Abstract
Both plant and mammalian cells express glucuronosyltransferases that catalyze glucuronidation of polyphenols such as flavonoids and other small molecules. Oral administration of select polyphenolic compounds leads to the accumulation of the corresponding glucuronidated metabolites at μM and sub-μM concentrations in the brain, associated with amelioration of a range of neurological symptoms. Determining the mechanisms whereby botanical extracts impact cognitive wellbeing and psychological resiliency will require investigation of the modes of action of the brain-targeted metabolites. Unfortunately, many of these compounds are not commercially available. This article describes the latest approaches for the analysis and synthesis of glucuronidated flavonoids. Synthetic schemes include both standard organic synthesis, semi-synthesis, enzymatic synthesis and use of synthetic biology utilizing heterologous enzymes in microbial platform organisms.
Keywords: biosynthesis, flavonoid, glucuronide, neurological disorder, organic synthesis, synthetic biology
Graphical abstract

■ INTRODUCTION
Flavonoids represent one of the major classes of plant specialized metabolites, and are synthesized through the phenypropanoid/polymalonate pathways, with the aromatic B-ring being derived from L-phenylalanine and the aromatic A-ring derived from condensation of three molecules of malonyl CoA by the plant polyketide synthase known as chalcone synthase1. Enzymatic isomerization of chalcone to flavanone yields the first flavonoid with the characteristic central heterocyclic ring. All other classes of flavonoid are formed biosynthetically in plants through oxidation and reduction reaction reactions occurring on the central C-ring, and diversity within the various classes occurs through various types of modifications to the aromatic A and B rings and the C-ring 3-hydroxyl group2. Glycosylation is perhaps the most common of such modifications, but plants also contain families of enzymes capable of catalyzing hydroxylation, O-methylation, sulfation, acetylation, prenylation, and other modifications of the flavonoid nucleus2.
Glycosylated flavonoids are widespread in plants; in fact, sugar substitution of the flavonoid aglycone is generally a prerequisite for transport to and storage of the flavonoid in the central vacuole of the plant cell3. Flavonoids can be substituted by a single sugar residue on one hydroxyl group, by a group of linked sugars attached to a single hydroxyl group, or by substitution with two or more sugars at more than one position. In addition, C-glycosyl flavonoids occur in which a non-hydrolysable carbon-carbon bond links the sugar directly, usually to an A-ring carbon4. Within this broad diversity, sugar substitution at a single position is probably the most common, with glucose as the most prevalent sugar. However, glucuronic acid is also attached to plant-derived flavonoids, and whether a particular flavonoid is glucuronidated or glucosylated in a particular plant tissue will depend on the tissue-specific expression patterns of the plant’s suite of glycosyltransferase enzymes and their sugar donor and acceptor specificities. For example, whereas the major flavonoids found in the aerial parts of Medicago species are glucuronides of the flavones tricin, apigenin, chrysoeriol and luteolin5, the roots contain isoflavone derivatives, primarily glucosides and malonylated glucosides of formononetin and medicarpin6. However, ectopic expression of the gene encoding the entry point enzyme of isoflavone biosynthesis in transgenic alfalfa (Medicago sativa) results in accumulation of isoflavone glucosides, not glucuronides, in the leaves7.
Flavonoid glucuronides have been ascribed health-promoting activities. Examples include biacalein-7-O-β-glucuronide (wound healing promotion and anticancer activity)8,9, (3-O-methyl) quercetin-3-O-β-glucuronide (anti-inflammatory and neuroprotective activities)10–12, 3-methoxyflavonol-4′-O-glucuronides (anti-allergenic)13 and epicatechin glucuronide (promotion of vascular function)14. In some of the above cases, the glucuronidation of the flavonoid is the result of mammalian metabolism. Ingestion of flavonoids by animals generally results in the hydrolysis of pre-existing O-glycosidic bonds in the digestive system, with further metabolism of the aglycone through the pathways common for metabolism of endo- and xenobiotics, namely phase I modification, phase II conjugation and phase III elimination. Phase II conjugation of flavonoids in mammals commonly involves glucuronidation to generate metabolites that can diffuse into portal and lymphatic circulation15.
Glucuronidation significantly impacts the physiological properties of the flavonoid such as its solubility (increased), bioactivity (decreased or in some case increased), bioavailability (usually increased), and inter- and intra-cellular transport as well as excretion (usually increased). Not only does the conjugation of flavonoid compounds contribute to their uptake, but the position of glucuronidation also impacts the anti-oxidant and pro-oxidant properties of flavonoids16,17; for example, glucosides and the 3-O-glucuronide of the stilbene resveratrol exhibit stronger antioxidant activity than trans-resveratrol itself18.
Recent evidence suggests that glucuronidation and other types of metabolism of flavonoids in animals might, in addition to allowing for secretion, also target bioactive molecules to their sites of action. For example, previous studies by our research group have demonstrated that oral administration of a botanical supplement mixture from grapevine is effective in protecting against neuropathology and cognitive impairment in aging19,20. These studies identified 18 biologically available phenolic metabolites, including 16 polyphenol metabolites21 and two phenolic acids22 that are found to accumulate in the brain with the potential to protect against Alzheimer’s disease pathogenic mechanisms. Moreover, in ongoing studies, we have demonstrated that some of these brain-accumulating polyphenol metabolites, in particular, 3′-O-methyl-epicatechin-5-glucuronide19, quercetin-glucuronide12, as well as 3-hydroxybenzoic acid and 3-(3′-hydroxyphenyl) propionic acid22 are capable of contributing to the efficacy of these botanical supplements to interfere with the mechanisms associated with cognitive and psychological resilience.
We are presently characterizing the cellular/molecular mechanisms through which individual flavonoids may contribute to the efficacy of the botanical mixture to modulate, respectively, psychology and cognitive resilience. For example, our studies support the evidence that select phenolic metabolites can contribute to the efficacy of the botanical mixture to promote cognitive resilience by modulating neuronal synaptic plasticity (e.g., by the polyphenol metabolites 3′-O-methyl-epicatechin-5-glucuronide and quercetin-glucuronide) as well as c-Fos, Arc, and Erg cellular signaling pathways (e.g., by the phenolic acids homovanillic acid and 3,4-dihydroxyphenylacetic acid resulting from flavonoid catabolism).
Figure 1 shows the basic structures of three of the major classes of flavonoids found in grape seeds and juice; after ingestion, these compounds appear in the brain as glucuronidated and methylated derivatives. To further pursue the potential mechanisms of action of such glucuronidated flavonoids in animal systems, it is necessary to have authenticated standard compounds. However, several of the phase II derivatives of the compounds shown in Figure 1 are not commercially available, including 3′-O-methyl quercetin 3-O-glucuronide, the 5-O-glucuronides of (epi) catechin and 3′-O-methyl (epi)catechin, and the 3-O-glucuronides of cyanidin, delphinidin and malvidin. Also unavailable are positional isomers (e.g. with the glucuronide or other substituents located on different positions) necessary for a full understanding of structure-activity relationships and identification of target receptor sites. Plants have preferences for glycosylation that also preclude access to some of these compounds from plant sources; for example, anthocyanidins are generally glucosylated, not glucuronidated, at the 3-O-position.
Figure 1.

Structures and numbering convention of three classes of flavonoids that undergo glucuronidation in mammalian tissues. A, flavan-3-ol; the stereochemistry of the aromatic ring and 3-hydroxyl denoted with ~~~ determines the catechin (2,3-trans) and epicatechin (2,3-cis) series. B, anthocyanidin (R1 =OH, R2=H = cyanidin; R1=R2=OMe = malvidin). C, flavonol, R1= OH, R2 = H = quercetin). The glucuronic acid is commonly attached via linkage to the 5-OH group in the flavan-3-ols, but is usually attached to the 3-OH in anthocyanidins and flavonols.
Although there is limited understanding of polyphenol metabolism in mammals, we have previously demonstrated that oral administration of certain brain-bioavailable phenolic glucosides, such as malvidin-3-glucoside, cyanidin-3-glucoside, delphinidin-3-glucoside and peonidin-3-glucoside, as well as resveratrol, result in their intact delivery to the brain21. However, there is no mechanistic understanding of how specific flavonoid modifications (e.g glucuronidatation) may influence delivery of targeted metabolites to the brain.
Because we are aware that brain concentrations of flavonoid metabolites are too low to allow for extraction and purification in the multi-mg amounts necessary for mechanistic studies, production of flavonoid glucuronides must therefore rely either on chemical synthesis, or biochemical approaches using enzymes, either in vitro or through synthetic biology approaches in host organisms. These approaches, as well as the analytical tools necessary to ascribe structure to biologically modified flavonoids, are outlined and evaluated in the present review.
■ ANALYSIS OF GLUCURONIDATED FLAVONOIDS
The most common methods for the detection and quantification of flavonoid glucuronides in complex matrices are high-performance liquid chromatography (HPLC) or liquid chromatography-mass spectrometry (LC-MS), and the preferred methods for structural characterization are nuclear magnetic resonance (NMR) spectroscopy, tandem mass spectrometry (MS/MS), and X-ray crystallography.
A drawback of both NMR and X-ray crystallography is the requirement for quite large amounts of purified compounds. This problem is partially alleviated by the use of hyphenated mass spectrometry techniques, including continuous-flow fast-atom bombardment (CF-FAB)23–25, MALDI-ToF26,26–29, and electrospray (ES)23,30–33, which allow for elucidation of flavonoid glucuronide structures based on accurate molecular weights and diagnostic fragmentation patterns. Capillary electrophoresis23,34,35, HPLC or UPLC36–38 and, latterly, HPLC–MS39–47 are now routine and provide advanced separation methods that have facilitated the solution of previously intractable flavonoid structures. Although gas chromatography (GC) coupled to MS can been used for glycosylated flavonoid analysis, it is not widely applicable in the present case because glucuronides exhibit limited volatility, necessitating time-consuming derivatization; the fragmentation patterns of these derivatives are also often hard to interpret.
Mass spectrometry
MS is the preferred detection technique for analysis of flavonoid glucuronides48–59, primarily because it only requires very small quantities of analyte to generate accurate tandem mass (MS/MS) spectra. Characteristic molecular ions are formed; either protonated [M + H]+ or ammonium or alkali-metal ion adducts in the positive-ion full-scan mode, or the [M − H]− ion in the negative-ion mode60.
In the positive ion mode, the ions that are formed from cleavage of two bonds in the C-ring are denoted asi, jA+ and i, jB+, with ion A comprising the A-ring and ion B the B-ring61. The C-ring bonds that are broken are represented by the indices i and j. These ions are denoted as i, jA− and i, jB−, respectively, when using the negative ion mode. For glucuronidated flavonoids, it is important to avoid confusion with the Ai+ and Bi+ (i ≥ 1) labels that designate fragments containing a terminal (non-reducing) glucuronic unit (Figure 2)57,6; for this purpose, an additional subscript 0 is used to the right of the letter.
Figure 2.

Ion nomenclature used for flavonoid glucuronides (adapted from57).1,3A0 and1,3B0 refer to aglycone fragments containing A- and B-rings, respectively, and superscripts 1 and 3 indicate the broken C-ring bonds. Ai, Bi, and Ci refer to fragments containing glucuronide fragments, with charges retained on the carbohydrate moiety, where i represents the number of broken glucuronidic bonds, counted from the terminal sugar. Xj, Yj, and Zj refer to ions containing the aglycone and j is the number of the interglycosidic bond cleaved, counted from the aglycone.
A diagnostic fragment of [M +H − 176]+ is commonly observed on analysis of glucuronides carried out in the positive-ion mode. Because the negative charge is retained on the glucuronide moiety, an abundant glucuronate fragment (m/z 193) is often seen in the negative-ion mode. Subsequent dehydration, yielding a less abundant ion at m/z 175, followed by successive losses of CO2 and H2O (m/z 113) and CO (m/z 85) (Figure 3), are also seen62–64.
Figure 3.

