the intestinal sulfur amino acid metabolism leads to numerous complications and diseases. In the present review, we will discuss the evidence of sulfur amino acid metabolism in the GIT and its functional and nutritional importance in health and disease.

Isotopic approaches to study sulfur amino acid metabolism

The introduction of stable-isotope tracer methodology has enabled the measurement of sulfur amino acid metabolism in vivo in human subjects, including adults and infants. Storch et al. (Reference Storch, Wagner and Burke13, Reference Storch, Wagner and Burke14) were the first to explore sulfur amino acid metabolism in human subjects using a stable-isotope approach. Their model was based on the intravenous infusion of a doubly-labelled methionine isotopomer [1-13C; methyl-2H3]methionine to determine the whole-body methionine carboxyl and methyl fluxes which provided estimates of transmethylation, trans-sulfuration and remethylation rates (Fig. 2). These studies demonstrated the role of feeding, fasting and the sparing effect of cysteine on kinetics of methionine catabolism in man and they defined the relative recycling of methionine through the remethylation of homocysteine(Reference Storch, Wagner and Burke14). In their methionine kinetic calculations, Storch et al. used an estimate of the intracellular methionine isotopic enrichment based on the intracellular dilution of leucine estimated with plasma α-ketoisocaproate enrichments observed by Matthews et al. (Reference Matthews, Schwarz and Yang15), yet this assumption does not reflect the true physiological effect of methionine metabolism. More recently, this model has been modified by MacCoss et al. (Reference MacCoss, Fukagawa and Matthews7) and others using plasma homocysteine enrichments to correct the intracellular methionine enrichments, since plasma homocysteine can be derived only from intracellular homocysteine formed by methionine transmethylation. Consistent with other reports(Reference Tessari, Kiwanuka and Coracina16, Reference Tessari, Coracina and Kiwanuka17), the rates of methionine kinetics were substantially higher after homocysteine correction than previous estimates.

Fig. 2 Stable isotopic tracer model used by Storch et al. (Reference Storch, Wagner and Burke14). This model is based on the intravenous infusion of a methionine isotopomer [1-13C; methyl-2H3]methionine [M+4]. The [2H3]methyl group is lost during transmethylation from S-adenosylmethionine (SAM) to S-adenosylhomocysteine (SAH), whereas the [13C]carbon group is transferred to [1-13C]homocysteine [M+1], [1-13C]cystathionine [M+1] and finally, oxidised to 13CO2 via α-ketobutyrate in the tricarboxylic acid cycle.

Few other studies have used different isotopomers to study whole-body methionine metabolism. A complication of the measurement of homocysteine remethylation rate is that the reaction can be catalysed by two enzymic pathways. Homocysteine can be methylated back to methionine by the ubiquitously distributed methionine synthase using 5-methyltetrahydrofolate as a methyl donor, and also in a folate-independent pathway by betaine-homocysteine methyltransferase (Fig. 1), which is expressed only in the liver and the kidney(Reference Brosnan and Brosnan3). To distinguish between the methyl groups transferred by the two homocysteine methyltransferases, Davis et al. (Reference Davis, Stacpoole and Williamson18) used a combined isotopic approach of [U-13C]methionine and [3-13C]serine. With this tracer combination, the remethylated methionine tracer, i.e. [13C4]methionine, was measured separately from the [13C1]methionine generated by the folate-dependent homocysteine remethylation pathway from [3-13C]serine. Using this tracer model, they determined that serine is the predominant one-carbon donor for homocysteine remethylation in healthy females with adequate folate, vitamin B12 and vitamin B6 nutritional status. The steady-state model of Storch et al. was also discussed by Hoffer(Reference Hoffer19) who concluded that this model can not provide information about the type and the severity of metabolic blocks that cause, for instance, hyperhomocysteinaemia. Based on this appreciation, Shinohara et al. (Reference Shinohara, Hasegawa and Ogawa20) examined the remethylation of homocysteine after a single administration of [2H7]methionine in rats under folate and/or choline (a precursor of betaine) deficiency. With a 2H label in the methyl position, this model has the advantage to recognise directly the elimination of methionine and to determine the extent of homocysteine remethylation by measuring [2H7]methionine, [2H4]homocysteine, [2H4]methionine and unlabelled methionine by GC-MS. Their results showed that plasma levels of remethylated methionine were influenced by choline deficiency rather than folate deficiency.

Important contributions to our knowledge of the dietary requirement for sulfur amino acids and the effect of cysteine on methionine requirement in man have resulted from recent studies using tracer isotopic approaches(Reference Ball, Courtney-Martin and Pencharz12, Reference Di Buono, Wykes and Cole21Reference Fukagawa23). The ability of cysteine to provide a proportion of the total sulfur amino acid requirements in man, hence providing a sparing effect on methionine requirement, has been extensively debated(Reference Ball, Courtney-Martin and Pencharz12, Reference Di Buono, Wykes and Cole21, Reference Fukagawa23). Many studies performed in animals have shown that more than 40 % of the sulfur amino acid requirement can be met by dietary cysteine(Reference Ball, Courtney-Martin and Pencharz12, Reference Baker, Clausing and Harmon24Reference Baker28). However, the cysteine-sparing effect on methionine requirement in adult human subjects was not found in a set of studies using methionine and cysteine tracers(Reference Storch, Wagner and Burke13, Reference Hiramatsu, Fukagawa and Marchini29Reference Raguso, Regan and Young32). In these human studies, the methionine and cysteine intake was based on the 1985 FAO, WHO and United Nations University recommendation of 13 mg/kg per d, which has been recently found to be deficient for humans(Reference Di Buono, Wykes and Ball33, Reference Kurpad, Regan and Varalakshmi34). The failure to detect a cysteine-sparing effect in those studies(Reference Storch, Wagner and Burke13, Reference Hiramatsu, Fukagawa and Marchini29Reference Raguso, Regan and Young32) has been discussed by Ball et al. (Reference Ball, Courtney-Martin and Pencharz12) and appears to be related to the methionine intake below the minimum obligatory requirement, which has not been defined yet. More recent stable-isotope studies in human subjects, using indicator amino acid oxidation(Reference Di Buono, Wykes and Ball35) and 24 h indicator amino acid oxidation and indicator amino acid balance(Reference Kurpad, Regan and Varalakshmi22) techniques to determine the sulfur amino acid requirement, have found a sparing effect of methionine requirement by cysteine with graded levels of methionine in the absence and presence of excess cysteine in the diet. These studies(Reference Kurpad, Regan and Varalakshmi22, Reference Di Buono, Wykes and Ball35) combined with earlier N balance studies(Reference Albanese, Holt and Davis36Reference Reynolds, Steel and Jones38) provide evidence of a cysteine-sparing effect in man that ranges from 17 to 90 %. The sparing effect of cysteine on dietary methionine requirement appears to occur by reducing methionine catabolism through trans-sulfuration to cystathionine, which is related to a decrease in the concentration of cystathionase β-synthase (CBS), the enzyme that synthesises cystathionine from homocysteine (Fig. 1), and SAM (Fig. 1), a positive effector of the CBS enzyme(Reference Finkelstein, Martin and Harris39). Most recently, it has been shown that when total sulfur amino acid needs are provided as methionine at 14 mg/kg per d in the presence of an adequate protein intake of 1 g/kg per d(Reference Humayun, Elango and Ball40), extra cysteine does not increase the synthesis of erythrocyte glutathione, the major cellular antioxidant, in healthy adult humans(Reference Courtney-Martin, Rafii and Wykes41). Moreover, the cysteine-sparing effect on methionine requirement has been also estimated in parenterally and enterally fed piglets and the safe methionine intake with an excess of cysteine was determined to be 0·27 and 0·36 g/kg per d in parenteral and enteral feeding, respectively(Reference Shoveller, Brunton and House27). The significant difference in methionine requirement between parenteral and enteral feeding indicates that the gut utilises about 28 % of the total sulfur amino acid requirement(Reference Shoveller, Brunton and House27), suggesting an active metabolic role of the gut in the whole-body sulfur amino acid metabolism, although the cysteine-sparing effect is not dependent on first-pass metabolism in the gut.

Evidence of intestinal sulfur amino acid metabolism

In 1965, Mudd et al. (Reference Mudd, Finkelstein and Irreverre42) were the first to show significant levels of three enzymes involved in the trans-sulfuration pathway, i.e. methionine adenosyltransferase (MAT), CBS and cystathionine γ-lyase, in rat small intestine mucosa, but at lower activities for one or more enzymes compared with the liver, pancreas and kidney. In 1972, Stegink & Den Besten(Reference Stegink and Den Besten43) then showed that splanchnic organs are an important site of transmethylation and trans-sulfuration of dietary methionine for the synthesis of cysteine in human adults. They found a twofold higher plasma cystine concentration in adults given a continuous cyst(e)ine-free nutrient solution for 2 weeks via the nasogastric route compared with the intravenous route. Since then, others studies have shown that total parenteral nutrition without cysteine results in low circulating levels of cystine in newborn infants(Reference Zlotkin, Bryan and Anderson44) and neonatal piglets(Reference Shoveller, Brunton and Pencharz45). In agreement with the study from Stegink & Den Besten(Reference Stegink and Den Besten43), plasma cystine concentrations were higher in enterally than parenterally fed piglets administered methionine as the sole sulfur amino acid source(Reference Shoveller, Brunton and Pencharz45). These data suggest that the first-pass splanchnic metabolism is important for the synthesis of cysteine in neonates as well as adults.

