Next Article in Journal
Impact of Different Drying Methods on the Microbiota, Volatilome, Color, and Sensory Traits of Sea Fennel (Crithmum maritimum L.) Leaves
Next Article in Special Issue
Therapeutic Potential of Curcumin, a Bioactive Compound of Turmeric, in Prevention of Streptozotocin-Induced Diabetes through the Modulation of Oxidative Stress and Inflammation
Previous Article in Journal
Synthesis and Antiproliferative Activity of 2,6-Disubstituted Imidazo[4,5-b]pyridines Prepared by Suzuki Cross Coupling
Previous Article in Special Issue
An In Vivo Assessment of the Effect of Hexane Extract from Endlicheria paniculata Branches and Its Main Compound, Methyldehydrodieugenol B, on Murine Sponge-Induced Inflammation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Potential Benefits of Antioxidant Phytochemicals in Type 2 Diabetes

by
Arman Arabshomali
1,
Shadi Bazzazzadehgan
1,
Fakhri Mahdi
2 and
Zia Shariat-Madar
2,*
1
Department of Pharmacy Administration, School of Pharmacy, University of Mississippi, University, MS 38677, USA
2
Department of BioMolecular Sciences, Division of Pharmacology, School of Pharmacy, University of Mississippi, University, MS 38677, USA
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(20), 7209; https://doi.org/10.3390/molecules28207209
Submission received: 22 September 2023 / Revised: 18 October 2023 / Accepted: 19 October 2023 / Published: 21 October 2023

Abstract

:
The clinical relationship between diabetes and inflammation is well established. Evidence clearly indicates that disrupting oxidant-antioxidant equilibrium and elevated lipid peroxidation could be a potential mechanism for chronic kidney disease associated with type 2 diabetes mellitus (T2DM). Under diabetic conditions, hyperglycemia, especially inflammation, and increased reactive oxygen species generation are bidirectionally associated. Inflammation, oxidative stress, and tissue damage are believed to play a role in the development of diabetes. Although the exact mechanism underlying oxidative stress and its impact on diabetes progression remains uncertain, the hyperglycemia-inflammation-oxidative stress interaction clearly plays a significant role in the onset and progression of vascular disease, kidney disease, hepatic injury, and pancreas damage and, therefore, holds promise as a therapeutic target. Evidence strongly indicates that the use of multiple antidiabetic medications fails to achieve the normal range for glycated hemoglobin targets, signifying treatment-resistant diabetes. Antioxidants with polyphenols are considered useful as adjuvant therapy for their potential anti-inflammatory effect and antioxidant activity. We aimed to analyze the current major points reported in preclinical, in vivo, and clinical studies of antioxidants in the prevention or treatment of inflammation in T2DM. Then, we will share our speculative vision for future diabetes clinical trials.

1. Introduction

Type 2 diabetes mellitus (T2DM) is a common hormonal disorder in the adult population, presenting in 11.3% of the US population and remaining undiagnosed in approximately 23% of adults, according to the Center for Disease Control (CDC) National Diabetes Statistics Report for 2020 cases of diabetes [1]. T2DM is caused by insulin deficiency or insulin resistance. While the incidence of T2DM in youth is increasing [2], it is rare in the pediatric population. Although most of the identified prediabetes are incidental findings without the need for intervention, some may present as clinically significant because they have a higher than normal blood sugar level or cause symptoms from medication or co-morbidities, including hypertension, hyperlipidemia, obesity, or nephropathy.
The etiology of T2DM is diverse and appears to be due to genetic [3,4] causes and environmental factors [5]. In certain cases, however, germline, including insulin association with cancer [6,7] or somatic genetic defects [8], are associated with the development of T2DM. Risk alleles at numerous loci, such as MTNR1B, CGKAL1, and IGF2BP2, have been investigated as potentially involved in the pathogenesis, presentation of T2DM, and beta-cell function [9]. T2DM appears to be linked to dyslipidemia, atherosclerotic heart disease, and cancer [10]. Additional environmental or stochastic events in T2DM, like DNA-methylation, maternal environment [11], the impact of intrauterine [12], early infancy, and the composition and metabolic function of the gut microbiota in obese individuals [13], have also been investigated. Thus, the pathogenesis that underlies T2DM is complex and multifactorial. One of the most important components is the role of glucose autoactivation following persistently elevated circulating glucose, resulting in prooxidant production.
Oxidative stress is caused by a perturbation of prooxidants (reactive oxygen species (ROS)) and the antioxidant microecosystem that favors excess production of prooxidants relative to antioxidant defense [14]. The ROS are formed by multiple overlapping and interacting mechanisms that highlight their biological complexity and their effects on individuals’ genetic backgrounds. These enzymatic and non-enzymatic pathways include mainly oxidative phosphorylation, plasma membrane proteins such as nicotinamide adenine dinucleotide phosphate (NADPH) oxidase (NOXs), lipid metabolism within the peroxisomes, and cyclooxygenases. The abnormality of these two pathways leads to oxidative stress, which accelerates the development of diabetes complications, both microvascular and cardiovascular (Figure 1). Readers are invited to read recent reviews on this issue [15,16].
Thus, we aimed to evaluate the best evidence available on the effects of antioxidants in T2DM in order to clarify their benefits in a systemic review of the principal areas of research. PubMed, Scopus, and clinicaltrials.gov were searched using terms appropriate to each area of research. Randomized controlled trials evaluating the effect of any antioxidant on diabetes, inflammation, and antioxidant status were included. A key element of our search was to identify works that examined the impact of antioxidants, specifically in relation to oxidative stress, damage, or injury, and their capacity to scavenge ROS. Studies that reported alterations in glycemic parameters, anti-inflammatory biomarkers, oxidative stress biomarkers, and antioxidant enzyme levels were collected.

2. Diabetes

Diabetes is the most common chronic disease of childhood [17,18,19] and adulthood [20], and the most common type of diabetes is T2DM. While its exact cause is unknown, T2DM patients have elevated blood glucose levels, dyslipidemia, and impaired insulin receptor function [21]. There is a positive correlation between the progressive nature of T2DM and an increased risk of major cardiovascular diseases, including coronary artery disease, ischemic heart disease, stroke, and peripheral artery disease. Diabetic patients with silent cardiac left ventricular hypertrophy or renal disease are at a higher risk of developing both microvascular and macrovascular diseases [22], leading to organ and tissue damage. Diabetes-induced vascular aberration includes anatomical, structural, and functional alteration resulting in endothelial dysfunction, reduced vascular compliance, and atherosclerosis [23].
As mentioned earlier, the cause of T2DM is unknown. But the diagnosis is made on the basis of a spectrum of clinical manifestations such as impaired regulation of hepatic glucose production, incremental decline in β-cell function, destruction of pancreatic β-cells in a small subset of type 2 diabetic patients, insulin resistance (also known as impaired insulin sensitivity), genetic mutations, dyslipidemia, or hormonal diseases. Blood glucose levels change due to diet, medications, and diabetes, a chronic heterogeneous metabolic disorder with complex pathogenesis. The above-leading clinical features are initially characterized by monitoring blood glucose to identify glucose fluctuation patterns in response to diet. A fasting plasma glucose of 126 mg/dL or more on more than one occasion is found to be associated with an increased risk of microvascular complications [24]. Elevated glucose undergoes autoactivation, which promotes the accumulation of advanced glycation end (AGE) products (Figure 2). Formed AGE products generate free radicals, including peroxides, superoxides, hydroxyl radicals, and singlet oxygen, as described in the succeeding sections. Notably, both diabetic microvascular and macrovascular complications apparently have similar etiologic characteristics.
Insulin regulates the blood glucose levels at several key peripheral organs, highlighting a complex process. Insulin plays a major role in energy metabolism by promoting glucose uptake via glucose transporter 4 (GLUT4) translocation in skeletal muscle and adipose tissue. This increases hepatic glycogen storage and enhances lipogenesis. Evidence suggests that elevated oxidative stress blocks not only insulin secretion but also GLUT4 translocation, which leads to glucose toxicity. AGE-induced free radical production and elevated oxidative stress-induced reduction in insulin secretion and GLUT4 translocation could cause a futile cycle because it involves the continuous accumulation of glucose in the blood, which is also known as hyperglycemia (Figure 3).
Circulating levels of free fatty acids are increased in individuals exhibiting obesity, non-alcoholic fatty liver disease, insulin resistance, or T2DM. While both growth hormone and the sympathetic nervous system promote circulating levels of free fatty acid [25], insulin reduces free fatty acid levels through suppressing lipolysis and enhancing free fatty acid clearance [26]. Chronically elevated plasma free fatty acid levels may trigger the initiation of diabetic vascular complications via increased production of oxygen-derived free radicals and the exacerbation of dyslipidemia.
Oxidative stress plays an important role in the development of diabetes. Under physiological conditions, oxidants are neutralized by enzymatic and nonenzymatic antioxidants. Enzymatic antioxidants superoxide dismutase (SOD), catalase (CAT), and glutathione peroxidase (GPx) provide a first line of defense against free radicals to protect the integrity of biomolecules and tissues [27]. This highlights the importance of these enzymes’ expression in neutralizing excessive oxygen-derived free radicals and maintaining ROS homeostasis. While elevated levels of fatty acids cause an increase in oxidative stress through NADPH oxidative activation, the expression of the antioxidant enzymes is reduced. In addition, excessive free fatty acid liberation alters endothelium-dependent vasodilation by affecting the synthesis or degradation of nitric oxide in diabetes [28]. Thus, diabetic dyslipidemia contributes to reduced antioxidant enzyme expression and reduced nitric oxide bioavailability.
In summary, diabetes is basically a metabolic disorder characterized by hyperglycemia, elevated free fatty acids, and insulin resistance. Each of these major hallmarks of T2DM provokes molecular mechanisms that contribute to microvascular or macrovascular dysfunction. These mechanisms include decreased bioavailability of nitric oxide, increased oxidative stress, decreased antioxidants, and activation of receptors for AGEs. The prevailing model for T2DM states that diabetes is a major risk factor for cardiovascular diseases (CVD) and is responsible for widespread morbidity and mortality. A large number of physiologic and genetic studies support this model and the central role of vascular endothelial cell metabolism [29,30,31,32], which affects the metabolism of other cell types, contributing to the pathogenesis of diabetes.
ROS are formed as natural byproducts of cellular metabolism [33] and are tightly controlled through well-coordinated cellular mechanisms. Increased ROS formation can range from low to severely high levels. At low levels, ROS act as regulators of certain signaling pathways [34] in both cell biology and physiology. At high levels, ROS cause an array of diseases, including cardiovascular, neurodegenerative, metabolic disorders, cancer, inflammation, and aging (Figure 1). ROS are involved in the degradation of pathogens and serve as intracellular signaling molecules in steady state, indicating their protective roles in infection and normal biological processes. However, their prolonged effects result in abnormal cellular physiology, which exacerbates ROS production, leading to progressive oxidative damage and the subsequent initiation of cell death. Moreover, oxidative stress-induced cell death appears to be the most proximal event of diabetes, leading to diabetic micro- and macro-angiopathic complications. Diverse components of the oxidative stress microenvironment affect cell metabolism and cell-cell homeostasis [16]. Furthermore, damage from these stressors can include mutations in mitochondrial DNA (mDNA) in patients with diabetes mellitus [35], somatic point mutations in age-related diseases [36], and immune-related somatic point mutation genes in diabetes [37].
The literature overflows with studies involving antioxidants. The majority of scientists in the field of experimental therapeutics for diabetes have observed that many preclinical antioxidants that are claimed to be potent and selective are often not the case due to off-target effects in general. In rare cases, the effects of antioxidants on cell function are obscure, indicating a false-positive association. In some cases, the antioxidant activity of crude extract depends on a mechanism involved in the disease process that is unelucidated. Since a close link exists among T2DM, CVD, and oxidant-induced injury, this review focuses on the recent research conducted on the effects of antioxidants on the classical biomarkers of diabetes and the components of metabolic syndrome, such as insulin resistance, atherogenic dyslipidemia, hypertension, and proinflammatory and prothrombotic states, as well as on the current biomarkers of oxidative stress, which may be useful for the treatment of diabetes.

2.1. Chronic Inflammation as a Cause of Diabetes

The role of AGE/RAGE signaling in diabetes and their influence on proinflammatory signaling pathways in hyperglycemic conditions have been elaborated on in more detail elsewhere [38,39] (Figure 3). C-reactive protein (CRP) and interleukin-6 (IL-6) are significantly elevated in patients with diabetes, hypertension, and the metabolic syndrome. Those with diabetes have increased systemic levels of CRP, a protein in the blood. IL-6, a protein, is produced by numerous cells and is elevated with the severity of diabetes. Although tumor necrosis factor (TNF) alone induces the induction of CRP expression [40], the release of IL-6 in response to tissue damage is the main activator of CRP expression, which occurs in the hepatocytes. Notably, IL-1 synergistically enhances the effects of IL-6-induced CRP expression [40]. While IL-6 promotes differentiation of B-cells, it is capable of promoting cell growth in certain cells and inhibiting growth in others. This indicates that cytokines, particularly IL-6, have multiple roles to play within the body under both physiological and pathophysiological conditions (Figure 3). However, it is still unclear whether IL-6 is causing or contributing to diabetes. Both IL-6 and CRP independently predict the risk of T2DM. Interestingly, combined elevation levels of IL-6 and IL-1β in the blood are found to be the most significant predictive features for the development of T2DM [41]. Although the precise mechanism by which the proinflammatory state emerges in patients with diabetes remains to be determined, persistent hyperglycemia activates nuclear factor kappa B (NF-κB), which triggers the expression of various cytokines [42]. NF-kB not only has a prooxidant role by inducing the expression of genes such as NADPH oxidase NOX2 gp91phox [43], but it also has numerous anti-oxidant targets, such as manganese superoxide dismutase [44] and NF-kB-induced ferritin heavy chain upregulation via preventing iron-mediated generation of high reactive hydroxyl radicals from H2O2 [45,46]. The activation of NF-κB is mediated via ROS production [42]. Cytokines also activate NF-κB and promote the recruitment of monocytes producing M1 and M2 that cause β-cell destruction and insulin resistance [38]. The role of inflammation in diabetes pathophysiology has been reviewed in detail [47].
Regardless of the triggers, low-grade inflammation and vascular diseases are believed to be the principal causes of disability and premature deaths in diabetes mellitus. Dendritic cells are critical regulators of both immune and inflammatory responses. Dendritic cells sense cellular debris and modified metabolites, which are found in diabetes [48,49]. Once again, the double-edged effect of ROS raises significant concerns. Dendritic NOX2 produces low levels of ROS, which are important in promoting dendritic cell maturation and cross-presentation [50]. Damaged tissues containing dysfunctional mitochondria produce non-physiological levels of ROS. Patients with T2DM exhibit increased ROS production [31], which results in microvascular and macrovascular complications. Evidence indicates that elevated ROS inhibit cross-presentation after entering dendritic cells [51] (Figure 3). Inadequate function of dendritic cells in diabetes may feedback ROS and foster the generation of cytokines, which can contribute to the enhanced generation of ROS. Like other cells, dendritic cells are vulnerable to increased ROS levels.
Diabetes and its clinical manifestations are characterized by chronically elevated levels of glucose and lipid, which lead to glucolipotoxic conditions. Of note, glucolipotoxicity alters mitochondrial function and, in particular, electron transport chain activity. Impaired electron transport chain activity leads to reduced and inefficient fatty acid β-oxidation [52] and promotes an increase in ROS production, triggering inflammation in highly metabolic organs including the kidneys, liver, heart, and insulin-secreting β-cells (Figure 3). The major regulators of fatty acid oxidation in these tissues with high oxidative rates are peroxisome proliferator-activated receptor a (PPARa) [53], a regulator of intra- and extracellular lipid metabolism, and AMP-activated kinase (AMPK), a sensor of cellular energy change [54]. Since both PPAR and AMPK are linked to metabolic disorders, modulation of these two molecules provides therapeutic targets for diseases such as diabetes, dyslipidemia, and inflammation.
The current therapeutic approaches for the treatment of metabolic disease and the potential of immune-modulatory approaches have been reviewed in detail elsewhere [55]. Briefly, the oxidation of low-density lipoprotein (LDL) particles in the endothelial wall of arteries instigates monocyte infiltration [56,57]. Monocytes are then differentiated into macrophages, which accumulate oxidized lipids to form foam cells. Remarkably, oxidative stress augments intimal LDL retention and accumulation [57]. Formed foam cells stimulate macrophage proliferation and promote the recruitment of T-lymphocytes, which induce smooth muscle proliferation in the arterial walls (Figure 3). The sustained inflammatory response continues to recruit mononuclear leukocytes, both monocytes and lymphocytes. The overactive immune system and inflammatory processes [58] are driving forces in a wide variety of vascular physiological mechanisms such as endothelial activation, the loss of the endothelial monolayer due to thrombus formation and neointimal thickening [59], increased oxidative stress [57], reduced endothelial trophic effect, and disrupted blood vessel function [57,60]. It has become apparent that oxidative stress and inflammation are closely linked events, and persistent oxidative stress and chronic inflammation are now known to play predominant roles in diabetes mellitus and atherosclerosis.

