Elsevier

Food Chemistry

Volume 278, 25 April 2019, Pages 443-451
Food Chemistry

Stereochemistry and glycosidic linkages of C3-glycosylations affected the reactivity of cyanidin derivatives

https://doi.org/10.1016/j.foodchem.2018.11.076Get rights and content

Highlights

  • Stereochemistry and glycosidic linkage of cyanidin-3-glycosides impacted reactivity.

  • Cyanidin-3-galactoside bleached & hydrated more than cyanidin-3-glucoside.

  • Cyanidin (Cy) with 1 → 6 disaccharides had higher t1/2 than with 1 → 2 dissacharides.

  • 1 → 2 disaccharides resisted hydration & bleaching more than their 1 → 6 counterparts.

  • Cy with 1 → 2 & 1 → 6 trisaccharide showed greatest stability & intermediate bleaching.

Abstract

The impact of glycosylation on anthocyanin stability has largely been associated with sugar type, site, and size, with glycosyl stereochemistry being under-explored. Seven cyanidin-3-glycosides were isolated by HPLC, diluted in pH 1–9, mixed with bisulfite or ascorbic acid at pH 3, and stored for 8 weeks (25 °C, dark). Spectral changes, half-lives, and bleaching rates were determined. Cyanidin-3-galactoside was more reactive (susceptible to hydration and bleaching) than cyanidin-3-glucoside. The 1 → 2 disaccharides exhibited greater λvis-max (≤16 nm), resistance to hydration, and bleaching compared to 1 → 6 disaccharides.The 1 → 6 disaccharides had similar λvis-max (∼2 nm) to the monosaccharides but slightly improved resistance to hydration and bleaching. The tri-glycosylated anthocyanin had the greatest stability and its spectral and bleaching characteristics was intermediate to 1 → 2 and 1 → 6 disaccharides. The 1 → 2 disaccharides generally exhibited lower half-lives compared to monosaccharides; whereas, 1 → 6 disaccharides exhibited higher stability. These findings highlight the role of glycosyl assembly on anthocyanin reactivity and stability.

Introduction

The reactivity and stability of anthocyanins are known to be affected by the natural chemical composition of the pigment. Anthocyanins are composed of a basic structure of 2-phenylbenzopyrylium of flavylium salt, with varying degrees of substitution. Anthocyanidins are inherently unstable and are primarily found glycosylated in nature; the glycosyl addition is believed to support anthocyanin stability through several mechanisms such as formation of H-bonding, formation of van der Waals interactions with the chromophore while exhibiting a hydrophobic face, enhanced streric hindrance, and decreased electron delocalization (Cabrita et al., 2014, Jing et al., 2008, Zhao et al., 2014). The most common glycosylation site is C3, followed by C3 and C5, and least common are C3 with either C4 or C7 substitution. Anthocyanins are typically mono-glycosylated, but attachments as large as tri-saccharides have been reported (Andersen & Jordheim, 2010).

It is frequently reported that increasing the number of glycosyls at C3 results in increased stability of the molecules (Zhao et al., 2014). Several sources have provided evidence of the enhanced stability of 3-trisaccharide anthocyanins in comparison to 3-monosaccharides (Bonerz et al., 2007, Eiro and Heinonen, 2002, García-Viguera et al., 1998, Sui et al., 2014). Evidence regarding stability for anthocyanin 3-disaccharides, however, is generally more conflicting. For example, cyanidin-3-rutinoside (rhamnosyl(1 → 6)glucoside) was reported as being more stable to heat treatment as compared to trisaccharide derivative cyanidin-3-glucosyl(1 → 2)-rutinoside in sour cherry juice (Szalóki-Dorkó, Végvári, Ladányi, Ficzek, & Stéger-Máté, 2015) and also cyanidin-3-glucoside (Sui et al., 2014). In contrast, cyanidin-3-xylosyl(1 → 2)-rutinoside was reported as being less stable than cyanidin-3-rutinoside and cyanidin-3-sophoroside (glucosyl(1 → 2)glucoside) (Bonerz et al., 2007). In the same study mentioned by Bonerz et al. (2007), cyanidin-3-glucosyl(1 → 2)-rutinoside always exhibited superior stability to the disaccharides present in sour cherry juice even when cyanidin-3-xylosyl(1 → 2)-rutinoside did not. The differences in stability among C3 di- and tri-saccharides may be influenced by primary sugar type attached to the anthocyanidin or how subsequent sugars are bonded in the glycosylations. Susceptibility to acid hydrolysis of anthocyanin monosaccharides has been reported in the following order arabinoside > galactoside > glucoside (Ichiyanagi, Oikawa, Tateyama, & Konishi, 2001).

