You are on page 1of 14

Critical Reviews in Food Science and Nutrition

2023, VOL. 63, NO. 22, 5953–5966


https://doi.org/10.1080/10408398.2022.2026291

Review

Current knowledge of anthocyanin metabolism in the digestive tract:


absorption, distribution, degradation, and interconversion
Hailong Guia,b, Lijun Sunc, Ruihai Liua,d, Xu Sia,b, Dongnan Lia,b , Yuehua Wanga,b, Chi Shua,b, Xiyun
Suna,b, Qiao Jianga,b, Yanyan Qiaoa,b, Bin Lia,b and Jinlong Tiana,b
a
College of Food Science, Shenyang Agricultural University, Shenyang, Liaoning, China; bKey Laboratory of Healthy Food Nutrition and
Innovative Manufacturing, Shenyang, Liaoning, China; cCollege of Food Science and Engineering, Northwest A & F University, Yangling,
Shaanxi, China; dDepartment of Food Science, Cornell University, Ithaca, NY, USA

ABSTRACT KEYWORDS
Potential roles for anthocyanins in preventing various chronic diseases have been reported. These Anthocyanins;
compounds are highly sensitive to external conditions and are susceptible to degradation, which metabolism;
increases the complexity of their metabolism in vivo. This review discusses anthocyanin metabolism absorption;
in the digestive tract, phase I and II metabolism, and enterohepatic circulation (EHC), as well as distribution;
interconversion
their distribution of anthocyanins in blood, urine, and several organs. In the oral cavity, anthocyanins
are partly hydrolyzed by microbiota into aglycones which are then conjugated by glucuronidase.
In stomach, anthocyanins are absorbed without deglycosylation via specific transporters, such as
sodium-dependent glucose co-transporter 1 and facilitative glucose transporters 1, while in small
intestine, they are mainly absorbed as aglycones. High polymeric anthocyanins are easily degraded
into low-polymeric forms or smaller phenolic acids by colonic microbiota, which improves their
absorption. Anthocyanins and their derivatives are modified by phase I and II metabolic enzymes
in cells and are released into the blood via the gastrovascular cavity into EHC. Notably, interconversion
can be occurred under the action of enzymes such as catechol-O-methyltransferase. Taking together,
differences in anthocyanin absorption, distribution, metabolism, and excretion largely depend on
their glycoside and aglycone structures.

1. Introduction clarified. This review aims to summarize the existing evi-


dence and provide a comprehensive knowledge-base of
As secondary metabolites, anthocyanins are found in many anthocyanin metabolism, in order to improve the under-
flowers, fruits, roots, stalks, leaves and seeds, accounting for standing of these compounds and guide future research.
a large proportion of natural blue, purple, orange, and red
hues. The types and concentrations of anthocyanins vary
between different locations, seasons, and specific conditions 2. Structural features of anthocyanins
(De Rosas et al. 2019). Anthocyanins are known to possess
health benefits, including antioxidation, anti-inflammatory, Anthocyanins are derived from the oxidation of
anti-obesity, antidiabetic, antihypertensive, anti-cardiovascular 2-phenylbenzopyrylium or flavylium salts of polyhydroxyl-
disease, anticancer, and neuroprotective properties, mediated ated or polymethoxylated glycosides or acylglycosides (Mazza
via effects on gene expression, cell signaling, hormone secre- and Miniati 1993). Anthocyanins mainly exist as glycosides
tion, and enzyme activity (Tian et al. 2019a; Si et al. 2021). linked with aglycone anthocyanidin-chromophores and com-
Although the daily intake of anthocyanins is generally mon sugar moieties (including glucose, galactose, arabinose,
sufficient, many studies have demonstrated their low bio- rutinose, rhamnose and xylose) bound as mono-, di-, or
availability which means that they cannot be fully absorbed trisaccharides and acylated with phenolic or aliphatic sub-
or utilized. The properties and applications of anthocyanins stituents (Mazza and Miniati 1993). Anthocyanidins with
have been thoroughly investigated for nearly a century (De different hydroxyl and methoxyl groups on the B-ring have
Rosas et al. 2019; Rusishvili et al. 2019; Zang et al. 2021). been found in nature, the six most common being cyanidin
Researchers have studied the metabolic processing of antho- (Cy), pelargonidin (Pg), peonidin (Pn), delphinidin (Dp),
cyanins in the human body to explain the cause of their petunidin (Pt), and malvidin (Mv) (Xie et al. 2018). These
low bioavailability and mechanisms of their health benefits, B-ring substituents account for the natural pigmentation of
aiming to determine and maximize these advantageous prop- the aglycones, with increased hydroxyl and methoxyl groups
erties (Lila et al. 2016; Kalt 2019; Li et al. 2020). However, favoring the blue and red ends of the spectrum, respectively
many aspects of anthocyanin metabolism remain to be (Rusishvili et al. 2019) (Figure 1a). In liquid, anthocyanins

CONTACT Bin Li libinsyau@163.com; Jinlong Tian 2017500015@syau.edu.com


Hailong Gui and Lijun Sun made the equal contribution to the article.
© 2022 Taylor & Francis Group, LLC
5954 H. GUI ET AL.

Figure 1. (a) Six common anthocyanidins and their glucoside forms with different substituent groups (in red dotted frames). The corresponding colors of the
anthocyanidins are shown in the background. (b) Anthocyanin metabolism in the human oral cavity according to Mallery et al. (2011). Anthocyanins and
derivatives: C3S, cyanidin-3 sambubioside; C3XR, cyanidin 3-xylosylrutinoside; PCA-G, glucuronidated protocatechuic acid; C3S-G, glucuronidated cyanidin-3
sambubioside. Enzymes: β-glucosidase, LPH, lactase phlorizin hydrolase; UGTs, UDP-glucuronosyl transferases; COMT, catechol-O-methyltransferase; UDP-Glu-DH,
UDP-glucose-dehydrogenase. Transporters: BCRP, breast cancer resistance protein; SGLT1 sodium dependent glucose co-transporter1. Supra enzymes and trans-
porters distribute granual layers, apinous layers, oral muscosa, and terminal ducts of the salivary glands.

present as the flavylium cation under acidic conditions, with microbiota. In oral models (including oral cells, saliva and
the hemiketal, cis-, and trans-chalcone species appearing at mucosal pellicle), anthocyanins had greater interaction with
higher pH when the hydroxyl substituents deprotonate and oral constituents than flavanols (Soares et al. 2020). Following
yield the respective anionic base. In basic conditions the acetonitrile extraction, additional anthocyanins (1–3% initial
different forms can coexist and interconvert to maintain an amounts) were detected in oral proteins, but much less than
equilibrium at different pH levels (Oliveira et al. 2019a). the total loss (>10%) (Kamonpatana et al. 2012). After heat
inactivation or cell-free treatment, anthocyanin degradation
in saliva decreased significantly compared with a control
3. Metabolism of anthocyanins group (Kamonpatana et al. 2012). Therefore, most anthocy-
anins were lost via hydrolysis by β-glycosidase secreted from
3.1. Oral cavity: the initial site of anthocyanin
bacteria and oral epithelial cells, as proven by antimicrobial
metabolism
treatment (Mallery et al. 2011; Kamonpatana et al. 2012).
Digestion of food, including anthocyanins, begins in the oral Aglycones decreased to varying degrees in enzyme-free arti-
cavity. During the incubation of human saliva, the degradation ficial saliva: Pt (10%), Dp (30%), Cy (40%), Pn (50%), and
of anthocyanins in various berries (blueberry, chokeberry, and Mv (50%), depending on their structural instability
strawberry) ranged from 8 to 90%, with the mean exceeding (Kamonpatana et al. 2012). The -OH and -OCH3 substituents
10% within 5 min (Kamonpatana et al. 2012, 2014). Losses of in the B-ring of aglycones determine their stability and affect
anthocyanins from fluids were relatively low while semisolid their degradation in the oral cavity. The degradation rate of
and solid foods with high viscosity, which are retained for Pg3G was much higher than Pg3R and C3G in saliva, while
longer in the oral cavity, had higher losses. In a recent study, C3R was higher than trisaccharide (Kamonpatana et al.
intact anthocyanins and their metabolites were detected in 2014). As for the glucoside moiety of anthocyanins, degra-
blood a few minutes after administration, implying they were dation of monosaccharide conjugation was significantly
absorbed from the oral cavity (Braga et al. 2018). Some clinical higher than for di- and trisaccharides (Kamonpatana et al.
studies have demonstrated that absorption of drugs is more 2014). Moreover, hexose conjugates had a higher degradation
efficient from the oral cavity than the intestinal tract (Cury rate than pentose conjugates in saliva (Kamonpatana et al.
et al. 2021). The metabolism of anthocyanins in the oral cavity 2012). In summary, anthocyanin degradation in saliva appears
requires clarification. to be affected by the aglycone and glycoside moieties.
Factors impacting anthocyanin metabolism in the oral Anthocyanins had been detected at high levels in saliva
cavity may include interference by proteins, cells, and in the oral cavity and can cross the buccal layer to reach
Critical Reviews in Food Science and Nutrition 5955

