Next Article in Journal
Amelioration of Obesity-Related Disorders in High-Fat Diet-Fed Mice following Fecal Microbiota Transplantation from Inulin-Dosed Mice
Next Article in Special Issue
Integration of Deep Learning and Sequential Metabolism to Rapidly Screen Dipeptidyl Peptidase (DPP)-IV Inhibitors from Gardenia jasminoides Ellis
Previous Article in Journal
Microwave-Assisted Synthesis of Aminophosphonic Derivatives and Their Antifungal Evaluation against Lomentospora prolificans
Previous Article in Special Issue
Anthocyanins: Potential Therapeutic Approaches towards Obesity and Diabetes Mellitus Type 2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Food Polyphenols and Type II Diabetes Mellitus: Pharmacology and Mechanisms

1
Department of Pharmacology, Faculty of Pharmacy, Bahauddin Zakariya University, Multan 60000, Pakistan
2
Department of Medical Lab Sciences, Faculty of Allied Medical Sciences, Zarqa University, Zarqa 13110, Jordan
3
Department of Chemistry, The University of Jordan, Amman 11942, Jordan
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(10), 3996; https://doi.org/10.3390/molecules28103996
Submission received: 3 April 2023 / Revised: 4 May 2023 / Accepted: 7 May 2023 / Published: 10 May 2023
(This article belongs to the Special Issue Novel Natural Compounds in Treatment of Diabetes)

Abstract

:
Type II diabetes mellitus and its related complications are growing public health problems. Many natural products present in our diet, including polyphenols, can be used in treating and managing type II diabetes mellitus and different diseases, owing to their numerous biological properties. Anthocyanins, flavonols, stilbenes, curcuminoids, hesperidin, hesperetin, naringenin, and phenolic acids are common polyphenols found in blueberries, chokeberries, sea-buckthorn, mulberries, turmeric, citrus fruits, and cereals. These compounds exhibit antidiabetic effects through different pathways. Accordingly, this review presents an overview of the most recent developments in using food polyphenols for managing and treating type II diabetes mellitus, along with various mechanisms. In addition, the present work summarizes the literature about the anti-diabetic effect of food polyphenols and evaluates their potential as complementary or alternative medicines to treat type II diabetes mellitus. Results obtained from this survey show that anthocyanins, flavonols, stilbenes, curcuminoids, and phenolic acids can manage diabetes mellitus by protecting pancreatic β-cells against glucose toxicity, promoting β-cell proliferation, reducing β-cell apoptosis, and inhibiting α-glucosidases or α-amylase. In addition, these phenolic compounds exhibit antioxidant anti-inflammatory activities, modulate carbohydrate and lipid metabolism, optimize oxidative stress, reduce insulin resistance, and stimulate the pancreas to secrete insulin. They also activate insulin signaling and inhibit digestive enzymes, regulate intestinal microbiota, improve adipose tissue metabolism, inhibit glucose absorption, and inhibit the formation of advanced glycation end products. However, insufficient data are available on the effective mechanisms necessary to manage diabetes.

1. Introduction

Phytochemicals and polyphenols in fruits and vegetables have antidiabetic effects [1]. Plant-based nutrients such as vegetables (onion, cabbage, and especially broccoli), fruits (apples, grapes, cherries, pears, and various berries), and grains contain hundreds of different polyphenols [2,3,4]. In this context, some vegetables such as beans, cabbage, onions, and cereals also contain anthocyanidins, whereas red fruits are the primary source of these polyphenols [5]. The plant kingdom contains a large number of polyphenols that fall under the categories of tannins, lignans, stilbenes, phenolic acids, and flavonoids, among others [6]. On the other hand, fruits, spices, grains, vegetables, and other phenolic-rich plant products contain phenolic acids (hydroxycinnamic acids and hydroxybenzoic acid), stilbenes, and lignans [3,4,7]. Phenolics are crucial to fruit quality because they impact the fruit’s taste, appearance, and nutritional value [8]. For example, flavonoids may lessen the risk of developing diabetes [6] by maintaining glucose uptake, blood glucose points, and insulin secretion, controlling immune function [9,10]. In this respect, dietary flavonoids demonstrated a significant anti-hyperglycemic-like effect through glucose absorption control [11], a reserve of digestive enzymes [12,13], regulation of intestinal microbiota [14], inhibition of the formation of innovative glycation end products [15], and other mechanisms. Polyphenols may also influence the signaling pathways and ensuing alterations in gene expression [16,17]. By controlling the events of glucose metabolism, hepatic enzymes, and lipid profiles, flavonoids reduce the pathogenesis of diabetes and its complications [18]. Flavone C-glycosides, which can also hinder digestive enzymes and activate insulin signaling, can lessen the production of advanced glycation end products (AGEs) [19]. Accordingly, the consumption of purple carrots, high in anthocyanins (flavonoids) and low in carotenoids, was linked to a decrease in impaired glucose tolerance [20]. Quercetin, a flavonoid, has received the most research attention for its in vivo and cellular anti-diabetic properties in animal and cell models [21], followed by kaempferol [22], luteolin [23], myricetin [24], and naringenin [25]. The most well-known sources of the stilbenes class of polyphenols, including resveratrol, are mulberries, grape skin, and peanuts [26]. The numerous and diverse phytochemicals known as polyphenols contain phenolic rings [9]. In this regard, two aromatic rings are joined by a 3-carbon chain to form an oxygenated heterocyclic ring, and this structure makes up a class of phenolic compounds known as flavonoids [27]. Anthocyanins, flavonols, flavones, isoflavonoids, and syringic acid are flavonoid subclasses connected to diabetes because the consumption of food that contains these compounds lowers the risk of type II diabetes [28].
According to estimates, there will likely be over 300 million cases of type II diabetes worldwide by 2030 [29]. Therefore, medical professionals, academics, and policymakers are taking note of the rising number of fatalities brought on by diabetes, related illnesses, and physiological disorders to promote healthy eating habits [1]. Currently, preventing and treating metabolic syndrome and type II diabetes involves increasing physical activity and decreasing calorie intake [30]. Hyperglycemia is a metabolic disease with multiple underlying origins that necessitate lifetime medication therapy and dietary adjustments. In diabetes management and prevention, herbal supplements are now supported by a growing body of scientific research. Nutritional polyphenols, the most common phytochemical in human diets, have drawn much interest due to growing evidence of their positive effects on humans. Dietary polyphenols aid in the management of type II diabetes and lessen the severity of diabetic complications in animals. The anti-diabetic effects of resveratrol [31,32], curcumin [33], and anthocyanins [34] have been demonstrated in humans. Studies validate that these polyphenols conducted in vitro and in vivo compounds have anti-inflammatory, antioxidant, chemopreventive, and neuroprotective properties. Accordingly, and because of the wide range of preventive and therapeutic and preventive options of food polyphenols and their involvement in managing and preventing type II diabetes mellitus, this review discusses the chemopreventive and therapeutic ability of these natural polyphenols in treating and managing type II diabetes mellitus. In addition, the current work discusses the numerous mechanisms of action through which these polyphenols exert their antidiabetic effects.

2. Results

2.1. Pathogenesis of Type II Diabetes Mellitus

Over 400 million people worldwide have type II diabetes (T2D), regarded as a multifactorial and complex metabolic disorder [35,36,37,38]; T2D is a chronic inflammatory disease [37]. Insulin resistance, deficiency of insulin secretion, and reduction of its anabolic activity on target tissues alter the metabolism, and its reflected chronic metabolic disorder can lead to death [39]. Through its numerous organ complications, diabetes lowers the quality of life [40] and affects whole-body physiology [41]. In this regard, hormones such as insulin and glucagon [42,43], adipokines/lipokines (adiponectin [44], leptin [45], and adipsin [46]), metabolites (amino acids [42,47], such as alanine [48,49]), lipids, free fatty acids [49,50,51], and glucagon-like peptide-1 are known metabolic regulators that disturb metabolism by signaling to various nerves and are crucial for T2D [52]. Even though T2D is most frequently passed down through families, it does so because of the interaction between risk genes primarily expressed in insulin resistance in target organs and β-cells, many other forms of hyperglycemia have nongenetic causes [53]. Depicted in Figure 1 are the essential factors attenuating type II diabetes mellitus.

2.1.1. Adipokine and Pro-Inflammatory Cytokine Roles in Diabetes

An adipokine called adiponectin stimulates AMP-activated kinase (AMPK), which reduces gluconeogenesis and improves insulin sensitivity in the liver [54]. In addition to the liver, adiponectin also affects the muscles by triggering AMPK, increasing acetyl CoA carboxylase (ACC) phosphorylation, fatty acid oxidation, and glucose uptake [55,56]; adipokine aids in maintaining the homeostasis of energy [57,58]. In this context, inflammatory and metabolic diseases are complicated by the presence of molecules such as retinol-binding protein 4 (RBP4) [59], TNF-α [60,61,62], and others that interfere with homeostasis [58,59]. By producing myokines, skeletal muscles contribute significantly to the endocrine response and T2D [63]. The most well-known myokine with various functions in numerous tissues is IL-6, which is frequently linked to inflammatory processes. In a murine model, IL-6 enhanced insulin signaling via AKT while inhibiting the expression of gluconeogenic genes [64]. In addition, IL-6 increased fat oxidation and lipolysis in adipose tissue by activating AMPK [65]. IL-15 aids in enhancing insulin action and lowering visceral adipose tissue [66]. TNF-α plays a significant role in this situation because of the buildup of fat in adipose tissue due to its production and release during inflammation, which promotes insulin resistance and increases lipolysis [67,68]. To further reduce insulin sensitivity, TNF-α inhibits IRS1 and downregulates PPAR-c in adipose tissues [69,70]. The cytokines generated by NF-kB activation can stimulate JNK, which causes insulin resistance and self-activates NF-kB in a feedback loop [37]. The macrophage initiates pro-inflammatory pathways and releases TNF, IL-1b, and IL-6 [71,72,73,74,75]. The recruitment of macrophages to tissues is mediated by elevated levels of chemoattractant protein-1 (MCP1), which is part of the inflammatory response [76]. The production of monocyte chemoattractant protein-1 (MCP1) by pancreatic islets is associated with pathophysiological conditions of pancreatic dysfunction [77]. Additionally, the inflammatory response is triggered by prostaglandins and leukotrienes, which are produced from arachidonic acid. Many factors contribute to inflammation, including pro-inflammatory cytokines, ROS, and environmental factors that release eicosanoids [78,79].

2.1.2. Insulin and β-Cell Involvement in Diabetes

β-Cells are stimulated to produce and secrete insulin when the plasma glucose levels are physiological, which helps the liver, brain, muscles, and adipocyte tissue absorb glucose. Insulin prevents the breakdown of fat and promotes the synthesis of proteins, lipogenesis, and glycogen while inhibiting hepatic gluconeogenesis [80]. This proves that insulin has generalized hormonal effects in addition to its well-known ability to lower blood sugar, which explains why diabetes affects various tissues. The hormone’s binding to the insulin receptor initiates a sequence of phosphorylation events that make up the insulin signal transduction pathway. Thus, the activation of intracellular protein substrates starts signaling cascades. Afterward, phosphatidylinositol 3-kinase (PI 3-kinase) activates protein kinase B (PKB), also known as AKT. GLUT4 is then translocated to the plasma membrane, except hepatocytes, which primarily express the non-insulin-regulated glucose transporter 2 (GLUT2), where it is activated by insulin in target cells along with several other enzymes, including glycogen synthase. The mitogen-activated protein kinase pathway is also responsive to insulin signaling, which controls gene expression, protein translocation, and cell growth [81]. Because insulin is a central regulator of lipid, protein, and carbohydrate metabolism regulator, an imbalance in metabolic paths directly affects how insulin behaves. The liver’s abilities to induce glucose uptake and glycolysis, which produce the building blocks for fatty acid synthesis, are just two of the numerous mechanisms contributing to lipogenesis [82]. Production of the pancreatic enzyme is dysregulated in T2D because of the close functional connections between the endocrine and exocrine pancreas [83]. Insulin resistance develops before insulin hypersecretion, which is viewed as a step to meet high insulin requirements [84]. In this respect, insulin resistance would result in hyperinsulinism. Whatever the underlying cause of hyperinsulinemia, the result is a reduction in glucose uptake by the muscles and an increase in the production of liver glycogen, which aids in the progress of T2D [36], [85], [86,87,88,89]. Furthermore, high glucose levels can cause β-cells to express the proapoptotic receptor FAS, which can produce IL-1b [90]. Insulin and glucagon functions associated with diabetes are shown in Figure 2.
By phosphorylating FOXO1 and SREBP1, AKT2 mediates the transcriptional activation of lipogenic genes induced by insulin [91]. Nucleotides are cofactors in crucial metabolic processes in addition to carbohydrates [92], and they may be connected to metabolic diseases [93]. For example, glyoxylic acid, trimethylamine, and uridine are all upregulated in T2D [94,95,96]. Interestingly, IMP, GMP, AMP, GTP, inosine, guanosine, and adenosine levels were elevated in T2D [97,98].

2.1.3. Free Fatty Acids and Type II Diabetes

Fatty acids have been linked to the risk of T2D [87]. In the blood, with increased insulin levels and insulin resistance in the liver and tissues, free fatty acids (FFAs) contribute to fat buildup, oxidative stress, inflammation, and hyperglycemia [85,86,99]. Furthermore, increased levels of FFAs prevent the lipolysis of adipose tissue induced by insulin [85]. Abnormal de novo lipogenesis and increased FFA levels are the root causes of several metabolic diseases [85,100,101,102,103]. As T2D progresses, one metabolic change occurs, which is an increase in FAAs. This change may open additional pathways that could help the disease progress. For instance, the lipid mediator palmitic acid has toxic effects in the islets, which activate the toll-like receptor to cause decreased insulin secretion and target organs’ insulin resistance [104,105]. In the liver and white adipose tissue (WAT), saturated fatty acids also cause the pro-inflammatory response via TLR4 [105,106,107]; NF-kB activation results in inflammation [108] and endoplasmic reticulum (ER) stress in immune cells and metabolic organs, which leads to insulin resistance [109,110]. Furthermore, there is a strong correlation between impaired insulin secretion and fatty acids. It has long been thought to be an aspect of the progress of type II diabetes, even though the molecular mechanisms relating to insulin resistance and fatty acids are still unknown [31]. FFAs modify islets in various techniques and accelerate the onset of T2D [36,111].
Phospholipids and triglycerides (TGs) are hydrolyzed to produce FFA and mono- and diacylglycerols (DAG), and TGs are inhaled as free fatty acids. Short- and medium-chain FFAs can be seen in the intestines, are carried to the bloodstream by serum albumin, and are stored in the liver and adipose tissues [112]. Moreover, lipogenesis is an additional source of FFAs [113]. FFAs’ high levels activate numerous pathways that may work together to affect the consequences of T2D and insulin resistance. Elevated palmitate levels induce a pro-inflammatory response by promoting IL-1 and IL-I8 secretion and maturation [114]. The serine phosphorylation of the insulin receptor substrate-1 (IRS1) in an NK/IKK-dependent fashion results in insulin resistance induced by pro-inflammatory cytokines [115]. Furthermore, high FAA concentration stresses cells because lipotoxicity causes apoptosis, ROS production, and ER stress [116]. However, sustained high-level exposure to FFAs causes lipotoxicity, which causes β-cell dysfunction and, ultimately, type II diabetes (T2D) [117]. Similarly, continual FFAs are due to the reserve of glucose-stimulated insulin (GSIS) release, changes in gene appearance, and promotion of apoptosis caused by stimulation of inaccessible pancreatic islets with stimulatory glucose concentrations [118]. ER stress, which can lead to β-cells apoptosis, can be brought on by saturated fatty acids. In β-cells, the ER stress and unfolded protein response are incredibly sensitive [119]. T2D is consequently developed in pancreatic islets exposed to FFAs over an extended period. Adipocytes can store adipose tissue more effectively when high FFA concentrations are present, but an increase in adipocyte fat content may cause inflammation and hypoxia in the tissue and cell [120].
Adipocytes develop insulin resistance and chronic low-grade inflammation, which help in the pathogenesis of T2D [116,121]. According to the most widely accepted theory, β-cells secrete too much hormone to counteract insulin resistance [117]. In this case, myocytes frequently take in more FFAs and store them as TGs because T2D increases the flux of TGs and FFAs. For metabolic energy, skeletal muscles primarily use glucose and FFAs [120,122]. FFA buildup in myocytes causes the synthesis of toxic ceramides and DAG, which can cause cell damage, lipotoxicity, inflammation, and insulin resistance [120].
Inflammation and metabolic disorders are frequently associated with metabolic dysregulation in the liver and muscles [123,124] because the accumulation of DAG promotes PKC activation while inhibiting insulin receptor activation, resulting in muscle and liver insulin resistance [125,126,127,128]. Insulin resistance positively correlates with irregular lipid buildup in the muscle and liver [129]. In this regard, elevated plasma FFA levels cause fat to build up in the WAT, liver, and muscle by regulating long-chain acyl-CoA, TGs, and DAG [128]. However, insulin resistance appears to lead to augmented lipid accumulation in these tissues [130]. Activating PKC isoforms, DAG, a precursor to TGs, regulates the phosphorylation of molecules in the insulin pathway [131]. In the development of T2D, DAG buildup appears to be a significant lipid mediator, inhibiting insulin sensitivity in the liver and muscle [130,132,133]. Furthermore, the activation of phosphatase 2A, which dephosphorylates AKT, reduces the translocation of the PIP3–PDK1 complex and inhibits insulin-stimulated AKT at the plasma membrane of target cells [134,135,136,137,138]. In addition to these mechanisms, ceramide buildup in membrane domains activates caspase, releasing pro-inflammatory cytokines, generating ROS, and leading to cell death [139].
There has been evidence linking higher levels of FFAs in people with high plasma-free radical levels to the production of ROS by NADPH oxidase in adipocytes, which led to the release of pro-inflammatory cytokines from WAT [140,141]. ROS are essential for inflammation and signaling [142]. Two tissues where pro-inflammatory cytokines may be produced and released are adipose tissue and the liver. These cytokines may affect other tissues due to blood circulation, resulting in tissue damage, cell death, and an intensified pro-inflammatory response [37]. IRS1 in the liver and adipose tissue is inhibited by lipid mediators, TNF-α, ROS, hypoxia-activated IKKb, and JNK [143,144,145]. IKK and JNK1 phosphorylate IRS1 and IRS2 on the serine residue, which causes activation of the gene linked to insulin resistance and inflammation [146,147]. On the other hand, pro-inflammatory cytokines such as IL1b, MCP1, TNF-a, and IL-6 can be produced and released when the NF-kB pathway is stimulated by high FFA concentrations [147]. Free fatty acids can bring on insulin resistance in several different ways; increased lipid metabolism caused by FFAs is linked to insulin resistance [148,149] because it inhibits the insulin receptor [150,151]. Additionally, high FFA levels cause ER stress in β-cells and the liver [152,153], as well as in adipocytes [154,155], which activates JNK and results in insulin resistance [155]. In T2D and obesity, FFAs are also necessary for the activation of the NLRP3 and the production of IL-1b [120]. IL-1 and IL-18 are released by the NLRP3 inflammasome, which promotes inflammation [154,155,156].

2.2. Polyphenols

A growing body of evidence from in vivo and in vitro studies points to a substantial role for dietetic polyphenols in treating type II diabetes (T2D) through insulin-dependent tactics, such as protecting pancreatic islet cells, reducing cell apoptosis, promoting islet cell proliferation, attenuating oxidative stress, activating insulin signaling, and stimulating insulin secretion [33]. This can also be achieved through insulin-independent approaches including the modification of the inflammatory response, inhibition of digestive enzymes, regulation of intestinal microbiota, and prevention of advanced glycation end products from forming [120]. Plant-based foods are increasingly used in dietary guidelines for people at the hazard of T2D. These may affect glucose breakdown through several mechanisms, including carbohydrate digestion inhibition and intestinal glucose absorption, stimulation of pancreatic β-cells insulin secretion, glucose release from the liver, initiation of insulin receptors and glucose acceptance in the insulin-sensitive tissues, and modification of hepatic glucose output [2,5]. Below are details about the role of documented polyphenols in T2D. The chemical structures of some important polyphenols are shown in Figure 3.

2.2.1. Resveratrol

Baur and coworkers reported that resveratrol increases the lifespan in high-caloric diet mice by reducing glucose and improving insulin levels. It increased insulin sensitivity in diabetic mice and homeostatic model assessment during glucose tolerance tests [157]. Research findings showed that resveratrol lowers blood insulin levels in animals with hyperinsulinemia and insulin resistance. Rodents with diet-induced hyperinsulinemia were used to demonstrate this effect [51,52,53,54,86]. On the other hand, resveratrol seems to raise blood insulin levels in rodent models of type II diabetes with reduced-cell mass and hypoinsulinemia, as demonstrated in db/db mice [57,61]. The improvement in insulin action lowers blood glucose levels, which prevents glucotoxicity, the harmful effects of hyperglycemia on β-cells [120]. In addition, resveratrol alleviates steatosis and lowers hepatic lipid buildup. Decreased expression of acetyl-CoA carboxylase (ACC) and fatty acid synthase (FAS) is linked to these effects [37,53,54,61,92,93,94,95]. It also reduces the expression of fatty acid synthase [156]. According to some published research, resveratrol’s effects on FAS and ACC may be mediated by the AMPK/SIRT1 axis [97,98]. It also decreases plasma amylase levels, which increases pancreatic damage. Thus, it prevents pancreatic damage.
In addition, resveratrol increases mitochondrial numbers and citrate synthase activity [158] with reduced caloric and exercise [158,159]. Furthermore, in liver tissue, resveratrol decreases the appearance of pro-inflammatory cytokines [83,94] and increases glutathione peroxidase activity, which decreases oxidative liver damage [96]. Furthermore, resveratrol decreases inflammatory markers, which protect pancreatic β-cells [103]. Findings also demonstrated that resveratrol lessens oxidative stress; reduces islet fibrosis and destruction; restores islet architecture; enhances islet structure and function; and attenuates other worsening changes in db/db mice, a type II diabetes animal model with diminished β-cell mass. Moreover, resveratrol increases the β-cell mass and partially stops β-cell failure [57,61]. Parametric analysis of gene set enrichment (PAGE) showed that resveratrol alters glycolysis, TCA cycle, classic and alternative complement pathways, butanoate, propanoate metabolism, and sterol biosynthesis [157]. In insulin-resistant rodents, resveratrol promotes intracellular glucose transport in rats fed a high-cholesterol and high-fructose diet and given resveratrol larger than those animals not given this supplement [160]. Resveratrol enhances skeletal muscle’s ability to absorb insulin-stimulated glucose [161,162].

Resveratrol Effect on Diabetes via GLUT4 Elevation

In insulin-resistant rodents, intracellular glucose transport increases by resveratrol. Within this context, Deng and colleagues indicated that when rats fed on a high-fructose and high-cholesterol diet are given resveratrol in the initial animal studies, they show greater soleus muscle glucose uptake than animals not given this supplement [160]. Similar results were obtained and showed that resveratrol increases skeletal muscle glucose uptake in rats nourished on a high-fat diet [161,163]. Resveratrol increases intracellular glucose transportation in insulin-resistant animals via two GLUT4-related mechanisms. It is well recognized that resveratrol expedites the translocation of GLUT4 to the muscle cells’ plasma membranes [160,161], and GLUT4 expression is also increased in animals with insulin resistance in their skeletal muscle [164] and in db/db mice [165]. Moreover, research findings showed improved insulin action by increased intracellular glucose transportation in resveratrol-consuming insulin-resistant animals. In skeletal muscle, resveratrol reduces insulin resistance through various mechanisms, including alterations in metabolism and lipid buildup. In addition, resveratrol encourages mitochondrial biogenesis in rats with diet-induced insulin resistance in their skeletal muscles [166] and improves mitochondrial β-oxidation [162]. Coen and Goodpaster reported that type II diabetes and insulin resistance are exacerbated by increased intramyocellular lipid accumulation, affecting how well insulin works [167].

Resveratrol Effect on Diabetes via SIRT1 Involvement

Kitada et al. [168] reported that variations in the expression and activities of two intracellular controllers are closely related to the beneficial effects of resveratrol on the muscle tissue of insulin-resistant rodents, i.e., SIRT1 and AMPK. The NAD+-dependent histone deacetylase SIRT1 (silent information regulator 1) involves several processes, including inflammation, mitochondrial biogenesis, stress resistance, intracellular metabolism, glucose homeostasis, apoptosis, and others. Since type II diabetic patients have decreased SIRT1 activity and expression, SIRT1 is considered a target for anti-diabetic medications [168,169]. In addition, scientists showed that resveratrol triggers SIRT1 in mammalian tissues [170] and triggers muscle SIRT1 in animals with diet-induced insulin resistance [162]. An increase in the NAD+/NADH ratio is related to this enzyme’s activation [166]. Findings also revealed that resveratrol raises the SIRT1 level in the muscle in rodents with genetically stimulated insulin resistance [56]. Deacetylation and activation of PGC-1α are linked to resveratrol-induced upregulation of AMPK in skeletal muscle, possibly via SIRT1-dependent mechanisms [164,168].

Resveratrol Effect on Diabetes via AMPK Activation

Another enzyme involved in the action of resveratrol, besides SIRT1, is AMP-activated protein kinase (AMPK). AMPK controls various physiological functions, such as mitochondrial function, energy metabolism, insulin secretion, and biogenesis [171]. In this regard, McCart reported that AMPK promotes insulin sensitivity and fatty acid oxidation [172]. Furthermore, resveratrol activates AMPK by phosphorylation and acetyl-coA carboxylase [158]. Insulin resistance induced by the diet in animal models is preceded by decreased AMPK activity [82], and insulin resistance is genetically determined [54]. The insulin-sensitizing medicines thiazolidinediones and metformin usually stimulate AMPK in various tissues, even though a direct connection between AMPK initiation and the reduction of insulin resistance in humans has not been established [171]. Resveratrol activates AMPK to these drugs in insulin-resistant animals. Resveratrol also reverses diet-induced insulin resistance in rodents by restoring AMPK phosphorylation [51] and makes AMPK active in skeletal muscle [165].

Resveratrol Effect on Diabetes Involving Mitochondria

Resveratrol reduced the acetylation status of PGC-1α [157], a transcriptional co-activator that regulates the mitochondrial biogenesis mediated by SIRT1 deacetylation [173,174]. In addition, it is believed that in humans, mitochondrial muscle dysfunction speeds up intramuscular lipid deposition and reduces insulin action [64]. Therefore, resveratrol action in muscle tissues appears to depend on the rise in mitochondrial biogenesis caused by a concurrent reduction in intramuscular lipid level [168,169].

Resveratrol Effect on Diabetes via FFA Reduction

Increased release of free fatty acids is identified as a significant factor in the emergence of insulin resistance [100,101] in rodents [50,54,83] with diet-induced insulin resistance. In this respect, resveratrol has been shown to lower pancreatic triglyceride levels in animals fed with high-fat diets [52]. The anti-obesity properties of resveratrol may be connected to its anti-diabetic properties [13,14], with decreased action of lipogenic enzymes (acetyl-CoA carboxylase, glucose-6-P-dehydrogenase, and lipoprotein liPase) [92]. It is well known that having more body fat reduces the effectiveness of insulin and increases the risk of developing type II diabetes in humans [2,44]. Without causing appreciable changes in adiposity, resveratrol may enhance insulin action [55] or decrease body weight [56,83]. By increasing insulin receptor phosphorylation, resveratrol may also enhance insulin signaling in animals with insulin resistance in their skeletal muscles [39] and increased protein levels of IRS-1 [56]. Table 1 shows the antidiabetic activity of resveratrol from molecular mechanisms to in vivo studies.

