Hostname: page-component-7c8c6479df-995ml Total loading time: 0 Render date: 2024-03-28T08:45:03.674Z Has data issue: false hasContentIssue false

Colonic metabolites of berry polyphenols: the missing link to biological activity?

Published online by Cambridge University Press:  19 October 2010

Gary Williamson*
Affiliation:
School of Food Science and Nutrition, University of Leeds, LeedsLS2 9JT, UK
Michael N. Clifford
Affiliation:
Division of Nutritional Sciences, Faculty of Health and Medical Sciences, University of Surrey, Guildford, SurreyGU2 7XH, UK
*
*Corresponding author: G. Williamson, fax +44 113 343 2982, email g.williamson@leeds.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

The absorption of dietary phenols, polyphenols and tannins (PPT) is an essential step for biological activity and effects on health. Although a proportion of these dietary bioactive compounds are absorbed intact, depending on their chemical structure and the nature of any attached moiety (e.g. sugar, organic acid), substantial amounts of lower molecular weight catabolites are absorbed after biotransformation by the colon microflora. The main products in the colon are (a) benzoic acids (C6–C1), especially benzoic acid and protocatechuic acid; (b) phenylacetic acids (C6–C2), especially phenylacetic acid per se; (c) phenylpropionic acids (C6–C3), where the latter are almost entirely in the dihydro form, notably dihydrocaffeic acid, dihydroferulic acid, phenylpropionic acid and 3-(3′-hydroxyphenyl)-propionic acid. As a result of this biotransformation, some of these compounds can each reach mm concentrations in faecal water. Many of these catabolites are efficiently absorbed in the colon, appear in the blood and are ultimately excreted in the urine. In the case of certain polyphenols, such as anthocyanins, these catabolites are major products in vivo; protocatechuic acid is reported to represent a substantial amount of the ingested dose of cyanidin-3-O-glucoside. The major catabolites of berries, and especially blackcurrants, are predicted based on compositional data for polyphenols from berries and other sources. Since microbial catabolites may be present at many sites of the body in higher concentration than the parent compound, it is proposed that at least a part of the biological activities ascribed to berry polyphenols and other PPT are due to their colonic catabolites.

Type
Full Papers
Copyright
Copyright © The Authors 2010

Dietary polyphenols are found in fruits and vegetables, and in products derived from plants such as fruit beverages, tea, coffee and dark chocolate. Epidemiological evidence shows that consumption of polyphenol-rich foods reduces the risk of CVD and associated conditions(Reference Hooper, Kroon and Rimm1), and human intervention studies have supported this association(Reference Williamson and Manach2). The absorption and metabolism of flavonoids is quite well described for some compounds, such as quercetin and (–)-epicatechin(Reference Scholz and Williamson3). However, although the absorption of certain other phenols, polyphenols and tannins (PPT) such as procyanidins, chlorogenic acids and anthocyanins has been described in the literature, the levels of the parent compounds in blood after a high dose, or a large amount of PPT-rich food, are very low compared with other flavonoids(Reference Vitaglione, Donnarumma and Napolitano4, Reference Nurmi, Mursu and Heinonen5). In contrast, intervention studies show that procyanidin-rich, chlorogenic acid-rich or anthocyanin-rich foods affect certain biomarkers, but because the concentration of parent compounds to be expected in blood after consumption is too low to affect these biomarkers(Reference Jensen, Wu and Patterson6), other bioactive substances such as metabolites may be responsible and must be identified. It is proposed that these ‘missing’ components are the colonic catabolites, and that they are potentially important compounds mediating some of the biological activities and health benefits of polyphenol-rich foods.

In excess of 8000 phenolic compounds have been reported, and they are widely dispersed throughout the plant kingdom. The nature of those occurring in foods and beverages has been reviewed(Reference Crozier, Jaganath and Clifford7, Reference Crozier, Clifford and Ashihara8). Briefly, they can be subdivided into flavonoids (anthocyanins, chalcones and dihydrochalcones, flavanols or catechins, flavanones, flavones, flavonols, isoflavones and proanthocyanidins or condensed tannins) and non-flavonoids (benzoic acids, cinnamic acids and cinnamic acid conjugates, such as the chlorogenic acids, and gallic acid esters or hydrolysable tannins). To these must be added the ‘derived polyphenols’, substances that are characteristic of many traditionally processed foods and beverages (coffee, black tea, matured red wine, etc.) but which do not occur in healthy intact plant tissue. Collectively, these three major groups will in this review be referred to as PPT. In addition, this review will discuss the aromatic and phenolic acids produced from these PPT by the gut microflora, many of which do not occur preformed in the diet. These acids can be classified by the number of carbons in the side chain (one to five), and whether the side chain is saturated or unsaturated, and whether it carries an aliphatic hydroxyl. Specimen structures are shown in Fig. 1.

Fig. 1 Structures of selected phenols, polyphenols and tannins, and selected gut microflora catabolites. A and B indicate the A- and B-ring, respectively, of a typical flavonoid. Typical structures are shown for flavanols (catechins) (I); flavanones (II); flavones (III); dihydrochalcones (IV); chlorogenic acids (cinnamate conjugates) (V); cinnamic acids (C6–C3) (VI); anthocyanidins (VII); phenylvaleric acids (C6–C5-γ-OH) (VII); phenylpropionic acids (C6–C3) (IX. XIII); phenylacetic acids (C6–C2) (X, XIV); flavonols (XI); benzoic acids (C6–C1) (XII, XV). Note: it is not possible to show all possible substrates, intermediates and pathways but those shown are representative.

Polyphenol breakdown in the colon

Bacterial composition of the human gut

The first part of this review is concerned primarily with the capability of the microflora in the distal gastro-intestinal tract (GIT) to transform PPT provided by the diet. A detailed discussion of the complex nature and composition of this microflora is beyond the scope of this review, but some basic statistics are instructive. The GIT is home to between some 10 and 100 trillion microbes(Reference Bäckhed, Ley and Sonnenburg9), approximately 10-times as many cells as found in the human body(Reference Savage10). The corresponding microbiome is said to contain more than 100-times as many genes as the human genome(Reference Bäckhed, Ley and Sonnenburg9). Each portion of the GIT (mouth, oesophagus, stomach, small intestine and large intestine) has a distinctive flora, but the majority of the organisms are located in the distal GIT, where the concentration of some 500 bacterial species may reach 1011–1012 colony forming units/g, and account for some 35–50 % of the contents(Reference Salminen, Isolauri and Salminen11). The flora contains bacteria, archaea and eukarya, and the bacterial component has been studied the most extensively. Three genera, Bacteroides spp., Clostridium spp. and Eubacterium spp., each account for approximately 30 % of the organisms present(Reference Bäckhed, Ley and Sonnenburg9), but the colon also has a significant content of Fusobacterium spp., Peptostreptococcus spp. and Bifidobacterium spp.(Reference Salminen, Isolauri and Salminen11, Reference Tuohy, Gibson, Crozier, Clifford and Ashihara12) These bacteria are accompanied by a single methanogenic archaeon, Methanobrevibacter smithii (Reference Bäckhed, Ley and Sonnenburg9, Reference Gill, Pop and Deboy13). The eukarya include protozoa, but those found in the GIT seem never to have been studied with reference to PPT catabolism.

It is generally accepted that colonisation of the GIT of neonates starts immediately after birth. The type of delivery (passage through the birth canal v. caesarean section), type of diet (breast v. formula feeding) and environment affect the colonisation pattern(Reference Gronlund, Lehtonen and Eerola14, Reference Harmsen, Wildeboer-Veloo and Raangs15). The composition of the gut flora of a 2-year-old child is essentially the same as that of an adult(Reference Conway, Gibson and Macfarlane16), but the variations in microflora composition associated with age, diet and health status are still not fully determined(Reference Eckburg, Bik and Bernstein17). A study of faecal samples from 230 volunteers, aged 20–50 years, from four European countries has demonstrated marked country–age interactions. The proportions of bifidobacteria were 2- to 3-fold higher in the Italian study population than in any other study group, and this effect was independent of age. Higher proportions of enterobacteria were found in all elderly volunteers independent of the location(Reference Mueller, Saunier and Hanisch18). There is good evidence that comparatively minor changes in the diet of rats can produce significant changes in the catabolites eventually produced from dietary PPT(Reference Phipps, Stewart and Wright19), but the extent to which the human microflora can be modulated by diet is unknown.

Models used to study colonic catabolism

The distal GIT is conveniently viewed as an anaerobic fermentation vessel in which polysaccharides, proteins, lipids and xenobiotics such as PPT can be transformed. Several strategies have been used to investigate these transformations. In vitro studies frequently use either a defined bacterial strain(Reference Wang, Hur and Lee20, Reference Raimondi, Roncaglia and De Lucia21), ileostomy fluid(Reference Knaup, Kahle and Erk22) or a flora from freshly voided human(Reference Aura, Martin-Lopez and O'Leary23Reference Gonthier, Remesy and Scalbert27) or animal faeces(Reference Labib, Hummel and Richling28Reference Forester and Waterhouse30) cultured in a suitable medium to which a defined substrate has been added, and monitor (i) disappearance of substrate, (ii) formation of catabolites or (iii) changes in the microflora(Reference Lee, Jenner and Low31, Reference Tzounis, Vulevic and Kuhnle32). Rarely all three parameters are studied at the same time. Typically controls would include (i) incubations containing test substrate but lacking the microbial culture to investigate purely chemical transformation, (ii) incubations containing test substrate and chloroform-inactivated culture to investigate substrate losses associated with cell binding and transformations associated with enzymes released after lysis of the micro-organisms and (iii) incubations containing basal medium and microbial culture in order to monitor catabolites arising from basal metabolism and to distinguish these from catabolites produced from the test substance. The culture may be static(Reference Gonthier, Remesy and Scalbert27) or dynamic(Reference Gao, Xu and Krul33). The various types of study used to examine microbial transformation of PPTs are shown in Table 1.

Table 1 Studies concerned with the transformation of phenols, polyphenols and tannins (PPT) by gut flora micro-organisms

It is impossible to know how well these in vitro models reflect the human GIT in vivo but studies using gnotobiotic rats that have been associated with a specific GIT micro-organism do not produce the same catabolites as that of the organism incubated in an artificial medium(Reference Peppercorn and Goldman34) and one must assume that the models are imperfect. In particular, micro-organisms that bind to the GIT surface might be very different to those found in the lumen, and it is these ‘unbound’ organisms that might predominate in the voided stool from which the culture is prepared(Reference Eckburg, Bik and Bernstein17, Reference Hooper and Gordon35). The micro-organisms that survive the culturing process will depend on how rigorously oxygen was excluded and the nature of the medium employed. Media for anaerobic micro-organisms are typically rich in protein or polypeptides and these can be incompatible with the recovery of PPT that bind strongly to protein thus preventing proper analysis of the transformations. Studies in the authors’ laboratories suggest that the catabolism of PPT that are not strongly bound to protein is not significantly different from their catabolism in protein-free media, and protein-free media have been used to investigate the catabolism of procyanidins that would bind irreversibly to protein(Reference Stoupi, Williamson and Drynan36). One advantage of these models is the requirement for comparatively small amounts of a scarce and costly pure substrate.

A second approach has been the use of laboratory animals, commonly rodents, but occasionally pigs. The rodents might be either (i) normal laboratory animals with their typical flora(Reference He, Magnuson and Giusti37), (ii) animals treated with antibiotics (typically neomycin) to destroy the flora, (iii) gnotobiotic animals delivered under sterile conditions, and housed in sterile conditions and fed only on γ-irradiated diets(Reference Peppercorn and Goldman34, Reference Scheline and Midtvedt38) or (iv) gnotobiotic animals associated with a human-type flora(Reference Tamura and Saitoh39). The use of such animals has allowed the effects of the rodent flora and human-type flora to be compared, and the effects of the flora and mammalian metabolism to be distinguished. Radio-labelled test substances can be used and full pharmacokinetic studies can be performed with these animals, and the contents (microbial and chemical) can be compared at different sites in the GIT(Reference He, Magnuson and Giusti37). However, gnotobiotic animal facilities, are very expensive to maintain. An extensive review of the in vitro and animal models is available(Reference Rumney and Rowland40). The early studies using these procedures, but focusing on flavonoids, have been reviewed(Reference de Eds, Florkin and Stotz41Reference Berry, Francis and Bollag43).

A third approach is to use volunteers, either free-living or following prescribed diets, and to collect faeces(Reference Knust, Erben and Spiegelhalder44), faecal water(Reference Jenner, Rafter and Halliwell45) and urine for analysis(Reference Vitaglione, Donnarumma and Napolitano4). Normal volunteers can be compared with volunteers who have undergone an ileostomy but who are otherwise healthy enabling metabolism in the small intestine to be distinguished from that in the large intestine. These are discussed in more detail as follows.

Transformation of polyphenols by colonic microflora

The transformations of which the microflora are capable include O- and C-deglycosylation, the hydrolysis of esters and amides, and deglucuronidation of excreted mammalian metabolites. The aglycones are susceptible to aromatic dehydroxylation, demethoxylation and demethylation, and hydrogenation, α-oxidation and β-oxidation of the aliphatic elements generated following rupture of the aromatic ring(Reference Peppercorn and Goldman34, Reference Scheline and Midtvedt38, Reference de Eds, Florkin and Stotz41Reference Berry, Francis and Bollag43, Reference Hattori, Shu and El Sedawy46Reference Kroon, Faulds and Ryden49). Most investigations have focused on phenolic catabolites, i.e. those that retain at least one phenolic hydroxyl, but it is clear that complete dehydroxylation can occur producing aromatic catabolites(Reference Phipps, Stewart and Wright19, Reference Goodwin, Ruthven and Sandler50). Benzoic acid, however, can also be produced in human subjects by GIT microflora aromatisation of quinic acid(Reference Adamson, Bridges and Evans51). It should also be noted that there is a significant, but often unquantified, yield of non-aromatic products, including oxaloacetate, CO2, etc.(Reference Walle, Walle and Halushka52). A key step in the catabolism of flavonoids is the fission of the A-ring and loss of carbons C5 to C8 as oxaloacetate that is ultimately metabolised to CO2(Reference Das and Griffiths53). Similarly, an oral-dosing volunteer study using [4-14C]-quercetin demonstrated that approximately 50 % of the radioactivity was eliminated as CO2(Reference Walle, Walle and Halushka52), but the fate of the A-ring and B-ring was not investigated.

Dehydroxylation of ortho-dihydroxy and ortho-trihydroxy substrates can occur at either a meta or para position, but it is accepted that the hydroxyl at the meta position is removed less readily. Accordingly, ortho-dihydroxy and ortho-trihydroxy substrates yield predominantly meta-hydroxy and meta-dihydroxy catabolites, respectively(Reference Griffiths and Smith54, Reference Smith and Griffiths55). Absence of a para-hydroxyl in the B-ring of flavonoids is considered to slow the degradation by the GIT microflora(Reference Labib, Hummel and Richling28, Reference Griffiths and Smith56). Studies using 2H-labelled substrates and crude human gut flora have established that with 3-(4′-hydroxyphenyl)-propionic acid and 4-hydroxyphenyl-lactic acid, the aliphatic side chain may be moved about the aromatic ring, thus converting a 4-hydroxy substrate to the corresponding 3-hydroxy isomer(Reference Curtius, Mettler and Ettlinger57, Reference Fuchs-Mettler, Curtius and Baerlocher58). The enzymes/micro-organisms responsible have not been characterised and their substrate specificity is unknown. Although it is possible that such an isomerisation might explain the apparent preference for removal of the para-hydroxyl from a 3,4-dihydroxy substrate, it would not explain the tendency for 3,4,5-trihydroxy substrates to be converted to 3,5-dihydroxy products.

Aromatic demethylation by rodent, pig and human GIT microflora has been observed in vitro, for example demethylation of 3′-O-methyl-(+)-catechin(Reference Gott and Griffiths59), 4′-O-methylgenistein (angolensin)(Reference Blaut, Schoefer and Braune60), 6-methoxyapigenin(Reference Labib, Hummel and Richling28) and phenolic acids formed from the B-ring of anthocyanins(Reference Keppler and Humpf24). In contrast, the gut microflora of the rat, mouse and marmoset are unable to degrade 3-O-methyl-(+)-catechin(Reference Hackett and Griffiths61), suggesting that the they cannot attack the C-ring aliphatic O-methyl group. Similarly, there is no evidence for demethylation of this compound when given orally to volunteers(Reference Hackett, Griffiths and Wermeille62).

