p-Cresyl Sulfate

Tessa Gryp, Raymond Vanholder, Mario Vaneechoutte, Griet Glorieux, Tessa Gryp, Raymond Vanholder, Mario Vaneechoutte, Griet Glorieux

Abstract

If chronic kidney disease (CKD) is associated with an impairment of kidney function, several uremic solutes are retained. Some of these exert toxic effects, which are called uremic toxins. p-Cresyl sulfate (pCS) is a prototype protein-bound uremic toxin to which many biological and biochemical (toxic) effects have been attributed. In addition, increased levels of pCS have been associated with worsening outcomes in CKD patients. pCS finds its origin in the intestine where gut bacteria metabolize aromatic amino acids, such as tyrosine and phenylalanine, leading to phenolic end products, of which pCS is one of the components. In this review we summarize the biological effects of pCS and its metabolic origin in the intestine. It appears that, according to in vitro studies, the intestinal bacteria generating phenolic compounds mainly belong to the families Bacteroidaceae, Bifidobacteriaceae, Clostridiaceae, Enterobacteriaceae, Enterococcaceae, Eubacteriaceae, Fusobacteriaceae, Lachnospiraceae, Lactobacillaceae, Porphyromonadaceae, Staphylococcaceae, Ruminococcaceae, and Veillonellaceae. Since pCS remains difficult to remove by dialysis, the gut microbiota could be a future target to decrease pCS levels and its toxicity, even at earlier stages of CKD, aiming at slowing down the progression of the disease and decreasing the cardiovascular burden.

Keywords: chronic kidney disease; intestinal microbiota; p-cresyl sulfate.

Conflict of interest statement

The authors declare no conflict of interest.

Figures

Figure 1
Figure 1
Conversion pathway of tyrosine and phenylalanine into p-cresyl sulfate. Full arrows: process through bacterial fermentation; dotted arrows: process through enzymatic reactions of the host; EC: enzyme commission number. l-tyrosine, derived from diet and endogenous proteins and peptides, can be converted to phenol and 4-hydroxyphenylpyruvate. Tyrosine phenol-lyase (EC 4.1.99.2.), previously named β–tyrosinase, is responsible for the reversible deamination of l-tyrosine, requiring pyridoxyl phosphate as a cofactor, into phenol ammonia and pyruvate [40,41,42]. This reaction is also reversible by the same enzyme using l-serine and phenol as substrates [43]. In addition, the reversible reaction of l-tyrosine with 2-oxoglutarate in 4-hydroxyphenylpyruvate and L-glutamate is catalysed by tyrosine transaminase (EC 2.6.1.5.) or by aromatic-amino-acid transaminase (EC 2.6.1.57.) [44,45,46]. To a small extent, 4-hydroxyphenylpyruvate and ammonia can also be formed by the enzyme phenylalanine dehydrogenase (EC 1.4.1.20.) from l-tyrosine [47]. 4-Hydroxyphenylpyruvate is the precursor of 4-hydroxyphenylacetate, catalysed by p-hydroxyphenylpyruvate oxidase (EC 1.2.3.13.) [44,46], and can subsequently lead to the formation of p-cresol by p-hydroxyphenylacetate decarboxylase (EC 4.1.1.83.) [48,49]. In the gut mucosa and in the liver, the majority of p-cresol will be conjugated into the uremic toxin p-cresyl sulfate by aryl sulfotransferases (EC 2.8.2.1.) [50] and a small fraction will be metabolized to p-cresyl glucuronide by UDP-glucuronyltransferases (EC 2.4.1.17.) [51]. Another aromatic amino acid, phenylalanine, also plays a role in the production of p-cresyl sulfate through the hydroxylation reaction to tyrosine by phenylalanine 4-monooxygenase, also referred as phenylalanine hydroxylase (EC 1.14.16.1.) [52]. This metabolic process is carried out by bacteria as well by liver cells, transforming excess diet phenylalanine to tyrosine [53]. In addition, phenylalanine is converted into 3-phenylpyruvate by either phenylalanine dehydrogenase (EC 1.4.1.20.) [47,54], branched-chain-amino-acid transaminase (EC 2.6.1.42.) [55] or by aromatic-amino-acid transaminase (EC 2.6.1.57.) [45]. Then 3-phenylpyruvate can be transformed in 3-phenylacetaldehyde by phenylpyruvate decarboxylase (EC 4.1.1.43.) [56] or in 3-phenyllactate by l-lactate dehydrogenase (EC 1.1.1.27.) [46]. Finally, 3-phenylacetaldehyde can be converted to 3-phenylacetate by phenylacetaldehyde dehydrogenase (EC 1.2.1.39.) [56,57]. All of these reactions of the phenylalanine metabolic pathway are reversible, which can lead, in the end, to p-cresyl sulfate generation.

References

    1. Duranton F., Cohen G., de Smet R., Rodriguez M., Jankowski J., Vanholder R., Argiles A. Normal and pathologic concentrations of uremic toxins. J. Am. Soc. Nephrol. 2012;23:1258–1270. doi: 10.1681/ASN.2011121175.
    1. Vanholder R., de Smet R., Glorieux G., Argiles A., Baurmeister U., Brunet P., Clark W., Cohen G., de Deyn P.P., Deppisch R., et al. Review on uremic toxins: Classification, concentration, and interindividual variability. Kidney Int. 2003;63:1934–1943. doi: 10.1046/j.1523-1755.2003.00924.x.
    1. Vanholder R., Glorieux G., de Smet R., Lameire N. New insights in uremic toxins. Kidney Int. Suppl. 2003:S6–S10. doi: 10.1046/j.1523-1755.63.s84.43.x.
    1. Meyer T.W., Hostetter T.H. Uremia. N. Engl. J. Med. 2007;357:1316–1325. doi: 10.1056/NEJMra071313.
    1. Vanholder R., de Smet R. Pathophysiologic effects of uremic retention solutes. J. Am. Soc. Nephrol. 1999;10:1815–1823.
    1. Vanholder R., Massy Z., Argiles A., Spasovski G., Verbeke F., Lameire N. Chronic kidney disease as cause of cardiovascular morbidity and mortality. Nephrol. Dial. Transplant. 2005;20:1048–1056. doi: 10.1093/ndt/gfh813.
    1. Vanholder R., Baurmeister U., Brunet P., Cohen G., Glorieux G., Jankowski J. A bench to bedside view of uremic toxins. J. Am. Soc. Nephrol. 2008;19:863–870. doi: 10.1681/ASN.2007121377.
    1. Evenepoel P., Meijers B.K., Bammens B.R., Verbeke K. Uremic toxins originating from colonic microbial metabolism. Kidney Int. Suppl. 2009:S12–S19. doi: 10.1038/ki.2009.402.
    1. Lesaffer G., de Smet R., Lameire N., Dhondt A., Duym P., Vanholder R. Intradialytic removal of protein-bound uraemic toxins: Role of solute characteristics and of dialyser membrane. Nephrol. Dial. Transplant. 2000;15:50–57. doi: 10.1093/ndt/15.1.50.
    1. Meert N., Beerenhout C., Schepers E., Glorieux G., Kooman J., Vanholder R. Evolution of protein-bound uraemic solutes during predilution haemofiltration. J. Nephrol. 2009;22:352–357.
    1. Deltombe O., Van Biesen W., Glorieux G., Massy Z., Dhondt A., Eloot S. Exploring protein binding of uremic toxins in patients with different stages of chronic kidney disease and during hemodialysis. Toxins (Basel) 2015;7:3933–3946. doi: 10.3390/toxins7103933.
    1. Pretorius C.J., McWhinney B.C., Sipinkoski B., Johnson L.A., Rossi M., Campbell K.L., Ungerer J.P. Reference ranges and biological variation of free and total serum indoxyl- and p-cresyl sulphate measured with a rapid UPLC fluorescence detection method. Clin. Chim. Acta. 2013;419:122–126. doi: 10.1016/j.cca.2013.02.008.
