Polyphenols are a structurally diverse class of naturally occurring secondary metabolites synthesized exclusively by plants, characterized by the presence of multiple phenol units—specifically, aromatic rings with at least two hydroxyl groups attached.[1] These compounds play essential roles in plant physiology, including defense against ultravioletradiation, pathogens, and oxidative stress.[2] Over 8,000 distinct polyphenols have been identified, making them one of the most abundant groups of bioactive substances in the humandiet.[3]Polyphenols are broadly classified based on their chemical structure, primarily into two main categories: flavonoids (the largest subgroup, including flavonols, flavones, flavanols, flavanones, isoflavones, and anthocyanins) and non-flavonoids (encompassing phenolic acids, stilbenes, lignans, and tannins).[4] This classification reflects variations in the number of phenol rings and the linking structural elements, such as heterocyclic rings in flavonoids or simpler chains in phenolic acids.[2] Their biosynthesis occurs via the phenylpropanoid pathway in plants, leading to a wide array of derivatives with differing solubility and stability.[4]In human nutrition, polyphenols are primarily obtained from plant-based foods and beverages, with rich sources including fruits (e.g., berries, apples, grapes), vegetables (e.g., onions, broccoli), whole grains, nuts, tea, coffee, red wine, and cocoa.[5] Dietary intake varies widely but averages 800–1,000 mg per day in typical Western diets, though higher in Mediterranean-style eating patterns.[3]Bioavailability is generally low (often <5%) due to poor absorption in the gut, rapid metabolism, and microbial transformation by gut microbiota, which can produce bioactive metabolites.[6]The health benefits of polyphenols stem largely from their potent antioxidant and anti-inflammatory properties, which help neutralize free radicals, modulate cellular signaling pathways, and reduce oxidative stress.[2] Epidemiological and clinical studies link regular polyphenol consumption to reduced risk of chronic diseases, including cardiovascular disorders (via improved endothelial function and lowered blood pressure), type 2 diabetes (through enhanced insulin sensitivity), certain cancers (by inhibiting tumor growth), and neurodegenerative conditions (such as Alzheimer's, via neuroprotection).[3] However, excessive intake from supplements may pose risks, such as interference with iron absorption or potential pro-oxidant effects at high doses.[3]
Terminology
Etymology
The term "polyphenol" originates from the Ancient Greek word polús (πολύς), meaning "many" or "much," combined with "phenol," referring to a chemical structure featuring a hydroxyl group attached to an aromatic hydrocarbon ring. This nomenclature reflects the compounds' characteristic multiple phenolic units, distinguishing them from simple phenols. The term first appeared in scientific literature in 1894, in the context of chemical studies on phenolic compounds.[7] It emerged more prominently in early 20th-century plant chemistry, where researchers like Maximilian Nierenstein investigated complex phenolic substances such as tannins and catechins in plants, laying foundational work on their structures and properties without yet using the specific term "polyphenol" in its modern botanical sense.The modern application of the term "polyphenol" to describe a broad class of plant-derived polyphenolic materials previously known collectively as "vegetable tannins," emphasizing their role in leather tanning and plant defense, was proposed in 1957 by industrial chemist Theodore White.[8] This marked a shift from narrower terms like "tannins" (referring to astringent proteins-precipitating phenolics) and "catechins" (specific flavonoid monomers) toward a more inclusive category encompassing polymers with multiple phenolic rings. White's definition focused on experimental observations of their physico-chemical behavior, such as solubility and reactivity.In 1962, E. C. Bate-Smith and Tony Swain refined the term in their seminal chapter on flavonoid compounds, defining polyphenols as "water-soluble phenolic compounds having molecular weights from 500 to 3000, possessing the typical phenolic nucleus and other ring systems."[9] This White–Bate-Smith–Swain–Haslam (WBSSH) framework, later expanded by Edwin Haslam in 1966 to include a molecular weight range up to 4000 Da and 12–16 phenolic hydroxy groups per 1000 Da, solidified the term's usage in scientific literature, particularly for flavonoid polymers with tanning properties. The 1960s–1970s saw further evolution with advances in chromatographic techniques, enabling separation and identification of diverse polyphenols beyond tannins, broadening the term to include non-tannin phenolics in fruits and beverages. Concurrently, the French term "polyphénols" gained traction in wine chemistry, influenced by studies on grape and oak-derived compounds contributing to color and astringency.[10]
Definition
Polyphenols are secondary metabolites produced primarily by plants, defined as organic compounds featuring multiple phenol units—aromatic rings bearing one or more hydroxyl groups—and involved in defense mechanisms against environmental stresses such as ultraviolet radiation and pathogens.[10] This core definition excludes primary metabolites like lignins, which serve structural roles in plant cell walls rather than secondary biosynthetic functions.[11]Definitional ambiguities arise from varying interpretations of structural criteria and scope, including the inclusion of tannins (polyphenolic polymers with astringent properties) and lignans (dimers of phenylpropane units), while excluding synthetic analogs that mimic natural structures but lack biological origin.[11] These debates impact standardization in research, with some classifications emphasizing biosynthetic pathways over mere structural multiplicity.The Quideau definition (2011) addresses these issues by focusing on phenolic oxidation products derived from shikimate-derived precursors, prioritizing reactivity and natural assembly: the term "polyphenol" applies exclusively to plant-origin compounds with at least two distinct phenolic rings linked by C–C, O–C, or glycosidic bonds (excluding ester linkages), encompassing up to six rings and including oligomeric/polymeric forms such as proanthocyanidins, ellagitannins, and gallotannins.[12] In contrast, the historical WBSSH framework (White 1957, Bate-Smith and Swain 1962, Haslam 1966) characterizes polyphenols as water-soluble, non-glycemic compounds with molecular weights of 500–4000 Da, multiple phenolic rings, and the ability to precipitate proteins, distinguishing them from broader, less precise uses in nutritional contexts and excluding carbohydrate-linked phenolics.[10]Reviews from the 2020s, including those synthesizing these perspectives, advocate for unified criteria to facilitate reproducible research, resolving ambiguities around boundary cases like high-molecular-weight tannins while maintaining emphasis on plant-derived, hydroxyl-rich aromatics.[11]
