全部 标题 作者
关键词 摘要

OALib Journal期刊
ISSN: 2333-9721
费用:99美元

查看量下载量

相关文章

更多...

Action of Chicory Fructooligosaccharides on Biomimetic Membranes

DOI: 10.1155/2014/186109

Full-Text   Cite this paper   Add to My Lib

Abstract:

Fructooligosaccharides from chicory (FOSC) are functional prebiotic foods recognized to exert several well-being effects in human health and animal production, as decreasing blood lipids, modulating the gut immune system, enhancing mineral bioavailability, and inhibiting microbial growth, among others. Mechanisms of actions directly on cell metabolism and structure are however little known. In this sense this work was targeted to investigate the interaction of FOSC with biomimetic membranes (liposomes and supported bilayer membrane; s-BLM) through cyclic voltammetry, impedance spectroscopy, spectrofluorimetry, and microscopy. FOSC was able to disrupt the membrane structure of liposomes and s-BLM from the onset of molecular pores induced on it. The mechanism of interaction of fructans with biomimetic membranes suggests hydrogen bonding between the polyhydroxylated structure of the oligosaccharides and the negative polar group of L--phosphatidylcholine (PC) present in both liposomes and s-BLM. 1. Introduction As stated almost two decades ago, prebiotics are considered a nondigestible food ingredient that beneficially affects the host by selectively stimulating the growth and/or activity of one or a limited number of bacteria in the colon and thus improving host health [1]. A prebiotic group largely studied is the chicory fructooligosaccharides (FOSC). FOSC are fructans (carbohydrates with a great extent of fructosylfructose links) extracted on a commercial basis from the chicory root, namely, the Compositae family (Cichorium intybus) [2]. FOSC are also present in several fruits and vegetables species and are produced by transfructosylation of sucrose. These compounds comprise a functional food group containing mixed -D-fructans with two to four (2-1) linked fructosyl units displaying a terminal -D-glucose residue, as kestose, nystose, fructosylnystose, and fructofuranosylnystose, among others [3]. FOSC differs from inulin, a well-known fructan of a high degree of polymerization (DP), as well as oligofructose, a small FOS (DP about 5) produced during endoglycolitic hydrolysis of inulin. FOSC are considered to arrive the human gastrointestinal tract almost without hydrolysis, being a carbon source for short-chain fatty acids by bifidobacteria and lactobacilli living into the lumen [4]. Some properties of FOSC, as their nondigestible and fermentable nature, as well as their sweetening power and low caloric value, make them attractive to be used in pastry, confectionery, and dairy industries [5]. Moreover, both FOSC and inulin are also known as uniquely (2-1)