    1. Boelaert J., Lynen F., Glorieux G., Eloot S., Van Landschoot M., Waterloos M.A., Sandra P., Vanholder R. A novel UPLC-MS-MS method for simultaneous determination of seven uremic retention toxins with cardiovascular relevance in chronic kidney disease patients. Anal. Bioanal. Chem. 2013;405:1937–1947. doi: 10.1007/s00216-012-6636-9.
    1. Lin C.J., Chen H.H., Pan C.F., Chuang C.K., Wang T.J., Sun F.J., Wu C.J. p-Cresylsulfate and indoxyl sulfate level at different stages of chronic kidney disease. J. Clin. Lab. Anal. 2011;25:191–197. doi: 10.1002/jcla.20456.
    1. Cuoghi A., Caiazzo M., Bellei E., Monari E., Bergamini S., Palladino G., Ozben T., Tomasi A. Quantification of p-cresol sulphate in human plasma by selected reaction monitoring. Anal. Bioanal. Chem. 2012;404:2097–2104. doi: 10.1007/s00216-012-6277-z.
    1. Hooper L.V., Gordon J.I. Commensal host-bacterial relationships in the gut. Science. 2001;292:1115–1118. doi: 10.1126/science.1058709.
    1. Flint H.J., Scott K.P., Louis P., Duncan S.H. The role of the gut microbiota in nutrition and health. Nat. Rev. Gastroenterol. Hepatol. 2012;9:577–589. doi: 10.1038/nrgastro.2012.156.
    1. Frank D.N., St Amand A.L., Feldman R.A., Boedeker E.C., Harpaz N., Pace N.R. Molecular-phylogenetic characterization of microbial community imbalances in human inflammatory bowel diseases. Proc. Natl. Acad. Sci. USA. 2007;104:13780–13785. doi: 10.1073/pnas.0706625104.
    1. Francescone R., Hou V., Grivennikov S.I. Microbiome, inflammation, and cancer. Cancer J. 2014;20:181–189. doi: 10.1097/PPO.0000000000000048.
    1. Ley R.E., Turnbaugh P.J., Klein S., Gordon J.I. Microbial ecology: Human gut microbes associated with obesity. Nature. 2006;444:1022–1023. doi: 10.1038/4441022a.
    1. Dunne J.L., Triplett E.W., Gevers D., Xavier R., Insel R., Danska J., Atkinson M.A. The intestinal microbiome in type 1 diabetes. Clin. Exp. Immunol. 2014;177:30–37. doi: 10.1111/cei.12321.
    1. Qin J., Li Y., Cai Z., Li S., Zhu J., Zhang F., Liang S., Zhang W., Guan Y., Shen D., et al. A metagenome-wide association study of gut microbiota in type 2 diabetes. Nature. 2012;490:55–60. doi: 10.1038/nature11450.
    1. Rajendhran J., Shankar M., Dinakaran V., Rathinavel A., Gunasekaran P. Contrasting circulating microbiome in cardiovascular disease patients and healthy individuals. Int. J. Cardiol. 2013;168:5118–5120. doi: 10.1016/j.ijcard.2013.07.232.
    1. Ramezani A., Raj D.S. The gut microbiome, kidney disease, and targeted interventions. J. Am. Soc. Nephrol. 2014;25:657–670. doi: 10.1681/ASN.2013080905.
    1. Backhed F., Ley R.E., Sonnenburg J.L., Peterson D.A., Gordon J.I. Host-bacterial mutualism in the human intestine. Science. 2005;307:1915–1920. doi: 10.1126/science.1104816.
    1. Guarner F. Enteric flora in health and disease. Digestion. 2006;73(Suppl. 1):5–12. doi: 10.1159/000089775.
    1. Yao C.K., Muir J.G., Gibson P.R. Review article: Insights into colonic protein fermentation, its modulation and potential health implications. Aliment. Pharmacol. Ther. 2016;43:181–196. doi: 10.1111/apt.13456.
    1. Chacko A., Cummings J.H. Nitrogen losses from the human small bowel: Obligatory losses and the effect of physical form of food. Gut. 1988;29:809–815. doi: 10.1136/gut.29.6.809.
    1. Smith E.A., Macfarlane G.T. Enumeration of human colonic bacteria producing phenolic and indolic compounds: Effects of pH, carbohydrate availability and retention time on dissimilatory aromatic amino acid metabolism. J. Appl. Bacteriol. 1996;81:288–302. doi: 10.1111/j.1365-2672.1996.tb04331.x.
    1. Macfarlane G.T., Cummings J.H., Allison C. Protein degradation by human intestinal bacteria. J. Gen. Microbiol. 1986;132:1647–1656. doi: 10.1099/00221287-132-6-1647.
    1. Gibson S.A., McFarlan C., Hay S., MacFarlane G.T. Significance of microflora in proteolysis in the colon. Appl. Environ. Microbiol. 1989;55:679–683.
    1. Smith E.A., Macfarlane G.T. Dissimilatory amino Acid metabolism in human colonic bacteria. Anaerobe. 1997;3:327–337. doi: 10.1006/anae.1997.0121.
    1. Ramakrishna B.S., Gee D., Weiss A., Pannall P., Roberts-Thomson I.C., Roediger W.E. Estimation of phenolic conjugation by colonic mucosa. J. Clin. Pathol. 1989;42:620–623. doi: 10.1136/jcp.42.6.620.
    1. Schepers E., Glorieux G., Vanholder R. The gut: The forgotten organ in uremia? Blood Purif. 2010;29:130–136. doi: 10.1159/000245639.
    1. Poesen R., Evenepoel P., de Loor H., Kuypers D., Augustijns P., Meijers B. Metabolism, protein binding, and renal clearance of microbiota-derived p-cresol in patients with CKD. Clin. J. Am. Soc. Nephrol. 2016;11:1136–1144. doi: 10.2215/CJN.00160116.
    1. de Loor H., Bammens B., Evenepoel P., de Preter V., Verbeke K. Gas chromatographic-mass spectrometric analysis for measurement of p-cresol and its conjugated metabolites in uremic and normal serum. Clin. Chem. 2005;51:1535–1538. doi: 10.1373/clinchem.2005.050781.
    1. Meyer T.W., Hostetter T.H. Uremic solutes from colon microbes. Kidney Int. 2012;81:949–954. doi: 10.1038/ki.2011.504.
    1. Suchy-Dicey A.M., Laha T., Hoofnagle A., Newitt R., Sirich T.L., Meyer T.W., Thummel K.E., Yanez N.D., Himmelfarb J., Weiss N.S., et al. Tubular secretion in CKD. J. Am. Soc. Nephrol. 2016;27:2148–2155. doi: 10.1681/ASN.2014121193.
    1. Vanholder R., Bammens B., de Loor H., Glorieux G., Meijers B., Schepers E., Massy Z., Evenepoel P. Warning: The unfortunate end of p-cresol as a uraemic toxin. Nephrol. Dial. Transplant. 2011;26:1464–1467. doi: 10.1093/ndt/gfr056.
    1. Kumagai H., Matsui H., Yamada H. Formation of tyrosine phenol-lyase by bacteria. Agric. Biol. Chem. 1970;34:1259–1261. doi: 10.1080/00021369.1970.10859762.
    1. Brot N., Smit Z., Weissbach H. Conversion of l-tyrosine to phenol by Clostridium tetanomorphum. Arch. Biochem. Biophys. 1965;112:1–6. doi: 10.1016/0003-9861(65)90002-0.
    1. Chandel M., Azmi W. Optimization of process parameters for the production of tyrosine phenol lyase by Citrobacter freundii MTCC 2424. Bioresour. Technol. 2009;100:1840–1846. doi: 10.1016/j.biortech.2008.09.044.
    1. Enei H., Matsui H., Yamashita K. Distribution of tyrosine phenol lyase in microorganisms. Agric. Biol. Chem. 1972;36:1861–1868. doi: 10.1080/00021369.1972.10860505.
