Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Skip to main content
1181 Views
532 Downloads
21 Pages
Download PDF
Table of Contents
Table of Contents
Rowaiye, Emenyonu, Nwonu, Okpalefe, Ogugua, Akinseye, Ibeanu, Kolesnikov, and Lewallen: Bioactive molecules from soybeans (Glycine max) with anti-type 2 diabetes activity: a systematic review
Review article

Abstract

The hallmarks of type 2 diabetes (T2D) include peripheral insulin resistance and insulin insufficiency. Given the significance of T2D as a major public health concern, the goal of this review article is to highlight the role of bioactive properties of soybeans in the prevention, control, and management of the disease. This study examined how bioactive molecules from soybeans modulate key enzymes that affect T2D. A systematic search of electronic databases such as PubMed and Google Scholar was conducted to find relevant original studies or non-original peer-reviewed papers published on the subject. Keyword combinations such as “Soy OR soybeans”, “Soybeans AND anti-inflammatory”, “Soybeans AND anti-oxidative”, and “Soybeans AND antidiabetic” were used in the search. A search was also conducted on all the key enzymes and bioactive molecules mentioned in the review. A total of 194 articles published in English were selected. Both in vivo and in vitro experiments, and human clinical trials have proven a marked efficacy of these bioactive molecules in suppressing the key enzyme biomarkers that modulate T2D pathogenesis. Based on empirical research, the consumption of soybean products and ingredients is associated with a lower incidence of T2D. These findings will contribute to the present understanding of the therapeutic potential of soy-derived compounds. However, this study does not capture the individual variances to these biomolecules; hence, there is the need for more future pharmacokinetic studies to better understand potential interactions, safety, and more efficacy concerns.

1. Introduction

Diabetes is a metabolic disease characterized by hyperglycemia, abnormal lipid, and protein metabolism, and in the long term, it affects the retina, kidney, liver, and the nervous system [1]. There are type 1 and type 2 diabetes, gestational diabetes, and some other uncommon forms. Also called non-insulin-dependent diabetes mellitus (NIDDM), T2D is characterized by insulin deficiency and peripheral insulin resistance resulting from the failure of pancreatic β-cell adaptation to increase in function and mass orchestrated by nutrient metabolism, hormones, and cytokines [2]. Reduction in insulin resistance and improvement of β-cell function and its mass are, therefore, important in the reduction of progression of T2D [2].

Advances in modern drug discovery have been able to uncover several antidiabetic strategies by evaluating the efficacy and safety of many bioactive scaffolds as modulators of experimentally confirmed molecular targets. These targets include enzymes involved in the sugar metabolism pathways such as glycolytic, hexosamine, advanced glycation end products, protein kinase C, polyol, and insulin signaling pathways [3] (Figure 1).

The mechanisms through which these compounds interact with these pathways are complex and involve multiple molecular targets [3]. In the glycolytic pathway, soy isoflavones, particularly genistein and daidzein, modulate enzymes such as adenosine monophosphate-activated protein kinase (AMPK), potentially enhancing glucose uptake and use by cells [4, 5]. In the hexosamine biosynthesis pathway (HBP), which is implicated in insulin resistance, soy bioactives, particularly isoflavones, may be modulated contributing to improved insulin sensitivity [6]. Specifically, it has been shown that genistein can inhibit the stimulation of the glutamine fructose-6-phosphate amidotransferase (GFAT) promoter. GFAT is a key enzyme in the HBP that catalyzes the conversion of fructose-6-phosphate to glucosamine-6-phosphate [7]. Soy compounds may have anti-glycation properties, reducing the formation of advanced glycation end products (AGEs) associated with complications of diabetes. Genistein can inhibit AGE-RAGE (receptor advanced glycation end products) binding, potentially reducing the downstream effects of RAGE activation [7]. Soy bioactives may influence the insulin signaling pathway in several ways. First, soy isoflavones, specifically genistein, have been reported to activate the insulin receptor and insulin receptor substrate (IRS) proteins. This activation enhances insulin signaling, leading to improved glucose uptake and metabolism [8]. Second, certain soy bioactives may inhibit tyrosine kinase, an enzyme that negatively regulates insulin signaling. The inhibition of tyrosine kinase can enhance insulin signaling [9]. Third, soy compounds may modulate the PI3K/Akt pathway which affects insulin signaling and can positively influence downstream signaling events involved in glucose homeostasis [10].

Figure 1

Molecular pathogenesis of T2D showing different signaling pathways. Created with BioRender.com.

media/image3.png

Most T2D chemotherapeutic agents are replete with side effects, and many of these are contraindicated during pregnancy. This has resulted in the search for natural products that could treat the disease while keeping the safety of the patient [11]. Therefore, attention is being focused on edible plants that have therapeutic molecules to offer effective alternatives to the treatment of T2D with little or no side effects [11, 12]. Among the plants believed to have the potential to treat diabetes is the soybean. Soybean has been used globally for medicinal purposes, and several studies have proven its ability to lower hyperglycemia [13]. This plant has been an important protein source, and among Asians who consume the grain routinely, the incidence of T2D is said to be reduced [14]. The antidiabetic effects of soybeans and their phytochemicals are mediated through various mechanisms which include improving insulin secretion and sensitivity, regulating glucose metabolism, exerting antioxidant and anti-inflammatory effects, and modulating gut microbiota [15].

Soybean consists of isoflavones, protein, carbohydrate, and lipid fractions, and it has been reported to play a significant role in controlling metabolic disorders such as obesity and hyperglycemia [16, 17]. There are numerous methods of preparing soybeans, and the meals made from them can either be fermented or unfermented. Tofu, soymilk, edamame, soy nuts, and sprouts are examples of non-fermented foods. Fermented soy products include miso, tempeh, natto, and soy sauce [18].

Fermentation has been shown to improve the anti-oxidative and antidiabetic potentials of soybeans. Fermentation leads to the activation of certain enzymes which include proteases and peptidases, which results in the production of bioactive peptides [19]; β-glucosidase and α-galactosidase, which break down complex carbohydrates and oligosaccharides, making nutrients more bioavailable and potentially enhancing antidiabetic effects [20]; and lipase which influences the breakdown of lipids and the production of bioactive lipid compounds with potential health benefits [21]. Other enzymes that could be triggered by fermentation are polyphenol oxidase which is involved in the oxidation of polyphenols, affecting their levels of polyphenols and their antioxidant capacities in soybeans [22], and phytase which improves the mineral bioavailability and potentially enhances the nutritional quality of soybeans [23]. Fermentation can also ease the deglycosylation of certain compounds, making them more bioactive. For example, isoflavone glycosides in soybeans may be converted into their aglycone forms during fermentation, potentially enhancing their health-promoting effects [24].

When fermented, the soybean contents such as fermented soybean products, isoflavonoids, and soy peptides improve glucose metabolism and insulin secretion, attenuate insulin resistance, and increase β-cell mass [25]. Soybean, therefore, may prevent T2D as well as delay its progression [2]. This review article aims at highlighting the bioactive properties of soybeans with respect to the prevention, control, and management of diabetes given its importance as a disease of major public health concern.

2. Antidiabetic potential

The extracts of soybeans have been linked with the inhibition of key enzymes associated with T2D, hypertension, and other related bioactivities (see Table 1) such as anti-inflammation and antioxidation [26].

2.1. Anti-oxidative activity

Oxidation is a chemical reaction that results in the production of reactive oxygen species (ROS) and reactive nitrogen species (RNS). In living things, free radicals are byproducts of metabolic processes in aerobic cells and exist in homeostasis with antioxidants. Accumulation of these highly unstable reactive species can result in oxidative stress, a phenomenon that occurs when there is an imbalance between the free radicals produced and the antioxidants that inhibit oxidation [27]. Oxidative stress has been remarkably connected to diabetes [28]. In diabetic patients, oxidative stress arises from an increase in the production of free radicals or a suppression of the activity of endogenous antioxidants, or both, and it induces dysfunction of β-cells as well as insulin resistance [29]. Hyperglycemia has been implicated in oxidative stress through the modulation of several molecular pathways involved in the metabolism of glucose and lipids such as glycolysis, hexosamine (HBP), advanced glycation end products (AGE), protein kinase C (PKC), polyol, and insulin signaling pathways [30] (Figure 1).

Table 1

The soy constituents, class of molecules, and their antidiabetic properties

Soy-constituent Class of molecule Anti-T2D effect
Genistein Polyphenol Anti-oxidative [31]
Daidzein Polyphenol Anti-oxidative [31]
Glycitein Polyphenol Anti-oxidative [31]
Equol Polyphenol Anti-oxidative [32]
Bowman-Birk inhibitor (BBI) Protein Anti-oxidative [33]
Lunasin Peptide Anti-oxidative [34]
GNPDIEHPE Peptide Anti-oxidative [35]
SVIKPPTDE Peptide Anti-oxidative [35]
VIKPPTDE Peptide Anti-oxidative [35]
GNPDIEHPET Peptide Anti-oxidative [35]
FEEPQQPQ Peptide Anti-oxidative [35]
D-10-R Peptide Anti-oxidative [36]
S-10-S Peptide Anti-oxidative [36]
V-11-R Peptide Anti-oxidative [36]
I-11-N Peptide Anti-oxidative [36]
ASP-3A Polysaccharide Anti-oxidative [37]
Raffinose Polysaccharide Anti-oxidative [38]
Phosphatidylcholine Polyunsaturated fatty acid Anti-oxidative [39]
Soybean germ phytosterols Lipids Anti-oxidative [40]
Cyanidin-3-O-glucoside Polyphenol Anti-inflammatory [41]
Daidzein Polyphenol Anti-inflammatory [42]
FLV Peptide Anti-inflammatory [43]
Val-Pro-Tyr Peptide Anti-inflammatory [44]
Lunasin Peptide Anti-inflammatory [45]
SSPS1 Polysaccharide Anti-inflammatory [46]
MET-SSPS-NPs Polysaccharide Anti-inflammatory [47]
O-3 PUFA Polyunsaturated fatty acid Anti-inflammatory [48]
Glucosylceramide Sphingolipids Anti-inflammatory [49]
Myricetin Polyphenol α-Amylase inhibition [50]
Glycitein Polyphenol α-Amylase inhibition [51]
Genistein Polyphenol α-Amylase inhibition [51]
Daidzein Polyphenol α-Amylase inhibition [51]
Peptide fractions Peptides α-Amylase inhibition [52]
SSPS Polysaccharides α-Amylase inhibition [53]
Insoluble dietary fibers Polysaccharides α-Amylase inhibition [54]
Soyasaponins Saponins α-Amylase inhibition [55]
Coumestrol Polyphenol α-Glucosidase inhibition [56]
Luteolin Polyphenol α-Glucosidase inhibition [57]
Glyceollin Polyphenol α-Glucosidase inhibition [57]
Daidzein Polyphenol α-Glucosidase inhibition [57]
Genistein Polyphenol α-Glucosidase inhibition [57]
1-Deoxynojirimycin Alkaloid α-Glucosidase inhibition [58]
(ESPro1) Protein α-Glucosidase inhibition [59]
Gly-Ser-Arg Peptide α-Glucosidase inhibition [60]
Glu-Ala-Lys Peptide α-Glucosidase inhibition [60]
LLPLPVLK Peptide α-Glucosidase inhibition [61]
SWLRL Peptide α-Glucosidase inhibition [61]
WLRL Peptide α-Glucosidase inhibition [61]
Soyasaponin Ba Saponins α-Glucosidase inhibition [62]
Soyasaponin Bb Saponins α-Glucosidase inhibition [62]
SSPS Polysaccharides α-Glucosidase inhibition [63]
Morin Polyphenol Aldose reductase inhibition [64]
Quercitrin Polyphenol Aldose reductase inhibition [64]
Naringenin Polyphenol Aldose reductase inhibition [64]
Rutin Polyphenol Aldose reductase inhibition [64]
Genistein Polyphenol Aldose reductase inhibition [65]
Daidzein Polyphenol Aldose reductase inhibition [66]
8-Hydroxydaid-zein Polyphenol Aldose reductase inhibition [67]
Glycitein Polyphenol Aldose reductase inhibition [68]
Stigmasterol Lipid DPP-4 inhibition [69]
VVNPDNNEN Peptide DPP-4 inhibition [69]
EEPQQPQQ Peptide DPP-4 inhibition [70]
QEPQESQQ Peptide DPP-4 inhibition [70]
SQRPQDRHQ Peptide DPP-4 inhibition [70]
PETMQQQQQQ Peptide DPP-4 inhibition [70]
SSPDIYNPQAGSVT Peptide DPP-4 inhibition [70]
APFL Peptide DPP-4 inhibition [70]
TPVGM Peptide DPP-4 inhibition [71]
EPAAV Peptide DPP-4 inhibition [71]
IAVPTGVA Peptide DPP-4 inhibition [72]
Daidzein Polyphenol Pancreatic lipase inhibition [73]
Glycitein Polyphenol Pancreatic lipase inhibition [73]
Genistein Polyphenol Pancreatic lipase inhibition [73]
PIL Protein Pancreatic lipase inhibition [74]
Lipoxygenase-1 Protein Pancreatic lipase inhibition [75]
56.0 kDa α-Amylase Protein Pancreatic lipase inhibition [76]
FPFPRPPHQ, Peptide Pancreatic lipase inhibition [77]
QCCAFEM Peptide Pancreatic lipase inhibition [77]
FAPEFLK Peptide Pancreatic lipase inhibition [77]
SFFFPFELPRE Peptide Pancreatic lipase inhibition [77]
FMYL Peptide Pancreatic lipase inhibition [77]
PFLL Peptide Pancreatic lipase inhibition [77]
FPLL Peptide Pancreatic lipase inhibition [77]
LPHF Peptide Pancreatic lipase inhibition [77]
Galactomannan Polysaccharide Pancreatic lipase inhibition [78]
Chicoric acid Polyphenol ACE inhibition [79]
Beta-sitosterol Lipid ACE inhibition [80]
Nicotianamine Metal chelator ACE inhibition [81]
Val-Leu-Ile-Val-Pro Peptide ACE inhibition [82]
LPYP Peptide ACE inhibition [82]
IAVPTGVA Peptide ACE inhibition [82]
Leu-Val-Gln-Gly-Ser Peptide ACE inhibition [83]
Genistein Polyphenol PEPCK inhibition [84]
Daidzein Polyphenol PEPCK inhibition [85]
Genistein Polyphenol PPARγ activation [85]
Daidzein Polyphenol PPARγ activation [85]
Formononetin Polyphenol PPARγ activation [85]
Biochanin A Polyphenol PPARγ activation [85]

Elevated levels of oxidized DNA, lipids, and proteins suggest that enhanced oxidative stress plays a major role in the pathogenesis of diabetes issues [86]. Several pathological mechanisms could trigger oxidative stress in diabetes mellitus. Excessive generation of superoxide anions happens when the mitochondria’s electron transport chain is disrupted by excessive glucose levels [87]. As advanced glycation end products (AGEs) form and accumulate, ROS are produced [88, 89]. As AGEs bind with their receptors (RAGE), ROS is further produced [90]. Moreover, the deactivation of antioxidant enzymes by glycation, as seen by the glycation of superoxide dismutase (SOD), compromises the antioxidant defense [91, 92]. An important additional avenue through which high glucose levels might cause oxidative damage is the polyol pathway [93]. Previous studies using aldose reductase (AR)–deficient mice models have demonstrated that the polyol pathway plays a critical role in diabetes-induced oxidative stress [94, 95].

