Next Article in Journal
Integrated System Pharmacology Approaches to Elucidate Multi-Target Mechanism of Solanum surattense against Hepatocellular Carcinoma
Next Article in Special Issue
GLP-1R Signaling and Functional Molecules in Incretin Therapy
Previous Article in Journal
Biogenic Synthesis of Silver Nanoparticles Using Catharanthus roseus and Its Cytotoxicity Effect on Vero Cell Lines
Previous Article in Special Issue
Isolation of Chalcomoracin as a Potential α-Glycosidase Inhibitor from Mulberry Leaves and Its Binding Mechanism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Mechanistic Pathways Targeted by Natural Compounds in the Prevention and Treatment of Diabetic Kidney Disease

College of Pharmacy, Changchun University of Chinese Medicine, Changchun 130117, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(19), 6221; https://doi.org/10.3390/molecules27196221
Submission received: 6 September 2022 / Revised: 18 September 2022 / Accepted: 19 September 2022 / Published: 21 September 2022
(This article belongs to the Special Issue Antidiabetic Natural Products)

Abstract

:
Diabetic kidney disease (DKD) is one of the most common complications of diabetes, and its prevalence is still growing rapidly. However, the efficient therapies for this kidney disease are still limited. The pathogenesis of DKD involves glucotoxicity, lipotoxicity, inflammation, oxidative stress, and renal fibrosis. Glucotoxicity and lipotoxicity can cause oxidative stress, which can lead to inflammation and aggravate renal fibrosis. In this review, we have focused on in vitro and in vivo experiments to investigate the mechanistic pathways by which natural compounds exert their effects against the progression of DKD. The accumulated and collected data revealed that some natural compounds could regulate inflammation, oxidative stress, renal fibrosis, and activate autophagy, thereby protecting the kidney. The main pathways targeted by these reviewed compounds include the Nrf2 signaling pathway, NF-κB signaling pathway, TGF-β signaling pathway, NLRP3 inflammasome, autophagy, glycolipid metabolism and ER stress. This review presented an updated overview of the potential benefits of these natural compounds for the prevention and treatment of DKD progression, aimed to provide new potential therapeutic lead compounds and references for the innovative drug development and clinical treatment of DKD.

1. Introduction

Diabetic kidney disease (DKD) is one of the most common complications of diabetes, and by 2040, more than 600 million people worldwide are expected to have diabetes, of whom 30–40% will develop DKD. It has become the major cause of chronic kidney disease in many developed and developing countries [1,2]. DKD is the most common cause of end-stage renal disease (ESRD) worldwide, chronic kidney disease is the dominant contributor to excess mortality in patients with type 2 diabetes [3]. Recently, sodium-glucose cotransporter-2 inhibitors (SGLT2i), glucagon-like peptide-1 receptor agonists (GLP-1 RAs), and dipeptidyl peptidase-4 inhibitors (DPP-4i) have been considered as new therapeutic options for DKD in the guidelines and consensus of some countries [4,5,6,7]. They have been proven to significantly improve the control of blood glucose and hemoglobin A1c (HbA1c), and prevent the onset of severe increased albuminuria [8]. But the use of these drugs has some limitations and adverse reactions. The adverse reactions of SGLT2i include genital mycotic infections and diabetic ketoacidosis [9,10]. The adverse reactions of GLP-1 RAs mainly include gastrointestinal damage and several cases of acute tubular injury [11]. DPP-4i causes sympathetic activation and enhanced Ca++/calmodulin-dependent protein kinase II signaling, which increases the risk of heart failure and arrhythmia in diabetic patients [12].
In recent years, some studies have shown that some natural compounds have the effects of lowering blood glucose, lowering blood lipid and anti-oxidation, and have good effects in the treatment of diabetes complications [13,14]. In this review, we searched the PubMed database and Google scholar with “(Diabetes nephropathy) OR (Diabetic kidney disease)”, found a recently published article on the treatment of DKD with a natural compound, and then searched with “((Diabetes nephropathy) OR (Diabetic kidney disease)) AND (compound)” to find all the studies on the treatment of DKD with this compound, and selected articles with in-depth mechanism research for review.
In this review, we aim to provide updated and comprehensive insights into the in vitro and in vivo experimental of natural compounds for DKD treatment and emphasize the potential mechanisms and molecular targets, especially those signaling pathways involved in metabolism regulation, anti-oxidation, anti-inflammatory, and anti-fibrosis, provide new potential therapeutic lead compounds and references for the innovative drug development and clinical treatment of DKD.

2. Signaling Pathways for DKD Progression

DKD is considered to be one of the serious long-term complications of diabetes affecting microvascular. Although DKD is a multifactorial disease with complex mechanisms, its pathogenesis could be initially explained by glucose and lipid metabolism disorders [15,16]. In diabetes, reactive oxygen species (ROS) is one of the important factors leading to renal injury. NADPH oxidase is widely expressed in the kidney and is the main source of oxidative stress in the kidney [17]. NADPH oxidase 4 (NOX4) is the main NADPH isoform in the kidney, which produces H2O2 that regulates physiological functions [18]. Hyperglycemia, hyperlipidemia, and other stimuli upregulate NOX4 expression in renal cells, leading to excessive ROS production [19,20]. Oxidative stress is related to the recruitment of inflammatory cells, inducing the production of proinflammatory factors, which leads to the activation of a variety of inflammatory pathways, such as nuclear factor-kappaB (NF-κB) signaling pathway, NACHT, LRR, and PYD domains-containing protein 3 (NLRP3) inflammasome signaling pathway, and other inflammatory pathways [21,22,23]. Proinflammatory factors can stimulate the expression of cytokines such as transforming growth factor-beta (TGF-β) from multiple pathways [24]. TGF-β is a key factor leading to fibrosis in most chronic kidney diseases, and the Smad signal is the key downstream regulator. Smad3 is strongly activated during fibrosis formation, while Smad7 is down-regulated, thereby promoting the expression of collagen I, collagen IV, and fibronectin (FN), resulting in increased extracellular matrix (ECM) production and decreased ECM degradation [25,26]. Under oxidative stress, the increase of ROS levels will lead to mitochondrial DNA breakage, causing a decrease in ATP production, which will affect the physiological function of the kidney [27]. In addition to oxidative stress-mediated kidney damage, endoplasmic reticulum (ER) stress also plays an important role in the development of DKD. Hyperglycemia and free fatty acids disturb the proteostasis, which leads to the accumulation of unfolded/misfolded proteins in the ER lumen, under ER stress, activating transcription factor 4 (ATF4), inositol requiring enzyme 1alpha (IRE1α) and protein kinase R-like endoplasmic reticulum kinase (PERK) are activated, which leads to the activation of downstream pro-apoptotic genes (caspase-4 or caspase-12), resulting in apoptosis [28]. Nuclear factor erythroid 2-related factor 2 (Nrf2) is a transcription factor that regulates genes that encode antioxidants, and it works in the nucleus. Kelch-like ECH-associated protein 1 (Keap1) binds to Nrf2 in the cytoplasm, preventing it from entering the nucleus. In the process of kidney diseases such as DKD, excessive ROS causes renal function damage, Keap1 increases, and Nrf2 activation decreases, reduce the production of antioxidant related proteins, such as superoxide dismutase(SOD)-1, heme oxygenase (HO)-1, and catalase (CAT) [29]. Autophagy can suppress inflammasome activity and clear damaged organelles in cells [30]. When kidney cells are exposed to oxidative stress and ER stress, autophagy activation plays a vital role in cell survival [31]. Autophagy also participates in the intracellular degradation of collagen, prevents the continuous accumulation of ECM, and plays a crucial role in inhibiting renal fibrosis [32]. Autophagy is mainly regulated by the mechanistic target of rapamycin (mTOR) signaling pathway, and its upstream is managed by AMP-activated protein kinase (AMPK). In DKD, renal AMPK activation is inhibited, which leads to the reduction of autophagy and the inability to clear damaged organelles [33].
The development of DKD can be roughly described as follows: in the early stage of DKD, hyperglycemia and the disorder of glucose and lipid metabolism caused by hyperglycemia gradually promote the structural and functional changes in the kidney, such as hyperfiltration, basement membrane thickening, glomerular mesangial matrix expansion, glomerular and tubular hypertrophy, and microalbuminuria [34,35,36]. Hyperglycemia is frequently accompanied by hyperlipidemia in the development of diabetes [37], hyperglycemia and hyperlipidemia can increase oxidative stress in the kidney, which is related to the reduction of antioxidant enzyme activity and the excessiveness of ROS [38,39]. ROS can directly impair mitochondrial function and can also regulate the expression of multiple genes associated with inflammation by activating NF-κB, ultimately leading to cellular inflammatory damage [40,41]. Oxidative stress and inflammation further lead to kidney damage, resulting in enhanced synthesis, weakened degradation, and excessive deposition of ECM, leading to renal fibrosis. Renal fibrosis is considered to be one of the most critical processes for DKD from a metabolic disorder, oxidative stress, and inflammation to ESRD [42,43] (Figure 1).

3. Natural Compounds

3.1. Phenolics

Phenolics are a large group of phytochemicals, with one or more aromatic rings and with one or more hydroxyl functional groups attached. They are present in fruits, cereals, vegetables, spices, teas, flowers and medical plants [44]. In the past decades, phenolics have been studied for their potential involvement in many areas including cancer, inflammation and microbial diseases [45].
Oleuropein is the most prevalent polyphenol in olive, which is a natural antioxidant molecule with a variety of biological activities and has many positive effects on human health, including anti-dyslipidemia, antidiabetic, anti-inflammatory, and antiatherogenic [46]. Studies have shown that oleuropein protected islet beta-cells from H2O2-induced cytotoxicity and promoted insulin secretion [47,48]. It has also been proven to treat gestational diabetes mellitus by activating AMPK signaling to improve lipid metabolism and inflammation [49]. In the treatment of DKD, Liu et al. found that administration of oleuropein can reduce renal injury, oxidative stress, and inflammation in db/db mice, oleuropein inhibits renal cell apoptosis by regulating the expression of mitogen-activated protein kinases (MAPK) signaling pathway and its downstream targets caspase-3, Bcl-2, and Bax [50]. In a clinical investigation, olive leaf extract (containing 136 mg oleuropein) effectively lowered plasma total cholesterol, low-density lipoprotein cholesterol, and triglyceride levels in prehypertensive patients. Furthermore, toxicology experiments revealed that olive leaf extract had no toxic effects at high doses [51].
Resveratrol is a natural polyphenol compound found in fruits such as grapes, mulberries, raspberries, and blueberries [52]. Many studies have indicated that resveratrol improves insulin sensitivity in insulin resistance animal models and has an anti-diabetic effect [53,54]. Clinical studies have shown that resveratrol can significantly reduce inflammation and oxidative stress, improve blood glucose control, and improve renal function in diabetes patients [55,56]. Resveratrol attenuates high-glucose (HG)-induced ER stress on the NRK 52E cells injured by hyperglycemia in vitro. The mechanism is related to inhibiting the increase of glucose-regulated protein 78 (GRP78) and C/EBP-homologous protein (CHOP) expression levels in cells, and alleviating ER stress-induced cell apoptosis [57]. Resveratrol can also improve renal injury by inhibiting podocyte apoptosis in DKD mediated by oxidative stress, the therapeutic effect is related to activating the AMPK signaling pathway [58]. Excessive generation of mitochondrial ROS is considered to be initiating event in the development of DKD. Resveratrol can activate sirtuin1 (SIRT1), increase the expression of SIRT1 and peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α) in renal tissue of DKD mice, reduce the production of mitochondrial ROS, and increase mitochondrial membrane potential, thus improving podocyte damage in diabetic mice [59]. Nrf2 is an important endogenous antioxidant transcription factor that protects cells from ROS injury, several studies have indicated that Nrf2 activators such as resveratrol have protective effects against DKD, resveratrol reduces oxidative stress, cell hyperproliferation, and ECM accumulation in mouse glomerular mesangium by activating the Keap1/Nrf2 signaling pathway [60]. In addition, Qiao et al. found that resveratrol inhibits HG-induced FN expression and reduce renal fibrosis by reducing MAPK activation and TGF-β1 expression in mesangial cells [61]. Gu et al. found that resveratrol can reduce the lipid accumulation in the kidney of DKD and improve kidney injury, the therapeutic effect of resveratrol is related to SIRT1 signaling pathway [62]. Another study found that resveratrol improves lipid metabolism in STZ-induced DKD by inducing AMPK/mTOR-mediated autophagy, at the same time, resveratrol alleviates lipid dysregulation by increasing the level of lipid oxidation-related protein such as peroxisome proliferator-activated receptor-alpha (PPARα) and carnitine palmitoyltransferase I (CPT1), and reducing the level of lipid production related protein such as sterol regulatory element-binding protein(SREBP)-1c and acyl-CoA synthetases (ACS) in DKD [63]. Some studies have demonstrated that resveratrol is a well-tolerated and safe compound in humans, but others have noted the harmful effects of resveratrol, which exhibited inhibition of P450 cytochromes when large doses were provided, while low doses of resveratrol are usually associated with beneficial effects, which needs to be considered in the research [64,65].
Gastrodin is a phenolic glycoside, which is the main active component of a traditional Chinese herbal medicine called Gastrodia elata [66]. Gastrodin has antioxidant, anti-inflammatory, hypolipidemic, anti-fatty liver, and therapeutic diabetes pharmacological activities. It has a therapeutic effect on the complications caused by diabetes, which can improve cognitive dysfunction, reduce blood glucose, reduce blood lipids, and increase insulin sensitivity in diabetic animals [67,68,69,70]. Huang et al. found that gastrodin increased the activity of antioxidant enzymes and reduced the level of inflammatory factors, thus reducing the inflammation, oxidative stress, and apoptosis of MPC-5 cells induced by HG. The mechanism is related to activating the AMPK/Nrf2 signaling pathway and reducing the formation of NLRP3 inflammasome [71]. Gastrodin is relatively safe to use. In mice, oral administration of gastrodin at a dose of 5000 mg/kg caused no mortality or obvious toxic effects [72]. At present, gastrodin is mainly used in clinics for neurasthenia, cephalagra, and other diseases, which has the effect of improving blood circulation [73]. With the deepening of research, gastrodin preparation is gradually applied to the treatment of diabetes, which improves the symptoms of patients [74]. Table 1 summarizes the available data on phenolics’ activities and the mechanisms of their action.