2.2.2. Curcumin

Curcumin (Figure 3) exhibits anti-inflammatory properties that may aid in controlling diabetes. Curcumin analogs have been identified and are currently the subject of extensive research for their potential roles in diabetes. In this regard, numerous studies on the effectiveness of curcumin in regulating blood glucose in various rodent models have been published. According to Arun and Nalini, curcumin lowers blood sugar, hemoglobin (Hb), and glycosylated hemoglobin levels (HbA1C) [188] and recovers insulin sensitivity [189]. Similarly, Abu-Taweel and coworkers reported that curcumin improves diabetes pathology through various mechanisms, including the control of lipid metabolism; antioxidant activity; and other activities such as antiapoptotic, anti-inflammatory, and antihyperglycemic activities [190]. Research findings indicated that curcumin extract reduces insulin resistance, prevents cell death, delays the onset of diabetes, and enhances cell functions in animal models [191]. Similar results were obtained when 250 mg curcuminoids were used for nine months in pre-diabetic patients not diagnosed with diabetes. Furthermore, Chuengsamarn et al. [33] reported that curcumin improves the overall performance of β-cells with higher homeostasis model assessment (HOMA-β) and lower C reactive protein (CRP). Those who received curcumin experienced higher levels of adiponectin and lower levels of insulin resistance. In the meantime, Wickenberg reported that postprandial serum insulin concentrations increased by 6 g turmeric ingestion without having an appreciable impact on plasma glucose levels [192]. A paper by Gutierres and colleagues showed that giving curcumin for 31 days to STZ-induced diabetic rats reduced the hyperlipidemic and hyperglycemic effects [193]. On the other hand, a different study found curcumin (90 mg/kg BW) with insulin (1 U/day vs. 4 U/day) in STZ-induced rats decreased hyperglycemia, hypercholesterolemia, and biochemical markers of kidney and liver damage while increasing the activity of glutathione peroxidase and superoxide dismutase (hepatic antioxidants) [194].
In addition, curcumin has excellent wound-healing qualities due to its capacity to reduce oxidative stress by removing free radicals [195]; many people with diabetes experience difficulties with wound healing [196]. In this context, Yang and coworkers showed that curcumin can prevent retinal attenuation by enhancing the retina’s ultrastructure [197]. By promoting the superoxide dismutase enzyme’s expression, curcumin can reduce oxidative stress [198] and the reduction of ROS production, both of which are crucial for treating diseases such as diabetes caused by oxidative stress and inflammation [199]. Oxidative stress is thought to make diabetes worse, whereas ROS have been proposed to be crucial in diabetes pathogenesis. Curcumin’s chemical makeup and anti-oxidative strength allow it to function naturally as a free radical scavenger. Fasting blood glucose (FBG), hemoglobin A1c (HbA1C), estimated average glucose (EAG), and body mass index (BMI) levels were all improved by curcumin in diabetic patients [200]. In this respect, Panahi et al. reported that curcuminoid supplementation has an antioxidant effect in T2DM patients because it reduced malondialdehyde (MDA) and raised serum SOD activity and total antioxidant capacity [201]. Similarly, Jain reported that curcumin diet supplements (50 or 100 mg/kg BW) decrease hyperglycemia and inflammatory processes in STZ-induced diabetic rats by preventing McP-1, HbA1c, TNf-α, IL-6, and lipid peroxidation and suppressing the NF-kB signaling pathway; protecting against inflammation [202]; and restoring normal antioxidant enzymes levels, including catalase, glutathione peroxidase, and SOD [203].
He et al. [204] also reported that curcumin prevents the NF-kB signaling cascade and inflammation. Reduced levels of IL-6 and TNF-a were assessed in STZ-induced diabetic rats with heart damage in a study by Abo-Salem et al. [205]. On the other hand, Arafa showed that curcumin could increase insulin sensitivity by decreasing cholesterol and blood glucose levels [206]. A high curcumin supplement (100 mg/kg) improved insulin intolerance and glucose in gestational diabetes mice by triggering the AMPK pathway [207]. Findings also showed that curcumin treatment significantly decreased superoxide production and NADPH oxidase subunit expression (p67phox, p22phox, and gp91phox) in diabetic rats. This effect may have been caused by curcumin inhibiting the protein kinase C (PKC)-MAPK signaling pathway [208]. Oxidative stress and endoplasmic reticulum (ER) were protected from diabetes by the novel curcumin analog C66, which inhibited JNK activation in diabetes [209]. Additionally, results showed that curcumin significantly increased mitochondrial permeability and decreased palmitate-induced oxidative stress. It did this by causing pancreatic β-cells to secrete more insulin when glucose was present [210]. Pathological complications of diabetes include diabetic nephropathy, diabetic neuropathy, vessel damage, and cardiovascular diseases [211]. In contrast, Panahi et al. [212] reported that taking curcumin (1 g daily) for three months reduces leptin levels and the leptin/adiponectin ratio (an indicator of atherosclerosis) in patients with atherosclerosis; it also increased adiponectin. Table 2 shows data related to the antidiabetic activity of curcumin.

2.2.3. Quercetin

Quercetin (Figure 3) has been proven useful in treating T2D [230]. Research by Pereira and coworkers showed that quercetin interacts with molecular marks in the adipose tissue, liver, skeletal muscle, pancreas, and small intestine to maintain glucose homeostasis [231]. Other studies reported that quercetin treats T2D by reducing hyperglycemia, enzyme levels, liver glucose content, high blood pressure, serum cholesterol levels, and hyperlipidemia, as well as by encouraging weight loss [230,232], lowering blood sugar levels [233,234,235], improving glucose tolerance [233,236] and hepatic glucokinase activity [236], and enhancing the subsequent release of insulin and pancreatic cell regeneration [237,238]. In this respect, research findings revealed that quercetin activates AMPK, which inhibits glycogenic isoenzymes such as phosphoenolpyruvate carboxylase (PEPCK) and glucose-6-phosphatase (G6Pase) to reduce glucose synthesis [235,239] and stimulate protein kinase B (Akt) and skeletal muscle GLUT4 receptors, which in turn activates AMPK in the cell membrane [240]. Pereira confirmed that the GLUT4 transporter controls blood sugar levels by controlling glucose entrance into the cells [231]. In another study, Borghi indicated that by encouraging the GLUT4 translocation to the cell membrane, quercetin administration, GLUT2 expression, and intestinal-sodium-dependent glucose uptake are reduced, thus lowering gastrointestinal absorption of glucose and controlling blood sugar levels [241].
Similarly, Spínola et al. showed that the inhibition of pancreatic-amylase and intestinal-glucosidase decreases starch hydrolysis, slows postprandial hyperglycemia progression, and diminishes the rate of glucose absorption by quercetin usage [242,243]. Another study reported that quercetin improves dyslipidemia caused by a high-fat diet (HFD) in Swiss albino mice [244]. By controlling the levels of c-peptide and HbA1c, quercetin reduced the harm to pancreatic β-cells [245] and decreased lipid levels and insulin resistance [246], thus increasing pancreatic β-cell functions and exerting anti-hyperglycemic activity in diabetic rats [247]. In this respect, 20 µM of quercetin induced a significant increase in insulin secretion by increasing intracellular calcium ions through interaction with L-type Ca2+ ion channels in INS-1 β-cells [248], as well as simultaneous transient inhibition of KATP channels [249]. According to these results, quercetin controls glucose metabolism by enhancing glycolysis and reducing gluconeogenesis [250]. Moreover, published research showed that fat accumulation, reduced body weight, dyslipidemia, hyperglycemia, and hyperinsulinemia were significantly improved by quercetin treatment due to improved gene-associated glucose or lipid metabolism in high-fat-fed obese mice [246,251]. In addition to lowering blood sugar and HbA1c levels, Wang et al. found that oral administration of quercetin in multiple doses improved glycogen synthesis, decreased insulin resistance, and lowered glucosidase activity. Furthermore, it decreased oxidative stress, which enhanced pancreatic insulin secretion and helped diabetic patients control their blood glucose levels [209]. In addition, quercetin helps in alleviating diabetic complications by blocking AR [252].
The protein expression of insulin-signaling molecules such as phosphatidylinositol 3-kinases (PI3K) and insulin receptor substrate-1 (IRS-1) can be increased by quercetin, according to studies on STZ-induced diabetic rats; this results in an increase in insulin-mediated glucose uptake [231]. A survey by Ashraf and colleagues showed that quercetin lowers oxidative stress by scavenging ROS and improving the AMP/ATP ratio in clonal pancreatic cells [253]. On the other hand, obesity-related T2DM is associated with fat buildup in the muscles and liver, which triggers the nuclear transcription factor NF-B (NF-B) and Jun N-terminal kinase (JNK) inflammatory pathways [254]; both of these pathways are suppressed by quercetin [255]. In addition, brown adipose tissue releases pro-inflammatory mediators such as IL-8, IL-4, IL-1, IL-6, TNF-α, and histamine in response to high blood glucose levels and improved insulin resistance [256]. These mediators are inhibited by quercetin, which also reduces oxidative stress [257]. Blocking the enzymes lipoxygenase and cyclooxygenase prevents the release of pro-inflammatory mediators such as prostaglandins and leukotrienes [258]. Yao et al. reported in a clinical survey conducted among the Chinese population an inverse relationship between quercetin consumption and the prevalence of T2D [259]. Table 3 lists the antidiabetic activity of quercetin and its mechanisms of action.

2.2.4. Catechins

Kim and colleagues reported that catechins stimulate either GLUT4 transcription or translocation to the plasma membrane in muscle cells and glucose uptake in peripheral tissues. Furthermore, catechins inhibit lipogenesis, glycogen synthesis, and glucose oxidation in liver cells [290]. Similar results were reported by several studies [291,292,293,294,295]. Catechins can also impair glucose transporters on the plasma membrane of intestinal cells, Similarly, epicatechin gallate inhibits the Na+-dependent glucose transporter in rabbit intestinal brush-border membrane vesicles (SGLT1), demonstrating that epicatechin gallate inhibits SGLT1 [296,297]. Moreover, researchers showed that catechins prevent weight gain and the start of chronic illnesses such as T2D or metabolic syndrome when consumed regularly [298,299]. Similarly, other researchers indicated that epigallocatechin gallate inhibits pancreatic glucosidase in a noncompetitive manner that is reversible [300,301,302]. Moreover, galloylated catechins are more potent than nongalloylated catechins at inhibiting glucosidase and amylase. Depending on their chemical composition, catechins have varying levels of inhibitory power [303].

2.2.5. Isoflavones

Findings showed that the consumption of isoflavone decreased the risk of diabetes [304] via glucose uptake inhibition and negligible intestinal carbohydrate absorption [305]. In addition, isoflavones enhance insulin sensitivity and resistance, safeguarding pancreatic β-cells, acting as an anti-inflammatory agent, reducing oxidative stress, and preventing the formation of the Maillard reaction and advanced glycation end products [306]. In this context, Rockwood et al. reported that genistein significantly lowers hyperglycemia in T2D [307,308], increases cell proliferation while decreasing apoptosis [309], and reduces oxidative stress and cardiac inflammation [310]. In contrast, daidzein’s preventive effect on reducing hyperglycemia, dyslipidemia, obesity, insulin resistance, inflammation, and other T2D complications has been thoroughly studied. It causes an immunomodulatory effect in mice with diabetes [311,312]. To incorporate several methods to increase flavonoids’ antidiabetic activity, numerous strategies have been developed in recent years to use flavonoids in vitro and in vivo models.

2.2.6. Hydroxycinnamic Acids

Ferulic Acid

Published research revealed that ferulic acid (FA) lowers hyperglycemia, the lipid profile, creatinine, urea, serum glutamic oxaloacetate transaminases, and serum glutamic pyruvic transaminases while maintaining islet mass in STZ-induced diabetic rats over the course of three weeks [313]. At doses of 0.01 and 0.1% of the standard diet, FA lowered blood glucose levels in STZ-induced diabetic mice. In KK-Ay mice, 0.05% FA significantly lowered blood glucose levels [314]. Similarly, oral administration of FA (10 and 50 mg/kg BW) into STZ-induced diabetic rats demonstrated antioxidant activity; it decreased the levels of lipid peroxidation indicators in the serum, liver, pancreas, and kidney [315]. In this respect, several food items such as tomatoes, berries (such as strawberries), rice husks, and other fruits and vegetables commonly contain FA [316,317]. By increasing plasma insulin levels, glucokinase activity, and liver glycogen synthesis in diabetic rats, FA and sinapic acid effectively decreased blood glucose levels [318,319].

Gallic Acid

Gandhi et al. reported that gallic acid (GA) exhibits antidiabetic properties in animal models lacking insulin or are resistant to insulin [320] by significantly reducing blood sugar, triglyceride, total cholesterol, urea, uric acid, low-density lipoprotein cholesterol, and creatinine while simultaneously raising plasma levels of insulin (16.3 U/mL), C-peptide, and glucose tolerance [321]. Other researchers showed that GA reduces gluconeogenesis and increases glycolysis, ultimately decreasing hyperglycemia in STZ-induced diabetic rats [322]. Fruits such as grapes and berries contain GA [323,324]; in this regard, researchers found that apple juice and berries might help improve short-term glycemic control [9].

Protocatechuic Acid

Protocatechuic acid (PCA) showed reduced levels of hepatic gluconeogenic enzymes such as fructose-1,6-bisphosphatase, glucose 6-phosphatase (G6Pase), and sorbitol dehydrogenase, as well as increased levels of glucose-6-phosphate dehydrogenase and hexokinase in STZ-induced diabetic rats [325]. These results show that PCA can enhance GLUT4 translocation, adiponectin secretion, and glucose uptake [326]; prodigious amounts of PCA are found in gooseberry, raspberry, blueberry, mulberry, honey, soybeans, and loquat fruit [325].

Ellagic Acid

Ellagic acid (EA) might be a useful dietary supplement to lessen the metabolic changes associated with HFD feeding animals in combination with STZ injection [327]. EA reduces glycation stress, hyperglycemia, inflammation, and hyperinsulinemia and aggravates renal function dose-dependently. In this respect, research findings showed that EA (3.12–50 M) increases the expression of PPAR in L6 myotubes and GLUT4 [328].

Salicylic Acid

Blackberries, cantaloupes, blueberries, dates, grapes, apricots, kiwis, olives, green peppers, radishes, tomatoes, and mushrooms are among the foods that contain salicylic acid in high concentrations. This acid lowers blood concentrations in diabetic Goto-Kakizaki rats [329].

Caffeic Acid

Numerous fruits and vegetables, including blueberries, kiwis, cherries, plums, apples, pears, potatoes, artichokes, cider, and coffee, contain caffeic acid (CA), a phenolic acid [7]. Researchers reported that dietary supplements with CA (0.02% in the diet for five weeks) decrease blood glucose, G6Pase, and phosphoenolpyruvate carboxy kinase activities, accompanied by a decrease in the liver GLUT2 expression and enhanced insulin levels, glucokinase, catalase, glutathione peroxidase, and SOD activities in db/db mice [330]. Additionally, CA significantly lowered the levels of plasma HbA1c [331]. In insulin-resistant rats undergoing a glucose test, administration of CA reduced the elevation of plasma glucose levels. CA also increases the isolated adipocytes’ ability to absorb glucose. Moreover, the reduction in plasma glucose appears to be caused by CA’s increased glucose utilization [332].

p-Coumaric Acid

Another phytochemical, p-coumaric acid, is prevalent in fruits and vegetables, including apples, pears, beans, potatoes, tomatoes, tea, and pineapple [333,334,335]. By changing glucose and lipids’ metabolism, p-coumaric acid can potentially prevent or treat insulin resistance and T2D [336].

Chlorogenic Acid

Chlorogenic acid (CGA) increases GLUT in skeletal muscle by phosphorylating AKP-activated protein kinase, which enhances the metabolism of lipids and glucose, thus reducing the hazard of diabetes [337]. Evidence suggests that CGA reduces intestinal-sodium-gradient-driven glucose transport and inhibits G6Pase. It increased AMPK phosphorylation and favorable metabolic changes linked to AMPK activation while improving skeletal muscle glucose uptake and lipid profiles [338]. In addition, Bassoli and coworkers reported that inhibiting G6Pase activity prevents the production of hepatic gluconeogenesis [339]. Moreover, it reduced hepatic steatosis and inhibited the expression and activity of G6Pase in the liver [340]. Cherries, apples, kiwis, artichokes, eggplants, plums, and coffee are just a few of the foods that contain CGA, one of the most prevalent phenolic compounds [7]. CGA reduces the effects of retinopathy and other diabetic complications in animals by preventing retinal neo-angiogenesis [341]. Furthermore, enzymes that break down carbohydrates are weakly inhibited by chlorogenic acid [342]. Research findings indicated that CGA inhibits glucosidase activity [343].

trans-Cinnamic Acid

trans-Cinnamic acid (t-CA) is found in numerous food-related plants, fruits, and herbs [344]. Through the involvement of GLUT4, t-CA (1 ng/mL) isolated from Cinnamomum cassia activates insulin-mediated glucose transport [345]. In isolated islets, it significantly increased glucose-enhanced insulin secretion [346]. Daily oral administration of t-CA (80 mg/kg BW) for four weeks decreased hyperglycemia in male albino rats with diabetes induced by alloxan [347]. These results demonstrate that treatment with t-CA (80 M) increases AMPK activation and adiponectin secretion. Additionally, the inhibitory effect of paclitaxel suggests that t-CA-stimulated signaling in 3T3-L1 adipocytes involves a G-protein-coupled receptor and enhances insulin sensitivity [348].

2.2.7. Anthocyanins/Anthocyanidins

Zhou and coworkers reported that anthocyanidins (ACNs) promote health through their antioxidant, anti-inflammatory, and blood-sugar-regulating properties [235]. In this regard, AMPK/ACC/mTOR pathway helps anthocyanin-rich mulberry extract prevent hyperglycemia [349]. Other researchers showed that by managing blood lipid and triglyceride levels, lowering cholesterol, and having low-density cholesterol while raising high-density cholesterol and apolipoprotein, ACNs might reduce insulin resistance [350]. Moreover, anthocyanins stimulated the release of insulin by increasing the appearance of the intracellular Ca2+ signaling pathway and the glucose-transport-related gene (Glut2) in mouse islet β-cells. Along this line, purple potato extract with added cyanidin increased insulin secretion [351]. Delphinidin 3-arabinoside anthocyanidins, found in fermented berry beverages, controlled DPPIV and its substrate GLP-1, boosted insulin secretion, and increased the mRNA expression of genes related to insulin receptors [352]. Published work by Graf et al. showed that ACN-rich grape-bilberry juice (AGBJ) supplementation improved several risk factors for diseases linked to obesity in male Fischer rats for ten weeks. Results revealed that AGBJ intervention successfully reduced serum levels of triglycerides and leptin while having no impact on the release of adipokines, adiponectin, glucose, insulin, or non-esterified fatty acids. In addition, AGBJ increased plasma levels of polyunsaturated fatty acids while lowering levels of saturated fatty acids. Overall, the findings suggested that AGBJ might effectively combat metabolic diseases linked to obesity [353]. In STZ-induced T2DM rats, ACNs from purple root vegetables reduced liver damage and oxidative stress and enhanced lipid and blood glucose levels [354].
ACNs act as anti-inflammatory agents by suppressing the expressions of a few inflammatory cytokines crucial to the inflammatory response, including TNF-, IL-6, and IL-1 [355,356,357,358]. Monocyte chemoattractant protein 1 (MCP-1), a chemokine, plays a role in developing diabetes mellitus by controlling leukocyte migration and infiltration [359]. Numerous studies demonstrated that ACNs can lower MCP-1 expression [358,360]. In addition, research findings showed that ACNs could be a potent therapeutic agent to prevent obesity and diabetes because of the changes in AMP-activated protein kinase activation. ACNs decreased the AMP/ATP ratio, which strongly correlated with ACN supplementation. [361]. AMP-activated protein kinase (AMPK) is a critical molecule in the control of glucose metabolism in the liver, white adipose tissue, and skeletal muscle, which is activated by ACNs [354,362,363,364,365]. Activation of AMPK induces GLUT4, thus improving glucose utilization and uptake [365,366]. Moreover, the production of the liver’s glucose is decreased when AMPK is activated [367]. Findings confirmed that ACNs could help with obesity, as well as impaired glucose tolerance, insulin resistance, and DM prevention. Cyanidin-3-glucoside (C3G) improved glucose tolerance (GT) and reduced body weight gain in mice fed with a high-fat diet [368]. In this regard, numerous studies demonstrated that ACN-rich blueberries can decrease body weight, enhance lipid profiles, suppress the countenance of inflammatory factors, and increase insulin sensitivity in animal models fed with a high-fat diet [356,369,370,371]. Black elderberry [360], raspberry [372], Aronia melano-carpa [373,374], and black rice [375] are rich in ACN and could improve insulin resistance and lipid metabolism in the liver or serum in obese mice.
Takikawa et al. [362] reported that bilberry extract containing an increased ACN level significantly decreases blood glucose levels in T2DM mice and improves insulin sensitivity. Feeding T2DM mice a diet containing 0%, 5%, or 10% buckwheat sprouts revealed that as the number of buckwheat sprouts in the diet increases, lipids levels and blood glucose improve more noticeably [376]. Similarly, ACNs from the black soybean seed coat could also lessen the harm done to the liver, kidney, and pancreas in STZ-induced T2DM mice [377]. In a different experiment involving animals, giving blueberry ACN extract to T2DM mice improved glucose tolerance and blood glucose levels; reduced polydipsia and polyuria symptoms; and reduced TC, TG, and insulin levels [378]. Ye and colleagues reported that C3G intervention reduces blood sugar and insulin resistance and improves blood sugar and lipid parameters in db/db mice [379]. Furthermore, diabetic db/db mice supplemented with dietary C3G for 5 weeks showed reduced hepatic triglyceride content and steatosis and decreased inflammatory cytokine concentration in the serum [380].
On the other hand, malvidin and ACNs were used in combination with metformin in the treatment of STZ-induced diabetic rats, and the outcomes demonstrated that the combination therapy has more significant relief from insulin resistance, decreased fasting blood glucose, and improved lipid metabolism and serum insulin compared to single therapy [381]. After receiving combined treatment with fenofibrate and ACNs in T2DM patients with postprandial hyperlipidemia, the serum postprandial triglyceride level and LDL cholesterol concentration were pointedly reduced (from black soybeans) [382]. Several studies showed that ACNs can decrease the initiation of pro-inflammatory factors and improve insulin resistance [367,383]. ACNs prevent the stimulation of JNK and NF-B, which lowers the phosphorylation of IRS-1 serine residues and improves insulin resistance [367,371]. Additionally, it has been demonstrated that ACN can trigger the production of adiponectin, which can potentially reduce insulin resistance [358,384,385]. ACNs increase the efficiency of two enzymatic antioxidants called SOD and catalase (CAT), which shield cells from oxidative damage by catalyzing the conversion of free radicals into hydrogen peroxide [358,386]. Furthermore, the inflammatory response may accelerate the development of DM complications and contribute to insulin resistance, eventually resulting in T2D complications [387]. Cranberries, blackberries, chokeberries, black grapes, gooseberries, bilberries, red raspberries, blueberries, blackcurrants, and strawberries are rich sources of ACNs. Other sources include a variety of other fruits such as peaches, grapes, nectarines, pomegranates, plums, cherries, seeds, and vegetables, i.e., red onions and red lettuce [388]. Table 4 lists the anthocyanins’ role as potential antidiabetic agents along with their molecular mechanisms.

2.2.8. Kaempferol

Kaempferol exhibits anti-oxidative stress anti-hyperglycemic [394], anti-inflammatory [395], and hypolipidemic [396] effects. Inflammatory cytokines, including TNF-α and IL-6, stimulate the c-Jun amino-terminal kinase (JNK) and I-kB kinase-b/nuclear factor-kB (NF-kB) paths in insulin-sensitive organs and inhibit insulin signaling [397]. Similar to an insulin secretagogue, kaempferol enhances insulin secretion. Kaempferol increased plasma insulin levels while lowering the blood glucose level in STZ-induced diabetic rats [398]. Kaempferol directly activates mitochondrial calcium uptake (MCU) in a concentration-dependent manner. An amount of 1 µM can trigger the pancreatic β-cell secretion/metabolism/coupling and closely dual the uptake of mitochondrial Ca2+ [399,400]. With an increase in cAMP, Ca2+, and glutathione (GSH) levels, kaempferol raises glucagon-like peptide 1 (GLP-1) and insulin levels [401]. In this respect, Fang et al. showed that in 3T3-L1 adipocytes, kaempferol enhances insulin-dependent glucose uptake [402]. Kaempferol also lowers blood glucose levels by boosting GCK levels and enhancing glycogen synthesis [22].
An imbalance in the making and utilization of glucose leads to disorders of glucose metabolism. Hepatic IR plays a significant role in fasting hyperglycemia. In this regard, abnormal glucose-metabolism-regulating enzyme levels, such as phosphoenolpyruvate carboxykinase, PC, glucokinase (GCK), and glucose-6-phosphatase, are a hallmark of hepatic IR (PEPCK). Blood sugar levels directly affect how GCK is activated and inactivated. Activation of GCK is thus a probable target for diabetes treatment [403]. Kaempferol (50 mg/kg/day), administered orally to mice, significantly reduces hyperglycemia by reactivating hexokinase and inhibiting PC and gluconeogenesis [394]. A direct rise in the activity of Akt and inhibition of PC are additional components of the mechanism by which kaempferol inhibits hepatic gluconeogenesis [22], as Akt phosphorylates and suppresses FOXO1 transcription when insulin signaling is activated, ultimately suppressing PEPCK and G6P expression [404,405]. As part of its anti-inflammatory effects, kaempferol prevents the hepatic inhibitor IkB kinase/NF-kB pathway and restores Akt activity [406]. To create phosphatidylinositol (3,4,5)-triphosphate, insulin first binds to the insulin receptor on the cell’s outer surface, causing tyrosine phosphorylation of the insulin receptor substrate (PIP3). Protein kinase C (PKC) and P70 ribosomal S6 kinase (S6K) are both activated by PIP3 after Akt, a 3-phospholipid-dependent protein kinase I, is activated [407].
The physiological effects of insulin are significantly influenced by Akt-dependent phosphorylation. GSK3a/b is first inactivated by Akt-induced phosphorylation, which then causes dephosphorylation and activation of glycogen synthase [408]. To control the intracellular GLUT4 vesicle movement to the cell membrane and boost glucose uptake, Akt phosphorylates the 160 kDa TBC1D4/AS160 substrate [409,410]. To have an anti-inflammatory effect, kaempferol constrains the hepatic Ik-B kinase/NF-kB pathway and increases Akt activity [406]. Adipose tissues, the liver, and the muscles exhibit increased AMPK and ACC phosphorylation in response to kaempferol [411,412]. For the treatment of diabetes, AMPK activation is an important pharmacological target. In this context, thiazolidinediones (TZDs) and metformin have been recognized as AMPK activators [413]. Foods high in kaempferol can lower postprandial glucose levels and decrease carbohydrate absorption. Changes in the intestinal microbiota play a significant role in metabolic syndrome, type II diabetes, and obesity [414]. Additionally, kaempferol decreases the relative richness of thick-walled flora, boosts bacteroides, lowers blood lipid and glucose levels, and enhances IR in C57BL/6 obese mice [415]. The excellent autophagy enhancer kaempferol reduces ER stress, promotes intracellular lipid degradation, and guards against lipotoxic damage to β-cells [416]. To maintain intracellular balance, autophagy is well-defined as an intracellular lysosomal degradation process of defective proteins, macromolecules, damaged organelles, and toxic aggregates [417]; disorders of autophagy are linked to IR, obesity, and T2DM [418]. In another study, Varshney and coworkers reported that through AMPK mTOR signaling, treatment with 10 µM kaempferol increased lipid droplet co-localization with lysosomes and autophagosomes in cells and decreased ectopic lipid buildup and ER stress [419]. Chronic hyperglycemia in diabetes eventually destroys the mitochondrial function, activates nicotinamide adenine dinucleotide phosphate oxidase, and increases the production of ROS [420]. The excellent antioxidant effect of kaempferol can prevent excessive ROS from damaging β-cells. Kaempferol protects pancreatic β-cells from oxidative damage in diabetes [421]. In the kidney, liver, heart tissues, and erythrocytes of diabetic rats, kaempferol significantly increases membrane-bound ATPase activity [422]. This is yet another way that kaempferol protects β-cells. Natural plants such as ginkgo biloba, galangal, and pueraria have been used for a long time, especially in Asia, and are good sources of kaempferol. In addition, it can be found in foods such as tomatoes, beans, gooseberries, grapes, cabbage, cauliflower, and strawberries [423]. Listed in Table 5 are data pertaining to the role of kaempferol as a potential antidiabetic agent from molecular mechanisms to in vivo studies.

