Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2312 2023-01-09 10:54:40 |
2 format correct -8 word(s) 2304 2023-01-10 07:20:28 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Bojarczuk, A.;  Dzitkowska-Zabielska, M. Properties of Polyphenol-Based Dietary Supplements. Encyclopedia. Available online: https://encyclopedia.pub/entry/39906 (accessed on 08 July 2024).
Bojarczuk A,  Dzitkowska-Zabielska M. Properties of Polyphenol-Based Dietary Supplements. Encyclopedia. Available at: https://encyclopedia.pub/entry/39906. Accessed July 08, 2024.
Bojarczuk, Aleksandra, Magdalena Dzitkowska-Zabielska. "Properties of Polyphenol-Based Dietary Supplements" Encyclopedia, https://encyclopedia.pub/entry/39906 (accessed July 08, 2024).
Bojarczuk, A., & Dzitkowska-Zabielska, M. (2023, January 09). Properties of Polyphenol-Based Dietary Supplements. In Encyclopedia. https://encyclopedia.pub/entry/39906
Bojarczuk, Aleksandra and Magdalena Dzitkowska-Zabielska. "Properties of Polyphenol-Based Dietary Supplements." Encyclopedia. Web. 09 January, 2023.
Properties of Polyphenol-Based Dietary Supplements
Edit

Antioxidants in sports exercise training remain a debated research topic. Plant-derived polyphenol supplements are frequently used by athletes to reduce the negative effects of exercise-induced oxidative stress, accelerate the recovery of muscular function, and enhance performance. 

dietary polyphenols athletes antioxidant status

1. Polyphenols, Oxidative Stress, and Inflammation

As mentioned above, polyphenols are antioxidants. This activity is ascribable to hydroxyl groups operating as electron donors and stabilizing free radicals through the delocalization of unpaired electrons [1], and is termed a chain-breaking function [2]. Polyphenols also demonstrate metal chelation activity. Because transition metals can produce reactive oxygen species (ROS) that damage the genome, polyphenols play an antioxidative and DNA-protective role [3]. In addition, they inhibit certain enzymes involved in ROS production, i.e., xanthine oxidase and NADPH oxidase, whilst upregulating other endogenous antioxidant enzymes, e.g., superoxide dismutase (SOD), catalase, and glutathione peroxidase (GPX) [4]. The antioxidative property of polyphenols is important for athletes, because it is well established that exercise generates ROS formation and prolonged exercise can promote oxidative damage to active myofibers [5][6][7]. Oxidative stress presents an imbalance between ROS-generating and -scavenging systems. ROS contain at least one oxygen atom and one or more unpaired electrons. ROS include free radicals (e.g., superoxide anion (O2), hydroxyl radical ·OH) and non-radicals (e.g., singlet oxygen (1O2), hydrogen peroxide (H2O2)). ROS overproduction can lead to tissue injury and initiate inflammation [4]. Superoxide (O2) is relatively membrane-impermeable and unreactive [8] and has a short half-life in the cell [9]. Superoxide dismutase (SOD) spontaneously converts O2 to non-radical hydrogen peroxide (H2O2) [10] which is further reduced to water by GPX [11]. H2O2 molecule is considered stable and can therefore diffuse considerable distances in or out of the cell [12]. O2 and H2O2 are products of NADPH oxidase [13]. In contrast to H2O2, the hydroxyl radical (·OH) is a strong oxidizing agent and is considered the most damaging [14][15]. Another free radical is nitric oxide (NO·), which is a short-lived and weak oxidant produced by the action of nitric oxide synthase (NOS) [16]. It has been shown that oxidative stress and inflammation are related, because inflammatory response occurs after exercise [17]. During inflammation, phagocytes undergo a respiratory burst that imposes high oxidative stress on the engulfed pathogens [18]. The oxidative burst is mediated by NADPH oxidase, which produces O2- and H2O2 [19], and these products achieve their cytotoxic effects due to reactions with other antimicrobial systems to generate ROS [8][20]. In skeletal muscle cells in culture, two primary NADPH oxidase isoforms (NOX2 and NOX4) [21] were found to be the main sources of ROS [22][23]. Thus, the role of NADPH oxidase-derived ROS is important in pathogen-mediated and sterile inflammation (i.e., without pathogen involvement). Importantly, ROS can exacerbate the inflammatory process [24]. Nonetheless, free radicals are crucial for proper cell physiology and function actively as intracellular messengers [9][24]. This means that ROS have both positive and negative physiological effects. ROS are by-products of cellular metabolism, and constitutive antioxidant defenses neutralize the housekeeping production of reactive species, e.g., superoxide dismutase (SOD) converts O2− to H2O2 [25]. Another antioxidant strategy is the scavenging of ROS by numerous molecules of low molecular weight that exist in the extracellular space and also within cells. A further method is to minimize the availability of pro-oxidants via metal-binding proteins. These chelating molecules avert these transition metals from participation in ROS formation [12][26]. The toxic effects of ROS and RNS are also counteracted by non-enzymatic reactions, e.g., ascorbic acid or vitamin E [27]. However, when the production of radicals overwhelms the antioxidant defenses, oxidative stress proceeds. This can lead to changes in cell membranes and other biological structures including lipoproteins, lipids, proteins, DNA [28], etc. Polyphenols are of particular interest because they can also protect against oxidative damage through various mechanisms. For instance, resveratrol, which is a stilbenoid and a type of natural phenol found in red grapes, reduced the expression of NADPH oxidase 4 (NOX4) in mice [29]. Quercetin, a natural flavonoid found in onions, green tea, and apples, decreases lipid peroxidation and inhibits cellular oxidation in erythrocytes in type 2 diabetic patients [30]. Another flavonoid, myricetin also protected healthy human erythrocytes from oxidative stress in vitro [31]. The same effect was observed for tea catechins in human diabetic erythrocytes [32]. Conversely, others showed that catechins provided only modest protection from oxidative damage in healthy men [33] and showed no effect in soccer players [34]. Notwithstanding their high training demands, athletes’ diets may not contain sufficient antioxidants to support their physical activity [35][36]. This raises the question of whether they should take dietary polyphenols. Given that the health benefits of exercise are well documented, it seems unlikely that exercise-generated free radicals should negatively impact performance in the long term [35][37].

