Next Article in Journal
Palmatine, a Bioactive Protoberberine Alkaloid Isolated from Berberis cretica, Inhibits the Growth of Human Estrogen Receptor-Positive Breast Cancer Cells and Acts Synergistically and Additively with Doxorubicin
Next Article in Special Issue
Stereoselective Synthesis of the Di-Spirooxindole Analogs Based Oxindole and Cyclohexanone Moieties as Potential Anticancer Agents
Previous Article in Journal
Deficiency in Androgen Receptor Aggravates Traumatic Brain Injury-Induced Pathophysiology and Motor Deficits in Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phytochemistry, Ethnopharmacological Uses, Biological Activities, and Therapeutic Applications of Cassia obtusifolia L.: A Comprehensive Review

1
Department of Physiology and Pharmacology, Hotchkiss Brain Institute and Alberta Children’s Hospital Research Institute, Cumming School of Medicine, University of Calgary, Calgary, AB T2N 4N1, Canada
2
Department of Prescriptionology, College of Korean Medicine, Kyung Hee University, 26, Kyunghee dae-ro, Dongdaemun-gu, Seoul 02447, Korea
3
Qgenetics, Seoul Bio Corporation Center, 504, 23 Kyunghee Dae-ro, Dongdaemun-gu, Seoul 02447, Korea
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(20), 6252; https://doi.org/10.3390/molecules26206252
Submission received: 23 September 2021 / Revised: 12 October 2021 / Accepted: 13 October 2021 / Published: 15 October 2021

Abstract

:
Cassia obtusifolia L., of the Leguminosae family, is used as a diuretic, laxative, tonic, purgative, and natural remedy for treating headache, dizziness, constipation, tophobia, and lacrimation and for improving eyesight. It is commonly used in tea in Korea. Various anthraquinone derivatives make up its main chemical constituents: emodin, chrysophanol, physcion, obtusifolin, obtusin, au rantio-obtusin, chryso-obtusin, alaternin, questin, aloe-emodin, gluco-aurantio-obtusin, gluco-obtusifolin, naphthopyrone glycosides, toralactone-9-β-gentiobioside, toralactone gentiobioside, and cassiaside. C. obtusifolia L. possesses a wide range of pharmacological properties (e.g., antidiabetic, antimicrobial, anti-inflammatory, hepatoprotective, and neuroprotective properties) and may be used to treat Alzheimer’s disease, Parkinson’s disease, and cancer. In addition, C. obtusifolia L. contributes to histamine release and antiplatelet aggregation. This review summarizes the botanical, phytochemical, and pharmacological features of C. obtusifolia and its therapeutic uses.

1. Introduction

Cassia (family Caesalpiniaceae) is a large tropical genus with ~600 species of herbs, shrubs, and trees. Cassia obtusifolia (sicklepod) Linn., a member of the genus Cassia (Leguminosae), is a well-known traditional Chinese medicinal plant. It belongs to the medically and economically important family Leguminosae (syn. Fabaceae; subfamily Caesalpinioideae). C. obtusifolia L. is found mainly in China, Korea, India, and the western tropical regions. It is an annual semi-shrubby herb that ranges in height from ~0.5 to 2 m. It has two or three pairs of round-tipped leaflets with one to three flowers on a short axillary peduncle with pedicels up to 2 cm; the yellow petals (0.8–1.5 cm) wilt by midday. The pods are linear (up to 20 cm in length), curve gently downward, and contain numerous shiny, dark brown seeds (~0.5 cm in length). The seeds of C. obtusifolia L. are rhomboidal or slightly flat, with linear concave ramps on each side. Cassia tora L. is considered synonymous with C. obtusifolia L., but differs in its botanical and morphological characteristics [1,2]. The main distinguishing morphological feature between the two is the seed coat, which is marked with an obliquely symmetrical dented line on each side of the rib (C. obtusifolia L.) or has broad bands on both sides of the rib (C. tora L.).
Cassia species are of medicinal interest because of their therapeutic value in traditional medicine. The dry seeds are processed as a crude drug for clinical use or as a dietary supplement. The cultured plants are important sources of Semen Cassiae-derived commercial products in the market. C. obtusifolia L. seeds are a well-known medicinal plant in East Asia and are consumed as food to clear liver heat, sharpen vision, lubricate the intestines, and promote bowel movement [3]. In Korea, dried and roasted Cassia seeds are frequently used in brewing tea. In traditional oriental and Chinese (Juemingzi in Chinese) medicine, C. obtusifolia L. has been used to treat lacrimation, headaches, dizziness, and constipation [3,4]. C. obtusifolia L. has several pharmacological properties, including antiplatelet aggregation, antidiabetic, antimicrobial, anti-inflammatory, hepatoprotective, and neuroprotective activities, and may be used to treat Alzheimer’s disease, Parkinson’s disease, and cancer [5,6,7,8,9,10,11,12]. It also contributes to histamine release and antiplatelet aggregation. The whole plant, as well as its roots, flowers, leaves, seeds, and pods, possesses medicinal properties. A summary of the ethnomedicinal uses of different parts of the plant is provided in Table 1. This review herein summarizes progress regarding the chemical analysis of C. obtusifolia L., primarily focusing on the development of the phytochemistry, botanical aspects, ethnopharmacological, and pharmacological effects of C. obtusifolia L. C. obtusifolia L. species are rich sources of different types of anthraquinones and naphthopyrone derivatives that exhibit a number of biological activities and may potentially impact human health. Unfortunately, C. obtusifolia L. has not been developed as a pharmaceutical agent. The main objective of this review is to present a summary of the studies published to date on this promising plant, with a solid platform to design and conduct clinical studies. This paper reviews the phytochemical and pharmacological activities of C. obtusifolia L. and discusses its potential uses as a human food source and/or a pharmacological agent.

2. Phytochemistry

Several classes of bioactive metabolites have been identified from C. obtusifolia L., including anthraquinones, terpenoids, flavonoids, and lipids [1,10,19]. The main plant chemicals include anthraquinone, emodin, chrysophanol, physcion, obtusifolin, obtusin, aurantio-obtusin, chryso-obtusin, alaternin, questin, aloe-emodin, gluco-aurantio-obtusin, gluco-obtusifolin, chrysophanol-2-O-tetraglucoside, chrysophanol-2-O-triglucosides, and chryso-obtusin-2-glucoside [2,5,6,7,8,9,10,11,12,19]. Other components include naphthopyrone glycosides, toralactone-9-β-gentiobioside, toralactone gentiobioside, cassiaside, rubrofusarin-6-O-gentiobiosideol, rubrofusarin-6-β-gentiobioside, cassiaside C, cassiaside B2, cassiaside C2, xanthones (1,8-dihydroxy-3-methoxy-6-methylxanthone, isogentisin, 1,7-dihydroxy-3-methylxanthone, euxanthone, 1,3,6-trihydroxy-8-methylxanthone), triterpenoids (lupeol, betulinic acid, α-amyrin, sterols, polyketide, steroids, fatty esters), and toralactone [1,17]. The chemical structures of the main compounds are presented in Figure 1. Research on C. obtusifolia L. reveals that the nature and number of phytochemicals vary according to climate. Researchers have found that the whole C. obtusifolia L. plant (seeds, twigs, leaves, and roots) is rich in free and bound anthraquinones, although the quantities differ markedly. In general, anthraquinone content is higher in seeds and less abundant in other components. The following section discusses the phytochemical contents of the various plant parts.

2.1. The Whole Plant

Analysis of the whole C. obtusifolia L. plant indicates the presence of various anthraquinones and naphthopyrones: aloe-emodin, emodin, 1,2-dihydroxyanthraquinone, obtusin, chryso-obtusin, aurantio-obtusin, gluco-obtusifolin, gluco-aurantio-obtusin, gluco-chryso-obtusin, 1-desmethylaurantio-obtusin-2-O-β-d-glucopyranoside, 1-desmethyl-obtusin, aurantio-obtusin-6-O-β-d-glucopyranoside, 1-desmethylaurantio-obtusin, alaternin-1-O-β-d-glucopyranoside, chryso-obtusin-2-O-β-d-glucopyranoside, physicon-8-O-β-d-glucoside, obtusifolin, O-methyl-chrysophanol, emodin-1-O-β-gentio-bioside, chrysophanol-1-O-β-gentiobioside, chrysophanol-1-O-β-d-glucopyranosyl-(13)-β-d-glucopyranosyl-(1→6)-β-d-glucopyranoside, physcion-8-O-β-glucoside, 1,3-dihydroxy-8-methylanthraquinone, torosachrysone, 1-methylaurantio-obtusin-2-O-β-d-glucopyranoside, 1-desmethylchryso-obtusin, chrysophanic, acid, physcion, chrysophanol-10,10′-bianthrone, physcion-8-O-β-gentiobioside, and questin [20].

2.2. Seeds

Cassia obtusifolia seeds are composed of 1–2% anthraquinones, 5–7% fats, 14–19% protein, and 66–99% carbohydrates [21]. In addition to proteins and fats, the seeds also contain a gum of commercial interest [22]. As much as 41% of the seed is extractable [23]. Several anthraquinone compounds and glycosides have been isolated from the methanol extract of the seeds; examples include anthraquinone, chrysophanol, physcion, emodin, obtusifolin, obtusin, questin, chryso-obtusin, gluco-obtusifolin, aloe-emodin, alaternin, aurantio-obtusin, gluco-aurantio obtusin, chrysophanol tetraglucoside, 2-hydroxyemodin-1 methylether, chryso-obtusin-2-glucoside, chrysophanol triglucoside, 1,2-dihydroxyanthraquinone, 1,4-dihydroxyanthraquinone, 1,8-dihydroxyanthraquinone, 1,8-dihydroxy-3-methylanthraquinone, naphthopyrone glycoside, toralactone gentiobioside, cassiaside, and the naphthalene glycoside cassitoroside [7,10]. Torosachrysone and naphthalenic lactones, isotoralactone, cassialactone, three benzyl-β-resorcylates (2-benzyl-4,6-dihydroxy benzoic acid, 2-benzyl-4,6-dihydroxy benzoic acid-6-O-β-d-glucopyranoside, and 2-benzyl-4,6-dihydroxy benzoic acid-4-O-β-d-glucopyranoside), a new sodium salt of anthraquinone (sodium emodin-1-O-β-gentiobioside), chrysophanol-1-O-β-d-glucopyranosyl-(1–3)-β-d-glucopyranosyl-(1–6)-β-d-glucopyranoside, rubrofusarin-6-O-β-d-gentiobioside, obtusifolin-2-O-β-d-glucopyranoside, aurantio-obtusin-6-O-β-d-glucopyranoside, physcion-8-O-β-d-glucopyranoside,1-hydroxyl-2-acetyl-3,8-dimethoxy-6-O-β-d-apiofuranosyl-(1–2)-β-d-glucosylnaphthalene, toralactone-9-O-β-d-gentiobioside, and rubrofusarin-6-O-β-d-apiofuranosyl-(1–6)-O-β-d-glucopyranoside have also been isolated from C. obtusifolia L. seeds [24,25,26]. In addition, three acetylated anthraquinone glycosides (obtusifoline-2-O-β-d-2,6-di-O-acetylglucopyranoside, obtusifoline-2-O-β-d-3,6-di-O-acetylglucopyranoside, and obtusifoline-2-O-β-d-4,6-di-O-acetylglucopyranoside) have been isolated from the ethanolic extract of the seeds [27]. Recently, Pang et al. [28,29] have isolated four new compounds from the seeds of C. obtusifolia obtusifolin-2-O-β-d-(6′-O-α, β-unsaturated butyryl)-glucopyranoside, epi-9-dehydroxyeurotinone-β-d-glucopyranoside, obtusinaphthalenside A, and obtusinaphthalenside B. Feng et al. [30] also purified various monosaccharides, and polysaccharides from the water extract of C. obtusifolia L.