    1. Blakley E.R. The catabolism of l-tyrosine by an Arthrobacter sp. Can. J. Microbiol. 1977;23:1128–1139. doi: 10.1139/m77-169.
    1. Powell J.T., Morrison J.F. The purification and properties of the aspartate aminotransferase and aromatic-amino-acid aminotransferase from Escherichia coli. Eur. J. Biochem. 1978;87:391–400. doi: 10.1111/j.1432-1033.1978.tb12388.x.
    1. Gummalla S., Broadbent J.R. Tyrosine and phenylalanine catabolism by Lactobacillus cheese flavor adjuncts. J. Dairy Sci. 2001;84:1011–1019. doi: 10.3168/jds.S0022-0302(01)74560-2.
    1. Seah S.Y., Britton K.L., Rice D.W., Asano Y., Engel P.C. Single amino acid substitution in Bacillus sphaericus phenylalanine dehydrogenase dramatically increases its discrimination between phenylalanine and tyrosine substrates. Biochemistry. 2002;41:11390–11397. doi: 10.1021/bi020196a.
    1. Selmer T., Andrei P.I. p-Hydroxyphenylacetate decarboxylase from Clostridium difficile. A novel glycyl radical enzyme catalysing the formation of p-cresol. Eur. J. Biochem. 2001;268:1363–1372. doi: 10.1046/j.1432-1327.2001.02001.x.
    1. Yokoyama M.T., Carlson J.R. Production of skatole and para-cresol by a rumen Lactobacillus sp. Appl. Environ. Microbiol. 1981;41:71–76.
    1. Brix L.A., Barnett A.C., Duggleby R.G., Leggett B., McManus M.E. Analysis of the substrate specificity of human sulfotransferases SULT1A1 and SULT1A3: Site-directed mutagenesis and kinetic studies. Biochemistry. 1999;38:10474–10479. doi: 10.1021/bi990795q.
    1. King C.D., Rios G.R., Green M.D., Tephly T.R. UDP-glucuronosyltransferases. Curr. Drug Metab. 2000;1:143–161. doi: 10.2174/1389200003339171.
    1. Erlandsen H., Kim J.Y., Patch M.G., Han A., Volner A., Abu-Omar M.M., Stevens R.C. Structural comparison of bacterial and human iron-dependent phenylalanine hydroxylases: Similar fold, different stability and reaction rates. J. Mol. Biol. 2002;320:645–661. doi: 10.1016/S0022-2836(02)00496-5.
    1. Fitzpatrick P.F. Mechanism of aromatic amino acid hydroxylation. Biochemistry. 2003;42:14083–14091. doi: 10.1021/bi035656u.
    1. Asano Y., Nakazawa A., Endo K., Hibino Y., Ohmori M., Numao N., Kondo K. Phenylalanine dehydrogenase of Bacillus badius. Purification, characterization and gene cloning. Eur. J. Biochem. 1987;168:153–159. doi: 10.1111/j.1432-1033.1987.tb13399.x.
    1. Lee-Peng F.C., Hermodson M.A., Kohlhaw G.B. Transaminase B from Escherichia coli: Quaternary structure, amino-terminal sequence, substrate specificity, and absence of a separate valine-alpha-ketoglutarate activity. J. Bacteriol. 1979;139:339–345.
    1. Asakawa T., Wada H., Yamano T. Enzymatic conversion of phenylpyruvate to phenylacetate. Biochim. Biophys. Acta. 1968;170:375–391. doi: 10.1016/0304-4165(68)90017-2.
    1. Ferrandez A., Prieto M.A., Garcia J.L., Diaz E. Molecular characterization of PadA, a phenylacetaldehyde dehydrogenase from Escherichia coli. FEBS Lett. 1997;406:23–27. doi: 10.1016/S0014-5793(97)00228-7.
    1. Bammens B., Verbeke K., Vanrenterghem Y., Evenepoel P. Evidence for impaired assimilation of protein in chronic renal failure. Kidney Int. 2003;64:2196–2203. doi: 10.1046/j.1523-1755.2003.00314.x.
    1. Bammens B., Evenepoel P., Verbeke K., Vanrenterghem Y. Impairment of small intestinal protein assimilation in patients with end-stage renal disease: Extending the malnutrition-inflammation-atherosclerosis concept. Am. J. Clin. Nutr. 2004;80:1536–1543.
    1. Qureshi A.R., Alvestrand A., Danielsson A., Divino-Filho J.C., Gutierrez A., Lindholm B., Bergstrom J. Factors predicting malnutrition in hemodialysis patients: A cross-sectional study. Kidney Int. 1998;53:773–782. doi: 10.1046/j.1523-1755.1998.00812.x.
    1. Cianciaruso B., Brunori G., Kopple J.D., Traverso G., Panarello G., Enia G., Strippoli P., de Vecchi A., Querques M., Viglino G., et al. Cross-sectional comparison of malnutrition in continuous ambulatory peritoneal dialysis and hemodialysis patients. Am. J. Kidney Dis. 1995;26:475–486. doi: 10.1016/0272-6386(95)90494-8.
    1. Evenepoel P., Claus D., Geypens B., Hiele M., Geboes K., Rutgeerts P., Ghoos Y. Amount and fate of egg protein escaping assimilation in the small intestine of humans. Am. J. Physiol. 1999;277:G935–G943.
    1. Evenepoel P., Claus D., Geypens B., Maes B., Hiele M., Rutgeerts P., Ghoos Y. Evidence for impaired assimilation and increased colonic fermentation of protein, related to gastric acid suppression therapy. Aliment. Pharmacol. Ther. 1998;12:1011–1019. doi: 10.1046/j.1365-2036.1998.00377.x.
    1. Geypens B., Claus D., Evenepoel P., Hiele M., Maes B., Peeters M., Rutgeerts P., Ghoos Y. Influence of dietary protein supplements on the formation of bacterial metabolites in the colon. Gut. 1997;41:70–76. doi: 10.1136/gut.41.1.70.
    1. McCullough A.J., Mullen K.D., Tavill A.S., Kalhan S.C. In vivo differences between the turnover rates of leucine and leucine’s ketoacid in stable cirrhosis. Gastroenterology. 1992;103:571–578. doi: 10.1016/0016-5085(92)90849-T.
    1. Halvatsiotis P.G., Turk D., Alzaid A., Dinneen S., Rizza R.A., Nair K.S. Insulin effect on leucine kinetics in type 2 diabetes mellitus. Diabetes Nutr. Metab. 2002;15:136–142.
    1. Mitch W.E., Price S.R., May R.C., Jurkovitz C., England B.K. Metabolic consequences of uremia: Extending the concept of adaptive responses to protein metabolism. Am. J. Kidney Dis. 1994;23:224–228. doi: 10.1016/S0272-6386(12)80976-0.
    1. Lim V.S., Yarasheski K.E., Flanigan M.J. The effect of uraemia, acidosis, and dialysis treatment on protein metabolism: A longitudinal leucine kinetic study. Nephrol. Dial. Transplant. 1998;13:1723–1730. doi: 10.1093/ndt/13.7.1723.
    1. Tremaroli V., Backhed F. Functional interactions between the gut microbiota and host metabolism. Nature. 2012;489:242–249. doi: 10.1038/nature11552.
    1. Van der Meulen R., Camu N., Van Vooren T., Heymans C., de Vuyst L. In vitro kinetic analysis of carbohydrate and aromatic amino acid metabolism of different members of the human colon. Int. J. Food Microbiol. 2008;124:27–33. doi: 10.1016/j.ijfoodmicro.2008.02.013.
    1. Ikeda N., Saito Y., Shimizu J., Ochi A., Mizutani J., Watabe J. Variations in concentrations of bacterial metabolites, enzyme activities, moisture, pH and bacterial composition between and within individuals in faeces of seven healthy adults. J. Appl. Bacteriol. 1994;77:185–194. doi: 10.1111/j.1365-2672.1994.tb03063.x.
