Next Article in Journal
β-carotene in Obesity Research: Technical Considerations and Current Status of the Field
Next Article in Special Issue
Correction: Fang et al. In Vivo Rodent Models of Type 2 Diabetes and Their Usefulness for Evaluating Flavonoid Bioactivity. Nutrients 2019, 11, 530
Previous Article in Journal
The Impact of Maternal Eating Disorders on Dietary Intake and Eating Patterns during Pregnancy: A Systematic Review
Previous Article in Special Issue
In Vivo Rodent Models of Type 2 Diabetes and Their Usefulness for Evaluating Flavonoid Bioactivity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Selective Role of Dietary Anthocyanins and Flavan-3-ols in Reducing the Risk of Type 2 Diabetes Mellitus: A Review of Recent Evidence

1
Department of Food Science and Nutrition, Center for Nutrition Research, Institute for Food Safety and Health, Illinois Institute of Technology, Chicago, IL 60616, USA
2
Department of Public Health and Social Medicine, Medical University of Gdansk, 80-210 Gdansk, Poland
*
Author to whom correspondence should be addressed.
Nutrients 2019, 11(4), 841; https://doi.org/10.3390/nu11040841
Submission received: 24 February 2019 / Revised: 29 March 2019 / Accepted: 10 April 2019 / Published: 13 April 2019
(This article belongs to the Special Issue Flavonoid Intake and Human Health)

Abstract

:
Type 2 diabetes mellitus (T2DM) is the most common form of DM and its prevalence is increasing worldwide. Because it is a progressive disease, prevention, early detection and disease course modification are possible. Diet plays a critical role in reducing T2DM risk. Therapeutic dietary approaches routinely recommend diets high in plant foods (i.e., vegetables, fruits, whole-grains). In addition to essential micronutrients and fiber, plant-based diets contain a wide-variety of polyphenols, specifically flavonoid compounds. Evidence suggests that flavonoids may confer specific benefits for T2DM risk reduction through pathways influencing glucose absorption and insulin sensitivity and/or secretion. The present review assesses the relationship between dietary flavonoids and diabetes risk reduction reviewing current epidemiology and clinical research. Collectively, the research indicates that certain flavonoids, explicitly anthocyanins and flavan-3-ols and foods rich in these compounds, may have an important role in dietary algorithms aimed to address diabetes risk factors and the development of T2DM.

1. Introduction

Diabetes Mellitus (DM) is a complex metabolic disease characterized by hyperglycemia resulting from impairments in insulin secretion, insulin action, or both [1]. The number of individuals with DM has increased nearly four-fold since 1980 from 108 million to ~422 million worldwide [2]. The health care liability of DM in the United States of America (USA) and other parts of the world has increased accordingly [3,4]. People with DM have an increased risk of developing a number of health issues. Persistently elevated blood glucose causes generalized vascular damage affecting the heart and blood vessels, eyes, kidney and nerves [5]. While Type 1 DM cannot be prevented with existing knowledge, prevention is possible for Type 2 DM (T2DM), and management of any DM reduces risk and severity of complications and premature death. Identifying early indicators of metabolic disruption and intervening with effective affordable approaches is critical in preventing T2DM and reducing the associated individual, national and global health care burden.

2. Path to T2DM

T2DM is the most common form of DM. Although both genetic and environmental factors contribute to the development and progression of T2DM [6], growing prevalence seems to be the result of changing dietary habits and lifestyles observed in modern societies [7]. Many of these changes promote obesity development, an established risk factor for T2DM [7,8]. Increasing body mass index (BMI) is associated with central adiposity, low-grade chronic inflammation, and cellular redox imbalances that lead to impaired metabolic processing. Impairments in metabolism are largely the result of tissue resistance to insulin’s actions [9,10], and may be considered the first stage along the disease path. This first stage is characterized by a long period of insulin resistance and compensatory hyperinsulinemia to control blood glucose within normal ranges (Figure 1A→B). In the progressive path to T2DM, insulin resistance continues to increase until insulin secretion from the pancreatic β-cell fails to compensate fully resulting in elevated blood glucose concentrations that eventually meet criteria for pre-diabetes (100–125 mg/dL, Figure 1C, [1]). Pre-diabetes represents a critical crossroads as ~50% of those with pre-diabetes will develop T2DM within seven years and 83% will convert over a lifetime (Figure 1D) [11]. Although the path to T2DM begins early with insulin resistance (Figure 1A→B), preventative care does not usually start until pre-diabetes is identified (Figure 1C). Intervention strategies typically focus on diet and lifestyle modification [12,13,14,15], yet, arguably, these strategies could begin sooner when metabolic impairments first appear, such as with insulin resistance (Figure 1B) [9]. Identifying specific dietary flavonoids that act on these early indicators is opportunistic and would deliver targeted dietary modification for overall disease risk reduction.

3. Diet and T2DM

The Western/modern day diet has been a major factor in the increased incidence of T2DM [12]. In addition to excess calorie intake, food choices are nutrient poor and comprised of highly refined, readily available carbohydrates and fats [14]. Modification of diet and lifestyle has proven to be a principal approach in T2DM risk reduction: the clinical evidence demonstrating these modifications were more effective than medication (metformin) in preventing conversion from pre-diabetes to T2DM in the Diabetes Prevention Program clinical trial (n = 3234) [16]. Dietary patterns, such as the Mediterranean diet and the Dietary Approaches to Stop Hypertension (DASH) diet also have shown favorable outcomes in T2DM risk reduction [17,18]. The foundation of these diets is plant foods (i.e., vegetables, fruits and whole-grains), which contain varied essential nutrients and fiber [19,20,21], along with an array of phytochemicals with biological activity. Some of these bioactive compounds are suggested to have glucose modulating and insulin sensitizing effects at the cellular level. Polyphenols, and specifically flavonoid compounds, are among the most studied for their biological activity and some may have an important role in preventing or delaying the onset of T2DM.

4. Dietary Flavonoids and T2DM

4.1. Flavonoids: General Chemistry and Intake

Flavonoid compounds are widely distributed in plant foods and represent ~2/3 of the total dietary polyphenols consumed [22]. Flavonoids are secondary plant metabolites with a characteristic C6–C3–C6 structural backbone. They are classified into six main categories: flavone, isoflavone, flavan-3-ol, flavanone, anthocyanidin and flavonol. More than 9000 flavonoids have been reported [23] differing in degree of hydroxylation and methylation of A and B rings along with the presence of various glycosylation patterns (O- or C-glycosides). Intake of dietary flavonoids ranges from 209 to 1017 mg/d (mean 435 mg/d) in Australian, European, and USA adult populations [24]. Figure 2 provides a summary of flavonoid compounds, dietary sources, and ranges of intake [24,25,26].

4.2. Epidemiology of Dietary Flavonoids and T2DM Incidence

The collective epidemiological research suggests an inverse association between dietary flavonoids and T2DM incidence [27,28,29,30,31]. This association appears to be driven by a few select flavonoids [28,29,30,31]. Dietary anthocyanins and flavan-3-ols have been consistently identified as having a strong association with T2DM risk reduction [30,31], and flavonols and isoflavones have shown risk reduction associations intermittently and with less strength. These assertions are further supported by observational data linking foods relatively high in these compounds, such as berries, tea, and chocolate, and reduced incidence of T2DM [28,32,33,34,35,36,37,38,39,40,41]. Dose–response meta-analyses have suggested that 7.5 mg/day increment of dietary anthocyanin intake or 17 g/d berry intake could decrease the risk of T2DM by 5% [28] and ≥3 cups of tea/day [33] or 1–6 servings of chocolate per week were related to significant risk reduction [40].
How flavonoids, and specifically those with the strongest relationship, anthocyanins and flavan-3-ols, reduce the incidence of T2DM is not clear. Risk factor modulation is one hypothesis. Insulin resistance is a major risk factor of T2DM, and the links between excess adiposity, chronic low grade inflammation and cellular oxidative stress in insulin resistance pathogenesis are well established [42,43,44]. Flavonoids are suggested to have anti-obesity, anti-oxidant and anti-inflammatory effects. Increased consumption of anthocyanins, flavan-3-ols, flavonols, and flavonoid polymers was inversely associated with weight gain over four-year time intervals in men and women of USA [45] and, after adjustment for dietary fiber, only anthocyanins and total flavonoid polymers remained significant. In a study using data from the Twins United Kingdom (UK) registry, flavonoid subclasses including anthocyanins, flavan-3-ols, and flavonols, were associated with a more favorable fat mass distribution as measured by dual-energy X-ray absorptiometry [46]. Importantly, these effects were independent of genetic and common environmental factors, as well as after controlling for total fruit and vegetable, and fiber intake. Also using the UK Twins registry, Jennings et al. examined the impact of individual flavonoid subclasses on various indicators of insulin action and inflammation [47]. They reported an inverse relationship between dietary anthocyanins and flavones and insulin resistance, fasting insulin and markers of inflammation in women 18–76 years old [47]. In another cross-sectional analysis of 2375 Framingham Heart Study Offspring Cohort participants, higher anthocyanin and flavonol intakes were associated with significantly lower composite inflammation scores [48], with anthocyanins having markedly stronger (>2 fold) association than flavonols in the fully adjusted model. When sub-scores were analyzed further, higher anthocyanin intake was significantly and inversely associated with all sub-scores, including biomarker concentrations contributing to acute inflammation, oxidative stress and cytokine levels. Higher flavonol intake was associated with lower cytokine and oxidative stress biomarker concentrations, and flavan-3-ols and total flavonoids were inversely associated with oxidative stress markers [48]. Overall, the epidemiological findings strongly support a role of certain flavonoids in reducing the incidence of T2DM, and the effect may be achieved through reducing inflammation, oxidative stress and insulin resistance.