2.3. Leaves

The leaves of C. obtusifolia L. contain anthraquinones, xanthones, polyketide, steroids, triterpenoids, and fatty esters [17]. The methanol extract of the leaves contains aloe emodin, emodin, 1,8-dihydroxy-3-methoxy-6-methylxantone, euxanthone, chrysophanol, physcion, 1,2,8-trihydroxy-6,7-dimethoxyanthraquinone,1,7-dihydroxy-3-methoxyxanthone,1,5-dihydroxy-3-methoxy-7-methylanthraquinone,3,7-dihydroxy-1-methoxyxanthone,1-O-methylchrysophanol, 8-O-methylchrysophanol, 1,3,6-trihydroxy-8-methylxanthone, 1-hydroxy-7-methoxy-3-methylanthraquinone, and obtusifolin. The ethyl acetate extract contains (4R*,5S*,6E,8Z)-ethyl-4-([E]-but-1-enyl)-5-hydroxypentdeca-6,8-dienoate, (24S)-24-ethylcholesta-5,22(E),25-trien-3β-ol, (–)-acetoxy-9,10-dimethyl-1,5-octacosanolide, friedelin, stigmasterol, lupeol, and (E)-eicos-14-enoic acid [17]. A single phytoalexin was isolated and purified from 12- to 14-day-old leaves [31].

2.4. Roots

The hairy roots of C. obtusifolia L. contain betulinic acid, sitosterol, stigmasterol, anthraquinones, chrysophanol, physicon, 1-hydroxy-7-methoxy-3-methylanthraquinone, 8-O-methylchrysophanol, 1-O-methylchrysophanol, 1,2,8-trihydroxy-6,7-dimethoxyanthraquinone, emodin, iso-landicin, helminthosporin, obtusifolin, aloe-emodin, and xanthorin [20,32].

3. Bioactivity

Numerous researchers have investigated the pharmacological activities of various C. obtusifolia L. extracts. Table 2 summarizes the pharmacological features that have been observed. They include: antidiabetic, anti-inflammatory, antimicrobial, antioxidant, hepatoprotective, neuroprotective, immune-modulatory, anti-Parkinson’s disease, anti-Alzheimer’s disease, and larvicidal properties. The anthraquinones and naphthopyrones isolated from C. obtusifolia L. are structurally diverse and exhibit multiple pharmacological properties, which suggests that these compounds contribute to its therapeutic effects (Table 3). C. obtusifolia L. and its major constituents display a vast number of biological activities (Figure 2). Natural products are highly promising sources for antioxidant and anti-inflammatory agents. A wide range of bioactive constituents of plants have antioxidant and anti-inflammatory activities. Based on various assay methods and activity indices, antioxidant or anti-inflammatory activities and nutraceutical and therapeutic effects of traditional Chinese medicines as well as the mechanisms underlying such activities and effects have been investigated. The generation of free radicals can result in damage to the cellular machinery. The seeds of C. obtusifolia L. are widely used in Chinese folk medicine and have been demonstrated to exhibit significant antioxidant and anti-inflammatory. Over the past century, natural products, especially anthraquinone compounds, have become valuable products for achieving chemical diversity in the molecules used for inflammation relief. In addition, COE has traditionally been used in Korea to treat eye inflammation, photophobia, and lacrimation.

3.1. Neuroprotective Activity

Various studies have demonstrated the direct neuroprotective activities of the C. obtusifolia L. seed extract (COE) and its major constituents (anthraquinones). More detailed studies are required to clarify the compositional features and neuroprotective activities of the anthraquinones. The ethanolic COE (25, 50, or 100 mg/kg) ameliorates scopolamine or bilateral common carotid artery occlusion (2VO)-induced memory impairment by inhibiting acetylcholinesterase [8]. COE (10 or 50 mg/kg/day) reduced memory impairment and neuronal damage caused by 2VO in a mouse model of transient global ischemia; it was suggested that the neuroprotective effects of COE are attributable to its anti-inflammatory properties resulting in decreased expression of inducible nitric oxide synthase (iNOX) and cyclooxygenase-2 (COX-2) and increased expression of the neurotrophic factors pCREB and BDNF [33]. Alaternin, the active compound in C. obtusifolia L., exhibits neuroprotective activity after transient cerebral hypoperfusion induced by bilateral common carotid artery occlusion. Administration of alaternin (10 mg/kg) prevented or reduced nitrotyrosine and lipid peroxidation, bilateral common carotid artery occlusion (BCCAO)-induced iNOS expression, and microglial activation [48]. Drever et al. [11] reported that ethanolic COE is neuroprotective against NMDA-induced calcium dysregulation and 3-nitropropionic acid-induced cell death in mouse hippocampal cultures. Recently, Paudel et al. [56] also reported that four major compounds (cassiaside, rubrofusarin gentiobioside, aurantio-obtusin, and 2-hydroxyemodin 1-methylether) exhibited neuroprotective effects; among them, aurantio-obtusin showed promising neuroprotective effects via targeting various G-protein-coupled receptors and transient brain ischemia/reperfusion injury C57BL/6 mice model.

3.1.1. Anti-Alzheimer’s Disease Activity

The effects of the ethanolic extract of COE in Aβ-induced anti-Alzheimer’s disease (anti-AD) models have been reported. The mechanism of COE ameliorated Aβ-induced LTP impairment in acute hippocampal slices and prevented Aβ-induced GSK-3β activation [35]. Moreover, COE prevented microglial activation as well as iNOS and COX activation induced by Aβ in the hippocampus, and in vivo studies have indicated that COE ameliorated Aβ-induced object recognition memory impairment [35]. Two anthraquinones from C. obtusifolia L., obtusifolin and gluco-obtusifolin, improved scopolamine-induced learning and memory impairment in mice based on the passive avoidance and Morris water maze tests [49]. Obtusifolin (0.25, 0.5, and 2 mg/kg) and gluco-obtusifolin (1, 2, and 4 mg/kg) significantly reversed scopolamine-induced cognitive impairment on the passive avoidance test; obtusifolin (0.5 mg/kg) and gluco-obtusifolin (2 mg/kg) improved escape latencies, swimming times in the target quadrant, and crossing numbers in the zone where the platform previously existed on the Morris water maze test [49]. The anti-AD properties of COE may be attributed to its constituents, such as anthraquinones and naphthopyrone glycosides. The methanolic seed extract and its solvent-soluble fractions from C. obtusifolia L. were tested for their acetylcholinesterase (AChE) and butyrylcholinesterase (BChE) inhibitory activities using Elman’s method. Ethyl acetate and butanol fractions significantly inhibited AChE activity at a final concentration of 100 µg/mL, with IC50 values of 9.45 ± 0.44 and 9.87 ± 0.70 μg/mL, respectively. Butanol (IC50 = 7.58 ± 0.51 μg/mL) and ethyl acetate (IC50 = 16.09 ± 0.16 μg/mL) fractions exhibited potent inhibitory activity against BChE. Furthermore, butanol fraction (IC50 = 26.19 ± 0.72 μg/mL) significantly inhibited the β-secretase (BACE1) activity [10]. In addition, several anthraquinones (emodin, chrysophanol, physcion, obtusifolin, alaternin, questin, aloe-emodin) that displayed strong anti-AD activity by inhibiting AChE, BChE, and BACE1 enzymes were isolated from this plant [10]. Recently, Shrestha et al. [55] observed anti-AD effects of naphthopyrone and its glycosides including rubrofusarin, rubrofusarin 6-O-β-d-glucopyranoside, rubrofusarin 6-O-β-d-gentiobioside, nor-rubrofusarin 6-O-β-d-glucoside, isorubrofusarin 10-O-β-d-gentiobioside, and rubrofusarin 6-O-β-d-triglucoside by inhibiting AChE, BChE, and BACE1 enzymes. The use of AChE, BChE, and BACE1 inhibitors has been a promising treatment strategy for AD; therefore, C. obtusifolia may be an effective agent for treating AD.

3.1.2. Prevention and Treatment of Parkinson’s Disease

A neuroprotective effect of COE was observed in both in vitro and in vivo models of Parkinson’s disease [34]. In PC12 cells, COE reduced cell damage induced by 100 µM 6-hydroxydopamine and inhibited the overproduction of reactive oxygen species, glutathione depletion, mitochondrial membrane depolarization, and caspase-3 activation at 0.1 to 10 µg/mL. In addition, COE displayed radical scavenging effects in DPPH and ABTS assays, which suggests that COE may be useful for treating Parkinson’s disease [34].

3.2. Hepatoprotective Activity

Few studies have demonstrated the hepatoprotective activities of COE [15]. Further studies are required to establish the hepatoprotective mechanisms of major COE anthraquinones. The protective effects of ethanolic COE against the cytotoxicity induced by CCl4 liver in mice were evaluated by assessing aminotransferase activities, histopathological changes, hepatic and mitochondrial antioxidant indices, and cytochrome P450 2E1(CYP2E1) activity. Administration of COE (0.5, 1, 2 g/kg) markedly reduced ALT and AST release, Ca2+-induced mitochondria membrane permeability transition, and CYP2E1 activity. In addition, COE significantly reduced hepatic and mitochondrial malondialdehyde levels, increased hepatic and mitochondrial glutathione levels, and restored superoxide dismutase, glutathione reductase, and glutathione S-transferase activities [15]. Meng et al. [38] reported the hepatoprotective effects of ethanolic COE on non-alcoholic fatty liver disease (NAFLD). Administration of COE (0.5, 1, 2 g/kg) markedly reduced the levels of AST, ALT, TG, TC, TNF-a, IL-6, IL-8, and MDA. COE treatments also increased the levels of SOD, GSH, and the expression of LDL-R mRNA [38]. Seo et al. [12] observed hepatoprotective effects of ethanolic COE and its components (e.g., toralactone glycoside) in t-BHP-induced cell death in HepG2 cells. Cassia anthraquinones, aurantio-obtusin, and obtusifolin also protected against tacrine-induced cytotoxicity in HepG2 cells [36]. Recently, Ali et al. [37] investigated the hepatoprotective effects of different soluble fractions of methanolic derived COE and its active components in t-BHP-induced oxidative stress in HepG2 cells. The possible mechanism was that alaternin, aloe emodin, and cassiaside potently scavenge ROS in t-BHP-induced HepG2 cells and the decrease in ROS generation parallels the up-regulation of glutathione (GSH). Very recently, Paudel et al. [57] investigated the hepatoprotective activity of an anthraquinone (1-desmethylaurantio-obtusin 2-O-β-d-glucopyranoside) and two naphthopyrone glycosides (rubrofusarin 6-O-β-d-apiofuranosyl-(1→6)-O-β-d-glucopyranoside and rubrofusarin 6-O-β-gentiobioside) isolated from the butanol fraction of COE in the t-BHP-induced oxidative stress in HepG2 cells through up-regulated HO-1 via the nuclear factor erythroid-2-related factor 2 (Nrf2) activation and modulation of the JNK/ERK/MAPK signaling pathway.