The polyol pathway may cause oxidative stress by one of three mechanisms. Firstly, in a state of hyperglycemia, about 30% of the glucose is directed into the AR-dependent polyol pathway, which depletes NADPH and reduces glutathione (GSH) levels [96]. Secondly, when sorbitol dehydrogenase (SDH) converts sorbitol into fructose, oxidative stress is triggered. At this point, SDH converts the co-factor NAD + into NADH. Superoxide anions are produced by NADH oxidase using NADH as a substrate [97]. Thirdly, glucose is converted to fructose through the polyol pathway. Fructose can be further broken down to yield fructose-3-phosphate and 3-deoxyglucosone, which are two stronger non-enzymatic glycation agents than glucose [98, 99]. This implies that more AGEs would be produced as more glucose goes via the polyol route, which would lead to the generation of ROS [100].

Small molecule natural products

Soybean has polyphenolic antioxidants which exist in various forms determined by the color of the seed coat. Like other plant extracts, the quality and antioxidant potential of a soybean extract are dependent on both its source and method of extraction. Isoflavones are the polyphenols mostly found in soybean and are mainly made up of genistein, daidzein, and glycitein, and these are made more bioavailable by incorporating the hydroxyl group through fermentation by microorganisms [31, 101, 102]. In three different antioxidant assays (DPPH, FRAP, and ABTS), daidzein, glycitein, and genistein have shown the highest antioxidant capacity in human intestinal Caco-2 cells, out of the 12 major soybean isoflavones [9]. Equol, a metabolite formed from daidzein, has the highest estrogenic and antioxidant activities of all isoflavone-derived compounds. Equol has been shown to have even greater antioxidant power than 17-estradiol, alpha-tocopherol, and ascorbic acid [32].

The antioxidant and anti-hyperglycemic activities of black soybean (Glycine max (L) Merr) seed coat treated with either ethyl acetate or hydro-ethanol both in vitro and in vivo using diabetes-induced Wistar rats have been investigated in a recent study. The results showed that in vitro, both ethyl acetate and hydroethanolic extracts of black soybean scavenged both 2,2-diphenyl-1-picrylhydrazyl (DPPH) free radical and hydrogen peroxide in a dose-dependent manner, and significantly reduced blood glucose level. In addition, the in vivo study reveals that the levels of GSH and activity of SOD in the pancreas and brain were elevated [13]. Interestingly, the wild soybean, Glycine soja, which has been a source of traditional food for the Chinese has isoflavones at levels significantly higher than its cultivated counterpart G. max [103]. In a recent study, the phenolic and flavonoid composition as well as the antioxidant activity of G. max and its wild relative G. soja was examined. The results of DPPH and 2,2′-azino-bis (3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) assays showed that the scavenging activity of soybean extracts obtained using 70% methanol was notably more elevated in wild soybean than the cultivated one [104].

Compared to non-fermented soyfoods, fermented soyfoods—especially those that are long-fermented—have a greater potential for antioxidants. The foods with the greatest H-ORAC values are tempeh, natto, and soy sauce, and then Soybean-koji miso. Given the high antioxidant content of fermented soy products, this may be related to the powerful radical scavenging properties of aglycone isoflavones [105]. Certain bacteria are involved in the fermentation process of soybeans, which helps change the substances found in soybeans, such as isoflavones. These microbes are essential to the fermentation process because they break down complicated molecules and produce beneficial metabolites. They include Bacillus spp. [106], Streptococcus spp. [107], Aspergillus spp. [108], Lactobacillus spp. [109], Enterococcus spp. [110], and Rhizopus spp. [111].

The microbes involved in soybeans fermentation produce the β-glucosidase enzyme, which transforms the glycosides and polyphenols in these products into their aglycone form. These changes result in increased therapeutic benefits and anti-cancer, anti-obesity, neuroprotective, and antidiabetic activities by increasing the bioavailability of these phenolic compounds in the human gut [31]. In a 12-week, randomized, controlled, crossover clinical trial (NCT03429920) involving 27 participants, it was discovered that a non-probiotic fermented soy product reduced total and LDL cholesterol levels [112]. Also, in another 12-week, randomized, multi-center, placebo-controlled, double-blind clinical trial, it was proven that Lactobacillus plantarum C29–fermented soybean (DW2009) has an ameliorative effect on mild cognitive impairment (MCI). After DW2009 was administered, there was an increase in serum brain-derived neurotrophic factor (BDNF) levels. This finding implies the involvement of the gut–brain axis in improving cognitive deficits in MCI and may offer some preliminary insight into the underlying effects of cognitive improvement. The clinical trial’s findings show that DW2009 can be given to MCI patients in a way that is safe and effective at improving their cognitive function [113].

Proteins/peptides

Soybean has bioactive peptides with antioxidant and antidiabetic activities. The Bowman-Birk inhibitor (BBI) is an 8 kDa natural soybean-derived protein that has been shown to have antioxidant properties. BBI inhibits superoxide anion radical generation and reduces oxidative stress, which can be beneficial for overall health [33]. Nuclear factor erythroid 2-related factor 2 (Nrf2) is typically sequestered in the cytoplasm by Kelch-like ECH-associated protein 1 (Keap1). Under conditions of oxidative stress or exposure to specific compounds, Nrf2 is released from Keap1, translocates to the nucleus, and induces the expression of genes coding antioxidant enzymes such as SOD or catalase, heme oxygenase-1 (HO-1), NAD(P)H quinone oxidoreductase 1 (NQO1), and glutathione S-transferase (GST). These enzymes play a crucial role in cellular defense against oxidative stress [114]. The peptides of BBI exhibit strong binding with Keap1 at the binding site of Nrf2. This makes Nrf2 more available to combat oxidative stress [115].

Also, Lunasin, a 5 KDa soy peptide, inhibits the oxidation of linoleic acid and acts as an ABTS radical scavenger [34, 116]. According to published reports, lunasin possesses several antioxidant qualities, including the ability to block signaling pathways triggered by oxidative stress, activate the Nrf2 pathway, and alter the activity of antioxidant enzymes including SOD and catalase. By neutralizing ROS, these enzymes are essential for the body’s defense against oxidative stress [117]. Lunasin also has metal-chelating properties. By chelating metals, lunasin may help prevent oxidative damage. Metal chelation is relevant to oxidative stress, as transition metals can catalyze the production of ROS [115].

To test the anti-oxidative potential of soybean peptides, Zhang et al. [35] hydrolyzed soybean using alcalase, and the resultant protein hydrolysate was digested with pepsin and pancreatin under simulated gastrointestinal conditions. A significant free-radical scavenging activity was seen in DPPH and ABTS antioxidant assays as well as in Caco-2cells with H2O2-induced oxidative stress, and was enhanced by digestion. Moreover, the peptides GNPDIEHPE, SVIKPPTDE, VIKPPTDE, GNPDIEHPET, and FEEPQQPQ remained intact after LC-MS absorption and could show antioxidant properties after being absorbed by the intestinal epithelia [35]. In another study, fermented soy products obtained using Lactobacillus delbrueckii subsp. bulgaricus and Streptococcus thermophilus and sequentially digested with pepsin and pancreatin were investigated for their antioxidant activity. The peptides produced D-10-R, S-10-S, V-11-R, and I-11-N had weak antioxidant activity in vitro (ABTS test), but in Caco-2 cells, they strongly inhibited the binding of Keap1 protein to nuclear factor erythroid 2-related factor 2 (Nrf2). Thus, the levels of Nrf2 were elevated suggesting that they offered protection against oxidative stress induced by tert-butyl hydroperoxide [36].

Polysaccharides

Dietary polysaccharides also have an antioxidant role in combating diabetes. Their antioxidant potential has been attributed to their ability to enhance the activity of antioxidant enzymes, as well as the possible conjugation of small moieties of peptides, flavones, polyphenols, and proteins on the polysaccharides [118]. Soy hulls are primarily made up of cellulose, hemicellulose, and pectin, and these polysaccharides have been suggested to play a role in regulating blood glucose and scavenging free radicals based on their varying physicochemical properties [119]. A study by Song et al. [37] using superoxide radical, hydroxy radical, ABTS, and DPPH scavenging tests revealed that ASP-3A, a pectic polysaccharide, displayed dose-dependent antioxidant activities, activated the NRF2/ARE pathway, and thus improved the antioxidant activity of hepatocytes in mice with CCl4-induced oxidative stress. In another study, polysaccharides from small black soybean were found to change the abundance of Oscillospira, Ruminococcus, Mucispirillum, and Dorea bacteria in the gut of STZ-induced diabetic rats, and control the metabolism of linoleic acid in addition to improving the symptoms of both hyperglycemia and lipidemia. Also, the soybean bran conferred stronger cytoprotection to hepatocytes [120]. Soybeans have raffinose, and this polysaccharide has been shown to improve the oxidative status of the in Nile Tilapia (Oreochromis niloticus). Feeding raffinose increased SOD activity at all levels, while glutathione peroxidase (GPX) and SOD improved by 1 to 1.5 g/kg above the control. Raffinose, on the other hand, reduced the level of malondialdehyde (MDA), and the lowest concentrations were found at 1 or 1.5 g/kg levels [38, 121].

Lipids

It has been proven that the polyunsaturated fatty acid phosphatidylcholine (PC), which is present in soybeans, has antioxidant properties. PC may be a useful functional material for hepatic and renal injuries involving oxidative stress generated by AGE during diabetes conditions since cells treated with PC showed a considerable reduction in cytotoxicity and oxidative stress [39]. Soybeans are rich in saponins [122], and in T2D mouse models, these compounds have demonstrated the potential to restore body weight, suppress hyperglycemia and hyperlipidemia, restore insulin sensitivity, improve adipocyte thermogenesis, reduce the underlying hepatic damage caused by steatosis, and increase the ratio of hepatic phosphorylated adenosine monophosphate-activated protein kinase (p-AMPK) to the total protein [123]. A study was conducted utilizing DPPH˙ and OH˙ radical scavenging assays, a beta-carotene protection assay, and a heating oil system to examine the antioxidant ability of soybean germ and its oil from Northeast (NE-SG) and Shandong Province (SD-SG) of China. According to the findings, soybean germ phytosterols were more effective antioxidants in oil systems than in non-oil systems. Since it was more effective at lower temperatures (60°C) with longer times as opposed to higher temperatures (120°C and 180°C) with shorter times, these phytosterols’ antioxidant capacity was temperature- and time-dependent. These findings prove that soybean germ phytosterols can act as antioxidants to stop lipid oxidation in foods that have been kept at low temperatures for an extended period [40].

2.2. Anti-inflammatory activity

The critical role of inflammation in the pathophysiology of T2D has been proven. Chronic inflammation is often seen in individuals with T2D, and it may lead to insulin resistance which is a hallmark of the disease [124]. High sugar levels can lead to oxidative stress and inflammation which can contribute to the complications associated with T2D such as cardiovascular disease and nerve damage [125]. The function of tumor necrosis factor-alpha (TNF-α) in obesity, notably in insulin resistance and diabetes, is revealed by empirical research. Epidemiologic links between inflammation and obesity and T2D have been set up because of elevated levels of circulating pro-inflammatory markers and mediators which include fibrinogen, C-reactive protein, interleukin (IL)-6, plasminogen activator inhibitor-1, sialic acid, and white blood cells, in these conditions [126].

Small molecule natural products

The anti-inflammatory properties of soybeans also have antidiabetic effects. In a study conducted by Matsukawa et al. [41], black soybean seed coat extract (BSSCE), together with its active ingredient cyanidin-3-O-glucoside (Cy3G), has been shown to have antidiabetic benefits on db/db mice via enhancing adipocyte development. Additionally, it was found that 3T3-Ll cells treated with BSSCE and Cy3G differentiated into smaller adipocytes, which was associated with upregulated PPARγ and C/EBP gene expressions, upregulated adiponectin secretion, downregulated TNF production, activated insulin signaling, and upregulated glucose uptake. PGC-1, SIRT1, and UCP-3 gene expressions were all noticeably elevated in C2C12 myotubes cultivated in conditioned media made from 3T3-L1 adipocyte cultures treated with Cy3G.

Beneficial bioactive substances found in soybeans that have been proven to be effective against chronic inflammatory diseases such as T2D include isoflavones, peptides, fibers, polysaccharides, and lipids [127]. The isoflavone, daidzein, obtained from fermented soybean can prevent inflammation associated with T2D [42]. A study investigated the inhibitory effect of daidzein on the expression of pro-inflammatory cytokines, CCL2 and IL-6, in a co-culture of 3T3L1 adipocytes and RAW264 macrophages. The results revealed that treatment with daidzein significantly suppressed mRNA expression of CCL2 and IL6 in both the adipocytes and macrophages. In addition, daidzein inhibited the transcription of pro-inflammatory genes and reduced the mRNA levels of CXCL2 in TNFα-treated cells [42].

