Open access peer-reviewed chapter

Anthocyanins: Dietary Sources, Bioavailability, Human Metabolic Pathways, and Potential Anti-Neuroinflammatory Activity

Written By

Ruth Hornedo-Ortega, Zuriñe Rasines-Perea, Ana B. Cerezo, Pierre-Louis Teissedre and Michael Jourdes

Submitted: 04 August 2021 Reviewed: 15 August 2021 Published: 20 December 2021

DOI: 10.5772/intechopen.99927

Chapter metrics overview

561 Chapter Downloads

View Full Metrics

Abstract

The objectives of this chapter are to summarize and discuss (i) the anthocyanins structure and content in foodstuffs and their dietary intake (ii) the anthocyanins bioavailability and human metabolic pathways and (iii) the in vitro and in vivo potent anti-neuroinflammatory effects of anthocyanins and their metabolites. Indeed, anthocyanins are polyphenolic compounds belonging to the group of flavonoids, and are one of the most commonly consumed polyphenols in a normal diet. They are responsible of red, blue and purple color of several fruits and vegetables and their intake has been related with several human health benefits. The anthocyanins structures diversities as well as their content in various fruits, vegetables and cereals is addressed. Moreover, despite the growing evidence for the protective effects of anthocyanins, it is important to highlight that the in vivo bioavailability of these compounds is relatively low in comparison to their more stable metabolites. Indeed, after consumption, these bioactives are subjected to substantial transformations in human body. Phase I and II metabolites generated by intestinal and hepatic enzymatic reactions, and phenolic acids produced by gut microbiota and their metabolized forms, are the most important metabolic anthocyanins forms. For this reason, the study of the biological properties of these circulating metabolites represents a more in vivo realistic situation. Although the anthocyanin bioavailability researches in humans are limited, they will be discussed together with a global metabolic pathway for the main anthocyanins. Moreover, several works have demonstrated that anthocyanins can cross the blood brain barrier, and accumulate in brain endothelial cells, brain parenchymal tissue, striatum, hippocampus, cerebellum and cortex. Consequently, the study of anthocyanins as potent therapeutic agents in neurodegenerative diseases has gained relevance and the principal and the most recent studies are also discussed in the book chapter.

Keywords

  • anthocyanin
  • metabolites
  • neuroinflammation
  • phenolic acids
  • bioactives

1. Introduction

Anthocyanins (deriving from the Greek anthos means flower, and kyanos means blue) are one of the most important pigments in the plant kingdom after chlorophyll. Anthocyanins belong to the widespread family of flavonoid polyphenolic compounds and are responsible of red, purple and blue colors of a great numbers of vegetables and fruits [1]. Although several hundred of natural anthocyanins has been identified (more than 600), they all derived from 31 naturally known anthocyanidins (anthocyanins aglycone) [2, 3]. When looking at the human diet (fruits, vegetables and cereals), the number of anthocyanidins can be reduce to only six different anthocyanidins which are: pelargonidin, cyanidin, peonidin, delphinidin, petunidin, and malvidin. Among these, cyanidin represents the most widespread anthocyanidin in plants (50%). Cyanidin, delphinidin and pelargonidin are the non-methylated anthocyanidins whereas peonidin, malvidin and peonidin possess O-methylation. However, as free aglycones are considerable unstable, anthocyanins (the glycosylated forms) are more usually present in natural sources [1, 4].

Apart of being responsible for the color of many foods and beverages, anthocyanins also have numerous health benefits resulting of their antioxidant and anti-inflammatory activities, among others. Although the dietary intake of anthocyanins depends on the nutritional habits [5], they have received less attention than other flavonoids compounds. This may be due to the fact that anthocyanins are poorly absorbed, highly metabolized, and rapidly excreted in the urine [6]. In addition, their bioavailability and the metabolites formed by intestinal, hepatic enzymatic reactions, and gut microbiota depend on the chemical structure of anthocyanin.

This book chapter will summarize and discuss (i) the anthocyanins structure and content found in fruits, vegetables and cereals as well as the global dietary intake (ii) the anthocyanin bioavailability and human metabolic pathways and (iii) the in vitro and in vivo anti-neuroinflammatory effects of anthocyanins and their metabolites.

Advertisement

2. Anthocyanins: chemistry, intake and dietary sources

From a structural point of view, anthocyanins are glycosylated, polyhydroxy or polymethoxy derivatives of 2-phenylbenzopyrylium (or flavylium cation) containing two benzoyl rings (A and B) separated by a heterocyclic (C). The number of hydroxyl groups and their degree of methylation, the nature and number of the sugar and the position of the attachment, as well as the nature and number of aliphatic or aromatic acids attached to the sugars, determine their different structural variations [3, 7].

Regarding sugars, they can be attached at different positions: 3-monoglycosides, 3-diglycosides, or 3-triglycosides, 3,5-diglycosides and to a lesser extent 3,7-diglycosides. Glucose is the most common sugar moiety but other monosaccharides as rutinoside, rhamnose, galactose, arabinose, xylose are found. Furthermore, the disaccharides as sambubioside or sophoroside and as well as trisaccharides like as xylosilrutinoside or glucosylrutinoside can also be present [8, 9]. The linking of acyl substituents to sugars make possible a further degree of complexity of anthocyanins. Among them, aliphatic (acetic, malonic, succinic, malic) and cinnamic acids (p-coumaric, ferulic, sinapic) are the most predominant [10]. Both glycosylation and acylation affects the chemical and physical properties of anthocyanins. Thus, glycosylation improves water solubility whereas acylation have the contrary effect [11].

Anthocyanins are sensible to different factors such as temperature, light, oxygen or enzymes but pH represent one of the most important factors affecting them. Four different equilibrium species can co-exist, the flavylium cation (red; pH 1), the quinonoidal base (bleu, pH 4), the carbinol pseudobase (colorless or pale yellow, pH 5) and the chalcone (−cis and -trans) (colorless; pH 6). At pH values higher than 7, anthocyanis are degraded. Generally, anthocyanins are more stable and more soluble at low pH [12]. Among anthocyanidins, pelargonidin is the most stable compounds because of its B ring substituents and the presence of hydroxyl or methoxyl groups decrease the stability. However, the glycosylation confers to the molecules a higher stability at neutral pH, since the presence of sugar avoid the degradation into phenolic acid and aldehyde compounds [13].

The determination of the dietary intake of flavonoids, and among them, the mean consumption of anthocyanins has been the subject of several studies over the last two decades. In United States the daily consumption of these compounds in adults has been estimated in 12.5 mg/day, representing cyanidin anthocyanins the 44.7% of the total intake followed by delphinidin, malvidin, petunidin, peonidin and pelargonidin anthocyanidins [14]. Another study in adults (17900 individuals) showed a lower anthocyanidin intake, 9.20 ± 0.79 mg/day. In addition, they stated differences among anthocyanin consumption according to gender (women’s consume higher anthocyanins than men’s) and sociodemographic and lifestyle factors such as education, alcohol consumption and activity levels [15]. Concerning European data, the European Prospective Investigation into Cancer and Nutrition (EPIC) study estimated a mean of anthocyanin intake of 31 mg/day. At the same time, they also observed that these values vary according to the country, age, sex, body mass index (BMI), level of education, smoking status and physical activity level [5]. Between European countries, significant differences were reported. Indeed, Italy, France and Germany displays the greater mean values, from 35.1 to 42.3 mg/day, whereas Netherlands and Sweeden are the countries with a lower anthocyanin consumption (22.6 and 20.9 mg/day, respectively). More recently, after the study of the dietary habits of 30000 subjects in 14 European countries the mean intake of anthocyanins was estimated to be 19 mg/day [16]. In other continents and countries such as Australia (12153 subjects) or China (1393 subjects) the estimated anthocyanin mean intake was calculated at 24.2 mg/day and 27.6 mg/day, respectively, which are very closed values to European levels [17, 18].

Apart of being present in many colored fruits and vegetables they appeared also in beverages as red wine or juices and in processed foods as jams. Both, the type and the concentrations of anthocyanins are influenced by genetics (cultivar, species), cultivation, climate, soil, and processing [19]. However, one of the best sources of anthocyanins are berries. Among them, bilberries, blueberries and blackcurrants can be reach values greater than 1000 mg/100 g of fresh weight (FW) (Table 1). Among vegetables and cereals, red cabbage, cauliflower and colored corn and rice represent good sources of anthocyanins. The most common anthocyanins are cyanidin glucosides, but some fruits contain other predominant anthocyanin (Table 1). For example, pelargonidin-3-O-glucoside is the principal anthocyanin of strawberries, whereas malvidin-3-O-glucoside predominates in grapes and cyanidin-3,5-diglucoside is the major one in pomegranates [29, 30, 31]. Generally, the main anthocyanins in vegetables and cereals are chemically more complex in comparison with fruits. In fact, acylated and diglucosylated anthocyanins such as cyanidin-3-(p-coumaroyl)-diglucoside-5-glucoside can be found [33]. In addition, others less conventional sugars like sophoroside (cauliflower and radish) or laminiaribioside (red onions) can be present [35, 36, 39]. Interestingly, the most common anthocyanin type in sorghum, are the 3-deoxyanthocyanidins luteolinidin and apigeninidin (characterized by the lack of hydroxyl group at C3 position) and their derivatives, which are not commonly found in higher plants [44].