4.3. Mechanisms of Action of Flavonoids with Anti-Diabetic Effects

The anti-diabetic effects of compounds, foods, and diets culminate to support glucose homeostasis through a number of possible mechanisms. Figure 3 summarizes purported mechanisms of action of flavonoids proposed to have anti-diabetic effects. In general, their bioactivity can be attributed to modulation of insulin dependent or independent pathways to decrease blood glucose. Insulin dependent pathways involve modifying cellular redox status and cell signaling paths that influence activities such as insulin synthesis and secretion (i.e., pancreatic β-cell function) and/or peripheral insulin sensitivity via Phosphoinositide 3-kinases (PI3K)/ Protein kinase B (AKT) or Peroxisome proliferator-activated receptor gamma (PPAR-γ) activation in tissues such as muscle, adipose and others [49,50,51,52]. Modulation of insulin-independent pathways includes activation of energy sensing molecules, such as AMP-activated protein kinase (AMPK) in liver, muscle and adipose [51,52]. Interference with carbohydrate digestion and glucose absorption in the small intestine has also been described via inhibition of α-amylase and/or α-glucosidase activity [53] and/or interfering with glucose transport [54]. While present knowledge has yet to identify the dominant mechanism or the relative potency of individual flavonoids, current epidemiological research suggests that anthocyanins and flavan-3-ols are leading flavonoids underpinning the flavonoid-diabetes risk reduction relationship. Other flavonoids with possible modulatory effects in diabetes pathways include flavonols and isoflavones, and many of the same mechanisms have been identified in vitro and in vivo animal studies [49,55,56].

5. Selecting Flavonoids That Reduce the Risk of Developing T2DM

In the general sequence of using epidemiology to generate hypotheses for testing in clinical trials, the current research points to anthocyanins and flavan-3-ols as having a selective role in reducing development of T2DM. Subsequent sections provide a summary of the clinical evidence, and specific preclinical research for these two flavonoid classes, and explore the evidence of flavonols and isoflavones in T2DM risk reduction.

5.1. Dietary Anthocyanins and Reducing Development of T2DM

5.1.1. General Background

Anthocyanins are water soluble pigment compounds responsible for the blue, red and purple colors in most fruits and vegetables. They are commonly found in red to purplish blue-colored fruits, vegetables, grains and beans and some roots. The human diet includes six main anthocyanidins (aglycone): cyanidin, delphinidin, malvidin, pelargonidin, peonidin, and petunidin. Glycosylation (i.e., anthocyanin) adds to their chemical complexity and influences their stability, absorption patterns and metabolic fate [57]. Anthocyanin content and composition varies greatly in plant foods [26] with berries being the most anthocyanin dense in the diet. Color is a general guide to which anthocyanins are present, which may be important for dietary guidance as research identifies specific effects of certain anthocyanins and/or their metabolites in T2DM risk reduction. Average daily intake of anthocyanin is ~10.3 mg/day in the USA [24]. A higher intake of anthocyanins has been reported in Italy (44.1–64.9 mg/d) [38] and Finland (47 mg/d) [58], which may be explained by higher berry consumption [59] compared with berry intake in the USA [60].

5.1.2. Preclinical Research

Proof of concept and mechanistic studies have been conducted in varied cell culture and animal models to understand the role of anthocyanins in T2DM. Figure 3 illustrates many of the mechanisms ascribed to anthocyanins and their metabolites. This area has been intensively reviewed previously [50,61,62,63]; therefore, the current paper is focused on work since 2010.
Extending earlier research to include the effects of metabolites, Scazzocchio et al. reported cyanidin-3-O-β-glucoside and its metabolite protocatechuic acid upregulated PPAR-γ activity, glucose transporter type 4 (GLUT4) translocation, and enhanced glucose uptake in human omental adipocytes, as well as murine 3T3-L1 differentiated adipocytes [54]. Using human HepG2 cells, a model system to study liver metabolism, including xenobiotic metabolism, an anthocyanin-rich mulberry extract was shown to alleviate insulin resistance and increase glucose uptake and glycogen synthesis via activation of PI3K/AKT pathways [64]. Improvements in metabolic parameters were verified in db/db mice supplemented with the mulberry extract [64]. Anthocyanins may also reduce absorption of glucose through glucose transporter 2 (GLUT2) and the sodium-glucose linked transporter 1 (SGLT1) as suggested by some studies [65,66,67]. They have also been reported to stimulate insulin secretion through activation of l-type voltage-dependent Ca2+ channels [68] and by the activation of the free fatty acid receptor-1 [69].
Feeding studies in rodents have provided further insight into the role of anthocyanins in T2DM. Kurimoto et al. reported that dietary black soybean seed coat extract rich in anthocyanins improved glycemia and insulin sensitivity in a T2DM mouse model, and effects were related to the activation of AMPK [70]. The activation of AMPK in the skeletal muscle and liver was accompanied by upregulation of GLUT4 in skeletal muscle and downregulation of gluconeogenesis in the liver. Xing et al. [71] also observed increased insulin sensitivity consistent with increased AMPK phosphorylation in white adipose tissue of obese mice supplemented with 5% freeze dried raspberry powder for 12 weeks. In another study, 80 days supplementation of blackberry extract in standard rat diet compared to standard diet alone increased insulin sensitivity, and decreased concentrations of glucose and insulin; however, the changes were greater in female than in male rats [72]. Tani et al. performed an intraperitoneal glucose tolerance test in six-week-old male Sprague–Dawley rats after oral administration of blackcurrant extract vs. control diet (no blackcurrant) and reported increased glucagon like peptide-1 (GLP-1) and insulin secretion [73]. Including purified blueberry anthocyanins in drinking water of mice fed a high fat diet for 72 days significantly lowered fasting glucose concentrations and corrected β-cell function compared to control mice fed a high fat diet only [68]. In contrast, including blueberry juice did not show the same effect as the purified anthocyanins in the same study setting, which could have been a result of differences in sugar intake in juice vs. anthocyanins supplemented water, lower dose of anthocyanins consumed in drinking water and or other components in blueberry juice that may interfere with anthocyanin mechanism of action [74].

5.1.3. Clinical Research

Available literature on the effect of dietary anthocyanins to reduce risk of diabetes in human clinical trials is growing. Supplementation with a juice made from dried strawberry and cranberry polyphenol extracts for six weeks resulted in significantly improved insulin sensitivity measured by hyperinsulinaemic-euglycemic clamp in subjects with insulin resistance [75]. This is the second trial supplementing with an anthocyanin rich food [76] or extract [75] where improvements in insulin sensitivity are reported using clamp methodology, but significant improvements on fasting glucose and insulin concentrations were not observed. These results may explain, at least in part, why other reports supplementing daily for 6–8 weeks with blueberry, bilberry, strawberry or pomegranate did not find improvements on fasting indices [77,78,79,80], but improvements were reported on postprandial indices of glycemic control when strawberry, blackberry, bilberry, blackcurrant or pomegranate, although not blueberries, were consumed with bread [81,82], with a meal [83,84,85] or a sugar drink [86,87,88,89]. Postprandial trials from our lab with strawberries suggest simultaneous intake with a meal or within 2 h before a meal may be required to elicit an effect on glucose and/or insulin metabolism, providing support for a role of early phase strawberry/berry metabolites in peripheral glucose regulation [83,90,91]. Likewise, the effects of dietary anthocyanins are subject to the amount of anthocyanins consumed relative to the population phenotype. A modest post-meal reduction in insulin concentrations was reported with 10 g (~ 1 cup fresh weight equivalent) of freeze-dried strawberry powder compared to control (0 g) in overweight hyperlipidemic individuals [83]. However, in a dose response study in obese individuals with insulin resistance, a beverage containing 40 g of freeze-dried strawberry powder significantly reduced the post-meal demand for insulin compared to a control beverage devoid of polyphenols, but matched for fiber [92]. Furthermore, statistical evaluation of the dose-dependent strawberry metabolite profiles relative to clinical outcomes indicated an inverse relationship between the primary anthocyanin metabolites of strawberry and insulin responses and glucose clearance [92]. In individuals with pre-diabetes and insulin resistance, intake of 250 g of frozen red raspberry (~2 cups) in a breakfast meal significantly reduced peak and postprandial (2 h) glucose concentrations compared to control (0 g of red raspberry), whereas inclusion of either 125 g or 250 g raspberry in the breakfast reduced postprandial insulin concentrations compared to control [93]. The aforementioned findings are corroborated by two recent meta-analyses of randomized controlled trials testing anthocyanin intake/anthocyanin rich foods on cardio-metabolic risk factors. The meta analyses indicated that the effects of anthocyanins on glucose homeostasis measures, including fasting glucose and hemoglobin A1c (HbA1c, an indicator of postprandial and fasting glucose control), and insulin sensitivity/ resistance (homeostasis model assessment index-insulin resistance, HOMA-IR) depends on the dose of anthocyanins, body mass index (BMI)/population phenotype, and anthocyanin source [94,95], specifically highlighting the beneficial effects of berries as a source of anthocyanins.
Intervention studies examining the effects of anthocyanin intake in people with T2DM are limited but demonstrate benefits. In one study, 58 patients with T2DM were given 160 mg of anthocyanins twice daily or placebo (n = 29/group) for 24 weeks in a randomized, placebo-controlled, double-blind trial [96]. Supplementation with anthocyanins decreased fasting glucose and insulin resistance (measured by HOMA-IR) compared to placebo [96]. Moazen et al. reported that daily intake of 50 g strawberry powder for six weeks significantly decreased HbA1c in newly diagnosed T2DM patients (n = 19) compared to matched subjects in control group (n = 17) [97].
Overall, there appears to be complementary data from the collective epidemiological and human clinical trial investigations suggesting that dietary anthocyanins have a strong potential to modulate the risk of T2DM in humans. These findings are further supported by a number of preclinical studies in animals and cell culture models indicating that anthocyanins can work through various cellular signaling pathways, many of which are redox sensitive, to achieve glucose homeostasis.