2.2.9. Hesperetin

Hesperidin effectively reduces pancreatic β-cell dysfunction and programmed cell death in diabetic rat models, as well as the expression of the 78-kDa glucose-regulated protein (GRP78) [429]. Additionally, by upregulating the anti-apoptotic cell lymphoma extra-large (Bcl-xL) and downregulating the BCL2-linked X-protein, hesperidin as an apoptosis regulator successfully modulated the expressions of apoptosis regulatory proteins (Bax) [429]. Additionally, by controlling AMPK-mediated p300 inactivation, hesperetin and naringenin protected pancreatic β-cells in both in vitro and in vivo models [430]. The apoptosis of pancreatic β-cells is influenced by the initiation of the MAPK and FoxO1/PPAR signaling pathways [431] and may accelerate the development of type II diabetes and insulin resistance [432]. Furthermore, phosphorylation of the MAPK activates NF-kB, causing the release of pro-inflammatory cytokines [433]. Research findings indicated that hesperetin metabolites reduce inflammation by preventing the phosphorylation of NF-B and MAPK. Finally, it is worth mentioning that hesperidin is most prevalent in citrus fruit [434].

3. Discussion

A diet high in vegetables and fruits offers several nutritional advantages. Vegetables and fruits contain polyphenols in addition to minerals, vitamins, and fiber [435,436]. Flavonoids are polyphenols, which include flavonols, flavanols, flavones, flavanones, anthocyanidins, and isoflavones. They are found in the human diet, such as in citrus fruits, which have the highest concentration of flavanones [437,438]. Increasing the intake of foods high in flavonoids has been linked to positive health effects and a decline in the incidence of chronic ailments such as type II diabetes (T2D), cardiovascular illnesses, and dyslipidemias [437,439]. By lowering oxidative stress, increasing insulin secretion, and enhancing insulin sensitivity, flavonoids protect against high glucose levels [440]. Previous research claimed that flavonoids prevent pancreatic β-cells from undergoing apoptosis [441] and that they engaged in anti-inflammatory, anti-apoptotic, and antioxidant-like activities. Flavonoids regulate these effects by modulating the activity of signaling cascades such as nuclear factor kappa-B (NF-kB) and protein mitogen-activated kinases (MAPKs) [442]. Flavonoids in these functional foods and phytomedicine have beneficial effects on immune function, blood sugar levels, glucose metabolism, and insulin secretion [9]. Numerous controlled studies showed that dietary phenolic consumption reduces diabetes risk factors by regulating the major pathways for carbohydrate metabolism and hepatic glucose homeostasis. Consuming many polyphenols is linked with a decreased risk of developing diabetes mellitus [6]. One of the phenolic acid’s best-known effects on the metabolism of carbohydrates is its ability to inhibit the key enzymes, glucosidase, and amylase, which convert dietetic carbohydrates to glucose [9,443]. Despite a genetic predisposition, dietary changes and augmented physical activity may delay the onset of type II diabetes [444,445].
Diets high in polyphenols can help in managing type II diabetes. The prevention of diabetes in various models of insulin resistance is recognized from changes in the liver, adipose tissue, and skeletal muscle, and animal studies consistently show that resveratrol improves insulin action. Resveratrol alters established pathways for aging, transforms obese mice’s physiology into that of mice on a standard diet, and enhances health, as demonstrated by various indicators such as survival, motor function, organ pathology, insulin sensitivity, PGC-1 activity, and mitochondrial number. Notably, none of these changes arose in tandem with a significant loss in body weight [157]. This is significant and indicates the potential of resveratrol to treat various diseases such as type II diabetes linked to impaired insulin action. However, human studies are needed to assess resveratrol’s therapeutic value given that type II diabetic patients may use it. It is important to note that resveratrol’s positive effects on β-cells were also observed in type II diabetic patients, significantly lowering blood insulin levels in those with hyperinsulinemia. A concurrent decline in the homeostasis model of assessment for β-cell function (HOMA-) was observed in conjunction with this effect [108]. Although resveratrol had some positive effects on type II diabetic patients, other studies showed that it did not affect blood insulin levels or HOMA-B [106]. Reduced demand for insulin is a benefit of resveratrol-induced reduction in insulin resistance. Consequently, β-cell failure is also decreased because they secrete less insulin.
Curcumin is potentially used for the treatment of diabetes and associated complications. It is an inexpensive drug and relatively safe, and it reduced hyperlipidemia and glycemia in rodent models of diabetes. Due to deficiencies in insulin secretion and its action, diabetics cannot effectively metabolize glucose, and curcumin can have a therapeutic effect by playing a crucial role in β-cell functions. A rise in blood glucose levels is a hallmark of T2DM, a heterogeneous and chronic metabolic sickness caused by insulin resistance in target tissues and pancreatic β-cell dysfunction. Preclinical research using animal models and clinical trials found that curcumin pointedly lowers fasting plasma glucose and glycated hemoglobin (HbA1c) levels, according to T2DM results. In the treatment of metabolic syndrome, curcumin successfully lowers triglycerides and LDL-C (low-density lipoprotein cholesterol); enhances fasting blood sugar levels and insulin resistance (HOMA-IR); and reduces AST levels, body weight, and aminotransferase levels [223]. Curcumin has been shown in preclinical studies to lessen inflammation by preventing and regulating the tissue release of pro-inflammatory cytokines, such as IL-4, IL-8, IL-6, and TNF-α [446].
Oral glucose tolerance and insulin secretion by pancreatic β-cells are both enhanced by quercetin. Due to its inhibition of glucosidase and DPP-IV enzymes, glucagon-like peptide-1 (GLP-1) and glucose-dependent insulinotropic polypeptides have a longer half-life (GIP). Additionally, quercetin inhibits the production of pro-inflammatory molecules such as IL-4, IL-6, IL-1, and TNF-α. Through the hangup of glucosidase and interference with glucose transport across intestinal cells’ plasma membrane, catechins regulate glucose absorption through two distinct mechanisms. In particular, catechins help to recover insulin sensitivity, lower blood lipid levels, decrease white fat depots, and reduce blood sugar and lipid levels. In vivo tests using substances such as streptozotocin and alloxan or diets (high fructose and fat diets) inducing T2D in animal models have demonstrated vital anti-hyperglycemic activity for several significant hydroxycinnamic acids, including p-coumaric acid, cinnamic acid, ferulic acid, caffeic acid, chlorogenic acid, and rosmarinic acid [447].
Anthocyanidins are of great nutritional interest because they have demonstrated antidiabetic activity primarily through inhibition of oxidative stress, insulin secretion promotion, insulin resistance improvement, lipid and glucose metabolism, and antioxidant and anti-inflammatory functions. One of the reasons anthocyanins have an anti-T2D outcome is because of their antioxidant properties. This is because oxidative-stress-related cell damage is a significant factor in the development of T2D. To decrease lipo-toxicity, kaempferol regulates lipid metabolism, enhances IR, and improves insulin signaling. It also restores the equilibrium between glucose production and consumption, reducing glucose toxicity. To protect β-cells, kaempferol corrects the imbalance in autophagy and apoptosis. Flavanones can improve health by changing the expression of genes and proteins in pancreatic cells. However, little is known about how flavanones work in pancreatic β-cells underneath high glycemic stress in physiologically relevant concentrations or how they affect the expression of all proteins. Citrus flavonoid hesperetin (Hst), which is effective in preventing diabetes and its complications, has recently attracted the attention of researchers. Novel methods with few side effects are urgently needed to treat diabetes and its complications. New monomeric molecules derived from herbal medicine, a type of complementary medicine, are being sought after for the cure of diabetes as well as its complications.

4. Materials and Methods

4.1. Literature Search and Methodology

In the current review on food polyphenols and type II diabetes mellitus, relevant references published between 2000 and 2022 were obtained from different bibliographical databases such as Google Scholar, PubMed, Web of Science, Science Direct, and Scopus. In our search, we used keywords related to food polyphenols (fruits and vegetables) and their pharmacologic profiling including “nutritional polyphenols”, “traditional medicinal uses”, “in vivo, in vitro anti-diabetic activities“, and “preclinical and clinical studies”. In this work, articles were chosen on the basis of the following criteria: fruits and vegetables containing polyphenols in the evaluation of in vitro/in vivo antidiabetic activity. After the selection of raw material, the pharmacology of anti-diabetic polyphenols was provided. We did not impose language restrictions in our search; however, we only included articles published in English for further consideration.

4.2. Illustrations and Figures

The chemical structures were drawn in ChemDraw 22.0.0 with the help of Pubchem (the mechanistic illustrated figures were drawn in Biorender (https://biorender.com/, accessed on 18 March 2023). Previously published literature data were used to draw the illustrated Figures.

5. Conclusions

T2D, which has a multifactorial pathology, affects millions of people around the world. Treatment of this disease includes lifestyle modifications, dietary adjustments, physical activity, and therapies involving medications for the rest of one’s life. This review article has summarized most of the in vivo and in vitro studies conducted so far to show how food polyphenols affect T2D. Recognized benefits of resveratrol in experimentally insulin-deficient diabetic animals include anti-hyperglycemic action and pancreatic β-cell protection. Curcumin is a safe and cost-effective natural anti-inflammatory and anti-diabetic property that provides a treatment option for this condition, according to several in vivo and in vitro studies, because it is pharmacologically safe, efficient, and with few side effects. In addition to their capacity to influence gene expression and glucose metabolism pathways such as AMPK, anthocyanins also have beneficial effects on insulin resistance; lipid metabolism; glucose metabolism; the immune system; and the ability to modulate hyperlipidemia, hyperglycemia, overweight, obesity, and cardiovascular diseases. Kaempferol may significantly improve how diabetes and its complications are managed. Consequently, dietary polyphenols could be used to prevent and treat diabetes. In addition, results obtained from this review show that natural ingredients are crucial for maintaining good health. Moreover, clinical studies and early research showed that polyphenols can reduce insulin resistance, blood glucose levels, and dyslipidemia in diabetic patients. More preclinical and clinical trials along with cytotoxicity tests should be conducted before these phenolic compounds hit the market as antidiabetic agents. “Over-the-counter” (OTC) polyphenol supplements for diabetics will be clinically effective because they are safe and reduce inflammation and diabetes stress.

Author Contributions

Conceptualization, F.S., M.W. and M.S.M.; methodology, R.N., M.W. and S.A.; writing—original draft preparation, R.N., M.F.L., I.I. and M.W.; writing—review and editing, F.S., S.A. and M.S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

Abbreviations

ACCAcetyl-CoA carboxylase
ACNsAnthocyanins
AGBJAnthocyanins-rich grape-bilberry juice
AGEsAdvanced glycation end products
AKTProtein kinase B
ALTAlanine aminotransferase
AMPAdenosine monophosphate
AMPKAMP-activated kinase
Apo A1Apolipoprotein AI
Apo BApolipoprotein B
ASTAspartate aminotransferase
Bcl-2B-cell lymphoma 2
BMIBody mass index
C3GCyanidin-3-glucoside
CACaffeic acid
CGAChlorogenic acid
ChREBPCarbohydrate-responsive element-binding protein
COX-2Cyclooxygenase-2
CPT1Carnitine palmitoyltransferase I
DAGDiacylglycerol
DPPIVDipeptidyl peptidase-4
EAEllagic acid
EAGEstimated average glucose
ER Endoplasmic reticulum
FAFerulic acid
FASFatty acid synthase
FBGFasting blood glucose
FFAsFree fatty acids
FOXO1Forkhead transcription factor FKHR
GAGalic acid
G6PaseGlucose 6-phosphatase
GCKGlucokinase
GLP-1Glucagon-like peptide-1
GLUT2Glucose transporter type 2
GLUT4Glucose transporter type 4
GMPGuanosine monophosphate
GSHGlutathione
GSISGlucose-stimulated insulin
GTGlucose tolerance
GTPGuanosine triphosphate
HbA1cHemoglobin A1C
HDLHigh-density lipoprotein
HOMA-IRHomeostatic Model Assessment for Insulin Resistance
ICAM-1Intercellular adhesion molecule 1
IKKInhibitor of nuclear factor-κB (IκB) kinase (IKK)
IKKbInhibitor of nuclear factor kappa-B kinase
IL-6Interleukin-6
IMPInosine monophosphate
IRInsulin resistance
IRS1Insulin receptor substrate 1
IRS-1Insulin receptor substrate 1
JNKC-Jun N-terminal kinase
LDLLow-density lipoprotein
MCP1Monocyte chemoattractant protein-1
MDAMalondialdehyde
mTORMammalian target of rapamycin
NADPHNicotinamide adenine dinucleotide phosphate
NF-kBNuclear factor kappa- B
NMNot mentioned
PCPyruvate carboxylase
PCAProtocatechuic acid
PDK1 3-Phosphoinositide-dependent protein kinase-1
PEPCKPhosphoenolpyruvate carboxykinase
PGC-1αPeroxisome-proliferator-activated receptor-gamma coactivator (PGC)-1alpha
PI 3-kinasePhosphatidylinositol 3-kinase
PIP3Phosphatidylinositol (3,4,5)-trisphosphate
PKCProtein kinase C
PPAR-cPeroxisome proliferator-activated receptor-C
PPAR-γPeroxisome proliferator-activated receptor gamma
RBP4Retinol-binding protein 4
ROSReactive oxygen species
S6KS6 kinase
SGLT1Sodium-glucose transporter 1
SIRT1Silent information regulator 1
SODSuperoxide dismutase
SREBP1Sterol regulatory element-binding proteins
SREBP-1Sterol regulatory element-binding protein 1
STZStreptozotocin
T2DType II diabetes
TAGTriacylglycerol
TCTotal cholesterol
TCA Tricarboxylic acid
TGTriglycerides
TGF-βTransforming growth factor-beta
TLR4Toll-like receptor 4
TNF-αTumor necrosis factor α
VCAM-1Vascular cell adhesion molecule 1
VLDLVery low density lipoprotein
WATWhite adipose tissue