oral cells with a downward trend. (Ugalde et al. 2009). In confirmed in IHC analysis, revealing metabolites appearing
alkaline conditions, anthocyanins transform into hemiacetal within 5 min. Thus, anthocyanin metabolites are directly
and chalcone forms (Fernandes et al. 2015). Chalcone glu- generated in the oral cavity by a combination of oral cells
cosides were detected in saliva during the incubation of and microbiota. The transmission of gut metabolites in bio-
C3G, demonstrating the cleavage of the heterocyclic C-ring logical fluids needs to be verified. The metabolic pathway
of anthocyanin glycosides prior to their deglycosylation of anthocyanins is complicated and, although many products
(Kamonpatana et al. 2012). After deglycosylation by β-D-ga- have been detected, further research in this area is
lactosidase and β-D-glucosidase from microbiota, anthocy- still needed.
anins produced aglycones which were automatically
degraded into phloroglucinol aldehyde (PGA) and proto-
catechuic acid (PCA) (Han et al. 2021). Derived from Cy 3.2. Stomach: a route for intact anthocyanin
rather than C3G, PCA was readily detected in saliva and absorption
its glucuronide form appeared after absorption by oral epi-
thelium, as cyanidin-3 sambubioside (Mallery et al. 2011). The acidic gastric juices in the stomach are a favorable
In IHC (immunohistochemistry) analyses, catechol-O-meth- environment for anthocyanin stability and absorption.
yltransferase (COMT), UDP-glucuronosyl transferases Anthocyanins could be rapidly absorbed without hydrolysis
(UGTs), UDP-glucose-dehydrogenase (UDP-Glu-DH), breast (20–25%) in the stomach and appear in the plasma within
cancer resistance protein (BCRP), β-glucosidase, lactase 30 min of intake (Mueller et al. 2017). Peak plasma concen-
phlorizin hydrolase (LPH), and sodium-dependent glucose trations were reached within 0.5–2 h, which was consistent
co-transporter 1 (SGLT 1) were found at the end of the with the time for emptying the stomach, implying that slow-
salivary ducts and stratified squamous surface epithelia ing the gastric emptying rate could significantly improve
(Mallery et al. 2011), suggesting that the glucuronidation, anthocyanin absorption (Passamonti et al. 2009). Although
methylation or glutathionylation of anthocyanins might p-hydroxybenzoic acid was detected in rat stomach, this
occur in the mucosa. Additionally, arylsulfatase and was due to unstable pelargonidin rather than the action of
β-glucuronidase, two key enzymes responsible for deglucu- the stomach (El Mohsen et al. 2006). In a gastric cell model
ronidation in phase II metabolism of anthocyanins, were (MKN-28, moderately differentiated adenocarcinoma stom-
also detected in human saliva (Mallery et al. 2011). The ach cell), the absorption of anthocyanins was independent
identification of these transporters and enzymes shows that of pH variation and a proton gradient (Fernandes et al.
anthocyanin metabolism in the oral cavity is more compli- 2012). In the same cell model, anthocyanin transport rates
cated than previously thought and involves a series of com- from jambo, jamelão, and jaboticaba varied over 3 h and
pounds. The possible pathway involved is schemed in the diglucoside of Cy was in line with Mv but higher than
Figure 1b. Dp and Pt, due to differences in stability and initial con-
The degradation of anthocyanins in the oral cavity is centration (Marques Peixoto et al. 2016). Among the agly-
primarily attributed to enzymatic hydrolysis (especially bac- cones, Cy is more stable than Dp and Pt due to its fewer
terial) and varies with their aglycone/glycoside structures. hydroxyl groups, which enhances transportation. The sugar
After enzymatic hydrolysis, anthocyanin aglycones are more moiety might play a crucial role in anthocyanin transport
lipophilic and orally absorbable, and automatically break- in a gastric cell model, but not the position of additional
down into PCA due to their instability. LPH and β-glucosidase glucose residue (Oliveira et al. 2019b). This also explains
(enzymes distributed in oral epithelium, mucosa, and saliva) the insignificant difference in transport rates between
catalyze the transformation of anthocyanins into aglycones malvidin-3,5-O-diglc and malvidin- 3,5-O-diglcCoum, as
except for cyanidin-3 sambubioside whose absorption well as between malvidin-3-O-glc and malvidin-3,5-O-diglc
depends on transportation by SGLT1. In oral cells, UGTs (Han et al. 2020). In addition, the lower transport efficiency
induce the glucuronidation of aglycones, anthocyanins, and of malvidin-3-O-glucur showed that the carboxyl group at
PCA, and some products return into the cavity via BCRP the C6” position of the glucose could impede the recognition
efflux. These glucuronidated compounds undergo deglucu- of transporters for sugar moiety (Han et al. 2020).
ronidation by arylsulfatase and β-glucuronidase in saliva, Interestingly, the transport of anthocyanins in extracts was
producing compounds for next circulation. The existence of significantly lower than the purified forms, and the addi-
these various transporters and enzymes implies that oral tional sugar slowed down the transport of anthocyanin. This
anthocyanins undergo not just degradation but also absorp- demonstrated that there was a competitive relationship
tion, transportation, derivatization, and EHC. Although sep- between anthocyanins (or with other compounds) for com-
arated by the oral-gut barrier, oral microbiota could reach mon transporters (Oliveira et al. 2015; Han et al. 2020).
the intestinal mucosa when the barrier malfunctions, while Fernandes et al. (2012) concluded that the transport of
gut-to-oral microbial transmission could happen due to anthocyanin in MKN-28 cells was time-consuming and
inter- and intrapersonal manners (Lira-Junior and Boström saturable.
2018; Park et al. 2021). Thus, the oral and gut microbial Due to hydrophilicity and size (Mw ≈ 500 g/mol),
profiles are likely to have some similarities, leading to sim- anthocyanins are unable to cross the gastric epithelium
ilar anthocyanin metabolism in the oral cavity and intestinal via para-cellular absorption. At pH 2, anthocyanins mainly
tract. Furthermore, the prerequisite of anthocyanin metab- existed as polar flavylium cations, which impeded their
olites (enzymes and transporters) in oral cells has been passive diffusion in the gastric mucosa (Mueller et al.
5956 H. GUI ET AL.

2017). Bilitranslocase, an organic anion carrier, is widely could hydrolyze anthocyanins, releasing the more hydro-
expressed in the epithelia of the stomach, small intestine, phobic aglycones, which facilitated their absorption by
liver, kidney, and vascular endothelium. This carrier binds passive diffusion (Rodriguez-Mateos et al. 2014). The two
to glycosides of anthocyanins, improving the direct absorp- main hexose transporters (SGLT1 and GLUT2) had also
tion of the intact molecules. The transport efficiency via been proven to be involved in the absorption of C3G by
bilitranslocase was related to its affinity with the antho- Caco-2 cells (Zou et al. 2014). Following treatment of
cyanin, which particularly depends on the structure of Caco-2 cells with red grape skin anthocyanins, the expres-
the anthocyanin’s glycoside portion (Passamonti et al. sion of GLUT2 increased and glucose uptake decreased,
2003, 2009). Except for acylated anthocyanins, mono- and which illustrated the competition between anthocyanins
di-glucosylated forms were good ligands for bilitranslocase and glucose for the same transporter (Faria et al. 2009).
(Novotny, Clevidence, and Kurilich 2012). Anthocyanin Hence, enzymes and transporters are both involved in
absorption from the stomach clearly involves glucose anthocyanin absorption in intestinal models, facilitating
transporters, as evidenced by the transport rate decreasing their entry into enterocytes.
following the genetic silencing of facilitative glucose trans- Having entered the enterocytes, aglycones underwent
porters 1 and 3 (GLUT1, 3) (Han et al. 2020). Other phase II metabolism, mainly producing glucuro- and
related transporters had also been detected, including sulpho-conjugates (Manach et al. 2005) with decreased
SGLT1, organic cation transporter 1 (OCT1), organic transmembrane capability. In a study by Chen et al. (2014),
anion transporter 2 (OAT2), and sodium/monocarboxylate associated transporters mediated the translocation of fla-
transporters (SMCT1, 2) (Fernandes et al. 2012; Oliveira vonoids across the cell membrane (Figure 2). ATP-binding
et al. 2015). cassette (ABC) transporters played a vital role in the dis-
Anthocyanins are stable in the acidic gastric juices (pH tribution of many conjugated flavonoids, and ABCC2, sit-
1.5–4), except for the most active types like pelargonidin. uated on the enterocytes, transported conjugates back to
Their absorption is mediated by transporters with rates the gut lumen. Some well-known drug transporters, includ-
depending on the interaction which is influenced by the ing multidrug resistance-associated proteins (MRPs), BCRP,
anthocyanin structure. In addition, absorption from the organic anion transporters (OATs), organic anion trans-
stomach is time-dependent, which means the mitigation of porting polypeptides (OATPs), P-glycoprotein (P-gp), and
gastric emptying is a feasible way to increase anthocyanin peptide transporters (PEPTs), were extensively distributed
absorption and improve their bioavailability. in the small intestine and liver (DeGorter et al. 2012).
Following the translocation of MRPs and P-gp, some
metabolites are returned to the lumen while others travel
3.3. Small intestine: the major location of anthocyanin
to the liver via the bloodstream for further phase II metab-
absorption
olism (Dreiseitel et al. 2009). As a result, anthocyanin
As the longest part of the digestive tract, the small intestine conjugates are transported in different directions with spe-
plays a major role in the absorption of nutrients from food. cific transporters in enterocytes. It should be noted that
According to pharmacokinetic data, anthocyanin absorption
was greatest from the jejunum, less from the duodenum,
and negligible from the ileum (Talavéra et al. 2004). In an
intestinal cell model (Caco-2 cells), anthocyanin transport
rates varied among various fruits and anthocyanins, and
bioavailability was low (Liu et al. 2014).
The absorption of anthocyanins from the small intestine
is related to their stability and hydrophobicity, which are
determined by their structure. Differences in their B-ring
substituents (Figure 1) are reflected in their absorption:
Dp has two -OH groups, Mv has two -OCH 3 groups and
Pn has one, which makes Dp more hydrophilic than the
others. The transport rate of hydrophobic anthocyanins
was higher than hydrophilic ones (Liu et al. 2014). The
sugar moiety was crucial for anthocyanin absorption in
cell models, but the position of the glucoside residue was
not (Oliveira et al. 2019b). For instance, the transport
rates of Cy-3,5-O-diglu and C3G in Caco-2 cells were not
significantly different after 2 h (Marques Peixoto et al.
2016), as were malvidin-3-O-glucoside and malvidin-3,
5-O-diglucoside from red wine (Han et al. 2020). The
hydrophilic flavonoid glycosides were hydrolyzed to more
hydrophobic aglycones, which could cross the phospholipid Figure 2. Schematic representation of anthocyanin derivatization and transport
bilayer cell membranes via passive diffusion or specific in the stomach, small intestine, colon, liver, and kidney based on the compre-
transporter (Chen et al. 2014). LPH from enterocytes hensive knowledge of anthocyanin metabolism.
Critical Reviews in Food Science and Nutrition 5957