Significant advancements in our understanding of splanchnic amino acid metabolism were derived from in vivo measurements of splanchnic organ balance(Reference Stoll and Burrin46). Pig models have been extensively utilised to study human nutrition since they have similar physiology to man, especially gut physiology. Our studies in piglets showed that the net portal absorption of several indispensable amino acids, including methionine, was significantly less than 100 % of the dietary intake, ranging from 40 to 70 %(Reference Stoll, Henry and Reeds47). The importance of the gut was also demonstrated in studies where the whole-body methionine requirement was 30 % greater in enterally fed than parenterally fed piglets(Reference Shoveller, Brunton and Pencharz45, Reference Shoveller, House and Brunton48). Our recent study in infant piglets using the Storch et al. isotopic model indicates that the GIT metabolises 20 % of dietary methionine intake which is mainly transmethylated to homocysteine and trans-sulfurated to cysteine(Reference Riedijk, Stoll and Chacko49). The GIT accounts for about 25 % of whole-body transmethylation and trans-sulfuration(Reference Riedijk, Stoll and Chacko49). More direct evidence of intestinal sulfur amino acid metabolism was found in studies in rat by Finkelstein(Reference Finkelstein50, Reference Finkelstein51), demonstrating that the small intestine possesses the enzymes necessary to metabolise methionine to cysteine, albeit at significantly lower activities than the liver. In our most recent piglet study, we also found significant enzymic activities of MAT, CBS and methionine synthase in the small intestine (jejunum and ileum)(Reference Bauchart-Thevret, Stoll and Chacko52), enzymes that are involved in the methionine transmethylation, trans-sulfuration and remethylation pathways, respectively. All these studies provide compelling evidence of intestinal sulfur amino acid metabolism. However, recent studies in pigs found negligible catabolism of methionine in enterocytes(Reference Chen, Yin and Jobgen53, Reference Chen, Li and Wang54), suggesting that the metabolism of methionine in the intestine may result from the action of non-epithelial cells or luminal microbes in the intestinal mucosa.

Homocysteine, an intermediate metabolite formed during the metabolism of methionine to cysteine (Fig. 1), plays an important physiological role, since a dysregulation in homocysteine metabolism, leading to an elevated homocysteine plasma concentration, has been associated with an increased risk of CVD, stroke, Alzheimer's disease and osteoporosis(Reference Zhou and Austin8Reference Herrmann, Peter and Umanskaya10). In our study in infant piglets, we demonstrated that the GIT is a site of net homocysteine release into the body(Reference Riedijk, Stoll and Chacko49). A recent study performed on female rats investigated tissue homocysteine metabolism using stable-isotope approaches and tried to assess the contribution of individual tissues to plasma homocysteine by measuring the isotopic enrichment(Reference Wilson, van den Borne and Calder55). Interestingly, in contrast with previous studies(Reference Stead, Brosnan and Brosnan56, Reference Mudd, Brosnan and Brosnan57), they did not find a dominant role for the liver as a source of plasma homocysteine. Other tissues with higher intracellular homocysteine enrichments than in the plasma, including the small intestine, may also contribute to the homocysteinaemia(Reference Wilson, van den Borne and Calder55).

As regards cysteine in the GIT, studies in pigs indicate that less than 100 % of dietary cysteine appears in the portal blood, suggesting intestinal utilisation of cysteine(Reference Stoll, Henry and Reeds47, Reference Bos, Stoll and Fouillet58). The first step in cysteine catabolism is its conversion to cysteine sulfinate via the enzyme cysteine dioxygenase (CDO) (Fig. 1). The cysteine sulfinate is then either decarboxylated via cysteine sulfinate decarboxylase to produce hypotaurine, which is further oxidised to taurine via a poorly understood mechanism, or transaminated to the putitative intermediate β-sulfinylpyruvate that spontaneously decomposes to pyruvate and sulfate(Reference Stipanuk, Ueki and Dominy59) (Fig. 1). Rodent studies with 14C-labelled cysteine showed significantly higher oxidation when given via the intragastric (70 %) than the intraperitoneal (41 %) route, suggesting that nearly half of the whole-body cysteine oxidation occurs in splanchnic tissues(Reference Stipanuk and Rotter60). More importantly, subsequent work demonstrated that rat enterocytes extensively metabolise cysteine via CDO to cysteine sulfinate(Reference Coloso and Stipanuk61). This is in accordance with the expression of the murine CDO gene in the small intestine(Reference Hirschberger, Daval and Stover62). In contrast to rodent studies, our recent studies in infant pigs showed that the fractional whole-body oxidation of enteral cysteine (22 %) is lower than parenteral cysteine (35 %)(Reference Cottrell, Stoll and Burrin63). We also found that splanchnic tissues (liver and gut) utilise about 40 % of dietary cysteine intake in first-pass metabolism(Reference Cottrell, Stoll and Burrin63). Intestinal absorption was the major metabolic fate of dietary cysteine, representing 75 % intake, indicating that the gut utilises 25 % of the dietary cysteine intake(Reference Cottrell, Stoll and Burrin63). We also determined that cysteine gut uptake represents about 65 % of the splanchnic first-pass uptake(Reference Cottrell, Stoll and Burrin63). In addition, our results in pigs suggest that gut tissues consume a substantial proportion of dietary splanchnic cysteine metabolism via non-oxidative pathways(Reference Cottrell, Stoll and Burrin63). We postulate that glutathione synthesis is a major non-oxidative metabolic fate for cysteine in the gut. Consistent with this, in vivo rodent studies with intravenous infusion of Reference Matthews, Schwarz and Yang[15N]cysteine indicate that an important metabolic fate of cysteine in the gut is incorporation into glutathione(Reference Malmezat, Breuille and Capitan64).

Taurine is also another endproduct of cellular cysteine catabolism(Reference Stipanuk1) (Fig. 1) and we recently found that taurine is produced by the small intestine in piglets(Reference Bauchart-Thevret, Stoll and Chacko52). Along with glutathione and cysteine, taurine plays an important role in the antioxidant function in the body(Reference Atmaca65), especially in the intestine which is constantly exposed to endogenous and exogenous diet-derived oxidants. Moreover, among the numerous other physiological functions of taurine (i.e. osmoregulation, retinal and cardiac function, diabetes modulator, hepatoprotection, neuroprotection(Reference Huxtable66, Reference Bouckenooghe, Remacle and Reusens67)), taurine is also involved in bile acid conjugation(Reference Bouckenooghe, Remacle and Reusens67). Bile acids are essential for intestinal digestion and absorption of lipids. In a recent study in adult patients with short-bowel syndrome(Reference Schneider, Joly and Gehrardt68), long-term parenteral nutrition was associated with an impaired tauro-conjugation of bile acids (enterohepatic pool) and a low biliary taurine concentration despite the long-term taurine intravenous supplementation, suggesting a partition between the systemic and the enterohepatic taurine pools. This finding suggests that the intestine plays an important role in providing taurine to the liver. Besides the fact that the majority of the intestinal taurine pool is derived from the diet and absorbed by the gut, some taurine might also come from endogenous synthesis in the gut from cysteine since CDO expression has been found in rodent small intestine(Reference Stipanuk, Ueki and Dominy59, Reference Hirschberger, Daval and Stover62). Moreover, we also found a decrease in taurine concentration in the ileum of piglets enterally fed a sulfur amino acid-free diet compared with a complete diet(Reference Bauchart-Thevret, Stoll and Chacko52), suggesting an intestinal synthesis of taurine from methionine and cysteine.

Catabolism of cysteine can also occur by desulfuration reactions that cleave the sulfur from cysteine to produce H2S (Fig. 1) via multiple pathways, mainly through CBS and cystathionine γ-lyase activity(Reference Stipanuk1, Reference Li, Bazer and Gao69). H2S has been found to be produced by the ileum(Reference Zhao, Ndisang and Wang70) and the colon(Reference Linden, Sha and Mazzone71) in rodents. H2S is a gas that has been reported to exert anti-inflammatory(Reference Wallace72), relaxant(Reference Teague, Asiedu and Moore73, Reference Gallego, Clave and Donovan74), prosecretory(Reference Schicho, Krueger and Zeller75) and antinociceptive(Reference Distrutti, Sediari and Mencarelli76) effects in the GIT. Factors that influence sulfur amino acid metabolism can also affect the production of H2S. Sulfur amino acid metabolism is regulated through the balance of production and disposal of cellular homocysteine and cysteine. Therefore, the release of homocysteine and H2S by the gut depends on the methionine utilisation by the gut, which then may have an impact on peripheral physiology. For example, a recent study in rats demonstrated that liver endothelial dysfunction caused by hyperhomocysteinaemia was reversed by exogenous H2S(Reference Distrutti, Mencarelli and Santucci77).