2.2. Understanding the Relationship between Hypertension and Type 2 Diabetes Mellitus

Statistics from the CDC and the National Health and Nutritional Examination Survey (NHANES) database show that the majority of individuals with diabetes have hypertension [61]. Dysregulation of the renin-angiotensin-aldosterone system has not only been implicated in the pathophysiology of hypertension, but it also favors T2DM [62] (Figure 3). The inhibitors of the renin-angiotensin-aldosterone system prevent diabetes in patients with hypertension. Diabetes-induced dyslipidemia and hypertension play a major role in the initiation and progression of macrovascular disease. There is a substantial overlap between diabetes and hypertension in disease mechanisms such as upregulation of the renin-angiotensin-aldosterone system and inflammation [63,64]. Should we consider diabetes and hypertension as chronic inflammatory diseases?
The dogma suggests that inflammation is a symptom of many chronic diseases, such as diabetes, hypertension, arthritis, or Alzheimer’s disease. However, systemic reviews and meta-analyses indicate unwaveringly that chronic inflammation increases the risk of various age-related diseases, such as diabetes mellitus, hypertension, kidney disease, and rheumatoid arthritis [65,66,67,68].
Although it is not well clear what causes age-associated chronic inflammation, it appears that cellular senescence is a major endogenous risk factor causing chronic inflammation. Senescent cells are characterized by an arrest of cell proliferation associated with changes in cell morphology, physiology, and chromatin organization, which leads to the development of a unique phenotype known as the senescence-associated secretory phenotype [69]. An endogenous contributor to this phenotype is oxidative stress, which will be reviewed in the following section. Senescent cells are capable of secreting pro-inflammatory cytokines, chemokines, and other pro-inflammatory molecules [70]. This production leads to enhanced infiltration of immune cells and activation of other inflammatory systems, such as nitric oxide and prostaglandin. In diabetes, hyperglycemic spikes affect cytokine concentrations [71].
An increasing body of evidence suggests that diabetes-induced ROS production or excessive generation of ROS [72], independent of diabetes, plays a significant role in the pathogenesis of hypertension (Figure 1) [64]. Sustainedly elevated ROS levels contribute to both hypertension and diabetes. Hypertensive patients exhibit a substantial increase in plasma H2O2 levels [73]. Studies revealed the presence of decreased endogenous antioxidant enzyme activity in peripheral mononuclear cells in hypertensive patients [74] as well as in animal models of hypertension, including angiotensin II-induced hypertension [75] and nephron-vascular hypertension [76].
NOXs and mitochondria produce ROS under physiological and pathological conditions, as described in the following section. Angiotensin II, a vasoactive peptide, not only controls blood pressure but is also implicated in vascular ROS production. It stimulates NADH/NADPH oxidase and activates kinases, including Rho/Rho kinase [77], protein kinase C (PKC) [78], and mitogen-activated protein kinase (MAPK) [79]. Thus, ROS-induced vasoconstriction contributes to the pathogenesis of hypertension. The interplay between ROS imbalance and hypertension has been described and reviewed in detail elsewhere [80].
Nitric oxide (NO), a major vasodilator, is catalyzed by nitric oxide synthase. Angiotensin II and NO have opposite effects on one another [81]. The bioavailability of NO is highly dependent on its redox status [82]. The two reactive nitrogen species (RNS), namely peroxynitrite and S-nitrosoglutathione, are derived from the NO radical. Like ROS, RNS is prone to uncontrolled overproduction under stressful conditions, leading to cellular nitro-oxidative damage. The generation of ROS takes place in different compartments than that of NO generation, suggesting compartmentalization in which these molecules are tightly regulated.
In addition, other health conditions that have a key role in the development of hypertension are kidney disease, glomerulonephritis, and atherosclerosis. Remarkably, both diabetes mellitus and hypertension are responsible for chronic kidney disease [83] and adversely alter the lumen of small arteries and arterioles. Diabetic patients manifest accelerated premature vascular aging characterized by enhanced vascular smooth muscle contraction and increased vascular stiffness [84]. These structural and functional vascular changes contribute to the development of hypertension. These multiple interactive mechanisms perpetuate diabetes over time and may even cause a vicious cycle. Thus, effective management of diabetes should include a multifaceted approach to ensure adequate control of blood pressure, lipids with appropriate glycemic control [85], ROS production, RNS production, and inflammation.

3. The Landscape of ROS

ROS are constantly produced and inactivated by antioxidant enzymes such as catalase, superoxide dismutase, and glutathione peroxidase in biological systems. Oxidative stress has been defined as an imbalance between oxidants and antioxidants and, more recently, as a disruption of redox signaling and control [86]. It is generally accepted that oxidative stress has a fundamental role in vascular dysfunction [86]. In diabetic retinopathy, oxidative stress-induced decreased nitric oxide bioavailability contributes to endothelial cell dysfunction, basement membrane thickening, and pericyte apoptosis, resulting in altered microvasculature [87]. Lipid peroxidation is one of the major mechanisms of oxidative damage [88]. There is considerable evidence that hyperglycemia represents the main cause of complications of DM, and oxidative stress resulting from increased generation of ROS plays a crucial role in their pathogenesis [88]. The disruption of antioxidant equilibrium and elevated lipid peroxidation mechanisms for vascular damage associated with T2DM have been extensively studied.

3.1. Sources of Free Radicals in the Cells

ROS and RNS are free radicals [89] derived from both exogenous and cellular sources. The history, chemistry, and biochemistry of both ROS and RNS are reviewed in detail elsewhere [90]. ROS are reactive molecules that are derived from oxygen and are typically involved in a biological process as free radicals. These free radicals include superoxide (O2•−), hydroxyl radical (OH), or non-radicals (H2O2). Like ROS, RNS are products of normal cellular metabolism. However, RNS are derived from nitrogen and can be free radicals (peroxynitrite (ONOO-) or nitric oxide (NO)). While hydroxyl radical is the most reactive species that is responsible for the induction of cytotoxicity, both nitric oxide and hydrogen peroxide are less reactive. Overproduction of hydroxyl radicals and peroxynitrite can lead to impaired normal physiological redox-regulated functions and subsequent unregulated cell signaling, followed by apoptosis via lipid peroxidation, thus disrupting the integrity of the cell membrane [91] and altering lipoprotein structures.
In diabetes, although free radicals are mainly produced in mitochondria, they are also produced by other cellular organs such as the peroxisome and endoplasmic reticulum [92]. Deciphering the molecular mechanisms through which ROS directly interact with pivotal signaling molecules and the physiology of cellular ROS by mitochondria and their signaling molecules in cell proliferation and survival are of utmost importance to understand how elevated levels of ROS may result in irreparable cell damage. ROS-modulating approaches have been used to lower the excessive production of mitochondrial ROS, which is considered to be a central mechanism for the development of diabetes complications. Readers are invited to read recent reviews on this issue [15,93].

3.1.1. Mitochondria

The biological significance of mitochondria, with special emphasis on the ROS, has already been summarized [94]. Compared to other cellular organelles, mitochondrial functioning is determined by mitochondrial DNA and the genome in the nucleus. Mitochondria have now emerged as multifaceted centers responsible for the site of energy production, the formation of ROS, and, notably, as an intracellular sink for ROS [95]. Needless to say, the mechanism by which the mitochondria perform a sink function remains unresolved. Growing evidence indicates mitochondria-related mechanisms in the pathologic process. Mitochondrial defects are also found in pathological studies of chronic diseases and major neurodegenerative diseases. The homeostasis of mitochondria is disrupted in several diseases, including atherosclerosis, renal diseases, autoimmune disorders, pancreatitis, diabetes, and chronic inflammation. In these diseases, dysfunction of mitochondrial complex I could potentially favor overproduction of ROS [96] beyond mitochondrial antioxidant capacity. The mitochondrial defects found in diabetes include fragmentation [97], fusion and fission [98], morphological changes [99], elevated mutation rates in mitochondrial DNA [100], changes in permeability of mitochondrial membranes [101], fluctuations in redox potential [102], the buildup of mutant proteins [103], and impaired oxidative phosphorylation [104]. Mitochondrial dysfunction and oxidative stress have been implicated in the pathogenesis of diabetes [97]. While autophagy is important for maintaining a balanced cell life, dysregulated or reduced autophagy [105,106,107] has been linked to diabetes-induced abnormal fuel metabolism. Based on these findings, a series of therapeutic methods targeting mitochondrial proteins and processes have been developed from bench to bedside. Mitochondria plays a pivotal role in ROS metabolism. However, mitochondrial dysfunction is not considered a hallmark of T2DM, diabetic nephropathy, or diabetic neuropathy.

3.1.2. Peroxisome

Peroxisomes serve as a sensing mechanism to sense the accumulated cellular free fatty acids and activate the signaling pathway and coupling of free fatty acid degradation with that of ROS [108]. While both mitochondria and peroxisomes are involved in hydrogen peroxide formation, hydrogen peroxide is produced by leaks of electrons from donor redox centers of the mitochondrial electron transport chain [109]. Unlike mitochondria, peroxisomes transfer electrons from various metabolites to oxygen, leading to the production of H2O2 within their respiratory pathway. Since this pathway is not coupled to oxidative phosphorylation to produce ATP, its free energy is released in the form of heat. In addition, peroxisomes are an important source of O2•−, OH, and NO via various ROS-generating enzymes [110], which play a key role in signaling pathways and are responsible for regulating essential processes in the cells. These free radicals are produced in a variety of metabolic pathways, such as fatty acid oxidation, nucleic acid, and polyamine catabolism. Of note, the discovery of ROS-metabolizing enzymes in peroxisomes has been reported [110], highlighting the importance of peroxisomes in scavenging ROS in cells.
As mentioned earlier, diabetes causes lipid accumulation. An array of isoforms of PPAR-α, PPAR-γ, and PPAR-σ play an essential role in the regulation of lipid and glucose metabolism. The PPAR-γ isoform is expressed ubiquitously in all tissues [111]. This isoform is the molecular target of a class of insulin-sensitizing drugs known as thiazolidinediones [112]. The PPAR-γ isotypes are capable of reducing inflammatory responses, which are intimately connected to ROS generation in addition to insulin resistance and atherosclerosis. Like mitochondria, peroxisomal ROS/RNS is not considered a hallmark of T2DM. However, it appears that there is a positive correlation between dysfunctional peroxisomes and the elevation of cellular ROS levels [113] and dysregulated lipid metabolism [114]. Since the removal of damaged peroxisomes is not currently possible, antioxidants have been used to maintain a correct redox balance in the cell. The structure and function of peroxisomes and peroxisomal enzymes involved in the production of different ROS are reviewed elsewhere [110].

3.1.3. Endoplasmic Reticulum

Hyperglycemia causes endoplasmic reticulum (ER) stress, which has a key role in a number of pathophysiological processes, including endothelial dysfunction, apoptosis in renal cells [115], and pancreatic β-cell failure [116]. The communication networks between ER stress and oxidative stress are by no means unidirectional, as previously described [117].
ER not only regulates the folding and post-translational maturation of a third of all proteins [118] and most secreted proteins, but it also plays a critical role in lipid biosynthesis, detoxification, and reduction-oxidation balance. Numerous endoplasmic reticulum-localized enzymes, including cytochrome p-450 and b5 enzymes and diamine oxidase, contribute to the formation of ROS [119], suggesting the ER has a major influence on maintaining cellular integrity and viability. Another important thiol oxidase enzyme of the oxidative protein-folding cycle, endoplasmic reticulum oxidoreductin 1 (ERO1), catalyzes the transfer of electrons from dithiols to molecular oxygen, resulting in the formation of H2O2 [120]. During the protein folding process, activation of NOX 4, NADPH-P450 reductase, and glutathione appears to have a causal role in the formation of ROS [121]. The enzymes of ER are responsible for improving protein folding and coping with ongoing oxidative stress. However, the accumulation of unfolded or misfolded proteins within the lumen of the ER causes the induction of ER stress, which impairs normal cellular function via excessive buildup of oxidant H2O2. Antidiabetic drugs are used as the first line of defense to improve glucose metabolism. Antioxidants, as the second line of defense, that scavenge the active radicals have been considered as an additional molecular mechanism for clinical T2DM treatment.

4. Visiting the Most Recent Antioxidants of Medicinal Plants and Their Therapeutic Potential and Mechanisms in Diabetes

Many plants that possess phenolic compounds are found to have an important protective role in scavenging free radicals during oxidative stress [122]. These studies assess the antioxidant properties of polyphenolic compounds in both in vitro and in vivo models for studying diabetes, as summarized in Figure 4. Here, we provide a brief review of the most frequently used antioxidants to assess their beneficial pharmacological effects on the animal body and humans.

4.1. Flavonoids

Polyphenols are a diverse group of natural compounds that fall into four distinct classes: flavonoids, phenolic acids, stilbenes, and lignans. The effects of the most common polyphenols on the biomarkers of oxidative stress are summarized in Table 1. Flavonoids are the most abundant polyphenols found in an optimal human diet. This class of compounds is further grouped into flavones, flavonols, flavan-3ols, isoflavones, anthocyanidins, and flavanones [123]. Phenolic acids are divided into two sub-groups called hydroxyl benzoic and hydroxyl cinnamic acids. The anti-inflammatory [124] and antioxidant properties of flavonoids [125] have been reviewed (Table 1 and Table 2). Though the mechanisms underlying the benefits of flavonoids are complex and remain incompletely understood, flavonoids have been shown to reduce glycemia and related complications [126,127]. In contrast, others think the bioavailability and effectiveness of flavonoids are relatively low [123]. So, it was suggested that nanoparticle systems should be used to prolong circulation and flavonoid efficacy and potentially reduce their non-specific bindings.
Seaweed and tropical papaya have been utilized as traditional remedies and passed down through generations. Based on animal research, both papaya leaves and seaweed exhibit beneficial effects in diabetes treatment. They protect against diabetes-induced β cell damage, reduce fasting plasma glucose levels, decrease A1C, increase the expression of antioxidative enzymes, and reduce ROS production [128]. Both seaweed and papaya also contain a package of antioxidants (vitamins A, C, and E complexes) and other substances such as polysaccharides, phenolic compounds, crucial fatty acids, saponins, fucoidans, and phlorotannins found naturally in other fruits and vegetables. It is unlikely that flavonoids alone can accomplish all diabetes outcomes.
Linarin, a natural flavonoid compound, is capable of counteracting oxidative stress and exerting an anti-inflammatory effect in diabetic mice [129]. Inhibition of aldo-keto reductase (AKR)1B, an NADP(H)-oxidoreductase, by linarin appears to serve as a mechanism for reducing oxidative stress and inflammation in a high-glucose and high-palmitic acid-induced hepatocyte injury model and a T2DM rat model [130].
Isorhamnetin, a methylated derivative of quercetin, belongs to the flavonoid group of phenolic compounds [131]. Isorhamnetin appears to hold great promise against diabetes via aldose reductase inhibition [132]. The important role that isorhamnetin plays in lowering glucose concentrations, improving oxidative status, reducing inflammation, and adjusting lipid metabolism in both in vitro and in vivo models is reviewed [133], suggesting that isorhamnetin may be a useful compound for the treatment of diabetes.
Amomum tsao-ko Crevost et Lemarié (A. tsao-ko), a very common dietary spice, is rich in flavonoids [134]. A. tsao-ko methanol extracts appear to exert remarkable antioxidant and antidiabetic effects in both in vitro and in vivo studies. Nevertheless, further investigations are needed to confirm the detailed mechanisms underlying A. tsao-ko-mediated control of diabetes and oxidative stress.
With the importance of apigenin in numerous physiological functions, there has been great interest in its strong antioxidant and anti-inflammatory actions [135]. Its action is mediated through neutralizing superoxide, singlet oxygen, and hydroxyl radicals, enhancing the function of the PPARγ signaling, and suppressing CD38 [136,137,138].
Licochalcone A, a flavonoid derived from licorice, has been shown to have potential preventative and therapeutic effects on diabetic nephropathy in both in vitro and in vivo studies [139,140]. The antidiabetic effects of the most recent common polyphenols on diabetes are summarized in Table 3.
Myricitrin, a flavone isolated from the bark of Myrica esculenta, has demonstrated a substantial reduction in blood glucose levels in type 2 diabetic mouse and rat models [141]. It facilitates glucose absorption by skeletal muscles through the activation of IRS-1/PI3K/Akt/GLUT4 signaling, as evidenced by both in vitro and in vivo studies. Additionally, myricitrin is shown to mitigate oxidative stress by scavenging and neutralizing oxidative radicals and enhancing the body’s natural oxidative defense through nuclear factor erythroid 2-related factor 2 (Nrf-2) activation in both laboratory and live organism studies [142,143].
Biochanin A is a well-known isoflavone for its anti-inflammatory, antihyperlipidemic [144], antioxidant [145], and anti-cancer health benefits [146]. Biochanin A appears to have an anti-diabetic effect beyond its function as an antioxidant. In animal models of T2DM, it causes increased insulin sensitivity [146], reduced glucose tolerance [146], and a reduction in glycohemoglobin A1C formation. Biochanin A maintains steady blood glucose concentrations within a normal range in a T2DM rat model [147]. A meta-analysis shows that increased transforming growth factor-β (TGF-β) levels were associated with a high risk of nephropathy [148]. All 4 subtypes of protease-activated receptors (PAR1-4) are mainly expressed in renal epithelial, endothelial, and podocyte cells [149]. PAR-2 inhibition improves autophagy and prevents fibrosis and inflammation [149]. Surprisingly, biochanin A reduces diabetic nephropathy via suppression of TGF-β1 and PAR-2 gene expression [144,145].
Formononetin is an isoflavone from the group of phytoestrogens that induces cell apoptosis via the intrinsic apoptosis pathway, which causes the permeabilization of the mitochondrial outer membrane [150,151]. Formononetin has a wide range of biological activities, including eliciting antioxidant [152] (Table 1) and antidiabetic properties in vitro and in vivo (Table 3). Sirtuin 1 (SIRT1), a nicotinamide adenine dinucleotide (NAD)-dependent histone deacetylase, protects cells from ROS. In addition, SIRT1 regulates hepatic lipid metabolism by increasing AMP-activated protein kinase, leading to the inhibition of hepatic lipogenic pathways in favor of fatty acid oxidation [153]. Formononetin reduces hyperglycemia by increasing SIRT1 expression in pancreatic cells [154]. Sirtuin-induced fatty acid oxidation is important because it lowers cytoplasmic lipid accumulation. However, elevated fatty acid oxidation could interfere with glucose metabolism in the muscle [155]. In addition, formononetin is considered a strong apoptotic inducer [151]. Several other mechanisms of action for formononetin have been described [152,156,157]. It is suggested to be used as adjunct therapy for diabetic neuropathy and nephropathy. Clearly, further research and investigation are needed in order to gain a better understanding of how precisely formononetin improves multiple aspects of metabolic syndrome, including diabetes.
Hesperetin, a derivative of hesperidin, is a bioflavonoid compound found in citrus fruits [158]. Like formononetin, hesperetin has been shown to improve diabetes (Table 3) by regulating SIRT1, alleviating inflammation [158], combating oxidative stress [159], and reducing insulin resistance [159].
Naringenin, a flavonoid compound found in propolis, exhibits potent anti-hyperglycemic and anti-hyperlipidemic properties in diabetic rat models [160]. The actions of naringenin include improving hyperglycemia, insulinemia, insulin sensitivity, pancreatic cell performance, and lipid profile [137,160].
Table 1. A summary of the effects of polyphenols in oxidative stress [152,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188].
Table 1. A summary of the effects of polyphenols in oxidative stress [152,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188].
PolyphenolsROSSODCatalaseNOLipid PeroxidaseGSHRNSDPPHRef.
Phenolic acids
 Hexane extract of Eryngium carlinae NE NE [162]
 Cinnamic acid [163,164,165,166]
 p-coumaric acid [167]
 Caffeic acid [168]
 Ferulic acid [169]
 Sinapic acid [170]
 Gentisic acid [171]
 Vanillic acid NE [171]
 Gallic acid [172]
 Syringic acid [173]
 Protocatechuic acid NE [171]
Flavonoids
1. 
van-3-ols
  Catechin (pumpkin pulp extract) [174]
  Epicatechin
  Epigallocatechin [175]
2. 
Isoflavones
[176]
  Biochanin A [177]
  Formononetin [153]
3. 
Flavonones
[153]
  Hesperetin
  Naringenin [179]
4. 
Flavonols
WE,↓[180,181]
  Quercetin
  Quercetin + Naringenin WE[180]
  Kaempferol [180]
  Galangin [182]
  Fisetin [183]
  Myricetin [184]
5. 
Flavononol
[185,186]
  Taxifolin
6. 
Flavanones
[187]
  Eriodictyol [188]
Stilbenes
  Resveratrol