3.3. Anti-Inflammatory and Antioxidant Activity

COE has traditionally been used in Korea to treat eye inflammation, photophobia, and lacrimation. Pretreatment with the aqueous extract of C. obtusifolia L. inhibited interleukin (IL)-6 and cyclooxygenase-2 (COX-2) and reduced the activation of transcription nuclear factor-kB p65 in colon tissues treated with dextran sulfate sodium [40]. Two major anthraquinones from C. obtusifolia, obtusifolin and gluco-obtusifolin, reduced neuropathic and inflammatory pain [40]. Pro-inflammatory cytokines (e.g., TNF-α, IL-1β, IL-6) and activation of NF-kB have been strongly implicated in the initiation and development of inflammatory and neuropathic pain, and the administration of obtusifolin and gluco-obtusifolin (1 and 2 mg/kg) significantly inhibited this upregulation. This finding suggests that obtusifolin and gluco-obtusifolin inhibited the overexpression of spinal TNF-α, IL-1β, IL-6, and NF-κB p65 associated with inflammatory and neuropathic pain, which involves the regulation of neuroinflammatory processes and the neuroimmune system [51]. In another study, water-extracted polysaccharides (CP) from the whole seeds of C. obtusifolia L. and its two subfractions CP-30 and CP-40 were obtained. CP, CP-30, and CP-40 possessed immunomodulation activity by promoting phagocytosis and stimulating the production of nitric oxide (NO) and cytokines TNF-α and IL-6 [41]. Methanolic COE was investigated for antioxidant and health-relevant functionality. The extract exhibited 1292 mM Fe[II] per 1 mg/mL extract of antioxidant power, 49.92% inhibition of β-carotene degradation, 65.79% of scavenging activity against DPPH, and 50.78% of superoxide radicals (at a concentration 1 mg/mL). These antioxidant properties may be attributed to the total free phenolic content of the raw seeds, which was 13.33 ± 1.73 g catechin equivalent/100 g extract [14]. Recently, Kwon et al. [58] investigated the anti-inflammatory activity of major anthraquinone derivatives; among them, aurantio-obtusin inhibited iNOS expression without affecting iNOS enzyme activity and down-regulation mechanisms included interruption of the JNK/IKK/NF-κB activation and proinflammatory cytokine production from the lung-related cells. Additionally, aurantio-obtusin also dose-dependently (10 and 100 mg/kg) inhibited the inflammatory responses in a mouse model of airway inflammation, LPS-induced acute lung injury. Very recently, Hou et al. [54] reported anti-inflammatory activity by decreasing the production of NO, PGE2, and inhibiting iNOS, COX-2, TNF-α, and IL-6. Additionally, there was a reduction in the LPS-induced activation of nuclear factor-κB in RAW264.7 cells [54].

3.4. Antimicrobial Activity

Because many bacterial and fungal strains are resistant to a wide variety of antibiotics, medicinal plants have been studied for their potential antimicrobial properties. COE was active against several different microbes (Bifidobacterium adolescentis, B. bifidum, B. longum, B. breve, Clostridium perfringens, Escherichia coli, Lactobacillus casei). Isolated 1,2-dihydroxyanthraquinone strongly inhibited the growth of C. perfringens and E. coli and promoted the growth of B. bifidum [7]. The C. obtusifolia L. leaf extract in petroleum ether and chloroform showed sensitivity against E. faecalis (minimal inhibitory concentration [MIC] 0.2725 mg/mL), whereas ethanol extracts showed sensitivity against A. fumigatus (MIC 0.3116 mg/mL). Similarly, stem extracts of C. obtusifolia L. in petroleum ether showed sensitivity against E. faecalis (MIC 0.407 mg/mL), ethanol extracts showed sensitivity against E. faecalis (MIC 0.3009 mg/mL), and chloroform extracts showed sensitivity against E. faecalis MIC 0.4946 mg/mL [18]. The whole plant extract of C. obtusifolia significantly inhibited the growth of Staphyloccocus aureus MRSA8 (MIC 64 μg/mL), E. coli AG100 (MIC 256 μg/mL), Pseudomonas aeruginosa PA01 (MIC 256 μg/mL), Enterobacter aerogenes EA289 (MIC 289 μg/mL), and Klebsiella pneumoniae KP55 MIC 256 μg/mL [42]. Phytoalexin 2-(phydroxyphenoxy)-5,7-dihydroxychromone isolated from C. obtusifolia L. exhibited strong antifungal activity [31]. The C. obtusifolia L. root extract and its constituents exhibited strong antibacterial activity. Emodin, 2,5-dimethoxybenzoquione, questin, isotoralactone, and toralactone exhibited strong antibacterial activity against S. aureus 209P (MICs 4.5, 19, 25, and 3 µg/mL, respectively) and E. coli NIHJ MICs 25, 50, 50, 12, and 5.5 µg/mL, respectively [46].

3.5. Antidiabetic Activity

Two key enzymes, protein tyrosine phosphatase 1B (PTP1B) and α-glucosidase, are effective in treating diabetes mellitus. The effects of methanolic COE revealed inhibitory activities against PTP1B and α-glucosidase. Out of 15 anthraquinones from the extract, compounds with alaternin, physcion, chrysophanol, emodin, obtusin, questin, chryso-obtusin, aurantio-obtusin, 2-hydroxyemodin-1 methylether, gluco-obtusifolin, gluco-aurantio obtusin, and naphthalene glycoside aloe-emodin exhibited the highest inhibitory activities against PTP1B and α-glucosidase in vitro [9]. The effects of alaternin and emodin on the stimulation of glucose uptake by insulin-resistant human HepG2 cells were examined at concentrations ranging from 12.5 to 50 µM and 3.12 to 12.5 µM, respectively. In another study, five new anthraquinones were isolated from ethanol seed extracts of C. obtusifolia L. and evaluated for their antidiabetic activities through the inhibition of α-glucosidase in vitro [39]. Obtusifolin isolated from C. obtusifolia L. may have an antihyperlipidemic effect; an intraperitoneal obtusifolin injection reduced blood lipid levels in streptozotocin-induced diabetic rats [59]. Results from another study indicated that oral administration of obtusifolin significantly reversed the changes induced by hyperlipidemia in body weight, total cholesterol, triglycerides, low-density lipoprotein cholesterol, and high-density lipoprotein cholesterol; increased serum superoxide dismutase, and nitric oxide, and reduced malondialdehyde [50].
Recently, two new naphthalenic lactone glycosides(3S)-9,10-dihydroxy-7-methoxy-3-methyl-1-oxo-3,4-dihydro-1H-benzo[g]isochromene-3-carboxylic acid 9-O-β-d-glucopyranoside and (3R)-cassialactone 9-O-β-d-glucopyranoside were isolated from seeds of C. obtusifolia L. that showed significant inhibitory activities against the formation of advanced glycation end-products (AGEs) with IC50 values of 11.63 and 23.40 µM, respectively [60].

3.6. Antiplatelet Aggregation Inhibitory Activity

Ethanolic COE and three major anthraquinones (aurantio-obtusin, chryso-obtusin, and emodin) demonstrated inhibitory activity against ADP (adenosine 5′-diphosphate), arachidonic acid (AA), or collagen-induced platelet aggregation [47]. Methanolic COE and different solvent soluble fractions, including normal butanol (n-BuOH) and dicloromethane (CH2Cl2), exhibited antiplatelet aggregation activities. Furthermore, 17 anthraquinones, including gluco-obtusifolin, gluco-aurantio-obtusin, obtusifolin, and gluco-chryso-obtusin, were identified as active antiplatelet aggregation components [5].

3.7. Anticancer Activity

Polysaccharide COB1B1S2 and its sulfated derivative COB1B1S2-Sul were isolated from an alkaline COE. Human hepatocellular carcinoma cell lines Bel7402, SMMC7721, and Huh7, as well as HT-29 and Caco-2, were used to evaluate the anticancer effects of COB1B1S2 and COB1B1S2-Sul [61]. COB1B1S2 had a weak inhibitory effect on Bel7402, Huh7, HT-29, as well as Caco-2 cells. By contrast, COB1B1S2-Sul significantly inhibited the growth of all cell lines, particularly Bel7402 cells at 250 µg/mL; the inhibition ratio was 61.7% [62]. Three acetylated benzyl-beta-resorcylate glycosides (2-benzyl-4,6-dihydroxy benzoic acid-6-O-[2,6-O-diacetyl]-d-glucopyranoside, 2-benzyl-4,6-dihydroxy benzoic acid-6-O-[3,6-O-diacetyl]-d-glucopyranoside, and 2-benzyl-4,6-dihydroxy benzoic acid-6-O-[4,6-O-diacetyl]-d-glucopyranoside) were isolated from seeds of C. obtusifolia and exhibited significant cytotoxicity against a human hepatoblastoma cell line, with IC50 values of 4.6, 5.0, and 4.3 µg/mL, respectively [62]. In addition, 12 compounds were isolated from seeds of obtusifolia and their anticancer activities evaluated in multiple cancer cell lines [52]. 8-Hydroxy-1,7-dimethoxy-3-methylanthracene-9,10-dione-2-O-β-d-glucoside was active against HCT-116, A549, HepG2, SGC7901, and LO2 cell lines, with IC50 values of 4.5, 7.6, 22.8, 20.7, and 18.1 µg/mL, respectively. 6,8-Dihydroxy-1,7-dimethoxy-3-methylanthracene-9,10-dione-2-O-β-d-glucoside was only weakly active against HCT-116 (IC50, 43.0 µg/mL). 1-Desmethylobtusin had moderate cytotoxicity against HCT-116, A549, and SGC7901cell lines, with IC50 values of 5.1, 10, and 25.4 µg/mL, respectively. Chryso-obtusin showed significant cytotoxic activity against HCT-116, A549, SGC7901, and LO2 cell lines, with IC50 values of 10.5 to 15.8 µg/mL. Obtusin was moderately active against HCT-116, A549, and SGC7901 cell lines, with IC50 values of 13.1, 29.2, and 15.2 µg/mL, respectively. Aurantio-obtusin was moderately active against HCT-116, A549, SGC7901, and LO2 cell lines, with IC50 values of 18.9 to 22.0 µg/mL. Chryso-obtusin-2-O-β-d-glucopyranoside was selectively cytotoxic against HCT-116, A549, HepG2, SGC7901, and LO2 cell lines, with IC50 values of 5.8 to 14.6 µg/mL. Finally, aurantio-obtusin-6-O-β-d-glucopyranoside was weakly cytotoxic against HCT-116 and SGC7901, with IC50 values of 31.1 and 23.3 µg/mL, respectively [52].