    1. Aronov P.A., Luo F.J., Plummer N.S., Quan Z., Holmes S., Hostetter T.H., Meyer T.W. Colonic contribution to uremic solutes. J. Am. Soc. Nephrol. 2011;22:1769–1776. doi: 10.1681/ASN.2010121220.
    1. Wikoff W.R., Anfora A.T., Liu J., Schultz P.G., Lesley S.A., Peters E.C., Siuzdak G. Metabolomics analysis reveals large effects of gut microflora on mammalian blood metabolites. Proc. Natl. Acad. Sci. USA. 2009;106:3698–3703. doi: 10.1073/pnas.0812874106.
    1. Kikuchi M., Ueno M., Itoh Y., Suda W., Hattori M. Uremic toxin-producing gut microbiota in rats with chronic kidney disease. Nephron. 2017;135:51–60. doi: 10.1159/000450619.
    1. Roager H.M., Hansen L.B., Bahl M.I., Frandsen H.L., Carvalho V., Gobel R.J., Dalgaard M.D., Plichta D.R., Sparholt M.H., Vestergaard H., et al. Colonic transit time is related to bacterial metabolism and mucosal turnover in the gut. Nat. Microbiol. 2016;1:16093. doi: 10.1038/nmicrobiol.2016.93.
    1. Elsden S.R., Hilton M.G., Waller J.M. The end products of the metabolism of aromatic amino acids by Clostridia. Arch. Microbiol. 1976;107:283–288. doi: 10.1007/BF00425340.
    1. Mead G.C. The amino acid-fermenting Clostridia. J. Gen. Microbiol. 1971;67:47–56. doi: 10.1099/00221287-67-1-47.
    1. Giesel H., Simon H. On the occurrence of enoate reductase and 2-oxo-carboxylate reductase in clostridia and some observations on the amino acid fermentation by Peptostreptococcus anaerobius. Arch. Microbiol. 1983;135:51–57. doi: 10.1007/BF00419482.
    1. Moss C.W., Lambert M.A., Goldsmith D.J. Production of hydrocinnamic acid by Clostridia. Appl. Microbiol. 1970;19:375–378.
    1. Bone E., Tamm A., Hill M. The production of urinary phenols by gut bacteria and their possible role in the causation of large bowel cancer. Am. J. Clin. Nutr. 1976;29:1448–1454.
    1. Walker A.W., Duncan S.H., Louis P., Flint H.J. Phylogeny, culturing, and metagenomics of the human gut microbiota. Trends Microbiol. 2014;22:267–274. doi: 10.1016/j.tim.2014.03.001.
    1. Vaziri N.D., Wong J., Pahl M., Piceno Y.M., Yuan J., DeSantis T.Z., Ni Z., Nguyen T.H., Andersen G.L. Chronic kidney disease alters intestinal microbial flora. Kidney Int. 2013;83:308–315. doi: 10.1038/ki.2012.345.
    1. Strid H., Simren M., Stotzer P.O., Ringstrom G., Abrahamsson H., Bjornsson E.S. Patients with chronic renal failure have abnormal small intestinal motility and a high prevalence of small intestinal bacterial overgrowth. Digestion. 2003;67:129–137. doi: 10.1159/000071292.
    1. Simenhoff M.L., Saukkonen J.J., Burke J.F., Wesson L.G., Jr., Schaedler R.W., Gordon S.J. Bacterial populations of the small intestine in uremia. Nephron. 1978;22:63–68. doi: 10.1159/000181424.
    1. Hida M., Aiba Y., Sawamura S., Suzuki N., Satoh T., Koga Y. Inhibition of the accumulation of uremic toxins in the blood and their precursors in the feces after oral administration of Lebenin, a lactic acid bacteria preparation, to uremic patients undergoing hemodialysis. Nephron. 1996;74:349–355. doi: 10.1159/000189334.
    1. Fukuuchi F., Hida M., Aiba Y., Koga Y., Endoh M., Kurokawa K., Sakai H. Intestinal bacteria-derived putrefactants in chronic renal failure. Clin. Exp. Nephrol. 2002;6:99–104. doi: 10.1007/s101570200016.
    1. Yoshifuji A., Wakino S., Irie J., Tajima T., Hasegawa K., Kanda T., Tokuyama H., Hayashi K., Itoh H. Gut Lactobacillus protects against the progression of renal damage by modulating the gut environment in rats. Nephrol. Dial. Transplant. 2016;31:401–412. doi: 10.1093/ndt/gfv353.
    1. Barrios C., Beaumont M., Pallister T., Villar J., Goodrich J.K., Clark A., Pascual J., Ley R.E., Spector T.D., Bell J.T., et al. Gut-microbiota-metabolite axis in early renal function decline. PLoS ONE. 2015;10:e0134311. doi: 10.1371/journal.pone.0134311.
    1. Wang I.K., Lai H.C., Yu C.J., Liang C.C., Chang C.T., Kuo H.L., Yang Y.F., Lin C.C., Lin H.H., Liu Y.L., et al. Real-time PCR analysis of the intestinal microbiotas in peritoneal dialysis patients. Appl. Environ. Microbiol. 2012;78:1107–1112. doi: 10.1128/AEM.05605-11.
    1. Bourke E., Milne M.D., Stokes G.S. Caecal pH and ammonia in experimental uraemia. Gut. 1966;7:558–561. doi: 10.1136/gut.7.5.558.
    1. Wong J., Piceno Y.M., Desantis T.Z., Pahl M., Andersen G.L., Vaziri N.D. Expansion of urease- and uricase-containing, indole- and p-cresol-forming and contraction of short-chain fatty acid-producing intestinal microbiota in ESRD. Am. J. Nephrol. 2014;39:230–237. doi: 10.1159/000360010.
    1. Mayrand D., Bourgeau G. Production of phenylacetic acid by anaerobes. J. Clin. Microbiol. 1982;16:747–750.
    1. Russell W.R., Duncan S.H., Scobbie L., Duncan G., Cantlay L., Calder A.G., Anderson S.E., Flint H.J. Major phenylpropanoid-derived metabolites in the human gut can arise from microbial fermentation of protein. Mol. Nutr. Food Res. 2013;57:523–535. doi: 10.1002/mnfr.201200594.
    1. Elsden S.R., Hilton M.G. Amino acid utilization patterns in clostridial taxonomy. Arch. Microbiol. 1979;123:137–141. doi: 10.1007/BF00446812.
    1. Giesel H., Machacek G., Bayerl J., Simon H. On the formation of 3-phenylpropionate and the different stereo-chemical course of the reduction of cinnamate by Clostridium sporogenes and Peptostreptococcus anaerobius. FEBS Lett. 1981;123:107–110. doi: 10.1016/0014-5793(81)80030-0.
    1. Ohhiral I., Kuwaki S., Morita H., Suzuki T., Tomita S., Hisamatsu S., Sonoki S., Shinoda S. Identification of 3-phenyllactic acid as a possible antibacterial substance produced by Enterococcus faecalis TH10. Biocontrol Sci. 2004;9:77–81. doi: 10.4265/bio.9.77.
    1. Valerio F., Lavermicocca P., Pascale M., Visconti A. Production of phenyllactic acid by lactic acid bacteria: An approach to the selection of strains contributing to food quality and preservation. FEMS Microbiol. Lett. 2004;233:289–295. doi: 10.1111/j.1574-6968.2004.tb09494.x.
    1. Makras L., Triantafyllou V., Fayol-Messaoudi D., Adriany T., Zoumpopoulou G., Tsakalidou E., Servin A., de Vuyst L. Kinetic analysis of the antibacterial activity of probiotic lactobacilli towards Salmonella enterica serovar Typhimurium reveals a role for lactic acid and other inhibitory compounds. Res. Microbiol. 2006;157:241–247. doi: 10.1016/j.resmic.2005.09.002.
    1. Magnusson J., Strom K., Roos S., Sjogren J., Schnurer J. Broad and complex antifungal activity among environmental isolates of lactic acid bacteria. FEMS Microbiol. Lett. 2003;219:129–135. doi: 10.1016/S0378-1097(02)01207-7.