FruitContentMain anthocyaninRef
Apples (Red)0.1–315 mg/Kg peelCy-3-gal[20]
Apricot1.9–230.4 mg/100 g FWCy-3-rut[21]
Bilberry933–1017 mg/100 g FWDelp-3-gluc/ Delp-3-ara/ Delp-3-gal[22, 23]
Blackberry84–201 mg/100 g FWCy-3-gluc/ Cy-3-rut[24]
Blueberry232–438 mg/100 g FWMv-3-gluc/ Mv-3-gal/ Delp-3-gal/Delp-3-ara[22]
Cherry6.3–60 mg/100 g FWCy-3-gluc/ Cy-3-rut[25, 26]
Cranberry12.4–207.3 mg/100 g FWCy-3-gal / Cy-3-ara/ Peo-3-gal / Peo-3-ara[27]
Blackcurrant146.15–403.66 mg/100 g FWCy-3-rut/Cy-3-gluc/Delp-3-rut/ Delp-3-gluc[28]
Red grapes11.5–29.8 g/Kg DMMv-3-gluc/ Mv-3-acetylgluc (V. vinifera)[29]
Mv-3,5-digluc (other than V. vinifera)
Pomegranate juice8.9–346.6 mg/LCy-3,5-digluc/ Cy-3-gluc/ Delp-3,5-digluc/Delp-3-gluc/ Pel-3-gluc[30]
Strawberry8.5–66 mg/100 g FWPel-3-gluc/ Cy-3-gluc/ Pel-3-rut[31]
VegetablesContentMain anthocyaninRef
Black beans32 mg/g DWDelp-3-gluc/ Pet-3-gluc/ Mv-3-gluc[32]
Red cabbage2.32 mg/g DWCy-3-digluc-5-gluc/ Cy-3-coumaroyldigluc-5-gluc/Cy-3-sinapoyldigluc-5-gluc[33]
Purple carrot168.7 mg/100 g FWCy-3-xylosyl-coumaroylglucosyl-gal/ Cy-3-xylosyl-feruloylglucosyl-gal/Cy-3-xylosyl-gal[34]
Purple cauliflower7.18–201 mg/100 g FWCy-3-coumarylsoph-5-gluc/Cy-3-coumarylsoph-5-sinapylgluc[35, 36]
Eggplant (skin)12.1 mg/ 100 g DWDelp-3-rut[37]
Delp-3-coumaroylrut-5-gluc
Colored potatoes14.42–25.79 mg/g DWPel-3-coumaroylrut-5-gluc/ Pel-3-feruloylrut-5-gluc (red)[38]
Pet-Pe and Mv-3-coumaroylrut-5-gluc (blue–purple)
Red onions48.5 mg/ 100 g FWCy-3-gluc/ Cy-3-laminaribioside/Cy-3-malonylgluc/Cy-3-malonyllaminaribioside[14]
Radish32 mg/100 g FWPel-3-coumaroylsoph-5-gluc/Pel-3-feruloylsoph-5-gluc/Pel-3-feruloylsoph-5-malonylgluc/Pel-3-coumaroylsoph-5-malonylgluc[39]
CerealsContentMain anthocyaninRef
Colored Barley8–679 mg/Kg DWCy-3-gluc/ Peo-3-gluc (purple and blue)[40, 41]
Dep-3-gluc/Peo-3-gluc/ Mv-3-gluc (purple)
Purple, blue, Red, black corn27–1439 mg/Kg DWCy-3-gluc/Cy-3-malonylgluc/Cy-3-dimalonylgluc[41, 42]
Purple, red, black rice68–5101 mg/KgCy-3-gluc/Peo-3-gluc (black)/Mv (red)[43]
Cy-3-gluc/Peo-3-gluc/Cy-3-gal/Cy-3-rut (purple)
Black and red sorghum32–680 μg/g DW3-deoxyanthocyanins (Luteolinidin and apigeninidin)[44]
Purple, blue, black wheat10–212 mg/Kg DWCy-3-gluc/Peo-3-gluc/ Cy-malonylgluc/Cy-succinylgluc[45]

Table 1.

Content and main anthocyanins in foodstuffs.

Cy: cyanidin; Delp: delphinidin; Mv: malvidin; Peo: peonidin; Pel: pelargonidin; Pet: petunidin; gluc: glucoside; digluc: diglucoside; sam: sambubioside; gal: galactoside; ara: arabinoside; rut: rutinoside; soph: sophoroside; DW: dry weight; FW: fresh weight.

Advertisement

3. Anthocyanins bioavailability and human metabolic pathways

To validate the prominent health-promoting effects revealed in many in vitro and in vivo models, it is necessary to consider the anthocyanin bioavailability. Anthocyanin bioavailability has been reported to be very low, with recoveries of less than 1% of the ingested anthocyanin dose. However, higher values have been reported reaching recoveries values of 12.4% [46, 47]. As will be described later, anthocyanin can be absorbed from the stomach and small intestine, but a non-negligible part of them can reach the large intestine where they undergo also an extensive catabolism resulting in several metabolites (phenolic acids, propionic acids). For this reason, anthocyanin bioavailability is estimated much greater taking into account not only the phase I and phase II metabolites but also the microbiota catabolites [6]. Although the currently anthocyanin bioavailability researches in humans are limited, they will be discussed below.

3.1 Anthocyanins absorption

Despite having different molecular sizes and types of sugars or acetylated groups attached, anthocyanins can be absorbed intact [48, 49]. Moreover, anthocyanins were found in the blood stream within minutes of consumption in humans [6] suggesting that they can be quickly absorbed from the stomach. This fact is supported by the fact that anthocyanin urine concentrations were fivefold higher when introduced through nasal tubes into the stomach as opposed to the jejunum in patients with colorectal liver metastases after administration of a bilberry extract [50]. In fact, thanks to the low stomach pH (1.5–4) the anthocyanin stability increase permitting their absorption under their glycoside forms. Because anthocyanins are hydrophilic molecules, an organic anion membrane carrier named bilitranslocase, which is expressed in the gastric mucosa has been proposed to mediate anthocyanin transport [51]. Another hypothesis is the involvement of glucose transporter 1 in the transport of anthocyanin glucosides [52]. However, the main site of anthocyanin absorption is the small intestine. They undergo deglycosylation mediated by β-glucosidase in the intestinal lumen and lactasephloridzin hydrolase in the brush border of the intestinal epithelial cells. Alternatively, anthocyanins can enter the enterocyte without deglycosylation via the sodium-coupled glucose transporter after which deglycosylation can occur by cytosol β-glucosidase [51]. These proposed mechanisms are based, in contrast, on in vitro studies. Thus, more studies are required in order to gain insight in human anthocyanin absorption.

3.2 Anthocyanins metabolism

Anthocyanin aglycones that enter the intestinal epithelial are metabolized before reaching portal circulation. This metabolism includes oxidation, reduction, and hydrolysis reactions (phase I metabolism) and conjugation reactions (phase II metabolism). In the intestine, anthocyanins can undergo methylation, sulfation, and glucuronidation by catechol-O-methyltransferase, sulfotransferase, and uridine-5′-diphospho-glucuronosyltransferase enzymes [53]. These reactions can also take place in the liver and the kidneys.

Anthocyanin aglycones can alternatively undergo degradation rendering different phenolic compounds within the intestinal lumen or epithelial cells. Anthocyanin fragmentation can also be a result of the colonic microbiota activity. The microbiota gut can release many deglycosylation enzymes giving rise to aglycones that further undergo ring-opening to produce different benzoic acids or aldehydes such as gallic, vanillic, protocatechuic and syringic acids or aldehydes [46, 54]. Consequently, the phenolics acids portion increases whereas ingested anthocyanin forms portion decreases along the gastrointestinal tract. These products of anthocyanin degradation may be absorbed from the intestine and be transported and further metabolized in the liver and kidneys [55]. The specific anthocyanins metabolism will be described below.

3.3 Anthocyanin’s distribution

The protective effects of flavonoids have been associated with diseases occurring in various tissues, but such claims are mainly based on in vitro evidence using different types of cell lines.

Anthocyanin distribution in tissues has been evaluated in rodents and pig models but never in humans [56, 57, 58, 59]. In a study in which Wistar rats were fed during 15 days with blackberry extract (370 nmol anthocyanin/day), total averaged anthocyanins concentrations were found in jejunum (605 nmol/g), in stomach (68.6 nmol/g), in kidney (3.27 nmol/L), in liver (0.38 nmol/g) and in brain (0.25 nmol/g) [60]. In pigs, anthocyanins were identified in the liver (1.30 pmol/g), in eyes (1.58 pmol/g), in cortex (0.878 pmol/g) and in cerebellum (0.664 pmol/g) after being supplemented with 0, 1, 2, or 4% w/w blueberries for 4 weeks [61]. In anesthetized rats received cyanidin-3-O-glucoside by intravenous injection, this compound has been detected within 15 seconds in the brain tissue and a concentration comparable to that in serum [62]. The results suggested that anthocyanins may provide protection for brain and eye tissues after crossing the blood–brain and blood-retinal barriers.

3.4 Anthocyanin excretion

Anthocyanins can be excreted in urine, bile and even though in air. Around 5% of 13C-label was recovered from urine after the [13C]-cyanidin-3-O-glucoside administration in humans [46]. The urinary excretion of pelargonodin-3-O-glucoside seems to be higher than that of cyanidin-3-O-glucoside [63, 64]. This may be related to the stability of pelargonidin-3-O-glucoside than its real higher absorption. Furthermore, anthocyanins can undergo extensive bile secretion in their original forms or as their phase II metabolites. In human studies enterohepatic recycling of a several xenobiotic could be revealed by a second peak on the plasma concentration versus time curve; This phenomenon can be observed in the literature for several anthocyanins (cyanidin-3-O-glucoside, peonidin-3-O-glucoside, delphinidin-3-O-glucoside) [49, 65].

Finally, volatile metabolites produced from [13C]-cyanidin-3-O-glucoside have also been found in large quantities in breath (6.9% of the administrated dose) following oral administration of [13C]-cyanidin-3-O-glucoside [46].

