Next Article in Journal
Development of a Modified Global Physical Activity Questionnaire and Its Construct Validity among Adults in Kerala, India
Previous Article in Journal
Nutritional Modulation of Dietary Sugars as a Strategy to Improve Insulin Resistance and Energy Balance in Diabetes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Diabetes Mellitus Management: An Extensive Review of 37 Medicinal Plants

by
Khwaja Zohura Zanzabil
1,
Md. Sabbir Hossain
2 and
Md. Kamrul Hasan
2,3,*
1
Biotechnology Program, Department of Mathematics and Natural Sciences, School of Data & Sciences, BRAC University, Dhaka 1212, Bangladesh
2
Department of Biochemistry and Molecular Biology, Tejgaon College, National University, Gazipur 1700, Bangladesh
3
Department of Public Health, North South University, Dhaka 1229, Bangladesh
*
Author to whom correspondence should be addressed.
Diabetology 2023, 4(2), 186-234; https://doi.org/10.3390/diabetology4020019
Submission received: 25 October 2022 / Revised: 8 June 2023 / Accepted: 9 June 2023 / Published: 12 June 2023

Abstract

:
Plants have been used as sources of medicine since ancient times. Natural products have been used extensively in Chinese, ayurvedic and folk medicine. In addition, a significant portion of the world’s population still utilizes herbal medicine. Diabetes is a common ailment affecting almost 463 million people in the world. However, current medications exert harmful after-effects on patients, while herbal medicines have fewer adverse effects. Plants possess secondary metabolites, such as alkaloids, flavonoids, tannins, steroids, etc., which exert numerous beneficial effects on health. Extensive research has been conducted over the years investigating and proving the hypoglycemic potential of various plants. The present paper reviews 37 such plants that are rich in phytoconstituents that possess a variety of pharmacological activities and have been experimentally proven to possess potentially hypoglycemic properties in animal models: Ficus racemosa, Agremone mexicana, Bombax ceiba, Cajanus cajan, Coccinia cordifolia, Momordica charantia, Syzygium cumini, Neolamarckia cadamba, Mangifera indica, Cocos nucifera, Tamarindus indica, Punica granatum, Azadirachta indica, Costus speciosus, Moringa oleifera, Andrographis paniculata, Ficus benghalensis, Anacardium occidentale, Annona squamosa, Boerhaavia diffusa, Catharanthus roseus, Cocculus hirsutus, Ficus hispida, Terminalia chebula, Terminalia catappa, Amaranthus tricolor, Blumea lacera, Piper betle leaves, Achyranthes aspera, Kalanchoe pinnata, Nelumbo nucifera, Mikania cordata, Wedelia chinensis, Murraya koenigii, Aloe barbadensis, Bryophyllum pinnatum and Asparagus racemosus. These 37 plant extracts exhibit antidiabetic activities through different mechanisms, including α-amylase and α-glucosidase inhibition, increases in glucose uptake and the stimulation of insulin secretion.

Graphical Abstract

1. Introduction

Diabetes mellitus, commonly known as diabetes, is a group of metabolic disorders characterized by hyperglycemia due to the faulty production, action or secretion of insulin [1]. Although no successful cure has been found for DM, it can be managed using insulin, diet modification and conventional or synthetic medicines. Despite their efficacy, synthetic drugs often present adverse side effects to patients and are relatively expensive and difficult to obtain. Therefore, there has been increasing research regarding traditional medicinal plants to search for alternative hypoglycemic drugs. Many synthetic medicines are derived from compounds originally found in plants, such as miglitol, an iminosugar derivative. Iminosugars and sugar derivatives are promising antidiabetic agents that regulate blood sugar levels by inhibiting specific enzymes involved in glucose metabolism [2]. For instance, N-hydroxyethyl-1-deoxynojirimycin (Miglitol), a derivative of 1-deoxynojirimycin iminosugar, is already used against type 2 diabetes as an α-glucosidase inhibitor [3].
Chronic hyperglycemia resulting from diabetes is associated with long-term damage to and the dysfunction of multiple organs, as well as complications such as retinopathy, nephropathy, neuropathy, peripheral vascular diseases, coronary heart disease, etc. [4]. The most common types of diabetes are type 1, type 2 and gestational diabetes. Type 1 diabetes is an autoimmune disease, wherein beta cells in the pancreas are destroyed by the immune system, resulting in absolute insulin deficiency, while Type 2 diabetes results from insulin resistance and the progressive loss of insulin secretion. In contrast, gestational diabetes occurs during pregnancy [5]. Type 2 DM is the most common health burden, particularly in the elderly population, affecting approximately 25% of people over the age of 65 years as of 2021, with the number of affected people expected to increase drastically over the next decades owing to increased life expectancy [6].
Plants produce secondary metabolites, such as alkaloids, flavonoids, tannins, terpenoids, ferulic acid, etc., which have proven to possess hypoglycemic activity. Alkaloids inhibit alpha-glucosidase and decrease glucose transport through the intestinal epithelium, lowering blood glucose. Flavonoids suppress glucose levels and increase hepatic glucose activity, most likely by enhancing insulin release from pancreatic islets. Saponins, triterpenoids and steroidal glycosides stimulate insulin release and block the formation of glucose in the bloodstream. Polysaccharides increase serum insulin levels, improve glucose tolerance and reduce blood glucose levels. Compounds such as ferulic acid stimulate insulin secretion [7].
Additionally, herbal medicine is cheaper and safer, with fewer adverse effects, than synthetic medicine. Furthermore, in herbal medicine, the whole plant or parts of the plant are treated instead of a single isolated active ingredient, as in conventional medicine. The use of the whole plant is believed to produce a better synergistic effect due to all the chemicals in the plant working together to produce a combined and more beneficial effect. Herbalists also claim that the synergistic effect reduces the toxicity of plants due to buffering [7,8]. Plants have significant potential as alternative hypoglycemic medicines due to their safety, efficacy, affordability and availability, as well as the rich array of potential hypoglycemic phytocompounds. This article attempts to give a brief overview of some plants that contain hypoglycemic phytoconstituents and have exhibited anti-diabetic activity in animal models.

2. Retrieval of Published Research

An exhaustive scientific-publication search was conducted for this review article covering the years up to 2022 on different databases, including PubMed, Scopus, Google Scholar, Science Direct, Web of Science and ResearchGate. During the search, no time limits or filters were applied to the publications, but recent publications relevant to this review’s scope were given priority during screening. The purpose of the initial search was to determine the medicinal plants with antidiabetic properties using different combinations of the following keywords: “medicinal plant”, “plant”, “herbal medicine”, “diabetes”, “hyperglycemia”, “antidiabetic”, “antihyperglycemic”, “hypoglycemic”, and “insulin”. In total, 37 plants with antidiabetic properties were finalized as subjects of this review. Further, 37 plants were searched individually for their antidiabetic activity, mechanism, relevant bioactive compounds and toxicity, using keywords such as “diabetes”, “hyperglycemia”, “antidiabetic”, “antihyperglycemic”, “hypoglycemic”, “insulin sensitivity”, “insulin secretion”, “mechanism”, “pathway”, “pancreatic beta cells”, “insulin mimetics”, “glucosidase”, “amylase”, “phytochemistry”, “bioactive compound”, “toxic”, “toxicity” and “cytotoxic” in different combination with the individual scientific names of 37 plants. Boolean operators such as AND, OR and NOT were used for different combinational searches. The full scientific names of 37 plants were used as keywords. A manual search was also conducted for additional studies in the reference lists of relevant articles and reviews. In total, 239 articles were identified, which were further screened based on their relevance to the topic of interest; 198 articles were considered relevant to the scope of this review. Only publications about antidiabetic activity, bioactive compounds and toxicity of the selected 37 plants with full texts available in English are considered in this review.

3. Plants in the Management of Diabetes Mellitus

3.1. Achyranthes aspera

Achyranthes aspera (Figure 1; all figures are taken from authors from different areas of Bangladesh) is an Amaranthaceae-family plant commonly known as Devil’s horsewhip. The plant’s root is known to be used as a cure for jaundice [9]. The plant contains phytochemicals such as alkaloids, oleanolic acid, β-sitosterol, saponins, D-glucuronic acid, quercetin-3-O-β-D-galactoside dihydroxy ketones, aliphatic alcohol, benzoquinone, hydroquinone, asarone, eugenol, etc. It is reported to have hepatoprotective, laxative, antiasthmatic and anti-allergic properties [10]. In a study conducted by Vijayaraj et al. (2016), the administration of 300 mg/kg and 600 mg/kg body weight of the ethanolic seed extract of A. aspera in STZ-induced diabetic rats for 28 days significantly reduced blood glucose levels (p < 0.001) compared to the standard drug, glibenclamide (5 mg/kg body weight) [11]. Furthermore, A. aspera may have antidiabetic effects by either reducing glucose absorption from the gut or increasing glucose transport from the blood [12]. Sadashiv and Krishna (2011) investigated the acute toxicity of leaf powder and a methanolic extract from the leaves of A. aspera in Swiss albino mice according to OECD guidelines. The powder was administered to the rats in doses of 2, 4, 6 and 8 gm/kg body weight and the extract was administered in doses of 2, 4, 6 and 8 mg/kg body weight. The rats were observed for signs of toxicity or mortality for the first 24 h and the next 14 days. No signs of toxicity or mortality were observed, which confirmed that the leaf powder and methanolic extract were safe [13].

3.2. Aloe barbadensis

Aloe barbadensis (Figure 2), commonly known as Aloe, lily of the desert or burn plant, is a perennial succulent plant of the Asphodelaceae family and is used for a variety of medicinal purposes, including in the treatment of allergic reactions, wounds, burns, rheumatoid arthritis, rheumatic fever, diabetes, inflammation, etc. The plant has phytoactive compounds, such as flavonoids, terpenoids and polysaccharides, including pectins, hemicelluloses, glucomannan, sterols β-sitosterol, lupeol, tannins, etc., with analgesic, antioxidant, anti-cancer and anti-inflammatory properties [14]. Lanjhiyana et al. (2011) demonstrated that the administration of the methanolic extract of 300 mg/kg body weight of Aloe vera gel in alloxan-induced diabetic rats for 21 days significantly reduced blood glucose levels (p < 0.05) compared to the reference drug, glibenclamide (0.25 mg/kg body weight). The acute oral toxicity of the extract was examined according to the OECD guidelines. The extract was administered to overnight-fasted rats in single oral doses of 5, 50, 300 and 2000 mg/kg body weight and observed for signs of mortality or toxicity for the next 30 min and the subsequent 14 days. The extract conferred no mortality or toxicity upon the rats [15]. Chysalodin, an anthraquinone dimer isolated from A. barbadensis, is a potent α-glucosidase inhibitor and can thus be used as an antidiabetic agent [16]. Again, the mechanism of the antidiabetic effect of A. barbadensis involves dipeptidyl peptidase (DPP)-IV-enzyme inhibition in type 2 diabetes. This DPP-IV inhibition may be caused by a dipyrrole derivative, 3,6-dioxo-3,3a,6,6a-tetrahydropyrrolo[3,4-c]pyrrole1,4-dicarboxamid, isolated from A. barbadensis leaves [17].

3.3. Amaranthus tricolor

Amaranthus tricolor (Figure 3) belongs to the family Amaranthaceae and is commonly known as edible amaranth. The plant is used to cure leprosy, bronchitis, piles, leucorrhea and constipation [18]. The plant is known to have anti-hyperglycemic, antihyperlipidemic, anti-diabetic and antioxidant properties. It contains flavonoids, amino acids, alkaloids, carbohydrates, saponins, phenolic compounds and tannins [19]. The plant also contains phytic acid, with diabetes-related enzyme-inhibition properties [20]. According to a study by Clemente et al. (2011), 200 mg/kg and 400 mg/kg body weight of the aqueous leaf extract of the plant showed reduced serum glucose levels in a dose-dependent manner in diabetic rats [21]. Further studies are needed to determine the exact molecular mechanism of the antidiabetic effect. Aneja et al. (2013) conducted research in which they administered a single dose of 2000 mg/kg body weight of aqueous extract of A. tricolor roots to Wistar albino rats after fasting overnight. The animals were closely observed for signs of toxicity or mortality during the first 30 min post-treatment, periodically for the next 24 h and once daily for the next 14 days. No lethality was conferred by the extract at 2000 mg/kg body weight [19].

3.4. Anacardium occidentale

Locally known as kaju badam and commonly known as cashew nut, Anacardium occidentale (Figure 4) is a plant in the sumac family. The bark of this plant has been reported to have hypoglycemic and antihypertensive properties and was thus used in African communities to treat diabetes mellitus. The herb leaves’ consumption has also been associated with improvements in diabetes-associated kidney problems by reducing the amount of mucopolysaccharide [22]. The plant’s stem, leaves and bark are rich in phenolics, saponin, flavonoids, vitamins and selenium. The A. occidentale plant contains alkaloids, anthocyanidins, tannins, essential oils, glycosides, myricetin, pentoside, lactone, quercetin hexoside, xanthones, chalcones, 4-hydroxy dodecanoic acid, palmitate, sitosterol, stigmasterol, 3-O-βD-galactopyranoside, etc., which are responsible for its antibacterial, antimutagenic and antifungal activities [23]. According to a study conducted by Ukwenya et al. (2012), the administration of 300 mg/kg body weight of methanolic extract of A. occidentale leaves to STZ-induced Wistar rats significantly reduced blood glucose (p < 0.05), with 1 I U/kg body weight of insulin also significantly reducing the blood glucose (p < 0.05) [24]. Anacardium occidentale seeds showed potent antidiabetic efficacy. Seed extracts enhance glucose uptake in skeletal muscle cells via AMPK activation, increases in Glut4 synthesis and translocation to the plasma membrane. Additionally, they may increase glycolysis by impairing mitochondrial oxidative phosphorylation [25]. Konan et al. (2007) administered a hydroethanolic extract of A. occidentale leaves to rats with a dose limit of 2000 mg/kg. They monitored the rats closely for the next 3 h and, subsequently, for 14 days, to observe any signs of mortality or toxicity. The animals were sacrificed after 14 days. There was no mortality or apparent changes in the animals’ vital organs or behavioral patterns, even at the highest dose of 2000 mg/kg body weight [26].

3.5. Andrographis paniculata

Commonly known as green chiretta, Andrographis paniculata (Figure 5) is a plant in the Acanthaceae family, native to Southern Asia. The plant has been reported to have anti-inflammatory, antioxidant, anti-diabetic and anti-infective properties. It is deemed a natural antibiotic, containing compounds such as terpenoids, flavonoids, iridoids, ferulic acid and, in particular, diterpenoid lactones, such as andrographolide, 14-deoxyandrographolide and 14-Deoxy-11,12-dehydroandrographolide, which are ferulic acids [27]. Andrographolide has been reported to have anti-diabetic properties. In a study conducted by Akter et al. (2013), the administration of 1000 mg/kg of ethanolic extract of A. paniculata leaves in Alloxan-induced diabetic Wistar rats significantly reduced their elevated blood glucose (p < 0.01). To test the acute toxicity of the extract, it was administered to rats orally and intraperitoneally at 4000 mg/kg body weight; the animals were observed for the next 24 h and then for 10 days for signs of mortality. The extract conferred no toxicity [28]. A study by Augustine et al. (2014) suggested that A. paniculata lowers blood glucose levels by regulating glucose uptake and oxidation, restoring insulin-signaling molecules in the liver and reducing serum lipid levels [29].

3.6. Annona squamosa

Annona squamosa (Figure 6), also known as custard apple, is a tropical plant in the Annonaceae family, cultivated in India, Brazil, Egypt, the West Indies, Ecuador, South and Central America and Brazil. The leaves of the plant possess a wide variety of phytochemicals, such as flavonoids, alkaloids, phenols, saponins, tannins, glycosides, diterpenes, etc., which confer anti-diabetic, antioxidant, antimicrobial, antiviral and anti-tumor properties [30]. According to a study conducted by Rout et al. (2013), the administration of 200 mg/kg and 400 mg/kg body weight of hydroethanolic extract from A. squamosa leaves to STZ-induced hyperglycemic rats brought about a significant decrease (p < 0.05) in blood sugar compared to the standard drug, metformin (300 mg/kg body weight) (Table 1) [31]. Extracts from A. squamosa leaves show antidiabetic properties by enhancing insulin secretion or action, inhibiting starch digestion and protein glycation, delaying glucose absorption and suppressing DPP-IV enzyme activity [32]. In an acute-toxicity study by Onwusonye et al. (2014), a methanolic extract of A. squamosa was administered orally to mice up to a dose limit of 5000 mg/kg as a single dose for 24 h, after which the animals were monitored for 30 days. A sub-acute study was also carried out by orally administering the extracts at doses of 200, 400, 600, 800 and 1000 mg/kg body weight to each of the five groups of mice for 30 days. No mortality or pathological alterations were found, with the LD50 of the A. squamosa leaves greater than 5000 mg/kg body weight, indicating that the extract was well tolerated [33].

3.7. Argemone mexicana

Argemone mexicana (Figure 7), commonly known as Mexican prickly poppy and locally known as Sialkanta, is a herb in the Papaveraceae family. The plant’s roots, stem and latex are used as diuretics and in treating skin diseases, jaundice and diabetes [69]. The roots’ stems, curry, latex and juice are consumed for medicinal purposes [70]. The plant contains several isoquinoline alkaloids and phenolic compounds, such as berberine, cheilanthifoline, coptisine, cryptopine, sanguinarine, stylopine, coptisine, tetrahydroberberine, protopine, benzophenathridines, sanguinarine and dihydrosanguinarine, as well as palmitic acid, oleic acid, myristic acid, linoleic acids and β-sitosterol [71,72]. It also contains amino acids, fatty acids, tannins, saponins, flavonoids and phytosterols [73]. It has been reported to have analgesic, anti-diabetic, anthelmintic, antioxidant, anti-inflammatory, antimicrobial, antimutagenic and anti-cancer properties [74]. Although the roots and other parts of the plant are generally non-toxic, A. mexicana has been proven to be toxic and lethal [75] and the consumption of its seeds causes epidemic dropsy, as characterized by a case in Delhi, India [76]. The administration of argemone seeds in the diet of the roof rat (Rattus rattus L.) for 10 days leads to sedation, abdominal jerking, sluggishness, corneal opacity, hind leg edema, etc., with 14 of the 16 rats dying at the end of the experiment [75]. However, the root extract of the plant was found to be non-toxic, as the administration of root powder in the aqueous slurry in Swiss albino mice demonstrated no signs of toxicity on cage observation, even at the highest dose level (7.0 gm/kg body weight), thereby proving that it had no significant toxic effects on mice [77]. According to a study conducted by Nayak et al. (2011), the administration of 200 mg/kg and 400 mg/kg body weight of ethanolic and aqueous extracts of the aerial parts of Argemone mexicana had a significant anti-diabetic effect (p < 0.05) on alloxan-induced hyperglycemic rats [68]. Therefore, A. mexicana has significant therapeutic potential, but caution must be undertaken regarding the usage of its seeds.

3.8. Asparagus racemosus

Asparagus racemosus (Figure 8), commonly known as Satamuli, is a plant in the Liliaceae family. The roots and whole parts of the plants are used to treat diabetes, urinary diseases, jaundice and urinary disease. The juice of the plant’s roots is used to treat diabetes, jaundice and diarrhea [69,70]. The plant contains several bioactive phytoconstituents, which are responsible alone or in conjunction for its various pharmacological properties, such as its anti-diabetic, antioxidant, neuroprotective and anti-ulcer activities. The main phytoconstituents of the plant are saponins and flavonoids. Furthermore, A. racemosa has triterpenoids, steroidal saponins such as shatavarin IV (asparanin B), shatavarin V, shatavarin VI, shatavarin VII, shatavarin VIII, immunoside, schidigerasaponin D5 (asparanin A), racemoside A, racemoside B and racemoside C, etc., alkaloids such as asparagamine A, quercetin, quercetin glycosides such as quercetin-3-O-rutinoside and quercetin 3-O-galactoside, isoflavones and sapogenins [78]. The flavonoids and triterpenoids in the plant may be responsible for inhibiting α-amylase and α-glucosidase [79]. According to a study by Vadivelan (2012), the ethanolic extract of Asparagus racemosus significantly alleviated the blood glucose in Wistar rats with STZ-induced diabetes when administered at dosages of 200 mg/kg and 400 mg/kg body weight for 21 days [35]. The ethanolic extracts of A. racemosus demonstrate more potent hypoglycemic activity than other plant extracts. The plant has also decreased serum glucose levels in in vitro anti-diabetic studies [80]. Asparagus racemosus exhibits antihyperglycemic activity by inhibiting carbohydrate digestion and absorption, boosting insulin secretion and enhancing its action in peripheral tissues [81]. The intracellular Ca2+ levels are increased by A. racemosus root extracts, which stimulates the physiological pathways that induce insulin release. Furthermore, A. racemosus extracts stimulate insulin release in β-cells through a secretory mechanism involving an inhibitory effect on KATP channels [82]. In another study by Kumar et al. (2010), aqueous extracts of A. racemosus did not produce any fatalities in rats, even when used at high doses of 3200 mg/kg body weight [83]. In addition, A. racemosus was found to be non-toxic, even after long-term administration [84].