3.8. Larvicidal Activity

The larvicidal activity of methanol COE against early fourth-stage larvae of Aedes aegypti and Culex pipiens pallens was investigated [43]. At 200 ppm, extracts of C. obtusifolia L. caused more than 90% mortality in larvae of Ae. aegypti and Cx. pipiens pallens. At 40 ppm, extracts of C. obtusifolia L. caused 51.4% and 68.5% mortality in fourth-stage larvae of Ae. aegypti and Cx. pipiens pallens, respectively. Larvicidal activity of C. obtusifolia extract at 20 ppm was significantly reduced [43]. In another study, COE obtained in different fractions showed mosquito larvicidal activity against fourth instar larvae of A. aegypti, Aedes togoi, and Cx. pipiens pallens [44]. However, the chloroform fraction of C. obtusifolia extracts exhibited a strong larvicidal activity of 100% mortality (at a concentration 25 mg/L), and the isolated active compound emodin showed strong larvicidal activity, with LC50 values of 1.4, 1.9, and 2.2 mg/L against C. pipiens pallens, A. aegypti, and A. togoi, respectively [44]. The ethanolic leaf extract of C. obtusifolia L. was also investigated for larvicidal and oviposition deterrence effects against late third instar larvae of Anopheles stephensi [45]. Extracts from the leaf displayed significant larvicidal activity, with LC50 and LC90 values of 52.2 and 108.7 mg/L, respectively (at concentrations of 25 mg/L). In addition, the oviposition study indicated that different concentrations of leaf extract greatly reduced the number of eggs deposited by gravid A. stephensi. At concentrations of 100, 200, 300, and 400 mg/L, the maximum percentages of effective repellency against oviposition were 75.5%, 83.0%, 87.2%, and 92.5%, respectively [45].

3.9. Other Activities

The methanol extract of C. obtusifolia L. and its isolated naphthopyrones cassiaside B2 and cassiaside C2 inhibited histamine release from rat peritoneal exudate mast cells induced by antigen–antibody reaction [6]. The anti-angiogenic activity of two polysaccharides, COB1B1S2 and COB1B1S2-Sul, from C. obtusifolia L. seeds was evaluated by tube formation of HMEC-1 cells on Matrigel. COB1B1S2 at 50 or 100 µg/mL did not impair tube formation, but COB1B1S2-Sul at 50 or 100 µg/mL significantly disrupted tube formation; even at 50 µg/mL, COB1B1S2-Sul could potentially completely inhibit tube formation in HMEC-1 cells [61]. Water-soluble polysaccharides (WSPs) from C. obtusifolia L. (pectic polysaccharides and hemicellulose) were identified. These WSPs reduced pancreatic α-amylase activity by 20.5% and 28.9% (at concentrations of 20 and 80 mg/mL, respectively), reduced pancreatic lipase activity by about 18.9% (at a concentration of 80 mg/mL), and increased protease activity 5- to 7-fold (at concentrations of 20 and 80 mg/mL, respectively). These WSPs were also able to bind bile acids and reduce the amount of cholesterol available for absorption [63]. The simultaneous determination and pharmacokinetic study of seven anthraquinones (chrysophanol, emodin, aloe-emodin, rhein, physcion, obtusifolin, and aurantio-obtusin) in rat plasma after oral administration of C. obtusifolia L. extract was investigated and may help to explain the bioactivity and clinical applications of C. obtusifolia L. [64]. The effects of COE and its anthraquinones on muscle mitochondrial function were evaluated in vivo in rats and in vitro using mitochondrial energy metabolism models. The organic extract of C. obtusifolia L. and emodin significantly inhibited NADH: cytochrome c oxidoreductase activity of bovine heart mitochondrial particles and NADH: coenzyme Q oxidoreductase activity of porcine heart mitochondrial NADH dehydrogenase and exhibited protective effects of coenzyme Q against enzyme inhibition by anthraquinones [65]. Inhibition of trypsin activity by C. obtusifolia L. seeds was investigated [66]. A Kunitz-type trypsin inhibitor showed strong resistance against the midgut trypsin-like protease of Pieris rapae. In addition, a trypsin inhibitor gene (CoTI1) was isolated from C. obtusifolia L. and exhibited dominant inhibitory activities against trypsin and trypsin-like proteases from Helicoverpa armigera, Spodoptera exigua, and Spodoptera litura [67]. Moreover, Dong et al. [68], has been also reported that Cassia semen (C. obtusifolia and C. tora) and its major constituents possesses a wide spectrum of pharmacological properties.

4. Conclusions and Perspectives

As presented in this review, pharmacological studies on C. obtusifolia L. and its putative active compounds, especially anthraquinones and naphthopyrone, support that several biological activities of C. obtusifolia can potentially impact human health. Anthraquinones and naphthopyrone can be effectively isolated and purified from C. obtusifolia seeds, leaves, root and its whole plant with various extraction analytical methods, mainly separation-based methods using TLC, HPLC, high-speed counter-current chromatography (HSCCC), and column chromatography (silica gel, reverse-phase, and Sephadex). The semi-shrubby herb C. obtusifolia L., which belongs to the family Leguminosae, has gained popularity because of its medicinal and historical importance. It has been widely used in traditional medicine to treat headaches, dizziness, dysentery, and eye disease. In addition, C. obtusifolia L. is important to the food industry and possesses a wide spectrum of pharmacological properties (e.g., anti-allergic, antidiabetic, anti-inflammatory, antimicrobial, antioxidant, hepatoprotective, neuroprotective, anti-Alzheimer’s disease, antiplatelet aggregation, and larvicidal activities) that are associated with its diverse chemical constituents (e.g., anthraquinones, naphthopyrone, terpenoid, flavonoid, polysaccharides, and lipids). The number of modern studies on bioactive compounds is increasing in biomedicine, suggesting that these compounds might have great medical significance in the future. Although the bioactivities of seed extracts or compounds isolated from C. obtusifolia L. have been substantiated using in vitro and in vivo studies, the mechanisms of action remain unknown. Thus, there are still opportunities and challenges for research of seed extracts or compounds. Therefore, additional studies are required before C. obtusifolia L. and its components can be considered for further clinical use. In conclusion, C. obtusifolia L. is an edible medicinal plant that is important to the food industry and has a wide range of potential pharmacological uses. This review presents a summary of studies published to date on this promising plant.

Author Contributions

Conceptualization, M.Y.A.; data curation, M.Y.A.; writing—original draft preparation, M.Y.A.; review and editing, M.Y.A., S.P. and M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research did not receive any specific grant from funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

All authors agree to the authorship and submission of the manuscript for peer review.