Chemistry
Molecular Structure
Polyphenols are a class of organic compounds characterized by the presence of multiple phenol structural units, consisting of one or more aromatic rings (typically benzene or heterocyclic rings) with attached hydroxyl (-OH) groups, often in ortho or para positions relative to each other. These structures can form through polymerization via carbon-carbon (C-C) or ether (C-O) linkages, enabling a range of monomeric to polymeric forms.[1][13]Key structural motifs in polyphenols include the phenolic hydroxyl groups, which facilitate hydrogen bonding and contribute to their reactivity and solubility, as well as extended conjugated π-electron systems within the aromatic rings that impart characteristic UV-visible absorption spectra around 280 nm. A representative building block is catechol (1,2-dihydroxybenzene), featuring two adjacent hydroxyl groups on a benzene ring, which serves as a core motif in many polyphenol subclasses such as flavonoids and tannins.[1][14]The structural diversity of polyphenols spans from simple monomers, such as phenolic acids (e.g., caffeic acid with molecular weight around 180 Da), to complex oligomers and polymers like proanthocyanidins, which can consist of up to 50 flavan-3-ol units linked by C-C bonds, resulting in molecular weights exceeding 10,000 Da. This diversity arises from variations in ring substitution, linkage types, and degrees of polymerization, with some structures incorporating chiral centers—such as in the C-ring of flavonoids—leading to stereoisomers that influence biological activity.[13][1]Physically, polyphenols exhibit polarity due to their multiple hydroxyl groups, enhancing solubility in polar solvents like water and ethanol while reducing solubility in non-polar solvents; this property is crucial for their extraction and bioavailability. Their molecular weight typically ranges from 100 Da for basic monomers to 10,000 Da or more for high-molecular-weight polymers, affecting their diffusion and interaction with biological macromolecules.[14][1]
Classification
Polyphenols are classified primarily based on their chemical structures, particularly the arrangement of phenolic rings and the carbon skeleton connecting them, as well as substitution patterns such as hydroxylation and glycosylation.[15] This structural taxonomy originates from biosynthetic pathways but emphasizes molecular architecture for categorization, with updates reflected in databases like Phenol-Explorer, which organizes compounds into hierarchical classes and subclasses as of the 2020s.[16] Over 8,000 polyphenol structures have been identified to date, predominantly from plant sources, though microbial and synthetic analogs exist in limited contexts.[17]The largest class is flavonoids, accounting for approximately 60% of all known polyphenols, characterized by a core C6-C3-C6 carbon skeleton consisting of two phenyl rings (A and B) linked by a heterocyclic pyran ring (C).[18] Subgroups include flavonols (e.g., quercetin, found in onions and apples), flavones (e.g., apigenin, in parsley and chamomile), and others like anthocyanins and isoflavonoids, differentiated by oxidation levels and substitutions on the rings.[15] Phenolic acids form another major group, divided into hydroxybenzoic acids (e.g., gallic acid) and hydroxycinnamic acids (e.g., caffeic and ferulic acids), featuring a single phenolic ring with carboxylic acid side chains.[19]Stilbenes, with a simpler C6-C2-C6 skeleton, include resveratrol (prominent in grapes and red wine), while lignans possess a C6-C3-C3-C6 framework derived from two phenylpropane units, such as secoisolariciresinol in flaxseeds.[20] Tannins represent complex polyphenols, subclassified into condensed tannins (proanthocyanidins, polymers of flavonoids) and hydrolyzable tannins (gallotannins and ellagitannins, esters of phenolic acids with sugars).[15] Additional groups encompass coumarins (benzopyrones like umbelliferone) and xanthones (tricyclic structures such as mangostin), often grouped under "other polyphenols" in databases.[16]While the focus remains on plant-derived polyphenols due to their prevalence in nature and dietary relevance, non-plant sources include microbial metabolites like those produced by gut bacteria from dietary precursors, and synthetic polyphenols engineered for industrial applications, though these are not central to natural classification schemes.[20]
Reactivity
Polyphenols exhibit reactivity primarily through their phenolic hydroxyl (-OH) groups, which enable them to act as antioxidants by scavenging free radicals via hydrogen atom transfer (HAT). In this mechanism, a phenolic compound (ArOH) donates a hydrogen atom to a reactive oxygen species (ROS) such as an alkoxyl radical (RO•), forming a relatively stable phenoxyl radical (ArO•) and the corresponding alcohol (ROH), as represented by the equation:\ce{ArOH + RO^\bullet -> ArO^\bullet + ROH}[4][21][15]This process is facilitated by the resonance stabilization of the resulting ArO•, particularly in polyphenols with multiple phenolic rings. Additionally, polyphenols contribute to antioxidant activity through metal chelation, where their ortho-dihydroxy (catechol) or adjacent hydroxyl groups bind transition metals like iron (Fe²⁺/Fe³⁺) or copper, preventing Fenton reactions that generate highly reactive hydroxyl radicals (•OH). This chelation alters the redox potential of the metals, inhibiting their catalytic role in oxidative damage.[22][23][24]Oxidation of polyphenols can lead to polymerization, often initiated by auto-oxidation under alkaline conditions, where molecular oxygen oxidizes the phenolic rings to form reactive o-quinones. These quinones are electrophilic intermediates that can further react with nucleophiles, such as thiols or amines, resulting in cross-linking and pigment formation. Enzymatic oxidation, mediated by polyphenol oxidase (PPO), accelerates this process in the presence of oxygen, converting monophenols or o-diphenols to quinones and causing enzymatic browning in plant tissues. PPO, a copper-containing enzyme, catalyzes the hydroxylation of phenols and their subsequent oxidation, leading to melanin-like polymers.