some anthocyanins in the upper gastrointestinal tract (pH only detectable after consumption of a high dose flavonoids
5.0– 6.5) converted to a combination of hemiketal, qui- or when administered intravenously, which could saturate
nonoidal, and chalcone forms, which were mainly degraded the phase II enzymes (Chen et al. 2014).
to PCA and 2, 4, 6-trihydroxybenzaldehyde via ring fission After absorption, anthocyanins and their aglycones were
(Chen et al. 2019). However, the metabolites of bilberry conjugated in phase I and II metabolism by various enzymes,
anthocyanins in ileal effluent were phloroglucinol aldehyde, changing their hydrophilicity for elimination or accumula-
syringic acid, vanillic acid, and 3-O-methylgallic acid tion. These conjugations are affected by the saturation level
(Mueller et al. 2017). C3G loss and its degradation prod- of anthocyanin. The quantity and quality of anthocyanin
ucts were not notably different in the presence and absence derivatives are related to both their concentration and mode
of Caco-2 cells, indicating that the major loss of C3G was of administration. Derivatives from anthocyanins commonly
attributable to chemical degradation rather than cellular appeared as glucuronide, methyl- and sulfoconjugates in
action (Faria et al. 2009). urine, blood, intestinal epithelia, liver, and kidney (Kay,
Anthocyanins are absorbed in their intact forms from Mazza, and Holub 2005; Di el al 2017; Tian et al. 2019b;
the small intestine via glucose transporters or hydrolyzed Netzel, Wright, and Sultanbawa 2020). Conjugates were gen-
into aglycones to facilitate their entry into enterocytes. Some erally more abundant than their native compounds in vivo,
may be degraded into small phenolic acids under alkaline showing their health benefits were associated with antho-
conditions. In the enterocytes, anthocyanins and their agly- cyanins (Williamson and Clifford 2010). Studies of this
cones undergo phase II metabolism and their conjugates aspect of anthocyanin metabolism are limited and further
are partly pumped out by efflux transporters (ABCC2, exploration of these mechanisms are needed.
MRPs, and P-gp) while other hydrophilic and hydrophobic
metabolites move with the help of transporters (ABC, BCRP,
OATs, OATPs, and PEPTs) into capillaries and/or lymph 3.4.2. Enterohepatic circulation of anthocyanins
vessels, for further transport (Kalt 2019). As described above, a proportion of anthocyanin conjugates
within enterocytes are pumped out by efflux transporters
and then are transported to the liver via the bloodstream
3.4. Phase I and II metabolism and enterohepatic
where they undergoe EHC. Radiolabeling techniques have
circulation of anthocyanins
demonstrated the participation of anthocyanins in EHC by
3.4.1. Phase I and II metabolism of anthocyanins detecting them in bile. Furthermore, the jejunum (the main
Most compounds including anthocyanins, undergo metabolic site of EHC) had also been shown to be a major site of
transformation in vivo (particularly in the liver), involving anthocyanin absorption (Verediano et al. 2021).
phase I (demethylation, hydrolysis) and phase II (glucuro- Pharmacokinetic studies show that EHC was crucial for
nidation, sulfation, methylation) metabolism. Metabolic compounds to amplify their bioactivity by increasing their
enzymes, including cytochrome P450 1 A (CYP1A), UGTs, distribution and prolonging the half-life of their glucuronide
COMT, and sulfotransferases (SULTs) are highly expressed and sulfated metabolites (Zhou et al. 2019). Bile acids could
in the liver and small intestine. UGTs produced glucuronic be recycled over 20 times via EHC and anthocyanins in
conjugates with improved hydrophilicity facilitating their bile could be retained for a long time, repeatedly undergoing
excretion in urine and bile (Rowland, Miners, and Mackenzie phase II metabolism (Lila et al. 2016). Some anthocyanins
2013). SULTs also increased the hydrophilicity of substrates generate multiple derivatives are retained for varying lengths
via sulfation and were regulated by low concentrations of of time. Di-, tri-, and higher conjugated derivatives have
flavonoids in vivo (while UGTs are not) (Huang et al. 2009). been detected in many studies, indicating that they have
Unlike the above enzymes, COMT increased the lipophilicity been conjugated and recycled during repeated EHC
of its substrates, enhancing their accumulation in tissues (Fornasaro et al. 2016). Anthocyanins are translocated into
(Chen et al. 2014). Conjugations via glucuronidation was enterocytes via transporters, are conjugated by enzymes like
very common for anthocyanins, with products being iden- UGTs, and some products are returned to the lumen via
tified at a high concentration in the circulatory system. efflux transporters (this is termed enteroenteric circulation).
Methylation could convert C3G into Pn3G (the methylic The remaining metabolites enter the venous system along
form of C3G) after a high intake of C3G (Talavéra et al. with anthocyanins from the stomach, bind with serum albu-
2003; Fornasaro et al. 2016). In contrast to glucuronidation min, and travel to the liver for further conjugation (Cahyana
and methylation, sulfation had been less reported, with fewer and Gordon 2013). Following conjugation, anthocyanin
sulfated metabolites being identified (Straßmann, Passon, derivatives could leave the hepatocytes with the help of
and Schieber 2021). Phase I metabolism is less important efflux transporters, and were excreted into the bile. Some
than phase II in the distribution of flavonoids. Mediated by were hydrolyzed to aglycones which could be directly reab-
CYPs, phase I metabolism oxidized flavonoids via demeth- sorbed by the small intestine or degraded to low molecular
ylation and hydroxylation, creating additional reactive sites weight aromatic acids under the action of colonic microbiota
for phase II conjugation (Chen et al. 2014). CYPs were (Banaszewski et al. 2013). The rest are ultimately excreted
widely spread in tissues, especially in liver. CYP1A2 mainly via the urine. This process is illustrated in Figure 3.
modulated the hydroxylation and demethylation of flavo- Under EHC, different doses of anthocyanins activate
noids at positions 3′ and 4′ of the B-ring, respectively phase I and II enzymes to differing degrees, resulting in
(DeGorter et al. 2012). However, this CYP metabolism was different products and conjugates after recycling many times.
5958 H. GUI ET AL.

mono-glycoside forms, then into aglycones which were auto-


matically transformed into quinoid bases and to an aldehyde
and a phenolic acid via an α-diketone intermediate (Kay,
Kroon, and Cassidy 2009; Nagar, Okun, and Shpigelman
2020). In vivo, gut microbiota not only hydrolyzed the gly-
cosylated group of anthocyanins, but also cleaved the antho-
cyanidin heterocycle (from C-ring), producing benzoic acids
(from B-ring) and phloroglucinol derivatives (from A-ring)
(Fernandes et al. 2015). Thus, different combinations of
anthocyanin monomer could decompose into various phenolic
acids and aldehyde metabolites (Tian et al. 2019b). The main
metabolites produced by intestinal probiotics from cyanidin
glycosides were 2,4,6-trihydroxybenzaldehyde, ρ-coumaric,
protocatechuic, vanillic acids, and phenolic conjugates (e.g.,
hippuric, phenylacetic, and phenylpropenoic acids) (Ludwig
et al. 2015; Mueller et al. 2017), while delphinidin-based
anthocyanins yielded 2,4,6-trihydroxybenzaldehyde, gallic, and
syringic acids. The main metabolites of malvidin, pelargonidin,
and peonidin glucosides were syringic, 4-hydroxybenzoic, and
vanillic acids, respectively (Eker et al. 2019; Dharmawansa,
Hoskin, and Rupasinghe 2020). Some phenolic acids could
be converted further by gut microbiota. For example, esters
of caftaric acid, chlorogenic acid, and caffeic acid were con-
verted to 3-(3′-hydroxyphenyl) propionic acid and benzoic
acid (Ozdal et al. 2016). These microbial metabolites could be
reabsorbed for further metabolism or excreted via the urine.
Large numbers of anthocyanin metabolites have greater
chemical and microbial stability than their parent com-
pounds, along with antioxidant and other physiological
Figure 3. Simple schematic representation of anthocyanin metabolism in the properties. It had been proven that some anthocyanin
digestive system.
metabolites produced by colonic microbiota possessed higher
bioavailability and bioactivity than their parents (Chen et al.
Compared with pharmaceutical drugs undergoing EHC, 2017; Nagar, Okun, and Shpigelman 2020). Additionally, the
anthocyanins are more susceptible to degradation and deri- interaction between gut microbiota and anthocyanins and
vatization, which increases the analytical challenge. However, their derivatives is a “win-win” situation—nutrient metab-
EHC is indispensable for anthocyanins to perform their olism is enhanced and the composition of the microflora
functions at low concentrations over extended periods. is modulated by the promotion of beneficial bacteria and
suppression of detrimental bacteria. For instance, anthocy-
anins from black raspberry induced the growth of
3.5. Colon: the primary site of anthocyanin Eubacterium rectale, Faecalibacterium prausnitzii and
degradation Lactobacillus, and the inhibition of Desulfovibrio spp. and
Enterococcus spp. (Chen et al. 2017). Moreover, gut metab-
The colon (the distal gastrointestinal tract) is the major olites from anthocyanins stimulated macrophage reverse
habitat for microbiota in the human body, with more than cholesterol transport and prevented atherosclerotic lesions
a trillion microorganisms having a symbiotic relationship in gut, indicating that some bio-functions of anthocyanins
with their host (Kataoka 2016). Equipped with large amounts were mediated by the microbiota (Tian et al. 2019b).
of enzymes, the gut microbiota has a powerful capacity to Consequently, the metabolism of colonic microbiota is vital
degrade and absorb nutrients. Some anthocyanins in food for the bioavailability and bioactivity of anthocyanins and
matrices are hard to degrade and/or absorb in the upper their health benefits.
gastrointestinal tract but easy for gut microbiota to decom- However, the health benefits of anthocyanins can be
pose into small, absorbable phenolic acids. It has long been unpredictable due to several uncontrollable points including
speculated that the colon had a major role in anthocyanin human individuality (genetic makeup, physiological status,
metabolism (Selma et al. 2020). etc.), lifestyle (physical activity, dietary pattern, etc.), meth-
The initial action of gut microbiota on anthocyanins was odological approach, evaluation system (trial designs, ana-
deglycosylation (producing aglycones) catalyzed by α-L-rham- lytical protocols, measurement procedure, etc.), and the
nosidase and β-D-glucosidase and was dependent upon the interactions of microflora and anthocyanins (Cortés-Martín
type of glycoside (Ozdal et al. 2016). At neutral pH, antho- et al. 2020). These factors combine to affect the quality and
cyanins with multiple glycosides were first degraded into quantity of gut metabolites, resulting in variable health
Critical Reviews in Food Science and Nutrition 5959