Peptides

Enzymatic hydrolysis, fermentation, food processing, and gastrointestinal digestion of soy protein give rise to soybean peptides (SPs), which aid in the absorption of amino acids in the body among other physiological activities [127]. Earlier studies have reported that administering synthesized soy peptide FLV caused a decrease in the production and effect of inflammatory molecules and increased the sensitivity of adipocytes to insulin [43]. An in vitro study using a Caco-2 cell line reported the anti-inflammatory effect of soy tripeptide Val-Pro-Tyr which decreased the expression of TNF-α, IL-6, IL-1, and IL-17 in those cells [44]. Another soy peptide, lunasin, also showed anti-inflammatory properties by decreasing the expression of COX-2 in mice colon induced by dextran sodium sulfate (DSS) [45].

Polysaccharides

Through underlying molecular pathways connected to inflammatory factors, numerous studies have proven a link between dietary polysaccharides and the alleviation of diabetes. The use of these polysaccharides has been shown to have hypoglycemic, hypolipidemic, antioxidant, and anti-inflammatory properties, which increase pancreatic β-cell mass and reduce β-cell dysfunction in in vivo and in vitro studies. Dietary polysaccharides activate the PI3K/Akt pathway, enhance insulin signaling pathways via insulin receptors, and eventually modulate the ERK/JNK/MAPK pathway [128]. Specifically, soluble soybean polysaccharide (SSPS1) with a molecular weight of 2,737 kDa obtained from soy whey enhanced the growth rate and phagocytic activity of RAW 264.7 macrophages and modulate the secretion of IL-1β, IL-6, INF-β, TNF-α, and nitric oxide [46]. The addition of soluble soybean polysaccharides to dairy products as a source of dietary fiber has been shown to reduce blood cholesterol, improve laxation, and reduce the risk of diabetes [129]. Also, He et al. [47] discovered that in a high-fat-diet animal model, metformin-soluble soybean polysaccharide nanoparticles (MET-SSPS-NPs) had a lipid-lowering effect. MET-SSPS-NPs can significantly downregulate PCSK9 and fatty acid-binding proteins (FABP5 and FABP7), and upregulate peroxisome proliferator-activated receptor gamma (PPARγ) expressions, consequently increasing the level of adipokines.

Lipids

Soybeans have omega-3 polyunsaturated fatty acids (O-3 PUFAs). Clinical and empirical data associates these compounds with decreased pro-inflammatory mediators in chronic inflammatory diseases like diabetes [130]. Bioactive metabolites produced by fatty acid oxygenases mediate the biological activities of O-3 PUFAs [48]. These compounds can alter the diversity and abundance of the gut microbiota, which in turn modulates T2DM [48, 131]. Specifically, O-3 PUFA reduces inflammation by inhibiting the production of pro-inflammatory cytokines, such as IL-17, in monocytes that have been exposed to harmful lipopolysaccharides (LPS) and stimulating the production of anti-inflammatory factors such as IL-10 [48]. In addition, the role of dietary sphingolipids in the prevention of metabolic syndrome has been studied. Dietary glucosylceramide (a sphingolipid) has anti-inflammatory properties which include suppressing the mRNA expression of pro-inflammatory cytokines such as IL-1β and IL-6, downregulating the activation of TNF-α, and suppressing tissue edema and leukocyte infiltration to the inflammatory site [49].

2.3. Alpha-amylase inhibition

Many enzymes are involved in the T2D pathogenesis (Figure 2). α-Amylase is a calcium metalloenzyme that plays a critical role in the breakdown of complex polysaccharides into smaller units such as glucose and maltose to ease their absorption into the body [132]. The enzyme also elevates blood sugar levels and produces postprandial hyperglycemia, and one of the strategies for regulating glucose levels in diabetes is the use of drugs or substances that can block the hydrolysis of carbohydrates by α-amylase [70, 133].

Figure 2

Enzyme targets for management of T2D. ↑ means activation. ↓ means inhibition. Created with BioRender.com

media/image4.png

Small molecule natural products

Isoflavones obtained from the ethanolic extract of Tempe (made from un-germinated soybeans) and Tempe flour (made from germinated soybeans) have been proven to be α-amylase inhibitors [134]. This study showed that the ability of the soybeans to inhibit α-amylase activity increased with germination, and this is evidenced by the IC50 values which were significantly lower in un-germinated soy Tempe. The results suggest that germination of soybeans helps increase the bioactive compound contents in soybeans, and this is due to various biochemical and enzymatic reactions such as the activation of protease enzymes, changes in pH, the breakdown of storage compounds, and changes in gene expression patterns. These changes occur in the seeds during germination, thereby improving their nutritional value and digestibility unlike un-germinated seeds [134].

In a study by Tan and Chang [50], high-performance liquid chromatography (HPLC) analysis of fractions of phenolic extract of black soybean was performed. Results showed that black soybean had phenolic acids such as syringic, gallic, p-hydroxybenzoic, vanillic, ferulic, protocatechuic, caffeic, sinapic, chlorogenic, and salicylic acids. Other flavonoids contained in black soybeans include catechin, quercetin-3-O-glucoside, myricetin, epicatechin, quercetin-3-O-glucoside, kaempferol-3-O-rutinoside, and kaempferol-3-O-glucoside. All the pure phenolic acids and flavonoids had anti-α-amylase inhibitory activity with myricetin as the strongest enzyme inhibitor [50].

Tiss et al. [51] proved that unfermented soy milk (USM) and kefir fermented soymilk (FSM) inhibit α-amylases with IC50 values of 72.60 and 52.71 μg/mL, respectively, with the strongest α-amylase inhibition activity recorded at 16 hours of fermentation by kefir. The strong inhibition of α-amylase might be due to increased amounts of isoflavone aglycones, including genistein, daidzein, and glycitein in the soy after fermentation by kefir. It was also observed that feeding high-fat fructose diet (HFFD) to rats with FSM led to a significant decrease in intestinal and pancreatic α-amylase activities by 26% and 31%, respectively. This, in turn, led to reduced hydrolysis of starch and a decrease in blood glucose level by 26%. This shows that FSM is more beneficial than USM in controlling glucose levels due to the increased bioavailability of soy isoflavones after fermentation [51].

Peptides

The benefits of soy peptides in the management of diabetes and insulin resistance have also been described by earlier studies [135]. Germinated SPs have been shown in vitro to control hyperglycemia through the inhibition of α-amylase in the saliva [135]. Bioactive peptides in soybeans are released through the activity of proteolytic enzymes which include proteases and peptidases produced by microorganisms during a fermentation process. These peptides are abundant in fermented soymilk, and when soymilk is fermented by kefir, it inhibits α-amylase. Studies in both animals and humans have revealed that fermented products of soybeans have better antidiabetic properties, unlike unfermented products. In a study of streptozotocin (STZ)-induced diabetic rats fed with fermented soybean, α-amylase was found to be inhibited [135]. Also, González-Montoya et al. [52] saw that six days germinated soybean protein digest (6GSPD) can inhibit salivary α-amylase with a lower IC50 of 1.70 compared to the positive control Acarbose with an IC50 of 0.16. Different peptide fractions were obtained from 6GSPD by ultrafiltration and evaluated. They showed varying levels of α-amylase inhibitory activity. Peptides that show anti-α-amylase activity have been found to contain high-molecular-weight amino acids and aromatic residues like phenylalanine, tryptophan, tyrosine, and arginine [52]. Apart from amino acid composition, peptides’ structures may modulate the inhibition of α-amylase through steric hindrance obstructing the active site of the enzyme; charge interactions affecting the binding affinity; secondary structures influencing the overall shape and flexibility of the peptide; hydrophobic interactions influencing the stability of the enzyme–peptide complex; and inducing conformational changes after binding with the enzyme [136].

Polysaccharides

Water-soluble soybean polysaccharide (SSPS) that has been derived from the leftover residue of soy processing is an acidic polysaccharide made up of arabinan and galactan as side chains and the primary backbone of galacturonan and rhamnogalacturonan. SSPS has a relatively low viscosity and great stability in an aqueous solution, and is primarily composed of the dietary fiber found in soybean cotyledons [137, 138]. Lin et al. [53] proved that SSPS has an α-amylase-inhibiting activity which is significantly increased by fermentation and microwave treatment. When compared to unfermented SSPS, the inhibitory activity of α-amylase in fermented SSPS and microwaved SSPS was 2.69 and 2.41 times greater at 11.69% and 10.47%, respectively. Also, insoluble dietary fibers from soybean hull demonstrated an α-amylase inhibition activity and a significant effect on the fecal microbiota increasing the abundance of Lactobacillales and Bifidobacteriales. In this regard, the citric acid extract of the insoluble dietary fiber from soy hull performed better than the oxalic acid extract [54].

The intricate interplay of molecular forces and structural properties characterizes the physical and chemical interactions between α-amylase and soy polysaccharides. By occupying the active site of α-amylase, soy polysaccharides could interact with it through non-specific binding and competitive inhibition [139]. Soy polysaccharides can also alter α-amylase’s kinetic properties, such as substrate binding affinity and catalytic rate, which can impact how the enzyme interacts with polysaccharides. This modification could lead to variations in the enzyme’s starch hydrolysis effectiveness [140]. α-Amylase can also undergo structural modifications because of binding with soy polysaccharides. These modifications could have an impact on the geometry or general structure of the enzyme’s active site, which would affect how well it interacts with starch molecules [141]. Moreover, there may be a pH-dependent interaction between α-amylase and soy polysaccharides. The charge distribution on the surfaces of the polysaccharides and the enzyme can change due to pH changes, which can affect how the two interact [142].

Lipids

Quan et al. [55] investigated the hypoglycemic effects of soyasaponins (SS) in GK/Jcl T2D rats. Through colorimetric methods, it was discovered that soybean hypocotyl extract fed to the diabetic rats showed low α-amylase inhibitory activity at the concentration of 1 g/L.

The distinctive structure of soyasaponins is derived from a triterpene or steroidal aglycone that is linked to one or more sugar moieties. The potential interactions and inhibitory effects of soyasaponins on α-amylase may be limited in comparison to other compounds found in soybeans [143]. This limitation might result from soyasaponins’ overall hydrophilic nature due to the presence of sugar moieties. This could cause steric hindrance when interacting with the hydrophobic active site of α-amylase. Furthermore, the sugar moieties could increase molecular weight, which may affect binding kinetics [144].

2.4. Alpha-glucosidase inhibition

The mammalian α-glucosidase is synthesized in the mucosal brush border of the small intestine, and its major physiological function is to catalyze the digestion of disaccharides and starch which make up a huge percentage of the human diet [145]. The enzyme selectively catalyzes the endonucleolytic or exonucleolytic hydrolysis of various sugars producing monosaccharides, oligosaccharides, or glycosaminoglycans, which increases postprandial blood glucose [146148]. This enzyme is considered a crucial clinical target for the treatment of T2D [149].

Small molecule natural products

Evidence has shown the availability of α-glucosidase inhibitors in soybeans. Both the wild soybean (G. soja; WS) and the cultivated soybean (G. max; CS) extracts displayed dose-dependent α-glucosidase inhibitory activities. The α-glucosidase inhibitory activity of the extracts from WS (84.82 0.34%) was higher than that of the extracts from CS (56.06 0.24%). This could be because 42 of the 79 flavonoid compounds that were found to be present in these plants after metabolomic analysis were elevated in WS. The most distinct compounds in WS were protocatechuic aldehyde, garbanzol, 2′-hydroxygenistein, ligustilide, and resveratrol [104].

The most prevalent polyphenol in soy leaves, coumestrol, proves strong α-glucosidase inhibitory activity, with inhibition levels rising as the plant develops [56]. Other polyphenols in soybeans – which include luteolin, glyceollin, daidzein, and genistein – have been shown to exert α-glucosidase inhibitory activity [57]. Daidzein is converted into glyceollin in soybeans during fermentation due to the activity of Aspergillus sojae. Glyceollin has been revealed to be a competitive inhibitor of α-glucosidase, while genistein is noncompetitive – luteolin showed a mixed mode of activity. Glyceollin and luteolin present a synergistic effect on α-glucosidase inhibition [57]. Also, through a molecular docking study, it was discovered that 6-gingerdiol and stigmasterol have a potent inhibitory effect on α-glucosidase activity [57]. Gao et al. [58] reported that the Japanese traditional fermented soybean dish, Okara, has a remarkable α-glucosidase inhibitory capacity due to the activity of a newly isolated strain of Bacillus amyloliquefaciens SY07. 1-Deoxynojirimycin (DNJ) was shown to be one of the bioactive constituents responsible for the anti-α-glucosidase activity of the fermented broth. The good α-glucosidase inhibitory activity of the Okara broth could make soybean a novel potential microbial source of antidiabetic products.

Protein

A soybean protein (ESPro1) with a significant α-glucosidase inhibitory activity was discovered in a recent study when it was extruded at 160°C and 30 rpm. The fraction (ESPro1 F3) with the highest inhibitory activity was isolated using digestion, ultrafiltration, and gel filtration chromatography. The ESPro1 F3 fraction was then used to screen and produce six peptides with α-glucosidase inhibitory activity, with LLRPPK displaying the highest inhibitory activity (46.98 0.63%) [59]. In another study, soybean protein hydrolysate was produced by hydrolyzing soybean protein with trypsin after being pre-treated with ultrasound. Ultrafiltration, ion exchange chromatography, gel filtration chromatography, and reversed-phase HPLC were used to separate the soybean protein hydrolysate. The fasting blood glucose levels of experimental mice were significantly lowered by an ultrafiltration-derived, highly active part with a molecular weight of less than 5 kDa. The reversed-phase HPLC fraction C-III-2a had the highest α-glucosidase-inhibitory activity, and the half-inhibitory concentration (IC50) was 0.049 mg/mL. By using ion trap tandem mass spectrometry, the amino acid sequences of two tripeptides from fraction C-III-2a were determined to be Gly-Ser-Arg and Glu-Ala-Lys [60]. Also, three peptides (LLPLPVLK, SWLRL, and WLRL) isolated from soy protein with alkaline proteinase digestion, purified by size-exclusion and anion-exchange chromatography, and found using tandem MS have been shown to possess high α-glucosidase inhibitory [61].