Comparatively few of the enzymes associated with these transformations have been characterised in microbes associated with the GIT, and the subtleties of substrate specificity, enzyme inhibition, etc., remain largely unknown. At least with regard to polysaccharide catabolism there appears to be few structures that cannot be degraded(Reference Bäckhed, Ley and Sonnenburg9), and this situation might well apply also to PPT.

It has been demonstrated that some strains of the GIT micro-organism Eubacterium oxidoreducens can convert pyrogallol to dihydrophloroglucinol and 3-hydroxy-5-oxohexanoate. The phloroglucinol reductase and pyrogallol–phloroglucinol isomerase have been isolated and characterised(Reference Krumholz and Bryant63Reference Haddock and Ferry66). This isomerase can be viewed as creating a new hydroxyl at C2 in the phloroglucinol (1,3,5-trihydroxybenzene) equivalent to creating a hydroxyl at either C6 or C8 in the flavonoid A-ring. Such a hydroxylation of the flavonoid A-ring has been demonstrated with Pseudomonas spp.(Reference Jeffrey, Jerina and Self67, Reference Jeffrey, Knight and Evans68) but the general relevance of this observation to GIT transformation of PPT is uncertain because these facultative anaerobes are not typical of the GIT microflora.

Eubacterium ramulus has been reported at 4·4 × 10Reference Crozier, Jaganath and Clifford7 to 2·0 × 109 colony forming units/g (n 20) of human faecal dry mass(Reference Simmering, Kleessen and Blaut69), and its growth is stimulated in vivo by dietary flavonoids(Reference Simmering, Pforte and Jacobasch70). Eubacterium spp. are well known for their ability to degrade flavonoids anaerobically, variously possessing deglycosylating activity, the ability to split the C-ring, chalcone isomerase and phloretin hydrolase activity(Reference Braune, Engst and Blaut26, Reference Schneider and Blaut71Reference Herles, Braune and Blaut76). These chalcone isomerase(Reference Herles, Braune and Blaut76) and phloretin hydrolase(Reference Braune, Engst and Blaut26, Reference Schoefer, Braune and Blaut75) enzymes convert flavanones to the isomeric dihydrochalcones and the dihydrochalcones to a C6–C3 acid and phloroglucinol, respectively. The phloroglucinol will be rapidly converted to aliphatic products as described earlier. It is not known whether the retro-chalcone associated with anthocyanins in media with pH>7(Reference Clifford77) are substrates for the phloretin hydrolase.

Eubacterium limosum demethylates isoflavones such as biochanin A, formononetin and glycitein(Reference Hur and Rafii78), and Eubacterium A-44 has a novel arylsulfotransferase but whether mammalian flavonoid sulphates can serve as donors is not known(Reference Kim, Konishi and Kobashi79). Eubacterium sp. SDG-2 and Peptostreptococcus sp. SDG-1 convert seco-isolariciresinol diglucoside to mammalian lignans(Reference Wang, Meselhy and Li80), and this Eubacterium degrades various (3R)- and (3S)-flavanols and their 3-O-gallates to diaryl-propan-2-ols, 3,5-dihydroxyphenylvalerolactone and 3-hydroxyphenyl-valerolactone. Interestingly, this organism can remove the para-hydroxyl from 3R flavanols such as (–)-epicatechin and (–)-catechin but not from the 3S flavanols such as (+)-catechin and (+)-epicatechin(Reference Wang, Meselhy and Li81), and it has been demonstrated in vitro that the conversion of (+)-catechin to C6–C5 and C6–C3 catabolites required its initial conversion to (+)-epicatechin(Reference Tzounis, Vulevic and Kuhnle32).

Some Clostridium spp. can cleave the C-ring of flavonoids, converting flavonols to a phenylacetic acid and presumably phloroglucinol. Activity with flavanones was much weaker and could not be detected with flavanols(Reference Winter, Moore and Dowell82, Reference Winter, Popoff and Grimont83). Some Fusobacterium spp. and Bacteroides spp. possess α-l-rhamnosidase(Reference Jang and Kim84, Reference Bokkenheuser, Shackleton and Winter85), β-d-glucosidase(Reference Kim, Sohng and Kobashi86, Reference Kim, Kim and Park87), β-d-galactosidase(Reference Bokkenheuser, Shackleton and Winter85) and β-d-glucuronidase(Reference Sung, Kang and Yoon88) activities.

The PPT catabolites most frequently reported are aromatic/phenolic acids having 0, 1, 2 or 3 aromatic hydroxyls, 0, 1 or 2 methoxyls, and a side chain of between one and five carbons, which might be unsaturated and might bear an aliphatic hydroxyl. For most flavonoids, the acids produced retain the intact B-ring, but anthocyanins also yield 2,4,6-trihydroxybenzoic acid and the corresponding benzaldehyde (phloroglucinaldehyde) both derived from the A-ring(Reference Keppler and Humpf24, Reference Schneider, Schwiertz and Collins72, Reference Justesen, Arrigoni and Larsen89, Reference Kim, Jung and Sohng90). There is evidence that these phloroglucinol derivatives might form from anthocyanins purely by chemical degradation at the prevailing pH value in the GIT(Reference Kay, Kroon and Cassidy91). Whether of chemical or microbial origin, these aromatic/phenolic acids are common to most PPT substrates so far investigated, although some variation of ring substitution and side chain length are substrate specific as summarised in Table 2. Phenolic acids can be decarboxylated to the corresponding hydrocarbons(Reference Labib, Hummel and Richling28, Reference Peppercorn and Goldman34, Reference Griffiths and Smith56, Reference Justesen, Arrigoni and Larsen89, Reference Peppercorn and Goldman92, Reference Indahl and Scheline93), but the corresponding alcohols seem not to have been reported.

Table 2 The nature of the aromatic and phenolic acids produced by the gut microflora from pure phenols, polyphenols and tannins (PPT) substrates

* C6–C3 βOH (phenylhydracrylic acids) probably arise from mammalian metabolism of a microbial catabolite rather than directly by microbial catabolism.

May occur also as lactones.

In the case of isoflavones the aryl substituent is attached to C2 of the C6–C3 dihydro side chain. For all other PPT, it is attached to C3.

§ C6–C1 acids may form also by a purely chemical mechanism at the prevailing pH value.

Certain larger mass intermediates have been characterised only during the catabolism of flavanols and proanthocyanidins (e.g. diaryl-propan-2-ols)(Reference Groenewoud and Hundt94, Reference Meselhy, Nakamura and Hattori95), and catabolites such as S-equol, diarylbutanes and urolithins(Reference Doyle and Griffiths96Reference Seeram, Aronson and Zhang99) are specific to the isoflavones, lignans and ellagitannins, respectively. Faecal water from free-living volunteers also contains some flavonoid aglycones indicating that aglycone degradation is not necessarily complete(Reference Jenner, Rafter and Halliwell45).

Models to study absorption in the colon compared with the small intestine

Although absorption is traditionally associated with the small intestine, the colon is also very capable of absorbing compounds; normally it is the supply of nutrients that limits absorption in the colon. There are several different models for studying absorption at these different sites. The most commonly used is to give a single bolus dose of a food or compound to volunteers and collect blood and urine over the next 24–48 h period. Compounds which are absorbed in the small intestine normally appear in the plasma within an hour and the maximum concentration of a compound at any time point in a pharmacokinetic study is usually less than 2·5 h, although this depends on how full the stomach is at the start, i.e. whether a meal was given or not. Compounds that appear after 3 h with T max (a time point corresponding to maximum concentration of a compound at any time point in a pharmacokinetic study) >5 h are generally considered to have been absorbed mainly in the colon. The latter situation involves, almost without exception, a chemical or microbial transformation which converts the compound from non-absorbable to absorbable. The microflora can catalyse two basic steps at this stage: removal of a conjugated chemical moiety such as a sugar or organic acid, and breakdown of the PPT itself. These interactions are illustrated by the flavonol quercetin. Conjugation with a glucose moiety, such as quercetin-4′-O-glucoside found in onions, leads to absorption in the small intestine after deglycosylation by the brush border enzyme lactase phloridzin hydrolase(Reference Day, Cañada and Díaz100), and the T max is < 1 h(Reference Hollman, Buijsman and van Gameren101). In contrast, if the quercetin is conjugated with a rhamnose moiety as in rutin (quercetin-3-O-rhamnoglucoside), a sugar which is not a substrate for the brush border enzyme lactase phloridzin hydrolase, then the compound is untouched in the small intestine and reaches the colon intact, where colonic microflora remove the terminal rhamnose and the glucose moieties(Reference Hollman, Buijsman and van Gameren101). The immediate product, quercetin aglycone, is then either absorbed intact and appears in the plasma as methylated, sulfated and/or glucuronidated conjugates, or is broken down by microflora as described previously. The inferior absorption of the quercetin component of rutin is thus due to degradation by microflora rather than less efficient colon absorption per se. This illustrates the dual role of the colon microflora for absorption and catabolism.

A second approach is to use healthy ileostomists, i.e. subjects who do not have a colon, because it has been removed by colectomy as a result of ulcerative colitis, Crohn's disease, familial polyposis or colon cancer complications. The end of the ileum is brought through the abdominal wall to form a stoma, usually on the lower right side of the abdomen. The GIT contents pass out of the body through the stoma and are collected in an individually fitted drainable pouch, which is worn at all times. The ileal fluids are easily accessible, their nature makes them easier to analyse than some other biological fluids, and ileostomists are a good tool for comparing the catabolism of a given PPT presented in different foods or matrices. There are some limitations: the terminal ileum develops a (minimal) gut microflora, only disappearance is measured, and some of the subjects have increased gut permeability, electrolyte imbalance and a lower production of urine. Nevertheless, this model has been used successfully to examine the absorption of quercetin(Reference Walle, Otake and Walle102Reference Olthof, Hollman and Buijsman104), flavanols and hydroxycinnamic acids(Reference Olthof, Hollman and Buijsman104, Reference Auger, Mullen and Hara105), apple polyphenols(Reference Kahle, Kraus and Scheppach106), blueberry polyphenols(Reference Kahle, Kraus and Scheppach107), cider dihydrochalcones(Reference Marks, Mullen and Borges108) and coffee hydroxycinnamates(267).

An alternative approach is to use sections of rat intestine; these are everted and then suspended in tissue culture, where they exhibit functionality for a short time ( < 1 h). These rat everted sacs can then be tested for transport of various substances as required. The difference between absorption of phenolic acids in the small intestine and colon of the rat was examined using this model(Reference Poquet, Clifford and Williamson109). Ascending and descending colon transported dihydrocaffeic and dihydroferulic acids to the serosal (blood) side 1·5- and 3-fold faster than jejunal segments, implying a more efficient absorption in the colon for these catabolites, both major compounds derived from microbial catabolism of hydroxycinnamic acids and some flavonoids (Table 2). In addition, the two colon segments were much more efficient at effluxing glucuronides back to the luminal side(Reference Poquet, Clifford and Williamson109).

The Caco-2 cell model is very commonly used as a model of small intestinal absorption. This cell line is very well characterised(Reference Hayeshi, Hilgendorf and Artursson110Reference Sambuy, De Angelis and Ranaldi112) but, unlike the small intestine and especially the colon, Caco-2 cells do not secrete mucus. The cell line was derived originally from colonic cells but monolayers of the cells differentiate to produce a small intestine-like morphology, with expression of sucrase and maltase, markers of the small intestine. An alternative colon model using combinations of Caco-2 (76 %) and mucus-secreting HT-29 goblet cells (24 %) has been reported(Reference Poquet, Clifford and Williamson109). The presence of mucus in the co-culture system reduced ferulic acid transport by 23 %, and the HT29 cell component alone was responsible for glucuronidation of the supplied ferulic acid. The co-culture system also reduced the hydroxycinnamic acids to the respective dihydrocinnamic acids, and in this system, dihydroferulic acid was more efficiently transferred to the basolateral side than dihydrocaffeic acid, in contrast to the system using rat everted sacs(Reference Poquet, Clifford and Williamson109).

Polyphenol content of blackcurrants and most abundant predicted catabolites from colonic microflora

The PPT content of the diet has been reviewed(Reference Crozier, Jaganath and Clifford7, Reference Lindsay and Clifford113Reference Crozier, Yokota, Jaganath, Crozier, Clifford and Ashihara115). Berries are among the foods richest in polyphenols, many having an especially high content of anthocyanins, and some (e.g. blueberries) being rich also in chlorogenic acids. Blackcurrants contain various anthocyanins at a high-total concentration (approximately 600 mg/100 g with the 3-O-glucosides and 3-O-rutinosides of cyanidin and delphinidin usually dominant), a relatively high level of proanthocyanidins (approximately 140 mg/100 g, consisting of procyanidins and prodelphinidins), some flavonols (approximately 14 mg/100 g) and hydroxycinnamates (approximately 13 mg/100 g), and lower levels of hydroxybenzoic acids (approximately 1·5 mg/100 g) and catechins (approximately 1·2 mg/100 g)(Reference Maatta, Kamal-Eldin and Torronen116, Reference Gu, Kelm and Hammerstone117) (Figs. 2 and 3).

Fig. 2 Content of anthocyanidins and flavanols in blackcurrants (http://www.phenol-explorer.eu/, accessed October 2009).

Fig. 3 Content of flavonols, cinnamates, benzoic and gallic acid derivatives in blackcurrants (http://www.phenol-explorer.eu/, accessed October 2009).

The next section of this review on the GIT transformation of dietary PPT is concerned with the potential for blackcurrants and blackcurrant-derived products to influence human health. Much of the information is derived from studies of similar PPT from other food sources. Table 2 shows that the classes of aromatic/phenolic acids that might be expected to form in the GIT tract is largely independent of the classes of PPT consumed. The dominance in blackcurrants of anthocyanins and proanthocyanidins will result primarily in the production of C6–C1 and C6–C3 dihydro acids, with some C6–C5-γ-OH products to be expected early in the degradation of the proanthocyanidins and an increasing yield of C6–C1 acids at later time points as the side chain is progressively shortened either by the microflora or by the mammalian enzyme systems. A percentage of the cinnamic acids might be absorbed as such but these also are reduced to the C6–C3 dihydro acids: the flavonols will yield C6–C2 acids, and these might also form from α-oxidation of C6–C3 dihydro acids and yield C6–C1 acids by the same mechanism.

The ring (C6) hydroxylation pattern depends primarily on the B-ring hydroxylation and the precise biochemical competence of the gut microflora. As summarised in Table 3, the dominance of cyanidin and delphinidin glycosides suggests that initially protocatechuic and gallic acid are likely to be the dominant C6–C1 acids. However, in human subjects, gallic acid consumed as flavanol-3-O-gallates is rapidly converted to 3-O-methylgallic acid, 4-O-methylgallic acid and 3,4-O-dimethylgallic acid(Reference Hodgson, Morton and Puddey118, Reference Shahrzad and Bitsch119). Although free gallic acid has been detected in human plasma after an oral dose (50 mg)(Reference Shahrzad and Bitsch119), this mammalian metabolism is consistent with the failure to observe gallic acid in plasma after volunteers had consumed delphinidin(Reference Nurmi, Mursu and Heinonen5).