    1. Li X., Jiang B., Pan B., Mu W., Zhang T. Purification and partial characterization of Lactobacillus species SK007 lactate dehydrogenase (LDH) catalyzing phenylpyruvic acid (PPA) conversion into phenyllactic acid (PLA) J. Agric. Food Chem. 2008;56:2392–2399. doi: 10.1021/jf0731503.
    1. Vermeulen N., Ganzle M.G., Vogel R.F. Influence of peptide supply and cosubstrates on phenylalanine metabolism of Lactobacillus sanfranciscensis DSM20451(T) and Lactobacillus plantarum TMW1.468. J. Agric. Food Chem. 2006;54:3832–3839. doi: 10.1021/jf052733e.
    1. Armaforte E., Carri S., Ferri G., Caboni M.F. High-performance liquid chromatography determination of phenyllactic acid in MRS broth. J. Chromatogr. A. 2006;1131:281–284. doi: 10.1016/j.chroma.2006.07.095.
    1. Martinez A.W., Recht N.S., Hostetter T.H., Meyer T.W. Removal of P-cresol sulfate by hemodialysis. J. Am. Soc. Nephrol. 2005;16:3430–3436. doi: 10.1681/ASN.2005030310.
    1. Schepers E., Meert N., Glorieux G., Goeman J., Van der Eycken J., Vanholder R. P-cresylsulphate, the main in vivo metabolite of p-cresol, activates leucocyte free radical production. Nephrol. Dial. Transplant. 2007;22:592–596. doi: 10.1093/ndt/gfl584.
    1. Pletinck A., Glorieux G., Schepers E., Cohen G., Gondouin B., Van Landschoot M., Eloot S., Rops A., Van de Voorde J., de Vriese A., et al. Protein-bound uremic toxins stimulate crosstalk between leukocytes and vessel wall. J. Am. Soc. Nephrol. 2013;24:1981–1994. doi: 10.1681/ASN.2012030281.
    1. Meijers B.K., Van Kerckhoven S., Verbeke K., Dehaen W., Vanrenterghem Y., Hoylaerts M.F., Evenepoel P. The uremic retention solute p-cresyl sulfate and markers of endothelial damage. Am. J. Kidney Dis. 2009;54:891–901. doi: 10.1053/j.ajkd.2009.04.022.
    1. Gross P., Massy Z.A., Henaut L., Boudot C., Cagnard J., March C., Kamel S., Drueke T.B., Six I. Para-cresyl sulfate acutely impairs vascular reactivity and induces vascular remodeling. J. Cell. Physiol. 2015;230:2927–2935. doi: 10.1002/jcp.25018.
    1. Han H., Zhu J., Zhu Z., Ni J., Du R., Dai Y., Chen Y., Wu Z., Lu L., Zhang R. p-Cresyl sulfate aggravates cardiac dysfunction associated with chronic kidney disease by enhancing apoptosis of cardiomyocytes. J. Am. Heart Assoc. 2015;4:e001852. doi: 10.1161/JAHA.115.001852.
    1. Sun C.Y., Chang S.C., Wu M.S. Suppression of Klotho expression by protein-bound uremic toxins is associated with increased DNA methyltransferase expression and DNA hypermethylation. Kidney Int. 2012;81:640–650. doi: 10.1038/ki.2011.445.
    1. Sun C.Y., Chang S.C., Wu M.S. Uremic toxins induce kidney fibrosis by activating intrarenal renin-angiotensin-aldosterone system associated epithelial-to-mesenchymal transition. PLoS ONE. 2012;7:e34026. doi: 10.1371/journal.pone.0034026.
    1. Watanabe H., Miyamoto Y., Honda D., Tanaka H., Wu Q., Endo M., Noguchi T., Kadowaki D., Ishima Y., Kotani S., et al. p-Cresyl sulfate causes renal tubular cell damage by inducing oxidative stress by activation of NADPH oxidase. Kidney Int. 2013;83:582–592. doi: 10.1038/ki.2012.448.
    1. Poveda J., Sanchez-Nino M.D., Glorieux G., Sanz A.B., Egido J., Vanholder R., Ortiz A. p-Cresyl sulphate has pro-inflammatory and cytotoxic actions on human proximal tubular epithelial cells. Nephrol. Dial. Transplant. 2014;29:56–64. doi: 10.1093/ndt/gft367.
    1. Mutsaers H.A., Caetano-Pinto P., Seegers A.E., Dankers A.C., van den Broek P.H., Wetzels J.F., van den Brand J.A., van den Heuvel L.P., Hoenderop J.G., Wilmer M.J., et al. Proximal tubular efflux transporters involved in renal excretion of p-cresyl sulfate and p-cresyl glucuronide: Implications for chronic kidney disease pathophysiology. Toxicol. In Vitro. 2015;29:1868–1877. doi: 10.1016/j.tiv.2015.07.020.
    1. Koppe L., Pillon N.J., Vella R.E., Croze M.L., Pelletier C.C., Chambert S., Massy Z., Glorieux G., Vanholder R., Dugenet Y., et al. p-Cresyl sulfate promotes insulin resistance associated with CKD. J. Am. Soc. Nephrol. 2013;24:88–99. doi: 10.1681/ASN.2012050503.
    1. Shiba T., Kawakami K., Sasaki T., Makino I., Kato I., Kobayashi T., Uchida K., Kaneko K. Effects of intestinal bacteria-derived p-cresyl sulfate on Th1-type immune response in vivo and in vitro. Toxicol. Appl. Pharmacol. 2014;274:191–199. doi: 10.1016/j.taap.2013.10.016.
    1. Shiba T., Makino I., Kawakami K., Kato I., Kobayashi T., Kaneko K. p-Cresyl sulfate suppresses lipopolysaccharide-induced anti-bacterial immune responses in murine macrophages in vitro. Toxicol. Lett. 2016;245:24–30. doi: 10.1016/j.toxlet.2016.01.009.
    1. Chen T.C., Wang C.Y., Hsu C.Y., Wu C.H., Kuo C.C., Wang K.C., Yang C.C., Wu M.T., Chuang F.R., Lee C.T. Free p-cresol sulfate is associated with survival and function of vascular access in chronic hemodialysis patients. Kidney Blood Press. Res. 2012;35:583–588. doi: 10.1159/000339709.
    1. Chiu C.A., Lu L.F., Yu T.H., Hung W.C., Chung F.M., Tsai I.T., Yang C.Y., Hsu C.C., Lu Y.C., Wang C.P., et al. Increased levels of total p-Cresylsulphate and indoxyl sulphate are associated with coronary artery disease in patients with diabetic nephropathy. Rev. Diabet. Stud. 2010;7:275–284. doi: 10.1900/RDS.2010.7.275.
    1. Hsu H.J., Yen C.H., Wu I.W., Hsu K.H., Chen C.K., Sun C.Y., Chou C.C., Chen C.Y., Tsai C.J., Wu M.S., et al. The association of uremic toxins and inflammation in hemodialysis patients. PLoS ONE. 2014;9:e102691. doi: 10.1371/journal.pone.0102691.
    1. Lin C.J., Pan C.F., Liu H.L., Chuang C.K., Jayakumar T., Wang T.J., Chen H.H., Wu C.J. The role of protein-bound uremic toxins on peripheral artery disease and vascular access failure in patients on hemodialysis. Atherosclerosis. 2012;225:173–179. doi: 10.1016/j.atherosclerosis.2012.07.012.
    1. Lin C.J., Pan C.F., Chuang C.K., Sun F.J., Wang D.J., Chen H.H., Liu H.L., Wu C.J. p-cresyl sulfate is a valuable predictor of clinical outcomes in pre-ESRD patients. BioMed Res. Int. 2014;2014:526932. doi: 10.1155/2014/526932.
    1. Lu L.F., Tang W.H., Hsu C.C., Tsai I.T., Hung W.C., Yu T.H., Wu C.C., Chung F.M., Lu Y.C., Lee Y.J., et al. Associations among chronic kidney disease, high total p-cresylsulfate and left ventricular systolic dysfunction. Clin. Chim. Acta. 2016;457:63–68. doi: 10.1016/j.cca.2016.03.012.