Lipids

Ha et al. [62] investigated the α-glucosidase inhibitory activity of soyasaponin derivatives. Through simple reversible slow-binding inhibition, two derivatives, soyasaponin Ba and soyasaponin Bb, significantly inhibited the activity with the IC50 of ca. 25 and 32 μM. In a different study, soyasaponins extracted from soybean hypocotyls using ODS column chromatography and HPLC showed noncompetitive inhibition of α-glucosidase with IC50 values ranging from 10 to 40 μmol/L. The findings imply that soyasaponins, which have α-glucosidase-inhibiting properties, may be physiologically helpful in reducing postprandial hyperglycemia in diabetic patients [150].

Polysaccharides

Acidic polysaccharides, known as soluble soybean polysaccharides (SSPS), are isolated from soybean cotyledons and have a high degree of stability and low viscosity in an aqueous solution. SSPS have a structure like pectin. The major backbone of SSPS was shown to be formed of galacturonan and rhamnogalacturonan [63]. Soluble polysaccharides have been proven to inhibit α-glucosidase in several plant species [151153].

2.5. Aldose reductase inhibition

AR is an enzyme in the polyol pathway involved in glucose metabolism by catalyzing the conversion of glucose into sorbitol (which is after oxidized to fructose). Sorbitol can accumulate in various tissues, leading to oxidative stress and tissue damage. The activation of AR has been implicated in the development of diabetic complications, such as neuropathy, nephropathy, and retinopathy. Inhibiting this enzyme in T2D can delay or prevent the pathogenic process [154]. Even in normoglycemic states, AR plays a critical role in promoting inflammatory and cytotoxic conditions. Its activation under hyperglycemic settings leads to the development of chronic diabetes complications. As a result, it is an established antidiabetic drug target [155].

Small molecule natural products

AR inhibitors are substances that can block the activity of this enzyme, potentially reducing the accumulation of sorbitol and preventing related tissue damage [155]. Some studies have shown that certain compounds present in soybeans, such as isoflavones, may possess AR inhibitory activity. These findings suggest that consuming soybeans or soy-derived products could potentially help regulate blood sugar levels and reduce the risk of diabetic complications [64]. Using NMR, UV, and IV spectroscopy, the flavonoids genistein, kaempherol, naringenin, quercetin, morin, and rutin were identified and extracted from the UFV-5′ cultivar of G. max (soya). At a concentration of 10−4 M, quercitrin and morin showed the greatest inhibitory activity. But at a concentration of 10−5 M, quercetrin, rutin, and naringenin had the strongest effects [64].

Found in Korean fermented soybean pastes (Doen-jang), genistein was shown to have an AR inhibiting activity. With an IC50 value of 20 μM, genistein inhibited the pig lens’ AR [156]. In another study, genistein has been shown to inhibit AR activity in human lens epithelial cells (HLE-B3). HLE-B3 cells cultivated in high-glucose conditions show downregulated expression of TGF-β2, αB-crystallin, and fibronectin mRNAs when exposed to purified genistein derived from Pueraria lobate roots. Furthermore, there was a decrease in thiobarbituratic acid-reactive compounds and glutathione (GSH) levels. These findings prove that genistein protects HLE-B3 cells against lens opacity and prevents high glucose–mediated harmful effects. Since genistein is probably able to inhibit AR and cellular oxidation to achieve these effects, it could be a promising therapeutic agent for the treatment and prevention of problems related to T2D, including diabetic cataracts [65].

Daidzein is another isoflavone found in soybeans that has been shown to have AR inhibitory activity. In experimental animals, it has been proven to ameliorate diabetic retinopathy protecting against retinal damage in hyperglycemic conditions. In this study, it was discovered that AR and sorbitol dehydrogenase levels were markedly lowered (p < 0.001 and p < 0.01, respectively) upon treatment with 100 mg/kg of daidzein. The results of the 28-day daidzein treatment showed a considerable reduction in oxidative stress as well. Also, retinal thickness decreased with daidzein administration, according to histopathological studies [66]. An analog of daidzein, 8-hydroxydaid-zein, has also been proven to be an AR inhibitor. 8-Hydroxydaid-zein was isolated and found from a methanolic extract of okara (soybean pulp) fermented with the fungal strain Aspergillus sp. HK-388 and showed noncompetitive inhibition of human recombinant AR with respect to DL-glyceraldehyde, its Ki value being evaluated as 7.0 μM [67]. Daidzein can be transformed into ortho-dihydroxy isoflavones and glycitein by the bacteria Bacillus subtilis Roh-1. Glycitein, which is also known to be an inhibitor of AR, can be produced from natural-fermented soybean paste Doenjang [68].

2.6. DPP-4 inhibition

The DPP4 is a transmembrane exopeptidase which is involved in the selective degradation of Incretins, and increased levels and activity of this enzyme have been implicated in T2D [157, 158]. Secreted from the intestines, Incretins such as glucagon-like peptide-1 and -2 (GLP-1 and GLP-2) and glucose-dependent insulinotropic peptide (GIP) are metabolic hormones that induce the production of insulin by the pancreatic β cells to reverse elevated glycaemia [159]. The “incretin effect” is caused by GLP-1 and GIP binding to their respective receptors, GLP-1R and GIPR. This initiates an intracellular signaling cascade that results in changes to ion channel function, increased levels of cytosolic calcium that help fuse insulin granules to the plasma membrane, and improved exocytosis of granules that carry insulin. This culminates in an increase in insulin secretion in a glucose-dependent way. Other insulinotropic activities include increased β-cell proliferation and improved resistance to apoptosis [160]. Consequently, targeting DPP4 has become the focal point of many recent studies seeking diverse approaches in the management of T2D and other metabolic disorders [159].

Small molecule natural products

In a recent study, the antidiabetic potential of soybean and ginger (Zingiber officinale) was investigated in silico and from the ethanolic extracts from both plants. The results revealed that stigmasterol from soybeans and 12-shogaol from ginger were strongly potent against DPP4. Comparatively, the soybean extract was found to be a more potent inhibitor of DPP4 than ginger [69].

Hydrolysates and peptides

Other studies have explored the inhibition of DPP4 using natural alternatives such as hydrolysates and bioactive peptides from foods rich in protein. The bioactivity of hydrolysates and biopeptides is dependent on the composition, length, and sequence of amino acids. DPP4 inhibitory peptides in soybean can be released by enzymatic hydrolysis in the gut, germination, and fermentation [52], and can be found using approaches such as liquid chromatography-mass spectrometry (LC-MS) and fractionation driven by bioactivity [161]. To investigate the potential of germinated soybean hydrolysate and biopeptides to inhibit DPP4, pepsin and pancreatin were used to hydrolyze soybean germinate for six days [70]. Results obtained show that the hydrolysate mostly impaired the activity of DPP4 at an inhibitory concentration (IC50) value of 1.49 mg/mL, and peptides ranging from 5 to 10 and >10 kDa had an IC50 value of 0.91 and 1.18 mg/mL, respectively. Their research also showed that peptides with proline found between the first and fourth positions at the N-terminal (VVNPDNNEN, EEPQQPQQ, QEPQESQQ, SQRPQDRHQ, PETMQQQQQQ, and SSPDIYNPQAGSVT) displayed higher inhibitory properties [70].

Based on their evaluation, Rivero-Pino et al. [162] reported that treating soybean with subtilisin, trypsin, and flavorzyme yielded soybean hydrolysate as the highest DPP4 inhibitor at a minimal concentration as compared to chickpea, lentil, and pea with an IC50 value of 2.68 mg/mL. Furthermore, APFL, TPVGM, and EPAAV were identified as biopeptides from soybean which had DPP4 inhibitory activity, with EPAAV as the most potent. This differential attribute of soybean was suggested to be due to the higher proportion of proline and alanine in its amino acid composition [71]. The DPP4 inhibitory activity of Soy1 (IAVPTGVA) peptide was studied in situ using a Caco-2 cell line from the human intestine and ex vivo using human serum from venous blood. Lammi et al. [72] report that Soy1 inhibited the activity of DPP4 both in situ (IC50 value = 223.2 μM) and ex vivo (27.7% and 35.0% at 100.0 and 300.0 μM concentrations) in a dose-dependent manner.

Besides, results from assays conducted by Lammi et al. [163] using pepsin and trypsin showed that both enzymes yielded protein hydrolysates and biopeptides from soybean which inhibited the activity of DPP4 both in vitro and in situ. For the in vitro tests, the human recombinant enzyme (H-Gly-Pro-AMC) was used as substrate, while the in situ assays were performed on human Caco-2 cells. In vitro and at concentrations of 1.0 and 2.5 mg/mL, pepsin reduced the activity of DPP4 by 16.3% and 31.4%, respectively, and trypsin by 15.3% and 11.0%, respectively. Using Caco-2 cells, the activity of DPP4 was also impaired in a dose-dependent trend (pepsin: 15.4%, 19.6 %, and 37.0%; trypsin: 17.0%, 25.2%, and 43.3%; both enzymes at concentrations of 1.0, 2.5, and 5.0 mg/mL, respectively).

2.7. Pancreatic lipase inhibition

Pancreatic lipase (PL) enzyme mainly hydrolyzes triacylglycerols into monoacylglycerols and free fatty acids, its inhibition being therefore considered as a potential treatment for obesity and T2D [77]. A lipase inhibitor may work through at least three different primary ways. The first is by directly attaching to lipase, the second is by acting as a substrate analog that either permanently or reversibly affects the active site, and the third method is by acting on the substrate to reduce the substrate’s availability. [164]. The crude extracts of legumes could inhibit PL, and they can be used to produce nutritional supplements to fight obesity. These beans could also be used to create nutritionally dense meal replacements and substitute foods that could reduce the rate at which lipids are absorbed [165]. In an extensive comparative study carried out to examine the PL inhibitory effects of 13 commonly consumed dietary legumes grown in China, it was discovered that yellow soybean saponin extract demonstrated a 34.1% inhibitory activity, and it could be a target for more research on the anti-obesity and anti-diabetes properties using cell and animal models [166].

Small molecule natural products

Natural compounds such as saponins, alkaloids, polyphenols, and other substances have been found as effective anti-obesity agents [166]. It is well recognized that the presence of isoflavone aglycones, such as daidzein, glycitein, and genistein, is what gives soybeans their health benefits because they are absorbed more quickly by humans than isoflavone glucosides. Furthermore, according to numerous research, the bioconversion of isoflavone aglycones by bacteria in fermented soybean meals such soy milk, soy yogurt, Cheonggukjang, and Doenjang results in larger amounts of these compounds than isoflavone glucosides [167].

The inhibitory activities of unfermented soy-powder milk (UFSPM) and soy-powder milk fermented using L. plantarum P1201 (FSPM) on PL have been set up. PL showed a dose-dependent inhibitory activity in response to UFSPM and FSPM. In comparison to the control group, FSPM, but not UFSPM, inhibited adipogenesis in 3T3-L1 cells and decreased their triglyceride content by 23.1% after treatment with 1,000 μg/mL of FSPM. The anti-obesity effect of FSPM can be attributed to conjugated linoleic acid (CLA) and isoflavone aglycones, which downregulated lipoprotein lipase (LPL), adiponectin, adipocyte fatty acid-binding protein (aP2), fatty acid synthase (FAS), and acetyl CoA carboxylase (ACC) mRNA. Additionally, FSPM increased the anti-obesity activity and pancreatic enzymes’ inhibitory function [168]. These findings suggest that the fermented hydrolysate milks may have helped to increase the value of the soybean [168]. Hwang et al. [169] also revealed that in soy-powder yogurts (SPYs) made using Lactobacillus brevis and L. plantarum, the total average isoflavones decreased while there was an increase in aglycones. Also, the highest antioxidant potential and enzyme-inhibiting activities were seen in SPY employing mixed strains.

In another study, isoflavone-enriched soybean leaves (IESLs) were solid-lactic acid fermented using starters Lactiplantibacillus plantarum P1201 and Levilactobacillus brevis BMK184. This solid-lactic acid fermentation resulted in alterations in the isoflavone levels and the inhibition of digestive enzymes. The amount of isoflavones in total declined throughout fermentation, while the amount of isoflavone aglycones increased. This fermentation process using P1201 + BMK184 made the IESL attain a PL inhibition activity value of 37.3% [170]. Suh et al. [73] used ultra-performance liquid chromatography with quadrupole-time-of-flight mass spectrometry (UPLC + QTFMS) to examine the changes in metabolites during the fermentation of soybeans. The results showed that the soybean’s isoflavone glycosides and soyasaponins were reduced, but their aglycones, like daidzein, glycitein, and genistein, were elevated. These aglycones also proved PL inhibition activities. Specifically, genistein has been shown to lower hyperlipidemia by preventing the production of hepatic lipids. Because genistein prevents PL activity, which promotes increased dietary fat excretion without absorption and reduces adipocyte development, it can be used as an effective treatment for obesity. Large amounts of proteins are discharged through the urine because of glomerulonephritis, which causes hypoalbuminemia and secondary hyperlipidemia that are brought on by the liver producing more lipid and lipoproteins. By preventing the production of hepatic lipids, genistein can lower hyperlipidemia [171].

Proteins

A protein molecule obtained from the crude aqueous extract of soybean meal, called the protein inhibitor of lipase (PIL), has been shown to inhibit the activity of PL on tributyrin or triolein. The activity of PIL, which has a molecular weight of approximately 70kDa, is remarkably similar to those of hydrophobic proteins such as serum albumin and lactoglobulin which inhibit PL in the absence of colipase and bile salt [74]. The ability of PIL to interact with lipids (by penetrating the monomolecular films of different glycerides and phospholipids) and alter the nature of the substrate–water interface is a factor that enhances its ability to inhibit PL [74]. The binding of PIL to substrate micelles stops lipase from acting normally on the interfacial region between the micelle surface and the aqueous medium [172]. PIL, which also inhibits the lipase activity of powdered porcine pancreas, has been proved to be a protein by the fact that it’s rendered inactive by heat and pronase treatment [173, 174].