[187]
Phytoestrogen
  Lignans


[188]
↑—represents a significantly elevated level; ↓—refers to a significantly reduced level; NE—no effect; WE—denotes weak effect.
Like other flavonoids, kaempferol exerts both antioxidant (Table 1) and anti-inflammatory effects (Table 2). Kaempferol produces anticancer effects through inhibition of epidermal growth factor receptor (EGFR)-dependent Src proto-oncogene, nonreceptor trosin kinase (SRC), ERK1/2, and AKT serine/threonine kinase (AKT) pathways [189], and cleavage of poly(ADP-ribose) polymerase (PARP) [190], an endogenous substrate of caspase [191]. Thus, kaempferol is an antiproliferative, anti-metastatic, and apoptotic drug. Kaempferol exerts an antidiabetic effect (Table 3) through targeting multiple pathways, including improving glycolysis, glucose uptake, glycogen synthesis, AMPK activity, and GLUT4 expression [192]. Additionally, several in vivo studies reported the antidiabetic effect of kaempferol through decreasing plasma glucose level [193], increasing plasma insulin level [193], decreasing glucose synthesis [194], and increasing glucagon-like peptide 1 (GLP-1) and insulin release [195]. Readers are invited to a recent review on all the pharmacological mechanisms of kaempferol in diabetes [196].
While galangin, a flavonoid, exerts antioxidant (Table 1) and anti-inflammatory activity [197,198], it elicits potent antitumor activity in diverse cancers [199]. It is a potent inhibitor of dipeptidyl peptidase-4 (DPP-4) [200], an integral membrane protein expressed in cells. DPP-4 cleaves a large number of bioactive molecules. However, its major physiological substrate is the incretin hormone GLP-1, which is responsible for the maintenance of normal glucose homeostasis [201]. Thus, galangin is an anti-hyperglycemic agent. Although the mechanism of action is not yet completely understood, recent evidence suggests that galangin improves the lipid profile and plasma insulin level [202].
Fisetin, a bioactive flavonol molecule, has antiproliferative [203], apoptotic [204], and antioxidant [205] activities (Table 1). Fisetin shows an antidiabetic effect, which is mediated by two different mechanisms [206]. Firstly, fisetin inhibits gluconeogenesis by inhibiting the transport of pyruvate into the mitochondria and reducing the cytosolic NADH/NAD(+) potential redox [206]. Secondly, fisetin inhibits glycogen breakdown, leading to a reduction in blood glucose levels [207,208]. Fisetin represents a promising therapeutic strategy that may synergize with other antidiabetic therapies.
Myricetin, a flavonoid, is widely distributed in different types of fruits, herbs, and tea [209]. Recent studies have reported myricetin’s mechanisms of action in diabetes, such as inhibiting DPP4 [210], being an inactivator of GLP-1, or serving as a GLP-1 receptor agonist [211]. Surprisingly, myricetin appears to normalize the intestinal flora of type 2 diabetic mice [212].
Anthocyanins, polyphenolic compounds of the flavonoid group, regulate digestive enzymes (α-amylase and α-glucosidase), GLUT-4, GLP-1, glucose-6-phsphatase (G6Pase), phosphoenolpyruvate carboxykinase (PEPCK), or PPARγ. These compounds also manage blood glucose levels by normalizing insulin secretion and insulin resistance [213]. Additional significant mechanisms involve the protection of pancreatic β cells through their anti-inflammatory and antioxidant properties [214]. It is unlikely that anthocyanins alone can regulate various classes of enzymes. Further investigations are required to confirm the clinical utility of anthocyanins for the treatment of diabetes, and studies should be standardized and quantified to draw universal conclusions regarding their true use, as suggested by the authors [213].
The extract of Delonix regia shows hypoglycemic, antioxidant, and hypolipidemic properties [215]. Of note, the antidiabetic effect of this extract was found to be similar to that of glibenclamide, a well-known antidiabetic agent, which lowers blood glucose levels by stimulating the amount of insulin produced by the pancreas. However, the constituents in Delonix regia extracts that exert antioxidant, hypoglycemic, and adverse effects remain uncharacterized.
Mulberry (Morus alba L.) leaves, used extensively as an effective traditional Chinese medicine for blood glucose management, have been reported to have beneficial effects on skeletal muscle function [216]. It appears that the flavonoids present in mulberry leaves markedly ameliorate skeletal muscle insulin resistance and enhance mitochondrial function in diabetic mice through the AMPK-PGC-1α signaling pathway [216,217,218]. Flavonoids in mulberry leaves show hypoglycemic effects via inhibiting the TGF-β1 pathway [219] and increasing antioxidase activity. Although very interesting, more randomized controlled trials should be required to warrant the findings of this study.
Vaccarin (VAC), a potent flavonoid glycoside extracted from Vaccariae Semen, has been discovered to improve blood glucose levels and insulin resistance, reduce oxidative stress, and enhance endothelium-dependent vasodilation in T2DM mouse models [220,221] (Table 1). Thus, Vaccariae Semen may serve as an inhibitor of ROS and RNS, which are instigators of several illnesses, as mentioned earlier (Table 2). It was further shown that vaccarin improves glucose metabolism and vascular endothelial function through the inhibition of the ROS/AMPK/miRNA-34a/eNOS signaling cascade [220]. The protective effects of vaccarin against ROS and RNS as a dietary supplement could be used as adjuvant therapy for T2DM.
Recent research has shown that the flavonoid extract of fenugreek restores the antioxidant enzyme activities of both SOD and catalase and causes a decrease in malondialdehyde content [222]. This extract also alleviates hyperglycemia in a streptozotocin (STZ)-induced T2DM mouse model.
Taxifolin, a recognized flavonoid, is found to be most active towards in inhibiting α-amylase. Thus, taxifolin has the ability to manage post-meal hyperglycemia [223]. It also prevents diabetic cardiomyopathy via inhibition of oxidative stress [224]. This function, in conjunction with its anti-inflammatory and antioxidant properties, can contribute to the treatment of diabetes mellitus.
A flavonoid-rich fraction derived from Trichilia emetica exhibits prominent radical scavenging and antidiabetic activities [225]. The binding of free ferrous to oxygen yields ferric iron and superoxide, thus generating hydrogen peroxide. Formed hydrogen peroxide reacts with ferrous iron and gives rise to hydroxyl radical formation. Trichilia emetica flavonoid-rich fractions are found to be capable of chelating ferrous ions. Although this plant could potentially be a valuable agent exhibiting strong antioxidant activity for controlling high blood sugar levels, the constituents in Trichilia emetica extracts remain uncharacterized.
2,3-dihydroxybenzoic acid (DHBA), a phenol, is a weak iron chelator and radical scavenger [226] (Table 2). This colonic-derived flavonoid metabolite is found to regulate glucose uptake and production in renal tubular NRK-52E cells [227]. Additional studies are needed to confirm whether or not this phenol offers a treatment benefit in a reliable large animal diabetes model.
Table 2. A summary of polyphenol-mediated protection against inflammation and associated oxidative stress [152,228,229,230,231,232,233,234].
Table 2. A summary of polyphenol-mediated protection against inflammation and associated oxidative stress [152,228,229,230,231,232,233,234].
PolyphenolsROSNOX-4ICAM-1p-ERKVCAMTNF-αIL-6MCP-1MAPKsMetal ion ChelatorRef.
Phenolic acids
 Epicatechin (EC)
 2,3-dihydroxybenzoic acid (DHBA)



















[228]
[228]
Flavonoids
 Catechins
 Formononetin










[229]
[152]
Stilbenes
 Resveratrol







[230,231,232]
Phytoestrogens
 Lignans
 Isoflavone



[233]
[234]
↑ represents a significantly elevated level; ↓ refers to a significantly reduced level.
Table 3. A summary of polyphenols as biomarkers of oxidative stress in diabetes [152,161,235,236,237,238,239,240].
Table 3. A summary of polyphenols as biomarkers of oxidative stress in diabetes [152,161,235,236,237,238,239,240].
PolyphenolsGlucoseA1CWater IntakeVolume of UrineNO2/NO3Heme Oxygenase-1 (HO-1)Nrf2Insulin
Resistance
LipoperoxidationRef.
Phenolic acids
 Phenethyl ester of caffeic acid
 Ferulic acid (FA)
 Hexane extract of Eryngium carlinae















[235]
[236]
[161]
Flavonoids
 Catechins
 Formononetin








[237]
[152]
Stilbenes
 Resveratrol



[238]
Phytoestrogen
 Lignans





[239,240]
↑ represents a significantly elevated level; ↓ refers to a significantly reduced level.

4.2. Catechins

Catechin, a flavonoid, is found in a variety of foods and herbs, including tea. As direct antioxidants, catechins are ROS scavengers and metal ion chelators (Table 1) [241]. As indirect antioxidants, they induce the activation of antioxidant enzymes [229], although catechins become pro-oxidants at micromolar levels in cells supplemented with iron and linoleic acid [241]. The antioxidant enzyme pathway is a key player in maintaining the delicate balance between ROS and RNS generation and antioxidants in both animal cells and plant cells. Antioxidant enzyme levels do not appear to be significantly changed in various tissues of aging animals [242,243,244]. However, stimulating this pathway to activate the expression [245] and the delivery of catalytic mimetic of the antioxidant enzyme pathway [246] has been considered important for preventing diseases caused by oxidative stress and various exogenous substances [247,248,249]. Various pumpkin species contain large amounts of catechin and kaempferol [250]. The pumpkin pulp shows antioxidant effects based on DPPH radicals and ferric-reducing antioxidant power (FRAP) radicals. However, the antioxidant potential of these polyphenols cannot be estimated based on the content of individual bioactive compounds, as described in detail in the paper.
Epigallocatechin Gallate (EGCG), a sub-class of flavonoids, is an antioxidant (Table 1). EGCG is identified not only as a potential autophagy regulator, but it is also capable of diminishing the expressions of inflammation-associated genes in peripheral leukocytes and adipose tissue of non-obese type 2 diabetic Goto-Kakizaki rats [225,251,252]. Evidence indicates that EGCG has pro-oxidant activity due to its instability and autoxidation [253].
Procyanidins are oligomeric compounds composed of catechin and epicatechin molecules. Research suggests that plants rich in procyanidins could have preventive effects on hyperglycemia and T2DM [254]. The readers can find additional information, including the structure, classification, and underlying mechanisms of catechins in regulating diabetes, by Wen et al., 2022 [237].
Although catechins have been considered beneficial in protecting against various diseases caused by oxidative stress and ROS, such as CVD and diabetes [237], and in hematology [173], their effects on CVD biomarkers such as the levels of C-reactive protein, B-type natriuretic peptide, N-terminal pro-atrial natriuretic peptide, aldosterone, renin, fibrinogen, D-dimer, plasminogen-activator inhibitor type 1, homocysteine, and the urinary albumin-to-creatinine ratio are not fully assessed.

4.3. Phytoestrogens

Lignan, a phytoestrogen, and polyphenols isolated from Linum usitatissimum show antidiabetic properties in STZ-induced diabetic rats [239]. Daily treatment with flaxseed extract improves HbA1c levels and blood glucose levels. It also reduces the total cholesterol, high-density lipoprotein (HDL), LDL, and triglyceride levels significantly in diabetic rats.

5. Clinical Trial of Antioxidant Therapy in Patients with Diabetes

Inflammation is intimately tied to higher glycated hemoglobin (HbA1c) levels, a good indicator of chronic hyperglycemia. Setting the treatment target for HbA1c below 6% in high-risk diabetic patients leads to reduced 5-year nonfatal myocardial infarctions in the Diabetes (ACCORD) Study Group. A new model of treating patients with antioxidants with inflammatory actions is emerging as a promising therapeutic option for patients with T2DM. The evidence provided by these various studies supports that ginger, resveratrol, and rutin flavonoid could reduce the HbA1c level, blood glucose level, and homeostatic model evaluation of insulin resistance (HOMA)-IR, an indicator of long-term glycemic control, while allium sativum, olea europaea oil, and astaxanthin cause a decrease in LDL and serum total cholesterol (Table 4).
Mudan granules possess neuroprotective effects in diabetic patients. Apparently, there is a new opportunity for these antioxidants and anti-inflammatory agents like resveratrol and rutin flavonoid to demonstrate their activity in combination with mudan granules. Notably, while ginger causes an increase in NO production, resveratrol and rutin flavonoid reduce CRP and IL-6, respectively. Thus, antioxidants may be effective for the treatment of cytokine storms in diabetes. The fact that these compounds were able to pass preclinical research and enter human studies indicates their potential effectiveness. The above data show the clinical utility of antioxidants with anti-inflammatory effects in T2DM. However, the application of these antioxidants requires further research investigations to evaluate their efficacy against T2DM-related complications, including hyperlipidemia, pro-inflammatory mediators, and insulin insensitivity. There is also a need to select diabetic patients at various stages of the disease to guarantee that the anti-inflammatory and antidiabetic effects of the compound are evident.

6. Conclusions and Future Implications

ROS mediates many important signaling functions in different cellular and developmental processes in human cells. The level and sources of intracellular ROS are critically important for their actions in influencing different complex signaling and biochemical pathways under both physiological and pathophysiological conditions.
Diabetes, a metabolic disorder, is associated with reduced levels of antioxidants, which can reduce the susceptibility of pancreatic islets to oxidative stress. Oxidative stress impairs insulin-mediated intracellular signaling pathways. Moreover, while insulin resistance is a complex metabolic disorder, oxidative stress can induce insulin resistance.
Diabetes triggers changes in endothelial function, which may lead to increased oxidative stress mediators and a reduction in antioxidants. Oxidative stress can be the result of mitochondrial dysfunction or inflammation, not the cause. Evidence indicates that oxidative stress accelerates the development of complications in diabetes. The diseases that accompany T2DM include myocardial infarction, stroke, peripheral artery disease, angina pectoris, and periodontal disease, a widespread condition that affects a large percentage of the population. These pathologies seem to arise from the same risk factors and display a similar inflammation profile. Identifying novel biomarkers and therapeutic strategies to prevent diabetes-induced endothelial dysfunction is of great interest.
Chronic production of both proinflammatory and proatherogenic mediators leads to systemic inflammatory diathesis, causing a decrease in beta cell insulin secretion and an increase in insulin resistance. The resulting chronic hyperglycemia and hyperlipidemia promote the inflammatory process, leading to a vicious cycle. Antidiabetic drugs such as metformin, pioglitazone, insulin, glucagon-like peptide-1 agonists, and dipeptidyl peptidase-4 inhibitors are found to reduce inflammation. Statin is capable of ameliorating inflammation to some extent. While anti-inflammatory drugs such as IL-1 and TNF-α antagonists are in trials, salsalate, an NSAID, appears to improve insulin sensitivity. Anti-inflammatory drugs are associated with a broad range of side effects, such as cardiovascular events and the worsening of congestive heart failure. Sirtuin and 12-lipo-oxygenase are considered new therapeutic targets for T2DM and inflammatory diseases. Unfortunately, some of the antidiabetic and anti-inflammatory drugs are linked to renal toxicity. The prevalence of total diabetes is expected to increase over the next ten years. More than ever before, it is imperative that research focus on developing new anti-inflammatory drugs with improved efficacy and safety as well as antioxidants, which serve as “safety switches” to prevent unwanted inflammatory responses via neutralizing free radicals. In this review, we highlighted that a number of therapeutic medicines derived from plants are promising sources of natural antioxidants and exhibit anti-inflammatory activities.
The United Kingdom Prospective Diabetes Study clearly verifies the importance of glycemic control in reducing the risk of microvascular complications. While the efficacy of antioxidative treatments remains elusive, recent clinical reports indicate that antioxidants have the ability to counteract inflammation and improve hyperglycemia, as well as beneficial effects on hepatocyte lipid metabolism through different mechanisms in diabetes. The mechanisms by which the antioxidants reduce hyperglycemia are only partly known. Since diabetes is a progressive disease, T2DM patients’ need for different treatments changes as well. It remains to be shown whether antioxidant treatments administered together with current antidiabetic or anti-inflammatory drugs can prevent the vascular complications of diabetes. Further studies are required to clarify the role of antioxidant therapy in the management of type 2 diabetes.