benefits. Polyphenol-related metabotypes may well reflect diet including blueberry, anthocyanins were detected in
the character of the microflora impacting on health. pig plasma, urine, and tissues (eyes, liver, cortex, and cer-
Metabotype was an effective way to predict the status of ebellum at 1.58, 1.30, 0.88, and 0.66 pmol/g, respectively)
gut microbiota and achieve the clustering of subjects, thus (Kalt et al. 2008). Unexpectedly, anthocyanins were also
avoiding the interference from individual variations in the detected in the control group on a grain diet (equivalent
verification of bioactive compounds (Cortés-Martín et al. to 0.0002% blueberry powder). The glucuronide forms of
2020). Although the major phenolic acids arising from some Cy, Mv, Pn, Dp, and Pt were detected in different regions
aglycones had been confirmed, these may also be derived of the pig brain after eight weeks on the blueberry diet
from other compounds in vivo. Anthocyanins are susceptible (Milbury and Kalt 2010). Pn3G, C3G and galactosides of
to external derivers and give rise to a huge number of Cy, Mv, Dp, and Pt accumulated in young swine tissues in
metabolites which are subject to mutual transformation by a dose-dependent manner during a three-week bilberry diet
enzymes and microbiota (Figure 4). The metabolic pathway (Chen et al. 2015). Following a fifteen-day blackberry diet,
of anthocyanins is, therefore, more complex than other phe- anthocyanins appeared in various organs of rats (jejunum,
nolic acids. Some complex interactions between microbiota stomach, kidney, liver, brain, and even bone tissues), and
and phenolic metabolites may be clarified by the application their methylated and glucuronidated forms were partic-
of “big data” analyses and developing “-omics” approaches. ularly concentrated in the jejunum (Talavéra et al. 2005;
However, it is almost certain that conversion pathways have Janle et al. 2010). It is noteworthy that the anthocyanins
been oversimplified (Williamson, Kay, and Crozier 2018). were most concentrated in the kidney and bladder, then
Multiple factors influence the production of metabolites, in the liver, even cross the blood-brain barrier (some in
from the profile of the gut microbiota to the chemical struc- glycoside forms), but were not found in the retroperito-
tures of the anthocyanins, and much research is still required neal fat (Kirakosyan et al. 2015). When equilibrium of
to clearly elucidate these pathways. anthocyanin levels in plasma and liver was achieved, con-
centrations in kidney were almost three times higher than
in plasma, indicating the reabsorption of anthocyanins by
the kidney (Vanzo et al. 2008). Cy was the most abundant
3.6. Distribution of anthocyanins and derivatives in
anthocyanidin, and C3G was more widely distributed than
tissues
other anthocyanins in nature (Bendokas et al. 2020). C3G
3.6.1. Distribution in tissues preferred peripheral spaces to systemic circulation and had
Although the bioavailability of anthocyanins is very low, a long residence time in rats. Its derivatives include Cy,
their accumulation in tissues is detectable. After a four-week Pn, Pn3G, and their glucuronide and methyl conjugates

Figure 4. Proposed pathways for the conversion of cyanidin-based anthocyanins to phenolic acids and related compounds. Compounds in red are major
components in 0–24 h urine. Based on published data (Borges et al., 2018; Czank et al. 2013; De Ferrars et al. 2014a; González-Barrio, Edwards, and Crozier
2011; Jaganath et al. 2006; Ludwig et al. 2015; Nurmi et al. 2009; Ottaviani et al. 2016; Pereira-Caro et al. 2014a, 2014b, 2015, 2016, 2017).
5960 H. GUI ET AL.

(Czank et al. 2013; Fornasaro et al. 2016). Mv contains After administrating 13C-labeled C3G, over half the
two methoxy groups and is the largest and most hydro- labeled compounds remained in the human body after 48 h
phobic aglycone of the six common anthocyanidins. After with the rest being excreted in urine (5.4%), breath (6.9%
drinking blueberry juice, 75% of the anthocyanins in vol- exhaled as CO2), and feces (32.1%) (Czank et al. 2013). The
unteer urine were glycosides of Mv while the rest were profile of C3G metabolites distributed in serum, urine, and
methyl and glucuronide conjugates of common aglycones feces is shown in Table S1 (Czank et al. 2013). Phenolic
(Kalt et al. 2017). Compared with other aglycones, Dp has products in blood were more likely to be conjugated via
the highest antioxidant activity due to its three hydroxyl glucuronidation or sulfation than their parent anthocyanins.
groups on the B-ring (Ali, Almagribi, and Al-Rashidi 2016). The glucuronide and methylated anthocyanins in urine
Glucuronidation of anthocyanins is common in vivo. Mv demonstrate their participation in phase II metabolism. In
and Dp glucuronides were 6–48 times more abundant than feces, the disappearance of some conjugates and the appear-
Cy, Pn, and Pt glucuronides in various regions of the brain ance of phenolic acids derived from breakdown of the A-
(Milbury and Kalt 2010). Dp3G was more prone to con- and B-rings suggest that some metabolites may be reabsorbed
version to glucuronic forms and accumulation in the brain via EHC or degraded by gut microbiota in the colon. For
(Milbury and Kalt 2010). instance, after oral administration of Pg3R, demethylated
The accumulation of anthocyanins and conjugates in vivo Pg3R (from the action of gut microflora) was much higher
follows time- and dose-dependent patterns. Generally, the than Pg3R-monoglucuronide and hydroxylated Pg3R in
order of anthocyanin concentrations in tissues is consist plasma (Xu et al. 2021).
with metabolic pathways (Figure 5a). At equilibrium, con- In a study by De Ferrars et al. (2014b), some acids
centrations of anthocyanin in plasma are proportional to appeared in urine rather than plasma while some were
those in kidney, liver, and brain. Moreover, anthocyanin detected in plasma but not urine, indicating they under-
distribution (especially glucuronides and methyl conjugates) went further metabolism in tissues or were eliminated via
varies with aglycone status, due to their different binding the biliary route or catabolized by colonic microbiota to
capabilities in vivo. be reabsorbed or excreted. Although some metabolites,
such as hippuric acid, 2-, 3- and 4-hydroxybenzoic acid,
and benzoic acid-4-glucuronide, may be derived from other
3.6.2. Concentrations in circulating blood endogenous sources, they are still regarded as urinary bio-
Generally, anthocyanins appear rapidly in the blood. markers in anthocyanin metabolism (Khanal, Howard, and
Maximum concentrations were reached within 0.5 h and Prior 2014; Liu et al. 2020). Parts of the metabolic pathway
ranged from 1.4 to 592 nmol/L over 0.5 to 4 h. The max- of certain anthocyanins may differ due to specific struc-
imum in urine (202.74 ± 85.06 µg) appeared after tural differences. Based on the current information, the in
3.72 ± 0.83 h (Kay, Mazza, and Holub 2005). Almost one vivo pathways of anthocyanin derivatization are summa-
third of anthocyanins in blood and urine were in their rized in Figure 4. Significantly, some anthocyanin deriva-
parent form and the rest were conjugated forms (Kay, tives have more potential as health treatments than their
Mazza, and Holub 2005), including aglycones, methylated, parent forms due to their superior absorption, size, and
sulfated, and glucuronidated forms (Czank et al. 2013; stability. Considering the powerful antioxidant properties
Wiczkowski, Szawara-Nowak, and Romaszko 2016; Kasote of anthocyanins and the advantages of their being safe,
et al. 2019). natural pigments, researchers improve anthocyanin

Figure 5. (a) The order of anthocyanin elimination and accumulation in vivo according to published data. (b) Schematic representation of anthocyanin inter-
conversion in vivo based on Fornasaro et al. (2016).
Critical Reviews in Food Science and Nutrition 5961

bioavailability with food ingredients (protein, peptides, optimizing culture and induction conditions, obtaining a
polysaccharides and lipids) via non-covalent bonding. final C3G titer of 350 mg/L. Anthocyanin biosynthesis using
engineered bacteria is a promising route, but has, so far,
been limited to production of derivatives by E. coli. (Zha
3.6.3. Anthocyanin interconversion in vivo et al. 2018). There is a long way to go to achieve
The addition and removal of -OH and -OCH3 substituents industrial-scale production of anthocyanin monomers. The
are common in vivo and can facilitate the interconversion second challenge is the various factors that may impact
of anthocyanins. The production of Pn from Cy is com- anthocyanin metabolism. Human studies offer limited sam-
mon in many studies but other anthocyanin interconver- ple types (blood, urine, and feces) and anthocyanin bio-
sions are rare. Fornasaro et al. (2016) gave rats 668 nmol availability was affected by food matrix, food processing,
C3G intravenously and observed the appearance of Pn3G metabolizing enzymes and gut microbiota (Eker et al. 2019).
and M3G after a rapid decrease of C3G in plasma. Dp3G Furthermore, methodological approaches and evaluation
appeared transiently after 2 min with small amounts of systems can be incompatible resulting high inter and
Pt3G at 5 min while Pg3G remained throughout the trial. intra-individual variation, and even contradictory results in
The AUC ranking was C3G > M3G > Pn3G > Pg3G in some studies. General pharmaceutical studies demonstrate
plasma, and C3G > Pt3G > Pn3G in brain. Despite these that greater numbers of comparable subjects are required
differing profiles, C3G concentration in plasma correlated in treatment groups to avoid these drawbacks. In vivo ani-
linearly with brain. In another study, metabolites of Pg, mal experiments allow greater control of anthocyanin metab-
which was not common in blueberry, were detected in olism, but although input dosages are lower than in human
urine after drinking blueberry juice (Kalt et al. 2017). In studies, it is still difficult to recover practical quantities of
a metabolic study of pelargonidin-based anthocyanins, Xu anthocyanin monomers. Furthermore, animal studies can
et al. (2021) detected higher levels of demethylated metab- be expensive, time-consuming, and laborious. Model systems
olites than their hydroxylated and monoglucuronide forms. utilizing cell lines such as MKN-28 cells and Caco-2 cells
These results imply the existence of anthocyanin inter- are now commonly applied in anthocyanin studies to replace
conversion in vivo, which could be achieved via enzymatic in vivo tissues. Japanese scientists have also established a
reaction. The corresponding products appear so rapidly repeatable way to create multi-cellular human liver organoids
that gut microflora with the capacity for dehydroxylation that can be used for in vitro drugs efficiency testing (Ouchi
and demethylation may not be responsible for this inter- et al. 2019). Notably, human embryonic stem cells have the
conversion. Except for the role of microbiota in upper potential to differentiate into every cell type found in the
gastrointestinal tract, the mammalian CYP enzymes dis- body. Researchers have utilized this unique potential to
cussed above are responsible for the hydroxylation and replace damaged or missing cells, aiding the healing of some
demethylation of flavonoids at positions 3′ and 4′ of the diseases (Klimanskaya, Kimbrel, and Lanza 2014).
B-ring and COMT is responsible for the methylation of Furthermore, 3 D printing has made cell culture possible in
anthocyanins. This may be sufficient to account for antho- complex three-dimensional biomimetic structures that
cyanin interconversion. A possible pathway of intercon- enhanced cell behavior and function (Zhu et al. 2016). This
version is illustrated in Figure 5b. technology has been successfully applied in the simulation
of human intestinal tissues (Madden et al. 2018) and bone
(Bandyopadhyay, Mitra, and Bose 2020). The combination
4. Challenges and opportunities for metabolic of cell culture and 3 D printing is a promising route for the
studies urgently needed construction of simulated human tissues
and organs. In vitro experiments using commercially avail-
Anthocyanin metabolism has been studied for a hundred able digestive enzymes and buffers are a suitable alternative
years, but several challenges still remain. Firstly, availability approach for the study of low concentration compounds
of anthocyanin monomer is limited. To date, more than 700 like the anthocyanins. For the development of in vitro exper-
anthocyanins have been identified in nature and the mix- iments that avoid low physiological relevance and incom-
tures commonly encountered in research studies severely parable results, a standardized digestion protocol is available
hinder the elucidation of their mechanisms. Anthocyanins based on available physiological data, covering parameters
are mainly extracted and isolated from plant tissues, which such as electrolytes, enzymes, bile, dilutions, pH, and diges-
is time consuming, labor intensive, and difficult on an tion time (Brodkorb et al. 2019). Researchers are continually
industrial scale. In metabolic studies, a large number of extending the capabilities of the techniques used to inves-
anthocyanin monomer is required, but only the common tigate the mechanisms and maximize the efficacy of these
anthocyanin C3G can satisfy this. Engineered bacteria could functional substances in the hope of matching the break-
be an alternative source of anthocyanins offering fast, throughs made in related fields.
high-yield, controllable, and stable production (Zha et al. The third challenge for metabolic studies of the antho-
2018). For example, Cress et al. (2017) optimized the pro- cyanins is that they are highly sensitive to physiological
duction of up to 56 mg/L peonidin 3-O-glucoside in E. coli conditions, including pH, chemicals, enzymes, and other
by increasing the availability of S-adenosyl-L-methionine. bioactive compounds. The different stages of digestion with
Lim et al. (2015) maximized production of (+)-catechin and their varied pH impact anthocyanin stability and degrada-
UDP-glucose (a key metabolic limiting factor) and tion. Numerous anthocyanin derivatives are widely dispersed
5962 H. GUI ET AL.