References

  1. Sob, S.V.T.; Wabo, H.K.; Tchinda, A.T.; Tane, P.; Ngadjui, B.T.; Ye, Y. Anthraquinones, sterols, triterpenoids and xanthones from Cassia obtusifolia. Biochem. Syst. Ecol. 2010, 38, 342–345. [Google Scholar] [CrossRef]
  2. Tang, L.; Wu, H.; Zhou, X.; Xu, Y.; Zhou, G.; Wang, T.; Kou, Z.; Wang, Z. Discrimination of Semen cassiae from two re-lated species based on the multivariate analysis of high-performance liquid chromatography fingerprints. J. Sep. Sci. 2015, 38, 2431–2438. [Google Scholar] [CrossRef] [PubMed]
  3. Yanjun, H.; Yuli, S.; Yuqing, Z. The advancement of the studies on the seeds of Cassia obtusifolia. Chin. Tradit. Herbal. Drugs. 2001, 32, 858–859. [Google Scholar]
  4. Li, Y.T.; Wang, Z.J.; Fu, M.H.; Yan, H.; Wei, H.W.; Lu, Q.H. A new anthraquinone glycoside from seeds of Cassia obtusifolia. Chin. Chem. Lett. 2008, 19, 1083–1085. [Google Scholar]
  5. Yun-Choi, H.S.; Kim, J.H.; Takido, M. Potential Inhibitors of Platelet Aggregation from Plant Sources, V. Anthraquinones from Seeds of Cassia obtusifolia and Related Compounds. J. Nat. Prod. 1990, 53, 630–633. [Google Scholar] [CrossRef]
  6. Kitanaka, S.; Nakayama, T.; Shibano, T.; Ohkoshi, E.; Takido, M. Antiallergic Agent from Natural Sources. Structures and Inhibitory Effect of Histamine Release of Naphthopyrone Glycosides from Seeds of Cassia obtusifolia L. Chem. Pharm. Bull. 1998, 46, 1650–1652. [Google Scholar] [CrossRef] [Green Version]
  7. Sung, B.K.; Kim, M.K.; Lee, W.H.; Lee, D.H.; Lee, H.S. Growth responses of Cassia obtusifolia toward human intestinal bac-teria. Fitoterapia 2004, 75, 505–509. [Google Scholar] [CrossRef]
  8. Kim, D.H.; Yoon, B.H.; Kim, Y.-W.; Lee, S.; Shin, B.Y.; Jung, J.W.; Kim, H.J.; Lee, Y.S.; Choi, J.S.; Kim, S.Y.; et al. The seed extract of Cassia obtusifolia ameliorates learning and memory impairments induced by scopolamine or transient cerebral hypoperfusion in mice. J. Pharmacol. Sci. 2007, 105, 82–93. [Google Scholar] [CrossRef] [Green Version]
  9. Jung, H.A.; Ali, M.Y.; Choi, J.S. Promising Inhibitory Effects of Anthraquinones, Naphthopyrone, and Naphthalene Glycosides, from Cassia obtusifolia on α-Glucosidase and Human Protein Tyrosine Phosphatases 1B. Molecules 2017, 22, 28. [Google Scholar] [CrossRef]
  10. Jung, H.A.; Ali, M.Y.; Jung, H.J.; Jeong, H.O.; Chung, H.Y.; Choi, J.S. Inhibitory activities of major anthraquinones and other constituents from Cassia obtusifolia against β-secretase and cholinesterases. J. Ethnopharmacol. 2016, 191, 152–160. [Google Scholar] [CrossRef]
  11. Drever, B.D.; Anderson, W.G.; Riedel, G.; Kim, D.H.; Ryu, J.H.; Choi, D.-Y.; Platt, B. The seed extract of Cassia obtusifolia offers neuroprotection to mouse hippocampal cultures. J. Pharmacol. Sci. 2008, 107, 380–392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Seo, Y.; Song, J.-S.; Kim, Y.-M.; Jang, Y.P. Toralactone glycoside in Cassia obtusifolia mediates hepatoprotection via an Nrf2-dependent anti-oxidative mechanism. Food Res. Int. 2017, 97, 340–346. [Google Scholar] [CrossRef] [PubMed]
  13. Doughari, J.H.; El-mahmood, A.M.; Tyoyina, I. Antimicrobial activity of leaf extracts of Senna obtusifolia (L). Afr. J. Pharm. Pharmacol. 2008, 2, 7–13. [Google Scholar]
  14. Vadivel, V.; Kunyanga, C.N.; Biesalski, H.K. Antioxidant Potential and Type II Diabetes-Related Enzyme Inhibition of Cassia obtusifolia L.: Effect of Indigenous Processing Methods. Food Bioprocess Technol. 2012, 5, 2687–2696. [Google Scholar] [CrossRef]
  15. Xie, Q.; Guo, F.F.; Zhou, W. Protective effects of cassia seed ethanol extract against carbon tetrachloride-induced liver in-jury in mice. Acta Biochem. Pol. 2012, 59, 265–270. [Google Scholar]
  16. Kirtikar, K.R.; Basu, B.D. Indian Medicinal Plant; Lalit Mohan Basu Press: Allahbad, India, 2006. [Google Scholar]
  17. Sob, S.V.T.; Wabo, H.K.; Tane, P.; Ngadjui, B.T.; Ma, D. A xanthone and a polyketide derivative from the leaves of Cassia obtusifolia (Leguminosae). Tetrahedron 2008, 64, 7999–8002. [Google Scholar] [CrossRef]
  18. Deshpande, S.R.; Naik, B.S. Evaluation of in vitro antimicrobial activity of extracts from Cassia obtusifolia L. and Senna so-phera (L.) Roxb against pathogenic organisms. J. Appl. Pharm. Sci. 2016, 6, 83–85. [Google Scholar] [CrossRef] [Green Version]
  19. Paudel, P.; Seong, S.H.; Shrestha, S.; Jung, H.A.; Choi, J.S. In vitro and in silico human monoamine oxidase inhibitory po-tential of anthraquinones, naphthopyrones, and naphthalenic lactones from Cassia obtusifolia Linn seeds. ACS Omega 2019, 4, 16139–16152. [Google Scholar] [CrossRef] [Green Version]
  20. Dave, H.; Ledwani, L. A review on anthraquinones isolated from Cassia species and their applications. Indian, J. Nat. Prod. Resour. 2012, 3, 291–319. [Google Scholar]
  21. Harry-O’Kuru, R.E.; Mohamed, A. Processing scale-up of sciklepod (Senna obtusifolia L.) seed. J. Agri. Food Chem. 2009, 57, 2726–2731. [Google Scholar] [CrossRef] [PubMed]
  22. Wu, Y.V.; Abbott, T.P. Gum and protein enrichment from sciklepod (Cassia obtusifolia) seed by fine grinding and sieving. Ind. Crops. Prod. 2005, 21, 387–390. [Google Scholar] [CrossRef]
  23. Abbott, T.P.; Vaughu, S.F.; Dowd, P.F.; Mojtahedi, H.; Wilson, R.F. Potential uses ofsciklepod (Cassia obtusifolia). Ind. Crops. Prod. 1998, 8, 77–82. [Google Scholar] [CrossRef]
  24. Kitanaka, S.; Takido, M. Studies on the constituents of the seeds of Cassia obtusifolia: The structures of two new lactones, isotoralactone and cassialactone. Phytochemistry 1981, 20, 1951–1953. [Google Scholar] [CrossRef]
  25. Wu, X.-H.; Ruan, J.-L.; Cheng, C.-R.; Wu, Z.-Y.; Guan, S.-H.; Tao, S.-J.; Xu, P.-P.; Guo, D.-A. Benzyl-β-resorcylates from Cassia obtusifolia. Fitoter. 2010, 81, 617–620. [Google Scholar] [CrossRef]
  26. Zhang, C.; Wang, R.; Liu, B.; Tu, G. Structure elucidation of a sodium salified anthraquinone from the seeds of Cassia obtusifolia by NMR technique assisted with acid-alkali titration. Magn. Reson. Chem. 2011, 49, 529–532. [Google Scholar] [CrossRef]
  27. Wu, X.-H.; Cai, J.-J.; Ruan, J.-L.; Lou, J.-S.; Duan, H.-Q.; Zhang, J.; Cheng, C.-R.; Guo, D.-A.; Wu, Z.-Y.; Zhang, Y.-W. Acetylated anthraquinone glycosides from Cassia obtusifolia. J. Asian Nat. Prod. Res. 2011, 13, 486–491. [Google Scholar] [CrossRef]
  28. Pang, X.; Li, N.-N.; Yu, H.-S.; Kang, L.-P.; Yu, H.-Y.; Song, X.-B.; Fan, G.-W.; Han, L.-F. Two new naphthalene glycosides from the seeds of Cassia obtusifolia. J. Asian Nat. Prod. Res. 2018, 21, 970–976. [Google Scholar] [CrossRef] [PubMed]
  29. Pang, X.; Wang, L.M.; Zhang, Y.C.; Kang, L.P.; Yu, H.S.; Fan, G.W.; Han, L.F. New anthraquinone and eurotinone ana-logue from the seeds of Senna obtusifolia and their inhibitory effects on human organic anion transporters 1 and 3. Nat. Prod. Res. 2018, 4, 1–8. [Google Scholar]
  30. Feng, L.; Yin, J.Y.; Nie, S.P.; Wan, Y.Q.; Xiea, M.Y. Enzymatic purification and structure characterization of glucuronoxy-lan from water extract of Cassia obtusifolia seeds. Int. J. Biol. Macromol. 2018, 107, 1438–1446. [Google Scholar] [CrossRef] [PubMed]
  31. Sharon, A.; Ghirlando, R.; Gressel, J. Isolation, Purification, and Identification of 2-(p-Hydroxyphenoxy)-5,7-Dihydroxychromone: A Fungal-Induced Phytoalexin from Cassia obtusifolia. Plant Physiol. 1992, 98, 303–308. [Google Scholar] [CrossRef] [Green Version]
  32. Guo, H.; Chang, Z.; Yang, R.; Guo, D.; Zheng, J. Anthraquinones from hairy root cultures of Cassia obtusifolia. Phytochemistry 1998, 49, 1623–1625. [Google Scholar] [CrossRef]
  33. Kim, D.H.; Kim, S.; Jung, W.Y.; Park, S.J.; Park, D.H.; Kim, J.M.; Cheong, J.H.; Ryu, J.H. The neuroprotective effects of the seeds of Cassia obtusifolia on transient cerebral global ischemia in mice. Food Chem. Toxicol. 2009, 47, 1473–1479. [Google Scholar] [CrossRef] [PubMed]
  34. Ju, M.S.; Kim, H.G.; Choi, J.G.; Ryu, J.H.; Hur, J.; Kim, Y.J.; Oh, M.S. Cassiae semen, a seed of Cassia obtusifolia, has neu-roprotective effects in Parkinson’s disease models. Food Chem. Toxicol. 2010, 48, 2037–2044. [Google Scholar] [CrossRef] [PubMed]
  35. Yi, J.H.; Park, H.J.; Lee, S.; Jung, J.W.; Kim, B.C.; Lee, Y.C.; Ryu, J.H.; Kim, D.H. Cassia obtusifolia seed ameliorates amyloid β-induced synaptic dysfunction through anti-inflammatory and Akt/GSK-3β pathways. J. Ethnopharmacol. 2016, 178, 50–57. [Google Scholar] [CrossRef] [PubMed]
  36. Byun, E.; Jeong, G.S.; An, R.B.; Li, B.; Lee, D.S.; Ko, E.K.; Yoon, K.H.; Kim, Y.C. Hepatoprotective compounds of Cassiae Semen on tacrine-induced cytotoxicity in HepG2 cells. Korean, J. Pharmacogn. 2007, 38, 400–402. [Google Scholar]
  37. Ali, M.Y.; Jannat, S.; Jung, H.A.; Min, B.-S.; Paudel, P.; Choi, J.S. Hepatoprotective effect of Cassia obtusifolia seed extract and constituents against oxidative damage induced by tert -butyl hydroperoxide in human hepatic HepG2 cells. J. Food Biochem. 2018, 42, e12439. [Google Scholar] [CrossRef] [Green Version]