Characteristic fragmentation of glucuronides in negative MS/MS spectra (adapted from63).
Absolute structural characterization of the sites of conjugation of positional isomers of flavonoid glucuronides is always difficult by MS. However, flavonoids are good chelating agents towards metal ions, and this has led to novel approaches for differentiating positional isomers by the formation of metal adducts with characteristic fragmentation65–67. The favored sites of chelation by iron, cobalt and copper are catechol groups, hydroxyl groups adjacent to oxo groups, and 1-oxo-3-hydroxyl-containing moieties68. This approach enhances the capabilities of MS69–71, allowing isomeric metabolites to be differentiated under CID conditions.
Nuclear magnetic resonance spectroscopy
Although NMR spectroscopy is a powerful technique for determining the structures of flavonoid glucuronides72–78, it is limited by poor sensitivity, low throughput, and difficulties resolving components in mixtures. It is, however, possible to completely assign all proton and carbon signals for most flavonoids using a few mg of sample79–82, based on chemical shifts (δ) and spin-spin couplings (coupling constants (J)) with comparison with compiled data. In addition to identifying the type of aglycone and substituent groups, NMR analysis also identifies the number and anomeric configurations of the attached glucuronide moieties. 1H NMR data can be complemented with results from 13C NMR experiments; however, 13C NMR is much less sensitive due to the low abundance of 13C (1.1%) compared to 1H (99.9%)83.
Unequivocal structural elucidation of flavonoid glucuronides by NMR requires various 2-D approaches. These yield contour maps showing the correlations between different nuclei in the molecules, either between the same (homonuclear) or diffferent (heteronuclear) elements83. Homonuclear 1H–1H correlated NMR techniques include double-quantum filtered COSY (1H–1H DQF-COSY), 1H–1H TOCSY, 1H–1H NOESY and rotating frame Overhaüser effect spectroscopy (1H–1H ROESY). Homonuclear experiments generate spectra in which 1H chemical shifts are correlated with each other along two axes. In contrast, heteronuclear NMR experiments such as 1H−13C HSQC and heteronuclear multiple bond correlation (1H–13C HMBC) show 1H–13C correlations as crosspeaks in the spectrua83–85.
A comparison of the chemical shifts of the glucuronide to those of the aglycone can reveal the site of glucuronidation, with the largest changes in chemical shifts in the glucuronide being found in the atoms near the site of conjugation79,80. A recent study on the isolation and characterization of several new flavonol glucuronides from the flower buds of Syzygium aromaticum (clove) well illustrates the use of combined NMR and MS approaches79. The compounds characterized were rhamnetin-3-O-β-D-glucuronide (1), rhamnazin-3-O-β-D-glucuronide (2), rhamnazin-3-O-β-D-glucuronide-6″-methyl ester (3), and rhamnocitrin-3-O-β-D-glucuronide-6″-methyl ester (4). As an example, Figure 4 shows HMBC correlations for compound 1 reported in79.
Figure 4.

A, Chemical structure of compounds 1–4 from Jang et al. (79). B, HMBC correlations of compound 1.
High-Performance Liquid Chromatography–Nuclear Magnetic Resonance Spectroscopy
Although LC–UV–MS and LC–MS–MS can often provide sufficient information to enable the identification flavonoids and their glycosides, additional analytical power is provided by the combined approach of LC–NMR (Figure 5). Generally, LC–UV–MS and LC–UV–NMR are run separately. Their coupling provides a very powerful approach in which the separation and structural elucidation of unknown compounds in even quite complex mixtures are combined40,86,–91.
Figure 5.

Schematic representation of the instrumentation used for LC–UV–MS (1) and LC–UV–NMR (2) analyses (adapted from64).
Recent progress in pulse field gradients, solvent suppression, probe design, and construction of high-field magnets have significantly improved the technique. 1H NMR spectra are obtained from selected peaks in the HPLC chromatogram, complementing the LC–MS data, from which the nature of substituent groups can be deduced from the fragmentation pattern but their exact positions cannot be determined. For simple flavonoids, such as apigenin, the 1H NMR component alone will reveal the substitution position because of the unique splitting pattern for each possible location of the B-ring hydroxyl group92.
■ STRATEGIES FOR CHEMICAL SYNTHESIS OF GLUCURONIDATED FLAVONOIDS
Both glycosyl donors of the glucuronic acid type and phenolic acceptors pose problems for synthetic coupling. Construction of the correct regiospecific phenol glucuronidic linkage requires, before conjugation with the glucuronyl donors 5–9 (Figure 6), that the phenolic compounds be converted into appropriately protected precursors.
Figure 6.

Glucoronyl donors 5–9
Protection of hydroxyl groups of flavonoids
To achieve glucuronidation of flavonoids, partial or complete protection of the hydroxyl groups is necessary. An approach was developed for selective protection of each hydroxyl group of quercetin 10 by the groups of Rolando93,94 and He95 following the scheme in Figure 7. Briefly, the free hydroxyl groups of 10 were benzylated using benzyl bromide and potassium carbonate in DMF at room temperature, leading to a mixture of 3,7,3′,4′-O-tetrabenzylquercetin 11 and 3,7,4′-O-tribenzylquercetin 12, which were recovered with 60% and 20% yield, respectively (Figure 7). Alternatively, selective protection of the north-east catechol of quercetin 10 by dichlorodiphenylmethane led to the ketal 13 with 86% yield (Figure 7) to give entry into the series substituted at the 3 position94,96,97. Finally, benzylation of quercetin pentaacetate 14 at the 4′ and 7-positions with benzyl chloride/sodium bicarbonate/benzyl(triethyl)ammonium chloride using microwave irradiation (545 W, 160 °C) for 10 minutes gave compound 1595.
Figure 7.

Selective protection of hydroxyl groups of quercetin 10.
To allow the regiospecific glucuronidation of a single hydroxyl group of epicatechin with protection of the remaining groups (Figure 8)98, the specific hydroxyl group to be glucuronidated is protected with a methoxymethyl (MOM) group, while the remaining hydroxyls are protected as benzyl ethers.
Figure 8.

Synthesis of the protected (R, R)-epicatechin derivatives 18.
Basic Glucoronidation
Glucuronidation of phenols by the Koenig–Knorr method using glycosyl bromide donors is probably the most reliable approach, although yields can be relatively low99–101. For synthesis of quercetin glucuronide by this approach, 4′,7-dibenzylquercetin 15102 was treated with methyl 2,3,4-tri-O-acetyl-1-bromo-α-D-glucuronate 5/silver oxide (Ag2O) at 0°C to give a 52% yield of glucuronidated product; the final deprotection using Na2CO3 in aqueous MeOH was more efficient103,104. The overall yield was considerably increased by reaction below room temperature104. A similar approach has also been attempted for the glucuronidation of unprotected catechin, but the results are a mixture of glucuronidated catechins105. Other bases have been used as catalysts in this reaction such as LiOH, K2CO3, Ag2CO3, AgClO4, AgOTf, Hg(CN)2 or CdCO3106,107. A by-product, a 2-acyloxyglycal 20 (Figure 9) arising from HBr elimination from 5, has been frequently observed when using the Koenig–Knorr reaction104,108. This most likely arises from use of basic catalysts such as Ag2O.
Figure 9.

The elimination by-product (2-acyloxyglycal, 20) from the Koenig–Knorr reaction.
The synthesis of quercetin 3-glucuronide 23 was first reported by Wagner in 1970109. Subsequently, Needs and Kroon104 carried out a selective glucuronidation of 15 with methyl (2,3,4-tri-O-acetyl-α-D-glucopyranosyl) uronate bromide 5 in the presence of pyridine and Ag2O at 0°C, using 3 Å molecular sieves to ensure anhydrous conditions. The reaction gave 22a and 22b in a combined yield of 52% (Figure 10). A three step debenzylation and ester hydrolysis afforded 23 in 40% overall yield from 15.
Figure 10.

Synthesis of quercetin-3-O-β-D-glucuronide 23 under basic glucuronidation.
To synthesize malvidin-3-O-β-glucuronide via the Koenigs–Knorr reaction, the reaction proceeded via α-hydroxyacetosyringone 25 that was formed in three steps from acetosyringone 24 by the method reported by Luis and Andres110. The glucuronidation reaction to form the ‘Eastern part’ 26 employed the Koenigs–Knorr reaction106 between 25 and bromo-2,3,4-tri-O-acetyl-α-D-glucopyranuronic acid methyl ester 5, refluxing with silver carbonate as base in dry toluene. The reaction products were a mixture of the mono- and di-glucuronic acetophenone derivatives (26a, 26b). The C ring was generated by an aldol-type condensation between compounds 27, Western part, and 26a, Eastern part, following the deprotection steps to afford malvidin-3-O-β-glucuronide 28 (Figure 11)111.
Figure 11.

Synthesis strategy to obtain malvidin 3-O-β-glucuronide 28.
Acid glucuronidation
Glucuronidation of compounds with phenolic hydroxyl groups often utilizes methyl (2,3,4-tri-O-acetyl-D-glucopyranosyl trichloroacetimidate) uronate 6, with activation of the coupling step by Lewis acids such as BF3ˑOEt2, TMSOTf or ZnCl299,102,103,112–115. Use of benzyl uronate counterpart 8 should facilitate the final release of the carboxylic acid function following hydrogenolysis under neutral conditions116. This acid glucuronidation reaction generally requires full or partial protection of the phenolic hydroxyls104,116. Glycosyl trifluoroacetimidates 7, 9 are valuable alternatives to the corresponding trichloroacetimidates 6, 8114,116–119 and have shown advantages in synthesis of flavanone glucuronides120.
For synthesis of quercetin 3′-O-glucuronide under selective acid glucuronidation, treatment of 15 with methyl 2,3,4-tri-O-acetyl-α-D-glucopyranosyluronate trichloroacetimidate 6 in the presence of 3 Å molecular sieves and dry CH2Cl2 gave 29 and recovered 15 (Figure 12). Glucuronidation at the 3-O-position was not observed. Debenzylation and de-esterification afforded, after purification, 30 in 11% yield. It is important to note the differences of regioselectivity between the reactions with glucuronyl donors 5 and 6104.
Figure 12.

Synthetic route to obtain quercetin 3′-O-glucuronide 30
For the chemical synthesis of epicatechin glucuronides, compounds 18a and 18b are suitably protected for O-glucuronidation specifically at positions 3′ and 4′, respectively. This was achieved using the glucuronic acid donor 7, under BF3•OEt2 catalysis. Mild alkaline hydrolysis, followed by hydrogenolysis over Pd(OH)2/C, yielded the glucuronides 32a and 32b (Figure 13)113,121.
Figure 13.

Syntheses of (−)-epicatechin 3′- and 4′-O-β-D-glucuronides 32a–b.
Regio- and stereo-selective synthesis of quercetin O-β-D-glucuronidated derivatives via selective and non-selective glucosylation of quercetin
Glucuronidation provides additional challenges when compared with the more usually performed glucosylation of natural products113. This is highlighted in the case of polyphenols, for which even glucosylation can be problematic. For example, the yield of quercetin-7-O-glucuronide was only 8% following alkylation of 3,3′,4′,5-tetrabenzoylquercetin by glucuronic acid methyl ester bromide triacetate 5122. In contrast, efficient glucosylation of flavonoids occurs under mild conditions using a phase transfer catalyst such as tetrabutylammonium bromide123. Synthetic procedures have been developed for the formation of quercetin glucuronides based on the sequential and selective protections of the hydroxyl functions to allow selective glucosylation, followed by TEMPO-mediated oxidation of the glucoside to the glucuronide. These technologies make it possible to synthesize the four O-β-D-glucuronides of quercetin93.
The most common glycosidation of quercetin is at position 3 of the C ring. Quercetin 3-O-β-D-glucuronide 23 can be synthesized from compound 13 in five steps (Figure 14): (i) selective glucosylation of the 3-hydroxyl group; (ii) protection of the remaining free 5 and 7 hydroxyl groups with benzyl bromide in excess K2CO3 in dimethylformamide at room temperature); (iii) deprotection of the sugar residue by removal of the acetoxy group with sodium methylate; (iv) selective oxidation by NaOCl (catalyzed by TEMPO) of the sugar of quercetin-3-O-β-D-glucoside 35 with the phenol groups still protected, with solubility of 35 ensured by phase transfer catalysis between CH2Cl2 and saturated sodium hydrogencarbonate with tetrabutylammonium; and finally (v) deprotection of the hydroxyl groups of the flavonoid by catalytic hydrogenation using 30% palladium on charcoal to yield the 3-O-β-D- glucuronide 23 (25%)93,94.
Figure 14.