2. Polyphenols Bioavailability

The bioavailability of a substance is the fractional extent to which the active moiety is absorbed and rate at which the dosage reaches the therapeutic site of action [38][39]. The issue of bioavailability presents a major challenge for clarifying the therapeutic effects of polyphenols and may account for large inter-individual variations in clinical trials [40]. Many factors influence the bioavailability of dietary polyphenols in humans including external factors (sun exposure, degree of ripeness of plants), factors related to food processing (homogenization, lyophilization), interactions with other compounds (bonds with proteins (e.g., albumin), polyphenol-related factors (i.e., chemical structure, concentration in food)) [41][42].
The in vitro approach to studying polyphenols’ absorption, metabolism, and bioavailability has mostly used the Caco-2 cell line [43][44][45][46], because the line reliably reflects biological barriers and the ability to perform phase II biotransformation by adding hydrophilic groups. Nonetheless, other cells have also been used [47]. The in vivo approach mostly uses rodents [48][49][50][51] and humans [52][53][54][55]. Absorption of polyphenols in mammals occurs in the small intestine after cleaving off sugar moieties [56]. This process occurs by either passive diffusion or with the use of specific transporters localized in the enterocyte membrane [42]. Only approximately 5–10% is absorbed, and the remainder accumulates in the large intestine with the bile conjugates, being thereby subjected to digestion [57]. Importantly, polyphenols of low molecular weight such as gallic or caffeic acids are partially absorbed into the body directly or after conversion. In contrast, polyphenols of higher molecular weight, e.g., proanthocyanidins, are very poorly absorbed [58]. Phenolic compounds absorbed in the small intestine are further transported to the colon, where they are transformed by the gut microbiota. These microorganisms perform a myriad of activities including hydrolysis, dihydroxylation, demethylation, and decarboxylation [42] thereby converting phenolic compounds into bioavailable metabolites. Meanwhile, polyphenol-rich extracts can also modulate gut microflora [59][60]. Importantly, the position of the hydroxyl groups might influence the polyphenols’ breakdown. For instance, flavonoids without hydroxyl groups at the C5, C7, and C4′ positions are degraded more slowly [56][61]. This implies that the slow-degrading compounds might be more bioavailable because their absorption can occur more quickly than those degraded more rapidly at the colon level [62]. When polyphenol sub-products are absorbed by the small intestine or colon, phase I and II transformation begins within enterocytes and in the liver to be distributed through the blood [63][64][65]. However, the metabolites may encounter enterohepatic circulation, whereby they are returned to the small intestine and transported back to the liver [40][42][66]. Polyphenols can stay in the blood or be delivered to other tissues. Indeed, polyphenol metabolites have been found in various tissues, e.g., glucuronidated and methyl-glucuronidated derivatives of catechin and epicatechin in the muscle of rats fed with grape pomace extract [67] or catechin-glucuronide, methyl catechin-glucuronide and methyl catechin-sulfate in kidneys, intestine, lungs, spleen, and thymus of rats receiving hazelnut extract [68]. Nonetheless, their targeted delivery is a major problem relating to polyphenol-rich supplements and their low bioavailability. So far, many systems have been developed to address the issue. However, these approaches mainly concentrate on achieving delivery that minimizes systemic diffusion and degradation of phenolic compounds, because some phenolics are poorly absorbed in the digestive tract while others are extensively metabolized to derivatives with a lower activity or are degraded [42][69]. This means that when polyphenols are orally administered, their therapeutic concentration might not be achieved [70]. Temperature, light, oxygen, acidic pH, and enzymatic activity in the digestive system [70] can all affect the beneficial effects of polyphenols [71]. As mentioned above, polyphenols follow phase I and II metabolic pathways. The major pathway is phase II, during which glucuronidation and sulfation lead to especially hydrophilic conjugates, and methylation to similar if not slightly more lipophilic metabolites. The methylated products are conjugated into glucuronides and sulfates, as long as there is a functional group available for conjugation into more hydrophilic metabolites [63]. Moreover, unabsorbed and hydrophilic polyphenol conjugates can undergo transformation triggered by gut microflora, which can even result in ring fission. Bacteria can dramatically reduce polyphenols’ bioactivity [72]. Therefore, various research efforts have been focused on the problem of polyphenol delivery. Of note, bioavailability is a key determinant of polyphenols’ potential health-promoting applications [42], and is also important in establishing dietary reference intake (RDI) [73].

3. The Intake of Phenolic Compounds in an Average Daily Diet

The biodiversity of polyphenols found in food is wide. Therefore, it is extremely difficult to determine their content within food products and measure their daily intake [74]. Several research articles have extrapolated the daily dietary intake of particular flavonoids among the European population, based on the US database of flavonoid concentrations in particular foods. Many authors have referred to the data published by Kühnau et al., almost 46 years ago, where a daily intake of 1 g of total phenols was established in the US population [75]. There have been attempts to define the daily consumption of polyphenols within the diets of different populations. So far, Hertog et al., have done this for flavonols and flavones [76], and Reinli and Block for isoflavones in the Dutch population [77]. Researchers established the intake of flavonols (mainly quercetin) and flavones as 21 and 2 mg/d, respectively [76]. The average dietary consumption of isoflavones in the Japanese population was determined to be 30–40 mg/d [78][79]. Interestingly, soy products are less popular in Western countries, which is reflected in the lower polyphenol content of the Western diet [80], explaining why the daily dietary intake of quercetin and genistein does not exceed 2–4% of total polyphenols [74]. Some authors report that daily polyphenol consumption above 650 mg decreases risk of death in comparison with those whose daily polyphenols intake is below 500 mg [81]. Other authors report health benefits ranging from intakes of 500 mg to 1500 mg per day [74]. Further sources recommend a daily dose of 0.1–1.0 g of polyphenols. Fruits such as grapes, apples, pears, cherries, and blueberries contain up to 200–300 mg of polyphenols per 100 g of fresh weight. Interestingly, a glass of red wine or a cup of tea contains about 100 mg of polyphenols, and the presence of these products in the diet may reduce the likelihood of chronic diseases. In Europe, the main sources of polyphenols are onions, black tea, red wine, and apples [81][82].

