Hostname: page-component-7c8c6479df-995ml Total loading time: 0 Render date: 2024-03-28T14:05:15.342Z Has data issue: false hasContentIssue false

Berries modify the postprandial plasma glucose response to sucrose in healthy subjects

Published online by Cambridge University Press:  24 November 2009

Riitta Törrönen*
Affiliation:
Department of Clinical Nutrition, Food and Health Research Centre, University of Kuopio, PO Box 1627, FI-70211Kuopio, Finland
Essi Sarkkinen
Affiliation:
Foodfiles Ltd, Kuopio, Finland
Niina Tapola
Affiliation:
Foodfiles Ltd, Kuopio, Finland
Elina Hautaniemi
Affiliation:
Foodfiles Ltd, Kuopio, Finland
Kyllikki Kilpi
Affiliation:
Finnsugar Ltd, Kantvik, Finland
Leo Niskanen
Affiliation:
Institute of Medicine, Internal Medicine, Kuopio University Hospital, Kuopio, Finland
*
*Corresponding author: Dr Riitta Törrönen, fax +358 17 162785, email riitta.torronen@uku.fi
Rights & Permissions [Opens in a new window]

Abstract

Sucrose increases postprandial blood glucose concentrations, and diets with a high glycaemic response may be associated with increased risk of obesity, type 2 diabetes and CVD. Previous studies have suggested that polyphenols may influence carbohydrate digestion and absorption and thereby postprandial glycaemia. Berries are rich sources of various polyphenols and berry products are typically consumed with sucrose. We investigated the glycaemic effect of a berry purée made of bilberries, blackcurrants, cranberries and strawberries, and sweetened with sucrose, in comparison to sucrose with adjustment of available carbohydrates. A total of twelve healthy subjects (eleven women and one man, aged 25–69 years) with normal fasting plasma glucose ingested 150 g of the berry purée with 35 g sucrose or a control sucrose load in a randomised, controlled cross-over design. After consumption of the berry meal, the plasma glucose concentrations were significantly lower at 15 and 30 min (P < 0·05, P < 0·01, respectively) and significantly higher at 150 min (P < 0·05) compared with the control meal. The peak glucose concentration was reached at 45 min after the berry meal and at 30 min after the control meal. The peak increase from the baseline was 1·0 mmol/l smaller (P = 0·002) after ingestion of the berry meal. There was no statistically significant difference in the 3 h area under the glucose response curve. These results show that berries rich in polyphenols decrease the postprandial glucose response of sucrose in healthy subjects. The delayed and attenuated glycaemic response indicates reduced digestion and/or absorption of sucrose from the berry meal.

Type
Short Communication
Copyright
Copyright © The Authors 2009

Foods or meals high in available carbohydrate such as sucrose increase postprandial blood glucose concentrations. Regular consumption of diets with high glycaemic impact may increase risk for obesity, type 2 diabetes and CVD by promoting excessive food intake, pancreatic β cell dysfunction, dyslipidaemia and endothelial dysfunction(Reference Ludwig1). Several in vitro and in vivo studies have suggested that polyphenols may influence carbohydrate digestion and absorption and thereby postprandial glycaemia. Polyphenols have inhibited intestinal α-glucosidase (maltase and sucrase) activity(Reference Welsch, Lachance and Wasserman2Reference Schäfer and Högger8) and glucose transport(Reference Welsch, Lachance and Wasserman9Reference Hanamura, Mayama and Aoki14)in vitro. They have also suppressed the elevation of blood glucose concentration after oral administration of glucose or maltose in animal models(Reference Matsui, Tanaka and Tamura7, Reference Song, Kwon and Chen11, Reference Hanamura, Mayama and Aoki14, Reference Matsui, Ebuchi and Kobayashi15). In human studies, beverages rich in polyphenolic compounds have shown beneficial effects on postprandial glycaemia. Delayed absorption of glucose after consumption of apple juice(Reference Johnston, Clifford and Morgan16) and coffee(Reference Johnston, Clifford and Morgan17) and attenuated glycaemic response to sucrose consumed in chlorogenic acid-enriched coffee(Reference Thom18) have been reported.