[25][26][27][28][29][30]The stability of polyphenols is influenced by environmental factors, including pH, light, and oxygen exposure, which can promote degradation via oxidation or hydrolysis. In alkaline pH (>7), auto-oxidation rates increase dramatically, leading to quinone formation and loss of bioactivity, whereas acidic conditions enhance stability. Exposure to light, particularly UV, and oxygen sensitizes polyphenols to photodegradation and peroxidation, respectively, reducing their half-life in solutions or foods. Glycosylation, the attachment of sugar moieties to phenolic hydroxyls, improves stability by sterically hindering oxidation sites and enhancing solubility, thereby protecting against enzymatic and non-enzymatic degradation. Reactivity varies among structural classes; for instance, catechols (ortho-dihydroxybenzenes) are more prone to oxidation and radical scavenging than resorcinols (meta-dihydroxybenzenes) due to lower oxidation potentials and greater electron-donating ability of the adjacent -OH groups.[31][32][33][34][35][34][36][37][38]Polyphenols also interact with biomolecules, binding to proteins through hydrogen bonding and hydrophobic interactions, which precipitate salivary proline-rich proteins and contribute to astringency in foods and beverages. This binding forms complexes that alter protein conformation and sensory perception. Furthermore, polyphenols inhibit lipid peroxidation by scavenging peroxyl radicals (LOO•) in cell membranes and chelating pro-oxidant metals, thereby preventing chain reactions that damage unsaturated fatty acids and maintaining membrane integrity.[39][40][41][42][43][44]
Analytical Methods
Extraction
Polyphenols are typically isolated from plant materials such as fruits, vegetables, and herbs using a variety of extraction techniques designed to maximize yield while preserving their structural integrity.[45]Conventional solventextraction remains the most widely used method, employing polar solvents like methanol or ethanol, often in 50-80% aqueous solutions, at temperatures of 50-80°C to enhance solubility and diffusion.[46]Maceration involves soaking the plant matrix in solvent for extended periods, typically 24-50 hours, allowing passive diffusion but requiring larger solvent volumes and longer times compared to dynamic methods.[45] In contrast, Soxhlet extraction uses a continuous percolation process, recycling hot solvent through the sample for 4-8 hours, which improves efficiency and reduces solvent use, though it may degrade heat-sensitive compounds.[46] Yields from these techniques generally range from 10-50% of the dry matrix weight, varying with the plant source and solvent polarity; for example, 80% ethanolmaceration of citrus peels achieves up to 18.5% yield.[47]Advanced methods have gained prominence for their efficiency and reduced environmental impact. Ultrasound-assisted extraction (UAE) applies ultrasonic waves at 20-40 kHz to generate cavitation bubbles that disrupt cell walls, accelerating solvent penetration and typically completing in 15-60 minutes with yields 20-50% higher than conventional approaches.[48]Microwave-assisted extraction (MAE) uses electromagnetic waves at 300-900 W to rapidly heat the solvent-sample mixture, often in 1-10 minutes, enhancing mass transfer and achieving comparable or superior polyphenol recovery, such as 30% more phenolics from apple pomace than maceration. Supercritical CO2 extraction targets less polar polyphenols using CO2 above its critical point (31°C, 73.8 bar), often with ethanol as a co-solvent, and is ideal for heat-labile compounds but less effective for highly polar ones without modifiers.[49]Green solvents like ionic liquids, such as 1-butyl-3-methylimidazolium-based variants, offer tunable polarity and low volatility, enabling higher selectivity and yields in microwave or ultrasound combinations compared to traditional organic solvents.[50]Optimization strategies address polyphenol variability and stability. Acidic pH adjustment (e.g., to 3-5) during extraction stabilizes sensitive structures like catechins by minimizing oxidation and hydrolysis, improving recovery by up to 25% in green tea extracts.[51] Enzyme pretreatment with cellulases, pectinases, or tannases hydrolyzes cell walls and releases bound polyphenols, boosting yields by 15-40% in materials like grape pomace, though it requires controlled conditions to avoid over-degradation.[52] Challenges arise with thermolabile polyphenols, such as quercetin glycosides, which degrade above 60°C, necessitating lower-temperature methods like UAE or enzyme-assisted approaches to retain bioactivity.[53]Post-extraction processing ensures purity and concentration. Crude extracts are first filtered through Whatman paper or microfiltration to remove particulate matter, followed by solvent removal via rotary evaporation under reduced pressure (40-60°C) to yield a concentrated residue, often 10-20 times the original volume reduction without significant polyphenol loss.[54]
Detection Techniques
Detection of polyphenols typically follows extraction procedures to isolate these compounds from complex matrices such as plant tissues or food samples.[55]Spectroscopic methods provide foundational tools for initial identification based on characteristic absorption or resonance patterns. Ultraviolet-visible (UV-Vis) spectroscopy exploits the conjugated aromatic systems in polyphenols, which exhibit strong absorption bands generally between 250 and 370 nm, enabling rapid screening of phenolic compounds.[55] Fourier-transform infrared (FTIR) spectroscopy identifies functional groups through vibrational modes, such as the broad O-H stretching band at 3200–3600 cm⁻¹ indicative of phenolic hydroxyl groups.[56] Nuclear magnetic resonance (NMR) spectroscopy, particularly ¹H and ¹³C NMR, elucidates structural features like aromatic ring protons and carbon environments, offering detailed confirmation of polyphenol identity in purified samples.[57]Chromatographic techniques separate polyphenols for subsequent identification, often using reversed-phase columns. High-performance liquid chromatography (HPLC) with C18 stationary phases and gradient elution systems (e.g., water-acetonitrile with formic acid) resolves diverse phenolic classes based on polarity and hydrophobicity.[58] Gas chromatography-mass spectrometry (GC-MS) suits volatile or derivatized polyphenols, providing retention time and mass spectral data for compound matching.