3.9. Azadirachta indica

Azadirachta indica (Figure 9), locally known as Neem, is a plant in the Meliaceae family [85]. The tree is primarily found in Southern Asia and Africa. The leaves, bark, oil, flowers, fruit and gum have been used to treat heart disease, cancer, diabetes and hypertension [86]. The A. indica plant consists of phenols, flavonoids, saponins, tannins, alkaloids, glycosides, carbohydrates, triterpenoids, β-sitosterol, ferulic acid, etc.; it has been reported to have antimicrobial, antioxidant, hepatoprotective and other properties [87,88]. The medicinal characteristics of the plant can be attributed to its DNA-repair, detoxification, cell-cycle-alteration, anti-inflammatory and anti-metastatic activities [89]. In an acute toxicity study conducted by Kanagasanthosh, Shanmugapriyan and Kavirajan (2015), ethanolic extracts of A. indica leaves were administered to Swiss albino mice at 20 mg/kg, 200 mg/kg and 2000 mg/kg body weight. The mice were observed for signs of mortality within the first 24 h and for other signs of toxicity for 14 subsequent days. No adverse effects were observed, even at 2000 mg/kg body weight of the ethanolic extract, indicating the safety of the usage of the extract for medicinal purposes [90]. A study by Bisht and Sisodia (2010) involved the administration of 200 mg/kg body weight of ethanolic extract of A. Indica to STZ (STZ)-induced diabetic rats. It was demonstrated that the extract normalized glucose levels after STZ-induced hyperglycemia (p < 0.05) [48]. Azadirachta indica leaf extracts exhibit antihyperglycemic activity by increasing insulin-receptor protein expression, insulin-receptor substrate-1 and its tyrosine phosphorylation (Tyr632), as well as the activity of AKT and GLUT4 proteins [91]. Meliacinolin, a tetranortriterpenoid extracted from A. indica, can inhibit α-glucosidase and α-amylase activities, improving insulin resistance [92].

3.10. Blumea lacera

Blumea lacera (Figure 10), commonly known as lettuce-leaf blumea, is a plant in the family Asteraceae [93]. The plant has significant medicinal value and has been used as a stimulant, astringent, anthelmintic drug and diuretic, as well as in the treatment of hemorrhoids [94]. The B. lacera plant consists of several phytochemicals, such as β-sitosterol, cineol, artemisinin, lupeol, β-caryophyllene and protocatechuic acid, as well as other alkaloids, flavonoids, tannins, terpenoids, flavones, triterpenes, etc. [94]. According to a study by Hassan et al. (2015), the oral administration of 50–400 mg/kg body weight of leaf methanolic extract to Swiss albino mice significantly reduced blood glucose compared to the reference drug, glibenclamide (10 mg/kg body weight) (p < 0.0001). The possible mechanisms include an increase in pancreatic insulin secretion [60]. The exact mechanism should be explored to consider this plant as a potent antidiabetic drug. Hossen et al. (2021) conducted a study on the acute oral toxicity of a methanolic extract of B. lacera leaves. The extract was administered to overnight-fasted rats in a single oral dose (500–5000 mg/kg body weight). The rats were kept under close observation for the first 30 min after the treatment and then for the next 72 h to check for any signs of toxicity, such as behavioral changes, allergenic symptoms or mortality. The extract did not demonstrate toxicity, even at the highest dose of 5000 mg/kg body weight [95].

3.11. Boerhaavia diffusa

Boerhaavia diffusa (Figure 11), also known as hogweed, is a species of flowering plant in the Nyctaginaceae family. Various parts of the plant, particularly the roots, have been associated with anti-diabetic, anti-cancer, gastroprotective and hepatoprotective properties. The plant contains steroids such as ecdysteroids, alkaloids, lignan glycosides, phenolic glycosides, flavonoids, isoflavonoids such as rotenoids, etc. [96]. According to a study conducted by Nalamolu (2003) et al., the chloroform extract of B. diffusa leaves reduced the blood glucose in STZ-induced diabetic rats in a dose-dependent manner when used in daily doses of 50 mg/kg, 100 mg/kg and 200 mg/kg body weight compared to the standard drug, glibenclamide. The blood glucose was probably reduced by the rejuvenation of pancreatic β-cells [54]. The ethanol extract of Boerhaavia diffusa exhibits therapeutic potential against type 2 diabetes by inhibiting small intestinal glucose absorption and stimulating muscle glucose uptake [97]. In an acute-toxicity study by Hiruma-Lima (2000), extracts obtained from lyophilized decoction and the juice of fresh leaves of B. diffusa were administered in doses of up to 5000 mg/kg body weight in mice after 12 h fasting. The animals were subsequently observed for 14 days. No signs of toxicity or mortality were observed until the dose of 5000 mg/kg body weight [98].

3.12. Bombax ceiba

Bombax ceiba (Figure 12), locally known as Shimul and commonly known as the red silk-cotton tree, is a plant in the Bombacaceae family. The bark and roots of the plants are used to treat excessive menstrual discharge, dysentery and diabetes [69]. The B. ceiba plant contains alkaloids, flavonoids, glycosides, β-sitosterol, kaempferol, hentriacontane, hentriacontanol, quercetin, shamimin (2-(2,4,5-trihydroxyphenyl)-3,5,7-trihydroxy-6-C-glucopyranosyloxy-4H-1-benzopyran-4-one), ceiba naphthoquinone, etc. [99]. In a study by Wanjari et al. (2016), the acute toxicity of the aqueous extract of B. ceiba flowers was investigated by administering the extract to healthy Wistar rats up to a dose of 2000 mg/kg body weight, according to the OECD 423 guidelines. The rats were observed for signs of toxicity for the first 2 h, 24 h, 72 h and 14 days. The extract did not confer toxicity or mortality up to a dose of 2000 mg/kg body weight [100]. According to a study conducted by Bhavsar and Talele (2013), the administration of 600 mg/kg of B. ceiba bark extract to STZ-induced diabetic rats demonstrated significant hypoglycemic activity (p < 0.001). This hypoglycemic activity may be attributed to the presence of triterpenoid compounds in the extract (Table 2) [36]. One possible mechanism of hypoglycemic activity is α-amylase inhibition, as ethyl-acetate extracts of Bombax ceiba leaves showed a 66.69 ± 0.95% α-amylase inhibition [101]. Another possible mechanism, suggested by Guang-Kai et al. (2017), is that the hypoglycemic effect may be due to an increase in insulin levels by protecting pancreatic β-cells and stimulating insulin secretion from the remaining β-cell [102]. A network-pharmacology analysis found several pathways, such as PPAR signaling, the activation of the oncogenic chemical receptor and fatty-acid metabolism, through which Bombax ceiba flowers help to treat type 2 diabetes mellitus [103]. The proposed possible mechanisms should be further explored to determine the exact mechanism of action of this plant’s antihyperglycemic effect.

3.13. Bryophyllum pinnatum

Commonly known as Goethe, Bryophyllum pinnatum (Figure 13) is a plant in the Crassulaceae family and grows in the temperate regions of Asia, the West Indies, Pacific, Hawaii, Australia, etc. The plant is rich in phytochemicals, such as flavonoids, alkaloids, reducing sugars, tannins, bufadienolides, glycosaponins, polyphenols and steroidal glycosides, which are responsible for its antioxidant, anti-tumor, anti-inflammatory, analgesic, anti-diabetic and hepatoprotective properties [136]. According to a study by Ezeagu et al. (2017), 200 mg/kg and 400 mg/kg body weight of ethanolic leaf extract from B. pinnatum significantly reduced the blood glucose levels in alloxan-induced diabetic rats compared to the standard drug (2.5 mg/kg body weight of glibenclamide) [67]. The serum-glucose-lowering mechanism involves α-amylase and α-glucosidase inhibition by kaempferol and quercetin isolated from B. pinnatum leaves [137]. In a study conducted by Afzal, Kazmi and Anwar (2013), the acute toxicity of ethanolic and aqueous extracts of B. pinnatum was conducted according to OECD guideline 420 and administered to Swiss albino mice orally in doses of 5 mg/kg, 50 mg/kg and 500 mg/kg body weight. The animals were observed for the first 2 h and the next 24 h for signs of toxicity or mortality. The extract was non-toxic, weighing up to 2000 mg/kg [138].

3.14. Cajanus cajan

Cajanus cajan (Figure 14), locally known as Arhar and commonly known as pigeon pea, is a shrub in the Fabaceae family. The juice from the plant’s roots is used to treat diabetes [69]. The plant contains over 40 bioactive compounds, with flavonoids and stilbenoids the chief bioactive constituents [139]. The plant contains β-sitosterol, flavonoids such as luteolin, vitexin, apigenin, genistic, ononin, sissotrin, 2′-hydroxygenistein, etc., as well as stilbenes such as cajanusin A, cajanusin B, cajanusin C, cajanusin D, cajanstilbene H, cajanolactone A, cajanonic acid A, canjanotone, etc. [105]. Tang et al. (2017) administered aqueous and ethanolic extracts of C. cajan in Sprague–Dawley rats in doses of 1.5, 3.0 and 6.0 g/kg body weight for four weeks and observed them for the two weeks after the treatment to check for symptoms of toxicity. No mortality, behavioral changes or weight alterations were observed. The study demonstrated the lack of acute toxicity from C. cajan in rats [139]. Extracts from C. cajan are used for the treatment of diabetes, dysentery, hepatitis, measles, wounds, bedsores, malaria, etc. According to a study conducted by Nahar et al. (2014), the administration of 200 mg/kg or 400 mg/kg of methanolic extract from the plant’s roots in alloxan-induced diabetic rats for two weeks significantly lowered the fasting serum glucose (p < 0.001) and blood glucose levels (p < 0.001) [37]. Increased insulin secretion may mediate the extract’s hypoglycemic effect [140]. However, there is still insufficient evidence of the glucose-lowering mechanism, so more studies are required to consider the significance of Cajanus cajan as a potent antidiabetic agent.

3.15. Catharanthus roseus Linn.

Locally known as Nayantara, Catharanthus roseus Linn. (Figure 15) belongs to the Apocynaceae family. The whole plant, is reported to have shown anti-cancer and anti-diabetic properties, especially the leaves [126]. The plant contains alkaloids, catharanthine, tetrahydroalstonine, vindoline, kaempferol, lochnerine and flavonoids, as well as pharmacological properties such as anti-microbial, antioxidant, anti-ulcer and wound-healing effects [124]. Research conducted by Nammi et al. (2003) demonstrated that 0.5 mL/kg, 0.75 mL/kg and 1.0 mL/kg body weight of leaf juice from C. roseus Linn. lowered the blood glucose in alloxan-induced diabetic rats, with 1.0 mL/kg body weight of glibenclamide used as a control. The maximum reduction in blood glucose was produced by 1.0 mL/kg body weight of C. roseus at 31.9% (p < 0.01) [55]. Several mechanisms of the antihyperglycemic effect have been reported [141]. The most significant mechanism is the enhancement of the glucose-transport gene (GLUT-2 and GLUT-4) expression, through which Catharanthus roseus seeds showed a potent antihyperglycemic effect [142]. In a study conducted by Vutukuri et al. (2017), the acute toxicity of an ethanolic extract from C. roseus leaves was assessed, according to the OECD 420 guidelines, in female albino Wistar rats. The extract was administered to groups of five rats in single oral doses of 5, 50, 300 and 2000 mg/kg body weight. The rats were kept under observation for the subsequent 14 days for signs of toxicity. No mortality was observed at the highest dose of 2000 mg/kg body weight, although slight tremors and restlessness were observed at the highest dose [143].

3.16. Coccinia cordifolia

Coccinia cordifolia (Figure 16), locally known as Telakucha and commonly known as Ivy gourd, is a plant in the Cucurbitaceae family [69]. The Cucurbitaceae family is a plant family consisting of 960 species and is considered to be one of the essential plant families, possessing potent hypoglycemic properties. Plants in this family were applied in the treatment of diabetes mellitus in both modern and ancient Ayurvedic medicine [144]. In a study conducted by Islam et al. (2011), 150 mg/kg body weight of ethanolic extract from Coccinia cordifolia leaves administered to STZ-induced diabetic rats for 24 h reduced blood glucose levels by 50.39% and 50% at the 10th and 24th hours, respectively (p < 0.001). Many of the hypoglycemic activities of the plant may be related to the phytochemicals saponin and glycoside, alkaloid compounds (catharanthin, leurosine, lochnerine, vindoline and vindolinine) and flavonoids [38]. The hypoglycemic activities of different parts of this plant are mediated through various mechanisms, including β-cell regeneration, insulin secretion stimulation, glucose-uptake enhancement, antioxidant-enzyme restoration, metabolic-enzyme regulation, digestive-enzyme inhibition and lipid-profile improvement [145]. A chemical analysis showed that the plant is rich in nutrients, mainly antioxidant compounds, such as total phenol, vitamin C and β-carotene. Phytochemical screening showed that the methanolic extract contains bioactive constituents, such as tannins, saponins, phenols, flavonoids and terpenoids [106]. In a study by Jha (2010), the acute oral toxicity of aqueous extracts from C. cordifolia was carried out according to the OPPTS (Office of Prevention, Pesticides and Toxic Substances) guidelines. The extracts were administered orally to albino Wistar rats at a dose of 2000 mg/kg. When no toxicity or mortality were observed, the dose was increased to 5000 mg/kg body weight, which also did not confer toxicity [146].

3.17. Cocculus hirsutus Linn.

Locally known as Daikhai or Jalajmani and commonly known as Broom creeper, Cocculus hirsutus (Figure 17) belongs to the family Menispermaceae. Its root and leaf extracts have been reported to show laxative and antiperiodic properties and alleviate symptoms of eczema and gonorrhea. The plant contains alkaloids, β-sitosterol, ginnol, flavonoids such as luteolin, kaempferol, quercetin, glycosides, carbohydrates, tannins, saponins and steroids, among others and possesses anti-diabetic, antibacterial, anti-cancer and antifungal activities [125]. A study conducted by Badole et al. (2006) demonstrated that the administration of 250 mg/kg, 500 mg/kg and 1000 mg/kg body weight of aqueous extract from C. hirsutus leaves in alloxan-induced diabetic mice for 28 days significantly reduced serum glucose levels (p < 0.01), with 10 mg/kg body weight p.o. of glyburide used as the standard drug [56]. The exact mechanism of the glucose-lowering effect of this plant extract is yet to be established. Bothara, Marya and Saluja (2011) performed an acute toxicity test for C. hirsutus methanolic and aqueous extracts in male Wistar albino rats, in which the LD50 of the extract was found to be 3000 mg/kg body weight [147].

3.18. Cocos nucifera

Cocos nucifera (Figure 18), commonly known as coconut, belongs to the Arecaceae family. The plant’s fruit has been used in the treatment of syphilis, jaundice, diabetes and cholera [69]. The plant has multiple phytoconstituents, such as phenols, tannins, β-sitosterol, flavonoids, nicotinic acid, folic acid, riboflavin, biotin, triterpenes, alkaloids, steroids, saponins, tannins, catechins, epicatechins, etc. [115]. Cocus nucifera can improve insulin secretion and blood glucose utilization, which is used in managing diabetes [148]. In a study conducted by Naskar et al. (2011), 250 mg/kg and 500 mg/kg of hydromethanol extract from Cocus nucifera was administered to STZ (STZ)-induced diabetic rats for 14 days (Table 1). The hydromethanolic extract from C. nucifera significantly lowered fasting blood glucose in the diabetic rats and had an effect comparable to 0.5 mg/kg glibenclamide (p < 0.0001) [44]. Salil et al. (2011) found that coconut-kernel protein (CKP) contains arginine in significant amounts and that it demonstrated potent anti-diabetic activity. As proposed by Salil et al. (2011), the mechanisms of anti-diabetic activity of CKP may include increasing glycogen and insulin levels while reducing serum glucose levels, the efficient degradation of carbohydrates by β-D-galactosidase and the inhibition of the regeneration of β-pancreatic cells by arginine, resulting in increased insulin secretion and subsequent increases in glycogen levels in blood serum [149]. Paul et al. (2012) administered petroleum ether, chloroform and methanol extracts from C. nucifera leaf extracts to Swiss albino mice to investigate the extracts’ acute toxicity. No mortality or toxic effects were observed at doses of up to 2000 mg/kg body weight in the acute toxicity studies, demonstrating that C. nucifera is non-toxic in adult albino mice [150].

3.19. Costus speciosus

Commonly known as Crepe Ginger, Costus speciosus (Figure 19) is a Costaceae-family plant with anti-diabetic, anti-microbial, antioxidant and anti-inflammatory activities [151]. The C. speciosus plant contains phytochemicals such as alkaloids, glycosides, flavonoids, steroids, polyphenols, tannins, etc. It also contains compounds such as β-sitosterol, gracillin, dioscin, diosgenin, etc. Diosgenin possesses anti-diabetic properties [118]. The rhizomes of C. speciosus are the primary sources of diosgenin, along with leaves. The plant’s leaves possess significant hypoglycemic properties and can lower blood glucose [152]. In a study conducted by Revathy, Abdullah and Kumar (2014), the administration of 200 mg/kg per body weight of ethanolic extract of C. speciosus enhanced insulin secretion in alloxan-induced diabetic rats [49]. Another study, conducted by Bavarva and Narasimhacharya (2008), demonstrated that doses of 300 mg/kg and 450 mg/kg body weight of ethanolic root extracts from C. speciosus lowered the concentration of blood glucose, decreased glyconeogenesis and increased glycogenesis, reinstating normal levels of glucose in the blood when administered to alloxan-induced diabetic rats [50]. An oligosaccharide called raffinose, which was isolated from the C. speciosus rhizome, increases glucose uptake in a dose-dependent manner. Through an increase in Glut4 translocation via IRβ/PI3K/Akt phosphorylation, its ability to increase insulin sensitivity was observed. Raffinose’s capacity to inhibit the activation of GSK3 for glycogen synthesis provides additional proof that it has therapeutic potential for treating diabetes [153]. In a study conducted by Mar (2019), 70% ethanolic extracts of C. speciosus rhizome were administered to albino mice, according to the OECD 423 guidelines. No mortality or toxicity were observed in the animal models, even with a maximum dose of 5000 mg/kg body weight over an observation period of 14 days, indicating the safe usage of the C. specious [57].

3.20. Ficus benghalensis L.

Ficus benghalensis (Figure 20), commonly known as Banyan, is a large evergreen semi-deciduous tree in the Moraceae family. All the parts of the plant contain latex, which is used in the treatment of toothache, dysentery and diarrhea. The tree’s bark is also used to treat diabetes [22]. The tree has been reported to possess anti-helminthic, anti-inflammatory, anti-diabetic, antimicrobial and analgesic properties. The phytochemicals derived from the plant include anthocyanidin derivatives, aliphatic long-chain ketones and glycosides [122]. It also contains phytoconstituents, such as flavonoids, amino acids, steroids, saponins, carbohydrates and tannins [123].
A study conducted by Saraswathi et al. (2013) demonstrated that the administration of 200 mg/kg and 400 mg/kg body weight of ethanolic extract from F. benghalensis leaves in alloxan-induced diabetic albino rats reduced blood glucose, cholesterol and triglyceride levels [52]. In another study, by Kasireddy et al. (2021), it was reported that an ethanolic extract of F. benghalensis lowered blood glucose in STZ-induced diabetic rats by stimulating insulin secretion from the Islets of Langerhans p < 0.01 [53]. The bark of F. benghalensis shows an anti-diabetic effect through a mechanism that includes the inhibition of the carbohydrate-hydrolyzing enzymes, α-amylase and α-glucosidase [154]. A network-pharmacology study suggested that F. benghalensis bark extract increases glucose uptake and insulin secretion through the PI3K/Akt signaling pathway, supporting other wet-laboratory research [155]. Gabhe, Tatke and Khan (2006) investigated the acute toxicity of the methanolic extract of F. benghalensis by orally administering a single dose of 2000 mg/kg body weight of the extract to rats and observing them for signs of toxicity for the next 14 days. The study’s results demonstrated that the LD50 of the methanolic extract of F. benghalensis was more than 2000 mg/kg body weight [123].