Studies on ruminants also showed kinetic evidence of methionine metabolism, i.e. transmethylation, remethylation and trans-sulfuration in the GIT(Reference Lobley, Connell and Revell78, Reference Lobley, Shen and Le79). Recent studies in lambs using stable-isotopic approaches demonstrated that the GIT is capable of methionine synthesis from the dietary methionine hydroxy analogue, namely 2-hydroxy-(4-methylthio)butanoic acid (HMB)(Reference Lobley, Wester and Calder80, Reference Lobley, Wester and Holtrop81). The methionine hydroxy analogue HMB is a synthetic source of methionine that is commonly added to commercial animal diets to ensure that the nutritional requirement for this essential amino acid is satisfied. The biochemical conversion of HMB to methionine involves two enzymic steps (Fig. 3). The first reaction is a stereospecific oxidation involving two enzymes: a peroxisomal l-2-hydroxy acid oxidase (l-HAOX) and a mitochondrial d-2-hydroxy acid dehydrogenase (d-HADH), which catalyses the oxidation of l-HMB and d-HMB, respectively, into 2-keto-(4-methylthio)butanoic acid(Reference Dibner and Knight82). The second reaction is the transamination of 2-keto-(4-methylthio)butanoic acid to l-methionine. This reaction is ubiquitous and does not constitute a limiting step in the metabolic conversion of dl-HMB to l-methionine(Reference Rangel-Lugo and Austic83). However, the specific l-HAOX activity has been found in chicken liver and kidney(Reference Dibner and Knight82, Reference Ferjancic-Biagini, Dupuis and De Caro84), rat liver(Reference Langer85), pig kidney(Reference Robinson, Keay and Molinari86) and also in liver, kidney and ruminal epithelia in sheep(Reference McCollum, Vazquez-Anon and Dibner87), whereas the d-HADH has been detected in numerous tissues in chickens, including liver, kidney, skeletal muscle, intestine, pancreas, spleen and brain(Reference Dibner and Knight82), and in liver, kidney and ruminal epithelia in sheep(Reference McCollum, Vazquez-Anon and Dibner87). Studies in lambs showed that more than 85 % of the [1-13C]HMB infused directly into the abomasum was absorbed into the portal vein and substantial enrichment of [1-13C]methionine occurred in the small- and large-intestinal tissues(Reference Lobley, Wester and Calder80, Reference Lobley, Wester and Holtrop81). Moreover, recent studies in chickens have also demonstrated the capacity of the small intestine for significant HMB conversion to methionine(Reference Martin-Venegas, Soriano-Garcia and Vinardell88, Reference Martin-Venegas, Geraert and Ferrer89). It is now well established that HMB is a safe and efficacious precursor of methionine in chicks(Reference Dilger, Kobler and Weckbecker90, Reference Stipanuk91). Although it still remains unknown whether the human GIT is capable of methionine synthesis from HMB, dl-HMB is currently used as an enteral product (Ketosteril®, Fresenius Kabi), a mixture of ketoacids and amino acids, in conjunction with a low-protein diet, for patients suffering from renal insufficiency(Reference Stipanuk91, Reference Feiten, Draibe and Watanabe92).

Fig. 3 Metabolic pathway for the conversion of 2-hydroxy-(4-methylthio)butanoic acid (HMB) to l-methionine. KMB, 2-keto-(4-methylthio)butanoic acid.

Sulfur amino acids and intestinal mucosal growth and function

Methionine, cysteine and glutathione

The intestinal epithelium is one of the most dynamic sites of cell turnover in mammals(Reference Johnson93). Its growth status can be modulated by various stimuli, accompanied by corresponding changes in subepithelial mucosa. Total parenteral nutrition is commonly associated with a decrease in crypt cell proliferation and epithelial renewal rate(Reference Burrin, Stoll and Jiang94). Our results in piglets indicate that the minimal enteral nutrient intake necessary to increase mucosal mass was 40 % of total nutrient intake, whereas 60 % of enteral nutrition was necessary to sustain normal mucosal proliferation and growth(Reference Burrin, Stoll and Jiang95). We recently demonstrated that a sulfur amino acid-free diet administered enterally to piglets for 7 d led to a reduced intestinal mucosal growth associated with villus atrophy, reduced epithelial cell proliferation, lower goblet cell number and diminished small-intestinal redox capacity(Reference Bauchart-Thevret, Stoll and Chacko52). Moreover, using [1-13C-methyl-2H3]methionine and [15N]cysteine tracer approaches, we also found an alteration in methionine metabolism under sulfur amino acid-deficient conditions at both whole-body and intestinal levels, where the methionine pool was preserved for protein synthesis by up-regulation of homocysteine remethylation and suppression of trans-sulfuration(Reference Bauchart-Thevret, Stoll and Chacko52). The suppression of trans-sulfuration contributed to the diminished cellular cysteine and glutathione concentrations and increased oxidant stress, which seemed to preferentially affect intestinal growth, especially in the jejunum(Reference Bauchart-Thevret, Stoll and Chacko52).

Evidence suggests that intracellular redox status plays an important role in the control of intestinal epithelial cell proliferation and apoptosis where increased oxidant stress and redox imbalance suppress cell proliferation and induce apoptosis(Reference Pias and Aw96Reference Aw99). Glutathione (γ-Glu-Cys-Gly) is the major low-molecular-weight thiol in the cells that controls cellular thiol/disulfide redox state(Reference Jones100) and serves as a major reservoir for cysteine(Reference Stipanuk1). Reduced glutathione (GSH) is a ubiquitous tripeptide present in high concentration in tissue, especially in the intestine(Reference Aw101). Cellular GSH homeostasis is maintained through de novo synthesis from precursor sulfur amino acids methionine and cysteine, regeneration from its oxidised form glutathione disulfide (GSSG), and uptake of extracellular intact GSH via Na+-dependent transport systems(Reference Aw101). Maintaining normal GSH concentration is essential to most tissues, especially the intestine, which is constantly challenged by luminal toxins and oxidants derived from the diet as well as endogenous generated reactive oxygen species. Indeed, a marked depletion of GSH, induced by buthionine sulfoximine, was shown to result in severe degeneration of jejunal and colonic epithelial cells in mice and this intestinal damage seemed to be prevented by concomitant GSH administration(Reference Martensson, Jain and Meister102). However, the notion of an essential need of GSH for cell growth appeared to be controversial(Reference Tian, Washizawa and Gu103). Changes in the extracellular cysteine/cystine redox status per se have been shown to mediate proliferative signalling that is independent of the intracellular GSH/GSSG in colon cancer cells(Reference Noda, Iwakiri and Fujimoto104, Reference Anderson, Iyer and Ziegler105). Thus, all these findings strongly suggest that sulfur amino acids methionine and cysteine play a key role in intestinal mucosal growth associated with a regulated redox status, and intestinal epithelial cell function.

S-adenosylmethionine in methylation reactions and polyamine synthesis

SAM has emerged as a key regulator of metabolism, proliferation, differentiation, apoptosis and cell death(Reference Martinez-Lopez, Varela-Rey and Ariz106). SAM is synthesised from methionine and ATP only by MAT (Fig. 4), which exists through three distinct isoforms (i.e. MAT I, MAT II and MAT III)(Reference Mato, Corrales and Lu4) that are the products of two different genes (i.e. MAT1A and MAT2A)(Reference Kotb, Mudd and Mato107). The gene MAT1A encodes the α1 catalytic subunit, which organises into dimers (MAT III) or tetramers (MAT I). The gene MAT2A encodes for the α2 catalytic subunit in the MAT II isoform. A third gene, MAT2β, encodes for a β regulatory subunit that regulates the activity of MAT II by lowering the inhibition constant (K i) for SAM and the Michaelis constant (K m) for methionine(Reference Halim, LeGros and Geller108). MAT1A is expressed mostly in adult liver, whereas MAT2A is widely distributed. MAT2A is also expressed by fetal liver but is replaced by MAT1A during development(Reference Lu and Mato109). Although MAT isoenzymes catalyse the same reaction, they are differentially regulated by their product, SAM. SAM maintains MAT1A expression in hepatocytes, but inhibits MAT2A expression(Reference Mato, Corrales and Lu4). In mammalian liver, MAT1A is a marker for the differentiated or mature liver phenotype, while MAT2A is a marker for rapid growth and dedifferentiation(Reference Lu and Mato109). The properties of mammalian MAT isoforms are summarised in Table 1. Few studies have examined the expression of MAT in intestinal tissues, yet the expression of the MAT2A gene and MAT II protein has been confirmed in human and mouse colonic tissue(Reference Ito, Ikeda and Kojima110, Reference Chen, Xia and Lin111). We recently confirmed MAT activity in piglet small intestine(Reference Bauchart-Thevret, Stoll and Chacko52). MAT2A expression has been shown to be up-regulated by growth factors, such as epidermal growth factor, insulin-like growth factor-1 and leptin, in colorectal cancer cells HT-29 and RKO and its inhibition using RNAi led to cell death(Reference Chen, Xia and Lin111). In addition, Chen et al. (Reference Chen, Xia and Lin111) also demonstrated that SAM and its metabolite methylthioadenosine (MTA) (Fig. 4) are able to down-regulate MAT2A gene expression at high doses. These findings clearly indicate that MAT2A gene expression is critical for cell survival where SAM appears as a key role in the regulation of intestinal cell proliferation.

Fig. 4 S-adenosylmethionine (SAM) metabolic pathways: transmethylation and polyamine synthesis. dcSAM, decarboxylated S-adenosylmethionine; MAT, methionine adenosyltransferase; MTA, methylthioadenosine; SAH, S-adenosylhomocysteine; SAMDC, S-adenosylmethionine decarboxylase; ODC, ornithine decarboxylase; PAO, polyamine oxidase; SSAT, spermidine/spermine N1-acetyltransferase; , methionine salvage pathway.

Table 1 Properties of mammalian methionine adenosyltransferase (MAT) isoforms (from Lu & Mato(Reference Lu and Mato145))

SAM, S-adenosylmethionine; HCC, hepatocellular carcinoma.

* K m values differ depending on the purity of the enzyme.

Besides its role as a constitutive precursor for protein synthesis, the role of methionine as a precursor for SAM synthesis may have greater regulatory importance for cell function and survival given its singular role in methylation reactions and control of gene expression. Once SAM is produced, it can be processed via two pathways, namely transmethylation or polyamine synthesis (Fig. 4). SAM is the major biological methyl group donor for a variety of methyltransferases, resulting in the methylation of substrates such as nucleic acids (DNA, RNA), proteins and lipids(Reference Chiang, Gordon and Tal112), as well as the synthesis of small molecules (for example, creatine, phosphatidylcholine, adrenaline), modification of xenobiotics (for example, thiols, arsenite) and inactivation of neurotransmitters (for example, adrenaline, noradrenaline, dopamine)(Reference Brosnan, da Silva and Brosnan113). DNA methylation is a major epigenetic modification of the genome that regulates crucial aspects of its function during development and in adults, including imprinting and X-chromosome inactivation(Reference Reik and Dean114). In mammalian cells, DNA methylation occurs mainly on cytosine residues at the C5 position within CpG dinucleotides to form 5-methylcytosine and this reaction is carried out by two important classes of DNA methyltransferases (DNMT). DNMT1 is essential for maintaining DNA-methylation patterns in proliferating cells, whereas DNMT3a and DNMT3b, two members of the second class of methyltransferases, are required for de novo methylation during embryonic development(Reference Patra, Patra and Rizzi115). Few studies have examined the epigenetic mechanisms underlying normal gut development. Most of our knowledge on epigenetic involvement, especially methylation reactions, in intestinal metabolism and physiology comes from epigenetic dysregulation in gastrointestinal carcinogenesis, such as in colorectal cancer. The role of sulfur amino acid metabolism in colorectal cancer will be discussed further in the present review.