  38. Meng, Y.; Liu, Y.; Fang, N.; Guo, Y. Hepatoprotective effects of Cassia semen ethanol extract on non-alcoholic fatty liver disease in experimental rat. Pharm. Biol. 2019, 57, 98–104. [Google Scholar] [CrossRef] [Green Version]
  39. Xu, Y.-L.; Tang, L.-Y.; Zhou, X.-D.; Zhou, G.-H.; Wang, Z.-J. Five new anthraquinones from the seed of Cassia obtusifolia. Arch. Pharmacal Res. 2014, 38, 1054–1058. [Google Scholar] [CrossRef] [PubMed]
  40. Kim, S.-J.; Kim, K.-W.; Kim, D.-S.; Kim, M.-C.; Jeon, Y.-D.; Kim, S.-G.; Jung, H.-J.; Jang, H.-J.; Lee, B.-C.; Chung, W.-S.; et al. The Protective Effect of Cassia obtusifolia on DSS-Induced Colitis. Am. J. Chin. Med. 2011, 39, 565–577. [Google Scholar] [CrossRef] [PubMed]
  41. Feng, L.; Yin, J.; Nie, S.; Wan, Y.; Xie, M. Fractionation, physicochemical property and immunological activity of poly-saccharides from Cassia obtusifolia. Int. J. Biol. Macromol. 2016, 91, 946–953. [Google Scholar] [CrossRef] [PubMed]
  42. Voukeng, I.K.; Beng, V.P.; Kuete, V. Antibacterial activity of six medicinal Cameroonian plants against Gram-positive and Gram-negative multidrug resistant phenotypes. BMC Complement. Altern. Med. 2016, 16, 388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Jang, Y.-S.; Baek, B.-R.; Yang, Y.-C.; Kim, M.-K.; Lee, H.-S. Larvicidal activity of leguminous seeds and grains against Aedes aegypti and Culex pipiens pallens. J. Am. Mosq. Control. Assoc. 2002, 18, 210–213. [Google Scholar]
  44. Yang, Y.C.; Lim, M.Y.; Lee, H.S. Emodin isolated from Cassia obtusifolia (Leguminosae) seed shows larvicidal activity against three mosquito species. J. Agri. Food. Chem. 2003, 51, 7629–7631. [Google Scholar] [CrossRef] [PubMed]
  45. Rajkumar, S.; Jebanesan, A. Larvicidal and oviposition activity of Cassia obtusifolia Linn (Family: Leguminosae) leaf ex-tract against malarial vector, Anopheles stephensi Liston (Diptera: Culicidae). Parasitol. Res. 2009, 104, 337–340. [Google Scholar] [CrossRef] [PubMed]
  46. Kitanaka, S.; Takido, M. Studies on the constituents in the roots of Cassia obtusifolia L. and the antimicrobial activities of aonstituents of the roots and the seeds. Yakugaku Zasshi 1986, 106, 302–306. [Google Scholar] [CrossRef] [Green Version]
  47. Yun-Choi, H.S.; Lee, J.R.; Kim, J.H.; Kim, Y.H.; Kim, T.H. Potential inhibitors oplatelet aggregation from plant sources, IV. Anthraquinones from seeds of Cassia obtusifolia and related compounds. Korean. J. Pharmacogn. 1987, 18, 203–206. [Google Scholar]
  48. Shin, B.Y.; Kim, D.H.; Hyun, S.K.; Jung, H.A.; Kim, J.M.; Park, S.J.; Kim, S.Y.; Cheong, J.H.; Choi, J.S.; Ryu, J.H. Alater-nin attenuates delayed neuronal cell death induced by transient cerebral hypoperfusion in mice. Food Chem. Toxicol. 2010, 48, 1528–1536. [Google Scholar] [CrossRef] [PubMed]
  49. Kim, N.H.; Hyun, S.K.; Yoon, B.H.; Seo, J.-H.; Lee, K.-T.; Cheong, J.H.; Jung, S.Y.; Jin, C.; Choi, J.S.; Ryu, J.H. Gluco-obtusifolin and its aglycon, obtusifolin, attenuate scopolamine-induced memory impairment. J. Pharmacol. Sci. 2009, 111, 110–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Zhuang, S.-Y.; Wu, M.-L.; Wei, P.-J.; Cao, Z.-P.; Xiao, P.; Li, C.-H. Changes in Plasma Lipid Levels and Antioxidant Activities in Rats after Supplementation of Obtusifolin. Planta Medica 2016, 82, 539–543. [Google Scholar] [CrossRef]
  51. He, Z.-W.; Wei, W.; Li, S.-P.; Ling, Q.; Liao, K.-J.; Wang, X. Anti-allodynic effects of obtusifolin and gluco-obtusifolin against inflammatory and neuropathic pain: Possible mechanism for neuroinflammation. Biol. Pharm. Bull. 2014, 37, 1606. [Google Scholar] [CrossRef] [Green Version]
  52. Shi, B.-J.; Zhang, W.-D.; Jiang, H.-F.; Zhu, Y.-Y.; Chen, L.; Zha, X.-M.; Lu, Y.-Y. A new anthraquinone from seed of Cassia obtusifolia. Nat. Prod. Res. 2016, 30, 35–41. [Google Scholar] [CrossRef]
  53. Vishnuprasad, C.N.; Tsuchiya, T.; Kanegasaki, S.; Kim, J.H.; Han, S.S. Aurantio-Obtusin stimulates chemotactic migra-tion and differentiation of MC3T3-E1 osteoblast cells. Planta Med. 2014, 80, 544–549. [Google Scholar]
  54. Hou, J.; Gu, Y.; Zhao, S.; Huo, M.; Wang, S.; Zhang, Y.; Qiao, Y.; Li, X. Anti-Inflammatory Effects of Aurantio-Obtusin from Seed of Cassia obtusifolia L. through Modulation of the NF-κB Pathway. Molecules 2018, 23, 3093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Shrestha, S.; Seong, S.H.; Paudel, P.; Jung, A.H.; Choi, J.S. Structure related inhibition of enzyme systems in cholinester-ases and BACE1 in vitro by naturally occurring naphthopyrone and its glycosides isolated from Cassia obtusifolia. Molecules 2018, 23, 69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Paudel, P.; Kim, D.; Jeon, J.; Park, S.; Seong, S.; Jung, H.; Choi, J. Neuroprotective Effect of Aurantio-Obtusin, a Putative Vasopressin V1A Receptor Antagonist, on Transient Forebrain Ischemia Mice Model. Int. J. Mol. Sci. 2021, 22, 3335. [Google Scholar] [CrossRef]
  57. Paudel, P.; Jung, H.A.; Choi, J.S. Anthraquinone and naphthopyrone glycosides from Cassia obtusifolia seeds mediate hepatoprotection via Nrf2-mediated HO-1 activation and MAPK modulation. Arch. Pharm. Res. 2018, 41, 677–689. [Google Scholar] [CrossRef] [PubMed]
  58. Kwon, K.S.; Lee, J.H.; So, K.S.; Park, B.K.; Lim, H.; Choi, J.S.; Kim, H.P. Aurantio-obtusin, an anthraquinone from cassiae semen, ameliorates lung inflammatory responses. Phytother. Res. 2018, 32, 1537–1545. [Google Scholar] [CrossRef] [PubMed]
  59. Tang, Y.; Zhong, Z. Obtusifolin Treatment Improves Hyperlipidemia and Hyperglycemia: Possible Mechanism Involving Oxidative Stress. Cell Biophys. 2014, 70, 1751–1757. [Google Scholar] [CrossRef] [PubMed]
  60. Shrestha, S.; Paudel, P.; Seong, S.H.; Min, B.S.; Seo, E.K.; Jung, H.A.; Choi, J.S. Two new naphthalenic lactone glyco-sides from Cassia obtusifolia L. seeds. Arch. Pharm. Res. 2018, 41, 737–742. [Google Scholar] [CrossRef] [PubMed]
  61. Cong, Q.; Shang, M.; Dong, Q.; Liao, W.; Xiao, F.; Ding, K. Structure and activities of a novel heteroxylan from Cassia ob-tusifolia seeds and its sulfated derivative. Carbohydr. Res. 2014, 393, 43–50. [Google Scholar] [CrossRef] [PubMed]
  62. Wu, X.; Ruan, J.; Yang, V.C.; Wu, Z.; Lou, J.; Duan, H.; Zhang, Y.J.; Zhang, A.; Guo, D. Three new acetylated ben-zyl-beta-resorcylate glycosides from Cassia obtusifolia. Fitoterapia 2012, 83, 166–169. [Google Scholar] [CrossRef] [PubMed]
  63. Huang, Y.-L.; Chow, C.-J.; Tsai, Y.-H. Composition, characteristics, and in-vitro physiological effects of the water-soluble polysaccharides from Cassia seed. Food Chem. 2012, 134, 1967–1972. [Google Scholar] [CrossRef]
  64. Yang, C.; Wang, S.; Guo, X.; Sun, J.; Liu, L.; Wu, L. Simultaneous determination of seven anthraquinones in rat plasma by ultra-high-performance liquid chromatography–tandem mass spectrometry and pharmacokinetic study after oral admin-istration of semen cassiae extract. J. Ethnopharmacol. 2015, 169, 305–313. [Google Scholar] [CrossRef] [PubMed]
  65. Lewis, D.C.; Shibamoto, T. Effects of Cassia obtusifolia (sicklepod) extracts and anthraquinones on muscle mitochondrial function. Toxicon 1989, 27, 519–529. [Google Scholar] [CrossRef]
  66. Liao, H.; Ren, W.; Kang, Z.; Jiang, J.-H.; Zhao, X.-J.; Du, L.-F. A trypsin inhibitor from Cassia obtusifolia seeds: Isolation, characterization and activity against Pieris rapae. Biotechnol. Lett. 2007, 29, 653–658. [Google Scholar] [CrossRef]
  67. Liu, Z.; Zhu, Q.; Li, J.; Zhang, G.; Jiamahate, A.; Zhou, J.; Liao, H. Isolation, structure modeling and function characteri-zation of a trypsin inhibitor from Cassia obtusifolia. Biotechnol. Lett. 2015, 37, 863–869. [Google Scholar] [CrossRef]
  68. Dong, X.; Fu, J.; Yin, X.; Yang, C.; Zhang, X.; Wang, W.; Du, X.; Wang, Q.; Ni, J. Cassiae semen: A review of its phyto-chemistry and pharmacology. Mol. Med. Rep. 2017, 16, 2331–2346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Chemical structures of the main compounds present in Cassia obtusifolia L.
Figure 1. Chemical structures of the main compounds present in Cassia obtusifolia L.
Molecules 26 06252 g001aMolecules 26 06252 g001b
Figure 2. Different biological activities displayed by Cassia obtusifolia.
Figure 2. Different biological activities displayed by Cassia obtusifolia.
Molecules 26 06252 g002
Table 1. Ethnomedicinal importance of Cassia obtusifolia.
Table 1. Ethnomedicinal importance of Cassia obtusifolia.
Sr. No.Plant Part UsedEthnomedicinal Use
1Whole plantIn traditional Oriental medicine, the whole plant of C. obtusifolia has been used for treatment of Laxative, eye infections, diarrhea, urinary tract infections, gingivitis, fever, and cough remedy [13].
2RootsRoot is considered bitter, tonic, stomachic and is antidote against snake bite. Other uses are in treatment of fungal diseases, worm infection, abdominal tumors, bronchitis, and asthma. The roots of C. obtusifolia are also usually crushed, mixed with lime juice, and applied to ringworms [14].