at low concentrations (ng–μg) and most are undetectable disciplines offers unique benefits to scientific progress. The
by current detection technologies. Radioisotope tracer tech- utilization of engineered bacteria, the combining of embry-
niques were first applied to anthocyanin metabolism by onic stem cell and 3 D printing technologies, and improve-
Czank et al. (2013). However, the complex reaction pathways ments to detection techniques will greatly expand our
in vivo makes the data difficult to interpret. Recently devel- knowledge of anthocyanin metabolism, and help to maxi-
oped comprehensive analysis called “omics” (genomics, tran- mize the use and health benefits of these, and similar,
scriptomics, proteomics, metabolomics, etc.) have been compounds.
applied to the analysis of complicate “big data”, making
many remarkable advances in bioinformatics, especially in
the analysis of gut microbiota. In summary, the study of Acknowledgments
anthocyanins (and other similar compounds) faces several
challenges: shortage of anthocyanin monomers, many exter- We are grateful to Dr. Jian Dai for her valuable comments and careful
nal and internal factors influencing their metabolism, and examination on this manuscript.
limitations of the detection technology. The current pace of
technological development suggests these problems can
be solved. Author’s contributions
Hailong Gui contributed to writing – original draft, Lijun Sun contributed
to writing – original draft, Ruihai Liu contributed to conceptualization,
5. Conclusions and perspectives Xu Si contributed to editing, Qiao Jiang contributed to editing and super-
vision, Yanyan Qiao contributed to grammar, Bin Li contributed to con-
The stability of anthocyanins varies with their structure ceptualization, funding acquisition, and writing – review and editing, and
(aglycone or glycoside), environmental pH (7.0 in the oral Jinlong Tian contributed to writing – review and editing.
cavity, 2–3 in the stomach, 7–8 in the small intestine), and
catalytic impact of the microbiota. Major time-dependent
breakdown of the anthocyanins by microbiota occurs in the Disclosure statement
oral cavity. Unexpectedly, their glucuronide forms, which
The authors declare no conflicts of interest.
involve the enzymes UGTs, arylsulfatase and β-glucuronidase
are also detectable in the oral cavity. Prior to absorption in
vivo, some anthocyanins are deglycosylated by LPH and Funding
β-glucosidase enzymes while others are absorbed directly
via specific transporters (bilitranslocase, SGLT1, GLUT1, −2 This work was supported by the National Natural Science Foundation
and −3, OCT1, OAT2, and SMCT1 and −2), depending on of China (U21A20273), China Agriculture Research System of MOF
and MARA (CARS-29), and the First Batch of Liaoning “Unveiling
their glycoside moiety. Anthocyanin aglycones with more
Leader” Scientific and Technological Projects (2021JH1/10400036).
–OH rather than –CH3 groups have lower hydrophobicity
and stability, resulting in low absorption and high degra-
dation. After passing through the acidic stomach and ORCID
alkalescent small intestine, unabsorbed anthocyanins are
metabolized by gut microbiota into small phenolic acids Dongnan Li http://orcid.org/0000-0001-8499-5022
Bin Li http://orcid.org/0000-0002-7393-2111
and absorbed in the colon. After absorption, anthocyanins
and their derivatives are conjugated by cellular enzymes
(UGTs, SULTs, COMT, and CYP1A), pumped in and out
by transporters (ABC, BCRP, OATs, OATPs, PEPTs, ABCC2, Abbreviations
MRPs, and P-gp), and begin EHC after being transported ABC ATP-binding cassette
the liver via the bloodstream. During EHC the compounds BCRP Breast cancer resistance protein
are further conjugated and maintained low concentrations C3G Cyanidin-3-O-glucoside
over extended periods. As a result of this circulation, antho- C3Gal Cyanidin-3-O- galactoside
C3R Cyanidin-3-O- rutinoside
cyanin derivatives are detectable and unevenly distributed
COMT Catechol-O-methyltransferase
in nondigestive tissues (eyes, liver, cortex, cerebellum, kid- Cy Cyanidin
ney, and even brain). Interconversion of anthocyanins occurs Cy3R Cyanidin-3-O-rutinoside
and is modulated by enzymes (COMT and CYP) whose CYP1A Cytochrome P450 1A
mechanisms need further exploration. Anthocyanin metab- Dp Delphinidin
olism is dependent on their chemical structure and is Dp3G Delphinidin-3-O-glucoside
reflected in their derivatives, absorption, distribution, metab- Dp3R Delphinidin-3-O-rutinoside
EHC Enterohepatic circulation
olism, and excretion. Based on recent findings concerning
GLUT 1, 3 Facilitative glucose transporters 1 and 3
competition during transportation, conjugation sequencing, LPH Lactase phlorizin hydrolase
variations in microflora and the effects of other food com- MRPs Multidrug resistance-associated proteins
ponents, factors including anthocyanin dosage, intake pat- Mv Malvidin
tern, duration, digestion environment, and food matrices Mv3G Malvidin-3-O-glucoside
should also be considered. Cooperation between research OAT2 Organic anion transporter 2
Critical Reviews in Food Science and Nutrition 5963