[59] Hyphenated methods like liquid chromatography-tandem mass spectrometry (LC-MS/MS) enhance specificity through fragmentation patterns, where precursor ions are selected and collided to yield diagnostic product ions for structural elucidation.[60]Emerging instrumental approaches improve sensitivity and throughput for polyphenol characterization. High-resolution mass spectrometry (HRMS), such as Orbitrap systems, determines exact masses (typically <5 ppm error) to confirm molecular formulas and aid in untargeted profiling of complex mixtures.[61] Electrochemical sensors, often based on modified electrodes (e.g., carbon nanomaterials), detect polyphenols via oxidation-reduction currents, enabling rapid, portable screening in real-time applications like food quality assessment.[62]Standard practices ensure method reliability, with gallic acid commonly used as a reference for calibrating spectroscopic detections of phenolics. Validation adheres to AOAC International guidelines, emphasizing parameters like limit of detection, linearity, and specificity for accurate polyphenol identification.[63]
Quantification
Quantification of polyphenols typically involves a range of analytical assays that measure total content or antioxidant capacity, as well as chromatographic techniques for individual compounds.[60]The Folin-Ciocalteu (FC) assay is a widely used colorimetric method for estimating total phenolic content (TPC), where the reagent reacts with phenolic compounds in alkaline medium to form a blue molybdenum-tungsten complex, measured spectrophotometrically at 765 nm.[64] Gallic acid is commonly employed as the standard for expressing results in gallic acid equivalents (GAE).[64] However, the assay's non-specificity leads to overestimation due to interference from reducing agents such as ascorbic acid, sugars, and proteins.[64]For assessing antioxidant capacity linked to polyphenols, the DPPH and ABTS assays are frequently applied; the DPPH method evaluates the ability of polyphenols to scavenge 2,2-diphenyl-1-picrylhydrazyl radicals, resulting in decolorization quantified by IC50 values (the concentration inhibiting 50% of the radical), while ABTS measures scavenging of 2,2'-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) radicals with similar IC50 metrics.[15] These assays are particularly suitable for lipophilic polyphenols like quercetin and resveratrol.[15]High-performance liquid chromatography (HPLC) enables precise quantification of individual polyphenols through separation and detection, typically using diode array or mass spectrometry detectors, with calibration curves constructed from standards to determine concentrations, alongside limits of detection (LOD) and quantification (LOQ) for method sensitivity.[60]The total phenolic content is calculated using the linear equation from the gallic acidcalibration curve:\text{TPC} = \frac{\text{absorbance} - \text{intercept}}{\text{slope}} \times \text{dilution factor}expressed in units of mg GAE/g sample.[64]Recent advances include near-infrared (NIR) spectroscopy for non-destructive quantification of polyphenols in foods, where chemometric models such as partial least squares regression achieve high predictive accuracy with coefficients of determination (R²) exceeding 0.9 in 2020s applications for items like olive oil and tea.[65]
Natural Occurrence
In Plants
Polyphenols are ubiquitous secondary metabolites in the plant kingdom, present in virtually all vascular plants, where they serve essential physiological and ecological functions. They are particularly abundant in protective tissues such as bark, leaves, and fruits, often comprising a significant portion of the dry weight in these organs. For instance, in tea leaves (Camellia sinensis), polyphenols constitute 25–35% of the dry weight, primarily as catechins that contribute to the plant's resilience.[66][67] These compounds originate from phenylpropanoid biosynthetic pathways and accumulate variably across plant species and tissues to support adaptation and survival.[68]In plants, polyphenols play critical ecological roles, primarily in defense mechanisms against biotic and abiotic stresses. They deter herbivores and pathogens by acting as toxins or feeding inhibitors, with tannins and flavonoids binding to proteins to reduce digestibility in potential predators. Flavonoids also function as UV protectants, absorbing harmful radiation and preventing DNA damage in exposed tissues like leaves. Additionally, certain phenolic acids mediate allelopathy by inhibiting the growth of neighboring plants through root exudates that disrupt germination or nutrient uptake. Polyphenols further facilitate symbiotic interactions, such as signaling in legume-rhizobia nitrogen fixation or arbuscular mycorrhizal associations, where they regulate microbial colonization in roots.[69][70][71][72]The distribution and concentration of polyphenols exhibit species-specific variation, influenced by genetic factors and environmental cues. Grapes (Vitis vinifera), for example, are notably rich in resveratrol, a stilbene polyphenol concentrated in skins and seeds, which enhances fungal resistance in this species. Environmental stresses, such as drought, can significantly elevate polyphenol levels as a protective response; studies show increases of 2–3-fold in total polyphenols and flavonoids under progressive water deficit, aiding in osmotic adjustment and antioxidant defense. This upregulation underscores polyphenols' role in stress acclimation across diverse plant taxa.[73][74]
In Foods and Beverages