5.2. Dietary Flavan-3-ols and Reducing Development of T2DM

5.2.1. General Background

The structure of flavan-3-ols is similar to other flavonoids, and like other flavonoids have subclasses that include flavan-3-ol monomers (catechin, epicatechin, epigallocatechin, epicatechin 3-gallate, epigallocatechin 3-gallate, gallocatechin, and catechin 3-gallate), and proanthocyanidins, which range in complexity, but principally represent polymerization of monomeric flavan-3-ols (dimers, trimers, 4–6 mers, 7–10 monomers, polymers) and theaflavins [98,99]. Flavan-3-ols are found in a number of plant foods, including broad beans (average concentration of 154.5 mg total flavan-3-ols/100 g fresh weight); some fruits (ranging from 10 mg to 50 mg/100 g fresh weight) such as plum, apple, custard apple, strawberry-tree fruit, berry fruits such as blueberry and cranberry, cherry, grapes, red wines, cocoa/chocolate (185 mg/100 g of cocoa powder); and green tea and black tea (43.8 and 26.8 mg/100 ml of infusion, respectively) [100,101]. Very high concentrations of flavan-3-ols (polymers) can be found in nuts (hazelnuts—500 mg/100 g; pecans—494 mg/100 g; pistachios—237 mg/100 g; almonds—184 mg/100 g) [102]. Epidemiological research suggests intake of flavan-3-ols ranges from 22–28 mg/d in the USA [32,34].

5.2.2. Preclinical Research

Similar to anthocyanins, several epidemiological studies suggest a role of flavon-3-ols in reducing T2DM risk. The preclinical research provides insight into potential mechanisms of action, which includes free radical scavenging activity, mediation of inflammatory responses and enzymes involved in glucose metabolism. Some of these activities are noted in Figure 3. Briefly, cell culture studies and animal studies have shown flavan-3-ols, such as epigallocatechin-3-gallate (EGCG) and extracts from cacao containing multiple flavan-3-ol subtypes, influence oxidative stress, inflammation and glycemic control pathways [103,104,105]. For example, EGCG inhibited Cd2+ induced apoptosis of human liver cells acting as a reactive oxygen species scavenger [103], and attenuated inflammation-induced insulin resistance in 3T3-L1 adipocytes [104]. Also with EGCG, insulin sensitivity was increased, c-Jun N-terminal kinases (JNK) phosphorylation (p-JNK) was suppressed and GLUT4 expression was increased in the adipose tissue of EGCG supplemented obese KK-ay mice and high-fat diet-induced obese rats [104]. The effects in these animal models coincided with reduced glucose concentrations and improved glucose tolerance [104]. Supplementing cacao extract in the diet of high fat fed mice promoted GLUT4 translocation and increased activation of AMPK in the plasma membrane of brown adipose and skeletal muscle, consistent with reduced hyperglycemia, glucose intolerance and fat accumulation [52]. Thus, in vitro and in vivo models suggest that both monomeric and mixtures of flavan-3-ols subtypes influence mechanisms important in glucose control.

5.2.3. Human Studies

Extending the preclinical research, there is considerable evidence supporting the effects of dietary flavan-3-ols on metabolic indices of glucose and/or insulin metabolism. These data are found in recent meta-analyses of randomized clinical trials examining the effects of cocoa product intake on cardio-metabolic risk factors in varied population groups [106,107,108]. Reported findings suggest modest but significant improvements on insulin-related outcomes, including decreased fasting insulin and improved insulin sensitivity assessed by HOMA-IR. Furthermore, intake of > 200 mg per day flavan-3-ols achieved benefits for insulin endpoints, whereas doses between 200–600 mg flavan-3-ols per day were associated with reduced glucose concentrations [108]. In postmenopausal women with T2DM, 1-year dietary intervention with flavonoid-enriched chocolate (850 mg flavan-3-ols/d) combined with isoflavones (100 mg/d) resulted in significant improvements in insulin sensitivity and reduced insulin concentrations, demonstrating additional benefits of a flavan-3-ols/isoflavone mixture to standard drug therapy on various cardio-metabolic risk endpoints [109]. Another study demonstrated insulin-specific benefits in adults with essential hypertension and impaired glucose tolerance after 15-day supplementation with 100 g flavan-3-ol-rich dark chocolate [110]. Using grape seed extract (GSE) as a source of flavan-3-ols, individuals with pre-hypertension who drank beverages containing 300 mg GSE flavan-3-ols vs. control (0 mg) daily for 6 weeks showed trends in decreased fasting insulin and increased insulin sensitivity; a benefit that regressed after discontinuation of the beverages for four weeks [111]. Tea, as a source of EGCG, has also been investigated for effects on T2DM endpoints revealing positive effects, including decreased HbA1c values compared to baseline measures in adults with impaired glucose tolerance [112]. Improved HbA1c and insulin sensitivity in individuals with T2DM have been reported [113,114]. In contrast, other clinical studies demonstrated that EGCG had no effect on insulin sensitivity [115] or glucose tolerance in obese adults with metabolic syndrome [116]. Overall, including foods rich in flavan-3-ols in the diet as a strategy to reduce risk of and manage T2DM seems advantageous. Similar to results from the dietary anthocyanin literature, the collective findings warrant follow-up testing for longer duration and dose finding trials to develop clear dietary recommendations for individuals with diabetes concerns.

5.3. Other Possible Flavonoids That Reduce Development of T2DM

Flavonols and isoflavones have surfaced as having possible anti-diabetic actions. In a recent meta-analysis, data analyzed from seven studies (nine cohorts) identified flavonols (as a class but not individual flavonols i.e., querecetin, kaempferol) correlated with lower incidence of T2DM, whereas analyses on isoflavones using six studies/eight cohorts indicated both the class (isoflavones) and individual subclass compounds (i.e., genistein and daidzein) were inversely associated with diabetes risk [31]. Preclinical data, including animal and in vitro studies, have demonstrated effects of flavonols, mainly quercetin, on a number of diabetes-related processes [55,117,118,119] and mechanisms related to flavonoid action as shown in Figure 3. However, the human clinical data are limited and have largely not substantiated the preclinical work to date [117].
In vitro studies with genistein [56] have shown effects on pancreatic β-cell proliferation, glucose-stimulated insulin secretion and protection against apoptosis, and these effects are independent of genistein’s activity as an estrogen receptor agonist, antioxidant, or tyrosine kinase inhibitor [56]. However, evidence of the effects of purified genistein intervention in humans with T2DM is extremely limited, despite studies indicating it’s safe for human consumption [120]. Isoflavone mixtures or soy are often studied instead, although still limited in establishing a clear relationship between isoflavones and T2DM risk reduction. For example, isoflavone supplementation (100 mg) for six months in a placebo controlled study in pre-diabetic or early untreated T2DM women showed no benefit on fasting glucose, 2 h glucose or HbA1c compared to placebo control, or soy protein with isoflavones [121]. Whereas earlier work (2002) in post-menopausal women with a slightly higher dose of isoflavones (132 mg) showed decreased HbA1c, fasting insulin and reduced insulin resistance as measured by HOMA-IR [122]. A meta-analysis of randomized control trials investigating phytoestrogen supplementation and body composition in postmenopausal women suggested that phytoestrogen supplementation is associated with reduced weight in healthy postmenopausal women, but increased body weight in postmenopausal women with pre-existing conditions, such as pre-diabetes [123]. With the association between body weight and T2DM risk, dietary supplementations that increase body weight would be counterproductive.

6. Conclusions

Interest in flavonoid health benefits has rapidly increased over the last decade. Advancements in instrumentation have enhanced the field’s ability to identify and quantify these compounds in foods, linking their intake with disease risk and possible mechanisms of actions in T2DM. The collective epidemiological research suggests that select flavonoids, specifically dietary anthocyanins and flavan-3-ols, have an important role in T2DM risk reduction. Controlled clinical trials testing purified compounds, extracts, and foods rich in anthocyanins and flavan-3-ols have provided additional evidence verifying beneficial bioactivity on endpoints important in T2DM development. Future work pursuing metabolite characterization and kinetic profiling linked with clinical biomarkers of disease risk after dietary flavonoid ingestion will help illuminate the preventative and therapeutic role of these compounds, particularly in understanding responses in the context of population characteristics, dose, and food vs. supplement effects. These data will be critical for devising dietary guidance that is targeted and efficacious for long-term metabolic health.
Overall, the research indicates that dietary flavonoids do not universally influence diabetes risk. Instead, selective flavonoids, explicitly anthocyanins and flavan-3-ols and foods rich in these compounds, are biologically active on mechanisms underlying risk factors of T2DM. Early intervention in at risk individuals with diets focused on increasing anthocyanin and flavan-3-ol intake may be particularly opportunistic in reversing or reducing the disease risk trajectory preventing advancement to pre-diabetes and further to T2DM.