OATPs Organic anion transporting polypeptides anthocyanin monomers by rat gut microflora. Food Chemistry
OATs Organic anion transporters 237:887–94. doi: 10.1016/j.foodchem.2017.06.054.
OCT 1 Organic cation transporter 1 Chen, Z., S. Zheng, L. Li, and H. Jiang. 2014. Metabolism of flavonoids
PCA Protocatechuic acid in human: A comprehensive review. Current Drug Metabolism 15
PEPTs Peptide transporters (1):48–61. doi: 10.2174/138920021501140218125020.
Pg Pelargonidin Cortés-Martín, A., M. V. Selma, F. A. Tomás-Barberán, A.
Pg3G Pelargonidin-3-O-glucoside González-Sarrías, and J. C. Espín. 2020. Where to look into the
Pg3R Pelargonidin-3-O-rutinoside puzzle of polyphenols and health? The postbiotics and gut micro-
PGA Phloroglucinol aldehyde biota associated with human metabotypes. Molecular Nutrition &
Food Research 64 (9):e1900952. doi: 10.1002/mnfr.201900952.
P-gp P-glycoprotein
Cress, B. F., Q. D. Leitz, D. C. Kim, T. D. Amore, J. Y. Suzuki, R. J.
Pn Peonidin
Linhardt, and M. A. Koffas. 2017. CRISPRi-mediated metabolic
Pn3G Peonidin-3-O-glucoside
engineering of E. coli for O-methylated anthocyanin production.
Pt Petunidin Microbial Cell Factories 16 (1):10. doi: 10.1186/s12934-016-0623-3.
Pt3G Petunidin -3-O-glucoside Cury, B. S. F., A. B. Meneguin, F. G. Prezotti, F. I. Boni, and V. M.
SGLT 1 Sodium-dependent glucose co-transporter 1 de Oliveira Cardoso. 2021. Nanocarriers for oral drug delivery. In
SMCT 1, 2 Sodium/monocarboxylate transporters 1 and 2 Nanocarriers for drug delivery, nanomedicine and nanotoxicology.
SULTs Sulfotransferases concepts and applications, ed. J. O. Eloy, J. P. Abriata, and J. M.
UDP-Glu-DH UDP-glucose-dehydrogenase Marchetti, 127–51. Switzerland, Cham: Springer.
UGTs UDP-glucuronosyl transferases 500 Czank, C., A. Cassidy, Q. Zhang, D. J. Morrison, T. Preston, P. A.
Kroon, N. P. Botting, and C. D. Kay. 2013. Human metabolism and
elimination of the anthocyanin, cyanidin-3-glucoside: A (13)C-tracer
study. The American Journal of Clinical Nutrition 97 (5):995–1003.
References doi: 10.3945/ajcn.112.049247.
Ali, H. M., W. Almagribi, and M. N. Al-Rashidi. 2016. Antiradical De Ferrars, R., A. Cassidy, P. Curtis, and C. Kay. 2014a. Phenolic
and reductant activities of anthocyanidins and anthocyanins, metabolites of anthocyanins following a dietary intervention study
structure-activity relationship and synthesis. Food Chemistry in post-menopausal women. Molecular Nutrition & Food Research
194:1275–82. doi: 10.1016/j.foodchem.2015.09.003. 58 (3):490–502. doi: 10.1002/mnfr.201300322.
Banaszewski, K., E. Park, I. Edirisinghe, J. C. Cappozzo, and B. M. De Ferrars, R., C. Czank, Q. Zhang, N. P. Botting, P. Kroon, A. Cassidy,
Burton-Freeman. 2013. A pilot study to investigate bioavailability and C. Kay. 2014b. The pharmacokinetics of anthocyanins and their
of strawberry anthocyanins and characterize postprandial plasma metabolites in humans. British Journal of Pharmacology 171
polyphenols absorption patterns by q-tof lc/ms in humans. Journal (13):3268–82. doi: 10.1111/bph.12676.
of Berry Research 3 (2):113–26. doi: 10.3233/JBR-130048. De Rosas, M. I., L. Deis, L. Martínez, M. Durán, E. Malovini, and J.
Bandyopadhyay, A., I. Mitra, and S. Bose. 2020. 3D printing for bone B. Cavagnaro. 2019. Anthocyanins in nutrition: Biochemistry and
regeneration. Current Osteoporosis Reports 18 (5):505–14. doi: health benefits. In Psychiatry and neuroscience update: from trans-
10.1007/s11914-020-00606-2. lational research to a humanistic approach-volume III, ed. P. Á.
Bendokas, V., V. Stanys, I. Mažeikienė, S. Trumbeckaite, R. Baniene, Gargiulo and H. L. Mesones Arroyo, 143–52. Cham: Springer
and J. Liobikas. 2020. Anthocyanins: From the field to the antiox- International Publishing.
idants in the body. Antioxidants 9 (9):819. doi: 10.3390/an- DeGorter, M. K., C. Xia, J. J. Yang, and R. B. Kim. 2012. Drug trans-
tiox9090819. porters in drug efficacy and toxicity. Annual Review of Pharmacology
Borges, G., J. I. Ottaviani, J. J. Van Der Hooft, H. Schroeter, and A. and To x i c o l o g y 52:249–73. doi: 10.1146/
Crozier. 2018. Absorption, metabolism, distribution and excretion annurev-pharmtox-010611-134529.
of (−)-epicatechin: A review of recent findings. Molecular Aspects Dharmawansa, K., D. W. Hoskin, and H. Rupasinghe. 2020.
of Medicine 61:18–30. doi:10.1016/j.mam.2017.11.002. Chemopreventive effect of dietary anthocyanins against gastrointes-
Braga, A., D. Murador, L. de Souza, and V. V. de Rosso. 2018. tinal cancers: A review of recent advances and perspectives.
Bioavailability of anthocyanins: Gaps in knowledge, challenges and International Journal of Molecular Sciences 21 (18):6555. doi: 10.3390/
future research. Journal of Food Composition and Analysis 68:31–40. ijms21186555.
doi: 10.1016/j.jfca.2017.07.031. Dreiseitel, A., B. Oosterhuis, K. V. Vukman, P. Schreier, A. Oehme, S.
Brodkorb, A., L. Egger, M. Alminger, P. Alvito, R. Assunção, S. Locher, G. Hajak, and P. G. Sand. 2009. Berry anthocyanins and
Ballance, T. Bohn, C. Bourlieu-Lacanal, R. Boutrou, F. Carrière, anthocyanidins exhibit distinct affinities for the efflux transporters
et al. 2019. INFOGEST static in vitro simulation of gastrointestinal BCRP and MDR1. British Journal of Pharmacology 158 (8):1942–50.
food digestion. Nature Protocols 14 (4):991–1014. doi: 10.1038/ doi: 10.1111/j.1476-5381.2009.00495.x.
s41596-018-0119-1. Edwards, C. A., M. Serafini, A. Crozier, and W. Mullen. 2008.
Cahyana, Y., and M. Gordon. 2013. Interaction of anthocyanins with Bioavailability of pelargonidin-3-O-glucoside and its metabolites in
human serum albumin: Influence of pH and chemical structure on humans following the ingestion of strawberries with and without
binding. Food Chemistry 141 (3):2278–85. doi: 10.1016/j.food- cream. Journal of Agricultural and Food Chemistry 56 (3):713–9.
chem.2013.05.026. doi: 10.1021/jf072000p.
Chen, Y., H. Chen, W. Zhang, Y. Ding, T. Zhao, M. Zhang, G. Mao, Eker, M. E., K. Aaby, I. Budic-Leto, S. R. Brnčić, S. N. El, S. Karakaya,
W. Feng, X. Wu, and L. Yang. 2019. Bioaccessibility and biotrans- S. Simsek, C. Manach, W. Wiczkowski, and S. Pascual-Teresa. 2019.
formation of anthocyanin monomers following in vitro simulated A Review of factors affecting anthocyanin bioavailability: Possible
gastric-intestinal digestion and in vivo metabolism in rats . Food & implications for the inter-individual variability. Foods 9 (1):2. doi:
Function 10 (9):6052–61. doi: 10.1039/C9FO00871C. 10.3390/foods9010002.
Chen, T.-Y., J. Kritchevsky, K. Hargett, K. Feller, R. Klobusnik, B. J. El Mohsen, M. A., J. Marks, G. Kuhnle, K. Moore, E. Debnam, S.
Song, B. Cooper, Z. Jouni, M. G. Ferruzzi, and E. M. Janle. 2015. Kaila Srai, C. Rice-Evans, and J. P. E. Spencer. 2006. Absorption,
Plasma bioavailability and regional brain distribution of polyphenols tissue distribution and excretion of pelargonidin and its metabolites
from apple/grape seed and bilberry extracts in a young swine mod- following oral administration to rats. The British Journal of Nutrition
el. Molecular Nutrition & Food Research 59 (12):2432–47. doi: 95 (1):51–8. doi: 10.1079/BJN20051596.
10.1002/mnfr.201500224. Faria, A., D. Pestana, J. Azevedo, F. Martel, V. de Freitas, I. Azevedo,
Chen, Y., Q. Li, T. Zhao, Z. Zhang, G. Mao, W. Feng, X. Wu, and L. N. Mateus, and C. Calhau. 2009. Absorption of anthocyanins
Yang. 2017. Biotransformation and metabolism of three mulberry through intestinal epithelial cells - Putative involvement of GLUT2
5964 H. GUI ET AL.