Author Contributions

Conceptualization, Z.S.-M. and F.M.; writing—original draft preparation, A.A., S.B. and Z.S.-M.; writing—review and editing, Z.S.-M., A.A., S.B. and F.M.; supervision, Z.S.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the University of Mississippi to Z.S.-M.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

A. tsao-koAmomum tsao-ko Crevost et Lemarié
AGEAdvanced glycation end
AKRAaldo-keto reductase
AKTProtein kinase
AMPKAMP-activated protein kinase
ATPAdenosine triphosphate
CATCatalase
CDCCenters for Disease Control and Prevention
CRPC-reactive protein
CVDCardiovascular diseases
DHBA2,3-dihydroxybenzoic acid
DPP-4Dipeptidyl peptidase-4
EGCGEpigallocatechin Gallate
EGFREpidermal Growth Factor Receptor
EREndoplasmic reticulum
ERK1/2Extracellular Signal-Regulated Kinase 1/2
ERO1Endoplasmic reticulum oxireductin 1
FRAPFerric reducing antioxidant power
G6PaseGlucose-6-phosphatase
GLP-1Glucagon-like peptide 1
GLUT4Glucose transporter 4
GPxGlutathione peroxidase
H2O2Hydrogen Peroxide
HbA1cHemoglobin A1c
HDLHigh-density lipoprotein
HOMA-IRHomeostatic model evaluation of insulin resistance
IL-1Interleukin-1
IL-1βInterleukin-1 beta
IL-6Interleukin-6
IRS-1Insulin Receptor Substrate 1
LDLLow-density lipoprotein
MAPKMitogen-activated protein kinase
mDNAMitochondrial DNA
NADNicotinamide Adenine Dinucleotide
NADHNicotinamide Adenine Dinucleotide
NADPHNicotinamide adenine dinucleotide phosphate
NF-kBNuclear factor kappa B
NHANESNational Health and Nutritional Examination Survey
NONitric Oxide
NOXsNicotinamide adenine dinucleotide phosphate oxidase
PARProtease-activated receptor
PARPPoly(ADP-Ribose) Polymerase
PEPCKPhosphoenolpyruvate carboxykinase
PGC-1αPeroxisome Proliferator-Activated Receptor Gamma Coactivator 1 alpha
PI3KPhosphoinositide 3-Kinase
PKCProtein kinase C
PPAR αPeroxisome proliferator-activated receptor alpha
PPAR γPeroxisome proliferator-activated receptor gamma
RAGEReceptor for advanced glycation end products
RNSReactive nitogen species
ROSReactive oxygen species
SDGSecoisolariciresinol diglucoside
SIRT1Sirtuin1
SODSuperoxide dismutase
STZStreptozotocin
T2DMType 2 diabetes mellitus
TGF-βTransforming growth factor-beta
VACVaccarin