3.21. Ficus hispida Linn.

Commonly known as hairy fig, Ficus hispida (Figure 21) is a plant in the Moraceae family. The plant is rich in phytochemicals, such as flavonoids, saponins, steroids, glycosides, alkaloids, alkanes, etc. It has been reported for wound healing, anti-inflammatory, sedative, anti-ulcer and anti-diabetic activity, as well as being used to alleviate symptoms of dysentery, ulcers, dysentery, ulcers, psoriasis, anemia, piles, jaundice, etc. [126,127]. In a study by Ghosh et al. (2004), the ethanolic extract of F. hispida was administered in doses of 1.25 g/kg body weight to normal and alloxan-induced diabetic albino rats with 0.5 mg/kg body weight of glibenclamide as a standard drug (Table 1). The blood glucose was estimated by the glucose-oxidase method 2 h prior to and 2 h post-treatment. The F. hispida extract demonstrated significant hypoglycemic activity, as it significantly reduced blood glucose in both the normal (p < 0.01) and diabetic (p < 0.001) rats, but the effect was weaker than that of glibenclamide. The extract also increased the blood glucose uptake by the hemidiaphragm, increased the glycogen content in the liver and demonstrated other hypoglycemic activities [57]. In a study by Swathi, Sreedevi and Bharathi (2011), an acute-toxicity study of a methanolic extract from the fresh fruits of F. hispida was carried out following CPCSEA guideline 420. The extracts were administered to Wistar albino rats in doses of up to 1000 mg/kg body weight and conferred no mortality or toxicity [156].

3.22. Ficus racemosa

Ficus racemosa (Figure 22), commonly known as cluster fig, or locally as Dumur, is a plant in the Moraceae family. The plant is used to treat diabetes, dry cough, kidney and spleen diseases, bronchitis, menorrhagia and hemoptysis. The plant’s unripe fruit is often cooked in curries and consumed by diabetic patients, in addition to the use of its sap as a tonic [22]. According to a study by Mandal et al. (1997), 200 mg/kg and 400 mg/kg petroleum-ether extract from Ficus racemosa leaves exhibited hypoglycemic activity in Streptozotocin-induced diabetic rats when dispensed orally (Table 2) [34]. Isolated from F. racemosa, β-sitosterol is a compound with potential anti-diabetic activity. Other potential pharmacological products isolated from the plant include flavonoids, triterpenes such as gluanol acetate and racemosic acid, alkaloids and tannins. In a study conducted by Kushwaha et al. (2015), β-sitosterol, stigmasterol and lanosterol isolated from the petroleum-ether extract of F. racemosa leaves significantly reduced the blood glucose in STZ-induced diabetic rats with an action comparable to that of glibenclamide [104]. Ficus racemosa bark extracts exhibited potent antidiabetic effects through multiple mechanisms, such as inhibiting α-amylase, adsorbing and trapping glucose in its fiber matrix to prevent glucose diffusion and enhancing glucose transport across cell membranes [157]. In a study by Bhaskara Rao et al. (2002), the acute toxicity of the methanolic extract of F. racemosa bark was determined according to the methodology of Lorke (1983). The extracts did not confer any mortality or toxicity up to 3.2 g/kg body weight, indicating that the extract was non-toxic [158].

3.23. Kalanchoe pinnata

Kalanchoe pinnata (Figure 23), commonly known as cathedral bells, life plant, etc., belongs to the family Crassulaceae [159]. The plant contains phytochemicals such as flavonoids, kaempferol, quercetin, alkaloids, tannins, phenolic compounds, caffeic acid, syringic acid, luteolin, rutin, para-coumaric acid, ferulic acid, β-sitosterol, stigmasterol, astragalin, campesterol, etc. The plant’s extracts have been used in ethnomedicine to cure respiratory-tract infections, insect bites, fever, constipation, piles, etc. Furthermore, it is also known to have anti-cancer, antifungal and anti-diabetic activities [131]. In a study conducted by Indah (2016), an ethanolic extract of K. pinnata leaves was administered to alloxan-induced diabetic rats in doses of 5.8 mg/kg, 11.6 mg/kg and 33.2 mg/kg body weight, with glibenclamide (1.35 mg/kg body weight) and acarbose (13.5 mg/kg body weight) as standard drugs. After the treatment, the fasting blood glucose of the rats was recorded every day for 5 days; the animals were sacrificed at the end of the study for histopathological examination. The extract significantly lowered the fasting blood glucose (p < 0.05) and increased the number of pancreatic beta Langerhans cells, albeit not in a dose-dependent manner [62]. Kalanchoe pinnata has anti-oxidative, as well as α-amylase- and α-glucosidase-inhibitory activities, which enhance glucose uptake, thereby indicating its antidiabetic potential [160]. Quazi et al. (2011) investigated the acute toxicity of K. pinnata leaves by orally administering a single dose of 2000 mg/kg body weight of the ethanolic leaf extract of the plant to Wistar rats, which were kept under subsequent observation for signs of toxicity/mortality for the next 14 days. No signs of mortality, morbidity or histopathological changes were noted with the administration of the extract. The acute-toxicity study revealed that the extract was safe for use in doses greater than 2000 mg/kg of the rats’ body weight [131].

3.24. Mangifera indica

Mangifera indica (Figure 24), commonly known as mango and locally known as Aam, is a plant from the Anacardiaceae family. The plant has been used as an antidote to poison, as well as in the treatment of edema, dysentery and diabetes [85]. It is native to tropical Asia and commonly used in Ayurvedic medicine. The plant consists of Mangiferin, a polyphenolic antioxidant, as well as glucosyl xanthone, with hypotensive, antidegenerative and anti-diabetic properties [113]. The most abundant phytoconstituents of the plant are flavonoids, phenolic acids and xanthones. The major polyphenolic compounds found in the plant include gallic acid, catechins, kaempferol, ellagic acids, anthocyanins and ascorbic acid, carotenoids such as luteoxanthine, zeaxantine, β-carotene, terpenoids such as careen, myrcene, terpinoline, terpenoid and saponins such as indicoside A and B, etc. [114]. John et al. (2012) administered aqueous ethanolic extracts from M. indica stem bark orally to rats with doses of 1000, 1500, 3000, 4000 and 5000 mg/kg, respectively. They observed signs of toxicity over the next 48 h. The LD50 of the extract was determined at >5000 mg/kg body weight. The extract showed low toxicity at 5000 mg/kg body weight, as there were no clinical indications of toxicity or mortality after 48 h of administration (Table 3) [161]. In a study conducted by Madhuri and Mohanvelu (2017), 200 mg/kg and 400 mg/kg of aqueous extracts of Mangifera indica leaves were administered to alloxan-induced diabetic rats for 15 days. The experiment was conducted using the aqueous extract of M. indica alone, gliclazide (a hypoglycemic drug) and a combination of both. When used alone, both the M. indica extract and the gliclazide significantly lowered the fasting blood glucose of the rats (p < 0.0001). However, the most significant lowering of fasting blood glucose was observed when the extract and drug were used in conjunction [43]. This commercially available plant has been studied widely and several mechanisms have been proposed for its anti-diabetic effect. Insulin-stimulated processes, such as glucose transport and glycogen synthesis, which stimulate glucose uptake, are regulated by the activation of the PI3K pathway [162]. Ethyl-acetate extracts and 3β-taraxerol from Mangifera indica stimulate insulin-stimulated glucose uptake through PI3K activation in 3T3-L1 adipocyte cells. This PI3K-pathway activation leads to the activation of the glucose transporter GLUT4 and of PKB, along with the suppression of GSK3β, which is responsible for increased glycogen synthesis [163].

3.25. Mikania cordata

Mikania cordata (Figure 25), commonly known as bitter vine, heartleaf hempvine, etc., is a wild plant in the Asteraceae family. The plant has been reported to have antibacterial, anti-cancer and anti-inflammatory properties and contains saponins, alkaloids, flavonoids, tannins and steroids [134]. In a study conducted by Nasrin et al. (2015), 200 mg/kg and 400 mg/kg body weight of ethanolic leaf extracts from M. cordata were administered to alloxan-induced diabetic rats for 14 days and demonstrated significant hypoglycemic activity compared to the control drug, metformin hydrochloride (100 mg/kg body weight) (p < 0.05 and p < 0.01). The extracts (200 mg/kg and 400 mg/kg body weight) reduced the blood glucose levels of the mice by 38.04% and 47.72%, respectively, demonstrating that M. cordata has anti-diabetic properties. The acute-toxicity study of the extract was carried out by following OECD-423 guidelines and administering the extract at doses of 5, 50, 300 and 2000 mg/kg body weight. The female mice were observed for the next 24 h and then for 14 days for signs of toxicity or mortality. The extract did not confer toxicity or mortality, even at doses up to 2000 mg/kg body weight [64].

3.26. Momordica charantia

Momordica charantia (Figure 26), locally known as Korola or bitter gourd, is a member of the Cucurbitaceae family. It is a climber, with its flowering time spanning from January to December. The juice made from the plant’s leaves is used to treat chickenpox and rheumatism, while the unripe fruit is used to treat diabetes [69]. The plant contains biologically active phytoconstituents, such as saponins, alkaloids, triterpenes, proteins and steroids. In addition, it contains several phytochemicals, such as diosgenin (a compound that alleviates hypoglycemia), momorcharins, momordenol, momordicin, charantin, charine, cucurbitin, gentisic acid, etc. Since the hypoglycemic chemicals in the plant are concentrated in its fruits, these fruits have significant hypoglycemic properties [107]. According to a study conducted by Sathishsekar et al. (2005), 150 mg/kg body weight of aqueous extract of M. charantia seeds significantly (p < 0.05) reduced plasma glucose levels when administered to STZ (STZ)-induced diabetic rats for 30 days [173]. The anti-diabetic effects of M. charantia involve several mechanisms, such as the stimulation of insulin resistance, the suppression of MAPKs and NF-κβ in pancreatic cells, the activation of the AMPK pathway, the promotion of glucose and fatty-acid catabolism, the enhancement of fatty acid absorption, the inducement of insulin production and the inhibition of glucose-metabolism enzymes such as fructose-1,6-bisphosphate and glucose-6-phosphatase [174]. Husna et al. (2013) administered single doses of 300 mg/kg and 2000 mg/kg body weight of ethanolic extract from M. charantia in Sprague–Dawley rats to determine the acute toxicity. The treated rats demonstrated dizziness and depression for the first 30 min after the dose’s administration. However, no marked differences were observed between the behavioral patterns or body weights thereafter. However, in the treatment group, which received 2000 mg/kg body weight of the extract, the hemoglobin counts and relative liver weights were significantly decreased compared to the control. The study demonstrated that the extract was safe to use at 2000 mg/kg body weight [136] and below.

3.27. Moringa oleifera

Moringa oleifera (Figure 27), commonly known as the Drumstick tree, is a tree in the Moringaceae family. It is a valued medicinal plant indigenous to Southeast Asia that has been reported to have analgesic, anti-diabetic, gastroprotective, antimicrobial and wound-healing properties. It contains β-sitosterol, saponins, steroids, alkaloids such as moringine and moringinine, flavonoids such as rhamnetin, isoquercitrin, kaempferitrin, saccharides, phenolic acids, tannins, terpenoid and alpha-carotene [119,120,121]. In a study by Kasolo et al. (2019), ethanolic and aqueous extracts from M. oleifera roots were administered to Swiss albino mice. The aqueous extracts were administered at 10, 15, 20, 25 and 35 g/kg body weight and the ethanolic extracts were administered ats 10, 15, 20, 25 and 30 g/kg body weight to different groups. The animals were observed for behavioral changes for the subsequent 12 h and for mortality in the 24 h following treatment. The LD50 levels of the ethanolic extract and aqueous extract were 17.8 g/kg body weight and 15.9 g/kg body weight, respectively. This demonstrated that the plant is relatively non-toxic (Table 3) [167]. According to a study conducted by Bamagous et al. (2018), 200 mg/kg body weight of ethyl-acetate-extract fraction from M. oleifera leaves administered to STZ-induced Sprague–Dawley rats for 30 days significantly reversed the effects of STZ on serum glucose and insulin, demonstrating anti-diabetic activity (p < 0.05) [51]. Extracts from M. oleifera leaves reduce hyperglycemia in type 2 diabetes by modulating hyperinsulinemia, PPAR-γ and inflammatory cytokines [175]. The inhibition of α-amylase and α-glucosidase activity, increases in glucose uptake in the muscles and liver, the inhibition of glucose uptake from the intestine, decreases in gluconeogenesis in the liver and increases in insulin secretion and sensitivity are some of the other proposed mechanisms for the antidiabetic effect of M. oleifera [176].

3.28. Murraya koenigii

Commonly known as the curry tree, Murraya koenigii (Figure 28) is a tropical tree native to Asia and belongs to the family Rutaceae. The leaves and roots of the plant are rich in bioactive phytocompounds, including carbazole alkaloids such as mahanimbine, flavonoids, sterols, koenimbine, koenine, girinimbine and mukoeic acid, which have antioxidant, antimicrobial, anti-helminthic, anti-diabetic and hepatoprotective properties. They have been used to treat night blindness, diarrhea, dysentery, etc. [135]. According to a study conducted by Husna et al. (2018), the administration of 200 mg/kg body weight and 400 mg/kg body weight of M. koenigii extract leaves in STZ- and nicotinamide-induced hyperglycemic rats led to a significant decrease (p < 0.05) in blood glucose levels compared to the standard drug glibenclamide (1 mg/kg body weight). This antidiabetic effect may have been due to its antioxidant properties and reduced insulin resistance [66]. Darvekar et al. (2011) conducted a study on the acute toxicity of petroleum ether, chloroform and ethanol extracts of Murraya koenigii according to OECD guidelines. The extracts were suspended in saline and administered orally in single doses of 500 mg/kg, 1000 mg/kg and 2500 mg/kg body weight to different groups of rats each; the rats were monitored for the next 24 h. The LD50 of the extracts was found to be 2500 mg/kg body weight [172].

3.29. Nelumbo nucifera

Nelumbo nucifera (Figure 29), commonly known as Indian Lotus, is a plant in the family Nymphaeaceae. It is an aquatic crop grown throughout Asia. The plant has been used in Ayurvedic and Chinese traditional medicine. Its rhizome extract possesses asteroidal triterpenoid, which gives it anti-diabetic and anti-inflammatory properties. It has also been used to treat hematuria, hemoptysis, diarrhea, cholera, fever, etc. [132]. The plant consists of phytoconstituents, such as alkaloids, phenolic bases, flavonoids, quercetin, glycoside and kaempferol. Its other compounds include nuciferine, roemerine, armepavine, β-sitosterol glucopyranoside, etc. [133]. A study conducted by Sakuljaitrong et al. (2013) investigated the hypoglycemic, hypolipidemic and possible acute toxicity of ethanolic extracts from N. nucifera flowers. On the acute-toxicity test, the extract was administered orally in doses of 0, 500, 1000, 1500 and 2000 mg/kg body weight to healthy adult male Wistar rats, which were observed for a subsequent 24 h and then 14 days for signs of toxicity or morbidity. The extract was also administered in 250-mg/kg-body-weight doses to STZ-induced male Wistar rats with 0.25 mg/kg body weight glibenclamide as a reference drug to investigate its possible hypolipidemic and hypoglycemic activities. The extract significantly (p < 0.05) lowered fasting blood glucose, total cholesterol, triglycerides and low-density lipoprotein. The acute-toxicity test revealed that the extract did not confer any toxicity or mortality, even at 2000 mg/kg body weight [63]. Nelumbo nucifera leaf extract was evaluated for its antidiabetic efficacy by Sharma BR et al. (2016). The results suggested that leaf extracts reduce the toxicity of pancreatic β-cells in type 1 diabetes, followed by the increased secretion of insulin from pancreatic β-cells [177]. This mechanism of action clearly indicates its antidiabetic potential.

3.30. Neolamarckia cadamba

Neolamarckia cadamba (Figure 30), also known as the Burflower tree and locally known as Kodom, is an evergreen tropical tree from the Rubiaceae family. It is found in Bangladesh, India, Sri Lanka, Nepal and Myanmar. The phenolics and alkaloids, cadambine, chlorogenic acid and dihydrocinchonine, present in the plant give it anti-diabetic properties [110,111]. Its other phytoconstituents include flavonoids, saponins, triterpenes, indole alkaloids, triterpenoid glycosides, 3α-dihydrocadambine, isodihydrocadambine, isocadamine, etc. Dubey et al. (2011) investigated the acute toxicity of N. cadamba barks by administering methanolic extracts from the plant in mouse models. The mice had no mortality for doses up to 3000 mg/kg, with acute toxicity demonstrated only at doses above 3000 mg/kg body weight [112]. According to a study by Ahmed et al. (2011), a methanolic extract from Neolamarckia cadamba reduced glucose-induced hyperglycemic mice’s elevated blood glucose. In the study, the methanolic extract of Neolamarckia cadamba was administered to glucose-loaded mice at 200 mg/kg and 400 mg/kg body weight. The dosage of 400 mg/kg body weight demonstrated the greatest anti-hyperglycemic activity, which was comparable to 10 mg/kg of glibenclamide (p < 0.05) [42].

3.31. Piper betle Leaves

Piper betle (Figure 31) is a Piperaceae-family plant commonly known as the betel leaf. Betel leaves are respiratory depressants and have anti-diabetic, antitumor, antimutagenic, carminative and stomachic properties. They are also commonly used as mouth fresheners [178]. The plant consists of several phytochemicals, such as glycosides, alkaloids, saponins, steroids, tannins, diterpenes, eugenol, quercetin, chavibetol, flavonoids, hydroxychavicol, etc. Many toxicity tests have been carried out on Piper betel extracts, confirming their safety for human use [130]. In a study by Avijit et al. (2005), 100, 200, 300 and 1500 mg/kg body weight of hot and cold ethanolic extracts from P. betle leaves were administered to STZ-induced diabetic rats, with the drug tolbutamide (22.5 mg/kg body weight) used as the standard drug. Both extracts significantly (p < 0.05) reduced the fasting blood glucose in the rats, suggesting that the extract had hypoglycemic activity [61]. The inhibition of α-glucosidase and α-amylase inhibition and the insulin-mimetic properties of Piper betle extract constitute the proposed mechanism behind the hypoglycemic effect [61,179]. In a study conducted by Al-Adhroey et al. (2010), following the OECD 423 guideline, “Acute oral toxicity–acute toxic class method”, the acute toxicity of methanolic extracts from P. betel leaves was studied. The extracts were administered to mice at doses of up to 5000 mg/kg body weight. The mice were observed for mortality or signs of toxicity twice, 1, 3 and 4 h after administration. They were then observed twice daily for 14 days, after which they were sacrificed for necropsy and histopathological study. The extract demonstrated no toxicity at 5000 mg/kg body weight, demonstrating that it is safe for use [170].

3.32. Punica granatum

Punica granatum (Figure 32), commonly known as pomegranate, is a plant in the Lythraceae family. The leaves, fruits and seeds of the plant have been used in the treatment of dysentery, diabetes, heart diseases, etc. [180]. The P. granatum plant consists of flavonols, triterpenoids, fatty acids, organic acids such as citric acid, malic acid, ascorbic acid, tannins such as gallagic acid, punicalin, punicalagin, flavonoids such as luteolin, quercetin, kaempferol, magnesium and alkaloids such as pelletierine, etc. [117]. In an acute toxicity study by Bhandary et al. (2013), 2000-mg/kg-body-weight ethanolic extracts from the whole fruit and seeds of P. granatum were orally administered to mice. There were no behavioral changes and no significant changes in biochemical or hematological parameters, with no mortality in the treated groups. The LD50 was greater than 2000 mg/kg body weight, so the extracts’ oral consumption was deemed non-toxic [166]. According to a study conducted by Pottahil et al. (2020), a methanolic extract from Punica granatum leaves was administered to STZ-induced type 2 diabetic rats at doses of 100, 200, 400 and 600 mg/kg body weight daily for 45 days. The P. granatum extract demonstrated significant anti-diabetic activity (p < 0.05) due to the presence of the antioxidants gallic acid, ellagic acid and apigenin [46]. In another study, conducted by Gharib and Kouhsari (2019), an aqueous extract of P. granatum fruit noticeably reduced fasting blood glucose compared to a diabetic control in alloxan-induced diabetic Wistar rats (p < 0.001) [47]. The understanding of the mechanism of action of Punica granatum flower extracts has improved. Punica granatum flowers contain gallic acid, which activates the PPAR-γ pathway to improve insulin-receptor sensitivity [181]. Punica granatum flowers’ phenolic extracts can lower glucose levels in type 2 diabetes by activating the Akt/GSK3β signaling pathway and inhibiting endoplasmic reticulum (ER) stress. The mechanism of ER-stress inhibition involves a decrease in IRE1 (inositol-requiring kinase1) phosphorylation and XBP-1 (X-box-binding protein-1) splicing. This mechanism leads to improved insulin sensitivity [182].