4. Polyphenols Delivery

Various delivery systems for polyphenols have been developed to improve their efficiency and combat their bioactivity problems. One major issue is to find a way to enhance the penetration of active substances and bring hydrophilic compounds into the tissues. This concept has been explored in numerous studies demonstrating the use of formulations with simple emulsions [83][84][85], cyclic glucan oligosaccharides known as cyclodextrins [86][87][88], gels [89][90][91], nanoemulsion [92][93][94], or liposomes [95][96][97]. Other solutions are also available, such as micelles, nanocomposites, metal oxide nanoparticles, etc. [70]. The main reason for encapsulating polyphenols is to tackle the problem of their stability. For instance, maltodextrins preserve the integrity of anthocyanins [98]. Encapsulation increases biocompatibility and prevents degradation caused by the external environment. It also minimizes interactions with other components of the human body [99]. In general, coating by microencapsulation in particles up to 1000 μm protects active substances and preserves their antioxidant properties [100][101][102]. However, nanotechnology is apparently more effective, because nanoencapsulated polyphenols have been found to increase the protection of active substances and bioavailability as well as improving controlled targeted release [103]. Particle size is generally seen as a factor strongly related to bioavailability. The development of a nanoscale delivery system is aimed at achieving improved site-selective targeting [104]. This is possible due to the small molecular size and active incorporation into cells by different endocytic pathways [105][106]. However, targeted delivery is complicated. It can be achieved actively or passively [107]; active targeting involves the therapeutic agent (in this case a polyphenol) being loaded into a carrier and this conjugate attached to tissue or cell-specific ligands [108]. By contrast, passive targeting requires loading the therapeutic agent into a nanomolecule that passively reaches the target tissue or organ. This leads to the accumulation of a drug delivery system with a specific size, molecular mass, and charge [109]. Given that a passive targeting system might utilize specific conditions in the diseased tissues or cells (e.g., low pH) [109], it is reasonable to assume that this approach is suitable to treat oxidative stress generated by exercise. After all, the higher the exercise intensity, the lower the muscle pH [110]. Nonetheless, the active approach also seems effective since it is possible to target specific cells or the inside of cells (for targeting intracellular organelles) [109].
Importantly, approaches should be compatible with another target, that of achieving the concentrations required for systemic therapies. In these circumstances, the hydrophilic part of a phenol carrier must be in balance with the lipophilic part [111], because the affinity of polyphenols for lipid bilayers partially determines their biological activity in vitro [112] and is of great significance in the biomedical and dietary fields. Polyphenolic extracts interact with the cell membrane by creating a protective coat around the lipid membrane, through their location on the membrane surface. This effect was shown using liposomes as models of lipid membranes, wherein trans-stilbenes and flavonoids interacted at the hydrophilic interface [113]. This finding is in agreement with another study reporting that three different types of blueberry extracts changed the arrangement of the hydrophilic region of the liposome membranes [114]. Liposomes themselves possess an aqueous central section as well as hydrophobic and hydrophilic components comprising a lipid bilayer. The aqueous cores typically encapsulate hydrophilic compounds. By contrast, hydrophobic substances favor lipid bilayers. Thus, liposomes might be used for the delivery of diverse substances, such as hydrophilic and hydrophobic compounds [115].

References

  1. Quideau, S.; Deffieux, D.; Douat-Casassus, C.; Pouységu, L. Plant Polyphenols: Chemical Properties, Biological Activities, and Synthesis. Angew. Chem. Int. Ed. Eng. 2011, 50, 586–621.
  2. Viglianisi, C.; Menichetti, S. Chain Breaking Antioxidant Activity of Heavy (S, Se, Te) Chalcogens Substituted Polyphenols. Antioxidants 2019, 8, 487.
  3. Lakey-Beitia, J.; Burillo, A.M.; La Penna, G.; Hegde, M.L.; Rao, K.S. Polyphenols as Potential Metal Chelation Compounds against Alzheimer’s Disease. J. Alzheimers Dis. 2021, 82, S335–S357.
  4. Yahfoufi, N.; Alsadi, N.; Jambi, M.; Matar, C. The Immunomodulatory and Anti-Inflammatory Role of Polyphenols. Nutrients 2018, 10, 1618.
  5. Alessio, H.M.; Goldfarb, A.H.; Cutler, R.G. MDA Content Increases in Fast- and Slow-Twitch Skeletal Muscle with Intensity of Exercise in a Rat. Am. J. Physiol. 1988, 255, C874–C877.
  6. Davies, K.J.; Quintanilha, A.T.; Brooks, A.G.; Packer, L. Free Radicals and Tissue Damage Produced by Exercise. Biochem. Biophys. Res. Commun. 1982, 107, 1198–1205.
  7. Duthie, G.G.; Robertson, J.D.; Maughan, R.J.; Morrice, P.C. Blood Antioxidant Status and Erythrocyte Lipid Peroxidation Following Distance Running. Arch. Biochem. Biophys. 1990, 282, 78–83.
  8. Powers, S.K.; Ji, L.L.; Kavazis, A.N.; Jackson, M.J. Reactive Oxygen Species: Impact on Skeletal Muscle. Compr. Physiol. 2011, 1, 941–969.
  9. Phaniendra, A.; Jestadi, D.B.; Periyasamy, L. Free Radicals: Properties, Sources, Targets, and Their Implication in Various Diseases. Indian J. Clin. Biochem. 2015, 30, 11–26.
  10. Fukai, T.; Ushio-Fukai, M. Superoxide Dismutases: Role in Redox Signaling, Vascular Function, and Diseases. Antioxid. Redox Signal 2011, 15, 1583–1606.
  11. Wang, Y.; Hekimi, S. Understanding Ubiquinone. Trends Cell Biol. 2016, 26, 367–378.
  12. Powers, S.K.; Deminice, R.; Ozdemir, M.; Yoshihara, T.; Bomkamp, M.P.; Hyatt, H. Exercise-Induced Oxidative Stress: Friend or Foe? J. Sport Health Sci. 2020, 9, 415–425.
  13. Amanso, A.; Lyle, A.N.; Griendling, K.K. Nadph Oxidases and Measurement of Reactive Oxygen Species. Methods Mol. Biol. 2017, 1527, 219–232.