References

  1. National Diabetes Statistics Report 2020, Estimates of Diabetes and Its Burden in the United States; U.S. Department of Health and Human Services: Washington, DC, USA, 2020.
  2. Perng, W.; Conway, R.; Mayer-Davis, E.; Dabelea, D. Youth-Onset Type 2 Diabetes: The Epidemiology of an Awakening Epidemic. Diabetes Care 2023, 46, 490–499. [Google Scholar] [CrossRef] [PubMed]
  3. Willemsen, G.; Ward, K.J.; Bell, C.G.; Christensen, K.; Bowden, J.; Dalgård, C.; Harris, J.R.; Kaprio, J.; Lyle, R.; Magnusson, P.K.; et al. The Concordance and Heritability of Type 2 Diabetes in 34,166 Twin Pairs From International Twin Registers: The Discordant Twin (DISCOTWIN) Consortium. Twin Res. Hum. Genet. 2015, 18, 762–771. [Google Scholar] [CrossRef]
  4. Almgren, P.; Lehtovirta, M.; Isomaa, B.; Sarelin, L.; Taskinen, M.R.; Lyssenko, V.; Tuomi, T.; Groop, L. Heritability and familiality of type 2 diabetes and related quantitative traits in the Botnia Study. Diabetologia 2011, 54, 2811–2819. [Google Scholar] [CrossRef] [PubMed]
  5. Mambiya, M.; Shang, M.; Wang, Y.; Li, Q.; Liu, S.; Yang, L.; Zhang, Q.; Zhang, K.; Liu, M.; Nie, F.; et al. The Play of Genes and Non-genetic Factors on Type 2 Diabetes. Front. Public Health 2019, 7, 349. [Google Scholar] [CrossRef] [PubMed]
  6. Udler, M.S. Type 2 Diabetes: Multiple Genes, Multiple Diseases. Curr. Diabetes Rep. 2019, 19, 55. [Google Scholar] [CrossRef]
  7. Caulfield, J.I.; Aizenbud, L.; Perdigoto, A.L.; Meffre, E.; Jilaveanu, L.; Michalek, D.A.; Rich, S.S.; Aizenbud, Y.; Adeniran, A.; Herold, K.C.; et al. Germline genetic variants are associated with development of insulin-dependent diabetes in cancer patients treated with immune checkpoint inhibitors. J. Immunother. Cancer 2023, 11, e006570. [Google Scholar] [CrossRef]
  8. Kamiya, J.; Aoki, Y. Associations between hyperglycaemia and somatic transversion mutations in mitochondrial DNA of people with diabetes mellitus. Diabetologia 2003, 46, 1559–1566. [Google Scholar] [CrossRef]
  9. Voight, B.F.; Scott, L.J.; Steinthorsdottir, V.; Morris, A.P.; Dina, C.; Welch, R.P.; Zeggini, E.; Huth, C.; Aulchenko, Y.S.; Thorleifsson, G.; et al. Twelve type 2 diabetes susceptibility loci identified through large-scale association analysis. Nat. Genet. 2010, 42, 579–589. [Google Scholar] [CrossRef]
  10. Billings, L.K.; Florez, J.C. The genetics of type 2 diabetes: What have we learned from GWAS? Ann. N. Y. Acad. Sci. 2010, 1212, 59–77. [Google Scholar] [CrossRef]
  11. Heijmans, B.T.; Tobi, E.W.; Stein, A.D.; Putter, H.; Blauw, G.J.; Susser, E.S.; Slagboom, P.E.; Lumey, L.H. Persistent epigenetic differences associated with prenatal exposure to famine in humans. Proc. Natl. Acad. Sci. USA 2008, 105, 17046–17049. [Google Scholar] [CrossRef]
  12. Seki, Y.; Williams, L.; Vuguin, P.M.; Charron, M.J. Minireview: Epigenetic programming of diabetes and obesity: Animal models. Endocrinology 2012, 153, 1031–1038. [Google Scholar] [CrossRef]
  13. Ley, R.E.; Bäckhed, F.; Turnbaugh, P.; Lozupone, C.A.; Knight, R.D.; Gordon, J.I. Obesity alters gut microbial ecology. Proc. Natl. Acad. Sci. USA 2005, 102, 11070–11075. [Google Scholar] [CrossRef] [PubMed]
  14. Pérez-Torres, I.; Castrejón-Téllez, V.; Soto, M.E.; Rubio-Ruiz, M.E.; Manzano-Pech, L.; Guarner-Lans, V. Oxidative Stress, Plant Natural Antioxidants, and Obesity. Int. J. Mol. Sci. 2021, 22, 1786. [Google Scholar] [CrossRef]
  15. Giacco, F.; Brownlee, M. Oxidative Stress and Diabetic Complications. Circ. Res. 2010, 107, 1058–1070. [Google Scholar] [CrossRef] [PubMed]
  16. Pitocco, D.; Tesauro, M.; Alessandro, R.; Ghirlanda, G.; Cardillo, C. Oxidative Stress in Diabetes: Implications for Vascular and Other Complications. Int. J. Mol. Sci. 2013, 14, 21525–21550. [Google Scholar] [CrossRef]
  17. Kelsey, M.M.; Zaepfel, A.; Bjornstad, P.; Nadeau, K.J. Age-related consequences of childhood obesity. Gerontology 2014, 60, 222–228. [Google Scholar] [CrossRef] [PubMed]
  18. Afkarian, M. Diabetic kidney disease in children and adolescents. Pediatr. Nephrol. 2015, 30, 65–74. [Google Scholar] [CrossRef] [PubMed]
  19. Bjerregaard, L.G.; Jensen, B.W.; Angquist, L.; Osler, M.; Sorensen, T.I.A.; Baker, J.L. Change in Overweight from Childhood to Early Adulthood and Risk of Type 2 Diabetes. N. Engl. J. Med. 2018, 378, 1302–1312. [Google Scholar] [CrossRef] [PubMed]
  20. Bianco-Miotto, T.; Craig, J.M.; Gasser, Y.P.; van Dijk, S.J.; Ozanne, S.E. Epigenetics and DOHaD: From basics to birth and beyond. J. Dev. Orig. Health Dis. 2017, 8, 513–519. [Google Scholar] [CrossRef]
  21. Abraham, M.A.; Filippi, B.M.; Kang, G.M.; Kim, M.S.; Lam, T.K. Insulin action in the hypothalamus and dorsal vagal complex. Exp. Physiol. 2014, 99, 1104–1109. [Google Scholar] [CrossRef]
  22. Cade, W.T. Diabetes-related microvascular and macrovascular diseases in the physical therapy setting. Phys. Ther. 2008, 88, 1322–1335. [Google Scholar] [CrossRef] [PubMed]
  23. Rahman, S.; Rahman, T.; Ismail, A.A.; Rashid, A.R. Diabetes-associated macrovasculopathy: Pathophysiology and pathogenesis. Diabetes Obes. Metab. 2007, 9, 767–780. [Google Scholar] [CrossRef]
  24. Chow, L.S.; Chen, H.; Miller, M.E.; Marcovina, S.M.; Seaquist, E.R. Biomarkers related to severe hypoglycaemia and lack of good glycaemic control in ACCORD. Diabetologia 2015, 58, 1160–1166. [Google Scholar] [CrossRef] [PubMed]
  25. Fanelli, C.; Calderone, S.; Epifano, L.; De Vincenzo, A.; Modarelli, F.; Pampanelli, S.; Perriello, G.; De Feo, P.; Brunetti, P.; Gerich, J.E.; et al. Demonstration of a critical role for free fatty acids in mediating counterregulatory stimulation of gluconeogenesis and suppression of glucose utilization in humans. J. Clin. Investig. 1993, 92, 1617–1622. [Google Scholar] [CrossRef]
  26. Cao, C.; Koh, H.E.; Van Vliet, S.; Patterson, B.W.; Reeds, D.N.; Laforest, R.; Gropler, R.J.; Mittendorfer, B. Increased plasma fatty acid clearance, not fatty acid concentration, is associated with muscle insulin resistance in people with obesity. Metabolism 2022, 132, 155216. [Google Scholar] [CrossRef]
  27. Ighodaro, O.M.; Akinloye, O.A. First line defence antioxidants-superoxide dismutase (SOD), catalase (CAT) and glutathione peroxidase (GPX): Their fundamental role in the entire antioxidant defence grid. Alex. J. Med. 2018, 54, 287–293. [Google Scholar] [CrossRef]
  28. King, G.L. The role of hyperglycaemia and hyperinsulinaemia in causing vascular dysfunction in diabetes. Ann. Med. 1996, 28, 427–432. [Google Scholar] [CrossRef]
  29. 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. 2004, 114, 1752–1761. [Google Scholar] [CrossRef]
  30. Leyane, T.S.; Jere, S.W.; Houreld, N.N. Oxidative Stress in Ageing and Chronic Degenerative Pathologies: Molecular Mechanisms Involved in Counteracting Oxidative Stress and Chronic Inflammation. Int. J. Mol. Sci. 2022, 23, 7273. [Google Scholar] [CrossRef]
  31. Tangvarasittichai, S. Oxidative stress, insulin resistance, dyslipidemia and type 2 diabetes mellitus. World J. Diabetes 2015, 6, 456–480. [Google Scholar] [CrossRef]
  32. Galicia-Garcia, U.; Benito-Vicente, A.; Jebari, S.; Larrea-Sebal, A.; Siddiqi, H.; Uribe, K.B.; Ostolaza, H.; Martín, C. Pathophysiology of Type 2 Diabetes Mellitus. Int. J. Mol. Sci. 2020, 21, 6275. [Google Scholar] [CrossRef]
  33. Thannickal, V.J.; Fanburg, B.L. Reactive oxygen species in cell signaling. Am. J. Physiol. Lung Cell. Mol. Physiol. 2000, 279, L1005–L1028. [Google Scholar] [CrossRef] [PubMed]
  34. Sinenko, S.A.; Starkova, T.Y.; Kuzmin, A.A.; Tomilin, A.N. Physiological Signaling Functions of Reactive Oxygen Species in Stem Cells: From Flies to Man. Front. Cell Dev. Biol. 2021, 9, 714370. [Google Scholar] [CrossRef]
  35. Kakimoto, M.; Inoguchi, T.; Sonta, T.; Yu, H.Y.; Imamura, M.; Etoh, T.; Hashimoto, T.; Nawata, H. Accumulation of 8-hydroxy-2′-deoxyguanosine and mitochondrial DNA deletion in kidney of diabetic rats. Diabetes 2002, 51, 1588–1595. [Google Scholar] [CrossRef] [PubMed]
  36. Nekhaeva, E.; Bodyak, N.D.; Kraytsberg, Y.; McGrath, S.B.; Van Orsouw, N.J.; Pluzhnikov, A.; Wei, J.Y.; Vijg, J.; Khrapko, K. Clonally expanded mtDNA point mutations are abundant in individual cells of human tissues. Proc. Natl. Acad. Sci. USA 2002, 99, 5521–5526. [Google Scholar] [CrossRef] [PubMed]
  37. Wang, K.; Zhou, W.; Meng, P.; Wang, P.; Zhou, C.; Yao, Y.; Wu, S.; Wang, Y.; Zhao, J.; Zou, D.; et al. Immune-related somatic mutation genes are enriched in PDACs with diabetes. Transl. Oncol. 2019, 12, 1147–1154. [Google Scholar] [CrossRef]
  38. Baker, R.G.; Hayden, M.S.; Ghosh, S. NF-κB, inflammation, and metabolic disease. Cell Metab. 2011, 13, 11–22. [Google Scholar] [CrossRef]
  39. Kay, A.M.; Simpson, C.L.; Stewart, J.A., Jr. The Role of AGE/RAGE Signaling in Diabetes-Mediated Vascular Calcification. J. Diabetes Res. 2016, 2016, 6809703. [Google Scholar] [CrossRef]
  40. Agrawal, A.; Cha-Molstad, H.; Samols, D.; Kushner, I. Overexpressed nuclear factor-κB can participate in endogenous C-reactive protein induction, and enhances the effects of C/EBPβ and signal transducer and activator of transcription-3. Immunology 2003, 108, 539–547. [Google Scholar] [CrossRef] [PubMed]
  41. Spranger, J.; Kroke, A.; Möhlig, M.; Hoffmann, K.; Bergmann, M.M.; Ristow, M.; Boeing, H.; Pfeiffer, A.F.H. Inflammatory Cytokines and the Risk to Develop Type 2 Diabetes: Results of the Prospective Population-Based European Prospective Investigation into Cancer and Nutrition (EPIC)-Potsdam Study. Diabetes 2003, 52, 812–817. [Google Scholar] [CrossRef]
  42. Suryavanshi, S.V.; Kulkarni, Y.A. NF-κβ: A Potential Target in the Management of Vascular Complications of Diabetes. Front. Pharmacol. 2017, 8, 798. [Google Scholar] [CrossRef]
  43. Malhotra, J.D.; Kaufman, R.J. Endoplasmic reticulum stress and oxidative stress: A vicious cycle or a double-edged sword? Antioxid. Redox Signal 2007, 9, 2277–2293. [Google Scholar] [CrossRef]
  44. Kairisalo, M.; Korhonen, L.; Blomgren, K.; Lindholm, D. X-linked inhibitor of apoptosis protein increases mitochondrial antioxidants through NF-kappaB activation. Biochem. Biophys. Res. Commun. 2007, 364, 138–144. [Google Scholar] [CrossRef]
  45. Pham, C.G.; Bubici, C.; Zazzeroni, F.; Papa, S.; Jones, J.; Alvarez, K.; Jayawardena, S.; De Smaele, E.; Cong, R.; Beaumont, C.; et al. Ferritin heavy chain upregulation by NF-kappaB inhibits TNFalpha-induced apoptosis by suppressing reactive oxygen species. Cell 2004, 119, 529–542. [Google Scholar] [CrossRef] [PubMed]
  46. Lingappan, K. NF-κB in Oxidative Stress. Curr. Opin. Toxicol. 2018, 7, 81–86. [Google Scholar] [CrossRef]
  47. 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, 50–59. [Google Scholar] [CrossRef]
  48. Liu, C.; Dong, W.; Li, J.; Kong, Y.; Ren, X. O-GlcNAc Modification and Its Role in Diabetic Retinopathy. Metabolites 2022, 12, 725. [Google Scholar] [CrossRef] [PubMed]
  49. Wu, T.; Qiao, S.; Shi, C.; Wang, S.; Ji, G. Metabolomics window into diabetic complications. J. Diabetes Investig. 2018, 9, 244–255. [Google Scholar] [CrossRef] [PubMed]
  50. Mantegazza, A.R.; Savina, A.; Vermeulen, M.; Pérez, L.; Geffner, J.; Hermine, O.; Rosenzweig, S.D.; Faure, F.; Amigorena, S. NADPH oxidase controls phagosomal pH and antigen cross-presentation in human dendritic cells. Blood 2008, 112, 4712–4722. [Google Scholar] [CrossRef] [PubMed]
  51. Chougnet, C.A.; Thacker, R.I.; Shehata, H.M.; Hennies, C.M.; Lehn, M.A.; Lages, C.S.; Janssen, E.M. Loss of Phagocytic and Antigen Cross-Presenting Capacity in Aging Dendritic Cells Is Associated with Mitochondrial Dysfunction. J. Immunol. 2015, 195, 2624–2632. [Google Scholar] [CrossRef]
  52. Gao, Z.; Chen, X. Fatty Acid β-Oxidation in Kidney Diseases: Perspectives on Pathophysiological Mechanisms and Therapeutic Opportunities. Front. Pharmacol. 2022, 13, 805281. [Google Scholar] [CrossRef] [PubMed]
  53. Pawlak, M.; Lefebvre, P.; Staels, B. Molecular mechanism of PPARα action and its impact on lipid metabolism, inflammation and fibrosis in non-alcoholic fatty liver disease. J. Hepatol. 2015, 62, 720–733. [Google Scholar] [CrossRef]
  54. Ke, R.; Xu, Q.; Li, C.; Luo, L.; Huang, D. Mechanisms of AMPK in the maintenance of ATP balance during energy metabolism. Cell Biol. Int. 2018, 42, 384–392. [Google Scholar] [CrossRef]
  55. Rohm, T.V.; Meier, D.T.; Olefsky, J.M.; Donath, M.Y. Inflammation in obesity, diabetes, and related disorders. Immunity 2022, 55, 31–55. [Google Scholar] [CrossRef]
  56. Nordestgaard, B.G. Triglyceride-Rich Lipoproteins and Atherosclerotic Cardiovascular Disease: New Insights From Epidemiology, Genetics, and Biology. Circ. Res. 2016, 118, 547–563. [Google Scholar] [CrossRef]
  57. Borén, J.; Chapman, M.J.; Krauss, R.M.; Packard, C.J.; Bentzon, J.F.; Binder, C.J.; Daemen, M.J.; Demer, L.L.; Hegele, R.A.; Nicholls, S.J.; et al. Low-density lipoproteins cause atherosclerotic cardiovascular disease: Pathophysiological, genetic, and therapeutic insights: A consensus statement from the European Atherosclerosis Society Consensus Panel. Eur. Heart J. 2020, 41, 2313–2330. [Google Scholar] [CrossRef]
  58. Mussbacher, M.; Schossleitner, K.; Kral-Pointner, J.B.; Salzmann, M.; Schrammel, A.; Schmid, J.A. More than Just a Monolayer: The Multifaceted Role of Endothelial Cells in the Pathophysiology of Atherosclerosis. Curr. Atheroscler. Rep. 2022, 24, 483–492. [Google Scholar] [CrossRef]
  59. Kornowski, R.; Hong, M.K.; Tio, F.O.; Bramwell, O.; Wu, H.; Leon, M.B. In-Stent Restenosis: Contributions of Inflammatory Responses and Arterial Injury to Neointimal Hyperplasia. J. Am. Coll. Cardiol. 1998, 31, 224–230. [Google Scholar] [CrossRef] [PubMed]
  60. Che Man, R.; Sulaiman, N.; Ishak, M.F.; Bt Hj Idrus, R.; Abdul Rahman, M.R.; Yazid, M.D. The Effects of Pro-Inflammatory and Anti-Inflammatory Agents for the Suppression of Intimal Hyperplasia: An Evidence-Based Review. Int. J. Environ. Res. Public Health 2020, 17, 7825. [Google Scholar] [CrossRef] [PubMed]
  61. Centers for Disease Control and Prevention. National Diabetes Statistics Report, 2020; Centers for Disease Control and Prevention, U.S. Dept of Health and Human Services: Atlanta, GA, USA, 2020.
  62. Favre, G.A.; Esnault, V.L.M.; Obberghen, E.V. Modulation of glucose metabolism by the renin-angiotensin-aldosterone system. Am. J. Physiol. Endocrinol. Metab. 2015, 308, E435–E449. [Google Scholar] [CrossRef]
  63. Petrie, J.R.; Guzik, T.J.; Touyz, R.M. Diabetes, Hypertension, and Cardiovascular Disease: Clinical Insights and Vascular Mechanisms. Can. J. Cardiol. 2018, 34, 575–584. [Google Scholar] [CrossRef] [PubMed]
  64. Cheung, B.M.; Li, C. Diabetes and hypertension: Is there a common metabolic pathway? Curr. Atheroscler. Rep. 2012, 14, 160–166. [Google Scholar] [CrossRef] [PubMed]
  65. Dregan, A.; Charlton, J.; Chowienczyk, P.; Gulliford, M.C. Chronic inflammatory disorders and risk of type 2 diabetes mellitus, coronary heart disease, and stroke: A population-based cohort study. Circulation 2014, 130, 837–844. [Google Scholar] [CrossRef] [PubMed]
  66. Armstrong, A.W.; Harskamp, C.T.; Armstrong, E.J. Psoriasis and the risk of diabetes mellitus: A systematic review and meta-analysis. JAMA Dermatol. 2013, 149, 84–91. [Google Scholar] [CrossRef] [PubMed]
  67. Zhang, J.; Fu, L.; Shi, J.; Chen, X.; Li, Y.; Ma, B.; Zhang, Y. The risk of metabolic syndrome in patients with rheumatoid arthritis: A meta-analysis of observational studies. PLoS ONE 2013, 8, e78151. [Google Scholar] [CrossRef]
  68. Furman, D.; Campisi, J.; Verdin, E.; Carrera-Bastos, P.; Targ, S.; Franceschi, C.; Ferrucci, L.; Gilroy, D.W.; Fasano, A.; Miller, G.W.; et al. Chronic inflammation in the etiology of disease across the life span. Nat. Med. 2019, 25, 1822–1832. [Google Scholar] [CrossRef]
  69. He, S.; Sharpless, N.E. Senescence in Health and Disease. Cell 2017, 169, 1000–1011. [Google Scholar] [CrossRef]
  70. Coppé, J.P.; Desprez, P.Y.; Krtolica, A.; Campisi, J. The senescence-associated secretory phenotype: The dark side of tumor suppression. Annu. Rev. Pathol. 2010, 5, 99–118. [Google Scholar] [CrossRef]
  71. Esposito, K.; Nappo, F.; Marfella, R.; Giugliano, G.; Giugliano, F.; Ciotola, M.; Quagliaro, L.; Ceriello, A.; Giugliano, D. Inflammatory Cytokine Concentrations Are Acutely Increased by Hyperglycemia in Humans. Circulation 2002, 106, 2067–2072. [Google Scholar] [CrossRef]
  72. Rodrigo, R.; González, J.; Paoletto, F. The role of oxidative stress in the pathophysiology of hypertension. Hypertens. Res. 2011, 34, 431–440. [Google Scholar] [CrossRef]
  73. Lacy, F.; Kailasam, M.T.; O’Connor, D.T.; Schmid-Schönbein, G.W.; Parmer, R.J. Plasma hydrogen peroxide production in human essential hypertension: Role of heredity, gender, and ethnicity. Hypertension 2000, 36, 878–884. [Google Scholar] [CrossRef] [PubMed]
  74. Redón, J.; Oliva, M.R.; Tormos, C.; Giner, V.; Chaves, J.; Iradi, A.; Sáez, G.T. Antioxidant activities and oxidative stress byproducts in human hypertension. Hypertension 2003, 41, 1096–1101. [Google Scholar] [CrossRef] [PubMed]