Author Contributions

Conceptualization—B.B.F., M.B., I.E.; research and writing of original draft, specific sections—B.B.F., M.B., I.E., E.P., D.X., A.S.; Figure development, B.B.F., M.B., I.E., A.S.; writing—review and editing—B.B.F., M.B., I.E., E.P., D.X., A.S.; writing—Final edits B.B.F., M.B., I.E.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. American Diabetes Association. 2. Classification and Diagnosis of Diabetes: Standards of Medical Care in Diabetes—2018. Diabetes Care 2018, 41, S13–S27. [Google Scholar] [CrossRef]
  2. Mathers, C.D.; Loncar, D. Projections of Global Mortality and Burden of Disease from 2002 to 2030. PLoS Med. 2006, 3, e442. [Google Scholar] [CrossRef]
  3. Gregg, E.W.; Li, Y.; Wang, J.; Rios Burrows, N.; Ali, M.K.; Rolka, D.; Williams, D.E.; Geiss, L. Changes in Diabetes-Related Complications in the United States, 1990–2010. N. Engl. J. Med. 2014, 370, 1514–1523. [Google Scholar] [CrossRef] [PubMed]
  4. Cho, N.; Shaw, J.; Karuranga, S.; Huang, Y.; Fernandes, J.D.R.; Ohlrogge, A.; Malanda, B. IDF Diabetes Atlas: Global estimates of diabetes prevalence for 2017 and projections for 2045. Diabetes Res. Clin. Pract. 2018, 138, 271–281. [Google Scholar] [CrossRef]
  5. World Health Organisation. Mortality and Burden of Disease Attributable to Selected Major Risks; World Health Organisation: Geneva, Switzerland, 2009. [Google Scholar]
  6. Gillett, M.; Royle, P.; Snaith, A.; Scotland, G.; Poobalan, A.; Imamura, M.; Black, C.; Boroujerdi, M.; Jick, S.; Wyness, L.; et al. Non-pharmacological interventions to reduce the risk of diabetes in people with impaired glucose regulation: A systematic review and economic evaluation. Health Technol. Assess. 2012, 16, 1. [Google Scholar] [CrossRef] [PubMed]
  7. Qi, L.; Hu, F.; Hu, G. Genes, Environment, and Interactions in Prevention of Type 2 Diabetes: A Focus on Physical Activity and Lifestyle Changes. Curr. Mol. Med. 2008, 8, 519–532. [Google Scholar] [CrossRef]
  8. Colditz, G.A.; Willett, W.C.; Rotnitzky, A.; Manson, J.E. Weight Gain as a Risk Factor for Clinical Diabetes Mellitus in Women. Ann. Intern. Med. 1995, 122, 481–486. [Google Scholar] [CrossRef] [PubMed]
  9. Mechanick, J.I.; Garber, A.J.; Grunberger, G.; Handelsman, Y.; Timothy Garvey, W. Position Statement dysglycemia-based chronic disease: An American association of clinical endocrinologists position Statement. Endocr. Pract. 2018, 24, 995. [Google Scholar] [CrossRef] [PubMed]
  10. Tabák, A.G.; Herder, C.; Rathmann, W.; Brunner, E.J.; Kivimäki, M. Prediabetes: A high-risk state for diabetes development. Lancet 2012, 379, 2279–2290. [Google Scholar] [CrossRef]
  11. Herman, W.H.; Hoerger, T.J.; Brändle, M.; Hicks, K.; Sorensen, S.; Zhang, P.; Hamman, R.F.; Ackermann, R.T.; Engelgau, M.M.; Ratner, R.E.; et al. The Cost-Effectiveness of Lifestyle Modification or Metformin in Preventing Type 2 Diabetes in Adults with Impaired Glucose Tolerance. Ann. Intern. Med. 2005, 142, 323–332. [Google Scholar] [CrossRef] [Green Version]
  12. WHO. Diet, Nutrition and the Prevention of Chronic Diseases Report of the Joint WHO/FAO Expert Consultation; WHO: Geneva, Switzerland, 2016. [Google Scholar]
  13. American Diabetes Association. 4. Lifestyle Management: Standards of Medical Care in Diabetes—2018. Diabetes Care 2018, 41, S38–S50. [Google Scholar] [CrossRef]
  14. Evert, A.B.; Boucher, J.L.; Cypress, M.; Dunbar, S.A.; Franz, M.J.; Mayer-Davis, E.J.; Neumiller, J.J.; Nwankwo, R.; Verdi, C.L.; Urbanski, P.; et al. Nutrition Therapy Recommendations for the Management of Adults with Diabetes. Diabetes Care 2013, 36, 3821–3842. [Google Scholar] [CrossRef]
  15. American Diabetes Association. 5. Prevention or Delay of Type 2 Diabetes: Standards of Medical Care in Diabetes—2018. Diabetes Care 2018, 41, S51–S54. [Google Scholar] [CrossRef]
  16. Knowler, W.C.; Barrett-Connor, E.; Fowler, S.E.; Hamman, R.F.; Lachin, J.M.; Walker, E.A.; Nathan, D.M. Diabetes Prevention Program Research Group (D.P.P.R). Reduction in the Incidence of Type 2 Diabetes with Lifestyle Intervention or Metformin. N. Engl. J. Med. 2002, 346, 393–403. [Google Scholar]
  17. Salas-Salvado, J.; Bullo, M.; Babio, N.; Martinez-Gonzalez, M.A.; Ibarrola-Jurado, N.; Basora, J.; Estruch, R.; Covas, M.I.; Corella, D.; Aros, F.; et al. Reduction in the Incidence of Type 2 Diabetes with the Mediterranean Diet: Results of the PREDIMED-Reus nutrition intervention randomized trial. Diabetes Care 2011, 34, 14–19. [Google Scholar] [CrossRef]
  18. Campbell, A.P. DASH Eating Plan: An Eating Pattern for Diabetes Management. Diabetes Spectr. 2017, 30, 76–81. [Google Scholar] [CrossRef]
  19. MacLeod, J.; Franz, M.J.; Handu, D.; Gradwell, E.; Brown, C.; Evert, A.; Reppert, A.; Robinson, M. Academy of Nutrition and Dietetics Nutrition Practice Guideline for Type 1 and Type 2 Diabetes in Adults: Nutrition Intervention Evidence Reviews and Recommendations. J. Acad. Nutr. Diet. 2017, 117, 1637–1658. [Google Scholar] [CrossRef]
  20. Rinaldi, S.; Campbell, E.E.; Fournier, J.; O’Connor, C.; Madill, J. A Comprehensive Review of the Literature Supporting Recommendations from the Canadian Diabetes Association for the Use of a Plant-Based Diet for Management of Type 2 Diabetes. Can. J. Diabetes 2016, 40, 471–477. [Google Scholar] [CrossRef]
  21. Handelsman, Y.; Bloomgarden, Z.T.; Grunberger, G.; Umpierrez, G.; Zimmerman, R.S.; Yehuda Handelsman, C.; Zachary Bloomgarden, F.T.; George Grunberger, M.; Guillermo Umpierrez, F.; Robert Zimmerman, F.S.; et al. AACE/ACE Guidelines AMERICAN ASSOCIATION OF CLINICAL ENDOCRINOLOGISTS AND AMERICAN COLLEGE OF ENDOCRINOLOGY-CLINICAL PRACTICE GUIDELINES FOR DEVELOPING A DIABETES MELLITUS COMPREHENSIVE CARE PLAN-2015 AACE TAsk ForCE for DEvEloping A DiAbETEs ComprEhEnsiv. Endocr. Pract. 2015, 21, 1–87. [Google Scholar] [CrossRef]
  22. Scalbert, A.; Williamson, G. Dietary Intake and Bioavailability of Polyphenols. J. Nutr. 2000, 130, 2073S–2085S. [Google Scholar] [CrossRef]
  23. Qiu, T.; Wu, D.; Yang, L.; Ye, H.; Wang, Q.; Cao, Z.; Tang, K. Exploring the Mechanism of Flavonoids Through Systematic Bioinformatics Analysis. Front. Pharmacol. 2018, 9, 918. [Google Scholar] [CrossRef] [PubMed]
  24. Peterson, J.J.; Dwyer, J.T.; Jacques, P.F.; McCullough, M.L. Improving the estimation of flavonoid intake for study of health outcomes. Nutr. Rev. 2015, 73, 553–576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Escobar-Cévoli, R.; Castro-Espín, C.; Béraud, V.; Buckland, G.; Zamora-Ros, R. An Overview of Global Flavonoid Intake and its Food Sources. In Flavonoids—From Biosynthesis to Human Health; InTech: London, UK, 2017. [Google Scholar] [Green Version]
  26. Haytowitz, D.B.; Bhagwat, S.; Harnly, J.; Holden, J.M.; Gebhardt, S.E. Sources of Flavonoids in the U.S. Diet Using USDA’s Updated Database on the Flavonoid Content of Selected Foods. USDA. Available online: https://www.ars.usda.gov/ARSUserFiles/80400525/Articles/AICR06_flav.pdf (accessed on 29 March 2019).
  27. Liu, Y.-J.; Zhan, J.; Liu, X.-L.; Wang, Y.; Ji, J.; He, Q.-Q. Dietary flavonoids intake and risk of type 2 diabetes: A meta-analysis of prospective cohort studies. Clin. Nutr. 2014, 33, 59–63. [Google Scholar] [CrossRef] [PubMed]