3.2. Alkaloids

Alkaloids are usually colorless and bitter basic nitrogen compounds (mainly heterocyclic), which mainly exist in the plant kingdom and usually have physiological activities [75]. Alkaloids exhibit different biological activities, such as anti-tumor, anti-inflammatory, antibacterial, and antiviral [76].
Trigonelline is an alkaloid in the extract of Trigonella foenum-graecum, Coffea sp., Glycine max, and Lycopersicon esculentum, which has a variety of biological activities such as the treatment of hyperglycemia, hypercholesterolemia, hormonal disorders, and cancers [77]. Studies showed that trigonelline significantly alleviated the oxidative stress and pathological changes in the kidneys and reduced the expression of FN and collagen ΙV in the mesangial ECM in DKD rats [78]. Human mesangial cells (HMCs) were stimulated with HG and treated with trigonelline. The results showed that trigonelline significantly inhibited the hyperproliferation of HMCs induced by HG and suppressed the levels of FN and collagen ΙV. Furthermore, trigonelline inhibited the activation of the Wnt/β-catenin signaling pathway to suppress cell-cycle progression and reduce apoptosis [79]. Li et al. found that trigonelline increased peroxisome proliferator-activated receptor-gamma (PPARγ) and glucose transporter type 4 (GLUT4) protein expression while suppressing leptin and tumour necrosis factor alpha (TNF-α) protein expression in the kidneys of DKD rats, thereby reduce inflammation, oxidative stress, and kidney cell apoptosis [80]. Chen et al. found that trigonelline upregulated the expression of miR-5189-5p, decreased hypoxia inducible factor 1 subunit alpha inhibitor (HIF1AN), and then activated the AMPK signaling pathway, increased autophagy level, and protected renal mesangial cells [81]. Toxicological studies showed that the lethal dose (LD50) of trigonelline in rats was around 5000 mg/kg after oral and subcutaneous administration [77]. In mice, trigonelline was fed 50 mg/kg daily for 21 days, and there was no change in the weight of the liver, kidney, thymus, thyroid, or adrenal gland [82]. Furthermore, trigonelline has good absorption and bioavailability. However, the existing data is insufficient to recommend trigonelline as a new medication; further clinical trials are needed to evaluate its adverse effects, pharmacokinetic characteristics, and mechanism of action.
Berberine is an isoquinoline alkaloid that occurs in Coptis chinensis. It has various pharmacological properties such as antioxidant, cardioprotective, anti-inflammatory, antibacterial, antidiabetic, and anticancer [83,84,85,86]. A meta-analysis of clinical studies shows that berberine significantly reduced fasting blood glucose, HbA1c, and triglyceride (TG), and improved insulin resistance in patients [87]. Studies have shown that berberine can significantly reduce fasting blood glucose, improve renal function, and alleviate podocyte injury in DKD rats. The mechanism may be related to berberine reducing the expression of phosphoinositide 3-kinase (PI3K) and protein kinase B (Akt) in the kidneys of DKD rats and increasing the expression of podocyte functional proteins (nephrin, podocin, and α3β1) in podocytes stimulated by high glucose [88]. In addition, berberine can also reduce the expression of p-AS160 and the level of membrane glucose transporter type 1 (GLUT1) by inhibiting the PI3K/Akt signaling pathway, thus reducing the glucose uptake of glomerular mesangial cells (GMCs) [89]. But another study found that in hypoxia/HG-induced NRK 52E and HK-2 cells, berberine promoted the activation of the PI3K/Akt signaling pathway and increased the expression of hypoxia-inducible factor-1alpha (HIF-1α), which led to a reduction in the apoptosis of cells [90].This may be related to the treatment of hypoxia. When glycolipid metabolism is out of balance, the lipid accumulation of renal tubular epithelial cells increases, leading to their dysfunction and tubulointerstitial fibrosis. Sun et al. study showed that berberine ameliorated apoptosis and decreased lipid accumulation in palmitate (PA)-induced HK-2 cells [91]. Rong et al. studied type 2 diabetic db/db mice and HG-induced HK-2 cells and found that berberine decreased the expression of alpha smooth muscle actin (α-SMA), collagen I, collagen IV, FN, and TGF-β1, thus reducing renal tubulointerstitial injury and renal fibrosis in diabetic db/db mice. Berberine increased the expression of CPT1, acyl-CoA oxidase 1 (ACOX1), and PPARα levels, thereby reducing lipid accumulation in the DKD models. The imbalance of glycolipid metabolism impairs mitochondrial morphology and mitochondrial function. PGC-1α is a transcriptional coactivator and has been shown to regulate mitochondrial functions, berberine enhanced AMPK activation and promoted PGC-1α expression in tubular epithelial cells [92]. In addition, berberine activated the CCAAT enhancer-binding protein beta (C/EBPβ) expression in HG-induced HK-2 cells. The C/EBPβ could combine with the reaction element on the promoter of lncRNA Gas5 to promote its expression, thereby inhibiting the miR-18a-5p expression, the expression level of miR-18a-5p is positively correlated with the ratio of apoptosis and mitochondrial ROS level. Meanwhile, C/EBPβ can activate the expression of PGC-1α and improve the mitochondrial energy metabolism. Berberine inhibited the generation of ROS, regulated the energy metabolism of mitochondria, and reduced apoptosis by activating the C/EBPβ/Gas5/miR-18a-5p signaling pathway and the C/EBPβ/PGC-1α signaling pathway [93]. Qin et al. found that in diabetic kidneys and PA-induced podocytes, berberine increased the protein levels of p-AMPK, PGC-1α, CPT1, and p-ACC while downregulating the expression of CD36. Therefore, it improves lipid accumulation, excessive generation of mitochondrial ROS, and mitochondrial dysfunction [94]. Another study showed that berberine also reduced the expression of dynamin-related protein 1 (Drp1) and consequently decreased the mitochondrial fission protein (MFF), mitochondrial fission protein 1 (Fis1), and mitochondrial dynamics proteins (Mid49, Mid51), thus to improving ROS generation, apoptosis, and mitochondrial dysfunction in diabetic kidneys and PA-induced podocytes [95]. Early studies have shown that Berberine inhibited the activation of the Notch/snail signaling pathway and upregulated α-SMA and E-cadherin levels in the DKD models, thus inhibiting renal tubular epithelial-mesenchymal transition (EMT) and renal fibrosis [96]. Berberine can also regulate the mTOR/P70S6K/4EBP1 signaling pathway and the toll-like receptor 4 (TLR4)/NF-κB signaling pathway in HG-induced podocytes, which activates autophagy and reduces inflammation and apoptosis [97,98]. Berberine has been demonstrated in animal experiments to have very minimal toxicity and adverse effects [99]. Some clinical trials on the safety of berberine in humans have found only mild gastrointestinal effects, such as diarrhea and constipation [100]. However, berberine has low intestinal absorption and oral bioavailability due to self-aggregation in the acidic environment of the stomach and intestinal first-pass elimination [101,102]. Therefore, it is necessary to change the dosage form to improve the bioavailability of berberine.
Sinomenine is a morphinane-type isoquinoline-derived alkaloid that is extracted from the roots and stems of Sinomenium diversifolius (Miq.) [103]. Sinomenine has an anti-diabetes effect on gestational diabetes rats and STZ-induced diabetes rats, reducing fasting blood glucose and inflammation levels in animals [104]. Studies found that sinomenine has a protective effect on human renal glomerular endothelial cells (HRGEs) induced by HG. When in a high glucose environment, intracellular ROS increases, which activates the Rho-associated protein kinase (ROCK) signaling pathway, destroys the expression of zonula occludens-1 (ZO-1)/occludin, and increases cell permeability. Sinomenine can reduce ROS production by activating the Nrf2 signaling pathway, thus protecting kidney cells [105]. Zhang et al. found that sinomenine can also activate the C/EBPα/claudin-5 pathway, alleviating HG-induced dysfunction of HRGEs and reducing inflammation, which has also been verified in DKD rats [106]. Zhu et al. found that sinomenine increased intracellular antioxidant enzymes, protected HK-2 cells from H2O2 damage, and reduced apoptosis. In DKD rats, sinomenine regulated the janus kinase (JAK)/signal transducer and activator of transcription (STAT) signaling pathway, thereby inhibiting the expression levels of fibrosis related proteins and inflammation related proteins [107]. In clinical studies, sinomenine has a potent anti-inflammatory effect. It has been formulated into tablets and injections for the treatment of knee osteoarthritis [108,109]. With the application of sinomenine, its adverse reactions, such as allergic reactions, nausea, and vomiting, have gradually attracted attention, which may be related to the bidirectional regulation of histamine release by sinomenine [110]. Huang et al. found that the safety of sinomenine was related to sexual distinction. The LD50 of male rats was 72.29 mg/kg, while that of female rats was 805.69 mg/kg [111]. Sinomenine may be used to treat kidney diseases in the future, but its safety is an issue that must be resolved. Table 2 summarizes the available data on alkaloids’ activities and the mechanisms of their action.

3.3. Flavonoids

Flavonoids are found almost everywhere in plants. They are rich in seeds, fruits, flowers, and medicinal plants [112]. They are low molecular weight compounds having a basic 15-carbon flavone skeleton, C6-C3-C6, with two benzene rings (A and B) linked by a three-carbon pyran ring (C) [113]. Flavonoids are categorized into six primary types based on their structure: anthocyanins, flavan-3-ols, flavones, flavanones, isoflavones, and flavonols [114].
Naringenin is a flavanone found mainly in citrus fruits, it has revealed promising pharmacological activities including cardiovascular diseases, anti-diabetic, antimicrobial, antiviral, anticancer, and anti-inflammatory [115,116]. On the NRK 52E cells injured by hyperglycemia in vitro and the DKD model in vivo, naringenin treatment markedly reduced the excessive production of intracellular ROS and downregulated the expression of endoplasmic reticulum (ER) stress marker proteins, including p-PERK, eukaryotic initiation factor 2 alpha (eIF2α), X-box-binding protein 1 (XBP1s), ATF4, and CHOP, anti-ER stress to reduce apoptosis of renal cells in diabetes [117]. Studies have also shown that, naringenin markedly reduced the proliferation and alleviated the morphological changes of NRK 52E cells induced by HG in a dose-dependent manner, naringenin ameliorates the renal damage of DKD mice, reduces glomeruli and renal tubular lesions through modulation of peroxisome proliferators-activated receptors (PPARs) with subsequent normalization of cytochrome P450 4A (CYP4A) expression and increasing 20-hydroxyeicosatetraenoic acid (20-HETE) [118]. In addition, Yan et al. found that MicroRNA let-7a was down expressed in both DKD rats and 293T mesangial cells under high glucose conditions, naringenin can inhibit TGF-β1/Smad signaling pathway by increasing the expression of MicroRNA let-7a, repressing glomerular mesangial cells proliferation and accumulation of ECM, thereby preventing renal fibrosis [119]. A clinical study showed a single dose of less than 900 mg of naringenin was safe and well tolerated in humans [120]. But another study found that naringenin has a dose-dependent inhibitory effect on the reproductive function of adult male mice and shows a pro-oxidative effect in testicular tissue [121]. This requires further clinical research to solve the safety and effectiveness of naringenin in humans.
Quercetin is present in numerous fruits and vegetables, recent studies have shown that quercetin has beneficial therapeutic effects in improving inflammation, blood lipids, and diabetes [122]. A meta-analysis of DKD animal experiments showed that after quercetin treatment, renal function index (such as urinary protein, uric acid, urinary albumin and serum creatinine levels) improved significantly [123]. The hyperglycemic environment in diabetes patients leads to the increase of advanced glycation end products (AGEs) production, AGEs can bind to the collagen that makes up the glomerular basement membrane, disrupting the glomerular barrier [124]. A study showed that quercetin could significantly reduce AGEs levels in renal tissue and serum levels of TNF-α and interleukin (IL)-6, increasing the level of SOD and glutathione peroxidase (GSH-Px) in serum [125]. Podocytes injury is one of the leading causes of proteinuria in patients with DKD, Liu et al. found that quercetin can prevent glomerular damage in diabetic mice and repress podocyte apoptosis by inhibiting the EGFR signaling pathway [126]. A clinical study showed that the expression of miR-485-5p in peripheral blood of patients with DKD decreased significantly, through cell experiment, it was found that quercetin inhibited the expression of YAP1 by regulating the increase of miR-485-5p, thereby inhibiting HG-induced HMCs hyperproliferation, inflammation, and oxidative stress [127]. In addition, Du et al. found that quercetin can also reduce the expression of yes-associated protein 1 (YAP1) by activating Hippo signaling pathway [128]. Dyslipidemia is one of the most serious and frequently occurring complications in DKD patients, lipid accumulation in the kidney has also been considered to play a role in the pathogenesis of DKD [129], Jiang et al. found that there were a large number of lipid droplets of different sizes in the renal cortex of diabetes mice, and quercetin could effectively reverse the lipid accumulation in both glomerulus and renal tubular cells by SREBP cleavage activating protein (SCAP)-SREBP2-low-density lipoprotein receptor (LDLr) signaling pathway [130]. Wang et al. found that quercetin can reduce renal lipid accumulation by regulating the expression of PPARα, CPT1, organic cation/carnitine transporter 2 (OCTN2), and acetyl-CoA carboxylase 2 (ACC2), it can also reduce renal inflammation by regulating the activation of renal NLRP3 inflammasome/caspase-1/IL-1β/IL-18 signaling pathway [131]. Currently, there have been some clinical experiments to investigate the effect of quercetin on diabetic patients. Oral quercetin (250 mg/day) for 8 weeks could significantly improve the antioxidant status of participants. Single oral administration of quercetin (400 mg) can effectively inhibit postprandial hyperglycemia after maltose loading in T2DM patients [132]. However, its oral bioavailability is minimal, limiting its therapeutic use. This is a major issue that must be addressed in future research.
Icariin is an active ingredient extracted from the traditional Chinese medicine Epimedium, some studies have shown that icariin can significantly inhibit cell apoptosis and oxidative stress [133]. In DKD rats and HG-induced MPC-5 cells, icariin could upregulate Sesn2 expression to induce mitophagy and activate the Keap1-Nrf2/HO-1 axis to inhibit NLRP3-related inflammation [134], icariin can also lighten renal inflammation by suppressing the TLR4/NF-κB signaling pathway, thus reducing renal fibrosis [135]. Icariin reduces the accumulation of collagen and FN in mesangial cells induced by high glucose by inhibiting the production of TGF-β1 and inhibiting Smad and extracellular signal-regulated kinase (ERK) signals in a G protein-coupled estrogen receptor (GPER)-dependent manner [136]. Jia et al. found that icariin can induce autophagy and reduce renal fibrosis in the DKD models, and its mechanism is related to the reduction of miR-192-5p, its overexpression inhibited glucagon-like peptide 1 receptor (GLP-1R), induced p-mTOR expression, and increasing the expression of collagen I, α-SMA, and FN. Icariin can alleviate DKD renal fibrosis by restoring autophagy through the miR-192-5p/GLP-1R pathway [137]. In addition, Zang et al. found that miR-122-5p also plays a role in the development of DKD, miR-122-5p inhibited forkhead box protein P2 (FOXP2) transcription, resulting in decreased cell viability, and inhibit E-cadherin expression, increasing α-SMA expression. Icariin can inhibit the expression of miR-122-5p and promote FOXP2 transcription, thus reducing the renal injury of DKD [138]. Icariin is a diglycoside, which indicates it is difficult to absorb. The bioavailability of icariin can be improved by pharmaceutical methods such as inclusion with β-cyclodextrin and propyleneglycol (PG)-liposomes [139]. At present, icariin lacks strong clinical evidence to prevent and treat DKD. In-depth mechanism research and evaluation of its safety are still needed.
Cardamonin is a flavonoid found in Alpinia, which can resist oxidative damage and apoptosis in vitro [140,141]. Cardamonin can reduce the blood glucose level of T2DM mice, improve hepatocyte lipid deposition, and inhibit sodium/glucose cotransporter 1 (SGLT1) [142,143]. Methylglyoxal (MGO) is capable of combining with proteins to form AGEs, which promote the production of ROS and induce cell apoptosis [144,145]. Gao et al. research showed that cardamonin reduced apoptosis, inflammation, oxidative stress, and renal fibrosis of MGO-treated NRK 52E cells and diabetic rats, its mechanism is related to regulating PI3K/AKT and JAK/STAT signaling pathway, reducing caspase-3, Bax, NF-κB, FN, α-SMA, and TGF-β1 protein expression, and increasing Bcl-2 and vimentin expression [146].
Morin is a flavonoid existing in the Moraceae family, which can improve oxidative stress, inflammation, and lipid metabolism [147,148]. Morin can reduce H2O2-induced Madin-Darby canine kidney cells oxidative stress and DNA oxidative damage [149]. Mathur et al. found that high glucose exposure in DKD models upregulated pleckstrin homology domain leucine-rich repeat protein phosphatases 1 (PHLPP1) and promoted the nuclear retention of forkhead box protein O1 (FoxO1) through double minute 2 protein (MDM2), thus leading to aberration in renal gluconeogenesis and activation of the apoptotic cascade. On the contrary, PHLPP1 gene silencing enhanced Nrf2 expression and weakened FoxO1-regulated apoptosis. Treatment with Morin can effectively down-regulate the expression of PHLPP1, relieve oxidative stress, and reduce renal cell apoptosis [150]. Another study showed morin inhibited HG-induced ECM expression, ROS generation, and NOX4 expression in glomerular mesangial cells. The mechanism is related to the inhibition of p38 MAPK and JNK signaling pathways [151]. Morin has good safety. No mortality or abnormal manifestations were found in rats given large doses of morin (about 300–2400 mg/kg) for 13 weeks [152]. The majority of the glycosylated morin administered orally was not absorbed in the small intestine and flowed into the colon, where it is then metabolized by colonic microorganisms to morin aglycones, which are readily absorbed [153,154].
Hesperetin is a flavanone that is present in the peels of several citrus fruits, and research found it has a variety of biological activities such as antioxidation, anti-inflammation, and improve glycolipid metabolism [155,156,157]. Early studies discovered that hesperetin had a hypoglycemic impact due to α-glucosidase inhibition [158,159]. Recent research has found that hesperetin has a therapeutic effect on DKD. It can increase antioxidant enzymes, such as thiobarbituric acid reactive substances (TBARS), GSH-Px, and CAT, reduce inflammatory cytokines (TNF-α, IL-6) expression, and inhibit TGF-β and glycogen synthase kinase-3beta (GSK-3β) expression, thereby reducing renal oxidative stress, inflammation, and fibrosis [160]. Chen et al. found that hesperetin can increase Glo-1 expression by activating the Nrf2 signaling pathway, thus accelerating AGEs clearance and decreasing inflammatory cytokines expression in the kidney. Furthermore, hesperetin reduces collagen ΙV and FN expression in the kidney and improves renal fibrosis [161]. Hesperetin also has good safety and has no mutagenic, toxic, or carcinogenic effects on pregnant mice [162]. Although hesperetin has obtained positive findings on diabetes treatment in animal studies, its functional mechanism in humans remains to be elucidated.
Fisetin is a natural dietary flavonoid that mainly exists in various fruits and vegetables, such as apples, grapes, cucumbers, and onions [163]. Through kinetic and molecular docking studies, it was found that fisetin had a potential inhibitory effect on α-glucosidase, which was verified by in vitro experiments [164,165]. Fisetin also has antioxidant, anti-inflammatory, and lipid metabolism-regulating activities [166,167], which can prevent the development of diabetes cardiomyopathy, diabetic neuropathy, and diabetes encephalopathy [168,169,170]. Research has found that fisetin reduced the EMT process, alleviated HG-induced podocyte injury and STZ-induced DKD, which is related to the restoration of cyclin-dependent kinase inhibitor 1B (CDKN1B)/P70S6K mediated autophagy and the inhibition of the NLRP3 inflammasome [171]. Obesity-induced hyperlipidemia is an important factor in DKD injury. Ge et al. studied high-fat diet (HFD) mice and PA-treated HK-2 cells and found that fisetin can regulate the insulin receptor signaling pathway to improve insulin sensitivity, inhibit the NF-κB signaling pathway, and receptor-interacting serine-threonine kinase 3 (RIP3)/NLRP3 signaling pathway to reduce inflammation [172]. Table 3 summarizes the available data on flavonoids’ activities and the mechanisms of their action.

3.4. Terpenoids

Terpenoids have very diverse physical and chemical properties as well as numerous biological activities, which are characterized by a carbon number multiple of five [173]. According to the amount of carbon, terpenoids can be divided into monoterpenes (C10), sesquiterpenes (C15), diterpenes (C20), sesterterpenes (C25), triterpenes (C30), tetraterpenes (C40), and higher homologs.
Sclareol is a natural diterpene that is an antifungal specialized metabolite produced by clary sage, Salvia sclarea [174]. Sclareol has been proven in studies to have anti-cancer, antioxidant, and anti-inflammatory effects [175,176,177], as well as the ability to enhance insulin sensitivity and glucose tolerance, hence improving metabolism in obese mice [175]. Han et al. found that sclareol treatment significantly alleviated renal dysfunction, fibrosis, and the levels of inflammatory cytokines in DKD mice, and sclareol treatment was dose-dependent. Sclareol inhibited the inflammatory reactions via the MAPK-mediated NF-κB pathway [178].
Ponicidin, a tetracyclic diterpenoid active ingredient extracted from the phytomedicine Rabdosia rubescens, has a positive effect on the treatment of a variety of cancers by inhibiting pro-inflammatory cytokine TNF-α induced angiogenesis and EMT [179,180,181]. An et al. found that ponicidin treatment can effectively decrease the levels of ROS and MDA in the serum of DKD rats, improve lipid metabolism in animals, reduce renal fibrosis and reduce the expression of inflammatory factors TNF-α, IL-1β, IL-6 and NF-κB [182].
Triptolide is a diterpenoid extracted from Tripterygium wilfordii Hook. f. which has a variety of biological activities such as antioxidation, immunomodulatory, and anticancer [183,184,185]. Some studies have shown that triptolide plays a significant role in the treatment of DKD, and a meta-analysis showed that triptolide significantly reduced albuminuria, blood urea nitrogen, serum creatinine, and urinary albumin/creatinine ratio in DKD animals [186,187]. T-helper (Th) cells is an immune cell populations, which can be divided into Th1/Th2 cells according to the cytokines. Th1 and Th2 cells cooperate to maintain the relative balance of immune response, and the activation of Th1 cells induces the secretion of proinflammatory cytokines, leading to the aggravation of DKD [188]. Guo et al. found that triptolide regulates the expression of pro-inflammatory (Interferon-γ, IL-12, and TNF-α) and anti-inflammatory (IL-4 and IL-10) cytokines and restores the balance of Th1/Th2 cells, thus reducing macrophage infiltration and the expression of inflammatory cytokines in the kidney [189]. Autophagy plays a positive role in the treatment of DKD, studies have shown that triptolide down-regulates the expression of miR-141-3p and miR-188-5p, causing the up-regulation of phosphatase and tensin homolog (PTEN) expression, affecting the expression of downstream pyruvate dehydrogenase kinase isoform 1 (PDK1), Akt, and mTOR, thus improving the level of autophagy, which in turn reduces the proliferation and fibrosis of mesangial cells [190,191,192]. Other studies have shown that triptolide can inhibit the Wnt3α/β-catenin signaling pathway to improve HG-induced EMT of podocytes [193]. Triptolide can also downregulate the protein expression levels of NLRP3 and apoptosis associated speck-like protein containing a CARD (ASC), inhibit the activation of the NLRP3 inflammasome, reducing the downstream protein expression such as caspase-1, IL-1β, and IL-18, thus preventing podocyte inflammatory injury [194]. Ren et al. found that triptolide can inhibit the TGF-β/Smad signaling pathway to downregulate p-Smad3, and inactivate kindlin-2, thus preventing podocyte EMT and protecting the kidney [195]. Clinically, triptolide is mainly used in inflammation and autoimmune diseases. It has high oral bioavailability but low safety. It has serious cytotoxicity to the heart, liver, kidney, and other organs [196]. Further research is needed to closely examine the relationship between the function and toxicology of triptolide to reduce its toxicity. Table 4 summarizes the available data on terpenoids’ activities and the mechanisms of their action.