Table 3 The expected B-ring fragments for the common anthocyanidins and their known mammalian metabolites

Evidence for microbial biotransformations from animal and human intervention studies

Anthocyanins

The absorption and catabolism of anthocyanins from blackcurrants in rats and human subjects have been studied(Reference Matsumoto, Inaba and Kishi120). Four anthocyanins, delphinidin-3-O-rutinoside, cyanidin-3-O-rutinoside, delphinidin-3-O-glucoside and cyanidin-3-O-glucoside, were absorbed and excreted as the glycosylated forms in both rats and human subjects. In human subjects, only 0·05 % of the ingested anthocyanin dose was found in urine. Although, unexpectedly, the 3-O-rutinosyl anthocyanins were directly absorbed, the amount found in the urine was very low(Reference Matsumoto, Inaba and Kishi120). Co-administration of anthocyanins with phytic acid enhances plasma concentration significantly by an unknown mechanism(Reference Matsumoto, Ito and Yonekura121). In all studies on healthy subjects consuming anthocyanin-rich food, it is now well established that < 2 % of the total anthocyanin dose is absorbed intact, as estimated from the amount of anthocyanin and conjugates either in urine(Reference Felgines, Talavera and Gonthier122) or in blood(Reference Mullen, Edwards and Serafini123). Using ileostomist subjects, a high proportion of the anthocyanins from blueberry (up to 85 %, depending on the attached sugar moiety) traversed the small intestine unchanged and were found in the ileostomy bags; this amount would therefore reach the colon under physiological conditions and be subject to microbial degradation(Reference Kahle, Kraus and Scheppach107). In vitro studies using human microflora in an anaerobic chamber have indicated that protocatechuic acid is one of the most likely major degradation products from anthocyanins having a 3,4-dihydroxy B-ring(Reference Aura, Martin-Lopez and O'Leary23). In a human intervention study on blood orange juice, the C max in blood for cyanidin-3-O-glucoside was 1·9 nm, whereas the value for protocatechuic acid was approximately 250-fold higher (492 nM). It was calculated that the main product, protocatechuic acid, accounted for up to 70 % of the anthocyanin intake. Only 1·2 % of the anthocyanin dose ultimately appeared in urine, whereas urinary protocatechuic acid represented 28 % of the total anthocyanin dose. The protocatechuic acid appeared to be unconjugated(Reference Vitaglione, Donnarumma and Napolitano4). Protocatechuic acid was also found in rat plasma after feeding cyanidin-3-O-glucoside(Reference Tsuda, Horio and Osawa124) and after deglycosylation, cyanidin can breakdown spontaneously to give protocatechuic acid (very pronounced at physiological pH)(Reference Kay, Kroon and Cassidy91). After ingestion of black raspberries by pigs, the profile of compounds in the gastrointestinal tract was analysed. In the entire gut, protocatechuic acid was the major phenolic, followed by 4-hydroxycinnamic, caffeic, ferulic and 3-hydroxybenzoic acids. These accounted for approximately 6 % of the ingested anthocyanin(Reference Wu, PittmanIii and Hager125) which is considerably less than the human study described previously.

After consumption of a high dose of strawberries by volunteers, the main phenolic acids in urine after 5 h were 4-hydroxybenzoic acid (10·4 mg/l), protocatechuic acid (0·7 mg/l), vanillic acid (1 mg/l) and genistic acid (1 mg/l)(Reference Russell, Scobbie and Labat126). Strawberry anthocyanins are primarily pelargonidin glycosides with lesser amounts of cyanidin glycosides(Reference Felgines, Talavera and Gonthier122, Reference Hollands, Brett and Dainty127) that would be expected to yield 4-hydroxybenzoic acid and protocatechuic acid (Table 3), but these acids are normal preformed constituents of strawberries(Reference Stohr and Herrmann128).

Consumption of oats added to a purée of bilberries (glycosides of delphinidin, accompanied by lesser amounts of malvidin, peonidin and petunidin glycosides)(Reference Ichiyanagi, Kashiwada and Ikeshiro129Reference Ichiyanagi, Hatano and Matsugo131) and lingonberries (cyanidin glycosides)(Reference Ek, Kartimo and Mattila132) by volunteers resulted in urinary excretion of 3-methoxy-4-hydroxyphenylacetic (homovanillic) and vanillic acid, low amounts of syringic acid (from malvidin glycosides) but no gallic acid (which would have been expected from delphinidin glycosides). Urinary excretion of these acids was maximal at 4–6 h and intact urinary anthocyanins comprised < 0·01 % of the dose(Reference Nurmi, Mursu and Heinonen5).

Procyanidins

When rats were fed procyanidin dimer B3, the major urinary catabolites were 3-(3′-hydroxyphenyl)-propionic, 3-hydroxycinnamic, 4-hydroxybenzoic and vanillic acids (total 6·5 % of intake). Feeding of the procyanidin trimer C3 or a mixture of procyanidin polymers gave in addition 3-hydroxyphenylvaleric acid (C6–C5)(Reference Gonthier, Donovan and Texier133). When rats were fed 14C-labelled procyanidin B2(Reference Stoupi, Williamson and Viton134), approximately 80 % of the 14C-radiolabel was absorbed and appeared in the urine. The 14C-radiolabel appeared rapidly in the blood at low levels, but did not reach a maximum until 6 h, indicative of a major contribution of catabolism by colonic microflora. These two studies indicate that the majority (>70 %) of the colonic catabolites of procyanidin dimers are unknown. In human subjects, after consumption of procyanidin and catechin-rich chocolate, the main urinary catabolites were 3-(3′-hydroxyphenyl)-propionic acid, ferulic acid, 3,4-dihydroxyphenylacetic acid, 3-hydroxyphenylacetic acid, vanillic acid and 3-hydroxybenzoic acid(Reference Rios, Gonthier and Remesy135).

Hydroxycinnamic acids

Hydroxycinnamic acids are most commonly found linked to a quinic acid moiety in fruits and foods. These compounds are called chlorogenic acids and the most common are caffeoyl-quinic acids, found at high levels in coffee. Fibre such as wheat bran is another source, where the hydroxycinnamic acids, particularly ferulic acid, are covalently linked to an arabinose sugar unit. Much of the understanding of the metabolism of hydroxycinnamates is derived from work on coffee. Chlorogenic acids are poorly hydrolysed by conditions in the stomach or small intestine. When coffee is given, there is a relatively small absorption of caffeic and ferulic acids in the small intestine and a low absorption of intact chlorogenic acids. The major absorption occurs in the colon, where dihydroferulic and dihydrocaffeic acids, products of microbial biotransformation, are the major products(Reference Stalmach, Mullen and Barron136, Reference Renouf, Guy and Marmet137). The important involvement of the colon is also supported by studies on ileostomists, where approximately 67 % of the dose of chlorogenic acids reaches the colon(Reference Olthof, Hollman and Katan138). Dihydrocaffeic acid is one of the major phenolic acids in human faecal water(Reference Jenner, Rafter and Halliwell45) and has been detected in the plasma of coffee drinkers(Reference Goldstein, Stull and Markey139), in urine as the free form and mainly conjugated in human plasma after ingestion of artichoke leaf extracts(Reference Wittemer, Ploch and Windeck140), in human urine after chocolate intake(Reference Rios, Gonthier and Remesy135) and in rat urine after ingestion of polyphenol-rich wine extract(Reference Gonthier, Cheynier and Donovan141). In summary, the major products from ingestion of chlorogenic acids are dihydrocinnamic acids in the plasma and urine. Studies on rats showed that the major 5-caffeoylquinic acid-derived phenolic acids, 3-hydroxybenzoic, 3-hydroxybenzoylglycine (3-hydroxyhippuric) and 3-hydroxycinnamic acids were from the caffeic acid moiety but that a significant portion of the hippuric acid (benzoylglycine) was produced from the quinic acid moiety(Reference Gonthier, Verny and Besson142).

Some older work on the tissue distribution of radiolabelled cinnamic acids is worth noting. Injection of 14C-cinnamic acid to rats gave distribution in organs as shown in Fig. 4, with the remaining radioactivity in urine (48 %), faeces (25 %) and exhaled CO2 (0·3 %)(Reference Teuchy and van Sumere143). There was a substantial amount in skin and gonads. A later study showed that 82–90 % of orally administered trans-(3-14C)-cinnamic acid was absorbed in rats as indicated by the amount present in urine after 24 h(Reference Nutley, Farmer and Caldwell144).

Fig. 4 Distribution of radiolabel in rat tissues after injection of 14C-cinnamic acid to rats. The remaining radioactivity was in urine (48 %), faeces (25 %) and exhaled CO2 (0·3 %)(Reference Teuchy and van Sumere143).

Once formed by microbial biotransformation, compounds must pass the colonic epithelium and enter the bloodstream. Mammalian enzymes will further transform the microbial products, mainly by conjugation, but also by β-oxidation. Conjugation can involve methylation (catalysed by the enzyme catechol-O-methyl transferase), sulfation (sulfotransferases), β-glucuronidation (UDP-glucuronosyl transferases), glycinylation (via a CoA thioester) and glutaminylation (by conjugation with glutamine). Many of the enzymes involved exist as multiple isoforms with different but overlapping specificities, typical of those acting on xenobiotics.

Properties of benzoate derivatives (C6–C1) (e.g. protocatechuic and 4-hydroxybenzoic acids)

Absorption and metabolism of C6–C1

Both protocatechuic acid and phloroglucinaldehyde are transported by Caco-2 cells, and then metabolised to sulphate and glucuronide conjugates by Caco-2 cells(Reference Kay, Kroon and Cassidy91) (Fig. 5). Studies with Caco-2 cells have shown that benzoic acid and the three mono-hydroxybezoic acid isomers are substrates for the monocarboxylate transporter(Reference Haughton, Clifford and Sharp145). 3-Methoxy-4-hydroxyphenylacetic (homovanillic) acid is a substrate for the rat organic anion transporter(Reference Mori, Takanaga and Ohtsuki146). Following absorption, catechol-O-methyl transferase methylates protocatechuic acid to vanillic acid(Reference Cao, Zhang and Ma147). Of the conjugating enzymes, UDP-glucuronosyl transferases 1A6 in human liver microsomes is active on protocatechuic aldehyde(Reference Liu, Liu and Zhang148). In addition, protocatechuic aldehyde is converted to approximately 70 % protocatechuic acid by guinea pig liver slices by the enzymes aldehyde oxidase, xanthine oxidase and aldehyde dehydrogenase(Reference Panoutsopoulos and Beedham149), but approximately 20 % was unidentified polar conjugates of protocatechuic acid. The conversion of caffeic acid to protocatechuic acid did not occur in the absence of a gut microflora in germ-free rats(Reference Peppercorn and Goldman34).

Fig. 5 Summary of absorption and metabolic pathways of C6–C1 and C6–C3 compounds in the gastrointestinal tract.

Biological effects of C6–C1

In a rat model of induced carcinogenesis, protocatechuic acid prevented 4-nitroquinoline-1-oxide-induced oral carcinogenesis, N-methyl-N-nitrosourea-induced carcinogenesis in the glandular stomach, azoxymethane-induced carcinogenesis in the colon and diethylnitrosamine-induced carcinogenesis in the liver(Reference Tanaka, Kawamori and Ohnishi150, Reference Tanaka, Kojima and Kawamori151). The doses of protocatechuic acid were expressed only as a percentage of the diet, but can be estimated as approximately 10–40 mg/kg body weight per d assuming that a 200 g rat typically consumes 20 g diet per d.

Several papers have reported in vitro activity of protocatechuic acid on cultured cells, and in some investigations, this microbial catabolite is more potent than cyanidin glycoside, an anthocyanin that gives rise to substantial levels of protocatechuic acid in vivo (Reference Vitaglione, Donnarumma and Napolitano4). For example, in human neuronal SH-SY5Y cells, protocatechuic acid (100 μm) is more potent than cyanidin-3-O–glucoside (100 μm) at inhibiting hydrogen peroxide-induced apoptotic events, including mitochondrial function and DNA fragmentation. However, neither was effective below 25 μm(Reference Tarozzi, Morroni and Hrelia152). Protocatechuic acid induces c-jun N-terminal kinase-dependent hepatocellular carcinoma cell death at concentrations>50 μm(Reference Yip, Chan and Pang153).

Protocatechuic acid at concentrations above 500 μm promoted time-dependent and concentration-dependent migration of human adipose-tissue derived stromal cells in vitro possibly via effects on matrix metalloproteinase-2(Reference Wang, Liu and Guan154), and promoted cell proliferation of cultured rat neural stem cells, with reduced apoptosis possibly via repression of caspase-3 activation(Reference Guan, Ge and Liu155). Significantly lower production of reactive oxygen species was seen following treatment with protocatechuic acid (6 μm for 7 d or 30 μm for 4 d) but suppression of caspase-3 required 30 μm for both durations. The relative effectiveness of lower doses longer term is encouraging and suggests that dietary exposure over a long period might perhaps promote better health over the subsequent years.

Several investigations have addressed lipid and cholesterol metabolism. Protocatechuic acid, the major component of the Chinese functional medicine, Danshen, has reported anti-angina efficacy(Reference Cao, Zhang and Ma147). Protocatechuic acid is methylated and then diffuses into mitochondria where it is conjugated with CoA. The result is that fatty acid oxidation is decreased, as shown by a lower acyl CoA/CoA ratio in heart, which in turn activates pyruvate dehydrogenase, a key and irreversible step in carbohydrate oxidation. This could switch heart energy substrate preference from fatty acid to glucose that would be beneficial for ischaemic heart conditions. There was no detectable accumulation of protocatechuic acid in tissues(Reference Cao, Zhang and Ma147). Protocatechuic acid at 0·5 g/kg diet given to rats (estimated as approximately 10 mg/kg body weight per d) produced changes in cholesterol and lipid metabolism. The serum total cholesterol, HDL-cholesterol and VLDL-cholesterol were all lower in the protocatechuic acid-treated group, suggested to be partly as a result of induction (estimated by mRNA) of hepatic LDL receptor, apoB, apoE, lecithin-cholesterol acyltransferase and hepatic TAG lipase(Reference Tamura, Fukushima and Shimada156).

Properties of phenylacetate derivatives (C6–C2)

Absorption and metabolism of C6–C2

Phenylacetate is an endogenous product of phenylalanine metabolism, is present at low levels in the mammalian circulation and is conjugated with glutatmine during metabolism. Its pharmacokinetic parameters have been well studied in a clinical setting in patients owing to its proposed effects on cancer. When phenylacetate was administered as an intravenous infusion as part of a phase I trial in children with refractory cancer, the half life was approximately 1 h, and phenylacetate was conjugated with glutatmine to form phenylacetylglutatmine(Reference Thompson, Balis and Serabe157). After oral consumption of phenylacetate, the concentration of phenylacetate peaked at 2 h and 40 % was excreted in the urine after 40 h(Reference Wajngot, Chandramouli and Schumann158). Intravenous radiolabelled 14C-phenylacetate was rapidly taken up by rat brain and converted into 14C-acetate(Reference Inoue, Hosoi and Momosaki159). Phenylacetate was well tolerated when infused into patients twice per d for weeks at a high dose (125 mg/kg body weight), and at these levels, phenylacetate induced its own clearance by 27 % during this period(Reference Thibault, Samid and Cooper160). Both phenylketonuric and normal subjects eliminated an oral dose (80 mg) of [14C]-phenylacetic acid in the urine almost entirely as phenacetylglutamine, showing that the glutamine conjugation mechanism is not defective in phenylketonuria and that it is able to cope with the large amounts of phenylacetic acid produced in this disorder(Reference James and Smith161). The contents of the colon, examined as human faecal water, contained phenylacetic acid, 3-phenylpropionic acid, 3-hydroxyphenylacetic acid, 3,4-dihydroxyphenylacetic acid, 3-(4′-hydroxyphenyl)-propionic acid and 4-hydroxy-3-methoxycinnamic (ferulic) acid at high levels, up to 400 μm for phenylacetic acid and 3-phenylpropionic acid in one individual. The mean levels were 188, 197, 110, 64, 61 and 10 μm, respectively(Reference Karlsson, Huss and Jenner162).

Biological effects of C6–C2

Phenylacetic acid (administered as sodium phenylacetate) is one component of a drug (together with sodium benzoate) which is given for the treatment of acute hyperammonemia(Reference MacArthur, Altincatal and Tuchman163). As discussed previously, the pharmacokinetic behaviour is well characterised(Reference MacArthur, Altincatal and Tuchman163), and phenylacetylglutatmine and hippuric acids are products excreted in the urine after an intravenous dose. Phenylbutyrate, a prodrug that is metabolised into its active component, phenylacetate, in vivo, has been reported to extend lifespan in Drosophila (Reference Kang, Benzer and Min164). Large doses of phenylbutyrate can have some toxic side effects(Reference Kasumov, Brunengraber and Comte165), and it has been shown that phenylacetate and phenylbutyrate modulate medulloblastoma(Reference Li, Parikh and Shu166). Some C6–C2 compounds may modulate cholesterol metabolism (see below). When given orally, a single high dose of phenylacetate did not affect plasma glucose concentration nor gluconeogenesis in type II diabetic patients(Reference Wajngot, Chandramouli and Schumann158). Phenylacetate affects cell growth and proliferation(Reference Li, Parikh and Shu166Reference Shibahara, Onishi and Franco168), inhibits prenylation and has been tested in both phase I and II clinical trials. 3,4-Dihydroxyphenylacetic acid exhibited antiproliferative activity on prostate and colon cancer cells(Reference Gao, Xu and Krul33). 3-Phenylpropionic acid, 3-hydroxyphenylacetic acid and 3-(4′-hydroxyphenyl)-propionic acid, found at high concentrations in faecal water, decreased the expression of cyclo-oxygenase-2 in HT29 cells at physiologically relevant concentrations(Reference Karlsson, Huss and Jenner162).