Another lipase-inhibiting protein in the soybean seed is lipoxygenase-1 (LOX-1). A study revealed that the amount of LOX-1 needed to achieve a 50% reduction in lipase activity was 3.2 × 10−7 M, and it reduced the release of fatty acids from a soybean oil emulsion in a dose-dependent manner. The soybean LOX-1 gene was transfected into Escherichia coli and the recombinant LOX-1 product also inhibited the activity of the lipase [75, 174]. From soybean seeds lacking in lipoxygenase (LOX), another 56.0 kDa lipase-inhibiting protein has been discovered. The estimated masses of the hypothetically expected lysyl endopeptidase–treated peptides of soybean seed protein were nearly identical to the masses of the lysyl endopeptidase–digested peptides of the 56.0 kDa inhibitory protein. The protein identified as an α-amylase prevented PL from hydrolyzing soybean oil though to a lesser extent than by LOX-1 [76].

The pancreatic lipase-inhibiting activity of mature (MSPHs) and young (YSPHs) soybean enzymatic protein hydrolysates was investigated in a recent study. Using LC-MS QTOF and molecular binding studies, PHs with strong anti-lipidemic effects were then chosen for sequencing. It was predicted that the peptides FPFPRPPHQ, QCCAFEM, and FAPEFLK from MSPHs and SFFFPFELPRE, FMYL, PFLL, FPLL, and LPHF from YSPHs would have strong inhibitory effects on PL. The findings suggest that MSPHs and YSPHs beans could be identified as possible components in the treatment of hypercholesterolemia [77]. In another study, in vitro and in silico analyses have also identified PL-inhibiting low-molecular-mass (LMMH) and high-molecular-mass (HMMH) peptides obtained from a protein hydrolysate from soybean protein concentrate produced by papain [175].

Polysaccharides

Dietary fibers, which include a range of gums, mucilages, and algal polysaccharides, are mostly indigestible complex carbohydrates found in plant cell walls (cellulose, hemicellulose, and pectin). Plant cell walls include lignin, a non-carbohydrate part of dietary fiber [176]. Soluble dietary fibers (SDFs) have proven their ability to control blood cholesterol and sugar levels due to their structural and functional properties [177]. Soybean hull galactomannan has also been shown to inhibit PL activity. Following pre-incubation of galactomannan with the enzyme, the impact of galactomannan on PL was investigated at various concentrations (50, 100, 150, and 200 ppm). Data showed that pre-incubating lipase with galactomannan resulted in a concentration-dependent suppression of enzyme activity, with the inhibition percentages ranging between 5.06% and 19.68% for the lowest concentration (50 ppm) and the highest concentration (200 ppm), respectively [78]. The inhibitory activity of galactomannan may result from one or more of the following mechanisms: the direct effect of galactomannan on enzyme, the reduction of enzyme–substrate binding due to a relatively high viscosity of the reaction medium (galactomannan), changing the conformation of the enzyme, interaction between galactomannan and enzyme (as protein), and interaction between galactomannan and enzyme substrate. These findings suggest that galactomannan can be used to inhibit lipase activity [78].

2.8. Angiotensin-converting enzyme inhibition

Angiotensin-converting enzyme (ACE) has been extensively investigated in relation to glucose metabolism and renal injury in diabetes [178]. Studies show that most diabetic complications such as nephropathy and hypertension are of vascular origin with the involvement of the renin–angiotensin–aldosterone system (RAAS). ACE is an important enzyme that plays a crucial role in the RAAS, which is involved in regulating blood pressure, fluid balance, and electrolyte homeostasis in the body. ACE is responsible for converting angiotensin I to angiotensin II, a potent vasoconstrictor that also stimulates the release of aldosterone [179]. ACE inhibitor therapy reduces both microvascular and macrovascular complications in diabetes and appears to improve insulin sensitivity and glucose metabolism [180].

In the context of diabetes, ACE and the RAAS have been implicated in the development and progression of diabetic complications [178]. Some of these include diabetic nephropathy ACE contributes to the pathogenesis of diabetic nephropathy through renal vasoconstriction, inflammation, oxidative stress, and the deposition of extracellular matrix proteins, ultimately causing kidney damage [181] and diabetic retinopathy where ACE promotes retinal vascular permeability and inflammation, contributing to the development of retinopathy [182]. Cardiovascular complication is another condition where ACE is associated with an increased risk of cardiovascular diseases such as hypertension, coronary artery disease, and heart failure [183].

ACE inhibitors are drugs commonly used for the treatment of hypertension and heart failure. They work by inhibiting the activity of ACE, an enzyme involved in the production of angiotensin II, which causes blood vessel constriction and increases blood pressure [183]. ACE inhibitors have also shown potential benefits in managing diabetes by enhancing insulin sensitivity and improving glucose metabolism [184]. By reducing angiotensin II levels, ACE inhibitors may trigger insulin activity and promote glucose uptake in tissues [185].

Small molecule natural products

Soy consumption has been associated with improved endothelial function and reduced markers of inflammation, which may have beneficial effects on diabetes and cardiovascular health [186]. This may be because soybeans contain phenolic-rich extracts that can inhibit ACE linked with T2D and hypertension [13]. A compound found in soybeans, chicoric acid, a polyphenol, has also been implicated in enhancing insulin secretion and increasing glucose uptake, suggesting its potential in diabetes management [79]. When ACE-inhibitory compounds from soybeans were analyzed in silico, beta-sitosterol showed the most potential for inhibiting ACE I when compared to 33 other soy phytomolecules [80]. Nicotianamine, found in soybeans, is a new inhibitor of ACE 2 [81].

Protein and peptides

Glycinin is the major storage protein of soybean. Different enzymatic hydrolysates of glycinin have shown ACE-inhibitory properties. The protease P hydrolysate, having a pentapeptide with the sequence Val-Leu-Ile-Val-Pro, demonstrated the greatest ACE-inhibitory activity (1.69 ± 0.17 μM). This peptide, which is produced from glycinin, has antihypertensive properties that may be useful in the creation of functional foods for medicinal purposes [82]. ACE-inhibitory fractions, namely, fractions 2, 3, and 4, were obtained from the hydrolysis of soybean protein isolate using pepsin and pancreatin. The values of the hydrolysate fraction that inhibited 50% of ACE activity were 1.09, 0.42, and 0.25 mg/mL, respectively, for fractions 2, 3, and 4. As revealed by kinetic studies, fractions 2, 3, and 4 are noncompetitive, competitive, and mixed ACE inhibitors, respectively. The ACE-inhibitory activity is due to the scavenging properties of cationic peptides in the fractions. The strongest antioxidant was fraction 3, whereas fraction 4 hardly showed any scavenging activity. Because fraction 3 has significant ACE-inhibitory and free-radical scavenging qualities, we determined that it had the greatest potential for application as a bioactive component [187]. Another soy storage protein, β-conglycinin, has also been implicated in regulating insulin sensitivity and ACE activity in rats fed with a high-salt diet, potentially slowing the progression of diabetic nephropathy [188].

Two SPs, LPYP and IAVPTGVA (Soy1), have been shown to have hypoglycemic and hypocholesterolemic properties in vitro. Examining the effects of Soy1 and LPYP in vitro, the results show that the intestinal and renal ACE activities were reduced, with IC50 values of 14.7 ± 0.28 and 5.0 ± 0.28 μM (Caco-2 cells) and 6.0 ± 0.35 and 6.8 ± 0.20 μM (HK-2 cells), respectively [189]. ACE-inhibitory peptides have also been found in fermented soybean extract (FSE). After 3 days of aging, FSE exhibited ACE-inhibitory activity with an IC50 value of 1.46 mg/mL. Ultrafiltration and consecutive chromatographic methods revealed that FSE has a novel ACE-inhibitory peptide with the sequence Leu-Val-Gln-Gly-Ser. The peptide had an IC50 value of 22 μg/mL (43.7 μM) which is a 66-fold increase in ACE-inhibitory activity compared to that of FSE [83].

2.9. Phosphoenolpyruvate carboxykinase inhibition

The phosphoenolpyruvate carboxykinase (PEPCK) is an enzyme involved in gluconeogenesis, a metabolic pathway that generates glucose from non-carbohydrate sources such as amino acids, lactate, and glycerol. In T2D, patients produce excess glucose through gluconeogenesis with the help of PEPCK [190].

In HepG2 cells and in mice with alloxan-induced diabetes, the impact of genistein on PEPCK and glucose synthesis has been investigated. The expression of cytosolic PEPCK (PEPCK-C) and the signaling pathways were examined in HepG2 cells after exposure to various concentrations of genistein in the presence or absence of modulators. Genistein boosted AMPK, MEK½, ERK½, and CRTC2 phosphorylation states while decreasing PEPCK-C expression and glucose production in HepG2 cells [84]. Genistein-induced MEK½ and ERK½ activities were increased after treatment with the AMPK inhibitor (compound C), suggesting that the two signaling pathways may interact. In mice with alloxan-induced diabetes, genistein treatment also decreased fasting glucose levels, along with decreased PEPCK-C expression and enhanced AMPK and ERK12 phosphorylation states in the liver. By lowering glucose production and inhibiting PEPCK-C expression in HepG2 cells as well as in mice that have been given alloxan to cause diabetes, genistein satisfies the requirements for an effective antidiabetic drug. These findings suggest that genistein is a strong candidate for preventing T2DM by altering the AMPK-CRTC2 and MEK/ERK signaling pathways, which may open the door to a novel strategy for changing dysfunction in hepatic gluconeogenesis in T2DM [191].

Small molecule natural products

It has been discovered that the soy isoflavone daidzein has a remarkable therapeutic potential for treating the pathophysiology of T2D. The main source of daidzein is fermented soybean, but it can also be made by consuming the glycosylated moiety of daidzein and then having an intestine bacterial enzyme hydrolyze it. Daidzein has been shown in numerous studies to have a preventative effect on the reduction of hyperglycemia, insulin resistance, dyslipidemia, obesity, inflammation, and other T2D complications [42].

In a study by Choi et al. [191], genistein and daidzein supplements enhanced the insulin/glucagon ratio and C-peptide level in non-obese diabetic (NOD) mice while keeping the insulin-staining cell of the pancreas. When compared to the control group, genistein and daidzein supplements decreased the activities of the lipogenic enzymes glucose-6-phosphatase (G6Pase) and PEPCK, and increased the activities of malic enzyme and glucose-6-phosphate dehydrogenase (G6PD). Significantly, genistein and daidzein supplementation decreased the activities of carnitine palmitoyltransferase (CPT) and fatty acid oxidation in these animals. In comparison to the control group, genistein and daidzein enhanced plasma triglyceride and free fatty acid (FFA) concentrations. These findings imply that genistein and daidzein regulate glucose homeostasis by decreasing G6Pase, PEPCK, fatty acid oxidation, and CPT activities, while increasing malic enzyme and G6PD activities in the liver and keeping pancreatic beta-cells [191].

Another strategic target for T2D control is the peroxisome proliferator-activated receptors (PPAR) which regulate insulin sensitivity and blood glucose homeostasis. Isoflavones such as genistein, daidzein, formononetin, and biochanin A bind and activate PPARγ and PPARδ. This isoflavone mechanism of action is of special importance in the context of diabetes [85].

3. Conclusion

Empirical data reveal that the consumption of soybeans reduced the risk of T2D by 17% [192]. Small molecules, polysaccharides, lipids, and peptides derived from soybeans have demonstrated modulatory properties on key enzymes that affect T2D pathogenesis. Interestingly, isoflavones such as genistein and daidzein, and glycitein affect multiple targets. Overall, soybeans decrease in fasting blood glucose levels, improve insulin sensitivity, control lipid profile, and offer promising health benefits, particularly in relation to T2D. While soybeans have been shown to have good impacts on diabetes management, individual differences may exist in responses. However, more research is needed to determine specific dose–response effects and any interactions with other drugs or diets.

Abbreviations

ABTS:

2,2’-Azino-bis (3-ethylbenzothiazoline-6-sulfonic acid)

ACC:

Acetyl CoA carboxylase

ACE:

Angiotensin-converting enzyme

AGE:

Advanced glycation end products

Akt:

RAC (Rho family)-alpha serine/threonine-protein kinase

AMPK:

Adenosine Monophosphate-Activated Protein Kinase

aP2:

Adiponectin, adipocyte fatty acid-binding protein

ARE:

Antioxidant responsive element

BBI:

Bowman-Birk Inhibitor

BDNF:

Brain-derived neurotrophic factor

BSSCE:

Black soybean seed coat extract.

C/EBP:

CCAAT-enhancer-binding proteins

CCL2:

Chemokine (C-C motif) ligand 2

CLA:

Conjugated linoleic acid

CPT:

Carnitine palmitoyltransferase

CRTC2:

CREB-regulated transcription coactivator 2

CXCL2:

C-X-C motif chemokine ligand 2

Cy3G:

Cyanidin-3-O-glucoside

DNA:

Deoxyribonucleic acid

DPPH:

2,2-Diphenyl-1-picrylhydrazyl

ERK:

Extracellular signal-regulated kinase

FABP:

Fatty acid-binding protein

FAS:

Fatty acid synthase

FFA:

Free fatty acid

FRAP:

Ferric-reducing ability of plasma

FSE:

Fermented soybean extract.