    1. Rossi M., Campbell K.L., Johnson D.W., Stanton T., Vesey D.A., Coombes J.S., Weston K.S., Hawley C.M., McWhinney B.C., Ungerer J.P., et al. Protein-bound uremic toxins, inflammation and oxidative stress: A cross-sectional study in stage 3–4 chronic kidney disease. Arch. Med. Res. 2014;45:309–317. doi: 10.1016/j.arcmed.2014.04.002.
    1. Shafi T., Meyer T.W., Hostetter T.H., Melamed M.L., Parekh R.S., Hwang S., Banerjee T., Coresh J., Powe N.R. Free levels of selected organic solutes and cardiovascular morbidity and mortality in hemodialysis patients: Results from the Retained Organic Solutes and Clinical Outcomes (ROSCO) investigators. PLoS ONE. 2015;10:e0126048. doi: 10.1371/journal.pone.0126048.
    1. Tang W.H., Wang C.P., Yu T.H., Hung W.C., Chung F.M., Lu Y.C., Hsu C.C., Lu L.F., Huang L.L., Lee Y.J., et al. Serum total p-cresylsulfate level is associated with abnormal QTc interval in stable angina patients with early stage of renal failure. Clin. Chim. Acta. 2014;437:25–30. doi: 10.1016/j.cca.2014.07.002.
    1. Wang C.P., Lu L.F., Yu T.H., Hung W.C., Chiu C.A., Chung F.M., Yeh L.R., Chen H.J., Lee Y.J., Houng J.Y. Serum levels of total p-cresylsulphate are associated with angiographic coronary atherosclerosis severity in stable angina patients with early stage of renal failure. Atherosclerosis. 2010;211:579–583. doi: 10.1016/j.atherosclerosis.2010.03.036.
    1. Wang C.P., Lu L.F., Yu T.H., Hung W.C., Chiu C.A., Chung F.M., Hsu C.C., Lu Y.C., Lee Y.J., Houng J.Y. Associations among chronic kidney disease, high total p-cresylsulfate and major adverse cardiac events. J. Nephrol. 2013;26:111–118. doi: 10.5301/jn.5000111.
    1. Wu I.W., Hsu K.H., Hsu H.J., Lee C.C., Sun C.Y., Tsai C.J., Wu M.S. Serum free p-cresyl sulfate levels predict cardiovascular and all-cause mortality in elderly hemodialysis patients—a prospective cohort study. Nephrol. Dial. Transplant. 2012;27:1169–1175. doi: 10.1093/ndt/gfr453.
    1. Poesen R., Viaene L., Verbeke K., Augustijns P., Bammens B., Claes K., Kuypers D., Evenepoel P., Meijers B. Cardiovascular disease relates to intestinal uptake of p-cresol in patients with chronic kidney disease. BMC Nephrol. 2014;15:87. doi: 10.1186/1471-2369-15-87.
    1. Lin C.J., Wu V., Wu P.C., Wu C.J. Meta-Analysis of the Associations of p-Cresyl Sulfate (PCS) and Indoxyl Sulfate (IS) with Cardiovascular Events and All-Cause Mortality in Patients with Chronic Renal Failure. PLoS ONE. 2015;10:e0132589. doi: 10.1371/journal.pone.0132589.
    1. Wu I.W., Hsu K.H., Lee C.C., Sun C.Y., Hsu H.J., Tsai C.J., Tzen C.Y., Wang Y.C., Lin C.Y., Wu M.S. p-Cresyl sulphate and indoxyl sulphate predict progression of chronic kidney disease. Nephrol. Dial. Transplant. 2011;26:938–947. doi: 10.1093/ndt/gfq580.
    1. Wang C.P., Lu Y.C., Tsai I.T., Tang W.H., Hsu C.C., Hung W.C., Yu T.H., Chen S.C., Chung F.M., Lee Y.J., et al. Increased Levels of Total p-Cresylsulfate Are Associated with Pruritus in Patients with Chronic Kidney Disease. Dermatology. 2016;232:363–370. doi: 10.1159/000445429.
    1. Liabeuf S., Barreto D.V., Barreto F.C., Meert N., Glorieux G., Schepers E., Temmar M., Choukroun G., Vanholder R., Massy Z.A. Free p-cresylsulphate is a predictor of mortality in patients at different stages of chronic kidney disease. Nephrol. Dial. Transplant. 2010;25:1183–1191. doi: 10.1093/ndt/gfp592.
    1. Montemurno E., Cosola C., Dalfino G., Daidone G., de Angelis M., Gobbetti M., Gesualdo L. What would you like to eat, Mr CKD Microbiota? A Mediterranean Diet, please! Kidney Blood Press. Res. 2014;39:114–123. doi: 10.1159/000355785.
    1. Walker A.W., Ince J., Duncan S.H., Webster L.M., Holtrop G., Ze X., Brown D., Stares M.D., Scott P., Bergerat A., et al. Dominant and diet-responsive groups of bacteria within the human colonic microbiota. ISME J. 2011;5:220–230. doi: 10.1038/ismej.2010.118.
    1. Ling W.H., Hanninen O. Shifting from a conventional diet to an uncooked vegan diet reversibly alters fecal hydrolytic activities in humans. J. Nutr. 1992;122:924–930.
    1. Patel K.P., Luo F.J., Plummer N.S., Hostetter T.H., Meyer T.W. The production of p-cresol sulfate and indoxyl sulfate in vegetarians versus omnivores. Clin. J. Am. Soc. Nephrol. 2012;7:982–988. doi: 10.2215/CJN.12491211.
    1. Cummings J.H., Hill M.J., Bone E.S., Branch W.J., Jenkins D.J. The effect of meat protein and dietary fiber on colonic function and metabolism. II. Bacterial metabolites in feces and urine. Am. J. Clin. Nutr. 1979;32:2094–2101.
    1. Birkett A., Muir J., Phillips J., Jones G., O’Dea K. Resistant starch lowers fecal concentrations of ammonia and phenols in humans. Am. J. Clin. Nutr. 1996;63:766–772.
    1. Rossi M., Johnson D.W., Xu H., Carrero J.J., Pascoe E., French C., Campbell K.L. Dietary protein-fiber ratio associates with circulating levels of indoxyl sulfate and p-cresyl sulfate in chronic kidney disease patients. Nutr. Metab. Cardiovasc. Dis. 2015;25:860–865. doi: 10.1016/j.numecd.2015.03.015.
    1. Sirich T.L., Plummer N.S., Gardner C.D., Hostetter T.H., Meyer T.W. Effect of increasing dietary fiber on plasma levels of colon-derived solutes in hemodialysis patients. Clin. J. Am. Soc. Nephrol. 2014;9:1603–1610. doi: 10.2215/CJN.00490114.
    1. Kieffer D.A., Piccolo B.D., Vaziri N.D., Liu S., Lau W.L., Khazaeli M., Nazertehrani S., Moore M.E., Marco M.L., Martin R.J., et al. Resistant starch alters gut microbiome and metabolomic profiles concurrent with amelioration of chronic kidney disease in rats. Am. J. Physiol. Ren. Physiol. 2016;310:F857–F871. doi: 10.1152/ajprenal.00513.2015.
    1. Naylor H.L., Jackson H., Walker G.H., Macafee S., Magee K., Hooper L., Stewart L., MacLaughlin H.L. British Dietetic Association evidence-based guidelines for the protein requirements of adults undergoing maintenance haemodialysis or peritoneal dialysis. J. Hum. Nutr. Diet. 2013;26:315–328. doi: 10.1111/jhn.12052.
    1. Jiang Z., Zhang X., Yang L., Li Z., Qin W. Effect of restricted protein diet supplemented with keto analogues in chronic kidney disease: A systematic review and meta-analysis. Int. Urol. Nephrol. 2016;48:409–418. doi: 10.1007/s11255-015-1170-2.