Properties of phenylpropionate derivatives (C6–C3) (e.g. dihydrocaffeic and dihydroferulic acids)

Absorption and metabolism of C6–C3

As summarised in Table 2, C6–C3 acids can be formed from many PPT substrates including proanthocyanidins. However, information on the absorption and metabolism of C6–C3 acids has come largely from studies of chlorogenic acids, conjugates of cinnamic acids with quinic acid. Such conjugates are widespread in foods and beverages, but coffee, artichoke, apples and blueberries are particularly rich(Reference Crozier, Jaganath and Clifford7, Reference Clifford169, Reference Clifford170) Approximately, 1 h after the consumption of coffee, ferulic, caffeic and isoferulic acids appear in the plasma at low levels indicating absorption in the small intestine(Reference Renouf, Guy and Marmet137). This early appearance was characterised in more detail, and the circulating compounds included caffeic acid-3-O-sulphate and ferulic acid-4-O-sulphate, and sulphates of 3- and 4-caffeoylquinic acid lactones(Reference Stalmach, Mullen and Barron136). These lactones are formed from the chlorogenic acids during coffee roasting(Reference Clifford169) and are rarely found in other foods or beverages. Several hours later (T max >4 h), greater concentrations of dihydroferulic acid, dihydroferulic acid-4-O-sulphate and dihydrocaffeic acid-3-O-sulphate appear in blood, indicating colonic absorption after microbial transformation(Reference Renouf, Guy and Marmet137). Dihydroferulic acid, dihydroferulic acid-4-O-sulphate, dihydrocaffeic acid-3-O-sulphate, ferulic acid-4-O-sulphate, dihydroferulic acid-4-O-glucuronide and feruloyl-glycine were major urinary components. These catabolites collectively represented approximately 29 % of the chlorogenic acid dose, indicating good absorption of the microbial catabolites of chlorogenic acids(Reference Stalmach, Mullen and Barron136).

The colon pH value ranges from pH 5·7 to 6·7(Reference Fallingborg171), and unless there are localised pockets of significantly lower pH value, these phenolic acids with pK values in the range pH 4–5(Reference Jovanovic, Steenken and Tosic172) will be extensively ionised. The absorption of C6–C3 catabolites has been examined using Caco-2 cells and rat everted intestinal sacs (Fig. 5). Dihydrocaffeic acid is absorbed mainly by the passive transcellular route and enters the circulation mainly as the free form, although a small proportion is glucuronidated(Reference Poquet, Clifford and Williamson109). The dihydrocaffeic acid is either rapidly sulphated, or methylated then sulphated, by the liver(Reference Poquet, Clifford and Williamson173), so that the major metabolites in the circulation are free dihydroferulic acid, dihydrocaffeic acid-3-O-sulphate or dihydroferulic acid-4-O-sulphate. Dihydroferulic acid in the gut is absorbed partially by the passive transcellular route, but with some involvement of a monocarboxylic transporter(Reference Poquet, Clifford and Williamson109, Reference Konishi and Kobayashi174). 3-Hydroxycinnamic acid and its dihydroform, 3-(3′-hydroxyphenyl)-propionic acid, are both absorbed at least partially by the monocarboxylic transporter in Caco-2 cells(Reference Konishi and Kobayashi175).

Once absorbed in the colon, sulphate conjugation occurs primarily in the liver. In human subjects, sulphotransferase 1A1 is the most active isoform for sulfation of caffeic, isoferulic and dihydrocaffeic acids, whereas sulphotransferase 1E1 is most active towards the other methylated forms, i.e. ferulic and dihydroferulic acids(Reference Wong, Meinl and Glatt176). Glucuronidation occurs to a much lesser extent. The levels of dihydrocaffeic acid-3-O-sulphate were approximately 45-fold greater than the corresponding glucuronides, and the amount of dihydroferulic acid-4-O-sulphate in urine was much greater than the amount of dihydroferulic acid-4-O-glucuronide. The only cinnamic acid for which the glucuronides predominated was isoferulic acid(Reference Wong, Meinl and Glatt176).

Cinnamic acids, as distinct from the dihydrocinnamic acids, have rarely been reported as gut flora catabolites of flavonoids. Rats excrete 4-hydroxycinnamic acid in urine after the consumption of apigenin, phloridzin and naringenin(Reference Griffiths and Smith56).

Biological effects of C6–C3

Excess solar UV radiation produces damage and initiates immune response and inflammation in skin, sometimes leading to cancer. Dihydrocaffeic and caffeic acids, but not dihydroferulic and ferulic acids, reduced the cytotoxicity and pro-inflammatory cytokine production (IL-6 and -8) in HaCaT cells, a keratinocyte model, following UV radiation(Reference Poquet, Clifford and Williamson177), and dihydrocaffeic and 4-hydroxycinnamic acids inhibited UV-B damage in human conjunctival cells as assessed using the DNA damage marker, 7,8-dihydro-8-oxo-2′-deoxyguanosine(Reference Larrosa, Lodovici and Morbidelli178). 4-Hydroxycinnamic acid, but not protocatechuic acid, decreases basal oxidative DNA damage in rat colonic mucosa(Reference Guglielmi, Luceri and Giovannelli179). In CCD-18 colon fibroblast cells stimulated with IL-1β, dihydrocaffeic, dihydroferulic and dihydroxyphenylacetic acids attenuated PG E2 demonstrating an anti-inflammatory effect in this system. In addition, dihydrocaffeic acid diminished the expression of the cytokines IL-1β, IL-8 and TNF-α, reduced malonyldialdehyde levels and reduced oxidative DNA damage (measured as 7,8-dihydro-8-oxo-2′-deoxyguanosine) in distal colon mucosal tissue in the dextran sodium sulphate-induced colitis model in rats(Reference Larrosa, Luceri and Vivoli180). Animal and studies in vitro also suggest that some C6–C2(Reference Bhat and Ramasarma181) and especially C6–C3 (Reference Lee, Choi and Jeon182) catabolites interfere with various enzymes in the mevalonate pathway including 3-hydroxy-3-methylglutaryl-CoA reductase, the rate limiting enzyme in hepatic cholesterol biosynthesis, albeit at concentrations unlikely to occur in plasma. However, these observations are of interest since commodities rich in PPT that would yield such catabolites, and the catabolites when given per os, have been shown to inhibit platelet aggregation(Reference Yasuda, Takasawa and Nakazawa183) or to have cholesterol-lowering activity in animal(Reference Lee, Choi and Jeon182, Reference Bok, Lee and Park184Reference Yamakoshi, Kataoka and Koga188) and human studies(Reference Kurowska, Spence and Jordan189), and such gut flora catabolites may have contributed to the in vivo effect. Interference in the mevalonate pathway, particularly 3-hydroxy-3-methylglutaryl-CoA reductase inhibition, may have broader human significance(Reference Mo and Elson190).

Properties of phenylvalerate derivatives (C6–C5)

Following the consumption of green tea flavanols or proanthocyanidin-rich extracts, several C6–C5 catabolites have been detected in human plasma and urine as methyl and/or glucuronide/sulphate conjugates(Reference Lee, Maliakal and Chen191Reference Duweler and Rohdewald193).

Prebiotic effects of polyphenols on the microflora

Although the fact that biotransformation of dietary PPT by the gut microflora occurs has been known for many years, the possibility that either the untransformed PPT or the phenolic acid catabolites might modify the composition and biochemical competence of the gut microflora has attracted comparatively little attention. Changes over 20 d in the catabolites produced from punicalagin have been attributed to changes in the GIT microflora(Reference Cerda, Llorach and Ceron97). There is evidence from in vitro models(Reference Lee, Jenner and Low31, Reference Tzounis, Vulevic and Kuhnle32) and in vivo studies using human subjects, pigs and sheep that PPT and/or their catabolites influence the composition of the gut microflora, for example lowering the colonic pH value, suppressing bacteroides and pathogenic Clostridium perfringens and Clostridium difficile, and increasing the proportion of bifidobacteria and eubacteria without inhibiting lactic acid bacteria(Reference Lee, Jenner and Low31, Reference Goto, Kanaya and Nishikawa194Reference Okubo, Ishihara and Oura196), but the exact mechanisms are uncertain(Reference Min, Attwood and Reilly197). The concentrations of benzoic acid, phenylacetic acid, phenylpropionic acid and 3-(3′-hydroxyphenyl)-propionic acid in human faecal water can each reach millimolar concentrations(Reference Knust, Erben and Spiegelhalder44, Reference Jenner, Rafter and Halliwell45) which has the obvious potential to modulate bacterial growth.

Caffeic acid, a comparatively minor component, can reach concentrations (approximately 50 μm)(Reference Jenner, Rafter and Halliwell45) well in excess of the concentration shown in vitro to produce 50 % inhibition in the growth of the opportunistic pathogen Listeria monocytogenes (Reference Ramos-Niño, Ramìrez-Rodriguez and Clifford198). Various flavonoid aglycones are also found in faecal water, but at low concentrations ( < 3 μm). There is some evidence also that some pathogenic protozoa are inhibited by some flavonoids (IC50 or EC50 values in the range 15–50 μm)(Reference Gargala, Baishanbo and Favennec199, Reference Kerboeuf, Riou and Guegnard200) but the effect of the aromatic/phenolic acid catabolites does not seem to have been investigated. Although the yield of phenolic/aromatic acids is variable (up to × 10) between individuals and is not necessarily normally distributed(Reference Knust, Erben and Spiegelhalder44, Reference Jenner, Rafter and Halliwell45, Reference Li, Lee and Sheng201, Reference Clifford, Copeland and Bloxsidge202), the potential for suppressing some pathogens and for a prebiotic effect in some individuals clearly exists.

Summary, conclusions and recommendations for future research

The colonic microbiota transform a very complex range of PPT substrates that will vary between individuals and occasions. These transfomations can be extensive, and while some microbial catabolites are substrate-specific (equol, urolithins, mammalian lignans), certain catabolites are common to many of the major substrates, and this implies that the spectrum of catabolites produced is less complex and qualitatively less variable than the spectrum of substrates consumed. The catabolites most likely to dominate are the C6–C1, C6–C2 and C6–C3 dihydro acids derived from cinnamates, anthocyanins and most other flavonoids. These microbial catabolites are often better absorbed than the parent compounds, because of the mechanism of absorption, the large absorptive area available in the colon and the high concentrations (approaching millimolar) in the colonic lumen.

Therefore, we propose that these microbial catabolites could be responsible for a significant proportion of the biological activity derived from consumption of fruits, vegetables and other plant derived products such as fruit drinks, wine, coffee and tea.

The biological activity of these microbial catabolites would be in addition to the biological activity of any absorbed parent compound and its mammalian metabolites, and the potential for synergy between microbial catabolites and the parent compound and its mammalian metabolites could be important but has not been studied. Because of the potential importance of the microbial catabolites, we need to understand better the factors controlling (i) their production and whether this can be modulated advantageously, (ii) their absorption and mammalian metabolism and (iii) their biological activities both in vitro and in vivo.

The existing epidemiological data must be revisited and extended so as to take account of substances not present in the diet per se. For a given substrate consumption, there is often a large inter-individual variation (approximately ×  10) in catabolite yields, subsequent plasma concentrations and also variation in the concentration–time profile. It is essential to define the factors responsible, whether manifest in the microflora, in the host or in both. Similarly it is important to determine the potential of PPT and their catabolites to modulate the gut microflora and hence potentially their own production. The answers to these are crucial if we are to understand and exploit the effect of PPT consumption on human health.

Acknowledgements

We acknowledge funding from GlaxoSmithKline to G. W. and M. N. C. for writing this review. There are no conflicts of interest. G. W. and M. N. C. contributed 50 % each to the writing of the present paper.