  14. Ward, J.F.; Evans, J.W.; Limoli, C.L.; Calabro-Jones, P.M. Radiation and Hydrogen Peroxide Induced Free Radical Damage to DNA. Br. J. Cancer Suppl. 1987, 55, 105–112.
  15. Ransy, C.; Vaz, C.; Lombès, A.; Bouillaud, F. Use of H2O2 to Cause Oxidative Stress, the Catalase Issue. Int. J. Mol. Sci. 2020, 21, 9149.
  16. Radi, R. Oxygen Radicals, Nitric Oxide, and Peroxynitrite: Redox Pathways in Molecular Medicine. Proc. Natl. Acad. Sci. USA 2018, 115, 5839–5848.
  17. Cerqueira, É.; Marinho, D.A.; Neiva, H.P.; Lourenço, O. Inflammatory Effects of High and Moderate Intensity Exercise—A Systematic Review. Front. Physiol. 2020, 10, 01550.
  18. Cavinato, L.; Genise, E.; Luly, F.R.; Domenico, E.G.D.; Del Porto, P.; Ascenzioni, F. Escaping the Phagocytic Oxidative Burst: The Role of SODB in the Survival of Pseudomonas Aeruginosa Within Macrophages. Front. Microbiol. 2020, 11, 326.
  19. Nauseef, W.M. Detection of Superoxide Anion and Hydrogen Peroxide Production by Cellular NADPH Oxidases. Biochim. Biophys. Acta 2014, 1840, 757–767.
  20. Vatansever, F.; de Melo, W.C.M.A.; Avci, P.; Vecchio, D.; Sadasivam, M.; Gupta, A.; Chandran, R.; Karimi, M.; Parizotto, N.A.; Yin, R.; et al. Antimicrobial Strategies Centered around Reactive Oxygen Species—Bactericidal Antibiotics, Photodynamic Therapy, and Beyond. FEMS Microbiol. Rev. 2013, 37, 955–989.
  21. Ferreira, L.F.; Laitano, O. Regulation of NADPH Oxidases in Skeletal Muscle. Free Radic. Biol. Med. 2016, 98, 18–28.
  22. Michaelson, L.P.; Shi, G.; Ward, C.W.; Rodney, G.G. Mitochondrial Redox Potential during Contraction in Single Intact Muscle Fibers. Muscle Nerve 2010, 42, 522–529.
  23. D’amico, A.; Cavarretta, E.; Fossati, C.; Borrione, P.; Pigozzi, F.; Frati, G.; Sciarretta, S.; Costa, V.; De Grandis, F.; Nigro, A.; et al. Platelet Activation Favours NOX2-Mediated Muscle Damage in Elite Athletes: The Role of Cocoa-Derived Polyphenols. Nutrients 2022, 14, 1558.
  24. Mittal, M.; Siddiqui, M.R.; Tran, K.; Reddy, S.P.; Malik, A.B. Reactive Oxygen Species in Inflammation and Tissue Injury. Antioxid. Redox Signal 2014, 20, 1126–1167.
  25. Snezhkina, A.V.; Kudryavtseva, A.V.; Kardymon, O.L.; Savvateeva, M.V.; Melnikova, N.V.; Krasnov, G.S.; Dmitriev, A.A. ROS Generation and Antioxidant Defense Systems in Normal and Malignant Cells. Oxid. Med. Cell Longev. 2019, 2019, 6175804.
  26. Powers, S.K.; Jackson, M.J. Exercise-Induced Oxidative Stress: Cellular Mechanisms and Impact on Muscle Force Production. Physiol. Rev. 2008, 88, 243–276.
  27. Landete, J.M. Dietary Intake of Natural Antioxidants: Vitamins and Polyphenols. Crit. Rev. Food Sci. Nutr. 2013, 53, 706–721.
  28. Rudrapal, M.; Khairnar, S.J.; Khan, J.; Dukhyil, A.B.; Ansari, M.A.; Alomary, M.N.; Alshabrmi, F.M.; Palai, S.; Deb, P.K.; Devi, R. Dietary Polyphenols and Their Role in Oxidative Stress-Induced Human Diseases: Insights into Protective Effects, Antioxidant Potentials and Mechanism(s) of Action. Front. Pharmacol. 2022, 13, 806470.
  29. Xian, Y.; Gao, Y.; Lv, W.; Ma, X.; Hu, J.; Chi, J.; Wang, W.; Wang, Y. Resveratrol Prevents Diabetic Nephropathy by Reducing Chronic Inflammation and Improving the Blood Glucose Memory Effect in Non-Obese Diabetic Mice. Naunyn Schmiedebergs Arch. Pharmacol. 2020, 393, 2009–2017.
  30. Rizvi, S.I.; Mishra, N. Anti-Oxidant Effect of Quercetin on Type 2 Diabetic Erythrocytes. J. Food Biochem. 2009, 33, 404–415.
  31. Maurya, P.K.; Rizvi, S.I. Protective Role of Tea Catechins on Erythrocytes Subjected to Oxidative Stress during Human Aging. Nat. Prod. Commun 2009, 4, 221–226.
  32. Rizvi, S.I.; Zaid, M.A.; Anis, R.; Mishra, N. Protective Role of Tea Catechins against Oxidation-Induced Damage of Type 2 Diabetic Erythrocytes. Clin. Exp. Pharmacol. Physiol. 2005, 32, 70–75.
  33. Jówko, E.; Sacharuk, J.; Balasińska, B.; Ostaszewski, P.; Charmas, M.; Charmas, R. Green Tea Extract Supplementation Gives Protection against Exercise-Induced Oxidative Damage in Healthy Men. Nutr. Res. 2011, 31, 813–821.
  34. Jówko, E.; Sacharuk, J.; Balasinska, B.; Wilczak, J.; Charmas, M.; Ostaszewski, P.; Charmas, R. Effect of a Single Dose of Green Tea Polyphenols on the Blood Markers of Exercise-Induced Oxidative Stress in Soccer Players. Int. J. Sport Nutr. Exerc. Metab. 2012, 22, 486–496.