    1. David C., Peride I., Niculae A., Constantin A.M., Checherita I.A. Very low protein diets supplemented with keto-analogues in ESRD predialysis patients and its effect on vascular stiffness and AVF Maturation. BMC Nephrol. 2016;17:131. doi: 10.1186/s12882-016-0347-y.
    1. De Preter V., Vanhoutte T., Huys G., Swings J., de Vuyst L., Rutgeerts P., Verbeke K. Effects of Lactobacillus casei Shirota, Bifidobacterium breve, and oligofructose-enriched inulin on colonic nitrogen-protein metabolism in healthy humans. Am. J. Physiol. Gastrointest. Liver Physiol. 2007;292:G358–G368. doi: 10.1152/ajpgi.00052.2006.
    1. Rossi M., Klein K., Johnson D.W., Campbell K.L. Pre-, pro-, and synbiotics: Do they have a role in reducing uremic toxins? A systematic review and meta-analysis. Int. J. Nephrol. 2012;673631 doi: 10.1155/2012/673631.
    1. Tohyama K., Kobayashi Y., Kan T., Yazawa K., Terashima T., Mutai M. Effect of lactobacilli on urinary indican excretion in gnotobiotic rats and in man. Microbiol. Immunol. 1981;25:101–112. doi: 10.1111/j.1348-0421.1981.tb00014.x.
    1. Fujiwara S., Seto Y., Kimura A., Hashiba H. Establishment of orally-administered Lactobacillus gasseri SBT2055SR in the gastrointestinal tract of humans and its influence on intestinal microflora and metabolism. J. Appl. Microbiol. 2001;90:343–352. doi: 10.1046/j.1365-2672.2001.01251.x.
    1. de Preter V., Vanhoutte T., Huys G., Swings J., Rutgeerts P., Verbeke K. Baseline microbiota activity and initial bifidobacteria counts influence responses to prebiotic dosing in healthy subjects. Aliment. Pharmacol. Ther. 2008;27:504–513. doi: 10.1111/j.1365-2036.2007.03588.x.
    1. Davis L.M., Martinez I., Walter J., Goin C., Hutkins R.W. Barcoded pyrosequencing reveals that consumption of galactooligosaccharides results in a highly specific bifidogenic response in humans. PLoS ONE. 2011;6:e25200. doi: 10.1371/journal.pone.0025200.
    1. Meijers B.K., de Preter V., Verbeke K., Vanrenterghem Y., Evenepoel P. p-Cresyl sulfate serum concentrations in haemodialysis patients are reduced by the prebiotic oligofructose-enriched inulin. Nephrol. Dial. Transplant. 2010;25:219–224. doi: 10.1093/ndt/gfp414.
    1. Evenepoel P., Bammens B., Verbeke K., Vanrenterghem Y. Acarbose treatment lowers generation and serum concentrations of the protein-bound solute p-cresol: A pilot study. Kidney Int. 2006;70:192–198. doi: 10.1038/sj.ki.5001523.
    1. Poesen R., Evenepoel P., de Loor H., Delcour J.A., Courtin C.M., Kuypers D., Augustijns P., Verbeke K., Meijers B. The influence of prebiotic arabinoxylan oligosaccharides on microbiota derived uremic retention solutes in patients with chronic kidney disease: A randomized controlled trial. PLoS ONE. 2016;11:e0153893. doi: 10.1371/journal.pone.0153893.
    1. Nakabayashi I., Nakamura M., Kawakami K., Ohta T., Kato I., Uchida K., Yoshida M. Effects of synbiotic treatment on serum level of p-cresol in haemodialysis patients: A preliminary study. Nephrol. Dial. Transplant. 2011;26:1094–1098. doi: 10.1093/ndt/gfq624.
    1. Guida B., Germano R., Trio R., Russo D., Memoli B., Grumetto L., Barbato F., Cataldi M. Effect of short-term synbiotic treatment on plasma p-cresol levels in patients with chronic renal failure: A randomized clinical trial. Nutr. Metab. Cardiovasc. Dis. 2014;24:1043–1049. doi: 10.1016/j.numecd.2014.04.007.
    1. Rossi M., Johnson D.W., Morrison M., Pascoe E.M., Coombes J.S., Forbes J.M., Szeto C.C., McWhinney B.C., Ungerer J.P., Campbell K.L. Synbiotics Easing Renal failure by improving Gut microbiology (SYNERGY): A randomized trial. Clin. J. Am. Soc. Nephrol. 2016;11:223–231. doi: 10.2215/CJN.05240515.
    1. Ling W.H., Korpela R., Mykkanen H., Salminen S., Hanninen O. Lactobacillus strain GG supplementation decreases colonic hydrolytic and reductive enzyme activities in healthy female adults. J. Nutr. 1994;124:18–23.
    1. Hyun H.S., Paik K.H., Cho H.Y. p-Cresyl sulfate and indoxyl sulfate in pediatric patients on chronic dialysis. Korean J. Pediatr. 2013;56:159–164. doi: 10.3345/kjp.2013.56.4.159.
    1. Wu M.J., Chang C.S., Cheng C.H., Chen C.H., Lee W.C., Hsu Y.H., Shu K.H., Tang M.J. Colonic transit time in long-term dialysis patients. Am. J. Kidney Dis. 2004;44:322–327. doi: 10.1053/j.ajkd.2004.04.048.
    1. Yasuda G., Shibata K., Takizawa T., Ikeda Y., Tokita Y., Umemura S., Tochikubo O. Prevalence of constipation in continuous ambulatory peritoneal dialysis patients and comparison with hemodialysis patients. Am. J. Kidney Dis. 2002;39:1292–1299. doi: 10.1053/ajkd.2002.33407.
    1. Kikuchi K., Itoh Y., Tateoka R., Ezawa A., Murakami K., Niwa T. Metabolomic search for uremic toxins as indicators of the effect of an oral sorbent AST-120 by liquid chromatography/tandem mass spectrometry. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2010;878:2997–3002. doi: 10.1016/j.jchromb.2010.09.006.
    1. Niwa T., Ise M., Miyazaki T., Meada K. Suppressive effect of an oral sorbent on the accumulation of p-cresol in the serum of experimental uremic rats. Nephron. 1993;65:82–87. doi: 10.1159/000187446.
    1. Velenosi T.J., Hennop A., Feere D.A., Tieu A., Kucey A.S., Kyriacou P., McCuaig L.E., Nevison S.E., Kerr M.A., Urquhart B.L. Untargeted plasma and tissue metabolomics in rats with chronic kidney disease given AST-120. Sci. Rep. 2016;6:22526. doi: 10.1038/srep22526.
    1. Yamamoto S., Kazama J.J., Omori K., Matsuo K., Takahashi Y., Kawamura K., Matsuto T., Watanabe H., Maruyama T., Narita I. Continuous reduction of protein-bound uraemic toxins with improved oxidative stress by using the oral charcoal adsorbent AST-120 in haemodialysis patients. Sci. Rep. 2015;5:14381. doi: 10.1038/srep14381.
    1. Lee C.T., Hsu C.Y., Tain Y.L., Ng H.Y., Cheng B.C., Yang C.C., Wu C.H., Chiou T.T., Lee Y.T., Liao S.C. Effects of AST-120 on blood concentrations of protein-bound uremic toxins and biomarkers of cardiovascular risk in chronic dialysis patients. Blood Purif. 2014;37:76–83. doi: 10.1159/000357641.
    1. Schulman G., Berl T., Beck G.J., Remuzzi G., Ritz E., Arita K., Kato A., Shimizu M. Randomized placebo-controlled EPPIC trials of AST-120 in CKD. J. Am. Soc. Nephrol. 2015;26:1732–1746. doi: 10.1681/ASN.2014010042.