In addition to its role in methylation, SAM is also involved in the synthesis of polyamines (Fig. 4) which are known to play a key role in cell proliferation by regulating the expression of various growth-related genes(Reference Moinard, Cynober and de Bandt116, Reference Bettuzzi, Davalli and Astancolle117). The intestinal epithelium is one of the most rapidly proliferating tissues in the body and, therefore, has a high demand for polyamines(Reference McCormack and Johnson118). Several studies have demonstrated that polyamines stimulate normal intestinal epithelial cell proliferation and are required for the maintenance of mucosal integrity(Reference McCormack and Johnson118, Reference Wang119). Very few studies have investigated the role of SAM in intestinal cell growth related to polyamine synthesis. SAM has been widely studied in hepatocytes since the liver is the major site of SAM synthesis and degradation(Reference Mato, Corrales and Lu4). A physiological increase in SAM has been reported to stimulate growth in liver cancer cells whereas a pharmalogical dose of SAM exerted a growth-inhibitory effect(Reference Ansorena, Garcia-Trevijano and Martinez-Chantar120). The effect of SAM at pharmalogical doses may be mediated by its metabolite MTA. SAM is very unstable and converts spontaneously to MTA in the polyamine synthesis (Fig. 4). In contrast to SAM, MTA has been shown to inhibit transmethylation and polyamine synthesis(Reference Clarke, Clarke and Tamanoi121). The removal of MTA by MTA phosphorylase is necessary both for the resynthesis of methionine (Fig. 5) and for polyamine synthesis, since MTA is a strong inhibitor of both spermine and spermidine synthases(Reference Grillo and Colombatto122). In colon cancer cells, 6 h of MTA treatment (1 mm) was shown to reduce polyamine level(Reference Chen, Xia and Lin111). In addition, it was recently reported that SAM and MTA are able to induce apoptosis in colon cancer cell lines but not in normal colon epithelial cells(Reference Li, Zhang and Oh123). These opposing effects of SAM and MTA on apoptosis in normal v. cancer cells have been also previously reported in hepatocytes(Reference Ansorena, Garcia-Trevijano and Martinez-Chantar120). However, the molecular mechanism of SAM and MTA on cell growth between colonocytes and hepatocytes may be different due to the presence of the MAT1A gene in hepatocytes that is absent in intestinal cells and which is differently regulated by SAM (Table 1)(Reference Mato, Corrales and Lu4). Interestingly, the proapototic effect of MTA was not blocked by increased doses of polyamines in colon cancer cells RKO(Reference Li, Zhang and Oh123). Polyamine depletion has been shown to both enhance and protect against apoptosis in a cell line- and stimulus-dependent manner(Reference Seiler and Raul124). Thus, all these data suggest that the concentration of SAM plays an important role in intestinal cell growth and apoptosis. A decrease in methionine intake or a deficiency in folate may disrupt methionine metabolism and have an impact on intestinal SAM pool and hence polyamine synthesis, which clearly emphasises the importance of dietary sulfur amino acid intake to maintain a normal intestinal growth.

Fig. 5 Methionine salvage pathway from 5′-methylthioadenosine (MTA). MTOB, 4-methylthio-2-oxobutanoic acid; MTR-1P, 5′-methylthioribose-1-P.

Sulfur amino acid metabolism and gastrointestinal diseases

Considerable epidemiological evidence suggests that a low-folate diet is associated with an increased risk of colorectal cancer(Reference Sanderson, Stone and Kim125). Like folate, choline and its oxidation product betaine are methyl-group donors that are involved in the remethylation of homocysteine to methionine (Fig. 1). Because homocysteine methylation by betaine via betaine-homocysteine methyltransferase activity is confined to the liver and the kidney(Reference Brosnan and Brosnan3), choline and betaine dietary intake may have less impact on the methylation of homocysteine in the colon than folate since methionine synthase that methylates homocysteine to methionine using 5-methyltetrahydrofolate as a methyl donor is ubiquitously distributed(Reference Brosnan and Brosnan3). One of the consequences of folate deficiency is SAM deficiency since folate is involved in the remethylation of homocysteine to methionine, the precursor of SAM (Fig. 1). However, folate supplementation can either prevent or exacerbate intestinal tumorigenesis, depending on the timing and dose of folate intervention(Reference Kim126). This may be explained by the function of folate in nucleotide synthesis (Fig. 1) where rapidly proliferating tissues, including tumours, have an increased requirement for nucleotides. At present, based on lack of compelling supportive evidence on the potential tumour-promoting effect and presence of some adverse effects(Reference Ulrich127, Reference Smith, Kim and Refsum128), a recent review concludes that folate supplementation should not be recommended as a chemopreventive measure against colorectal cancer(Reference Kim129). Interestingly, a low methionine intake, which decreases SAM levels and increases SAH levels, stimulates methylenetretrahydrofolate reductase (MTHFR) to irreversibly convert 5,10-methylenetetrahydrofolate to 5-methyltetrahydrofolate (Fig. 1)(Reference Giovannucci130). DNA methylation is an important epigenetic determinant in gene expression, in the maintenance of DNA integrity and stability, in chromatin modifications and in the development of mutations(Reference Jones and Baylin131). Aberrant DNA methylation, i.e. global genomic hypomethylation and hypermethylation of the promoter region of certain tumour-suppressor genes, has been implicated in colorectal carcinogenesis(Reference Duthie, Narayanan and Sharp132). MTHFR is an important regulatory enzyme in the metabolism of folate and the genetic polymorphism of MTHFR can affect DNA methylation and influence the risk of colorectal cancer, particularly in the presence of a low-folate diet(Reference Kim133). One of the most common functional polymorphisms of the MTHFR gene is the C677T polymorphism which results in an alanine-to-valine substitution at codon 222(Reference Frosst, Blom and Milos134). Individuals with the 677TT genotype (variant homozygotes) have no more than 30 % of normal enzyme activity, while heterozygotes (CT) have 65 % of normal enzyme activity(Reference Frosst, Blom and Milos134). A recent meta-analysis demonstrated that individuals with a homozygous variant of the MTHFR gene (677TT genotype) have a reduced risk of colorectal cancer(Reference Hubner and Houlston135). However, the beneficial effect of the 677TT genotype was reported to be abolished when folate or methionine intake is low and/or with a high alcohol intake(Reference Chen, Giovannucci and Kelsey136, Reference Ma, Stampfer and Giovannucci137).

Although there is an abundant literature on the role of folate in cancer risk, few studies have examined the role of SAM in the pathogenesis of colon cancer. One study examined MAT protein expression and activity in human colorectal cancer and found that MAT II activity is well correlated with the stage of colorectal tumour compared with normal tissue(Reference Ito, Ikeda and Kojima110). More recently, Chen et al. (Reference Chen, Xia and Lin111) found higher SAM levels in human colon cancer specimens and polyps from Min (multiple intestinal neoplasia) mice, associated with an increase in MAT2A gene expression and MAT II protein level. In addition, they demonstrated that SAM and its metabolite MTA, at pharmacological doses, are able to down-regulate MAT2A expression and to block the ability of growth factors (leptin, insulin-like growth factor-1 and epidermal growth factor) to induce MAT2A expression and exert their mitogenic response in colorectal cancer cells HT-29 and RKO(Reference Chen, Xia and Lin111). They also observed similar results in vivo in wild-type mice where SAM and MTA supplementation in drinking water for 6 d decreased MAT2A expression in proximal small intestine by 20 %(Reference Chen, Xia and Lin111). The increased levels of intracellular SAM in cancerous colonocytes are likely to support polyamine synthesis required for growth and differentiation. However, high-dose SAM treatment was shown to inhibit growth and the effect is mediated by MTA(Reference Chen, Xia and Lin111). Indeed, in contrast to SAM, MTA inhibits methylation and polyamine synthesis(Reference Avila, Garcia-Trevijano and Lu138). Therefore, like for folate, the dose of SAM supplementation may be a critical determinant in colorectal cancer susceptibility. In addition, due to their proapoptotic effect in colon cancer cells(Reference Li, Zhang and Oh123), SAM and MTA may be attractive agents in the chemoprevention and treatment of colon cancer.