3.5 Anthocyanin’s behavior in vivo

Researching the xenobiotic methylation and hydroxylation of anthocyanins is challenging based on MS/MS because anthocyanidins are themselves differentiated by hydroxyl and methyl groups on the B-ring. For example, 3’-O-methylation can convert cyanidin to peonidin, and delphinidin to petunidin and 5’-O-methylation converts petunidin to malvidin [66]. Moreover, the removal of functional groups will interconvert anthocyanidins. For example, if cyanidin loses the hydroxyl group in position 2″ from the B-ring, it gives rise to pelargonidin (Figure 1) [67]. As methylation and glucuronidation occurs on hydroxyl groups, abundant in anthocyanins, positional isomers of anthocyanin and anthocyanidin conjugates can be predicted and are indeed detected [64, 65, 66, 67, 68]. As a consequence, data on anthocyanins bioavailabitily in humans after ingestion is potentially more straight forward to interpret.

Figure 1.

Interconversion reactions between anthocyanins: (a) dehydroxylation reaction to arise pelargonidin from cyanidin; (b) methylation pathway that could be carried on by the action of catechol-O-methyltransferase enzyme. Reactions: dOH, dihydroxylation; COMT, catechol-O-methyltransferase.

3.5.1 Cyanidin metabolism

Cyanidin is the best-studied anthocynidin as it is the most widely distributed. Isotopically-labeled cyanidin-3-O-glucoside (C3g) was used to examine the absorption and metabolism of 13C cyanidin-3-O-glucoside in humans [46]. In this study, 44% of the 13C label has been excreted in urine (5.4%), breath (6.9%) and feces (32.1%) at 48 hours after intake. That implies also that more than 50% of the 13C label was still inside the body at that moment. The absorption, digestion, metabolism and excretion of cyanidin-3-O-glucoside concur that methylation and glucuronidation are major routes of cyanidin-3-O-glucoside conjugation in vivo [46, 67]. The metabolites detected in these studies included methyl and glucuronide conjugates of cyanidin-3-O-glucoside, methyl cyanidin-3-O-glucoside (peonidin-3-O-glucoside), and their aglycones cyanidin and peonidin.

Recently, a human study has been carried on to investigate the metabolic pathways and human bioavailability of anthocyanins of red-fleshed apple in which 22% of phenolic compounds are anthocyanins and the main is cyanidin-3-O-galactoside. As a result, cyanidin glycosides (galactoside and arabinose) have been detected in plasma and urine samples. Moreover, peonidin-3-O-galactoside as phase II metabolite of cyanidin-3-O-galactoside methylation by the action of catechol-O-methyl-transferase enzyme has been also detected [69]. Methylation, as one of the first metabolic reaction of cyanidin glycosides was also reported after the oral ingestion of 500 mg of 13C-labeled cyanidin-3-O-glucoside [55].

Protocatechuic acid (PCA) and dihydroxyphenylpropionic acid (dihydrocaffeic acid) were respectively detected in these studies [55, 69]. PCA has been observed at maximum concentrations of 147 nM, thus suggesting that it is not a major metabolite of anthocyanins. The A-ring-derived degradation product, phloroglucinolaldehyde, was present at concentrations greater than either cyanidin-3-O-glucoside or PCA in the serum [55].

Hippuric acid has been identified as the major metabolite of anthocyanins, reaching a maximum concentration of 1962 nM in serum [55]. The detection of 13C2-labeled hippuric acid in this study indicates that PCA and its conjugates are likely further metabolized to form benzoic acid, which is conjugated with glycine to form hippuric acid, or alternatively, formed from the α-oxidation and dihydroxylation of hydroxyphenylacetic acids [64]. PCA might have been formed by β-oxidation of dihydroxyphenylpropionic acid. Then, this phenolic acid could either be further degraded by the action of the gut microbiota to catechol metabolites (α-oxidation), pyrogallol metabolites (hydroxylation) and hydroxybenzoic acid (dehydroxylation), or methylated to vanillic acid [55, 69].

Colonic metabolism has long been speculated to be a major contributor to the overall metabolism of anthocyanins [70]. It has been proposed that phenylpropenoic acids arise from cyanidin-3-O-glucoside as a result of bacterial cleavage of the C-ring in the colon [71], which is supported by the detection of caffeic acid and its methyl metabolite, ferulic acid [55].

On the basis of the findings of these studies, the metabolic pathway of cyanidin-3-O-glucoside and peonidin-3-O-glucoside can be summarized as undergoing multiple biotransformation (Figure 2).

Figure 2.

Proposed metabolic pathway for cyanidin and peonidin glucosides. Reactions: dH, dehydrogenation; SULT, sulphotransferase; UGT, glucuronosyl-transferase; COMT, catechol-O-methyltransferase; dOH, dehydroxylation; dMe, demethylation; α-oxidation, one decarboxylation; β-oxidation, two decarboxylation.

3.5.2 Pelargonidin metabolism

As it was shown before, demethylation and dihydroxylation of highly substituted anthocyanins gives rise to pelargonidin, that helps to explain the high apparent recovery of pelargonidin-based metabolites [63]. Indeed, pelargonidin glucuronide has been detected in urine after the ingestion of boysenberry (rich in four cyanidin glycosides and without pelargonidin) in humans [67]. Furthermore, strawberry pelargonidin was found to be metabolized to 4-hydroxybenzoic acid in humans when 13 healthy volunteers consumed 300 g of fresh or stored strawberries [72]. In which 4-hydroxybenzoic acid plasma recovery was 23 and 17 mmol, corresponding to the percentages of 54 and 56% of pelargonidin-3-O-glucoside.

3.5.3 Delphinidin, petunidin and malvidin metabolism

After administration of Concord grape juice in humans, delphinidin-3-O-glucoside, petunidin-3-O-glucoside and malvidin-3-O-glucoside were found in blood or urine. Glucuronidated metabolites of aglycones have been identified as their major metabolites in urine [49]. In the urine of volunteers administered bilberry-lingonberry puree, a small amount of syringic acid, a potential metabolite of malvidin glycosides, was detected [73]. Recently, in a long-term study with humans consuming blueberry juice, 55 anthocyanin metabolites have identified. Among them, malvidin-3-O-glucoside, malvidin-3-O-galactoside and malvidin-3-O-arabinoside have been described representing around 5% of the total excretion [68]. In vitro experiments state that gallic acid is the major degradation product of delphinidin-3-O-glucoside. Moreover, syringic acid was described as the mean metabolite for malvidin-3-O-glucoside [13].

Advertisement

4. Anti-neuroinflammatory effects on anthocyanins and their metabolites

As it was discussed above, several works have demonstrated that anthocyanins can cross the blood brain barrier, and accumulate in brain endothelial cells, brain parenchymal tissue, striatum, hippocampus, cerebellum and cortex [74, 75, 76]. Consequently, the study of anthocyanins as therapeutic agents in neurodegenerative diseases has gained relevance.

Neuroinflammation is a common physiopathological hallmark in neurodegenerative diseases as Alzheimer, Parkinson or amyotrophic lateral sclerosis, among others. This process is mediated by microglial cells, the immune cells of central nervous system. Their functions are related with the host defense by destroying pathogens, promoting tissue repair and facilitating tissue homoeostasis [77]. Nowadays it is well establish that these cells can adopt different phenotypes depending on the brain environment to shift into pro-inflammatory/neurotoxic or anti-inflammatory/neuroprotective phenotypes. The stimulation agent will be the responsible of trigger one or another phenotype. Thus, when microglial cells are stimulated with lipopolysaccharide (LPS) and interferon gamma (IFN-γ), microglia develop a classically phenotype or M1, while when it is activated with IL-4 microglia show an alternative activated phenotype or M2 [78]. On the one hand, M1 microglia type is characterized by the production of nitric oxide (NO) by the inducible nitric oxide synthase (iNOS) [79, 80] and by the expression of inflammatory chemokines and cytokines, such as interleukin (IL)-6, IL-12, IL-1β, IL-23, and tumor necrosis factor (TNF)-α. All this culminates in the influx of new immune system cells to combat the infection. When neuroinflammation becomes chronic, it can ultimately lead to neuronal cell death. On the other hand, M2 microglia is characterized by a suppression of IL-12 secretion and an induction of the release of IL-10, transforming growth factor beta (TGB-β), IL-1R [81]. Furthermore, the expression of arginase-1 instead of iNOS, switching arginine metabolism from production of NO to ornithine, and also the increase of polyamines production for extracellular matrix and collagen synthesis, promotes the neuroregeneration and tissue repair [82].