References

  1. Halpin, H.A.; Morales-Suárez-Varela, M.M.; Martin—Moreno, J.M. Chronic disease prevention and the new public health. Public Health Rev. 2010, 32, 120–154. [Google Scholar] [CrossRef]
  2. Xiao, J.; Hogger, P. Dietary polyphenols and type 2 diabetes: Current insights and future perspectives. Curr. Med. Chem. 2015, 22, 23–38. [Google Scholar] [CrossRef] [PubMed]
  3. Williamson, G. Possible effects of dietary polyphenols on sugar absorption and digestion. Mol. Nutr. Food Res. 2013, 57, 48–57. [Google Scholar] [CrossRef] [PubMed]
  4. Scalbert, A.; Manach, C.; Morand, C.; Rémésy, C.; Jiménez, L. Dietary polyphenols and the prevention of diseases. Crit. Rev. Food Sci. Nutr. 2005, 45, 287–306. [Google Scholar] [CrossRef] [PubMed]
  5. Lee, J.; Durst, R.W.; Wrolstad, R.E.; Collaborators: Eisele T Giusti MM Hach J Hofsommer H Koswig S Krueger DA Kupina; S Martin SK Martinsen BK Miller TC Paquette F Ryabkova A Skrede G Trenn U Wightman JD. Determination of total monomeric anthocyanin pigment content of fruit juices, beverages, natural colorants, and wines by the pH differential method: Collaborative study. J. AOAC Int. 2005, 88, 1269–1278. [Google Scholar] [CrossRef] [PubMed]
  6. Pandey, K.B.; Rizvi, S.I. Plant polyphenols as dietary antioxidants in human health and disease. Oxid. Med. Cell. Longev. 2009, 2, 270–278. [Google Scholar] [CrossRef]
  7. Manach, C.; Scalbert, A.; Morand, C.; Rémésy, C.; Jiménez, L. Polyphenols: Food sources and bioavailability. AJCN 2004, 79, 727–747. [Google Scholar] [CrossRef]
  8. Cheynier, V. Polyphenols in foods are more complex than often thought. AJCN 2005, 81, 223S–229S. [Google Scholar] [CrossRef]
  9. Hanhineva, K.; Törrönen, R.; Bondia—Pons, I.; Pekkinen, J.; Kolehmainen, M.; Mykkänen, H.; Poutanen, K. Impact of dietary polyphenols on carbohydrate metabolism. Int. J. Mol. Sci. 2010, 11, 1365–1402. [Google Scholar] [CrossRef]
  10. Hajiaghaalipour, F.; Khalilpourfarshbafi, M.; Arya, A. Modulation of glucose transporter protein by dietary flavonoids in type 2 diabetes mellitus. Int. J. Mol. Sci. 2015, 11, 508–524. [Google Scholar] [CrossRef]
  11. Loureiro, G.; Martel, F. The effect of dietary polyphenols on intestinal absorption of glucose and fructose: Relation with obesity and type 2 diabetes. Food Rev. Int. 2019, 35, 390–406. [Google Scholar] [CrossRef]
  12. Xiao, J.; Kai, G.; Yamamoto, K.; Chen, X. Advance in dietary polyphenols as α—Glucosidases inhibitors: A review on structure—Activity relationship aspect. Crit. Rev. Food Sci. Nutr. 2013, 53, 818–836. [Google Scholar] [CrossRef] [PubMed]
  13. Xiao, J.; Ni, X.; Kai, G.; Chen, X. A review on structure–activity relationship of dietary polyphenols inhibiting α—Amylase. Crit. Rev. Food Sci. Nutr. 2013, 53, 497–506. [Google Scholar] [CrossRef]
  14. Gowd, V.; Karim, N.; Shishir, M.R.I.; Xie, L.; Chen, W. Dietary polyphenols to combat the metabolic diseases via altering gut microbiota. Trends Food Sci. Technol. 2019, 93, 81–93. [Google Scholar] [CrossRef]
  15. Xie, Y.; Chen, X. Structures required of polyphenols for inhibiting advanced glycation end products formation. Curr. Drug. Metab. 2013, 14, 414–431. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, P.C.; Wheeler, D.S.; Malhotra, V.; Odoms, K.; Denenberg, A.G.; Wong, H.R. A green tea—Derived polyphenol, epigallocatechin—3—Gallate, inhibits IκB kinase activation and IL—8 gene expression in respiratory epithelium. Inflammation 2002, 26, 233–241. [Google Scholar] [CrossRef]
  17. Pfeilschifter, J.; Eberhardt, W.; Beck, K.F.; Huwiler, A. Redox signaling in mesangial cells. Nephron. Exp. Nephrol. 2003, 93, e23–e26. [Google Scholar] [CrossRef]
  18. Al—Ishaq, R.K.; Abotaleb, M.; Kubatka, P.; Kajo, K.; Büsselberg, D. Flavonoids and their anti—Diabetic effects: Cellular mechanisms and effects to improve blood sugar levels. Biomolecules 2019, 9, 430. [Google Scholar] [CrossRef]
  19. Xiao, J.; Capanoglu, E.; Jassbi, A.R.; Miron, A. Advance on the flavonoid C—Glycosides and health benefits. Crit. Rev. Food Sci. Nutr. 2016, 56 (Suppl. 1), S29–S45. [Google Scholar] [CrossRef]
  20. Poudyal, H.; Panchal, S.; Brown, L. Comparison of purple carrot juice and β-carotene in a high-carbohydrate, high-fat diet-fed rat model of the metabolic syndrome. Br. J. Nutr. 2010, 104, 1322–1332. [Google Scholar] [CrossRef]
  21. 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]
  22. Alkhalidy, H.; Moore, W.; Wang, A.; Luo, J.; McMillan, R.P.; Wang, Y.; Zhen, W.; Hulver, M.W.; Liu, D. Kaempferol ameliorates hyperglycemia through suppressing hepatic gluconeogenesis and enhancing hepatic insulin sensitivity in diet—Induced obese mice. J. Nutr. Biochem. 2018, 58, 90–101. [Google Scholar] [CrossRef] [PubMed]
  23. Sangeetha, R. Luteolin in the management of type 2 diabetes mellitus. Curr. Res. Nutr. Food Sci. 2019, 7, 393–398. [Google Scholar] [CrossRef]
  24. Li, Y.; Zheng, X.; Yi, X.; Liu, C.; Kong, D.; Zhang, J.; Gong, M. Myricetin: A potent approach for the treatment of type 2 diabetes as a natural class B GPCR agonist. FASEB J. 2017, 31, 2603–2611. [Google Scholar] [CrossRef] [PubMed]
  25. Den Hartogh, D.J.; Tsiani, E. Antidiabetic properties of naringenin: A citrus fruit polyphenol. Biomolecules 2019, 9, 99. [Google Scholar] [CrossRef]
  26. Burns, J.; Yokota, T.; Ashihara, H.; Lean, M.E.; Crozier, A. Plant foods and herbal sources of resveratrol. J. Agric. Food Chem. 2002, 50, 3337–3340. [Google Scholar] [CrossRef]
  27. Arts, I.C.; Hollman, P.C. Polyphenols and disease risk in epidemiologic studies. AJCN 2005, 81, 317S–325S. [Google Scholar] [CrossRef]
  28. Da Silva Dias, J.C.; Imai, S. Vegetable consumption and its benefits on diabetes. J. Nutr. Ther. 2017, 6, 1–10. [Google Scholar] [CrossRef]
  29. Wild, S.; Roglic, G.; Green, A.; Sicree, R.; King, H. Global prevalence of diabetes: Estimates for the year 2000 and projections for 2030. Diabetes Care 2004, 27, 1047–1053. [Google Scholar] [CrossRef]
  30. Tuomilehto, J.; Lindstrom, J.; Eriksson, J.; Valle, T.; Hamalainen, H.; Ilanne-Parikka, P.; Keinanen-Kiukaanniemi, S.; Laakso, M.; Louheranta, A.; Rastas, M. Finnish Diabetes Prevention Study. Group 2001, 344, 1343–1350. [Google Scholar]
  31. Paolisso, G.; Tataranni, P.; Foley, J.; Bogardus, C.; Howard, B.; Ravussin, E. A high concentration of fasting plasma non—Esterified fatty acids is a risk factor for the development of NIDDM. Diabetologia 1995, 38, 1213–1217. [Google Scholar] [CrossRef] [PubMed]
  32. Knop, F.K.; Konings, E.; Timmers, S.; Schrauwen, P.; Holst, J.J.; Blaak, E. Thirty days of resveratrol supplementation does not affect postprandial incretin hormone responses, but suppresses postprandial glucagon in obese subjects. Diabet. Med. 2013, 30, 1214–1218. [Google Scholar] [CrossRef] [PubMed]
  33. Chuengsamarn, S.; Rattanamongkolgul, S.; Luechapudiporn, R.; Phisalaphong, C.; Jirawatnotai, S. Curcumin extract for prevention of type 2 diabetes. Diabetes Care 2012, 35, 2121–2127. [Google Scholar] [CrossRef] [PubMed]
  34. Nikbakht, E.; Singh, I.; Vider, J.; Williams, L.T.; Vugic, L.; Gaiz, A.; Kundur, A.R.; Colson, N. Potential of anthocyanin as an anti—Inflammatory agent: A human clinical trial on type 2 diabetic, diabetic at—Risk and healthy adults. Inflamm. Res. 2021, 70, 275–284. [Google Scholar] [CrossRef]
  35. Adams, S.H.; Hoppel, C.L.; Lok, K.H.; Zhao, L.; Wong, S.W.; Minkler, P.E.; Hwang, D.H.; Newman, J.W.; Garvey, W.T. Plasma acylcarnitine profiles suggest incomplete long—Chain fatty acid β—Oxidation and altered tricarboxylic acid cycle activity in type 2 diabetic African—American women. J. Nutr. 2009, 139, 1073–1081. [Google Scholar] [CrossRef] [PubMed]
  36. Chatterjee, S.; Khunti, K.; Davies, M.J. Type 2 diabetes. Lancet 2017, 389, 2239–2251. [Google Scholar] [CrossRef]
  37. Donath, M.Y.; Shoelson, S.E. Type 2 diabetes as an inflammatory disease. Nat. Rev. Immunol. 2011, 11, 98–107. [Google Scholar] [CrossRef]
  38. Zheng, Y.; Ley, S.H.; Hu, F.B. Global etiology and epidemiology of type 2 diabetes mellitus and its complications. Nat. Rev. Endocrinol. 2018, 14, 88–98. [Google Scholar] [CrossRef]
  39. Mooradian, A.D. Dyslipidemia in type 2 diabetes mellitus. Nat. Rev. Endocrinol. 2009, 5, 150–159. [Google Scholar] [CrossRef]
  40. Hu, F.B.; Satija, A.; Manson, J.E. Curbing the diabetes pandemic: The need for global policy solutions. Jama 2015, 313, 2319–2320. [Google Scholar] [CrossRef]
  41. Rizza, R.A. Pathogenesis of fasting and postprandial hyperglycemia in type 2 diabetes: Implications for therapy. Diabetes 2010, 59, 2697–2707. [Google Scholar] [CrossRef] [PubMed]
  42. Solloway, M.J.; Madjidi, A.; Gu, C.; Eastham-Anderson, J.; Clarke, H.J.; Kljavin, N.; Zavala-Solorio, J.; Kates, L.; Friedman, B.; Brauer, M. Glucagon couples hepatic amino acid catabolism to mTOR-dependent regulation of α-cell mass. Cell Rep. 2015, 12, 495–510. [Google Scholar] [CrossRef] [PubMed]
  43. Knudsen, J.G.; Hamilton, A.; Ramracheya, R.; Tarasov, A.I.; Brereton, M.; Haythorne, E.; Chibalina, M.V.; Spegel, P.; Mulder, H.; Zhang, Q. Dysregulation of glucagon secretion by hyperglycemia—Induced sodium—Dependent reduction of ATP production. Cell Metab. 2019, 29, 430–442.e4. [Google Scholar] [CrossRef] [PubMed]
  44. Straub, L.; Scherer, P. Metabolic messengers: Adiponectin. Nat. Metab. 2019, 1, 334–339. [Google Scholar] [CrossRef]
  45. Ravussin, Y.; Leibel, R.L.; Ferrante, A.W. A missing link in body weight homeostasis: The catabolic signal of the overfed state. Cell Metab. 2014, 20, 565–572. [Google Scholar] [CrossRef]
  46. Cohen, P.; Levy, J.D.; Zhang, Y.; Frontini, A.; Kolodin, D.P.; Svensson, K.J.; Lo, J.C.; Zeng, X.; Ye, L.; Khandekar, M.J. Ablation of PRDM16 and beige adipose causes metabolic dysfunction and a subcutaneous to visceral fat switch. Cell 2014, 156, 304–316. [Google Scholar] [CrossRef]
  47. Holst, J.J.; Wewer Albrechtsen, N.J.; Pedersen, J.; Knop, F.K. Glucagon and amino acids are linked in a mutual feedback cycle: The liver-α-cell axis. Diabetes 2017, 66, 235–240. [Google Scholar] [CrossRef]
  48. Winther—Sørensen, M.; Galsgaard, K.D.; Santos, A.; Trammell, S.A.; Sulek, K.; Kuhre, R.E.; Pedersen, J.; Andersen, D.B.; Hassing, A.S.; Dall, M. Glucagon acutely regulates hepatic amino acid catabolism and the effect may be disturbed by steatosis. Mol. Metab. 2020, 42, 101080. [Google Scholar] [CrossRef]
  49. Gar, C.; Haschka, S.J.; Kern—Matschilles, S.; Rauch, B.; Sacco, V.; Prehn, C.; Adamski, J.; Seissler, J.; Wewer Albrechtsen, N.J.; Holst, J.J. The liver–alpha cell axis associates with liver fat and insulin resistance: A validation study in women with non—Steatotic liver fat levels. Diabetologia 2021, 64, 512–520. [Google Scholar] [CrossRef]
  50. Haber, E.; Ximenes, H.; Procópio, J.; Carvalho, C.R.O.D.; Curi, R.; Carpinelli, A.R. Pleiotropic effects of fatty acids on pancreatic β-cells. J. Cell. Physiol. 2003, 194, 1–12. [Google Scholar] [CrossRef]
  51. Jimenez—Feltstrom, J.; Salehi, A.; Abaraviciene, S.M.; Henningsson, R.; Lundquist, I. Abnormally decreased NO and augmented CO production in islets of the leptin—Deficient ob/ob mouse might contribute to explain hyperinsulinemia and islet survival in leptin—Resistant type 2 obese diabetes. Regul. Pept. 2011, 170, 43–51. [Google Scholar] [CrossRef]
  52. Holst, J.J. Incretin therapy for diabetes mellitus type 2. Current Opinion in Endocrinology. Diabetes Obes. Metab. 2020, 27, 2–10. [Google Scholar]
  53. Feingold, K.R. Atypical forms of diabetes. In Endotext [Internet]; MDText.com, Inc.: South Dartmouth, MA, USA, 2022. [Google Scholar]
  54. Combs, T.P.; Pajvani, U.B.; Berg, A.H.; Lin, Y.; Jelicks, L.A.; Laplante, M.; Nawrocki, A.R.; Rajala, M.W.; Parlow, A.F.; Cheeseboro, L. A transgenic mouse with a deletion in the collagenous domain of adiponectin displays elevated circulating adiponectin and improved insulin sensitivity. Endocrinology 2004, 145, 367–383. [Google Scholar] [CrossRef] [PubMed]
  55. Tomas, E.; Tsao, T.-S.; Saha, A.K.; Murrey, H.E.; Zhang, C.C.; Itani, S.I.; Lodish, H.F.; Ruderman, N.B. Enhanced muscle fat oxidation and glucose transport by ACRP30 globular domain: Acetyl–CoA carboxylase inhibition and AMP-activated protein kinase activation. Proc. Natl. Acad. Sci. USA 2002, 99, 16309–16313. [Google Scholar] [CrossRef]
  56. Yamauchi, T.; Kamon, J.; Minokoshi, Y.A.; Ito, Y.; Waki, H.; Uchida, S.; Yamashita, S.; Noda, M.; Kita, S.; Ueki, K. Adiponectin stimulates glucose utilization and fatty—Acid oxidation by activating AMP—Activated protein kinase. Nat. Med. 2002, 8, 1288–1295. [Google Scholar] [CrossRef] [PubMed]
  57. Kubota, N.; Yano, W.; Kubota, T.; Yamauchi, T.; Itoh, S.; Kumagai, H.; Kozono, H.; Takamoto, I.; Okamoto, S.; Shiuchi, T. Adiponectin stimulates AMP—Activated protein kinase in the hypothalamus and increases food intake. Cell Metab. 2007, 6, 55–68. [Google Scholar] [CrossRef]
  58. Stern, J.H.; Rutkowski, J.M.; Scherer, P.E. Adiponectin, leptin, and fatty acids in the maintenance of metabolic homeostasis through adipose tissue crosstalk. Cell Metab. 2016, 23, 770–784. [Google Scholar] [CrossRef]
  59. Fasshauer, M.; Blüher, M.A. Adipokines in health and disease. Trends Pharmacol. Sci. 2015, 36, 461–470. [Google Scholar] [CrossRef] [PubMed]
  60. Dunmore, S.J.; Brown, J. The role of adipokines in b—Cell failure of type 2 diabetes. J. Endocrinol. 2013, 216, 37–45. [Google Scholar] [CrossRef]
  61. Hotamisligil, G.S.; Shargill, N.S.; Spiegelman, B.M. Adipose expression of tumor necrosis factor—α: Direct role in obesity—Linked insulin resistance. Science 1993, 259, 87–91. [Google Scholar] [CrossRef] [PubMed]
  62. Peraldi, P.; Hotamisligil, G.S.; Buurman, W.A.; White, M.F.; Spiegelman, B.M. Tumor necrosis factor (TNF)—α inhibits insulin signaling through stimulation of the p55 TNF receptor and activation of sphingomyelinase. J. Biol. Chem. 1996, 271, 13018–13022. [Google Scholar] [CrossRef] [PubMed]
  63. Severinsen, M.C.K.; Pedersen, B.K. Muscle-organ crosstalk: The emerging roles of myokines. Endocr. Rev. 2020, 41, 594–609. [Google Scholar] [CrossRef] [PubMed]
  64. Peppler, W.T.; Townsend, L.K.; Meers, G.M.; Panasevich, M.R.; MacPherson, R.E.; Rector, R.S.; Wright, D.C. Acute administration of IL-6 improves indices of hepatic glucose and insulin homeostasis in lean and obese mice. Am. J. Physiol. Gastrointest. Liver Physiol. 2019, 316, G166–G178. [Google Scholar] [CrossRef] [PubMed]
  65. Pedersen, B.K.; Febbraio, M.A. Muscle as an endocrine organ: Focus on muscle-derived interleukin-6. Physiol. Rev. 2008, 88, 1379–1406. [Google Scholar] [CrossRef]
  66. Oh, K.-J.; Lee, D.S.; Kim, W.K.; Han, B.S.; Lee, S.C.; Bae, K.-H. Metabolic adaptation in obesity and type II diabetes: Myokines, adipokines and hepatokines. Int. J. Mol. Sci. 2016, 18, 8. [Google Scholar] [CrossRef]
  67. Hotamisligil, G.S.; Murray, D.L.; Choy, L.N.; Spiegelman, B.M. Tumor necrosis factor alpha inhibits signaling from the insulin receptor. Proc. Natl. Acad. Sci. USA 1994, 91, 4854–4858. [Google Scholar] [CrossRef]
  68. Zhang, H.H.; Halbleib, M.; Ahmad, F.; Manganiello, V.C.; Greenberg, A.S. Tumor necrosis factor-α stimulates lipolysis in differentiated human adipocytes through activation of extracellular signal-related kinase and elevation of intracellular cAMP. Diabetes 2002, 51, 2929–2935. [Google Scholar] [CrossRef]
  69. Dalziel, B.; Gosby, A.K.; Richman, R.M.; Bryson, J.M.; Caterson, I.D. Association of the TNF-α- 308 G/A promoter polymorphism with insulin resistance in obesity. Obes. Res. 2002, 10, 401–407. [Google Scholar] [CrossRef]
  70. Navarro—Gonzalez, J.F.; Mora—Fernandez, C. The role of inflammatory cytokines in diabetic nephropathy. J. Am. Soc. Nephrol. 2008, 19, 433–442. [Google Scholar] [CrossRef]
  71. Arkan, M.C.; Hevener, A.L.; Greten, F.R.; Maeda, S.; Li, Z.-W.; Long, J.M.; Wynshaw-Boris, A.; Poli, G.; Olefsky, J.; Karin, M. IKK-β links inflammation to obesity-induced insulin resistance. Nat. Med. 2005, 11, 191–198. [Google Scholar] [CrossRef]
  72. Cai, D.; Yuan, M.; Frantz, D.F.; Melendez, P.A.; Hansen, L.; Lee, J.; Shoelson, S.E. Local and systemic insulin resistance resulting from hepatic activation of IKK-β and NF-κB. Nat. Med. 2005, 11, 183–190. [Google Scholar] [CrossRef] [PubMed]
  73. Sabio, G.; Das, M.; Mora, A.; Zhang, Z.; Jun, J.Y.; Ko, H.J.; Barrett, T.; Kim, J.K.; Davis, R.J. A stress signaling pathway in adipose tissue regulates hepatic insulin resistance. Science 2008, 322, 1539–1543. [Google Scholar] [CrossRef] [PubMed]
  74. Tuncman, G.; Hirosumi, J.; Solinas, G.; Chang, L.; Karin, M.; Hotamisligil, G.S. Functional in vivo interactions between JNK1 and JNK2 isoforms in obesity and insulin resistance. Proc. Natl. Acad. Sci. USA 2006, 103, 10741–10746. [Google Scholar] [CrossRef] [PubMed]
  75. Solinas, G.; Vilcu, C.; Neels, J.G.; Bandyopadhyay, G.K.; Luo, J.-L.; Naugler, W.; Grivennikov, S.; Wynshaw-Boris, A.; Scadeng, M.; Olefsky, J.M. JNK1 in hematopoietically derived cells contributes to diet—Induced inflammation and insulin resistance without affecting obesity. Cell Metab. 2007, 6, 386–397. [Google Scholar] [CrossRef]
  76. Rollins, B.J.; Walz, A.; Baggiolini, M. Recombinant human MCP-1/JE induces chemotaxis, calcium flux, and the respiratory burst in human monocytes. Blood 1991, 78, 1112–1116. [Google Scholar] [CrossRef]
  77. Al—Amily, I.M.; Dunér, P.; Groop, L.; Salehi, A. The functional impact of G protein—Coupled receptor 142 (Gpr142) on pancreatic β—Cell in rodent. Arch. Eur. J. Phys. 2019, 471, 633–645. [Google Scholar] [CrossRef]
  78. Serhan, C.N.; Savill, J. Resolution of inflammation: The beginning programs the end. Nat. Immunol. 2005, 6, 1191–1197. [Google Scholar] [CrossRef]
  79. Von Moltke, J.; Trinidad, N.J.; Moayeri, M.; Kintzer, A.F.; Wang, S.B.; van Rooijen, N.; Brown, C.R.; Krantz, B.A.; Leppla, S.H.; Gronert, K. Rapid induction of inflammatory lipid mediators by the inflammasome in vivo. Nature 2012, 490, 107–111. [Google Scholar] [CrossRef]
  80. Saltiel, A.R.; Kahn, C.R. Insulin signaling and the regulation of glucose and lipid metabolism. Nature 2001, 414, 799–806. [Google Scholar] [CrossRef]
  81. Niswender, K.D. Basal insulin: Physiology, pharmacology, and clinical implications. Postgrad. Med. J. 2011, 123, 17–26. [Google Scholar]
  82. Manning, B.D.; Toker, A. AKT/PKB signaling: Navigating the network. Cell 2017, 169, 381–405. [Google Scholar] [PubMed]
  83. Hardt, P.; Krauss, A.; Bretz, L.; Porsch—Oezcueruemez, M.; Schnell—Kretschmer, H.; Mäser, E.; Bretzel, R.; Zekorn, T.; Klör, H. Pancreatic exocrine function in patients with type 1 and type 2 diabetes mellitus. Acta Diabetol. 2000, 37, 105–110. [Google Scholar] [CrossRef] [PubMed]
  84. Prentki, M.; Corkey, B.E.; Madiraju, S.M. Lipid-associated metabolic signalling networks in pancreatic beta cell function. Diabetologia 2020, 63, 10–20. [Google Scholar] [CrossRef] [PubMed]
  85. Bedi, O.; Aggarwal, S.; Trehanpati, N.; Ramakrishna, G.; Krishan, P. Molecular and pathological events involved in the pathogenesis of diabetes—Associated nonalcoholic fatty liver disease. J. Clin. Exp. Hepatol. 2019, 9, 607–618. [Google Scholar] [CrossRef]
  86. Finck, B.N. Targeting metabolism, insulin resistance, and diabetes to treat nonalcoholic steatohepatitis. Diabetes 2018, 67, 2485–2493. [Google Scholar] [CrossRef]
  87. Sun, Y.; Gao, H.-Y.; Fan, Z.-Y.; He, Y.; Yan, Y.-X. Metabolomics signatures in type 2 diabetes: A systematic review and integrative analysis. J. Clin. Endocrinol. Metab. 2020, 105, 1000–1008. [Google Scholar] [CrossRef]
  88. Hu, M.; Phan, F.; Bourron, O.; Ferré, P.; Foufelle, F. Steatosis and NASH in type 2 diabetes. Biochimie 2017, 143, 37–41. [Google Scholar] [CrossRef]
  89. Gerst, F.; Wagner, R.; Kaiser, G.; Panse, M.; Heni, M.; Machann, J.; Bongers, M.N.; Sartorius, T.; Sipos, B.; Fend, F. Metabolic crosstalk between fatty pancreas and fatty liver: Effects on local inflammation and insulin secretion. Diabetologia 2017, 60, 2240–2251. [Google Scholar] [CrossRef]
  90. Maedler, K.; Spinas, G.A.; Lehmann, R.; Sergeev, P.; Weber, M.; Fontana, A.; Kaiser, N.; Donath, M.Y. Glucose induces β—Cell apoptosis via upregulation of the Fas receptor in human islets. Diabetes 2001, 50, 1683–1690. [Google Scholar] [CrossRef]
  91. Gross, D.N.; Wan, M.; Birnbaum, M.J. The role of FOXO in the regulation of metabolism. Curr. Diabetes Rep. 2009, 9, 208–214. [Google Scholar] [CrossRef]
  92. Park, S.; Sadanala, K.C.; Kim, E.-K. A metabolomic approach to understanding the metabolic link between obesity and diabetes. Mol. Cells 2015, 38, 587. [Google Scholar] [CrossRef] [PubMed]
  93. Sparks, D.L.; Doelle, H.; Chatterjee, C. Circulating nucleotides in health and disease. Recept. Clin. Investig. 2014, 1, e344. [Google Scholar]
  94. Salek, R.M.; Maguire, M.L.; Bentley, E.; Rubtsov, D.V.; Hough, T.; Cheeseman, M.; Nunez, D.; Sweatman, B.C.; Haselden, J.N.; Cox, R. A metabolomic comparison of urinary changes in type 2 diabetes in mouse, rat, and human. Physiol. Genom. 2007, 29, 99–108. [Google Scholar]
  95. Fiehn, O.; Garvey, W.T.; Newman, J.W.; Lok, K.H.; Hoppel, C.L.; Adams, S.H. Plasma metabolomic profiles reflective of glucose homeostasis in non-diabetic and type 2 diabetic obese African-American women. PLoS ONE 2010, 5, e15234. [Google Scholar] [CrossRef] [PubMed]
  96. Guan, M.; Xie, L.; Diao, C.; Wang, N.; Hu, W.; Zheng, Y.; Jin, L.; Yan, Z.; Gao, H. Systemic perturbations of key metabolites in diabetic rats during the evolution of diabetes studied by urine metabonomics. PLoS ONE 2013, 8, e60409. [Google Scholar] [CrossRef]
  97. Dudzinska, W. Purine nucleotides and their metabolites in patients with type 1 and 2 diabetes mellitus. J. Biomed. Sci. Eng. 2014, 2014, 42427. [Google Scholar] [CrossRef]
  98. Huang, Q.; Yin, P.; Wang, J.; Chen, J.; Kong, H.; Lu, X.; Xu, G. Method for liver tissue metabolic profiling study and its application in type 2 diabetic rats based on ultra performance liquid chromatography–mass spectrometry. J. Chromatogr. B 2011, 879, 961–967. [Google Scholar]
  99. Buzzetti, E.; Pinzani, M.; Tsochatzis, E.A. The multiple-hit pathogenesis of non-alcoholic fatty liver disease (NAFLD). Metab. 2016, 65, 1038–1048. [Google Scholar]
  100. Ameer, F.; Scandiuzzi, L.; Hasnain, S.; Kalbacher, H.; Zaidi, N. De novo lipogenesis in health and disease. Metabolism 2014, 63, 895–902. [Google Scholar] [CrossRef]
  101. Bugianesi, E.; McCullough, A.J.; Marchesini, G. Insulin resistance: A metabolic pathway to chronic liver disease. Hepatology 2005, 42, 987–1000. [Google Scholar]
  102. Qian, M.; Hu, H.; Yao, Y.; Zhao, D.; Wang, S.; Pan, C.; Duan, X.; Gao, Y.; Liu, J.; Zhang, Y. Coordinated changes of gut microbiome and lipidome differentiates nonalcoholic steatohepatitis (NASH) from isolated steatosis. Liver Int. 2020, 40, 622–637. [Google Scholar] [CrossRef] [PubMed]
  103. Suhre, K.; Meisinger, C.; Döring, A.; Altmaier, E.; Belcredi, P.; Gieger, C.; Chang, D.; Milburn, M.V.; Gall, W.E.; Weinberger, K.M. Metabolic footprint of diabetes: A multiplatform metabolomics study in an epidemiological setting. PLoS ONE 2010, 5, e13953. [Google Scholar] [CrossRef] [PubMed]
  104. Maedler, K.; Spinas, G.; Dyntar, D.; Moritz, W.; Kaiser, N.; Donath, M.Y. Distinct effects of saturated and monounsaturated fatty acids on β—Cell turnover and function. Diabetes 2001, 50, 69–76. [Google Scholar] [CrossRef] [PubMed]
  105. Shi, H.; Kokoeva, M.V.; Inouye, K.; Tzameli, I.; Yin, H.; Flier, J.S. TLR4 links innate immunity and fatty acid–induced insulin resistance. J. Clin. Investig. 2006, 116, 3015–3025. [Google Scholar] [CrossRef]
  106. Chavez, J.A.; Summers, S.A. Characterizing the effects of saturated fatty acids on insulin signaling and ceramide and diacylglycerol accumulation in 3T3-L1 adipocytes and C2C12 myotubes. Arch. Biochem. Biophys. 2003, 419, 101–109. [Google Scholar] [CrossRef]
  107. Holland, W.L.; Bikman, B.T.; Wang, L.-P.; Yuguang, G.; Sargent, K.M.; Bulchand, S.; Knotts, T.A.; Shui, G.; Clegg, D.J.; Wenk, M.R. Lipid-induced insulin resistance mediated by the proinflammatory receptor TLR4 requires saturated fatty acid–induced ceramide biosynthesis in mice. J. Clin. Investig. 2011, 121, 1858–1870. [Google Scholar] [CrossRef]
  108. Wahid, M.; Ali, A.; Saqib, F.; Aleem, A.; Bibi, S.; Afzal, K.; Ali, A.; Baig, A.; Khan, S.A.; Bin Asad, M.H.H. Pharmacological exploration of traditional plants for the treatment of neurodegenerative disorders. Phytother. Res. 2020, 34, 30893112. [Google Scholar] [CrossRef]
  109. Hotamisligil, G.S.; Davis, R.J. Cell signaling and stress responses. Cold Spring Harb. Perspect. Biol. 2016, 8, a006072. [Google Scholar] [CrossRef]