Polyphenols are abundant in various foods and beverages derived from plant sources, serving as key contributors to their nutritional profiles. Berries represent one of the richest dietary sources, with total polyphenol contents ranging from approximately 250 to 1,200 mg per 100 g fresh weight, primarily in the form of anthocyanins. For instance, elderberries contain about 1,191 mg/100 g, blueberries around 525 mg/100 g, and blackcurrants up to 560 mg/100 g, making them standout examples among fruits. In beverages, tea provides significant catechins, with a typical cup of green tea delivering 100–200 mg of epigallocatechin gallate (EGCG) alone, and total catechins often exceeding 200 mg per 240 mL serving. Coffee similarly contributes chlorogenic acids, with light roast varieties offering up to 188 mg per cup, while total phenolic content in brewed coffee averages around 200–500 mg per serving. Red wine stands out for its diverse polyphenols, with total contents ranging from 1,800 to 4,000 mg/L, including trace amounts of resveratrol (typically 0.1–10 mg/L, higher in varieties like Pinot noir). These values are documented in comprehensive databases such as the USDA Database for the Flavonoid Content of Selected Foods (Release 3.3, 2018), integrated into the Food and Nutrient Database for Dietary Studies (FNDDS) for NHANES cycles up to 2021–2023 as of 2024.[75]Food processing significantly alters polyphenol levels and bioaccessibility. Fermentation and oxidation alter polyphenol profiles in products like black tea, where enzymatic oxidation converts catechins to theaflavins and thearubigins, potentially improving bioavailability of these metabolites compared to unprocessed forms, though total catechin content decreases relative to green tea. Enzymatic activities during processing transform complex structures into more absorbable compounds in some cases. In contrast, thermal processing such as cooking reduces heat-sensitive polyphenols by 30–50%, with losses observed in anthocyanins from berries and phenolic acids in vegetables due to degradation and leaching into cooking water; for example, boiling can diminish chlorogenic acid in potatoes by similar margins. These changes are influenced by factors like temperature, duration, and medium volume, with minimal processing (e.g., steaming) preserving more compounds than prolonged high-heat methods.Average daily polyphenol intake in Western diets is estimated at around 1 g per person, varying by 300–1,200 mg based on consumption patterns, with coffee, tea, and fruits as primary contributors. This figure aligns with epidemiological data from European and North American cohorts, where flavonoids from beverages account for over 50% of total intake. The USDA Flavonoid Database provides analytical values for hundreds of foods, supporting precise intake estimations in dietary surveys like What We Eat in America (NHANES).Beyond nutrition, polyphenols influence sensory attributes in foods and beverages, imparting color through anthocyanins (e.g., red hues in berries and wine) and contributing to bitterness and astringency via catechins and proanthocyanidins. In tea and wine, these compounds elicit a puckering astringency and bitter aftertaste, which enhance perceived quality but can vary with concentration and processing; for instance, higher polyphenol levels in dark-roast coffee amplify bitterness from chlorogenic acids.
Biosynthesis and Metabolism
Biosynthesis Pathways
Polyphenols in plants are synthesized primarily through the integration of the shikimate pathway and the phenylpropanoid pathway, with contributions from the acetate-malonate pathway for certain branches. The shikimate pathway initiates with the condensation of phosphoenolpyruvate (derived from glycolysis) and erythrose-4-phosphate (from the pentose phosphate pathway) to form 3-deoxy-D-arabino-heptulosonate 7-phosphate (DAHP), catalyzed by DAHPsynthase; this seven-step process, involving enzymes such as 3-dehydroquinate synthase and shikimate kinase, culminates in chorismate formation. Chorismate is then channeled into the arogenate pathway to produce phenylalanine, the key amino acid precursor for most phenolic compounds.[76]From phenylalanine, the phenylpropanoid pathway begins with the deamination by phenylalanine ammonia-lyase (PAL) to yield trans-cinnamic acid, which is subsequently hydroxylated by cinnamate 4-hydroxylase (C4H) to p-coumaric acid. This core pathway branches into various polyphenol classes; for instance, p-coumaroyl-CoA is condensed with three molecules of malonyl-CoA by chalcone synthase (CHS), the first committed enzyme of the flavonoid branch, to form chalcone, which isomerizes to flavanone and further diversifies into flavonols, anthocyanins, and proanthocyanidins via enzymes like flavanone 3-hydroxylase and dihydroflavonol 4-reductase. Other branches lead to lignins, coumarins, and stilbenes, such as resveratrol produced by stilbene synthase in response to stress.[77][76]The biosynthesis of polyphenols is tightly regulated at transcriptional, post-transcriptional, and enzymatic levels to respond to developmental cues and environmental stresses. MYB transcription factors, particularly R2R3-MYB proteins like AtMYB12 in Arabidopsis, activate promoters of key genes such as PAL, CHS, and chalcone isomerase, thereby coordinating flux through the pathway for flavonoid accumulation. Elicitors like jasmonic acid (JA) induce polyphenol production by activating JA-responsive signaling cascades that upregulate MYB factors and downstream enzymes, enhancing defense against pathogens and herbivores. Genetic engineering has demonstrated the potential for pathway manipulation; for example, CRISPR/Cas9-mediated knockout of the Vitis davidii CHS2 gene redirects metabolic flux from flavonoids to stilbenoids, resulting in elevated resveratrol levels in cell cultures.[78]Evolutionarily, polyphenol biosynthesis traces back to ancient origins, with PAL-like genes present in algae as early as two copies per genome, indicating a primitive capacity for phenylpropanoid metabolism in non-vascular organisms. Gene duplications and expansions during the transition to land plants and diversification in angiosperms led to increased PAL family members—up to dozens in eudicots—enabling specialized polyphenol production for structural reinforcement, UV protection, and biotic interactions in complex terrestrial environments.[79][80]
Metabolism in Organisms
Polyphenols are primarily absorbed in the small intestine, where low molecular weight compounds (<500 Da), such as aglycones, undergo passive diffusion across the intestinal epithelium.14380-4/fulltext) Higher molecular weight or glycosylated forms, like glucosides, require deconjugation by gut microbiota or intestinal enzymes (e.g., lactase-phlorizin hydrolase) to release absorbable aglycones before uptake.[81] Only about 5-10% of ingested polyphenols are absorbed in the small intestine, with the remainder reaching the colon for further microbial transformation.[82]Following absorption, polyphenols undergo phase I and II metabolism primarily in the intestinal mucosa and liver. Phase I reactions involve cytochrome P450 (CYP450) enzymes, such as CYP1A2 and CYP3A4, which perform oxidation to introduce or modify hydroxyl groups, enhancing reactivity for subsequent conjugation.[83] Phase II metabolism, more dominant for polyphenols, includes glucuronidation, sulfation, and methylation, mediated by UDP-glucuronosyltransferases (UGTs), sulfotransferases (SULTs), and catechol-O-methyltransferases (COMTs), respectively, to increase water solubility.