  28. Tan, J.; Jiang, J.; Li, D.; Guo, X.; Yang, B. Associations of dietary intakes of anthocyanins and berry fruits with risk of type 2 diabetes mellitus: A systematic review and meta-analysis of prospective cohort studies. Eur. J. Clin. Nutr. 2016, 70, 1360–1367. [Google Scholar]
  29. Guo, X.-F.; Ruan, Y.; Li, Z.-H.; Li, D. Flavonoid subclasses and type 2 diabetes mellitus risk: A meta-analysis of prospective cohort studies. Crit. Rev. Food Sci. Nutr. 2018, 1–13. [Google Scholar] [CrossRef]
  30. Xu, H.; Luo, J.; Huang, J.; Wen, Q. Flavonoids intake and risk of type 2 diabetes mellitus: A meta-analysis of prospective cohort studies. Medicine (Baltimore) 2018, 97, e0686. [Google Scholar] [CrossRef]
  31. Rienks, J.; Barbaresko, J.; Oluwagbemigun, K.; Schmid, M.; Nöthlings, U. Polyphenol exposure and risk of type 2 diabetes: Dose-response meta-analyses and systematic review of prospective cohort studies. Am. J. Clin. Nutr. 2018, 108, 49–61. [Google Scholar] [CrossRef]
  32. Jacques, P.F.; Cassidy, A.; Rogers, G.; Peterson, J.J.; Meigs, J.B.; Dwyer, J.T. Higher Dietary Flavonol Intake Is Associated with Lower Incidence of Type 2 Diabetes. J. Nutr. 2013, 143, 1474–1480. [Google Scholar] [CrossRef]
  33. Yang, W.-S.; Wang, W.-Y.; Fan, W.-Y.; Deng, Q.; Wang, X. Tea consumption and risk of type 2 diabetes: A dose–response meta-analysis of cohort studies. Br. J. Nutr. 2014, 111, 1329–1339. [Google Scholar] [CrossRef]
  34. Wedick, N.; Pan, A.; Cassidy, A.; Rimm, E.; Sampson, L.; Rosner, B.; Willett, W.; Hu, F.; Sun, Q.; Dam, R. Dietary flavonoid intakes and risk of type 2 diabetes in US men and women. Am. J. Clin. Nutr. 2012, 95, 925–933. [Google Scholar] [CrossRef]
  35. Montonen, J.; Järvinen, R.; Heliövaara, M.; Reunanen, A.; Aromaa, A.; Knekt, P. Food consumption and the incidence of type II diabetes mellitus. Eur. J. Clin. Nutr. 2005, 59, 441–448. [Google Scholar] [CrossRef] [Green Version]
  36. Muraki, I.; Imamura, F.; E Manson, J.; Hu, F.B.; Willett, W.C.; Van Dam, R.M.; Sun, Q. Fruit consumption and risk of type 2 diabetes: Results from three prospective longitudinal cohort studies. BMJ 2013, 347, 5001. [Google Scholar] [CrossRef]
  37. Grosso, G.; Stepaniak, U.; Micek, A.; Kozela, M.; Stefler, D.; Bobak, M.; Pajak, A. Dietary polyphenol intake and risk of type 2 diabetes in the Polish arm of the Health, Alcohol and Psychosocial factors in Eastern Europe (HAPIEE) study. Br. J. Nutr. 2017, 118, 60–68. [Google Scholar] [CrossRef] [Green Version]
  38. Zamora-Ros, R.; Knaze, V.; Lujan-Barroso, L.; Slimani, N.; Romieu, I.; Fedirko, V.; De Magistris, M.S.; Ericson, U.; Amiano, P.; Trichopoulou, A.; et al. Estimated dietary intakes of flavonols, flavanones and flavones in the European Prospective Investigation into Cancer and Nutrition (EPIC) 24 hour dietary recall cohort. Br. J. Nutr. 2011, 106, 1915–1925. [Google Scholar] [CrossRef] [Green Version]
  39. Mursu, J.; Virtanen, J.K.; Tuomainen, T.-P.; Nurmi, T.; Voutilainen, S. Intake of fruit, berries, and vegetables and risk of type 2 diabetes in Finnish men: The Kuopio Ischaemic Heart Disease Risk Factor Study. Am. J. Clin. Nutr. 2014, 99, 328–333. [Google Scholar] [CrossRef]
  40. Yuan, S.; Li, X.; Jin, Y.; Lu, J. Chocolate Consumption and Risk of Coronary Heart Disease, Stroke, and Diabetes: A Meta-Analysis of Prospective Studies. Nutrients 2017, 9, 688. [Google Scholar] [CrossRef]
  41. Yang, J.; Mao, Q.-X.; Xu, H.-X.; Ma, X.; Zeng, C.-Y. Tea consumption and risk of type 2 diabetes mellitus: A systematic review and meta-analysis update. BMJ Open 2014, 4, 005632. [Google Scholar] [CrossRef]
  42. Hotamisligil, G.; Shargill, N.; Spiegelman, B. Adipose expression of tumor necrosis factor-alpha: Direct role in obesity-linked insulin resistance. Science 1993, 259, 87–91. [Google Scholar] [CrossRef]
  43. Feinstein, R.; Kanety, H.; Papa, M.Z.; Lunenfeld, B.; Karasik, A. Tumor necrosis factor-alpha suppresses insulin-induced tyrosine phosphorylation of insulin receptor and its substrates. J. Biol. Chem. 1993, 268, 26055–26058. [Google Scholar]
  44. Shoelson, S.E.; Lee, J.; Goldfine, A.B. Inflammation and insulin resistance. J. Clin. Investig. 2006, 116, 2308. [Google Scholar] [CrossRef]
  45. Bertoia, M.L.; Rimm, E.B.; Mukamal, K.J.; Hu, F.B.; Willett, W.C.; Cassidy, A. Dietary flavonoid intake and weight maintenance: Three prospective cohorts of 124,086 US men and women followed for up to 24 years. BMJ 2016, 352, i17. [Google Scholar] [CrossRef]
  46. Jennings, A.; MacGregor, A.; Spector, T.; Cassidy, A. Higher dietary flavonoid intakes are associated with lower objectively measured body composition in women: Evidence from discordant monozygotic twins12. Am. J. Clin. Nutr. 2017, 105, 626–634. [Google Scholar] [CrossRef]
  47. Jennings, A.; Welch, A.A.; Spector, T.; Macgregor, A.; Cassidy, A. Intakes of Anthocyanins and Flavones Are Associated with Biomarkers of Insulin Resistance and Inflammation in Women. J. Nutr. 2014, 144, 202–208. [Google Scholar] [CrossRef]
  48. Cassidy, A.; Rogers, G.; Peterson, J.J.; Dwyer, J.T.; Lin, H.; Jacques, P.F. Higher dietary anthocyanin and flavonol intakes are associated with anti-inflammatory effects in a population of US adults1. Am. J. Clin. Nutr. 2015, 102, 172–181. [Google Scholar] [CrossRef] [Green Version]
  49. Fang, J.-Y.; Lin, C.-H.; Huang, T.-H.; Chuang, S.-Y. In Vivo Rodent Models of Type 2 Diabetes and Their Usefulness for Evaluating Flavonoid Bioactivity. Nutrients 2019, 11, 530. [Google Scholar] [CrossRef]
  50. Różańska, D.; Regulska-Ilow, B. The significance of anthocyanins in the prevention and treatment of type 2 diabetes. Adv. Clin. Exp. Med. 2018, 27, 135–142. [Google Scholar] [CrossRef] [Green Version]
  51. Edirisinghe, I.; Burton-Freeman, B. Anti-diabetic actions of Berry polyphenols – Review on proposed mechanisms of action. J. Berry Res. 2016, 6, 237–250. [Google Scholar] [CrossRef] [Green Version]
  52. Yamashita, Y.; Okabe, M.; Natsume, M.; Ashida, H. Prevention mechanisms of glucose intolerance and obesity by cacao liquor procyanidin extract in high-fat diet-fed C57BL/6 mice. Arch. Biochem. Biophys. 2012, 527, 95–104. [Google Scholar] [CrossRef]
  53. Sancho, R.A.S.; Pastore, G.M. Evaluation of the effects of anthocyanins in type 2 diabetes. Res. Int. 2012, 46, 378–386. [Google Scholar] [CrossRef]
  54. Scazzocchio, B.; Varì, R.; Filesi, C.; D’Archivio, M.; Santangelo, C.; Giovannini, C.; Iacovelli, A.; Silecchia, G.; Volti, G.L.; Galvano, F.; et al. Cyanidin-3-O-β-glucoside and protocatechuic acid exert insulin-like effects by upregulating PPARγ activity in human omental adipocytes. Diabetes 2011, 60, 2234–2244. [Google Scholar] [CrossRef]
  55. Shi, G.-J.; Li, Y.; Cao, Q.-H.; Wu, H.-X.; Tang, X.-Y.; Gao, X.-H.; Yu, J.-Q.; Chen, Z.; Yang, Y. In vitro and in vivo evidence that quercetin protects against diabetes and its complications: A systematic review of the literature. Biomed. Pharmacother. 2019, 109, 1085–1099. [Google Scholar] [CrossRef] [PubMed]
  56. Gilbert, E.R.; Liu, D. Anti-diabetic functions of soy isoflavone genistein: Mechanisms underlying its effects on pancreatic β-cell function. Food Funct. 2013, 4, 200–212. [Google Scholar] [CrossRef]
  57. Lila, M.A.; Burton-Freeman, B.; Grace, M.; Kalt, W. Unraveling Anthocyanin Bioavailability for Human Health. Annu. Rev. Food Sci. Technol. 2016, 7, 375–393. [Google Scholar] [CrossRef]