  110. Frakes, A.E.; Dillin, A. The UPRER: Sensor and coordinator of organismal homeostasis. Mol. Cell 2017, 66, 761–771. [Google Scholar] [CrossRef]
  111. Zimmet, P.; Alberti, K.G.; Magliano, D.J.; Bennett, P.H. Diabetes mellitus statistics on prevalence and mortality: Facts and fallacies. Nat. Rev. Endocrinol. 2016, 12, 616–622. [Google Scholar] [CrossRef]
  112. Ramırez, M.; Amate, L.; Gil, A. Absorption and distribution of dietary fatty acids from different sources. Early Hum. Dev. 2001, 65, S95–S101. [Google Scholar] [CrossRef] [PubMed]
  113. Rui, L. Energy metabolism in the liver. Compr. Physiol. 2014, 4, 177. [Google Scholar] [PubMed]
  114. Bonacina, F.; Baragetti, A.; Catapano, A.L.; Norata, G.D. The interconnection between immuno-metabolism, diabetes, and CKD. Curr. Diabetes Rep. 2019, 19, 1–8. [Google Scholar] [CrossRef]
  115. Capurso, C.; Capurso, A. From excess adiposity to insulin resistance: The role of free fatty acids. Vasc. Pharmacol. 2012, 57, 91–97. [Google Scholar] [CrossRef] [PubMed]
  116. Legrand—Poels, S.; Esser, N.; L’homme, L.; Scheen, A.; Paquot, N.; Piette, J. Free fatty acids as modulators of the NLRP3 inflammasome in obesity/type 2 diabetes. Biochem. Pharmacol. 2014, 92, 131–141. [Google Scholar] [CrossRef]
  117. Sharma, R.B.; Alonso, L.C. Lipotoxicity in the pancreatic beta cell: Not just survival and function, but proliferation as well? Curr. Diabetes Rep. 2014, 14, 492. [Google Scholar] [CrossRef]
  118. Shimabukuro, M.; Zhou, Y.-T.; Levi, M.; Unger, R.H. Fatty acid—Induced β cell apoptosis: A link between obesity and diabetes. Proc. Natl. Acad. Sci. USA 1998, 95, 2498–2502. [Google Scholar] [CrossRef]
  119. Biden, T.J.; Boslem, E.; Chu, K.Y.; Sue, N. Lipotoxic endoplasmic reticulum stress, β cell failure, and type 2 diabetes mellitus. Trends Endocrinol. Metab. 2014, 25, 389–398. [Google Scholar] [CrossRef]
  120. IS Sobczak, A.; Blindauer, C.A.; Stewart, A.J. Changes in plasma free fatty acids associated with type-2 diabetes. Nutrients 2019, 11, 2022. [Google Scholar] [CrossRef]
  121. Suganami, T.; Ogawa, Y. Adipose tissue macrophages: Their role in adipose tissue remodeling. J. Leukoc. Biol. 2010, 88, 33–39. [Google Scholar] [CrossRef]
  122. Tumova, J.; Andel, M.; Trnka, J. Excess of free fatty acids as a cause of metabolic dysfunction in skeletal muscle. Physiol. Res. 2016, 65, 193. [Google Scholar] [CrossRef] [PubMed]
  123. Goodpaster, B.H.; He, J.; Watkins, S.; Kelley, D.E. Skeletal muscle lipid content and insulin resistance: Evidence for a paradox in endurance—Trained athletes. J. Clin. Endocrinol. Metab. 2001, 86, 5755–5761. [Google Scholar] [CrossRef] [PubMed]
  124. Stefan, N.; Kantartzis, K.; Häring, H.-U. Causes and metabolic consequences of fatty liver. Endocr. Rev. 2008, 29, 939–960. [Google Scholar] [CrossRef] [PubMed]
  125. Szendroedi, J.; Yoshimura, T.; Phielix, E.; Koliaki, C.; Marcucci, M.; Zhang, D.; Jelenik, T.; Müller, J.; Herder, C.; Nowotny, P. Role of diacylglycerol activation of PKCθ in lipid—Induced muscle insulin resistance in humans. Proc. Natl. Acad. Sci. USA 2014, 111, 9597–9602. [Google Scholar] [CrossRef]
  126. Erion, D.M.; Shulman, G.I. Diacylglycerol—Mediated insulin resistance. Nat. Med. 2010, 16, 400–402. [Google Scholar] [CrossRef]
  127. Samuel, V.T.; Petersen, K.F.; Shulman, G.I. Lipid—Induced insulin resistance: Unraveling the mechanism. Lancet 2010, 375, 2267–2277. [Google Scholar] [CrossRef]
  128. Boden, G.; Lebed, B.; Schatz, M.; Homko, C.; Lemieux, S. Effects of acute changes of plasma free fatty acids on intramyocellular fat content and insulin resistance in healthy subjects. Diabetes 2001, 50, 1612–1617. [Google Scholar] [CrossRef]
  129. Shulman, G.I. Ectopic fat in insulin resistance, dyslipidemia, and cardiometabolic disease. N. Engl. J. Med. 2014, 371, 1131–1141. [Google Scholar] [CrossRef]
  130. Yang, Q.; Vijayakumar, A.; Kahn, B.B. Metabolites as regulators of insulin sensitivity and metabolism. Nat. Rev. Mol. Cell. Biol. 2018, 19, 654–672. [Google Scholar] [CrossRef]
  131. Boden, G. 45Obesity, insulin resistance and free fatty acids. Curr. Opin. Endocrinol. Diabetes Obes. 2011, 18, 139. [Google Scholar] [CrossRef]
  132. Kishimoto, A.; Takai, Y.; Mori, T.; Kikkawa, U.; Nishizuka, Y. Activation of calcium and phospholipid—Dependent protein kinase by diacylglycerol, its possible relation to phosphatidylinositol turnover. J. Biol. Chem. 1980, 255, 2273–2276. [Google Scholar] [CrossRef] [PubMed]
  133. Gray, S.; Idris, I.; Davis, K.; Donnelly, R. Increased skeletal muscle expression of PKC-θ but not PKC-α mRNA in type 2 diabetes: Inverse relationship with in-vivo insulin sensitivity. Eur. J. Clin. Investig. 2003, 33, 983–987. [Google Scholar] [CrossRef] [PubMed]
  134. Chavez, J.A.; Knotts, T.A.; Wang, L.-P.; Li, G.; Dobrowsky, R.T.; Florant, G.L.; Summers, S.A. A role for ceramide, but not diacylglycerol, in the antagonism of insulin signal transduction by saturated fatty acids. J. Biol. Chem. 2003, 278, 10297–10303. [Google Scholar] [CrossRef] [PubMed]
  135. Teruel, T.; Hernandez, R.; Lorenzo, M. Ceramide mediates insulin resistance by tumor necrosis factor-α in brown adipocytes by maintaining Akt in an inactive dephosphorylated state. Diabetes 2001, 50, 2563–2571. [Google Scholar] [CrossRef] [PubMed]
  136. Stratford, S.; Dewald, D.B.; Summers, S.A. Ceramide dissociates 3′—Phosphoinositide production from pleckstrin homology domain translocation. Biochem. J. 2001, 354, 359–368. [Google Scholar] [CrossRef]
  137. Salinas, M.; López-Valdaliso, R.; Martín, D.; Alvarez, A.; Cuadrado, A. Inhibition of PKB/Akt1 by C2-ceramide involves activation of ceramide-activated protein phosphatase in PC12 cells. Mol. Cell. Neurosci. 2000, 15, 156–169. [Google Scholar] [CrossRef]
  138. Hajduch, E.; Balendran, A.; Batty, I.; Litherland, G.; Blair, A.; Downes, C.; Hundal, H. Ceramide impairs the insulin—Dependent membrane recruitment of protein kinase B leading to a loss in downstream signaling in L6 skeletal muscle cells. Diabetologia 2001, 44, 173–183. [Google Scholar] [CrossRef]
  139. Zhang, Y.; Li, X.; Becker, K.A.; Gulbins, E. Ceramide-enriched membrane domains-structure and function. Biochim. Biophys. Acta Biomembr. 2009, 1788, 178–183. [Google Scholar] [CrossRef]
  140. Paolisso, G.; Gambardella, A.; Tagliamonte, M.R.; Saccomanno, F.; Salvatore, T.; Gualdiero, P.; D’Onofrio, M.; Howard, B.V. Does free fatty acid infusion impair insulin action also through an increase in oxidative stress? J. Clin. Endocrinol. Metab. 1996, 81, 4244–4248. [Google Scholar] [CrossRef]
  141. Furukawa, S.; Fujita, T.; Shimabukuro, M.; Iwaki, M.; Yamada, Y.; Nakajima, Y.; Nakayama, O.; Makishima, M.; Matsuda, M.; Shimomura, I. Increased oxidative stress in obesity and its impact on metabolic syndrome. J. Clin. Investig. 2017, 114, 1752–1761. [Google Scholar] [CrossRef]
  142. Halim, M.; Halim, A. The effects of inflammation, aging and oxidative stress on the pathogenesis of diabetes mellitus (type 2 diabetes). Diabetes & metabolic syndrome: J. Med. Clin. Res. Rev. 2019, 13, 1165–1172. [Google Scholar]
  143. Ye, J. Emerging role of adipose tissue hypoxia in obesity and insulin resistance. Int. J. Obes. 2009, 33, 54–66. [Google Scholar] [CrossRef] [PubMed]
  144. Gao, Z.; Hwang, D.; Bataille, F.; Lefevre, M.; York, D.; Quon, M.J.; Ye, J. Serine phosphorylation of insulin receptor substrate 1 by inhibitor κB kinase complex. J. Biol. Chem. 2002, 277, 48115–48121. [Google Scholar] [CrossRef]
  145. Aguirre, V.; Uchida, T.; Yenush, L.; Davis, R.; White, M.F. The c—Jun NH2—Terminal kinase promotes insulin resistance during association with insulin receptor substrate-1 and phosphorylation of Ser307. J. Biol. Chem. 2000, 275, 9047–9054. [Google Scholar] [CrossRef]
  146. Rui, L.; Aguirre, V.; Kim, J.K.; Shulman, G.I.; Lee, A.; Corbould, A.; Dunaif, A.; White, M.F. Insulin/IGF—1 and TNF—α stimulate phosphorylation of IRS-1 at inhibitory Ser 307 via distinct pathways. J. Clin. Investig. 2001, 107, 181–189. [Google Scholar] [CrossRef] [PubMed]
  147. Boden, G.; She, P.; Mozzoli, M.; Cheung, P.; Gumireddy, K.; Reddy, P.; Xiang, X.; Luo, Z.; Ruderman, N. Free fatty acids produce insulin resistance and activate the proinflammatory nuclear factor-κB pathway in rat liver. Diabetes 2005, 54, 3458–3465. [Google Scholar] [CrossRef]
  148. Boden, G.; Chen, X. Effects of fat on glucose uptake and utilization in patients with non-insulin-dependent diabetes. J. Clin. Investig. 1995, 96, 1261–1268. [Google Scholar] [CrossRef]
  149. Dresner, A.; Laurent, D.; Marcucci, M.; Griffin, M.E.; Dufour, S.; Cline, G.W.; Slezak, L.A.; Andersen, D.K.; Hundal, R.S.; Rothman, D.L. Effects of free fatty acids on glucose transport and IRS-1–associated phosphatidylinositol 3-kinase activity. J. Clin. Investig. 1999, 103, 253–259. [Google Scholar] [CrossRef]
  150. Wei, Y.; Wang, D.; Topczewski, F.; Pagliassotti, M.J. Saturated fatty acids induce endoplasmic reticulum stress and apoptosis independently of ceramide in liver cells. Am. J. Physiol. Endocrinol. Metab. 2006, 291, E275–E281. [Google Scholar] [CrossRef]
  151. Karaskov, E.; Scott, C.; Zhang, L.; Teodoro, T.; Ravazzola, M.; Volchuk, A. Chronic palmitate but not oleate exposure induces endoplasmic reticulum stress, which may contribute to INS-1 pancreatic β-cell apoptosis. Endocrinology 2006, 147, 3398–3407. [Google Scholar] [CrossRef]
  152. Guo, W.; Wong, S.; Xie, W.; Lei, T.; Luo, Z. Palmitate modulates intracellular signaling, induces endoplasmic reticulum stress, and causes apoptosis in mouse 3T3-L1 and rat primary preadipocytes. Am. J. Physiol. Endocrinol. Metab. 2007, 293, E576–E586. [Google Scholar] [CrossRef] [PubMed]
  153. Urano, F.; Wang, X.; Bertolotti, A.; Zhang, Y.; Chung, P.; Harding, H.P.; Ron, D. Coupling of stress in the ER to activation of JNK protein kinases by transmembrane protein kinase IRE1. Science 2000, 287, 664–666. [Google Scholar] [CrossRef] [PubMed]
  154. Davis, B.K.; Wen, H.; Ting, J.P.-Y. The inflammasome NLRs in immunity, inflammation, and associated diseases. Annu. Rev. Immunol. 2011, 29, 707–735. [Google Scholar] [CrossRef] [PubMed]
  155. Platnich, J.M.; Muruve, D.A. NOD-like receptors and inflammasomes: A review of their canonical and non-canonical signaling pathways. Arch. Biochem. Biophys. 2019, 670, 4–14. [Google Scholar] [CrossRef] [PubMed]
  156. Kim, Y.K.; Shin, J.-S.; Nahm, M.H. NOD-like receptors in infection, immunity, and diseases. Yonsei Med. J. 2016, 57, 5–14. [Google Scholar] [CrossRef] [PubMed]
  157. Baur, J.A.; Pearson, K.J.; Price, N.L.; Jamieson, H.A.; Lerin, C.; Kalra, A.; Prabhu, V.V.; Allard, J.S.; Lopez—Lluch, G.; Lewis, K. Resveratrol improves health and survival of mice on a high—Calorie diet. Nature 2006, 444, 337–342. [Google Scholar] [CrossRef]
  158. Nisoli, E.; Tonello, C.; Cardile, A.; Cozzi, V.; Bracale, R.; Tedesco, L.; Falcone, S.; Valerio, A.; Cantoni, O.; Clementi, E. Calorie restriction promotes mitochondrial biogenesis by inducing the expression of eNOS. Science 2005, 310, 314–317. [Google Scholar] [CrossRef]
  159. López—Lluch, G.; Hunt, N.; Jones, B.; Zhu, M.; Jamieson, H.; Hilmer, S.; Cascajo, M.; Allard, J.; Ingram, D.K.; Navas, P. Calorie restriction induces mitochondrial biogenesis and bioenergetic efficiency. Proc. Natl. Acad. Sci. USA 2006, 103, 1768–1773. [Google Scholar] [CrossRef]
  160. Deng, J.-Y.; Hsieh, P.-S.; Huang, J.-P.; Lu, L.-S.; Hung, L.-M. Activation of estrogen receptor is crucial for resveratrol-stimulating muscular glucose uptake via both insulin-dependent and-independent pathways. Diabetes 2008, 57, 1814–1823. [Google Scholar] [CrossRef]
  161. Tan, Z.; Zhou, L.-J.; Mu, P.-W.; Liu, S.-P.; Chen, S.-J.; Fu, X.-D.; Wang, T.-H. Caveolin-3 is involved in the protection of resveratrol against high-fat-diet-induced insulin resistance by promoting GLUT4 translocation to the plasma membrane in skeletal muscle of ovariectomized rats. J. Nutr. Biochem. 2012, 23, 1716–1724. [Google Scholar] [CrossRef]
  162. Chen, L.-L.; Zhang, H.-H.; Zheng, J.; Hu, X.; Kong, W.; Hu, D.; Wang, S.-X.; Zhang, P. Resveratrol attenuates high-fat diet–induced insulin resistance by influencing skeletal muscle lipid transport and subsarcolemmal mitochondrial β-oxidation. Metabolism 2011, 60, 1598–1609. [Google Scholar] [CrossRef] [PubMed]
  163. Kim, S.; Jin, Y.; Choi, Y.; Park, T. Resveratrol exerts anti—Obesity effects via mechanisms involving down—Regulation of adipogenic and inflammatory processes in mice. Biochem. Pharmacol. 2011, 81, 1343–1351. [Google Scholar] [CrossRef] [PubMed]
  164. Do, G.M.; Jung, U.J.; Park, H.J.; Kwon, E.Y.; Jeon, S.M.; McGregor, R.A.; Choi, M.S. Resveratrol ameliorates diabetes-related metabolic changes via activation of AMP-activated protein kinase and its downstream targets in db/db mice. Mol. Nutr. Food Res. 2012, 56, 1282–1291. [Google Scholar] [CrossRef] [PubMed]
  165. Burgess, T.A.; Robich, M.P.; Chu, L.M.; Bianchi, C.; Sellke, F.W. Improving glucose metabolism with resveratrol in a swine model of metabolic syndrome through alteration of signaling pathways in the liver and skeletal muscle. Arch. Surg. 2011, 146, 556–564. [Google Scholar] [CrossRef] [PubMed]
  166. Um, J.-H.; Park, S.-J.; Kang, H.; Yang, S.; Foretz, M.; McBurney, M.W.; Kim, M.K.; Viollet, B.; Chung, J.H. AMP-activated protein kinase–deficient mice are resistant to the metabolic effects of resveratrol. Diabetes 2010, 59, 554–563. [Google Scholar] [CrossRef]
  167. Coen, P.M.; Goodpaster, B.H. Role of intramyocelluar lipids in human health. Trends Endocrinol. Metab. 2012, 23, 391–398. [Google Scholar] [CrossRef]
  168. Kitada, M.; Koya, D. SIRT1 in type 2 diabetes: Mechanisms and therapeutic potential. Diabetes Metab. J. 2013, 37, 315–325. [Google Scholar] [CrossRef]
  169. Kitada, M.; Kume, S.; Kanasaki, K.; Takeda—Watanabe, A.; Koya, D. Sirtuins as possible drug targets in type 2 diabetes. Curr. Drug. Targets 2013, 14, 622–636. [Google Scholar] [CrossRef]
  170. Baur, J.A. Biochemical effects of SIRT1 activators. Biochim. Biophys. Acta—Proteins Proteom. 2010, 1804, 1626–1634. [Google Scholar] [CrossRef]
  171. Ruderman, N.B.; Carling, D.; Prentki, M.; Cacicedo, J.M. AMPK, insulin resistance, and the metabolic syndrome. J. Clin. Investig. 2013, 123, 2764–2772. [Google Scholar] [CrossRef]
  172. McCarty, M.F. Chronic activation of AMP-activated kinase as a strategy for slowing aging. Med. Hypotheses 2004, 63, 334–339. [Google Scholar] [CrossRef] [PubMed]
  173. Rodgers, J.T.; Lerin, C.; Haas, W.; Gygi, S.P.; Spiegelman, B.M.; Puigserver, P. Nutrient control of glucose homeostasis through a complex of PGC—1α and SIRT1. Nature 2005, 434, 113–118. [Google Scholar] [CrossRef] [PubMed]
  174. Lerin, C.; Rodgers, J.T.; Kalume, D.E.; Kim, S.-h.; Pandey, A.; Puigserver, P. GCN5 acetyltransferase complex controls glucose metabolism through transcriptional repression of PGC—1α. Cell Metab. 2006, 3, 429–438. [Google Scholar] [CrossRef] [PubMed]
  175. Brasnyó, P.; Molnár, G.A.; Mohás, M.; Markó, L.; Laczy, B.; Cseh, J.; Mikolás, E.; Szijártó, I.A.; Mérei, A.; Halmai, R. Resveratrol improves insulin sensitivity, reduces oxidative stress and activates the Akt pathway in type 2 diabetic patients. Br. J. Nutr. 2011, 106, 383–389. [Google Scholar] [CrossRef] [PubMed]
  176. Mahjabeen, W.; Khan, D.A.; Mirza, S.A. Role of resveratrol supplementation in regulation of glucose hemostasis, inflammation and oxidative stress in patients with diabetes mellitus type 2: A randomized, placebo-controlled trial. Complement. Ther. Med. 2022, 66, 102819. [Google Scholar] [CrossRef]
  177. Yoshino, J.; Conte, C.; Fontana, L.; Mittendorfer, B.; Imai, S.-I.; Schechtman, K.B.; Gu, C.; Kunz, I.; Fanelli, F.R.; Patterson, B.W. Resveratrol supplementation does not improve metabolic function in nonobese women with normal glucose tolerance. Cell Metab. 2012, 16, 658–664. [Google Scholar] [CrossRef]
  178. Bashmakov, Y.K.; Assaad-Khalil, S.H.; Abou Seif, M.; Udumyan, R.; Megallaa, M.; Rohoma, K.H.; Zeitoun, M.; Petyaev, I.M. Resveratrol promotes foot ulcer size reduction in type 2 diabetes patients. Int. Sch. Res. Not. 2014, 2014, 816307. [Google Scholar] [CrossRef]
  179. Timmers, S.; Konings, E.; Bilet, L.; Houtkooper, R.H.; van de Weijer, T.; Goossens, G.H.; Hoeks, J.; van der Krieken, S.; Ryu, D.; Kersten, S. Calorie restriction—Like effects of 30 days of resveratrol supplementation on energy metabolism and metabolic profile in obese humans. Cell Metab. 2011, 14, 612–622. [Google Scholar] [CrossRef]
  180. Timmers, S.; De Ligt, M.; Phielix, E.; Van De Weijer, T.; Hansen, J.; Moonen-Kornips, E.; Schaart, G.; Kunz, I.; Hesselink, M.K.; Schrauwen-Hinderling, V.B. Resveratrol as add-on therapy in subjects with well-controlled type 2 diabetes: A randomized controlled trial. Diabetes Care 2016, 39, 2211–2217. [Google Scholar] [CrossRef]
  181. Bhatt, J.K.; Thomas, S.; Nanjan, M.J. Resveratrol supplementation improves glycemic control in type 2 diabetes mellitus. Nutr. Res. 2012, 32, 537–541. [Google Scholar] [CrossRef]
  182. Kumar, B.J.; Joghee, N.M. Resveratrol supplementation in patients with type 2 diabetes mellitus: A prospective, open label, randomized controlled trial. Int. Res. J. Pharm. 2013, 4, 246–249. [Google Scholar]
  183. Olesen, J.; Gliemann, L.; Biensø, R.; Schmidt, J.; Hellsten, Y.; Pilegaard, H. Exercise training, but not resveratrol, improves metabolic and inflammatory status in skeletal muscle of aged men. J. Physiol. 2014, 592, 1873–1886. [Google Scholar] [CrossRef] [PubMed]
  184. Sattarinezhad, A.; Roozbeh, J.; Yeganeh, B.S.; Omrani, G.; Shams, M. Resveratrol reduces albuminuria in diabetic nephropathy: A randomized double—Blind placebo—Controlled clinical trial. Diabetes Metab. J. 2019, 45, 53–59. [Google Scholar] [CrossRef]
  185. Movahed, A.; Nabipour, I.; Lieben Louis, X.; Thandapilly, S.J.; Yu, L.; Kalantarhormozi, M.; Rekabpour, S.J.; Netticadan, T. Antihyperglycemic effects of short term resveratrol supplementation in type 2 diabetic patients. Evid. -Based Complement. Altern. Med. 2013, 2013, 851267. [Google Scholar] [CrossRef]
  186. Poulsen, M.M.; Vestergaard, P.F.; Clasen, B.F.; Radko, Y.; Christensen, L.P.; Stødkilde-Jørgensen, H.; Møller, N.; Jessen, N.; Pedersen, S.B.; Jørgensen, J.O.L. High-dose resveratrol supplementation in obese men: An investigator-initiated, randomized, placebo-controlled clinical trial of substrate metabolism, insulin sensitivity, and body composition. Diabetes 2013, 62, 1186–1195. [Google Scholar] [CrossRef] [PubMed]
  187. Goh, K.P.; Lee, H.Y.; Lau, D.P.; Supaat, W.; Chan, Y.H.; Koh, A.F.Y. Effects of resveratrol in patients with type 2 diabetes mellitus on skeletal muscle SIRT1 expression and energy expenditure. Int. J. Sport. Nutr. Exerc. Metab. 2014, 24, 2–13. [Google Scholar] [CrossRef]
  188. Arun, N.; Nalini, N. Efficacy of turmeric on blood sugar and polyol pathway in diabetic albino rats. Plant. Foods Hum. Nutr. 2002, 57, 41–52. [Google Scholar] [CrossRef]
  189. Murugan, P.; Pari, L. Influence of tetrahydrocurcumin on hepatic and renal functional markers and protein levels in experimental type 2 diabetic rats. Basic Clin. Pharmacol. Toxicol. 2007, 101, 241–245. [Google Scholar] [CrossRef]
  190. Abu-Taweel, G.M.; Attia, M.F.; Hussein, J.; Mekawi, E.M.; Galal, H.M.; Ahmed, E.I.; Allam, A.A.; El-Naggar, M.E. Curcumin nanoparticles have potential antioxidant effect and restore tetrahydrobiopterin levels in experimental diabetes. Biomed. Pharmacother. 2020, 131, 110688. [Google Scholar] [CrossRef]
  191. Pivari, F.; Mingione, A.; Brasacchio, C.; Soldati, L. Curcumin and type 2 diabetes mellitus: Prevention and treatment. Nutrients 2019, 11, 1837. [Google Scholar] [CrossRef]
  192. Wickenberg, J.; Ingemansson, S.L.; Hlebowicz, J. Effects of Curcuma longa (turmeric) on postprandial plasma glucose and insulin in healthy subjects. J. Nutr. 2010, 9, 43. [Google Scholar] [CrossRef] [PubMed]
  193. Gutierres, V.O.; Pinheiro, C.M.; Assis, R.P.; Vendramini, R.C.; Pepato, M.T.; Brunetti, I.L. Curcumin-supplemented yoghurt improves physiological and biochemical markers of experimental diabetes. Br. J. Nutr. 2012, 108, 440–448. [Google Scholar] [CrossRef] [PubMed]
  194. Gutierres, V.O.; Assis, R.P.; Arcaro, C.A.; Oliveira, J.O.; Lima, T.F.O.; Beretta, A.L.R.Z.; Costa, P.I.; Baviera, A.M.; Brunetti, I.L. Curcumin improves the effect of a reduced insulin dose on glycemic control and oxidative stress in streptozotocin-diabetic rats. Phytother. Res. 2019, 33, 976–988. [Google Scholar] [CrossRef] [PubMed]
  195. Liu, J.; Chen, Z.; Wang, J.; Li, R.; Li, T.; Chang, M.; Yan, F.; Wang, Y. Encapsulation of curcumin nanoparticles with MMP9—Responsive and thermos—Sensitive hydrogel improves diabetic wound healing. ACS Appl. Mater. Interfaces 2018, 10, 16315–16326. [Google Scholar] [CrossRef]
  196. Matei, A.-M.; Caruntu, C.; Tampa, M.; Georgescu, S.R.; Matei, C.; Constantin, M.M.; Constantin, T.V.; Calina, D.; Ciubotaru, D.A.; Badarau, I.A. Applications of nanosized-lipid-based drug delivery systems in wound care. Appl. Sci. 2021, 11, 4915. [Google Scholar] [CrossRef]
  197. Yang, F.; Yu, J.; Ke, F.; Lan, M.; Li, D.; Tan, K.; Ling, J.; Wang, Y.; Wu, K.; Li, D. Curcumin alleviates diabetic retinopathy in experimental diabetic rats. Ophthalmic Res. 2018, 60, 43–54. [Google Scholar] [CrossRef]
  198. Munir, D.; Maria, A.; Bashiruddin, J. The antioxidant effect of curcumin on cochlear fibroblasts in rat models of diabetes mellitus. Iran. J. Otorhinolaryngol. 2017, 29, 197. [Google Scholar]
  199. Liang, Y.; Zhu, B.; Li, S.; Zhai, Y.; Yang, Y.; Bai, Z.; Zeng, Y.; Li, D. Curcumin protects bone biomechanical properties and microarchitecture in type 2 diabetic rats with osteoporosis via the TGFβ/Smad2/3 pathway. Exp. Ther. Med. 2020, 20, 2200–2208. [Google Scholar] [CrossRef]
  200. Rahimi, H.R.; Mohammadpour, A.H.; Dastani, M.; Jaafari, M.R.; Abnous, K.; Mobarhan, M.G.; Oskuee, R.K. The effect of nano—Curcumin on HbA1c, fasting blood glucose, and lipid profile in diabetic subjects: A randomized clinical trial. Avicenna J. Phytomedicine 2016, 6, 567. [Google Scholar]
  201. Panahi, Y.; Khalili, N.; Sahebi, E.; Namazi, S.; Karimian, M.S.; Majeed, M.; Sahebkar, A. Antioxidant effects of curcuminoids in patients with type 2 diabetes mellitus: A randomized controlled trial. Inflammopharmacology 2017, 25, 25–31. [Google Scholar] [CrossRef]
  202. Jain, S.K.; Rains, J.; Croad, J.; Larson, B.; Jones, K. Curcumin supplementation lowers TNF-α, IL-6, IL-8, and MCP-1 secretion in high glucose-treated cultured monocytes and blood levels of TNF-α, IL-6, MCP-1, glucose, and glycosylated hemoglobin in diabetic rats. Antioxid. Redox Signal. 2009, 11, 241–249. [Google Scholar] [CrossRef] [PubMed]
  203. Jiménez-Flores, L.M.; López-Briones, S.; Macías-Cervantes, M.H.; Ramírez-Emiliano, J.; Pérez-Vázquez, V. A PPARγ, NF-κB and AMPK-dependent mechanism may be involved in the beneficial effects of curcumin in the diabetic db/db mice liver. Molecules 2014, 19, 8289–8302. [Google Scholar] [CrossRef] [PubMed]
  204. He, Y.; Yue, Y.; Zheng, X.; Zhang, K.; Chen, S.; Du, Z. Curcumin, inflammation, and chronic diseases: How are they linked? Molecules 2015, 20, 9183–9213. [Google Scholar] [CrossRef] [PubMed]
  205. Abo-Salem, O.; Harisa, G.; Ali, T.; El-Sayed, E.; Abou-Elnour, F. Curcumin ameliorates streptozotocin-induced heart injury in rats: Curcumin attenuates diabetic heart injury. J. Biochem. Mol. Toxicol. 2014, 28, 263–270. [Google Scholar] [CrossRef]
  206. Arafa, H. Curcumin attenuates diet—Induced hypercholesterolemia in rats. Med. Sci. monitor: Inter. Med. J. Exp. Clin. Res. 2005, 11, BR228–BR234. [Google Scholar]
  207. Lu, X.; Wu, F.; Jiang, M.; Sun, X.; Tian, G. Curcumin ameliorates gestational diabetes in mice partly through activating AMPK. Pharm. Biol. 2019, 57, 250–254. [Google Scholar] [CrossRef]
  208. Soetikno, V.; Sari, F.R.; Sukumaran, V.; Lakshmanan, A.P.; Mito, S.; Harima, M.; Thandavarayan, R.A.; Suzuki, K.; Nagata, M.; Takagi, R. Curcumin prevents diabetic cardiomyopathy in streptozotocin-induced diabetic rats: Possible involvement of PKC–MAPK signaling pathway. Eur. J. Pharm. Sci. 2012, 47, 604–614. [Google Scholar] [CrossRef]
  209. Wang, Y.; Zhou, S.; Sun, W.; McClung, K.; Pan, Y.; Liang, G.; Tan, Y.; Zhao, Y.; Liu, Q.; Sun, J. Inhibition of JNK by novel curcumin analog C66 prevents diabetic cardiomyopathy with a preservation of cardiac metallothionein expression. Am. J. Physiol. Endocrinol. Metab. 2014, 306, E1239–E1247. [Google Scholar] [CrossRef]
  210. Song, J.-Q.; Teng, X.; Cai, Y.; Tang, C.-S.; Qi, Y.-F. Activation of Akt/GSK-3β signaling pathway is involved in intermedin1–53 protection against myocardial apoptosis induced by ischemia/reperfusion. Apoptosis 2009, 14, 1299–1307. [Google Scholar] [CrossRef]