FSM:

Fermented soymilk

G6Pase:

Glucose-6-phosphatase

G6PD:

Glucose-6-phosphate dehydrogenase

GFAT:

Glutamine fructose-6-phosphate amidotransferase

GIP:

Glucose-dependent insulinotropic peptide

GLP:

Glucagon-like peptide

GPX:

Glutathione peroxidase

GSH:

Glutathione

GST:

Glutathione S-transferase

HBP:

Hexosamine biosynthesis pathway

HFFD:

High-fat fructose diet

HO-1:

Heme oxygenase-1

HORAC:

Hydrophilic oxygen radical absorbance capacity

HPLC:

High-performance liquid chromatography

IL:

Interleukin

IRS:

Insulin receptor substrate

JNK:

c-Jun N-terminal kinases

Keap 1:

Kelch-like ECH-associated protein 1

LC-MS QTOF:

Liquid chromatograph-mass spectrometry. quadrupole time-of-flight

LDL:

Low-density lipoprotein

LOX-1:

Lipoxygenase-1

LPL:

Lipoprotein lipase

MAPK:

Mitogen-activated protein kinases

MCI:

Mild cognitive impairment

MDA:

Malondialdehyde

MEK:

Mitogen-activated protein kinase kinase

mRNA:

Messenger RNA

NAD+:

Nicotinamide adenine dinucleotide

NADH:

Nicotinamide adenine dinucleotide hydrogen

NADPH:

Nicotinamide adenine dinucleotide phosphate

NIDDM:

Non-insulin-dependent diabetes mellitus

NOD:

Non-obese Diabetic

NQO1:

NAD(P)H quinone oxidoreductase 1

Nrf2:

Nuclear factor erythroid 2-related factor 2

O-3 PUFA:

Omega-3 polyunsaturated fatty acids

p-AMPK:

Phosphorylated adenosine monophosphate-activated protein kinase

PC:

Phosphatidylcholine

PEPCK:

Phosphoenolpyruvate carboxykinase

PGC-1:

Peroxisome proliferator-activated receptor gamma coactivator

PI3K:

Phosphoinositide 3-kinase

PIL:

Protein Inhibitor of Lipase

PKC:

Protein kinase C

PL:

Pancreatic lipase

PPAR:

Peroxisome proliferator-activated receptors

RAAS:

Renin–angiotensin–aldosterone system

RAGE:

Receptor for advanced glycation end products

RNS:

Reactive nitrogen species

ROS:

Reactive oxygen species

SDF:

Soluble dietary fibers

SDH:

Sorbitol dehydrogenase

SIRT1;

Sirtuin 1

SOD:

Superoxide dismutase

SSPS:

Soybean soluble polysaccharides

STZ:

Streptozocin

T2D:

Type 2 diabetes

T2DM:

Type 2 diabetes mellitus

TGF-β2:

Transforming growth factor beta-2

TNF-α:

Tumor necrosis factor-alpha

UCP-3:

Mitochondrial uncoupling protein 3

UPLC + QTFMS:

Ultra-performance liquid chromatography with quadrupole-time-of-flight mass spectrometry

USM:

Unfermented soy milk

Funding

The authors declare no financial support for the research, authorship, or publication of this article.

Author contributions

Conceptualization, A.B.R.; methodology, A.B.R, G.C.I. and J.A.O.; validation, G.C.I., J.A.O. and V.O.A.; formal analysis, A.B.R., G.C.I., J.A.O. and V.O.A.; investigation, A.B.R., L.C.E., E.J.N., O.A.O., J.A.O., V.O.A. and G.C.I.; resources, A.B.R., L.C.E., E.J.N., O.A.O., J.A.O., V.O.A. and G.C.I.; data curation, A.B.R., L.C.E., E.J.N., O.A.O., J.A.O., V.O.A. and G.C.I.; writing—original draft preparation, A.B.R., L.C.E., E.J.N., O.A.O., J.A.O., V.O.A. and G.C.I.; writing—review and editing, A.B.R., G.C.I., J.A.O. and V.O.A.; visualization, A.B.R.; supervision, A.B.R.; project administration, A.B.R. and G.C.I. All authors have read and agreed to the published version of the manuscript.

Conflict of interest

The authors declare no conflict of interest.

Data availability statement

Data supporting these findings are available within the article, at https://doi.org/10.20935/AcadBiol6177, or upon request.

Institutional review board statement

Not applicable.

Informed consent statement

Not applicable.

Sample availability

The authors declare no physical samples were used in the study.

Publisher’s note

Academia.edu stays neutral with regard to jurisdictional claims in published maps and institutional affiliations. All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

1

El-Sayed MI. Effects of Portulaca oleracea L. seeds in treatment of type-2 diabetes mellitus patients as adjunctive and alternative therapy. J Ethnopharmacol. 2011;137(1):643–51.

2

Kwon DY, Daily III JW, Kim HJ, Park S. Antidiabetic effects of fermented soybean products on type 2 diabetes. Nutr Res. 2010;30(1):1–3.

3

Babel RA, Dandekar MP. A review on cellular and molecular mechanisms linked to the development of diabetes complications. Curr Diabetes Rev. 2021;17(4):457–73.

4

Tuli HS, et al. Molecular mechanisms of action of genistein in cancer: recent advances. Front Pharmacol. 2019;10:1336.

5

Yu G, et al. Daidzein improved glucose homeostasis via PI3K/AKT and modulated the communities of gut microbiota in turbot (Scophthalmus maximus L.). Aquaculture. 2023;577:739949.

6

Lee SH, Park SY, Choi CS. Insulin resistance: from mechanisms to therapeutic strategies. Diabetes Metab J. 2022;46(1):15–37.

7

Peng Q, et al. Effects of genistein on common kidney diseases. Nutrients. 2022;14(18):3768.

8

Abbasi E, Khodadadi I. Antidiabetic effects of genistein: mechanism of action. Endocr Metab Immune Disord Drug Targets. 2023;23(13):1599–610.

9

Kim MS, et al. Antioxidant capacity of 12 major soybean isoflavones and their bioavailability under simulated digestion and in human intestinal Caco-2 cells. Food Chem. 2022;374:131493.

10

Arjunan S, Thangaiyan R, Balaraman D. Biochanin A, a soy isoflavone, diminishes insulin resistance by modulating insulin-signalling pathway in high-fat diet-induced diabetic mice: biochanin A diminishes insulin resistance in diabetic mice. Arch Physiol Biochem. 2023;129(2):316–22.

11

Wang GG, Lu XH, Li W, Zhao X, Zhang C. Protective effects of luteolin on diabetic nephropathy in STZ-induced diabetic rats. Evid Based Complement Alternat Med. 2011;2011:323171.

12

Suba V, Murugesan T, Arunachalam G, Mandal SC, Saha BP. Anti-diabetic potential of Barleria lupulina extract in rats. Phytomedicine. 2004;11(2–3):202–5.

13

Chauhan S, Gupta S, Yasmin S, Saini M. Antihyperglycemic and antioxidant potential of plant extract of Litchi chinensis and Glycine max. Int J Nutr Pharmacol. Neurol Dis. 2021;11(3):225–33.

14

Villegas R, et al. Legume and soy food intake and the incidence of type 2 diabetes in the Shanghai Women’s Health Study. Am J Clin Nutr. 2008;87(1):162–7.

15

de Paulo Farias D, de Araujo FF, Neri-Numa IA, Pastore GM. Antidiabetic potential of dietary polyphenols: a mechanistic review. Food Res Int. 2021;145:110383.

16

Chatterjee C, Gleddie S, Xiao CW. Soybean bioactive peptides and their functional properties. Nutrients. 2018;10(9):1211.

17

Basson AR, Ahmed S, Almutairi R, Seo B, Cominelli F. Regulation of intestinal inflammation by soybean and soy-derived compounds. Foods. 2021;10(4):774.

18

Liu K. Food use of whole soybeans. In: Johnson LA, White PJ, Galloway R, editors. Soybeans. Urbana (IL): AOCS Press; 2008. p. 441–81.

19

Zhao G, et al. Extracellular proteome analysis and flavor formation during soy sauce fermentation. Front Microbiol. 2018;9:1872.

20

Undhad Trupti J, Das S, Solanki D, Kinariwala D, Hati S. Bioactivities and ACE-inhibitory peptides releasing potential of lactic acid bacteria in fermented soy milk. Food Prod Process Nutr. 2021;3:1–14.

21

do Nascimento FV, de Castro AM, Secchi AR, Coelho MAZ. Insights into media supplementation in solid-state fermentation of soybean hulls by Yarrowia lipolytica: impact on lipase production in tray and insulated packed-bed bioreactors. Biochem Eng J. 2021;166:107866.

22

Schulman P, et al. A microbial fermentation product induces defense-related transcriptional changes and the accumulation of phenolic compounds in Glycine max. Phytopathology. 2022;112(4):862–71.

23

Nkhata SG, et al. Fermentation and germination improve nutritional value of cereals and legumes through activation of endogenous enzymes. Food Sci Nutr. 2018;6(8):2446–58.

24

do Prado FG, Pagnoncelli MGB, de Melo Pereira GV, Karp SG, Soccol CR. Fermented soy products and their potential health benefits: a review. Microorganisms. 2022;10(8):1606.

25

Choi SB, Jang JS, Park S. Estrogen and exercise may enhance β-cell function and mass via insulin receptor substrate 2 induction in ovariectomized diabetic rats. Endocrinology. 2005;146(11):4786–94.

26

Ademiluyi AO, Oboh G. Soybean phenolic-rich extracts inhibit key-enzymes linked to type 2 diabetes (α-amylase and α-glucosidase) and hypertension (angiotensin I converting enzyme) in vitro. Exp Toxicol Pathol. 2013;65(3):305–9.

27

Salehi B, et al. Antioxidants: positive or negative actors? Biomolecules. 2018;8(4):124.

28

Neha K, Haider MR, Pathak A, Yar MS. Medicinal prospects of antioxidants: a review. Eur J Med Chem. 2019;178:687–704.

29

Zhang P, et al. Oxidative stress and diabetes: antioxidative strategies. Front Med. 2020;14:583–600.

30

Ighodaro OM. Molecular pathways associated with oxidative stress in diabetes mellitus. Biomed Pharmacother. 2018;108:656–62.

31

Khosravi A, Razavi SH. Therapeutic effects of polyphenols in fermented soybean and black soybean products. J Funct Foods. 2021;81:104467.

32

Zhang X, et al. Potential protective effects of equol (soy isoflavone metabolite) on coronary heart diseases—from molecular mechanisms to studies in humans. Nutrients. 2021;13(11):3739.

33

Guan J, et al. Effects of dietary supplements on the space radiation-induced reduction in total antioxidant status in CBA mice. Radiat Res. 2006;165(4):373–8.

34

Krishnan HB, Wang TT. An effective and simple procedure to isolate abundant quantities of biologically active chemopreventive Lunasin Protease Inhibitor Concentrate (LPIC) from soybean. Food Chem. 2015;177:120–6.

35

Zhang Q, et al. Changes in antioxidant activity of Alcalase-hydrolyzed soybean hydrolysate under simulated gastrointestinal digestion and transepithelial transport. J Funct Foods. 2018;42:298–305.

36

Tonolo F, et al. Fermented soy-derived bioactive peptides selected by a molecular docking approach show antioxidant properties involving the Keap1/Nrf2 pathway. Antioxidants. 2020;9(12):1306.

37

Song H, et al. Structural properties and bioactivities of pectic polysaccharides isolated from soybean hulls. LWT. 2022;170:114079.

38

Jha P, Kumar V, Rani A, Kumar A. Genomic regions governing the biosynthesis of sucrose and raffinose family oligosaccharides in soybean. J Plant Biochem Biotechnol. 2022;31:637–47.

39

Choi J, Song I, Lee S, You M, Kwon J. Protective effects of phosphatidylcholine against hepatic and renal cell injury from advanced glycation end products. Medicina. 2022;58(11):1519.

40

Chen J, Tang G, Zhou J, Liu W, Bi Y. The characterization of soybean germ oil and the anti-oxidative activity of its phytosterols. RSC Adv. 2019;9(68):40109–17.

41

Matsukawa T, Inaguma T, Han J, Villareal MO, Isoda H. Cyanidin-3-glucoside derived from black soybeans ameliorate type 2 diabetes through the induction of differentiation of preadipocytes into smaller and insulin-sensitive adipocytes. J Nutr Biochem. 2015;26(8):860–7.

42

Das D, et al. Daidzein, its effects on impaired glucose and lipid metabolism and vascular inflammation associated with type 2 diabetes. Biofactors. 2018;44(5):407–17.

43

de Campos Zani SC, Wu J, Chan CB. Egg and soy-derived peptides and hydrolysates: a review of their physiological actions against diabetes and obesity. Nutrients. 2018;10(5):549.

44

Juárez-Chairez MF, Meza-Márquez OG, Márquez-Flores YK, Jiménez-Martínez C. Potential anti-inflammatory effects of legumes: a review. Br J Nutr. 2022;128(11):2158–69.

45

Kusmardi K, Nessa N, Estuningtyas A, Tedjo A. The effect of lunasin from Indonesian soybean extract on histopatologic examination and cox-2 expression in dextran sodium sulfate-induced mice colon. Int J Physiol Pathophysiol Pharmacol. 2018;10(6):154.

46

Guan X, et al. Structural characterization of a soluble polysaccharide SSPS1 from soy whey and its immunoregulatory activity in macrophages. Int J Biol Macromol. 202;217:131–41.

47

He H, Liu M, He R, Zhao W. Lipid-lowering activity of metformin-soluble soybean polysaccharide nanoparticles. Food Funct. 2022;13(19):10265–74.

48

Ishihara T, Yoshida M, Arita M. Omega-3 fatty acid-derived mediators that control inflammation and tissue homeostasis. Int Immunol. 2019;31(9):559–67.

49

Sugawara T. Sphingolipids as functional food components: benefits in skin improvement and disease prevention. J Agric Food Chem. 2022;70(31):9597–609.

50

Tan Y, Chang SK. Digestive enzyme inhibition activity of the phenolic substances in selected fruits, vegetables and tea as compared to black legumes. J Funct Foods. 2017;38:644–55.

51

Tiss M, et al. Fermented soy milk prepared using kefir grains prevents and ameliorates obesity, type 2 diabetes, hyperlipidemia and Liver-Kidney toxicities in HFFD-rats. J Funct Foods. 2020;67:103869.

52

González-Montoya M, Hernández-Ledesma B, Silván JM, Mora-Escobedo R, Martínez-Villaluenga C. Peptides derived from in vitro gastrointestinal digestion of germinated soybean proteins inhibit human colon cancer cells proliferation and inflammation. Food Chem. 2018;242:75–82.

53

Lin D, et al. Study on the functional properties and structural characteristics of soybean soluble polysaccharides by mixed bacteria fermentation and microwave treatment. Int J Biol Macromol. 2020;157:561–8.

54

Yang L, et al. Insoluble dietary fiber from soy hulls regulates the gut microbiota in vitro and increases the abundance of bifidobacteriales and lactobacillales. J Food Sci Technol. 2020;57:152–62.

55

Quan J, Yin X, Tanaka M, Kanazawa T. The hypoglycemic effects of soybean hypocotyl extract in diabetic rats and their mechanism. Acta Nutrimenta Sinica. 1956;0(03).