. Molecular Nutrition & Food Research 53 (11):1430–7. doi: 10.1002/ Kataoka, K. 2016. The intestinal microbiota and its role in human
mnfr.200900007. health and disease. The Journal of Medical Investigation: JMI 63
Felgines, C., O. Texier, P. Garcin, C. Besson, J. L. Lamaison, and A. (1–2):27–37. doi: 10.2152/jmi.63.27.
Scalbert. 2009. Tissue distribution of anthocyanins in rats fed a Kay, C., P. Kroon, and A. Cassidy. 2009. The bioactivity of dietary
blackberry anthocyanin-enriched diet. Molecular Nutrition & Food anthocyanins is likely to be mediated by their degradation products.
Research 53 (9):1098–103. doi: 10.1002/mnfr.200800323. Molecular Nutrition & Food Research 53 (S1):S92–S101. doi: 10.1002/
Fernandes, I., A. Faria, V. de Freitas, C. Calhau, and N. Mateus. 2015. mnfr.200800461.
Multiple-approach studies to assess anthocyanin bioavailability. Kay, C., G. Mazza, and B. Holub. 2005. Anthocyanins exist in the
Phytochemistry Reviews 14 (6):899–919. doi: 10.1201/9781351069700. circulation primarily as metabolites in adult men. The Journal of
Fernandes, I., F. Nave, R. Gonçalves, V. Freitas, and N. Mateus. 2012. Nutrition 135 (11):2582–8. doi: 10.1093/jn/135.11.2582.
On the bioavailability of flavanols and anthocyanins: Khanal, R., L. R. Howard, and R. L. Prior. 2014. Urinary excretion of
Flavanol-anthocyanin dimers. Food Chemistry 135 (2):812–8. doi: phenolic acids in rats fed cranberry, blueberry, or black raspberry
10.1016/j.foodchem.2012.05.037. powder. Journal of Agricultural and Food Chemistry 62 (18):3987–96.
Fornasaro, S., L. Ziberna, M. Gasperotti, F. Tramer, U. Vrhovsek, F. doi: 10.1021/jf403883r.
Mattivi, and S. Passamonti. 2016. Determination of cyanidin Kirakosyan, A., E. M. Seymour, J. Wolforth, R. McNish, P. Kaufman,
3-glucoside in rat brain, liver and kidneys by UPLC/MS-MS and and S. Bolling. 2015. Tissue bioavailability of anthocyanins from
its application to a short-term pharmacokinetic study. Scientific whole tart cherry in healthy rats. Food Chemistry 171:26–31. doi:
Reports 6 (1):22815. doi: 10.1038/srep22815. 10.1016/j.foodchem.2014.08.114.
González-Barrio, R., C. Edwards, and A. Crozier. 2011. Colonic ca- Klimanskaya, I., E. A. Kimbrel, and R. Lanza. 2014. Embryonic stem
tabolism of ellagitannins, ellagic acid, and raspberry anthocyanins: cells. In Principles of tissue engineering, ed. R. Lanza, R. Langer,
In vivo and in vitro studies. Drug Metabolism and Disposition 39 and J. Vacanti 4th ed., 565–79. Salt Lake: Academic Press.
(9):1680–8. doi: 10.1124/dmd.111.039651. Li, B., Z. Cheng, X. Sun, X. Si, E. Gong, Y. Wang, J. Tian, C. Shu, F.
Han, H., C. Liu, W. Gao, Z. Li, G. Qin, S. Qi, H. Jiang, X. Li, M. Liu, Ma, D. Li, et al. 2020. Lonicera caerulea L. polyphenols alleviate
F. Yan, et al. 2021. Anthocyanins are converted into anthocyanidins oxidative stress-induced intestinal environment imbalance and
and phenolic acids and effectively absorbed in the jejunum and lipopolysaccharide-induced liver injury in HFD-fed rats by regulat-
ileum. J Agric Food Chem 69 (3):992–1002. doi: 10.1021/acs. ing the Nrf2/HO-1/NQO1 and MAPK pathways. Molecular Nutrition
jafc.0c07771. & Food Research 64 (10):1901315. doi: 10.1002/mnfr.201901315.
Han, F., H. Oliveira, N. F. Brás, I. Fernandes, L. Cruz, V. De Freitas, Lila, M., B. Burton-Freeman, M. Grace, and W. Kalt. 2016. Unraveling
and N. Mateus. 2020. In vitro gastrointestinal absorption of red anthocyanin bioavailability for human health. Annual Review of
wine anthocyanins - Impact of structural complexity and phase II Food Science and Technology 7:375–93. doi: 10.1146/
metabolization. Food Chemistry 317:126398. doi: 10.1016/j.food- annurev-food-041715-033346.
chem.2020.126398. Lim, C. G., L. Wong, N. Bhan, H. Dvora, P. Xu, S. Venkiteswaran,
Huang, C., Y. Chen, T. Zhou, and G. Chen. 2009. Sulfation of dietary and M. A. Koffas. 2015. Development of a recombinant escherich-
flavonoids by human sulfotransferases. Xenobiotica; the Fate of ia coli strain for overproduction of the plant pigment anthocyanin.
Foreign Compounds in Biological Systems 39 (4):312–22. doi: Applied and Environmental Microbiology 81 (18):6276–84. doi:
10.1080/00498250802714915. 10.1128/AEM.01448-15.
Jaganath, I., W. Mullen, C. Edwards, and A. Crozier. 2006. The relative Lira-Junior, R., and E. A. Boström. 2018. Oral-gut connection: One
contribution of the small and large intestine to the absorption and step closer to an integrated view of the gastrointestinal tract?
metabolism of rutin in man. Free Radical Research 40 (10):1035–46. Mucosal Immunology 11 (2):316–8. doi: 10.1038/mi.2017.116.
doi: 10.1080/10715760600771400. Liu, H., T. J. Garrett, Z. Su, C. Khoo, S. Zhao, and L. Gu. 2020.
Janle, E. M., M. A. Lila, M. Grannan, L. Wood, A. Higgins, G. G. Modifications of the urinary metabolome in young women after
Yousef, R. B. Rogers, H. Kim, G. S. Jackson, L. Ho, et al. 2010. cranberry juice consumption were revealed using the
Pharmacokinetics and tissue distribution of 14C-labeled grape poly- UHPLC-Q-orbitrap-HRMS-based metabolomics approach. Food &
phenols in the periphery and the central nervous system following Function 11 (3):2466–76. doi: 10.1039/c9fo02266j.
oral administration. Journal of Medicinal Food 13 (4):926–33. doi: Liu, Y., D. Zhang, Y. Wu, D. Wang, Y. Wei, J. Wu, and B. Ji. 2014.
10.1089/jmf.2009.0157. Stability and absorption of anthocyanins from blueberries subjected
Kalt, W. 2019. Anthocyanins and their C6-C3-C6 metabolites in hu- to a simulated digestion process. International Journal of Food Sciences
mans and animals. Molecules 24 (22):4024. doi: 10.3390/mole- and Nutrition 65 (4):440–8. doi: 10.3109/09637486.2013.869798.
cules24224024. Ludwig, I. A., P. Mena, L. Calani, G. Borges, G. Pereira-Caro, L.
Kalt, W., J. B. Blumberg, J. E. McDonald, M. R. Vinqvist-Tymchuk, S. Bresciani, D. Del Rio, M. E. Lean, and A. Crozier. 2015. New in-
A. E. Fillmore, B. A. Graf, J. M. O’Leary, and P. E. Milbury. 2008. sights into the bioavailability of red raspberry anthocyanins and
Identification of anthocyanins in the liver, eye, and brain of ellagitannins. Free Radical Biology & Medicine 89:758–69. doi:
blueberry-fed pigs. Journal of Agricultural and Food Chemistry 56 10.1016/j.freeradbiomed.2015.10.400.
(3):705–12. doi: 10.1021/jf071998l. Madden, L. R., T. V. Nguyen, S. Garcia-Mojica, V. Shah, A. V. Le, A.
Kalt, W., J. McDonald, Y. Liu, and S. A. E. Fillmore. 2017. Flavonoid Peier, R. Visconti, E. M. Parker, S. C. Presnell, D. G. Nguyen, et al.
metabolites in human urine during blueberry anthocyanin intake. 2018. Bioprinted 3D primary human intestinal tissues model aspects
Journal of Agricultural and Food Chemistry 65 (8):1582–91. doi: of native physiology and ADME/tox functions. iScience 2:156–67.
10.1021/acs.jafc.6b05455. doi: 10.1016/j.isci.2018.03.015.
Kamonpatana, K., M. Failla, P. Kumar, and M. Giusti. 2014. Anthocyanin Mallery, S. R., D. E. Budendorf, M. P. Larsen, P. Pei, M. Tong, A. S.
structure determines susceptibility to microbial degradation and Holpuch, P. E. Larsen, G. D. Stoner, H. W. Fields, K. K. Chan, et al.
bioavailability to the buccal mucosa. Journal of Agricultural and 2011. Effects of human oral mucosal tissue, saliva, and oral micro-
Food Chemistry 62 (29):6903–10. doi: 10.1021/jf405180k. flora on intraoral metabolism and bioactivation of black raspberry
Kamonpatana, K., M. Giusti, C. Chitchumroonchokchai, M. anthocyanins. Cancer Prevention Research (Philadelphia, Pa.) 4
MorenoCruz, K. Riedl, P. Kumar, and M. Failla. 2012. Susceptibility (8):1209–21. doi: 10.1158/1940-6207.CAPR-11-0040.
of anthocyanins to ex vivo degradation in human saliva. Food Manach, C., G. Williamson, C. Morand, A. Scalbert, and C. Rémésy.
Chemistry 135 (2):738–47. doi: 10.1016/j.foodchem.2012.04.110. 2005. Bioavailability and bioefficacy of polyphenols in human. I.
Kasote, D. M., G. J. Duncan, M. Neacsu, and W. R. Russell. 2019. Review of 97 bioavailability studies. The American Journal of Clinical
Rapid method for quantification of anthocyanidins and anthocyanins Nutrition 81 (1):230S–42S. doi: 10.1093/ajcn/81.1.230S.
in human biological samples. Food Chemistry 290:56–63. doi: Marques Peixoto, F., I. Fernandes, A. C. M. S. Gouvêa, M. C. P. A.
10.1016/j.foodchem.2019.03.109. Santiago, R. Galhardo Borguini, N. Mateus, V. Freitas, R. L. O.
Critical Reviews in Food Science and Nutrition 5965