  211. Lawson, T.B.; Scott-Drechsel, D.E.; Chivukula, V.K.; Rugonyi, S.; Thornburg, K.L.; Hinds, M.T. Hyperglycemia alters the structure and hemodynamics of the developing embryonic heart. J. Cardiovasc. Dev. Dis. 2018, 5, 13. [Google Scholar] [CrossRef]
  212. Panahi, Y.; Khalili, N.; Sahebi, E.; Namazi, S.; Atkin, S.L.; Majeed, M.; Sahebkar, A. Curcuminoids plus piperine modulate adipokines in type 2 diabetes mellitus. Curr. Clin. Pharmacol. 2017, 12, 253–258. [Google Scholar] [CrossRef] [PubMed]
  213. Chen, H.; Yang, X.; Lu, K.; Lu, C.; Zhao, Y.; Zheng, S.; Li, J.; Huang, Z.; Huang, Y.; Zhang, Y. Inhibition of high glucose—Induced inflammation and fibrosis by a novel curcumin derivative prevents renal and heart injury in diabetic mice. Toxicol. Lett. 2017, 278, 48–58. [Google Scholar] [CrossRef] [PubMed]
  214. Ren, J.; Sowers, J.R. Application of a novel curcumin analog in the management of diabetic cardiomyopathy. Diabetes 2014, 63, 3166–3168. [Google Scholar] [CrossRef] [PubMed]
  215. Aziz, M.T.A.; El Ibrashy, I.N.; Mikhailidis, D.P.; Rezq, A.M.; Wassef, M.A.A.; Fouad, H.H.; Ahmed, H.H.; Sabry, D.A.; Shawky, H.M.; Hussein, R.E. Signaling mechanisms of a water-soluble curcumin derivative in experimental type 1 diabetes with cardiomyopathy. Diabetol. Metab. Syndr. 2013, 5, 13. [Google Scholar] [CrossRef] [PubMed]
  216. Jang, E.-M.; Choi, M.-S.; Jung, U.J.; Kim, M.-J.; Kim, H.-J.; Jeon, S.-M.; Shin, S.-K.; Seong, C.-N.; Lee, M.-K. Beneficial effects of curcumin on hyperlipidemia and insulin resistance in high-fat–fed hamsters. Metabolism 2008, 57, 1576–1583. [Google Scholar] [CrossRef]
  217. Na, L.-X.; Zhang, Y.-L.; Li, Y.; Liu, L.-Y.; Li, R.; Kong, T.; Sun, C.-H. Curcumin improves insulin resistance in skeletal muscle of rats. Diabetes Nutr. Metab. 2011, 21, 526–533. [Google Scholar]
  218. El-Moselhy, M.A.; Taye, A.; Sharkawi, S.S.; El-Sisi, S.F.; Ahmed, A.F. The antihyperglycemic effect of curcumin in high fat diet fed rats. Role of TNF-α and free fatty acids. Food Chem. Toxicol. 2011, 49, 1129–1140. [Google Scholar] [CrossRef]
  219. Karthikesan, K.; Pari, L.; Menon, V. Antihyperlipidemic effect of chlorogenic acid and tetrahydrocurcumin in rats subjected to diabetogenic agents. Chem.-Biol. Interact. 2010, 188, 643–650. [Google Scholar] [CrossRef]
  220. Kaur, G. Amelioration of obesity, glucose intolerance, and oxidative stress in high-fat diet and low-dose streptozotocin-induced diabetic rats by combination consisting of “curcumin with piperine and quercetin”. Int. Sch. Res. Notices. 2012, 2012, 957283. [Google Scholar] [CrossRef]
  221. Yu, W.; Wu, J.; Cai, F.; Xiang, J.; Zha, W.; Fan, D.; Guo, S.; Ming, Z.; Liu, C. Curcumin alleviates diabetic cardiomyopathy in experimental diabetic rats. PLoS ONE 2012, 7, e52013. [Google Scholar] [CrossRef]
  222. Rastogi, M.; Ojha, R.P.; Rajamanickam, G.; Agrawal, A.; Aggarwal, A.; Dubey, G. Curcuminoids modulates oxidative damage and mitochondrial dysfunction in diabetic rat brain. Free. Radical Res. 2008, 42, 999–1005. [Google Scholar] [CrossRef] [PubMed]
  223. Al-Ali, K.; Fatah, H.S.A.; El-Badry, Y.A.-M. Dual effect of curcumin–zinc complex in controlling diabetes mellitus in experimentally induced diabetic rats. Biol. Pharm. Bull. 2016, 39, 1774–1780. [Google Scholar] [CrossRef] [PubMed]
  224. Seo, K.I.; Choi, M.S.; Jung, U.J.; Kim, H.J.; Yeo, J.; Jeon, S.M.; Lee, M.K. Effect of curcumin supplementation on blood glucose, plasma insulin, and glucose homeostasis related enzyme activities in diabetic db/db mice. Mol. Nutr. Food Res. 2008, 52, 995–1004. [Google Scholar] [CrossRef] [PubMed]
  225. Pan, Y.; Wang, Y.; Zhao, Y.; Peng, K.; Li, W.; Wang, Y.; Zhang, J.; Zhou, S.; Liu, Q.; Li, X. Inhibition of JNK phosphorylation by a novel curcumin analog prevents high glucose–induced inflammation and apoptosis in cardiomyocytes and the development of diabetic cardiomyopathy. Diabetes 2014, 63, 3497–3511. [Google Scholar] [CrossRef] [PubMed]
  226. Usharani, P.; Mateen, A.; Naidu, M.; Raju, Y.; Chandra, N. Effect of NCB-02, atorvastatin and placebo on endothelial function, oxidative stress and inflammatory markers in patients with type 2 diabetes mellitus. Drugs R D 2008, 9, 243–250. [Google Scholar] [CrossRef]
  227. Panahi, Y.; Khalili, N.; Sahebi, E.; Namazi, S.; Simental-Mendía, L.E.; Majeed, M.; Sahebkar, A. Effects of curcuminoids plus piperine on glycemic, hepatic and inflammatory biomarkers in patients with type 2 diabetes mellitus: A randomized double-blind placebo-controlled trial. Drug Res. 2018, 68, 403–409. [Google Scholar] [CrossRef]
  228. Neerati, P.; Devde, R.; Gangi, A.K. Evaluation of the effect of curcumin capsules on glyburide therapy in patients with type-2 diabetes mellitus. Phytother. Res. 2014, 28, 1796–1800. [Google Scholar] [CrossRef]
  229. Patil, V.M.; Das, S.; Balasubramanian, K. Quantum chemical and docking insights into bioavailability enhancement of curcumin by piperine in pepper. J. Phys. Chem. A 2016, 120, 3643–3653. [Google Scholar] [CrossRef]
  230. Talirevic, E.; Jelena, S. Quercetin in the treatment of dyslipidemia. Med. Arch. 2012, 66, 87–88. [Google Scholar] [CrossRef]
  231. Pereira, D.F.; Cazarolli, L.H.; Lavado, C.; Mengatto, V.; Figueiredo, M.S.R.B.; Guedes, A.; Pizzolatti, M.G.; Silva, F.R.M.B. Effects of flavonoids on α-glucosidase activity: Potential targets for glucose homeostasis. Nutrition 2011, 27, 1161–1167. [Google Scholar] [CrossRef]
  232. Lin, T.-Y.; Liu, Y.-C.; Jheng, J.-R.; Tsai, H.-P.; Jan, J.-T.; Wong, W.-R.; Horng, J.-T. Anti-enterovirus 71 activity screening of Chinese herbs with anti-infection and inflammation activities. Am. J. Chin. Med. 2009, 37, 143–158. [Google Scholar] [CrossRef]
  233. Oboh, G.; Ademosun, A.O.; Ayeni, P.O.; Omojokun, O.S.; Bello, F. Comparative effect of quercetin and rutin on α-amylase, α-glucosidase, and some pro-oxidant-induced lipid peroxidation in rat pancreas. Comp. Clin. Pathol. 2015, 24, 1103–1110. [Google Scholar] [CrossRef]
  234. Chen, S.; Jiang, H.; Wu, X.; Fang, J. Therapeutic effects of quercetin on inflammation, obesity, and type 2 diabetes. Mediat. Inflamm. 2016, 2016, 9340637. [Google Scholar] [CrossRef]
  235. Zhou, M.; Wang, S.; Zhao, A.; Wang, K.; Fan, Z.; Yang, H.; Liao, W.; Bao, S.; Zhao, L.; Zhang, Y. Transcriptomic and metabonomic profiling reveal synergistic effects of quercetin and resveratrol supplementation in high fat diet fed mice. J. Proteome Res. 2012, 11, 4961–4971. [Google Scholar] [CrossRef] [PubMed]
  236. Yang, D.K.; Kang, H.-S. Anti-diabetic effect of cotreatment with quercetin and resveratrol in streptozotocin-induced diabetic rats. Biomol. Ther. 2018, 26, 130. [Google Scholar] [CrossRef] [PubMed]
  237. Spencer, J.P.; Vauzour, D.; Rendeiro, C. Flavonoids and cognition: The molecular mechanisms underlying their behavioural effects. Arch. Biochem. Biophys. 2009, 492, 1–9. [Google Scholar] [CrossRef]
  238. Ay, M.; Luo, J.; Langley, M.; Jin, H.; Anantharam, V.; Kanthasamy, A.; Kanthasamy, A.G. Molecular mechanisms underlying protective effects of quercetin against mitochondrial dysfunction and progressive dopaminergic neurodegeneration in cell culture and MitoPark transgenic mouse models of Parkinson’s Disease. J. Neurochem. 2017, 141, 766–782. [Google Scholar] [CrossRef]
  239. Vafadar, A.; Shabaninejad, Z.; Movahedpour, A.; Fallahi, F.; Taghavipour, M.; Ghasemi, Y.; Akbari, M.; Shafiee, A.; Hajighadimi, S.; Moradizarmehri, S. Quercetin and cancer: New insights into its therapeutic effects on ovarian cancer cells. Cell. Biosci. 2020, 10, 83. [Google Scholar] [CrossRef]
  240. Dhanya, R.; Arun, K.; Syama, H.; Nisha, P.; Sundaresan, A.; Kumar, T.S.; Jayamurthy, P. Rutin and quercetin enhance glucose uptake in L6 myotubes under oxidative stress induced by tertiary butyl hydrogen peroxide. Food Chem. 2014, 158, 546–554. [Google Scholar] [CrossRef]
  241. Borghi, S.M.; Mizokami, S.S.; Pinho-Ribeiro, F.A.; Fattori, V.; Crespigio, J.; Clemente-Napimoga, J.T.; Napimoga, M.H.; Pitol, D.L.; Issa, J.P.; Fukada, S.Y. The flavonoid quercetin inhibits titanium dioxide (TiO2)-induced chronic arthritis in mice. J. Nutr. Biochem. 2018, 53, 81–95. [Google Scholar] [CrossRef]
  242. Spínola, V.; Llorent-Martínez, E.J.; Castilho, P.C. Inhibition of α-amylase, α-glucosidase and pancreatic lipase by phenolic compounds of Rumex maderensis (Madeira sorrel). Influence of simulated gastrointestinal digestion on hyperglycaemia-related damage linked with aldose reductase activity and protein glycation. Lwt 2020, 118, 108727. [Google Scholar]
  243. Gong, L.; Feng, D.; Wang, T.; Ren, Y.; Liu, Y.; Wang, J. Inhibitors of α-amylase and α-glucosidase: Potential linkage for whole cereal foods on prevention of hyperglycemia. Food Sci. Nutr. 2020, 8, 6320–6337. [Google Scholar] [CrossRef] [PubMed]
  244. Dhanya, R.; Arya, A.; Nisha, P.; Jayamurthy, P. Quercetin, a lead compound against type 2 diabetes ameliorates glucose uptake via AMPK pathway in skeletal muscle cell line. Front. Pharmacol. 2017, 8, 336. [Google Scholar] [CrossRef] [PubMed]
  245. Kulkarni, C.R.; Joglekar, M.M.; Patil, S.B.; Arvindekar, A.U. Antihyperglycemic and antihyperlipidemic effect of Santalum album in streptozotocin induced diabetic rats. Pharma Biol. 2012, 50, 360–365. [Google Scholar] [CrossRef]
  246. Vessal, M.; Hemmati, M.; Vasei, M. Antidiabetic effects of quercetin in streptozocin—Induced diabetic rats. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2003, 135, 357–364. [Google Scholar] [CrossRef]
  247. Yim, S.; Malhotra, A.; Veves, A. Antioxidants and CVD in diabetes: Where do we stand now? Curr. Diabetes Rep. 2007, 7, 8–13. [Google Scholar] [CrossRef]
  248. Bardy, G.; Virsolvy, A.; Quignard, J.F.; Ravier, M.A.; Bertrand, G.; Dalle, S.; Cros, G.; Magous, R.; Richard, S.; Oiry, C. Quercetin induces insulin secretion by direct activation of L-type calcium channels in pancreatic beta cells. Brit. J. Pharmacol. 2013, 169, 1102–1113. [Google Scholar] [CrossRef]
  249. Kittl, M.; Beyreis, M.; Tumurkhuu, M.; Fürst, J.; Helm, K.; Pitschmann, A.; Gaisberger, M.; Glasl, S.; Ritter, M.; Jakab, M. Quercetin stimulates insulin secretion and reduces the viability of rat INS-1 beta-cells. Cell. Physiol. Biochem. 2016, 39, 278–293. [Google Scholar] [CrossRef]
  250. Wang, S.; Yao, J.; Zhou, B.; Yang, J.; Chaudry, M.T.; Wang, M.; Xiao, F.; Li, Y.; Yin, W. Bacteriostatic effect of quercetin as an antibiotic alternative in vivo and its antibacterial mechanism in vitro. J. Food Prot. 2018, 81, 68–78. [Google Scholar] [CrossRef]
  251. Saisho, Y.; Kou, K.; Tanaka, K.; Abe, T.; Kurosawa, H.; Shimada, A.; Meguro, S.; Kawai, T.; Itoh, H. Postprandial serum C—Peptide to plasma glucose ratio as a predictor of subsequent insulin treatment in patients with type 2 diabetes. Endocr. J. 2011, 58, 315–322. [Google Scholar] [CrossRef]
  252. Shetty, A.; Rashmi, R.; Rajan, M.; Sambaiah, K.; Salimath, P. Antidiabetic influence of quercetin in streptozotocin—Induced diabetic rats. Nutr. Res. 2004, 24, 373–381. [Google Scholar] [CrossRef]
  253. Ashraf, J.M.; Shahab, U.; Tabrez, S.; Lee, E.J.; Choi, I.; Ahmad, S. Quercetin as a finer substitute to aminoguanidine in the inhibition of glycation products. Int. J. Biol. Macromol. 2015, 77, 188–192. [Google Scholar] [CrossRef] [PubMed]
  254. Shoelson, S.E.; Lee, J.; Goldfine, A.B. Inflammation and insulin resistance. J. Clin. Investig. 2006, 116, 1793–1801. [Google Scholar] [CrossRef] [PubMed]
  255. Tsalamandris, S.; Antonopoulos, A.S.; Oikonomou, E.; Papamikroulis, G.-A.; Vogiatzi, G.; Papaioannou, S.; Deftereos, S.; Tousoulis, D. The role of inflammation in diabetes: Current concepts and future perspectives. Eur. Cardiol. Rev. 2019, 14, 50. [Google Scholar] [CrossRef]
  256. Tziomalos, K.; Athyros, V.G. Diabetic nephropathy: New risk factors and improvements in diagnosis. Rev. Diabetes Stud. 2015, 12, 110. [Google Scholar] [CrossRef]
  257. Cermak, R.; Landgraf, S.; Wolffram, S. Quercetin glucosides inhibit glucose uptake into brush—Border—Membrane vesicles of porcine jejunum. Br. J. Nutr. 2004, 91, 849–855. [Google Scholar] [CrossRef]
  258. Kwon, O.; Eck, P.; Chen, S.; Corpe, C.P.; Lee, J.H.; Kruhlak, M.; Levine, M. Inhibition of the intestinal glucose transporter GLUT2 by flavonoids. FASEB J. 2007, 21, 366–377. [Google Scholar] [CrossRef]
  259. Yao, Z.; Gu, Y.; Zhang, Q.; Liu, L.; Meng, G.; Wu, H.; Xia, Y.; Bao, X.; Shi, H.; Sun, S. Estimated daily quercetin intake and association with the prevalence of type 2 diabetes mellitus in Chinese adults. Eur. J. Nutr. 2019, 58, 819–830. [Google Scholar] [CrossRef]
  260. Gupta, S.; Burman, S.; Nair, A.B.; Chauhan, S.; Sircar, D.; Roy, P.; Dhanwat, M.; Lahiri, D.; Mehta, D.; Das, R. Brassica oleracea Extracts Prevent Hyperglycemia in Type 2 Diabetes Mellitus. Prev. Nutr. Food Sci. 2022, 27, 50. [Google Scholar] [CrossRef]
  261. Shah, M.A.; Sarker, M.; Gousuddin, M. Antidiabetic potential of Brassica Oleracea Var. Italica in type 2 diabetic sprague dawley (sd) rats. Int. J. Pharmacogn. Phytochem. Res. 2016, 8, 462–469. [Google Scholar]
  262. Anjaneyulu, M.; Chopra, K. Quercetin, an anti-oxidant bioflavonoid, attenuates diabetic nephropathy in rats. Clin. Exp. Pharmacol. Physiol. 2004, 31, 244–248. [Google Scholar] [CrossRef] [PubMed]
  263. Kermani, J.; Goodarzi, N.; Bakhtiari, M. An experimental study to evaluate the protective effects of Solanum lycopersicum seed essential oil on diabetes-induced testicular injuries. Medicina 2019, 55, 499. [Google Scholar] [CrossRef] [PubMed]
  264. Ojo, O.O.; Olorunsogo, O.O. Quercetin and vitamin E attenuate diabetes-induced testicular anomaly in Wistar rats via the mitochondrial-mediated apoptotic pathway. Andrologia 2021, 53, e14185. [Google Scholar] [CrossRef] [PubMed]
  265. Fard, M.H.; Naseh, G.; Lotfi, N.; Hosseini, S.M.; Hosseini, M. Effects of aqueous extract of turnip leaf (Brassica rapa) in alloxan-induced diabetic rats. Avicenna J. Phytomed. 2015, 5, 148. [Google Scholar]
  266. Abdelmoaty, M.A.; Ibrahim, M.; Ahmed, N.; Abdelaziz, M. Confirmatory studies on the antioxidant and antidiabetic effect of quercetin in rats. Indian J. Clin. Biochem. 2010, 25, 188. [Google Scholar] [CrossRef]
  267. Elekofehinti, O.O.; Onunkun, A.T.; Olaleye, T.M. Cymbopogon citratus (DC.) Stapf mitigates ER-stress induced by streptozotocin in rats via down-regulation of GRP78 and up-regulation of Nrf2 signaling. J. Ethnopharmacol. 2020, 262, 113130. [Google Scholar] [CrossRef]
  268. Ahmed, N.Z.; Ibrahim, S.R.; Ahmed-Farid, O.A. Quercetin and Apigenin of Cymbopogon citratus mediate inhibition of HCT-116 and PC-3 cell cycle progression and ameliorate Doxorubicin-induced testicular dysfunction in male rats. Biomed. Res. Ther. 2018, 5, 2466–2479. [Google Scholar] [CrossRef]
  269. Srinivasan, P.; Vijayakumar, S.; Kothandaraman, S.; Palani, M. Anti-diabetic activity of quercetin extracted from Phyllanthus emblica L. fruit: In silico and in vivo approaches. J. Pharm. Anal. 2018, 8, 109–118. [Google Scholar] [CrossRef]
  270. Ansari, P.; Hannon-Fletcher, M.P.; Flatt, P.R.; Abdel-Wahab, Y.H. Effects of 22 traditional anti-diabetic medicinal plants on DPP-IV enzyme activity and glucose homeostasis in high-fat fed obese diabetic rats. Biosci. Rep. 2021, 41, BSR20203824. [Google Scholar] [CrossRef]
  271. Atal, S.; Atal, S.; Vyas, S.; Phadnis, P. Bio—Enhancing effect of piperine with metformin on lowering blood glucose level in alloxan induced diabetic mice. Pharmacogn. Res. 2016, 8, 56. [Google Scholar] [CrossRef]
  272. Oršolić, N.; Gajski, G.; Garaj-Vrhovac, V.; Đikić, D.; Prskalo, Z.Š.; Sirovina, D. DNA-protective effects of quercetin or naringenin in alloxan-induced diabetic mice. Eur. J. Pharmacol. 2011, 656, 110–118. [Google Scholar] [CrossRef] [PubMed]
  273. Rasheed, R.A.; Elshikh, M.S.; Mohamed, M.O.; Darweesh, M.F.; Hussein, D.S.; Almutairi, S.M.; Embaby, A.S. Quercetin mitigates the adverse effects of high fat diet on pancreatic and renal tissues in adult male albino rats. J. King Saud Univ. Sci. 2022, 34, 101946. [Google Scholar] [CrossRef]
  274. Eidi, M.; Eidi, A.; Saeidi, A.; Molanaei, S.; Sadeghipour, A.; Bahar, M.; Bahar, K. Effect of coriander seed (Coriandrum sativum L.) ethanol extract on insulin release from pancreatic beta cells in streptozotocin-induced diabetic rats. Phytother. Res. 2009, 23, 404–406. [Google Scholar] [CrossRef] [PubMed]
  275. Tang, L.; Li, K.; Zhang, Y.; Li, H.; Li, A.; Xu, Y.; Wei, B. Quercetin liposomes ameliorate streptozotocin—Induced diabetic nephropathy in diabetic rats. Sci. Rep. 2020, 10, 2440. [Google Scholar] [CrossRef] [PubMed]
  276. Das, S.; Chaware, S.; Narkar, N.; Tilak, A.V.; Raveendran, S.; Rane, P. Antidiabetic activity of Coriandrum sativum in streptozotocin induced diabetic rats. Int. J. Basic Clin. Pharmacol. 2019, 8, 925–929. [Google Scholar] [CrossRef]
  277. Chadchan, K.S.; Jargar, J.G.; Das, S.N. Anti-diabetic effects of aqueous prickly lettuce (Lactuca scariola Linn.) leaves extract in alloxan-induced male diabetic rats treated with nickel (II). J. Basic Clin. Physiol. Pharmacol. 2016, 27, 49–56. [Google Scholar] [CrossRef]
  278. Ismail, H.; Gillespie, A.L.; Calderwood, D.; Iqbal, H.; Gallagher, C.; Chevallier, O.P.; Elliott, C.T.; Pan, X.; Mirza, B.; Green, B.D. The health promoting bioactivities of Lactuca sativa can be enhanced by genetic modulation of plant secondary metabolites. Metabolites 2019, 9, 97. [Google Scholar] [CrossRef]
  279. Nabi, R.K.; Abdullah, M.A. Effect of Quercetin on the Biochemical Parameters of the Alloxan Induced Diabetes in Male Rats. Bas. J. Vet. Res. 2019, 18, 158–170. [Google Scholar]
  280. Hafizur, R.M.; Kabir, N.; Chishti, S. Asparagus officinalis extract controls blood glucose by improving insulin secretion and β—Cell function in streptozotocin—Induced type 2 diabetic rats. Br. J. Nutr. 2012, 108, 1586–1595. [Google Scholar] [CrossRef]
  281. Xie, J.; Song, W.; Liang, X.; Zhang, Q.; Shi, Y.; Liu, W.; Shi, X. Protective effect of quercetin on streptozotocin—Induced diabetic peripheral neuropathy rats through modulating gut microbiota and reactive oxygen species level. Biomed. Pharmacother. 2020, 127, 110147. [Google Scholar] [CrossRef]
  282. Faienza, M.F.; Corbo, F.; Carocci, A.; Catalano, A.; Clodoveo, M.L.; Grano, M.; Wang, D.Q.-H.; D’Amato, G.; Muraglia, M.; Franchini, C. Novel insights in health-promoting properties of sweet cherries. J. Funct. Foods 2020, 69, 103945. [Google Scholar] [CrossRef] [PubMed]
  283. Mahesh, T.; Menon, V.P. Quercetin allievates oxidative stress in streptozotocin-induced diabetic rats. Phytother. Res. 2004, 18, 123–127. [Google Scholar] [CrossRef] [PubMed]
  284. Xiong, Y.; Zhu, G.-H.; Wang, H.-N.; Hu, Q.; Chen, L.-L.; Guan, X.-Q.; Li, H.-L.; Chen, H.-Z.; Tang, H.; Ge, G.-B. Discovery of naturally occurring inhibitors against SARS-CoV-2 3CLpro from Ginkgo biloba leaves via large-scale screening. Fitoterapia 2021, 152, 104909. [Google Scholar] [CrossRef] [PubMed]
  285. Lu, Q.; Hao, M.; Wu, W.; Zhang, N.; Isaac, A.T.; Yin, J.; Zhu, X.; Du, L.; Yin, X. Antidiabetic cataract effects of GbE, rutin and quercetin are mediated by the inhibition of oxidative stress and polyol pathway. Acta Biochim. Pol. 2018, 65, 35–41. [Google Scholar] [CrossRef] [PubMed]
  286. Iskender, H.; Dokumacioglu, E.; Sen, T.M.; Ince, I.; Kanbay, Y.; Saral, S. The effect of hesperidin and quercetin on oxidative stress, NF-κB and SIRT1 levels in a STZ-induced experimental diabetes model. Biomed. Pharmacother. 2017, 90, 500–508. [Google Scholar] [CrossRef]
  287. Ozougwu, J.C. Anti-diabetic effects of Allium cepa (onions) aqueous extracts on alloxan-induced diabetic Rattus novergicus. J. Med. Plants Res. 2011, 5, 1134–1139. [Google Scholar]
  288. Khaki, A.; Fathi, A.F.; Ahmadi, A.H.; Rezazadeh, S.; Rastegar, H.; Imani, A. Compartments of quercetin & Allium cepa (onion) on blood glucose in diabetic rats. J. Med. Plants 2010, 9, 107–112. [Google Scholar]
  289. Campos, K.; Diniz, Y.; Cataneo, A.; Faine, L.; Alves, M.; Novelli, E. Hypoglycaemic and antioxidant effects of onion, Allium cepa: Dietary onion addition, antioxidant activity and hypoglycaemic effects on diabetic rats. Int. J. Food Sci. Nutr. 2003, 54, 241–246. [Google Scholar] [CrossRef]
  290. Kim, J.J.; Tan, Y.; Xiao, L.; Sun, Y.-L.; Qu, X. Green tea polyphenol epigallocatechin-3-gallate enhance glycogen synthesis and inhibit lipogenesis in hepatocytes. BioMed Res. Int. 2013, 2013, 920128. [Google Scholar] [CrossRef]
  291. Ashida, H.; Furuyashiki, T.; Nagayasu, H.; Bessho, H.; Sakakibara, H.; Hashimoto, T.; Kanazawa, K. Anti-obesity actions of green tea: Possible involvements in modulation of the glucose uptake system and suppression of the adipogenesis-related transcription factors. Biofactors 2004, 22, 135–140. [Google Scholar] [CrossRef]
  292. Li, Y.; Zhao, S.; Zhang, W.; Zhao, P.; He, B.; Wu, N.; Han, P. Epigallocatechin-3-O-gallate (EGCG) attenuates FFAs-induced peripheral insulin resistance through AMPK pathway and insulin signaling pathway in vivo. Diabetes Res. Clin. Pract. 2011, 93, 205–214. [Google Scholar] [CrossRef] [PubMed]
  293. Takagaki, A.; Yoshioka, Y.; Yamashita, Y.; Nagano, T.; Ikeda, M.; Hara-Terawaki, A.; Seto, R.; Ashida, H. Effects of microbial metabolites of (−)-epigallocatechin gallate on glucose uptake in l6 skeletal muscle cell and glucose tolerance in icr mice. Biol. Pharm. Bull. 2019, 42, 212–221. [Google Scholar] [CrossRef] [PubMed]
  294. Ueda-Wakagi, M.; Hayashibara, K.; Nagano, T.; Ikeda, M.; Yuan, S.; Ueda, S.; Shirai, Y.; Yoshida, K.-I.; Ashida, H. Epigallocatechin gallate induces GLUT4 translocation in skeletal muscle through both PI3K-and AMPK-dependent pathways. Food Funct. 2018, 9, 4223–4233. [Google Scholar] [CrossRef] [PubMed]
  295. Ueda, M.; Nishiumi, S.; Nagayasu, H.; Fukuda, I.; Yoshida, K.-i.; Ashida, H. Epigallocatechin gallate promotes GLUT4 translocation in skeletal muscle. Biochem. Biophys. Res. Commun. 2008, 377, 286–290. [Google Scholar] [CrossRef] [PubMed]
  296. Kobayashi, Y.; Suzuki, M.; Satsu, H.; Arai, S.; Hara, Y.; Suzuki, K.; Miyamoto, Y.; Shimizu, M. Green tea polyphenols inhibit the sodium-dependent glucose transporter of intestinal epithelial cells by a competitive mechanism. J. Agric. Food Chem. 2000, 48, 5618–5623. [Google Scholar] [CrossRef]
  297. Shimizu, M.; Kobayashi, Y.; Suzuki, M.; Satsu, H.; Miyamoto, Y. Regulation of intestinal glucose transport by tea catechins. Biofactors 2000, 13, 61–65. [Google Scholar] [CrossRef]
  298. Thielecke, F.; Boschmann, M. The potential role of green tea catechins in the prevention of the metabolic syndrome–a review. Phytochemistry 2009, 70, 11–24. [Google Scholar] [CrossRef]
  299. Park, J.-H.; Bae, J.-H.; Im, S.-S.; Song, D.-K. Green tea and type 2 diabetes. Integr. Med. Res. 2014, 3, 4–10. [Google Scholar] [CrossRef]
  300. Xu, H.; Barnes, G.T.; Yang, Q.; Tan, G.; Yang, D.; Chou, C.J.; Sole, J.; Nichols, A.; Ross, J.S.; Tartaglia, L.A. Chronic inflammation in fat plays a crucial role in the development of obesity—Related insulin resistance. J. Clin. Investig. 2003, 112, 1821–1830. [Google Scholar] [CrossRef]
  301. Li, X.; Li, S.; Chen, M.; Wang, J.; Xie, B.; Sun, Z. (−)-Epigallocatechin-3-gallate (EGCG) inhibits starch digestion and improves glucose homeostasis through direct or indirect activation of PXR/CAR-mediated phase II metabolism in diabetic mice. Food Funct. 2018, 9, 4651–4663. [Google Scholar] [CrossRef]
  302. Li, F.; Gao, C.; Yan, P.; Zhang, M.; Wang, Y.; Hu, Y.; Wu, X.; Wang, X.; Sheng, J. EGCG reduces obesity and white adipose tissue gain partly through AMPK activation in mice. Front. Pharmacol. 2018, 9, 1366. [Google Scholar] [CrossRef] [PubMed]
  303. Kamiyama, O.; Sanae, F.; Ikeda, K.; Higashi, Y.; Minami, Y.; Asano, N.; Adachi, I.; Kato, A. In vitro inhibition of α—Glucosidases and glycogen phosphorylase by catechin gallates in green tea. Food Chem. 2010, 122, 1061–1066. [Google Scholar] [CrossRef]
  304. Konishi, K.; Wada, K.; Yamakawa, M.; Goto, Y.; Mizuta, F.; Koda, S.; Uji, T.; Tsuji, M.; Nagata, C. Dietary soy intake is inversely associated with risk of type 2 diabetes in Japanese women but not in men. J. Nutr. 2019, 149, 1208–1214. [Google Scholar] [CrossRef] [PubMed]
  305. Jin, M.; Shen, M.-H.; Jin, M.-H.; Jin, A.-H.; Yin, X.-Z.; Quan, J.-S. Hypoglycemic property of soy isoflavones from hypocotyl in Goto-Kakizaki diabetic rats. J. Clin. Biochem. Nutr. 2018, 62, 148–154. [Google Scholar] [CrossRef]
  306. Chen, X.; Yu, J.; Shi, J. Management of diabetes mellitus with puerarin, a natural isoflavone from Pueraria lobata. Am. J. Chin. Med. 2018, 46, 1771–1789. [Google Scholar] [CrossRef]