Several in vitro and in vivo studies have shown that anthocyanins, overall rich anthocyanins extracts, are able to be neuroprotective and counteract neuroinflammation [83, 84]. Regarding in vitro studies, a blueberry extract (25–50 μg/mL) have demonstrated to be able to diminish the release of NO, TNF-α, iNOS and cyclooxygenase-2 (COX-2) protein expression in LPS-stimulated BV2 cells [85, 86] and in LPS or IFNγ-stimulated N9 cells [87]. In addition, they proved that this effect is mediated by NF-ĸB signaling pathway, via the inhibition of Nuclear Factor Kappa B (NF-ĸB) nuclear translocation [88]. NF-κB is a key inflammation regulator located in the cell cytoplasm and their nuclear translocation trigger the expression of inflammation-related genes. Likewise, the anti-inflammatory effect of blueberry extract has been related with the activation of janus kinase/signal transducer and activator of transcription (JAK/STAT) signaling (pathway activated after IFN-γ stimulation) [87]. Other study that evaluated the potential anti-neuroinflammatory effect of a large variety berries extracts (blackberry, black raspberry, blueberry, cranberry, red raspberry, and strawberry), showed that the cranberry extract (20 μg/mL) was the most active diminishing the NO production and inhibiting the fibrillation of amyloid-β peptide (peptide responsible of the formation of senile plaques in brain Alzheimer’s patients) [89]. Moreover, elderberry extracts (400 μg/mL, ethanol or ethyl acetate extracts) has also been proposed as a potent suppressor of NO release [90]. Mitogen-activated protein kinases (MAPKs) are a family of serine/threonine protein kinases that mediate fundamental cellular responses to external stress signals. In particular, p38 MAPK, is involve in the regulation of the synthesis of inflammation mediators being for that a potential target for anti-inflammatory therapeutics. In this context, an anthocyanin-enriched extract of acai berry and a mixture of anthocyanins isolated from black soybean seed coats (cyanidin-3-O-glucoside (72%), delphinidin-3-O-glucoside (20%) and petunidin-3-O-glucoside (6%)) have demonstrated that MAPK pathways can be also implicated in the decrease in inflammatory mediators and cytokines [91, 92]. Finally, this year, an article have been published showing that a black raspberry extract reduced the production of IL-18, IL-1β and reactive oxygen species (ROS) in LPS-induced BV2 microglia by down-regulating the level of NADPH oxidase 2 (NOX2) and its downstream factors, including thioredoxin-interacting protein (TXNIP) and NOD-like receptor protein 3 (NLRP3) inflammasome [93]. The complexity of these extracts containing several structurally diverse anthocyanins makes difficult the interpretation of results. For this, some papers have been published concerning the evaluation of the activity of pure anthocyanins. This type of studies provide insight into the plausible mechanism of single compounds facilitating the understanding. Some (although very few) studies, have been performed with pure anthocyanins. An interesting work published by Miraeles and collaborators demonstrated that cyanidin-3-O-glucoside, (1 μM) and also cyanidin-3-O-glucoside and a mixture of 3′-methyl-cyanidin-3-O-glucoside and 4′-methyl-cyanidin-3-O-glucoside, were able to decrease a great number of pro-inflammatory mediators. Indeed, TNF-α and IL-6 mRNA expression was decrease by and methyl-cyanidin-3-O-glucoside. Moreover, cyanidin reverted the IL-1β expression. This paper also shows that even though cyanidin and theirs different chemical forms, are not able to shift microglia to an M2, they can interact with microglia biology increasing CX3C Motif Chemokine Ligand 1 (CX3CL1) expression [94]. Neurons can express this chemokine, which mediates microglial activation via interacting with its sole receptor CX3CR1 in microglia (axis CX3CL1/CX3CR1). Comparable results have been recently published, showing that the underlying responsible anti-neuroinflammatory mechanism of cynidin-3-O-glucoside is related with suppression of NF-κB and p38 MAPK signaling pathways [95]. Other pure anthocyanins as delphinidin-3-O-glucoside, malvidin-3-O-glucoside (20 μM) [86] and pelargonidin-3-O-glucoside (100 μM) [96] are also shown to be able to suppress the LPS/IFN-γ -induced phosphorylation of p38, p42/44 and MAPKs in BV2 cells and mouse C8-4B microglial cells.

Concerning in vivo studies, only around ten papers have been published about the effect of anthocyanins extracts/pure compounds in microglia-related diseases. The first paper published in 2015, evaluate the effect of a blackberry extract consumption at a dose of 25 mg/Kg in an standard or in a high fat diet, during 17 weeks in Wistar rats. The results showed that the intake of this fruit, in both dietary conditions, modulates CX3CL1 expression and the thymus chemokine TCK-1. In addition, they also found that blueberry can ameliorate synapse connectivity by regulating platelet-derived growth factor (PDGF)-AA, activin, vascular endothelial growth factor (VEGF) and agrin [97]. Another three works proved that the consumption of anthocyanins extracted of Korean black soybean (24–100 mg/Kg) inhibited the activation of astrocytes and neuroinflammation via suppression NF-κB, iNOS and TNF-α in the hippocampus and cortex regions of D-galactose and LPS treated rats brain [98, 99, 100].

Not only the reduction of IL-1β and TNF-α but also the reduction of IL-10 induced by LPS was observed after the treatment with 100 mg/Kg of anthocyanin obtained from V. vinifera grapes in mice [101]. Moreover, the addition of an enriched anthocyanin extract from purple corn in water (mean of 53 mg/Kg body weight) has proved to be able to reduce microglia size and Iba1 staining (marker of microglia activation) and IL-6, TNF-α, IL-1β, MCP-1 and iNOS. Interestingly, this papers showed that purple corn anthocyanins not only inhibit microglia activation but also promote their shift towards the production of anti-inflammatory mediators, such as arginase-1, IL-10, Fizz1, IL-13 and YM-1 (a marker of M2 microglia phenotype) [102]. In agreement, a diet based on anthocyanin-rich wheat during 6 months on Alzheimer and Parkinson disease mouse models, reduced the α-synuclein accumulation (protein responsible of the formation of Lewy bodies in Parkinson patients) [103].

Other rich anthocyanins fruits as bilberry has exhibited promising results. In fact, the administration in food or in water of an bilberry extract (20 mg/Kg day) on APP/PSEN1 mice and their littermates downregulates the expression of several inflammatory factors (TNF-α, NF-κβ, IL-1β, IL-6, COX-2, iNOS and cluster of differentiation 33 (CD33), the chemokine receptor CX3CR1, but also and for the first time, the microglia homeostatic factors (TREM2 and TYROBP) and the Toll-like receptors (TLR2 and TLR4) [104].

As was explained above, circulating concentrations of phenolic acid metabolites derived from anthocyanin degradation such as protocatechuic, gallic, syringic and ferulic acids have been observed at up to eight times to that of the parent anthocyanins [72]. Two papers have been very recently published showing that a mixture of anthocyanin metabolites can have anti-neuroainflammatory activities. Indeed, an in vitro digested blueberry and raspberry extracts (1.25–10 μg/mL) proved be able to reduce some key inflammatory markers (TNF-α and NO) and ROS in N9 cell line exposure to LPS and IFN-γ. This bioactivity has been related with the NF-κB and STAT1 molecular pathways [87, 105]. By using pure compounds, ferulic, caffeic and protocatechuic acids have been the most studied metabolites on neurodegenerative diseases with an inflammatory component. The pre-treatment of BV2 microglial cells (1 and 4 hours) with PCA (2.5–10 μM) attenuated microglial activation by suppressing TLR4-mediated NF-κB and MAPKs (JNK, p38, ERK) activation and SIRT1 pathway [106, 107]. Other interesting paper displayed that PCA (3,4-dihydroxybenzoic acid), ant not 4-hydroxybenzoic acid can reduced NO production of BV2 cells, however, in this case, PCA concentrations are ten times higher (100 μM) [108]. Furthermore, Koga and their co-workers demonstrated that caffeic acid-treated mice exhibited significantly lower levels of 4-hydroxynonenal (oxidative stress marker) and fewer activated microglia [109]. A long-term treatment (4-weeks) with ferulic acid (in drinking water (0.006%)) for male mice prevented the Aβ1–42-induced activation of microglia [110]. Ferulic acid has also demonstrated interfered with TLR4 interaction sites in mouse hippocampus and in BV2 cells by down streaming iNOS, COX-2, TNF-α, and IL-1β via JNK and NF-κB phosphorylation [111]. Furthermore, the intra-peritoneal injection of 30 mg/Kg of vanillic acid reversed LPS-induced glial cells activation, neuroinflammation (TNF-α, IL1-β, and COX-2) and amyloidogenic markers (β-site amyloid precursor protein (APP)–cleaving enzyme 1 (BACE1) and amyloid-β [112]. Finally, concerning gallic acid, two articles can be highlighted. This compound (at 5–50 μM concentration) in a co-culture system consisted on BV2 and Neuro-2A cells and in primary microglia resulted on the diminution of cytokine production induced by the Aβ peptide [113]. After the orally administration of gallic acid (100 mg/Kg) 1 hour prior to the LPS infusion and daily afterwards for 7 days, an attenuation of LPS-induced elevation in heme oxygenase-1 level and α-synuclein aggregation was observed. Moreover, this same work revealed that gallic acid diminished the iNOS gene expression and the NO production in vitro [114].

However, any anti-neuroinflammatory activity has been reported for glucuronidated, sulfated and O-methylated anthocyanins and their corresponding metabolites. This lack of studies can be explained due to the lack of commercial compounds which makes the chemical synthesis or hemy-synthesis as the only alternative available. Even if it is a challenge to obtain glucuronidated, sulfated and O-methylated anthocyanins some methodologies can be used and are reported in the literature. For example methylation of cyandin-3-O-glucoside can be carried out by the reaction with dimethylcarbonate [115]. Regarding the hemisynthesis of sulfated derivatives several approaches are possible by bringing the anthocyanins into contact with chlorosulfonic acid [116] or even with sulfur trioxide-N-triethylamine [117]. Finally, glucuronidated anthocyanin can be obtain by acidic aldo-condensation between trihydroybenzalhaldehyde and acetophenone which have been previously functionalized with the expected OH and OMe group as well as the glucuronic acid at the proper position [118].