Homocysteine has been associated with the pathophysiology of inflammatory bowel disease (IBD). The two major forms of IBD, Crohn's disease and ulcerative colitis, represent chronic inflammatory disorders characterised by progressive destructive inflammation in the GIT. Hyperhomocysteinaemia has been consistently found in patients with IBD, as well as an increased level of homocysteine in colonic mucosa(Reference Morgenstern, Raijmakers and Peters139, Reference Danese, Sgambato and Papa140). Elevated plasma and intestinal mucosal homocysteine concentration can result from either genetic defects affecting either the trans-sulfuration or remethylation pathways of homocysteine such as CBS deficiency or a polymorphism in MTHFR, or nutritional deficiencies of the vitamin cofactors (folate, vitamin B12 and vitamin B6) required for homocysteine metabolism (Fig. 1)(Reference Morgenstern, Raijmakers and Peters139, Reference Danese, Sgambato and Papa140). However, no direct link has been established between homocysteine and IBD. Several studies suggest that homocysteine may promote inflammatory processes through oxidative stress(Reference Papatheodorou and Weiss141) and contribute to the inflammatory state of the mucosal endothelium through endoplasmic reticulum stress(Reference Zhang, Cai and Adachi142). Although the aetiology of Crohn's disease and ulcerative colitis remains unknown, gut tissue injury is the result of an abnormal immune response and involves multiple non-immune cellular systems, including intestinal microvascular endothelial cells(Reference Papa, Scaldaferri and Danese143, Reference Deban, Correale and Vetrano144). Whether increased circulating or local homocysteine is involved in these phenomena warrants further study.

Conclusion

An increasing body of evidence supports the metabolic significance of the GIT in sulfur amino acid metabolism. Recent in vivo studies show that the intestinal metabolism of dietary sulfur amino acids methionine and cysteine is nutritionally important for normal intestinal mucosal growth. However, it still remains unknown how sulfur amino acid metabolism regulates gut growth beyond its role as a precursor of protein biosynthesis. Methionine and cysteine also are the precursors of important molecules that may contribute to intestinal mucosal growth and gut function. SAM appears to be a key molecule that links three important metabolic pathways, i.e. transmethylation, trans-sulfuration and polyamine synthesis with MTA production. Epigenetic regulation of gene expression via modulation of DNA methylation may be the critical role of SAM. Further studies are warranted to determine the role of SAM and its metabolite MTA in methylation reactions in intestinal epithelial cells and how the regulation of their synthesis can affect intestinal cell survival and proliferation. Also, glutathione plays an important role in intestinal gut function due to, in part, its role in redox status. Thus, the availability of cysteine, the rate-limiting amino acid in glutathione synthesis, appears to be important for the maintenance of epithelial cell glutathione concentration and hence in the regulation of intestinal cell redox status. However, further investigation needs to be done to determine whether the increase in dietary cysteine intake may be an effective strategy to treat diseases related to oxidant injury in the GIT. Another potentially important metabolite of methionine cycle activity is homocysteine. As regards to IBD, further studies are warranted to establish whether the association between homocysteine and IBD is a causal link or a consequence of a poor nutritional status, mainly vitamin deficiencies (folate, vitamin B12 and B6). In the case of SAM and MTA, these metabolites may have important roles in both colonic epithelial cell survival and inflammation that could also be relevant to IBD and cancer.

Acknowledgements

The present review is a publication of the US Department of Agriculture, Agricultural Research Service (USDA/ARS) Children's Nutrition Research Center, Department of Pediatrics, Baylor College of Medicine and Texas Children's Hospital, Houston, TX. It was supported by federal funds from the USDA/ARS under cooperative agreement number 58-6250-6-001. The contents of this publication do not necessarily reflect the views or policies of the USDA, nor does mention of trade names, commercial products, or organisations imply endorsement by the US Government. C. B.-T. was supported by grants from Association pour le Développement de la Recherche en Nutrition (ADREN)/Société Francophone Nutrition Clinique et Métabolisme (SFNEP), Nutricia Research Foundation, and Société Française de Nutrition (SFN).

We acknowledge the contribution of all authors in the writing of the present review.

The authors declare that they have no conflicts of interest.