[84] For example, quercetin is rapidly conjugated in the liver to form quercetin-3-glucuronide via UGT1A1:\text{Quercetin} + \text{UDPGA} \xrightarrow{\text{UGT1A1}} \text{Quercetin-3-glucuronide} + \text{UDP}These conjugates predominate in plasma, with minimal free aglycone detected.[84]Metabolized polyphenols are excreted mainly via urine and feces, with phase II conjugates facilitating renal clearance and biliary elimination.[85] Plasma half-lives vary from 1 to 24 hours, with a median of approximately 2.8 hours across polyphenol metabolites, influenced by conjugation efficiency and enterohepatic recirculation.[86] Gut microbiota contribute to inter-individual variation; for instance, only 30-50% of individuals convert the isoflavone daidzein to equol due to specific bacterial strains like those in the Adlercreutzia genus.[87] Recent 2024 studies highlight enterohepatic recirculation, where biliary-excreted conjugates are reabsorbed in the intestine after deconjugation, prolonging systemic exposure and potentially enhancing bioefficacy.[88]
Applications
Nutritional and Health Uses
Polyphenols lack an established Recommended Dietary Allowance (RDA) due to their diverse sources and variable bioavailability, but observational and interventional studies suggest that daily intakes of 500–1000 mg may support antioxidant and anti-inflammatory benefits.[89][90] This range aligns with typical consumption in polyphenol-rich diets, such as the Mediterranean diet, where extra virgin olive oil contributes substantially—up to 11% of total polyphenol intake—through compounds like hydroxytyrosol and oleuropein.[91] Foods like fruits, vegetables, tea, and coffee naturally provide these levels, with coffee alone delivering 500–1000 mg per day for regular drinkers.[90]Polyphenol supplements, often derived from concentrated plant extracts, are widely used to achieve targeted intakes beyond diet alone. Grape seed extract, standardized for proanthocyanidins, is a common example, with typical dosages ranging from 100–300 mg daily to support vascular and skin health.[92]Bioavailability remains a challenge for many polyphenols due to rapid metabolism, but enhancers like piperine from black pepper can increase absorption; for instance, piperine boosts the bioavailability of tea polyphenols such as epigallocatechin gallate by inhibiting glucuronidation in animal models.[93]Functional foods fortified with polyphenols have gained popularity as convenient delivery vehicles, including enriched juices from sources like pomegranate or grapes, which retain high levels of ellagitannins and anthocyanins for enhanced nutritional profiles.[94] The global market for polyphenols, encompassing these functional food applications, was valued at approximately $1.4 billion in 2024 and is projected to reach $2.1 billion by 2030, driven by consumer demand for health-promoting products.[95]Historically, polyphenols have featured prominently in traditional medicine, with green tea—rich in catechins—used in ancient Asian practices for promoting longevity, mental clarity, and digestive health since at least the Tang Dynasty in China.[96] Such uses underscore the long-recognized role of polyphenol-containing plants in preventive health strategies across cultures.
Industrial and Environmental Applications
In the food industry, polyphenols serve as natural preservatives to extend shelf life by inhibiting lipid oxidation and microbial growth. Rosemary extracts, rich in phenolic compounds like carnosic acid and carnosol, effectively prevent oxidation in meat products and baked goods, offering a synthetic-free alternative to butylated hydroxytoluene (BHT).[97] Anthocyanins, another class of polyphenols, function as natural colorants, imparting stable red, purple, and blue hues to beverages, confectionery, and dairy products; sources such as grape skin and elderberry extracts are commercially preferred for their pH-dependent stability and regulatory approval as food additives.[98][99]In green chemistry applications, polyphenols contribute to sustainable material development, particularly through tannins derived from plant sources like oak bark and grape pomace. These condensed tannins are incorporated into biodegradable polymers, such as polyhydroxyalkanoate/tannin composites, enhancing mechanical strength, barrier properties, and antioxidant activity for eco-friendly food packaging films that reduce plastic waste.[100][101] In cosmetics, polyphenols from green tea and grape seed extracts act as potent antioxidants, stabilizing formulations against oxidative degradation and providing anti-aging benefits; they are commonly formulated into creams and serums at effective concentrations to neutralize free radicals without irritation.[102][103]Environmentally, polyphenols play a key role in phytoremediation by chelating heavy metals in contaminated soils, aiding hyperaccumulatorplants in uptake and detoxification. Such processes promote soil rehabilitation in metal-polluted sites. The European Union's 2025 ban on bisphenols in food contact materials under Regulation (EU) 2024/3190 further incentivizes bio-based polyphenols as safer, sustainable substitutes for synthetic antioxidants in industrial applications.[104]Despite these advantages, industrial adoption of polyphenol extracts faces challenges in scalability and cost, with production prices ranging from $10 to $50 per kg depending on extraction methods and purity, limiting widespread use compared to cheaper synthetics.[105][106] Advanced techniques like supercritical fluid extraction are being explored to improve yield and reduce expenses, but variability in plant sourcing remains a barrier.[107]
Health Research
Cardiovascular Effects