3.5. Saponins

Saponins are high molecular weight amphiphilic compounds with a lipophilic moiety of triterpenoid or steroid aglycon and a hydrophilic moiety of sugars (usually glucose, arabinose, rhamnose, and xylose) [197]. Saponins usually have protective effects on the cardiovascular system and therapeutic effects on diabetes [198].
Dioscin is a natural steroidal saponin that is isolated from the Dioscoreaceae family [199], in 2015, DA-9801 containing dioscin completed the Phase II clinical trial for the treatment of diabetic neuropathy in the United States [200], some experimental studies showed that dioscin has therapeutic effect on some complications caused by diabetes mellitus, which can reduce the vascular damage in the retina of db/db mice and alleviate glycolipid metabolic disorder of T2DM [201,202]. In the DKD studies, dioscin significantly ameliorated renal damage via antagonizing the activation of the TLR4/NF-κB pathway and the production of inflammatory cytokines [203]. In addition, Zhong et al. found that dioscin reduced ROS levels, enhanced antioxidant enzyme (SOD, CAT) activities, and reduced inflammatory cytokine (IL-1β, IL-6, TNF-α, NF-κB) expressions. Dioscin could significantly inhibit the increase of p-PERK, IRE1, p-IRE1, ATF4, CHOP, and Caspase-12 expression levels in kidneys, and alleviate ER stress-induced cell apoptosis. Dioscin could also regulate the expression of the AMPK/mTOR pathway to promote autophagy in the DKD. Mitophagy and mitochondrial fission/fusion belong to the process of mitochondrial quality and quantity control, dioscin improves the expression of PTEN-induced putative kinase 1 (PINK1), Drp1, p-Drp1, and mitofusin 2(MFN2) to relieve the disorder of mitochondrial [204]. In the clinic, some drugs with dioscin as the main component are used to treat cardiovascular diseases [205]. However, diosgenin is a poorly soluble drug, and its oral absolute bioavailability is only 0.2% [206]. Some researchers have formulated dioscin as a new nano-drug delivery system, which can improve its oral bioavailability and drug loading [207]. Dioscin may have potential hepatotoxicity; when 300 mg/kg of dioscin was administered to rats for 90 days, levels of alanine aminotransferase increased significantly [208]. Clinical experimental research should be given special consideration.
Ginsenoside Rb1 is a tetracyclic triterpene saponin extracted from Panax L., which has a variety of biological activities such as antioxidation, anti-inflammation, anti-arrhythmia, anti-shock, and anti-diabetic [209,210]. Ginsenoside Rb1 regulates glucose and lipid metabolism by improving insulin and leptin sensitivity [211]. Studies have shown that ginsenoside Rb1 significantly alleviated the oxidative stress and pathological changes in the kidneys and reduced the expression of FN and collagen ΙV in the mesangial ECM in DKD rats. The mechanism is related to downregulating miR-3550 expression and inhibiting the Wnt/β-catenin signaling pathway [78]. He et al. study showed that ginsenoside Rb1 improves mitochondrial damage, oxidative stress, and apoptosis of renal podocytes in DKD models. High glucose can increase the expression of NOX4, a member of the NADPH oxidase family, and induce excessive ROS production. Aldose reductase (AR), an NADPH-dependent oxidoreductase, is a key rate-limiting enzyme in the polyol pathway of glucose metabolism. Ginsenoside Rb1 can bind to AR and inhibit AR activity, thereby reducing the expression of downstream NOX4, inhibiting the generation of ROS, and preventing the activation of caspase-9 [212]. The clinical experiment showed that healthy people orally had red ginseng extract containing 75 mg of ginsenoside Rb1 once or red ginseng extract containing 23 mg of ginsenoside Rb1 for two weeks without experiencing any abnormal effects [213,214]. At present, there are many studies showing the positive effect of ginsenoside Rb1 in the treatment of diabetes, which needs to be systematically studied in the clinic.
Platycodin D is a deglycosylated triterpene saponin, which is found in Platycodon grandiflorum, a traditional Chinese medicinal herb with medicine and food homology [215]. Platycodin D has potent anti-inflammatory effects and anti-organ fibrosis effects, and it is also a new AMPK activator that reduces obesity in db/db mice through regulating adipogenesis and thermogenesis [216,217,218]. Studies have shown that platycodin D can inhibit the glomerular basement membrane thickening and fibrosis in DKD rats, which is related to the regulation of the PI3K/Akt signaling pathway to reduce oxidative stress and inflammation in the kidney [219]. Ferroptosis is involved in the regulation of cell death in a variety of diseases, including ischemia-reperfusion injury, cancer, and kidney disease. Excess iron in cells inhibits cell function by producing ROS, eventually leading to cell death [220]. Huang et al. found that HG increased oxidative stress and iron levels in HK-2 cells, thus causing ferroptosis. Platycodin D alleviated HG-induced ferroptosis in HK cells by upregulating glutathione peroxidase 4 (GPX4), ferritin heavy chain (FTH1), solute carrier family 7 member 11 (SLC7A11), and down-regulating the expression of acyl-coA synthetase long chain family member 4 (ACSL4) and transferrin receptor protein 1 (TFR1) [221]. Toxicological studies showed that platycodin D at a single oral dose of 2000 mg/kg body weight showed no toxic effects on mice [222]. An in vitro hemolysis assay showed that platycodin D had no obvious hemolytic effect on rabbit erythrocytes at concentrations ranging from 2.5~10 μM [223]. Platycodin D has poor bioavailability. After oral administration of 10 mg/kg platycodin D in rats, the bioavailability was only 0.079% [224]. This is the main problem restricting its application. Table 5 summarizes the available data on saponins’ activities and the mechanisms of their action.

3.6. Other Compounds

Caffeoylisocitric acid is a rare cinnamic acid derivative, it is a condensed ester of an isocitric acid and a caffeoylic acid, which is first found in Amaranthus cruentus [225]. Studies have shown that in the range of 0.01 to 200 µM, caffeoylisocitric acid has no significant cytotoxicity on normal HMCs, it can activates Nrf2 signaling pathway and inactivates MAPK signaling pathway to attenuate oxidative stress, inflammation and accumulation of ECM in mesangial cells under high glucose [226].
Crocin is a water-soluble carotenoid extracted from saffron (Crocus sativus L.), which has anti-inflammatory and antioxidant effects [227]. Clinical trials suggest that crocin may regulate the serum lipid profile in patients with metabolic disorders [228]. Many studies have shown that crocin has a therapeutic effect on DKD. Crocin can reduce oxidative stress and apoptosis to protect renal tubular epithelial cells from high glucose damage, which is related to the activation of the SIRT1/Nrf2 pathway [229]. In vivo experiments, crocin significantly reduced fasting blood glucose and blood lipid levels in db/db mice, and decreased the production of ROS and the expression of inflammatory factors in the kidney. The mechanism was related to the increase in the expression levels of Nrf2, SOD-1, HO-1, and CAT, and inhibition of the NF-κB signaling pathway [230]. Zhang et al. research showed that crocin reduced renal oxidative stress and inflammatory factors (TNF-α, IL-1β, and IL-18) by inhibiting the NLRP3 inflammasome, thereby reducing the expression of renal fibrosis proteins (TGF-β, collagen I, and collagen IV) and protecting the kidney [231]. A clinical study showed that after three months of crocin treatment, the fasting blood glucose and glycosylated hemoglobin of patients with diabetes were significantly reduced compared with the placebo group [232]. Hosseinzadeh et al. reported no animal mortality after single-dose administration of crocin (tolerated dose of 3 g/kg, i.v. or i.p.) in mice. Rats were administered crocin daily (15–180 mg/kg, i.p.) for 21 days, and no abnormal changes were observed [233]. These results suggest that crocin could be a promising natural product for the treatment of DKD.
Fraxin is the main active component of Fraxinus rhynchophylla Hance and belongs to the coumarin family, which has the functions of scavenging radicals and antioxidation [234]. The study found that fraxin inhibits the expression of inflammatory fibrosis factors by increasing the antioxidant enzymes in HG-induced primary glomerular mesangial cells, Cx43 interacted with AKT and consequently regulated the Nrf2 signaling pathway, fraxin could activate the Nrf2 pathway by regulating the interaction between connexin 43 (Cx43) and Akt to reduce intracellular oxidative stress and ROS generation. In addition, fraxin reduced the degree of renal fibrosis in db/db mice by inhibiting the protein expression of FN and ICAM-1 [235]. Table 6 summarizes the available data on other compounds’ activities and the mechanisms of their action.

4. Discussion

As the prevalence of DKD has increased over the decades, it has become one of the most common kidney diseases worldwide, to date, there are no sufficiently effective drug treatments to reverse the onset of DKD or prevent it from progressing to more severe stages of the disease. At present, the treatment of DKD mainly focuses on reducing blood glucose. However, the development of DKD is the result of a variety of injury factors. Lowering blood glucose can control the progression of the disease in the early stages, but in the middle and late stages, it is difficult to successfully prevent the development of the disease by only regulating blood glucose, and other targeted drugs are required to improve kidney damage. Indeed, many natural compounds have been extensively researched and have demonstrated good pharmacological action in the treatment of DKD.
The mechanisms of natural compounds reviewed in this article are mainly studied through the Nrf2 signaling pathway, NF-κB signaling pathway, TGF-β signaling pathway, NLRP3 inflammasome, autophagy, glycolipid metabolism and ER stress. Some compounds have multiple pathways and targets in the treatment of DKD. However, it is unclear if this effect just influences the expression of a specific protein before affecting the expression of other pathway proteins, or whether it regulates the expression of many pathway proteins and plays a therapeutic role together. This is also a problem in the current research of most natural compounds. It is necessary to deeply and systematically study the mechanism of natural compounds through transcriptomics, proteomics, pathway inhibitors, and other methods.
Here, we reviewed the selected natural compounds with beneficial effects in DKD reported by preclinical studies. They are all natural monomer compounds from plants. Some of these compounds have undergone relatively comprehensive mechanism studies and shown positive effects on DKD treatment, but there is no clinical evidence, such as oleopein, trigonelline, naringenin, icariin, hesperetin, ginsenoside Rb1, and platycodin D. Some natural compounds have comprehensive mechanism studies and are applied to other diseases clinically. Whether they can also be applied to the treatment of DKD needs further discussion, such as gastrodin, sinomenine, triptolide, and dioscin. Some natural compounds have been applied to the treatment of DKD in clinical preliminary studies, such as resveratrol, berberine, quercetin, and crocin. Some natural compounds only show a preliminary positive effect on the treatment of DKD, and their specific effects need to be further studied, such as cardamonin, morin, fisetin, sclareol, ponicidin, caffeoylisocitric acid, and fraxin. At the same time, it should be noted that some natural compounds have low bioavailability or poor safety, which are the reasons hindering clinical application and need to be solved by researchers.
This review provided an updated overview of the potential benefits of these natural compounds for the prevention and treatment of DKD progression, aiming to provide new potential therapeutic lead compounds and references for the innovative drug development and clinical treatment of DKD. Many natural compounds have shown potential effects on the treatment of DKD. At present, most studies focus on the antioxidant and anti-inflammatory capabilities of these compounds, which can reduce the deterioration of DKD. In the future, it may be possible to combine natural compounds with existing hypoglycemic drugs to exert the antioxidant and anti-inflammatory capabilities of natural compounds while reducing glucose, so as to achieve better protection of the kidney.

5. Methodology

In this review, we searched the PubMed database (https://pubmed.ncbi.nlm.nih.gov/) and Google scholar (https://scholar.google.com/). In the methodology, the name of the natural compound is replaced by “NC”. Search keywords include “(Diabetes nephropathy) OR (Diabetic kidney disease)”, “((Diabetes nephropathy) OR (Diabetic kidney disease)) AND (NC)”, “(Diabetes) AND (NC)”, “(Clinical) AND (NC)”, “(Toxicology) AND (NC)”, “(Pharmacokinetics) AND (NC)”.