Synthesis of quercetin-3-O-β-D-glucuronide 23.
It is also possible to carry out the non-selective glycosylation of the A-ring 5-position of quercetin, which is less reactive than the 3-position. In this case, the protocol includes a first protection of the 3, 3′, 4′ and 7 hydroxyl groups, which are not to be glycosylated. The glycosylation is achieved on the 3, 3′, 4′, 7-tetrabenzylated quercetin 11 (Figure 15). The synthesis of quercetin-5-O-β-D-glucuronide 38 proceeds in four steps: (i) tetrabenzylated quercetin 11 is reacted with acetobromoglucose in the presence of potassium carbonate; (ii) the glucoside moiety is deprotected as described above; (iii) oxidation of the primary alcohol of quercetin 5-O-β-D-glucoside with protected phenol groups 37 is performed to form the corresponding protected glucuronide. Finally, deprotection by removal of the benzyl groups generates quercetin 5-O-β-D-glucuronide 38 with 25% yield93.
Figure 15.

Synthesis of quercetin-5-O-β-D-glucuronide 38.
Selective glycosylation of the B ring of quercetin can be performed starting from the tribenzylated quercetin 12; the coupling reaction is carried out under phase transfer conditions with acetobromoglucose, as discussed earlier (Figure 16). Position 5 is not protected as it is not reactive. Deprotection of the glucoside moiety is performed as described above. Finally, selective oxidation of the primary alcohol and deprotection of the hydroxyl groups results in quercetin 3′-O-β-D-glucuronide 30 (24% yield)93.
Figure 16.

Synthesis of quercetin-3′-O-β-D-glucuronide 30.
■ ENZYMATIC GLUCURONIDATION OF FLAVONOIDS
Plants, animals and microorganisms possess enzymes capable of glycosylating a range of plant-derived flavonoid compounds. These enzymes, members of the uridine diphosphate (UDP)-glycosyltransferase (UGT) superfamily, generally possess a common protein structure as well as a 44 amino acid residue signature sequence (the PSPG box) for binding to the UDP moiety of the UDP-sugar that serves as the sugar donor (UDP-glucuronic acid in the case of the UDP-glucuronosyltransferases)124. Regioselectivity (i.e. the position of conjugation of the sugar on the flavonoid) depends on the type of flavonoid and the nature of the enzyme catalyzing the conjugation reaction. For example, glucuronidation of the flavone luteolin and the flavonol quercetin in mammals does not follow the same pattern, with regioselectivity depending on the individual flavonoid and the class of UDP-glucuronosyltransferase isoenzyme involved16. Because of this specificity, UGTs provide excellent catalysts for biochemical synthesis of sugar conjugates, using simple reaction conditions that do not require protection of non-reacting hydroxyl groups.
Animal enzymes
Mammals, including humans, have evolved a wide range of UGT enzymes and isozymes for glucuronidating compounds, with varying degrees of catalytic efficiency and promiscuity in terms of substrate preference. In humans, there are 27 UGT gene products identified, and these are key phase II drug metabolizing enzymes that play central roles in metabolizing and detoxifying foreign chemicals such as carcinogens and hydrophobic drugs125. Mammalian UGT1A1 (expressed in liver), UGT1A8 (intestine), UGT1A9 (liver) and UGT1A3 are highly active in conjugating flavonoids (e.g. quercetin and luteolin), whereas UGT1A4 and UGT1A10 and the isoenzymes from the UGTB family, UGT2B7 and UGT2B15, are less efficient16,126.
The presence of this wide range of UGTs in mammals can be attributed in part to herbivore (mammal): plant: microbe co-evolution as animal herbivores have had to deal with ingestion of toxic phytoanticipins (pre-formed antimicrobial substances) and phytoalexins (inducible antimicrobial substances produced in plants in response to microbial pathogens)127. The relative promiscuity of the enzymes allows a range of mammalian tissues, including intestines, liver, and kidney, to effectively detoxify phytoalexins, phytoanticipins, and drugs124.
The sugar acceptor specificities of mammalian UDP-glucuronosyltransferases vary considerably. UGT1A1 is perhaps the most important drug-conjugating and xenobiotic detoxifying UGT because of its broad substrate specificity. Table 1 shows the tissue location, substrate preferences and regiospecificities for flavonoids of the human UDP-glucuronosyltransferases. Because of the differential tissue distributions of enzymes with different substrate- and regio-specificities, different tissues function to detoxify flavonoid compounds in different manners. For example, glucuronidation of prunetin (a methylated derivative of the isoflavone genistein and a potential prodrug for cancer prevention) by liver UGT1A7, UGT1A8, and UGT1A9 yielded prunetin-5-O-glucuronide whereas intestinal UGT1A1, UGT1A8, and UGT1A10 produced prunetin-4′-O-glucuronide128.
Table 1.
Properties of Human Glucuronosyltransferases. Regioselectivity shows products formed from luteolin (L), quercetin (Q), or epicatechin (E).
| Isoenzyme | Tissue localization | Preferred substrates | Regioselectivity | Reference |
|---|---|---|---|---|
| UGT1A1 | Liver, intestine | Bilrubin; anthraquinones; oripavin opiods (e.g.buprenorphine); estrogens; phenols and flavonoids (e.g chrysin, apigenin, baicelin, luteolin, quercetin, fisetin, genistein, narigenin). | L-4′-O-GlcA > L-3′-O-GlcA>L-7-O-GlcA Q-3′-O-GlcA > Q-4′-O-GlcA> Q-7-O-GlcA |
16, 126, 157, 158 |
| UGT1A3 | Liver | Certain estrogens; flavonoids; coumarin; amines;anthraquinones. | L-7-O-GlcA > L-4′-O-GlcA > L-3′-O-GlcA Q-3′-O-GlcA > Q-7-O-GlcA>Q-3-O-GlcA > Q-4′-O-GlcA |
16, 159 |
| UGT1A4 | Liver | Primary, secondary, and tertiary amines; monoterpenoid alcohols; sapogenins;androstanediol; progestins; certain flavonoids. | L-7-O-GlcA>L-4′-O-GlcA>L-3′-O-GlcA Q-4′-O-GlcA>Q-3′-O-GlcA |
16, 160 |
| UGT1A6 | Liver, intestine, kidney | Few planar phenolic compounds and some flavonoids. | L-7-O-GlcA Q-4′-O-GlcA > Q-7-O-GlcA>Q-3′-O-GlcA > Q-3-O-GlcA |
16, 161 |
| UGT1A8 | Kidney, colon, intestine, liver | Flavonoids including apigenin, luteolin, narigenin, daizdein. | L-7-O-GlcA>L-3′-O-GlcA>L-4′-O-GlcA Q-3′-O-GlcA> Q-7-O-GlcA>Q-4′-O-GlcA > Q-3-O-GlcA |
16, 126 |
| UGT1A9 | Liver, kidney | Flavonoids; anthraquinones; bulky phenols; certain aliphatic alcohols; nonsteroidal anti-inflammatory drugs. | L-4′-O-GlcA > L-3′-O-GlcA>L-7-O-GlcA Q-3′-O-GlcA>Q-4′-O-GlcA > Q-7-O-GlcA EC-3′-O-GlcA > EC-5-O-GlcA 3′-O-Me-EC-5-O-GlcA |
16, 161 131, 132 |
| UGT1A10 | Intestine, liver | mycophenolic acid; some flavonoids; antineoplastic and immunosuppressive agents. | L-7-O-GlcA > L-4′-O-GlcA> L-3′-O-GlcA Q-7-O-GlcA> Q-3-O-GlcA > Q-4′-O-GlcA |
16, 162 |
| UGT2B7 | Kidney, liver | Some flavonoids | L-3′-O-GlcA Q-7-O-GlcA > Q-3′-O-GlcA > Q-3-O-GlcA |
16 |
| UGT2B15 | Intestine, bone marrow and immune system, liver | Some flavonoids | L-7-O-GlcA > L-4′-O-GlcA > L-3′-O-GlcA Q-7-O-GlcA > Q-4′-O-GlcA > Q-3′-O-GlcA |
16 |
Although glucuronidation, sulfation, and methylation of compounds such as flavonoids are now well established features of phase II endo- and xeno-biotic metabolism, the underlying mechanisms for flavonoid uptake into portal and lymphatic circulation still require elucidation129.
One major challenge in using mammalian UDP-glucuronosyltranferases in biotechnology applications for the synthesis of flavonoid glucuronides is that the mammalian enzymes are membrane bound proteins. Heterologous expression of these enzymes for novel applications has therefore often proven difficult. Commercial preparations of these enzymes are usually available as supersomes or microsomes prepared from mammalian sources, or alternatively can be obtained by transfecting cDNAs encoding human UGTs into mammalian or insect cell lines, as first demonstrated more than 25 years ago130. Such preparations have been used for the biochemical synthesis of glucuronides of epicatechin, catechin, and their 3′-O-methyl esters131,132. However, relatively large amounts of enzyme are required. These factors of difficulty, expense, and possibly safety, weigh against the use of mammalian enzymes for the biochemical synthesis of flavonoid glucuronides for therapeutic applications such as in the treatment of neurological disorders. At the same time, the purely chemical approaches to the glucuronidation of these compounds as reviewed above are complex and time-consuming. An alternative approach is necessary, and the UDP-glucuronosyltransferase enzymes from plants provide several advantages for safe, rapid and efficient synthesis of flavonoid glucuronides.
Plant enzymes
Glycosylation of phytochemicals mediated by UGTs is one of the major factors determining plant natural product bioactivity and bioavailability133,134, and many UGTs have evolved for glycosylating plant natural products. For example, about 107 UGT genes have been identified in the genome of Arabidopsis thaliana135 and over 300 UGT genes are present in the genome of the model legume Medicago truncatula136. Although a large proportion of the enzymes in these two species identified to date are glucosyl transferases with different acceptor specificities, both species do make some glucuronidated flavonoids. To date, a small number of UDP-glucuronosyltransferases have been identified and functionally characterized from a few plant species; these include flavonoid 7-O-glucuronosyltransferases (F7GATs that comprise UGT88D1, UGT88D4, UGT88D6 and UGT88D7) from members of the Lamiales 137; BpUGT94B1, which is a glucuronosyltransferase of cyanidin-derived flavonoids from red daisy138; UBGAT (UGT88D1), purified from cultured cells of Scutellaria baicalensis Georgi, that conjugates the 7-OH of the 5-deoxy flavonoid baicalein with glucuronic acid139; and the flavonol-3-O-glucuronosyltransferase VvGT5 from grapevine (Vitis vinifera)140.
In contrast to human enzymes that are membrane-bound proteins, all plant UDP-glucuronosyltransferases discovered to date are soluble, thus making them prime candidates for developing novel phytotherapeutic flavonoid glucuronides. More effort should be devoted to extending the repertoire of natural plant UDP-glucuronosyltransferases capable of modifying flavonoids beneficial to human health, as only a small number of plants that accumulate glucuronidated flavonoids have been investigated in this respect to date. The idea of broadening the repertoire of available biocatalysts by structure based engineering of plant UGTs is discussed in the final section of this review.
Microbial enzymes
Glucuronidation is one of the mechanisms through which some microorganisms naturally metabolize phenolic compounds such as flavonoids, probably as a detoxification mechanism to make the compounds more soluble141. The extent of microbial conjugation of compounds with sugars is subject to multiple parameters including the pH, medium composition, temperature, and concentration of the substrate142.
Streptomyces sp. strain M52104 can bring about the transformation of flavanone (naringenin), flavonol (quercetin) and several stilbenoids (trans-resveratrol, rhapontigenin, deoxy-rhapontigenin) into their O-β-D-glucuronide derivatives. The bioconversions were always β-stereospecific, but not completely regioselective141. Beauveria bassiana ATCC 7159 and Cunninghamella echinulata ATCC 9244 respectively convert quercetin and its disaccharide derivative rutin into their respective glucuronide derivatives143. Generally, however, microbes require engineering with a specific glycosyltransferase if they are to be used as effective catalysts for bioconversion of flavonoids, as discussed further below.
■ BIOTECHNOLOGICAL APPROACHES TO FLAVONOID GLUCURONIDATION
A number of studies have addressed the potential use of microbial biotransformation to generate flavonoid glycosides from the corresponding aglycone. In most cases, the microorganism has been Escherichia coli, and the sugar attached has been glucose, derived from the host’s endogenous pools of uridine diphosphate glucose (UDPG). Examples include the formation of glucosides of the flavone luteolin, isoflavones genistein and biochanin A, and flavonols kaempferol and quercetin by E. coli expressing UGT71G1 or UGT73C8 from Medicago truncatula144. Yields of glucuronidated products following feeding of the bacterial cultures with the aglycone were in the range of 10–20 mg/L.
Fewer studies have used this approach for the formation of flavonoid glucuronides. Because of the smaller pool of UDP-glucuronic acid than that of UDP-glucose in E. coli, the level of sugar donor can limit the overall yield of glucuronidated product. To overcome this limitation, expression of heterologous glucuronidation enzymes has recently been coupled with engineering of the host to increase the pool of endogenous UDP-glucuronic acid. The manipulation of UDP-glucuronic acid biosynthesis was in two steps; firstly, the araA gene encoding UDP-4-deoxy-4-formamido-L-arabinose formyltransferase/UDP-glucuronic acid C-4″ decarboxylase, an enzyme that consumes UDP-glucuronic acid as substrate for capsular polysaccharide biosynthesis, was knocked out in E. coli, and secondly, the E. coli UDP-glucose dehydrogenase (ugd) gene that produces UDP-glucuronic acid145 was overexpressed. Finally, the respective flavonoid glucuronosyl transferases were expressed in the modified E. coli strain; AmUGT10 from Antirrhinum majus for luteolin and VvUGT from Vitis vinifera for quercetin146. Using this strategy, luteolin-7-O-glucuronide and quercetin-3-O-glucuronide were biosynthesized to levels as high as 300 mg/L and 687 mg/L respectively146. A similar approach has been used to manipulate the endogenous upstream biosynthetic pathway for sugar donor (UDP-glucuronic acid) accumulation in concert with the heterologous expression of downstream UDP-dependent glycosyltransferase (SbUBGAT) isolated from Scutellaria baicalensis Georgi, which catalyzes the glucuronidation of baicalein. As a result, about 797 mg L−1 of baicalein-7-O-glucuronide were biosynthesized in engineered E. coli147.
Systems-based engineering for flavonoid glucuronide biosynthesis has the potential of becoming a powerful approach to making health promoting flavonoid glucuronides on a scale that will support further research and development. However, because the currently identified plant enzymes do not cover the full range of substrate- and regio-specificities necessary for generation of all mammalian-derived flavonoid conjugates, for this strategy to have its broadest utility, it will be necessary to either discover more UGTs with different specificities, or else broaden the existing substrate- and regio- specificities of available UGTs by protein engineering.
■ STRUCTURE-BASED DESIGN OF NOVEL GLYCOSYLTRANSFERASES
Identification and characterization of new UDP-glucuronosyltransferases, followed by protein modelling, design and engineering, are strategies that can be used to facilitate the synthesis of currently investigated or novel flavonoid glucuronides. Understanding the mechanism of catalysis by UDP-glucuronosyltransferases is critical for structure-based protein design, but as yet no complete UDP-glucuronosyltransferase crystal structure has been solved. Progress to date has focused on plant glucosyltransferases, the crystal structures of six of which have been reported, namely Medicago truncatula UGT71G1, UGT85H2, and UGT78G1, which were determined in our lab148–150, and grape (Vitis vinifera) VvGT1151, Arabidopsis thaliana UGT72B1152, and Clitoria ternatea UGT78K6153. A crystal structure of the C-terminal domain of human UGT2B7 has also been reported154. These studies have provided structural bases for understanding the catalytic mechanism(s) and specificity of UGTs, and also provide a framework for beginning to address the specific features of flavonoid glucuronosyltransferases155.
The first insight into the mechanism of catalysis by a plant UDP-glucuronosyltransferase used a combination of protein modeling and site-directed mutagenesis followed by analysis of the substrate specificity of wild-type and mutated forms of the enzyme. BpUGT94B1 from red daisy (Bellis perennis) is a sugar-sugar/branch forming glucuronosyltransferase that catalyzes glucuronidation of a sugar already attached to a flavonoid such as cyanidin. Modeling and biochemical studies showed that an arginine residue (Arg25) in the N-terminus near the catalytic histidine is crucial for sugar donor specificity for UDP-glucuronic acid (Figure. 17)156.
Figure 17.