Polyphenols have been extensively studied for their potential to mitigate cardiovascular disease (CVD) risk through multiple mechanisms, primarily involving the improvement of endothelial function and the attenuation of oxidative stress. A key mechanism is the enhancement of nitric oxide (NO) production in endothelial cells, where polyphenols such as quercetin and resveratrol activate endothelial nitric oxide synthase (eNOS) via phosphorylation pathways, leading to vasodilation and reduced vascular stiffness.[108][109] Another critical pathway is the inhibition of low-density lipoprotein (LDL) oxidation; polyphenols like those in green tea catechins scavenge reactive oxygen species and chelate metal ions, thereby preventing the formation of oxidized LDL, a primary initiator of atherosclerotic plaque development.[110][111] These actions collectively contribute to lowered blood pressure, improved lipid profiles, and reduced inflammation in the vascular system.[112]Dietary sources rich in polyphenols, such as berries and red wine, demonstrate specific cardiovascular benefits in human studies. Berries, containing high levels of anthocyanins, have been shown to reduce systolic blood pressure by approximately 3-5 mmHg and improve HDL cholesterol levels in randomized trials, attributed to their antioxidant capacity.[113][114] Similarly, cocoa flavanols from dark chocolate supplementation lower systolic blood pressure by 4-5 mmHg in normotensive and hypertensive individuals, as evidenced by meta-analyses of over 40 trials involving thousands of participants.[115] Red wine polyphenols, particularly resveratrol, exhibit anti-atherosclerotic effects by modulating inflammatory markers like TNF-α, though benefits are confounded by alcohol's J-shaped dose-response curve, where moderate intake (1-2 glasses daily) correlates with 20-30% lower CVD risk compared to abstinence or heavy consumption.[116][117]Large-scale meta-analyses reinforce these findings, indicating substantial CVD risk reductions with higher polyphenol intake. A 2021 analysis of prospective cohort studies reported a 46% lower risk of CVD events associated with elevated total polyphenol consumption, driven largely by flavonoids and lignans from plant sources.[118] For resveratrol specifically, a 2022 systematic review of randomized controlled trials in CVD patients found significant reductions in inflammatory cytokines and improved endothelial function, supporting its role in slowing atherosclerosis progression.[119] A 2024 meta-analysis of 284 studies (n = 17,613) on antioxidant polyphenol supplementation confirmed improvements in cardiometabolic markers, including subtype-specific reductions such as -1.56 mmHg in systolic blood pressure with catechins and -0.18 mmol/L (~7 mg/dL) in LDL cholesterol with anthocyanins, particularly with multi-polyphenol formulations like those from grapes or berries.[120] In May 2025, the European Food Safety Authority (EFSA) confirmed a health claim for olive oil polyphenols (≥5 mg hydroxytyrosol and derivatives daily) in protecting LDL particles from oxidative damage, thereby contributing to the control of blood LDL-cholesterol levels and reduction of systolic blood pressure, further supporting cardiovascular benefits.[121]Despite these associations, evidence from long-term randomized controlled trials remains limited prior to 2025, with most studies spanning only 4-12 weeks and focusing on surrogate endpoints rather than hard CVD outcomes like myocardial infarction.[122] Ongoing challenges include variability in polyphenol bioavailability due to gut metabolism and the need for larger, diverse population trials to confirm causality and optimal dosing.[123] Future research prioritizing whole-food interventions over isolates may better elucidate sustained benefits.[124]
Anticancer Potential
Polyphenols, particularly flavonoids and isoflavones, demonstrate anticancer potential through multiple molecular mechanisms, including the induction of apoptosis and inhibition of the NF-κB signaling pathway, which regulates inflammation and cell survival in tumors.[125][126] For instance, epigallocatechin gallate (EGCG), a catechinflavonoid from green tea, induces apoptosis in cancer cells by blocking epidermal growth factor receptor (EGFR) autophosphorylation and downstream signaling pathways such as MAPK and PI3K, as observed in prostate cancer cell lines.[127] Similarly, curcumin, a polyphenolic compound from turmeric, promotes apoptosis and sensitizes breast cancer xenografts to mitomycin Cchemotherapy in animal models by enhancing drug-induced cell death and reducing tumor regression resistance.[128]Isoflavones like genistein further contribute by modulating estrogen receptor pathways and inhibiting cell proliferation in hormone-dependent cancers.[129]Epidemiological studies support a protective role for polyphenol-rich diets against cancer development, with high intake of flavonoids linked to reduced colorectal cancerrisk. In a large cohort analysis, flavonol consumption was associated with approximately a 35% lower incidence of colorectal cancer, while flavone intake correlated with a 22-23% risk reduction in Italian populations.[130][131] Animal and in vitro evidence reinforces these findings; for example, quercetin, a flavonoid, inhibits NF-κB activation and induces apoptosis in colon cancer cell lines, leading to reduced tumor growth in rodent models.[132]Isoflavones from soy have shown chemopreventive effects in preclinical studies by suppressing prostate tumor progression through epigenetic modulation.[129]Clinical trials exploring polyphenols as chemopreventives have yielded promising but mixed results, often limited by bioavailability and dosing challenges. A phase II randomized trial of soy isoflavones (80 mg/day) in men with localized prostate cancer demonstrated modulation of prostate-specific antigen levels, suggesting potential as a chemopreventive agent, though larger studies are needed for confirmation.[133] In breast cancer, a 2024 meta-analysis of green tea polyphenol interventions reported modest reductions in tumor markers and a 44% lower recurrence risk with prediagnostic consumption in early-stage patients, attributed to EGCG's anti-proliferative effects.[134][135] However, achieving therapeutic plasma concentrations remains difficult due to rapid metabolism, necessitating strategies like nanoparticle formulations to enhance efficacy in future trials.[136]
Neurological and Cognitive Benefits