Author Contributions

K.Z. contributed to conception and design. X.Z., J.S. and Q.Z. searched the literature. K.Z. wrote the manuscript. J.L., H.B. and L.L. critically viewed, edited the manuscript. All authors listed have read and approved it for publication. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Jilin Province (No. YDZJ202201ZYTS199) and the National College Students Innovation and Entrepreneurship Training Program (No. 202210199020).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Zhao, L.; Han, Q.; Zhou, L.; Bai, L.; Wang, Y.; Wu, Y.; Ren, H.; Zou, Y.; Li, S.; Su, Q.; et al. Addition of glomerular lesion severity improves the value of anemia status for the prediction of renal outcomes in Chinese patients with type 2 diabetes. Ren. Fail. 2022, 44, 346–357. [Google Scholar] [CrossRef] [PubMed]
  2. Hashemi, L.; Hsiung, J.T.; Arif, Y.; Soohoo, M.; Jackson, N.; Gosmanova, E.O.; Budoff, M.; Kovesdy, C.P.; Kalantar-Zadeh, K.; Streja, E. Serum Low-Density Lipoprotein Cholesterol and Cardiovascular Disease Risk Across Chronic Kidney Disease Stages (Data from 1.9 Million United States Veterans). Am. J. Cardiol. 2022, 170, 47–55. [Google Scholar] [CrossRef]
  3. Wu, T.H.; Chang, L.H.; Chu, C.H.; Hwu, C.M.; Chen, H.S.; Lin, L.Y. Soluble tumor necrosis factor receptor 2 is associated with progressive diabetic kidney disease in patients with type 2 diabetes mellitus. PLoS ONE 2022, 17, e0266854. [Google Scholar] [CrossRef] [PubMed]
  4. Chinese Diabetes Society. Guideline for the prevention and treatment of type 2 diabetes mellitus in China (2020 edition). Chin. J. Diabetes Mellit. 2021, 13, 315–409. [Google Scholar]
  5. Navaneethan, S.D.; Zoungas, S.; Caramori, M.L.; Chan, J.C.N.; Heerspink, H.J.L.; Hurst, C.; Liew, A.; Michos, E.D.; Olowu, W.A.; Sadusky, T.; et al. Diabetes Management in Chronic Kidney Disease: Synopsis of the 2020 KDIGO Clinical Practice Guideline. Ann. Intern. Med. 2021, 174, 385–394. [Google Scholar] [CrossRef]
  6. Association, A.D. 9. Pharmacologic Approaches to Glycemic Treatment: Standards of Medical Care in Diabetes-2021. Diabetes Care 2021, 44 (Suppl. S1), S111–S124. [Google Scholar] [CrossRef] [PubMed]
  7. Au, P.C.M.; Tan, K.C.B.; Cheung, B.M.Y.; Wong, I.C.K.; Li, H.L.; Cheung, C.L. Association between SGLT2 Inhibitors vs DPP4 Inhibitors and Renal Outcomes among Patients with Type 2 Diabetes. J. Clin. Endocrinol. Metab. 2022, 107, e2962–e2970. [Google Scholar] [CrossRef] [PubMed]
  8. de Boer, I.H.; Caramori, M.L.; Chan, J.C.N.; Heerspink, H.J.L.; Hurst, C.; Khunti, K.; Liew, A.; Michos, E.D.; Navaneethan, S.D.; Olowu, W.A.; et al. Executive summary of the 2020 KDIGO Diabetes Management in CKD Guideline: Evidence-based advances in monitoring and treatment. Kidney Int. 2020, 98, 839–848. [Google Scholar] [CrossRef] [PubMed]
  9. McMurray, J.J.V.; Solomon, S.D.; Inzucchi, S.E.; Køber, L.; Kosiborod, M.N.; Martinez, F.A.; Ponikowski, P.; Sabatine, M.S.; Anand, I.S.; Bělohlávek, J.; et al. Dapagliflozin in Patients with Heart Failure and Reduced Ejection Fraction. N. Engl. J. Med. 2019, 381, 1995–2008. [Google Scholar] [CrossRef]
  10. Packer, M.; Anker, S.D.; Butler, J.; Filippatos, G.; Pocock, S.J.; Carson, P.; Januzzi, J.; Verma, S.; Tsutsui, H.; Brueckmann, M.; et al. Cardiovascular and Renal Outcomes with Empagliflozin in Heart Failure. N. Engl. J. Med. 2020, 383, 1413–1424. [Google Scholar] [CrossRef] [PubMed]
  11. Chan, A.T.P.; Tang, S.C.W. Advances in the management of diabetic kidney disease: Beyond sodium-glucose co-transporter 2 inhibitors. Korean J. Nephrol. 2022, 21, 285. [Google Scholar] [CrossRef] [PubMed]
  12. Thotamgari, S.R.; Grewal, U.S.; Sheth, A.R.; Babbili, A.; Dominic, P. Can glucagon-like peptide-1 receptor agonists and dipeptidyl peptidase-4 inhibitors help in mitigating the risk of atrial fibrillation in patients with diabetes? Cardiovasc. Endocrinol. Metab. 2022, 11, e0265. [Google Scholar] [CrossRef] [PubMed]
  13. Pan, Y.; Liu, T.; Wang, X.; Sun, J. Research progress of coumarins and their derivatives in the treatment of diabetes. J. Enzym. Inhib. Med. Chem. 2022, 37, 616–628. [Google Scholar] [CrossRef] [PubMed]
  14. Ajayi, A.M.; Badaki, V.; Adebayo, O.G.; Ben-Azu, B. Plukenetia conophora seed oil ameliorates streptozotocin-induced hyperglycaemia and oxidative stress in rats. Biomark. Biochem. Indic. Expo. Response Susceptibility Chem. 2022, 27, 240–246. [Google Scholar] [CrossRef]
  15. Molitch, M.E.; Adler, A.I.; Flyvbjerg, A.; Nelson, R.G.; So, W.Y.; Wanner, C.; Kasiske, B.L.; Wheeler, D.C.; de Zeeuw, D.; Mogensen, C.E. Diabetic kidney disease: A clinical update from Kidney Disease: Improving Global Outcomes. Kidney Int. 2015, 87, 20–30. [Google Scholar] [CrossRef]
  16. Gillard, P.; Schnell, O.; Groop, P.H. The nephrological perspective on SGLT-2 inhibitors in type 1 diabetes. Diabetes Res. Clin. Pract. 2020, 170, 108462. [Google Scholar] [CrossRef] [PubMed]
  17. Su, H.; Wan, C.; Song, A.; Qiu, Y.; Xiong, W.; Zhang, C. Oxidative Stress and Renal Fibrosis: Mechanisms and Therapies. Adv. Exp. Med. Biol. 2019, 1165, 585–604. [Google Scholar] [PubMed]
  18. Yang, Q.; Wu, F.R.; Wang, J.N.; Gao, L.; Jiang, L.; Li, H.D.; Ma, Q.; Liu, X.Q.; Wei, B.; Zhou, L.; et al. Nox4 in renal diseases: An update. Free Radic. Biol. Med. 2018, 124, 466–472. [Google Scholar] [CrossRef]
  19. Wan, C.; Su, H.; Zhang, C. Role of NADPH Oxidase in Metabolic Disease-Related Renal Injury: An Update. Oxidative Med. Cell. Longev. 2016, 2016, 7813072. [Google Scholar] [CrossRef] [PubMed]
  20. Rhee, E.P. NADPH Oxidase 4 at the Nexus of Diabetes, Reactive Oxygen Species, and Renal Metabolism. J. Am. Soc. Nephrol. JASN 2016, 27, 337–339. [Google Scholar] [CrossRef]
  21. Li, K.; Li, Q. LINC00323 mediates the role of M1 macrophage polarization in diabetic nephropathy through PI3K/AKT signaling pathway. Hum. Immunol. 2021, 82, 960–967. [Google Scholar] [CrossRef] [PubMed]
  22. Calle, P.; Hotter, G. Macrophage Phenotype and Fibrosis in Diabetic Nephropathy. Int. J. Mol. Sci. 2020, 21, 2806. [Google Scholar] [CrossRef]
  23. Tang, S.C.W.; Yiu, W.H. Innate immunity in diabetic kidney disease. Nat. Rev. Nephrol. 2020, 16, 206–222. [Google Scholar] [CrossRef] [PubMed]
  24. Frangogiannis, N.G. Transforming growth factor-β in myocardial disease. Nat. Rev. Cardiol. 2022, 19, 435–455. [Google Scholar] [CrossRef] [PubMed]
  25. Gu, Y.Y.; Liu, X.S.; Huang, X.R.; Yu, X.Q.; Lan, H.Y. Diverse Role of TGF-β in Kidney Disease. Front. Cell Dev. Biol. 2020, 8, 123. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, Y.; Meng, X.M.; Huang, X.R.; Lan, H.Y. The preventive and therapeutic implication for renal fibrosis by targetting TGF-β/Smad3 signaling. Clin. Sci. 2018, 132, 1403–1415. [Google Scholar] [CrossRef]
  27. Bhargava, P.; Schnellmann, R.G. Mitochondrial energetics in the kidney. Nat. Rev. Nephrol. 2017, 13, 629–646. [Google Scholar] [CrossRef]
  28. Victor, P.; Umapathy, D.; George, L.; Juttada, U.; Ganesh, G.V.; Amin, K.N.; Viswanathan, V.; Ramkumar, K.M. Crosstalk between endoplasmic reticulum stress and oxidative stress in the progression of diabetic nephropathy. Cell Stress Chaperones 2021, 26, 311–321. [Google Scholar] [CrossRef] [PubMed]
  29. Ruiz, S.; Pergola, P.E.; Zager, R.A.; Vaziri, N.D. Targeting the transcription factor Nrf2 to ameliorate oxidative stress and inflammation in chronic kidney disease. Kidney Int. 2013, 83, 1029–1041. [Google Scholar] [CrossRef] [PubMed]
  30. Kimura, T.; Isaka, Y.; Yoshimori, T. Autophagy and kidney inflammation. Autophagy 2017, 13, 997–1003. [Google Scholar] [CrossRef]
  31. Yang, D.; Livingston, M.J.; Liu, Z.; Dong, G.; Zhang, M.; Chen, J.K.; Dong, Z. Autophagy in diabetic kidney disease: Regulation, pathological role and therapeutic potential. Cell. Mol. Life Sci. CMLS 2018, 75, 669–688. [Google Scholar] [CrossRef] [PubMed]
  32. Liang, S.; Wu, Y.S.; Li, D.Y.; Tang, J.X.; Liu, H.F. Autophagy and Renal Fibrosis. Aging Dis. 2022, 13, 712–731. [Google Scholar] [CrossRef]
  33. Tang, C.; Livingston, M.J.; Liu, Z.; Dong, Z. Autophagy in kidney homeostasis and disease. Nat. Rev. Nephrol. 2020, 16, 489–508. [Google Scholar] [CrossRef] [PubMed]
  34. Neumiller, J.J.; Alicic, R.Z.; Tuttle, K.R. Therapeutic Considerations for Antihyperglycemic Agents in Diabetic Kidney Disease. J. Am. Soc. Nephrol. JASN 2017, 28, 2263–2274. [Google Scholar] [CrossRef]
  35. Bae, J.H.; Park, E.G.; Kim, S.; Kim, S.G.; Hahn, S.; Kim, N.H. Comparative Renal Effects of Dipeptidyl Peptidase-4 Inhibitors and Sodium-Glucose Cotransporter 2 Inhibitors on Individual Outcomes in Patients with Type 2 Diabetes: A Systematic Review and Network Meta-Analysis. Endocrinol. Metab. 2021, 36, 388–400. [Google Scholar] [CrossRef] [PubMed]
  36. Hammoud, S.H.; AlZaim, I.; Mougharbil, N.; Koubar, S.; Eid, A.H.; Eid, A.A.; El-Yazbi, A.F. Peri-renal adipose inflammation contributes to renal dysfunction in a non-obese prediabetic rat model: Role of anti-diabetic drugs. Biochem. Pharmacol. 2021, 186, 114491. [Google Scholar] [CrossRef] [PubMed]
  37. Yang, Y.Q.; Tan, H.B.; Zhang, X.Y.; Zhang, Y.Z.; Lin, Q.Y.; Huang, M.Y.; Lin, Z.Y.; Mo, J.Z.; Zhang, Y.; Lan, T.; et al. The Chinese medicine Fufang Zhenzhu Tiaozhi capsule protects against renal injury and inflammation in mice with diabetic kidney disease. J. Ethnopharmacol. 2022, 292, 115165. [Google Scholar] [CrossRef] [PubMed]
  38. Lee, J.; Chung, J.O.; Park, S.Y.; Rajamohan, N.; Singh, A.; Kim, J.; Lowe, V.J.; Lee, S. Natural COA water inhibits mitochondrial ROS-mediated apoptosis through Plk3 downregulation under STZ diabetic stress in pancreatic β-cell lines. Biochem. Biophys. Rep. 2022, 30, 101247. [Google Scholar] [CrossRef] [PubMed]
  39. Fang, D.; Wan, X.; Deng, W.; Guan, H.; Ke, W.; Xiao, H.; Li, Y. Fufang Xue Shuan Tong capsules inhibit renal oxidative stress markers and indices of nephropathy in diabetic rats. Exp. Ther. Med. 2012, 4, 871–876. [Google Scholar] [CrossRef]
  40. Yao, L.; Liang, X.; Qiao, Y.; Chen, B.; Wang, P.; Liu, Z. Mitochondrial dysfunction in diabetic tubulopathy. Metab. Clin. Exp. 2022, 131, 155195. [Google Scholar] [CrossRef]
  41. Yuan, T.; Yang, T.; Chen, H.; Fu, D.; Hu, Y.; Wang, J.; Yuan, Q.; Yu, H.; Xu, W.; Xie, X. New insights into oxidative stress and inflammation during diabetes mellitus-accelerated atherosclerosis. Redox Biol. 2019, 20, 247–260. [Google Scholar] [CrossRef] [PubMed]
  42. Aboolian, A.; Urner, S.; Roden, M.; Jha, J.C.; Jandeleit-Dahm, K. Diabetic Kidney Disease: From Pathogenesis to Novel Treatment Possibilities. In Handbook of Experimental Pharmacology; Springer: Berlin/Heidelberg, Germany, 2022; pp. 269–307. [Google Scholar]
  43. Wu, X.M.; Gao, Y.B.; Xu, L.P.; Zou, D.W.; Zhu, Z.Y.; Wang, X.L.; Yao, W.J.; Luo, L.T.; Tong, Y.; Tian, N.X.; et al. Tongxinluo Inhibits Renal Fibrosis in Diabetic Nephropathy: Involvement of the Suppression of Intercellular Transfer of TGF-β1-Containing Exosomes from GECs to GMCs. Am. J. Chin. Med. 2017, 45, 1075–1092. [Google Scholar] [CrossRef] [PubMed]
  44. Simón, J.; Casado-Andrés, M.; Goikoetxea-Usandizaga, N.; Serrano-Maciá, M.; Martínez-Chantar, M.L. Nutraceutical Properties of Polyphenols against Liver Diseases. Nutrients 2020, 12, 3517. [Google Scholar] [CrossRef]
  45. Li, A.N.; Li, S.; Zhang, Y.J.; Xu, X.R.; Chen, Y.M.; Li, H.B. Resources and biological activities of natural polyphenols. Nutrients 2014, 6, 6020–6047. [Google Scholar] [CrossRef] [PubMed]
  46. Ahamad, J.; Toufeeq, I.; Khan, M.A.; Ameen, M.S.M.; Anwer, E.T.; Uthirapathy, S.; Mir, S.R.; Ahmad, J. Oleuropein: A natural antioxidant molecule in the treatment of metabolic syndrome. Phytother. Res. PTR 2019, 33, 3112–3128. [Google Scholar] [CrossRef]
  47. Marrano, N.; Spagnuolo, R.; Biondi, G.; Cignarelli, A.; Perrini, S.; Vincenti, L.; Laviola, L.; Giorgino, F.; Natalicchio, A. Effects of Extra Virgin Olive Oil Polyphenols on Beta-Cell Function and Survival. Plants 2021, 10, 286. [Google Scholar] [CrossRef] [PubMed]
  48. Chaari, A. Inhibition of human islet amyloid polypeptide aggregation and cellular toxicity by oleuropein and derivatives from olive oil. Int. J. Biol. Macromol. 2020, 162, 284–300. [Google Scholar] [CrossRef] [PubMed]
  49. Zhang, Z.; Zhao, H.; Wang, A. Oleuropein alleviates gestational diabetes mellitus by activating AMPK signaling. Endocr. Connect. 2021, 10, 45–53. [Google Scholar] [CrossRef]
  50. Liu, Y.; Dai, W.; Ye, S. The olive constituent oleuropein exerts nephritic protective effects on diabetic nephropathy in db/db mice. Arch. Physiol. Biochem. 2022, 128, 455–462. [Google Scholar] [CrossRef]
  51. Acar-Tek, N.; Ağagündüz, D. Olive Leaf (Olea europaea L. folium): Potential Effects on Glycemia and Lipidemia. Ann. Nutr. Metab. 2020, 76, 10–15. [Google Scholar] [CrossRef] [PubMed]
  52. Michno, A.; Grużewska, K.; Ronowska, A.; Gul-Hinc, S.; Zyśk, M.; Jankowska-Kulawy, A. Resveratrol Inhibits Metabolism and Affects Blood Platelet Function in Type 2 Diabetes. Nutrients 2022, 14, 1633. [Google Scholar] [CrossRef] [PubMed]
  53. Szkudelski, T.; Szkudelska, K. Resveratrol and diabetes: From animal to human studies. Biochim. Biophys. Acta 2015, 1852, 1145–1154. [Google Scholar] [CrossRef]
  54. Choudhury, H.; Pandey, M.; Hua, C.K.; Mun, C.S.; Jing, J.K.; Kong, L.; Ern, L.Y.; Ashraf, N.A.; Kit, S.W.; Yee, T.S.; et al. An update on natural compounds in the remedy of diabetes mellitus: A systematic review. J. Tradit. Complement. Med. 2018, 8, 361–376. [Google Scholar] [CrossRef] [PubMed]
  55. Mahjabeen, W.; Khan, D.A.; Mirza, S.A. Role of resveratrol supplementation in regulation of glucose hemostasis, inflammation and oxidative stress in patients with diabetes mellitus type 2: A randomized, placebo-controlled trial. Complement. Ther. Med. 2022, 66, 102819. [Google Scholar] [CrossRef] [PubMed]
  56. Ma, N.; Zhang, Y. Effects of resveratrol therapy on glucose metabolism, insulin resistance, inflammation, and renal function in the elderly patients with type 2 diabetes mellitus: A randomized controlled clinical trial protocol. Medicine 2022, 101, e30049. [Google Scholar] [CrossRef]
  57. Zhang, J.; Dong, X.J.; Ding, M.R.; You, C.Y.; Lin, X.; Wang, Y.; Wu, M.J.; Xu, G.F.; Wang, G.D. Resveratrol decreases high glucose-induced apoptosis in renal tubular cells via suppressing endoplasmic reticulum stress. Mol. Med. Rep. 2020, 22, 4367–4375. [Google Scholar] [CrossRef]
  58. Wang, F.; Li, R.; Zhao, L.; Ma, S.; Qin, G. Resveratrol ameliorates renal damage by inhibiting oxidative stress-mediated apoptosis of podocytes in diabetic nephropathy. Eur. J. Pharm. 2020, 885, 173387. [Google Scholar] [CrossRef] [PubMed]