  58. Koponen, J.M.; Happonen, A.M.; Mattila, P.H.; Törrönen, A.R. Contents of Anthocyanins and Ellagitannins in Selected Foods Consumed in Finland. J. Agric. Food Chem. 2007, 55, 1612–1619. [Google Scholar] [CrossRef] [PubMed]
  59. Pojer, E.; Mattivi, F.; Johnson, D.; Stockley, C.S. The Case for Anthocyanin Consumption to Promote Human Health: A Review. Compr. Rev. Sci. Saf. 2013, 12, 483–508. [Google Scholar] [CrossRef] [Green Version]
  60. Burton-Freeman, B.M.; Guenther, P.M.; Oh, M.; Stuart, D.; Jensen, H.H. Assessing the consumption of berries and associated factors in the United States using the National Health and Nutrition Examination Survey (NHANES), 2007–2012. Food Funct. 2018, 9, 1009–1016. [Google Scholar] [CrossRef]
  61. Turrini, E.; Ferruzzi, L.; Fimognari, C. Possible Effects of Dietary Anthocyanins on Diabetes and Insulin Resistance. Curr. Drug Targets 2017, 18, 629–640. [Google Scholar] [CrossRef]
  62. Tsuda, T.; Klimis-Zacas, D. Recent Progress in Anti-Obesity and Anti-Diabetes Effect of Berries. Antioxidants 2016, 5, 13. [Google Scholar] [CrossRef]
  63. Cassidy, A. Berry anthocyanin intake and cardiovascular health. Mol. Asp. Med. 2018, 61, 76–82. [Google Scholar] [CrossRef]
  64. Yan, F.; Dai, G.; Zheng, X. Mulberry anthocyanin extract ameliorates insulin resistance by regulating PI3K/AKT pathway in HepG2 cells and db/db mice. J. Nutr. Biochem. 2016, 36, 68–80. [Google Scholar] [CrossRef]
  65. Castro-Acosta, M.; Hall, W.; Corpe, C. Polyphenol-rich blackcurrant and apple extracts inhibit glucose uptake in in vitro models of intestinal sugar transport, but individual anthocyanins have no effect. Proc. Nutr. Soc. 2016, 75, 75. [Google Scholar] [CrossRef]
  66. Moser, S.; Lim, J.; Chegeni, M.; Wightman, J.D.; Hamaker, B.R.; Ferruzzi, M.G. Concord and Niagara Grape Juice and Their Phenolics Modify Intestinal Glucose Transport in a Coupled in Vitro Digestion/Caco-2 Human Intestinal Model. Nutrients 2016, 8, 414. [Google Scholar] [CrossRef] [PubMed]
  67. Kamiloglu, S.; Capanoglu, E.; Grootaert, C.; Van Camp, J.; Battino, M. Anthocyanin Absorption and Metabolism by Human Intestinal Caco-2 Cells—A Review. Int. J. Mol. Sci. 2015, 16, 21555–21574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Suantawee, T.; Elazab, S.; Hsu, W.; Yao, S.; Cheng, H.; Adisakwattana, S. Cyanidin Stimulates Insulin Secretion and Pancreatic β-Cell Gene Expression through Activation of l-type Voltage-Dependent Ca2+ Channels. Nutrients 2017, 9, 814. [Google Scholar] [CrossRef]
  69. Luna-Vital, D.A.; De Mejia, E.G. Anthocyanins from purple corn activate free fatty acid-receptor 1 and glucokinase enhancing in vitro insulin secretion and hepatic glucose uptake. PLoS ONE 2018, 13, e0200449. [Google Scholar] [CrossRef]
  70. Kurimoto, Y.; Shibayama, Y.; Inoue, S.; Soga, M.; Takikawa, M.; Ito, C.; Nanba, F.; Yoshida, T.; Yamashita, Y.; Ashida, H.; et al. Black Soybean Seed Coat Extract Ameliorates Hyperglycemia and Insulin Sensitivity via the Activation of AMP-Activated Protein Kinase in Diabetic Mice. J. Agric. Food Chem. 2013, 61, 5558–5564. [Google Scholar] [CrossRef] [PubMed]
  71. Xing, T.; Kang, Y.; Xu, X.; Wang, B.; Du, M.; Zhu, M.-J. Raspberry Supplementation Improves Insulin Signaling and Promotes Brown-Like Adipocyte Development in White Adipose Tissue of Obese Mice. Mol. Nutr. Res. 2018, 62, 1701035. [Google Scholar] [CrossRef]
  72. Bispo, K.; Amusquivar, E.; García-Seco, D.; Ramos-Solano, B.; Gutiérrez-Mañero, J.; Herrera, E. Supplementing Diet with Blackberry Extract Causes a Catabolic Response with Increments in Insulin Sensitivity in Rats. Plant Foods Hum. Nutr. 2015, 70, 170–175. [Google Scholar] [CrossRef] [PubMed]
  73. Tani, T.; Nishikawa, S.; Kato, M.; Tsuda, T. Delphinidin 3-rutinoside-rich blackcurrant extract ameliorates glucose tolerance by increasing the release of glucagon-like peptide-1 secretion. Food Sci. Nutr. 2017, 5, 929–933. [Google Scholar] [CrossRef]
  74. Prior, R.L.; Wilkes, S.E.; Rogers, T.R.; Khanal, R.C.; Wu, X.; Howard, L.R. Purified Blueberry Anthocyanins and Blueberry Juice Alter Development of Obesity in Mice Fed an Obesogenic High-Fat Diet†. J. Agric. Food Chem. 2010, 58, 3970–3976. [Google Scholar] [CrossRef]
  75. Paquette, M.; Larqué, A.S.M.; Weisnagel, S.J.; Desjardins, Y.; Marois, J.; Pilon, G.; Dudonné, S.; Marette, A.; Jacques, H. Strawberry and cranberry polyphenols improve insulin sensitivity in insulin-resistant, non-diabetic adults: A parallel, double-blind, controlled and randomised clinical trial. Br. J. Nutr. 2017, 117, 519–531. [Google Scholar] [CrossRef]
  76. Stull, A.J.; Cash, K.C.; Johnson, W.D.; Champagne, C.M.; Cefalu, W.T. Bioactives in Blueberries Improve Insulin Sensitivity in Obese, Insulin-Resistant Men and Women. J. Nutr. 2010, 140, 1764–1768. [Google Scholar] [CrossRef]
  77. Basu, A.; Du, M.; Leyva, M.J.; Sanchez, K.; Betts, N.M.; Wu, M.; Aston, C.E.; Lyons, T.J. Blueberries Decrease Cardiovascular Risk Factors in Obese Men and Women with Metabolic Syndrome123. J. Nutr. 2010, 140, 1582–1587. [Google Scholar] [CrossRef] [Green Version]
  78. Moazzen, H.; Alizadeh, M. Effects of Pomegranate Juice on Cardiovascular Risk Factors in Patients with Metabolic Syndrome: A Double-Blinded, Randomized Crossover Controlled Trial. Plant Foods Hum. Nutr. 2017, 5, 115–133. [Google Scholar] [CrossRef] [PubMed]
  79. Ellis, C.L.; Edirisinghe, I.; Kappagoda, T.; Burton-Freeman, B. Attenuation of Meal-Induced Inflammatory and Thrombotic Responses in Overweight Men and Women After 6-Week Daily Strawberry (Fragaria) Intake. J. Atheroscler. Thromb. 2011, 18, 318–327. [Google Scholar] [CrossRef] [Green Version]
  80. De Mello, V.D.F.; Lankinen, M.A.; Lindström, J.; Puupponen-Pimiä, R.; Laaksonen, D.E.; Pihlajamäki, J.; Lehtonen, M.; Uusitupa, M.; Tuomilehto, J.; Kolehmainen, M.; et al. Fasting serum hippuric acid is elevated after bilberry (Vaccinium myrtillus) consumption and associates with improvement of fasting glucose levels and insulin secretion in persons at high risk of developing type 2 diabetes. Mol. Nutr. Food Res. 2017, 61, 1700019. [Google Scholar] [CrossRef] [PubMed]
  81. Kerimi, A.; Nyambe-Silavwe, H.; Gauer, J.S.; Tomás-Barberán, F.A.; Williamson, G. Pomegranate juice, but not an extract, confers a lower glycemic response on a high–glycemic index food: Randomized, crossover, controlled trials in healthy subjects. Am. J. Clin. Nutr. 2017, 106, 1384–1393. [Google Scholar] [CrossRef]
  82. Törrönen, R.; Kolehmainen, M.; Sarkkinen, E.; Poutanen, K.; Mykkänen, H.; Niskanen, L. Berries Reduce Postprandial Insulin Responses to Wheat and Rye Breads in Healthy Women. J. Nutr. 2013, 143, 430–436. [Google Scholar] [CrossRef] [Green Version]
  83. Edirisinghe, I.; Banaszewski, K.; Cappozzo, J.; Sandhya, K.; Ellis, C.L.; Tadapaneni, R.; Kappagoda, C.T.; Burton-Freeman, B.M. Strawberry anthocyanin and its association with postprandial inflammation and insulin. Br. J. Nutr. 2011, 106, 913–922. [Google Scholar] [CrossRef] [Green Version]
  84. Solverson, P.M.; Rumpler, W.V.; Leger, J.L.; Redan, B.W.; Ferruzzi, M.G.; Baer, D.J.; Castonguay, T.W.; Novotny, J.A. Blackberry Feeding Increases Fat Oxidation and Improves Insulin Sensitivity in Overweight and Obese Males. Nutrients 2018, 10, 1048. [Google Scholar] [CrossRef] [PubMed]