  75. Rajagopalan, S.; Kurz, S.; Münzel, T.; Tarpey, M.; Freeman, B.A.; Griendling, K.K.; Harrison, D.G. Angiotensin II-mediated hypertension in the rat increases vascular superoxide production via membrane NADH/NADPH oxidase activation. Contribution to alterations of vasomotor tone. J. Clin. Investig. 1996, 97, 1916–1923. [Google Scholar] [CrossRef]
  76. Jung, O.; Schreiber, J.G.; Geiger, H.; Pedrazzini, T.; Busse, R.; Brandes, R.P. gp91phox-containing NADPH oxidase mediates endothelial dysfunction in renovascular hypertension. Circulation 2004, 109, 1795–1801. [Google Scholar] [CrossRef]
  77. Terada, Y.; Yayama, K. Angiotensin II-Induced Vasoconstriction via Rho Kinase Activation in Pressure-Overloaded Rat Thoracic Aortas. Biomolecules 2021, 11, 1076. [Google Scholar] [CrossRef] [PubMed]
  78. Malhotra, A.; Kang, B.P.; Cheung, S.; Opawumi, D.; Meggs, L.G. Angiotensin II promotes glucose-induced activation of cardiac protein kinase C isozymes and phosphorylation of troponin I. Diabetes 2001, 50, 1918–1926. [Google Scholar] [CrossRef]
  79. Tian, Y.; Smith, R.D.; Balla, T.; Catt, K.J. Angiotensin II activates mitogen-activated protein kinase via protein kinase C and Ras/Raf-1 kinase in bovine adrenal glomerulosa cells. Endocrinology 1998, 139, 1801–1809. [Google Scholar] [CrossRef]
  80. Togliatto, G.; Lombardo, G.; Brizzi, M.F. The Future Challenge of Reactive Oxygen Species (ROS) in Hypertension: From Bench to Bed Side. Int. J. Mol. Sci. 2017, 18, 1988. [Google Scholar] [CrossRef]
  81. Millatt, L.J.; Abdel-Rahman, E.M.; Siragy, H.M. Angiotensin II and nitric oxide: A question of balance. Regul. Pept. 1999, 81, 1–10. [Google Scholar] [CrossRef]
  82. Chen, K.; Pittman, R.N.; Popel, A.S. Nitric oxide in the vasculature: Where does it come from and where does it go? A quantitative perspective. Antioxid. Redox Signal. 2008, 10, 1185–1198. [Google Scholar] [CrossRef]
  83. Kawanami, D.; Takashi, Y.; Muta, Y.; Oda, N.; Nagata, D.; Takahashi, H.; Tanabe, M. Mineralocorticoid Receptor Antagonists in Diabetic Kidney Disease. Front. Pharmacol. 2021, 12, 754239. [Google Scholar] [CrossRef]
  84. Hill, M.A.; Yang, Y.; Zhang, L.; Sun, Z.; Jia, G.; Parrish, A.R.; Sowers, J.R. Insulin resistance, cardiovascular stiffening and cardiovascular disease. Metabolism 2021, 119, 154766. [Google Scholar] [CrossRef] [PubMed]
  85. Gaede, P.; Vedel, P.; Larsen, N.; Jensen, G.V.; Parving, H.H.; Pedersen, O. Multifactorial intervention and cardiovascular disease in patients with type 2 diabetes. N. Engl. J. Med. 2003, 348, 383–393. [Google Scholar] [CrossRef] [PubMed]
  86. Sena, C.M.; Leandro, A.; Azul, L.; Seica, R.; Perry, G. Vascular Oxidative Stress: Impact and Therapeutic Approaches. Front. Physiol. 2018, 9, 1668. [Google Scholar] [CrossRef] [PubMed]
  87. Tekin, K.; Tekin, M.I. Chapter 3—Oxidative stress and diabetic retinopathy. In Pathology; Preedy, V.R., Ed.; Academic Press: Cambridge, MA, USA, 2020; pp. 29–37. [Google Scholar]
  88. Davi, G.; Falco, A.; Patrono, C. Lipid peroxidation in diabetes mellitus. Antioxid. Redox Signal. 2005, 7, 256–268. [Google Scholar] [CrossRef]
  89. Pham-Huy, L.A.; He, H.; Pham-Huy, C. Free radicals, antioxidants in disease and health. Int. J. Biomed. Sci. 2008, 4, 89–96. [Google Scholar]
  90. Valko, M.; Leibfritz, D.; Moncol, J.; Cronin, M.T.; Mazur, M.; Telser, J. Free radicals and antioxidants in normal physiological functions and human disease. Int. J. Biochem. Cell Biol. 2007, 39, 44–84. [Google Scholar] [CrossRef] [PubMed]
  91. Pizzino, G.; Irrera, N.; Cucinotta, M.; Pallio, G.; Mannino, F.; Arcoraci, V.; Squadrito, F.; Altavilla, D.; Bitto, A. Oxidative Stress: Harms and Benefits for Human Health. Oxidative Med. Cell. Longev. 2017, 2017, 8416763. [Google Scholar] [CrossRef]
  92. Burgos-Moron, E.; Abad-Jimenez, Z.; Maranon, A.M.; Iannantuoni, F.; Escribano-Lopez, I.; Lopez-Domenech, S.; Salom, C.; Jover, A.; Mora, V.; Roldan, I.; et al. Relationship Between Oxidative Stress, ER Stress, and Inflammation in Type 2 Diabetes: The Battle Continues. J. Clin. Med. 2019, 8, 1385. [Google Scholar] [CrossRef]
  93. Ray, P.D.; Huang, B.W.; Tsuji, Y. Reactive oxygen species (ROS) homeostasis and redox regulation in cellular signaling. Cell. Signal. 2012, 24, 981–990. [Google Scholar] [CrossRef]
  94. Bolisetty, S.; Jaimes, E.A. Mitochondria and reactive oxygen species: Physiology and pathophysiology. Int. J. Mol. Sci. 2013, 14, 6306–6344. [Google Scholar] [CrossRef]
  95. Venditti, P.; Di Stefano, L.; Di Meo, S. Mitochondrial metabolism of reactive oxygen species. Mitochondrion 2013, 13, 71–82. [Google Scholar] [CrossRef]
  96. Guo, C.; Sun, L.; Chen, X.; Zhang, D. Oxidative stress, mitochondrial damage and neurodegenerative diseases. Neural Regen. Res. 2013, 8, 2003–2014. [Google Scholar] [CrossRef] [PubMed]
  97. Kowluru, R.A.; Mohammad, G. Mitochondrial Fragmentation in a High Homocysteine Environment in Diabetic Retinopathy. Antioxidants 2022, 11, 365. [Google Scholar] [CrossRef] [PubMed]
  98. Nadalutti, C.A.; Ayala-Pena, S.; Santos, J.H. Mitochondrial DNA damage as driver of cellular outcomes. Am. J. Physiol. Cell Physiol. 2022, 322, C136–C150. [Google Scholar] [CrossRef] [PubMed]
  99. Duraisamy, A.J.; Mohammad, G.; Kowluru, R.A. Mitochondrial fusion and maintenance of mitochondrial homeostasis in diabetic retinopathy. Biochim. Biophys. Acta Mol. Basis Dis. 2019, 1865, 1617–1626. [Google Scholar] [CrossRef]
  100. So, S.; Lee, S.; Lee, Y.; Han, J.; Kang, S.; Choi, J.; Kim, B.; Kim, D.; Yoo, H.J.; Shim, I.K.; et al. Dysfunctional pancreatic cells differentiated from induced pluripotent stem cells with mitochondrial DNA mutations. BMB Rep. 2022, 55, 453–458. [Google Scholar] [CrossRef]
  101. Neginskaya, M.A.; Pavlov, E.V.; Sheu, S.S. Electrophysiological properties of the mitochondrial permeability transition pores: Channel diversity and disease implication. Biochim. Biophys. Acta Bioenerg. 2021, 1862, 148357. [Google Scholar] [CrossRef]
  102. Scialo, F.; Fernandez-Ayala, D.J.; Sanz, A. Role of Mitochondrial Reverse Electron Transport in ROS Signaling: Potential Roles in Health and Disease. Front. Physiol. 2017, 8, 428. [Google Scholar] [CrossRef]
  103. Manford, A.G.; Mena, E.L.; Shih, K.Y.; Gee, C.L.; McMinimy, R.; Martinez-Gonzalez, B.; Sherriff, R.; Lew, B.; Zoltek, M.; Rodriguez-Perez, F.; et al. Structural basis and regulation of the reductive stress response. Cell 2021, 184, 5375–5390.e5316. [Google Scholar] [CrossRef]
  104. Zhang, H.; Shen, Y.; Kim, I.M.; Weintraub, N.L.; Tang, Y. The Impaired Bioenergetics of Diabetic Cardiac Microvascular Endothelial Cells. Front. Endocrinol. 2021, 12, 642857. [Google Scholar] [CrossRef] [PubMed]
  105. Barnes, P.J. Pulmonary Diseases and Ageing. Subcell. Biochem. 2019, 91, 45–74. [Google Scholar] [CrossRef] [PubMed]
  106. Qian, L.; Mehrabi Nasab, E.; Athari, S.M.; Athari, S.S. Mitochondria signaling pathways in allergic asthma. J. Investig. Med. 2022, 70, 863–882. [Google Scholar] [CrossRef]
  107. Saxena, S.; Mathur, A.; Kakkar, P. Critical role of mitochondrial dysfunction and impaired mitophagy in diabetic nephropathy. J. Cell Physiol. 2019, 234, 19223–19236. [Google Scholar] [CrossRef] [PubMed]
  108. Ding, L.; Sun, W.; Balaz, M.; He, A.; Klug, M.; Wieland, S.; Caiazzo, R.; Raverdy, V.; Pattou, F.; Lefebvre, P.; et al. Peroxisomal β-oxidation acts as a sensor for intracellular fatty acids and regulates lipolysis. Nat. Metab. 2021, 3, 1648–1661. [Google Scholar] [CrossRef] [PubMed]
  109. Wong, H.-S.; Dighe, P.A.; Mezera, V.; Monternier, P.-A.; Brand, M.D. Production of superoxide and hydrogen peroxide from specific mitochondrial sites under different bioenergetic conditions. J. Biol. Chem. 2017, 292, 16804–16809. [Google Scholar] [CrossRef]
  110. Schrader, M.; Fahimi, H.D. Peroxisomes and oxidative stress. Biochim. Biophys. Acta Mol. Cell Res. 2006, 1763, 1755–1766. [Google Scholar] [CrossRef]
  111. Desvergne, B.; Wahli, W. Peroxisome proliferator-activated receptors: Nuclear control of metabolism. Endocr. Rev. 1999, 20, 649–688. [Google Scholar] [CrossRef]
  112. Lehmann, J.M.; Moore, L.B.; Smith-Oliver, T.A.; Wilkison, W.O.; Willson, T.M.; Kliewer, S.A. An antidiabetic thiazolidinedione is a high affinity ligand for peroxisome proliferator-activated receptor gamma (PPAR gamma). J. Biol. Chem. 1995, 270, 12953–12956. [Google Scholar] [CrossRef] [PubMed]
  113. Pascual-Ahuir, A.; Manzanares-Estreder, S.; Proft, M. Pro- and Antioxidant Functions of the Peroxisome-Mitochondria Connection and Its Impact on Aging and Disease. Oxidative Med. Cell. Longev. 2017, 2017, 9860841. [Google Scholar] [CrossRef]
  114. Kim, J.; Bai, H. Peroxisomal Stress Response and Inter-Organelle Communication in Cellular Homeostasis and Aging. Antioxidants 2022, 11, 192. [Google Scholar] [CrossRef] [PubMed]
  115. Wu, J.; Zhang, R.; Torreggiani, M.; Ting, A.; Xiong, H.; Striker, G.E.; Vlassara, H.; Zheng, F. Induction of Diabetes in Aged C57B6 Mice Results in Severe Nephropathy: An Association with Oxidative Stress, Endoplasmic Reticulum Stress, and Inflammation. Am. J. Pathol. 2010, 176, 2163–2176. [Google Scholar] [CrossRef] [PubMed]
  116. Cunard, R.; Sharma, K. The endoplasmic reticulum stress response and diabetic kidney disease. Am. J. Physiol. Ren. Physiol. 2011, 300, F1054–F1061. [Google Scholar] [CrossRef] [PubMed]
  117. Cao, S.S.; Kaufman, R.J. Endoplasmic reticulum stress and oxidative stress in cell fate decision and human disease. Antioxid. Redox Signal. 2014, 21, 396–413. [Google Scholar] [CrossRef] [PubMed]
  118. Huh, W.K.; Falvo, J.V.; Gerke, L.C.; Carroll, A.S.; Howson, R.W.; Weissman, J.S.; O’Shea, E.K. Global analysis of protein localization in budding yeast. Nature 2003, 425, 686–691. [Google Scholar] [CrossRef] [PubMed]
  119. Snezhkina, A.V.; Kudryavtseva, A.V.; Kardymon, O.L.; Savvateeva, M.V.; Melnikova, N.V.; Krasnov, G.S.; Dmitriev, A.A. ROS Generation and Antioxidant Defense Systems in Normal and Malignant Cells. Oxidative Med. Cell. Longev. 2019, 2019, 6175804. [Google Scholar] [CrossRef] [PubMed]
  120. Zito, E. ERO1: A protein disulfide oxidase and H2O2 producer. Free Radic. Biol. Med. 2015, 83, 299–304. [Google Scholar] [CrossRef]
  121. Zeeshan, H.M.; Lee, G.H.; Kim, H.R.; Chae, H.J. Endoplasmic Reticulum Stress and Associated ROS. Int. J. Mol. Sci. 2016, 17, 327. [Google Scholar] [CrossRef]
  122. Khojasteh, A.; Mirjalili, M.H.; Alcalde, M.A.; Cusido, R.M.; Eibl, R.; Palazon, J. Powerful Plant Antioxidants: A New Biosustainable Approach to the Production of Rosmarinic Acid. Antioxidants 2020, 9, 1273. [Google Scholar] [CrossRef]
  123. Hussain, T.; Tan, B.; Murtaza, G.; Liu, G.; Rahu, N.; Saleem Kalhoro, M.; Hussain Kalhoro, D.; Adebowale, T.O.; Usman Mazhar, M.; Rehman, Z.U.; et al. Flavonoids and type 2 diabetes: Evidence of efficacy in clinical and animal studies and delivery strategies to enhance their therapeutic efficacy. Pharmacol. Res. 2020, 152, 104629. [Google Scholar] [CrossRef]
  124. González, R.; Ballester, I.; López-Posadas, R.; Suárez, M.D.; Zarzuelo, A.; Martínez-Augustin, O.; Sánchez de Medina, F. Effects of flavonoids and other polyphenols on inflammation. Crit. Rev. Food Sci. Nutr. 2011, 51, 331–362. [Google Scholar] [CrossRef]
  125. Panche, A.N.; Diwan, A.D.; Chandra, S.R. Flavonoids: An overview. J. Nutr. Sci. 2016, 5, e47. [Google Scholar] [CrossRef] [PubMed]
  126. Zou, J.; Yu, X.; Qu, S.; Li, X.; Jin, Y.; Sui, D. Protective effect of total flavonoids extracted from the leaves of Murraya paniculata (L.) Jack on diabetic nephropathy in rats. Food Chem. Toxicol. 2014, 64, 231–237. [Google Scholar] [CrossRef]
  127. Jin, Y.; Arroo, R. The protective effects of flavonoids and carotenoids against diabetic complications-A review of in vivo evidence. Front. Nutr. 2023, 10, 1020950. [Google Scholar] [CrossRef]
  128. Nyakundi, B.B.; Yang, J. Uses of Papaya Leaf and Seaweed Supplementations for Controlling Glucose Homeostasis in Diabetes. Int. J. Mol. Sci. 2023, 24, 6846. [Google Scholar] [CrossRef]
  129. Han, X.; Wu, Y.C.; Meng, M.; Sun, Q.S.; Gao, S.M.; Sun, H. Linarin prevents LPS-induced acute lung injury by suppressing oxidative stress and inflammation via inhibition of TXNIP/NLRP3 and NF-kappaB pathways. Int. J. Mol. Med. 2018, 42, 1460–1472. [Google Scholar] [CrossRef]
  130. Wang, T.; Shan, M.Y.; Tang, C.Y.; Cheng, M.Y.; Chen, B.; Yan, J.; Xu, Z.H. Linarin Ameliorates Diabetic Liver Injury by Alleviating Oxidative Stress and Inflammation through the Inhibition of AKR1B1. Comb. Chem. High Throughput Screen. 2023. [Google Scholar] [CrossRef]
  131. Lee, J.; Mitchell, A.E. Quercetin and Isorhamnetin Glycosides in Onion (Allium cepa L.): Varietal Comparison, Physical Distribution, Coproduct Evaluation, and Long-Term Storage Stability. J. Agric. Food Chem. 2011, 59, 857–863. [Google Scholar] [CrossRef]
  132. Abdel Motaal, A.; Salem, H.H.; Almaghaslah, D.; Alsayari, A.; Bin Muhsinah, A.; Alfaifi, M.Y.; Elbehairi, S.E.I.; Shati, A.A.; El-Askary, H. Flavonol Glycosides: In Vitro Inhibition of DPPIV, Aldose Reductase and Combating Oxidative Stress are Potential Mechanisms for Mediating the Antidiabetic Activity of Cleome droserifolia. Molecules 2020, 25, 5864. [Google Scholar] [CrossRef] [PubMed]
  133. Kalai, F.Z.; Boulaaba, M.; Ferdousi, F.; Isoda, H. Effects of Isorhamnetin on Diabetes and Its Associated Complications: A Review of In Vitro and In Vivo Studies and a Post Hoc Transcriptome Analysis of Involved Molecular Pathways. Int. J. Mol. Sci. 2022, 23, 704. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, X.F.; Tang, Y.J.; Guan, X.X.; Lu, X.; Li, J.; Chen, X.L.; Deng, J.L.; Fan, J.M. Flavonoid constituents of Amomum tsao-ko Crevost et Lemarie and their antioxidant and antidiabetic effects in diabetic rats—in vitro and in vivo studies. Food Funct. 2022, 13, 437–450. [Google Scholar] [CrossRef] [PubMed]
  135. Yan, X.; Qi, M.; Li, P.; Zhan, Y.; Shao, H. Apigenin in cancer therapy: Anti-cancer effects and mechanisms of action. Cell Biosci. 2017, 7, 50. [Google Scholar] [CrossRef] [PubMed]
  136. Alam, W.; Rocca, C.; Khan, H.; Hussain, Y.; Aschner, M.; De Bartolo, A.; Amodio, N.; Angelone, T.; Cheang, W.S. Current Status and Future Perspectives on Therapeutic Potential of Apigenin: Focus on Metabolic-Syndrome-Dependent Organ Dysfunction. Antioxidants 2021, 10, 1643. [Google Scholar] [CrossRef]
  137. Laaroussi, H.; Bakour, M.; Ousaaid, D.; Aboulghazi, A.; Ferreira-Santos, P.; Genisheva, Z.; Teixeira, J.A.; Lyoussi, B. Effect of antioxidant-rich propolis and bee pollen extracts against D-glucose induced type 2 diabetes in rats. Food Res. Int. 2020, 138, 109802. [Google Scholar] [CrossRef]
  138. Bakour, M.; Laaroussi, H.; Ferreira-Santos, P.; Genisheva, Z.; Ousaaid, D.; Teixeira, J.A.; Lyoussi, B. Exploring the Palynological, Chemical, and Bioactive Properties of Non-Studied Bee Pollen and Honey from Morocco. Molecules 2022, 27, 5777. [Google Scholar] [CrossRef] [PubMed]
  139. Luo, Z.; Fu, C.; Li, T.; Gao, Q.; Miao, D.; Xu, J.; Zhao, Y. Hypoglycemic Effects of Licochalcone A on the Streptozotocin-Induced Diabetic Mice and Its Mechanism Study. J. Agric. Food Chem. 2021, 69, 2444–2456. [Google Scholar] [CrossRef]
  140. Luo, Z.; Li, T.; Gao, Q.; Chen, Y.; Su, G.; Zhao, Y. Impact of licochalcone A on the progression of diabetic nephropathy in type 2 diabetes mellitus of C57BL/6 mice. Food Funct. 2021, 12, 10676–10689. [Google Scholar] [CrossRef]
  141. Zhang, B.; Chen, Y.; Shen, Q.; Liu, G.; Ye, J.; Sun, G.; Sun, X. Myricitrin Attenuates High Glucose-Induced Apoptosis through Activating Akt-Nrf2 Signaling in H9c2 Cardiomyocytes. Molecules 2016, 21, 880. [Google Scholar] [CrossRef]
  142. Zhang, B.; Shen, Q.; Chen, Y.; Pan, R.; Kuang, S.; Liu, G.; Sun, G.; Sun, X. Myricitrin Alleviates Oxidative Stress-induced Inflammation and Apoptosis and Protects Mice against Diabetic Cardiomyopathy. Sci. Rep. 2017, 7, 44239. [Google Scholar] [CrossRef]
  143. Ahangarpour, A.; Oroojan, A.A.; Khorsandi, L.; Kouchak, M.; Badavi, M. Solid Lipid Nanoparticles of Myricitrin Have Antioxidant and Antidiabetic Effects on Streptozotocin-Nicotinamide-Induced Diabetic Model and Myotube Cell of Male Mouse. Oxidative Med. Cell. Longev. 2018, 2018, 7496936. [Google Scholar] [CrossRef]
  144. Amri, J.; Alaee, M.; Babaei, R.; Salemi, Z.; Meshkani, R.; Ghazavi, A.; Akbari, A.; Salehi, M. Biochanin-A has antidiabetic, antihyperlipidemic, antioxidant, and protective effects on diabetic nephropathy via suppression of TGF-β1 and PAR-2 genes expression in kidney tissues of STZ-induced diabetic rats. Biotechnol. Appl. Biochem. 2022, 69, 2112–2121. [Google Scholar] [CrossRef]
  145. Ram, C.; Gairola, S.; Verma, S.; Mugale, M.N.; Bonam, S.R.; Murty, U.S.; Sahu, B.D. Biochanin A Ameliorates Nephropathy in High-Fat Diet/Streptozotocin-Induced Diabetic Rats: Effects on NF-kB/NLRP3 Axis, Pyroptosis, and Fibrosis. Antioxidants 2023, 12, 1052. [Google Scholar] [CrossRef]
  146. Oza, M.J.; Kulkarni, Y.A. Biochanin A improves insulin sensitivity and controls hyperglycemia in type 2 diabetes. Biomed. Pharmacother. 2018, 107, 1119–1127. [Google Scholar] [CrossRef] [PubMed]
  147. Sadri, H.; Goodarzi, M.T.; Salemi, Z.; Seifi, M. Antioxidant effects of biochanin A in streptozotocin induced diabetic rats. Braz. Arch. Biol. Technol. 2017, 60. [Google Scholar] [CrossRef]