  307. Fu, Z.; Gilbert, E.R.; Pfeiffer, L.; Zhang, Y.; Fu, Y.; Liu, D. Genistein ameliorates hyperglycemia in a mouse model of nongenetic type 2 diabetes. Appl. Physiol. Nutr. Metab. 2012, 37, 480–488. [Google Scholar] [CrossRef]
  308. Rockwood, S.; Mason, D.; Lord, R.; Lamar, P.; Prozialeck, W.; Al-Nakkash, L. Genistein diet improves body weight, serum glucose and triglyceride levels in both male and female ob/ob mice. Diabetes Metab. Syndr. Obes. Targets Ther. 2019, 12, 2011–2021. [Google Scholar] [CrossRef]
  309. 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]
  310. Gupta, S.K.; Dongare, S.; Mathur, R.; Mohanty, I.R.; Srivastava, S.; Mathur, S.; Nag, T.C. Genistein ameliorates cardiac inflammation and oxidative stress in streptozotocin—Induced diabetic cardiomyopathy in rats. Mol. Cell. Biochem. 2015, 408, 63–72. [Google Scholar] [CrossRef]
  311. Das, D.; Sarkar, S.; Bordoloi, J.; Wann, S.B.; Kalita, J.; Manna, P. Daidzein, its effects on impaired glucose and lipid metabolism and vascular inflammation associated with type 2 diabetes. Biofactors 2018, 44, 407–417. [Google Scholar] [CrossRef]
  312. Huang, G.; Xu, J.; Guo, T.L. Isoflavone daidzein regulates immune responses in the B6C3F1 and non—Obese diabetic (NOD) mice. Int. Immunopharmacol. 2019, 71, 277–284. [Google Scholar] [CrossRef] [PubMed]
  313. Prabhakar, P.K.; Prasad, R.; Ali, S.; Doble, M. Synergistic interaction of ferulic acid with commercial hypoglycemic drugs in streptozotocin induced diabetic rats. Phytomedicine 2013, 20, 488–494. [Google Scholar] [CrossRef] [PubMed]
  314. Ohnishi, M.; Matuo, T.; Tsuno, T.; Hosoda, A.; Nomura, E.; Taniguchi, H.; Sasaki, H.; Morishita, H. Antioxidant activity and hypoglycemic effect of ferulic acid in STZ—Induced diabetic mice and KK—A mice. Biofactors 2004, 21, 315–319. [Google Scholar] [CrossRef] [PubMed]
  315. Roy, S.; Metya, S.K.; Sannigrahi, S.; Rahaman, N.; Ahmed, F. Treatment with ferulic acid to rats with streptozotocin-induced diabetes: Effects on oxidative stress, pro-inflammatory cytokines, and apoptosis in the pancreatic β cell. Endocrine 2013, 44, 369–379. [Google Scholar] [CrossRef] [PubMed]
  316. Aaby, K.; Ekeberg, D.; Skrede, G. Characterization of phenolic compounds in strawberry (Fragaria× ananassa) fruits by different HPLC detectors and contribution of individual compounds to total antioxidant capacity. J. Agric. Food Chem. 2007, 55, 4395–4406. [Google Scholar] [CrossRef]
  317. Yogeeta, S.K.; Gnanapragasam, A.; Senthilkumar, S.; Subhashini, R.; Devaki, T. Synergistic salubrious effect of ferulic acid and ascorbic acid on membrane-bound phosphatases and lysosomal hydrolases during experimental myocardial infarction in rats. Life Sci. 2006, 80, 258–263. [Google Scholar] [CrossRef]
  318. Jung, E.H.; Ran Kim, S.; Hwang, I.K.; Youl Ha, T. Hypoglycemic effects of a phenolic acid fraction of rice bran and ferulic acid in C57BL/KsJ-db/db mice. J. Agric. Food Chem. 2007, 55, 9800–9804. [Google Scholar] [CrossRef]
  319. Cherng, Y.-G.; Tsai, C.-C.; Chung, H.-H.; Lai, Y.-W.; Kuo, S.-C.; Cheng, J.-T. Antihyperglycemic action of sinapic acid in diabetic rats. J. Agric. Food Chem. 2013, 61, 12053–12059. [Google Scholar] [CrossRef]
  320. Gandhi, G.R.; Jothi, G.; Antony, P.J.; Balakrishna, K.; Paulraj, M.G.; Ignacimuthu, S.; Stalin, A.; Al-Dhabi, N.A. Gallic acid attenuates high-fat diet fed-streptozotocin-induced insulin resistance via partial agonism of PPARγ in experimental type 2 diabetic rats and enhances glucose uptake through translocation and activation of GLUT4 in PI3K/p—Akt signaling pathway. Eur. J. Pharmacol. 2014, 745, 201–216. [Google Scholar] [CrossRef]
  321. Latha, R.C.R.; Daisy, P. Insulin-secretagogue, antihyperlipidemic and other protective effects of gallic acid isolated from Terminalia bellerica Roxb. in streptozotocin-induced diabetic rats. Chem.-Biol. Interact. 2011, 189, 112–118. [Google Scholar] [CrossRef]
  322. Punithavathi, V.R.; Prince, P.S.M.; Kumar, R.; Selvakumari, J. Antihyperglycaemic, antilipid peroxidative and antioxidant effects of gallic acid on streptozotocin induced diabetic Wistar rats. Eur. J. Pharmacol. 2011, 650, 465–471. [Google Scholar] [CrossRef] [PubMed]
  323. Ma, J.; Luo, X.-D.; Protiva, P.; Yang, H.; Ma, C.; Basile, M.J.; Weinstein, I.B.; Kennelly, E.J. Bioactive novel polyphenols from the fruit of Manilkara zapota (Sapodilla). J. Nat. Prod. 2003, 66, 983–986. [Google Scholar] [CrossRef] [PubMed]
  324. Singh, J.; Rai, G.; Upadhyay, A.; Kumar, R.; Singh, K. Antioxidant phytochemicals in tomato (Lycopersicon esculentum). Indian J. Agric. Sci. 2004, 74, 3–5. [Google Scholar]
  325. Harini, R.; Pugalendi, K.V. Antihyperglycemic effect of protocatechuic acid on streptozotocin—Diabetic rats. J. Basic Clin. Physiol. Pharmacol. 2010, 21, 79–92. [Google Scholar] [CrossRef]
  326. Scazzocchio, B.; Varì, R.; Filesi, C.; D’Archivio, M.; Santangelo, C.; Giovannini, C.; Iacovelli, A.; Silecchia, G.; Volti, G.L.; Galvano, F. 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]
  327. Panchal, S.K.; Ward, L.; Brown, L. Ellagic acid attenuates high-carbohydrate, high-fat diet-induced metabolic syndrome in rats. Eur. J. Nutr. 2013, 52, 559–568. [Google Scholar] [CrossRef]
  328. Nankar, R.P.; Doble, M. Ellagic acid potentiates insulin sensitizing activity of pioglitazone in L6 myotubes. J. Funct. Foods 2015, 15, 1–10. [Google Scholar] [CrossRef]
  329. Cao, Y.; DuBois, D.C.; Almon, R.R.; Jusko, W.J. Pharmacokinetics of salsalate and salicylic acid in normal and diabetic rats. Biopharm. Drug Dispos. 2012, 33, 285–291. [Google Scholar] [CrossRef]
  330. Jung, U.J.; Lee, M.-K.; Park, Y.B.; Jeon, S.-M.; Choi, M.-S. Antihyperglycemic and antioxidant properties of caffeic acid in db/db mice. J. Pharmacol. Exp. Ther. 2006, 318, 476–483. [Google Scholar] [CrossRef]
  331. Chao, C.Y.; Mong, M.C.; Chan, K.C.; Yin, M.C. Anti-glycative and anti-inflammatory effects of caffeic acid and ellagic acid in kidney of diabetic mice. Mol. Nutr. Food Res. 2010, 54, 388–395. [Google Scholar] [CrossRef]
  332. Hsu, F.-L.; Chen, Y.-C.; Cheng, J.-T. Caffeic acid as active principle from the fruit of xanthiumstrumarium to lower plasma glucose in diabetic rats. Planta Med. 2000, 66, 228–230. [Google Scholar] [CrossRef] [PubMed]
  333. Mahmood, T.; Anwar, F.; Abbas, M.; Saari, N. Effect of maturity on phenolics (phenolic acids and flavonoids) profile of strawberry cultivars and mulberry species from Pakistan. Int. J. Mol. Sci. 2012, 13, 4591–4607. [Google Scholar] [CrossRef] [PubMed]
  334. Fuentes, E.; Forero—Doria, O.; Carrasco, G.; Maricán, A.; Santos, L.S.; Alarcón, M.; Palomo, I. Effect of tomato industrial processing on phenolic profile and antiplatelet activity. Molecules 2013, 18, 11526–11536. [Google Scholar] [CrossRef] [PubMed]
  335. Kang, S.-I.; Shin, H.-S.; Kim, H.-M.; Hong, Y.-S.; Yoon, S.-A.; Kang, S.-W.; Kim, J.-H.; Ko, H.-C.; Kim, S.-J. Anti-obesity properties of a Sasa quelpaertensis extract in high-fat diet-induced obese mice. Biosci. Biotechnol. Biochem. 2012, 76, 755–761. [Google Scholar] [CrossRef]
  336. Yoon, S.-A.; Kang, S.-I.; Shin, H.-S.; Kang, S.-W.; Kim, J.-H.; Ko, H.-C.; Kim, S.-J. p-Coumaric acid modulates glucose and lipid metabolism via AMP-activated protein kinase in L6 skeletal muscle cells. Biochem. Biophys. Res. Commun. 2013, 432, 553–557. [Google Scholar] [CrossRef]
  337. Jin, S.; Chang, C.; Zhang, L.; Liu, Y.; Huang, X.; Chen, Z. Chlorogenic acid improves late diabetes through adiponectin receptor signaling pathways in db/db mice. PLoS ONE 2015, 10, e0120842. [Google Scholar] [CrossRef]
  338. McCarty, M.F. A chlorogenic acid—Induced increase in GLP—1 production may mediate the impact of heavy coffee consumption on diabetes risk. Med. Hypotheses 2005, 64, 848–853. [Google Scholar] [CrossRef]
  339. Bassoli, B.K.; Cassolla, P.; Borba-Murad, G.R.; Constantin, J.; Salgueiro-Pagadigorria, C.L.; Bazotte, R.B.; da Silva, R.S.d.S.F.; de Souza, H.M. Chlorogenic acid reduces the plasma glucose peak in the oral glucose tolerance test: Effects on hepatic glucose release and glycemia. Cell Biochem. Funct. 2008, 26, 320–328. [Google Scholar] [CrossRef]
  340. Ong, K.W.; Hsu, A.; Tan, B.K.H. Anti-diabetic and anti-lipidemic effects of chlorogenic acid are mediated by ampk activation. Biochem. Pharmacol. 2013, 85, 1341–1351. [Google Scholar] [CrossRef]
  341. Mei, X.; Zhou, L.; Zhang, T.; Lu, B.; Sheng, Y.; Ji, L. Chlorogenic acid attenuates diabetic retinopathy by reducing VEGF expression and inhibiting VEGF—Mediated retinal neoangiogenesis. Vasc. Pharmacol. 2018, 101, 29–37. [Google Scholar] [CrossRef]
  342. Nyambe-Silavwe, H.; Williamson, G. Chlorogenic and phenolic acids are only very weak inhibitors of human salivary α-amylase and rat intestinal maltase activities. Food Res. Int. 2018, 113, 452–455. [Google Scholar] [CrossRef] [PubMed]
  343. Ishikawa, A.; Yamashita, H.; Hiemori, M.; Inagaki, E.; Kimoto, M.; Okamoto, M.; Tsuji, H.; Memon, A.N.; Mohammadi, A.; Natori, Y. Characterization of inhibitors of postprandial hyperglycemia from the leaves of Nerium indicum. J. Nutr. Sci. Vitaminol. 2007, 53, 166–173. [Google Scholar] [CrossRef] [PubMed]
  344. Zhang, H.; Zhou, Q.; Cao, J.; Wang, Y. Mechanism of cinnamic acid-induced trypsin inhibition: A multi-technique approach. Spectrochim. Acta Part A 2013, 116, 251–257. [Google Scholar] [CrossRef] [PubMed]
  345. Lakshmi, B.S.; Sujatha, S.; Anand, S.; Sangeetha, K.N.; Narayanan, R.B.; Katiyar, C.; Kanaujia, A.; Duggar, R.; Singh, Y.; Srinivas, K. Cinnamic acid, from the bark of Cinnamomum cassia, regulates glucose transport via activation of GLUT4 on L6 myotubes in a phosphatidylinositol 3-kinase-independent manner. J. Diabetes 2009, 1, 99–106. [Google Scholar] [CrossRef]
  346. Hafizur, R.M.; Hameed, A.; Shukrana, M.; Raza, S.A.; Chishti, S.; Kabir, N.; Siddiqui, R.A. Cinnamic acid exerts anti—Diabetic activity by improving glucose tolerance in vivo and by stimulating insulin secretion in vitro. Phytomedicine 2015, 22, 297–300. [Google Scholar] [CrossRef]
  347. Wang, H.; Li, Q.; Deng, W.; Omari-Siaw, E.; Wang, Q.; Wang, S.; Wang, S.; Cao, X.; Xu, X.; Yu, J. Self-nanoemulsifying drug delivery system of trans-cinnamic acid: Formulation development and pharmacodynamic evaluation in alloxan-induced type 2 diabetic rat model. Drugs Dev. Res. 2015, 76, 82–93. [Google Scholar] [CrossRef]
  348. Kopp, C.; Singh, S.P.; Regenhard, P.; Müller, U.; Sauerwein, H.; Mielenz, M. Trans-cinnamic acid increases adiponectin and the phosphorylation of AMP-activated protein kinase through G-protein-coupled receptor signaling in 3T3-L1 adipocytes. Int. J. Mol. Sci. 2014, 15, 2906–2915. [Google Scholar] [CrossRef]
  349. Yan, F.; Zheng, X. Anthocyanin-rich mulberry fruit improves insulin resistance and protects hepatocytes against oxidative stress during hyperglycemia by regulating AMPK/ACC/mTOR pathway. J. Funct. Foods 2017, 30, 270–281. [Google Scholar] [CrossRef]
  350. Shi, M.; Loftus, H.; McAinch, A.J.; Su, X.Q. Blueberry as a source of bioactive compounds for the treatment of obesity, type 2 diabetes and chronic inflammation. J. Funct. Foods 2017, 30, 16–29. [Google Scholar] [CrossRef]
  351. Sun, X.; Du, M.; Navarre, D.A.; Zhu, M.J. Purple potato extract promotes intestinal epithelial differentiation and barrier function by activating AMP-activated protein kinase. Mol. Nutr. Food Res. 2018, 62, 1700536. [Google Scholar] [CrossRef]
  352. Johnson, M.H.; De Mejia, E.G.; Fan, J.; Lila, M.A.; Yousef, G.G. Anthocyanins and proanthocyanidins from blueberry–blackberry fermented beverages inhibit markers of inflammation in macrophages and carbohydrate-utilizing enzymes in vitro. Mol. Nutr. Food Res. 2013, 57, 1182–1197. [Google Scholar] [CrossRef] [PubMed]
  353. Graf, D.; Seifert, S.; Jaudszus, A.; Bub, A.; Watzl, B. Anthocyanin-rich juice lowers serum cholesterol, leptin, and resistin and improves plasma fatty acid composition in fischer rats. PLoS ONE 2013, 8, e66690. [Google Scholar] [CrossRef] [PubMed]
  354. Jiang, T.; Shuai, X.; Li, J.; Yang, N.; Deng, L.; Li, S.; He, Y.; Guo, H.; Li, Y.; He, J. Protein—Bound anthocyanin compounds of purple sweet potato ameliorate hyperglycemia by regulating hepatic glucose metabolism in high—Fat diet/streptozotocin—Induced diabetic mice. J. Agric. Food Chem. 2020, 68, 1596–1608. [Google Scholar] [CrossRef] [PubMed]
  355. Qin, B.; Anderson, R.A. An extract of chokeberry attenuates weight gain and modulates insulin, adipogenic and inflammatory signalling pathways in epididymal adipose tissue of rats fed a fructose-rich diet. Br. J. Nutr. 2012, 108, 581–587. [Google Scholar] [CrossRef]
  356. Wu, T.; Jiang, Z.; Yin, J.; Long, H.; Zheng, X. Anti-obesity effects of artificial planting blueberry (Vaccinium ashei) anthocyanin in high-fat diet-treated mice. Int. J. Food Sci. Nutr. 2016, 67, 257–264. [Google Scholar] [CrossRef]
  357. Qin, Y.; Zhai, Q.; Li, Y.; Cao, M.; Xu, Y.; Zhao, K.; Wang, T. Cyanidin-3-O-glucoside ameliorates diabetic nephropathy through regulation of glutathione pool. Biomed. Pharmacother. 2018, 103, 1223–1230. [Google Scholar] [CrossRef]
  358. Nemes, A.; Homoki, J.R.; Kiss, R.; Hegedűs, C.; Kovács, D.; Peitl, B.; Gál, F.; Stündl, L.; Szilvássy, Z.; Remenyik, J. Effect of anthocyanin—Rich tart cherry extract on inflammatory mediators and adipokines involved in type 2 diabetes in a high fat diet induced obesity mouse model. Nutrients 2019, 11, 1966. [Google Scholar] [CrossRef]
  359. Mussa, B.M.; Srivastava, A.; Al-Habshi, A.; Mohammed, A.K.; Halwani, R.; Abusnana, S. Inflammatory biomarkers levels in T2DM Emirati patients with diabetic neuropathy. Diabetes Metab. Syndr. Obes. Targets Ther. 2021, 14, 3389–3397. [Google Scholar] [CrossRef]
  360. Farrell, N.J.; Norris, G.H.; Ryan, J.; Porter, C.M.; Jiang, C.; Blesso, C.N. Black elderberry extract attenuates inflammation and metabolic dysfunction in diet-induced obese mice. Br. J. Nutr. 2015, 114, 1123–1131. [Google Scholar] [CrossRef]
  361. Tsuda, T.; Ueno, Y.; Aoki, H.; Koda, T.; Horio, F.; Takahashi, N.; Kawada, T.; Osawa, T. Anthocyanin enhances adipocytokine secretion and adipocyte-specific gene expression in isolated rat adipocytes. Biochem. Biophys. Res. Commun. 2004, 316, 149–157. [Google Scholar] [CrossRef]
  362. Takikawa, M.; Inoue, S.; Horio, F.; Tsuda, T. Dietary anthocyanin-rich bilberry extract ameliorates hyperglycemia and insulin sensitivity via activation of AMP-activated protein kinase in diabetic mice. J. Nutr. Biochem. 2010, 140, 527–533. [Google Scholar] [CrossRef] [PubMed]
  363. Kurimoto, Y.; Shibayama, Y.; Inoue, S.; Soga, M.; Takikawa, M.; Ito, C.; Nanba, F.; Yoshida, T.; Yamashita, Y.; Ashida, H. 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]
  364. Choi, K.H.; Lee, H.A.; Park, M.H.; Han, J.-S. Mulberry (Morus alba L.) fruit extract containing anthocyanins improves glycemic control and insulin sensitivity via activation of AMP-activated protein kinase in diabetic C57BL/Ksj-db/db mice. J. Med. Food 2016, 19, 737–745. [Google Scholar] [CrossRef] [PubMed]
  365. Iizuka, Y.; Ozeki, A.; Tani, T.; Tsuda, T. Blackcurrant extract ameliorates hyperglycemia in type 2 diabetic mice in association with increased basal secretion of glucagon-like peptide-1 and activation of AMP-activated protein kinase. J. Nutr. Sci. Vitaminol. 2018, 64, 258–264. [Google Scholar] [CrossRef]
  366. Sasaki, R.; Nishimura, N.; Hoshino, H.; Isa, Y.; Kadowaki, M.; Ichi, T.; Tanaka, A.; Nishiumi, S.; Fukuda, I.; Ashida, H. Cyanidin 3—Glucoside ameliorates hyperglycemia and insulin sensitivity due to downregulation of retinol binding protein 4 expression in diabetic mice. Biochem. Pharmacol. 2007, 74, 1619–1627. [Google Scholar] [CrossRef]
  367. Daveri, E.; Cremonini, E.; Mastaloudis, A.; Hester, S.N.; Wood, S.M.; Waterhouse, A.L.; Anderson, M.; Fraga, C.G.; Oteiza, P.I. Cyanidin and delphinidin modulate inflammation and altered redox signaling improving insulin resistance in high fat—Fed mice. Redox Biol. 2018, 18, 16–24. [Google Scholar] [CrossRef]
  368. Tian, L.; Ning, H.; Shao, W.; Song, Z.; Badakhshi, Y.; Ling, W.; Yang, B.B.; Brubaker, P.L.; Jin, T. Dietary cyanidin-3-glucoside attenuates high-fat-diet–induced body-weight gain and impairment of glucose tolerance in mice via effects on the hepatic hormone FGF21. J. Nutr. 2020, 150, 2101–2111. [Google Scholar] [CrossRef]
  369. Seymour, E.M.; Tanone, I.I.; Urcuyo-Llanes, D.E.; Lewis, S.K.; Kirakosyan, A.; Kondoleon, M.G.; Kaufman, P.B.; Bolling, S.F. Blueberry intake alters skeletal muscle and adipose tissue peroxisome proliferator-activated receptor activity and reduces insulin resistance in obese rats. J. Med. Food 2011, 14, 1511–1518. [Google Scholar] [CrossRef]
  370. Seamon, B.; DeFranco, M.; Thigpen, M. Use of the Xbox Kinect virtual gaming system to improve gait, postural control and cognitive awareness in an individual with Progressive Supranuclear Palsy. Disabil. Rehabil. 2017, 39, 721–726. [Google Scholar] [CrossRef]
  371. Lee, S.; Keirsey, K.I.; Kirkland, R.; Grunewald, Z.I.; Fischer, J.G.; de La Serre, C.B. Blueberry supplementation influences the gut microbiota, inflammation, and insulin resistance in high-fat-diet–fed rats. J. Nutr. 2018, 148, 209–219. [Google Scholar] [CrossRef]
  372. Wu, T.; Yang, L.; Guo, X.; Zhang, M.; Liu, R.; Sui, W. Raspberry anthocyanin consumption prevents diet—Induced obesity by alleviating oxidative stress and modulating hepatic lipid metabolism. Food Funct. 2018, 9, 2112–2120. [Google Scholar] [CrossRef] [PubMed]
  373. Kim, N.-H.; Jegal, J.; Kim, Y.N.; Chung, D.-M.; Heo, J.-D.; Rho, J.-R.; Yang, M.H.; Jeong, E.J. Antiobesity effect of fermented chokeberry extract in high-fat diet-induced obese mice. J. Med. Food 2018, 21, 1113–1119. [Google Scholar] [CrossRef] [PubMed]
  374. Lim, S.-M.; Lee, H.S.; Jung, J.I.; Kim, S.M.; Kim, N.Y.; Seo, T.S.; Bae, J.-S.; Kim, E.J. Cyanidin-3-O-galactoside-enriched Aronia melanocarpa extract attenuates weight gain and adipogenic pathways in high-fat diet-induced obese C57BL/6 mice. Nutrients 2019, 11, 1190. [Google Scholar] [CrossRef]
  375. Song, H.; Shen, X.; Zhou, Y.; Zheng, X. Black rice anthocyanins alleviate hyperlipidemia, liver steatosis and insulin resistance by regulating lipid metabolism and gut microbiota in obese mice. Food Funct. 2021, 12, 10160–10170. [Google Scholar] [CrossRef] [PubMed]
  376. Watanabe, M.; Ayugase, J. Effects of buckwheat sprouts on plasma and hepatic parameters in type 2 diabetic db/db mice. J. Food Sci. 2010, 75, H294–H299. [Google Scholar] [CrossRef] [PubMed]
  377. Chen, Z.; Wang, C.; Pan, Y.; Gao, X.; Chen, H. Hypoglycemic and hypolipidemic effects of anthocyanins extract from black soybean seed coat in high fat diet and streptozotocin-induced diabetic mice. Food Funct. 2018, 9, 426–439. [Google Scholar] [CrossRef]
  378. Herrera—Balandrano, D.D.; Chai, Z.; Hutabarat, R.P.; Beta, T.; Feng, J.; Ma, K.; Li, D.; Huang, W. Hypoglycemic and hypolipidemic effects of blueberry anthocyanins by AMPK activation: In vitro and in vivo studies. Redox Biol. 2021, 46, 102100. [Google Scholar] [CrossRef]
  379. Ye, X.; Chen, W.; Tu, P.; Jia, R.; Liu, Y.; Tang, Q.; Chen, C.; Yang, C.; Zheng, X.; Chu, Q. Antihyperglycemic effect of an anthocyanin, cyanidin-3-O-glucoside, is achieved by regulating GLUT-1 via the Wnt/β-catenin-WISP1 signaling pathway. Food Funct. 2022, 13, 4612–4623. [Google Scholar] [CrossRef]
  380. Guo, H.; Xia, M.; Zou, T.; Ling, W.; Zhong, R.; Zhang, W. Cyanidin 3-glucoside attenuates obesity-associated insulin resistance and hepatic steatosis in high-fat diet-fed and db/db mice via the transcription factor FoxO1. J. Nutr. Biochem. 2012, 23, 349–360. [Google Scholar] [CrossRef]
  381. Zou, W.; Zhang, C.; Gu, X.; Li, X.; Zhu, H. Metformin in combination with malvidin prevents progression of non—Alcoholic fatty liver disease via improving lipid and glucose metabolisms, and inhibiting inflammation in type 2 diabetes rats. Drug Des. Dev. Ther. 2021, 15, 2565–2576. [Google Scholar] [CrossRef]
  382. Kusunoki, M.; Sato, D.; Tsutsumi, K.; Tsutsui, H.; Nakamura, T.; Oshida, Y. Black soybean extract improves lipid profiles in fenofibrate—Treated type 2 diabetics with postprandial hyperlipidemia. J. Med. Food 2015, 18, 615–618. [Google Scholar] [CrossRef] [PubMed]
  383. 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] [PubMed]
  384. Liu, Y.; Li, D.; Zhang, Y.; Sun, R.; Xia, M. Anthocyanin increases adiponectin secretion and protects against diabetes-related endothelial dysfunction. Am. J. Physiol. Endocrinol. Metab. 2014, 306, E975–E988. [Google Scholar] [CrossRef]
  385. 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] [PubMed]
  386. Ye, X.; Chen, W.; Tu, P.; Jia, R.; Liu, Y.; Li, Y.; Tang, Q.; Zheng, X.; Chu, Q. Food-derived cyanidin-3-O-glucoside alleviates oxidative stress: Evidence from the islet cell line and diabetic db/db mice. Food Funct. 2021, 12, 11599–11610. [Google Scholar] [CrossRef]
  387. Lontchi-Yimagou, E.; Sobngwi, E.; Matsha, T.E.; Kengne, A.P. Diabetes mellitus and inflammation. Curr. Diabetes Rep. 2013, 13, 435–444. [Google Scholar] [CrossRef]
  388. Cásedas, G.; Les, F.; Gómez-Serranillos, M.P.; Smith, C.; López, V. Anthocyanin profile, antioxidant activity and enzyme inhibiting properties of blueberry and cranberry juices: A comparative study. Food Funct. 2017, 8, 4187–4193. [Google Scholar] [CrossRef]
  389. Banihani, S.; Makahleh, S.; El-Akawi, Z.; Al-Fashtaki, R.; Khabour, O.; Gharibeh, M.; Saadah, N.; Al-Hashimi, F.; Al-Khasieb, N. Fresh pomegranate juice ameliorates insulin resistance, enhances β-cell function, and decreases fasting serum glucose in type 2 diabetic patients. Nutr. Res. 2014, 34, 862–867. [Google Scholar] [CrossRef]
  390. Alnajjar, M.; Barik, S.K.; Bestwick, C.; Campbell, F.; Cruickshank, M.; Farquharson, F.; Holtrop, G.; Horgan, G.; Louis, P.; Moar, K.-M. Anthocyanin-enriched bilberry extract attenuates glycaemic response in overweight volunteers without changes in insulin. J. Funct. Foods 2020, 64, 103597. [Google Scholar] [CrossRef]
  391. Yang, L.; Ling, W.; Yang, Y.; Chen, Y.; Tian, Z.; Du, Z.; Chen, J.; Xie, Y.; Liu, Z.; Yang, L. Role of purified anthocyanins in improving cardiometabolic risk factors in chinese men and women with prediabetes or early untreated diabetes-A randomized controlled trial. Nutrients 2017, 9, 1104. [Google Scholar] [CrossRef]
  392. 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]
  393. Kianbakht, S.; Abasi, B.; Dabaghian, F.H. Anti—Hyperglycemic effect of Vaccinium arctostaphylos in type 2 diabetic patients: A randomized controlled trial. J. Complement. Med. Res. 2013, 20, 17–22. [Google Scholar] [CrossRef] [PubMed]
  394. Alkhalidy, H.; Moore, W.; Wang, Y.; Luo, J.; McMillan, R.P.; Zhen, W.; Zhou, K.; Liu, D. The flavonoid kaempferol ameliorates streptozotocin-induced diabetes by suppressing hepatic glucose production. Molecules 2018, 23, 2338. [Google Scholar] [CrossRef]
  395. Crespo, I.; Garcia-Mediavilla, M.V.; Gutiérrez, B.; Sánchez-Campos, S.; Tunon, M.J.; González-Gallego, J. A comparison of the effects of kaempferol and quercetin on cytokine-induced pro-inflammatory status of cultured human endothelial cells. Br. J. Nutr. 2008, 100, 968–976. [Google Scholar] [CrossRef] [PubMed]
  396. Torres—Villarreal, D.; Camacho, A.; Castro, H.; Ortiz-Lopez, R.; De la Garza, A. Anti-obesity effects of kaempferol by inhibiting adipogenesis and increasing lipolysis in 3T3-L1 cells. J. Physiol. Biochem. 2019, 75, 83–88. [Google Scholar] [CrossRef]
  397. Martin, B.C.; Warram, J.H.; Krolewski, A.S.; Soeldner, J.; Kahn, C.; Bergman, R. Role of glucose and insulin resistance in development of type 2 diabetes mellitus: Results of a 25-year follow—Up study. Lancet 1992, 340, 925–929. [Google Scholar] [CrossRef]
  398. Al—Numair, K.S.; Chandramohan, G.; Veeramani, C.; Alsaif, M.A. Ameliorative effect of kaempferol, a flavonoid, on oxidative stress in streptozotocin—Induced diabetic rats. Redox Rep. 2015, 20, 198–209. [Google Scholar] [CrossRef]
  399. Montero, M.; Lobatón, C.D.; Hernández-Sanmiguel, E.; Santodomingo, J.; Vay, L.; Moreno, A.; Alvarez, J. Direct activation of the mitochondrial calcium uniporter by natural plant flavonoids. Biochem. J. 2004, 384, 19–24. [Google Scholar] [CrossRef]
  400. Bermont, F.; Hermant, A.; Benninga, R.; Chabert, C.; Jacot, G.; Santo-Domingo, J.; Kraus, M.R.; Feige, J.N.; De Marchi, U. Targeting mitochondrial calcium uptake with the natural flavonol kaempferol, to promote metabolism/secretion coupling in pancreatic β—Cells. Nutrients 2020, 12, 538. [Google Scholar] [CrossRef]