3SeedsThe seeds of C. obtusifolia are used to treat dizziness and to benefit the eyes by anchoring and nourishing the liver. The dried and roasted seeds are also used as brew a tea. Seeds of C. obtusifolia were also used for the treatment of headache, ophthalmic diseases, constipation, hypertension, and hyperlipidemia. In Korea, the hot extract of seeds is taken orally for protection of liver [10,15].
4LeavesC. obtusifolia leaves and pods have been widely used as purgatives and laxatives. In Indian traditional ayurveda system, the leaves and Pods are used as digestible, laxative, diuretic, stomachic, antipyretic, improves the appetite, biliousness, blood diseases, burning sensation, leprosy, bronchitis, piles, and leucorrhoea [16,17].
5Stem barkIn Indian traditional ayurveda system, Stem bark extract is used for various skin ailments, rheumatic diseases, and as laxative [18].
6Pods and fruitsPods are used in dysentery, in eye diseases and pains in the joints. The unripe fruits are also cooked and eaten [14].
Table 2. Pharmacological activities of Cassia obtusifolia extracts.
Table 2. Pharmacological activities of Cassia obtusifolia extracts.
Pharmacological ActivityPart of PlantType of ExtractIn Vivo/
In Vitro
ModelAdministration (In Vivo)Dose RangeActive ConcentrationReference
Neuroprotective ActivitySeeds85% EtOH ext.In vivoAmeliorate scopolamine or 2VO-induced memory impairments by inhibiting AChEOral25–100 mg/kg50 mg/kg[8]
Seeds85% EtOH ext.In vivoNeuroprotection by inhibition of pro-inflammatory genes iNOX, and COX-2, and increased neurotrophic factor expression of pCREB and BDNFOral10, 50 mg/kg50 mg/kg[33]
Seeds85% EtOH ext.In vitroReduced Aβ toxicity and maintenance of Ca2+ dysregulation and excitotoxicity, mitochondrial dysfunction in primary hippocampal cultures-0.1–10 µg/mL1, 10 µg/mL[11]
SeedsEtOH ex.In vivoprotected the dopaminergic cells against 6-OHDA- and MPP+-induced neurotoxicities in primary mesencephalic cultures and in a mouse model in PDIntraperitoneal injection0.1–10 µg/mL for DA, 50 mg/kg mouse0.1, 1 µg/mL
50 mg/kg
[34]
SeedsEtOH ext.In vitroInhibited cell loss against 6-OHDA-induced DA neural toxicity by an anti-oxidant and anti-mitochondrial-mediated apoptosis mechanism in PC12 cells.-0.1–10 µg/mL
1000 µg/mL for DPPH, ABTS
1 µg/mL ROS, 10 µg/mL GSH, 75% Casp-3, 92%-DPPH, 85% ABTS[35]
SeedsMeOH ext.
EtOAc fr.
CH2Cl2 fr.
BuOH fr.
In vitroInhibitory activity against MAO-A, and MAO-B-0.25–120 µg/mLEtOAc fr. exhibited greatest inhibitory IC50 = 20, and 56 µg/mL activity against MAO-A, and MAO-B[36]
SeedsMeOH ext.
EtOAc fr.
CH2Cl2 fr.
BuOH fr.
H2O fr.
In vitroInhibitory activity against AChE, BChE, BACE1-0.4–120 µg/mLIC50 = 9.45~29 µg/mL for AChE, IC50 = 7.58~49 µg/mL for BChE, IC50 = 26~96 µg/mL for BACE1[10]
Seeds85% EtOH ext.In vivoAmeliorate Aβ-induced LTP impairment in the acute hippocampal slices and regulates GSK-3β, Akt signaling pathways through the inhibition of iNOS, COX expression-1 and 10 µg/mL10 µg/mL[35]
Hepatoprotective ActivitySeedsMeOH ext.In vitroProtection against tacrine-induced hepatotoxicity in HepG2 cells-300 µg/mL300 µg/mL[36]
Seeds70% EtOH ext.
EtOAc, CH2Cl2, BuOH, H2O fr.
In vitroProtective effect against t-BHP-induced hepatotoxicity in HepG2 cells-10–100 µg/mLEtOAc fr. showed most potent hepatoprotective activity (30 µg/mL)[12]
SeedsEtOH ext.In vivoHepatoprotective effects against CCl4-induced liver injury in miceIntraperitoneal injection0.5, 1, 2 g/kgReduced ALT and AST, Ca2+, MDA, and increased GSH, SOD, GR, GPx, GST, CYP2E1 (2 g/kg)[15]
seedsEtOAc fr. CH2Cl2 fr.
BuOH fr.
H2O fr.
In vitroProtective effect against t-BHP-induced hepatotoxicity in HepG2 cells-12.5–50 µg/mLEtOAc fr. showed most potent hepatoprotective activity (50 µg/mL)[37]
Seeds70% EtOH ext.In vivo(a) Significantly decreased the levels of AST, ALT, TG, TC, TNF-a, IL-6, IL-8 and MDA; (b) Increased the levels of SOD and GSH; (c) Significantly increased the mRNA expression levels of LDL-ROral0.5–2 g/kg(a) Dose-dependently decreased biomarkers at 0.5–2 g/kg; (b) Dose-dependently decreased at 0.5–2 g/kg; (c) Significantly increased the levels of LDL-R at 2 g/kg[38]
Anti-diabetic ActivitySeedsMeOH ext.
EtOAc fr.
CH2Cl2 fr.
BuOH fr.
H2O fr.
In vitroInhibitory activity against PTP1B and α-glucosidase-0.4–400 µg/mL for PTP1B, 0.16–400 µg/mL
for α-glucosidase
MeOH ext. (IC50 = 14 µg/mL) and EtOAc fr. (IC50 = 74 µg/mL) exhibited greatest inhibitory activity against PTP1B and α-glucosidase[9]
SeedsEtOH ext.In vitroInhibitory activity against α-glucosidase-1000 µg/mL20% inhibition of α-glucosidase (1000 µg/mL)[39]
Anti-inflammatory, Antioxidant, and Immune-modulatory ActivitiesRoasted seedsHot H2O ext.In vivoProtection against dextran sulfate sodium (DSS)-induced colitis through the inhibition of (IL)-6, COX-2, NF-κBOral1 g/kgSignificantly reduced clinical signs and the levels of inflammatory mediators (at concentration 1 g/kg)[40]
SeedsH2O soluble polysaccharide fr.In vitroIncreased immune-modulatory activity by promoting phagocytosis and stimulating the production of NO and cytokines TNF- and IL-6 on macrophage cell line RAW264.7-62.5–500 µg/mLStimulates NO, TNF- and IL-6 expression (250 µg/mL) and promotes phagocytic activity (500 µg/mL)[41]
SeedsMeOH ext.In vitroDPPH, Fe [II], superoxide radicals scavenging activity and inhibit ß-carotene degradation-1 mg/mLInhibition 65.79% DPPH, 50.78% superoxide radical, 49.92% inhibit ß-carotene degradation,1292 mM Fe [II] inhibited (at 1 mg/mL)[14]
Antimicrobial ActivitySeedsMeOH ext.
Hexane fr.
EtOAc fr.
CH2Cl2 fr.
BuOH
fr.H2O fr.
In vitroBifidobacterium adolescentis, B. bifidum, B. longum, B. breve, Clostridium perfringens, Escherichia coli, Lactobacillus casei-5 mg discs−1CH2Cl2 fr, MeOH ext. and Hexane fr. exhibited the greatest antibacterial activity[7]
LeafPet ether ext.
EtOH ext.
Chloroform ext.
In vitroAspergilus fumigatus, Staphylococcus aureus, Enterococcus faecalis, E. coli, Klebsiella sp., Candia albicans-0.6–1 mg/mLPet ether, chloroform ext. active against C. albicans (MIC 0.3524, and 0.4239 mg/mL), ethanol E. faecalis (MIC 0.2738 mg/mL)[18]
stemPet ether ext.
EtOH ext.
Chloroform ext.
In vitroAspergilus fumigatus, Staphylococcus aureus, Enterococcus faecalis, E. coli, Klebsiella sp., Candia albicans 0.6–1 mg/mLEthanol, pet ether, chloroform ext. was more active against E.faecalis (MIC 0.298, 0.254, and 0.589 mg/mL, respectively)[18]
Whole plantMeOH ext.In vitroE. coli, P. aeruginosa, Enterobacter aerogenes Providencia stuartii, K.pneumoniae, Enterobacter cloacae, S. aureus-256 µg/mLinhibition of S. aureus, E. coli, P. aeruginosa, E. aerogenes, K. pneumoniae (MIC ranges of 64–289 μg/mL[42]
Larvicidal ActivitySeedsMeOH ext.In vitroLarvicidal activity against Aedes aegypti and Culex pipiens pallens-10–300 ppm40 ppm[43]
SeedsChloroform fr.In vitroLarvicidal activity against A. aegypti, Aedes togoi, and Cx. pipiens-25 mg/L100% Mortality (at concentration 25 mg/L)[44]
LeafEtOH ext.In vitroLarvicidal activity against Anopheles stephensi-25–125 mg/LLC50 = 52.2 mg/L, LC90 = 108.7 mg/L (at concentration 25 mg/L)[45]
LeafEtOH ext.In vitroAnti-oviposition activity against Anopheles stephensi-100–400 mg/L92.5% for 400 mg/L
87.2% for 300 mg/L
83.0% for 200 mg/L
[45]
Table 3. Major Phytochemicals in Cassia obtusifolia and their pharmacological activities.
Table 3. Major Phytochemicals in Cassia obtusifolia and their pharmacological activities.
CompoundsBiological ActivityIn Vivo/
In Vitro
ModelAdministration
(In Vivo)
Dose RangeActive ConcentrationReference
Anthraquinones
EmodinAnti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) Butyrylcholinesterase inhibitory activity
(c) β-secretase inhibitory activity
-0–100 µg/mL(a) IC50 = 9.17µg/mL
(b) IC50 = 157 µg/mL
(c) IC50 = 4.48 µg/mL
[10]
Antimicrobial activityIn vitroAntibacterial activity against
(a) Staphylococcus aureus 209P
(b) Escherichia coli NIHJ
-0–1 mg/mLMIC (a) 4.5 µg/mL
(b) 25 µg/mL
[46]
Antidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
(c) Stimulation of glucose uptake in HepG2 cells
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(c) 3.12–12.5 µM
(a) IC50 = 3.51 µg/mL
(b) IC50 = 1.02 µg/mL
(c) glucose uptake
[9]
Platelet anti-aggregatory activityIn vitro(a) Adenosine 5′-diphosphate inhibitory activity
(b) Arachidonic-acid inhibitory activity
(c) Collagen inhibitory activity
-0–1 mg/mL1 mg/mL[47]
Larvicidal activityIn vitroLarvicidal activity against (a) Culex pipiens pallens (b), Aedes aegypti (c) Aedes togoi-1–20 mg/L(a) LC50 = 1.4 mg/L
(b) LC50 = 1.9 mg/L
(c) LC50 = 2.2 mg/L
[44]
Hepatoprotective activityIn vitroProtection against t-BHP-induced hepatotoxicity in HepG2 cells-25 µMprotect cells damage[37]
Parkinson’s disease activityIn vitro(a) MAO-A inhibitory activity
(b) MAO-B inhibitory activity
-25 µM(a) IC50 = 23 µM
(b) IC50 = 54 µM
[19]