A modeled structure of the flavonoid 7-O-glucuronosyltransferase UGT88D7. The docked UDP-glucuronic acid (in yellow) and epicatechin (magenta) are shown as bond models. The key amino acids Arg350 in UGT88D7 and Arg25 in UGT94B1 are also shown as bond models in green and blue, respectively.
A modeling study of an F7GAT, UGT88D7, in combination with mutagenesis, showed that the key residue Arg350, which would form interactions with the anionic carboxylate of the glucuronic acid moiety of UDP-glucuronic acid, is crucial for defining the sugar donor-specificity of the enzyme for UDP-glucuronic acid137. Arginine-140 was shown to be the determinant for UDP-glucuronic acid specificity of grapevine Vv GT5. However, this amino acid residue did not corespond to any of the previously identified amino acid residues necessary for UDP-glucuronic acid specificity, suggesting independent convergent evolution of plant UDP-glucuronosyltranferases across plant species140.
Structure-based protein engineering has emerged as an attractive approach to manipulate biosynthetic enzymes to generate novel biocatalysts. Structure-based mutagenesis studies on UGTs have shown that it is possible to manipulate both the regioselectivity of glycosylation and the rate of substrate turnover by site-directed mutations. For example, the F148V and Y202A mutants of UGT71G1 glycosylated quercetin at the 3-O-position, compared to the wild-type enzyme that predominantly acts on the 3′-O-position149. The I305T mutation of UGT85H2 enhanced enzyme catalytic efficiency 37-or 19-fold with kaempferol or biochanin A as sugar acceptors, respectively148.
Structure-based modifications of plant UDP-glucuronosyltransferases to alter pocket topology, size and composition, presents a new approach for the design of biocatalysts for synthesizing a range of bioactive glucuronides for basic studies and treatment of metabolic syndrome, Alzheimer’s disease, and other neurological disorders.
Acknowledgments
FUNDING
This study was supported by Grant Number P50 AT008661-01 from the NCCIH and the ODS, and by the University of North Texas. Dr. Pasinetti holds a Senior VA Career Scientist Award. We acknowledge that the contents of this study do not represent the views of the NCCIH, the ODS, the NIH, the U.S. Department of Veterans Affairs, or the United States Government
References
- 1.Heller W, Hahlbrock K. Highly purified “flavanone synthase” from parsley catalyzes the formation of naringenin chalcone. Arch Biochem Biophys. 1980;200:617–619. doi: 10.1016/0003-9861(80)90395-1. [DOI] [PubMed] [Google Scholar]
- 2.Modolo LV, Reichert AI, Dixon RA. Introduction to the different classes of biosynthetic enzymes. In: Osbourn AE, Lanzotti V, editors. Plant-Derived Natural Products. Springer; Dordrecht: 2009. pp. 97–125. [Google Scholar]
- 3.Kitamura S. Transport of flavonoids. From cytosolic synthesis to vacuolar accumulation. In: Grotewald E, editor. The Science of Flavonoids. Springer; New York: 2006. pp. 123–146. [Google Scholar]
- 4.Brazier-Hicks M, Evans KM, Gershaater MC, Puschmann H, Steel PG, Edwards R. The C-glycosylation of flavonoids in cereals. J Biol Chem. 2009;284:17926–17934. doi: 10.1074/jbc.M109.009258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Kowalska I, Stochmal A, Kapusta I, Janda B, Pizza C, Piacente S, Oleszek W. Flavonoids from barrel medic (Medicago truncatula) aerial parts. J Agric Food Chem. 2007;55:2645–2652. doi: 10.1021/jf063635b. [DOI] [PubMed] [Google Scholar]
- 6.Farag MA, Huhman DV, Lei Z, Sumner LW. Metabolic profiling and systematic identification of flavonoids and isoflavonoids in roots and cell suspension cultures of Medicago truncatula using HPLC-UV-ESI-MS and GC-MS. Phytochemistry. 2007;68:342–354. doi: 10.1016/j.phytochem.2006.10.023. [DOI] [PubMed] [Google Scholar]
- 7.Deavours BE, Dixon RA. Metabolic engineering of isoflavonoid biosynthesis in alfalfa (Medicago sativa L.) Plant Physiol. 2005;138:2245–2259. doi: 10.1104/pp.105.062539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Kalaivanan P, Sivagnanam I, Rajamanickam M. Evaluation of wound healing activity of baicalein-7-O-β-D-glucuronide isolated from Leucas aspera. J Appl Pharm Sci. 2013:46–51. [Google Scholar]
- 9.Batra P, Sharma AK. Anti-cancer potential of flavonoids: recent trends and future perspectives. 3 Biotech. 2013;3:439–459. doi: 10.1007/s13205-013-0117-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Gong X, Zhou X, Zhao C, Chen H, Zhao Y. Anti-inflammatory properties of quercetin-3-O-β-D-glucuronide-methyl ester from Polygonum perfoliatum in mice. Int J Pharmacol. 2013;9:533–537. [Google Scholar]
- 11.Kawai Y. β-Glucuronidase activity and mitochondrial dysfunction: the sites where flavonoid glucuronides act as anti-inflammatory agents. J Clin Biochem Nutr. 2014;54:145–150. doi: 10.3164/jcbn.14-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Ho L, Ferruzzi MG, Janle EM, Wang J, Gong B, Chen T-Y, Lobo J, Cooper B, Wu QL, Talcott ST, Percival SS, Simon JE, Pasinetti GM. Identification of brain-targeted bioactive dietary quercetin-3-O-glucuronide as a novel intervention for Alzheimer’s disease. FASEB J. 2013;27:769–781. doi: 10.1096/fj.12-212118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Morishita Y, Saito E, Takemura E, Fujikawa R, Yamamoto R, Kuroyanagi M, Osamu Shirota O, Muto N. Flavonoid glucuronides isolated from spinach inhibit IgE mediated degranulation in basophilic leukemia RBL-2H3 cells and passive cutaneous anaphylaxis reaction in mice. Integr Mol Med. 2015;2:99–105. [Google Scholar]
- 14.Schroeter H, Heiss C, Balzer J, Kleinbongard P, Keen CL, Hollenberg NK, Sies H, Kwik-Uribe C, Schmitz HH, Kelm M. (−)-Epicatechin mediates beneficial effects of flavanol-rich cocoa on vascular function in humans. Proc Natl Acad Sci USA. 2006;103:1024–1029. doi: 10.1073/pnas.0510168103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Omiecinski CJ, Heuvel JPV, Perdew GH, Peters JM. Xenobiotic metabolism, disposition, and regulation by receptors: from biochemical phenomenon to predictors of major toxicities. Toxicol Sci. 2011;120(Suppl 1):S49–S75. doi: 10.1093/toxsci/kfq338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Boersma MG, van der Woude H, Bogaards J, Boeren S, Vervoort J, Cnubben NHP, van Iersel MLPS, van Bladeren PJ, Rietjens IMCM. Regioselectivity of phase II metabolism of luteolin and quercetin by UDP-glucuronosyl transferases. Chem Res Toxicol. 2002;15:662–670. doi: 10.1021/tx0101705. [DOI] [PubMed] [Google Scholar]
- 17.Spencer JPE, Chowrimootoo G, Choudhury R, Debnam ES, Srai SK, Rice-Evans C. The small intestine can both absorb and glucuronidate luminal flavonoids. FEBS Lett. 1999;458:224–230. doi: 10.1016/s0014-5793(99)01160-6. [DOI] [PubMed] [Google Scholar]
- 18.Mikulski D, Molski M. Quantitative structure-antioxidant activity relationship of trans-resveratrol oligomers, trans-4,4′-dihydroxystilbene dimer, trans-resveratrol-3-O-glucuronide, glucosides: trans-piceid, cis-piceid, trans-astringin and trans-resveratrol-4′-O-β-D-glucopyranoside. Eur J Med Chem. 2010;45:2366–2380. doi: 10.1016/j.ejmech.2010.02.016. [DOI] [PubMed] [Google Scholar]
- 19.Wang J, Ferruzzi MG, Ho L, Blount J, Janle E, Arrieta-Cruz I, Sharma V, Cooper B, Lobo J, Simon JE, Zhang C, Cheng A, Qian X, Pavlides C, Dixon RA, Pasinetti GM. Brain-targeted proanthocyanidin metabolites for Alzheimer’s disease treatment. J Neurosci. 2012;32:5144–5150. doi: 10.1523/JNEUROSCI.6437-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Wang J, Ho L, Zhao W, Ono K, Rosensweig C, Chen L, Humala N, Teplow DB, Pasinetti GM. Grape-derived polyphenolics prevent Aβ oligomerization and attenuate cognitive deterioration in a mouse model of Alzheimer’s disease. J Neurosci. 2008;28:6388–6392. doi: 10.1523/JNEUROSCI.0364-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Wang J, Bi W, Cheng A, Freire D, Vempati P, Zhao W, Gong B, Janle E, Chen TY, Ferruzzi M, Schmeidler J, Ho L, Pasinetti G. Targeting multiple pathogenic mechanisms with polyphenols for the treatment of Alzheimer’s disease-experimental approach and therapeutic implications. Front Aging Neurosci. 2014;6 doi: 10.3389/fnagi.2014.00042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Wang D, Ho L, Faith J, Ono K, Janle EM, Lachcik PJ, Cooper BR, Jannasch AH, D’Arcy BR, Williams BA, Ferruzzi MG, Levine S, Zhao W, Dubner L, Pasinetti GM. Role of intestinal microbiota in the generation of polyphenol-derived phenolic acid mediated attenuation of Alzheimer’s disease β-amyloid oligomerization. Mol Nutr Food Res. 2015;59:1025–1040. doi: 10.1002/mnfr.201400544. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Prasain JK, Wang CC, Barnes S. Mass spectrometric methods for the determination of flavonoids in biological samples. Free Radic Biol Med. 2004;37:1324–1350. doi: 10.1016/j.freeradbiomed.2004.07.026. [DOI] [PubMed] [Google Scholar]
- 24.Stobiecki M. Application of mass spectrometry for identification and structural studies of flavonoid glycosides. Phytochemistry. 2000;54:237–256. doi: 10.1016/s0031-9422(00)00091-1. [DOI] [PubMed] [Google Scholar]