56

Yuk HJ, et al. The most abundant polyphenol of soy leaves, coumestrol, displays potent α-glucosidase inhibitory activity. Food Chem. 2011;126(3):1057–63.

57

Son HU, et al. Effects of synergistic inhibition on α-glucosidase by phytoalexins in soybeans. Biomolecules. 2019;9(12):828.

58

Gao Y, et al. α-Glucosidase inhibitory activity of fermented okara broth started with the strain Bacillus amyloliquefaciens SY07. Molecules. 2022;27(3):1127.

59

Li W, et al. Isolation and identification of an α-glucosidase inhibitory peptide from extruded soybean protein and its hypoglycemic activity in T2DM mice. Food Funct. 2023;14(9):4288–301.

60

Jiang M, Yan H, He R, Ma Y. Purification and a molecular docking study of α-glucosidase-inhibitory peptides from a soybean protein hydrolysate with ultrasonic pretreatment. Eur Food Res Technol. 2018;244:1995–2005.

61

Wang R, et al. Preparation of bioactive peptides with antidiabetic, antihypertensive, and antioxidant activities and identification of α-glucosidase inhibitory peptides from soy protein. Food Sci Nutr. 2019;7(5):1848–56.

62

Ha TJ, et al. α-Glucosidase inhibitory activity of isoflavones and saponins from soybean (Glycine max L.) and comparisons of their constituents during heat treatments. J Korean Soc Food Sci Nutr. 2019;48(9):953–60.

63

Nakamura A, Furuta H, Maeda H, Takao T, Nagamatsu Y. Structural studies by stepwise enzymatic degradation of the main backbone of soybean soluble polysaccharides consisting of galacturonan and rhamnogalacturonan. Biosci Biotechnol Biochem. 2002;66(6):1301–13.

64

Oliveira TT, Nagem TJ, Miranda LC, Paula VF, Teixeira MA. Inhibitory action on aldose reductase by soybean flavonoids. J Braz Chem Soc. 1997;8:211–3.

65

Kim YS, et al. Genistein inhibits aldose reductase activity and high glucose-induced TGF-β2 expression in human lens epithelial cells. Eur J Pharmacol. 2008;594(1–3):18–25.

66

Laddha AP, Kulkarni YA. Daidzein ameliorates diabetic retinopathy in experimental animals. Life Sci. 2021;265:118779.

67

Fujita T, Funako T, Hayashi H. 8-Hydroxydaidzein, an aldose reductase inhibitor from okara fermented with Aspergillus sp. HK-388. Biosci Biotechnol Biochem. 2004;68(7):1588–90.

68

Roh C. Microbial transformation of bioactive compounds and production of ortho-dihydroxyisoflavones and glycitein from natural fermented soybean paste. Biomolecules. 2014;4(4):1093–101.

69

Purnomo Y, Taufiq M, Wijaya AN, Hakim R. Molecular docking of soybean (Glycine max) seed and ginger (Zingiber officinale) rhizome components as anti-diabetic through inhibition of dipeptidyl peptidase 4 (DPP-4) and alpha-glucosidase enzymes. Trop J Nat Prod Res. 2021;5(10):1735–42.

70

González-Montoya M, Hernández-Ledesma B, Mora-Escobedo R, Martínez-Villaluenga C. Bioactive peptides from germinated soybean with anti-diabetic potential by inhibition of dipeptidyl peptidase-IV, α-amylase, and α-glucosidase enzymes. Int J Mol Sci. 2018;19(10):2883.

71

Samaranayaka A. Lentil: revival of poor man’s meat. In: Nadathur S, Wanasundara JPD, Scanlin L, editors. Sustainable protein sources. Academic Press; 2017. p. 185–96.

72

Lammi C, et al. Soybean-and lupin-derived peptides inhibit DPP-IV activity on in situ human intestinal Caco-2 cells and ex vivo human serum. Nutrients. 2018;10(8):1082.

73

Suh DH, et al. Comparison of metabolites variation and antiobesity effects of fermented versus nonfermented mixtures of Cudrania tricuspidata, Lonicera caerulea, and soybean according to fermentation in vitro and in vivo. PloS One. 2016;11(2):e0149022.

74

Gargouri Y, Julien R, Pieroni G, Verger R, Sarda L. Studies on the inhibition of pancreatic and microbial lipases by soybean proteins. J Lipid Res. 1984;25(11):1214–21.

75

Satouchi K, et al. Lipoxygenase-1 from soybean seed inhibiting the activity of pancreatic lipase. Biosci Biotechnol Biochem. 1998;62(8):1498–503.

76

Satouchi K, et al. A lipase-inhibiting protein from lipoxygenase-deficient soybean seeds. Biosci Biotechnol Biochem. 2002;66(10):2154–60.

77

Alnuaimi A, et al. A comparative analysis of anti-lipidemic potential of soybean (Glycine max) protein hydrolysates obtained from different ripening stages: identification, and molecular interaction mechanisms of novel bioactive peptides. Food Chem. 2023;402:134192.

78

Kashef RK, Hassan HM, Afify AS, Ghabbour SI, Saleh NT. Effect of soybean galactomannan on the activities of α-amylase, trypsin, lipase and starch digestion. J Appl Sci Res. 2008;4(12):1893–7.

79

Kim JH, Choi C, Benveniste EN, Kwon D. TRAIL induces MMP-9 expression via ERK activation in human astrocytoma cells. Biochem Biophys Res Commun. 2008;377(1):195–9.

80

Ramlal A, et al. In silico analysis of angiotensin-converting enzyme inhibitory compounds obtained from soybean [Glycine max (L.) Merr.]. Front Physiol. 2023;14:1172684.

81

Takahashi S, Yoshiya T, Yoshizawa-Kumagaye K, Sugiyama T. Nicotianamine is a novel angiotensin-converting enzyme 2 inhibitor in soybean. Biomed Res. 2015;36(3):219–24.

82

Mallikarjun Gouda KG, Gowda LR, Rao AA, Prakash V. Angiotensin I-converting enzyme inhibitory peptide derived from glycinin, the 11S globulin of soybean (Glycine max). J Agric Food Chem. 2006;54(13):4568–73.

83

Rho SJ, Lee JS, Chung YI, Kim YW, Lee HG. Purification and identification of an angiotensin I-converting enzyme inhibitory peptide from fermented soybean extract. Process Biochem. 2009;44(4):490–3.

84

Dkhar B, et al. Genistein represses PEPCK-C expression in an insulin-independent manner in HepG2 cells and in alloxan-induced diabetic mice. J Cell Biochem. 2018;119(2):1953–70.

85

Kuryłowicz A. The role of isoflavones in type 2 diabetes prevention and treatment—a narrative review. Int J Mol Sci. 2020;22(1):218.

86

Wiernsperger NF. Oxidative stress as a therapeutic target in diabetes: revisiting the controversy. Diabetes Metab. 2003;29(6):579–85.

87

Nishikawa T, et al. Normalizing mitochondrial superoxide production blocks three pathways of hyperglycaemic damage. Nature. 2000;404(6779):787–90.

88

Kennedy AL, Lyons TJ. Glycation, oxidation, and lipoxidation in the development of diabetic complications. Metabolism. 1997;46:14–21.

89

Yim MB, Yim HS, Lee C, Kang SO, Chock PB. Protein glycation: creation of catalytic sites for free radical generation. Ann N Y Acad Sci. 2001;928(1):48–53.

90

Schmidt AM, et al. Cellular receptors for advanced glycation end products. Implications for induction of oxidant stress and cellular dysfunction in the pathogenesis of vascular lesions. Arterioscler Thromb. 1994;14(10):1521–8.

91

Kawamura N, et al. Increased glycated Cu, Zn-superoxide dismutase levels in erythrocytes of patients with insulin-dependent diabetes mellitus. J Clin Endocrinol Metab. 1992;74(6):1352–4.

92

Morgan PE, Dean RT, Davies MJ. Inactivation of cellular enzymes by carbonyls and protein-bound glycation/glycoxidation products. Arch Biochemistry Biophys. 2002;403(2):259–69.

93

Lee AY, Chung SS. Contributions of polyol pathway to oxidative stress in diabetic cataract. FASEB J. 1999;13(1):23–30.

94

Obrosova IG, et al. Aldose reductase inhibitor fidarestat prevents retinal oxidative stress and vascular endothelial growth factor overexpression in streptozotocin-diabetic rats. Diabetes. 2003;52(3):864–71.

95

Drel VR, Pacher P, Stevens MJ, Obrosova IG. Aldose reductase inhibition counteracts nitrosative stress and poly (ADP-ribose) polymerase activation in diabetic rat kidney and high-glucose-exposed human mesangial cells. Free Radic Biol Med. 2006;40(8):1454–65.

96

Cheng HM, González RG. The effect of high glucose and oxidative stress on lens metabolism, aldose reductase, and senile cataractogenesis. Metabolism. 1986;35(4):10–4.

97

Morre DM, Lenaz GI, Morre DJ. Surface oxidase and oxidative stress propagation in aging. J Exp Biol. 2000;203(10):1513–21.

98

Hamada Y, Araki N, Horiuchi S, Hotta N. Role of polyol pathway in nonenzymatic glycation. Nephrol Dial Transplant. 1996;11(Suppl 5):95–8.

99

Hamada Y, et al. Rapid formation of advanced glycation end products by intermediate metabolites of glycolytic pathway and polyol pathway. Biochem Biophys Res Commun. 1996;228(2):539–43.

100

Matoba K, et al. Targeting redox imbalance as an approach for diabetic kidney disease. Biomedicines. 2020;8(2):40.

101

Lüersen K, et al. Soy Extract, rich in hydroxylated isoflavones, exhibits antidiabetic properties in vitro and in Drosophila melanogaster in vivo. Nutrients. 2023;15(6):1392.

102

Rizzo G. The antioxidant role of soy and soy foods in human health. Antioxidants. 2020;9(7):635.

103

Nawaz MA, et al. Soyisoflavone diversity in wild soybeans (Glycine soja Sieb. & Zucc.) from the main centres of diversity. Biochem Syst Ecol. 2018;77:16–21.

104

Chen Q, et al. Comparison of phenolic and flavonoid compound profiles and antioxidant and α-glucosidase inhibition properties of cultivated soybean (Glycine max) and wild soybean (Glycine soja). Plants. 2021;10(4):813.

105

Chitisankul WT, Shimada K, Tsukamoto C. Antioxidative capacity of soyfoods and soy active compounds. Polish J Food Nutr Sci. 2022;72(1):101–8.

106

Chen X, et al. Quantitative analyses for several nutrients and volatile components during fermentation of soybean by Bacillus subtilis natto. Food Chem. 2022;374:131725.

107

Lovabyta NS, Jayus J, Nugraha AS. Bioconversion of isoflavones glycoside to aglycone during edamame (Glycine max) soygurt production using Streptococcus thermophillus FNCC40, Lactobacillus delbrueckii FNCC41, and L. plantarum FNCC26. Biodiversitas J Biol Divers. 2020; 21(4):1358–64.

108

Hyeon H, et al. Metabolic profiling-based evaluation of the fermentative behavior of Aspergillus oryzae and Bacillus subtilis for soybean residues treated at different temperatures. Foods. 2020;9(2):117.

109

Li S, et al. Effect of solid-state fermentation with Lactobacillus casei on the nutritional value, isoflavones, phenolic acids and antioxidant activity of whole soybean flour. LWT. 2020;125:109264.

110

Ma H, et al. Microbial population succession and community diversity and its correlation with fermentation quality in soybean meal treated with Enterococcus faecalis during fermentation and aerobic exposure. Microorganisms. 2022;10(3):530.

111

Zhang Y, et al. Solid-state fermentation with Rhizopus oligosporus RT-3 enhanced the nutritional properties of soybeans. Frontiers in Nutrition. 2022;9:972860.

112

Jung SM, et al. A non-probiotic fermented soy product reduces total and LDL cholesterol: a randomized controlled crossover trial. Nutrients. 2021;13(2):535.

113

Hwang YH, et al. Efficacy and safety of Lactobacillus plantarum C29-fermented soybean (DW2009) in individuals with mild cognitive impairment: a 12-week, multi-center, randomized, double-blind, placebo-controlled clinical trial. Nutrients. 2019;11(2):305.

114

Yu C, Xiao JH. The Keap1-Nrf2 system: a mediator between oxidative stress and aging. Oxid Med Cell Longev. 2021;2021;1–16.

115

Rebollo-Hernanz M, Kusumah J, Bringe NA, Shen Y, de Mejia EG. Peptide release, radical scavenging capacity, and antioxidant responses in intestinal cells are determined by soybean variety and gastrointestinal digestion under simulated conditions. Food Chem. 2023;405:134929.

116

Hernández-Ledesma B, Hsieh CC, de Lumen BO. Antioxidant and anti-inflammatory properties of cancer preventive peptide lunasin in RAW 264.7 macrophages. Biochem Biophys Res Commun. 2009;390(3):803–8.

117

Gu L, et al. Lunasin attenuates oxidant-induced endothelial injury and inhibits atherosclerotic plaque progression in ApoE−/− mice by up-regulating heme oxygenase-1 via PI3K/Akt/Nrf2/ARE pathway. FASEB J. 2019;33(4):4836–50.

118

Chen B, Xu M. Natural antioxidants in foods. In: Melton L, Shahidi F, Varelis P, editors. Encyclopedia of food chemistry. Oxford: Academic Press; 2019. p. 180–8.

119

Han L, et al. Effect of extraction method on the chemical profiles and bioactivities of soybean hull polysaccharides. Food Sci Nutr. 2021;9(11):5928–38.

120

Bai Z, et al. Polysaccharides from small black soybean alleviating type 2 diabetes via modulation of gut microbiota and serum metabolism. Food Hydrocoll. 2023;141:108670.

121

Abdel-Latif HM, et al. The influence of raffinose on the growth performance, oxidative status, and immunity in Nile tilapia (Oreochromis niloticus). Aquac Rep. 2020;18:100457.

122

Tsai CY, Chen YH, Chien YW, Huang WH, Lin SH. Effect of soy saponin on the growth of human colon cancer cells. World J Gastroenterol. 2010;16(27):3371.