    1. Cha R.H., Kang S.W., Park C.W., Cha D.R., Na K.Y., Kim S.G., Yoon S.A., Han S.Y., Chang J.H., Park S.K., et al. A Randomized, Controlled trial of oral intestinal sorbent AST-120 on Renal function deterioration in patients with advanced renal dysfunction. Clin. J. Am. Soc. Nephrol. 2016;11:559–567. doi: 10.2215/CJN.12011214.
    1. Meert N., Eloot S., Schepers E., Lemke H.D., Dhondt A., Glorieux G., Van Landschoot M., Waterloos M.A., Vanholder R. Comparison of removal capacity of two consecutive generations of high-flux dialysers during different treatment modalities. Nephrol. Dial. Transplant. 2011;26:2624–2630. doi: 10.1093/ndt/gfq803.
    1. Meert N., Eloot S., Waterloos M.A., Van Landschoot M., Dhondt A., Glorieux G., Ledebo I., Vanholder R. Effective removal of protein-bound uraemic solutes by different convective strategies: A prospective trial. Nephrol. Dial. Transplant. 2009;24:562–570. doi: 10.1093/ndt/gfn522.
    1. Krieter D.H., Hackl A., Rodriguez A., Chenine L., Moragues H.L., Lemke H.D., Wanner C., Canaud B. Protein-bound uraemic toxin removal in haemodialysis and post-dilution haemodiafiltration. Nephrol. Dial. Transplant. 2010;25:212–218. doi: 10.1093/ndt/gfp437.
    1. Meert N., Waterloos M.A., Van Landschoot M., Dhondt A., Ledebo I., Glorieux G., Goeman J., Van der Eycken J., Vanholder R. Prospective evaluation of the change of predialysis protein-bound uremic solute concentration with postdilution online hemodiafiltration. Artif. Organs. 2010;34:580–585. doi: 10.1111/j.1525-1594.2010.01005.x.
    1. Sirich T.L., Luo F.J., Plummer N.S., Hostetter T.H., Meyer T.W. Selectively increasing the clearance of protein-bound uremic solutes. Nephrol. Dial. Transplant. 2012;27:1574–1579. doi: 10.1093/ndt/gfr691.
    1. Marquez I.O., Tambra S., Luo F.Y., Li Y., Plummer N.S., Hostetter T.H., Meyer T.W. Contribution of residual function to removal of protein-bound solutes in hemodialysis. Clin. J. Am. Soc. Nephrol. 2011;6:290–296. doi: 10.2215/CJN.06100710.
    1. Meijers B.K., Weber V., Bammens B., Dehaen W., Verbeke K., Falkenhagen D., Evenepoel P. Removal of the uremic retention solute p-cresol using fractionated plasma separation and adsorption. Artif. Organs. 2008;32:214–219. doi: 10.1111/j.1525-1594.2007.00525.x.
    1. Brettschneider F., Tolle M., von der Giet M., Passlick-Deetjen J., Steppan S., Peter M., Jankowski V., Krause A., Kuhne S., Zidek W., et al. Removal of protein-bound, hydrophobic uremic toxins by a combined fractionated plasma separation and adsorption technique. Artif. Organs. 2013;37:409–416. doi: 10.1111/j.1525-1594.2012.01570.x.
    1. Kruse A., Tao X., Bhalani V., Handelman G., Levin N.W., Kotanko P., Thijssen S. Clearance of p-cresol sulfate and β-2-microglobulin from dialysate by commercially available sorbent technology. ASAIO J. 2011;57:219–224. doi: 10.1097/MAT.0b013e3182178c59.
    1. Tijink M.S., Wester M., Glorieux G., Gerritsen K.G., Sun J., Swart P.C., Borneman Z., Wessling M., Vanholder R., Joles J.A., et al. Mixed matrix hollow fiber membranes for removal of protein-bound toxins from human plasma. Biomaterials. 2013;34:7819–7828. doi: 10.1016/j.biomaterials.2013.07.008.
    1. Sandeman S.R., Howell C.A., Phillips G.J., Zheng Y., Standen G., Pletzenauer R., Davenport A., Basnayake K., Boyd O., Holt S., et al. An adsorbent monolith device to augment the removal of uraemic toxins during haemodialysis. J. Mater. Sci. Mater. Med. 2014;25:1589–1597. doi: 10.1007/s10856-014-5173-9.
    1. Tetali S.D., Jankowski V., Luetzow K., Kratz K., Lendlein A., Jankowski J. Adsorption capacity of poly(ether imide) microparticles to uremic toxins. Clin. Hemorheol. Microcirc. 2016;61:657–665. doi: 10.3233/CH-152026.
    1. Bohringer F., Jankowski V., Gajjala P.R., Zidek W., Jankowski J. Release of uremic retention solutes from protein binding by hypertonic predilution hemodiafiltration. ASAIO J. 2015;61:55–60. doi: 10.1097/MAT.0000000000000166.
    1. Krieter D.H., Devine E., Korner T., Ruth M., Wanner C., Raine M., Jankowski J., Lemke H.D. Haemodiafiltration at Increased Plasma Ionic Strength for Improved Protein-Bound Toxin Removal. Acta Physiol. (Oxf.) 2016 doi: 10.1111/apha.12730.
    1. Evenepoel P., Bammens B., Verbeke K., Vanrenterghem Y. Superior dialytic clearance of beta(2)-microglobulin and p-cresol by high-flux hemodialysis as compared to peritoneal dialysis. Kidney Int. 2006;70:794–799. doi: 10.1038/sj.ki.5001640.
    1. Pham N.M., Recht N.S., Hostetter T.H., Meyer T.W. Removal of the protein-bound solutes indican and p-cresol sulfate by peritoneal dialysis. Clin. J. Am. Soc. Nephrol. 2008;3:85–90. doi: 10.2215/CJN.02570607.
    1. Viaene L., Meijers B.K., Bammens B., Vanrenterghem Y., Evenepoel P. Serum concentrations of p-cresyl sulfate and indoxyl sulfate, but not inflammatory markers, increase in incident peritoneal dialysis patients in parallel with loss of residual renal function. Perit. Dial. Int. 2014;34:71–78. doi: 10.3747/pdi.2012.00276.
    1. Huang W.H., Hung C.C., Yang C.W., Huang J.Y. High correlation between clearance of renal protein-bound uremic toxins (indoxyl sulfate and p-cresyl sulfate) and renal water-soluble toxins in peritoneal dialysis patients. Ther. Apher. Dial. 2012;16:361–367. doi: 10.1111/j.1744-9987.2012.01068.x.
    1. Vanholder R., Meert N., Van Biesen W., Meyer T., Hostetter T., Dhondt A., Eloot S. Why do patients on peritoneal dialysis have low blood levels of protein-bound solutes? Nat. Clin. Pract. Nephrol. 2009;5:130–131. doi: 10.1038/ncpneph1023.
    1. Eloot S., Van Biesen W., Glorieux G., Neirynck N., Dhondt A., Vanholder R. Does the adequacy parameter Kt/V(urea) reflect uremic toxin concentrations in hemodialysis patients? PLoS ONE. 2013;8:e76838. doi: 10.1371/journal.pone.0076838.
    1. Liabeuf S., Desjardins L., Massy Z.A., Brazier F., Westeel P.F., Mazouz H., Titeca-Beauport D., Diouf M., Glorieux G., Vanholder R., et al. Levels of indoxyl sulfate in kidney transplant patients, and the relationship with hard outcomes. Circ. J. 2016;80:722–730. doi: 10.1253/circj.CJ-15-0949.
    1. Poesen R., Evenepoel P., de Loor H., Bammens B., Claes K., Sprangers B., Naesens M., Kuypers D., Augustijns P., Meijers B. The influence of renal transplantation on retained microbial-human co-metabolites. Nephrol. Dial. Transplant. 2016;31:1721–1729. doi: 10.1093/ndt/gfw009.
    1. Vanholder R., Glorieux G., Massy Z.A. Intestinal metabolites, chronic kidney disease and renal transplantation: Enigma Variations? Nephrol. Dial. Transplant. 2016;31:1547–1551. doi: 10.1093/ndt/gfw040.

Source: PubMed

3
購読する