- 25.Castaneda-Ovando A, de Lourdes Pacheco-Hernandez Ma, Elena Paez-Hernandez Ma, Rodriguez Jose A, Galan-Vidal Carlos Andres. Chemical studies of anthocyanins: a review. J Food Chem. 2009;113:859–871. [Google Scholar]
- 26.Marczak L, Kachlicki P, Koźniewski P, Skirycz A, Krajewski P, Stobiecki M. Matrix-assisted laser desorption/ionization time-of-flight mass spectrometry monitoring of anthocyanins in extracts from Arabidopsis thaliana leaves. Rapid Commun Mass Spectrom. 2008;22:3949–3956. doi: 10.1002/rcm.3819. [DOI] [PubMed] [Google Scholar]
- 27.Hölscher D, Shroff R, Knop K, Gottschaldt M, Crecelius A, Schneider B, Heckel DG, Schubert US, Svatos A. Matrix-free UV-laser desorption/ionization (LDI) mass spectrometric imaging at the single-cell level: distribution of secondary metabolites of Arabidopsis thaliana and Hypericum species. Plant J. 2009;60:907–918. doi: 10.1111/j.1365-313X.2009.04012.x. [DOI] [PubMed] [Google Scholar]
- 28.Zhang H, Cha S, Yeung ES. Colloidal graphite-assisted laser desorption/ionization MS and MS(n) of small molecules. 2. Direct profiling and MS imaging of small metabolites from fruits. Anal Chem. 2007;79:6575–6584. doi: 10.1021/ac0706170. [DOI] [PubMed] [Google Scholar]
- 29.Madeira PJ, Florêncio MH. Flavonoid-matrix cluster ions in MALDI mass spectrometry. J Mass Spectrom. 2009;44:1105–1113. doi: 10.1002/jms.1588. [DOI] [PubMed] [Google Scholar]
- 30.Li J, Wang YH, Smillie TJ, Khan IA. Identification of phenolic compounds from Scutellaria lateriflora by liquid chromatography with ultraviolet photodiode array and electrospray ionization tandem mass spectrometry. J Pharm Biomed Anal. 2012;63:120–127. doi: 10.1016/j.jpba.2012.01.027. [DOI] [PubMed] [Google Scholar]
- 31.Dehkharghanian M, Adenier H, Vijayalakshmi MA. Study of flavonoids in aqueous spinach extract using positive electrospray ionisation tandem quadrupole mass spectrometry. Food Chem. 2010;121:863–870. [Google Scholar]
- 32.Barnes JS, Schug KA. Structural characterization of cyanidin-3, 5-diglucoside and pelargonidin-3, 5-diglucoside anthocyanins: Multi-dimensional fragmentation pathways using high performance liquid chromatography-electrospray ionization-ion trap-time of flight mass spectrometry. Int J Mass Spectrom. 2011;308:71–80. [Google Scholar]
- 33.Orrego-Lagarón N, Vallverdú-Queralt A, Martínez-Huélamo M, Lamuela-Raventos RM, Escribano-Ferrer E. Metabolic profile of naringenin in the stomach and colon using liquid chromatography/electrospray ionization linear ion trap quadrupole-Orbitrap-mass spectrometry (LC-ESI-LTQ-Orbitrap-MS) and LC-ESI-MS/MS. J Pharm Biomed Anal. 2016;120:38–45. doi: 10.1016/j.jpba.2015.10.040. [DOI] [PubMed] [Google Scholar]
- 34.Suntornsuk L. Capillary electrophoresis of phytochemical substances. J Pharm Biomed Anal. 2002;27:679–698. doi: 10.1016/s0731-7085(01)00531-3. [DOI] [PubMed] [Google Scholar]
- 35.Hurtado-Fernández E, Gómez-Romero M, Carrasco-Pancorbo A, Fernández-Gutiérrez A. Application and potential of capillary electroseparation methods to determine antioxidant phenolic compounds from plant food material. J Pharm Biomed Anal. 2010;53:1130–1160. doi: 10.1016/j.jpba.2010.07.028. [DOI] [PubMed] [Google Scholar]
- 36.Padilha CV, Miskinis GA, de Souza ME, Pereira GE, de Oliveira D, Bordignon-Luiz MT, Lima MD. Rapid determination of flavonoids and phenolic acids in grape juices and wines by RP-HPLC/DAD: Method validation and characterization of commercial products of the new Brazilian varieties of grape. Food Chem. 2017;228:106–115. doi: 10.1016/j.foodchem.2017.01.137. [DOI] [PubMed] [Google Scholar]
- 37.Merken HM, Beecher GR. Measurement of food flavonoids by high-performance liquid chromatography: a review. J Agric Food Chem. 2000;48:577–599. doi: 10.1021/jf990872o. [DOI] [PubMed] [Google Scholar]
- 38.Careri M, Corradini C, Elviri L, Nicoletti I, Zagnoni I. Direct HPLC analysis of quercetin and trans-resveratrol in red wine, grape, and winemaking byproducts. J Agric Food Chem. 2003;51:5226–5231. doi: 10.1021/jf034149g. [DOI] [PubMed] [Google Scholar]
- 39.Valls J, Millán S, Martí MP, Borràs E, Arola L. Advanced separation methods of food anthocyanins, isoflavones and flavanols. J Chromatogr A. 2009;1216:7143–7172. doi: 10.1016/j.chroma.2009.07.030. [DOI] [PubMed] [Google Scholar]
- 40.Saldanha LL, Vilegas W, Dokkedal AL. Characterization of flavonoids and phenolic acids in Myrcia bella Cambess. using FIA-ESI-IT-MS(n) and HPLC-PAD-ESI-IT-MS combined with NMR. Molecules. 2013;18:8402–8416. doi: 10.3390/molecules18078402. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Petreska Stanoeva J, Stefova M. Assay of urinary excretion of polyphenols after ingestion of a cup of mountain tea (Sideritis scardica) measured by HPLC-DAD-ESI-MS/MS. J Agric Food Chem. 2013;61:10488–10497. doi: 10.1021/jf403052w. [DOI] [PubMed] [Google Scholar]
- 42.Lee J, Ebeler SE, Zweigenbaum JA, Mitchell AE. UHPLC-(ESI)QTOF MS/MS profiling of quercetin metabolites in human plasma postconsumption of applesauce enriched with apple peel and onion. J Agric Food Chem. 2012;60:8510–8520. doi: 10.1021/jf302637t. [DOI] [PubMed] [Google Scholar]
- 43.Kure A, Nakagawa K, Kondo M, Kato S, Kimura F, Watanabe A, Shoji N, Hatanaka S, Tsushida T, Miyazawa T. Metabolic fate of luteolin in rats: its relationship to anti-inflammatory effect. J Agric Food Chem. 2016;64:4246–4254. doi: 10.1021/acs.jafc.6b00964. [DOI] [PubMed] [Google Scholar]
- 44.Wu L, Liu J, Han W, Zhou X, Yu X, Wei Q, Liu S, Tang L. Time-dependent metabolism of luteolin by human UDP-glucuronosyltransferases and its intestinal first-pass glucuronidation in mice. J Agric Food Chem. 2015;63:8722–8733. doi: 10.1021/acs.jafc.5b02827. [DOI] [PubMed] [Google Scholar]
- 45.Hsu BY, Lin SW, Inbaraj BS, Chen BH. Simultaneous determination of phenolic acids and flavonoids in Chenopodium formosanum Koidz. (djulis) by HPLC-DAD-ESI-MS/MS. J Pharm Biomed Anal. 2017;132:109–116. doi: 10.1016/j.jpba.2016.09.027. [DOI] [PubMed] [Google Scholar]
- 46.Dueñas M, Mingo-Chornet H, Pérez-Alonso JJ, Di Paola-Naranjo R, González-Paramás AM, Santos-Buelga C. Preparation of quercetin glucuronides and characterization by HPLC–DAD–ESI/MS. Eur Food Res Technol. 2008;227:1069–1076. [Google Scholar]
- 47.Abu-Reidah IM, Ali-Shtayeh MS, Jamous RM, Arráez-Román D, Segura-Carretero A. HPLC-DAD-ESI-MS/MS screening of bioactive components from Rhus coriaria L. (Sumac) fruits. Food Chem. 2015;166:179–191. doi: 10.1016/j.foodchem.2014.06.011. [DOI] [PubMed] [Google Scholar]
- 48.Muzzio M, Huang Z, Hu SC, Johnson WD, McCormick DL, Kapetanovic IM. Determination of resveratrol and its sulfate and glucuronide metabolites in plasma by LC-MS/MS and their pharmacokinetics in dogs. J Pharm Biomed Anal. 2012;59:201–8. doi: 10.1016/j.jpba.2011.10.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Lacroix MZ, Puel S, Collet SH, Corbel T, Picard-Hagen N, Toutain PL, Viguié C, Gayrard V. Simultaneous quantification of bisphenol A and its glucuronide metabolite (BPA-G) in plasma and urine: applicability to toxicokinetic investigations. Talanta. 2011;85:2053–9. doi: 10.1016/j.talanta.2011.07.040. [DOI] [PubMed] [Google Scholar]
- 50.Schwaninger AE, Meyer MR, Huestis MA, Maurer HH. Development and validation of LC-HRMS and GC-NICI-MS methods for stereoselective determination of MDMA and its phase I and II metabolites in human urine. J Mass Spectrom. 2011;46:603–14. doi: 10.1002/jms.1929. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Warth B, Sulyok M, Berthiller F, Schuhmacher R, Fruhmann P, Hametner C, Adam G, Fröhlich J, Krska R. Direct quantification of deoxynivalenol glucuronide in human urine as biomarker of exposure to the Fusarium mycotoxin deoxynivalenol. Anal Bioanal Chem. 2011;401:195–200. doi: 10.1007/s00216-011-5095-z. [DOI] [PubMed] [Google Scholar]
- 52.He J, Feng Y, Ouyang HZ, Yu B, Chang YX, Pan GX, Dong GY, Wang T, Gao XM. A sensitive LC-MS/MS method for simultaneous determination of six flavonoids in rat plasma: application to a pharmacokinetic study of total flavonoids from mulberry leaves. J Pharm Biomed Anal. 2013;84:189–95. doi: 10.1016/j.jpba.2013.06.019. [DOI] [PubMed] [Google Scholar]
- 53.Sulaiman C, Balachandran I. LC/MS characterization of antioxidant flavonoids from Tragia involucrata L. Beni-Suef University J Basic Appl Sci. 2016;5:231–235. [Google Scholar]
- 54.Kachlicki P, Piasecka A, Stobiecki M, Marczak Ł. Structural characterization of flavonoid glycoconjugates and their derivatives with mass spectrometric techniques. Molecules. 2016;21:1494. doi: 10.3390/molecules21111494. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Corradini E, Foglia P, Giansanti P, Gubbiotti R, Samperi R, Lagana A. Flavonoids: chemical properties and analytical methodologies of identification and quantitation in foods and plants. Nat Prod Res. 2011;25:469–95. doi: 10.1080/14786419.2010.482054. [DOI] [PubMed] [Google Scholar]
- 56.Fang S, Qu Q, Zheng Y, Zhong H, Shan C, Wang F, Li C, Peng G. Structural characterization and identification of flavonoid aglycones in three Glycyrrhiza species by liquid chromatography with photodiode array detection and quadrupole time-of-flight mass spectrometry. J Sep Sci. 2016;39:2068–78. doi: 10.1002/jssc.201600073. [DOI] [PubMed] [Google Scholar]