References

[1]  G. R. Gibson and M. B. Roberfroid, “Dietary modulation of the human colonic microbiota: introducing the concept of prebiotics,” Journal of Nutrition, vol. 125, no. 6, pp. 1401–1412, 1995.
[2]  M. B. Roberfroid, “Chicory fructooligosaccharides and the gastrointestinal tract,” Nutrition, vol. 16, no. 7-8, pp. 677–679, 2000.
[3]  E. Biedrzycka and M. Bielecka, “Prebiotic effectiveness of fructans of different degrees of polymerization,” Trends in Food Science & Technology, vol. 15, no. 3-4, pp. 170–175, 2004.
[4]  T. Fukasawa, A. Kamei, Y. Watanabe, J. Koga, and K. Abe, “Short-chain fructooligosaccharide regulates hepatic peroxisome proliferator-activated receptor α and farnesoid X receptor target gene expression in Rats,” Journal of Agricultural and Food Chemistry, vol. 58, no. 11, pp. 7007–7012, 2010.
[5]  F. R. J. Bornet, F. Brouns, Y. Tashiro, and V. Duvillier, “Nutritional aspects of short-chain fructooligosaccharides: natural occurrence, chemistry, physiology and health implications,” Digestive and Liver Disease, vol. 34, supplement 2, pp. S111–S120, 2002.
[6]  A. Nzeusseu, D. Dienst, V. Haufroid, G. Depresseux, J.-P. Devogelaer, and D.-H. Manicourt, “Inulin and fructo-oligosaccharides differ in their ability to enhance the density of cancellous and cortical bone in the axial and peripheral skeleton of growing rats,” Bone, vol. 38, no. 3, pp. 394–399, 2006.
[7]  W. van den Ende, D. Peshev, and L. de Gara, “Disease prevention by natural antioxidants and prebiotics acting as ROS scavengers in the gastrointestinal tract,” Trends in Food Science and Technology, vol. 22, no. 12, pp. 689–697, 2011.
[8]  N. Matsukawa, M. Matsumoto, A. Shinoki, M. Hagio, R. Inoue, and H. Hara, “Nondigestible saccharides suppress the bacterial degradation of quercetin aglycone in the large intestine and enhance the bioavailability of quercetin glucoside in rats,” Journal of Agricultural and Food Chemistry, vol. 57, no. 20, pp. 9462–9468, 2009.
[9]  J. M. Laparra and Y. Sanz, “Interactions of gut microbiota with functional food components and nutraceuticals,” Pharmacological Research, vol. 61, no. 3, pp. 219–225, 2010.
[10]  T. Suzuki and H. Hara, “Various nondigestible saccharides open a paracellular calcium transport pathway with the induction of intracellular calcium signaling in human intestinal Caco-2 cells,” The Journal of Nutrition, vol. 134, no. 8, pp. 1935–1941, 2004.
[11]  T. Fukasawa, K. Murashima, T. Nemoto, et al., “Identification of marker genes for lipid-lowering effect of a short-chain fructooligosaccharide by DNA microarray analysis,” Journal of Dietary Supplements, vol. 6, no. 3, pp. 254–262, 2009.
[12]  W. Huang, Z. Zhang, X. Han et al., “Concentration-dependent behavior of nisin interaction with supported bilayer lipid membrane,” Biophysical Chemistry, vol. 99, no. 3, pp. 271–279, 2002.
[13]  Y. Ma, J. Wang, F. Hui, and S. Zang, “The reassembled behavior of bilayer lipid membranes supported by Pt electrode,” Journal of Membrane Science, vol. 286, no. 1-2, pp. 174–179, 2006.
[14]  X. Lu, T. Liao, L. Ding et al., “Interaction of quercetin with supported bilayer lipid membranes on glassy carbon electrode,” International Journal of Electrochemical Science, vol. 3, no. 7, pp. 797–805, 2008.
[15]  D. Jiang, P. Diao, R. Tong, D. Gu, and B. Zhong, “Ca2+ induced electron transfer at Pt supported BLM electrode,” Bioelectrochemistry and Bioenergetics, vol. 44, no. 2, pp. 285–288, 1998.
[16]  S. Basu and M. Basu, Liposome Methods and Protocols, Methods in Molecular Biology, Humana Press, New York, NY, USA, 2002.
[17]  H. Y. Wang, Y. Sun, and B. Tang, “Study on fluorescence property of dopamine and determination of dopamine by fluorimetry,” Talanta, vol. 57, no. 5, pp. 899–907, 2002.
[18]  R Development Core Team, R: A Language and Environment for Statistical Computing, R Foundation for Statisti cal Computing, Vienna, Austria, 2012.
[19]  J. J. Harris and M. L. Bruening, “Electrochemical and in situ ellipsometric investigation of the permeability and stability of layered polyelectrolyte films,” Langmuir, vol. 16, no. 4, pp. 2006–2013, 2000.
[20]  P. Diao, D. Jiang, X. Cui, D. Gu, R. Tong, and B. Zhong, “Cyclic voltammetry and a.c. impedance studies of Ca2+-induced ion channels on Pt-BLM,” Bioelectrochemistry and Bioenergetics, vol. 45, no. 2, pp. 173–179, 1998.
[21]  D. Pan, J. Chen, W. Tao, L. Nie, and S. Yao, “Phosphopolyoxomolybdate absorbed on lipid membranes/carbon nanotube electrode,” Journal of Electroanalytical Chemistry, vol. 579, no. 1, pp. 77–82, 2005.
[22]  N. Wilke and A. M. Baruzzi, “Comparative analysis of the charge transfer processes of the Ru(NH3)63+/Ru(NH3)62+ and Fe(CN)63?/Fe(CN)64? redox couples on glassy carbon electrodes modified by different lipid layers,” Journal of Electroanalytical Chemistry, vol. 537, no. 1-2, pp. 67–76, 2002.
[23]  J. Wang, L. Wang, S. Liu, X. Han, W. Huang, and E. Wang, “Interaction of K7Fe3+P2W17O62H2 with supported bilayer lipid membranes on platinum electrode,” Biophysical Chemistry, vol. 106, no. 1, pp. 31–38, 2003.
[24]  X. Liu, W. Huang, and E. Wang, “An electrochemical study on the interaction of surfactin with a supported bilayer lipid membrane on a glassy carbon electrode,” Journal of Electroanalytical Chemistry, vol. 577, no. 2, pp. 349–354, 2005.
[25]  L. Du, X. Liu, W. Huang, and E. Wang, “A study on the interaction between ibuprofen and bilayer lipid membrane,” Electrochimica Acta, vol. 51, no. 26, pp. 5754–5760, 2006.
[26]  J. Wang, B. Zeng, C. Fang, and X. Zhou, “Influence of surfactants on the electron-transfer reaction at self-assembled thiol monolayers modifying a gold electrode,” Journal of Electroanalytical Chemistry, vol. 484, no. 1, pp. 88–92, 2000.
[27]  J.-S. Ye, A. Ottova, H. T. Tien, and F.-S. Sheu, “Nanostructured platinum-lipid bilayer composite as biosensor,” Bioelectrochemistry, vol. 59, no. 1-2, pp. 65–72, 2003.
[28]  G. Favero, A. D'Annibale, L. Campanella, R. Santucci, and T. Ferri, “Membrane supported bilayer lipid membranes array: preparation, stability and ion-channel insertion,” Analytica Chimica Acta, vol. 460, no. 1, pp. 23–34, 2002.
[29]  K. Asaka, H. Ti Tien, and A. Ottova, “Voltammetric study of charge transfer across supported bilayer lipid membranes (s-BLMs),” Journal of Biochemical and Biophysical Methods, vol. 40, no. 1-2, pp. 27–37, 1999.
[30]  V. Kochev and M. Karabaliev, “Wetting films of lipids in the development of sensitive interfaces. An electrochemical approach,” Advances in Colloid and Interface Science, vol. 107, no. 1, pp. 9–26, 2004.
[31]  C. G. Zoski, Handbook of Electrochemistry, vol. 5, Elsevier, New York, NY, USA, 2007.
[32]  H. Haas, G. Lamura, and A. Gliozzi, “Improvement of the quality of self assembled bilayer lipid membranes by using a negative potential,” Bioelectrochemistry, vol. 54, no. 1, pp. 1–10, 2001.
[33]  C. Amatore, J. M. Savéant, and D. Tessier, “Charge transfer at partially blocked surfaces. A model for the case of microscopic active and inactive sites,” Journal of Electroanalytical Chemistry, vol. 147, no. 1-2, pp. 39–51, 1983.
[34]  N. Yang, Q. Wan, and X. Wang, “Voltammetry of Vitamin B12 on a thin self-assembled monolayer modified electrode,” Electrochimica Acta, vol. 50, no. 11, pp. 2175–2180, 2005.
[35]  H. Sato, H. Hakamada, Y. Yamazaki, M. Uto, M. Sugawara, and Y. Umezawa, “Ionophore incorporated bilayer lipid membranes that selectively respond to metal ions and induce membrane permeability changes,” Biosensors and Bioelectronics, vol. 13, no. 9, pp. 1035–1046, 1998.
[36]  W. Huang, Z. Zhang, X. Han et al., “Ion channel behavior of Amphotericin B in sterol-free and cholesterol- or ergosterol-containing supported phosphatidylcholine bilayer model membranes investigated by electrochemistry and spectroscopy,” Biophysical Journal, vol. 83, no. 6, pp. 3245–3255, 2002.
[37]  S. Schreier, S. V. P. Malheiros, and E. de Paula, “Surface active drugs: self-association and interaction with membranes and surfactants. Physicochemical and biological aspects,” Biochimica et Biophysica Acta: Biomembranes, vol. 1508, no. 1-2, pp. 210–234, 2000.
[38]  X. Liu, H. Bai, W. Huang, L. Du, X. Yang, and E. Wang, “Concentration and time dependant behavior of chlorpromazine interaction with supported bilayer lipid membrane,” Electrochimica Acta, vol. 51, no. 12, pp. 2512–2517, 2006.
[39]  Z. Oren and Y. Shai, “Mode of action of linear amphipathic α-helical antimicrobial peptides,” Biopolymers, vol. 47, no. 6, pp. 451–463, 1998.
[40]  L. M. Crowe, J. H. Crowe, and D. Chapman, “Interaction of carbohydrates with dry dipalmitoylphosphatidylcholine,” Archives of Biochemistry and Biophysics, vol. 236, no. 1, pp. 289–296, 1985.
[41]  J. Grdadolnik and D. Had?i, “FT infrared and Raman investigation of saccharide-phosphatidylcholine interactions using novel structure probes,” Spectrochimica Acta A: Molecular and Biomolecular Spectroscopy, vol. 54, no. 12, pp. 1989–2000, 1998.
[42]  L. Paasonen, T. Laaksonen, C. Johans, M. Yliperttula, K. Kontturi, and A. Urtti, “Gold nanoparticles enable selective light-induced contents release from liposomes,” Journal of Controlled Release, vol. 122, no. 1, pp. 86–93, 2007.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133