  85. Stote, K.; Corkum, A.; Sweeney, M.; Shakerley, N.; Kean, T.; Gottschall-Pass, K. Postprandial Effects of Blueberry (Vaccinium angustifolium) Consumption on Glucose Metabolism, Gastrointestinal Hormone Response, and Perceived Appetite in Healthy Adults: A Randomized, Placebo-Controlled Crossover Trial. Nutrients 2019, 11, 202. [Google Scholar] [CrossRef]
  86. Törrönen, R.; Sarkkinen, E.; Niskanen, T.; Tapola, N.; Kilpi, K.; Niskanen, L. Postprandial glucose, insulin and glucagon-like peptide 1 responses to sucrose ingested with berries in healthy subjects. Br. J. Nutr. 2012, 107, 1445–1451. [Google Scholar] [CrossRef] [PubMed]
  87. Törrönen, R.; Sarkkinen, E.; Tapola, N.; Hautaniemi, E.; Kilpi, K.; Niskanen, L. Berries modify the postprandial plasma glucose response to sucrose in healthy subjects. Br. J. Nutr. 2009, 103, 1. [Google Scholar] [CrossRef] [Green Version]
  88. Bell, L.; Lamport, D.J.; Butler, L.T.; Williams, C.M. A study of glycaemic effects following acute anthocyanin-rich blueberry supplementation in healthy young adults. Food Funct. 2017, 8, 3104–3110. [Google Scholar] [CrossRef]
  89. Castro-Acosta, M.L.; Smith, L.; Miller, R.J.; McCarthy, D.I.; Farrimond, J.A.; Hall, W.L. Drinks containing anthocyanin-rich blackcurrant extract decrease postprandial blood glucose, insulin and incretin concentrations. J. Nutr. Biochem. 2016, 38, 154–161. [Google Scholar] [CrossRef] [PubMed]
  90. Burton-Freeman, B.M.; Huang, Y.; Park, E.; Edirisinghe, I. Maximizing the health effects of strawberry anthocyanins: Understanding the influence of the consumption timing variable. Food Funct. 2016, 7, 4745–4752. [Google Scholar]
  91. Burton-Freeman, B.; Linares, A.; Hyson, D.; Kappagoda, T. Strawberry Modulates LDL Oxidation and Postprandial Lipemia in Response to High-Fat Meal in Overweight Hyperlipidemic Men and Women. J. Am. Coll. Nutr. 2010, 29, 46–54. [Google Scholar] [CrossRef]
  92. Park, E.; Edirisinghe, I.; Wei, H.; Vijayakumar, L.P.; Banaszewski, K.; Cappozzo, J.C.; Burton-Freeman, B. A dose-response evaluation of freeze-dried strawberries independent of fiber content on metabolic indices in abdominally obese individuals with insulin resistance in a randomized, single-blinded, diet-controlled crossover trial. Mol. Nutr. Food Res. 2016, 60, 1099–1109. [Google Scholar] [CrossRef]
  93. Xiao, D.; Zhu, L.; Edirisinghe, I.; Fareed, J.; Brailovsky, Y.; Burton-Freeman, B. Attenuation of Postmeal Metabolic Indices with Red Raspberries in Individuals at Risk for Diabetes: A Randomized Controlled Trial. Obesity 2019, 27, 542–550. [Google Scholar] [CrossRef] [PubMed]
  94. Yang, L.; Ling, W.; Du, Z.; Chen, Y.; Li, D.; Deng, S.; Liu, Z.; Yang, L. Effects of Anthocyanins on Cardiometabolic Health: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. Adv. Nutr. Int. Rev. J. 2017, 8, 684–693. [Google Scholar] [CrossRef] [Green Version]
  95. García-Conesa, M.-T.; Chambers, K.; Combet, E.; Pinto, P.; Garcia-Aloy, M.; Andrés-Lacueva, C.; De Pascual-Teresa, S.; Mena, P.; Ristic, A.K.; Hollands, W.J.; et al. Meta-Analysis of the Effects of Foods and Derived Products Containing Ellagitannins and Anthocyanins on Cardiometabolic Biomarkers: Analysis of Factors Influencing Variability of the Individual Responses. Int. J. Mol. Sci. 2018, 19, 694. [Google Scholar] [CrossRef] [PubMed]
  96. Li, D.; Zhang, Y.; Liu, Y.; Sun, R.; Xia, M. Purified Anthocyanin Supplementation Reduces Dyslipidemia, Enhances Antioxidant Capacity, and Prevents Insulin Resistance in Diabetic Patients. J. Nutr. 2015, 145, 742–748. [Google Scholar] [CrossRef] [Green Version]
  97. Moazen, S.; Amani, R.; Rad, A.H.; Ahmadi, K.; Jalali, M.-T.; Shahbazian, H. Effects of Freeze-Dried Strawberry Supplementation on Metabolic Biomarkers of Atherosclerosis in Subjects with Type 2 Diabetes: A Randomized Double-Blind Controlled Trial. Ann. Nutr. Metab. 2013, 63, 256–264. [Google Scholar] [CrossRef]
  98. Hammerbacher, A.; Paetz, C.; Wright, L.P.; Fischer, T.C.; Bohlmann, J.; Davis, A.J.; Fenning, T.M.; Gershenzon, J.; Schmidt, A. Flavan-3-ols in Norway Spruce: Biosynthesis, Accumulation, and Function in Response to Attack by the Bark Beetle-Associated Fungus Ceratocystis polonica. Plant Physiol. 2014, 164, 2107–2122. [Google Scholar] [CrossRef]
  99. Ferreira, D.; Slade, D. Oligomeric proanthocyanidins: Naturally occurring O-heterocycles. Nat. Prod. Rep. 2002, 19, 517–541. [Google Scholar] [CrossRef] [PubMed]
  100. De Pascual-Teresa, S.; Santos-Buelga, C.; Rivas-Gonzalo, J.-C. Quantitative Analysis of Flavan-3-ols in Spanish Foodstuffs and Beverages. J. Agric. Food Chem. 2000, 48, 5331–5337. [Google Scholar] [CrossRef] [PubMed]
  101. Aron, P.M.; Kennedy, J.A. Flavan-3-ols: Nature, occurrence and biological activity. Mol. Nutr. Food Res. 2008, 52, 79–104. [Google Scholar] [CrossRef]
  102. Calani, L.; Dall’Asta, M.; Bruni, R.; Rio, D. Del Flavonoid Occurrence, Bioavailability, Metabolism, and Protective Effects in Humans: Focus on Flavan-3-ols and Flavonols. In Recent Advances in Polyphenol Research; John Wiley & Sons, Ltd.: Chichester, UK, 2014; pp. 239–279. [Google Scholar]
  103. An, Z.; Qi, Y.; Huang, D.; Gu, X.; Tian, Y.; Li, P.; Li, H.; Zhang, Y. EGCG inhibits Cd(2+)-induced apoptosis through scavenging ROS rather than chelating Cd(2+) in HL-7702 cells. Toxicol. Mech. Methods 2014, 24, 259–267. [Google Scholar] [CrossRef]
  104. Yan, J.; Zhao, Y.; Suo, S.; Liu, Y.; Zhao, B. Green tea catechins ameliorate adipose insulin resistance by improving oxidative stress. Free Radic. Biol. Med. 2012, 52, 1648–1657. [Google Scholar] [CrossRef]
  105. Sür-Altiner D; Yenice B Effect of black tea on lipid peroxidation in carbon tetrachloride treated male rats. Drug Metab. Drug Interact. 2000, 16, 123–128.
  106. Hooper, L.; A Kroon, P.; Rimm, E.B.; Cohn, J.S.; Harvey, I.; A Le Cornu, K.; Ryder, J.J.; Hall, W.L.; Cassidy, A. Flavonoids, flavonoid-rich foods, and cardiovascular risk: A meta-analysis of randomized controlled trials. Am. J. Clin. Nutr. 2008, 88, 38–50. [Google Scholar] [CrossRef] [PubMed]
  107. Shrime, M.G.; Bauer, S.R.; McDonald, A.C.; Chowdhury, N.H.; Coltart, C.E.M.; Ding, E.L. Flavonoid-Rich Cocoa Consumption Affects Multiple Cardiovascular Risk Factors in a Meta-Analysis of Short-Term Studies. J. Nutr. 2011, 141, 1982–1988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Lin, X.; Zhang, I.; Li, A.; Manson, J.E.; Sesso, H.D.; Wang, L.; Liu, S. Cocoa Flavanol Intake and Biomarkers for Cardiometabolic Health: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. J. Nutr. 2016, 146, 2325–2333. [Google Scholar] [CrossRef] [PubMed]
  109. Curtis, P.J.; Sampson, M.; Potter, J.; Dhatariya, K.; Kroon, P.A.; Cassidy, A. Chronic ingestion of flavan-3-ols and isoflavones improves insulin sensitivity and lipoprotein status and attenuates estimated 10-year CVD risk in medicated postmenopausal women with type 2 diabetes: A 1-year, double-blind, randomized, controlled trial. Diabetes Care 2012, 35, 226–232. [Google Scholar] [CrossRef] [PubMed]
  110. Grassi, D.; Desideri, G.; Necozione, S.; Lippi, C.; Casale, R.; Properzi, G.; Blumberg, J.B.; Ferri, C. Blood Pressure Is Reduced and Insulin Sensitivity Increased in Glucose-Intolerant, Hypertensive Subjects after 15 Days of Consuming High-Polyphenol Dark Chocolate. J. Nutr. 2008, 138, 1671–1676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Park, E.; Edirisinghe, I.; Choy, Y.Y.; Waterhouse, A.; Burton-Freeman, B. Effects of grape seed extract beverage on blood pressure and metabolic indices in individuals with pre-hypertension: A randomised, double-blinded, two-arm, parallel, placebo-controlled trial. Br. J. Nutr. 2016, 115, 226–238. [Google Scholar] [CrossRef]