3.33. Syzygium cumini (L.)

Syzygium cumini (Figure 33), locally known as Kalojam and commonly known as Java plum, is a plant in the Myrtaceae family. The bark and seeds of the plant are used to treat wounds, dysentery and diabetes [69]. The plant consists of bioactive phytoconstituents, such as alkaloids, tannins, flavonoids, carotenoids, glycosides, saponins, steroids, triterpenoids, anthocyanins, phenols, oxalic acid, phytosterols, myricetin, gallic acid etc., which endow it with anti-inflammatory, antibacterial, antioxidant, hypoglycemic and hypolipidemic properties [108,109]. Deb et al. (2013) administered methanolic and aqueous extracts from the bark, roots, seeds and leaves of S. cumini in albino mice of both sexes, according to the OECD 423 guidelines. The median lethal dose (LD50) cut-off for all the extracts was 2000 mg/kg body weight, except for the methanolic seed extract, for which the LD50 cut-off was 200 mg/kg body weight (Table 3). No common side effects, such as mild diarrhea, weight loss or depression, were reported within 7 days of the administration of the extracts [164]. A study conducted by Gupta and Saxena (2011) demonstrated that a single oral dose of 200 mg/kg body weight of 95% ethanolic extract from the whole fruit of Syzygium cumini in STZ-induced diabetic rats led to a significant decrease in blood glucose concentration (p < 0.01), increases in muscle glycogen stores and degranulation in pancreatic β-cells, demonstrating hypoglycemic activity [40]. In another study, by Schoenfelder et al. (2008), a crude ethanolic extract from Syzygium cumini leaves was administered in normal, hyperglycemic and alloxan-induced diabetic rats. The extract significantly (p < 0.001) decreased blood glucose levels in the alloxan-induced diabetic rats at dosages of 125 mg/kg and 250 mg/kg [41]. There are several mechanisms of action in the antidiabetic activity of Syzygium cumini in type 2 diabetes, such as the improvement of insulin resistance and β-cell dysfunction through the modulation of PPARγ and the reduction in dyslipidemia, oxidative stress and TNF-α [183].

3.34. Tamarindus indica

Tamarindus indica (Figure 34), commonly known as Tamarind, is a leguminous tree in the Fabaceae family. The leaves and fruits of the plant have been used to treat diabetes [85]. The plant is native to the Indian subcontinent and most tropical countries. It has been used to treat diarrhea, dysentery, wound healing, bacterial and parasitic infections. It contains essential amino acids and phytochemicals and possesses antimicrobial, antivenom and anti-diabetic properties [184]. Furthermore, T. indica consists of phenolic compounds, β-sitosterol, proanthocyanidins, apigenin, tartaric acid, cardiac glycosides, mucilage, pectin, eicosanoic acid and β-sitosterol, as well as essential elements such as calcium, copper, iron, manganese, magnesium, palmitic acid, oleic acid, succinic acid, formic acid, β-amyrin, apigenin, epicatechins, catechins, taxifolin, eriodictyol, naringenin, etc. [116]. Abukakar et al. (2008) administered aqueous pulp extract to adult albino rats for acute-toxicity testing. No toxicity was observed in the tests up to 900–4500 mg/kg body weight. However, at higher doses of 2700–4500 mg/kg, mild behavioral changes, such as anorexia, restlessness, sensitivity to sound, etc., were observed. Nevertheless, no mortality was observed at these concentrations, with no apparent histopathological alterations. Therefore, the plant had no apparent toxic effects in animal models, which supported its safe usage as a medicine [165]. In a study conducted by Parvin et al. (2013), 1.25 g/kg of T. indica seed powder was administered to non-diabetic and STZ-induced diabetic Long–Evans rats and demonstrated anti-hyperglycemic properties (p < 0.03) (Table 1) [45]. The antidiabetic property of Tamarindus indica seeds may be due to the presence of phenols and flavonoids, as they have insulin-mimetic properties and stimulate glucose uptake [185]. A recent study by Costa et al. (2022) also suggested this insulin-mimetic effect [186]. The specific molecular mechanism involves increases in the expression of glucose transporter (GLUT-2 and GLUT-4) genes, which increase glucose uptake and help to maintain glucose homeostasis. Furthermore, the compounds in T. indica increase the concentration of Ca2+ in islets of Langerhans cells, stimulating β-pancreatic cell proliferation and enhancing insulin secretion [185].

3.35. Terminalia catappa

Terminalia catappa (Figure 35) belongs to the family Combretaceae and is locally known as Bangla badam [187]. The juice of its leaves is used to treat skin conditions, such as leprosy and scabies. It is also ingested to alleviate symptoms of stomach ache and headache. Furthermore, it has antimicrobial, anti-inflammatory, antioxidant and anti-cancer activities, alongside its anti-diabetic properties [129]. The plant contains tannins, saponins, phenolics, flavonoids, triterpenoids, punicalin, punicalagin, kaempferol, geraniin, punicalin, quercetin, gentisic acid, tercatain, tergallagin, β-carotene, cyanidin-3-glucoside, ellagic acid, gallic acid, etc. [59,129].
According to a study by Ahmed et al. (2005), aqueous and cold plant-leaf extracts reduced fasting blood sugar levels in alloxan-induced diabetes rats, with glibenclamide (10 mg/kg body weight) used as a control. The regeneration of the pancreas was confirmed by the results of histological tests [59]. The anti-diabetic properties of Terminalia catappa leaf extract involve the activation of the PI3K/AKT pathway, the reversal of insulin resistance and the enhancement of glucose transport in type 2 diabetes [188]. Azrul et al. (2013) conducted a toxicity study in which they investigated the primary and secondary toxicity of a crude aqueous extract from T. catappa in Sprague–Dawley rats by administering 0.5, 1.0 and 3.0 g/kg body weight of the extract to the rats for 14 days. No mortality or signs of toxicity were noted in the rats [169].

3.36. Terminalia chebula

Terminalia chebula (Figure 36), commonly known as Myrobalan, is a plant in the Combretaceae family [189]. The plant’s leaves are known to have antihyperglycemic properties and are thus reportedly used to treat diabetes by traditional health practitioners [180]. The plant contains flavonoids, β-sitosterol, tannins, sterols, gallic acid, chebulanin, corilagin, ellagic acid, chebulinic acid, amino acids, fructose, resin, triterpenoids, glycosides, etc. The plant has laxative, carminative, anti-diabetic, anti-cancer, antimutagenic and antiviral properties [58,128]. According to a study by Kumar et al. (2006), STZ-induced rats were orally administered with 200 mg/kg body weight of the fruit extract of the plant. Within 30 days of the administration, the rats showed lowered blood glucose levels. The fruit extract was linked with insulin stimulation and its effectiveness was comparable with the hypoglycemic drug, glibenclamide (600 μg/kg body weight) (p < 0.05) [58]. The bioactive compounds present in Terminalia chebula fruit exhibit insulin-like actions and inhibit the α2 receptors of pancreatic β-cells, increasing insulin secretion [190]. In an acute-toxicity study of T. chebula conducted by Panunto et al. (2010), aqueous extracts from dried fruits from the plant were orally administered to five male and five female rats at a single dose of 5000 mg/kg body weight. The extract did not confer any mortality, behavioral changes or changes in the internal organs, demonstrating that the extract was non-toxic [168].

3.37. Wedelia chinensis

Wedelia chinensis (Figure 37), also known as Wedelia, is a climbing wetland herb from the Asteraceae family native to India, China and Japan. Furthermore, is reported to have anti-cancer, antimicrobial, antioxidant and anti-inflammatory properties and contains flavonoids, alkaloids, saponins, phytosterols, mucilage, carbohydrates and tannins [171,191]. Bari et al. (2020) demonstrated that the administration of 100 mg/kg and 200 mg/kg body weight of methanolic extract of W. chinensis to alloxan-induced hyperglycemic Swiss albino mice significantly reduced blood glucose levels (p < 0.01), in addition to improving the lipid profiles (TG, LDL, TC, VLDL and HDL) compared to standard the drug (glibenclamide 5 mg/kg body weight) [65]. Compounds isolated from W. chinensis leaves, especially flavonoids, have significant α-amylase- and α-glucosidase-inhibitory activity, which is an effective mechanism for managing type 2 diabetes [192]. In a study conducted by Umasankar et al. (2010), the toxicity of the ethanolic extract from W. chinensis was investigated following the methods of Litchfield and Wilcoxon (1949). The extracts were each administered in doses of 100, 200, 400, 800 and 1600 mg/kg body weight to five different groups of mice. The mice were observed for mortality for the subsequent 72 h. The extract did not confer mortality or toxicity, indicating that the LD50 value of the extract was >1600 g/kg body weight [171].

4. Plant-Based Treatments: Current Progress and Future Prospects

Plants are natural sources of bioactive compounds, with various health-beneficial pharmacological activities [193]. The comparatively low or almost no toxicity of the 37 plants discussed above is the main advantage of plant-based treatments. With the progress in the understanding of the mechanisms through which plant extracts can be used to treat different conditions, researchers are conducting clinical trials. A randomized, double-blind, placebo-controlled clinical trial on sixty type 2 diabetes mellitus (T2DM) patients was conducted by Ghafouri et al. (2020) to evaluate the hypoglycemic effect of Rheum ribes extracts. The consumption of Rheum ribes extracts improved the diabetic patients’ insulin sensitivity, malondialdehyde levels and high-sensitivity C-reactive protein levels significantly [194]. In a clinical trial, the widely used synthetic antidiabetic drug metformin’s efficacy was improved by a Ginkgo biloba extract as an adjuvant [195]. Another recent development in the management of T2DM is the use of herbal nanoformulations of plant-based natural products, which have been proven to be better than crude extracts [196]. Plant-based vaccines have shown significant therapeutic potential to treat respiratory diseases in recent years [197]. This is evidence that in the future, plant-based treatments will greatly influence the discovery of novel drugs and treatments. However, the mechanism of action behind different pharmacological activities is yet to be established for many plants; this is one of the main limitations of plant-based treatments. A caveat to the use of medicinal plants is that approximately 15,000 species are in danger of becoming extinct due to habitat loss and overharvesting, highlighting the urgent need for conservation efforts [198]. The implementation of biotechnological techniques, such as tissue culture, micropropagation, synthetic seed technology and molecular-marker-based approaches can enhance the yield and modify the potency of medicinal plants. These advances can be complemented with conservation strategies to ensure the sustainable use of these resources [198]. The phytoconstituents responsible for the pharmacological activities and their mechanism of action with safety measures should be studied more extensively, followed by clinical trials. This will lead to the development of plant-based treatments that are accepted by medical science.

5. Conclusions

There is potential for plant medicine to become an established form of alternative medicine in Bangladesh due to its effectiveness, affordability and accessibility. It is encouraged to undertake further research and studies to maximize the potential of these plants for use in the management of diabetes mellitus, a widespread affliction in Bangladesh and the wider world. As the world is searching for new natural sources of drugs to treat diabetes mellitus, plants for which the mechanism behind their antidiabetic effect has already been established and clearly understood may be chosen for developing antidiabetic drugs. Further clinical trials can be conducted to prove their efficacy.

Author Contributions

M.K.H. and K.Z.Z. conceptualized the manuscript; K.Z.Z. and M.S.H. wrote the manuscript; M.K.H. critically reviewed and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The dataset used in this study is available upon request (mailing to the corresponding author).

Conflicts of Interest

The authors declare no conflict of interest. The authors alone are responsible for the content and writing of this article.

References

  1. El-Tantawy, W.H.; Temraz, A. Management of diabetes using herbal extracts: Review. Arch. Physiol. Biochem. 2017, 124, 383–389. [Google Scholar] [CrossRef]
  2. Nash, R.J.; Kato, A.; Yu, C.-Y.; Fleet, G.W. Iminosugars as therapeutic agents: Recent advances and promising trends. Futur. Med. Chem. 2011, 3, 1513–1521. [Google Scholar] [CrossRef] [PubMed]
  3. Chennaiah, A.; Dahiya, A.; Dubbu, S.; Vankar, Y.D. A Stereoselective Synthesis of an Imino Glycal: Application in the Synthesis of (-)-1-epi-Adenophorine and a Homoimindosugar. Eur. J. Org. Chem. 2018, 2018, 6574–6581. [Google Scholar] [CrossRef]
  4. Deyno, S.; Eneyew, K.; Seyfe, S.; Wondim, E. Efficacy, safety and phytochemistry of medicinal plants used for the management of diabetes mellitus in Ethiopia: A systematic review. Clin. Phytoscience 2021, 7, 16. [Google Scholar] [CrossRef]
  5. Malode, L.L.; Manwar, J.V.; Panchale, W.A.; Bartere, S.A.; Bakal, R.L. Potential of medicinal plants in management of diabetes: An updates. GSC Adv. Res. Rev. 2021, 8, 149–169. [Google Scholar] [CrossRef]
  6. Izzo, A.; Massimino, E.; Riccardi, G.; Della Pepa, G. A Narrative Review on Sarcopenia in Type 2 Diabetes Mellitus: Prevalence and Associated Factors. Nutrients 2021, 13, 183. [Google Scholar] [CrossRef] [PubMed]
  7. Mishra, S.B.; Rao, C.V.; Ojha, S.K.; Vijayakumar, A.; Verma, A.; Alok, S. An analytical review of plants for anti diabetic activity with their phytoconstituent & mechanism of action. Int. J. Pharm. Sci. Res. 2010, 1, 29–46. [Google Scholar]
  8. Abdel-Aziz, S.M.; Aeron, A.; Kahil, T.A. Health Benefits and Possible Risks of Herbal Medicine. Microbes Food Health 2016, 97–116. [Google Scholar] [CrossRef]
  9. Nawaz, A.H.M.M.; Hossain, M.; Karim, M.; Khan, M.; Jahan, R.; Rahamatullah, M. An ethnobotanical survey of Rajshahi district in Rajshahi division, Bangladesh. Am. Eurasian J. Sustain. Agric. 2009, 3, 143–150. [Google Scholar]
  10. Srivastav, S.; Singh, P.; Mishra, G.; Jha, K.K.; Khosa, R.L. Achyranthes aspera—An important medicinal plant: A review. J. Nat. Prod. Plant Resour. 2011, 1, 1–14. [Google Scholar]
  11. Vijayaraj, R.; Kumar, K.N.; Mani, P.; Senthil, J.; Jayaseelan, T.; Kumar, G.D. Hypoglycemic and antioxidant activity of Achyranthes aspera seed extract and its effect on Streptozotocin induced diabetic rats. Int. J. Biol. Pharm. Res. 2016, 7, 23–28. [Google Scholar]
  12. Kumar, S.S.; Gnananath, K.; Saibana, G.; Rajasekhar, G.E.; Rajesh, P.; Nagarjuna, S. Anti diabetic Activity of Ethanolic Extract of Achyranthes aspera Leaves in Streptozotocin induced diabetic rats. J. Pharm. Res. 2011, 4, 3124–3125. [Google Scholar]
  13. Sadashiv, P.; Krishna, A. Acute toxicity study for Achyranthes aspera leaves. J. Pharmacol. Res. 2011, 4, 2221–2222. [Google Scholar]
  14. Joseph, B.; Raj, S.J. Pharmacognostic and phytochemical properties of Aloe vera Linn—An overview. Int. J. Pharm. Sci. Rev. Res. 2010, 4, 106–110. [Google Scholar]
  15. Lanjhiyana, S.; Garabadu, D.; Ahirwar, D.; Bigoniya, P.; Rana, A.C.; Patra, K.C.; Lanjhiyana, S.K.; Karuppaih, M. Antihyperglycemic potential of Aloe vera gel in experimental animal model. Ann. Biol. Res. 2011, 2, 17–31. [Google Scholar]
  16. Kim, J.H.; Cho, C.W.; Lee, J.I.; Vinh, L.B.; Kim, K.T.; Cho, I.S. An investigation of the inhibitory mechanism of α-glucosidase by chysalodin from Aloe vera. Int. J. Biol. Macromol. 2020, 147, 314–318. [Google Scholar] [CrossRef]
  17. Prasannaraja, C.; Kamalanathan, A.S.; Vijayalakshmi, M.A.; Venkataraman, K. A dipyrrole derivative from Aloe vera inhibits an anti-diabetic drug target Dipeptidyl Peptidase (DPP)-IV in vitro. Prep. Biochem. Biotechnol. 2020, 50, 511–520. [Google Scholar] [CrossRef]
  18. Rahman, A.H.M.M.; Gulshana, M.I.A. Taxonomy and Medicinal Uses on Amaranthaceae Family of Rajshahi, Bangladesh. Appl. Ecol. Environ. Sci. 2014, 2, 54–59. [Google Scholar] [CrossRef] [Green Version]
  19. Aneja, S.; Vats, M.; Aggarwal, S.; Sardana, S. Phytochemistry and hepatoprotective activity of aqueous extract of Amaranthus tricolor Linn. roots. J. Ayurveda Integr. Med. 2013, 4, 211–215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Kunyanga, C.N.; Imungi, J.K.; Okoth, M.W.; Biesalski, H.K.; Vadivel, V. Antioxidant and Type 2 Diabetes Related Functional Properties of Phytic Acid Extract from Kenyan Local Food Ingredients: Effects of Traditional Processing Methods. Ecol. Food Nutr. 2011, 50, 452–471. [Google Scholar] [CrossRef] [PubMed]
  21. Clemente, A.; Desai, P. Evaluation of the Hematological, Hypoglycemic, Hypolipidemic and Antioxidant Properties of Amaranthus tricolor Leaf Extract in Rat. Trop. J. Pharm. Res. 2011, 10, 595–602. [Google Scholar] [CrossRef] [Green Version]
  22. Kumar, S.; Saini, M.; Kumar, V.; Prakash, O.; Arya, R.; Rana, M.; Kumar, D. Traditional medicinal plants curing diabetes: A promise for today and tomorrow. Asian J. Tradit. Med. 2012, 7, 178–188. [Google Scholar]
  23. De Sousa Leite, A.; Islam, T.; Gomes, A., Jr.; de Castro e Sousa, J.M.; de Alencar, M.V.A.B.; Fernanda Correia Jardim Paz, M.; Rolim, H.M.L.; das Graças Freire de Medeiros, M.; de Carvalho Melo-Cavalcante, A.A.; Lopes, J.A.D. Pharmacological properties of cashew (Anacardium occidentale). Afr. J. Biotechnol. 2016, 15, 1855–1863. [Google Scholar]
  24. Ukwenya, V.O.; Ashaolu, J.O.; Adeyemi, D.O.; Akinola, O.A.; Caxton-Martins, E.A. Antihyperglycemic activities of methanolic leaf extract of Anacardium occidentale (Linn.) on the pancreas of streptozotocin-induced diabetic rats. J. Cell Anim. Biol. 2012, 6, 207–212. [Google Scholar] [CrossRef]
  25. Tedong, L.; Madiraju, P.; Martineau, L.C.; Vallerand, D.; Arnason, J.T.; Desire, D.D.P.; Lavoie, L.; Kamtchouing, P.; Haddad, P.S. Hydro-ethanolic extract of cashew tree (Anacardium occidentale) nut and its principal compound, anacardic acid, stimulate glucose uptake in C2C12 muscle cells. Mol. Nutr. Food Res. 2010, 54, 1753–1762. [Google Scholar] [CrossRef]
  26. Konan, N.A.; Bacchi, E.M.; Lincopan, N.; Varela, S.D.; Varanda, E.A. Acute, subacute toxicity and genotoxic effect of a hydroethanolic extract of the cashew (Anacardium occidentale L.). J. Ethnopharmacol. 2007, 110, 30–38. [Google Scholar] [CrossRef]
  27. Jiang, M.; Sheng, F.; Zhang, Z.; Ma, X.; Gao, T.; Fu, C.; Li, P. Andrographis paniculata (Burm.f.) Nees and its major constituent andrographolide as potential antiviral agents. J. Ethnopharmacol. 2021, 272, 113954. [Google Scholar] [CrossRef]
  28. Akter, R.; Mahabub-uz-Zaman, M.; Rahman, S.; Abdullah, A.; Ahmed, N.U.; Islam, F. Comparative Studies on Antidiabetic effect with phytochemical screening of Azadirachta indicia and Andrographis paniculata. IOSR J. Pharm. Biol. Sci. 2013, 5, 122–128. [Google Scholar] [CrossRef] [Green Version]
  29. Augustine, A.W.; Narasimhan, A.; Vishwanathan, M.; Karundevi, B. Evaluation of antidiabetic property of Andrographis paniculata powder in high fat and sucrose-induced type-2 diabetic adult male rat. Asian Pac. J. Trop. Dis. 2014, 4, S140–S147. [Google Scholar] [CrossRef]
  30. Kumar, M.; Changan, S.; Tomar, M.; Prajapati, U.; Saurabh, V.; Hasan, M.; Sasi, M.; Maheshwari, C.; Singh, S.; Dhumal, S.; et al. Custard Apple (Annona squamosa L.) Leaves: Nutritional Composition, Phytochemical Profile, and Health-Promoting Biological Activities. Biomolecules 2021, 11, 614. [Google Scholar] [CrossRef]
  31. Rout, S.P.; Kar, D.M.; Mohapatra, S.B.; Swain, S.P. Anti-hyperglycemic effect Annona reticulata L. leaves on experimental diabetic rat model. Asian J. Pharm. Clin. Res. 2013, 6, 56–60. [Google Scholar]
  32. Ansari, P.; Flatt, P.R.; Harriott, P.; Abdel-Wahab, Y.H. Evaluation of the Antidiabetic and Insulin Releasing Effects of A. squamosa, Including Isolation and Characterization of Active Phytochemicals. Plants 2020, 9, 1348. [Google Scholar] [CrossRef]
  33. Onwusonye, J.C.; Uwakwe, A.; Patrick, A.; Iwuanyanwu, K.P. Acute and sub-acute toxicity studies of methanol leaf extracts of Annona squamosa Linn. In mice. Sky J. Biochem. Res. 2014, 3, 53–59. [Google Scholar]
  34. Mandal, S.C.; Mukherjee, P.K.; Saha, K.; Das, J.; Pal, M.; Saha, B.P. Hypoglycemic activity of Ficus racemosa L. (Moraceae) leaves in streptozotocin-induced diabetic rats. Nat. Prod. Sci. 1997, 3, 38–41. [Google Scholar]
  35. Vadivelan, R.; Dipanjan, M.; Umasankar, P.; Dhanabal, S.P.; Satishkumar, M.N.; Antony, S.; Elango, K. Hypoglycemic, antioxidant and hypolipidemic activity of Asparagus racemosus on streptozotocin-induced diabetic in rats. Adv. Appl. Sci. Res. 2011, 2, 179–185. [Google Scholar]
  36. Bhavsar, C.; Talele, G.S. Potential anti-diabetic activity of Bombax ceiba. Bangladesh J. Pharmacol. 2013, 8, 102–106. [Google Scholar] [CrossRef] [Green Version]
  37. Nahar, L.; Nasrin, F.; Zahan, R.; Haque, A.; Haque, E.; Mosaddik, A. Comparative study of antidiabetic activity of Cajanus cajan and Tamarindus indica in alloxan-induced diabetic mice with a reference to in vitro antioxidant activity. Pharmacogn. Res. 2014, 6, 180–187. [Google Scholar] [CrossRef] [Green Version]
  38. Islam, M.A.; Khan, M.D.; Hossain, M.S.; Alam, A.K.; Wahed, M.I.; Rahman, B.M.; Anisuzzaman, A.S.; Shaheen, S.; Ahmed, M. Antidiabetic and hypolipidemic effects of different fractions of Coccinia cordifolia L. on normal and streptozotocin-induced diabetic rats. Pak. J. Pharm. Sci. 2011, 24, 331–338. [Google Scholar]
  39. Husna, R.N.; Noriham, A.; Nooraain, H.; Azizah, A.H.; Amna, O.F. Acute oral toxicity effects of Momordica charantia in sprague dawley rats. Int. J. Biochem. Bioinforma. 2013, 3, 408. [Google Scholar]
  40. Gupta, R.; Saxena, A.M. Hypoglycemic and anti-hyperglycemic activities of Syzygium cumini (Linn.) skeels whole fruit, in normal and streptozotocin-induced diabetic rats. Asian J. Pharm. Biol. Res. 2011, 8, 88–93. [Google Scholar]
  41. Schoenfelder, T.; Warmlin, C.Z.; Manfredini, M.S.; Pavei, L.L.; Réus, J.V.; Tristão, T.C.; Fernandes, M.S.; Costa-Campos, L. Hypoglycemic and hypolipidemic effect of leaves from Syzygium cumini (L.) Skeels, Myrtaceae. in diabetic rats. Rev. Bras. Farm. 2010, 20, 222–227. [Google Scholar] [CrossRef] [Green Version]
  42. Ahmed, F.; Rahman, S.; Ahmed, N.; Hossain, M.; Biswas, A.; Sarkar, S.; Banna, H.; Khatun, A. Evaluation of Neolamarckia cadamba (Roxb.) bosser leaf extract on glucose tolerance in glucose-induced hyperglycemic mice. Afr. J. Tradit. Complement. Altern. Med. 2010, 8, 79–81. [Google Scholar] [CrossRef] [Green Version]
  43. Madhuri, A.S.; Mohanvelu, R. Evaluation of Antidiabetic Activity of Aqueous Extract of Mangifera Indica Leaves in Alloxan Induced Diabetic Rats. Biomed. Pharmacol. J. 2017, 10, 1029–1035. [Google Scholar] [CrossRef]
  44. Naskar, S.; Mazumder, U.K.; Pramanik, G.; Gupta, M.; Kumar, R.S.; Bala, A.; Islam, A. Evaluation of antihyperglycemic activity of Cocos nucifera Linn. on streptozotocin induced type 2 diabetic rats. J. Ethnopharmacol. 2011, 138, 769–773. [Google Scholar] [CrossRef] [PubMed]
  45. Parvin, A.; Alam, M.; Haque, A.; Bhowmik, A.; Ali, L.; Rokeya, B. Study of the Hypoglycemic Effect of Tamarindus indica Linn. Seeds on Non-Diabetic and Diabetic Model Rats. Br. J. Pharm. Res. 2013, 3, 1094–1105. [Google Scholar] [CrossRef]
  46. Pottathil, S.; Nain, P.; Morsy, M.A.; Kaur, J.; Al-Dhubiab, B.E.; Jaiswal, S.; Nair, A.B. Mechanisms of Antidiabetic Activity of Methanolic Extract of Punica granatum Leaves in Nicotinamide/Streptozotocin-Induced Type 2 Diabetes in Rats. Plants 2020, 9, 1609. [Google Scholar] [CrossRef]
  47. Gharib, E.; Kouhsari, S.M. Study of the antidiabetic activity of Punica granatum L. fruits aqueous extract on the allox-an-diabetic wistar rats. Iran. J. Pharm. Res. 2019, 18, 368. [Google Scholar]
  48. Bisht, S.; Sisodia, S.S. Anti-hyperglycemic And antidyslipidemic potential of Azadirachta indica leaf extract in STZ- induced diabetes mellitus. J. Pharm. Sci. Res. 2010, 2, 622. [Google Scholar]
  49. Revathy, J.; Abdullah, S.S.; Kumar, P.S. Antidiabetic effect of Costus speciosus rhizome Extract in alloxan induced albino rats. J. Chem. Biochem. 2014, 2, 13–22. [Google Scholar]
  50. Bavarva, J.H.; Narasimhacharya, A.V.R.L. Antihyperglycemic and hypolipidemic effects of Costus speciosus in alloxan induced diabetic rats. Phytotherapy Res. 2008, 22, 620–626. [Google Scholar] [CrossRef]
  51. Bamagous, G.A.; Al Ghamdi, S.S.; Ibrahim, I.A.A.; Mahfoz, A.M.; Afify, M.A.; Alsugoor, M.H.; Shammah, A.A.; Arulselvan, P.; Rengarajan, T. Antidiabetic and antioxidant activity of ethyl acetate extract fraction of Moringa oleifera leaves in streptozotocin-induced diabetes rats via inhibition of inflammatory mediators. Asian Pac. J. Trop. Biomed. 2018, 8, 320. [Google Scholar] [CrossRef]
  52. Saraswathi, S.; Senthamarai, R.; Sundari, S. Antidiabetic activity of leaves extract of Ficus benghalensis linn on alloxan induced diabeteic rats. Int. J. Pharmacol. Biol. Sci. 2013, 7, 47–51. [Google Scholar]
  53. Kasireddy, G.R.; Kumar, K.S.; Nadithe, L.R.; Chinnam, P. Experimental evaluation of hypoglycemic effect of bark extract of Ficus benghalensis in streptozotocin-induced diabetic rats. Natl. J. Physiol. Pharm. Pharmacol. 2021, 11, 320–325. [Google Scholar] [CrossRef]
  54. Nalamolu, R.K.; Boini, K.M.; Nammi, S. Effect of chronic administration of Boerhaavia diffusa Linn. leaf extract on experimental diabetes in rats. Trop. J. Pharm. Res. 2007, 3, 305–309. [Google Scholar] [CrossRef] [Green Version]
  55. Nammi, S.; Boini, M.K.; Lodagala, S.D.; Behara, R.B.S. The juice of fresh leaves of Catharanthus roseus Linn. reduces blood glucose in normal and alloxan diabetic rabbits. BMC Complement. Altern. Med. 2003, 3, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Bodhankar, S.; Jain, B.; Bhardwaj, S.; Badole, S.; Patel, N. Antihyperglycemic activity of aqueous extract of leaves of Cocculus hirsutus (L.) Diels in alloxan-induced diabetic mice. Indian J. Pharmacol. 2006, 38, 49. [Google Scholar] [CrossRef]
  57. Ghosh, R.; Sharatchandra, K.H.; Rita, S.; Thokchom, I.S. Hypoglycemic activity of Ficus hispida (bark) in normal and diabetic albino rats. Indian J. Pharmacol. 2004, 36, 222. [Google Scholar]
  58. Kumar, G.P.S.; Arulselvan, P.; Kumar, D.S.; Subramanian, S.P. Anti-Diabetic Activity of Fruits of Terminalia chebula on Streptozotocin Induced Diabetic Rats. J. Heal Sci. 2006, 52, 283–291. [Google Scholar] [CrossRef] [Green Version]
  59. Ahmed, S.M.; Swamy, V.B.M.; Gopkumar, P.; Dhanapal, R.; Chandrashekara, V.M. Anti-diabetic activity of Terminalia catappa Linn. leaf extracts in alloxan-induced diabetic rats. Iran. J. Pharmacol. Ther. 2005, 4, 36–39. [Google Scholar]
  60. Hasan, N.; Rahman, M.H.; Guo, R.; Hirashima, A. Hypoglycemic activity of methanolic leaf extract of Blumea lacera in Swiss-albino mice. Asian Pac. J. Trop. Dis. 2015, 5, 195–198. [Google Scholar] [CrossRef]
  61. Arambewela, L.; Arawwawala, L.; Ratnasooriya, W. Antidiabetic activities of aqueous and ethanolic extracts of Piper betle leaves in rats. J. Ethnopharmacol. 2005, 102, 239–245. [Google Scholar] [CrossRef]
  62. Dd, I. Antidiabetic activity of ethanolic extract of Kalanchoe pinnata leaves in alloxan induced hyperglycaemic rats. Indones. J. Pharm. 2016, 27, 139. [Google Scholar] [CrossRef] [Green Version]
  63. Sakuljaitrong, S.; Buddhakala, N.; Chomko, S.; Talubmook, C. Effects of flower extract from Lotus (Nelumbo nucifera) on hypoglycemic and hypolipidemic in streptozotocin-induced diabetic rats. Int. J. Sci. Eng. Res. 2013, 4, 1441–1446. [Google Scholar]
  64. Nasrin, F.; Hakim, M.L.; Hassan, M.R.; Afroz, N.; Ikbal, M.H.; Azam, S. Hypoglycemic study of ethanolic extract of Mikania cordata leaf. World J. Pharm. Res. 2016, 4, 1–9. [Google Scholar]
  65. Bari, W.; Islam, M.; Khatun, M.; Sultana, M.J.; Ahmed, R.; Islam, A.; Hossain, I.; Rahman, M.; Islam, M.A. Antidiabetic effect of Wedelia chinensis leaf extract in alloxan induced Swiss albino diabetic mice. Clin. Phytoscience 2020, 6, 58. [Google Scholar] [CrossRef]
  66. Husna, F.; Suyatna, F.D.; Arozal, W.; Poerwaningsih, E.H. Anti-Diabetic Potential of Murraya koenigii (L.) and its Antioxidant Capacity in Nicotinamide-Streptozotocin Induced Diabetic Rats. Drug Res. 2018, 68, 631–636. [Google Scholar] [CrossRef] [PubMed]
  67. Casmir, U.E.; Joshua, P.E.; Ukegbu, C.Y.; Ezeagu, C.U.; Nwodo, O.F.C. Antidiabetic potential of ethanol leaf extract of Bryophyllum pinnatum on alloxan-induced diabetic rats and their haematological profiles. Afr. J. Pharm. Pharmacol. 2017, 11, 526–533. [Google Scholar] [CrossRef] [Green Version]
  68. Nayak, P.; Kar, D.M.; Maharana, L. Antidiabetic activity of aerial parts of Argemone mexicana linn. In alloxan induced hyperglycaemic rats. Pharmacologyonline 2011, 1, 889–903. [Google Scholar]
  69. Rahman, A.H.; Biswas, M.C.; Islam, A.K.; Zaman, A.T. Assessment of traditional medicinal plants used by local people of monirampur thana under Jessore district of Bangladesh. Wudpecker J. Med. Plans 2013, 2, 99–109. [Google Scholar]
  70. Rahman, A.M.; Kabir, E.Z.M.F.; Islam, A.R.; Zaman, A.T.M.N. Medico-botanical investigation by the tribal people of Naogaon district, Bangladesh. J. Med. Plants 2013, 1, 147–196. [Google Scholar]
  71. Priya, C.L.; Rao, K.V.B. Ethanobotanical and current ethanopharmacological aspects of Argemone mexicana Linn: An overview. Int. J. Pharm. Sci. Res. 2012, 3, 2143. [Google Scholar]
  72. Chanda, P.; Gupta, N.; Kumari, A.; Bhattacharya, S. A review on pharmacological potential of Argemone mexicana in management of wound healing & antidiabetic activity. Pharmacol. Potential Med. Plants Manag. Differ. Dis. 2022, 11, 21–27. [Google Scholar]
  73. Prasad, M.; Venugopal, S.P. Preliminary phytochemical analysis and oral acute toxicity study of the root of Argemone mexicana linn. Int. J. Res. Dev. Pharm. Life Sci. 2016, 5, 2010–2017. [Google Scholar]
  74. Sarkar, K.K.; Mitra, T.; Acharyya, R.N.; Sadhu, S.K. Phytochemical screening and evaluation of the pharmacological activities of ethanolic extract of Argemone mexicana Linn. aerial parts. Orient. Pharm. Exp. Med. 2019, 19, 91–106. [Google Scholar] [CrossRef]
  75. Pahwa, R.; Chatterjee, V.C. The toxicity of Mexican poppy (Argemone mexicana L.) seeds to rats. Veter-Hum. Toxicol. 1989, 31, 555–558. [Google Scholar]
  76. Verma, S.; Dev, G.; Tyagi, A.; Goomber, S.; Jain, G. Argemone mexicana poisoning: Autopsy findings of two cases. Forensic Sci. Int. 2000, 115, 135–141. [Google Scholar] [CrossRef]
  77. Pingale, S.S. Toxicity study for Argemone mexicana L. World J. Pharm. Sci. 2013, 1, 151–155. [Google Scholar]
  78. Singh, R.; Geetanjali. Asparagus racemosus: A review on its phytochemical and therapeutic potential. Nat. Prod. Res. 2015, 30, 1896–1908. [Google Scholar] [CrossRef] [PubMed]
  79. Vadivelan, R.; Krishnan, R.G.; Kannan, R. Antidiabetic potential of Asparagus racemosus Willd leaf extracts through inhibition of α-amylase and α-glucosidase. J. Tradit. Complement. Med. 2018, 9, 1–4. [Google Scholar] [CrossRef]
  80. Selvaraj, K.; Sivakumar, G.; Veeraraghavan, V.P.; Dandannavar, V.S.; Veeraraghavan, G.R.; Rengasamy, G. Asparagus racemosus—A Review. Syst. Rev. Pharm. 2019, 10, 87–89. [Google Scholar]
  81. Hannan, J.M.A.; Ali, L.; Khaleque, J.; Akhter, M.; Flatt, P.R.; Abdel-Wahab, Y.H.A. Antihyperglycaemic activity of Asparagus racemosus roots is partly mediated by inhibition of carbohydrate digestion and absorption, and enhancement of cellular insulin action. Br. J. Nutr. 2012, 107, 1316–1323. [Google Scholar] [CrossRef] [Green Version]
  82. Hannan, J.M.A.; Marenah, L.; Ali, L.; Begum, R.; Peter, R.F.; Yasser, H.A.W. Insulin secretory actions of extracts of Asparagus racemosus root in perfused pancreas, isolated islets and clonal pancreatic beta-cells. J. Endocrinol. 2007, 192, 159–168. [Google Scholar] [CrossRef] [PubMed]
  83. Kumar, M.C.S.; Udupa, A.L.; Sammodavardhana, K.; Rathnakar, U.P.; Shvetha, U.; Kodancha, G.P. Acute toxicity and diuretic studies of the roots of Asparagus racemosus Willd in rats. West Indian Med. J. 2010, 59, 3–6. [Google Scholar] [PubMed]
  84. Goyal, R.K.; Singh, J.; Lal, H. Asparagus racemosus—An update. Indian J. Med. Sci. 2003, 57, 408–414. [Google Scholar] [PubMed]
  85. Rahmatullah, M.; Mollik, M.A.H.; Khatun, A.; Jahan, R.; Chowdhury, A.R.; Seraj, S.; Hossain, M.; Nasrin, D.; Khatun, Z. A survey on the use of medicinal plants by folk medicinal practitioners in five villages of Boalia sub-district, Rajshahi district, Bangladesh. Adv. Nat. Appl. Sci. 2010, 4, 39–44. [Google Scholar]
  86. Islas, J.F.; Acosta, E.; G-Buentello, Z.; Delgado-Gallegos, J.L.; Moreno-Treviño, M.G.; Escalante, B.; Moreno-Cuevas, J.E. An overview of Neem (Azadirachta indica) and its potential impact on health. J. Funct. Foods 2020, 74, 104171. [Google Scholar] [CrossRef]
  87. Rahmani, A.H.; Almatroudi, A.; Alrumaihi, F.; Khan, A.A. Pharmacological and therapeutic potential of neem (Azadirachta indica). Pharmacogn. Rev. 2018, 12, 250. [Google Scholar] [CrossRef]
  88. Pandey, G.; Verma, K.K.; Singh, M. Evaluation of phytochemical, antibacterial and free radical scavenging properties of Azadirachta indica (neem) leaves. Int. J. Pharm. Pharm. Sci. 2014, 6, 444–447. [Google Scholar]
  89. Arumugam, A.; Agullo, P.; Boopalan, T.; Nandy, S.; Lopez, R.; Gutierrez, C.; Narayan, M.; Rajkumar, L. Neem leaf extract inhibits mammary carcinogenesis by altering cell proliferation, apoptosis, and angiogenesis. Cancer Biol. Ther. 2014, 15, 26–34. [Google Scholar] [CrossRef] [Green Version]
  90. Kanagasanthosh, K.; Shanmugapriyan, S.; Kavirajan, V. Evaluation of acute toxicity, anti-inflammatory activity and phytochemical screening of ethanolic extract of Azadirachta indica leaves. Int. J. Res. Dev. Pharm. Life Sci. 2015, 4, 1737–1742. [Google Scholar]
  91. Satyanarayana, K.; Sravanthi, K.; Shaker, I.; Ponnulakshmi, R. Molecular approach to identify antidiabetic potential of Azadirachta indica. J. Ayurveda Integr. Med. 2015, 6, 165–174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Perez-Gutierrez, R.M.; Damian-Guzman, M. Meliacinolin: A Potent α-Glucosidase and α-Amylase Inhibitor Isolated from Azadirachta indica Leaves and In Vivo Antidiabetic Property in Streptozotocin-Nicotinamide-Induced Type 2 Diabetes in Mice. Biol. Pharm. Bull. 2012, 35, 1516–1524. [Google Scholar] [CrossRef] [Green Version]
  93. Rahman, A.H.M.M.; Hossain, M.M.; Islam, A.K.M.R. Taxonomy and Medicinal Uses of Angiosperm Weeds in the Wheat Field of Rajshahi, Bangladesh. Front. Biol. Life Sci. 2014, 2, 8. [Google Scholar] [CrossRef]
  94. Kundu, P.; Debnath, S.L.; Sadhu, S.K. Exploration of Pharmacological and Toxicological Properties of Aerial Parts of Blumea lacera, a Common Weed in Bangladesh. Clin. Complement. Med. Pharmacol. 2022, 2, 10038. [Google Scholar] [CrossRef]
  95. Hossen, A.; Reza, A.A.; Abu Ahmed, A.; Islam, K.; Jahan, I.; Hossain, R.; Khan, M.F.; Alam Maruf, M.R.; Haque, A.; Rahman, A. Pretreatment of Blumea lacera leaves ameliorate acute ulcer and oxidative stress in ethanol-induced Long-Evan rat: A combined experimental and chemico-biological interaction. Biomed. Pharmacother. 2021, 135, 111211. [Google Scholar] [CrossRef] [PubMed]
  96. Mishra, S.; Aeri, V.; Gaur, P.K.; Jachak, S.M. Phytochemical, Therapeutic, and Ethnopharmacological Overview for a Traditionally Important Herb:Boerhavia diffusa Linn. BioMed Res. Int. 2014, 2014, 808302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Oyebode, O.A.; Erukainure, O.L.; Chukwuma, C.I.; Ibeji, C.U.; Koorbanally, N.A.; Islam, S. Boerhaavia diffusa inhibits key enzymes linked to type 2 diabetes in vitro and in silico; and modulates abdominal glucose absorption and muscle glucose uptake ex vivo. Biomed. Pharmacother. 2018, 106, 1116–1125. [Google Scholar] [CrossRef]
  98. Hiruma-Lima, C.A.; Gracioso, J.S.; Bighetti, E.J.; Robineou, L.G.; Brito, A.S. The juice of fresh leaves of Boerhaavia diffusa L. (Nyctaginaceae) markedly reduces pain in mice. J. Ethnopharmacol. 2000, 71, 267–274. [Google Scholar] [CrossRef]
  99. Chaudhary, P.H.; Khadabadi, S.S. Bombax ceiba Linn.: Pharmacognosy, ethnobotany and phyto-pharmacology. Pharmacogn. Commun. 2012, 2, 2–9. [Google Scholar] [CrossRef] [Green Version]
  100. Wanjari, M.M.; Gangoria, R.; Dey, Y.N.; Gaidhani, S.N.; Pandey, N.K.; Jadhav, A.D. Hepatoprotective and antioxidant activity of Bombax ceiba flowers against carbon tetrachloride-induced hepatotoxicity in rats. Hepatoma Res. 2016, 2, 144. [Google Scholar] [CrossRef] [Green Version]
  101. Jan, H.T.U.; Zahra, S.S.; Nasir, B.; Baig, M.W.; Ahmed, M. Divulging the Antimicrobial and Antidiabetic Potential of Bombax ceiba L. J. Bioresour. Manag. 2017, 4, 1–6. [Google Scholar] [CrossRef]
  102. Xu, G.-K.; Qin, X.-Y.; Wang, G.-K.; Xie, G.-Y.; Li, X.-S.; Sun, C.-Y.; Liu, B.-L.; Qin, M.-J. Antihyperglycemic, antihyperlipidemic and antioxidant effects of standard ethanol extract of Bombax ceiba leaves in high-fat-diet- and streptozotocin-induced Type 2 diabetic rats. Chin. J. Nat. Med. 2017, 15, 168–177. [Google Scholar] [CrossRef] [PubMed]
  103. Yin, K.; Yang, J.; Wang, F.; Wang, Z.; Xiang, P.; Xie, X.; Sun, J.; He, X.; Zhang, X. A preliminary study of the chemical composition and bioactivity of Bombax ceiba L. flower and its potential mechanism in treating type 2 diabetes mellitus using ultra-performance liquid chromatography quadrupole-time-flight mass spectrometry and network pharmacology analysis. Front. Nutr. 2022, 9, 2519. [Google Scholar] [CrossRef]
  104. Kushwaha, P.S.; Raj, V.; Singh, A.K.; Keshari, A.K.; Saraf, S.A.; Mandal, S.C.; Yadav, R.K.; Saha, S. Antidiabetic effects of isolated sterols from Ficus racemosa leaves. RSC Adv. 2015, 5, 35230–35237. [Google Scholar] [CrossRef]
  105. Ca, J.; Dai, W.; Zhang, N. Advance on chemical constituents and pharmacological activities of Cajanus cajan (L.) Millsp. Nat. Prod. Res. Dev. 2020, 32, 515. [Google Scholar] [CrossRef]
  106. Khatun, S.; Pervin, F.; Karim, M.R.; Ashraduzzaman, M.; Rosma, A. Phytochemical screening and antimicrobial activity of Coccinia cordifolia L. plant. Pak. J. Pharm. Sci. 2012, 25, 25. [Google Scholar]
  107. Grover, J.; Yadav, S. Pharmacological actions and potential uses of Momordica charantia: A review. J. Ethnopharmacol. 2004, 93, 123–132. [Google Scholar] [CrossRef]
  108. Chhikara, N.; Kaur, R.; Jaglan, S.; Sharma, P.; Gat, Y.; Panghal, A. Bioactive compounds and pharmacological and food applications of Syzygium cumini—A review. Food Funct. 2018, 9, 6096–6115. [Google Scholar] [CrossRef] [PubMed]
  109. Bandiola, T.M.; Ignacio, G.B.; Yunson, E.G.; Bandiola, P.D. Syzygium cumini (L.) skeels: A review of its phytochemical constituents, toxicity studies, and traditional and pharmacological uses. Int. J. Appl. Pharm. Biol. Res. 2017, 2, 15–23. [Google Scholar]
  110. Gurjar, H.; Jain, S.K.; Irchhaiya, R.; Nandanwar, R.; Sahu, V.K.; Saraf, H. Hypoglycemic effects of methanolic extract of Anthocephalus cadamba bark in alloxan induced diabetic rats (Rox B) Miq. Int. J. Pharm. Sci. Res. 2010, 1, 79–83. [Google Scholar]
  111. Pandey, A.; Negi, P.S. Traditional uses, phytochemistry and pharmacological properties of Neolamarckia cadamba: A review. J. Ethnopharmacol. 2016, 181, 118–135. [Google Scholar] [CrossRef]
  112. Dubey, A.; Nayak, S.; Goupale, D.C. A review on phytochemical, pharmacological and toxicological studies on Neo-lamarckia cadamba. Pharm. Lett. 2011, 3, 45–54. [Google Scholar]
  113. Shah, K.; Patel, M.; Patel, R.; Parmar, P. Mangifera indica (Mango). Pharmacogn. Rev. 2010, 4, 42–48. [Google Scholar] [CrossRef]
  114. Ediriweera, M.K.; Tennekoon, K.H.; Samarakoon, S.R. A Review on Ethnopharmacological Applications, Pharmacological Activities, and Bioactive Compounds of Mangifera indica (Mango). Evid.-Based Complement. Altern. Med. 2017, 2017, 6949835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Lima, E.B.C.; Sousa, C.N.S.; Meneses, L.N.; Ximenes, N.C.; Júnior, M.A.S.; Vasconcelos, G.S.; Lima, N.B.C.; Patrocínio, M.C.A.; Macedo, D.; Vasconcelos, S.M.M. Cocos nucifera (L.) (Arecaceae): A phytochemical and pharmacological review. Braz. J. Med. Biol. Res. 2015, 48, 953–964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Meher, B.; Dash, D.K.; Roy, A. A review on: Phytochemistry, phamracology and tradiional uses of Tamarindus indica L. World J. Pharm. Pharm. Sci. 2014, 3, 229–240. [Google Scholar]
  117. Moga, M.; Dimienescu, O.; Bălan, A.; Dima, L.; Toma, S.; Bîgiu, N.; Blidaru, A. Pharmacological and Therapeutic Properties of Punica granatum Phytochemicals: Possible Roles in Breast Cancer. Molecules 2021, 26, 1054. [Google Scholar] [CrossRef] [PubMed]
  118. El-Far, A.H.; Shaheen, H.M.; Alsenosy, A.W.; El-Sayed, Y.S.; Al Jaouni, S.K.; Mousa, S. Costus speciosus: Traditional uses, phytochemistry, and therapeutic potentials. Pharmacogn. Rev. 2018, 12, 120–127. [Google Scholar] [CrossRef]
  119. Kumar, S.; Bhattacharya, A.; Tiwari, P.; Sahu, P. A review of the phytochemical and pharmacological characteristics of Moringa oleifera. J. Pharm. Bioallied Sci. 2018, 10, 181–191. [Google Scholar] [CrossRef]
  120. Xu, Y.B.; Chen, G.L.; Guo, M.Q. Antioxidant and anti-inflammatory activities of the crude extracts of Moringa oleifera from Kenya and their correlations with flavonoids. Antioxidants 2019, 8, 296–318. [Google Scholar] [CrossRef] [Green Version]
  121. Zainab, B.; Ayaz, Z.; Alwahibi, M.S.; Khan, S.; Rizwana, H.; Soliman, D.W.; Alawaad, A.; Abbasi, A.M. In-silico elucidation of Moringa oleifera phytochemicals against diabetes mellitus. Saudi J. Biol. Sci. 2020, 27, 2299–2307. [Google Scholar] [CrossRef]
  122. Gopukumar, S.T.; Praseetha, P.K. Ficus benghalensis Linn—The sacred indian medicinal tree with potent pharmacological remedies. Int. J. Pharm. Sci. Rev. Res. 2015, 32, 223–227. [Google Scholar]
  123. Gabhe, S.; Tatke, P.; Khan, T. Evaluation of the immunomodulatory activity of the methanol extract of Ficus benghalensis roots in rats. Indian J. Pharmacol. 2006, 38, 271. [Google Scholar] [CrossRef] [Green Version]
  124. Nisar, A.; Mamat, A.S.; Hatim, M.I.; Aslam, M.S.; Syarhabil, M. An updated review on Catharanthus roseus: Phytochemical and pharmacological analysis. Indian Res. J. Pharm. Sci. 2016, 3, 631–653. [Google Scholar]
  125. Logesh, R.; Das, N.; Adhikari-Devkota, A.; Devkota, H.P. Cocculus hirsutus (L.) W.Theob. (Menispermaceae): A Review on Traditional Uses, Phytochemistry and Pharmacological Activities. Medicines 2020, 7, 69. [Google Scholar] [CrossRef]
  126. Majumder, S.C.; Islam, K.; Hossain, M.M. Potential anti-diabetic medicinal plants in bangladesh: A comprehensive review. World J. Pharm. Res. 2019, 8, 140–150. [Google Scholar]
  127. Cheng, J.-X.; Zhang, B.-D.; Zhu, W.-F.; Zhang, C.-F.; Qin, Y.-M.; Abe, M.; Akihisa, T.; Liu, W.-Y.; Feng, F.; Zhang, J. Traditional uses, phytochemistry, and pharmacology of Ficus hispida L.f.: A review. J. Ethnopharmacol. 2019, 248, 112204. [Google Scholar] [CrossRef]
  128. Gupta, P.C. Biological and pharmacological properties of Terminalia chebula Retz. (Haritaki). Int. J. Pharm. Pharm. Sci. 2012, 4, 62–68. [Google Scholar]
  129. Anand, A.; Divya, N.; Kotti, P. An updated review of Terminalia catappa. Pharmacogn. Rev. 2015, 9, 93–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Azahar, N.I.; Mokhtar, N.M.; Arifin, M.A. Piper betle: A review on its bioactive compounds, pharmacological properties, and extraction process. IOP Conf. Series Mater. Sci. Eng. 2020, 991, 12044. [Google Scholar] [CrossRef]
  131. Majaz, A.Q.; Khurshid, M.; Nazim, S. The miracle plant (Kalanchoe pinnata ): A phytochemical and pharmacological review. Int. J. Res. Ayurveda Pharm. 2011, 2, 1478–1482. [Google Scholar]
  132. Paudel, K.R.; Panth, N. Phytochemical Profile and Biological Activity of Nelumbo nucifera. Evid.-Based Complement. Altern. Med. 2015, 2015, 789124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Mehta, N.R.; Patel, E.P.; Patani, P.V.; Shah, B. Nelumbo nucifera (Lotus): A Review on Ethanobotany, Phytochemistry and Pharmacology. Indian J. Pharm. Biol. Res. 2013, 1, 152–167. [Google Scholar] [CrossRef] [Green Version]
  134. Dewi, K.; Ridwanuloh, A.; Rahmat, J. Phytochemical screening, antibacterial and cytotoxic activity of Mikania cordata extracts. In Proceedings of the 2nd International Seminar on Chemistry; 2021; pp. 217–221. [Google Scholar]
  135. Gahlawat, D.K.; Jakhar, S.; Dahiya, P. Murraya koenigii (L.) Spreng: An ethnobotanical, phytochemical and pharmaco-logical review. J. Pharmacogn. Phytochem. 2014, 3, 109–119. [Google Scholar]
  136. Latif, A.; Ashiq, K.; Qayyum, M.; Ashiq, S.; Ali, E.; Anwer, I. Phytochemical and pharmacological profile of the medicinal herb: Bryophyllum pinnatum. J. Anim. Plant Sci. 2019, 29, 295–301. [Google Scholar]
  137. Ibitoye, O.B.; Olofinsan, K.A.; Teralı, K.; Ghali, U.M.; Ajiboye, T.O. Bioactivity-guided isolation of antidiabetic principles from the methanolic leaf extract of Bryophyllum pinnatum. J. Food Biochem. 2018, 42, e12627. [Google Scholar] [CrossRef]
  138. Afzal, M.; Kazmi, I.; Anwar, F. Antineoplastic potential of Bryophyllum pinnatum lam. on chemically induced hepatocarcinogenesis in rats. Pharmacogn. Res. 2013, 5, 247–253. [Google Scholar] [CrossRef] [Green Version]
  139. Tang, R.; Tian, R.-H.; Cai, J.-Z.; Wu, J.-H.; Shen, X.-L.; Hu, Y.-J. Acute and sub-chronic toxicity of Cajanus cajan leaf extracts. Pharm. Biol. 2017, 55, 1740–1746. [Google Scholar] [CrossRef] [Green Version]
  140. Ezike, A.C.; Akah, P.A.; Okoli, C.C.; Okpala, C.B. Experimental evidence for the antidiabetic activity of Cajanus cajan leaves in rats. J. Basic Clin. Pharm. 2010, 1, 81–84. [Google Scholar] [PubMed]
  141. Ohadoma, S.C.; Akpan, J.L.; Odey, P.A.; Okoro, E.P. Mechanistic considerations of Catharanthus roseus on the hypo-glycemic activity of alpha glucosidase inhibitors and biguanides: A review. J. Pharm. Adv. Res. 2021, 4, 1390–1398. [Google Scholar]
  142. Al-Shaqha, W.M.; Khan, M.; Salam, N.; Azzi, A.; Chaudhary, A.A. Anti-diabetic potential of Catharanthus roseus Linn. and its effect on the glucose transport gene (GLUT-2 and GLUT-4) in streptozotocin induced diabetic wistar rats. BMC Complement. Altern. Med. 2015, 15, 379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Vutukuri, V.R.; Das, M.C.; Reddy, M.; Prabodh, S.; Sunethri, P. Evaluation of Acute Oral Toxicity of Ethanol Leaves Extract of Catharanthus roseus in Wistar Albino Rats. J. Clin. Diagn. Res. 2017, 11, FF01–FF04. [Google Scholar] [CrossRef] [PubMed]
  144. Balaraman, A.K.; Singh, J.; Dash, S.; Maity, T.K. Antihyperglycemic and hypolipidemic effects of Melothria maderaspatana and Coccinia indica in Streptozotocin induced diabetes in rats. Saudi Pharm. J. 2010, 18, 173–178. [Google Scholar] [CrossRef] [Green Version]
  145. Putra, I.M.W.A.; Fakhrudin, N.; Nurrochmad, A.; Wahyuono, S. Antidiabetic activity of Coccinia grandis (L.) Voigt: Bioactive constituents, mechanisms of action, and synergistic effects. J. Appl. Pharm. Sci. 2021, 12, 41–54. [Google Scholar] [CrossRef]
  146. Jha, U.; Asad, M.; Asdaq, S.M.B.; Das, A.K.; Prasad, V.S. Fertility inducing effect of aerial parts of Coccinia cordifolia L. in female rats. J. Ethnopharmacol. 2010, 127, 561–564. [Google Scholar] [CrossRef] [PubMed]
  147. Bothara, S.; Marya, B.; Saluja, A.K. Antiarthritic activity of root extracts of Cocculus hirsutus. Int. J. Pharm. Pharm. Sci. 2011, 3, 175–177. [Google Scholar]
  148. Obidoa, O.; Joshua, P.E.; Eze, N.J. Phytochemical analysis of Cocos nucifera L. J. Pharm. Res. 2010, 3, 280–286. [Google Scholar]
  149. Salil, G.; Nevin, K.; Rajamohan, T. Arginine rich coconut kernel protein modulates diabetes in alloxan treated rats. Chem. Biol. Interact. 2011, 189, 107–111. [Google Scholar] [CrossRef]
  150. Paul, N.; Roy, R.; Bhattacharya, S.; Biswas, M. Acute and sub-chronic toxicity study of Cocos nucifera leaf extracts in mice. J. Adv. Pharm. Educ. Res. 2012, 2, 74–81. [Google Scholar]
  151. Waisundara, V.; Watawana, M.; Jayawardena, N. Costus speciosus and Coccinia grandis: Traditional medicinal remedies for diabetes. S. Afr. J. Bot. 2015, 98, 1–5. [Google Scholar] [CrossRef]
  152. Rani, A.S.; Sulakshana, G.; Patnaik, S. Costus speciosus, An antidiabetic plant-review. FS J. Pharm. Res. 2012, 1, 51–53. [Google Scholar]
  153. Muthukumaran, P.; Thiyagarajan, G.; Babu, R.A.; Lakshmi, B.S. Raffinose from Costus speciosus attenuates lipid synthesis through modulation of PPARs/SREBP1c and improves insulin sensitivity through PI3K/AKT. Chem. Interact. 2018, 284, 80–89. [Google Scholar] [CrossRef]
  154. Ahmed, F.; Chavan, S.; Satish, A.; Punith, K.R. Inhibitory activities of Ficus benghalensis bark against carbohydrate hydrolyzing enzymes—An in vitro study. Pharmacogn. J. 2011, 3, 33–37. [Google Scholar] [CrossRef] [Green Version]
  155. Khanal, P.; Patil, B.M. Consolidation of network and experimental pharmacology to divulge the antidiabetic action of Ficus benghalensis L. bark. 3 Biotech 2021, 11, 238. [Google Scholar] [CrossRef]
  156. Swathi, N.; Sreedevi, A.; Bharathi, K. Evaluation of Nephroprotective Activity of Fruits of Ficus hispida on Cisplatin-Induced Nephrotoxicity. Pharmacogn. J. 2011, 3, 62–68. [Google Scholar] [CrossRef] [Green Version]
  157. Ahmed, F.; Urooj, A. In vitro studies on the hypoglycemic potential of Ficus racemosa stem bark. J. Sci. Food Agric. 2009, 90, 397–401. [Google Scholar] [CrossRef] [PubMed]
  158. Rao, R.B.; Murugesan, T.; Sinha, S.; Saha, B.P.; Pal, M.; Mandal, S.C. Glucose lowering efficacy of Ficus racemosa bark extract in normal and alloxan diabetic rats. Phytotherapy Res. 2002, 16, 590–592. [Google Scholar] [CrossRef]
  159. Morshed, A.; Hossain, M.H.; Shakil, S.; Nahar, K.; Rahman, S.; Ferdausi, D.; Hossain, T.; Ahmad, I.; Chowdhury, M.H.; Rahmatullah, M. Evaluation of antinociceptive activity of two Bangladeshi medicinal plants, Kalanchoe pinnata (Lam.) Pers. and Lagerstroemia speciosa (L.). Pers. Adv. Nat. Appl. Sci. 2010, 4, 193–197. [Google Scholar]
  160. George, L.O.; Radha, H.R.; Somasekariah, B.V. In vitro anti-diabetic activity and GC-MS analysis of bioactive compounds present in the methanol extract of Kalanchoe pinnata. Indian J. Chem.-Sect. B 2019, 57, 1213–1221. [Google Scholar]
  161. John, O.R.; Yahaya, A.A.; Emmanuel, A. Aqueous ethanolic extract of Mangifera indica stem bark effect on the biochemical and haematological Parameters of Albino Rats. Arch. Appl. Sci. Res. 2012, 4, 1618–1622. [Google Scholar]
  162. Katagiri, H.; Asano, T.; Ishihara, H.; Inukai, K.; Shibasaki, Y.; Kikuchi, M.; Yazaki, Y.; Oka, Y. Overexpression of Catalytic Subunit p110α of Phosphatidylinositol 3-Kinase Increases Glucose Transport Activity with Translocation of Glucose Transporters in 3T3-L1 Adipocytes. J. Biol. Chem. 1996, 271, 16987–16990. [Google Scholar] [CrossRef] [Green Version]
  163. Sangeetha, K.N.; Sujatha, S.; Muthusamy, V.S.; Anand, S.; Nithya, N.; Velmurugan, D.; Balakrishnan, A.; Lakshmi, B.S. 3β-taraxerol of Mangifera indica, a PI3K dependent dual activator of glucose transport and glycogen synthesis in 3T3-L1 adipocytes. Biochim. Biophys. Acta-Gen. Subj. 2010, 1800, 359–366. [Google Scholar] [CrossRef]
  164. Deb, L.; Bhattacharjee, C.; Shetty, S.R.; Dutta, A. Evaluation of anti-diabetic potential of the Syzygium cumini (Linn) Skeels by reverse pharmacological approaches. Bull. Pharm. Res. 2012, 3, 135–145. [Google Scholar]
  165. Abukakar, M.G.; Ukwuani, A.N.; Shehu, R.A. An evaluation of the toxic effects of Tamarindus indica pulp extract in albino rats. J. Pharmacol. Toxicol. 2008, 3, 111–118. [Google Scholar] [CrossRef] [Green Version]
  166. Bhandary, B.S.K.; Sharmila, K.P.; Kumari, N.S.; Bhat, S.V. Acute and subacute toxicity study of the ethanol extracts of Punica granatum (linn). Whole fruit and seeds and synthetic ellagic acid in swiss albino mice. Asian J. Pharm. Clin. Res. 2013, 6, 192–198. [Google Scholar]
  167. Kasolo, J.N.; Bimenya, G.S.; Ojok, L.; Ogwal-okeng, J.W. Phytochemicals and acute toxicity of Moringa oleifera roots in mice. J. Pharmacogn. Phyther. 2011, 3, 38–42. [Google Scholar]
  168. Panunto, W.; Jaijoy, K.; Lerdvuthisopon, N.; Lertprasertsuke, N.; Jiruntanat, N.; Soonthornchareonnon, N.; Sireeratawong, S. Acute and chronic toxicity studies of the water extract from dried fruits of Terminalia chebula Rezt. in rats. Int. J. Appl. Res. Nat. Prod. 2010, 3, 36–43. [Google Scholar]
  169. Lokman, M.A.; Arshad, A.M.; Ahmad, W.M.A.W.; Effendy, A.W.M. Determination of toxicological effects of Terminalia catappa leaves on Sprague-Dawley white rats in short-term period. Int. J. Toxicol. Appl. Pharm. 2013, 3, 44–47. [Google Scholar]
  170. Al-Adhroey, A.H.; Nor, Z.M.; Al-Mekhlafi, H.M.; Amran, A.A.; Mahmud, R. Antimalarial Activity of Methanolic Leaf Extract of Piper betle L. Molecules 2010, 16, 107–118. [Google Scholar] [CrossRef] [PubMed]
  171. Umasankar, K.; Suresh, V.; Kumar, R.M.; Suresh, A.; Kumar, N.S.; Arunachalam, G. CNS Activity of Ethanol Extract of Wedelia chinensis in Experimental Animals. Int. J. Pharm. Sci. Nanotechnol. 2010, 3, 881–886. [Google Scholar] [CrossRef]
  172. Darvekar, V.M.; Patil, V.R.; Choudhari, A.B. Anti-inflammatory activity of Murraya koenigii Spreng on experimental animals. J. Nat. Prod. Plant Resour. 2011, 1, 65–69. [Google Scholar]
  173. Sathishsekar, D.; Subramanian, S. Beneficial Effects of Momordica charantia Seeds in the Treatment of STZ-Induced Diabetes in Experimental Rats. Biol. Pharm. Bull. 2005, 28, 978–983. [Google Scholar] [CrossRef] [Green Version]
  174. Oyelere, S.F.; Ajayi, O.H.; Ayoade, T.E.; Pereira, G.B.S.; Owoyemi, B.C.D.; Ilesanmi, A.O.; Akinyemi, O.A. A detailed review on the phytochemical profiles and anti-diabetic mechanisms of Momordica charantia. Heliyon 2022, 8, e09253. [Google Scholar] [CrossRef] [PubMed]
  175. Anwer, T.; Safhi, M.M.; Makeen, H.A.; Alshahrani, S.; Siddiqui, R.; Sivakumar, S.; Shaheen, E.S.; Alam, M.F. Antidiabetic potential of Moringa oleifera Lam. leaf extract in type 2 diabetic rats, and its mechanism of action. Trop. J. Pharm. Res. 2021, 20, 95–103. [Google Scholar] [CrossRef]
  176. Ahmad, J.; Khan, I.; Blundell, R. Moringa oleifera and glycemic control: A review of current evidence and possible mechanisms. Phytotherapy Res. 2019, 33, 2841–2848. [Google Scholar] [CrossRef] [PubMed]
  177. Sharma, B.R.; Kim, M.S.; Rhyu, D.Y. Nelumbo nucifera leaf extract attenuated pancreatic ß-cells toxicity induced by interleukin-1ß and interferon-γ, and increased insulin secrection of pancreatic ß-cells in streptozotocin-induced diabetic rats. J. Tradit. Chin. Med. 2016, 36, 71–77. [Google Scholar] [CrossRef] [Green Version]
  178. Avijit, B.; Zerin, T.; Rajia, S. Comparative Phytochemical and Antibacterial Properties of Piper betle Leave Extracts from Barguna and Moheshkhali, Bangladesh. Iran. J. Med. Microbiol. 2020, 14, 125–132. [Google Scholar] [CrossRef]
  179. Yogeswari, S.; Bindu, K.H.; Kamalraj, S.; Ashokkumar, V.; Jayabaskaran, C. Antidiabetic, Antithrombin and Cytotoxic bioactive compounds in five cultivars of Piper betle L. Environ. Technol. Innov. 2020, 20, 101140. [Google Scholar] [CrossRef]
  180. Ocvirk, S.; Kistler, M.; Khan, S.; Talukder, S.H.; Hauner, H. Traditional medicinal plants used for the treatment of diabetes in rural and urban areas of Dhaka, Bangladesh—An ethnobotanical survey. J. Ethnobiol. Ethnomedicine 2013, 9, 43. [Google Scholar] [CrossRef] [Green Version]
  181. Huang, T.H.; Peng, G.; Kota, B.P.; Li, G.; Yamahara, J.; Roufogalis, B.; Li, Y. Anti-diabetic action of Punica granatum flower extract: Activation of PPAR-γ and identification of an active component. Toxicol. Appl. Pharmacol. 2005, 207, 160–169. [Google Scholar] [CrossRef]
  182. Tang, D.; Liu, L.; Ajiakber, D.; Ye, J.; Xu, J.; Xin, X.; Aisa, H.A. Anti-diabetic Effect of Punica granatum Flower Polyphenols Extract in Type 2 Diabetic Rats: Activation of Akt/GSK-3β and Inhibition of IRE1α-XBP1 Pathways. Front. Endocrinol. 2018, 9, 586. [Google Scholar] [CrossRef] [PubMed]
  183. Sharma, A.K.; Bharti, S.; Kumar, R.; Krishnamurthy, B.; Bhatia, J.; Kumari, S.; Arya, D.S. Syzygium cumini Ameliorates Insulin Resistance and β-Cell Dysfunction via Modulation of PPARγ, Dyslipidemia, Oxidative Stress, and TNF-α in Type 2 Diabetic Rats. J. Pharmacol. Sci. 2012, 119, 205–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Kuru, P. Tamarindus indica and its health related effects. Asian Pac. J. Trop. Biomed. 2014, 4, 676–681. [Google Scholar] [CrossRef] [Green Version]
  185. Sole, S.S.; Srinivasan, B. Aqueous extract of tamarind seeds selectively increases glucose transporter-2, glucose transporter-4, and islets’ intracellular calcium levels and stimulates β-cell proliferation resulting in improved glucose homeostasis in rats with streptozotocin-induced diabetes mellitus. Nutr. Res. 2012, 32, 626–636. [Google Scholar] [CrossRef] [PubMed]
  186. Costa, I.; Lima, M.; Medeiros, A.; Bezerra, L.; Santos, P.; Serquiz, A.; Lima, M.; Oliveira, G.; Santos, E.; Maciel, B.; et al. An Insulin Receptor-Binding Multifunctional Protein from Tamarindus indica L. Presents a Hypoglycemic Effect in a Diet-Induced Type 2 Diabetes—Preclinical Study. Foods 2002, 11, 2207. [Google Scholar] [CrossRef]
  187. Mallik, J.; AI, F.A.; Kumar, B.R. A comprehensive review on Pharmacological activity of Terminalia Catappa (Com-bretaceae)—An Update. Asian J. Pharm. Res. Dev. 2010, 1, 65–70. [Google Scholar]
  188. Iheagwam, F.N.; Iheagwam, O.T.; Onuoha, M.K.; Ogunlana, O.O.; Chinedu, S.N. Terminalia catappa aqueous leaf extract reverses insulin resistance, improves glucose transport and activates PI3K/AKT signalling in high fat/streptozotocin-induced diabetic rats. Sci. Rep. 2022, 12, 10711. [Google Scholar] [CrossRef]
  189. Biswas, K.R.; Ishika, T.; Rahman, M.; Swarna, A.; Khan, T.; Monalisa, M.N.; Rahmatullah, M. Antidiabetic plants and formulations used by folk medicinal practitioners of two villages in Narail and Chuadanga districts, Bangladesh. Am. J. Sustain. Agric. 2011, 5, 158–167. [Google Scholar]
  190. Borgohain, R.A.; Lahon, K.I.; Das, S.W.; Gohain, K.A. Evaluation of mechanism of anti-diabetic activity of Terminalia chebula on alloxan and adrenaline-induced diabetic albino rats. Drugs 2012, 3, 256–266. [Google Scholar]
  191. Das, M.P.; Kumar, S.S. Preliminary phytochemical analysis of Illicium verum and Wedelia chinensis. Int. J. Pharmtech Res. 2013, 5, 324–329. [Google Scholar]
  192. Thao, N.P.; Binh, P.T.; Luyen, N.T.; Hung, T.M.; Dang, N.H.; Dat, N.T. α-Amylase and α-Glucosidase Inhibitory Activities of Chemical Constituents from Wedelia chinensis (Osbeck.) Merr. Leaves. J. Anal. Methods Chem. 2018, 2018, 2794904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Ansari, P.; Akther, S.; Hannan, J.M.A.; Seidel, V.; Nujat, N.J.; Abdel-Wahab, Y.H.A. Pharmacologically Active Phytomolecules Isolated from Traditional Antidiabetic Plants and Their Therapeutic Role for the Management of Diabetes Mellitus. Molecules 2022, 27, 4278. [Google Scholar] [CrossRef]
  194. Ghafouri, A.; Hajiluian, G.; Karegar, S.J.; Hosseini, S.; Shidfar, S.; Kamalinejad, M.; AghaHosseini, F.; Heydari, I.; Shidfar, F. The effect of Aqueous, Ethanolic extracts of Rheum ribeson insulin sensitivity, inflammation, oxidative stress in patients with type 2 diabetes mellitus: A Randomized, Double-Blind, Placebo-Controlled Trial. J. Herb. Med. 2020, 24, 100389. [Google Scholar] [CrossRef]
  195. Aziz, T.A.; Hussain, S.A.; Mahwi, T.O.; Ahmed, Z.A.; Rahman, H.S.; Rasedee, A. The efficacy and safety of Ginkgo biloba extract as an adjuvant in type 2 diabetes mellitus patients ineffectively managed with metformin: A double-blind, randomized, placebo-controlled trial. Drug. Des. Dev. Ther. 2018, 12, 735–742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Wickramasinghe, A.S.D.; Kalansuriya, P.; Attanayake, A.P. Nanoformulation of Plant-Based Natural Products for Type 2 Diabetes Mellitus: From Formulation Design to Therapeutic Applications. Curr. Ther. Res. 2022, 96, 100672. [Google Scholar] [CrossRef]
  197. Márquez-Escobar, V.A.; Rosales-Mendoza, S.; Beltrán-López, J.I.; González-Ortega, O. Plant-based vaccines against respiratory diseases: Current status and future prospects. Expert. Rev. Vaccines 2016, 16, 137–149. [Google Scholar] [CrossRef]
  198. Chen, S.L.; Yu, H.; Luo, H.M.; Wu, Q.; Li, C.F.; Steinmetz, A. Conservation and sustainable use of medicinal plants: Problems, progress, and prospects. Chinese Med. 2016, 11, 37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Achyranthes aspera.
Figure 1. Achyranthes aspera.
Diabetology 04 00019 g001
Figure 2. Aloe barbadensis.
Figure 2. Aloe barbadensis.
Diabetology 04 00019 g002
Figure 3. Amaranthus tricolor.
Figure 3. Amaranthus tricolor.
Diabetology 04 00019 g003
Figure 4. Anacardium occidentale.
Figure 4. Anacardium occidentale.
Diabetology 04 00019 g004
Figure 5. Andrographis paniculata.
Figure 5. Andrographis paniculata.
Diabetology 04 00019 g005
Figure 6. Annona squamosa.
Figure 6. Annona squamosa.
Diabetology 04 00019 g006
Figure 7. Argemone mexicana.
Figure 7. Argemone mexicana.
Diabetology 04 00019 g007
Figure 8. Asparagus racemosus.
Figure 8. Asparagus racemosus.
Diabetology 04 00019 g008
Figure 9. Azadirachta indica.
Figure 9. Azadirachta indica.
Diabetology 04 00019 g009
Figure 10. Blumera lacera.
Figure 10. Blumera lacera.
Diabetology 04 00019 g010
Figure 11. Boerhaavia diffusa.
Figure 11. Boerhaavia diffusa.
Diabetology 04 00019 g011
Figure 12. Bombax ceiba.
Figure 12. Bombax ceiba.
Diabetology 04 00019 g012
Figure 13. Bryophyllum pinnatum.
Figure 13. Bryophyllum pinnatum.
Diabetology 04 00019 g013
Figure 14. Cajanus cajan.
Figure 14. Cajanus cajan.
Diabetology 04 00019 g014
Figure 15. Catharanthus roseus Linn.
Figure 15. Catharanthus roseus Linn.
Diabetology 04 00019 g015
Figure 16. Coccinia cordifolia.
Figure 16. Coccinia cordifolia.
Diabetology 04 00019 g016
Figure 17. Cocculus hirsutus.
Figure 17. Cocculus hirsutus.
Diabetology 04 00019 g017
Figure 18. Cocos nucifera.
Figure 18. Cocos nucifera.
Diabetology 04 00019 g018
Figure 19. Costus speciosus.
Figure 19. Costus speciosus.
Diabetology 04 00019 g019
Figure 20. Ficus benghalensis.
Figure 20. Ficus benghalensis.
Diabetology 04 00019 g020
Figure 21. Ficus hispida.
Figure 21. Ficus hispida.
Diabetology 04 00019 g021
Figure 22. Ficus racemosa.
Figure 22. Ficus racemosa.
Diabetology 04 00019 g022
Figure 23. Kalanchoe pinnata.
Figure 23. Kalanchoe pinnata.
Diabetology 04 00019 g023
Figure 24. Mangifera indica.
Figure 24. Mangifera indica.
Diabetology 04 00019 g024
Figure 25. Mikania cordata.
Figure 25. Mikania cordata.
Diabetology 04 00019 g025
Figure 26. Momordica korola.
Figure 26. Momordica korola.
Diabetology 04 00019 g026
Figure 27. Moringa oleifera.
Figure 27. Moringa oleifera.
Diabetology 04 00019 g027
Figure 28. Murraya koenigii.
Figure 28. Murraya koenigii.
Diabetology 04 00019 g028
Figure 29. Nelumbo nucifera.
Figure 29. Nelumbo nucifera.
Diabetology 04 00019 g029
Figure 30. Neolamarckia cadamba.
Figure 30. Neolamarckia cadamba.
Diabetology 04 00019 g030
Figure 31. Piper betle.
Figure 31. Piper betle.
Diabetology 04 00019 g031
Figure 32. Punica granatum.
Figure 32. Punica granatum.
Diabetology 04 00019 g032
Figure 33. Syzygium cumini.
Figure 33. Syzygium cumini.
Diabetology 04 00019 g033
Figure 34. Tamarindus indica.
Figure 34. Tamarindus indica.
Diabetology 04 00019 g034
Figure 35. Terminalia catappa.
Figure 35. Terminalia catappa.
Diabetology 04 00019 g035
Figure 36. Terminalia chebula.
Figure 36. Terminalia chebula.
Diabetology 04 00019 g036
Figure 37. Wedelia chinensis.
Figure 37. Wedelia chinensis.
Diabetology 04 00019 g037
Table 1. Summary of the Bangladeshi diabetes plants with utilized parts, model organisms, solvents, dosages, reference drugs and others.
Table 1. Summary of the Bangladeshi diabetes plants with utilized parts, model organisms, solvents, dosages, reference drugs and others.
Scientific NameCommon Name FamilyParts UsedModel OrganismSolventDosageReference DrugResultReferences
Ficus racemosaCluster fig, DumurMoraceae Leaves STZ-induced diabetic rats Petroleum ether 200 mg/kg body weight and 400 mg/kg body weight Glibenclamide Reduction in blood glucose levels (p < 0.001) [34]
Asparagus racemosusSatamuliLiliaceaeWhole plant STZ-induced diabetic Wistar rats Ethanol 200 and 400 mg/kg body weight Glibenclamide 10 mg/kg body weightDecreased blood glucose level (p < 0.05)[35]
Bombax ceibaShimul, red silk-cotton treeBombacaceaeBark STZ-induced diabetic rats Ethyl-acetate extract600 mg/kg body weight Glibenclamide (10 mg/kg body weight) Displayed significant hypoglycemic activity (p < 0.001)[36]
Cajanus cajanArhar, pigeon peaFabaceaeRoot Alloxan-induced diabetic mice Methanol 200 mg/kg body weight or 400 mg/kg body weight Metformin (50 mg/kg body weight) Significantly lowered fasting serum glucose (p < 0.001) and blood glucose level (p < 0.001)[37]
Coccinia cordifoliaTelakucha, Ivy gourd CucurbitaceaeLeaves STZ-induced diabetic rats Petroleum ether, ethyl acetate150 mg/kg body weight Metformin HCl (150 mg/kg body weight) Reduced blood glucose levels by 50.39% and 50% at the 10th and 24th hours, respectively (p < 0.001) [38]
Momordica charantiaKorola, bitter gourdCucurbitaceaeSeed STZ-induced diabetic rats Water 150 mg/kg body weight Glibenclamide (600 μg/kg) Significantly reduced plasma glucose (p < 0.05)[39]
Syzygium cuminiKalojam, Java plumMyrtaceaeWhole fruit,
Leaves
STZ-induced diabetic albino Wistar rats
STZ-induced diabetic rats
Ethanol

Ethanol
200 mg/kg body weight
250 mg/kg body weight
Glibenclamide (200 mg/kg body weight)Decreased blood glucose levels (p < 0.01)
Significantly decreased blood glucose concentration, increased muscle glycogen store (p < 0.001)
[40,41]
Neolamarckia cadambaBurflower tree, KodomRubiaceaeLeaves Glucose-loaded hyperglycemic mice Methanol 200 mg/kg body weight and 400 mg/kg body weight10 mg/kg body weight of glibenclamide Reduced elevated blood glucose (p < 0.05)[42]
Mangifera indicaMango, Aam AnacardiaceaeLeaves Alloxan-induced diabetic rats Water 200 mg/kg body weight and 400 mg/kg body weightGliclazide 2 mg/kg body weight and 4 mg/kg body weight Significantly lowered the fasting blood glucose (p < 0.0001)[43]
Cocos nuciferaCoconut Arecaceae Spadix STZ (STZ)-induced diabetic ratsHydromethanol 250 mg/kg body weight and 500 mg/kg body weight 0.5 mg/kg body weight glibenclamide Significantly lowered fasting blood glucose (p < 0.001)[44]
Tamarindus indicaTamarind FabaceaeSeed STZ-induced diabetic Long–Evans rats -1.25 g/kg Glibenclamide (5 mg/kg body weight) Demonstrated anti-hyperglycemic properties (p < 0.03) [45]
Punica granatumPomegranate LythraceaeLeaves
Fruit
STZ-induced type 2 diabetic rats
Alloxan-induced diabetic Wistar rats
Methanol


Water
100, 200, 400 and 600 mg/kg body weight
100 mg/kg, 200 mg/kg and 350 mg/kg body weight
Glibenclamide (1 mg/kg body weight)
-
Demonstrated anti-diabetic activity(p < 0.05)

Reduced fasting blood glucose (p < 0.001)
[46,47]
Azadirachta indicaNeemMeliaceaeLeaves STZ (STZ)-induced diabetic rats Ethanol 200 mg/kg body weightGlibenclamide (0.25 mg/kg body weight) Normalized glucose levels after STZ-induced hyperglycemia (p < 0.05)[48]
Costus speciosusCrepe GingeCostaceae Rhizome
Root

Alloxan-induced diabetic rats
Ethanol 200 mg/kg body weight

300 and 450 mg/kg body weight
Glibenclamide (2.5 mg/kg body weight)
Glibenclamide (600 μg/kg)
Enhanced insulin secretion
lowered blood glucose concentration, decreased glyconeogenesis, increased glycogenesis
[49,50]
Moringa oleiferaDrumstick treeMoringaceaeLeaves STZ-induced Sprague-Dawley ratsEthyl acetate200 mg/kg body weightGlibenclamide (5 mg/kg body weight) Significantly reverses the effects of STZ on serum glucose and insulin, demonstrating anti-diabetic activity (p < 0.05)[51]
Andrographis paniculataGreen chirettaAcanthaceaeleavesAlloxan-induced diabetic Wistar ratsEthanol 1000 mg/kg body weight Glimepiride (4 mg/kg) body weightSignificantly reduced blood glucose (p < 0.01) [28]
Ficus benghalensisBanyan MoraceaeLeaves
Bark
Alloxan-induced diabetic albino rats
STZ-induced diabetic rats
Ethanol


Ethanol
200 mg/kg body weight and 400 mg/kg body weight
150 mg/kg body weight, 300 mg/kg body weight, 500 mg/kg body weight
Glibenclamide (5 mg/kg body weight)

Glibenclamide 0.5 mg/kg body weight
Reduced triglyceride, cholesterol and
glucose levels
Reduction in blood glucose levels by stimulating insulin secretion (p < 0.01)
[52,53]
Anacardium occidentaleCashew nut, kaju badam SumacLeavesSTZ-induced diabetic Wistar ratsMethanol 300 mg/kg body weight 1 IU/kg body weight insulin Significantly reduced blood glucose (p < 0.05) [24]
Annona squamosaCustard appleAnnonaceaeLeaves STZ-induced rats Water and ethanol 200 mg/kg body weight, 400 mg/kg body weightMetformin (300 mg/kg body weight) Reduced blood glucose (p < 0.05) [31]
Boerhaavia diffusaHogweedNyctaginaceae Leaves STZ-induced diabetic ratsChloroform50, 100, 200 mg/kg body weight Glibenclamide9 (25 µg/kg body weight) Reduced blood glucose, possibly through rejuvenation of pancreatic β-cells [54]
Catharanthus roseusNayantaraApocynaceaeLeaf Alloxan-induced diabetic rat Juice 0.5, 0.75, 1.0 mL/kg body weight Glibenclamide (40 μg/kg body weight) Lowered blood glucose (p < 0.01)[55]
Cocculus hirsutusBroom creeper, Daikhali, JalajmaniMenispermaceaeLeafAlloxan-induced diabetic rat Water 250, 500, 1000 mg/kg body weightGlyburide (10 mg/kg body weight) Reduced serum glucose level (p < 0.01)[56]
Ficus hispidaHairy fig MoraceaeBark Alloxan-induced diabetic albino rat Ethanol 1.25 g/kg body weight Glibenclamide (0.5 mg/kg body weight)Reduced blood glucose (p < 0.001)[57]
Terminalia chebulaMyrobalanCombretaceaeFruit STZ-induced diabetic rats Ethanol 200 mg/kg body weight Glibenclamide (600 μg/kg body weight)Stimulated insulin and lowered blood glucose (p < 0.05) [58]
Terminilia catappaBangla badamCombretaceaeLeaf Alloxan-induced diabetic Wistar albino ratsWater, cold extract Aqueous extract (43 mg/kg body weight)
Cold extract (46 mg/kg body weight)
Glibenclamide (10 mg/kg body weight)Regenerated pancreas [59]
Amaranthus tricolorEdible amaranthAmaranthaceaeLeaf Alloxan-induced diabetic male albino rats Water 400 mg/kg body weight-Reduced serum glucose levels (p < 0.001) [21]
Blumea laceraLettuce leaf blumeaAsteraceaeLeaf Swiss albino miceMethanol 50–400 mg/kg body weightGlibenclamide (10 mg/kg body weight)Reduced blood glucose (p < 0.0001)[60]
Piper betle leavesBetel leafPiperaceaeLeaf STZ-induced diabetic ratsEthanol 100, 200, 300, 1500 mg/kg body weightTolbutamide (22.5 mg/kg body weight)Reduced blood glucose significantly (p < 0.050)[61]
Achyranthes asperaDevil’s horsewhipAmaranthaceaeSeed STZ-induced diabetic rat Ethanol300, 600 mg/kg body weight for 24 days Glibenclamide (5 mg/kg body weight) Reduced blood glucose level (p < 0.001)[11]
Kalanchoe pinnataCathedral bells, life plantCrassulaceaeLeaf Alloxan-induced diabetic rats Ethanol 5.8, 11.6 and 33.2 mg/kg body weight Glibenclamide (1.35 mg/kg body weight), acarbose (13.5 mg/kg body weight) Significantly powered fasting blood glucose (p < 0.05), increased number of pancreatic β-cells [62]
Nelumbo nuciferaIndian lotusNymphaeaceaeFlower STZ-induced diabetic male Wistar ratEthanol 250 mg/kg body weight Glibenclamide (0.25 mg/kg body weight) Lowered fasting blood glucose [63]
Mikania cordataBitter vineAsteraceaeLeaf Alloxan-induced diabetic ratEthanol200 mg/kg, 400 mg/kg body weight Metformin hydrochloride (100 mg/kg body weight) Reduced blood glucose (p < 0.05, p < 0.01) [64]
Wedelia chinensisWedeliaAsteraceaeLeafAlloxan-induced diabetic Swiss albino miceMethanole 100 mg/kg, 200 mg/kg body weight Glibenclamide (5 mg/kg body weight)Reduced blood glucose levels (p < 0.01)[65]
Murraya koenigiiCurry treeRutaceaeLeaf STZ-induced and nicotinamide-induced hyperglycemic rats Ethaol 200 mg/kg, 400 mg/kg body weight Glibenclamide (1 mg/kg body weight) Decreased blood glucose (p < 0.01) [66]
Aloe barbadensisAloeAsphodelaceaeGel Alloxan-induced diabetic rat Methanol300 mg/kg body weight Glibenclamide (0.25 mg/kg body weight) Reduced blood glucose level (p < 0.05) [15]
Bryophyllum pinnatumGoetheCrassulaceaeLeaf Alloxan-induced diabetic rat Ethanol 200 mg/kg, 400 mg/kg body weight Glibenclamide (2.5 mg/kg body weight) Reduced blood glucose level (p < 0.05) [67]
Agremone mexicanMexican prickly poppy, SialkantaPapaveraceaeAerial parts Alloxan-induced diabetic rat Ethanol, water 200 mg/kg, 400 mg/kg body weight Glibenclamide (5 mg/kg body weight) Demonstrated significant anti-diabetic effect (p < 0.05) [68]
Table 2. List of plants with potential hypoglycemic compounds.
Table 2. List of plants with potential hypoglycemic compounds.
Name of PlantPotential Hypoglycemic CompoundsReferences
Ficus racemosaβ-Sitosterol, stigmasterol, lanosterol, flavonoids (gluanol acetate and racemosic acid)[104]
Asparagus racemosaFlavonoids, Saponins-shatavarin IV (asparanin B), shatavarin V, shatavarin VI, shatavarin VII, shatavarin VIII, immunoside, schidigerasaponin D5 (asparanin A), racemoside A, racemoside B and racemoside C, triterpenoids, alkaloids-asparagamine A, quercetin, quercetin glycosides- quercetin-3-O-rutinoside, quercetin 3-O-galactoside, isoflavones, sapogenins[78]
Bombax ceibaalkaloids, flavonoids, glycosides, β-sitosterol, kaempferol, hentriacontane, hentriacontanol, quercetin, shamimin (2-(2,4,5-trihydroxyphenyl)-3,5,7-trihydroxy-6-C-glucopyranosyloxy-4H-1-benzopyran-4-one)[99]
Cajanus cajanΒ-sitosterol, flavonoids- luteonin, vitexin, apigenin, genistic, ononin, sissiotrin, 2′-hydroxygenistein, stilbenes-cajanusin A, cajanusin B, cajanusin C, cajanusin D, cajanstilbene H, cajanolactone A, cajanonic acid A, canjanotone[105]
Coccinia cordifoliasaponin, glycoside, alkaloid compounds (catharanthin, leurosine, lochnerine, vindoline and vindolinine), flavonoids, tannins, phenols, terpenoids [38,106]
Momordica charantiaDiosgenin, saponins, alkaloids, trtiterpenes, proteins, steroids[107]
Syzygium cuminialkaloids, tanins, flavonoids, carotenoids, glycosides, saponins, steroids, triterpenoids, anthocyanins, phenols, oxalic acid, phytosterols, myrcetin, gallic acid [108,109]
Neolamarckia cadamaPhenolics, alkaloids, Cadambine, Chlorogenic acid, Dihydrocinchonine, flavonoids, saponins, triterpenes, indole alkaloids, triterpenoid glycosides, 3α-dihydrocadambine, isodihydrocadambine, isocadamine[110,111,112]
Mangifera indicaMangiferin, flavonoids, phenolic acids, xanthones, gallic acid, catechins, kaempferol, carotenoids-luteoxanthine, zeaxantine, β-carotene, terpenoids- careen, myrcene, terpinoline, terpenoid saponins- indicoside A and B[113,114]
Cocos nuciferaphenols, tannins, β-sitosterol, flavonoids, noctinic acid, folic acid, riboflavin, biotin, triterpenes, alkaloids, steroids, saponins, tannins, catechins, epicatechins[115]
Tamarindus indicaphenolic compounds, β-sitosterol, proanthocyanins, apigenins tartaric acid, cardiac glycosides, mucilage, pectin, eicosanoic acid, β-sitosterol, palmitic acid, oleic acid, succininc acid, formic acid, β-amyrin, apigenin, epicatechins, catechins, taxifolin, eriodictoyl, naringenin[116]
Punica granatumflavonols, triterpenoids, fatty acids, organic acids- citric acid, malic acid, ascorbic acid, tannins- gallagic acid, punicalin, punicalagin, flavonoids-luteolin, quercetin, kaempferol, magnesium, alkaloids, catechin, gallic acid[117]
Azadirachta indicaphenols, flavonoids, saponins, tannins, alkaloids, glycosides, carbohydrates, triterpenoids, β-sitosterol, ferulic acid[87,88]
Costus speciosusalkaloids, glycosides, flavonoids, steroids, polyphenols, tannins, β-sitosterol, gracillin, dioscin, diosgenin[118]
Moringa oleiferaβ-sitosterol, saponins, steroids, alkaloids- moringine and moringinine, flavonoids- rhamnetin, isoquercitrin, kaempferitrin, saccharides, phenolic acids, tannins, terpenoid, alpha-carotene [119,120,121]
Andrographis paniculataterpenoids, flavonoids, iridoids, ferulic acid, diterpenoid lactones, andrographolide, 14-deoxyandrographolide, 14-Deoxy-11,12-dehydroandrographolide, fersulic acids[27]
Ficus benghalensisanthocyanidin derivatives, aliphatic long-chain ketones, glycosides, flavonoids, amino acids, steroids, saponins, carbohydrates, tannins[122,123]
Anacardium occidentalephenolics, saponin, flavonoids, alkaloids, anthocyanidins, tannins, essential oils, glycosides, myricetin, pentoside, lactone, quercetin hexoside, canthones, chalcones, 4-hydroxydodecanoic acid, palmitate, sitosterol, stigmasterol, 3-O-βD-galactopyranoside[23]
Annona squamosaflavonoids, alkaloids, phenols, saponins, tannins, glycosides, diterpenes[30]
Boerhaavia diffusasteroids, ecdysteroid, alkaloids, lignan glycosides, phenolic glycosides, flavonoids, isoflavonoids, rotenoids[96]
Catharanthus roseusalkaloids, catharanthine, tetrahydroalstonine, vindoline, kaempferol, lochnerine, flavonoids[124]
Cocculus hirsutusβ-sitosterol, ginnol, flavonoids such as luteolin, kaempferol and quercetin, glycosides, carbohydrates, tannins, saponins, steroids[125]
Ficus hispidaflavonoids, saponins, steroids, glycosides, alkaloids, alkanes[126,127]
Terminalia chebulaΒ-sitosterol, flavonoids, tannins, sterols, gallic acid, chebulanin, corilagin, ellagic acid, chebulinic acid, amino acids, fructose, resin, triterpenoids, glycosides[128]
Terminilia catappatannins, saponins, phenolics, flavonoids, triterpenoids, kaempferol, geraniin, punicalin, quercetin, gentisic acid, tercatain, tergallagin, β-carotene, cyanidin-3-glucoside, ellagic acid, gallic acid[59,129]
Amaranthus tricolorflavonoids, amino acids, alkaloids, carbohydrates, saponins, phenolic compounds, tannins, phytic acid [19,20]
Blumea laceraβ-sitosterol, artemesinin, lupeol, β-caryophyllene, protocatechuic acid, alkaloids, flavonoids, tannins, terpenoids, flavones, triterpenes[94]
Piper betle leavesglycosides, quercitin, alkaloids, saponins, steroids, tannins, diterpenes, eugenol, chavibetol, flavonoids, hydroxychavicol[130]
Achyranthes asperaAlkaloids, oleanolic acid, saponins, D-Glucuronic Acid, dihydroxy ketones, quercitin, β-sitosterol, aliphatic alcohol, benzoquinone, hydroquinone, asarone and eugenol[10]
Kalanchoe pinnataβ-sitosterol, flavonoids, kaempferol, quercetin, alkaloids, tannins, phenolic compounds, caffeic acid, syringic acid, luteolin, rutin, para-coumaric acid, ferulic acid, stigmaterol, astragalin, campesterol[131]
Nelumbo nuciferaasteroidal triterpenoid, alkaloids, phenolic bases, flavonoids, quercetin, glycoside, kaempferol, nuciferin, roemerin, armepavine, β-sitosterol glucopyranoside[132,133]
Mikania cordatasaponins, alkaloids, flavonoids, tannins, steroids[134]
Wedelia chinensisflavonoids, alkaloids, saponins, phytosterols, mucilage, tannins
Murraya koenigiicarbazole alkaloids, mahanimbine, flavonoids, sterols, koenimbine, koenine, girinimbine [135]
Aloe barbadensisflavonoids, terpenoids, polysaccharides, pectins, hemicelluloses, glucomannan, sterols β-sitosterol, lupeol, tannins[14]
Bryophyllum pinnatumflavonoids, alkaloids, reducing sugars, tannins, bufadienolides, glycosaponins, polyphenols, steroidal glycosides[136]
Agremone mexicanaBerberine, cheilanthifoline, coptisine, cryptopine, sanguinarine, stylopine, tetrahydroberberine, protopine, sanguinarine, dihydrosanguinarine, palmitic acid, oleic acid, myristic acid, linoleic acids, β-sitosterol, amino acids, fatty acids, tannin, saponins, flavonoids, phytosterols[71,72,73]
Table 3. Summary of the key studies conducted with Bangladeshi diabetes plants.
Table 3. Summary of the key studies conducted with Bangladeshi diabetes plants.
Name of PlantSolventPart of PlantModelHighest Safe Dose with No Toxicity (mg/kg Body Weight)LD50 (Median Lethal Dose) (mg/kg Body Weight)References
Ficus racemosaMethanol Bark Normal and alloxan-induced diabetic rat 3200 mg/kg 3200 mg/kg [158]
Asparagus racemosusWaterRoots Rats 3200 mg/kg [83]
Bombax ceibaWaterFlower Wistar rats 2000 mg/kg [100]
Cajanus cajanWater, ethanol Leaf Sprague–Dawley rats 6000 mg/kg [139]
Coccinia cordifoliaWater Aerial parts Wistar rats 5000 mg/kg [146]
Momordica charantiaEthanol Fruit Sprague–Dawley rats 2000 mg/kg [39]
Syzygium cuminiMethanol, water Bark, root, seed, leaf Albino mice 2000 mg/kg
(seed extract-200 mg/kg)
[164]
Neolamarckia cadamaMethanol Bark Mouse 3000 mg/kg [112]
Mangifera indicaWater Bark Rat >5000 mg/kg[161]
Cocos nuciferaPetroleum ether, chloroform, methanol Leaf Swiss albino rat2000 mg/kg [150]
Tamarindus indicaWater Pulp Albino rats 4500 mg/kg [165]
Punica granatumEthanol Whole fruit, seeds Swiss albino mice 2000 mg/kg [166]
Azadirachta indicaEthanol Leaf Swiss albino mice 2000 mg/kg [90]
Costus speciosusEthanol Rhizome Albino mice 5000 mg/kg [57]
Moringa oleiferaEthanol, water Roots Swiss albino mice Ethanolic extract—1780 mg/kg; aqueous extract—1590 mg/kg [167]
Andrographis paniculataEthanol Leaf Wistar rats 4000 mg/kg [28]
Ficus benghalensisMethanol Roots Rats >2000 mg/kg[123]
Anacardium occidentaleWater, ethanol Leaf Rats 2000 mg/kg [26]
Annona squamosaMethanol Leaf Rat >5000 mg/kg [33]
Boerhaavia diffusaJuice Leaf Mice 5000 mg/kg [98]
Catharanthus roseusEthanol Leaf Wistar rat 2000 mg/kg [143]
Cocculus hirsutusWater, methanol Root Wistar albino rat >3000 mg/kg [147]
Ficus hispidaMethanol Fruit Wistar albino rat 1000 mg/kg [156]
Terminalia chebulaWater Fruit Rat 5000 mg/kg [168]
Terminilia catappaWater Leaf Sprague-Dawley rats 3000 mg/kg [169]
Amaranthus tricolorWater Root Wistar albino rat 2000 mg/kg [19]
Blumea laceraMethanol Leaf Rat 5000 mg/kg [95]
Piper betle leavesMethanol Leaf Mice 5000 mg/kg [170]
Achyranthes asperaMethanol Leaf Swiss albino mice8000 mg/kg [13]
Kalanchoe pinnataEthanol Leaf Wistar rat>2000 mg/kg [131]
Nelumbo nuciferaEthanol Flower Wistar rat2000 mg/kg [63]
Mikania cordataEthanol Leaf Rat2000 mg/kg [64]
Wedelia chinensisEthanol Whole plant Mice >1600 mg/kg [171]
Murraya koenigiiPetroleum ether, chloroform, ethanol Leaf Rat 2500 mg/kg [172]
Aloe barbadensisMethanol Gel Rat 2000 mg/kg [15]
Bryophyllum pinnatumEthanol, water Aerial parts Swiss albino mice 2000 mg/kg [138]
Agremone mexicanaAqueous slurry of root-bark powder Root bark Swiss albino mice 7000 mg/kg [77]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zanzabil, K.Z.; Hossain, M.S.; Hasan, M.K. Diabetes Mellitus Management: An Extensive Review of 37 Medicinal Plants. Diabetology 2023, 4, 186-234. https://doi.org/10.3390/diabetology4020019

AMA Style

Zanzabil KZ, Hossain MS, Hasan MK. Diabetes Mellitus Management: An Extensive Review of 37 Medicinal Plants. Diabetology. 2023; 4(2):186-234. https://doi.org/10.3390/diabetology4020019

Chicago/Turabian Style

Zanzabil, Khwaja Zohura, Md. Sabbir Hossain, and Md. Kamrul Hasan. 2023. "Diabetes Mellitus Management: An Extensive Review of 37 Medicinal Plants" Diabetology 4, no. 2: 186-234. https://doi.org/10.3390/diabetology4020019

Article Metrics

Back to TopTop