- 57.Vukics V, Guttman A. Structural characterization of flavonoid glycosides by multi-stage mass spectrometry. Mass Spectrom Rev. 2010;29:1–16. doi: 10.1002/mas.20212. [DOI] [PubMed] [Google Scholar]
- 58.Cuyckens F, Claeys M. Mass spectrometry in the structural analysis of flavonoids. J Mass Spectrom. 2004;39:1–15. doi: 10.1002/jms.585. [DOI] [PubMed] [Google Scholar]
- 59.Kalt W, McDonald JE, Liu Y, Fillmore SA. Flavonoid metabolites in human urine during blueberry anthocyanin intake. J Agric Food Chem. 2017;65:1582–1591. doi: 10.1021/acs.jafc.6b05455. [DOI] [PubMed] [Google Scholar]
- 60.Marczak Ł, Stobiecki M, Jasiński M, Oleszek W, Kachlicki P. Fragmentation pathways of acylated flavonoid diglucuronides from leaves of Medicago truncatula. Phytochem Anal. 2010;21:224–33. doi: 10.1002/pca.1189. [DOI] [PubMed] [Google Scholar]
- 61.Domon B, Costello CE. A systematic nomenclature for carbohydrate fragmentations in FAB-MS/MS spectra of glycoconjugates. Glycoconjugate J. 1988;5:397–409. [Google Scholar]
- 62.Nobilis M, Holcapek M, Kolárová L, Kopecký J, Kunes M, Svoboda Z, Kvetina J. Identification and determination of phase II nabumetone metabolites by high-performance liquid chromatography with photodiode array and mass spectrometric detection. J Chromatogr A. 2004;1031:229–36. doi: 10.1016/j.chroma.2004.01.031. [DOI] [PubMed] [Google Scholar]
- 63.Levsen K, Schiebel HM, Behnke B, Dötzer R, Dreher W, Elend M, Thiele H. Structure elucidation of phase II metabolites by tandem mass spectrometry: an overview. J Chromatogr A. 2005;1067:55–72. doi: 10.1016/j.chroma.2004.08.165. [DOI] [PubMed] [Google Scholar]
- 64.Karlsson ES, Johnson CH, Sarda S, Iddon L, Iqbal M, Meng X, Harding JR, Stachulski AV, Nicholson JK, Wilson ID, Lindon JC. High-performance liquid chromatography/mass spectrometric and proton nuclear magnetic resonance spectroscopic studies of the transacylation and hydrolysis of the acyl glucuronides of a series of phenylacetic acids in buffer and human plasma. Rapid Commun Mass Spectrom. 2010;24:3043–51. doi: 10.1002/rcm.4740. [DOI] [PubMed] [Google Scholar]
- 65.Davis BD, Brodbelt JS. Regioselectivity of human UDP-glucuronosyl-transferase 1A1 in the synthesis of flavonoid glucuronides determined by metal complexation and tandem mass spectrometry. J Am Soc Mass Spectrom. 2008;19:246–56. doi: 10.1016/j.jasms.2007.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Niemeyer ED, Brodbelt JS. Regiospecificity of human UDP-glucuronosyltransferase isoforms in chalcone and flavanone glucuronidation determined by metal complexation and tandem mass spectrometry. J Nat Prod. 2013;76:1121–32. doi: 10.1021/np400195z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Robotham SA, Brodbelt JS. Identification of flavone glucuronide isomers by metal complexation and tandem mass spectrometry: regioselectivity of uridine 5′-diphosphate-glucuronosyltransferase isozymes in the biotransformation of flavones. J Agric Food Chem. 2013;61:1457–63. doi: 10.1021/jf304853j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Ren J, Meng S, Lekka CE, Kaxiras E. Complexation of flavonoids with iron: structure and optical signatures. J Phys Chem B. 2008;112:1845–50. doi: 10.1021/jp076881e. [DOI] [PubMed] [Google Scholar]
- 69.Davis BD, Needs PW, Kroon PA, Brodbelt JS. Identification of isomeric flavonoid glucuronides in urine and plasma by metal complexation and LC-ESI-MS/MS. J Mass Spectrom. 2006;41:911–20. doi: 10.1002/jms.1050. [DOI] [PubMed] [Google Scholar]
- 70.Fernandez MT, Mira ML, Florêncio MH, Jennings KR. Iron and copper chelation by flavonoids: an electrospray mass spectrometry study. J Inorg Biochem. 2002;92:105–11. doi: 10.1016/s0162-0134(02)00511-1. [DOI] [PubMed] [Google Scholar]
- 71.Satterfield M, Brodbelt JS. Enhanced detection of flavonoids by metal complexation and electrospray ionization mass spectrometry. Anal Chem. 2000;72:5898–906. doi: 10.1021/ac0007985. [DOI] [PubMed] [Google Scholar]
- 72.Babu U, Bhandari S, Garg H. Barbacarpan, a pterocarpan from Crotalaria barbata. Phytochemistry. 1998;48:1457–1459. [Google Scholar]
- 73.Barragán-Huerta BE, Peralta-Cruz J, González-Laredo RF, Karchesy J. Neocandenatone, an isoflavan-cinnamylphenol quinone methide pigment from Dalbergia congestiflora. Phytochemistry. 2004;65:925–8. doi: 10.1016/j.phytochem.2003.11.011. [DOI] [PubMed] [Google Scholar]
- 74.Barrero AF, Cabrera E, Garcia IR. Pterocarpans from Ononis viscosa subsp breviflora. Phytochemistry. 1998;48:187–190. [Google Scholar]
- 75.Bashir A, Hamburger M, Msonthi JD, Hostettmann K. Isoflavones and xanthones from Polygala virgata. Phytochemistry. 1992;31:309–311. doi: 10.1016/s0031-9422(00)95164-1. [DOI] [PubMed] [Google Scholar]
- 76.Smolarz HD, Budzianowski J, Bogucka-Kocka A, Kocki J, Mendyk E. Flavonoid glucuronides with anti-leukaemic activity from Polygonum amphibium L. Phytochem Anal. 2008;19:506–13. doi: 10.1002/pca.1076. [DOI] [PubMed] [Google Scholar]
- 77.Lehtonen HM, Lindstedt A, Järvinen R, Sinkkonen J, Graça G, Viitanen M, Kallio H, Gil AM. 1H NMR-based metabolic fingerprinting of urine metabolites after consumption of lingonberries (Vaccinium vitis-idaea) with a high-fat meal. Food Chem. 2013;138:982–90. doi: 10.1016/j.foodchem.2012.10.081. [DOI] [PubMed] [Google Scholar]
- 78.Grace MH, Esposito D, Timmers MA, Xiong J, Yousef G, Komarnytsky S, Lila MA. Chemical composition, antioxidant and anti-inflammatory properties of pistachio hull extracts. Food Chem. 2016;210:85–95. doi: 10.1016/j.foodchem.2016.04.088. [DOI] [PubMed] [Google Scholar]
- 79.Ryu B, Kim HM, Lee JS, Lee CK, Sezirahiga J, Woo JH, Choi JH, Jang DS. New flavonol glucuronides from the flower buds of Syzygium aromaticum (Clove) J Agric Food Chem. 2016;64:3048–3053. doi: 10.1021/acs.jafc.6b00337. [DOI] [PubMed] [Google Scholar]
- 80.Ichiyanagi T, Kashiwada Y, Shida Y, Sekiya M, Hatano Y, Takaishi Y, Ikeshiro Y. Structural elucidation and biological fate of two glucuronyl metabolites of pelargonidin 3-O-β-D-glucopyranoside in rats. J Agric Food Chem. 2013;61:569–78. doi: 10.1021/jf3032793. [DOI] [PubMed] [Google Scholar]
- 81.Smara O, Julia A, Moral-Salmi C, Vigor C, Vercauteren J, Legseir B. Flavonoïds from Euphorbia guyoniana Boissier & Reuter. J Life Sci. 2014;8 [Google Scholar]
- 82.Tisserant LP, Hubert J, Lequart M, Borie N, Maurin N, Pilard S, Jeandet P, Aziz A, Renault JH, Nuzillard JM, Clément C, Boitel-Conti M, Courot E. (13)C NMR and LC-MS profiling of stilbenes from elicited grapevine hairy root cultures. J Nat Prod. 2016;79:2846–2855. doi: 10.1021/acs.jnatprod.6b00608. [DOI] [PubMed] [Google Scholar]
- 83.Claridge TD. High-resolution NMR techniques in organic chemistry. Vol. 27 Elsevier; 2016. [Google Scholar]
- 84.Krishnamurthy VV, Russell DJ, Hadden CE, Martin GE. 2J,(3)J-HMBC: A new long-range heteronuclear shift correlation technique capable of differentiating (2)J(CH) from (3)J(CH) correlations to protonated carbons. J Magn Reson. 2000;146:232–9. doi: 10.1006/jmre.2000.2141. [DOI] [PubMed] [Google Scholar]
- 85.Fossen T, Andersen ØM. Spectroscopic techniques applied to flavonoids. Flavonoids: Chemistry, Biochemistry and Applications. 2006:37–142. [Google Scholar]
- 86.Hansen SH, Jensen AG, Cornett C, Bjørnsdottir I, Taylor S, Wright B, Wilson ID. High-performance liquid chromatography on-line coupled to high-field NMR and mass spectrometry for structure elucidation of constituents of Hypericum perforatum L. Anal Chem. 1999;71:5235–5241. [Google Scholar]
- 87.Hilbert G, Temsamani H, Bordenave L, Pedrot E, Chaher N, Cluzet S, Delaunay JC, Ollat N, Delrot S, Mérillon JM, Gomès E, Richard T. Flavonol profiles in berries of wild Vitis accessions using liquid chromatography coupled to mass spectrometry and nuclear magnetic resonance spectrometry. Food Chem. 2015;169:49–58. doi: 10.1016/j.foodchem.2014.07.079. [DOI] [PubMed] [Google Scholar]
- 88.Seger C, Godejohann M, Tseng LH, Spraul M, Girtler A, Sturm S, Stuppner H. LC-DAD-MS/SPE-NMR hyphenation. A tool for the analysis of pharmaceutically used plant extracts: identification of isobaric iridoid glycoside regioisomers from Harpagophytum procumbens. Anal Chem. 2005;77:878–85. doi: 10.1021/ac048772r. [DOI] [PubMed] [Google Scholar]
- 89.Paudel L, Wyzgoski FJ, Scheerens JC, Chanon AM, Reese RN, Smiljanic D, Wesdemiotis C, Blakeslee JJ, Riedl KM, Rinaldi PL. Nonanthocyanin secondary metabolites of black raspberry (Rubus occidentalis L.) fruits: identification by HPLC-DAD, NMR, HPLC-ESI-MS, and ESI-MS/MS analyses. J Agric Food Chem. 2013;61:12032–43. doi: 10.1021/jf4039953. [DOI] [PubMed] [Google Scholar]
- 90.Cassani J, Nilsson M, Morris GA. Flavonoid mixture analysis by matrix-assisted diffusion-ordered spectroscopy. J Nat Pro. 2012;75:131–4. doi: 10.1021/np2005264. [DOI] [PubMed] [Google Scholar]
- 91.Piccinelli AL, García Mesa M, Armenteros DM, Alfonso MA, Arevalo AC, Campone L, Rastrelli L. HPLC-PDA-MS and NMR characterization of C-glycosyl flavones in a hydroalcoholic extract of Citrus aurantifolia leaves with antiplatelet activity. J Agric Food Chem. 2008;56:1574–81. doi: 10.1021/jf073485k. [DOI] [PubMed] [Google Scholar]