  59. Zhang, T.; Chi, Y.; Kang, Y.; Lu, H.; Niu, H.; Liu, W.; Li, Y. Resveratrol ameliorates podocyte damage in diabetic mice via SIRT1/PGC-1α mediated attenuation of mitochondrial oxidative stress. J. Cell. Physiol. 2019, 234, 5033–5043. [Google Scholar] [CrossRef]
  60. Du, L.; Wang, L.; Wang, B.; Wang, J.; Hao, M.; Chen, Y.B.; Li, X.Z.; Li, Y.; Jiang, Y.F.; Li, C.C.; et al. A novel compound AB38b attenuates oxidative stress and ECM protein accumulation in kidneys of diabetic mice through modulation of Keap1/Nrf2 signaling. Acta Pharmacol. Sin. 2020, 41, 358–372. [Google Scholar] [CrossRef]
  61. Qiao, Y.; Gao, K.; Wang, Y.; Wang, X.; Cui, B. Resveratrol ameliorates diabetic nephropathy in rats through negative regulation of the p38 MAPK/TGF-β1 pathway. Exp. Ther. Med. 2017, 13, 3223–3230. [Google Scholar] [CrossRef]
  62. Gu, W.; Wang, X.; Zhao, H.; Geng, J.; Li, X.; Zheng, K.; Guan, Y.; Hou, X.; Wang, C.; Song, G. Resveratrol ameliorates diabetic kidney injury by reducing lipotoxicity and modulates expression of components of the junctional adhesion molecule-like/sirtuin 1 lipid metabolism pathway. Eur. J. Pharm. 2022, 918, 174776. [Google Scholar] [CrossRef]
  63. Zhu, H.; Zhong, S.; Yan, H.; Wang, K.; Chen, L.; Zhou, M.; Li, Y. Resveratrol reverts Streptozotocin-induced diabetic nephropathy. Front. Biosci. 2020, 25, 699–709. [Google Scholar]
  64. Salehi, B.; Mishra, A.P.; Nigam, M.; Sener, B.; Kilic, M.; Sharifi-Rad, M.; Fokou, P.V.T.; Martins, N.; Sharifi-Rad, J. Resveratrol: A Double-Edged Sword in Health Benefits. Biomedicines 2018, 6, 91. [Google Scholar] [CrossRef]
  65. Shaito, A.; Posadino, A.M.; Younes, N.; Hasan, H.; Halabi, S.; Alhababi, D.; Al-Mohannadi, A.; Abdel-Rahman, W.M.; Eid, A.H.; Nasrallah, G.K.; et al. Potential Adverse Effects of Resveratrol: A Literature Review. Int. J. Mol. Sci. 2020, 21, 2084. [Google Scholar] [CrossRef]
  66. Dong, Z.; Bian, L.; Wang, Y.L.; Sun, L.M. Gastrodin protects against high glucose-induced cardiomyocyte toxicity via GSK-3β-mediated nuclear translocation of Nrf2. Hum. Exp. Toxicol. 2021, 40, 1584–1597. [Google Scholar] [CrossRef]
  67. Bai, Y.; Mo, K.; Wang, G.; Chen, W.; Zhang, W.; Guo, Y.; Sun, Z. Intervention of Gastrodin in Type 2 Diabetes Mellitus and Its Mechanism. Front. Pharmacol. 2021, 12, 710722. [Google Scholar] [CrossRef]
  68. Deng, C.-K.; Mu, Z.-H.; Miao, Y.-H.; Liu, Y.-D.; Zhou, L.; Huang, Y.-J.; Zhang, F.; Wang, Y.-Y.; Yang, Z.-H.; Qian, Z.-Y.; et al. Gastrodin Ameliorates Motor Learning Deficits Through Preserving Cerebellar Long-Term Depression Pathways in Diabetic Rats. Front. Neurosci. 2019, 13, 1239. [Google Scholar] [CrossRef]
  69. Xu, G.; Huang, K.; Zhou, J. Hepatic AMP Kinase as a Potential Target for Treating Nonalcoholic Fatty Liver Disease: Evidence from Studies of Natural Products. Curr. Med. Chem. 2018, 25, 889–907. [Google Scholar] [CrossRef] [PubMed]
  70. Ye, T.; Meng, X.; Zhai, Y.; Xie, W.; Wang, R.; Sun, G.; Sun, X. Gastrodin Ameliorates Cognitive Dysfunction in Diabetes Rat Model via the Suppression of Endoplasmic Reticulum Stress and NLRP3 Inflammasome Activation. Front. Pharmacol. 2018, 9, 1346. [Google Scholar] [CrossRef]
  71. Huang, L.; Shao, M.; Zhu, Y. Gastrodin inhibits high glucose-induced inflammation, oxidative stress and apoptosis in podocytes by activating the AMPK/Nrf2 signaling pathway. Exp. Ther. Med. 2022, 23, 168. [Google Scholar] [CrossRef]
  72. Liu, Y.; Gao, J.; Peng, M.; Meng, H.; Ma, H.; Cai, P.; Xu, Y.; Zhao, Q.; Si, G. A Review on Central Nervous System Effects of Gastrodin. Front. Pharmacol. 2018, 9, 24. [Google Scholar] [CrossRef] [PubMed]
  73. Lai, Y.; Wang, R.; Li, W.; Zhu, H.; Fei, S.; Shi, H.; Lu, N.; Ung, C.O.L.; Hu, H.; Han, S. Clinical and economic analysis of Gastrodin injection for dizziness or vertigo: A retrospective cohort study based on electronic health records in China. Chin. Med. 2022, 17, 6. [Google Scholar] [CrossRef]
  74. Liu, J.; Zhao, Y.; Zhu, L.; Han, C. Clinical Research Progress on Treatment of Diabetic Peripheral Neuropathy by Acupoint Injection of Traditional Chinese Medicine. Clin. J. Tradit. Chin. Med. 2018, 30, 1574–1577. [Google Scholar]
  75. Ponikvar-Svet, M.; Zeiger, D.N.; Liebman, J.F. Alkaloids and Selected Topics in Their Thermochemistry. Molecules 2021, 26, 6715. [Google Scholar] [CrossRef]
  76. Abookleesh, F.L.; Al-Anzi, B.S.; Ullah, A. Potential Antiviral Action of Alkaloids. Molecules 2022, 27, 903. [Google Scholar] [CrossRef] [PubMed]
  77. Mohamadi, N.; Sharififar, F.; Pournamdari, M.; Ansari, M. A Review on Biosynthesis, Analytical Techniques, and Pharmacological Activities of Trigonelline as a Plant Alkaloid. J. Diet. Suppl. 2018, 15, 207–222. [Google Scholar] [CrossRef]
  78. Shao, X.; Chen, C.; Miao, C.; Yu, X.; Li, X.; Geng, J.; Fan, D.; Lin, X.; Chen, Z.; Shi, Y. Expression analysis of microRNAs and their target genes during experimental diabetic renal lesions in rats administered with ginsenoside Rb1 and trigonelline. Pharmazie 2019, 74, 492–498. [Google Scholar] [PubMed]
  79. Chen, C.; Shi, Y.; Ma, J.; Chen, Z.; Zhang, M.; Zhao, Y. Trigonelline reverses high glucose-induced proliferation, fibrosis of mesangial cells via modulation of Wnt signaling pathway. Diabetol. Metab. Syndr. 2022, 14, 28. [Google Scholar] [CrossRef] [PubMed]
  80. Li, Y.; Li, Q.; Wang, C.; Lou, Z.; Li, Q. Trigonelline reduced diabetic nephropathy and insulin resistance in type 2 diabetic rats through peroxisome proliferator-activated receptor-γ. Exp. Ther. Med. 2019, 18, 1331–1337. [Google Scholar] [CrossRef]
  81. Chen, C.; Ma, J.; Miao, C.S.; Zhang, H.; Zhang, M.; Cao, X.; Shi, Y. Trigonelline induces autophagy to protect mesangial cells in response to high glucose via activating the miR-5189-5p-AMPK pathway. Phytomed. Int. J. Phytother. Phytopharm. 2021, 92, 153614. [Google Scholar] [CrossRef]
  82. Zhou, J.; Chan, L.; Zhou, S. Trigonelline: A plant alkaloid with therapeutic potential for diabetes and central nervous system disease. Curr. Med. Chem. 2012, 19, 3523–3531. [Google Scholar] [CrossRef]
  83. Jia, X.; Shao, W.; Tian, S. Berberine alleviates myocardial ischemia-reperfusion injury by inhibiting inflammatory response and oxidative stress: The key function of miR-26b-5p-mediated PTGS2/MAPK signal transduction. Pharm. Biol. 2022, 60, 652–663. [Google Scholar] [CrossRef] [PubMed]
  84. Akdad, M.; Ameziane, R.; Khallouki, F.; Bakri, Y.; Eddouks, M. Antidiabetic Phytocompounds Acting as Glucose Transport Stimulators. Endocr. Metab. Immune Disord. Drug Targets 2022. [Google Scholar] [CrossRef] [PubMed]
  85. Xia, S.; Ma, L.; Wang, G.; Yang, J.; Zhang, M.; Wang, X.; Su, J.; Xie, M. In vitro Antimicrobial Activity and the Mechanism of Berberine against Methicillin-Resistant Staphylococcus aureus Isolated from Bloodstream Infection Patients. Infect. Drug Resist. 2022, 15, 1933–1944. [Google Scholar] [CrossRef] [PubMed]
  86. Dian, L.; Xu, Z.; Sun, Y.; Li, J.; Lu, H.; Zheng, M.; Wang, J.; Drobot, L.; Horak, I. Berberine alkaloids inhibit the proliferation and metastasis of breast carcinoma cells involving Wnt/β-catenin signaling and EMT. Phytochemistry 2022, 200, 113217. [Google Scholar] [CrossRef] [PubMed]
  87. Guo, J.; Chen, H.; Zhang, X.; Lou, W.; Zhang, P.; Qiu, Y.; Zhang, C.; Wang, Y.; Liu, W.J. The Effect of Berberine on Metabolic Profiles in Type 2 Diabetic Patients: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. Oxidative Med. Cell. Longev. 2021, 2021, 2074610. [Google Scholar] [CrossRef]
  88. Ni, W.J.; Zhou, H.; Ding, H.H.; Tang, L.Q. Berberine ameliorates renal impairment and inhibits podocyte dysfunction by targeting the phosphatidylinositol 3-kinase-protein kinase B pathway in diabetic rats. J. Diabetes Investig. 2020, 11, 297–306. [Google Scholar] [CrossRef]
  89. Ni, W.J.; Guan, X.M.; Zeng, J.; Zhou, H.; Meng, X.M.; Tang, L.Q. Berberine regulates mesangial cell proliferation and cell cycle to attenuate diabetic nephropathy through the PI3K/Akt/AS160/GLUT1 signalling pathway. J. Cell. Mol. Med. 2022, 26, 1144–1155. [Google Scholar] [CrossRef]
  90. Zhang, X.; Guan, T.; Yang, B.; Chi, Z.; Wan, Q.; Gu, H.F. Protective effect of berberine on high glucose and hypoxia-induced apoptosis via the modulation of HIF-1α in renal tubular epithelial cells. Am. J. Transl. Res. 2019, 11, 669–682. [Google Scholar]
  91. Sun, J.; Chen, X.; Liu, T.; Jiang, X.; Wu, Y.; Yang, S.; Hua, W.; Li, Z.; Huang, H.; Ruan, X.; et al. Berberine Protects against Palmitate-Induced Apoptosis in Tubular Epithelial Cells by Promoting Fatty Acid Oxidation. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2018, 24, 1484–1492. [Google Scholar] [CrossRef]
  92. Rong, Q.; Han, B.; Li, Y.; Yin, H.; Li, J.; Hou, Y. Berberine Reduces Lipid Accumulation by Promoting Fatty Acid Oxidation in Renal Tubular Epithelial Cells of the Diabetic Kidney. Front. Pharmacol. 2022, 12, 729384. [Google Scholar] [CrossRef] [PubMed]
  93. Xu, J.; Liu, L.; Gan, L.; Hu, Y.; Xiang, P.; Xing, Y.; Zhu, J.; Ye, S. Berberine Acts on C/EBPβ/lncRNA Gas5/miR-18a-5p Loop to Decrease the Mitochondrial ROS Generation in HK-2 Cells. Front. Endocrinol. 2021, 12, 675834. [Google Scholar] [CrossRef] [PubMed]
  94. Qin, X.; Jiang, M.; Zhao, Y.; Gong, J.; Su, H.; Yuan, F.; Fang, K.; Yuan, X.; Yu, X.; Dong, H.; et al. Berberine protects against diabetic kidney disease via promoting PGC-1α-regulated mitochondrial energy homeostasis. Br. J. Pharmacol. 2020, 177, 3646–3661. [Google Scholar] [CrossRef] [PubMed]
  95. Qin, X.; Zhao, Y.; Gong, J.; Huang, W.; Su, H.; Yuan, F.; Fang, K.; Wang, D.; Li, J.; Zou, X.; et al. Berberine Protects Glomerular Podocytes via Inhibiting Drp1-Mediated Mitochondrial Fission and Dysfunction. Theranostics 2019, 9, 1698–1713. [Google Scholar] [CrossRef]
  96. Yang, G.; Zhao, Z.; Zhang, X.; Wu, A.; Huang, Y.; Miao, Y.; Yang, M. Effect of berberine on the renal tubular epithelial-to-mesenchymal transition by inhibition of the Notch/snail pathway in diabetic nephropathy model KKAy mice. Drug Des. Dev. Ther. 2017, 11, 1065–1079. [Google Scholar] [CrossRef]
  97. Zhu, L.; Han, J.; Yuan, R.; Xue, L.; Pang, W. Berberine ameliorates diabetic nephropathy by inhibiting TLR4/NF-κB pathway. Biol. Res. 2018, 51, 9. [Google Scholar] [CrossRef]
  98. Li, C.; Guan, X.M.; Wang, R.Y.; Xie, Y.S.; Zhou, H.; Ni, W.J.; Tang, L.Q. Berberine mitigates high glucose-induced podocyte apoptosis by modulating autophagy via the mTOR/P70S6K/4EBP1 pathway. Life Sci. 2020, 243, 117277. [Google Scholar] [CrossRef]
  99. Mohammadzadeh, N.; Mehri, S.; Hosseinzadeh, H. Berberis vulgaris and its constituent berberine as antidotes and protective agents against natural or chemical toxicities. Iran. J. Basic Med. Sci. 2017, 20, 538–551. [Google Scholar]
  100. Imenshahidi, M.; Hosseinzadeh, H. Berberine and barberry (Berberis vulgaris): A clinical review. Phytother. Res. PTR 2019, 33, 504–523. [Google Scholar] [CrossRef]
  101. Spinozzi, S.; Colliva, C.; Camborata, C.; Roberti, M.; Ianni, C.; Neri, F.; Calvarese, C.; Lisotti, A.; Mazzella, G.; Roda, A. Berberine and Its Metabolites: Relationship between Physicochemical Properties and Plasma Levels after Administration to Human Subjects. J. Nat. Prod. 2014, 77, 766–772. [Google Scholar] [CrossRef]
  102. Liu, Y.-T.; Hao, H.-P.; Xie, H.-G.; Lai, L.; Wang, Q.; Liu, C.-X.; Wang, G.-J. Extensive Intestinal First-Pass Elimination and Predominant Hepatic Distribution of Berberine Explain Its Low Plasma Levels in Rats. Drug Metab. Dispos. 2010, 38, 1779–1784. [Google Scholar] [CrossRef] [PubMed]
  103. Gao, X.; Sun, B.; Hou, Y.; Liu, L.; Sun, J.; Xu, F.; Li, D.; Hua, H. Anti-breast cancer sinomenine derivatives via mechanisms of apoptosis induction and metastasis reduction. J. Enzym. Inhib. Med. Chem. 2022, 37, 1870–1883. [Google Scholar] [CrossRef] [PubMed]
  104. Li, Y.; Xie, H.; Zhang, H. Protective effect of sinomenine against inflammation and oxidative stress in gestational diabetes mellitus in female rats via TLR4/MyD88/NF-κB signaling pathway. J. Food Biochem. 2021, 45, e13952. [Google Scholar] [CrossRef] [PubMed]
  105. Yin, Q.; Xia, Y.; Wang, G. Sinomenine alleviates high glucose-induced renal glomerular endothelial hyperpermeability by inhibiting the activation of RhoA/ROCK signaling pathway. Biochem. Biophys. Res. Commun. 2016, 477, 881–886. [Google Scholar] [CrossRef]
  106. Zhang, L.; Wang, J. Sinomenine alleviates glomerular endothelial permeability by activating the C/EBP-α/claudin-5 signaling pathway. Hum. Cell 2022, 35, 1453–1463. [Google Scholar] [CrossRef]
  107. Zhu, M.; Wang, H.; Chen, J.; Zhu, H. Sinomenine improve diabetic nephropathy by inhibiting fibrosis and regulating the JAK2/STAT3/SOCS1 pathway in streptozotocin-induced diabetic rats. Life Sci. 2021, 265, 118855. [Google Scholar] [CrossRef]
  108. Huang, Z.; Mao, X.; Chen, J.; He, J.; Shi, S.; Gui, M.; Gao, H.; Hong, Z. The Efficacy and Safety of Zhengqing Fengtongning for Knee Osteoarthritis: A Systematic Review and Meta-Analysis of Randomized Clinical Trials. Evid.-Based Complement. Altern. Med. 2022, 2022, 2768444. [Google Scholar] [CrossRef] [PubMed]
  109. Huang, Z.; Mao, X.; Chen, J.; He, J.; Shi, S.; Gui, M.; Gao, H.; Hong, Z. Sinomenine hydrochloride injection for knee osteoarthritis: A protocol for systematic review and meta-analysis. Medicine 2022, 101, e28503. [Google Scholar] [CrossRef]
  110. Zhang, Y.S.; Han, J.Y.; Iqbal, O.; Liang, A.H. Research Advances and Prospects on Mechanism of Sinomenin on Histamine Release and the Binding to Histamine Receptors. Int. J. Mol. Sci. 2018, 20, 70. [Google Scholar] [CrossRef]
  111. Huang, H.; Zhang, E.B.; Yi, O.Y.; Wu, H.; Deng, G.; Huang, Y.M.; Liu, W.L.; Yan, J.Y.; Cai, X. Sex-related differences in safety profiles, pharmacokinetics and tissue distribution of sinomenine hydrochloride in rats. Arch. Toxicol. 2022. [Google Scholar] [CrossRef]
  112. Middleton, E., Jr.; Kandaswami, C.; Theoharides, T.C. The effects of plant flavonoids on mammalian cells: Implications for inflammation, heart disease, and cancer. Pharmacol. Rev. 2000, 52, 673–751. [Google Scholar] [PubMed]
  113. Dias, M.C.; Pinto, D.; Silva, A.M.S. Plant Flavonoids: Chemical Characteristics and Biological Activity. Molecules 2021, 26, 5377. [Google Scholar] [CrossRef] [PubMed]
  114. Šamec, D.; Karalija, E.; Šola, I.; Vujčić Bok, V.; Salopek-Sondi, B. The Role of Polyphenols in Abiotic Stress Response: The Influence of Molecular Structure. Plants 2021, 10, 118. [Google Scholar] [CrossRef] [PubMed]