References

1 Hooper, L, Kroon, PA, Rimm, EB, et al. (2008) Flavonoids, flavonoid-rich foods, and cardiovascular risk: a meta-analysis of randomized controlled trials. Am J Clin Nutr 88, 3850.CrossRefGoogle ScholarPubMed
2 Williamson, G & Manach, C (2005) Bioavailability and bioefficacy of polyphenols in humans. II. Review of 93 intervention studies. Am J Clin Nutr 81, 243S255S.CrossRefGoogle ScholarPubMed
3 Scholz, S & Williamson, G (2007) Interactions affecting the bioavailability of dietary polyphenols in vivo. Int J Vitam Nutr Res 77, 224235.CrossRefGoogle ScholarPubMed
4 Vitaglione, P, Donnarumma, G, Napolitano, A, et al. (2007) Protocatechuic acid is the major human metabolite of cyanidin-glucosides. J Nutr 137, 20432048.CrossRefGoogle ScholarPubMed
5 Nurmi, T, Mursu, J, Heinonen, M, et al. (2009) Metabolism of berry anthocyanins to phenolic acids in humans. J Agric Food Chem 57, 22742281.CrossRefGoogle ScholarPubMed
6 Jensen, GS, Wu, X, Patterson, KM, et al. (2008) In vitro and in vivo antioxidant and anti-inflammatory capacities of an antioxidant-rich fruit and berry juice blend. Results of a pilot and randomized, double-blinded, placebo-controlled, crossover study. J Agric Food Chem 56, 83268333.CrossRefGoogle ScholarPubMed
7 Crozier, A, Jaganath, IB & Clifford, MN (2009) Dietary phenolics: chemistry, bioavailability and effects on health. Nat Prod Rep 26, 10011043.CrossRefGoogle ScholarPubMed
8 Crozier, A, Clifford, MN & Ashihara, H (2006) Plant Secondary Metabolites. Occurrence, Structure and Role in the Human Diet. Oxford: Blackwell.CrossRefGoogle Scholar
9 Bäckhed, F, Ley, RE, Sonnenburg, JL, et al. (2005) Host-bacterial mutualism in the human intestine. Science 307, 19151920.CrossRefGoogle ScholarPubMed
10 Savage, DC (1977) Microbial ecology of the gastrointestinal tract. Ann Rev Microbiol 31, 107133.CrossRefGoogle ScholarPubMed
11 Salminen, S, Isolauri, E & Salminen, E (1996) Clinical uses of probiotics for stabilizing the gut mucosal barrier: successful strains and future challenges. Antonie Van Leeuwenhoek 70, 347358.CrossRefGoogle ScholarPubMed
12 Tuohy, K & Gibson, GR (2006) Functions of the human intestinal flora: the use of probiotics and prebiotics. In Plant Secondary Metabolites. Occurrence, Structure and Role in the Human Diet, pp. 174207 [Crozier, A, Clifford, MN and Ashihara, H, editors]. Oxford: Blackwell.CrossRefGoogle Scholar
13 Gill, SR, Pop, M, Deboy, RT, et al. (2006) Metagenomic analysis of the human distal gut microbiome. Science 312, 13551359.CrossRefGoogle ScholarPubMed
14 Gronlund, MM, Lehtonen, OP, Eerola, E, et al. (1999) Fecal microflora in healthy infants born by different methods of delivery: permanent changes in intestinal flora after cesarean delivery. J Pediatr Gastroenterol Nutr 28, 1925.Google ScholarPubMed
15 Harmsen, HJ, Wildeboer-Veloo, AC, Raangs, GC, et al. (2000) Analysis of intestinal flora development in breast-fed and formula-fed infants by using molecular identification and detection methods. J Pediatr Gastroenterol Nutr 30, 6167.Google ScholarPubMed
16 Conway, PL (2009) Microbial ecology of the human large intestine. In Human Colonic Bacteria-Role in Nutrition, Physiology and Pathology, [Gibson, GR and Macfarlane, GT, editors]. Boca Raton, FL: CRC Press.Google Scholar
17 Eckburg, PB, Bik, EM, Bernstein, CN, et al. (2005) Diversity of the human intestinal microbial flora. Science 308, 16351638.CrossRefGoogle ScholarPubMed
18 Mueller, S, Saunier, K, Hanisch, C, et al. (2006) Differences in fecal microbiota in different European study populations in relation to age, gender, and country: a cross-sectional study. Appl Environ Microbiol 72, 10271033.CrossRefGoogle ScholarPubMed
19 Phipps, AN, Stewart, J, Wright, B, et al. (1998) Effect of diet on the urinary excretion of hippuric acid and other dietary-derived aromatics in rat. A complex interaction between diet, gut microflora and substrate specificity. Xenobiotica 28, 527537.CrossRefGoogle Scholar
20 Wang, XL, Hur, HG, Lee, JH, et al. (2005) Enantioselective synthesis of S-equol from dihydrodaidzein by a newly isolated anaerobic human intestinal bacterium. Appl Environ Microbiol 71, 214219.CrossRefGoogle ScholarPubMed
21 Raimondi, S, Roncaglia, L, De Lucia, M, et al. (2009) Bioconversion of soy isoflavones daidzin and daidzein by Bifidobacterium strains. Appl Microbiol Biotechnol 81, 943950.CrossRefGoogle ScholarPubMed
22 Knaup, B, Kahle, K, Erk, T, et al. (2007) Human intestinal hydrolysis of phenol glycosides – a study with quercetin and p-nitrophenol glycosides using ileostomy fluid. Mol Nutr Food Res 51, 14231429.CrossRefGoogle ScholarPubMed
23 Aura, AM, Martin-Lopez, P, O'Leary, KA, et al. (2005) In vitro metabolism of anthocyanins by human gut microflora. Eur J Nutr 44, 133142.CrossRefGoogle ScholarPubMed
24 Keppler, K & Humpf, HU (2005) Metabolism of anthocyanins and their phenolic degradation products by the intestinal microflora. Bioorg Med Chem 13, 51955205.CrossRefGoogle ScholarPubMed
25 Fleschhut, J, Kratzer, F, Rechkemmer, G, et al. (2006) Stability and biotransformation of various dietary anthocyanins in vitro. Eur J Nutr 45, 718.CrossRefGoogle ScholarPubMed
26 Braune, A, Engst, W & Blaut, M (2005) Degradation of neohesperidin dihydrochalcone by human intestinal bacteria. J Agric Food Chem 53, 17821790.CrossRefGoogle ScholarPubMed
27 Gonthier, MP, Remesy, C, Scalbert, A, et al. (2006) Microbial metabolism of caffeic acid and its esters chlorogenic and caftaric acids by human faecal microbiota in vitro. Biomed Pharmacother 60, 536540.CrossRefGoogle ScholarPubMed
28 Labib, S, Hummel, S, Richling, E, et al. (2006) Use of the pig caecum model to mimic the human intestinal metabolism of hispidulin and related compounds. Mol Nutr Food Res 50, 7886.CrossRefGoogle ScholarPubMed
29 Rafii, F, Jackson, LD, Ross, I, et al. (2007) Metabolism of daidzein by fecal bacteria in rats. Comp Med 57, 282286.Google ScholarPubMed
30 Forester, SC & Waterhouse, AL (2008) Identification of Cabernet Sauvignon anthocyanin gut microflora metabolites. J Agric Food Chem 56, 92999304.CrossRefGoogle ScholarPubMed
31 Lee, HC, Jenner, AM, Low, CS, et al. (2006) Effect of tea phenolics and their aromatic fecal bacterial metabolites on intestinal microbiota. Res Microbiol 157, 876884.CrossRefGoogle ScholarPubMed
32 Tzounis, X, Vulevic, J, Kuhnle, GG, et al. (2008) Flavanol monomer-induced changes to the human faecal microflora. Br J Nutr 99, 782792.CrossRefGoogle Scholar
33 Gao, K, Xu, A, Krul, C, et al. (2006) Of the major phenolic acids formed during human microbial fermentation of tea, citrus, and soy flavonoid supplements, only 3,4-dihydroxyphenylacetic acid has antiproliferative activity. J Nutr 136, 5257.CrossRefGoogle ScholarPubMed
34 Peppercorn, MA & Goldman, P (1972) Caffeic acid metabolism by gnotobiotic rats and their intestinal bacteria. Proc Natl Acad Sci U S A 69, 14131415.CrossRefGoogle ScholarPubMed
35 Hooper, LV & Gordon, JI (2001) Commensal host-bacterial relationships in the gut. Science 292, 11151118.CrossRefGoogle ScholarPubMed
36 Stoupi, S, Williamson, G, Drynan, JW, et al. (2010) A comparison of the in vitro biotransformation of ( − )-epicatechin and procyanidin B2 by human faecal microbiota. Mol Nutr Food Res 54, 747759.CrossRefGoogle ScholarPubMed
37 He, J, Magnuson, BA & Giusti, MM (2005) Analysis of anthocyanins in rat intestinal contents – impact of anthocyanin chemical structure on fecal excretion. J Agric Food Chem 53, 28592866.CrossRefGoogle ScholarPubMed
38 Scheline, RR & Midtvedt, T (1970) Absence of dehydroxylation of caffeic acid in germ-free rats. Experientia 26, 10681069.CrossRefGoogle ScholarPubMed
39 Tamura, M & Saitoh, H (2006) Comparison of the in vitro metabolism of isoflavones by fecal flora from human flora-associated mice and human. J Sci Food Agric 86, 15671570.CrossRefGoogle Scholar
40 Rumney, CJ & Rowland, IR (1992) In vivo and in vitro models of the human colonic flora. CRC Crit Rev Food Sci Nutr 31, 299331.CrossRefGoogle ScholarPubMed
41 de Eds, F (1968) Flavonoid metabolism. In Metabolism of Cyclic Compounds, pp. 127192 [Florkin, M and Stotz, EH, editors]. London: Elsevier.Google Scholar
42 Griffiths, LA (1982) Mammalian metabolism of flavonoids. In The Flavonoids: Advances in Research, pp. 681718 [Harborne, JB and Mabry, TJ, editors]. London: Chapman & Hall.CrossRefGoogle Scholar
43 Berry, DF, Francis, AJ & Bollag, J-M (1987) Microbial metabolism of homocyclic and heterocyclic aromatic compounds under anaerobic conditions. Microbiol Rev 51, 4359.CrossRefGoogle ScholarPubMed
44 Knust, U, Erben, G, Spiegelhalder, B, et al. (2006) Identification and quantitation of phenolic compounds in faecal matrix by capillary gas chromatography and nano-electrospray mass spectrometry. Rapid Commun Mass Spectrom 20, 31193129.CrossRefGoogle ScholarPubMed
45 Jenner, AM, Rafter, J & Halliwell, B (2005) Human fecal water content of phenolics: the extent of colonic exposure to aromatic compounds. Free Radic Biol Med 38, 763772.CrossRefGoogle ScholarPubMed
46 Hattori, M, Shu, YZ, El Sedawy, AI, et al. (1988) Metabolism of homoorientin by human intestinal bacteria. J Nat Prod 51, 874878.CrossRefGoogle ScholarPubMed
47 Daniel, EM, Ratnayake, S, Kinstle, T, et al. (1991) The effects of pH and rat intestinal contents on the liberation of ellagic acid from purified and crude ellagitannins. J Nat Prod 54, 946952.CrossRefGoogle ScholarPubMed
48 Kim, DH, Kang, HJ, Park, SH, et al. (1994) Characterization of β-glucosidase and β-glucuronidase of alkalotolerant intestinal bacteria. Biol Pharm Bull 17, 423426.CrossRefGoogle ScholarPubMed
49 Kroon, PA, Faulds, CB, Ryden, P, et al. (1996) Release of covalently bound ferulic acid from fiber in the human colon. J Agric Food Chem 45, 661667.CrossRefGoogle Scholar
50 Goodwin, BL, Ruthven, CRJ & Sandler, M (1994) Gut flora and the origin of some urinary aromatic phenolic compounds. Biochem Pharmacol 47, 22942297.CrossRefGoogle ScholarPubMed
51 Adamson, RH, Bridges, JW, Evans, ME, et al. (1970) Species differences in the aromatization of quinic acid in vivo and the role of gut bacteria. Biochem J 116, 437443.CrossRefGoogle ScholarPubMed
52 Walle, T, Walle, UK & Halushka, PV (2001) Carbon dioxide is the major metabolite of quercetin in humans. J Nutr 131, 26482652.CrossRefGoogle ScholarPubMed
53 Das, NP & Griffiths, LA (1969) Studies on flavonoid metabolism. Metabolism of (+)-[14C]catechin in the rat and guinea pig. Biochem J 115, 831836.CrossRefGoogle Scholar
54 Griffiths, LA & Smith, GE (1972) Metabolism of myricetin and related compounds in the rat. Metabolite formation in vivo and by the intestinal microflora in vitro. Biochem J 130, 141151.CrossRefGoogle ScholarPubMed
55 Smith, GE & Griffiths, LA (1970) Metabolism of myricitrin and 3,4,5-trihydroxyphenylacetic acid. Biochem J 118, 53P54P.CrossRefGoogle Scholar
56 Griffiths, LA & Smith, GE (1972) Metabolism of apigenin and related compounds in the rat. Biochem J 128, 901911.CrossRefGoogle ScholarPubMed
57 Curtius, HC, Mettler, M & Ettlinger, L (1976) Study of the intestinal tyrosine metabolism using stable isotopes and gas chromatography-mass spectrometry. J Chromatogr 126, 569580.CrossRefGoogle ScholarPubMed
58 Fuchs-Mettler, M, Curtius, HC, Baerlocher, K, et al. (1980) A new rearrangement reaction in tyrosine metabolism. Eur J Biochem 108, 527534.CrossRefGoogle ScholarPubMed
59 Gott, DM & Griffiths, LA (1987) Effects of antibiotic pretreatments on the metabolism and excretion [U14C](+)-catechin ([U14C]cyanidanol-3) and its metabolite, 3′-O-methyl-(+)-catechin. Xenobiotica 17, 423434.CrossRefGoogle Scholar
60 Blaut, M, Schoefer, L & Braune, A (2003) Transformation of flavonoids by intestinal microorganisms. Int J Vitam Nutr Res 73, 7987.CrossRefGoogle ScholarPubMed
61 Hackett, AM & Griffiths, LA (1981) The metabolism and excretion of 3-O-methyl-(+)-catechin in the rat, mouse, and marmoset. Drug Metab Dispos 9, 5459.Google Scholar
62 Hackett, AM, Griffiths, LA & Wermeille, M (1985) The quantitative disposition of 3-O-methyl-(+)-U-[-14C]catechin in man following oral administration. Xenobiotica 15, 907914.CrossRefGoogle ScholarPubMed
63 Krumholz, LR & Bryant, MP (1988) Characterization of the pyrogallol-phloroglucinol isomerase of Eubacterium oxidoreducens. J Bact 170, 24722479.CrossRefGoogle ScholarPubMed
64 Krumholz, LR, Crawford, RL, Hemling, ME, et al. (1987) Metabolism of gallate and phloroglucinol in Eubacterium oxidoreducens via 3-hydroxy-5-oxohexanoate. J Bact 169, 18861890.CrossRefGoogle ScholarPubMed
65 Haddock, JD & Ferry, JG (1993) Initial steps in the anaerobic degradation of 3,4,5-trihydroxybenzoate by Eubacterium oxidoreducens: characterization of mutants and role of 1,2,3,5-tetrahydroxybenzene. J Bacteriol 175, 669673.CrossRefGoogle Scholar
66 Haddock, JD & Ferry, JG (1989) Purification and properties of phloroglucinol reductase from Eubacterium oxidoreducens G-41. J Biol Chem 264, 44234427.CrossRefGoogle ScholarPubMed
67 Jeffrey, AM, Jerina, DM, Self, R, et al. (1972) The bacterial degradation of flavonoids. Oxidative fission of the A-ring of dihydrogossypetin by a Pseudomonas sp. Biochem J 130, 383390.CrossRefGoogle ScholarPubMed
68 Jeffrey, AM, Knight, M & Evans, WC (1972) The bacterial degradation of flavonoids. Hydroxylation of the A-ring of taxifolin by a soil pseudomonad. Biochem J 130, 373381.CrossRefGoogle ScholarPubMed
69 Simmering, R, Kleessen, B & Blaut, M (1999) Quantification of the flavonoid-degrading bacterium Eubacterium ramulus in human fecal samples with a species-specific oligonucleotide hybridization probe. Appl Environ Microbiol 65, 37053709.CrossRefGoogle ScholarPubMed
70 Simmering, R, Pforte, H, Jacobasch, G, et al. (2002) The growth of the flavonoid-degrading intestinal bacterium Eubacterium ramulus, is stimulated by dietary flavonoids in vivo. FEMS Microbiol Ecol 40, 243248.CrossRefGoogle ScholarPubMed
71 Schneider, H & Blaut, M (2000) Anaerobic degradation of flavonoids by Eubacterium ramulus. Arch Microbiol 173, 7175.CrossRefGoogle ScholarPubMed
72 Schneider, H, Schwiertz, A, Collins, MD, et al. (1999) Anaerobic transformation of quercetin-3-glucoside by bacteria from the human intestinal tract. Arch Microbiol 171, 8191.CrossRefGoogle ScholarPubMed