Berries are excellent sources of various polyphenols, such as anthocyanins, flavonols, phenolic acids, ellagitannins and proanthocyanidins(Reference Määttä-Riihinen, Kamal-Eldin and Mattila19Reference Ovaskainen, Törrönen and Koponen23). Polyphenol-rich extracts of blueberries, blackcurrants, strawberries and raspberries have been shown to inhibit α-glucosidase activity in vitro (Reference McDougall, Shpiro and Dobson24, Reference da Silva Pinto, Kwon and Apostolidis25). In addition, consumption of cranberry juice sweetened with high-fructose corn syrup resulted in a different (but not statistically significant) pattern of postprandial glycaemia compared with the similar amount of the sweetener in water(Reference Wilson, Singh and Vorsa26). In the present study we investigated the glycaemic effect of a berry purée made of bilberries, blackcurrants, cranberries and strawberries, and sweetened with sucrose, in reference to sucrose alone.

Experimental methods

Subjects

A total of twelve subjects (eleven women and one man) were recruited from the register of Foodfiles Ltd. At the screening visit, the health status of the subjects was checked by routine blood chemistry (fasting plasma glucose, blood count, serum thyroid-stimulating hormone, plasma creatinine, γ-glutamyl transferase and urate) and a structured interview on previous and current diseases, current medication, alcohol and tobacco consumption, physical activity and use of nutrient supplements. The mean age was 54·2 (sd 15·1; range 25–69) years, BMI 25·4 (sd 2·9; range 21·1–30·0) kg/m2, and fasting plasma glucose concentration 5·3 (sd 0·4; range 4·5–6·0) mmol/l.

The present study was conducted according to the guidelines laid down in the Declaration of Helsinki, and all procedures involving human subjects were approved by the Research Ethics Committee of the Hospital District of Northern Savo (Finland). Written informed consent was obtained from all subjects.

Study design

The randomised, controlled, cross-over study was carried out single-blinded for the study nurse. Each subject was studied in two 3 h meal tests, on separate days, at least 5 d apart. The test meals were administered in a randomised order in an open-label design. The subjects were advised to keep their medication, lifestyles and body weight constant and to follow their habitual diet throughout the study. In the evening before the test, the subjects were instructed to avoid berries, and to consume a meal of choice and repeat that meal before the second test.

The experiments began in the morning after a 12 h overnight fast. The fasting blood samples were obtained from a fingertip capillary blood drop using a lancing device. The subjects were advised to consume the test meal within 15 min. The first bite in the mouth was set as time 0 and the following blood samples were taken after 15, 30, 45, 60, 90, 120, 150 and 180 min. Blood glucose was analysed using a HemoCue Glucose 201+analyser (HemoCue AB, Angelholm, Sweden) calibrated to plasma-equivalent glucose concentrations.

Test meals

The test meal was a mixed berry purée (150 g) with 35 g sucrose. It consisted of equal amounts (37·5 g) of black currants (Ribes nigrum), bilberries (wild European blueberries, Vaccinium myrtillus), European cranberries (Vaccinium oxycoccos) and strawberries (Fragaria x ananassa). The natural sugar composition of the purée was 3·0 % glucose (4·5 g/portion) and 3·4 % fructose (5·1 g/portion), as analysed by HPLC. Water (120 ml) was served with the berry purée. The control meal included 250 ml water, 35 g sucrose, 4·5 g glucose and 5·1 g fructose, to achieve the similar profile and amounts of available carbohydrates.

Statistical analysis

The data were analysed with the SPSS 15·0 for Windows statistical program (SPSS, Inc., Chicago, IL, USA). Normal distribution of variables was checked with the Shapiro–Wilk test. The statistical significance of the overall difference in plasma glucose concentrations between the meals was assessed with a general linear model (GLM) for repeated measures followed by paired-samples t tests with Bonferroni correction to analyse the differences between the test meals at different timepoints. In addition, the maximum increase from baseline was calculated and the difference between the test meals was analysed by paired-samples t tests with Bonferroni correction. The 0–180 min areas under the glucose response curve were calculated using Canvas 8·0·2 (Deneba Software, Miami, FL, USA), ignoring the area below the baseline concentration, and the statistical significance was assessed with paired-samples t tests. Statistical significance was obtained at P < 0·05.