- 92.Exarchou V, Godejohann M, van Beek TA, Gerothanassis IP, Vervoort J. LC-UV-solid-phase extraction-NMR-MS combined with a cryogenic flow probe and its application to the identification of compounds present in Greek oregano. Anal Chem. 2003;75:6288–94. doi: 10.1021/ac0347819. [DOI] [PubMed] [Google Scholar]
- 93.Kajjout M, Rolando C. Regiospecific synthesis of quercetin O-β-D-glucosylated and O-β-D-glucuronidated isomers. Tetrahedron. 2011;67(25):4731–4741. [Google Scholar]
- 94.Bouktaib M, Atmani A, Rolando C. Regio- and stereoslective synthesis of the major metabolite of quercetin, quercetin-3-O-β-D-glucuronide. Tetrahedron Lett. 2002;43:6263–6266. [Google Scholar]
- 95.Zhou Z-h, Fang Z, Jin H, Chen Y, He L. Selective monomethylation of quercetin. Synthesis. 2010;2010(23):3980–3986. [Google Scholar]
- 96.Alluis B, Dangles O. Quercetin (= 2-(3, 4-dihydroxyphenyl)-3, 57-trihydroxy-4H1-benzopyran-4-one) glycosides and sulfates: chemical synthesis, complexation, and antioxidant properties. Helvetica Chimica Acta. 2001;84(5):1133–1156. [Google Scholar]
- 97.Li NG, Shi ZH, Tang YP, Yang JP, Duan JA. An efficient partial synthesis of 4′-O-methylquercetin via regioselective protection and alkylation of quercetin. Beilstein J Org Chem. 2009;5:60. doi: 10.3762/bjoc.5.60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Zhang M, Jagdmann GE, Jr, Van Zandt M, Sheeler R, Beckett P, Schroeter H. Chemical synthesis and characterization of epicatechin glucuronides and sulfates: bioanalytical standards for epicatechin metabolite identification. J Nat Prod. 2013;76:157–69. doi: 10.1021/np300568m. [DOI] [PubMed] [Google Scholar]
- 99.Needs PW, Williamson G. Syntheses of daidzein-7-yl beta-D-glucopyranosiduronic acid and daidzein-4′,7-yl di-beta-D-glucopyranosiduronic acid. Carbohydrate Res. 2001;330(4):511–515. doi: 10.1016/s0008-6215(00)00326-8. [DOI] [PubMed] [Google Scholar]
- 100.Florent JC, Dong X, Gaudel G, Mitaku S, Monneret C, Gesson JP, Jacquesy JC, Mondon M, Renoux B, Andrianomenjanahary S, Michel S, Koch M, Tillequin F, Gerken M, Czech JSR, Bosslet K. Prodrugs of anthracyclines for use in antibody-directed enzyme prodrug therapy. J Med Chem. 1998;41(19):3572–3581. doi: 10.1021/jm970589l. [DOI] [PubMed] [Google Scholar]
- 101.Oyama K-i, Yoshida K, Kondo T. Recent progress in the synthesis of flavonoids: from monomers to supra-complex molecules. Curr Org Chem. 2011;15(15):2567–2607. [Google Scholar]
- 102.Hasuoka A, Nakayama Y, Adachi M, Kamiguchi H, Kamiyama K. Development of a stereoselective practical synthetic route to indolmycin, a candidate anti-H. pylori agent. Chem Pharm Bull (Tokyo) 2001;49(12):1604–1608. doi: 10.1248/cpb.49.1604. [DOI] [PubMed] [Google Scholar]
- 103.Ferguson JR, Harding JR, Lumbard KW, Scheinmann F, Stachulski AV. Glucuronide and sulfate conjugates of ICI 182,780, a pure anti-estrogenic steroid. Order of addition, catalysis and substitution effects in glucuronidation. Tetrahedron Lett. 2000;41(3):389–392. [Google Scholar]
- 104.Needs PW, Kroon PA. Convenient syntheses of metabolically important quercetin glucuronides and sulfates. Tetrahedron. 2006;62(29):6862–6868. [Google Scholar]
- 105.Gonzalez-Manzano S, Gonzalez-Paramas A, Santos-Buelga C, Duenas M. Preparation and characterization of catechin sulfates, glucuronides, and methylethers with metabolic interest. J Agric Food Chem. 2009;57:1231–1238. doi: 10.1021/jf803140h. [DOI] [PubMed] [Google Scholar]
- 106.Stachulski AV, Meng X. Glucuronides from metabolites to medicines: a survey of the in vivo generation, chemical synthesis and properties of glucuronides. Nat Prod Rep. 2013;30(6):806–848. doi: 10.1039/c3np70003h. [DOI] [PubMed] [Google Scholar]
- 107.O’Neill PM, Scheinmann F, Stachulski AV, Maggs JL, Park BK. Efficient preparations of the beta-glucuronides of dihydroartemisinin and structural confirmation of the human glucuronide metabolite. J Med Chem. 2001;44(9):1467–1470. doi: 10.1021/jm001061a. [DOI] [PubMed] [Google Scholar]
- 108.Fan J, Brown SM, Tu Z, Kharasch ED. Chemical and enzyme-assisted syntheses of norbuprenorphine-3-β-D-glucuronide. Bioconjugate Chem. 2011;22(4):752–758. doi: 10.1021/bc100550u. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Wagner H, Danninger H, Seligmann O, Farkas L. Synthesis of glucuronides in the flavonoid series. I. The first synthesis of a naturally occurring flavonoid glucuronide (quercetin-3-beta-D-glucuronide) Chem Ber. 1970;103:3674–7. doi: 10.1002/cber.19701031131. [DOI] [PubMed] [Google Scholar]
- 110.Luis JG, Andrés LS. Synthesis of danielone (α-hydroxyacetosyringone) J Chem Res Synopses. 1999;1999(3):220–221. [Google Scholar]
- 111.Cruz L, Fernandes I, Évora A, de Freitas V, Mateus N. Synthesis of the main red wine anthocyanin metabolite: malvidin-3-O-β-glucuronide. Synlett. 2017;28(05):593–596. [Google Scholar]
- 112.Ferguson JR, Harding JR, Killick DA, Lumbard KW, Scheinmann F, Stachulski AV. Putative metabolites of fulvestrant, an estrogen receptor downregulator. Improved glucuronidation using trichloroacetimidates. J Chem Soc Perkin Trans. 2001;2001(22):3037–3041. [Google Scholar]
- 113.Stachulski AV, Jenkins GN. The synthesis of O-glucuronides. Nat Prod Rep. 1998;15(2):173–186. doi: 10.1039/a815173y. [DOI] [PubMed] [Google Scholar]
- 114.Li Y, Mo H, Lian G, Yu B. Revisit of the phenol O-glycosylation with glycosyl imidates, BF₃·OEt₂ is a better catalyst than TMSOTf. Carbohydrate Res. 2012;363:14–22. doi: 10.1016/j.carres.2012.09.025. [DOI] [PubMed] [Google Scholar]
- 115.Zhang Q, Raheem KS, Botting NP, Slawin AM, Kay CD, O’Hagan D. Flavonoid metabolism: the synthesis of phenolic glucuronides and sulfates as candidate metabolites for bioactivity studies of dietary flavonoids. Tetrahedron. 2012;68(22):4194–4201. [Google Scholar]
- 116.Wang P, Zhang Z, Yu B. Total synthesis of CRM646-A and -B, two fungal glucuronides with potent heparinase inhibition activities. J Org Chem. 2005;70(22):8884–8889. doi: 10.1021/jo051384k. [DOI] [PubMed] [Google Scholar]
- 117.Al-Maharik N, Botting NP. A facile synthesis of isoflavone 7-O-glucuronides. Tetrahedron Lett. 2006;47(49):8703–8706. [Google Scholar]
- 118.Zhang Z, Yu B. Total synthesis of the antiallergic naphtho-alpha-pyrone tetraglucoside, cassiaside C(2), isolated from cassia seeds. J Org Chem. 2003;68(16):6309–6313. doi: 10.1021/jo034223u. [DOI] [PubMed] [Google Scholar]
- 119.Adinolfi M, Iadonisi A, Ravidà A, Schiattarella M. Versatile use of ytterbium(III) triflate and acid washed molecular sieves in the activation of glycosyl trifluoroacetimidate donors. Assemblage of a biologically relevant tetrasaccharide sequence of Globo H. J Org Chem. 2005;70(13):5316–5319. doi: 10.1021/jo050301x. [DOI] [PubMed] [Google Scholar]
- 120.Boumendjel A, Blanc M, Williamson G, Barron D. Efficient synthesis of flavanone glucuronides. J Agric Food Chem. 2009;57(16):7264–7267. doi: 10.1021/jf9011467. [DOI] [PubMed] [Google Scholar]
- 121.Romanov-Michailidis F, Viton F, Fumeaux R, Lévèques A, Actis-Goretta L, Rein M, Williamson G, Barron D. Epicatechin B-ring conjugates: first enantioselective synthesis and evidence for their occurrence in human biological fluids. Org Lett. 2012;14(15):3902–3905. doi: 10.1021/ol3016463. [DOI] [PubMed] [Google Scholar]
- 122.O’Leary KA, Day AJ, Needs PW, Sly WS, O’Brien NM, Williamson G. Flavonoid glucuronides are substrates for human liver beta-glucuronidase. FEBS Lett. 2001;503(1):103–106. doi: 10.1016/s0014-5793(01)02684-9. [DOI] [PubMed] [Google Scholar]
- 123.Cao Z, Qu Y, Zhou J, Liu W, Yao G. Stereoselective synthesis of quercetin 3-O-glycosides of 2-amino-2-deoxy-D-glucose under phase transfer catalytic conditions. J Carbohydrate Chem. 2015;34(1):28–40. [Google Scholar]
- 124.Bock KW. The UDP-glycosyltransferase (UGT) superfamily expressed in humans, insects and plants: Animal–plant arms-race and co-evolution. Biochem Pharmacol. 2016;99:11–17. doi: 10.1016/j.bcp.2015.10.001. [DOI] [PubMed] [Google Scholar]
- 125.Guillemette C. Pharmacogenomics of human UDP-glucuronosyltransferase enzymes. Pharmacogenomics J. 2003;3:136–158. doi: 10.1038/sj.tpj.6500171. [DOI] [PubMed] [Google Scholar]
- 126.Cheng Z, Radominska-Pandya A, Tephly TR. Cloning and expression of human UDP-glucuronosyltransferase (UGT) 1A8. Arch Biochem Biophys. 1998;356:301–305. doi: 10.1006/abbi.1998.0781. [DOI] [PubMed] [Google Scholar]
- 127.VanEtten H, Mansfield JW, Bailey JA, Farmer EE. Two classes of plant antibiotics: phytoalexins versus “phytoanticipins”. Plant Cell. 1994;6:1191–1192. doi: 10.1105/tpc.6.9.1191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Joseph TB, Wang SWJ, Liu X, Kulkarni KH, Wang J, Xu H, Hu M. Disposition of flavonoids via enteric recycling: enzyme stability affects characterization of prunetin glucuronidation across species, organs, and UGT isoforms. Mol Pharm. 2007;4:883–894. doi: 10.1021/mp700135a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Williamson G, Day AJ, Plumb GW, Couteau D. Human metabolic pathways of dietary flavonoids and cinnamates. Biochem Soc Trans. 2000;28:16. doi: 10.1042/bst0280016. [DOI] [PubMed] [Google Scholar]