More recently, it was reported that how the molecules are assembled, primarily the location of glycosidic bonds, may influence the formation of pyranoanthocyanins; 1 → 6 disaccharides were reported to produce higher yields than monosaccharides and 1 → 2 disaccharides (Farr, Sigurdson, & Giusti, 2018). These findings suggest that the type of sugar attachment, their stereochemical organization around the chromophore, and thus location of glycosidic linkages within the di- or tri-saccharide all may affect anthocyanin stability and reactivity. Similarly these differences in glycosidic linkages may play roles in anthocyanin stability and reactivity; subsequent investigation could contribute further to understanding anthocyanin chemistry and ameliorate previously reported unexpected findings regarding di and tri-saccharide anthocyanin-3-glycosides.

Beyond being innately reactive, anthocyanins have been found to bleach (decrease in Chroma, increase in Lightness) in response to certain chemical stimuli. Bleaching is thought to occur as a result of nucleophilic species, such as bisulfite, hydrogen peroxide, or ascorbic acid, attacking the electrophilic sites of the aglycone. There is evidence specifically in regards to bisulfite addition suggesting these types of reactions are intitated with condensation between the reactive species and anthocyanin primarily at Carbon-4 and and less preferentially at Carbon-2 of the aglycone structure (Berké, Chèze, Vercauteren, & Deffieux, 1998). Bleaching behaviors are often investigated in regards to a plant source such as an anthocyanin mixture present in a juice, lacking a systematic comparison of the influence of the 3-substituted glycosylation. The differing number or stereochemical arrangement of glycosyl units could influence the extent and rate of bleaching of anthocyanins.

The objective of this study was to evaluate the possible effects of spatial organizations of glycosylations on the stability and reactivity of cyanidin derivatives. Significant differences in reactivity of anthocyanins bearing 1 → 2 and 1 → 6 disaccharides was hypothesized to occur with 1 → 6 anthocyanins being less stable and more prone to bleaching as suggested by pyranoanthocyanin yield as previously discovered.

Section snippets

Materials

Seven cyanidin-3-glycosides were obtained from the following plant materials: Cyanidin-3-glucoside from blackberry (Rubus sp.); cyanidin-3-galactoside from chokeberry (Aronia melanocarpa); cyanidin-3-xylosyl(1 → 2)galactoside, cyanidin-3-glucosyl(1 → 6)galactoside, and cyanidin-3-xylosyl(1 → 2)glucosyl(1 → 6)galactoside from black carrot (Daucus carota L.); cyanidin-3-rhamnosyl(1 → 6)glucoside (common name: cyanidin-3-rutinoside) from mulberry (Morus nigra); and

Initial λvis-max differences at pH 1–9

The glycosidic subsitution patterns of o-dihydroxylated anthocyanins have previously been demonstrated to uniquely impact their spectral and colorimetric properities, particularly when comparing 3-glycosylation to 3,5-glycosylation patterns (Sigurdson et al., 2018, Stintzing et al., 2002, Torskangerpoll and Andersen, 2005). To further explore the role of the stereochemistry of C3 glycosylations on the spectral characteristics of cyanidin derivatives, the absorbance spectra of the isolates were

Conclusion

Reactivity of mono-, di, and tri-substituted cyanidin-3 isolates were found to be impacted by the stereochemical properties or location of glycosidic linkages within the glycosyl moiety. The difference in the orientation of the hydroxyl group on the C4 of monosaccharides (axial vs. equatorial) influenced reactivity observed by cyanidin-3-galactoside bleaching and hydrating to a greater extent than cyanidin-3-glucoside. The position of the glycosidic bond (1 → 2 vs. 1 → 6) on disaccharides also

Declaration of interests

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

The authors declare the following financial interests/personal relationships which may be considered as potential competing interests.