References

1Stipanuk, MH (2004) Sulfur amino acid metabolism: pathways for production and removal of homocysteine and cysteine. Annu Rev Nutr 24, 539577.CrossRefGoogle ScholarPubMed
2Shoveller, AK, Stoll, B, Ball, RO, et al. . (2005) Nutritional and functional importance of intestinal sulfur amino acid metabolism. J Nutr 135, 16091612.CrossRefGoogle ScholarPubMed
3Brosnan, JT & Brosnan, ME (2006) The sulfur-containing amino acids: an overview. J Nutr 136, 1636S1640S.CrossRefGoogle ScholarPubMed
4Mato, JM, Corrales, FJ, Lu, SC, et al. . (2002) S-adenosylmethionine: a control switch that regulates liver function. FASEB J 16, 1526.CrossRefGoogle ScholarPubMed
5Kim, YI (2005) Nutritional epigenetics: impact of folate deficiency on DNA methylation and colon cancer susceptibility. J Nutr 135, 27032709.CrossRefGoogle ScholarPubMed
6Giovannucci, E, Rimm, EB, Ascherio, A, et al. . (1995) Alcohol, low-methionine–low-folate diets, and risk of colon cancer in men. J Natl Cancer Inst 87, 265273.CrossRefGoogle ScholarPubMed
7MacCoss, MJ, Fukagawa, NK & Matthews, DE (2001) Measurement of intracellular sulfur amino acid metabolism in humans. Am J Physiol Endocrinol Metab 280, E947E955.CrossRefGoogle ScholarPubMed
8Zhou, J & Austin, RC (2009) Contributions of hyperhomocysteinemia to atherosclerosis: causal relationship and potential mechanisms. Biofactors 35, 120129.CrossRefGoogle ScholarPubMed
9Schulz, RJ (2007) Homocysteine as a biomarker for cognitive dysfunction in the elderly. Curr Opin Clin Nutr Metab Care 10, 718723.CrossRefGoogle ScholarPubMed
10Herrmann, M, Peter, SJ, Umanskaya, N, et al. . (2007) The role of hyperhomocysteinemia as well as folate, vitamin B(6) and B(12) deficiencies in osteoporosis: a systematic review. Clin Chem Lab Med 45, 16211632.CrossRefGoogle Scholar
11Zlotkin, SH & Anderson, GH (1982) The development of cystathionase activity during the first year of life. Pediatr Res 16, 6568.CrossRefGoogle ScholarPubMed
12Ball, RO, Courtney-Martin, G & Pencharz, PB (2006) The in vivo sparing of methionine by cysteine in sulfur amino acid requirements in animal models and adult humans. J Nutr 136, 1682S1693S.CrossRefGoogle ScholarPubMed
13Storch, KJ, Wagner, DA, Burke, JF, et al. . (1990) [1-13C; methyl-2H3]methionine kinetics in humans: methionine conservation and cystine sparing. Am J Physiol 258, E790E798.Google ScholarPubMed
14Storch, KJ, Wagner, DA, Burke, JF, et al. . (1988) Quantitative study in vivo of methionine cycle in humans using [methyl-2H3]- and [1-13C]methionine. Am J Physiol 255, E322E331.Google ScholarPubMed
15Matthews, DE, Schwarz, HP, Yang, RD, et al. . (1982) Relationship of plasma leucine and α-ketoisocaproate during a l-[1-13C]leucine infusion in man: a method for measuring human intracellular leucine tracer enrichment. Metabolism 31, 11051112.CrossRefGoogle Scholar
16Tessari, P, Kiwanuka, E, Coracina, A, et al. . (2005) Insulin in methionine and homocysteine kinetics in healthy humans: plasma vs. intracellular models. Am J Physiol Endocrinol Metab 288, E1270E1276.CrossRefGoogle ScholarPubMed
17Tessari, P, Coracina, A, Kiwanuka, E, et al. . (2005) Effects of insulin on methionine and homocysteine kinetics in type 2 diabetes with nephropathy. Diabetes 54, 29682976.CrossRefGoogle ScholarPubMed
18Davis, SR, Stacpoole, PW, Williamson, J, et al. . (2004) Tracer-derived total and folate-dependent homocysteine remethylation and synthesis rates in humans indicate that serine is the main one-carbon donor. Am J Physiol Endocrinol Metab 286, E272E279.CrossRefGoogle ScholarPubMed
19Hoffer, LJ (2004) Homocysteine remethylation and trans-sulfuration. Metabolism 53, 14801483.CrossRefGoogle ScholarPubMed
20Shinohara, Y, Hasegawa, H, Ogawa, K, et al. . (2006) Distinct effects of folate and choline deficiency on plasma kinetics of methionine and homocysteine in rats. Metabolism 55, 899906.CrossRefGoogle ScholarPubMed
21Di Buono, M, Wykes, LJ, Cole, DE, et al. . (2003) Regulation of sulfur amino acid metabolism in men in response to changes in sulfur amino acid intakes. J Nutr 133, 733739.CrossRefGoogle ScholarPubMed
22Kurpad, AV, Regan, MM, Varalakshmi, S, et al. . (2004) Effect of cystine on the methionine requirement of healthy Indian men determined by using the 24-h indicator amino acid balance approach. Am J Clin Nutr 80, 15261535.CrossRefGoogle ScholarPubMed
23Fukagawa, NK (2006) Sparing of methionine requirements: evaluation of human data takes sulfur amino acids beyond protein. J Nutr 136, 1676S1681S.CrossRefGoogle Scholar
24Baker, DH, Clausing, WW, Harmon, BG, et al. . (1969) Replacement value of cystine for methionine for the young pig. J Anim Sci 29, 581584.CrossRefGoogle Scholar
25Graber, G & Baker, DH (1971) Sulfur amino acid nutrition of the growing chick: quantitative aspects concerning the efficacy of dietary methionine, cysteine and cystine. J Anim Sci 33, 10051011.CrossRefGoogle ScholarPubMed
26Aguilar, TS (1982) Studies in vivo on the methionine-sparing effect of cysteine in rats. Arch Latinoam Nutr 32, 130147.Google ScholarPubMed
27Shoveller, AK, Brunton, JA, House, JD, et al. . (2003) Dietary cysteine reduces the methionine requirement by an equal proportion in both parenterally and enterally fed piglets. J Nutr 133, 42154224.CrossRefGoogle ScholarPubMed
28Baker, DH (2006) Comparative species utilization and toxicity of sulfur amino acids. J Nutr 136, 1670S1675S.CrossRefGoogle ScholarPubMed
29Hiramatsu, T, Fukagawa, NK, Marchini, JS, et al. . (1994) Methionine and cysteine kinetics at different intakes of cystine in healthy adult men. Am J Clin Nutr 60, 525533.CrossRefGoogle ScholarPubMed
30Raguso, CA, Ajami, AM, Gleason, R, et al. . (1997) Effect of cystine intake on methionine kinetics and oxidation determined with oral tracers of methionine and cysteine in healthy adults. Am J Clin Nutr 66, 283292.CrossRefGoogle ScholarPubMed
31Fukagawa, NK, Yu, YM & Young, VR (1998) Methionine and cysteine kinetics at different intakes of methionine and cysteine in elderly men and women. Am J Clin Nutr 68, 380388.CrossRefGoogle ScholarPubMed
32Raguso, CA, Regan, MM & Young, VR (2000) Cysteine kinetics and oxidation at different intakes of methionine and cystine in young adults. Am J Clin Nutr 71, 491499.CrossRefGoogle ScholarPubMed
33Di Buono, M, Wykes, LJ, Ball, RO, et al. . (2001) Total sulfur amino acid requirement in young men as determined by indicator amino acid oxidation with l-[1-13C]phenylalanine. Am J Clin Nutr 74, 756760.CrossRefGoogle Scholar
34Kurpad, AV, Regan, MM, Varalakshmi, S, et al. . (2003) Daily methionine requirements of healthy Indian men, measured by a 24-h indicator amino acid oxidation and balance technique. Am J Clin Nutr 77, 11981205.CrossRefGoogle ScholarPubMed
35Di Buono, M, Wykes, LJ, Ball, RO, et al. . (2001) Dietary cysteine reduces the methionine requirement in men. Am J Clin Nutr 74, 761766.CrossRefGoogle ScholarPubMed
36Albanese, AA, Holt, LE Jr, Davis, VI, et al. . (1948) The sulfur amino acid requirement of the infant. Fed Proc 7, 141.Google ScholarPubMed
37Rose, WC & Wixom, RL (1955) The amino acid requirements of man. XIII. The sparing effect of cystine on the methionine requirement. J Biol Chem 216, 753773.Google ScholarPubMed
38Reynolds, MS, Steel, DL, Jones, EM, et al. . (1958) Nitrogen balances of women maintained on various levels of methionine and cystine. J Nutr 64, 99111.CrossRefGoogle ScholarPubMed
39Finkelstein, JD, Martin, JJ & Harris, BJ (1988) Methionine metabolism in mammals. The methionine-sparing effect of cystine. J Biol Chem 263, 1175011754.CrossRefGoogle ScholarPubMed
40Humayun, MA, Elango, R, Ball, RO, et al. . (2007) Reevaluation of the protein requirement in young men with the indicator amino acid oxidation technique. Am J Clin Nutr 86, 9951002.CrossRefGoogle ScholarPubMed
41Courtney-Martin, G, Rafii, M, Wykes, LJ, et al. . (2008) Methionine-adequate cysteine-free diet does not limit erythrocyte glutathione synthesis in young healthy adult men. J Nutr 138, 21722178.CrossRefGoogle Scholar
42Mudd, SH, Finkelstein, JD, Irreverre, F, et al. . (1965) Transsulfuration in mammals. Microassays and tissue distributions of three enzymes of the pathway. J Biol Chem 240, 43824392.CrossRefGoogle ScholarPubMed
43Stegink, LD & Den Besten, L (1972) Synthesis of cysteine from methionine in normal adult subjects: effect of route of alimentation. Science 178, 514516.CrossRefGoogle ScholarPubMed
44Zlotkin, SH, Bryan, MH & Anderson, GH (1981) Cysteine supplementation to cysteine-free intravenous feeding regimens in newborn infants. Am J Clin Nutr 34, 914923.CrossRefGoogle ScholarPubMed
45Shoveller, AK, Brunton, JA, Pencharz, PB, et al. . (2003) The methionine requirement is lower in neonatal piglets fed parenterally than in those fed enterally. J Nutr 133, 13901397.CrossRefGoogle ScholarPubMed
46Stoll, B & Burrin, DG (2006) Measuring splanchnic amino acid metabolism in vivo using stable isotopic tracers. J Anim Sci 84, Suppl., E60E72.CrossRefGoogle ScholarPubMed
47Stoll, B, Henry, J, Reeds, PJ, et al. . (1998) Catabolism dominates the first-pass intestinal metabolism of dietary essential amino acids in milk protein-fed piglets. J Nutr 128, 606614.CrossRefGoogle ScholarPubMed
48Shoveller, AK, House, JD, Brunton, JA, et al. . (2004) The balance of dietary sulfur amino acids and the route of feeding affect plasma homocysteine concentrations in neonatal piglets. J Nutr 134, 609612.CrossRefGoogle ScholarPubMed
49Riedijk, MA, Stoll, B, Chacko, S, et al. . (2007) Methionine transmethylation and transsulfuration in the piglet gastrointestinal tract. Proc Natl Acad Sci U S A 104, 34083413.CrossRefGoogle ScholarPubMed
50Finkelstein, JD (2000) Pathways and regulation of homocysteine metabolism in mammals. Semin Thromb Hemost 26, 219225.CrossRefGoogle ScholarPubMed
51Finkelstein, JD (1990) Methionine metabolism in mammals. J Nutr Biochem 1, 228237.CrossRefGoogle ScholarPubMed
52Bauchart-Thevret, C, Stoll, B, Chacko, S, et al. . (2009) Sulfur amino acid deficiency upregulates intestinal methionine cycle activity and suppresses epithelial growth in neonatal pigs. Am J Physiol Endocrinol Metab 296, E1239E1250.CrossRefGoogle ScholarPubMed
53Chen, LX, Yin, YL, Jobgen, WJS, et al. . (2007) In vitro oxidation of essential amino acids by jejunal mucosal cells of growing pigs. Livest Sci 109, 1923.CrossRefGoogle Scholar
54Chen, L, Li, P, Wang, J, et al. . (2009) Catabolism of nutritionally essential amino acids in developing porcine enterocytes. Amino Acids 37, 143152.CrossRefGoogle ScholarPubMed
55Wilson, FA, van den Borne, JJ, Calder, AG, et al. . (2009) Tissue methionine cycle activity and homocysteine metabolism in female rats: impact of dietary methionine and folate plus choline. Am J Physiol Endocrinol Metab 296, E702E713.CrossRefGoogle ScholarPubMed
56Stead, LM, Brosnan, ME & Brosnan, JT (2000) Characterization of homocysteine metabolism in the rat liver. Biochem J 350, 685692.Google Scholar
57Mudd, SH, Brosnan, JT, Brosnan, ME, et al. . (2007) Methyl balance and transmethylation fluxes in humans. Am J Clin Nutr 85, 1925.CrossRefGoogle ScholarPubMed
58Bos, C, Stoll, B, Fouillet, H, et al. . (2003) Intestinal lysine metabolism is driven by the enteral availability of dietary lysine in piglets fed a bolus meal. Am J Physiol Endocrinol Metab 285, E1246E1257.CrossRefGoogle ScholarPubMed
59Stipanuk, MH, Ueki, I, Dominy, JE Jr, et al. . (2009) Cysteine dioxygenase: a robust system for regulation of cellular cysteine levels. Amino Acids 37, 5563.CrossRefGoogle ScholarPubMed
60Stipanuk, MH & Rotter, MA (1984) Metabolism of cysteine, cysteinesulfinate and cysteinesulfonate in rats fed adequate and excess levels of sulfur-containing amino acids. J Nutr 114, 14261437.CrossRefGoogle ScholarPubMed
61Coloso, RM & Stipanuk, MH (1989) Metabolism of cyst(e)ine in rat enterocytes. J Nutr 119, 19141924.CrossRefGoogle ScholarPubMed
62Hirschberger, LL, Daval, S, Stover, PJ, et al. . (2001) Murine cysteine dioxygenase gene: structural organization, tissue-specific expression and promoter identification. Gene 277, 153161.CrossRefGoogle ScholarPubMed
63Cottrell, J, Stoll, B & Burrin, DG (2006) Kinetics of splanchnic 13C-cysteine metabolism in infant pigs. FASEB J 20, A9.CrossRefGoogle Scholar
64Malmezat, T, Breuille, D, Capitan, P, et al. . (2000) Glutathione turnover is increased during the acute phase of sepsis in rats. J Nutr 130, 12391246.CrossRefGoogle ScholarPubMed
65Atmaca, G (2004) Antioxidant effects of sulfur-containing amino acids. Yonsei Med J 45, 776788.CrossRefGoogle ScholarPubMed
66Huxtable, RJ (1992) Physiological actions of taurine. Physiol Rev 72, 101163.CrossRefGoogle ScholarPubMed
67Bouckenooghe, T, Remacle, C & Reusens, B (2006) Is taurine a functional nutrient? Curr Opin Clin Nutr Metab Care 9, 728733.CrossRefGoogle ScholarPubMed
68Schneider, SM, Joly, F, Gehrardt, MF, et al. . (2006) Taurine status and response to intravenous taurine supplementation in adults with short-bowel syndrome undergoing long-term parenteral nutrition: a pilot study. Br J Nutr 96, 365370.Google Scholar
69Li, X, Bazer, FW, Gao, H, et al. . (2009) Amino acids and gaseous signaling. Amino Acids 37, 6578.CrossRefGoogle ScholarPubMed
70Zhao, W, Ndisang, JF & Wang, R (2003) Modulation of endogenous production of H2S in rat tissues. Can J Physiol Pharmacol 81, 848853.Google Scholar
71Linden, DR, Sha, L, Mazzone, A, et al. . (2008) Production of the gaseous signal molecule hydrogen sulfide in mouse tissues. J Neurochem 106, 15771585.CrossRefGoogle ScholarPubMed
72Wallace, JL (2007) Hydrogen sulfide-releasing anti-inflammatory drugs. Trends Pharmacol Sci 28, 501505.CrossRefGoogle ScholarPubMed
73Teague, B, Asiedu, S & Moore, PK (2002) The smooth muscle relaxant effect of hydrogen sulphide in vitro: evidence for a physiological role to control intestinal contractility. Br J Pharmacol 137, 139145.Google Scholar
74Gallego, D, Clave, P, Donovan, J, et al. . (2008) The gaseous mediator, hydrogen sulphide, inhibits in vitro motor patterns in the human, rat and mouse colon and jejunum. Neurogastroenterol Motil 20, 13061316.CrossRefGoogle ScholarPubMed