Advertisement

5. Conclusion

Anthocyanins represents one the most consumed polyphenols in human diet. However, their type, complexity and quantities depends on the foodstuff. For example, anthocyanins in vegetables and cereals are chemically more complex in comparison with fruits, but berries are the major source of these compounds. Anthocyanin bioavailability has been reported to be very low, with recovery of less than 1% of the ingested anthocyanin dose. However, nowadays much greater bioavailability values have reported taking into account not only the phase I and phase II metabolites but also the microbiota catabolites. One of the peculiarities of anthocyanin metabolism is their capacity of interconversion between them. For example, dehydroxylation reaction can arise pelargonidin from cyanidin and methylation reactions can convert delphinidin into petunidin and malvidin. For this reason, metabolism data after anthocyanin ingestion is more straight forward to interpret. Regarding metabolism, cyanidin is the most studied anthocyanin due to ubiquitous character in the nature. However, more studies are necessary to better understand the similarities and differences with the other less studied anthocyanins. Even though several papers have reported the potential anti-neuroinflammatory effect of rich anthocyanin extracts, anthocyanins or their metabolites, the number of papers are very scarce. The most important limitation to study the activity of anthocyanin metabolites is the lack of commercial phase II and microbiota catabolites compounds. Thus, the chemical synthesis is the most employed technique to obtain standards although more developments are requires in order to obtain greater quantities. Moreover, little is known about the molecular mechanisms implicated in the observed effects. Furthermore, the majority of works are based on the study of the microglia M1 phenotype, so more studies are necessary to know if anthocyanins and their metabolites are able to induce an anti-inflammatory phenotype. To sum up, more research is necessary to stablish if anthocyanins and their metabolites are efficacious in slowing the progression of brain aging or of neurodegenerative diseases with an inflammatory component.