Acknowledgements

This work was supported in part by the USDA National Institute of Food and Agriculture, 455 Hatch Project OHO01423, Accession Number 1014136.

Conflicts of Interest

No conflicts of interest known.

References (37)

  • K. Torskangerpoll et al.

    Colour stability of anthocyanins in aqueous solutions at various pH values

    Food Chemistry

    (2005)
  • Ø.M. Andersen et al.

    Anthocyanins

  • D. Bonerz et al.

    Analytical characterization and the impact of ageing on anthocyanin composition and degradation in juices from five sour cherry cultivars

    European Food Research and Technology

    (2007)
  • R. Brouillard et al.

    Chemistry of anthocyanin pigments. 6. Kinetic and thermodynamic study of hydrogen sulfite addition to cyanin. Formation of a highly stable Meisenheimer-type adduct derived from a 2-phenylbenzopyrylium salt

    Journal of the American Chemical Society

    (1980)
  • L. Cabrita et al.

    On the thermal degradation of anthocyanidins: Cyanidin

    RSC Advances

    (2014)
  • V. de Freitas et al.

    Chemical transformations of anthocyanins yielding a variety of colours (Review)

    Environmental Chemistry Letters

    (2006)
  • M.J. Eiro et al.

    Anthocyanin color behavior and stability during storage: Effect of intermolecular copigmentation

    Journal of Agricultural and Food Chemistry

    (2002)
  • G. Elham et al.

    Isolation and structure charactisation of anthocyanin pigments in black carrot (Daucus carota L.)

    Pakistan Journal of Biological Sciences

    (2006)
  • Cited by (11)

    • Effects of organic acids on color intensification, thermodynamics, and copigmentation interactions with anthocyanins

      2022, Food Chemistry
      Citation Excerpt :

      Owing to this degradation, the color of anthocyanins can fade, and they loses bioactivity during food processing; therefore, their applications are limited. A variety of techniques including copigmentation (Qian, Liu, Zhao, Cai, & Jing, 2017), encapsulation (Lyu et al., 2021), anabolism (Shan et al., 2019), and acylation (Farr, Sigurdson, & Giusti, 2019) have been used for anthocyanin stabilization. Copigmentation is an efficient method for enhancing the stability of anthocyanins by combining electron-rich π systems containing copigments with the electron-insufficient flavylium cation to form complexes (Bąkowska, Kucharska, & Oszmiański, 2003).

    • A comprehensive review on innovative and advanced stabilization approaches of anthocyanin by modifying structure and controlling environmental factors

      2022, Food Chemistry
      Citation Excerpt :

      Acylation is another effective way to maintain the stability of anthocyanins and the intensity of colour. Acylation on core structure of anthocyanin can hide the active sites to improve stability even for the probability of anthocyanins passes through the double-layer phospholipid biofilm (Farr et al., 2019). Generally, organic acids can be linked and clustered to the glycosidic bonds by acylation.

    • Alterations of natural pigments

      2020, Chemical Changes During Processing and Storage of Foods: Implications for Food Quality and Human Health
    • Effects of sucrose and copigment sources on the major anthocyanins isolated from sour cherries

      2019, Food Chemistry
      Citation Excerpt :

      Furthermore, Eiro and Heinonen (2002) investigated the copigmentation tendency of cyanidin-3-xylosyl(1 → 2)glucosyl(1 → 6)galactoside, while we investigated that of cyanidin-3-rhamnosyl(1 → 6)glucosyl(1 → 2)glucoside. According to the results of the study conducted by Farr et al. (2019), the reactivity of “glucosyl(1 → 2)glucoside” was lower than that of “glucosyl(1 → 6)galactoside”. However, the reactivity of “xylosyl(1 → 2)glucosyl(1 → 6)galactoside” was more lower than “glucosyl(1 → 2)glucoside”.

    View all citing articles on Scopus
    View full text