  115. Singh, S.; Sharma, A.; Monga, V.; Bhatia, R. Compendium of naringenin: Potential sources, analytical aspects, chemistry, nutraceutical potentials and pharmacological profile. Crit. Rev. Food Sci. Nutr. 2022, 1–32. [Google Scholar] [CrossRef] [PubMed]
  116. Nyane, N.A.; Tlaila, T.B.; Malefane, T.G.; Ndwandwe, D.E.; Owira, P.M.O. Metformin-like antidiabetic, cardio-protective and non-glycemic effects of naringenin: Molecular and pharmacological insights. Eur. J. Pharmacol. 2017, 803, 103–111. [Google Scholar] [CrossRef]
  117. Khan, M.F.; Mathur, A.; Pandey, V.K.; Kakkar, P. Naringenin alleviates hyperglycemia-induced renal toxicity by regulating activating transcription factor 4-C/EBP homologous protein mediated apoptosis. J. Cell Commun. Signal. 2022, 16, 271–291. [Google Scholar] [CrossRef]
  118. Ding, S.; Qiu, H.; Huang, J.; Chen, R.; Zhang, J.; Huang, B.; Zou, X.; Cheng, O.; Jiang, Q. Activation of 20-HETE/PPARs involved in reno-therapeutic effect of naringenin on diabetic nephropathy. Chem. Biol. Interact. 2019, 307, 116–124. [Google Scholar] [CrossRef]
  119. Yan, N.; Wen, L.; Peng, R.; Li, H.; Liu, H.; Peng, H.; Sun, Y.; Wu, T.; Chen, L.; Duan, Q.; et al. Naringenin Ameliorated Kidney Injury through Let-7a/TGFBR1 Signaling in Diabetic Nephropathy. J. Diabetes Res. 2016, 2016, 8738760. [Google Scholar] [CrossRef]
  120. Rebello, C.J.; Beyl, R.A.; Lertora, J.J.L.; Greenway, F.L.; Ravussin, E.; Ribnicky, D.M.; Poulev, A.; Kennedy, B.J.; Castro, H.F.; Campagna, S.R.; et al. Safety and pharmacokinetics of naringenin: A randomized, controlled, single-ascending-dose clinical trial. Diabetes Obes. Metab. 2020, 22, 91–98. [Google Scholar] [CrossRef]
  121. Ranawat, P.; Bakshi, N. Naringenin; a bioflavonoid, impairs the reproductive potential of male mice. Toxicol. Mech. Methods 2017, 27, 417–427. [Google Scholar] [CrossRef]
  122. Yan, L.; Vaghari-Tabari, M.; Malakoti, F.; Moein, S.; Qujeq, D.; Yousefi, B.; Asemi, Z. Quercetin: An effective polyphenol in alleviating diabetes and diabetic complications. Crit. Rev. Food Sci. Nutr. 2022, 1–24. [Google Scholar] [CrossRef] [PubMed]
  123. Feng, X.; Bu, F.; Huang, L.; Xu, W.; Wang, W.; Wu, Q. Preclinical evidence of the effect of quercetin on diabetic nephropathy: A meta-analysis of animal studies. Eur. J. Pharm. 2022, 921, 174868. [Google Scholar] [CrossRef]
  124. Adeshara, K.A.; Bangar, N.; Diwan, A.G.; Tupe, R.S. Plasma glycation adducts and various RAGE isoforms are intricately associated with oxidative stress and inflammatory markers in type 2 diabetes patients with vascular complications. Diabetes Metab. Syndr. 2022, 16, 102441. [Google Scholar] [CrossRef] [PubMed]
  125. Tang, L.; Li, K.; Zhang, Y.; Li, H.; Li, A.; Xu, Y.; Wei, B. Quercetin liposomes ameliorate streptozotocin-induced diabetic nephropathy in diabetic rats. Sci. Rep. 2020, 10, 2440. [Google Scholar] [CrossRef] [PubMed]
  126. Liu, Y.; Li, Y.; Xu, L.; Shi, J.; Yu, X.; Wang, X.; Li, X.; Jiang, H.; Yang, T.; Yin, X.; et al. Quercetin Attenuates Podocyte Apoptosis of Diabetic Nephropathy Through Targeting EGFR Signaling. Front. Pharmacol. 2021, 12, 792777. [Google Scholar] [CrossRef] [PubMed]
  127. Wan, H.; Wang, Y.; Pan, Q.; Chen, X.; Chen, S.; Li, X.; Yao, W. Quercetin attenuates the proliferation, inflammation, and oxidative stress of high glucose-induced human mesangial cells by regulating the miR-485-5p/YAP1 pathway. Int. J. Immunopathol. Pharm. 2022, 36, 20587384211066440. [Google Scholar] [CrossRef]
  128. Lei, D.; Chengcheng, L.; Xuan, Q.; Yibing, C.; Lei, W.; Hao, Y.; Xizhi, L.; Yuan, L.; Xiaoxing, Y.; Qian, L. Quercetin inhibited mesangial cell proliferation of early diabetic nephropathy through the Hippo pathway. Pharm. Res. 2019, 146, 104320. [Google Scholar] [CrossRef]
  129. Yuan, Y.; Sun, H.; Sun, Z. Advanced glycation end products (AGEs) increase renal lipid accumulation: A pathogenic factor of diabetic nephropathy (DN). Lipids Health Dis. 2017, 16, 126. [Google Scholar] [CrossRef] [PubMed]
  130. Jiang, X.; Yu, J.; Wang, X.; Ge, J.; Li, N. Quercetin improves lipid metabolism via SCAP-SREBP2-LDLr signaling pathway in early stage diabetic nephropathy. Diabetes Metab. Syndr. Obes. Targets Ther. 2019, 12, 827–839. [Google Scholar] [CrossRef] [PubMed]
  131. Wang, C.; Pan, Y.; Zhang, Q.Y.; Wang, F.M.; Kong, L.D. Quercetin and allopurinol ameliorate kidney injury in STZ-treated rats with regulation of renal NLRP3 inflammasome activation and lipid accumulation. PLoS ONE 2012, 7, e38285. [Google Scholar] [CrossRef] [PubMed]
  132. Yi, H.; Peng, H.; Wu, X.; Xu, X.; Kuang, T.; Zhang, J.; Du, L.; Fan, G. The Therapeutic Effects and Mechanisms of Quercetin on Metabolic Diseases: Pharmacological Data and Clinical Evidence. Oxidative Med. Cell. Longev. 2021, 2021, 6678662. [Google Scholar] [CrossRef] [PubMed]
  133. Zhao, S.-Y.; Liao, L.-X.; Tu, P.-F.; Li, W.-W.; Zeng, K.-W. Icariin Inhibits AGE-Induced Injury in PC12 Cells by Directly Targeting Apoptosis Regulator Bax. Oxidative Med. Cell. Longev. 2019, 2019, 7940808. [Google Scholar] [CrossRef] [PubMed]
  134. Ding, X.; Zhao, H.; Qiao, C. Icariin protects podocytes from NLRP3 activation by Sesn2-induced mitophagy through the Keap1-Nrf2/HO-1 axis in diabetic nephropathy. Phytomed. Int. J. Phytother. Phytopharm. 2022, 99, 154005. [Google Scholar] [CrossRef] [PubMed]
  135. Qi, M.Y.; He, Y.H.; Cheng, Y.; Fang, Q.; Ma, R.Y.; Zhou, S.J.; Hao, J.Q. Icariin ameliorates streptozocin-induced diabetic nephropathy through suppressing the TLR4/NF-κB signal pathway. Food Funct. 2021, 12, 1241–1251. [Google Scholar] [CrossRef]
  136. Li, Y.C.; Ding, X.S.; Li, H.M.; Zhang, C. Icariin attenuates high glucose-induced type ΙV collagen and fibronectin accumulation in glomerular mesangial cells by inhibiting transforming growth factor-β production and signalling through G protein-coupled oestrogen receptor 1. Clin. Exp. Pharmacol. Physiol. 2013, 40, 635–643. [Google Scholar] [CrossRef]
  137. Jia, Z.; Wang, K.; Zhang, Y.; Duan, Y.; Xiao, K.; Liu, S.; Ding, X. Icariin Ameliorates Diabetic Renal Tubulointerstitial Fibrosis by Restoring Autophagy via Regulation of the miR-192-5p/GLP-1R Pathway. Front. Pharmacol. 2021, 12, 720387. [Google Scholar] [CrossRef] [PubMed]
  138. Zang, L.; Gao, F.; Huang, A.; Zhang, Y.; Luo, Y.; Chen, L.; Mao, N. Icariin inhibits epithelial mesenchymal transition of renal tubular epithelial cells via regulating the miR-122-5p/FOXP2 axis in diabetic nephropathy rats. J. Pharmacol. Sci. 2022, 148, 204–213. [Google Scholar] [CrossRef] [PubMed]
  139. He, C.; Wang, Z.; Shi, J. Chapter Seven—Pharmacological effects of icariin. In Advances in Pharmacology; Du, G., Ed.; Academic Press: Cambridge, MA, USA, 2020; Volume 87, pp. 179–203. [Google Scholar]
  140. Zhou, F.M.; Huang, J.J.; Hu, X.J.; Wang, J.; Zhu, B.Q.; Ding, Z.S.; Huang, S.; Fang, J.J. Protective effects of flavonoids from the leaves of Carya cathayensis Sarg. against H2O2-induced oxidative damage and apoptosis in vitro. Exp. Ther. Med. 2021, 22, 1443. [Google Scholar] [CrossRef]
  141. Sun, H.; Zhang, N.; Jin, Y.; Xu, H. Cardamonin Promotes the Apoptosis and Chemotherapy Sensitivity to Gemcitabine of Pancreatic Cancer Through Modulating the FOXO3a-FOXM1 Axis. Dose-Response Publ. Int. Hormesis Soc. 2021, 19, 15593258211042163. [Google Scholar] [CrossRef] [PubMed]
  142. Satsu, H.; Shibata, R.; Suzuki, H.; Kimura, S.; Shimizu, M. Inhibitory Effect of Tangeretin and Cardamonin on Human Intestinal SGLT1 Activity In Vitro and Blood Glucose Levels in Mice In Vivo. Nutrients 2021, 13, 3382. [Google Scholar] [CrossRef]
  143. Zhang, T.; Yamamoto, N.; Ashida, H. Chalcones suppress fatty acid-induced lipid accumulation through a LKB1/AMPK signaling pathway in HepG2 cells. Food Funct. 2014, 5, 1134–1141. [Google Scholar] [CrossRef] [PubMed]
  144. Schalkwijk, C.G.; Stehouwer, C.D.A. Methylglyoxal, a Highly Reactive Dicarbonyl Compound, in Diabetes, Its Vascular Complications, and Other Age-Related Diseases. Physiol. Rev. 2020, 100, 407–461. [Google Scholar] [CrossRef]
  145. Cha, S.H.; Hwang, Y.; Heo, S.J.; Jun, H.S. Diphlorethohydroxycarmalol Attenuates Methylglyoxal-Induced Oxidative Stress and Advanced Glycation End Product Formation in Human Kidney Cells. Oxidative Med. Cell. Longev. 2018, 2018, 3654095. [Google Scholar] [CrossRef] [PubMed]
  146. Gao, C.; Fei, X.; Wang, M.; Chen, Q.; Zhao, N. Cardamomin protects from diabetes-induced kidney damage through modulating PI3K/AKT and JAK/STAT signaling pathways in rats. Int. Immunopharmacol. 2022, 107, 108610. [Google Scholar] [CrossRef] [PubMed]
  147. Mohammadi, N.; Asle-Rousta, M.; Rahnema, M.; Amini, R. Morin attenuates memory deficits in a rat model of Alzheimer’s disease by ameliorating oxidative stress and neuroinflammation. Eur. J. Pharm. 2021, 910, 174506. [Google Scholar] [CrossRef] [PubMed]
  148. Miao, Y.; Zhang, C.; Yang, L.; Zeng, X.; Hu, Y.; Xue, X.; Dai, Y.; Wei, Z. The activation of PPARγ enhances Treg responses through up-regulating CD36/CPT1-mediated fatty acid oxidation and subsequent N-glycan branching of TβRII/IL-2Rα. Cell Commun. Signal. CCS 2022, 20, 48. [Google Scholar] [CrossRef]
  149. Issac, P.K.; Velayutham, M.; Guru, A.; Sudhakaran, G.; Pachaiappan, R.; Arockiaraj, J. Protective effect of morin by targeting mitochondrial reactive oxygen species induced by hydrogen peroxide demonstrated at a molecular level in MDCK epithelial cells. Mol. Biol. Rep. 2022, 49, 4269–4279. [Google Scholar] [CrossRef] [PubMed]
  150. Mathur, A.; Pandey, V.K.; Khan, M.F.; Kakkar, P. PHLPP1/Nrf2-Mdm2 axis induces renal apoptosis via influencing nucleo-cytoplasmic shuttling of FoxO1 during diabetic nephropathy. Mol. Cell. Biochem. 2021, 476, 3681–3699. [Google Scholar] [CrossRef]
  151. Ke, Y.Q.; Liu, C.; Hao, J.B.; Lu, L.; Lu, N.N.; Wu, Z.K.; Zhu, S.S.; Chen, X.L. Morin inhibits cell proliferation and fibronectin accumulation in rat glomerular mesangial cells cultured under high glucose condition. Biomed. Pharmacother. 2016, 84, 622–627. [Google Scholar] [CrossRef]
  152. Cho, Y.M.; Onodera, H.; Ueda, M.; Imai, T.; Hirose, M. A 13-week subchronic toxicity study of dietary administered morin in F344 rats. Food Chem. Toxicol. Int. J. Publ. Br. Ind. Biol. Res. Assoc. 2006, 44, 891–897. [Google Scholar] [CrossRef]
  153. Caselli, A.; Cirri, P.; Santi, A.; Paoli, P. Morin: A Promising Natural Drug. Curr. Med. Chem. 2016, 23, 774–791. [Google Scholar] [CrossRef]
  154. Mottaghi, S.; Abbaszadeh, H. The anticarcinogenic and anticancer effects of the dietary flavonoid, morin: Current status, challenges, and future perspectives. Phytother. Res. PTR 2021, 35, 6843–6861. [Google Scholar] [CrossRef] [PubMed]
  155. 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] [PubMed]
  156. Rabbani, N.; Xue, M.; Weickert, M.O.; Thornalley, P.J. Reversal of Insulin Resistance in Overweight and Obese Subjects by trans-Resveratrol and Hesperetin Combination-Link to Dysglycemia, Blood Pressure, Dyslipidemia, and Low-Grade Inflammation. Nutrients 2021, 13, 2374. [Google Scholar] [CrossRef]
  157. Li, J.; Wang, T.; Liu, P.; Yang, F.; Wang, X.; Zheng, W.; Sun, W. Hesperetin ameliorates hepatic oxidative stress and inflammation via the PI3K/AKT-Nrf2-ARE pathway in oleic acid-induced HepG2 cells and a rat model of high-fat diet-induced NAFLD. Food Funct. 2021, 12, 3898–3918. [Google Scholar] [CrossRef]
  158. Gong, Y.; Qin, X.Y.; Zhai, Y.Y.; Hao, H.; Lee, J.; Park, Y.D. Inhibitory effect of hesperetin on α-glucosidase: Molecular dynamics simulation integrating inhibition kinetics. Int. J. Biol. Macromol. 2017, 101, 32–39. [Google Scholar] [CrossRef] [PubMed]
  159. Rasouli, H.; Hosseini-Ghazvini, S.M.; Adibi, H.; Khodarahmi, R. Differential α-amylase/α-glucosidase inhibitory activities of plant-derived phenolic compounds: A virtual screening perspective for the treatment of obesity and diabetes. Food Funct. 2017, 8, 1942–1954. [Google Scholar] [CrossRef] [PubMed]
  160. Abdou, H.M.; Abd Elkader, H.A.E. The potential therapeutic effects of Trifolium alexandrinum extract, hesperetin and quercetin against diabetic nephropathy via attenuation of oxidative stress, inflammation, GSK-3β and apoptosis in male rats. Chem. Biol. Interact. 2022, 352, 109781. [Google Scholar] [CrossRef] [PubMed]
  161. Chen, Y.J.; Kong, L.; Tang, Z.Z.; Zhang, Y.M.; Liu, Y.; Wang, T.Y.; Liu, Y.W. Hesperetin ameliorates diabetic nephropathy in rats by activating Nrf2/ARE/glyoxalase 1 pathway. Biomed. Pharmacother. 2019, 111, 1166–1175. [Google Scholar] [CrossRef] [PubMed]
  162. Hajialyani, M.; Hosein Farzaei, M.; Echeverría, J.; Nabavi, S.M.; Uriarte, E.; Sobarzo-Sánchez, E. Hesperidin as a Neuroprotective Agent: A Review of Animal and Clinical Evidence. Molecules 2019, 24, 648. [Google Scholar] [CrossRef]
  163. Kim, H.J.; Kim, S.H.; Yun, J.M. Fisetin inhibits hyperglycemia-induced proinflammatory cytokine production by epigenetic mechanisms. Evid.-Based Complement. Altern. Med. 2012, 2012, 639469. [Google Scholar] [CrossRef] [PubMed]
  164. Shen, B.; Shangguan, X.; Yin, Z.; Wu, S.; Zhang, Q.; Peng, W.; Li, J.; Zhang, L.; Chen, J. Inhibitory Effect of Fisetin on α-Glucosidase Activity: Kinetic and Molecular Docking Studies. Molecules 2021, 26, 5306. [Google Scholar] [CrossRef]
  165. Jia, Y.; Ma, Y.; Cheng, G.; Zhang, Y.; Cai, S. Comparative Study of Dietary Flavonoids with Different Structures as α-Glucosidase Inhibitors and Insulin Sensitizers. J. Agric. Food Chem. 2019, 67, 10521–10533. [Google Scholar] [CrossRef]
  166. Althunibat, O.Y.; Al Hroob, A.M.; Abukhalil, M.H.; Germoush, M.O.; Bin-Jumah, M.; Mahmoud, A.M. Fisetin ameliorates oxidative stress, inflammation and apoptosis in diabetic cardiomyopathy. Life Sci. 2019, 221, 83–92. [Google Scholar] [CrossRef] [PubMed]
  167. Yan, L.; Jia, Q.; Cao, H.; Chen, C.; Xing, S.; Huang, Y.; Shen, D. Fisetin ameliorates atherosclerosis by regulating PCSK9 and LOX-1 in apoE-/- mice. Exp. Ther. Med. 2021, 21, 25. [Google Scholar] [CrossRef] [PubMed]
  168. JZ, A.L.; BinMowyna, M.N.; AlFaris, N.A.; Alagal, R.I.; El-Kott, A.F.; Al-Farga, A.M. Fisetin protects against streptozotocin-induced diabetic cardiomyopathy in rats by suppressing fatty acid oxidation and inhibiting protein kinase R. Saudi Pharm. J. SPJ Off. Publ. Saudi Pharm. Soc. 2021, 29, 27–42. [Google Scholar]
  169. Sandireddy, R.; Yerra, V.G.; Komirishetti, P.; Areti, A.; Kumar, A. Fisetin Imparts Neuroprotection in Experimental Diabetic Neuropathy by Modulating Nrf2 and NF-κB Pathways. Cell. Mol. Neurobiol. 2016, 36, 883–892. [Google Scholar] [CrossRef]