  148. Mou, X.; Zhou, D.-Y.; Zhou, D.-Y.; Ma, J.-R.; Liu, Y.-H.; Chen, H.-P.; Hu, Y.-B.; Shou, C.-M.; Chen, J.-W.; Liu, W.-H.; et al. Serum TGF-β1 as a Biomarker for Type 2 Diabetic Nephropathy: A Meta-Analysis of Randomized Controlled Trials. PLoS ONE 2016, 11, e0149513. [Google Scholar] [CrossRef] [PubMed]
  149. Bagang, N.; Gupta, K.; Singh, G.; Kanuri, S.H.; Mehan, S. Protease-activated receptors in kidney diseases: A comprehensive review of pathological roles, therapeutic outcomes and challenges. Chem. Biol. Interact. 2023, 377, 110470. [Google Scholar] [CrossRef] [PubMed]
  150. Lossi, L. The concept of intrinsic versus extrinsic apoptosis. Biochem. J. 2022, 479, 357–384. [Google Scholar] [CrossRef]
  151. Tay, K.-C.; Tan, L.T.-H.; Chan, C.K.; Hong, S.L.; Chan, K.-G.; Yap, W.H.; Pusparajah, P.; Lee, L.-H.; Goh, B.-H. Formononetin: A Review of Its Anticancer Potentials and Mechanisms. Front. Pharmacol. 2019, 10, 820. [Google Scholar] [CrossRef]
  152. Jain, P.G.; Nayse, P.G.; Patil, D.J.; Shinde, S.D.; Surana, S.J. The possible antioxidant capabilities of formononetin in guarding against streptozotocin-induced diabetic nephropathy in rats. Future J. Pharm. Sci. 2020, 6, 53. [Google Scholar] [CrossRef]
  153. Hou, X.; Xu, S.; Maitland-Toolan, K.A.; Sato, K.; Jiang, B.; Ido, Y.; Lan, F.; Walsh, K.; Wierzbicki, M.; Verbeuren, T.J.; et al. SIRT1 regulates hepatocyte lipid metabolism through activating AMP-activated protein kinase. J. Biol. Chem. 2008, 283, 20015–20026. [Google Scholar] [CrossRef]
  154. Oza, M.J.; Kulkarni, Y.A. Formononetin treatment in type 2 diabetic rats reduces insulin resistance and hyperglycemia. Front. Pharmacol. 2018, 9, 739. [Google Scholar] [CrossRef] [PubMed]
  155. Zhang, L.; Keung, W.; Samokhvalov, V.; Wang, W.; Lopaschuk, G.D. Role of fatty acid uptake and fatty acid beta-oxidation in mediating insulin resistance in heart and skeletal muscle. Biochim. Biophys. Acta 2010, 1801, 1–22. [Google Scholar] [CrossRef]
  156. Lv, J.; Zhuang, K.; Jiang, X.; Huang, H.; Quan, S. Renoprotective effect of formononetin by suppressing Smad3 expression in Db/Db mice. Diabetes Metab. Syndr. Obes. 2020, 13, 3313–3324. [Google Scholar] [CrossRef] [PubMed]
  157. Zhou, Z.; Zhou, X.; Dong, Y.; Li, M.; Xu, Y. Formononetin ameliorates high glucose-induced endothelial dysfunction by inhibiting the JAK/STAT signaling pathway. Mol. Med. Rep. 2019, 20, 2893–2901. [Google Scholar] [CrossRef]
  158. Lee, A.; Gu, H.; Gwon, M.-H.; Yun, J.-M. Hesperetin suppresses LPS/high glucose-induced inflammatory responses via TLR/MyD88/NF-κB signaling pathways in THP-1 cells. Nutr. Res. Pract. 2021, 15, 591–603. [Google Scholar] [CrossRef]
  159. Tian, M.; Han, Y.-B.; Zhao, C.-C.; Liu, L.; Zhang, F.-L. Hesperidin alleviates insulin resistance by improving HG-induced oxidative stress and mitochondrial dysfunction by restoring miR-149. Diabetol. Metab. Syndr. 2021, 13, 50. [Google Scholar] [CrossRef] [PubMed]
  160. Den Hartogh, D.J.; Tsiani, E. Antidiabetic Properties of Naringenin: A Citrus Fruit Polyphenol. Biomolecules 2019, 9, 99. [Google Scholar] [CrossRef]
  161. García-Cerrillo, D.; Noriega-Cisneros, R.; Peña-Montes, D.; Huerta-Cervantes, M.; Silva-Ríos, M.; Salgado-Garciglia, R.; Montoya-Pérez, R.; Saavedra-Molina, A. Antioxidant effects of Eryngium carlinae in diabetic rats. Asian J. Appl. Sci. 2018, 6. [Google Scholar] [CrossRef]
  162. Pagliari, S.; Forcella, M.; Lonati, E.; Sacco, G.; Romaniello, F.; Rovellini, P.; Fusi, P.; Palestini, P.; Campone, L.; Labra, M.; et al. Antioxidant and Anti-Inflammatory Effect of Cinnamon (Cinnamomum verum J. Presl) Bark Extract after In Vitro Digestion Simulation. Foods 2023, 12, 452. [Google Scholar] [CrossRef]
  163. Stanciauskaite, M.; Marksa, M.; Babickaite, L.; Majiene, D.; Ramanauskiene, K. Comparison of Ethanolic and Aqueous Populus balsamifera L. Bud Extracts by Different Extraction Methods: Chemical Composition, Antioxidant and Antibacterial Activities. Pharmaceuticals 2021, 14, 1018. [Google Scholar] [CrossRef]
  164. Stanciauskaite, M.; Marksa, M.; Rimkiene, L.; Ramanauskiene, K. Evaluation of Chemical Composition, Sun Protection Factor and Antioxidant Activity of Lithuanian Propolis and Its Plant Precursors. Plants 2022, 11, 3558. [Google Scholar] [CrossRef] [PubMed]
  165. Oladimeji, O.H.; Owere, P.O.; Anthony, P.C. Acetylation of Cinnamic Acid and Evaluation of Antioxidant Activity of the Resultant Derivative. Biomed. J. Sci. Tech. Res. 2021, 39, 31084–31088. [Google Scholar] [CrossRef]
  166. Kiliç, I.; Yeşiloğlu, Y. Spectroscopic studies on the antioxidant activity of p-coumaric acid. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2013, 115, 719–724. [Google Scholar] [CrossRef]
  167. Gülçin, I. Antioxidant activity of caffeic acid (3,4-dihydroxycinnamic acid). Toxicology 2006, 217, 213–220. [Google Scholar] [CrossRef]
  168. Bian, Y.Y.; Guo, J.; Majeed, H.; Zhu, K.X.; Guo, X.N.; Peng, W.; Zhou, H.M. Ferulic acid renders protection to HEK293 cells against oxidative damage and apoptosis induced by hydrogen peroxide. In Vitr. Cell. Dev. Biol. Anim. 2015, 51, 722–729. [Google Scholar] [CrossRef] [PubMed]
  169. Nenadis, N.; Tsimidou, M. Observations on the estimation of scavenging activity of phenolic compounds using rapid 1,1-diphenyl-2-picrylhydrazyl (DPPH•) tests. J. Am. Oil Chem. Soc. 2002, 79, 1191. [Google Scholar] [CrossRef]
  170. Chandrasekar, V.; Belur, P.D.; Regupathi, I. Effect of hydroxybenzoic acids antioxidants on the oxidative stability of sardine oil. Resour. Effic. Technol. 2016, 2, S114–S118. [Google Scholar] [CrossRef]
  171. Badhani, B.; Sharma, N.; Kakkar, R. Gallic acid: A versatile antioxidant with promising therapeutic and industrial applications. RSC Adv. 2015, 5, 27540–27557. [Google Scholar] [CrossRef]
  172. Li, Y.; Zhang, L.; Wang, X.; Wu, W.; Qin, R. Effect of Syringic acid on antioxidant biomarkers and associated inflammatory markers in mice model of asthma. Drug Dev. Res. 2019, 80, 253–261. [Google Scholar] [CrossRef] [PubMed]
  173. Madkhali, H.; Tarawneh, A.; Ali, Z.; Le, H.V.; Cutler, S.J.; Khan, I.A.; Shariat-Madar, Z. Identification of Human Kinin-Forming Enzyme Inhibitors from Medicinal Herbs. Molecules 2021, 26, 4126. [Google Scholar] [CrossRef]
  174. Youn, H.S.; Lee, J.Y.; Saitoh, S.I.; Miyake, K.; Kang, K.W.; Choi, Y.J.; Hwang, D.H. Suppression of MyD88- and TRIF-dependent signaling pathways of toll-like receptor by (−)-epigallocatechin-3-gallate, a polyphenol component of green tea. Biochem. Pharmacol. 2006, 72, 850–859. [Google Scholar] [CrossRef] [PubMed]
  175. 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]
  176. Hamza Sherif, S.; Gebreyohannes, B. Synthesis, Characterization, and Antioxidant Activities of Genistein, Biochanin A, and Their Analogues. J. Chem. 2018, 2018, 4032105. [Google Scholar] [CrossRef]
  177. Samie, A.; Sedaghat, R.; Baluchnejadmojarad, T.; Roghani, M. Hesperetin, a citrus flavonoid, attenuates testicular damage in diabetic rats via inhibition of oxidative stress, inflammation, and apoptosis. Life Sci. 2018, 210, 132–139. [Google Scholar] [CrossRef]
  178. Baranowska, M.; Koziara, Z.; Suliborska, K.; Chrzanowski, W.; Wormstone, M.; Namieśnik, J.; Bartoszek, A. Interactions between polyphenolic antioxidants quercetin and naringenin dictate the distinctive redox-related chemical and biological behaviour of their mixtures. Sci. Rep. 2021, 11, 12282. [Google Scholar] [CrossRef]
  179. Rashmi, R.; Bojan Magesh, S.; Mohanram Ramkumar, K.; Suryanarayanan, S.; Venkata SubbaRao, M. Antioxidant Potential of Naringenin Helps to Protect Liver Tissue from Streptozotocin-Induced Damage. Rep. Biochem. Mol. Biol. 2018, 7, 76–84. [Google Scholar]
  180. Tran, M.H.; Nguyen, H.D.; Kim, J.C.; Choi, J.S.; Lee, H.K.; Min, B.S. Phenolic glycosides from Alangium salviifolium leaves with inhibitory activity on LPS-induced NO, PGE(2), and TNF-alpha production. Bioorganic Med. Chem. Lett. 2009, 19, 4389–4393. [Google Scholar] [CrossRef]
  181. Aloud, A.A.; Veeramani, C.; Govindasamy, C.; Alsaif, M.A.; El Newehy, A.S.; Al-Numair, K.S. Galangin, a dietary flavonoid, improves antioxidant status and reduces hyperglycemia-mediated oxidative stress in streptozotocin-induced diabetic rats. Redox Rep. 2017, 22, 290–300. [Google Scholar] [CrossRef]
  182. Maher, P. Modulation of the Neuroprotective and Anti-inflammatory Activities of the Flavonol Fisetin by the Transition Metals Iron and Copper. Antioxidants 2020, 9, 1113. [Google Scholar] [CrossRef] [PubMed]
  183. Awad, H.M.; Boersma, M.G.; Vervoort, J.; Rietjens, I.M. Peroxidase-catalyzed formation of quercetin quinone methide–glutathione adducts. Arch. Biochem. Biophys. 2000, 378, 224–233. [Google Scholar] [CrossRef]
  184. Park, K.-S.; Chong, Y.; Kim, M.K. Myricetin: Biological activity related to human health. Appl. Biol. Chem. 2016, 59, 259–269. [Google Scholar] [CrossRef]
  185. Topal, F.; Nar, M.; Gocer, H.; Kalin, P.; Kocyigit, U.M.; Gülçin, İ.; Alwasel, S.H. Antioxidant activity of taxifolin: An activity–structure relationship. J. Enzym. Inhib. Med. Chem. 2016, 31, 674–683. [Google Scholar] [CrossRef]
  186. Buranasudja, V.; Muangnoi, C.; Sanookpan, K.; Halim, H.; Sritularak, B.; Rojsitthisak, P. Eriodictyol Attenuates H2O2-Induced Oxidative Damage in Human Dermal Fibroblasts through Enhanced Capacity of Antioxidant Machinery. Nutrients 2022, 14, 2553. [Google Scholar] [CrossRef] [PubMed]
  187. Hu, H.-C.; Lei, Y.-H.; Zhang, W.-H.; Luo, X.-Q. Antioxidant and Anti-inflammatory Properties of Resveratrol in Diabetic Nephropathy: A Systematic Review and Meta-analysis of Animal Studies. Front. Pharmacol. 2022, 13, 841818. [Google Scholar] [CrossRef]
  188. Lu, Y.; Xue, Y.; Chen, S.; Zhu, H.; Zhang, J.; Li, X.-N.; Wang, J.; Liu, J.; Qi, C.; Du, G.; et al. Antioxidant Lignans and Neolignans from Acorus tatarinowii. Sci. Rep. 2016, 6, 22909. [Google Scholar] [CrossRef] [PubMed]
  189. Lee, J.; Kim, J.H. Kaempferol Inhibits Pancreatic Cancer Cell Growth and Migration through the Blockade of EGFR-Related Pathway In Vitro. PLoS ONE 2016, 11, e0155264. [Google Scholar] [CrossRef]
  190. Wang, X.; Yang, Y.; An, Y.; Fang, G. The mechanism of anticancer action and potential clinical use of kaempferol in the treatment of breast cancer. Biomed. Pharmacother. 2019, 117, 109086. [Google Scholar] [CrossRef] [PubMed]
  191. Timmer, J.C.; Salvesen, G.S. Caspase substrates. Cell Death Differ. 2007, 14, 66–72. [Google Scholar] [CrossRef] [PubMed]
  192. Alkhalidy, H.; Moore, W.; Zhang, Y.; McMillan, R.; Wang, A.; Ali, M.; Suh, K.-S.; Zhen, W.; Cheng, Z.; Jia, Z.; et al. Small Molecule Kaempferol Promotes Insulin Sensitivity and Preserved Pancreatic β-Cell Mass in Middle-Aged Obese Diabetic Mice. J. Diabetes Res. 2015, 2015, 532984. [Google Scholar] [CrossRef]
  193. Chandramohan, G.; Al-Numair, K.S.; Alsaif, M.A.; Veeramani, C. Antidiabetic effect of kaempferol a flavonoid compound, on streptozotocin-induced diabetic rats with special reference to glycoprotein components. Prog. Nutr. 2015, 17, 50–57. [Google Scholar]
  194. 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] [PubMed]
  195. 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]
  196. Yang, Y.; Chen, Z.; Zhao, X.; Xie, H.; Du, L.; Gao, H.; Xie, C. Mechanisms of Kaempferol in the treatment of diabetes: A comprehensive and latest review. Front. Endocrinol. 2022, 13, 990299. [Google Scholar] [CrossRef]
  197. Abukhalil, M.H.; Althunibat, O.Y.; Aladaileh, S.H.; Al-Amarat, W.; Obeidat, H.M.; Alayn’Al-marddyah, A.; Hussein, O.E.; Alfwuaires, M.A.; Algefare, A.I.; Alanazi, K.M. Galangin attenuates diabetic cardiomyopathy through modulating oxidative stress, inflammation and apoptosis in rats. Biomed. Pharmacother. 2021, 138, 111410. [Google Scholar] [CrossRef]
  198. Al-Amarat, W.; Abukhalil, M.H.; Althunibat, O.Y.; Alfwuaires, M.A.; Alnamshan, M.M.; Alqosaibi, A.I.; Ahmeda, A.F.; Kamel, E.M.; Arab, H.H.; Mahmoud, A.M. Galangin attenuates liver injury, oxidative stress and inflammation, and upregulates Nrf2/HO-1 signaling in streptozotocin-induced diabetic rats. Processes 2021, 9, 1562. [Google Scholar] [CrossRef]
  199. Kong, Y.; Feng, Z.; Chen, A.; Qi, Q.; Han, M.; Wang, S.; Zhang, Y.; Zhang, X.; Yang, N.; Wang, J.; et al. The Natural Flavonoid Galangin Elicits Apoptosis, Pyroptosis, and Autophagy in Glioblastoma. Front. Oncol. 2019, 9, 942. [Google Scholar] [CrossRef]
  200. Kalhotra, P.; Chittepu, V.C.; Osorio-Revilla, G.; Gallardo-Velázquez, T. Discovery of galangin as a potential DPP-4 inhibitor that improves insulin-stimulated skeletal muscle glucose uptake: A combinational therapy for diabetes. Int. J. Mol. Sci. 2019, 20, 1228. [Google Scholar] [CrossRef] [PubMed]
  201. Deacon, C.F. Physiology and Pharmacology of DPP-4 in Glucose Homeostasis and the Treatment of Type 2 Diabetes. Front. Endocrinol. 2019, 10, 80. [Google Scholar] [CrossRef]
  202. Aloud, A.A.; Chinnadurai, V.; Govindasamy, C.; Alsaif, M.A.; Al-Numair, K.S. Galangin, a dietary flavonoid, ameliorates hyperglycaemia and lipid abnormalities in rats with streptozotocin-induced hyperglycaemia. Pharm. Biol. 2018, 56, 302–308. [Google Scholar] [CrossRef] [PubMed]
  203. Haddad, A.Q.; Venkateswaran, V.; Viswanathan, L.; Teahan, S.J.; Fleshner, N.E.; Klotz, L.H. Novel antiproliferative flavonoids induce cell cycle arrest in human prostate cancer cell lines. Prostate Cancer Prostatic Dis. 2006, 9, 68–76. [Google Scholar] [CrossRef]
  204. Sung, B.; Pandey, M.K.; Aggarwal, B.B. Fisetin, an inhibitor of cyclin-dependent kinase 6, down-regulates nuclear factor-kappaB-regulated cell proliferation, antiapoptotic and metastatic gene products through the suppression of TAK-1 and receptor-interacting protein-regulated IkappaBalpha kinase activation. Mol. Pharmacol. 2007, 71, 1703–1714. [Google Scholar] [CrossRef] [PubMed]
  205. Hou, D.X.; Fukuda, M.; Johnson, J.A.; Miyamori, K.; Ushikai, M.; Fujii, M. Fisetin induces transcription of NADPH:quinone oxidoreductase gene through an antioxidant responsive element-involved activation. Int. J. Oncol. 2001, 18, 1175–1179. [Google Scholar] [CrossRef] [PubMed]
  206. Constantin, R.P.; Constantin, J.; Pagadigorria, C.L.; Ishii-Iwamoto, E.L.; Bracht, A.; Ono, M.D.; Yamamoto, N.S. The actions of fisetin on glucose metabolism in the rat liver. Cell Biochem. Funct. 2010, 28, 149–158. [Google Scholar] [CrossRef]
  207. Vinayagam, R.; Xu, B. Antidiabetic properties of dietary flavonoids: A cellular mechanism review. Nutr. Metab. 2015, 12, 60. [Google Scholar] [CrossRef]
  208. Prasath, G.S.; Pillai, S.I.; Subramanian, S.P. Fisetin improves glucose homeostasis through the inhibition of gluconeogenic enzymes in hepatic tissues of streptozotocin induced diabetic rats. Eur. J. Pharmacol. 2014, 740, 248–254. [Google Scholar] [CrossRef] [PubMed]
  209. Li, Y.; Ding, Y. Minireview: Therapeutic potential of myricetin in diabetes mellitus. Food Sci. Hum. Wellness 2012, 1, 19–25. [Google Scholar] [CrossRef]
  210. Lalitha, N.; Sadashivaiah, B.; Ramaprasad, T.R.; Singh, S.A. Anti-hyperglycemic activity of myricetin, through inhibition of DPP-4 and enhanced GLP-1 levels, is attenuated by co-ingestion with lectin-rich protein. PLoS ONE 2020, 15, e0231543. [Google Scholar] [CrossRef] [PubMed]
  211. 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]
  212. Zhao, Z.; Chen, Y.; Li, X.; Zhu, L.; Wang, X.; Li, L.; Sun, H.; Han, X.; Li, J. Myricetin relieves the symptoms of type 2 diabetes mice and regulates intestinal microflora. Biomed. Pharmacother. 2022, 153, 113530. [Google Scholar] [CrossRef] [PubMed]
  213. Les, F.; Casedas, G.; Gomez, C.; Moliner, C.; Valero, M.S.; Lopez, V. The role of anthocyanins as antidiabetic agents: From molecular mechanisms to in vivo and human studies. J. Physiol. Biochem. 2021, 77, 109–131. [Google Scholar] [CrossRef] [PubMed]
  214. Casedas, G.; Les, F.; Gomez-Serranillos, M.P.; Smith, C.; Lopez, 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]
  215. Khan, D.; Sharif, A.; Zafar, M.; Akhtar, B.; Akhtar, M.F.; Awan, S. Delonix regia a Folklore Remedy for Diabetes; Attenuates Oxidative Stress and Modulates Type II Diabetes Mellitus. Curr. Pharm. Biotechnol. 2020, 21, 1059–1069. [Google Scholar] [CrossRef]
  216. Meng, Q.; Qi, X.; Fu, Y.; Chen, Q.; Cheng, P.; Yu, X.; Sun, X.; Wu, J.; Li, W.; Zhang, Q.; et al. Flavonoids extracted from mulberry (Morus alba L.) leaf improve skeletal muscle mitochondrial function by activating AMPK in type 2 diabetes. J. Ethnopharmacol. 2020, 248, 112326. [Google Scholar] [CrossRef] [PubMed]
  217. Thaipitakwong, T.; Numhom, S.; Aramwit, P. Mulberry leaves and their potential effects against cardiometabolic risks: A review of chemical compositions, biological properties and clinical efficacy. Pharm. Biol. 2018, 56, 109–118. [Google Scholar] [CrossRef] [PubMed]
  218. Thaipitakwong, T.; Supasyndh, O.; Rasmi, Y.; Aramwit, P. A randomized controlled study of dose-finding, efficacy, and safety of mulberry leaves on glycemic profiles in obese persons with borderline diabetes. Complement. Ther. Med. 2020, 49, 102292. [Google Scholar] [CrossRef] [PubMed]