75Schicho, R, Krueger, D, Zeller, F, et al. . (2006) Hydrogen sulfide is a novel prosecretory neuromodulator in the guinea-pig and human colon. Gastroenterology 131, 15421552.CrossRefGoogle ScholarPubMed
76Distrutti, E, Sediari, L, Mencarelli, A, et al. . (2006) Evidence that hydrogen sulfide exerts antinociceptive effects in the gastrointestinal tract by activating KATP channels. J Pharmacol Exp Ther 316, 325335.CrossRefGoogle ScholarPubMed
77Distrutti, E, Mencarelli, A, Santucci, L, et al. . (2008) The methionine connection: homocysteine and hydrogen sulfide exert opposite effects on hepatic microcirculation in rats. Hepatology 47, 659667.CrossRefGoogle ScholarPubMed
78Lobley, GE, Connell, A & Revell, D (1996) The importance of transmethylation reactions to methionine metabolism in sheep: effects of supplementation with creatine and choline. Br J Nutr 75, 4756.CrossRefGoogle ScholarPubMed
79Lobley, GE, Shen, X, Le, G, et al. . (2003) Oxidation of essential amino acids by the ovine gastrointestinal tract. Br J Nutr 89, 617630.Google Scholar
80Lobley, GE, Wester, TJ, Calder, AG, et al. . (2006) Absorption of 2-hydroxy-4-methylthiobutyrate and conversion to methionine in lambs. J Dairy Sci 89, 10721080.CrossRefGoogle ScholarPubMed
81Lobley, GE, Wester, TJ, Holtrop, G, et al. . (2006) Absorption and digestive tract metabolism of 2-hydroxy-4-methylthiobutanoic acid in lambs. J Dairy Sci 89, 35083521.CrossRefGoogle ScholarPubMed
82Dibner, JJ & Knight, CD (1984) Conversion of 2-hydroxy-4-(methylthio)butanoic acid to l-methionine in the chick: a stereospecific pathway. J Nutr 114, 17161723.Google Scholar
83Rangel-Lugo, M & Austic, RE (1998) Transamination of 2-oxo-4-[methylthio]butanoic acid in chicken tissues. Poult Sci 77, 98104.CrossRefGoogle ScholarPubMed
84Ferjancic-Biagini, A, Dupuis, L, De Caro, J, et al. . (1995) In vitro oxidative decarboxylation of l-2-hydroxy-4-methylthiobutanoic acid by l-2-hydroxy acid oxidase A from chicken liver. Biochimie 77, 249255.CrossRefGoogle ScholarPubMed
85Langer, BW Jr (1965) The biochemical conversion of 2-hydroxy-4-methylthiobutyric acid into methionine by the rat in vitro. Biochem J 95, 683687.CrossRefGoogle Scholar
86Robinson, JC, Keay, L, Molinari, R, et al. . (1962) l-α-Hydroxy acid oxidases of hog renal cortex. J Biol Chem 237, 20012010.CrossRefGoogle ScholarPubMed
87McCollum, MQ, Vazquez-Anon, M, Dibner, JJ, et al. . (2000) Absorption of 2-hydroxy-4-(methylthio)butanoic acid by isolated sheep ruminal and omasal epithelia. J Anim Sci 78, 10781083.Google Scholar
88Martin-Venegas, R, Soriano-Garcia, JF, Vinardell, MP, et al. . (2006) Oligomers are not the limiting factor in the absorption of dl-2-hydroxy-4-(methylthio)butanoic acid in the chicken small intestine. Poult Sci 85, 5663.CrossRefGoogle Scholar
89Martin-Venegas, R, Geraert, PA & Ferrer, R (2006) Conversion of the methionine hydroxy analogue dl-2-hydroxy-(4-methylthio) butanoic acid to sulfur-containing amino acids in the chicken small intestine. Poult Sci 85, 19321938.CrossRefGoogle ScholarPubMed
90Dilger, RN, Kobler, C, Weckbecker, C, et al. . (2007) 2-Keto-4-(methylthio)butyric acid (keto analog of methionine) is a safe and efficacious precursor of l-methionine in chicks. J Nutr 137, 18681873.CrossRefGoogle ScholarPubMed
91Stipanuk, MH (2007) The keto acid of methionine is a safe and efficacious substitute for dietary l-methionine: the answer from chick bioassays. J Nutr 137, 18441845.CrossRefGoogle ScholarPubMed
92Feiten, SF, Draibe, SA, Watanabe, R, et al. . (2005) Short-term effects of a very-low-protein diet supplemented with ketoacids in nondialyzed chronic kidney disease patients. Eur J Clin Nutr 59, 129136.CrossRefGoogle ScholarPubMed
93Johnson, LR (1988) Regulation of gastrointestinal mucosal growth. Physiol Rev 68, 456502.Google Scholar
94Burrin, DG, Stoll, B, Jiang, R, et al. . (2000) GLP-2 stimulates intestinal growth in premature TPN-fed pigs by suppressing proteolysis and apoptosis. Am J Physiol Gastrointest Liver Physiol 279, G1249G1256.Google Scholar
95Burrin, DG, Stoll, B, Jiang, R, et al. . (2000) Minimal enteral nutrient requirements for intestinal growth in neonatal piglets: how much is enough? Am J Clin Nutr 71, 16031610.CrossRefGoogle Scholar
96Pias, EK & Aw, TY (2002) Apoptosis in mitotic competent undifferentiated cells is induced by cellular redox imbalance independent of reactive oxygen species production. FASEB J 16, 781790.CrossRefGoogle ScholarPubMed
97Noda, T, Iwakiri, R, Fujimoto, K, et al. . (2001) Induction of mild intracellular redox imbalance inhibits proliferation of CaCo-2 cells. FASEB J 15, 21312139.CrossRefGoogle ScholarPubMed
98Jonas, CR, Ziegler, TR, Gu, LH, et al. . (2002) Extracellular thiol/disulfide redox state affects proliferation rate in a human colon carcinoma (Caco2) cell line. Free Radic Biol Med 33, 14991506.CrossRefGoogle Scholar
99Aw, TY (2003) Cellular redox: a modulator of intestinal epithelial cell proliferation. News Physiol Sci 18, 201204.Google Scholar
100Jones, DP (2002) Redox potential of GSH/GSSG couple: assay and biological significance. Methods Enzymol 348, 93112.CrossRefGoogle ScholarPubMed
101Aw, TY (1999) Molecular and cellular responses to oxidative stress and changes in oxidation–reduction imbalance in the intestine. Am J Clin Nutr 70, 557565.Google Scholar
102Martensson, J, Jain, A & Meister, A (1990) Glutathione is required for intestinal function. Proc Natl Acad Sci U S A 87, 17151719.CrossRefGoogle ScholarPubMed
103Tian, J, Washizawa, N, Gu, LH, et al. . (2007) Local glutathione redox status does not regulate ileal mucosal growth after massive small bowel resection in rats. J Nutr 137, 320325.CrossRefGoogle Scholar
104Noda, T, Iwakiri, R, Fujimoto, K, et al. . (2002) Exogenous cysteine and cystine promote cell proliferation in CaCo-2 cells. Cell Prolif 35, 117129.Google Scholar
105Anderson, CL, Iyer, SS, Ziegler, TR, et al. . (2007) Control of extracellular cysteine/cystine redox state by HT-29 cells is independent of cellular glutathione. Am J Physiol Regul Integr Comp Physiol 293, R1069R1075.Google Scholar
106Martinez-Lopez, N, Varela-Rey, M, Ariz, U, et al. . (2008) S-adenosylmethionine and proliferation: new pathways, new targets. Biochem Soc Trans 36, 848852.CrossRefGoogle ScholarPubMed
107Kotb, M, Mudd, SH, Mato, JM, et al. . (1997) Consensus nomenclature for the mammalian methionine adenosyltransferase genes and gene products. Trends Genet 13, 5152.CrossRefGoogle ScholarPubMed
108Halim, AB, LeGros, L, Geller, A, et al. . (1999) Expression and functional interaction of the catalytic and regulatory subunits of human methionine adenosyltransferase in mammalian cells. J Biol Chem 274, 2972029725.CrossRefGoogle ScholarPubMed
109Lu, SC & Mato, JM (2005) Role of methionine adenosyltransferase and S-adenosylmethionine in alcohol-associated liver cancer. Alcohol 35, 227234.CrossRefGoogle ScholarPubMed
110Ito, K, Ikeda, S, Kojima, N, et al. . (2000) Correlation between the expression of methionine adenosyltransferase and the stages of human colorectal carcinoma. Surg Today 30, 706710.Google Scholar
111Chen, H, Xia, M, Lin, M, et al. . (2007) Role of methionine adenosyltransferase 2A and S-adenosylmethionine in mitogen-induced growth of human colon cancer cells. Gastroenterology 133, 207218.CrossRefGoogle ScholarPubMed
112Chiang, PK, Gordon, RK, Tal, J, et al. . (1996) S-adenosylmethionine and methylation. FASEB J 10, 471480.CrossRefGoogle ScholarPubMed
113Brosnan, JT, da Silva, R & Brosnan, ME (2007) Amino acids and the regulation of methyl balance in humans. Curr Opin Clin Nutr Metab Care 10, 5257.CrossRefGoogle ScholarPubMed
114Reik, W & Dean, W (2001) DNA methylation and mammalian epigenetics. Electrophoresis 22, 28382843.3.0.CO;2-M>CrossRefGoogle ScholarPubMed
115Patra, SK, Patra, A, Rizzi, F, et al. . (2008) Demethylation of (cytosine-5-C-methyl) DNA and regulation of transcription in the epigenetic pathways of cancer development. Cancer Metastasis Rev 27, 315334.Google Scholar
116Moinard, C, Cynober, L & de Bandt, JP (2005) Polyamines: metabolism and implications in human diseases. Clin Nutr 24, 184197.CrossRefGoogle ScholarPubMed
117Bettuzzi, S, Davalli, P, Astancolle, S, et al. . (1999) Coordinate changes of polyamine metabolism regulatory proteins during the cell cycle of normal human dermal fibroblasts. FEBS Lett 446, 1822.CrossRefGoogle ScholarPubMed
118McCormack, SA & Johnson, LR (1991) Role of polyamines in gastrointestinal mucosal growth. Am J Physiol 260, G795G806.Google ScholarPubMed
119Wang, JY (2007) Polyamines and mRNA stability in regulation of intestinal mucosal growth. Amino Acids 33, 241252.CrossRefGoogle ScholarPubMed
120Ansorena, E, Garcia-Trevijano, ER, Martinez-Chantar, ML, et al. . (2002) S-adenosylmethionine and methylthioadenosine are antiapoptotic in cultured rat hepatocytes but proapoptotic in human hepatoma cells. Hepatology 35, 274280.CrossRefGoogle ScholarPubMed
121Clarke, SG (2006) Inhibition of mammalian protein methyltransferases by 5′-methylthioadenosine (MTA): a mechanism of action of dietary SAMe? In The Enzymes, vol. 24, pp. 467493 [Clarke, SG and Tamanoi, F, editors]. New York: Elsevier/Academic Press.Google Scholar
122Grillo, MA & Colombatto, S (2008) S-adenosylmethionine and its products. Amino Acids 34, 187193.CrossRefGoogle ScholarPubMed
123Li, TW, Zhang, Q, Oh, P, et al. . (2009) S-adenosylmethionine and methylthioadenosine inhibit cellular FLICE inhibitory protein expression and induce apoptosis in colon cancer cells. Mol Pharmacol 76, 192200.CrossRefGoogle ScholarPubMed
124Seiler, N & Raul, F (2005) Polyamines and apoptosis. J Cell Mol Med 9, 623642.CrossRefGoogle ScholarPubMed
125Sanderson, P, Stone, E, Kim, YI, et al. . (2007) Folate and colo-rectal cancer risk. Br J Nutr 98, 12991304.CrossRefGoogle ScholarPubMed
126Kim, YI (2003) Role of folate in colon cancer development and progression. J Nutr 133, 3731S3739S.CrossRefGoogle ScholarPubMed
127Ulrich, CM (2007) Folate and cancer prevention: a closer look at a complex picture. Am J Clin Nutr 86, 271273.CrossRefGoogle Scholar
128Smith, AD, Kim, YI & Refsum, H (2008) Is folic acid good for everyone? Am J Clin Nutr 87, 517533.CrossRefGoogle ScholarPubMed
129Kim, YI (2008) Folic acid supplementation and cancer risk: point. Cancer Epidemiol Biomarkers Prev 17, 22202225.CrossRefGoogle ScholarPubMed
130Giovannucci, E (2002) Epidemiologic studies of folate and colorectal neoplasia: a review. J Nutr 132, 2350S2355S.CrossRefGoogle ScholarPubMed
131Jones, PA & Baylin, SB (2002) The fundamental role of epigenetic events in cancer. Nat Rev Genet 3, 415428.CrossRefGoogle ScholarPubMed
132Duthie, SJ, Narayanan, S, Sharp, L, et al. . (2004) Folate, DNA stability and colo-rectal neoplasia. Proc Nutr Soc 63, 571578.CrossRefGoogle ScholarPubMed
133Kim, DH (2007) The interactive effect of methyl-group diet and polymorphism of methylenetetrahydrofolate reductase on the risk of colorectal cancer. Mutat Res 622, 1418.CrossRefGoogle ScholarPubMed
134Frosst, P, Blom, HJ, Milos, R, et al. . (1995) A candidate genetic risk factor for vascular disease: a common mutation in methylenetetrahydrofolate reductase. Nat Genet 10, 111113.CrossRefGoogle ScholarPubMed
135Hubner, RA & Houlston, RS (2007) MTHFR C677T and colorectal cancer risk: a meta-analysis of 25 populations. Int J Cancer 120, 10271035.CrossRefGoogle Scholar
136Chen, J, Giovannucci, E, Kelsey, K, et al. . (1996) A methylenetetrahydrofolate reductase polymorphism and the risk of colorectal cancer. Cancer Res 56, 48624864.Google ScholarPubMed
137Ma, J, Stampfer, MJ, Giovannucci, E, et al. . (1997) Methylenetetrahydrofolate reductase polymorphism, dietary interactions, and risk of colorectal cancer. Cancer Res 57, 10981102.Google ScholarPubMed
138Avila, MA, Garcia-Trevijano, ER, Lu, SC, et al. . (2004) Methylthioadenosine. Int J Biochem Cell Biol 36, 21252130.CrossRefGoogle ScholarPubMed
139Morgenstern, I, Raijmakers, MT, Peters, WH, et al. . (2003) Homocysteine, cysteine, and glutathione in human colonic mucosa: elevated levels of homocysteine in patients with inflammatory bowel disease. Dig Dis Sci 48, 20832090.Google Scholar
140Danese, S, Sgambato, A, Papa, A, et al. . (2005) Homocysteine triggers mucosal microvascular activation in inflammatory bowel disease. Am J Gastroenterol 100, 886895.CrossRefGoogle ScholarPubMed
141Papatheodorou, L & Weiss, N (2007) Vascular oxidant stress and inflammation in hyperhomocysteinemia. Antioxid Redox Signal 9, 19411958.CrossRefGoogle ScholarPubMed
142Zhang, C, Cai, Y, Adachi, MT, et al. . (2001) Homocysteine induces programmed cell death in human vascular endothelial cells through activation of the unfolded protein response. J Biol Chem 276, 3586735874.Google Scholar
143Papa, A, Scaldaferri, F, Danese, S, et al. . (2008) Vascular involvement in inflammatory bowel disease: pathogenesis and clinical aspects. Dig Dis 26, 149155.Google Scholar
144Deban, L, Correale, C, Vetrano, S, et al. . (2008) Multiple pathogenic roles of microvasculature in inflammatory bowel disease: a Jack of all trades. Am J Pathol 172, 14571466.CrossRefGoogle ScholarPubMed
145Lu, SC & Mato, JM (2008) S-adenosylmethionine in cell growth, apoptosis and liver cancer. J Gastroenterol Hepatol 23, Suppl. 1, S73S77.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1 Sulfur amino acid metabolism. AMT, aminotransferase; BHMT, betaine-homocysteine methyltransferase; CBS, cystathionine β-synthase; CDO, cysteine dioxygenase; CGL, cystathionine γ-lyase; CSD, cysteine sulfinate decarboxylase; DHFR, dihydrofolate reductase; DMG, dimethyl-glycine; dTMP, thymidylate; dUMP, deoxyuridylate; MAT, methionine adenosyltransferase; MS, methionine synthase; MT, methyl transferases; MTHFR, methylenetetrahydrofolate reductase, SAH, S-adenosylhomocysteine; SAHH, S-adenosylhomocysteine hydrolase; SAM, S-adenosylmethionine; SHMT, serine hydroxymethyltransferase; TS, thymidylate synthetase.