- 130.Rietjens IMCM, Vervoort J. Microsomal metabolism of fluoroanilines. Xenobiotica. 1989;19:1297–1305. doi: 10.3109/00498258909043181. [DOI] [PubMed] [Google Scholar]
- 131.Blount JW, Ferruzzi MG, Raftery D, Pasinetti GM, Dixon RA. Enzymatic synthesis of substituted epicatechins for bioactivity studies in neurological disorders. Biochem Biophys Res Comm. 2011;417:457–461. doi: 10.1016/j.bbrc.2011.11.139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Blount JW, Redan BW, Feruzzi MG, Reuhs BL, Cooper BR, Harwood JS, Shulaev V, Pasinetti G, Dixon RA. Synthesis and quantitative analysis of plasma-targeted metabolites of catechin and epicatechin. J Food Agric Chem. 2015;63:2233–2240. doi: 10.1021/jf505922b. [DOI] [PubMed] [Google Scholar]
- 133.Bowles D, Isayenkova J, Lim E-K, Poppenberger B. Glycosyltransferases: managers of small molecules. Curr Opin Plant Biol. 2005;8:254–263. doi: 10.1016/j.pbi.2005.03.007. [DOI] [PubMed] [Google Scholar]
- 134.Kren V, Martinkova L. Glycosides in medicine: The role of glycosidic residue in biological activity. Curr Med Chem. 2001;8:1303–28. doi: 10.2174/0929867013372193. [DOI] [PubMed] [Google Scholar]
- 135.Li Y, Baldauf S, Lim EK, Bowles DJ. Phylogenetic analysis of the UDP-glycosyltransferase multigene family of Arabidopsis thaliana. J Biol Chem. 2001;276:4338–4343. doi: 10.1074/jbc.M007447200. [DOI] [PubMed] [Google Scholar]
- 136.Young ND, Debellé F, Oldroyd G, Geurts R, Cannon SB, et al. The Medicago genome provides insight into evolution of rhizobial symbiosis. Nature. 2011;480:520–524. doi: 10.1038/nature10625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Noguchi A, Horikawa M, Fukui Y, Fukuchi-Mizutani M, Iuchi-Okada A, Ishiguro M, Kiso Y, Nakayama T, Ono E. Local differentiation of sugar donor specificity of flavonoid glycosyltransferase in Lamiales. Plant Cell. 2009;21:1556–1572. doi: 10.1105/tpc.108.063826. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Sawada S, Suzuki H, Ichimaida F, Yamaguchi MA, Iwashita T, Fukui Y, Hemmi H, Nishino T, Nakayama T. UDP-glucuronic acid:anthocyanin glucuronosyltransferase from red daisy (Bellis perennis) flowers. Enzymology and phylogenetics of a novel glucuronosyltransferase involved in flower pigment biosynthesis. J Biol Chem. 2005;280:899–906. doi: 10.1074/jbc.M410537200. [DOI] [PubMed] [Google Scholar]
- 139.Nagashima S, Hirotani M, Yoshikawa T. Purification and characterization of UDP-glucuronate: baicalein 7-O-glucuronosyltransferase from Scutellaria baicalensis Georgi. cell suspension cultures. Phytochemistry. 2000;53:533–538. doi: 10.1016/s0031-9422(99)00593-2. [DOI] [PubMed] [Google Scholar]
- 140.Ono E, Homma Y, Horikawa M, Kunikane-Doi S, Imai H, Takahashi S, Kawai Y, Ishiguro M, Fukui Y, Nakayama T. Functional differentiation of the glycosyltransferases that contribute to the chemical diversity of bioactive flavonol glycosides in grapevines (Vitis vinifera) Plant Cell. 2010;22(8):2856–2871. doi: 10.1105/tpc.110.074625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Marvalin C, Azerad R. Microbial glucuronidation of polyphenols. J Mol Catal B: Enzymatic. 2011;73:43–52. [Google Scholar]
- 142.Huynh NT, Van Camp J, Smagghe G, Raes K. Improved release and metabolism of flavonoids by steered fermentation processes: a review. Int J Mol Sci. 2014;15(11):19369–19388. doi: 10.3390/ijms151119369. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Araújo KCF, de MB Costa EM, Pazini F, Valadares MC, de Oliveira V. Bioconversion of quercetin and rutin and the cytotoxicity activities of the transformed products. Food Chem Toxicol. 2013;51:93–96. doi: 10.1016/j.fct.2012.09.015. [DOI] [PubMed] [Google Scholar]
- 144.He X-Z, Li W-S, Blount JW, Dixon RA. Regioselective synthesis of plant flavonoid glycosides in E. coli. Appl Microbiol Biotechnol. 2008;80:253–260. doi: 10.1007/s00253-008-1554-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Roman E, Roberts I, Lidholt K, Kusche-Gullberg M. Overexpression of UDP-glucose dehydrogenase in Escherichia coli results in decreased biosynthesis of K5 polysaccharide. Biochem J. 2003;374:767–72. doi: 10.1042/BJ20030365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Kim SY, Lee HR, Park K-s, Kim B-G, Ahn J-H. Metabolic engineering of Escherichia coli for the biosynthesis of flavonoid-O-glucuronides and flavonoid-O-galactoside. Appl Microbiol Biotechnol. 2015;99:2233–2242. doi: 10.1007/s00253-014-6282-6. [DOI] [PubMed] [Google Scholar]
- 147.Yang Y, Wang H-M, Tong Y-F, Liu M-Z, Cheng K-D, Wu S, Wang W. Systems metabolic engineering of Escherichia coli to enhance the production of flavonoid glucuronides. RSC Adv. 2016;6:33622–33630. [Google Scholar]
- 148.Li L, Modolo LV, Achnine L, Dixon RA, Wang X. Structure of UGT85H2, an (iso)flavonoid uridine diphosphate glycosyltransferase from the model legume Medicago truncatula. J Biol Chem. 2007;370:951–963. doi: 10.1016/j.jmb.2007.05.036. [DOI] [PubMed] [Google Scholar]
- 149.Shao H, He X, Achnine L, Blount JW, Dixon RA, Wang X. The structure of UGT71G1, a multifunctional triterpene/flavonoid uridine diphosphate glycosyltransferase from the model legume Medicago truncatula. Plant Cell. 2005;17:3141–3154. doi: 10.1105/tpc.105.035055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Modolo LV, Li L, Pan H, Blount JW, Dixon RA, Wang X. Crystal structures of glycosyltransferase UGT78G1 reveal the molecular basis for glycosylation and deglycosylation of (iso)flavonoids. J Mol Biol. 2009;392:1292–302. doi: 10.1016/j.jmb.2009.08.017. [DOI] [PubMed] [Google Scholar]
- 151.Offen W, Martinez-Fleites C, Yang M, Kiat-Lim E, Davis BG, Tarling CA, Ford CM, Bowles DJ, Davies GJ. Structure of a flavonoid glucosyltransferase reveals the basis for plant natural product modification. EMBO J. 2006;25:1396–1405. doi: 10.1038/sj.emboj.7600970. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Brazier-Hicks M, Offen WA, Gershater MC, Revett TJ, Lim E-K, Bowles DJ, Davies GJ, Edwards R. Characterization and engineering of the bifunctional N- and O-glucosyltransferase involved in xenobiotic metabolism in plants. Proc Natl Acad Sci USA. 2007;104:20238–20243. doi: 10.1073/pnas.0706421104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Hiromoto T, Honjo E, Noda N, Tamada T, Kazuma K, Suzuki M, Blaber M, Kuroki R. Structural basis for acceptor-substrate recognition of UDP-glucose: anthocyanidin 3-O-glucosyltransferase from Clitoria ternatea. Protein Sci. 2015;24:395–407. doi: 10.1002/pro.2630. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Miley MJ, Zielinska AK, Keenan JE, Bratton SM, Radominska-Pandya A, Redinbo MR. Crystal structure of the cofactor-binding domain of the human phase II drug-metabolism enzyme UDP-glucuronosyltransferase 2B7. J Mol Biol. 2007;369:498–511. doi: 10.1016/j.jmb.2007.03.066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Fujiwara R, Yokoi T, Nakajima M. Structure and protein–protein interactions of human UDP-glucuronosyltransferases. Frontiers Pharmacol. 2016;7 doi: 10.3389/fphar.2016.00388. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Osmani SA, Bak S, Imberty A, Olsen CE, Møller BL. Catalytic key amino acids and UDP-sugar donor specificity of a plant glucuronosyltransferase, UGT94B1: molecular modeling substantiated by site-specific mutagenesis and biochemical analyses. Plant Physiol. 2008;148:1295–1308. doi: 10.1104/pp.108.128256. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.King CD, Green MD, Rios GR, Coffman BL, Owens IS, Bishop WP, Tephly TR. The glucuronidation of exogenous and endogenous compounds by stably expressed rat and human UDP-glucuronosyltransferase 1.1. Arch Biochem Biophys. 1996;332:92–100. doi: 10.1006/abbi.1996.0320. [DOI] [PubMed] [Google Scholar]
- 158.Pritchett LE, Atherton KM, Mutch E, Ford D. Glucuronidation of the soyabean isoflavones genistein and daidzein by human liver is related to levels of UGT1A1 and UGT1A9 activity and alters isoflavone response in the MCF-7 human breast cancer cell line. J Nutr Biochem. 2008;19:739–745. doi: 10.1016/j.jnutbio.2007.10.002. [DOI] [PubMed] [Google Scholar]
- 159.Green MD, King CD, Mojarrabi B, Mackenzie PI, Tephly TR. Glucuronidation of amines and other xenobiotics catalyzed by expressed human UDP-glucuronosyltransferase 1A3 Drug Metab. Dispositon. 1998;26:507–512. [PubMed] [Google Scholar]
- 160.Green MD, Tephly TR. Glucuronidation of amines and hydroxylated xenobiotics and endobiotics catalyzed by expressed human UGT1. 4 protein Drug Metab. Disposition. 1996;24:356–363. [PubMed] [Google Scholar]
- 161.Ebner T, Burchell B. Substrate specificities of two stably expressed human liver UDP-glucuronosyltransferases of the UGT1 gene family Drug Metab. Disposition. 1993;21:50–55. [PubMed] [Google Scholar]
- 162.Mojarrabi B, Mackenzie PI. The human UDP glucuronosyltransferase, UGT1A10, glucuronidates mycophenolic acid. Biochem Biophys Res Commun. 1997;238:775–778. doi: 10.1006/bbrc.1997.7388. [DOI] [PubMed] [Google Scholar]