  401. Sharma, D.; Tekade, R.K.; Kalia, K. Kaempferol in ameliorating diabetes—Induced fibrosis and renal damage: An in vitro and in vivo study in diabetic nephropathy mice model. Phytomedicine 2020, 76, 153235. [Google Scholar] [CrossRef]
  402. Fang, X.-K.; Gao, J.; Zhu, D.-N. Kaempferol and quercetin isolated from Euonymus alatus improve glucose uptake of 3T3-L1 cells without adipogenesis activity. Life Sci. 2008, 82, 615–622. [Google Scholar] [CrossRef] [PubMed]
  403. Matschinsky, F.M.; Magnuson, M.A.; Zelent, D.; Jetton, T.L.; Doliba, N.; Han, Y.; Taub, R.; Grimsby, J. The network of glucokinase-expressing cells in glucose homeostasis and the potential of glucokinase activators for diabetes therapy. Diabetes 2006, 55, 1–12. [Google Scholar] [CrossRef] [PubMed]
  404. Haeusler, R.A.; Kaestner, K.H.; Accili, D. FoxOs function synergistically to promote glucose production. J. Biol. Chem. 2010, 285, 35245–35248. [Google Scholar] [CrossRef]
  405. Nakae, J.; Kitamura, T.; Silver, D.L.; Accili, D. The forkhead transcription factor Foxo1 (Fkhr) confers insulin sensitivity onto glucose—6—Phosphatase expression. J. Clin. Investig. 2001, 108, 1359–1367. [Google Scholar] [CrossRef]
  406. Luo, C.; Yang, H.; Tang, C.; Yao, G.; Kong, L.; He, H.; Zhou, Y. Kaempferol alleviates insulin resistance via hepatic IKK/NF—κB signal in type 2 diabetic rats. Int. Immunopharmacol. 2015, 28, 744–750. [Google Scholar] [CrossRef] [PubMed]
  407. Mora, A.; Komander, D.; van Aalten, D.M.; Alessi, D.R. PDK1, the Master Regulator of AGC Kinase Signal Transduction, Seminars in Cell & Developmental Biology; Elsevier: Amsterdam, The Netherlands, 2004; pp. 161–170. [Google Scholar]
  408. Cross, D.A.; Alessi, D.R.; Cohen, P.; Andjelkovich, M.; Hemmings, B.A. Inhibition of glycogen synthase kinase—3 by insulin mediated by protein kinase B. Nature 1995, 378, 785–789. [Google Scholar] [CrossRef] [PubMed]
  409. Donath, M.Y.; Ehses, J.A.; Maedler, K.; Schumann, D.M.; Ellingsgaard, H.; Eppler, E.; Reinecke, M. Mechanisms of β—Cell death in type 2 diabetes. Diabetes 2005, 54 (Suppl. 2), S108–S113. [Google Scholar] [CrossRef]
  410. Sano, Y.; Inamura, K.; Miyake, A.; Mochizuki, S.; Kitada, C.; Yokoi, H.; Nozawa, K.; Okada, H.; Matsushime, H.; Furuichi, K. A novel two-pore domain K+ channel, TRESK, is localized in the spinal cord. J. Biol. Chem. 2003, 278, 27406–27412. [Google Scholar] [CrossRef]
  411. Chen, Y.; Zhang, C.; Jin, M.-N.; Qin, N.; Qiao, W.; Yue, X.-L.; Duan, H.-Q.; Niu, W.-Y. Flavonoid derivative exerts an antidiabetic effect via AMPK activation in diet-induced obesity mice. Nat. Product. Res. 2016, 30, 1988–1992. [Google Scholar] [CrossRef]
  412. Qin, N.; Li, C.-B.; Jin, M.-N.; Shi, L.-H.; Duan, H.-Q.; Niu, W.-Y. Synthesis and biological activity of novel tiliroside derivants. Eur. J. Med. Chem. 2011, 46, 5189–5195. [Google Scholar] [CrossRef]
  413. Saha, A.K.; Avilucea, P.R.; Ye, J.-M.; Assifi, M.M.; Kraegen, E.W.; Ruderman, N.B. Pioglitazone treatment activates AMP-activated protein kinase in rat liver and adipose tissue in vivo. Biochem. Biophys. Res. Commun. 2004, 314, 580–585. [Google Scholar] [CrossRef] [PubMed]
  414. MICROBIOTA, G. Gut microbiota, obesity and metabolic disorders. Minerva Dietol. Gastroenterol. 2017, 63, 337–344. [Google Scholar]
  415. Wang, T.; Wu, Q.; Zhao, T. Preventive effects of kaempferol on high-fat diet-induced obesity complications in C57BL/6 mice. BioMed Res. Int. 2020, 2020, 4532482. [Google Scholar] [CrossRef] [PubMed]
  416. Ashrafizadeh, M.; Tavakol, S.; Ahmadi, Z.; Roomiani, S.; Mohammadinejad, R.; Samarghandian, S. Therapeutic effects of kaempferol affecting autophagy and endoplasmic reticulum stress. Phytother. Res. 2020, 34, 911–923. [Google Scholar] [CrossRef] [PubMed]
  417. Mizushima, N. Autophagy: Process and function. Genes Dev. 2007, 21, 2861–2873. [Google Scholar] [CrossRef] [PubMed]
  418. Codogno, P.; Meijer, A.J. Autophagy: A potential link between obesity and insulin resistance. Cell Metab. 2010, 11, 449–451. [Google Scholar] [CrossRef] [PubMed]
  419. Varshney, R.; Varshney, R.; Mishra, R.; Gupta, S.; Sircar, D.; Roy, P. Kaempferol alleviates palmitic acid-induced lipid stores, endoplasmic reticulum stress and pancreatic β-cell dysfunction through AMPK/mTOR—Mediated lipophagy. J. Nutr. Biochem. 2018, 57, 212–227. [Google Scholar] [CrossRef]
  420. Yaghoobi, Z.; Safahieh, A.; Ronagh, M.T.; Movahedinia, A.; Mousavi, S.M. Hematological changes in yellowfin seabream (Acanthopagrus latus) following chronic exposure to bisphenol A. Comp. Clin. Pathol. 2017, 26, 1305–1313. [Google Scholar] [CrossRef]
  421. Li, H.; Ji, H.-S.; Kang, J.-H.; Shin, D.-H.; Park, H.-Y.; Choi, M.-S.; Lee, C.-H.; Lee, I.-K.; Yun, B.-S.; Jeong, T.-S. Soy leaf extract containing kaempferol glycosides and pheophorbides improves glucose homeostasis by enhancing pancreatic β-cell function and suppressing hepatic lipid accumulation in db/db mice. J. Agric. Food Chem. 2015, 63, 7198–7210. [Google Scholar] [CrossRef]
  422. Al-Numair, K.S.; Veeramani, C.; Alsaif, M.A.; Chandramohan, G. Influence of kaempferol, a flavonoid compound, on membrane-bound ATPases in streptozotocin-induced diabetic rats. Pharm. Biol. 2015, 53, 1372–1378. [Google Scholar] [CrossRef]
  423. López-Lázaro, M.; Calderón-Montaño, J.; Burgos-Morón, E.; Pérez-Guerrero, C. A review on the dietary flavonoid kaempferol. Mini Rev. Med. Chem. 2011, 11, 298–344. [Google Scholar]
  424. Zhang, Y.; Liu, D. Flavonol kaempferol improves chronic hyperglycemia—Impaired pancreatic beta—Cell viability and insulin secretory function. Eur. J. Pharmacol. 2011, 670, 325–332. [Google Scholar] [CrossRef] [PubMed]
  425. Gómez-Zorita, S.; Lasa, A.; Abendaño, N.; Fernández-Quintela, A.; Mosqueda-Solís, A.; Garcia-Sobreviela, M.P.; Arbonés-Mainar, J.M.; Portillo, M.P. Phenolic compounds apigenin, hesperidin and kaempferol reduce in vitro lipid accumulation in human adipocytes. J. Transl. Med. 2017, 15, 237. [Google Scholar] [CrossRef] [PubMed]
  426. Lee, Y.-J.; Choi, H.-S.; Seo, M.-J.; Jeon, H.-J.; Kim, K.-J.; Lee, B.-Y. Kaempferol suppresses lipid accumulation by inhibiting early adipogenesis in 3T3-L1 cells and zebrafish. Food Funct. 2015, 6, 2824–2833. [Google Scholar] [CrossRef]
  427. Ochiai, A.; Othman, M.B.; Sakamoto, K. Kaempferol ameliorates symptoms of metabolic syndrome by improving blood lipid profile and glucose tolerance. Biosci. Biotech. Biochem. 2021, 85, 2169–2176. [Google Scholar] [CrossRef] [PubMed]
  428. Chang, C.J.; Tzeng, T.-F.; Liou, S.-S.; Chang, Y.-S.; Liu, I.-M. Kaempferol regulates the lipid-profile in high-fat diet-fed rats through an increase in hepatic PPARα levels. Planta Med. 2011, 77, 1876–1882. [Google Scholar] [CrossRef] [PubMed]
  429. Hanchang, W.; Khamchan, A.; Wongmanee, N.; Seedadee, C. Hesperidin ameliorates pancreatic β-cell dysfunction and apoptosis in streptozotocin-induced diabetic rat model. Life Sci. 2019, 235, 116858. [Google Scholar] [CrossRef]
  430. Wang, S.-W.; Sheng, H.; Bai, Y.-F.; Weng, Y.-Y.; Fan, X.-Y.; Zheng, F.; Fu, J.-Q.; Zhang, F. Inhibition of histone acetyltransferase by naringenin and hesperetin suppresses Txnip expression and protects pancreatic β cells in diabetic mice. Phytomedicine 2021, 88, 153454. [Google Scholar] [CrossRef]
  431. Pavlovic, D.; Andersen, N.A.; Mandrup-Poulsen, T.; Zizirik, D. Activation of extracellular signal-regulated kinase (ERK) 1/2 contributes to cytokine-induced apoptosis in purified rat pancreatic b—Cells. Eur. Cytokine Netw. 2000, 11, 267–274. [Google Scholar]
  432. Diamanti-Kandarakis, E.; Dunaif, A. Insulin resistance and the polycystic ovary syndrome revisited: An update on mechanisms and implications. Endocr. Rev. 2012, 33, 981–1030. [Google Scholar] [CrossRef]
  433. Catrysse, L.; van Loo, G. Inflammation and the metabolic syndrome: The tissue-specific functions of NF-κB. Trends Cell Biol. 2017, 27, 417–429. [Google Scholar] [CrossRef] [PubMed]
  434. Chen, X.; Wei, W.; Li, Y.; Huang, J.; Ci, X. Hesperetin relieves cisplatin-induced acute kidney injury by mitigating oxidative stress, inflammation and apoptosis. Chem.-Biol. Interact. 2019, 308, 269–278. [Google Scholar] [CrossRef] [PubMed]
  435. Santos—Marcos, J.A.; Perez—Jimenez, F.; Camargo, A. The role of diet and intestinal microbiota in the development of metabolic syndrome. J. Nutr. Biochem. 2019, 70, 1–27. [Google Scholar] [CrossRef]
  436. Crozier, A.; Borges, G.; Ryan, D. The glass that cheers: Phenolic and polyphenolic constituents and the beneficial effects of moderate red wine consumption. Biochemist 2010, 32, 4–9. [Google Scholar] [CrossRef]
  437. Dias, T.R.; Alves, M.G.; Casal, S.; Oliveira, P.F.; Silva, B.M. Promising potential of dietary (poly) phenolic compounds in the prevention and treatment of diabetes mellitus. Curr. Med. Chem. 2017, 24, 334–354. [Google Scholar]
  438. Williamson, G.; Kay, C.; Crozier, A. The bioavailability, transport, and bioactivity of dietary flavonoids: A review from a historical perspective. Compr. Rev. Food. Sci. Food. Saf. 2018, 17, 1054–1112. [Google Scholar] [CrossRef]
  439. Godos, J.; Castellano, S.; Ray, S.; Grosso, G.; Galvano, F. Dietary polyphenol intake and depression: Results from the mediterranean healthy eating, lifestyle and aging (meal) study. Molecules 2018, 23, 999. [Google Scholar] [CrossRef]
  440. Kang, G.G.; Francis, N.; Hill, R.; Waters, D.; Blanchard, C.; Santhakumar, A.B. Dietary polyphenols and gene expression in molecular pathways associated with type 2 diabetes mellitus: A Review. Int. J. Mol. Sci. 2019, 21, 140. [Google Scholar] [CrossRef]
  441. Ghorbani, A.; Rashidi, R.; Shafiee-Nick, R. Flavonoids for preserving pancreatic beta cell survival and function: A mechanistic review. Biomed. Pharmacother. 2019, 111, 947–957. [Google Scholar] [CrossRef]
  442. Jiao, D.; Jiang, Q.; Liu, Y.; Ji, L. Nephroprotective effect of wogonin against cadmium—Induced nephrotoxicity via inhibition of oxidative stress–induced MAPK and NF—kB pathway in Sprague Dawley rats. Hum. Exp. Toxicol. 2019, 38, 1082–1091. [Google Scholar] [CrossRef]
  443. Manzanaro, S.; Salvá, J.; de la Fuente, J.Á. Phenolic marine natural products as aldose reductase inhibitors. J. Nat. Prod. 2006, 69, 1485–1487. [Google Scholar] [CrossRef] [PubMed]
  444. Han, T.; Wu, F.; Lean, M. Obesity and weight management in the elderly: A focus on men. Best Pract. Res. Clin. Endocrinol. Metab. 2013, 27, 509–525. [Google Scholar]
  445. Pedersen, S.D. Metabolic complications of obesity. Best Pract. Res. Clin. Endocrinol. Metab. 2013, 27, 179–193. [Google Scholar]
  446. El-Azab, M.F.; Attia, F.M.; El-Mowafy, A.M. Novel role of curcumin combined with bone marrow transplantation in reversing experimental diabetes: Effects on pancreatic islet regeneration, oxidative stress, and inflammatory cytokines. Eur. J. Pharmacol. 2011, 658, 41–48. [Google Scholar] [CrossRef] [PubMed]
  447. Allam, M.A.; Subhan, N.; Hossain, H.; Hossain, M.; Reza, H.M.; Rahman, M.M.; Ullah, M.O. Hydroxycinnamic acid derivatives: A potential class of natural compounds for the management of lipid metabolism and obesity. Nutr. Metab. 2016, 13, 27. [Google Scholar] [CrossRef]
Figure 1. The increase or decrease in different physiological factors causes hyperglycemia.
Figure 1. The increase or decrease in different physiological factors causes hyperglycemia.
Molecules 28 03996 g001
Figure 2. Regulation of insulin release, functions of insulin and glucagon, and effect of insulin on the healthy; regulations of type 1 and type 2 diabetes.
Figure 2. Regulation of insulin release, functions of insulin and glucagon, and effect of insulin on the healthy; regulations of type 1 and type 2 diabetes.
Molecules 28 03996 g002
Figure 3. Chemical structures of polyphenols.
Figure 3. Chemical structures of polyphenols.
Molecules 28 03996 g003
Table 1. Antidiabetic activity of resveratrol in in vivo studies with its molecular mechanisms.
Table 1. Antidiabetic activity of resveratrol in in vivo studies with its molecular mechanisms.
Resveratrol DoseDurationModalMechanism of ActionRef.
5 mgTwice a day
4 weeks
T2D patientsDecreased insulin resistance[175]
10 mg/day4 weeksRCT double-blind
19 men with T2DM 55 ± 9 years
No changes in insulin levels,
Tendency to decrease HOMA-IR
[175]
50 mgTwice a day
60 days
T2D patientsNo change in insulin resistance
Decreased blood glucose levels
Decreased diabetic ulcer size
[112,176]
75 mg/day12 weeksNonobese women (with normal glucose tolerance)Does not cause any changes in insulin sensitivity, plasma inflammation markers, and systolic blood pressure [177]
100 mg/day 8 weeksRCT parallel-blind
24 subjects with diabetic food
Age: 56 ± 9 years old
Non-significant decrease in glucose in both study groups; no changes in HOMA-IR and insulin[178]
150 mg30 daysObese men Decreased systolic blood pressure, insulin resistance, plasma inflammation markers, and blood glucose levels[179]
150 mg/day30 daysObese men Decrease postprandial glucagon responses [32]
150 mg/day4 weeks16 subjects with T2DM
RCT double-blind cross-over
Non-significant changes in
glucose and insulin levels,
HbA1c level
[180]
200 mg/day 24 weeks110 subjects with T2DM
RCT double-blind
Significant decrease in
glucose and HbA1c (p = 0.005), and significantly reduced insulin and HOMA-IR levels (p = 0.001)
[176]
250 mg/day 3 months57 subjects with T2DM
RCT open-label
Significant decrease in HbA1c (p < 0.05) [181]
250 mg/day 6 months57 subjects with T2DM
RCT open-label
Nonsignificant decrease in HbA1c and glucose levels[182]
250 mg 3 monthsT2DP Decreased blood glucose levels and systolic blood pressures[181]
250 mg per day8 weeksHealthy aged men No changes in metabolic and inflammatory status in skeletal muscle[183]
500 mg/day 3 months60 subjects with T2DM and albuminuria
RCT double-blind
Improvement in HOMA-IR and a significant decrease in insulin, glucose, and HbA1c levels (p < 0.05)[184]
500 mg Twice a day
45 days
T2DPDecreased insulin resistance, blood glucose levels, HOMA-β, and systolic blood pressure [185]
500 mg 3 times a day4 weeksObese men No changes in insulin resistance, plasma inflammation markers, and systolic blood pressure [186]
500 mg 3 times a day90 daysPatients with metabolic syndrome Decreased insulin resistance, but did not cause changes in systolic blood pressure [31]
1 g/day 45 days64 subjects with T2DM
RCT double-blind
Caused a significant decrease in glucose, insulin, and HbA1c levels (p < 0.05), and improvement in HOMA-IR after RV administration[185]
First week 1 g/day
second
week 2 g/day
2 weeksObese men No change in insulin resistance and blood glucose levels
Caused a decrease in the production of intestinal and hepatic lipoprotein
[111]
1, 1.5, 2 g/day4 weeksOlder adults Decreased insulin resistance [110]
3 g/day8 weeksOverweight or obese men with nonalcoholic fatty liver disease and IRNo change in insulin resistance [113]
3 g/day 3 months10 subjects with TD2M
RCT double-blind
Caused a decrease in HbA1c
No significant changes in HOMA-IR
No changes in glucose and insulin levels
[187]
Table 2. Antidiabetic activity of curcumin along with molecular mechanisms.
Table 2. Antidiabetic activity of curcumin along with molecular mechanisms.
Curcumin Dose DurationModelMechanism of ActionReferences
0.01–1 µM24 hStreptozotocin-induced diabetic ratsDecreased TNF-α, IL-6, HbA1c, lipid peroxidation, and MCP-1 secretion [202]
2.5 or 10 Mfor 30 minHigh-glucose-treated H9C2 cardiomyocytesDecreased TNF-a and IL-6 (pro-inflammatory cytokines) and VCAM-1 and ICAM-1 (adhesion molecules) expressions
Inhibited the HG-induced increase in fibrotic genes (collagen-IV, TGF-b, and collagen-I), and decreased AKT phosphorylation
[213]
2.5, 5, or 10 µMonce every two days for 12 weeksPrimary cultures of neonatal rat cardiomyocytesDecreased JNK phosphorylation[214]
0.75%8 weeksdb/db miceDecreased PPAR-γ via AMPK activation and decreased lipid peroxidation[203]
10 mg/kg/day42 daysSTZ-induced diabetic C57BL/6 miceSuppressed hyperglycemia-induced inflammation, hypertrophy, and fibrosis, and decreased TNF-α and ICAM-1[213]
20 mg/kg45 daysStreptozotocin-induced rats fed with a high-cholesterol diet (HCD) Decreased glycemia and dyslipidemia[215]
30–90 mg/kg31 daysStreptozotocin-induced diabetic ratsAnti-hyperglycemic and anti-hyperlipidemic effect
Decreased blood glucose and lipid levels, and lowered levels of hepatic antioxidants
[193,194]
0.05 g/100 g diet10 weeksStreptozotocin-induced rats fed with a high-cholesterol diet (HCD)Decreased glycemia and dyslipidemia[216]
50, 150, or 250 mg/kg7 weeksStreptozotocin-induced rats fed with a high-cholesterol diet (HCD)Decreased glycemia and dyslipidemia[217]
80 mg/kg60–75 daysStreptozotocin-induced rats fed with a high-cholesterol diet (HCD)Decreased glycemia and dyslipidemia[218]
80 mg/kg45 daysSTZ-induced diabetic ratsDecreased blood glucose
Decrease antioxidant defenses
[219]
100 mg/kg28 daysStreptozotocin-induced rats fed with a high-cholesterol diet (HCD)Decreased glycemia and dyslipidemia[220]
100 or 200 mg/kg/day8 weeksSTZ-induced diabetic Wistar ratsDecreased inflammatory factors (TNF-α and IL-1β)
Activated AKT/GSK-3β signaling pathway
[221]
120 mg/kg 1 monthDiabetic male ratsDecreased glucose level and mitochondrial dysfunction
Increased antioxidant defense
[222]
150 mg/kg,45 daysDiabetic male ratsDecreased blood glucose and HbA1c
Increased plasma insulin, AST, and ALT
[223]
0.2 g/kg6 weeksDiabetic db/db miceDecreased SREBP1c, ChREBP, CPT1, and ACAT[224]
200 mg/kg/day6 weeksSTZ-induced diabetic Wistar ratsInhibited IL-6 and TNF-α levels [205]
200 mg/kg 16 weeksStreptozotocin-induced diabetic ratsDecreased Bcl-2
Increased Bax and caspase-3
[221]
250 mg/day 9 months240 prediabetic subjects
n = 120 placebo group
n = 120 curcuminoid group
0% T2DM incidence in the treated group vs. 16.4% incidence in the placebo group
Increased HOMA-β and adiponectin levels
Decreased HOMA-IR (insulin resistance)
Decreased C-peptide level
Improved β-cells function
[33,225]
300 mg 8 weeks67 T2DM patients:
n = 21 placebo group
n = 22 atorvastatin group
n = 23 NCB-02 group
Improved the endothelial function
Decreased malondialdehyde, endothelin-1, IL-6, and NF-α
[226]
500 mg/day
plus 5 mg/day for
3 months100 T2DM patients: n = 50 in the placebo group
n = 50 in the curcuminoids group
Decreased blood glucose level, C-peptide, HbA1c, alanine aminotransferase, and aspartate aminotransferase[227]
475 mg10 days8 T2DM patients treated with glyburide (5 mg)Decreased LDL, VLDL, and triglycerides
Increased HDL
Improved glycemic control (lower blood glucose levels after breakfast, lunch, and dinner)
[228]
1000 mg/day + 10 mg/day12 weeks100 T2DM patients:
n = 50 placebo group
n = 50 curcuminoids group
Decreased leptin and TNF-α
Decrease leptin/adiponectin ratio
Decreased adiponectin
[212]
300 mg/day3 months100 overweight/obese T2DM patients, n = 50 placebo group and n = 50 in the curcuminoid groupDecreased fasting glycemia
Decreased HOMA-IR (insulin resistance)
Decreased HbA1c
Increased lipoprotein lipase activity
Decreased FFA and triglycerides
[34,229]
Table 3. Antidiabetic activity of quercetin with its molecular mechanisms.
Table 3. Antidiabetic activity of quercetin with its molecular mechanisms.
Quercetin Dose DurationModelMechanism of ActionReferences
10 mg/kg4 weeksSTZ-induced diabetic ratsDecreased blood glucose and increased insulin secretion
Decreased blood glucose levels
Decreased creatinine and blood urea nitrogen levels
[260,261,262]
10 mg/kg28 daysSTZ-induced diabetic ratsIncreased insulin secretion
Decreased blood glucose levels
inhibited apoptosis
[263,264]
15 mg/kg25 daysSTZ-induced
diabetic rats
Decreased blood glucose levels and
Improved glucose tolerance
[265,266]
20–50 mg/kg 6 weeksSTZ-induced
diabetic rats
Decreased inflammation
Reduced blood glucose levels
Decreased fasting blood glucose
Decreased hypertension
Increased insulin secretion
Decreased ROS production
[267,268]
25–75 mg/kg28 daysSTZ-induced diabetic ratsIncreased insulin secretion and decreased blood glucose[269]
50 mg/kg30 days Alloxan-induced diabetic ratsInhibited α-glucosidase activity and reduced oxidative stress[270]
50 mg/kg 7 daysAlloxan-induced diabetic miceDecreased blood glucose
Increased insulin secretion
Decreased inflammation
[271,272]
50 mg/kg 12 weeksHFF obese ratsReduced oxidative stress [270,273]
50 mg/kg 8 weeksSTZ-induced diabetic ratsDecreased blood glucose
Decreased fasting blood glucose
Decreased inflammation
Suppressed IL-1β, TNF-α, and
production of AGEs
Increased insulin secretion
[274,275,276]
50 mg/kg 4 weeksAlloxan-induced diabetic ratsLowered blood glucose levels
Decreased inflammation
Decreased fasting blood glucose
Increased insulin secretion
Decreased creatinine, AST, ALT, and cholesterol levels
[277,278,279]
50 mg/kg12 weeksSTZ-induced
diabetic rats
Decreased the production of reactive oxygen species (ROS) and
improved glucose tolerance
[280,281]
50–80 mg/kg45 days STZ-induced diabetic ratsReduced blood glucose levels
Improved oxidative stress
Decreased LDL and VLDL cholesterol
Decreased blood glucose
Increased insulin secretion
[282,283]
90 mg/kg 10 weeksSTZ-induced diabetic ratsDecreased oxidative stress
Decreased lipid peroxidation
Reduced AGE product activity
[284,285]
100 mg/kg14 daysSTZ-induced diabetic ratsIncreased insulin secretion
Decreased fasting blood glucose
Decreased blood glucose
[286]
100–200 mg/kg6 weeks STZ-induced diabetic ratsImproved glucose tolerance
Decreased blood glucose
Increased insulin secretion
Increased HDL cholesterol
Decreased triglycerides, VLDL, LDL, and total cholesterol
[287,288,289]
1 g/kg 1 monthSTZ-induced diabetic Wistar ratsImproved insulin secretion insulin and increased glucose uptake
Decreased fasting blood sugar
[252]
Table 4. Antidiabetic activity of anthocyanins and their molecular mechanisms.
Table 4. Antidiabetic activity of anthocyanins and their molecular mechanisms.
Anthocyanins Dose DurationModel Mechanism of ActionReferences
320 mg/day4 weeks T2D patientsDecreased FBG, LDL-cholesterol, IL-6, IL-18, and TNF-a
Increased IL-10 and adiponectin (anti-inflammatory markers)
[38]
160 mg24 weeksT2D patientsIncreased antioxidant capacity and
decreased insulin resistance
[385]
1.5 mL/kg After 12 h of fasting condition T2D patientsDecreased FBG level, improved insulin resistance and β-cell functions[389,390]
0.47 g 3 weeksT2D patientsDecreased postprandial glycemia[385]
320 mg/day12 weeks160 pre-diabetics, double-blindCaused moderate reductions of LDL-c, HbA1c, apo A1, and apo B[391]
150, 300, or 600 mg/day4 weeks23 healthy subjects, double-blindDecreased glucose in the blood and hindered the secretion of insulin and incretins.[392]
1050 mg/day whortleberry
extract (9 mg anthocyanins)
2 months (every week 3 days)37 T2D, double-blindDecreased blood glucose levels and HbA1c[393]
Table 5. Antidiabetic activities of kaempferol, along with molecular mechanisms.
Table 5. Antidiabetic activities of kaempferol, along with molecular mechanisms.
Kaempferol DoseDurationModelMechanism of ActionReferences
0.01, 0.1, 1, and 10 µM4 daysHuman islet (CMRL-1066) cellsDecreased apoptosis and increased pancreatic β-cells[424]
1, 10, and 25 µMTreated on days 3, 8, and 12, and observed after 48 h of the last treatmentHuman mesenchymal stem cells (hMSCs)Decrease adipogenesis and
Increased lipolysis
[425]
5, 10, and 20 µM15 daysZebrafishDecreased triglyceride synthase[426]
5 mg/kg
15 mg/kg
6 weeksMale TSOD and TSNO miceDecreased lipid synthesis, decreased fatty acid oxidation, and increased liver cholesterol transport[427]
50 mg/kg12 weeksMale C57BL/6J miceDecreased hepatic gluconeogenesis, increased glycogen synthesis, and decreased blood glucose[22]
75, 150, or 300 mg/kg8 weeksMale Wistar ratsIncreased fatty acid oxidation[428]
100 mg/kg45 daysMale Wistar ratsIncreased membrane-bound ATPases, and increased antioxidants[398]
200 mg/kg8 weeksC57BL/6 mice Decreased blood glucose and insulin resistance
Regulated intestinal flora
[415]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Naz, R.; Saqib, F.; Awadallah, S.; Wahid, M.; Latif, M.F.; Iqbal, I.; Mubarak, M.S. Food Polyphenols and Type II Diabetes Mellitus: Pharmacology and Mechanisms. Molecules 2023, 28, 3996. https://doi.org/10.3390/molecules28103996

AMA Style

Naz R, Saqib F, Awadallah S, Wahid M, Latif MF, Iqbal I, Mubarak MS. Food Polyphenols and Type II Diabetes Mellitus: Pharmacology and Mechanisms. Molecules. 2023; 28(10):3996. https://doi.org/10.3390/molecules28103996

Chicago/Turabian Style

Naz, Rabia, Fatima Saqib, Samir Awadallah, Muqeet Wahid, Muhammad Farhaj Latif, Iram Iqbal, and Mohammad S. Mubarak. 2023. "Food Polyphenols and Type II Diabetes Mellitus: Pharmacology and Mechanisms" Molecules 28, no. 10: 3996. https://doi.org/10.3390/molecules28103996

Article Metrics

Back to TopTop