  170. Zhang, S.; Xue, R.; Geng, Y.; Wang, H.; Li, W. Fisetin Prevents HT22 Cells from High Glucose-Induced Neurotoxicity via PI3K/Akt/CREB Signaling Pathway. Front. Neurosci. 2020, 14, 241. [Google Scholar] [CrossRef]
  171. Dong, W.; Jia, C.; Li, J.; Zhou, Y.; Luo, Y.; Liu, J.; Zhao, Z.; Zhang, J.; Lin, S.; Chen, Y. Fisetin Attenuates Diabetic Nephropathy-Induced Podocyte Injury by Inhibiting NLRP3 Inflammasome. Front. Pharmacol. 2022, 13, 783706. [Google Scholar] [CrossRef]
  172. Ge, C.; Xu, M.; Qin, Y.; Gu, T.; Lou, D.; Li, Q.; Hu, L.; Nie, X.; Wang, M.; Tan, J. Fisetin supplementation prevents high fat diet-induced diabetic nephropathy by repressing insulin resistance and RIP3-regulated inflammation. Food Funct. 2019, 10, 2970–2985. [Google Scholar] [CrossRef]
  173. Couillaud, J.; Leydet, L.; Duquesne, K.; Iacazio, G. The Terpene Mini-Path, a New Promising Alternative for Terpenoids Bio-Production. Genes 2021, 12, 1974. [Google Scholar] [CrossRef] [PubMed]
  174. Chalvin, C.; Drevensek, S.; Gilard, F.; Mauve, C.; Chollet, C.; Morin, H.; Nicol, E.; Héripré, E.; Kriegshauser, L.; Gakière, B.; et al. Sclareol and linalyl acetate are produced by glandular trichomes through the MEP pathway. Hortic. Res. 2021, 8, 206. [Google Scholar] [CrossRef] [PubMed]
  175. Cerri, G.C.; Lima, L.C.F.; Lelis, D.F.; Barcelos, L.D.S.; Feltenberger, J.D.; Mussi, S.V.; Monteiro-Junior, R.S.; Santos, R.; Ferreira, L.A.M.; Santos, S.H.S. Sclareol-loaded lipid nanoparticles improved metabolic profile in obese mice. Life Sci. 2019, 218, 292–299. [Google Scholar] [CrossRef]
  176. Chen, H.L.; Gong, J.Y.; Lin, S.C.; Li, S.; Chiang, Y.C.; Hung, J.H.; Yen, C.C.; Lin, C.C. Effects of Sclareol Against Small Cell Lung Carcinoma and the Related Mechanism: In Vitro and In Vivo Studies. Anticancer Res. 2020, 40, 4947–4960. [Google Scholar] [CrossRef] [PubMed]
  177. Wong, J.; Chiang, Y.F.; Shih, Y.H.; Chiu, C.H.; Chen, H.Y.; Shieh, T.M.; Wang, K.L.; Huang, T.C.; Hong, Y.H.; Hsia, S.M. Salvia sclarea L. Essential Oil Extract and Its Antioxidative Phytochemical Sclareol Inhibit Oxytocin-Induced Uterine Hypercontraction Dysmenorrhea Model by Inhibiting the Ca2+-MLCK-MLC20 Signaling Cascade: An Ex Vivo and In Vivo Study. Antioxidants 2020, 9, 991. [Google Scholar] [CrossRef] [PubMed]
  178. Han, X.; Zhang, J.; Zhou, L.; Wei, J.; Tu, Y.; Shi, Q.; Zhang, Y.; Ren, J.; Wang, Y.; Ying, H.; et al. Sclareol ameliorates hyperglycemia-induced renal injury through inhibiting the MAPK/NF-κB signaling pathway. Phytother. Res. PTR 2022, 36, 2511–2523. [Google Scholar] [CrossRef]
  179. Du, J.; Chen, C.; Sun, Y.; Zheng, L.; Wang, W. Ponicidin suppresses HT29 cell growth via the induction of G1 cell cycle arrest and apoptosis. Mol. Med. Rep. 2015, 12, 5816–5820. [Google Scholar] [CrossRef]
  180. Zhang, Z.; Xu, J.; Liu, B.; Chen, F.; Li, J.; Liu, Y.; Zhu, J.; Shen, C. Ponicidin inhibits pro-inflammatory cytokine TNF-α-induced epithelial-mesenchymal transition and metastasis of colorectal cancer cells via suppressing the AKT/GSK-3β/Snail pathway. Inflammopharmacology 2019, 27, 627–638. [Google Scholar] [CrossRef]
  181. Islam, M.T.; Biswas, S.; Bagchi, R.; Khan, M.R.; Khalipha, A.B.R.; Rouf, R.; Uddin, S.J.; Shilpi, J.A.; Bardaweel, S.K.; Sabbah, D.A.; et al. Ponicidin as a promising anticancer agent: Its biological and biopharmaceutical profile along with a molecular docking study. Biotechnol. Appl. Biochem. 2019, 66, 434–444. [Google Scholar] [CrossRef]
  182. An, S.; Li, Y.; Jia, X.; Yang, Y.; Jia, X.; Jia, X.; Xue, W. Ponicidin attenuates streptozotocin-induced diabetic nephropathy in rats via modulating hyperlipidemia, oxidative stress, and inflammatory markers. J. Biochem. Mol. Toxicol. 2022, 36, e22988. [Google Scholar] [CrossRef]
  183. Peng, H.; You, L.; Yang, C.; Wang, K.; Liu, M.; Yin, D.; Xu, Y.; Dong, X.; Yin, X.; Ni, J. Ginsenoside Rb1 Attenuates Triptolide-Induced Cytotoxicity in HL-7702 Cells via the Activation of Keap1/Nrf2/ARE Pathway. Front. Pharmacol. 2021, 12, 723784. [Google Scholar] [CrossRef] [PubMed]
  184. Yang, J.; Tang, X.; Ke, X.; Dai, Y.; Shi, J. Triptolide Suppresses NF-κB-Mediated Inflammatory Responses and Activates Expression of Nrf2-Mediated Antioxidant Genes to Alleviate Caerulein-Induced Acute Pancreatitis. Int. J. Mol. Sci. 2022, 23, 1252. [Google Scholar] [CrossRef] [PubMed]
  185. Song, X.; He, H.; Zhang, Y.; Fan, J.; Wang, L. Mechanisms of action of triptolide against colorectal cancer: Insights from proteomic and phosphoproteomic analyses. Aging 2022, 14, 3084–3104. [Google Scholar] [CrossRef] [PubMed]
  186. Gao, Q.; Shen, W.; Qin, W.; Zheng, C.; Zhang, M.; Zeng, C.; Wang, S.; Wang, J.; Zhu, X.; Liu, Z. Treatment of db/db diabetic mice with triptolide: A novel therapy for diabetic nephropathy. Nephrol. Dial. Transplant. Off. Publ. Eur. Dial. Transpl. Assoc. Eur. Ren. Assoc. 2010, 25, 3539–3547. [Google Scholar] [CrossRef]
  187. Liang, D.; Mai, H.; Ruan, F.; Fu, H. The Efficacy of Triptolide in Preventing Diabetic Kidney Diseases: A Systematic Review and Meta-Analysis. Front. Pharmacol. 2021, 12, 728758. [Google Scholar] [CrossRef]
  188. Iwata, Y.; Furuichi, K.; Hashimoto, S.; Yokota, K.; Yasuda, H.; Sakai, N.; Kitajima, S.; Toyama, T.; Shinozaki, Y.; Sagara, A.; et al. Pro-inflammatory/Th1 gene expression shift in high glucose stimulated mesangial cells and tubular epithelial cells. Biochem. Biophys. Res. Commun. 2014, 443, 969–974. [Google Scholar] [CrossRef] [Green Version]
  189. Guo, H.; Pan, C.; Chang, B.; Wu, X.; Guo, J.; Zhou, Y.; Liu, H.; Zhu, Z.; Chang, B.; Chen, L. Triptolide Improves Diabetic Nephropathy by Regulating Th Cell Balance and Macrophage Infiltration in Rat Models of Diabetic Nephropathy. Exp. Clin. Endocrinol. Diabetes Off. J. Ger. Soc. Endocrinol. Ger. Diabetes Assoc. 2016, 124, 389–398. [Google Scholar] [CrossRef]
  190. Han, F.; Xue, M.; Chang, Y.; Li, X.; Yang, Y.; Sun, B.; Chen, L. Triptolide Suppresses Glomerular Mesangial Cell Proliferation in Diabetic Nephropathy Is Associated with Inhibition of PDK1/Akt/mTOR Pathway. Int. J. Biol. Sci. 2017, 13, 1266–1275. [Google Scholar] [CrossRef]
  191. Li, X.Y.; Wang, S.S.; Han, Z.; Han, F.; Chang, Y.P.; Yang, Y.; Xue, M.; Sun, B.; Chen, L.M. Triptolide Restores Autophagy to Alleviate Diabetic Renal Fibrosis through the miR-141-3p/PTEN/Akt/mTOR Pathway. Mol. Ther. Nucleic Acids 2017, 9, 48–56. [Google Scholar] [CrossRef]
  192. Xue, M.; Cheng, Y.; Han, F.; Chang, Y.; Yang, Y.; Li, X.; Chen, L.; Lu, Y.; Sun, B.; Chen, L. Triptolide Attenuates Renal Tubular Epithelial-mesenchymal Transition via the MiR-188-5p-mediated PI3K/AKT Pathway in Diabetic Kidney Disease. Int. J. Biol. Sci. 2018, 14, 1545–1557. [Google Scholar] [CrossRef]
  193. Shi, G.; Wu, W.; Wan, Y.G.; Hex, H.W.; Tu, Y.; Han, W.B.; Liu, B.H.; Liu, Y.L.; Wan, Z.Y. Low dose of triptolide ameliorates podocyte epithelial-mesenchymal transition induced by high dose of D-glucose via inhibiting Wnt3α/β-catenin signaling pathway activation. China J. Chin. Mater. Med. 2018, 43, 139–146. [Google Scholar]
  194. Wu, W.; Liu, B.H.; Wan, Y.G.; Sun, W.; Liu, Y.L.; Wang, W.W.; Fang, Q.J.; Tu, Y.; Yee, H.Y.; Yuan, C.C.; et al. Triptolide inhibits NLRP3 inflammasome activation and ameliorates podocyte epithelial-mesenchymal transition induced by high glucose. China J. Chin. Mater. Med. 2019, 44, 5457–5464. [Google Scholar]
  195. Ren, L.; Wan, R.; Chen, Z.; Huo, L.; Zhu, M.; Yang, Y.; Chen, Q.; Zhang, X.; Wang, X. Triptolide Alleviates Podocyte Epithelial-Mesenchymal Transition via Kindlin-2 and EMT-Related TGF-β/Smad Signaling Pathway in Diabetic Kidney Disease. Appl. Biochem. Biotechnol. 2022, 194, 1000–1012. [Google Scholar] [CrossRef] [PubMed]
  196. Xi, C.; Peng, S.; Wu, Z.; Zhou, Q.; Zhou, J. Toxicity of triptolide and the molecular mechanisms involved. Biomed. Pharmacother. 2017, 90, 531–541. [Google Scholar] [CrossRef]
  197. Singh, D.; Chaudhuri, P.K. Structural characteristics, bioavailability and cardioprotective potential of saponins. Integr. Med. Res. 2018, 7, 33–43. [Google Scholar] [CrossRef]
  198. Yang, W.Z.; Hu, Y.; Wu, W.Y.; Ye, M.; Guo, D.A. Saponins in the genus Panax L. (Araliaceae): A systematic review of their chemical diversity. Phytochemistry 2014, 106, 7–24. [Google Scholar] [CrossRef]
  199. Cai, J.; Liu, J.; Fan, P.; Dong, X.; Zhu, K.; Liu, X.; Zhang, N.; Cao, Y. Dioscin prevents DSS-induced colitis in mice with enhancing intestinal barrier function and reducing colon inflammation. Int. Immunopharmacol. 2021, 99, 108015. [Google Scholar] [CrossRef]
  200. Kang, K.B.; Ryu, J.; Cho, Y.; Choi, S.Z.; Son, M.; Sung, S.H. Combined Application of UHPLC-QTOF/MS, HPLC-ELSD and (1) H-NMR Spectroscopy for Quality Assessment of DA-9801, A Standardised Dioscorea Extract. Phytochem. Anal. PCA 2017, 28, 185–194. [Google Scholar] [CrossRef]
  201. Wang, J.; Yang, G.Y.; Sun, H.Y.; Meng, T.; Cheng, C.C.; Zhao, H.P.; Luo, X.L.; Yang, M.M. Dioscin Reduces Vascular Damage in the Retina of db/db Mice by Inhibiting the VEGFA Signaling Pathway. Front. Pharmacol. 2022, 12, 811897. [Google Scholar] [CrossRef]
  202. Xu, L.N.; Yin, L.H.; Jin, Y.; Qi, Y.; Han, X.; Xu, Y.W.; Liu, K.X.; Zhao, Y.Y.; Peng, J.Y. Effect and possible mechanisms of dioscin on ameliorating metabolic glycolipid metabolic disorder in type-2-diabetes. Phytomed. Int. J. Phytother. Phytopharm. 2020, 67, 153139. [Google Scholar] [CrossRef]
  203. Cai, S.; Chen, J.; Li, Y. Dioscin protects against diabetic nephropathy by inhibiting renal inflammation through TLR4/NF-κB pathway in mice. Immunobiology 2020, 225, 151941. [Google Scholar] [CrossRef] [PubMed]
  204. Zhong, Y.; Liu, J.; Sun, D.; Guo, T.; Yao, Y.; Xia, X.; Shi, C.; Peng, X. Dioscin relieves diabetic nephropathy via suppressing oxidative stress and apoptosis, and improving mitochondrial quality and quantity control. Food Funct. 2022, 13, 3660–3673. [Google Scholar] [CrossRef] [PubMed]
  205. Qu, L.; Li, D.; Gao, X.; Li, Y.; Wu, J.; Zou, W. Di’ao Xinxuekang Capsule, a Chinese Medicinal Product, Decreases Serum Lipids Levels in High-Fat Diet-Fed ApoE(-/-) Mice by Downregulating PCSK9. Front. Pharmacol. 2018, 9, 1170. [Google Scholar] [CrossRef]
  206. Li, K.; Tang, Y.; Fawcett, J.P.; Gu, J.; Zhong, D. Characterization of the pharmacokinetics of dioscin in rat. Steroids 2005, 70, 525–530. [Google Scholar] [CrossRef]
  207. Li, X.; Liu, S.; Qu, L.; Chen, Y.; Yuan, C.; Qin, A.; Liang, J.; Huang, Q.; Jiang, M.; Zou, W. Dioscin and diosgenin: Insights into their potential protective effects in cardiac diseases. J. Ethnopharmacol. 2021, 274, 114018. [Google Scholar] [CrossRef] [PubMed]
  208. Xu, T.; Zhang, S.; Zheng, L.; Yin, L.; Xu, L.; Peng, J. A 90-day subchronic toxicological assessment of dioscin, a natural steroid saponin, in Sprague–Dawley rats. Food Chem. Toxicol. 2012, 50, 1279–1287. [Google Scholar] [CrossRef]
  209. Zhou, P.; Xie, W.; He, S.; Sun, Y.; Meng, X.; Sun, G.; Sun, X. Ginsenoside Rb1 as an Anti-Diabetic Agent and Its Underlying Mechanism Analysis. Cells 2019, 8, 204. [Google Scholar] [CrossRef] [PubMed]
  210. Dong, C.; Liu, P.; Wang, H.; Dong, M.; Li, G.; Li, Y. Ginsenoside Rb1 attenuates diabetic retinopathy in streptozotocin-induced diabetic rats1. Acta Cir. Bras. 2019, 34, e201900201. [Google Scholar] [CrossRef] [PubMed]
  211. Tabandeh, M.R.; Hosseini, S.A.; Hosseini, M. Ginsenoside Rb1 exerts antidiabetic action on C2C12 muscle cells by leptin receptor signaling pathway. J. Recept. Signal Transduct. Res. 2017, 37, 370–378. [Google Scholar] [CrossRef]
  212. He, J.Y.; Hong, Q.; Chen, B.X.; Cui, S.Y.; Liu, R.; Cai, G.Y.; Guo, J.; Chen, X.M. Ginsenoside Rb1 alleviates diabetic kidney podocyte injury by inhibiting aldose reductase activity. Acta Pharmacol. Sin. 2022, 43, 342–353. [Google Scholar] [CrossRef] [PubMed]
  213. Kim, H.J.; Oh, T.K.; Kim, Y.H.; Lee, J.; Moon, J.M.; Park, Y.S.; Sung, C.M. Pharmacokinetics of Ginsenoside Rb1, Rg3, Rk1, Rg5, F2, and Compound K from Red Ginseng Extract in Healthy Korean Volunteers. Evid.-Based Complement. Altern. Med. 2022, 2022, 8427519. [Google Scholar] [CrossRef] [PubMed]
  214. Jin, S.; Jeon, J.-H.; Lee, S.; Kang, W.Y.; Seong, S.J.; Yoon, Y.-R.; Choi, M.-K.; Song, I.-S. Detection of 13 Ginsenosides (Rb1, Rb2, Rc, Rd, Re, Rf, Rg1, Rg3, Rh2, F1, Compound K, 20(S)-Protopanaxadiol, and 20(S)-Protopanaxatriol) in Human Plasma and Application of the Analytical Method to Human Pharmacokinetic Studies Following Two Week-Repeated Administration of Red Ginseng Extract. Molecules 2019, 24, 2618. [Google Scholar]
  215. Yang, R.; Pei, T.; Huang, R.; Xiao, Y.; Yan, J.; Zhu, J.; Zheng, C.; Xiao, W.; Huang, C. Platycodon grandiflorum Triggers Antitumor Immunity by Restricting PD-1 Expression of CD8+ T Cells in Local Tumor Microenvironment. Front. Pharmacol. 2022, 13, 774440. [Google Scholar] [CrossRef] [PubMed]
  216. Wu, Y.; Huang, D.; Wang, X.; Pei, C.; Xiao, W.; Wang, F.; Wang, Z. Suppression of NLRP3 inflammasome by Platycodin D via the TLR4/MyD88/NF-κB pathway contributes to attenuation of lipopolysaccharide induced acute lung injury in rats. Int. Immunopharmacol. 2021, 96, 107621. [Google Scholar] [CrossRef] [PubMed]
  217. Kim, H.L.; Park, J.; Jung, Y.; Ahn, K.S.; Um, J.Y. Platycodin D, a novel activator of AMP-activated protein kinase, attenuates obesity in db/db mice via regulation of adipogenesis and thermogenesis. Phytomed. Int. J. Phytother. Phytopharm. 2019, 52, 254–263. [Google Scholar] [CrossRef] [PubMed]
  218. Liu, Y.M.; Cong, S.; Cheng, Z.; Hu, Y.X.; Lei, Y.; Zhu, L.L.; Zhao, X.K.; Mu, M.; Zhang, B.F.; Fan, L.D.; et al. Platycodin D alleviates liver fibrosis and activation of hepatic stellate cells by regulating JNK/c-JUN signal pathway. Eur. J. Pharm. 2020, 876, 172946. [Google Scholar] [CrossRef] [PubMed]
  219. Wu, H.; Fu, L.; Zhao, Y.; Ke, W.; Zhuang, Y. Platycodin D improves renal injury in diabetic nephropathy model rats by regulating oxidative stress mediated PI3K/Akt/mTOR signaling pathway. Chin. J. Pharmacol. Toxicol. 2022, 36, 170–176. [Google Scholar]
  220. Shen, S.; Ji, C.; Wei, K. Cellular Senescence and Regulated Cell Death of Tubular Epithelial Cells in Diabetic Kidney Disease. Front. Endocrinol. 2022, 13, 924299. [Google Scholar] [CrossRef]
  221. Huang, J.; Chen, G.; Wang, J.; Liu, S.; Su, J. Platycodin D regulates high glucose-induced ferroptosis of HK-2 cells through glutathione peroxidase 4 (GPX4). Bioengineered 2022, 13, 6627–6637. [Google Scholar] [CrossRef]
  222. Lee, W.H.; Gam, C.O.; Ku, S.K.; Choi, S.H. Single oral dose toxicity test of platycodin d, a saponin from platycodin radix in mice. Toxicol. Res. 2011, 27, 217–224. [Google Scholar] [CrossRef] [PubMed]