  35. Higgins, M.R.; Izadi, A.; Kaviani, M. Antioxidants and Exercise Performance: With a Focus on Vitamin e and c Supplementation. Int. J. Environ. Res. Public Health 2020, 17, 8452.
  36. Powers, S.K.; Sollanek, K.J. Endurance Exercise and Antioxidant Supplementation: Sense or Nonsense?—Part 1. Sport Sci. Exch. 2014, 27, 1–4.
  37. Kawamura, T.; Muraoka, I. Exercise-Induced Oxidative Stress and the Effects of Antioxidant Intake from a Physiological Viewpoint. Antioxidants 2018, 7, 119.
  38. Kim, M.T.; Sedykh, A.; Chakravarti, S.K.; Saiakhov, R.D.; Zhu, H. Critical Evaluation of Human Oral Bioavailability for Pharmaceutical Drugs by Using Various Cheminformatics Approaches. Pharm. Res. 2014, 31, 1002–1014.
  39. Chow, S.-C. Bioavailability and Bioequivalence in Drug. Wiley Interdiscip. Rev. Comput. Stat. 2014, 6, 304–312.
  40. Teng, H.; Chen, L. Polyphenols and Bioavailability: An Update. Crit. Rev. Food Sci. Nutr. 2019, 59, 2040–2051.
  41. D’Archivio, M.; Filesi, C.; Varì, R.; Scazzocchio, B.; Masella, R. Bioavailability of the Polyphenols: Status and Controversies. Int. J. Mol. Sci. 2010, 11, 1321–1342.
  42. Cosme, P.; Rodríguez, A.B.; Espino, J.; Garrido, M. Plant Phenolics: Bioavailability as a Key Determinant of Their Potential Health-Promoting Applications. Antioxidants 2020, 9, 1263.
  43. CenciČ, A.; Langerholc, T. Functional Cell Models of the Gut and Their Applications in Food Microbiology—A Review. Int. J. Food Microbiol. 2010, 141 (Suppl. 1), S4–S14.
  44. Li, Q.; Wang, C.; Liu, F.; Hu, T.; Shen, W.; Li, E.; Liao, S.; Zou, Y. Mulberry Leaf Polyphenols Attenuated Postprandial Glucose Absorption: Via Inhibition of Disaccharidases Activity and Glucose Transport in Caco-2 Cells. Food Funct. 2020, 11, 1835–1844.
  45. Lingua, M.S.; Theumer, M.G.; Kruzynski, P.; Wunderlin, D.A.; Baroni, M.V. Bioaccessibility of Polyphenols and Antioxidant Properties of the White Grape by Simulated Digestion and Caco-2 Cell Assays: Comparative Study with Its Winemaking Product. Food Res. Intal. 2019, 122, 496–505.
  46. Zhang, H.; Hassan, Y.I.; Renaud, J.; Liu, R.; Yang, C.; Sun, Y.; Tsao, R. Bioaccessibility, Bioavailability, and Anti-Inflammatory Effects of Anthocyanins from Purple Root Vegetables Using Mono- and Co-Culture Cell Models. Mol. Nutr. Food Res. 2017, 61, 1600928.
  47. Grootaert, C.; Kamiloglu, S.; Capanoglu, E.; Van Camp, J. Cell Systems to Investigate the Impact of Polyphenols on Cardiovascular Health. Nutrients 2015, 7, 9229–9255.
  48. Olivero-David, R.; Ruiz-Roso, M.B.; Caporaso, N.; Perez-Olleros, L.; De las Heras, N.; Lahera, V.; Ruiz-Roso, B. In Vivo Bioavailability of Polyphenols from Grape By-Product Extracts, and Effect on Lipemia of Normocholesterolemic Wistar Rats. J. Sci. Food Agric. 2018, 98, 5581–5590.
  49. Peng, Y.; Meng, Q.; Zhou, J.; Chen, B.; Xi, J.; Long, P.; Zhang, L.; Hou, R. Nanoemulsion Delivery System of Tea Polyphenols Enhanced the Bioavailability of Catechins in Rats. Food Chem. 2018, 242, 527–532.
  50. Lambert, J.D.; Hong, J.; Kim, D.H.; Mishin, V.M.; Yang, C.S. Piperine Enhances the Bioavailability of the Tea Polyphenol (−)-Epigallocatechin-3-Gallate in Mice. J Nutr. 2004, 134, 1948–1952.
  51. Curti, V.; Zaccaria, V.; Sokeng, A.J.T.; Dacrema, M.; Masiello, I.; Mascaro, A.; D’antona, G.; Daglia, M. Bioavailability and in Vivo Antioxidant Activity of a Standardized Polyphenol Mixture Extracted from Brown Propolis. Int. J. Mol. Sci. 2019, 20, 1250.
  52. Scholz, S.; Williamson, G. Interactions Affecting the Bioavailability of Dietary Polyphenols in Vivo. Int. J. Vitam. Nutr. Res. 2007, 77, 224–235.
  53. Zhong, S.; Sandhu, A.; Edirisinghe, I.; Burton-Freeman, B. Characterization of Wild Blueberry Polyphenols Bioavailability and Kinetic Profile in Plasma over 24-h Period in Human Subjects. Mol. Nutr. Food Res. 2017, 61, 1700405.
  54. Vitaglione, P.; Barone Lumaga, R.; Ferracane, R.; Sellitto, S.; Morelló, J.R.; Reguant Miranda, J.; Shimoni, E.; Fogliano, V. Human Bioavailability of Flavanols and Phenolic Acids from Cocoa-Nut Creams Enriched with Free or Microencapsulated Cocoa Polyphenols. Br. J. Nutr. 2013, 109, 1832–1843.
  55. Martínez-Huélamo, M.; Vallverdú-Queralt, A.; Di Lecce, G.; Valderas-Martínez, P.; Tulipani, S.; Jáuregui, O.; Escribano-Ferrer, E.; Estruch, R.; Illan, M.; Lamuela-Raventós, R.M. Bioavailability of Tomato Polyphenols Is Enhanced by Processing and Fat Addition: Evidence from a Randomized Feeding Trial. Mol. Nutr. Food Res. 2016, 60, 1578–1589.
  56. Marín, L.; Miguélez, E.M.; Villar, C.J.; Lombó, F. Bioavailability of Dietary Polyphenols and Gut Microbiota Metabolism: Antimicrobial Properties. Biomed. Res. Int. 2015, 2015, 905215.
  57. Singh, A.K.; Cabral, C.; Kumar, R.; Ganguly, R.; Rana, H.K.; Gupta, A.; Lauro, M.R.; Carbone, C.; Reis, F.; Pandey, A.K. Beneficial Effects of Dietary Polyphenols on Gut Microbiota and Strategies to Improve Delivery Efficiency. Nutrients 2019, 11, 2216.
  58. Kawabata, K.; Yoshioka, Y.; Terao, J. Role of Intestinal Microbiota in the Bioavailability and Physiological Functions of Dietary Polyphenols. Molecules 2019, 24, 370.
  59. Zhu, Y.; Zhang, J.-Y.; Wei, Y.-L.; Hao, J.-Y.; Lei, Y.-Q.; Zhao, W.-B.; Xiao, Y.-H.; Sun, A.-D. The Polyphenol-Rich Extract from Chokeberry (Aronia Melanocarpa L.) Modulates Gut Microbiota and Improves Lipid Metabolism in Diet-Induced Obese Rats. Nutr. Metab. 2020, 17, 54.
  60. Istas, G.; Wood, E.; Le Sayec, M.; Rawlings, C.; Yoon, J.; Dandavate, V.; Cera, D.; Rampelli, S.; Costabile, A.; Fromentin, E.; et al. Effects of Aronia Berry (Poly)Phenols on Vascular Function and Gut Microbiota: A Double-Blind Randomized Controlled Trial in Adult Men. Am. J. Clin. Nutr. 2019, 110, 316–329.
  61. Winter, J.; Popoff, M.R.; Grimont, P.; Bokkenheuser, V.D. Clostridium Orbiscindens Sp. Nov., a Human Intestinal Bacterium Capable of Cleaving the Flavonoid C-Ring. Int. J. Syst. Bacteriol. 1991, 41, 355–357.
  62. Simons, A.L.; Renouf, M.; Hendrich, S.; Murphy, P.A. Human Gut Microbial Degradation of Flavonoids: Structure−Function Relationships. J. Agric. Food Chem. 2005, 53, 4258–4263.
  63. Liu, Z.; Hu, M. Natural Polyphenol Disposition via Coupled Metabolic Pathways. Expert Opin. Drug Metab. Toxicol. 2007, 3, 389–406.
  64. Van Duynhoven, J.; Vaughan, E.E.; Jacobs, D.M.; Kemperman, R.A.; Van Velzen, E.J.J.; Gross, G.; Roger, L.C.; Possemiers, S.; Smilde, A.K.; Doré, J.; et al. Metabolic Fate of Polyphenols in the Human Superorganism. Proc. Natl. Acad. Sci. USA 2011, 108 (Suppl. 1), 4531–4538.
  65. Luca, S.V.; Macovei, I.; Bujor, A.; Miron, A.; Skalicka-Woźniak, K.; Aprotosoaie, A.C.; Trifan, A. Bioactivity of Dietary Polyphenols: The Role of Metabolites. Crit. Rev. Food Sci. Nutr. 2020, 60, 626–659.
  66. Tu, Y.; Wang, L.; Rong, Y.; Tam, V.; Yin, T.; Gao, S.; Singh, R.; Hu, M. Hepatoenteric Recycling Is a New Disposition Mechanism for Orally Administered Phenolic Drugs and Phytochemicals in Rats. Elife 2021, 10, e58820.
  67. Andres-Lacueva, C.; Macarulla, M.T.; Rotches-Ribalta, M.; Boto-Ordóñez, M.; Urpi-Sarda, M.; Rodríguez, V.M.; Portillo, M.P. Distribution of Resveratrol Metabolites in Liver, Adipose Tissue, and Skeletal Muscle in Rats Fed Different Doses of This Polyphenol. J. Agric. Food Chem. 2012, 60, 4833–4840.
  68. Serra, A.; Macià, A.; Romero, M.-P.; Anglès, N.; Morelló, J.R.; Motilva, M.-J. Distribution of Procyanidins and Their Metabolites in Rat Plasma and Tissues after an Acute Intake of Hazelnut Extract. Food Funct. 2011, 2, 562–568.
  69. Makarewicz, M.; Drożdż, I.; Tarko, T.; Duda-Chodak, A. The Interactions between Polyphenols and Microorganisms, Especially Gut Microbiota. Antioxidants 2021, 10, 188.
  70. Enaru, B.; Socaci, S.; Farcas, A.; Socaciu, C.; Danciu, C.; Stanila, A.; Diaconeasa, Z. Novel Delivery Systems of Polyphenols and Their Potential Health Benefits. Pharmaceuticals 2021, 14, 946.
  71. Pimentel-Moral, S.; Teixeira, M.C.; Fernandes, A.R.; Arráez-Román, D.; Martínez-Férez, A.; Segura-Carretero, A.; Souto, E.B. Lipid Nanocarriers for the Loading of Polyphenols—A Comprehensive Review. Adv. Colloid Interface 2018, 260, 85–94.
  72. Miksits, M.; Maier-Salamon, A.; Aust, S.; Thalhammer, T.; Reznicek, G.; Kunert, O.; Haslinger, E.; Szekeres, T.; Jaeger, W. Sulfation of Resveratrol in Human Liver: Evidence of a Major Role for the Sulfotransferases SULT1A1 and SULT1E1. Xenobiotica 2005, 35, 1101–1119.
  73. Hambidge, K.M. Micronutrient Bioavailability: Dietary Reference Intakes and a Future Perspective. Am. J. Clin. Nutr. 2010, 91, 1430–1432.
  74. Scalbert, A.; Williamson, G. Dietary Intake and Bioavailability of Polyphenols. J. Nutr. 2000, 130 (Suppl. 8), 2073S–2085S.
  75. Kühnau, J. The Flavonoids. A Class of Semi-Essential Food Components: Their Role in Human Nutrition. World Rev. Nutr. Diet 1976, 24, 117–191.
  76. Hertog, M.G.L.; Hollman, P.C.H.; Hertog, M.G.L.; Katan, M.B. Content of Potentially Anticarcinogenic Flavonoids of 28 Vegetables and 9 Fruits Commonly Consumed in the Netherlands. J. Agric. Food Chem. 1992, 40, 2379–2383.
  77. Reinli, K.; Block, G. Phytoestrogen Content of Foods—A Compendium of Literature Values. Nutr. Cancer 1996, 26, 123–148.
  78. Kimira, M.; Arai, Y.; Shimoi, K.; Watanabe, S. Japanese Intake of Flavonoids and Isoflavonoids from Foods. J. Epidemiol. 1998, 8, 168–175.
  79. Wakai, K.; Egami, I.; Kato, K.; Kawamura, T.; Tamakoshi, A.; Lin, Y.; Nakayama, T.; Wada, M.; Ohno, Y. Dietary Intake and Sources of Isoflavones among Japanese. Nutr. Cancer 1999, 33, 139–145.
  80. Kirk, P.; Patterson, R.E.; Lampe, J. Development of a Soy Food Frequency Questionnaire to Estimate Isoflavone Consumption in US Adults. J. Am. Diet Assoc. 1999, 99, 558–563.
  81. GutiErrez-Grijalva, E.P.; Ambriz-Pere, D.L.; Leyva-Lopez, N.; Castillo-Lopez, R.I.; Heiedia, J.B. Review: Dietary Phenolic Compounds, Health Benefits and Bioaccessibility. Arch. Latinoam. Nutr. 2016, 66, 87–100.
  82. Del Bo, C.; Bernardi, S.; Marino, M.; Porrini, M.; Tucci, M.; Guglielmetti, S.; Cherubini, A.; Carrieri, B.; Kirkup, B.; Kroon, P.; et al. Systematic Review on Polyphenol Intake and Health Outcomes: Is There Sufficient Evidence to Define a Health-Promoting Polyphenol-Rich Dietary Pattern? Nutrients 2019, 11, 1355.
  83. Wang, J.; Ballon, A.; Schroën, K.; de Lamo-Castellví, S.; Ferrando, M.; Güell, C. Polyphenol Loaded W1/O/W2 Emulsions Stabilized with Lesser Mealworm (Alphitobius Diaperinus) Protein Concentrate Produced by Membrane Emulsification: Stability under Simulated Storage, Process, and Digestion Conditions. Foods 2021, 10, 2997.
  84. Zembyla, M.; Murray, B.S.; Sarkar, A. Water-In-Oil Pickering Emulsions Stabilized by Water-Insoluble Polyphenol Crystals. Langmuir 2018, 34, 10001–10011.
  85. Martins, C.; Higaki, N.T.F.; Montrucchio, D.P.; de Oliveira, C.F.; Gomes, M.L.S.; Miguel, M.D.; Miguel, O.G.; Zanin, S.M.W.; Dias, J.d.F.G. Development of W1/O/W2 Emulsion with Gallic Acid in the Internal Aqueous Phase. Food Chem. 2020, 314, 126174.
  86. Pralhad, T.; Rajendrakumar, K. Study of Freeze-Dried Quercetin–Cyclodextrin Binary Systems by DSC, FT-IR, X-Ray Diffraction and SEM Analysis. J. Pharm. Biomed. Anal. 2004, 34, 333–339.
  87. Hunt, L.E.; Bourne, S.A.; Caira, M.R. Inclusion of Hydroxycinnamic Acids in Methylated Cyclodextrins: Host-Guest Interactions and Effects on Guest Thermal Stability. Biomolecules 2021, 11, 45.
  88. Matencio, A.; Hernández-García, S.; García-Carmona, F.; López-Nicolás, J.M. An Integral Study of Cyclodextrins as Solubility Enhancers of α-Methylstilbene, a Resveratrol Analogue. Food Funct. 2017, 8, 270–277.
  89. Nguyen Thi, D.P.; Tran, D.L.; Le Thi, P.; Park, K.D.; Hoang Thi, T.T. Supramolecular Gels Incorporating Cordyline Terminalis Leaf Extract as a Polyphenol Release Scaffold for Biomedical Applications. Int. J. Mol. Sci. 2021, 22, 8759.
  90. Boyacı, D.; Kavur, P.B.; Gulec, S.; Yemenicioğlu, A. Physicochemical and Active Properties of Gelatine-Based Composite Gels Loaded with Lysozyme and Green Tea Polyphenols. Food Technol. Biotechnol. 2021, 59, 337–348.
  91. Stanciauskaite, M.; Marksa, M.; Ivanauskas, L.; Perminaite, K.; Ramanauskiene, K. Ophthalmic in Situ Gels with Balsam Poplar Buds Extract: Formulation, Rheological Characterization, and Quality Evaluation. Pharmaceutics 2021, 13, 953.
  92. Niknam, S.M.; Kashaninejad, M.; Escudero, I.; Sanz, M.T.; Beltrán, S.; Benito, J.M. Preparation of Water-in-Oil Nanoemulsions Loaded with Phenolic-Rich Olive Cake Extract Using Response Surface Methodology Approach. Foods 2022, 11, 279.
  93. Ragelle, H.; Crauste-Manciet, S.; Seguin, J.; Brossard, D.; Scherman, D.; Arnaud, P.; Chabot, G.G. Nanoemulsion Formulation of Fisetin Improves Bioavailability and Antitumour Activity in Mice. Int. J. Pharm. 2012, 427, 452–459.
  94. Rabelo, C.A.S.; Taarji, N.; Khalid, N.; Kobayashi, I.; Nakajima, M.; Neves, M.A. Formulation and Characterization of Water-in-Oil Nanoemulsions Loaded with Açaí Berry Anthocyanins: Insights of Degradation Kinetics and Stability Evaluation of Anthocyanins and Nanoemulsions. Food Res. Int. 2018, 106, 542–548.
  95. Seguin, J.; Brullé, L.; Boyer, R.; Lu, Y.M.; Romano Ramos, M.; Touil, Y.S.; Scherman, D.; Bessodes, M.; Mignet, N.; Chabot, G.G. Liposomal Encapsulation of the Natural Flavonoid Fisetin Improves Bioavailability and Antitumor Efficacy. Int. J. Pharm. 2013, 444, 146–154.
  96. Sultana, F.; Neog, M.K.; Rasool, M.K. Targeted Delivery of Morin, a Dietary Bioflavanol Encapsulated Mannosylated Liposomes to the Macrophages of Adjuvant-Induced Arthritis Rats Inhibits Inflammatory Immune Response and Osteoclastogenesis. Eur. J. Pharm. Biopharm. 2017, 115, 229–242.
  97. Wei, X.Q.; Zhu, J.F.; Wang, X.B.; Ba, K. Improving the Stability of Liposomal Curcumin by Adjusting the Inner Aqueous Chamber PH of Liposomes. ACS Omega 2020, 5, 1120–1126.
  98. Vega, E.N.; Molina, A.K.; Pereira, C.; Dias, M.I.; Heleno, S.A.; Rodrigues, P.; Fernandes, I.P.; Barreiro, M.F.; Stojković, D.; Soković, M.; et al. Anthocyanins from Rubus Fruticosus l. And Morus Nigra l. Applied as Food Colorants: A Natural Alternative. Plants 2021, 10, 1181.
  99. Yang, B.; Dong, Y.; Wang, F.; Zhang, Y. Nanoformulations to Enhance the Bioavailability and Physiological Functions of Polyphenols. Molecules 2020, 25, 4613.
  100. Shaygannia, S.; Eshaghi, M.R.; Fazel, M.; Hashemiravan, M. The Effect of Microencapsulation of Phenolic Compounds from Lemon Waste by Persian and Basil Seed Gums on the Chemical and Microbiological Properties of Mayonnaise. Prev. Nutr. Food Sci. 2021, 26, 82–91.
  101. Desai, N.M.; Haware, D.J.; Basavaraj, K.; Murthy, P.S. Microencapsulation of Antioxidant Phenolic Compounds from Green Coffee. Prep. Biochem. Biotechnol. 2019, 49, 400–406.
  102. de Meneses Costa Ferreira, L.M.; Pereira, R.R.; de Carvalho-Guimarães, F.B.; do Nascimento Remígio, M.S.; Barbosa, W.L.R.; Ribeiro-Costa, R.M.; Silva-Júnior, J.O.C. Microencapsulation by Spray Drying and Antioxidant Activity of Phenolic Compounds from Tucuma Coproduct. Polymers 2022, 14, 2905.
  103. Rosales, T.K.O.; Hassimotto, N.M.A.; Lajolo, F.M.; Fabi, J.P. Nanotechnology as a Tool to Mitigate the Effects of Intestinal Microbiota on Metabolization of Anthocyanins. Antioxidants 2022, 11, 506.
  104. Saifullah, M.; Shishir, M.R.I.; Ferdowsi, R.; Tanver Rahman, M.R.; Van Vuong, Q. Micro and Nano Encapsulation, Retention and Controlled Release of Flavor and Aroma Compounds: A Critical Review. Trends Food Sci. Technol. 2019, 86, 230–251.
  105. Behzadi, S.; Serpooshan, V.; Tao, W.; Hamaly, M.A.; Alkawareek, M.Y.; Dreaden, E.C.; Brown, D.; Alkilany, A.M.; Farokhzad, O.C.; Mahmoudi, M. Cellular Uptake of Nanoparticles: Journey Inside the Cell. Chem. Soc. Rev. 2017, 46, 4218–4244.
  106. Mosquera, J.; García, I.; Liz-Marzán, L.M. Cellular Uptake of Nanoparticles versus Small Molecules: A Matter of Size. Acc. Chem. Res. 2018, 51, 2305–2313.
  107. Rajesh, S.; Lillard, J.W., Jr. Nanoparticle-Based Targeted Drug Delivery. Exp. Mol. Pathol. 2009, 86, 215–223.
  108. Lamprecht, A.; Ubrich, N.; Yamamoto, H.; Schäfer, U.; Takeuchi, H.; Maincent, P.; Kawashima, Y.; Lehr, C.M. Biodegradable Nanoparticles for Targeted Drug Delivery in Treatment of Inflammatory Bowel Disease. J. Pharmacol. Exp. Ther. 2001, 299, 775–781.
  109. Majumder, J.; Taratula, O.; Minko, T. Nanocarrier-Based Systems for Targeted and Site Specific Therapeutic Delivery. Adv. Drug Deliv. Rev. 2019, 144, 57–77.
  110. Street, D.; Bangsbo, J.; Juel, C. Interstitial PH in Human Skeletal Muscle during and after Dynamic Graded Exercise. J. Physiol. 2001, 537, 993–998.
  111. Parisi, O.I.; Puoci, F.; Restuccia, D.; Farina, G.; Iemma, F.; Picci, N. Polyphenols in Human Health and Disease. In Polyphenols in Human Health and Disease; Watson, R.R., Preedy, V.R., Zibadi, S., Eds.; Academic Press-Elsevier Inc.: Amsterdam, The Netherlands, 2014.
  112. Nakayama, T.; Hashimoto, T.; Kajiya, K.; Kumazawa, S. Affinity of Polyphenols for Lipid Bilayers. Biofactors 2000, 13, 147–151.
  113. Phan, H.T.T.; Yoda, T.; Chahal, B.; Morita, M.; Takagi, M.; Vestergaard, M.C. Structure-Dependent Interactions of Polyphenols with a Biomimetic Membrane System. Biochim. Biophys. Acta 2014, 1838, 2670–2677.
  114. Pruchnik, H.; Bonarska-Kujawa, D.; Żyłka, R.; Oszmiański, J.; Kleszczyńska, H. Application of the DSC and Spectroscopy Methods in the Analysis of the Protective Effect of Extracts from the Blueberry Fruit of the Genus Vaccinium in Relation to the Lipid Membrane. J. Therm. Anal. Calorim. 2018, 134, 679–689.
  115. Malekar, S.A.; Sarode, A.L.; Bach, A.C.; Worthen, D.R. The Localization of Phenolic Compounds in Liposomal Bilayers and Their Effects on Surface Characteristics and Colloidal Stability. AAPS PharmSciTech 2016, 17, 1468–1476.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : ,
View Times: 465
Revisions: 2 times (View History)
Update Date: 10 Jan 2023
1000/1000
Video Production Service