73 Schneider, H, Simmering, R, Hartmann, L, et al. (2000) Degradation of quercetin-3-glucoside in gnotobiotic rats associated with human intestinal bacteria. J Appl Microbiol 89, 10271037.CrossRefGoogle ScholarPubMed
74 Schoefer, L, Mohan, R, Braune, A, et al. (2002) Anaerobic C-ring cleavage of genistein and daidzein by Eubacterium ramulus. FEMS Microbiol Lett 208, 197202.CrossRefGoogle ScholarPubMed
75 Schoefer, L, Braune, A & Blaut, M (2004) Cloning and expression of a phloretin hydrolase gene from Eubacterium ramulus and characterization of the recombinant enzyme. Appl Environ Microbiol 70, 61316137.CrossRefGoogle ScholarPubMed
76 Herles, C, Braune, A & Blaut, M (2004) First bacterial chalcone isomerase isolated from Eubacterium ramulus. Arch Microbiol 181, 428434.CrossRefGoogle ScholarPubMed
77 Clifford, MN (2000) Anthocyanins – nature, occurrence and dietary burden. J Sci Food Agric 80, 10631072.3.0.CO;2-Q>CrossRefGoogle Scholar
78 Hur, H & Rafii, F (2000) Biotransformation of the isoflavonoids biochanin A, formononetin, and glycitein by Eubacterium limosum. FEMS Microbiol Lett 192, 2125.CrossRefGoogle ScholarPubMed
79 Kim, DH, Konishi, L & Kobashi, K (1986) Purification, characterization and reaction mechanism of novel arylsulfotransferase obtained from an anaerobic bacterium of human intestine. Biochim Biophys Acta 872, 3341.CrossRefGoogle ScholarPubMed
80 Wang, LQ, Meselhy, MR, Li, Y, et al. (2000) Human intestinal bacteria capable of transforming secoisolariciresinol diglucoside to mammalian lignans, enterodiol and enterolactone. Chem Pharm Bull 48, 16061610.CrossRefGoogle ScholarPubMed
81 Wang, LQ, Meselhy, MR, Li, Y, et al. (2001) The heterocyclic ring fission and dehydroxylation of catechins and related compounds by Eubacterium sp. strain SDG-2, a human intestinal bacterium. Chem Pharm Bull 49, 16401643.CrossRefGoogle ScholarPubMed
82 Winter, J, Moore, LH, Dowell, VR, et al. (1989) C-ring cleavage of flavonoids by human intestinal bacteria. Appl Environ Microbiol 55, 12031208.CrossRefGoogle ScholarPubMed
83 Winter, J, Popoff, MR, Grimont, P, et al. (1991) Clostridium orbiscindens sp. nova., human intestinal bacterium capable of cleaving the flavonoid C-ring. Int J Syst Bacteriol 41, 355357.CrossRefGoogle Scholar
84 Jang, IS & Kim, DH (1996) Purification and characterization of alpha-L-rhamnosidase from Bacteroides JY-6, a human intestinal bacterium. Biol Pharm Bull 19, 15461549.CrossRefGoogle ScholarPubMed
85 Bokkenheuser, VD, Shackleton, CH & Winter, J (1987) Hydrolysis of dietary flavonoid glycosides by strains of intestinal Bacteroides from humans. Biochem J 248, 953956.CrossRefGoogle ScholarPubMed
86 Kim, D-H, Sohng, IS, Kobashi, K, et al. (1996) Purification and characterization of β-glucosidase from Bacteroides JY-6, a human intestinal bacterium. Biol Pharm Bull 19, 11211125.CrossRefGoogle ScholarPubMed
87 Kim, DH, Kim, SY, Park, SY, et al. (1999) Metabolism of quercitrin by human intestinal bacteria and its relation to some biological activities. Biol Pharm Bull 22, 749751.CrossRefGoogle ScholarPubMed
88 Sung, CK, Kang, GH, Yoon, SS, et al. (1996) Glycosidases that convert natural glycosides to bioactive compounds. Adv Exp Med Biol 404, 2336.CrossRefGoogle ScholarPubMed
89 Justesen, U, Arrigoni, E, Larsen, BR, et al. (2000) Degradation of flavonoid glycosides and aglycones during in vitro fermentation with human faecal flora. Lebensmittel Wissenschaft Technol 33, 424430.CrossRefGoogle Scholar
90 Kim, DH, Jung, EA, Sohng, IS, et al. (1998) Intestinal bacterial metabolism of flavonoids and its relation to some biological activities. Arch Pharm Res 21, 1723.CrossRefGoogle ScholarPubMed
91 Kay, CD, Kroon, PA & Cassidy, A (2009) The bioactivity of dietary anthocyanins is likely to be mediated by their degradation products. Mol Nutr Food Res 53, Suppl. 1, S92101.CrossRefGoogle ScholarPubMed
92 Peppercorn, MA & Goldman, P (1971) Caffeic acid metabolism by bacteria of the human gastrointestinal tract. J Bact 108, 9961000.CrossRefGoogle ScholarPubMed
93 Indahl, SR & Scheline, RR (1968) Decarboxylation of 4-hydroxycinnamic acids by Bacillus strains isolated from rat intestine. Appl Microbiol 16, 667.CrossRefGoogle ScholarPubMed
94 Groenewoud, G & Hundt, HKL (1984) The microbial metabolism of (+)-catechins to two novel diarylpropan-2-ol metabolites in vitro. Xenobiotica 14, 711717.CrossRefGoogle ScholarPubMed
95 Meselhy, MR, Nakamura, N & Hattori, M (1997) Biotransformation of ( − )-epicatechin 3-O-gallate by human intestinal bacteria. Chem Pharm Bull 45, 888893.CrossRefGoogle ScholarPubMed
96 Doyle, B & Griffiths, LA (1980) The metabolism of ellagic acid in the rat. Xenobiotica 10, 247256.CrossRefGoogle Scholar
97 Cerda, B, Llorach, R, Ceron, JJ, et al. (2003) Evaluation of the bioavailability and metabolism in the rat of punicalagin, an antioxidant polyphenol from pomegranate juice. Eur J Nutr 42, 1828.CrossRefGoogle ScholarPubMed
98 Larrosa, M, Gonzalez-Sarrias, A, Garcia-Conesa, MT, et al. (2006) Urolithins, ellagic acid-derived metabolites produced by human colonic microflora, exhibit estrogenic and antiestrogenic activities. J Agric Food Chem 54, 16111620.CrossRefGoogle ScholarPubMed
99 Seeram, NP, Aronson, WJ, Zhang, Y, et al. (2007) Pomegranate ellagitannin-derived metabolites inhibit prostate cancer growth and localize to the mouse prostate gland. J Agric Food Chem 55, 77327737.CrossRefGoogle Scholar
100 Day, AJ, Cañada, FJ, Díaz, JC, et al. (2000) Dietary flavonoid and isoflavone glycosides are hydrolysed by the lactase site of lactase phlorizin hydrolase. FEBS Lett 468, 166170.CrossRefGoogle ScholarPubMed
101 Hollman, PCH, Buijsman, MNCP, van Gameren, Y, et al. (1999) The sugar moiety is a major determinant of the absorption of the dietary flavonoid glycosides in man. Free Radic Res 31, 569573.CrossRefGoogle Scholar
102 Walle, T, Otake, Y, Walle, UK, et al. (2000) Quercetin glucosides are completely hydrolyzed in ileostomy patients before absorption. J Nutr 130, 26582661.CrossRefGoogle ScholarPubMed
103 Hollman, PCH, de Vries, JHM, van Leeuwen, SD, et al. (1995) Absorption of dietary quercetin glycosides and quercetin in healthy ileostomy volunteers. Am J Clin Nutr 62, 12761282.CrossRefGoogle ScholarPubMed
104 Olthof, MR, Hollman, PC, Buijsman, MN, et al. (2003) Chlorogenic acid, quercetin-3-rutinoside and black tea phenols are extensively metabolized in humans. J Nutr 133, 18061814.CrossRefGoogle ScholarPubMed
105 Auger, C, Mullen, W, Hara, Y, et al. (2008) Bioavailability of polyphenon E flavan-3-ols in humans with an ileostomy. J Nutr 138, 1535S1542S.CrossRefGoogle ScholarPubMed
106 Kahle, K, Kraus, M, Scheppach, W, et al. (2005) Colonic availability of apple polyphenols – a study in ileostomy subjects. Mol Nutr Food Res 49, 11431150.CrossRefGoogle ScholarPubMed
107 Kahle, K, Kraus, M, Scheppach, W, et al. (2006) Studies on apple and blueberry fruit constituents: do the polyphenols reach the colon after ingestion? Mol Nutr Food Res 50, 418423.CrossRefGoogle ScholarPubMed
108 Marks, SC, Mullen, W, Borges, G, et al. (2009) Absorption, metabolism, and excretion of cider dihydrochalcones in healthy humans and subjects with an ileostomy. J Agric Food Chem 57, 20092015.CrossRefGoogle ScholarPubMed
109 Poquet, L, Clifford, MN & Williamson, G (2008) Transport and metabolism of ferulic acid through the colonic epithelium. Drug Metab Dispos 36, 190197.CrossRefGoogle ScholarPubMed
110 Hayeshi, R, Hilgendorf, C, Artursson, P, et al. (2008) Comparison of drug transporter gene expression and functionality in Caco-2 cells from 10 different laboratories. Eur J Pharm Sci 35, 383396.CrossRefGoogle ScholarPubMed
111 Gutmann, H, Fricker, G, Torok, M, et al. (1999) Evidence for different ABC-transporters in Caco-2 cells modulating drug uptake. Pharm Res 16, 402407.CrossRefGoogle ScholarPubMed
112 Sambuy, Y, De Angelis, I, Ranaldi, G, et al. (2005) The Caco-2 cell line as a model of the intestinal barrier: influence of cell and culture-related factors on Caco-2 cell functional characteristics. Cell Biol Toxicol 21, 126.CrossRefGoogle Scholar
113 Lindsay, DG & Clifford, MN (2000) Special issue devoted to critical reviews produced within the EU Concerted Action ‘Nutritional Enhancement of Plant-based Food in European Trade’ (NEODIET). J Sci Food Agric 80, 7931137.3.0.CO;2-P>CrossRefGoogle Scholar
114 Crozier, A, Jaganath, IB & Clifford, MN (2006) Phenols, polyphenols and tannins: an overview. In Plant Secondary Metabolites. Occurrence, Structure and Role in the Human Diet, pp. 124 [Crozier, A, Clifford, MN and Ashihara, H, editors]. Oxford: Blackwell.CrossRefGoogle Scholar
115 Crozier, A, Yokota, T, Jaganath, IB, et al. (2006) Secondary metabolites in fruits, vegetables, beverages and other plant-based dietary components. In Plant Secondary Metabolites. Occurrence, Structure and Role in the Human Diet, pp. 208302 [Crozier, A, Clifford, MN and Ashihara, H, editors]. Oxford: Blackwell.CrossRefGoogle Scholar
116 Maatta, KR, Kamal-Eldin, A & Torronen, AR (2003) High-performance liquid chromatography (HPLC) analysis of phenolic compounds in berries with diode array and electrospray ionization mass spectrometric (MS) detection: Ribes species. J Agric Food Chem 51, 67366744.CrossRefGoogle ScholarPubMed
117 Gu, L, Kelm, MA, Hammerstone, JF, et al. (2004) Concentrations of proanthocyanidins in common foods and estimations of normal consumption. J Nutr 134, 613617.Google ScholarPubMed
118 Hodgson, JM, Morton, LW, Puddey, IB, et al. (2000) Gallic acid metabolites are markers of black tea intake in humans. J Agric Food Chem 48, 22762280.CrossRefGoogle ScholarPubMed
119 Shahrzad, S & Bitsch, I (1998) Determination of gallic acid and its metabolites in human plasma and urine by high-performance liquid chromatography. J Chrom B Biomed Sci Appl 705, 8795.CrossRefGoogle ScholarPubMed
120 Matsumoto, H, Inaba, H, Kishi, M, et al. (2001) Orally administered delphinidin 3-rutinoside and cyanidin 3-rutinoside are directly absorbed in rats and humans and appear in the blood as the intact forms. J Agric Food Chem 49, 15461551.CrossRefGoogle ScholarPubMed
121 Matsumoto, H, Ito, K, Yonekura, K, et al. (2007) Enhanced absorption of anthocyanins after oral administration of phytic acid in rats and humans. J Agric Food Chem 55, 24892496.CrossRefGoogle ScholarPubMed
122 Felgines, C, Talavera, S, Gonthier, MP, et al. (2003) Strawberry anthocyanins are recovered in urine as glucuro- and sulfoconjugates in humans. J Nutr 133, 12961301.CrossRefGoogle ScholarPubMed
123 Mullen, W, Edwards, CA, Serafini, M, et al. (2008) Bioavailability of pelargonidin-3-O-glucoside and its metabolites in humans following the ingestion of strawberries with and without cream. J Agric Food Chem 56, 713719.CrossRefGoogle ScholarPubMed
124 Tsuda, T, Horio, F & Osawa, T (1999) Absorption and metabolism of cyanidin 3-O-β-d-glucoside in rats. FEBS Lett 449, 179182.CrossRefGoogle ScholarPubMed
125 Wu, X, PittmanIii, HE, Hager, T, et al. (2009) Phenolic acids in black raspberry and in the gastrointestinal tract of pigs following ingestion of black raspberry. Mol Nutr Food Res 53, Suppl. 1, S76S84.CrossRefGoogle ScholarPubMed
126 Russell, WR, Scobbie, L, Labat, A, et al. (2009) Selective bio-availability of phenolic acids from Scottish strawberries. Mol Nutr Food Res 53, Suppl. 1, S85S91.CrossRefGoogle ScholarPubMed
127 Hollands, W, Brett, GM, Dainty, JR, et al. (2008) Urinary excretion of strawberry anthocyanins is dose dependent for physiological oral doses of fresh fruit. Mol Nutr Food Res 52, 10971105.CrossRefGoogle ScholarPubMed
128 Stohr, H & Herrmann, K (1975) The phenolics of strawberries and their changes during development and ripeness of the fruits. Z Lebensm Unters -Forsch 158, 341348.Google ScholarPubMed
129 Ichiyanagi, T, Kashiwada, Y, Ikeshiro, Y, et al. (2004) Complete assignment of bilberry (Vaccinium myrtillus L.) anthocyanins separated by capillary zone electrophoresis. Chem Pharm Bull (Tokyo) 52, 226229.CrossRefGoogle ScholarPubMed
130 Katsube, N, Iwashita, K, Tsushida, T, et al. (2003) Induction of apoptosis in cancer cells by bilberry (Vaccinium myrtillus) and the anthocyanins. J Agric Food Chem 51, 6875.CrossRefGoogle ScholarPubMed
131 Ichiyanagi, T, Hatano, Y, Matsugo, S, et al. (2004) Structural dependence of HPLC separation pattern of anthocyanins from Bilberry (Vaccinium myrtillus L.). Chem Pharm Bull (Tokyo) 52, 628630.CrossRefGoogle ScholarPubMed
132 Ek, S, Kartimo, H, Mattila, S, et al. (2006) Characterization of phenolic compounds from lingonberry (Vaccinium vitis-idaea). J Agric Food Chem 54, 98349842.CrossRefGoogle ScholarPubMed
133 Gonthier, MP, Donovan, JL, Texier, O, et al. (2003) Metabolism of dietary procyanidins in rats. Free Radic Biol Med 35, 837844.CrossRefGoogle ScholarPubMed
134 Stoupi, S, Williamson, G, Viton, S, et al. (2010) In vivo bioavailability, absorption, excretion and pharmacokinetics of [14C]-procyanidin B2 in male rats. Drug Metab Dispos 38, 287291.CrossRefGoogle ScholarPubMed
135 Rios, LY, Gonthier, MP, Remesy, C, et al. (2003) Chocolate intake increases urinary excretion of polyphenol-derived phenolic acids in healthy human subjects. Am J Clin Nutr 77, 912918.CrossRefGoogle ScholarPubMed
136 Stalmach, A, Mullen, W, Barron, D, et al. (2009) Metabolite profiling of methyl, glucuronyl and sulfate conjugates in plasma and urine derived from chlorogenic acids following the ingestion of coffee by humans: identification of biomarkers of coffee consumption. Drug Metab Dispos 37, 17491758.CrossRefGoogle Scholar
137 Renouf, M, Guy, PA, Marmet, C, et al. (2010) Measurement of caffeic and ferulic acid equivalents in plasma after coffee consumption: small intestine and colon are key sites for coffee metabolism. Mol Nutr Food Res 54, 760766.CrossRefGoogle ScholarPubMed
138 Olthof, MR, Hollman, PCH & Katan, MB (2001) Chlorogenic acid and caffeic acid are absorbed in humans. J Nutr 131, 6671.CrossRefGoogle ScholarPubMed
139 Goldstein, DS, Stull, R, Markey, SP, et al. (1984) Dihydrocaffeic acid: a common contaminant in the liquid chromatographic-electrochemical measurement of plasma catecholamines in man. J Chromatogr 311, 148153.CrossRefGoogle ScholarPubMed
140 Wittemer, SM, Ploch, M, Windeck, T, et al. (2005) Bioavailability and pharmacokinetics of caffeoylquinic acids and flavonoids after oral administration of artichoke leaf extracts in humans. Phytomed 12, 2838.CrossRefGoogle ScholarPubMed