  112. Fukino, Y.; Ikeda, A.; Maruyama, K.; Aoki, N.; Okubo, T.; Iso, H. Randomized controlled trial for an effect of green tea-extract powder supplementation on glucose abnormalities. Eur. J. Clin. Nutr. 2008, 62, 953–960. [Google Scholar] [CrossRef] [PubMed]
  113. Hsu, C.-H.; Liao, Y.-L.; Lin, S.-C.; Tsai, T.-H.; Huang, C.-J.; Chou, P. Does supplementation with green tea extract improve insulin resistance in obese type 2 diabetics? A randomized, double-blind, and placebo-controlled clinical trial. Altern. Med. Rev. 2011, 16, 157–163. [Google Scholar] [CrossRef]
  114. Liu, C.-Y.; Huang, C.-J.; Huang, L.-H.; Chen, I.-J.; Chiu, J.-P.; Hsu, C.-H. Effects of Green Tea Extract on Insulin Resistance and Glucagon-Like Peptide 1 in Patients with Type 2 Diabetes and Lipid Abnormalities: A Randomized, Double-Blinded, and Placebo-Controlled Trial. PLoS ONE 2014, 9, e91163. [Google Scholar] [CrossRef]
  115. Brown, A.L.; Lane, J.; Coverly, J.; Stocks, J.; Jackson, S.; Stephen, A.; Bluck, L.; Coward, A.; Hendrickx, H. Effects of dietary supplementation with the green tea polyphenol epigallocatechin-3-gallate on insulin resistance and associated metabolic risk factors: Randomized controlled trial. Br. J. Nutr. 2009, 101, 886–894. [Google Scholar] [CrossRef]
  116. Basu, A.; Du, M.; Sanchez, K.; Leyva, M.J.; Betts, N.M.; Blevins, S.; Wu, M.; Aston, C.E.; Lyons, T.J. Green Tea minimally affects Biomarkers of Inflammation in Obese Subjects with Metabolic Syndrome. Nutrition 2011, 27, 206–213. [Google Scholar] [CrossRef] [Green Version]
  117. Leo, C.-H.; Woodman, O.L. Flavonols in the Prevention of Diabetes-induced Vascular Dysfunction. J. Cardiovasc. Pharmacol. 2015, 65, 532–544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Chen, S.; Jiang, H.; Wu, X.; Fang, J. Therapeutic Effects of Quercetin on Inflammation, Obesity, and Type 2 Diabetes. Mediat. Inflamm. 2016, 2016, 1–5. [Google Scholar] [CrossRef] [PubMed]
  119. Bule, M.; Abdurahman, A.; Nikfar, S.; Abdollahi, M.; Amini, M. Antidiabetic effect of quercetin: A systematic review and meta-analysis of animal studies. Food Chem. Toxicol. 2019, 125, 494–502. [Google Scholar] [CrossRef] [PubMed]
  120. Bloedon, L.T.; Jeffcoat, A.R.; Lopaczynski, W.; Schell, M.J.; Black, T.M.; Dix, K.J.; Thomas, B.F.; Albright, C.; Busby, M.G.; A Crowell, J.; et al. Safety and pharmacokinetics of purified soy isoflavones: Single-dose administration to postmenopausal women. Am. J. Clin. Nutr. 2002, 76, 1126–1137. [Google Scholar] [CrossRef] [PubMed]
  121. Liu, Z.M.; Chen, Y.M.; Ho, S.C.; Ho, Y.P.; Woo, J. Effects of soy protein and isoflavones on glycemic control and insulin sensitivity: A 6-mo double-blind, randomized, placebo-controlled trial in postmenopausal Chinese women with prediabetes or untreated early diabetes. Am. J. Clin. Nutr. 2010, 91, 1394–1401. [Google Scholar] [CrossRef] [PubMed]
  122. Jayagopal, V.; Albertazzi, P.; Kilpatrick, E.S.; Howarth, E.M.; Jennings, P.E.; Hepburn, D.A.; Atkin, S. Beneficial Effects of Soy Phytoestrogen Intake in Postmenopausal Women with Type 2 Diabetes. Diabetes Care 2002, 25, 1709–1714. [Google Scholar] [CrossRef]
  123. Glisic, M.; Kastrati, N.; Musa, J.; Milic, J.; Asllanaj, E.; Fernandez, E.P.; Nano, J.; Rosales, C.O.; Amiri, M.; Kraja, B.; et al. Phytoestrogen supplementation and body composition in postmenopausal women: A systematic review and meta-analysis of randomized controlled trials. Maturitas 2018, 115, 74–83. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Path to Diabetes/T2DM: A. normal, non-disease/diabetes state (based on fasting glucose 70–99 mg/dL). B. declining insulin sensitivity/increasing insulin resistance, and increasing concentrations of insulin (hyperinsulinemia) to maintain normal glucose: Opportunity * for early detection and intervention to restore normal insulin sensitivity and decrease beta cell burden. C. pre-diabetes identified (fasting glucose 100–125 mg/dL, [1], hyperinsulinemia insufficient to overcome insulin resistance. Diet and lifestyle intervention prescribed. D. T2DM diagnosed (fasting glucose ≥126 mg/dL or 2-h post glucose ≥140 mg/dL, [1], diet and lifestyle and medications prescribed.
Figure 1. Path to Diabetes/T2DM: A. normal, non-disease/diabetes state (based on fasting glucose 70–99 mg/dL). B. declining insulin sensitivity/increasing insulin resistance, and increasing concentrations of insulin (hyperinsulinemia) to maintain normal glucose: Opportunity * for early detection and intervention to restore normal insulin sensitivity and decrease beta cell burden. C. pre-diabetes identified (fasting glucose 100–125 mg/dL, [1], hyperinsulinemia insufficient to overcome insulin resistance. Diet and lifestyle intervention prescribed. D. T2DM diagnosed (fasting glucose ≥126 mg/dL or 2-h post glucose ≥140 mg/dL, [1], diet and lifestyle and medications prescribed.
Nutrients 11 00841 g001
Figure 2. Flavonoid compounds, dietary sources and daily intake range. 1 Aglycone values [25].
Figure 2. Flavonoid compounds, dietary sources and daily intake range. 1 Aglycone values [25].
Nutrients 11 00841 g002
Figure 3. Purported mechanisms of action of flavonoids with effects on glucose metabolism: AMPK, AMP-activated protein kinase; CHO, carbohydrate; GK, Glucokinase; GLUT 4, Glucose transporter type 4; GNG, Gluconeogenesis; GS, Glutamine synthetase; IRS-1, Insulin receptor substrate 1; JNK, c-Jun N-terminal kinases; Ox, Oxidative; PI3/AKT, phosphatidylinositol 3-kinase (PI3K) and AKT/Protein Kinase B signaling pathway; SGLT, Sodium-dependent glucose transporter; (+), increase, activation; (−), decrease, inactivation.
Figure 3. Purported mechanisms of action of flavonoids with effects on glucose metabolism: AMPK, AMP-activated protein kinase; CHO, carbohydrate; GK, Glucokinase; GLUT 4, Glucose transporter type 4; GNG, Gluconeogenesis; GS, Glutamine synthetase; IRS-1, Insulin receptor substrate 1; JNK, c-Jun N-terminal kinases; Ox, Oxidative; PI3/AKT, phosphatidylinositol 3-kinase (PI3K) and AKT/Protein Kinase B signaling pathway; SGLT, Sodium-dependent glucose transporter; (+), increase, activation; (−), decrease, inactivation.
Nutrients 11 00841 g003

Share and Cite

MDPI and ACS Style

Burton-Freeman, B.; Brzeziński, M.; Park, E.; Sandhu, A.; Xiao, D.; Edirisinghe, I. A Selective Role of Dietary Anthocyanins and Flavan-3-ols in Reducing the Risk of Type 2 Diabetes Mellitus: A Review of Recent Evidence. Nutrients 2019, 11, 841. https://doi.org/10.3390/nu11040841

AMA Style

Burton-Freeman B, Brzeziński M, Park E, Sandhu A, Xiao D, Edirisinghe I. A Selective Role of Dietary Anthocyanins and Flavan-3-ols in Reducing the Risk of Type 2 Diabetes Mellitus: A Review of Recent Evidence. Nutrients. 2019; 11(4):841. https://doi.org/10.3390/nu11040841

Chicago/Turabian Style

Burton-Freeman, Britt, Michał Brzeziński, Eunyoung Park, Amandeep Sandhu, Di Xiao, and Indika Edirisinghe. 2019. "A Selective Role of Dietary Anthocyanins and Flavan-3-ols in Reducing the Risk of Type 2 Diabetes Mellitus: A Review of Recent Evidence" Nutrients 11, no. 4: 841. https://doi.org/10.3390/nu11040841

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop