Procyanidins are members of the proanthocyanidin (or condensed tannins) class of flavonoids. They are oligomeric compounds, formed from catechin and epicatechin molecules. They yield cyanidin when depolymerized under oxidative conditions.

Epicatechin (EC), one of the building blocks of procyanidins
Cyanidin, the anthocyanidin produced when procyanidin are depolymerized under oxidative conditions

See the box below entitled "Types of procyanidins" for links to articles on the various types.

Distribution in plants edit

Procyanidins, including the lesser bioactive / bioavailable polymers (4 or more catechines), represent a group of condensed flavan-3-ols that can be found in many plants, most notably apples, maritime pine bark, cinnamon, aronia fruit, cocoa beans, grape seed, grape skin,[1] and red wines of Vitis vinifera (the common grape).[2] However, bilberry, cranberry, black currant, green tea, black tea, and other plants also contain these flavonoids.[3] Procyanidins can also be isolated from Quercus petraea and Q. robur heartwood (wine barrel oaks).[4] Açaí oil, obtained from the fruit of the açaí palm (Euterpe oleracea), is rich in numerous procyanidin oligomers.[5]

Apples contain on average per serving about eight times the amount of procyanidin found in wine, with some of the highest amounts found in the Red Delicious and Granny Smith varieties.[6]

The seed testas of field beans (Vicia faba) contain procyanidins[7] that affect the digestibility in piglets[8] and could have an inhibitory activity on enzymes.[9] Cistus salviifolius also contains oligomeric procyanidins.[10]

Analysis edit

Condensed tannins can be characterised by a number of techniques including depolymerisation, asymmetric flow field flow fractionation or small-angle X-ray scattering. DMACA is a dye used for localization of procyanidin compounds in plant histology. The use of the reagent results in blue staining.[11] It can also be used to titrate procyanidins. Total phenols (or antioxidant effect) can be measured using the Folin-Ciocalteu reaction. Results are typically expressed as gallic acid equivalents (GAE).

Procyanidins from field beans (Vicia faba)[12] or barley[13] have been estimated using the vanillin-HCl method, resulting in a red color of the test in the presence of catechin or proanthocyanidins.

Procyanidins can be titrated using the Procyanidolic Index (also called the Bates-Smith Assay). It is a testing method that measures the change in color when the product is mixed with certain chemicals. The greater the color changes, the higher the PCOs content is. However, the Procyanidolic Index is a relative value that can measure well over 100. Unfortunately, a Procyanidolic Index of 95 was erroneously taken to mean 95% PCO by some and began appearing on the labels of finished products. All current methods of analysis suggest that the actual PCO content of these products is much lower than 95%.[14][unreliable medical source?]

An improved colorimetric test, called the Porter Assay or butanol-HCl-iron method, is the most common PCO assay currently in use.[15][self-published source?] The unit of measurement of the Porter Assay is the PVU (Porter Value Unit). The Porter Assay is a chemical test to help determine the potency of procyanidin containing compounds, such as grape seed extract. It is an acid hydrolysis, which splits larger chain units (dimers and trimers) into single unit monomers and oxidizes them. This leads to a colour change, which can be measured using a spectrophotometer. The greater the absorbance at a certain wavelength of light, the greater the potency. Ranges for grape seed extract are from 25 PVU for low grade material to over 300 for premium grape seed extracts.[16][unreliable medical source?]

Gel permeation chromatography (GPC) analysis allows to separate monomers from larger PCO molecules.

Monomers of procyanidins can be characterized by HPLC analysis. Condensed tannins can undergo acid-catalyzed cleavage in the presence of a nucleophile like phloroglucinol (reaction called phloroglucinolysis), thioglycolic acid (thioglycolysis), benzyl mercaptan or cysteamine (processes called thiolysis[17]) leading to the formation of oligomers that can be further analyzed.[18]

Phloroglucinolysis can be used for instance for procyanidins characterisation in wine[19] or in the grape seed and skin tissues.[20]

Thioglycolysis can be used to study procyanidins[21] or the oxidation of condensed tannins.[22] It is also used for lignin quantitation.[23] Reaction on condensed tannins from Douglas fir bark produces epicatechin and catechin thioglycolates.[24]

Condensed tannins from Lithocarpus glaber leaves have been analysed through acid-catalyzed degradation in the presence of cysteamine.[25]

Research edit

Procyanidin content in dietary supplements has not been well documented.[26] Pycnogenol is a dietary supplement derived from extracts from maritime pine bark that contains 70% procyanidins, and is marketed with claims it can treat many conditions. The medical evidence is insufficient to support its use for the treatment of seven different chronic disorders.[27]

See also edit

References edit

  1. ^ Souquet, J; Cheynier, Véronique; Brossaud, Franck; Moutounet, Michel (1996). "Polymeric proanthocyanidins from grape skins". Phytochemistry. 43 (2): 509–512. doi:10.1016/0031-9422(96)00301-9.
  2. ^ Yang, J; Xiao, YY (2013). "Grape phytochemicals and associated health benefits". Crit Rev Food Sci Nutr. 53 (11): 1202–1225. doi:10.1080/10408398.2012.692408. PMID 24007424. S2CID 19094002.
  3. ^ USDA, August 2004. USDA Database for the Proanthocyanidin Content of Selected Foods. PDF summary accessed from main USDA page here. Page accessed July 31, 2015
  4. ^ Vivas, N; Nonier, M; Pianet, I; Vivasdegaulejac, N; Fouquet, E (2006). "Proanthocyanidins from Quercus petraea and Q. robur heartwood: quantification and structures". Comptes Rendus Chimie. 9: 120–126. doi:10.1016/j.crci.2005.09.001.
  5. ^ Pacheco-Palencia LA, Mertens-Talcott S, Talcott ST (Jun 2008). "Chemical composition, antioxidant properties, and thermal stability of a phytochemical enriched oil from Acai (Euterpe oleracea Mart.)". J Agric Food Chem. 56 (12): 4631–4636. doi:10.1021/jf800161u. PMID 18522407.
  6. ^ Hammerstone, John F.; Lazarus, Sheryl A.; Schmitz, Harold H. (August 2000). "Procyanidin content and variation in some commonly consumed foods". The Journal of Nutrition. 130 (8S Suppl): 2086S–92S. doi:10.1093/jn/130.8.2086S. PMID 10917927. Figure 5
  7. ^ Merghem, R.; Jay, M.; Brun, N.; Voirin, B. (2004). "Qualitative analysis and HPLC isolation and identification of procyanidins fromvicia faba". Phytochemical Analysis. 15 (2): 95–99. doi:10.1002/pca.731. PMID 15116939.
  8. ^ Van Der Poel, A. F. B.; Dellaert, L. M. W.; Van Norel, A.; Helsper, J. P. F. G. (2007). "The digestibility in piglets of faba bean (Vicia faba L.) as affected by breeding towards the absence of condensed tannins". British Journal of Nutrition. 68 (3): 793–800. doi:10.1079/BJN19920134. PMID 1493141.
  9. ^ Griffiths, D. W. (1981). "The polyphenolic content and enzyme inhibitory activity of testas from bean (Vicia faba) and pea (Pisum spp.) varieties". Journal of the Science of Food and Agriculture. 32 (8): 797–804. doi:10.1002/jsfa.2740320808.
  10. ^ Qa’Dan, F.; Petereit, F.; Mansoor, K.; Nahrstedt, A. (2006). "Antioxidant oligomeric proanthocyanidins fromCistus salvifolius". Natural Product Research. 20 (13): 1216–1224. doi:10.1080/14786410600899225. PMID 17127512. S2CID 254865.
  11. ^ Bogs, J.; Jaffe, F. W.; Takos, A. M.; Walker, A. R.; Robinson, S. P. (2007). "The Grapevine Transcription Factor VvMYBPA1 Regulates Proanthocyanidin Synthesis during Fruit Development". Plant Physiology. 143 (3): 1347–1361. doi:10.1104/pp.106.093203. PMC 1820911. PMID 17208963.
  12. ^ Cabrera, A.; Martin, A. (2009). "Genetics of tannin content and its relationship with flower and testa colours in Vicia faba". The Journal of Agricultural Science. 113: 93–98. doi:10.1017/S0021859600084665. S2CID 84456901.
  13. ^ Kristensen, H.; Aastrup, S. (1986). "A non-destructive screening method for proanthocyanidin-free barley mutants". Carlsberg Research Communications. 51 (7): 509–513. doi:10.1007/BF02906893.
  14. ^ Grape Seed Extract, White paper, The Grape Seed Method Evaluation Committee, Under the Auspices of NNFA ComPli Archived 2002-12-24 at the Wayback Machine
  15. ^ The Truth About PCOs, Debasis Bagchi, Ph.D. on www.activin.com Archived 2013-10-19 at the Wayback Machine
  16. ^ Porter Assay on www.omegabiotech.com
  17. ^ Torres, J. L.; Lozano, C. (2001). "Chromatographic characterization of proanthocyanidins after thiolysis with cysteamine". Chromatographia. 54 (7–8): 523–526. doi:10.1007/BF02491211. S2CID 95355684.
  18. ^ Jorgensen, Emily M.; Marin, Anna B.; Kennedy, James A. (2004). "Analysis of the Oxidative Degradation of Proanthocyanidins under Basic Conditions". Journal of Agricultural and Food Chemistry. 52 (8): 2292–2296. doi:10.1021/jf035311i. PMID 15080635.
  19. ^ Analysis of Tannins in Red Wine Using Multiple Methods: Correlation with Perceived Astringency by mean of depolymerisation James A. Kennedy, Jordan Ferrier, James F. Harbertson and Catherine Peyrot des Gachons, Am. J. Enol. Vitic. 57:4, 2006, pp. 481–485
  20. ^ Kennedy, J. A.; Jones, G. P. (2001). "Analysis of Proanthocyanidin Cleavage Products Following Acid-Catalysis in the Presence of Excess Phloroglucinol". Journal of Agricultural and Food Chemistry. 49 (4): 1740–1746. doi:10.1021/jf001030o. PMID 11308320.
  21. ^ Sears, K. D.; Casebier, R. L. (1968). "Cleavage of proanthocyanidins with thioglycollic acid". Chemical Communications (22): 1437. doi:10.1039/C19680001437.
  22. ^ Vernhet, A.; Dubascoux, S. P.; Cabane, B.; Fulcrand, H. L. N.; Dubreucq, E.; Poncet-Legrand, C. L. (2011). "Characterization of oxidized tannins: Comparison of depolymerization methods, asymmetric flow field-flow fractionation and small-angle X-ray scattering". Analytical and Bioanalytical Chemistry. 401 (5): 1559–1569. doi:10.1007/s00216-011-5076-2. PMID 21573842. S2CID 4645218. Vernhet, A.; Dubascoux, S. P.; Cabane, B.; Fulcrand, H. L. N.; Dubreucq, E.; Poncet-Legrand, C. L. (2011). "Characterization of oxidized tannins: Comparison of depolymerization methods, asymmetric flow field-flow fractionation and small-angle X-ray scattering". Analytical and Bioanalytical Chemistry. 401 (5): 1559–1569. doi:10.1007/s00216-011-5076-2. PMID 21573842. S2CID 4645218.
  23. ^ Lange, B. M.; Lapierre, C.; Sandermann Jr, H. (1995). "Elicitor-Induced Spruce Stress Lignin (Structural Similarity to Early Developmental Lignins)". Plant Physiology. 108 (3): 1277–1287. doi:10.1104/pp.108.3.1277. PMC 157483. PMID 12228544.
  24. ^ Douglas-Fir Bark: Characterization of a Condensed Tannin Extract, by Hong-Keun Song, A thesis submitted to Oregon State University in partial fulfillment of the requirements for the degree of Master of Science, December 13, 1984
  25. ^ Zhang, L. L.; Lin, Y. M. (2008). "HPLC, NMR and MALDI-TOF MS Analysis of Condensed Tannins from Lithocarpus glaber Leaves with Potent Free Radical Scavenging Activity". Molecules. 13 (12): 2986–2997. doi:10.3390/molecules13122986. PMC 6245341. PMID 19052523.
  26. ^ USDA Database for the Proanthocyanidin Content of Selected Foods (PDF). August 2004. Archived from the original (PDF) on 2010-10-16. Retrieved 2013-01-09.[page needed]
  27. ^ Robertson, Nina U.; Schoonees, Anel; Brand, Amanda; Visser, Janicke (29 September 2020). "Pine bark (Pinus spp.) extract for treating chronic disorders". The Cochrane Database of Systematic Reviews. 2020 (9): CD008294. doi:10.1002/14651858.CD008294.pub5. ISSN 1469-493X. PMC 8094515. PMID 32990945.

External links edit