141 Gonthier, MP, Cheynier, V, Donovan, JL, et al. (2003) Microbial aromatic acid metabolites formed in the gut account for a major fraction of the polyphenols excreted in urine of rats fed red wine polyphenols. J Nutr 133, 461467.CrossRefGoogle Scholar
142 Gonthier, MP, Verny, MA, Besson, C, et al. (2003) Chlorogenic acid bioavailability largely depends on its metabolism by the gut microflora in rats. J Nutr 133, 18531859.CrossRefGoogle ScholarPubMed
143 Teuchy, H & van Sumere, CF (1971) The metabolism of [1-14C] phenylalanine, [3-14C] cinnamic acid and [2-14C] ferulic acid in the rat. Arch Int Physiol Biochim 79, 589618.Google Scholar
144 Nutley, BP, Farmer, P & Caldwell, J (1994) Metabolism of trans-cinnamic acid in the rat and the mouse and its variation with dose. Food Chem Toxicol 32, 877886.CrossRefGoogle ScholarPubMed
145 Haughton, E, Clifford, MN & Sharp, P (2007) Monocarboxylate transporter expression is associated with the absorption of benzoic acid in human intestinal epithelial cells. J Sci Food Agric 87, 239244.CrossRefGoogle Scholar
146 Mori, S, Takanaga, H, Ohtsuki, S, et al. (2003) Rat organic anion transporter 3 (rOAT3) is responsible for brain-to-blood efflux of homovanillic acid at the abluminal membrane of brain capillary endothelial cells. J Cereb Blood Flow Metab 23, 432440.CrossRefGoogle Scholar
147 Cao, YG, Zhang, L, Ma, C, et al. (2009) Metabolism of protocatechuic acid influences fatty acid oxidation in rat heart: new anti-angina mechanism implication. Biochem Pharmacol 77, 10961104.CrossRefGoogle ScholarPubMed
148 Liu, HX, Liu, Y, Zhang, JW, et al. (2008) UDP-glucuronosyltransferase 1A6 is the major isozyme responsible for protocatechuic aldehyde glucuronidation in human liver microsomes. Drug Metab Dispos 36, 15621569.CrossRefGoogle ScholarPubMed
149 Panoutsopoulos, GI & Beedham, C (2005) Enzymatic oxidation of vanillin, isovanillin and protocatechuic aldehyde with freshly prepared Guinea pig liver slices. Cell Physiol Biochem 15, 8998.CrossRefGoogle ScholarPubMed
150 Tanaka, T, Kawamori, T, Ohnishi, M, et al. (1994) Chemoprevention of 4-nitroquinoline 1-oxide-induced oral carcinogenesis by dietary protocatechuic acid during initiation and postinitiation phases. Cancer Res 54, 23592365.Google ScholarPubMed
151 Tanaka, T, Kojima, T, Kawamori, T, et al. (1995) Chemoprevention of digestive organs carcinogenesis by natural product protocatechuic acid. Cancer 75, 14331439.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
152 Tarozzi, A, Morroni, F, Hrelia, S, et al. (2007) Neuroprotective effects of anthocyanins and their in vivo metabolites in SH-SY5Y cells. Neurosci Lett 424, 3640.CrossRefGoogle ScholarPubMed
153 Yip, EC, Chan, AS, Pang, H, et al. (2006) Protocatechuic acid induces cell death in HepG2 hepatocellular carcinoma cells through a c-Jun N-terminal kinase-dependent mechanism. Cell Biol Toxicol 22, 293302.CrossRefGoogle ScholarPubMed
154 Wang, H, Liu, TQ, Guan, S, et al. (2008) Protocatechuic acid from Alpinia oxyphylla promotes migration of human adipose tissue-derived stromal cells in vitro. Eur J Pharmacol 599, 2431.CrossRefGoogle ScholarPubMed
155 Guan, S, Ge, D, Liu, TQ, et al. (2009) Protocatechuic acid promotes cell proliferation and reduces basal apoptosis in cultured neural stem cells. Toxicol In Vitro 23, 201208.CrossRefGoogle ScholarPubMed
156 Tamura, A, Fukushima, M, Shimada, K, et al. (2004) Cholesterol metabolism in rat is affected by protocatechuic acid. J Nutr Sci Vitaminol (Tokyo) 50, 1318.Google ScholarPubMed
157 Thompson, P, Balis, F, Serabe, BM, et al. (2003) Pharmacokinetics of phenylacetate administered as a 30-min infusion in children with refractory cancer. Cancer Chemother Pharmacol 52, 417423.CrossRefGoogle ScholarPubMed
158 Wajngot, A, Chandramouli, V, Schumann, WC, et al. (2000) A probing dose of phenylacetate does not affect glucose production and gluconeogenesis in humans. Metabolism 49, 12111214.CrossRefGoogle Scholar
159 Inoue, O, Hosoi, R, Momosaki, S, et al. (2006) Evaluation of [14C]phenylacetate as a prototype tracer for the measurement of glial metabolism in the rat brain. Nucl Med Biol 33, 985989.CrossRefGoogle ScholarPubMed
160 Thibault, A, Samid, D, Cooper, MR, et al. (1995) Phase I study of phenylacetate administered twice daily to patients with cancer. Cancer 75, 29322938.3.0.CO;2-P>CrossRefGoogle ScholarPubMed
161 James, MO & Smith, RL (2009) The conjugation of phenylacetic acid in phenylketonurics. Eur J Clin Pharmacol 5, 243246.CrossRefGoogle Scholar
162 Karlsson, PC, Huss, U, Jenner, A, et al. (2005) Human fecal water inhibits COX-2 in colonic HT-29 cells: role of phenolic compounds. J Nutr 135, 23432349.CrossRefGoogle ScholarPubMed
163 MacArthur, RB, Altincatal, A & Tuchman, M (2004) Pharmacokinetics of sodium phenylacetate and sodium benzoate following intravenous administration as both a bolus and continuous infusion to healthy adult volunteers. Mol Genet Metab 81, Suppl. 1, S67S73.CrossRefGoogle ScholarPubMed
164 Kang, HL, Benzer, S & Min, KT (2002) Life extension in Drosophila by feeding a drug. Proc Natl Acad Sci U S A 99, 838843.CrossRefGoogle ScholarPubMed
165 Kasumov, T, Brunengraber, LL, Comte, B, et al. (2004) New secondary metabolites of phenylbutyrate in humans and rats. Drug Metab Dispos 32, 1019.CrossRefGoogle ScholarPubMed
166 Li, XN, Parikh, S, Shu, Q, et al. (2004) Phenylbutyrate and phenylacetate induce differentiation and inhibit proliferation of human medulloblastoma cells. Clin Cancer Res 10, 11501159.CrossRefGoogle ScholarPubMed
167 Franco, OE, Onishi, T, Umeda, Y, et al. (2003) Phenylacetate inhibits growth and modulates cell cycle gene expression in renal cancer cell lines. Anticancer Res 23, 16371642.Google ScholarPubMed
168 Shibahara, T, Onishi, T, Franco, OE, et al. (2005) Down-regulation of Skp2 is correlated with p27-associated cell cycle arrest induced by phenylacetate in human prostate cancer cells. Anticancer Res 25, 18811888.Google ScholarPubMed
169 Clifford, MN (2000) Chlorogenic acids and other cinnamates – nature, occurrence, dietary burden, absorption and metabolism. J Sci Food Agric 80, 10331042.3.0.CO;2-T>CrossRefGoogle Scholar
170 Clifford, MN (1999) Chlorogenic acids and other cinnamates – nature, occurrence and dietary burden. J Sci Food Agric 79, 362372.3.0.CO;2-D>CrossRefGoogle Scholar
171 Fallingborg, J (1999) Intraluminal pH of the human gastrointestinal tract. Dan Med Bull 46, 183196.Google ScholarPubMed
172 Jovanovic, SV, Steenken, S, Tosic, M, et al. (1994) Flavonoids as antioxidants. J Am Chem Soc 116, 48464851.CrossRefGoogle Scholar
173 Poquet, L, Clifford, MN & Williamson, G (2008) Investigation of the metabolic fate of dihydrocaffeic acid. Biochem Pharmacol 75, 12181229.CrossRefGoogle ScholarPubMed
174 Konishi, Y & Kobayashi, S (2004) Microbial metabolites of ingested caffeic acid are absorbed by the monocarboxylic acid transporter (MCT) in intestinal Caco-2 cell monolayers. J Agric Food Chem 52, 64186424.CrossRefGoogle ScholarPubMed
175 Konishi, Y & Kobayashi, S (2004) Transepithelial transport of chlorogenic acid, caffeic acid, and their colonic metabolites in intestinal caco-2 cell monolayers. J Agric Food Chem 52, 25182526.CrossRefGoogle ScholarPubMed
176 Wong, CC, Meinl, W, Glatt, H-R, et al. (2010) In vitro and in vivo conjugation of dietary hydroxycinnamic acids by UDP-glucuronosyltransferases and sulfotransferases in humans. J Nutr Biochem, (Epublication ahead of print version).Google ScholarPubMed
177 Poquet, L, Clifford, MN & Williamson, G (2008) Effect of dihydrocaffeic acid on UV irradiation of human keratinocyte HaCaT cells. Arch Biochem Biophys 476, 196204.CrossRefGoogle ScholarPubMed
178 Larrosa, M, Lodovici, M, Morbidelli, L, et al. (2008) Hydrocaffeic and p-coumaric acids, natural phenolic compounds, inhibit UV-B damage in WKD human conjunctival cells in vitro and rabbit eye in vivo. Free Radic Res 42, 903910.CrossRefGoogle ScholarPubMed
179 Guglielmi, F, Luceri, C, Giovannelli, L, et al. (2003) Effect of 4-coumaric and 3,4-dihydroxybenzoic acid on oxidative DNA damage in rat colonic mucosa. Br J Nutr 89, 581587.CrossRefGoogle ScholarPubMed
180 Larrosa, M, Luceri, C, Vivoli, E, et al. (2009) Polyphenol metabolites from colonic microbiota exert anti-inflammatory activity on different inflammation models. Mol Nutr Food Res 53, 10441054.CrossRefGoogle ScholarPubMed
181 Bhat, CS & Ramasarma, T (1979) Effect of phenyl and phenolic acids on mevalonate-5-phosphate kinase and mevalonate-5-pyrophosphate decarboxylase of the rat brain. J Neurochem 32, 15311537.CrossRefGoogle ScholarPubMed
182 Lee, JS, Choi, MS, Jeon, SM, et al. (2001) Lipid-lowering and antioxidative activities of 3,4-di(OH)-cinnamate and 3,4-di(OH)-hydrocinnamate in cholesterol-fed rats. Clin Chim Acta 314, 221229.CrossRefGoogle ScholarPubMed
183 Yasuda, T, Takasawa, A, Nakazawa, T, et al. (2003) Inhibitory effects of urinary metabolites on platelet aggregation after orally administering Shimotsu-To, a traditional Chinese medicine, to rats. J Pharm Pharmacol 55, 239244.CrossRefGoogle ScholarPubMed
184 Bok, SH, Lee, SH, Park, YB, et al. (1999) Plasma and hepatic cholesterol and hepatic activities of 3-hydroxy-3-methyl-glutaryl-CoA reductase and acyl CoA: cholesterol transferase are lower in rats fed citrus peel extract or a mixture of citrus bioflavonoids. J Nutr 129, 11821185.CrossRefGoogle ScholarPubMed
185 Kim, HK, Jeong, TS, Lee, MK, et al. (2003) Lipid-lowering efficacy of hesperetin metabolites in high-cholesterol fed rats. Clin Chim Acta 327, 129137.CrossRefGoogle ScholarPubMed
186 Lee, CH, Jeong, TS, Choi, YK, et al. (2001) Anti-atherogenic effect of citrus flavonoids, naringin and naringenin, associated with hepatic ACAT and aortic VCAM-1 and MCP-1 in high cholesterol-fed rabbits. Biochem Biophys Res Commun 284, 681688.CrossRefGoogle ScholarPubMed
187 Matsumoto, N, Okushio, K & Hara, Y (1998) Effect of black tea polyphenols on plasma lipids in cholesterol-fed rats. J Nutr Sci Vitaminol 44, 337342.CrossRefGoogle ScholarPubMed
188 Yamakoshi, J, Kataoka, S, Koga, T, et al. (1999) Proanthocyanidin-rich extract from grape seeds attenuates the development of aortic atherosclerosis in cholesterol-fed rabbits. Atherosclerosis 142, 139149.CrossRefGoogle ScholarPubMed
189 Kurowska, EM, Spence, JD, Jordan, J, et al. (2000) HDL-cholesterol-raising effect of orange juice in subjects with hypercholesterolemia. Am J Clin Nutr 72, 10951100.CrossRefGoogle ScholarPubMed
190 Mo, H & Elson, CE (2004) Studies of the isoprenoid-mediated inhibition of mevalonate synthesis applied to cancer chemotherapy and chemoprevention. Exp Biol Med (Maywood) 229, 567585.CrossRefGoogle ScholarPubMed
191 Lee, MJ, Maliakal, P, Chen, L, et al. (2002) Pharmacokinetics of tea catechins after ingestion of green tea and ( − )-epigallocatechin-3-gallate by humans: formation of different metabolites and individual variability. Cancer Epidemiol Biomarkers Prev 11, 10251032.Google ScholarPubMed
192 Sun, CL, Yuan, JM, Lee, MJ, et al. (2002) Urinary tea polyphenols in relation to gastric and esophageal cancers: a prospective study of men in Shanghai, China. Carcinogen 23, 14971503.CrossRefGoogle ScholarPubMed
193 Duweler, KG & Rohdewald, P (2000) Urinary metabolites of French maritime pine bark extract in humans. Pharmazie 55, 364368.Google ScholarPubMed
194 Goto, K, Kanaya, S, Nishikawa, T, et al. (1998) The influence of tea catechins on fecal flora of elderly residents in long-term care facilities. Ann Long-Term Care 6, 4348.Google Scholar
195 Hara, Y (1997) Influence of tea catechins on the digestive tract. J Cell Biochem 27, Suppl. 5258.3.0.CO;2-N>CrossRefGoogle ScholarPubMed
196 Okubo, T, Ishihara, N, Oura, A, et al. (1992) In vivo effect of tea polyphenol intake on human intestinal microflora and metabolism. Biosci Biotechnol Biochem 56, 588591.CrossRefGoogle ScholarPubMed
197 Min, BR, Attwood, GT, Reilly, K, et al. (2002) Lotus corniculatus condensed tannins decrease in vivo populations of proteolytic bacteria and affect nitrogen metabolism in the rumen of sheep. Can J Microbiol 48, 911921.CrossRefGoogle ScholarPubMed
198 Ramos-Niño, ME, Ramìrez-Rodriguez, CA, Clifford, MN, et al. (1997) A comparison of quantitative structure-activity relationships for the effect of benzoic and cinnamic acids on Listeria monocytogenes using multiple linear regression, artificial neural networks and fuzzy systems. J Appl Microbiol 82, 168176.CrossRefGoogle ScholarPubMed
199 Gargala, G, Baishanbo, A, Favennec, L, et al. (2005) Inhibitory activities of epidermal growth factor receptor tyrosine kinase-targeted dihydroxyisoflavone and trihydroxydeoxybenzoin derivatives on Sarcocystis neurona, Neospora caninum, and Cryptosporidium parvum development. Antimicrob Agents Chemother 49, 46284634.CrossRefGoogle ScholarPubMed
200 Kerboeuf, D, Riou, M & Guegnard, F (2008) Flavonoids and related compounds in parasitic disease control. Mini Rev Med Chem 8, 116128.CrossRefGoogle ScholarPubMed
201 Li, C, Lee, M-J, Sheng, S, et al. (2000) Stuctural identification of two metabolites of catechins and their kinetics in human urine and blood after tea ingestion. Chem Res Toxicol 13, 177184.CrossRefGoogle Scholar
202 Clifford, MN, Copeland, EL, Bloxsidge, JP, et al. (2000) Hippuric acid is a major excretion product associated with black tea consumption. Xenobiotica 30, 317326.CrossRefGoogle Scholar
203 Krumholz, LR & Bryant, MP (1986) Eubacterium oxidoreducens sp. nov. requiring H2 or formate to degrade gallate, pyrogallol, phloroglucinol and quercetin. 144, 814.Google Scholar
204 Booth, AN & Williams, RT (1963) Dehydroxylation of caffeic acid by rat and rabbit caecal contents and sheep rumen liquor. Nature 684685.CrossRefGoogle ScholarPubMed
205 Dayman, J & Jepson, JB (1969) The metabolism of caffeic acid in humans: the dehydroxylating action of intestinal bacteria. Biochem J 113, 11P.CrossRefGoogle Scholar
206 Plumb, GW, García-Conesa, MT, Kroon, PA, et al. (1999) Metabolism of chlorogenic acid by human plasma, liver, intestine and gut microflora. J Sci Food Agric 79, 390392.3.0.CO;2-0>CrossRefGoogle Scholar
207 Couteau, D, McCartney, AL, Gibson, GR, et al. (2001) Isolation and characterization of human colonic bacteria able to hydrolyse chlorogenic acid. J Appl Microbiol 90, 873881.CrossRefGoogle ScholarPubMed