References

  1. 1. Kong JM, Chia LS, Goh NK, Chia TF, Brouillard R. Analysis and biological activities of anthocyanins. Phytochemistry. 2003;64:923-933. DOI: 10.1016/s0031-9422(03)00438-2. Erratum in: Phytochemistry. 2008;69:1939-1940
  2. 2. Anderson, OM & Jordheim, M. The anthocyanins. In Flavonoids: Chemistry, Biochemistry and Applications. Boca Raton, FL: CRC Press/Taylor & Francis Group. 2006
  3. 3. Mazza G, Miniati E. Anthocyanins in fruits, vegetables, and grains. 1st ed. Boca Raton: CRC Press; 1993. DOI: 10.1201/97813510697000
  4. 4. Smeriglio A, Barreca D, Bellocco E, Trombetta D. Chemistry, pharmacology and health benefits of anthocyanins. Phytotherapy Research. 2016;30:1265-1286. DOI: 10.1002/ptr.5642
  5. 5. Zamora-Ros R, Knaze V, Luján-Barroso L, Slimani N, Romieu I, Touillaud M, Kaaks R, Teucher B, Mattiello A, Grioni S, Crowe F, Boeing H, Förster J, Quirós JR, Molina E, Huerta JM, Engeset D, Skeie G, Trichopoulou A, Dilis V, Tsiotas K, Peeters PHM, Khaw K-T, Wareham N, Bueno-de-Mesquita B, Ocké MC, Olsen A, Tjønneland A, Tumino R, Johansson G, Johansson I, Ardanaz E, Sacerdote C, Sonestedt E, Ericson U, Clavel-Chapelon F, Boutron-Ruault M-C, Fagherazzi G, Salvini S, Amiano P, Riboli E, González CA. Estimation of the intake of anthocyanidins and their food sources in the European Prospective Investigation into Cancer and Nutrition (EPIC) study. The British Journal of Nutrition. 2011:1-10. DOI: 10.1017/ S0007114511001437
  6. 6. Fang, J. Bioavailability of anthocyanins. Drug Metabolism Reviews. 2014;46:508-520. DOI: 10.3109/036002532.2014.978080
  7. 7. Konczak I, Zhang W. Anthocyanins—more than nature’s colours. Journal of Biomedicine and Biotechnology. 2004;5:239-240. DOI: 10.1155/s1110724304407013
  8. 8. He J, Giusti MM. Anthocyanins: natural colorants with health-promoting properties. Annual Review of Food Science and Technology. 2010;1:163-187. DOI: 10.1146/annurev.food.080708.100754
  9. 9. Tulio AZ Jr, Reese RN, Wyzgoski FJ, Rinaldi PL, Fu R, Scheerens JC, Miller AR. Cyanidin 3-rutinoside and cyanidin 3-xylosylrutinoside as primary phenolic antioxidants in black raspberry. Journal of Agricultural and Food Chemistry. 2008;56:1880-1888. DOI: 10.1021/jf072313k
  10. 10. Clifford MN. Anthocyanins–nature, occurrence and dietary burden. Journal of Science of Food and Agriculture. 2000;80:1063-1072. DOI: 10.1002/(SICI)1097-0010(20000515)80:7<1063::AID-JSFA605>3.0.CO;2-Q
  11. 11. Borkowski T, Szymusiak H, Gliszczynska-Swiglo A, Tyrakowska B. The effect of 3-O-beta-glycosylation on structural transformations of anthocyanins. Food Research International. 2005; 38:1031-1037. DOI: 10.1016/j.foodres.2005.02.020
  12. 12. Fossen T, Cabrita L, Andersen OM. Colour and stability of pure anthocyanins influenced by pH including the alkaline region. Food Chemistry. 1998;63:435-440. DOI: 10.1016/s0308-8146(98)00065-x
  13. 13. Fleschut I, Kratzer F, Rechkemmer G, Kulling SE. Stability and biotransformation of various dietary anthocyanins in vitro. European Journal of Nutrition. 2006;45:7-18. DOI: 10.1007/s00394-005-0557-8
  14. 14. Wu X, Beecher GR, Holden JM, Haytowitz DB, Gebhardt SE, Prior RL. Concentrations of anthocyanins in common foods in the United States and estimation of normal consumption. Journal of Agricultural and Food Chemistry. 2006;54:4069-4075. DOI: 10.1021/jf060300l
  15. 15. Bai W, Wang C, Ren C. Intakes of total and individual flavonoids by US adults. International Journal of Food Science and Nutrition. 2014;65:9-20. DOI: 10.3109/09637486.2013.832170
  16. 16. Vogiatzoglou A, Mulligan AA, Lentjes MA, Luben RN, Spencer JP, Schroeter H, Khaw KT, Kuhnle GG. Flavonoid intake in European adults (18 to 64 years). PLoS One. 2015;10:e0128132-e0128154. DOI: 10.1371/journal.pone.0128132
  17. 17. Igwe EO, Charlton KE, Probst YC. Usual dietary anthocyanin intake, sources and their association with blood pressure in a representative sample of Australian adults. Journal of Human Nutrition Dietetics. 2019;32:578-590. DOI: 10.1111/jhn.12647
  18. 18. Li G, Zhu Y, Zhang Y, Lang J, Chen Y, Ling W. Estimated daily flavonoid and stilbene intake from fruits, vegetables, and nuts and associations with lipid profiles in Chinese adults. Journal of the Academy of Nutrition and Dietetics. 2013;113:786-794. DOI: 10.1016/j.jand.2013.01.018
  19. 19. Patras A, Brunton NP, Colm O'Donnell, B.K. Tiwari. Effect of thermal processing on anthocyanin stability in foods; mechanisms and kinetics of degradation. Trends in Food Science & Technology. 2010;21:3-11. DOI: 10.1016/j.tifs.2009.07.004
  20. 20. Bars-Cortina D, Macià A, Iglesias I, Romero MP, Motilva MJ. Phytochemical profiles of new red-fleshed apple varieties compared with traditional and new white-fleshed varieties. Journal of Agricultural and Food Chemistry. 2017;65:1684-1696. DOi: 10.1021/acs.jafc.6b02931
  21. 21. Bureau S, Renard CMGC, Reich M, Ginies C, Audergon JM. Change in anthocyanin concentrations in red apricot fruits during ripening. LWT - Food Science and Technology. 2009; 42:372-377. DOI: 10.1016/j.lwt.2008.03.010
  22. 22. Müller D, Schantz M, Richling E. High performance liquid chromatography analysis of anthocyanins in bilberries (Vaccinium myrtillus L.), blueberries (Vaccinium corymbosum L.), and corresponding juices. Journal of Food Science. 2012 Apr;77(4):C340-C345. DOI: 10.1111/j.1750-3841.2011.02605.x
  23. 23. Pires TCSP, Caleja C, Santos-Buelga C, Barros L, Ferreira ICFR. Vaccinium myrtillus L. Fruits as a novel source of phenolic compounds with health benefits and industrial applications - a Review. Current Pharmaceutical Design. 2020;26:1917-1928. DOI: 10.2174/1381612826666200317132507
  24. 24. Fan-Chiang H, Wrolstad RE. Anthocyanin pigment composition of blackberries. Journal of Food Science. 2005;70:C198-C202. DOI: 10.1111/j.1365-2621.2005
  25. 25. Gao L, Mazza G, Characterization, quantitation, and distribution of anthocyanins and colorless phenolics in sweet cherries. Journal of Agricultural and Food Chemistry.1995;43:343-346. DOI: 10.1021/jf00050a015
  26. 26. Grigoras C, Destandau E, Zubrzycki S, Elfakir C. Sweet cherries anthocyanins: An environmental friendly extraction and purification method. Separation and Purification Technology. 2012;100:51-58. DOI: 10.1016/j.seppur.2012.08.032
  27. 27. Cesoniene L, Jasutiene I, Sarkinas A. Phenolics and anthocyanins in berries of European cranberry and their antimicrobial activity. Medicina (Kaunas). 2009;45(12):992-999. PMID: 20173403
  28. 28. Chen X, Parker J, Krueger C, Shanmuganayagam D, Reed J. Validation of HPLC assay for the identification and quantification of anthocyanins in black currants. Analytical Methods. 2014;6:8141-8147. DOI: 10.1039/C4AY01500B
  29. 29. Ky I, Lorrain B, Kolbas N, Crozier A, Teissedre PL. Wine by-products: Phenolic characterization and antioxidant activity evaluation of grapes and grape pomaces from six different French grape varieties. Molecules. 2014;19:482-506. DOI: 10.3390/molecules19010482
  30. 30. Vardin H, Fenercioğlu H. Study on the development of pomegranate juice processing technology: clarification of pomegranate juice. Nahrung. 2003;47:300-303. DOI: 10.1002/food.200390070
  31. 31. Aaby K, Mazur S, Nes A, Skrede G. Phenolic compounds in strawberry (Fragaria x ananassa Duch.) fruits: Composition in 27 cultivars and changes during ripening. Food Chemistry. 2012;132:86-97. DOI: 10.1016/j.foodchem.2011.10.037
  32. 32. Mojica L, Berhow M, Gonzalez de Mejia E. Black bean anthocyanin-rich extracts as food colorants: Physicochemical stability and antidiabetes potential. Food Chemistry. 2017;229:628-639. DOI: 10.1016/j.foodchem.2017.02.124
  33. 33. Wiczkowski W, Dorota Szawara-Nowak, Joanna Topolska. Red cabbage anthocyanins: Profile, isolation, identification, and antioxidant activity. Food Research International. 2013;51:303-309. DOI: 10.1016/j.foodres.2012.12.015
  34. 34. Assous M, Abdel-Hady M, Medany GM. Evaluation of red pigment extracted from purple carrots and its utilization as antioxidant and natural food colorants. Annals of Agricultural Science. 2014;59:1-7. DOI: 10.1016/j.aoas.2014.06.001
  35. 35. Kapusta-Duch J, Szeląg-Sikora A, Sikora J, Niemiec M, Gródek-Szostak Z, Kuboń M, Leszczyńska T, Borczak B. Health-promoting properties of fresh and processed purple cauliflower. Sustainability. 2019;11:4008-4013. DOI: 10.3390/su11154008
  36. 36. Lo Scalzo R, Genna A, Branca F, Chedin M, Chassaigne H. Anthocyanin composition of cauliflower (Brassica oleracea L. var. botrytis) and cabbage (B. oleracea L. var. capitata) and its stability in relation to thermal treatments. Food Chemistry. 2008;107:136-144. DOI: 10.1016/j.foodchem.2007.07.072
  37. 37. Lo Scalzo R, Fibiani M, Francese G, D'Alessandro A, Rotino GL, Conte P, Mennella G. Cooking influence on physico-chemical fruit characteristics of eggplant (Solanum melongena L.). Food Chemistry. 2016;194:835-842. DOI: 10.1016/j.foodchem.2015.08.063
  38. 38. Ji X, Rivers L, Zielinski Z, Xu M, MacDougall E, Stephen J, Zhang S, Wang Y, Chapman RG, Keddy P, Robertson GS, Kirby CW, Embleton J, Worrall K, Murphy A, De Koeyer D, Tai H, Yu L, Charter E, Zhang J. Quantitative analysis of phenolic components and glycoalkaloids from 20 potato clones and in vitro evaluation of antioxidant, cholesterol uptake, and neuroprotective activities. Food Chemistry. 2012;133:1177-1187. DOI: 10.1016/j.foodchem.2011.08.065
  39. 39. Koponen JM, Happonen AM, Mattila PH, Törrönen AR. Contents of anthocyanins and ellagitannins in selected foods consumed in Finland. Journal of Agricultural and Food Chemistry. 2007;55:1612-1619. DOI: 10.1021/jf062897a
  40. 40. Bellido GG, Beta T. Anthocyanin composition and oxygen radical scavenging capacity (ORAC) of milled and pearled purple, black, and common barley. Journal of Agricultural and Food Chemistry. 2009;57:1022-1028. DOI: 10.1021/jf802846x
  41. 41. Francavilla A, Joye IJ. Anthocyanins in whole grain cereals and their potential effect on health. Nutrients. 2020;12:2922-2942. DOI: 10.3390/nu12102922
  42. 42. Harakotr, B., Suriharn, B., Tangwongchai, R., Scott, M. Paul, & Lertrat, K. (2014). Anthocyanins and antioxidant activity in coloured waxy corn at different maturation stages. Journal of Functional Foods, 9, 109-118. DOI: 10.1016/j.jff.2014.04.012