Polyphenols exert neuroprotective effects through their ability to cross the blood-brain barrier (BBB), where certain compounds like flavonoids and phenolic acids reach brain tissues in sufficient concentrations to influence neuronal function.[137] Once in the brain, polyphenols promote neurogenesis by upregulating brain-derived neurotrophic factor (BDNF), a key protein involved in synaptic plasticity and neuronal survival, particularly in the hippocampus.[138] For instance, anthocyanins from blueberries have been shown to enhance BDNF expression and improve memory performance in elderly individuals during randomized controlled trials.[139]Epidemiological evidence links higher polyphenol intake to reduced dementia risk, with a 2022 consensus on Mediterranean diet adherence—rich in polyphenols—indicating a 10-20% lower incidence of cognitive impairment and Alzheimer's disease across large cohorts.[140]Curcumin, a polyphenol from turmeric, has demonstrated potential in reducing amyloid-beta plaques and tau protein accumulation in Alzheimer's models and small human trials, correlating with memory improvements.[141]Flavonoids in dark chocolate, such as epicatechin, enhance mood by modulating serotonin and endorphin pathways, as evidenced by acute supplementation studies showing reduced stress and improved emotional well-being.[142]Resveratrol provides neuroprotection in stroke animal models by mitigating oxidative stress and inflammation, reducing infarct size by up to 40% in rodent ischemia experiments.[143]Despite promising preclinical and observational data, much human evidence remains from cohort studies rather than large-scale interventional trials, limiting causal inferences.[144] Ongoing neuroimaging research, including 2025 functional MRI studies, is beginning to elucidate polyphenol-induced changes in brainconnectivity and perfusion, potentially bridging these gaps.[145]
Polyphenols exhibit bidirectional interactions with the gut microbiome, acting as substrates for microbial metabolism while modulating bacterial composition and function. Gut bacteria transform complex polyphenols into bioavailable metabolites, and in turn, polyphenols promote the growth of beneficial microbes, enhancing overall microbial diversity and metabolic output. These interactions contribute to gut homeostasis by influencing bacterial populations and their fermentation products.[146]Polyphenols demonstrate prebiotic effects by selectively stimulating the proliferation of beneficial bacteria, such as Bifidobacterium species. For instance, procyanidins from apples have been shown to significantly increase Bifidobacterium populations in human fecal samples during in vitrofermentation, with studies reporting up to a twofold to threefold rise in their abundance compared to controls. This selective growth supports a healthier microbial ecosystem. Additionally, polyphenols enhance the production of short-chain fatty acids (SCFAs) like acetate, propionate, and butyrate through microbial fermentation, which nourish colonocytes and regulate immune responses.[147][148][149]A key aspect of these interactions involves the biotransformation of polyphenols by gut microbiota, where bacteria cleave glycosidic bonds and perform deglycosylation or ring fission to generate active metabolites. Ellagitannins, found in pomegranates and berries, are converted to urolithins via multi-step microbial processes, with genera such as Bacteroides contributing to initial cleavage steps; however, only about 40% of individuals produce significant urolithin A due to inter-individual microbiome variations. These metabolites exhibit enhanced bioavailability and bioactivity compared to parent compounds.[150][151]Recent randomized controlled trials (RCTs) from 2023 and 2024 underscore these benefits, demonstrating that polyphenol-rich interventions improve gut barrier integrity and reduce inflammation. In one 2024 RCT involving older adults, a polyphenol-enriched diet significantly lowered serum and fecal calprotectin levels—a marker of intestinal inflammation—while enhancing barrier function as measured by zonulin concentrations. Such interventions have also shown promise in alleviating irritable bowel syndrome (IBS) symptoms, with polyphenol blends combined with prebiotics reducing abdominal pain and bloating in IBS patients over 8 weeks.[152][153]High polyphenol intake further addresses dysbiosis by modulating key bacterial ratios, particularly decreasing the Firmicutes/Bacteroidetes ratio, which is often elevated in metabolic disorders. In a 2024 RCT with patients post-acute myocardial infarction, polyphenol supplementation reduced this ratio (p=0.04) while increasing SCFA-producers like Roseburia, thereby restoring microbial balance and mitigating inflammatory dysbiosis.[154][155]
Safety and Toxicity
Polyphenols present in foods are generally recognized as safe, with the U.S. Food and Drug Administration (FDA) granting GRAS status to various polyphenol-rich extracts, such as those from olive leaves and green tea catechins, when used at intended levels in food products.[156][157] However, for polyphenol supplements, safe upper intake levels are not uniformly established across all types due to variability in bioavailability and compound-specific effects, though intakes exceeding 1 g per day from supplements have been associated with potential risks.[158] The European Food Safety Authority (EFSA) has not issued broad 2025 guidelines for polyphenols but maintains specific assessments, such as for green tea catechins, recommending caution with high-dose supplements.[159]At high doses, particularly above 1 g per day, polyphenols can exhibit pro-oxidant activity, shifting from antioxidant benefits to generating reactive oxygen species that may damage cells.[158] Additionally, their iron-chelating properties can inhibit non-heme iron absorption in the gastrointestinal tract, potentially leading to iron deficiency anemia, especially in individuals with low iron stores, vegetarians, or pregnant women consuming high-polyphenol diets or supplements.[160][161] Certain polyphenols, such as soy isoflavones, act as phytoestrogens with weak estrogenic activity due to structural similarity to 17-β-estradiol, posing risks of hormonal disruption in sensitive populations, including infants, children, and those with hormone-sensitive conditions like thyroid disorders or estrogen-dependent cancers.[162][163]Polyphenols can interact with medications by inhibiting cytochrome P450 (CYP) enzymes; for instance, quercetin potently inhibits CYP3A4 and CYP2D6, potentially altering the metabolism and increasing plasma levels of drugs like statins, anticoagulants, and immunosuppressants, leading to enhanced toxicity or reduced efficacy.[164][165] Case reports have linked high-dose green tea extracts (rich in catechins like EGCG) to hepatotoxicity, including acute liver injury, prompting regulatory warnings; EFSA noted elevated liver enzyme levels in human studies at doses above 800 mg EGCG per day and advised against use in susceptible individuals.[166][159]In toxicology studies, polyphenols demonstrate low acute toxicity, with no observed adverse effects in animals at oral doses up to 3 g/kg body weight, corresponding to LD50 values exceeding 5 g/kg in rodents.[167] Subchronic and chronicanimal studies support safety at dietary levels, though EFSA emphasizes monitoring for compound-specific effects in updated risk assessments.[159]