208 Rechner, AR, Smith, MA, Kuhnle, G, et al. (2004) Colonic metabolism of dietary polyphenols: influence of structure on microbial fermentation products. Free Radic Biol Med 36, 212225.CrossRefGoogle ScholarPubMed
209 Gavaghan, CL, Nicholson, JK, Connor, SC, et al. (2001) Directly coupled high-performance liquid chromatography and nuclear magnetic resonance spectroscopic with chemometric studies on metabolic variation in Sprague–Dawley rats. Anal Biochem 291, 245252.CrossRefGoogle ScholarPubMed
210 Stalmach, A, Mullen, W, Barron, D, et al. (2009) Metabolite profiling of hydroxycinnamate derivatives in plasma and urine after the ingestion of coffee by humans: identification of biomarkers of coffee consumption. Drug Metab Dispos 37, 17491758.CrossRefGoogle ScholarPubMed
211 Kahle, K, Huemmer, W, Kempf, M, et al. (2007) Polyphenols are intensively metabolized in the human gastrointestinal tract after apple juice consumption. J Agric Food Chem 55, 1060510614.CrossRefGoogle ScholarPubMed
212 Das, NP (1969) Studies on flavonoid metabolism. Degradation of (+)-catechin by rat intestinal contents. Biochim Biophys Acta 177, 668670.CrossRefGoogle ScholarPubMed
213 Scheline, RR (1970) The metabolism of (+)-catechin to hydroxyphenylvaleric acids by the intestinal microflora. Biochim Biophys Acta 222, 228230.CrossRefGoogle Scholar
214 Oshima, Y & Watanabe, H (1958) The mechanisms of catechin metabolism. 2. Neutral substances in the urine of rabbits administered (+)-catechin. J Biochem 45, 973977.CrossRefGoogle Scholar
215 Das, NP & Griffiths, LA (1968) Studies on flavonoid metabolism. Metabolism of (+)-catechin in the guinea pig. Biochem J 110, 449456.CrossRefGoogle ScholarPubMed
216 Das, NP (1974) Studies on flavonoid metabolism. Excretion of m-hydroxyphenylhydracrylic acid from (+)-catechin in the monkey (Macaca iris sp.). Drug Metab Dispos 2, 209213.Google ScholarPubMed
217 Griffiths, LA & Barrow, A (1964) Metabolism of flavonoid compounds in germ-free rats. Biochem J 92, 173179.CrossRefGoogle Scholar
218 Hattori, S & Noguchi, I (1959) Microbial degradation of rutin. Nature 184, 11451146.CrossRefGoogle ScholarPubMed
219 Cheng, KJ, Jones, GA, Simpson, FJ, et al. (1969) Isolation and identification of rumen bacteria capable of anaerobic rutin degradation. Can J Microbiol 15, 13651371.CrossRefGoogle ScholarPubMed
220 Krishnamurty, HG, Cheng, KJ, Jones, GA, et al. (1970) Identification of products produced by the anaerobic degradation of rutin and related flavonoids by Butyrivibrio sp. C3. Can J Microbiol 16, 759767.CrossRefGoogle ScholarPubMed
221 Baba, S, Furuta, T, Fujioka, M, et al. (1983) Studies on drug metabolism by use of isotopes XXVII: urinary metabolites of rutin in rats and the role of intestinal microflora in the metabolism of rutin. J Pharm Sci 72, 11551158.CrossRefGoogle Scholar
222 Mullen, W, Rouanet, JM, Auger, C, et al. (2008) Bioavailability of [2-14C]quercetin-4′-glucoside in rats. J Agric Food Chem 56, 1212712137.CrossRefGoogle ScholarPubMed
223 Jaganath, IB, Mullen, W, Edwards, CA, et al. (2006) The relative contribution of the small and large intestine to the absorption and metabolism of rutin in man. Free Radic Res 40, 10351046.CrossRefGoogle Scholar
224 Coldham, NG, Darby, C, Hows, M, et al. (2002) Comparative metabolism of genistin by human and rat gut microflora: detection and identification of the end-products of metabolism. Xenobiotica 32, 4562.CrossRefGoogle ScholarPubMed
225 Tsangalis, D, Ashton, JE, McGill, AEJ, et al. (2002) Enzymic transformation of isoflavone phytoestrogens in soymilk by β-glucosidase-producing Bifidobacteria. J Food Sci 67, 31043113.CrossRefGoogle Scholar
226 Steer, TE, Johnson, IT, Gee, JM, et al. (2003) Metabolism of the soybean isoflavone glycoside genistin in vitro by human gut bacteria and the effect of prebiotics. Br J Nutr 90, 635642.CrossRefGoogle ScholarPubMed
227 Atkinson, C, Berman, S, Humbert, O, et al. (2004) In vitro incubation of human feces with daidzein and antibiotics suggests interindividual differences in the bacteria responsible for equol production. J Nutr 134, 596599.CrossRefGoogle ScholarPubMed
228 Liang, G, Zhang, T, Wang, J, et al. (2005) X-ray single-crystal analysis of ( − )-(S)-equol isolated from rat's feces. Chem Biodivers 2, 959963.CrossRefGoogle ScholarPubMed
229 Simons, AL, Renouf, M, Hendrich, S, et al. (2005) Metabolism of glycitein (7,4′-dihydroxy-6-methoxy-isoflavone) by human gut microflora. J Agric Food Chem 53, 85198525.CrossRefGoogle ScholarPubMed
230 Otieno, DO, Ashton, JF & Shah, NP (2006) Evaluation of enzymic potential for biotransformation of isoflavone phytoestrogen in soymilk by Bifidobacterium animalis, Lactobacillus acidophilus and Lactobacillus casei. Food Res Int 39, 394407.CrossRefGoogle Scholar
231 Tamura, M, Tsushida, T & Shinohara, K (2007) Isolation of an isoflavone-metabolizing, Clostridium-like bacterium, strain TM-40, from human faeces. Anaerobe 13, 3235.CrossRefGoogle ScholarPubMed
232 Bowey, E, Adlercreutz, H & Rowland, I (2003) Metabolism of isoflavones and lignans by the gut microflora: a study in germ-free and human flora associated rats. Food Chem Toxicol 41, 631636.CrossRefGoogle ScholarPubMed
233 Kelly, GE, Nelson, C, Waring, MA, et al. (1993) Metabolites of dietary (soya) isoflavones in human urine. Clin Chim Acta 223, 922.CrossRefGoogle ScholarPubMed
234 Xu, X, Harris, KS, Wang, HJ, et al. (1995) Bioavailability of soybean isoflavones depends upon gut microflora in women. J Nutr 125, 23072315.CrossRefGoogle ScholarPubMed
235 Heinonen, S, Wahala, K & Adlercreutz, H (1999) Identification of isoflavone metabolites dihydrodaidzein, dihydrogenistein, 6′-OH-O-DMA, and cis-4-OH-equol in human urine by gas chromatography–mass spectroscopy using authentic reference compounds. Anal Biochem 274, 211219.CrossRefGoogle Scholar
236 Rowland, IR, Wiseman, H, Sanders, TA, et al. (2000) Interindividual variation in metabolism of soy isoflavones and lignans: influence of habitual diet on equol production by the gut microflora. Nutr Cancer 36, 2732.CrossRefGoogle ScholarPubMed
237 Hoey, L, Rowland, IR, Lloyd, AS, et al. (2004) Influence of soya-based infant formula consumption on isoflavone and gut microflora metabolite concentrations in urine and on faecal microflora composition and metabolic activity in infants and children. Br J Nutr 91, 607616.CrossRefGoogle ScholarPubMed
238 Wiseman, H, Casey, K, Bowey, EA, et al. (2004) Influence of 10 wk of soy consumption on plasma concentrations and excretion of isoflavonoids and on gut microflora metabolism in healthy adults. Am J Clin Nutr 80, 692699.CrossRefGoogle ScholarPubMed
239 Walsh, KR, Haak, SJ, Bohn, T, et al. (2007) Isoflavonoid glucosides are deconjugated and absorbed in the small intestine of human subjects with ileostomies. Am J Clin Nutr 85, 10501056.CrossRefGoogle ScholarPubMed
240 Cheng, KJ, Krishnamurty, HG, Jones, GA, et al. (1971) Identification of products produced by the anaerobic degradation of naringin by Butyrivibrio sp. C3. Can J Microbiol 17, 129131.CrossRefGoogle ScholarPubMed
241 Yu, KU, Jang, IS, Kang, KH, et al. (1997) Metabolism of Saikosaponin C and naringin by human intestinal bacteria. Arch Pharmacal Res 20, 420424.CrossRefGoogle ScholarPubMed
242 Possemiers, S, Heyerick, A, Robbens, V, et al. (2005) Activation of proestrogens from hops (Humulus lupulus L.) by intestinal microbiota; conversion of isoxanthohumol into 8-prenylnaringenin. J Agric Food Chem 53, 62816288.CrossRefGoogle ScholarPubMed
243 Felgines, C, Texier, O, Morand, C, et al. (2000) Bioavailability of the flavanone naringenin and its glycosides in rats. Am J Physiol Gastrointest Liver Physiol 279, G1148G1154.CrossRefGoogle ScholarPubMed
244 Roowi, S, Mullen, W, Edwards, CA, et al. (2009) Yoghurt impacts on the excretion of phenolic acids derived from colonic breakdown of orange juice flavanones in humans. Mol Nutr Food Res 53, Suppl. 1, S68S75.CrossRefGoogle ScholarPubMed
245 Skjevrak, I, Solheim, E & Scheline, RR (1986) Dihydrochalcone metabolism in the rat: trihydroxylated derivatives related to phloretin. Xenobiotica 16, 3545.CrossRefGoogle ScholarPubMed
246 Borges, G, Roowi, S, Rouanet, JM, et al. (2007) The bioavailability of raspberry anthocyanins and ellagitannins in rats. Mol Nutr Food Res 51, 714725.CrossRefGoogle ScholarPubMed
247 Groenewoud, G & Hundt, HKL (1986) The microbial metabolism of condensed (+)-catechins by rat caecal microflora. Xenobiotica 16, 99107.CrossRefGoogle ScholarPubMed
248 Deprez, S, Brezillon, C, Rabot, S, et al. (2000) Polymeric proanthocyanidins are catabolized by human colonic microflora into low-molecular-weight phenolic acids. J Nutr 130, 27332738.CrossRefGoogle ScholarPubMed
249 Appeldoorn, MA, Vincken, JP, Aura, MA, et al. (2009) Procyanidin dimers are metabolized by human microbiota with 2-(3,4-dihydroxyphenyl)acetic acid and 5-(3,4-dihydroxyphenyl)-γ-valerolactone as the major metabolites. J Agric Food Chem 57, 10841092.CrossRefGoogle ScholarPubMed
250 Ward, NC, Croft, KD, Puddey, IB, et al. (2004) Supplementation with grape seed polyphenols results in increased urinary excretion of 3-hydroxyphenylpropionic acid, an important metabolite of proanthocyanidins in humans. J Agric Food Chem 52, 55455549.CrossRefGoogle ScholarPubMed
251 Rickard, SE, Orcheson, LJ, Seidl, MM, et al. (1996) Dose-dependent production of mammalian lignans in rats and in vitro from the purified precursor secoisolariciresinol diglycoside in flaxseed. J Nutr 126, 20122019.Google ScholarPubMed
252 Heinonen, S, Nurmi, T, Liukkonen, K, et al. (2001) In vitro metabolism of plant lignans: new precursors of mammalian lignans enterolactone and enterodiol. J Agric Food Chem 49, 31783186.CrossRefGoogle ScholarPubMed
253 Clavel, T, Henderson, G, Engst, W, et al. (2006) Phylogeny of human intestinal bacteria that activate the dietary lignan secoisolariciresinol diglucoside. FEMS Microbiol Ecol 55, 471478.CrossRefGoogle ScholarPubMed
254 Rickard, SE & Thompson, LU (1998) Chronic exposure to secoisolariciresinol diglycoside alters lignan disposition in rats. J Nutr 128, 615623.CrossRefGoogle ScholarPubMed
255 Glitso, LV, Mazur, WM, Adlercreutz, H, et al. (2000) Intestinal metabolism of rye lignans in pigs. Br J Nutr 84, 429437.CrossRefGoogle ScholarPubMed
256 Nesbitt, PD, Lam, Y & Thompson, LU (1999) Human metabolism of mammalian lignan precursors in raw and processed flaxseed. Am J Clin Nutr 69, 549555.CrossRefGoogle ScholarPubMed
257 Juntunen, KS, Mazur, WM, Liukkonen, KH, et al. (2000) Consumption of wholemeal rye bread increases serum concentrations and urinary excretion of enterolactone compared with consumption of white wheat bread in healthy Finnish men and women. Br J Nutr 84, 839846.CrossRefGoogle ScholarPubMed
258 Mulder, TP, Rietveld, AG & van Amelsvoort, JM (2005) Consumption of both black tea and green tea results in an increase in the excretion of hippuric acid into urine. Am J Clin Nutr 81, 256S260S.CrossRefGoogle Scholar
259 Griffiths, LA (1970) 3,5-Dihydroxyphenylpropionic acid, a further metabolite of sinapic acid. Experientia 26, 723724.CrossRefGoogle Scholar
260 Kroon, PA, Faulds, CB, Ryden, P, et al. (1996) Solubilisation of ferulic acid from plant cell wall materials in a model human gut system. Biochem Soc Trans 24, 384S.CrossRefGoogle Scholar
261 Das, NP & Sothy, SP (1971) Studies on flavonoid metabolism. Biliary and urinary excretion of metabolites of (+)-(U-14C)catechin. Biochem J 125, 417423.CrossRefGoogle Scholar
262 Booth, AN, Jones, FT & DeEds, F (1958) Metabolic fate of hesperidin, eriodictyol, homoeridictyol, and diosmin. J Biol Chem 230, 661668.CrossRefGoogle ScholarPubMed
263 Honohan, T, Hale, RL, Brown, JP, et al. (1976) Synthesis and metabolic fate of hesperetin-3-14C. J Agric Food Chem 24, 906911.CrossRefGoogle ScholarPubMed
264 Scheline, RR (1978) Mammalian Metabolism of Plant Xenobiotics. London: Academic Press.Google Scholar
265 Yasuda, T, Inaba, A, Ohmori, M, et al. (2000) Urinary metabolites of gallic acid in rats and their radical-scavenging effects on 1,1-diphenyl-2-picrylhydrazyl radical. J Nat Prod 63, 14441446.CrossRefGoogle Scholar
266 Muskiet, FA & Groen, A (1979) Urinary excretion of conjugated homovanillic acid, 3,4-dihydroxyphenylacetic acid, p-hydroxyphenylacetic acid, and vanillic acid by persons on their usual diet and patients with neuroblastoma. Clin Chem 25, 12811284.CrossRefGoogle ScholarPubMed
267 Stalmach, A, Steiling, H, Williamson, G, et al. (2010) Bioavailability of chlorogenic acids following acute ingestion of coffee by humans with an ileostomy. Arch Biochem Biophys 501, 98105.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1 Structures of selected phenols, polyphenols and tannins, and selected gut microflora catabolites. A and B indicate the A- and B-ring, respectively, of a typical flavonoid. Typical structures are shown for flavanols (catechins) (I); flavanones (II); flavones (III); dihydrochalcones (IV); chlorogenic acids (cinnamate conjugates) (V); cinnamic acids (C6–C3) (VI); anthocyanidins (VII); phenylvaleric acids (C6–C5-γ-OH) (VII); phenylpropionic acids (C6–C3) (IX. XIII); phenylacetic acids (C6–C2) (X, XIV); flavonols (XI); benzoic acids (C6–C1) (XII, XV). Note: it is not possible to show all possible substrates, intermediates and pathways but those shown are representative.

Figure 1

Table 1 Studies concerned with the transformation of phenols, polyphenols and tannins (PPT) by gut flora micro-organisms

Figure 2

Table 2 The nature of the aromatic and phenolic acids produced by the gut microflora from pure phenols, polyphenols and tannins (PPT) substrates

Figure 3

Fig. 2 Content of anthocyanidins and flavanols in blackcurrants (http://www.phenol-explorer.eu/, accessed October 2009).

Figure 4

Fig. 3 Content of flavonols, cinnamates, benzoic and gallic acid derivatives in blackcurrants (http://www.phenol-explorer.eu/, accessed October 2009).

Figure 5

Table 3 The expected B-ring fragments for the common anthocyanidins and their known mammalian metabolites

Figure 6

Fig. 4 Distribution of radiolabel in rat tissues after injection of 14C-cinnamic acid to rats. The remaining radioactivity was in urine (48 %), faeces (25 %) and exhaled CO2 (0·3 %)(143).

Figure 7

Fig. 5 Summary of absorption and metabolic pathways of C6–C1 and C6–C3 compounds in the gastrointestinal tract.