  43. 43. Chen XQ, Nagao N, Itani T, Irifune K. Anti-oxidative analysis, and identification and quantification of anthocyanin pigments in different coloured rice. Food Chemistry. 2012;135:2783-2788. DOI: 10.1016/j.foodchem.2012.06.098
  44. 44. Awika JM, Yang L, Browning JD, Faraj A. Comparative antioxidant, antiproliferative and phase II enzyme inducing potential of sorghum (Sorghum bicolor) varieties. LWT - Food Science and Technology. 2009;42:1041-1046. DOI: 10.1016/j.lwt.2009.02.003
  45. 45. Abdel-Aal el-SM, Young JC, Rabalski I. Anthocyanin composition in black, blue, pink, purple, and red cereal grains. Journal of Agricultural and Food Chemistry. 2006;54:4696-4704. DOI: 10.1021/jf0606609
  46. 46. Czank C, Cassidy A, Zhang Q, Morrison DJ, Preston T, Kroon PA, Botting NP, Kay CD. Human metabolism and elimination of the anthocyanin, cyanidin-3-glucoside: a (13)C-tracer study. The American Journal of Clinical Nutrition. 2013;97:995-1003. DOI: 10.3945/ajcn.112.049247
  47. 47. Lila MA, Burton-Freeman B, Grace M, Kalt W. Unraveling anthocyanin bioavailability for human health. Annual Review of Food Science and Technology. 2016;7:375-394. DOI: 10.1146/annurev-foof-041715-033346
  48. 48. Kurilich AC, Clevidence BA, Britz SJ, Simon PW, Novotny JA. Plasma and urine responses are lower for acylated versus nonacylated anthocyanins from raw and cooked purple carrots. Journal of Agricultural and Food Chemistry. 2005;53:6537-6542. DOI: 10.1021/jf050570o
  49. 49. Stalmach A, Edwards CA, Wightman JD, Crozier, A. Gastrointestinal stability and bioavailability of (poly)phenolic compounds following ingestion of Concord grape juice by humans. Molecular Nutrition & Food Research. 2012;56,497-509. DOI: 10.1002/mnfr.201100566
  50. 50. Cai H, Thomasset SC, Berry DP, Garcea G, Brown K, Steward WP, Gescher AJ. Determination of anthocyanins in the urine of patients with colorectal liver metastases after administration of bilberry extract. Biomedical Chromatography. 2011;25:660-663. DOI: 10.1002/bmc.1499
  51. 51. Hribar U, Ulrih NP. The Metabolism of Anthocyanins. Current Drug Metabolism. 2014;15:3−13. DOI: 10.2174/1389200214666131211160308
  52. 52. Oliveira H, Fernandes I, Brás NF, Faria A, De Freitas V, Calhau C, Mateus N. Experimental and theoretical data on the mechanism by which red wine anthocyanins are transported through a human MKN-28 gastric cell model. Journal of Agricultural and Food Chemistry. 2015;63:7685−7692. DOI: 10.1021/acs.jafc.5b00412
  53. 53. Zhong S, Sandhu A, Edirisinghe I, Burton-Freeman, B. Mint: Characterization of wild blueberry polyphenols bioavailability and kinetic profile in plasma over 24-h period in human subjects. Molecular Nutrition & Food Research. 2017;61:1700405-1700418. DOI: 10.1002/mnfr.201700405
  54. 54. Vitaglione P, Donnarumma G, Napolitano A, Galvano F, Gallo A, Scalfi L, Fogliano V. Protocatechuic acid is the major human metabolite of cyanidin-glucosides. The Journal of Nutrition. 2007;137:2043-2048. DOI: 10.1093/jn/137.9.2043
  55. 55. de Ferrars R, Czank C, Zhang Q, Botting N, Kroon P,Cassidy A, Kay C. The Pharmacokinetics of anthocyanins and their metabolites in humans. British Journal of Pharmacology. 2014;171:3268−3282. DOI: 10.1111/bph.12676
  56. 56. Sandoval-Ramírez BA, Catalán Ú, Fernández-Castillejo S, Rubió L, Macià A, Solà R. Anthocyanin tissue bioavailability in animals: Possible implications for human health. A systematic review. Journal Agricultural and Food Chemistry. 2018;66:11531-11543. DOI: 10.1021/acs.jafc.8b04014
  57. 57. Kirakosyan A, Seymour EM, Wolforth J, McNish R, Kaufman PB, Bolling SF. Tissue bioavailability of anthocyanins from whole tart cherry in healthy rats. Food Chemistry. 2015;171:26-31. DOI: 10.1016/j.foodchem.2014.08.114
  58. 58. Chen TY, Kritchevsky J, Hargett K, Feller K, Klobusnik R, Song BJ, Cooper B, Jouni Z, Ferruzzi MG, Janle EM. Plasma bioavailability and regional brain distribution of polyphenols from apple/grape seed and bilberry extracts in a young swine model. Molecular Nutrition & Food Research. 2015;59:2432-2447. DOI: 10.1002/mnfr.201500224
  59. 59. Aqil F, Vadhanam MV, Jeyabalan J, Cai J, Singh IP, Gupta RC. Detection of anthocyanins/anthocyanidins in animal tissues. Journal of Agricultural and Food Chemistry. 2014;62:3912-3918. DOI: 10.1021/jf500467b
  60. 60. Talavéra S, Felgines C, Texier O, Besson C, Gil-Izquierdo A, Lamaison J-L, Rémésy C. Anthocyanin metabolism in rats and their distribution to digestive area, kidney, and brain. Journal of Agricultural and Food Chemistry. 2005;53:3902-3908. DOI: 10.1021/jf050145v
  61. 61. Kalt W, Blumberg JB, McDonald JE, Vinqvist-Tymchuk MR, Fillmore SAE, Graf BA, O’Leary JM, Milbury PE. Identification of anthocyanins in the liver, eye, and brain of blueberry-fed pigs. Journal of Agricultural Food Chemitry. 2008 ;56:705-12. DOI : 10.1021/jf071998l
  62. 62. Fornasaro S, Ziberna L, Gasperotti M, Tramer F, Vrhovšek U, Mattivi F, Passamonti S. Determination of cyanidin 3-glucoside in rat brain, liver and kidneys by UPLC/MS-MS and its application to a short-term pharmacokinetic study. Scientific Reports. 2016;6:1-11. DOI: 10.1038/srep22815
  63. 63. Carkeet C, Clevidence BA, Novotny JA. Anthocyanin excretion by humans increases linearly with increasing strawberry dose. Journal ol Nutrition. 2008;138:897-902. DOI: 10.1093/jn/138.5.897
  64. 64. Mullen W, Edwards CA, Serafini M, Crozier A. Bioavailability of pelargonidin-3-O-glucoside and its metabolites in humans following the ingestion of strawberries with and without cream. Journal of Agricultural and Food Chemistry. 2008;56:713-719. DOI: 10.1021/jf072000p
  65. 65. Charron CS, Kurilich AC, Clevidence BA, Simon PW, Harrison DJ, Britz SJ, Baer DJ, Novotny JA. Bioavailability of anthocyanins from purple carrot juice: effects of acylation and plant matrix. Journal of Agricultural and Food Chemistry. 2009;57:1226-1230. DOI: 10.1021/jf802899s
  66. 66. Crozier A, Del Rio D, Clifford MN. Bioavailability of dietary flavonoids and phenolic compounds. Molecular Aspects of Medicine. 2010;31:446-467. DOI: 10.1016/j.mam.2010.09.007
  67. 67. Cooney JM, Jensen DJ, McGhie TK. LC-MS identification of anthocyanins in boysenberry extract and anthocyanin metabolites in human urine following dosing. Journal of the Science of Food and Agriculture. 2004;84:237-245. DOI: 10.1002/jsfa.1645
  68. 68. Kalt W, McDonald JE, Liu Y, Fillmore SAE. Flavonoid metabolites in human urine during blueberry anthocyanin intake. Journal of Agricultural and Food Chemistry. 2017;65:1582-1591. DOI: 10.1021/acs.jafc.6b05455
  69. 69. Yustea S, Ludwiga IA, Rubió L, Romero M-P, Pedret A, Valls R-M, Solà R, Motilva M-J, Macia A. In vivo biotransformation of (poly)phenols and anthocyanins of red-fleshed apple and identification of intake biomarkers. Journal of Functional Foods. 2019;55:146-155. DOI: 10.1016/j.jff.2019.02.013
  70. 70. Cardona F, Andrés-Lacueva C, Tulipani S, Tinahones FJ, Queipo-Ortuño MI. Benefits of polyphenols on gut microbiota and implications in human health. J Nutrition Biochemistry. 2013;24:1415-1422. DOI: 10.1016/j.jnutbio.2013.05.001
  71. 71. Gonzalez-Barrio R, Edwards C, Crozier A. Colonic catabolism of ellagitannins, ellagic acid, and raspberry anthocyanins: in vivo and in vitro studies. Drug Metabolism & Disposition. 2011;39:1680-1688. DOI: 10.1124/dmd.111.039651
  72. 72. Azzini E,Vitaglione P, Intorre F, Napolitano A, Durazzo A, Foddai MS, Fumagalli A, Catasta G, Rossi L, Venneria E, Raguzzini A, Palomba L, Fogliano V, Maiani G. Bioavailability of strawberry antioxidants in human subjects. British Journal of Nutrition. 2010;104:1165-1173. DOI: 10.1017/S000711451000187X
  73. 73. Nurmi T, Mursu J, Heinonen M, Nurmi A, Hiltunen R, Voutilainen S. Metabolism of berry anthocyanins to phenolic acids in humans. Journal of Agricultural and Food Chemistry. 2009;57:2274-2281. DOI: 10.1021/jf8035116
  74. 74. Andres-Lacueva C, Shukitt-Hale B, Galli RL, Jauregui O, Lamuela-Raventos RM, Joseph JA. Anthocyanins in aged blueberry-fed rats are found centrally and may enhance memory. Nutrition Neuroscience. 2005;8:111-120. DOI: 10.1080/10284150500078117
  75. 75. El Mohsen MA, Marks J, Kuhnle G, Moore K, Debnam E, Kaila Srai S, Rice-Evans C, Spencer JP. Absorption, tissue distribution and excretion of pelargonidin and its metabolites following oral administration to rats. British Journal of Nutrition. 2006;95:51-58. DOI: 10.1079/BJN20051596
  76. 76. Mullen W, Larcombe S, Arnold K, Welchman H, Crozier A. Use of accurate mass full scan mass spectrometry for the analysis of anthocyanins in berries and berry-fed tissues. Journal of Agricultural and Food Chemistry. 2010;58:3910-3915. DOI: 10.1021/jf902267v
  77. 77. Glass CK, Saijo K, Winner B, Marchetto MC, Gage FH. Mechanisms underlying inflammation in neurodegeneration. Cell. 2010;19;140(6):918-34. DOI: 10.1016/j.cell.2010.02.016
  78. 78. Martinez FO, and Gordon S. The M1 and M2 paradigm of macrophage activation: time for reassessment. F1000Prime Rep. 2014; 6:13. doi: 10.12703/P6-13
  79. 79. MacMicking J, Xie QW, Nathan C. Nitric oxide and macrophage function. Annual Review of Immunology. 1997;15:323-350. DOI: 10.1146/annurev.immunol.15. 1.323
  80. 80. Arnold CE, Whyte CS, Gordon P, Barker RN, Rees AJ, Wilson HM. A critical role for suppressor of cytokine signalling 3 in promoting M1 macrophage activation and function in vitro and in vivo. Immunology. 2014;141:96-110. DOI: 10.1111/imm.12173
  81. 81. Brancato S K, Albina JE. Wound macrophages as key regulators of repair: origin, phenotype, and function. The American Journal of Pathology. 2011;178:19-25. DOI: 10.1016/j.ajpath.2010.08.003
  82. 82. Gordon S, Martinez FO. Alternative activation of macrophages: mechanism and functions. Immunity. 2010;32:593-604. DOI: 10.1016/j.immuni.2010. 05.007
  83. 83. Salehi B, Sharifi-Rad J, Cappellini F, Reiner Ž, Zorzan D, Imran M, Sener B, Kilic M, El-Shazly M, Fahmy NM, Al-Sayed E, Martorell M, Tonelli C, Petroni K, Docea AO, Calina D, Maroyi A. The Therapeutic Potential of Anthocyanins: Current Approaches Based on Their Molecular Mechanism of Action. Frontiers in Pharmacology. 2020;11:1300. DOI: 10.3389/fphar.2020.01300