Figure 1

Fig. 2 Stable isotopic tracer model used by Storch et al.(14). This model is based on the intravenous infusion of a methionine isotopomer [1-13C; methyl-2H3]methionine [M+4]. The [2H3]methyl group is lost during transmethylation from S-adenosylmethionine (SAM) to S-adenosylhomocysteine (SAH), whereas the [13C]carbon group is transferred to [1-13C]homocysteine [M+1], [1-13C]cystathionine [M+1] and finally, oxidised to 13CO2 via α-ketobutyrate in the tricarboxylic acid cycle.

Figure 2

Fig. 3 Metabolic pathway for the conversion of 2-hydroxy-(4-methylthio)butanoic acid (HMB) to l-methionine. KMB, 2-keto-(4-methylthio)butanoic acid.

Figure 3

Fig. 4 S-adenosylmethionine (SAM) metabolic pathways: transmethylation and polyamine synthesis. dcSAM, decarboxylated S-adenosylmethionine; MAT, methionine adenosyltransferase; MTA, methylthioadenosine; SAH, S-adenosylhomocysteine; SAMDC, S-adenosylmethionine decarboxylase; ODC, ornithine decarboxylase; PAO, polyamine oxidase; SSAT, spermidine/spermine N1-acetyltransferase; , methionine salvage pathway.

Figure 4

Table 1 Properties of mammalian methionine adenosyltransferase (MAT) isoforms (from Lu & Mato(145))

Figure 5

Fig. 5 Methionine salvage pathway from 5′-methylthioadenosine (MTA). MTOB, 4-methylthio-2-oxobutanoic acid; MTR-1P, 5′-methylthioribose-1-P.