Godoy, and I. MPLVO. Ferreira. 2016. Simulation of in vitro diges- Absorption, metabolism, and colonic catabolism. Journal of
tion coupled to gastric and intestinal transport models to estimate Agricultural and Food Chemistry 65 (26):5365–74. doi: 10.1021/acs.
absorption of anthocyanins from peel powder of jabuticaba, jamelao jafc.7b01707.
and jamb fruits. Journal of Functional Foods 24:373–81. doi: Pereira-Caro, G., G. Borges, I. Ky, A. Ribas, L. Calani, D. Del Rio, M.
10.1016/j.jff.2016.04.021. N. Clifford, S. A. Roberts, and A. Crozier. 2014b. In vitro colonic
Mazza, G., and E. Miniati. 1993. Anthocyanins in Fruits. Vegetables catabolism of orange juice (poly)phenols. Molecular Nutrition &
and grains. London: CRC Press. Food Research 59 (3):465–75. doi: 10.1002/mnfr.201400779.
Milbury, P., and W. Kalt. 2010. Xenobiotic metabolism and berry flavo- Pereira-Caro, G., G. Borges, J. van der Hooft, M. N. Clifford, D. Del
noid transport across the blood-brain barrier. Journal of Agricultural Rio, M. E. J. Lean, S. A. Roberts, M. B. Kellerhals, and A. Crozier.
and Food Chemistry 58 (7):3950–6. doi: 10.1021/jf903529m. 2014a. Orange juice (poly)phenols are highly bioavailable in humans.
Mueller, D., K. Jung, M. Winter, D. Rogoll, R. Melcher, and E. Richling. The American Journal of Clinical Nutrition 100 (5):1378–84. doi:
2017. Human intervention study to investigate the intestinal acces- 10.3945/ajcn.114.090282.
sibility and bioavailability of anthocyanins from bilberries. Food Pereira-Caro, G., I. A. Ludwig, T. Polyviou, D. Malkova, A. García, J.
Chemistry 231:275–86. doi: 10.1016/j.foodchem.2017.03.130. M. Moreno-Rojas, and A. Crozier. 2016. Identification of plasma
Nagar, E., Z. Okun, and A. Shpigelman. 2020. Digestive fate of poly- and urinary metabolites and catabolites derived from orange juice
phenols: Updated view of the influence of chemical structure and (p oly)phenols: Analysis by high-p erformance liquid
the presence of cell wall material. Current Opinion in Food Science chromatography-high-resolution mass spectrometry. Journal of
31:38–46. doi: 10.1016/j.cofs.2019.10.009. Agricultural and Food Chemistry 64 (28):5724–35. doi: 10.1021/acs.
Netzel, G., O. Wright, and Y. Sultanbawa. 2020. Understanding the jafc.6b02088.
metabolic fate and bioactivity of dietary anthocyanins. Proceedings Pereira-Caro, G., C. M. Oliver, R. Weerakkody, T. Singh, M. Conlon,
36:64. doi: 10.3390/proceedings2019036064. G. Borges, L. Sanguansri, T. Lockett, S. A. Roberts, A. Crozier, et al.
Novotny, J. A., B. A. Clevidence, and A. C. Kurilich. 2012. Anthocyanin 2015. Chronic administration of a microencapsulated probiotic en-
kinetics are dependent on anthocyanin structure. The British Journal hances the bioavailability of orange juice flavanones in humans.
of Nutrition 107 (4):504–9. doi: 10.1017/S000711451100314X. Free Radical Biology & Medicine 84:206–14. doi: 10.1016/j.freerad-
Nurmi, T., J. Mursu, M. Heinonen, A. Nurmi, R. Hiltunen, and S. biomed.2015.03.010.
Voutilainen. 2009. Metabolism of berry anthocyanins to phenolic Rodriguez-Mateos, A., D. Vauzour, C. G. Krueger, D. Shanmuganayagam,
acids in humans. Journal of Agricultural and Food Chemistry 57 J. Reed, L. Calani, P. Mena, D. Del Rio, and A. Crozier. 2014.
(6):2274–81. doi: 10.1021/jf8035116. Bioavailability, bioactivity and impact on health of dietary flavonoids
Oliveira, H., N. Basílio, F. Pina, I. Fernandes, V. de Freitas, and N. and related compounds: An update. Archives of Toxicology 88
Mateus. 2019a. Purple-fleshed sweet potato acylated anthocyanins: (10):1803–53. doi: 10.1007/s00204-014-1330-7.
Equilibrium network and photophysical properties. Food Chemistry Rowland, A., J. Miners, and P. Mackenzie. 2013. The
288:386–94. doi: 10.1016/j.foodchem.2019.02.132. UDP-glucuronosyltransferases: Their role in drug metabolism and
Oliveira, H., I. Fernandes, N. Bras, A. Faria, V. Freitas, C. Calhau, and detoxification. The International Journal of Biochemistry & Cell
N. Mateus. 2015. Experimental and theoretical data on the mech- Biology 45 (6):1121–32. doi: 10.1016/j.biocel.2013.02.019.
anism by which red wine anthocyanins are transported through a Rusishvili, M., L. Grisanti, S. Laporte, M. Micciarelli, M. Rosa, R. J.
human MKN-28 gastric cell model. Journal of Agricultural and Food Robbins, T. Collins, A. Magistrato, and S. Baroni. 2019. Unraveling
Chemistry 63 (35):7685–92. doi: 10.1021/acs.jafc.5b00412. the molecular mechanisms of color expression in anthocyanins.
Oliveira, H., R. Perez-Gregório, V. de Freitas, N. Mateus, and I. Physical Chemistry Chemical Physics: PCCP 21 (17):8757–66. doi:
Fernandes. 2019b. Comparison of the in vitro gastrointestinal bio- 10.1039/C9CP00747D.
availability of acylated and non-acylated anthocyanins: Purple-fleshed Selma, M. V., F. A. 2020. Tomás-Barberán, M. Romo-Vaquero, A.
sweet potato vs red wine. Food Chemistry 276:410–8. doi: 10.1016/j. Cortés-Martín, and J. C. Espín. Understanding polyphenols’ health
foodchem.2018.09.159. effects through the gut microbiota. In Dietary polyphenols: Their
Ottaviani, J., G. Borges, T. Momma, J. Spencer, C. Keen, A. Crozier, metabolism and health effects, ed. F. A. Tomás-Barberán, A.
and H. Schroeter. 2016. The metabolome of [2-(14)C](-)-epicatechin González-Sarrías, & R. García-Villalba,. 497–531. Brisbane: Wiley.
in humans: Implications for the assessment of efficacy, safety, and Si, X., J. Bi, Q. Chen, H. Cui, Y. Bao, J. Tian, C. Shu, Y. Wang, H.
mechanisms of action of polyphenolic bioactives. Scientific Reports Tan, W. Zhang, et al. 2021. Effect of blueberry anthocyanin-rich
6:29034. doi: 10.1038/srep29034. extracts on peripheral and hippocampal antioxidant defensiveness:
Ouchi, R., S. Togo, M. Kimura, T. Shinozawa, M. Koido, H. Koike, The analysis of the serum fatty acid species and gut microbiota
W. Thompson, R. A. Karns, C. N. Mayhew, P. S. McGrath, et al. profile. Journal of Agricultural and Food Chemistry 69 (12):3658–66.
2019. Modeling steatohepatitis in humans with pluripotent stem doi: 10.1021/acs.jafc.0c07637.
cell-derived organoids. Cell Metab 30 (2):374–84.e6. doi: 10.1016/j. Soares, S., S. Soares, E. Brandão, C. Guerreiro, N. Mateus, and V. de
cmet.2019.05.007. Freitas. 2020. Oral interactions between a green tea flavanol extract
Ozdal, T., D. Sela, J. Xiao, D. Boyacioglu, F. Chen, and E. Capanoglu. and red wine anthocyanin extract using a new cell-based model:
2016. The Reciprocal Interactions between polyphenols and gut Insights on the effect of different oral epithelia. Scientific Reports
microbiota and effects on bioaccessibility. Nutrients 8 (2):1–36.78. 10 (1):12638. doi: 10.1038/s41598-020-69531-9.
doi: 10.3390/nu80200:. Straßmann, S., M. Passon, and A. Schieber. 2021. Chemical hemisyn-
Park, S. Y., B. O. Hwang, M. Lim, S. H. Ok, S. K. Lee, K. S. Chun, thesis of sulfated cyanidin-3-o-glucoside and cyanidin metabolites.
K. K. Park, Y. Hu, W. Y. Chung, and N. Y. Song. 2021. Oral-gut Molecules (Basel, Switzerland) 26 (8):2146. doi: 10.3390/mole-
microbiome axis in gastrointestinal disease and cancer. Cancers 13 cules26082146.
(9):2124. doi: 10.3390/cancers13092124. Talavéra, S., C. Felgines, O. Texier, C. Besson, A. Gil-Izquierdo, J.-L.
Passamonti, S., M. Terdoslavich, R. Franca, A. Vanzo, F. Tramer, E. Lamaison, and C. Rémésy. 2005. Anthocyanin metabolism in rats
Braidot, E. Petrussa, and A. Vianello. 2009. Bioavailability of flavo- and their distribution to digestive area, kidney, and brain. Journal
noids: A review of their membrane transport and the function of of Agricultural and Food Chemistry 53 (10):3902–8. doi: 10.1021/
bilitranslocase in animal and plant organisms. Current Drug jf050145v.
Metabolism 10 (4):369–94. doi: 10.2174/138920009788498950. Talavéra, S., C. Felgines, O. Texier, C. Besson, J. L. Lamaison, and C.
Passamonti, S., U. Vrhovsek, A. Vanzo, and F. Mattivi. 2003. The Rémésy. 2003. Anthocyanins are efficiently absorbed from the stom-
stomach as a site for anthocyanins absorption from food. 2003. ach in anesthetized rats. The Journal of nutrition 133 (12):4178–82.
FEBS Letters 544 (1–3):210–3. doi: 10.1016/S0014-5793(03)00504-0. doi: 10.1093/jn/133.12.4178.
Pereira Caro, G., J. Moreno-Rojas, N. Brindani, D. Del Rio, M. Lean, Talavéra, S., C. Felgines, O. Texier, C. Besson, C. Manach, J.-L.
Y. Hara, and A. Crozier. 2017. Bioavailability of black tea theaflavins: Lamaison, and C. Rémésy. 2004. Anthocyanins are efficiently ab-
5966 H. GUI ET AL.

sorbed from the small intestine in rats. The Journal of Nutrition Williamson, G., C. Kay, and A. Crozier. 2018. The bioavailability,
134 (9):2275–9. doi: 10.1093/jn/134.9.2275. transport, and bioactivity of dietary flavonoids: A review from a
Tian, J.-L., X.-J. Liao, Y.-H. Wang, X. Si, C. Shu, E.-S. Gong, X. Xie, historical perspective. Comprehensive Reviews in Food Science and
X.-L. Ran, and B. Li. 2019a. Identification of cyanidin-3-arabinoside Food Safety 17 (5):1054–110. doi: 10.1111/1541-4337.12351.
extracted from blueberry as a selective protein tyrosine phosphatase Xie, L., H. Su, C. Sun, X. Zheng, and W. Chen. 2018. Recent advanc-
1B inhibitor. Journal of Agricultural and Food Chemistry 67 es in understanding the anti-obesity activity of anthocyanins and
(49):13624–34. doi: 10.1021/acs.jafc.9b06155. their biosynthesis in microorganisms. Trends in Food Science &
Tian, L., Y. Tan, G. Chen, G. Wang, J. Sun, S. Ou, W. Chen, and W. Technology 72:13–24. doi: 10.1016/j.tifs.2017.12.002.
Bai. 2019b. Metabolism of anthocyanins and consequent effects on Xu, Y., Y. Li, J. Xie, L. Xie, J. Mo, and W. Chen. 2021. Bioavailability,
the gut microbiota. Critical Reviews in Food Science and Nutrition absorption, and metabolism of pelargonidin-based anthocyanins using
59 (6):982–91. doi: 10.1080/10408398.2018.1533517. sprague-dawley rats and Caco-2 cell monolayers. Journal of Agricultural
Ugalde, C. M., Z. Liu, C. Ren, K. K. Chan, K. A. Rodrigo, Y. Ling, and Food Chemistry 69 (28):7841–50. doi: 10.1021/acs.jafc.1c00257.
P. E. Larsen, G. E. Chacon, G. D. Stoner, R. J. Mumper, et al. 2009. Zang, Z., S. Chou, J. Tian, Y. Lang, Y. Shen, X. Ran, N. Gao, and B.
Distribution of anthocyanins delivered from a bioadhesive black Li. 2021. Effect of whey protein isolate on the stability and antiox-
raspberry gel following topical intraoral application in normal idant capacity of blueberry anthocyanins: A mechanistic and in
healthy volunteers. Pharmaceutical Research 26 (4):977–86. doi: vitro simulation study. Food Chemistry 336:127700. doi: 10.1016/j.
10.1007/s11095-008-9806-x. foodchem.2020.127700.
Vanzo, A., M. Terdoslavich, A. Brandoni, A. Torres, U. Vrhovsek, and Zha, J., Y. Zang, M. Mattozzi, J. Plassmeier, M. Gupta, X. Wu, S.
S. Passamonti. 2008. Uptake of grape anthocyanins into the rat Clarkson, and M. Koffas. 2018. Metabolic engineering of coryne-
kidney and the involvement of bilitranslocase. Molecular Nutrition bacterium glutamicum for anthocyanin production. Microbial Cell
& Food Research 52 (10):1106–16. doi: 10.1002/mnfr.200700505. Factories 17 (1):143. doi: 10.1186/s12934-018-0990-z.
Verediano, T. A., H. Stampini Duarte Martino, M. C. Dias Paes, and Zhou, X., K. C. Cassidy, L. Hudson, M. A. Mohutsky, G. A. Sawada,
E. Tako. 2021. Effects of anthocyanin on intestinal health: A sys- and J. Hao. 2019. Enterohepatic circulation of glucuronide metab-
tematic review. Nutrients 13 (4):1331. doi: 10.3390/nu13041331. olites of drugs in dog. Pharmacology Research & Perspectives 7
Wiczkowski, W., D. Szawara-Nowak, and J. Romaszko. 2016. The im- (4):e00502. doi: 10.1002/prp2.502.
pact of red cabbage fermentation on bioavailability of anthocyanins Zhu, W., X. Ma, M. Gou, D. Mei, K. Zhang, and S. Chen. 2016. 3D
and antioxidant capacity of human plasma. Food Chemistry 190:730– printing of functional biomaterials for tissue engineering. Current
40. doi: 10.1016/j.foodchem.2015.06.021. Opinion in Biotechnology 40:103–12. doi: 10.1016/j.copbio.2016.03.014.
Williamson, G., and M. N. Clifford. 2010. Colonic metabolites of Zou, T., D. Feng, G. Song, H. Li, H. Tang, and W. Ling. 2014. The
berry polyphenols: The missing link to biological activity? The British role of sodium-dependent glucose transporter 1 and glucose trans-
Journal of Nutrition 104 (Suppl 3):S48–S66. doi: 10.1017/ porter 2 in the absorption of cyanidin-3-O-β-glucoside in Caco-2
S0007114510003946. cells. Nutrients 6 (10):4165–77. doi: 10.3390/nu6104165.

You might also like