  84. 84. Henriques JF, Serra D, Dinis TCP, Almeida LM. The Anti-Neuroinflammatory Role of Anthocyanins and Their Metabolites for the Prevention and Treatment of Brain Disorders. International Journal of Molecular Sciences. 2020.21(22):8653. DOI: 10.3390/ijms21228653
  85. 85. Lau FC, Bielinski DF, Joseph JA. Inhibitory effects of blueberry extract on the production of inflammatory mediators in lipopolysaccharide-activated BV2 microglia. Journal of Neuroscience Research. 2007;85:1010-1017. DOI: 10.1002/jnr.21205
  86. 86. Carey AN, Fisher DR, Rimando AM, Gomes SM, Bielinski DF, Shukitt-Hale B. Stilbenes and anthocyanins reduce stress signaling in BV-2 mouse microglia. Journal of Agricultural and Food Chemistry. 2013;61:5979-5986. DOI: 10.1021/jf400342g
  87. 87. Serra D, Henriques JF, Serra T, Bento Silva A, Bronze MR, Dinis TCP, Almeida LM. An anthocyanin-rich extract obtained from portuguese blueberries maintains its efficacy in reducing microglia-driven neuroinflammation after simulated digestion. Nutrients. 2020;12:3670-3692. DOI: 10.3390/nu12123670
  88. 88. Lau FC, Joseph JA, McDonald JE, Kalt W. Attenuation of iNOS and COX2 by blueberry polyphenols is mediated through the suppression of NF-B activation. Journal of Functional Foods. 2009;1:274-283. DOi: 10.1016/j.jff.2009.05.001
  89. 89. Ma H, Johnson SL, Liu W, DaSilva NA, Meschwitz S, Dain JA, Seeram NP. Evaluation of polyphenol anthocyanin-enriched extracts of blackberry, black raspberry, blueberry, cranberry, red raspberry, and strawberry for free radical scavenging, reactive carbonyl species trapping, anti-glycation, anti-β-amyloid aggregation, and microglial neuroprotective effects. International Journal of Molecular Sciences. 2018;19:461-780. DOI: 10.3390/ijms19020461
  90. 90. Simonyi A, Chen Z, Jiang J, Zong Y, Chuang DY, Gu Z, Lu CH, Fritsche KL, Greenlief CM, Rottinghaus GE, Thomas AL, Lubahn DB, Sun GY. Inhibition of microglial activation by elderberry extracts and its phenolic components. Life Sciences. 2015;128:30-38. DOI: 10.1016/j.lfs.2015.01.037
  91. 91. Poulose SM, Fisher DR, Larson J, Bielinski DF, Rimando AM, Carey AN, Schauss AG, Shukitt-Hale B. Anthocyanin-rich acai (Euterpe oleracea Mart.) fruit pulp fractions attenuate inflammatory stress signaling in mouse brain BV-2 microglial cells. Journal of Agricultural and Food Chemistry. 2012;60:1084-1093. DOI: 10.1021/jf203989k
  92. 92. Jeong JW, Lee WS, Shin SC, Kim GY, Choi BT, Choi YH. Anthocyanins downregulate lipopolysaccharide-induced inflammatory responses in BV2 microglial cells by suppressing the NF-kappaB and Akt/MAPKs signaling pathways. International Journal of Molecular Sciences. 2013;14:1502-1515. DOI: 10.3390/ijms14011502
  93. 93. Mu T, Guan Y, Chen T, Wang S, Li M, Chang AK, Yang Z, Bi X. Black raspberry anthocyanins protect BV2 microglia from LPS-induced inflammation through down-regulating NOX2/TXNIP/NLRP3 signaling. Journal of Berry Research. 2021;2:333-347. DOI: 10.3233/jbr-200692
  94. 94. Meireles M, Marques C, Norberto S, Santos P, Fernandes I, Mateus N, Faria A, Calhau C. Anthocyanin effects on microglia M1/M2 phenotype: Consequence on neuronal fractalkine expression. Behavioural Brain Research. 2016;305:223-228. DOI: 10.1016/j.bbr.2016.03.010
  95. 95. Kaewmool C, Udomruk S, Phitak T, Pothacharoen P, Kongtawelert P. Cyanidin-3-O-glucoside protects PC12 cells against neuronal apoptosis mediated by LPS-stimulated BV2 microglial activation. Neurotoxicity Research. 2020;37:111-125. DOI: 10.1007/s12640-019-00102-1
  96. 96. Zhao L, Chen S, Liu T, Wang X, Huang H, Liu W. Callistephin enhances the protective effects of isoflurane on microglial injury through downregulation of inflammation and apoptosis. Molecular Medicine Reports. 2019;20:802-812. DOI: 10.3892/mmr.2019.10282
  97. 97. Meireles M, Marques C, Norberto S, Fernandes I, Mateus N, Rendeiro C, Spencer JP, Faria A, Calhau C. The impact of chronic blackberry intake on the neuroinflammatory status of rats fed a standard or high-fat diet. Journal of Nutritional Biochemistry. 2015;26:1166-1173. DOI: 10.1016/j.jnutbio.2015.05.008
  98. 98. Rehman SU, Shah SA, Ali T, Chung JI, Kim MO. Anthocyanins reversed D-galactose-induced oxidative stress and neuroinflammation mediated cognitive impairment in adult rats. Molecular Neurobiology. 2017;54:255-271. DOI: 10.1007/s12035-015-9604-5
  99. 99. Khan MS, Ali T, Kim MW, Jo MH, Jo MG, Badshah H, Kim MO. Anthocyanins protect against LPS-induced oxidative stress-mediated neuroinflammation and neurodegeneration in the adult mouse cortex. Neurochemistry International. 2016;100:1-10. DOI: 10.1016/j.neuint.2016.08.005
  100. 100. Khan MS, Ali T, Kim MW, Jo MH, Chung JI, Kim MO. Anthocyanins improve hippocampus-dependent memory function and prevent neurodegeneration via JNK/Akt/GSK3β signaling in LPS-Treated adult mice. Molecular Neurobiology. 2019;56:671-687. DOI: 10.1007/s12035-018-1101-1
  101. 101. Carvalho FB, Gutierres JM, Bueno A, Agostinho P, Zago AM, Vieira J, Frühauf P, Cechella JL, Nogueira CW, Oliveira SM, Rizzi C, Spanevello RM, Duarte MMF, Duarte T, Dellagostin OA, Andrade CM. Anthocyanins control neuroinflammation and consequent memory dysfunction in mice exposed to lipopolysaccharide. Molecular Neurobiology. 2017;54:3350-3367. DOI: 10.1007/s12035-016-9900-8
  102. 102. Magni G, Marinelli A, Riccio D, Lecca D, Tonelli C, Abbracchio MP, Petroni K, Ceruti S. Purple corn extract as anti-allodynic treatment for trigeminal pain: Role of microglia. Frontiers in Cellular Neuroscience. 2018;12:378-390. DOI: 10.3389/fncel.2018.00378
  103. 103. Tikhonova MA, Shoeva OY, Tenditnik MV, Ovsyukova MV, Akopyan AA, Dubrovina NI, Amstislavskaya TG, Khlestkina EK. Evaluating the effects of grain of isogenic wheat lines differing in the content of anthocyanins in mouse models of neurodegenerative disorders. Nutrients. 2020;12:3877-3898. DOI: 10.3390/nu12123877
  104. 104. Li J , Zhao R , Jiang Y , Xu Y , Zhao H , Lyu X , Wu T. Bilberry anthocyanins improve neuroinflammation and cognitive dysfunction in APP/PSEN1 mice via the CD33/TREM2/TYROBP signaling pathway in microglia. Food & Function. 2020;11:1572-1584. DOI: 10.1039/c9fo02103e
  105. 105. Garcia G, Nanni S, Figueira I, Ivanov I, McDougall GJ, Stewart D, Ferreira RB, Pinto P, Silva RF, Brites D, Santos CN. Bioaccessible (poly)phenol metabolites from raspberry protect neural cells from oxidative stress and attenuate microglia activation. Food Chemistry. 2017;215:274-283. DOI: 10.1016/j.foodchem.2016.07.128
  106. 106. Wang HY, Wang H, Wang JH, Wang Q, Ma QF, Chen YY. Protocatechuic acid inhibits inflammatory responses in LPS-Stimulated BV2 microglia via NF-κB and MAPKs signaling pathways. Neurochemical Research. 2015;40:1655-1660. DOI: 10.1007/s11064-015-1646-6
  107. 107. Kaewmool C, Kongtawelert P, Phitak T, Pothacharoen P, Udomruk S. Protocatechuic acid inhibits inflammatory responses in LPS-activated BV2 microglia via regulating SIRT1/NF-κB pathway contributed to the suppression of microglial activation-induced PC12 cell apoptosis. Journal of Neuroimmunology. 2020;341:577164-577175. DOI: 10.1016/j.jneuroim.2020.577164
  108. 108. Winter AN, Brenner MC, Punessen N, Snodgrass M, Byars C, Arora Y, Linseman DA. Comparison of the neuroprotective and anti-inflammatory effects of the anthocyanin metabolites, protocatechuic acid and 4-hydroxybenzoic acid. Oxidative Medicine and Cellular Longevity. 2017;2017:6297080-6297094. DOI: 10.1155/2017/6297080
  109. 109. Koga M, Nakagawa S, Kato A, Kusumi I. Caffeic acid reduces oxidative stress and microglial activation in the mouse hippocampus. Tissue and Cell. 2019;60:14-20. DOI: 10.1016/j.tice.2019.07.006
  110. 110. Cho JY, Kim HS, Kim DH, Yan JJ, Suh HW, Song DK. Inhibitory effects of long-term administration of ferulic acid on astrocyte activation induced by intracerebroventricular injection of beta-amyloid peptide (1-42) in mice. Prog Neuropsychopharmacol Biological Psychiatry. 2005;29:901-907. DOI: 10.1016/j.pnpbp.2005.04.022
  111. 111. Rehman SU, Ali T, Alam SI, Ullah R, Zeb A, Lee KW, Rutten BPF, Kim MO. Ferulic acid rescues LPS-induced neurotoxicity via modulation of the TLR4 receptor in the mouse hippocampus. Molecular Neurobiology. 2019;56:2774-2790. DOI: 10.1007/s12035-018-1280-9
  112. 112. Ullah R, Ikram M, Park TJ, Ahmad R, Saeed K, Alam SI, Rehman IU, Khan A, Khan I, Jo MG, Kim MO. Vanillic acid, a bioactive phenolic compound, counteracts LPS-induced neurotoxicity by regulating c-Jun N-terminal kinase in mouse brain. International Journal of Molecular Sciences. 2020;22:361-382. DOI: 10.3390/ijms22010361
  113. 113. Kim MJ, Seong AR, Yoo JY, Jin CH, Lee YH, Kim YJ, Lee J, Jun WJ, Yoon HG. Gallic acid, a histone acetyltransferase inhibitor, suppresses β-amyloid neurotoxicity by inhibiting microglial-mediated neuroinflammation. Molecular Nutrition & Food Research. 2011;55:1798-1808. DOI: 10.1002/mnfr.201100262
  114. 114. Liu YL, Hsu CC, Huang HJ, Chang CJ, Sun SH, Lin AM. Gallic acid attenuated LPS-induced neuroinflammation: Protein aggregation and necroptosis. Molecular Neurobiology. 2020;57:96-104. DOI: 10.1007/s12035-019-01759-7
  115. 115. Straßmann S, Brehmer T, Passon M, Schieber A. Methylation of cyanidin-3-O-glucoside with dimethyl carbonate. Molecules. 2021;26:1342. DOI: 10.3390/molecules26051342
  116. 116. Soidinsalo O, Wähälä K. Synthesis of phytoestrogenic isoflavonoid disulfates. Steroids. 2004;69:613-616. DOI: 10.1016/j.steroids.2004.03.01
  117. 117. Straßmann S, Passon M, Schieber A. Chemical hemisynthesis of sulfated cyanidin-3-O-glucoside and cyanidin metabolites. Molecules, 2021;26:2146. DOI: 10.3390/molecules26082146
  118. 118. Cruz L, Basílio N, Mateus N, Pina F, de Freitas V. Characterization of kinetic and thermodynamic parameters of cyanidin-3-glucoside methyl and glucuronyl metabolite conjugates. The Journal of Physical Chemistry B. 2015; 119:2010-2018 DOI:10.1021/jp511537e

Written By

Ruth Hornedo-Ortega, Zuriñe Rasines-Perea, Ana B. Cerezo, Pierre-Louis Teissedre and Michael Jourdes

Submitted: 04 August 2021 Reviewed: 15 August 2021 Published: 20 December 2021