123

Calderon Guzman D, et al. Consumption of cooked common beans or saponins could reduce the risk of diabetic complications. Diabetes, Metab Syndr Obes. 2020:3481–6.

124

Rohm TV, Meier DT, Olefsky JM, Donath MY. Inflammation in obesity, diabetes, and related disorders. Immunity. 2022;55(1):31–55.

125

Molehin OR, Adefegha SA, Adeyanju AA. Role of oxidative stress in the pathophysiology of type 2 diabetes and cardiovascular diseases. In: Maurya PK, Dua K, editors. Role of oxidative stress in pathophysiology of diseases. Singapore: Springer Nature Singapore; 2020. p. 277–97.

126

Tsalamandris S, et al. The role of inflammation in diabetes: current concepts and future perspectives. Eur Cardiol Rev. 2019;14(1):50.

127

Kim IS, Yang WS, Kim CH. Beneficial effects of soybean-derived bioactive peptides. Int J Mol Sci. 2021;22(16):8570.

128

Ganesan K, Xu B. Anti-diabetic effects and mechanisms of dietary polysaccharides. Molecules. 2019;24(14):2556.

129

Chen W, Duizer L, Corredig M, Goff HD. Addition of soluble soybean polysaccharides to dairy products as a source of dietary fiber. J Food Sci. 2010;75(6):C478–84.

130

Natto ZS, Yaghmoor W, Alshaeri HK, Van Dyke TE. Omega-3 fatty acids effects on inflammatory biomarkers and lipid profiles among diabetic and cardiovascular disease patients: a systematic review and meta-analysis. Sci Rep. 2019;9(1):18867.

131

Deol P, et al. Omega-6 and omega-3 oxylipins are implicated in soybean oil-induced obesity in mice. Sci Rep. 2017;7(1):12488.

132

Pradhan SP, Ashok GC, Joshi P, Pandey BP. Elemental analysis and enzymes inhibitory potential of the soybean and soy products available in Nepal. J Agric Food Res. 2023;12:100574.

133

Kaur N, et al. Alpha-amylase as molecular target for treatment of diabetes mellitus: a comprehensive review. Chem Biol Drug Des. 2021;98(4):539–60.

134

Astawan M, et al. Comparison between the potential of tempe flour made from germinated and nongerminated soybeans in preventing diabetes mellitus. HAYATI J Biosci. 2020;27(1):16.

135

Antony P, Vijayan R. Bioactive peptides as potential nutraceuticals for diabetes therapy: a comprehensive review. Int J Mol Sci. 2021;22(16):9059.

136

Shi Y, et al. Structural Understanding of Peptide-Bound G Protein-Coupled Receptors: Peptide–Target Interactions. J Med Chem. 2023;66(2):1083–111.

137

Maeda H, Nakamura A. Soluble soybean polysaccharide. In: Phillips GO, Williams PA, editors. Handbook of hydrocolloids. Cambridge (MA): Woodhead Publishing; 2021. p. 463–80.

138

Chen FP, Ou SY, Chen Z, Tang CH. Soy soluble polysaccharide as a nanocarrier for curcumin. J Agric Food Chem. 2017;65(8):1707–14.

139

Bhattarai RR. Effect of food structure on macronutrients digestion and fermentation kinetics [PhD thesis]. The University of Queensland; 2018.

140

Richard JP. Protein flexibility and stiffness enable efficient enzymatic catalysis. J Am Chem Soc. 2019;141(8):3320–31.

141

Wang M, Chen J, Ye X, Liu D. In vitro inhibitory effects of Chinese bayberry (Myrica rubra Sieb. et Zucc.) leaves proanthocyanidins on pancreatic α-amylase and their interaction. Bioorg Chem. 2020;101:104029.

142

da Rocha L, Baptista AM, Campos SR. Approach to study pH-dependent protein association using constant-pH molecular dynamics: application to the dimerization of β-Lactoglobulin. J Chem Theory Comput. 2022;18(3):1982–2001.

143

Oleszek M, Oleszek W. Saponins in food. In: Xiao J, Sarker SD, Asakawa Y, editors. Handbook of dietary phytochemicals. Singapore: Springer Nature Singapore; 2020. p. 1–40.

144

Malabed R. Atomistic mechanism of cholesterol-dependent membrane activity of saponins with different glycosylation positions [doctoral thesis]. Osaka University; 2018.

145

Assefa ST, et al. Alpha glucosidase inhibitory activities of plants with focus on common vegetables. Plants (Basel). 2019;9(1):2.

146

Daub CD, Mabate B, Malgas S, Pletschke BI. Fucoidan from Ecklonia maxima is a powerful inhibitor of the diabetes-related enzyme, α-glucosidase. Int J Biol Macromol. 2020;151:412–20.

147

Ismail GA, Gheda SF, Abo-Shady AM, Abdel-Karim OH. In vitro potential activity of some seaweeds as antioxidants and inhibitors of diabetic enzymes. Food Sci Technol. 2020;40(3):681–91.

148

Attjioui M, et al. Kinetics and mechanism of α-glucosidase inhibition by edible brown algae in the management of type 2 diabetes. Proc Nutr Soc. 2020;79(OCE2):E633.

149

Ye GJ, et al. Design and synthesis of novel xanthone-triazole derivatives as potential antidiabetic agents: α-Glucosidase inhibition and glucose uptake promotion. Eur J Med Chem. 2019;177:362–73.

150

Quan J, Yin X, Jin M, Shen M. Study on the inhibition of alpha-glucosidase by soyasaponins. J Chin Med Mater. 2003;26(9):654–6.

151

Lv QQ, Cao JJ, Liu R, Chen HQ. Structural characterization, α-amylase and α-glucosidase inhibitory activities of polysaccharides from wheat bran. Food Chem. 2021;341:128218.

152

Gu S, et al. Structural characterization and inhibitions on α-glucosidase and α-amylase of alkali-extracted water-soluble polysaccharide from Annona squamosa residue. Int J Biol Macromol. 2021;166:730–40.

153

Kasipandi M, et al. Effects of in vitro simulated gastrointestinal digestion on the antioxidant, α-glucosidase and α-amylase inhibitory activities of water-soluble polysaccharides from Opilia amentacea roxb fruit. LWT. 2019;111:774–81.

154

Chung SS, Chung SK. Aldose reductase in diabetic microvascular complications. Curr Drug Targets. 2005;6(4):475–86.

155

Quattrini L, La Motta C. Aldose reductase inhibitors: 2013-present. Expert Opin Ther Pat. 2019;29(3):199–213.

156

Choi SW, et al. Isolation and structural determination of aldose reductase inhibitor from Korean fermented soybean paste. Food Sci Biotechnol. 2005;14(3):344–9.

157

Röhrborn D, Wronkowitz N, Eckel J. DPP4 in diabetes. Front Immunol. 2015;6:386.

158

Nargis T, Chakrabarti P. Significance of circulatory DPP4 activity in metabolic diseases. IUBMB Life. 2018;70(2):112–9.

159

Gupta S, Sen U. More than just an enzyme: dipeptidyl peptidase-4 (DPP-4) and its association with diabetic kidney remodelling. Pharmacol Res. 2019;147:104391.

160

Chhabria S, et al. A review on phytochemical and pharmacological facets of tropical ethnomedicinal plants as reformed DPP-IV inhibitors to regulate incretin activity. Front Endocrinol. 2022;13:1027237.

161

Nongonierma AB, FitzGerald RJ. Prospects for the management of type 2 diabetes using food protein-derived peptides with dipeptidyl peptidase IV (DPP-IV) inhibitory activity. Curr Opin Food Sci. 2016;8:19–24.

162

Rivero-Pino F, Espejo-Carpio FJ, Guadix EM. Identification of dipeptidyl peptidase IV (DPP-IV) inhibitory peptides from vegetable protein sources. Food Chem. 2021;354:129473.

163

Lammi C, Arnoldi A, Aiello G. Soybean peptides exert multifunctional bioactivity modulating 3-hydroxy-3-methylglutaryl-coa reductase and dipeptidyl peptidase-IV targets in vitro. J Agric Food Chem. 2019;67(17):4824–30.

164

Huang AH, Wang SM. Proteinaceous inhibitors of lipase activities in soybean and other oil seeds. In: Linskens HF, Jackson JF, editors. Seed analysis. Berlin, Heidelberg: Springer Berlin Heidelberg; 1992. p. 263–71.

165

Lee SS, Mohd Esa N, Loh SP. In vitro inhibitory activity of selected legumes against pancreatic lipase. J Food Biochem. 2015;39(4):485–90.

166

Zhang WL, Zhu L, Jiang JG. Active ingredients from natural botanicals in the treatment of obesity. Obes Rev. 2014;15(12):957–67.

167

Lee JH, et al. Changes in conjugated linoleic acid and isoflavone contents from fermented soymilks using Lactobacillus plantarum P1201 and screening for their digestive enzyme inhibition and antioxidant properties. J Funct Foods. 2018;43:17–28.

168

Hwang CE, et al. Enhanced digestive enzyme activity and anti-adipogenic of fermented soy-powder milk with probiotic Lactobacillus plantarum P1201 through an increase in conjugated linoleic acid and isoflavone aglycone content. Korean J Food Preserv. 2018;25(4):461–70.

169

Hwang CE, et al. Enhancement of isoflavone aglycone, amino acid, and CLA contents in fermented soybean yogurts using different strains: Screening of antioxidant and digestive enzyme inhibition properties. Food Chem. 2021;340:128199.

170

Lee HY, et al. Changes of γ-aminobutyric acid, phytoestrogens, and biofunctional properties of the isoflavone-enriched soybean (Glycine max) leaves during solid lactic acid fermentation. Fermentation. 2022;8(10):525.

171

Islam A, Islam MS, Uddin MN, Hasan MM, Akanda MR. The potential health benefits of the isoflavone glycoside genistin. Arch Pharm Res. 2020;43:395–408.

172

Wang SM, Huang AH. Inhibitors of lipase activities in soybean and other oil seeds. Plant Physiol. 1984;76(4):929–34.

173

Mori T, Satouchi K, Matsushita S. A protein inhibiting pancreatic lipase activity in soybean seeds. Agric Biol Chem. 1973;37(5):1225–6.

174

Satouchi K, Mori T, Matsushita S. Characterization of inhibitor protein for lipase in soybean seeds. Agric Biol Chem. 1974;38(1):97–101.

175

Farias TC, et al. Bioactive properties of peptide fractions from Brazilian soy protein hydrolysates: In silico evaluation and experimental evidence. Food Hydrocoll Health. 2023;3:100112.

176

Subcommittee on the Tenth Edition of the RDAs, Food and Nutrition Board, Commission on Life Sciences, National Research Council. Recommended dietary allowances. 10th ed. Washington (DC): National Academy Press; 1989.

177

Zou X, et al. Properties of plant-derived soluble dietary fibers for fiber-enriched foods: a comparative evaluation. Int J Biol Macromol. 2022;223:1196–207.

178

Velasquez MT, Bhathena SJ, Striffler JS, Thibault N, Scalbert E. Role of angiotensin-converting enzyme inhibition in glucose metabolism and renal injury in diabetes. Metabolism. 1998;47:7–11.

179

Üstündağ B, Canatan H, Çinkilinç N, Halifeoğlu İ, Bahçecioğlu İH. Angiotensin converting enzyme (ACE) activity levels in insulin-independent diabetes mellitus and effect of ACE levels on diabetic patients with nephropathy. Cell Biochem Funct. 2000;18(1):23–8.

180

McFarlane SI, Kumar A, Sowers JR. Mechanisms by which angiotensin-converting enzyme inhibitors prevent diabetes and cardiovascular disease. Am J Cardiol. 2003;91(12):30–7.

181

Brenner BM, et al. Effects of losartan on renal and cardiovascular outcomes in patients with type 2 diabetes and nephropathy. N Engl J Med. 2001;345(12):861–9.

182

Antonetti DA, Klein R, Gardner TW. Mechanisms of disease diabetic retinopathy. N Engl J Med. 2012;366(13):1227–39.

183

McMurray JJ, et al. Dapagliflozin in patients with heart failure and reduced ejection fraction. N Engl J Med. 2019;381(21):1995–2008.

184

Chowdhury TA, Lasker SS. Complications and cardiovascular risk factors in South Asians and Europeans with early-onset type 2 diabetes. QJM. 2002;95(4):241–6.

185

Liu Y, et al. Angiotensin-converting enzyme is involved in glucose metabolism through the renin-angiotensin system in diabetes. Oxid Med Cell Longev. 2017;2017:5683961.

186

Azadbakht L, Esmaillzadeh A. Beneficial effects of a dietary approaches to stop hypertension eating plan on features of the metabolic syndrome: response to Esposito and Giugliano. Diabetes Care. 2006;29(4):954–5.

187

Farzamirad V, Aluko RE. Angiotensin-converting enzyme inhibition and free-radical scavenging properties of cationic peptides derived from soybean protein hydrolysates. Int J Food Sci Nutr. 2008;59(5):428–37.

188

Yeh WJ, Yang HY, Chen JR. Soy β-conglycinin retards progression of diabetic nephropathy via modulating the insulin sensitivity and angiotensin-converting enzyme activity in rats fed with high salt diet. Food Funct. 2014;5(11):2898–904.

189

Dellafiora L, et al. “Bottom-up” strategy for the identification of novel soybean peptides with angiotensin-converting enzyme inhibitory activity. J Agric Food Chem. 2020;68(7):2082–90.

190

Yu S, Meng S, Xiang M, Ma H. Phosphoenolpyruvate carboxykinase in cell metabolism: roles and mechanisms beyond gluconeogenesis. Mol Metab. 2021;53:101257.

191

Choi MS, Jung UJ, Yeo JK, Kim MJ, Lee MK. Genistein and daidzein prevent diabetes onset by elevating insulin level and altering hepatic gluconeogenic and lipogenic enzyme activities in non-obese diabetic (NOD) mice. Diabetes Metab Res Rev. 2008;24(1):74–81.

192

Zuo X, Zhao R, Wu M, Wan Q, Li T. Soy consumption and the risk of type 2 diabetes and cardiovascular diseases: a systematic review and meta-analysis. Nutrients. 2023;15(6):1358.