  223. Huang, M.Y.; Jiang, X.M.; Xu, Y.L.; Yuan, L.W.; Chen, Y.C.; Cui, G.; Huang, R.Y.; Liu, B.; Wang, Y.; Chen, X.; et al. Platycodin D triggers the extracellular release of programed death Ligand-1 in lung cancer cells. Food Chem. Toxicol. Int. J. Publ. Br. Ind. Biol. Res. Assoc. 2019, 131, 110537. [Google Scholar] [CrossRef] [PubMed]
  224. Li, Q.; Yang, T.; Zhao, S.; Zheng, Q.; Li, Y.; Zhang, Z.; Sun, X.; Liu, Y.; Zhang, Y.; Xie, J. Distribution, Biotransformation, Pharmacological Effects, Metabolic Mechanism and Safety Evaluation of Platycodin D: A Comprehensive Review. Curr. Drug Metab. 2022, 23, 21–29. [Google Scholar] [CrossRef] [PubMed]
  225. Dieter, S.; Paula, L.; Maria, B.; Victor, W.; Lutz, G. Hydroxycinnamic acid esters of isocitric acid: Accumulation and enzymatic synthesis in Amaranthus cruentus. Phytochemistry 1987, 26, 2919–2922. [Google Scholar] [CrossRef]
  226. Yao, H.; Zhang, W.; Yang, F.; Ai, F.; Du, D.; Li, Y. Discovery of caffeoylisocitric acid as a Keap1-dependent Nrf2 activator and its effects in mesangial cells under high glucose. J. Enzym. Inhib. Med. Chem. 2022, 37, 178–188. [Google Scholar] [CrossRef]
  227. Bastani, S.; Vahedian, V.; Rashidi, M.; Mir, A.; Mirzaei, S.; Alipourfard, I.; Pouremamali, F.; Nejabati, H.; Kadkhoda, J.; Maroufi, N.F.; et al. An evaluation on potential anti-oxidant and anti-inflammatory effects of Crocin. Biomed. Pharmacother. 2022, 153, 113297. [Google Scholar] [CrossRef]
  228. Roshanravan, B.; Samarghandian, S.; Ashrafizadeh, M.; Amirabadizadeh, A.; Saeedi, F.; Farkhondeh, T. Metabolic impact of saffron and crocin: An updated systematic and meta-analysis of randomised clinical trials. Arch. Physiol. Biochem. 2022, 128, 666–678. [Google Scholar] [CrossRef] [PubMed]
  229. Zhang, J.; Zhao, X.; Zhu, H.; Wang, J.; Ma, J.; Gu, M. Crocin protects the renal tubular epithelial cells against high glucose-induced injury and oxidative stress via regulation of the SIRT1/Nrf2 pathway. Iran. J. Basic Med. Sci. 2022, 25, 193–197. [Google Scholar]
  230. Qiu, Y.; Jiang, X.; Liu, D.; Deng, Z.; Hu, W.; Li, Z.; Li, Y. The Hypoglycemic and Renal Protection Properties of Crocin via Oxidative Stress-Regulated NF-κB Signaling in db/db Mice. Front. Pharmacol. 2020, 11, 541. [Google Scholar] [CrossRef]
  231. Zhang, L.; Jing, M.; Liu, Q. Crocin alleviates the inflammation and oxidative stress responses associated with diabetic nephropathy in rats via NLRP3 inflammasomes. Life Sci. 2021, 278, 119542. [Google Scholar] [CrossRef] [PubMed]
  232. Sepahi, S.; Golfakhrabadi, M.; Bonakdaran, S.; Lotfi, H.; Mohajeri, S.A. Effect of crocin on diabetic patients: A placebo-controlled, triple-blinded clinical trial. Clin. Nutr. ESPEN 2022, 50, 255–263. [Google Scholar] [CrossRef]
  233. Hosseinzadeh, H.; Shariaty, V.; Sameni, A.; Vahabzadeh, M. Acute and sub-acute toxicity of crocin, a constituent of Crocus sativus L. (saffron), in mice and rats. Pharmacologyonline 2010, 2, 943–951. [Google Scholar]
  234. Nam, P.C.; Thong, N.M.; Hoa, N.T.; Quang, D.T.; Hoang, L.P.; Mechler, A.; Vo, Q.V. Is natural fraxin an overlooked radical scavenger? RSC Adv. 2021, 11, 14269–14275. [Google Scholar] [CrossRef] [PubMed]
  235. Chen, R.; Zeng, J.; Li, C.; Xiao, H.; Li, S.; Lin, Z.; Huang, K.; Shen, J.; Huang, H. Fraxin Promotes the Activation of Nrf2/ARE Pathway via Increasing the Expression of Connexin43 to Ameliorate Diabetic Renal Fibrosis. Front. Pharmacol. 2022, 13, 853383. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Some pathophysiological mechanisms of DKD.
Figure 1. Some pathophysiological mechanisms of DKD.
Molecules 27 06221 g001
Table 1. Mechanisms of phenolics in the treatment of DKD.
Table 1. Mechanisms of phenolics in the treatment of DKD.
Natural CompoundModelFunctionMechanism/TargetReference
OleuropeinIn vivo: db/db mice DKDAntioxidant
Anti-inflammatory
Reduce apoptosis
MAPK signaling pathway, caspase-3, Bcl-2, Bax[50]
ResveratrolIn vitro: HG stimulated NRK 52E cellsAnti-ER stress
Reduce apoptosis
GRP78, CHOP[57]
In vitro: HG stimulated mouse podocytes
In vivo: db/db mice DKD
Antioxidant
Reduce apoptosis
AMPK signaling pathway[58]
In vitro: HG stimulated mouse podocytes
In vivo: db/db mice DKD
Antioxidant
Reduce apoptosis
SIRT1, PGC-1α, NRF1, TFAM[59]
In vitro: HG stimulated SV40-MES-13
In vivo: STZ-induced mice DKD
Antioxidant
Inhibit proliferation
Anti-fibrotic
Keap1/Nrf2 signaling pathway[60]
In vitro: HG stimulated CRL-2573
In vivo: STZ-induced rats DKD
Anti-fibroticMAPK/TGF-β1 signaling pathway[61]
In vivo: STZ-induced mice DKDReduce lipid accumulationSIRT1 signaling pathway[62]
In vivo: STZ-induced rats DKDReduce lipid accumulation
Improve autophagy
AMPK/mTOR signaling pathway
PPARα, CPT1, SREBP-1c, ACS
[63]
GastrodinIn vitro: HG stimulated MPC-5 cellsAntioxidant
Anti-inflammatory
Reduce apoptosis
AMPK/Nrf2 signaling pathway
NLRP3 inflammasome
[71]
Table 2. Mechanisms of alkaloids in the treatment of DKD.
Table 2. Mechanisms of alkaloids in the treatment of DKD.
Natural CompoundModelFunctionMechanism/TargetReference
TrigonellineIn vitro: HG stimulated HMCs
In vivo: STZ-induced rats DKD
Antioxidant
Reduce apoptosis
Anti-fibrotic
Wnt/β-catenin signaling pathway
FN, collagen IV
[78,79]
In vivo: STZ-induced rats DKDAntioxidant
Anti-inflammatory
Reduce apoptosis
PPARγ/GLUT4-leptin/TNF-α signaling pathway[80]
In vitro: HG stimulated HMCsImprove autophagymiR-5189-5p, HIF1AN
AMPK signaling pathway
[81]
BerberineIn vitro: HG stimulated rat podocytes
In vivo: STZ-induced rats DKD
Alleviate podocyte injuryPI3K/Akt signaling pathway
nephrin, podocin, α3β1
[88]
In vitro: HG stimulated GMCs
In vivo: STZ-induced mice DKD
Reduce glucose uptake
Inhibit proliferation
PI3K/Akt/AS160/GLUT1 signaling pathway[89]
In vitro: hypoxia/HG-stimulated NRK 52E and HK-2 cellsReduce apoptosisPI3K/Akt signaling pathway
HIF-1α
[90]
In vitro: PA stimulated HK-2 cellsReduce apoptosis
Reduce lipid accumulation
CPT1A, PPARα, PGC1α[91]
In vitro: HG stimulated HK-2 cells
In vivo: db/db mice DKD
Anti-fibrotic
Reduce lipid accumulation
Improve mitochondrial function
α-SMA, collagen I, collagen IV, FN, TGF-β1, CPT1, ACOX1, PPARα, AMPK, PGC-1α[92]
In vitro: HG stimulated HK-2 cells
In vivo: STZ-induced rats DKD
Antioxidant
Reduce apoptosis
Improve mitochondrial function
C/EBPβ/Gas5/miR-18a-5p signaling pathway
C/EBPβ/PGC-1α signaling pathway
[93]
In vitro: PA stimulated mouse podocytes
In vivo: db/db mice DKD
Antioxidant
Reduce lipid accumulation
Improve mitochondrial function
AMPK/PGC-1α signaling pathway
CPT1, ACC, CD36
[94]
In vitro: PA stimulated mouse podocytes
In vivo: db/db mice DKD
Antioxidant
Reduce apoptosis
Improve mitochondrial function
Drp1, MFF, Fis1, Mid49, Mid51[95]
In vitro: HG stimulated mouse podocytes
In vivo: STZ-induced rats DKD
Anti-inflammatory
Reduce apoptosis
TLR4/NF-κB signaling pathway[97]
In vitro: HG stimulated mRTEC
In vivo: KKAy mice DKD
Anti-fibroticNotch/snail signaling pathway
α-SMA, E-cadherin
[96]
In vitro: HG stimulated mouse podocytesReduce apoptosis
Improve autophagy
mTOR/P70S6K/4EBP1 signaling pathway[98]
SinomenineIn vitro: HG stimulated HRGEsAntioxidant
Decrease cell permeability
Nrf2 signaling pathway
ROCK signaling pathway
ZO-1, occludin
[105]
In vitro: HG stimulated HRGEs
In vivo: STZ-induced rats DKD
Anti-inflammatory
Decrease cell permeability
C/EBPα/claudin-5 signaling pathway[106]
In vitro: H2O2 stimulated HK-2 cells
In vivo: STZ-induced rats DKD
Antioxidant
Reduce apoptosis
Anti-inflammatory
Anti-fibrotic
JAK/STAT signaling pathway[107]
Table 3. Mechanisms of flavonoids in the treatment of DKD.
Table 3. Mechanisms of flavonoids in the treatment of DKD.
Natural CompoundModelFunctionMechanism/TargetReference
NaringeninIn vitro: HG stimulated NRK 52E cells
In vivo: STZ-induced rats DKD
Antioxidant
Anti-ER stress
Reduce apoptosis
p-PERK, eIF2α, XBP1s, ATF4, CHOP[117]
In vitro: HG stimulated NRK 52E cells
In vivo: STZ-induced mice DKD
Reduce renal tissue injuryPPARs, CYP4A, 20-HETE[118]
In vitro: HG stimulated 293T cells
In vivo: STZ-induced rats DKD
Anti-fibroticMicroRNA let-7a
TGF-β1/Smad signaling pathway
[119]
QuercetinIn vivo: STZ-induced rats DKDAntioxidant
Anti-inflammatory
AGEs, TNF-α, IL-6[125]
In vitro: HG stimulated mouse podocytes
In vivo: db/db mice DKD
Reduce apoptosisEGFR signaling pathway[126]
In vitro: HG stimulated HMCs
In vivo: DKD patients
Inhibit proliferation
Antioxidant
Anti-inflammatory
miR-485-5p, YAP1[127]
In vitro: HG stimulated SV40-MES-13
In vivo: db/db mice DKD
Inhibit proliferation
Antioxidant
Anti-inflammatory
Hippo signaling pathway[128]
In vivo: db/db mice DKDReduce lipid accumulationSCAP-SREBP2-LDLr signaling pathway[130]
In vivo: STZ-induced rats DKDReduce lipid accumulation
Anti-inflammatory
PPARα, CPT1, OCTN2, ACC2
NLRP3 inflammasome/caspase-1/IL-1β/IL-18 signaling pathway
[131]
IcariinIn vitro: HG stimulated MPC-5 cells
In vivo: STZ-induced rats DKD
Anti-inflammatory
Improve autophagy
Improve mitochondrial function
Sesn2, NLRP3
Nrf2 signaling pathway
[134]
In vivo: STZ-induced mice DKDAnti-inflammatory
Anti-fibrotic
TLR4/NF-κB signaling pathway[135]
In vitro: HG stimulated SV40-MES-13Anti-fibroticTGF-β1/Smad signaling pathway
ERK, GPER
[136]
In vitro: HG stimulated HK-2 and NRK 49F cells
In vivo: STZ-induced rats DKD
Improve autophagy
Anti-fibrotic
miR-192-5p/GLP-1R signaling pathway
p-mTOR, collagen I, α-SMA, FN
[137]
In vitro: HG stimulated NRK 52E cells
In vivo: STZ-induced rats DKD
Reduce apoptosis
Anti-fibrotic
miR-122-5p, FOXP2, E-cadherin, α-SMA [138]
CardamoninIn vitro: MGO stimulated NRK 52E cells
In vivo: STZ-induced rats DKD
Antioxidant
Anti-inflammatory
Reduce apoptosis
Anti-fibrotic
PI3K/AKT signaling pathway
JAK/STAT signaling pathway
caspase-3, Bcl-2, Bax, NF-κB, FN, α-SMA, TGF-β1, Vimentin
[146]
MorinIn vitro: HG stimulated NRK 52E cells
In vivo: STZ-induced rats DKD
Antioxidant
Reduce apoptosis
PHLPP1/FoxO1-Mdm2 signaling pathway
Nrf2
[150]
In vitro: HG stimulated primary rat GMCsAntioxidant
Anti-fibrotic
MAPK signaling pathway
JNK signaling pathway
NOX4
[151]
HesperetinIn vivo: STZ-induced rats DKDAntioxidant
Anti-inflammatory
Anti-fibrotic
TBARS, GSH-Px, CAT, TNF-α, IL-6, TGF-β, GSK-3β[160]
In vivo: STZ-induced rats DKDAnti-inflammatory
Anti-fibrotic
Nrf2 signaling pathway
Glo-1, collagen ΙV, FN
[161]
FisetinIn vitro: HG stimulated mouse podocytes
In vivo: STZ-induced mice DKD
Anti-inflammatory
Anti-fibrotic
Improve autophagy
CDKN1B/P70S6K signaling pathway
NLRP3 inflammasome
[171]
In vitro: PA stimulated HK-2 cells
In vivo: HFD-induced mice kidney injury
Improve insulin sensitivity
Anti-inflammatory
Insulin receptor signaling pathway
NF-κB signaling pathway
RIP3/NLRP3 signaling pathway
[172]
Table 4. Mechanisms of terpenoids in the treatment of DKD.
Table 4. Mechanisms of terpenoids in the treatment of DKD.
Natural CompoundModelFunctionMechanism/TargetReference
SclareolIn vitro: HG stimulated SV40-MES-13
In vivo: STZ-induced mice DKD
Antioxidant
Anti-inflammatory
Anti-fibrotic
MAPK/NF-κB signaling pathway [178]
PonicidinIn vivo: STZ-induced rats DKDAntioxidant
Anti-inflammatory
Improve lipid metabolism
Anti-fibrotic
TNF-α, IL-1β, IL-6, NF-κB[182]
TriptolideIn vivo: STZ-induced rats DKDAnti-inflammatory
Regulate Th1/Th2 cells balance
Interferon-γ, IL-12, TNF-α, IL-4, IL-10[189]
In vitro: HG stimulated HMCs and HK-2 cells
In vivo: STZ-induced rats DKD
Improve autophagy
Inhibit proliferation
Anti-fibrotic
miR-141-3p, miR-188-5p, PTEN, PDK1, Akt, mTOR[190,191,192]
In vitro: HG stimulated mouse podocytesAnti-inflammatory
Anti-fibrotic
Wnt3α/β-catenin signaling pathway
NLRP3, ASC, caspase-1, IL-1β, IL-18
[193,194]
In vitro: HG stimulated mouse podocytes
In vivo: db/db mice DKD
Anti-fibroticTGF-β/Smad signaling pathway
kindlin-2
[195]
Table 5. Mechanisms of saponins in the treatment of DKD.
Table 5. Mechanisms of saponins in the treatment of DKD.
Natural CompoundModelFunctionMechanism/TargetReference
DioscinIn vivo: STZ-induced mice DKDAnti-inflammatoryTLR4/NF-κB signaling pathway[203]
In vivo: STZ-induced rats DKDAntioxidant
Anti-inflammatory
Anti-ER stress
Reduce apoptosis
Improve autophagy
Improve mitochondrial function
IL-1β, IL-6, TNF-α, NF-κB, p-PERK, IRE1, p-IRE1, ATF4, CHOP, Caspase-12, PINK1, Drp1, p-Drp1, MFN2
AMPK/mTOR signaling pathway
[204]
Ginsenoside Rb1In vivo: STZ-induced rats DKDAntioxidant
Anti-fibrotic
Wnt/β-catenin signaling pathway
miR-3550
[78]
In vitro: HG stimulated mouse podocytes
In vivo: STZ-induced mice DKD
Antioxidant
Protect mitochondria
Reduce apoptosis
AR, NOX4, caspase-9[212]
Platycodin DIn vivo: STZ-induced rats DKDAntioxidant
Anti-inflammatory
Anti-fibrotic
PI3K/Akt signaling pathway[219]
In vitro: HG stimulated HMCs and HK-2 cellsAntioxidant
Reduce ferroptosis
GPX4, FTH1, SLC7A11, ACSL4, TFR1[221]
Table 6. Mechanisms of other compounds in the treatment of DKD.
Table 6. Mechanisms of other compounds in the treatment of DKD.
Natural CompoundModelFunctionMechanism/TargetReference
Caffeoylisocitric acidIn vitro: HG stimulated HMCsAntioxidant
Anti-inflammatory
Anti-fibrotic
Nrf2 signaling pathway
MAPK signaling pathway
[226]
CrocinIn vitro: HG stimulated HK-2 cellsAntioxidant
Reduce apoptosis
SIRT1/Nrf2 signaling pathway[229]
In vivo: db/db mice DKDAntioxidant
Anti-inflammatory
Nrf2, SOD-1, HO-1, CAT
NF-κB signaling pathway
[230]
In vivo: STZ-induced rats DKDAntioxidant
Anti-inflammatory
Anti-fibrotic
NLRP3 inflammasome
TNF-α, IL-1β, IL-18, TGF-β, collagen I, collagen IV
[231]
FraxinIn vitro: HG stimulated primary GMCs
In vivo: db/db mice DKD
Antioxidant
Anti-inflammatory
Anti-fibrotic
Nrf2 signaling pathway
Cx43, Akt, FN, ICAM-1
[235]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhou, K.; Zi, X.; Song, J.; Zhao, Q.; Liu, J.; Bao, H.; Li, L. Molecular Mechanistic Pathways Targeted by Natural Compounds in the Prevention and Treatment of Diabetic Kidney Disease. Molecules 2022, 27, 6221. https://doi.org/10.3390/molecules27196221

AMA Style

Zhou K, Zi X, Song J, Zhao Q, Liu J, Bao H, Li L. Molecular Mechanistic Pathways Targeted by Natural Compounds in the Prevention and Treatment of Diabetic Kidney Disease. Molecules. 2022; 27(19):6221. https://doi.org/10.3390/molecules27196221

Chicago/Turabian Style

Zhou, Kaixuan, Xue Zi, Jiayu Song, Qiulu Zhao, Jia Liu, Huiwei Bao, and Lijing Li. 2022. "Molecular Mechanistic Pathways Targeted by Natural Compounds in the Prevention and Treatment of Diabetic Kidney Disease" Molecules 27, no. 19: 6221. https://doi.org/10.3390/molecules27196221

Article Metrics

Back to TopTop