Results

The mean body weight of the study subjects remained stable during the study: 68·6 (sd 9·7) kg at screening, 68·5 (sd 9·9) kg at visit 1 and 68·3 (sd 9·9) kg at visit 2. The ingestion order of the berry meal and the control meal had no effect on the results. The fasting plasma glucose concentrations did not differ between the test meal occasions. The berry meal was ingested in 9·9 (sd 0·7) min and the control meal in 10·0 (sd 1·0) min.

The plasma glucose concentrations at 15 and 30 min after the berry meal were significantly lower than those after the control meal (P < 0·05, P < 0·01, respectively) (Fig. 1). In addition, the glucose response at 150 min after the berry meal was higher than after the control meal (P < 0·05). The peak glucose concentration was reached at 45 min after the berry meal and at 30 min after the control meal. The maximum increase in plasma glucose concentration from the baseline was smaller (P = 0·002) after ingestion of the berry meal (2·3 (sd 1·3) mmol/l) than after ingestion of the control meal (3·3 (sd 1·5) mmol/l). The 0–180 min areas under the glucose response curve tended to be lower after the berry meal (133 (sd 58) min × mmol/l) than after the control meal (149 (sd 74) min × mmol/l; P = 0·29).

Fig. 1 Plasma glucose concentrations after ingestion of the berry meal (●) and the control meal (○) in healthy subjects (n 12). Values are means, with standard deviations represented by vertical bars. Mean value was significantly different from that of the control meal: *P < 0·05, **P < 0·01 (paired-samples t test with Bonferroni correction).

Discussion

The present study shows that ingestion of sucrose with berries produced a different postprandial glycaemic response compared with the control without berries but with a comparable profile of available carbohydrates. The shape of the plasma glucose curve, with reduced concentrations in the early phase and a slightly elevated concentration in the later phase, indicates a delayed response due to berry consumption. Berries also significantly decreased the peak glucose increment.

We did not analyse the polyphenol content of the berry meal, but the contents of several classes of polyphenols have been extensively reported for Finnish berries(Reference Määttä-Riihinen, Kamal-Eldin and Mattila19Reference Ovaskainen, Törrönen and Koponen23). According to the data published previously(Reference Ovaskainen, Törrönen and Koponen23), the total amount of polyphenols in the berry meal was nearly 800 mg, with anthocyanins and proanthocyanidins as the major groups. Due to the dark-coloured bilberries and blackcurrants, the anthocyanin content was high. Based on our previous data(Reference Koponen, Happonen and Mattila22), we estimate that the berry meal provided approximately 300 mg anthocyanins. It has been reported that the extent of in vitro inhibition of α-glucosidase by berry extracts is related to their anthocyanin content(Reference McDougall, Shpiro and Dobson24). The two anthocyanins cyanidin-3-rutinoside(Reference Adisakwattana, Ngamrojanavanich and Kalampakorn27) and cyanidin-3-galactoside(Reference Adisakwattana, Charoenlertkul and Yibchok-Anun28) were in vitro inhibitors of α-glucosidase. Cyanidin-3-rutinoside is one of the major anthocyanins in blackcurrants(Reference Buchert, Koponen and Suutarinen29), and it showed inhibitory activity comparable with voglibose(Reference Adisakwattana, Ngamrojanavanich and Kalampakorn27). Cyanidin-3-galactoside is present in bilberries(Reference Buchert, Koponen and Suutarinen29) and cranberries(Reference Wilson, Singh and Vorsa26), and showed a synergistic effect with acarbose(Reference Adisakwattana, Charoenlertkul and Yibchok-Anun28). Acarbose and voglibose are inhibitors of α-glucosidase used in the treatment of diabetes. Also proanthocyanidins have shown potent α-glucosidase inhibitory activity(Reference Schäfer and Högger8). It is thus possible that at least part of the reduced postprandial glycaemia observed in the present study can be explained by inhibition of α-glucosidase, the enzyme responsible for the digestion of sucrose to glucose in the intestinal epithelium, by berry polyphenols.

Intestinal absorption of glucose is mediated by active Na-dependent transport via sodium glucose co-transporter 1 (SGLT1) and facilitated Na-independent transport via GLUT2(Reference Levin30). The Na+-dependent SGLT1-mediated glucose uptake was inhibited in a competitive manner by several phenolic acids (chlorogenic, ferulic and caffeic acids)(Reference Welsch, Lachance and Wasserman9) as well as by glucosides of quercetin(Reference Cermak, Landgraf and Wolffram12), whereas the galactoside and glucorhamnoside and the aglycone quercetin itself were ineffective. The glucose transport by GLUT2 was inhibited by the flavonols quercetin and myricetin(Reference Song, Kwon and Chen11, Reference Johnston, Sharp and Clifford13). These phenolic acids and flavonols with inhibitory activity against intestinal glucose uptake are common polyphenolic constituents of berries(Reference Määttä-Riihinen, Kamal-Eldin and Mattila19Reference Mattila, Hellström and Törrönen21) and were present in our berry meal.

A similar postprandial glucose response as observed in the present study has also been reported after consumption of a 25 g glucose load in commercial apple juices(Reference Johnston, Clifford and Morgan16). The mean plasma glucose concentrations were significantly lower at 15 and 30 min after ingestion of clear apple juice, and significantly lower at 15 min but significantly higher at 45 and 60 min after ingestion of cloudy apple juice compared with the control drink (glucose load). The effects of apple juices with high levels of polyphenols (chlorogenic acid and phloridzin) on plasma glucose, insulin, glucose-dependent insulinotropic peptide and glucagon-like peptide-1 concentrations were consistent with the delayed absorption of glucose.

Soluble dietary fibre attenuates postprandial glucose responses after carbohydrate-rich meals(Reference Wood31). Based on the Finnish food composition database(32), our berry meal contained 5·4 g dietary fibre, of which approximately 70 % was insoluble. Since the amount of soluble fibre provided by the berry meal was no more than 1·5 g, it is unlikely that the reduced glycaemic response could be solely explained by the fibre content of the berry meal.

The peak glucose increase was 1·0 mmol/l smaller after ingestion of the berry meal with sucrose compared with the control sucrose load. In addition, the difference in glucose values at 30 min was even bigger, being 1·2 mmol/l. The reduction in the postprandial glucose concentrations was of the same magnitude as previously detected with 4–11 g oat fibre, β-glucan(Reference Braaten, Wood and Scott33Reference Tappy, Gügolz and Würsch35), and it can be considered clinically significant.

In conclusion, a mixture of berries rich in polyphenols decreased the postprandial glucose response of a sucrose load in healthy subjects. Reduced rates of sucrose digestion and/or absorption from the gastrointestinal tract are the most probable mechanisms underlying the delayed and attenuated glycaemic response. For better understanding of the role of berries in the regulation of glucose metabolism, further studies assessing their effects on insulin and other hormonal responses are needed.

Acknowledgements

The study was funded by Finnsugar Ltd.

R. T., E. S., N. T. and K. K. contributed to the conception and design of the study. N. T. and E. H. carried out the study in practice, including data management and statistics. R. T. wrote the first draft of the manuscript, with help from N. T. and E. H., and E. S. and L. N. critically revised the manuscript.

None of the authors had any conflicts of interest.

References

1Ludwig, DS (2002) The glycemic index. Physiological mechanisms relating to obesity, diabetes, and cardiovascular disease. JAMA 287, 24142423.CrossRefGoogle ScholarPubMed
2Welsch, CA, Lachance, PA & Wasserman, BP (1989) Effects of native and oxidized phenolic compounds on sucrase activity in rat brush border membrane vesicles. J Nutr 119, 17371740.CrossRefGoogle ScholarPubMed
3Matsui, T, Ueda, T, Oki, T, et al. (2001) α-Glucosidase inhibitory action of natural acylated anthocyanins. 2. α-Glucosidase inhibition by isolated acylated anthocyanins. J Agric Food Chem 49, 19521956.CrossRefGoogle ScholarPubMed
4Hanamura, T, Hagiwara, T & Kawagishi, H (2005) Structural and functional characterization of polyphenols isolated from acerola (Malpighia emarginata DC.) fruit. Biosci Biotechnol Biochem 69, 280286.CrossRefGoogle ScholarPubMed
5Iwai, K, Kim, MY, Onodera, A, et al. (2006) α-Glucosidase inhibitory and antihyperglycemic effects of polyphenols in the fruit of Viburnum dilatatum Thunb. J Agric Food Chem 54, 45884592.CrossRefGoogle ScholarPubMed
6Tadera, K, Minami, Y, Takamatsu, K, et al. (2006) Inhibition of α-glucosidase and α-amylase by flavonoids. J Nutr Sci Vitaminol (Tokyo) 52, 149153.CrossRefGoogle ScholarPubMed
7Matsui, T, Tanaka, T, Tamura, S, et al. (2007) α-Glucosidase inhibitory profile of catechins and theaflavins. J Agric Food Chem 55, 99105.CrossRefGoogle ScholarPubMed
8Schäfer, A & Högger, P (2007) Oligomeric procyanidins of French maritime pine bark extract (Pycnogenol®) effectively inhibit α-glucosidase. Diabetes Res Clin Pract 77, 4146.CrossRefGoogle ScholarPubMed
9Welsch, CA, Lachance, PA & Wasserman, BP (1989) Dietary phenolic compounds: inhibition of Na+-dependent d-glucose uptake in rat intestinal brush border membrane vesicles. J Nutr 119, 16981704.CrossRefGoogle ScholarPubMed
10Kobayashi, Y, Suzuki, M, Satsu, H, et al. (2000) Green tea polyphenols inhibit the sodium-dependent glucose transporter of intestinal epithelial cells by a competitive mechanism. J Agric Food Chem 48, 56185623.CrossRefGoogle ScholarPubMed
11Song, J, Kwon, O, Chen, S, et al. (2002) Flavonoid inhibition of sodium-dependent vitamin C transporter 1 (SVCT1) and glucose transporter isoform 2 (GLUT2), intestinal transporters for vitamin C and glucose. J Biol Chem 277, 1525215260.CrossRefGoogle ScholarPubMed
12Cermak, R, Landgraf, S & Wolffram, S (2004) Quercetin glucosides inhibit glucose uptake into brush-border-membrane vesicles of porcine jejunum. Br J Nutr 91, 849855.CrossRefGoogle ScholarPubMed
13Johnston, K, Sharp, P, Clifford, M, et al. (2005) Dietary polyphenols decrease glucose uptake by human intestinal Caco-2 cells. FEBS Lett 579, 16531657.CrossRefGoogle ScholarPubMed
14Hanamura, T, Mayama, C, Aoki, H, et al. (2006) Antihyperglycemic effect of polyphenols from acerola (Malpighia emarginata DC.) fruit. Biosci Biotechnol Biochem 70, 18131820.CrossRefGoogle ScholarPubMed
15Matsui, T, Ebuchi, S, Kobayashi, M, et al. (2002) Anti-hyperglycemic effect of diacylated anthocyanin derived from Ipomoea batatas cultivar Ayamurasaki can be achieved through the α-glucosidase inhibitory action. J Agric Food Chem 50, 72447248.CrossRefGoogle ScholarPubMed
16Johnston, KL, Clifford, MN & Morgan, LM (2002) Possible role for apple juice phenolic compounds in the acute modification of glucose tolerance and gastrointestinal hormone secretion in humans. J Sci Food Agric 82, 18001805.CrossRefGoogle Scholar
17Johnston, KL, Clifford, MN & Morgan, LM (2003) Coffee acutely modifies gastrointestinal hormone secretion and glucose tolerance in humans: glycemic effects of chlorogenic acid and caffeine. Am J Clin Nutr 78, 728733.CrossRefGoogle ScholarPubMed
18Thom, E (2007) The effect of chlorogenic acid enriched coffee on glucose absorption in healthy volunteers and its effect on body mass when used long-term in overweight and obese people. J Int Med Res 35, 900908.CrossRefGoogle ScholarPubMed
19Määttä-Riihinen, KR, Kamal-Eldin, A, Mattila, PH, et al. (2004) Distribution and contents of phenolic compounds in eighteen Scandinavian berry species. J Agric Food Chem 52, 44774486.CrossRefGoogle ScholarPubMed
20Määttä-Riihinen, KR, Kamal-Eldin, A & Törrönen, AR (2004) Identification and quantification of phenolic compounds in berries of Fragaria and Rubus species (family Rosaceae). J Agric Food Chem 52, 61786187.CrossRefGoogle ScholarPubMed
21Mattila, P, Hellström, J & Törrönen, R (2006) Phenolic acids in berries, fruits, and beverages. J Agric Food Chem 54, 71937199.CrossRefGoogle ScholarPubMed
22Koponen, JM, Happonen, AM, Mattila, PH, et al. (2007) Contents of anthocyanins and ellagitannins in selected foods consumed in Finland. J Agric Food Chem 55, 16121619.CrossRefGoogle ScholarPubMed
23Ovaskainen, ML, Törrönen, R, Koponen, JM, et al. (2008) Dietary intake and major food sources of polyphenols in Finnish adults. J Nutr 138, 562566.CrossRefGoogle ScholarPubMed
24McDougall, GJ, Shpiro, F, Dobson, P, et al. (2005) Different polyphenolic components of soft fruits inhibit α-amylase and α-glucosidase. J Agric Food Chem 53, 27602766.CrossRefGoogle ScholarPubMed
25da Silva Pinto, M, Kwon, YI, Apostolidis, E, et al. (2008) Functionality of bioactive compounds in Brazilian strawberry (Fragaria x ananassa Duch.) cultivars: evaluation of hyperglycemia and hypertension potential using in vitro models. J Agric Food Chem 56, 43864392.CrossRefGoogle ScholarPubMed
26Wilson, T, Singh, AP, Vorsa, N, et al. (2008) Human glycemic response and phenolic content of unsweetened cranberry juice. J Med Food 11, 4654.CrossRefGoogle ScholarPubMed
27Adisakwattana, S, Ngamrojanavanich, N, Kalampakorn, K, et al. (2004) Inhibitory activity of cyanidin-3-rutinoside on α-glucosidase. J Enzyme Inhib Med Chem 19, 313316.CrossRefGoogle ScholarPubMed
28Adisakwattana, S, Charoenlertkul, P & Yibchok-Anun, S (2009) α-Glucosidase inhibitory activity of cyanidin-3-galactoside and synergistic effect with acarbose. J Enzyme Inhib Med Chem 24, 6569.CrossRefGoogle ScholarPubMed
29Buchert, J, Koponen, JM, Suutarinen, M, et al. (2005) Effect of enzyme-aided pressing on anthocyanin yield and profiles in bilberry and blackcurrant juices. J Sci Food Agric 85, 25482556.CrossRefGoogle Scholar
30Levin, RJ (1995) Digestion and absorption of carbohydrates – from molecules and membranes to humans. Am J Clin Nutr 59, Suppl., 690S698S.CrossRefGoogle Scholar
31Wood, PJ (2007) Cereal β-glucans in diet and health. J Cereal Sci 46, 230238.CrossRefGoogle Scholar
32National Institute for Health and Welfare (2009) Fineli – The Finnish Food Composition Database (website in Finnish; also available in English).http://www.fineli.fi/.Google Scholar
33Braaten, JT, Wood, PJ, Scott, FW, et al. (1991) Oat gum lowers glucose and insulin after an oral glucose load. Am J Clin Nutr 53, 14251430.CrossRefGoogle ScholarPubMed
34Wood, PJ, Braaten, JT, Scott, FW, et al. (1994) Effect of dose and modification of viscous properties of oat gum on plasma glucose and insulin following an oral glucose load. Br J Nutr 72, 731743.CrossRefGoogle ScholarPubMed
35Tappy, L, Gügolz, E & Würsch, P (1996) Effects of breakfast cereals containing various amounts of β-glucan fibers on plasma glucose and insulin responses in NIDDM subjects. Diabetes Care 19, 831834.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1 Plasma glucose concentrations after ingestion of the berry meal (●) and the control meal (○) in healthy subjects (n 12). Values are means, with standard deviations represented by vertical bars. Mean value was significantly different from that of the control meal: *P < 0·05, **P < 0·01 (paired-samples t test with Bonferroni correction).