  219. Zhang, Q.; Lu, Y.; Ma, Z.; Li, Y.; Guo, J.; Meng, Q.; Bian, H. A novel formula from mulberry leaf ameliorates diabetic nephropathy in rats via inhibiting the TGF-beta1 pathway. Food Funct. 2015, 6, 3307–3315. [Google Scholar] [CrossRef] [PubMed]
  220. Xu, F.; Liu, Y.; Zhu, X.; Li, S.; Shi, X.; Li, Z.; Ai, M.; Sun, J.; Hou, B.; Cai, W.; et al. Protective Effects and Mechanisms of Vaccarin on Vascular Endothelial Dysfunction in Diabetic Angiopathy. Int. J. Mol. Sci. 2019, 20, 4587. [Google Scholar] [CrossRef]
  221. Li, T.; Yu, X.; Zhu, X.; Wen, Y.; Zhu, M.; Cai, W.; Hou, B.; Xu, F.; Qiu, L. Vaccarin alleviates endothelial inflammatory injury in diabetes by mediating miR-570-3p/HDAC1 pathway. Front. Pharmacol. 2022, 13, 956247. [Google Scholar] [CrossRef]
  222. Li, X.Y.; Lu, S.S.; Wang, H.L.; Li, G.; He, Y.F.; Liu, X.Y.; Rong, R.; Li, J.; Lu, X.C. Effects of the fenugreek extracts on high-fat diet-fed and streptozotocin-induced type 2 diabetic mice. Anim. Model. Exp. Med. 2018, 1, 68–73. [Google Scholar] [CrossRef]
  223. Rehman, K.; Chohan, T.A.; Waheed, I.; Gilani, Z.; Akash, M.S.H. Taxifolin prevents postprandial hyperglycemia by regulating the activity of alpha-amylase: Evidence from an in vivo and in silico studies. J. Cell. Biochem. 2019, 120, 425–438. [Google Scholar] [CrossRef] [PubMed]
  224. Sun, X.; Chen, R.C.; Yang, Z.H.; Sun, G.B.; Wang, M.; Ma, X.J.; Yang, L.J.; Sun, X.B. Taxifolin prevents diabetic cardiomyopathy in vivo and in vitro by inhibition of oxidative stress and cell apoptosis. Food Chem. Toxicol. 2014, 63, 221–232. [Google Scholar] [CrossRef]
  225. Konate, K.; Yomalan, K.; Sytar, O.; Zerbo, P.; Brestic, M.; Patrick, V.D.; Gagniuc, P.; Barro, N. Free Radicals Scavenging Capacity, Antidiabetic and Antihypertensive Activities of Flavonoid-Rich Fractions from Leaves of Trichilia emetica and Opilia amentacea in an Animal Model of Type 2 Diabetes Mellitus. Evid. Based Complement. Altern. Med. 2014, 2014, 867075. [Google Scholar] [CrossRef] [PubMed]
  226. Xu, J.; Jordan, R.B. Kinetics and mechanism of the oxidation of 2,3-dihydroxybenzoic acid by iron(III). Inorg. Chem. 1988, 27, 4563–4566. [Google Scholar] [CrossRef]
  227. Álvarez-Cilleros, D.; Martín, M.; Ramos, S. (-)-Epicatechin and the Colonic 2,3-Dihydroxybenzoic Acid Metabolite Regulate Glucose Uptake, Glucose Production, and Improve Insulin Signaling in Renal NRK-52E Cells. Mol. Nutr. Food Res. 2018, 62, 1700470. [Google Scholar] [CrossRef] [PubMed]
  228. Alvarez Cilleros, D.; Lopez-Oliva, M.E.; Martin, M.A.; Ramos, S. (-)-Epicatechin and the colonic metabolite 2,3-dihydroxybenzoic acid protect against high glucose and lipopolysaccharide-induced inflammation in renal proximal tubular cells through NOX-4/p38 signalling. Food Funct. 2020, 11, 8811–8824. [Google Scholar] [CrossRef]
  229. Bernatoniene, J.; Kopustinskiene, D.M. The Role of Catechins in Cellular Responses to Oxidative Stress. Molecules 2018, 23, 965. [Google Scholar] [CrossRef]
  230. Meng, Z.; Jing, H.; Gan, L.; Li, H.; Luo, B. Resveratrol attenuated estrogen-deficient-induced cardiac dysfunction: Role of AMPK, SIRT1, and mitochondrial function. Am. J. Transl. Res. 2016, 8, 2641–2649. [Google Scholar]
  231. Zhang, L.X.; Li, C.X.; Kakar, M.U.; Khan, M.S.; Wu, P.F.; Amir, R.M.; Dai, D.F.; Naveed, M.; Li, Q.Y.; Saeed, M.; et al. Resveratrol (RV): A pharmacological review and call for further research. Biomed. Pharmacother. 2021, 143, 112164. [Google Scholar] [CrossRef]
  232. Abu-Amero, K.K.; Kondkar, A.A.; Chalam, K.V. Resveratrol and Ophthalmic Diseases. Nutrients 2016, 8, 200. [Google Scholar] [CrossRef]
  233. Michalak, B.; Filipek, A.; Chomicki, P.; Pyza, M.; Woźniak, M.; Żyżyńska-Granica, B.; Piwowarski, J.P.; Kicel, A.; Olszewska, M.A.; Kiss, A.K. Lignans From Forsythia x Intermedia Leaves and Flowers Attenuate the Pro-inflammatory Function of Leukocytes and Their Interaction With Endothelial Cells. Front. Pharmacol. 2018, 9, 401. [Google Scholar] [CrossRef]
  234. Kuryłowicz, A. The Role of Isoflavones in Type 2 Diabetes Prevention and Treatment-A Narrative Review. Int. J. Mol. Sci. 2020, 22, 218. [Google Scholar] [CrossRef] [PubMed]
  235. Sorrenti, V.; Raffaele, M.; Vanella, L.; Acquaviva, R.; Salerno, L.; Pittala, V.; Intagliata, S.; Di Giacomo, C. Protective Effects of Caffeic Acid Phenethyl Ester (CAPE) and Novel Cape Analogue as Inducers of Heme Oxygenase-1 in Streptozotocin-Induced Type 1 Diabetic Rats. Int. J. Mol. Sci. 2019, 20, 2441. [Google Scholar] [CrossRef]
  236. Kose, T.; Sharp, P.A.; Latunde-Dada, G.O. Phenolic Acids Rescue Iron-Induced Damage in Murine Pancreatic Cells and Tissues. Molecules 2023, 28, 4084. [Google Scholar] [CrossRef] [PubMed]
  237. Wen, L.; Wu, D.; Tan, X.; Zhong, M.; Xing, J.; Li, W.; Li, D.; Cao, F. The Role of Catechins in Regulating Diabetes: An Update Review. Nutrients 2022, 14, 4681. [Google Scholar] [CrossRef]
  238. Yang, J.; Zhu, C.; Ye, J.; Lv, Y.; Wang, L.; Chen, Z.; Jiang, Z. Protection of Porcine Intestinal-Epithelial Cells from Deoxynivalenol-Induced Damage by Resveratrol via the Nrf2 Signaling Pathway. J. Agric. Food Chem. 2019, 67, 1726–1735. [Google Scholar] [CrossRef] [PubMed]
  239. Draganescu, D.; Andritoiu, C.; Hritcu, D.; Dodi, G.; Popa, M.I. Flaxseed Lignans and Polyphenols Enhanced Activity in Streptozotocin-Induced Diabetic Rats. Biology 2021, 10, 43. [Google Scholar] [CrossRef]
  240. Fukumitsu, S.; Aida, K.; Shimizu, H.; Toyoda, K. Flaxseed lignan lowers blood cholesterol and decreases liver disease risk factors in moderately hypercholesterolemic men. Nutr. Res. 2010, 30, 441–446. [Google Scholar] [CrossRef]
  241. Sugihara, N.; Ohnishi, M.; Imamura, M.; Furuno, K. Differences in antioxidative efficiency of catechins in various metal-induced lipid peroxidations in cultured hepatocytes. J. Health Sci. 2001, 47, 99–106. [Google Scholar] [CrossRef]
  242. Sohal, R.S.; Arnold, L.A.; Sohal, B.H. Age-related changes in antioxidant enzymes and prooxidant generation in tissues of the rat with special reference to parameters in two insect species. Free Radic. Biol. Med. 1990, 9, 495–500. [Google Scholar] [CrossRef] [PubMed]
  243. Lopez-Torres, M.; Perez-Campo, R.; Barja de Quiroga, G. Aging in brown fat: Antioxidant defenses and oxidative stress. Mech. Ageing Dev. 1991, 59, 129–137. [Google Scholar] [CrossRef]
  244. Lopez-Torres, M.; Perez-Campo, R.; Barja de Quiroga, G. Effect of natural ageing and antioxidant inhibition on liver antioxidant enzymes, glutathione system, peroxidation, and oxygen consumption in Rana perezi. J. Comp. Physiol. B 1991, 160, 655–661. [Google Scholar] [CrossRef]
  245. Shih, P.-H.; Yeh, C.-T.; Yen, G.-C. Anthocyanins Induce the Activation of Phase II Enzymes through the Antioxidant Response Element Pathway against Oxidative Stress-Induced Apoptosis. J. Agric. Food Chem. 2007, 55, 9427–9435. [Google Scholar] [CrossRef]
  246. Renjini, R.; Gayathri, N.; Nalini, A.; Srinivas Bharath, M.M. Oxidative damage in muscular dystrophy correlates with the severity of the pathology: Role of glutathione metabolism. Neurochem. Res. 2012, 37, 885–898. [Google Scholar] [CrossRef] [PubMed]
  247. Samanta, L.; Roy, A.; Chainy, G.B. Changes in rat testicular antioxidant defence profile as a function of age and its impairment by hexachlorocyclohexane during critical stages of maturation. Andrologia 1999, 31, 83–90. [Google Scholar] [CrossRef] [PubMed]
  248. Giusti, L.; Gabriele, M.; Penno, G.; Garofolo, M.; Longo, V.; Del Prato, S.; Lucchesi, D.; Pucci, L. A Fermented Whole Grain Prevents Lipopolysaccharides-Induced Dysfunction in Human Endothelial Progenitor Cells. Oxid. Med. Cell Longev. 2017, 2017, 1026268. [Google Scholar] [CrossRef]
  249. Rammohan, A.; Zyryanov, G.V.; Bhagath, Y.B.; Manjula, K. Antioxidants: Structure-activity of plant polyphenolics. Vitam. Horm. 2023, 121, 395–411. [Google Scholar] [CrossRef] [PubMed]
  250. Stryjecka, M.; Krochmal-Marczak, B.; Cebulak, T.; Kiełtyka-Dadasiewicz, A. Assessment of Phenolic Acid Content and Antioxidant Properties of the Pulp of Five Pumpkin Species Cultivated in Southeastern Poland. Int. J. Mol. Sci. 2023, 24, 8621. [Google Scholar] [CrossRef] [PubMed]
  251. Uchiyama, Y.; Suzuki, T.; Mochizuki, K.; Goda, T. Dietary supplementation with a low dose of (-)-epigallocatechin-3-gallate reduces pro-inflammatory responses in peripheral leukocytes of non-obese type 2 diabetic GK rats. J. Nutr. Sci. Vitaminol. 2013, 59, 541–547. [Google Scholar] [CrossRef]
  252. Uchiyama, Y.; Suzuki, T.; Mochizuki, K.; Goda, T. Dietary supplementation with (-)-epigallocatechin-3-gallate reduces inflammatory response in adipose tissue of non-obese type 2 diabetic Goto-Kakizaki (GK) rats. J. Agric. Food Chem. 2013, 61, 11410–11417. [Google Scholar] [CrossRef]
  253. Elbling, L.; Weiss, R.M.; Teufelhofer, O.; Uhl, M.; Knasmueller, S.; Schulte-Hermann, R.; Berger, W.; Micksche, M. Green tea extract and (-)-epigallocatechin-3-gallate, the major tea catechin, exert oxidant but lack antioxidant activities. FASEB J. 2005, 19, 807–809. [Google Scholar] [CrossRef]
  254. Ahangarpour, A.; Afshari, G.; Mard, S.A.; Khodadadi, A.; Hashemitabar, M. Preventive effects of procyanidin A2 on glucose homeostasis, pancreatic and duodenal homebox 1, and glucose transporter 2 gene expression disturbance induced by bisphenol A in male mice. J. Physiol. Pharmacol. 2016, 67, 243–252. [Google Scholar]
  255. Memon, A.R.; Rajput, M.A.; Rizwan, F.; Akram, M.; Rizwan, M.; Iqbal, Z. Effect of Allium sativum and Olea europaea on serum lipids in patients with diabetes mellitus. J. Taibah Univ. Med. Sci. 2023, 18, 420–426. [Google Scholar] [CrossRef] [PubMed]
  256. Ciaraldi, T.P.; Boeder, S.C.; Mudaliar, S.R.; Giovannetti, E.R.; Henry, R.R.; Pettus, J.H. Astaxanthin, a natural antioxidant, lowers cholesterol and markers of cardiovascular risk in individuals with prediabetes and dyslipidaemia. Diabetes Obes. Metab. 2023, 25, 1985–1994. [Google Scholar] [CrossRef]
  257. Karimzadeh, L.; Sohrab, G.; Hedayati, M.; Ebrahimof, S.; Emami, G.; Razavion, T. Effects of concentrated beetroot juice consumption on glycemic control, blood pressure, and lipid profile in type 2 diabetes patients: Randomized clinical trial study. Ir. J. Med. Sci. 2023, 192, 1143–1153. [Google Scholar] [CrossRef]
  258. Rostamkhani, H.; Veisi, P.; Niknafs, B.; Jafarabadi, M.A.; Ghoreishi, Z. The effect of zingiber officinale on prooxidant-antioxidant balance and glycemic control in diabetic patients with ESRD undergoing hemodialysis: A double-blind randomized control trial. BMC Complement. Med. Ther. 2023, 23, 52. [Google Scholar] [CrossRef] [PubMed]
  259. Zhang, A.; Wang, Q.; Liu, M.; Tan, M.; Zhang, X.; Wu, R. Efficacy and safety of Mudan granules for painful diabetic peripheral neuropathy: A protocol for a double-blind randomized controlled trial. Medicine 2022, 101, e28896. [Google Scholar] [CrossRef]
  260. Garcia-Martinez, B.I.; Ruiz-Ramos, M.; Pedraza-Chaverri, J.; Santiago-Osorio, E.; Mendoza-Nunez, V.M. Effect of Resveratrol on Markers of Oxidative Stress and Sirtuin 1 in Elderly Adults with Type 2 Diabetes. Int. J. Mol. Sci. 2023, 24, 7422. [Google Scholar] [CrossRef]
  261. Fatima, S.; Khan, D.A.; Aamir, M.; Pervez, M.A.; Fatima, F. delta-Tocotrienol in Combination with Resveratrol Improves the Cardiometabolic Risk Factors and Biomarkers in Patients with Metabolic Syndrome: A Randomized Controlled Trial. Metab. Syndr. Relat. Disord. 2023, 21, 25–34. [Google Scholar] [CrossRef] [PubMed]
  262. Bazyar, H.; Moradi, L.; Zaman, F.; Zare Javid, A. The effects of rutin flavonoid supplement on glycemic status, lipid profile, atherogenic index of plasma, brain-derived neurotrophic factor (BDNF), some serum inflammatory, and oxidative stress factors in patients with type 2 diabetes mellitus: A double-blind, placebo-controlled trial. Phytother. Res. 2023, 37, 271–284. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Pathological ROS in diabetes.
Figure 1. Pathological ROS in diabetes.
Molecules 28 07209 g001
Figure 2. Diabetes is associated with vascular dysfunction in multiple ways. Uncontrolled free radicals are the central components of cell injury. Antioxidants may reduce or prevent ROS/RNS-induced vascular complications. ↑—indicates an increase in level.
Figure 2. Diabetes is associated with vascular dysfunction in multiple ways. Uncontrolled free radicals are the central components of cell injury. Antioxidants may reduce or prevent ROS/RNS-induced vascular complications. ↑—indicates an increase in level.
Molecules 28 07209 g002
Figure 3. Feed-forward (black arrows) and feed-back (red arrows) events where diabetes can contribute to ROS formation. Hyperglycemia, dyslipidemia, hypertension, kidney disease, and inflammation all lead to the formation of ROS. As a consequence, elevated levels of ROS might directly contribute to fibrosis, apoptosis, and dedifferentiation, or indirectly via enhanced inflammation, reduced FFA β-oxidation, or NO bioavailability. Diabetes itself may contribute to the generation of cytokines and vascular remodeling, which can also contribute to the generation of ROS. Under normal circumstances (non-diabetic response), ROS and RNS play vital roles in numerous biological processes. ↑—indicates an increase in level; ↓—represents a decrease in level.
Figure 3. Feed-forward (black arrows) and feed-back (red arrows) events where diabetes can contribute to ROS formation. Hyperglycemia, dyslipidemia, hypertension, kidney disease, and inflammation all lead to the formation of ROS. As a consequence, elevated levels of ROS might directly contribute to fibrosis, apoptosis, and dedifferentiation, or indirectly via enhanced inflammation, reduced FFA β-oxidation, or NO bioavailability. Diabetes itself may contribute to the generation of cytokines and vascular remodeling, which can also contribute to the generation of ROS. Under normal circumstances (non-diabetic response), ROS and RNS play vital roles in numerous biological processes. ↑—indicates an increase in level; ↓—represents a decrease in level.
Molecules 28 07209 g003
Figure 4. An illustration depicting how the antioxidant properties of polyphenols were measured.
Figure 4. An illustration depicting how the antioxidant properties of polyphenols were measured.
Molecules 28 07209 g004
Table 4. Summary of the acute effects of supplemental dietary antioxidant therapy for the prevention and treatment of diabetes in a double-blind placebo-controlled study [255,256,257,258,259,260,261,262].
Table 4. Summary of the acute effects of supplemental dietary antioxidant therapy for the prevention and treatment of diabetes in a double-blind placebo-controlled study [255,256,257,258,259,260,261,262].
Dietary SourceDose
(Daily)
T2DM
Patients
Inflammatory and Oxidative Stress MarkersDiabetes MarkersCardiovascular Risk MarkersRenal
Markers
Ref.
Allium sativum and Olea europaea oil 160 patients Decreased serum cholesterol, triglycerides (TGs), and low-density lipoprotein (LDL)
Increased High-density lipoprotein (HDL) levels
[255]
Astaxanthin12 mg24 patients Not a significant change in glucose and insulin levels Decreased low-density lipoprotein and total cholesterolReduced fibrinogen, L-selectin, and fetuin-A [256]
Beetroot juice (concentrated)24 ml46 patients No effect [257]
Ginger2000 mg44 hemodialysis patients with end-stage renal diseaseA higher nonsignificant increase nitric oxideReduced blood glucose levels, decreased homeostatic model evaluation of insulin resistance (HOMA)-IR Decreased creatinine[258]
Mudan granules3 times93 patients with painful diabetic neuropathyReduced pain and numbness in the extremities
Increased nerve conduction velocity
[259]
Resveratrol1000 mg97 older adults with T2DReduced lipoperoxides, isoprostanes and C-reactive protein levelsHbA1c level Hypolipemic effect [260]
Resveratrol plus delta-tocotrienol400 mg capsule (delta-tocotrienol 250 mg; resveratrol 150 mg)82 patients with metabolic syndromeReduced C-reactive protein, interleukin-6, tumor necrosis factor-alpha, malondialdehydeReduced fasting plasma glucose, serum triglyceride [261]
Rutin flavonoid 50 patientsDecreased interleukin 6, total antioxidant capacity, and malondialdehydeDecreased FBG, insulin, HbA1c, homeostasis model assessment of insulin resistance (HOMA)-IR; Decreased LDL-c, triglyceride (TG) [262]
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

Arabshomali, A.; Bazzazzadehgan, S.; Mahdi, F.; Shariat-Madar, Z. Potential Benefits of Antioxidant Phytochemicals in Type 2 Diabetes. Molecules 2023, 28, 7209. https://doi.org/10.3390/molecules28207209

AMA Style

Arabshomali A, Bazzazzadehgan S, Mahdi F, Shariat-Madar Z. Potential Benefits of Antioxidant Phytochemicals in Type 2 Diabetes. Molecules. 2023; 28(20):7209. https://doi.org/10.3390/molecules28207209

Chicago/Turabian Style

Arabshomali, Arman, Shadi Bazzazzadehgan, Fakhri Mahdi, and Zia Shariat-Madar. 2023. "Potential Benefits of Antioxidant Phytochemicals in Type 2 Diabetes" Molecules 28, no. 20: 7209. https://doi.org/10.3390/molecules28207209

Article Metrics

Back to TopTop