AlaterninNeuroprotective activityIn vivoPrevented nitrotyrosine and lipid peroxidation, as well as BCCAO induced-iNOS expression and significantly reduced microglial activationOrally1, 10 mg/kg10 mg/kg[48]
Antidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
(c) Stimulation of glucose uptake in HepG2 cells
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(c) 12.5–50 µM
(a) IC50 = 1.22 µg/mL
(b) IC50 = 0.99 µg/mL
(c) glucose uptake
[9]
Anti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) Butyrylcholinesterase inhibitory activity
(c) β-secretase inhibitory activity
-0–100 µg/mL(a) IC50 = 6.29 µg/mL
(b) IC50 = 113 µg/mL
(c) IC50 = 0.94 µg/mL
[10]
Hepatoprotective activityIn vitroProtection against t-BHP-induced hepatotoxicity in HepG2 cells-50, 100 µM(a) protect cells damage
(b) increased GSH level and reduce ROS level
[37]
Parkinson’s disease activityIn vitro(a) MAO-A inhibitory activity
(b) MAO-B inhibitory activity
-10 µM(a) IC50 = 5.35 µM
(b) IC50 = 4.55 µM
[19]
ObtusifolinNeuroprotective activityIn vivoSignificantly reversed scopolamine-induced cognitive impairments in the passive avoidance test, improved escape latencies, swimming times in the target quadrant, and crossing numbers in the zone in Morris water maze testOrally0.25–2 mg/kg0.5 mg/kg[49]
Hyperlipidemia and antioxidant activityIn vivoReduced body weight, TC, TG, LDL-C and increased HDL-C levels, as well as increased SOD and NO, and reduced MDA levels in hyperlipidemic rats.Orally5 and 20 mg/kg20 mg/kg[50]
Neuropathic and anti-inflammatory activityIn vivoInhibition of TNF-α, IL-1β, IL-6 and NF-kB up-regulation in the spinal cord in mice and rat modelsIntraperitoneal injection0.25–2 mg/kg1 and 2 mg/kg[51]
Anti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) Butyrylcholinesterase inhibitory activity
(c) β-secretase inhibitory activity
-0–100 µg/mL(a) IC50 = 18.5 µg/mL
(b) IC50 = 284 µg/mL
(c) IC50 = 64.8 µg/mL
[10]
Antidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(a) IC50 = 35.2 µg/mL
(b) IC50 = 142 µg/mL
[9]
Hepatoprotective activityIn vitroProtection against tacrine-induced hepatotoxicity in HepG2 cells-160 µMProtection ratio value 41.2% at 160 µM[36]
Parkinson’s disease activityIn vitro(a) MAO-A inhibitory activity; (b) MAO-B inhibitory activity-100 µM(a) IC50 = 31 µM
(b) IC50 ≥ 400 µM
[19]
Gluco-obtusifolinNeuropathic and anti-inflammatory activityIn vivoInhibition of TNF-α, IL-1β, IL-6 and NF-kB up-regulation in the spinal cord in mice and rat modelsIntraperitoneal injection0.25–2 mg/kg1 and 2 mg/kg[51]
Anti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) Butyrylcholinesterase inhibitory activity
(c) β-secretase inhibitory activity
-0–400 µg/mL(a) IC50 = 37.2 µg/mL
(b) IC50 = 172 µg/mL
(c) IC50 = 41.1 µg/mL
[10]
Neuroprotective activityIn vivoSignificantly reversed scopolamine-induced cognitive impairments in the passive avoidance test, improved escape latencies, swimming times in the target quadrant, and crossing numbers in the zone in the Morris water maze testOrally0.25–2 mg/kg0.5 mg/kg[49]
Antidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(a) IC50 = 53.35 µg/mL
(b) IC50 = 23.77 µg/mL
[9]
Platelet anti-aggregatory activityIn vitro(a) Adenosine 5′-diphosphate inhibitory activity
(b) Arachidonic-acid inhibitory activity
(c) Collagen inhibitory activity
-0–1 mg/mL(a) IC50 = 0.25 µg/mL
(b) IC50 = 0.05 µg/mL
(c) IC50 = 0.1 µg/mL
[5]
Parkinson’s disease activityIn vitro(a) MAO-A inhibitory activity
(b) MAO-B inhibitory activity
-500 µM(a) IC50 ≥ 400 µM
(b) IC50 ≥ 400 µM
[19]
Aurantio-obtusinHepatoprotective activityIn vitroProtection against tacrine-induced hepatotoxicity in HepG2 cells-160 µMProtection ratio value 55.3% at 160 µM[36]
Anti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) Butyrylcholinesterase inhibitory activity
(c) β-secretase inhibitory activity
-0–100 µg/mL(a) IC50 = 92.1 µg/mL
(b) IC50 = 314 µg/mL
(c) IC50 = 67.9 µg/mL
[10]
Platelet anti-aggregatory activityIn vitro(a) Adenosine 5′-diphosphate inhibitory activity
(b) Arachidonic-acid inhibitory activity
(c) Collagen inhibitory activity
-0–1 mg/mL1 mg/mL[48]
Antidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(a) IC50 = 27.19 µg/mL
(b) IC50 = 41.20 µg/mL
[9]
Anti-cancer activityIn vitroCytotoxicity against (a) HCT-116, (b) A549, (c) SGC7901 and (d) LO2 cell lines-0.4–50 µg/mL(a) IC50 = 18.9 µg/mL
(b) IC50 = 20.1 µg/mL
(c) IC50 = 22.0 µg/mL
(d) IC50 = 23.1 µg/mL
[52]
Prevention of bone diseaseIn vitroStimulates osteoblast migration, differentiation, and mineralization in a dose-dependent manner in MC3T3-E1 osteoblast cells-0.1–100 µM10 µM[53]
Anti-inflammatory activityIn vitro(a) Significantly decreased the production of NO, PGE2, and inhibited the iNOS, COX-2, TNF-α and IL-6.
(b) Reduced the LPS-induced activation of nuclear factor-κB in RAW264.7 cells.
-6.12–100 µM6.12–100 µM[54]
Parkinson’s disease activityIn vitro(a) MAO-A inhibitory activity
(b) MAO-B inhibitory activity
-200 µM(a) IC50 = 27.23 µM
(b) IC50 = 174.40 µM
[19]
ObtusinAntidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(a) IC50 = 6.44 µg/mL
(b) IC50 = 20.92 µg/mL
[9]
Anti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) Butyrylcholinesterase inhibitory activity
(c) β-secretase inhibitory activity
-0–100 µg/mL(a) IC50 = 82 µg/mL
(b) IC50 = 287 µg/mL
(c) IC50 = 61.9 µg/mL
[10]
Anti-cancer activityIn vitroCytotoxicity against (a) HCT-116, (b) A549, and (c) SGC7901 cell lines-0.4–50 µg/mL(a) IC50 = 13.1 µg/mL
(b) IC50 = 29.2 µg/mL
(c) IC50 = 15.2 µg/mL
[52]
Parkinson’s disease activityIn vitro(a) MAO-A inhibitory activity
(b) MAO-B inhibitory activity
-400 µM(a) IC50 = 11.12 µM
(b) IC50 ≥ 400 µM
[19]
Chryso-obtusinAnti-cancer activityIn vitroCytotoxicity against (a) HCT-116, (b) A549, (c) SGC7901 and (d) LO2 cell lines-0.4–50 µg/mL(a) IC50 = 10.5 µg/mL
(b) IC50 = 14.6 µg/mL
(c) IC50 = 12.0 µg/mL
(d) IC50 = 15.8 µg/mL
[52]
Anti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) Butyrylcholinesterase inhibitory activity
(c) β-secretase inhibitory activity
-0–100 µg/mL(a) IC50 = 68.6 µg/mL
(b) IC50 = 287 µg/mL
(c) IC50 = 49.9 µg/mL
[10]
Antidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(a) IC50 = 14.88 µg/mL
(b) IC50 = 36.1 µg/mL
[9]
Platelet anti-aggregatory activityIn vitro(a) Adenosine 5′-diphosphate inhibitory activity
(b) Arachidonic-acid inhibitory activity
(c) Collagen inhibitory activity
-0–1 mg/mL1 mg/mL[47]
Parkinson’s disease activityIn vitro(a) MAO-A inhibitory activity
(b) MAO-B inhibitory activity
-400 µM(a) IC50 = 327.67 µM
(b) IC50 ≥ 400 µM
[19]
QuestinAntimicrobial activityIn vitroAntibacterial activity against
(a) Staphylococcus aureus 209P and
(b) Escherichia coli NIHJ
-0–100 µg/mLMIC (a) 25 µg/mL
(b) 50 µg/mL
[48]
Anti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) Butyrylcholinesterase inhibitory activity
(c) β-secretase inhibitory activity
-0–100 µg/mL(a) IC50 = 34.0 µg/mL
(b) IC50 = 138 µg/mL
(c) IC50 = 32.8 µg/mL
[10]
Antidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(a) IC50 = 5.69 µg/mL
(b) IC50 = 136.1 µg/mL
[9]
Parkinson’s disease activityIn vitro(a) MAO-A inhibitory activity
(b) MAO-B inhibitory activity
-20 µM(a) IC50 = 0.17 µM
(b) IC50 = 10.58 µM
[19]
Gluco-aurantio-obtusinPlatelet anti-aggregatory activityIn vitro(a) Adenosine 5′-diphosphate inhibitory activity
(b) Arachidonic-acid inhibitory activity
(c) Collagen inhibitory activity
-0–1 mg/mL(a) IC50 = 0.25 µg/mL
(b) IC50 = 0.05 µg/mL
(c) IC50 = 0.1 µg/mL
[5]
Anti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) β-secretase inhibitory activity
-0–100 µg/mL(a) IC50 = 109 µg/mL
(b) IC50 = 50.9 µg/mL
[10]
Antidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(a) IC50 = 31.3 µg/mL
(b) IC50 = 142.1 µg/mL
[9]
Hepatoprotective activityIn vitroHepatoprotective efficacy against t-BHP-induced cell death in HepG2 cells-20 µMProtection ratio value 49.7% at 20 µM[12]
Parkinson’s disease activityIn vitro(a) MAO-A inhibitory activity
(b) MAO-B inhibitory activity
-400 µM(a) IC50 = 39.55 µM
(b) IC50 = 180.76 µM
[19]
Chrysophanol; Aloe-emodin; Physcion; Chrysophanol tri, Tetraglucoside; 2-hydroxyemodin-1methylether; Chryso-obtusin-2-O-β-d-glucosideAntidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(a) IC50 = 5~103 µg/mL
(b) IC50 = 5~228 µg/mL
[9]
Anti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) Butyrylcholinesterase inhibitory activity
(c) β-secretase inhibitory activity
-0–400 µg/mL(a) IC50 = 14~71 µg/mL
(b) IC50 ≥ 100 µg/mL
(c) IC50 = 13~59 µg/mL
[10]
Parkinson’s disease activityIn vitro(a) MAO-A inhibitory activity
(b) MAO-B inhibitory activity
-400 µM(a) IC50 = 2.47~400 µM
(b) IC50 ≥ 400 µM
[19]
DihydroxyanthraquinoneBacterial growth promoting and inhibiting activityIn vitro(a) Growth promoting activity against Bifidobacterium bifidum
(b) Growth inhibiting activity against Clostridium perfringens and Escherichia coli
-(a) 0.05–0.5 mg/d
(b) 0.1–5 mg/d
(a) GIR > 2.0 at 0.5 mg/disk
(b) Inhibitory zone diameter > 30 mm
[7]
Naphthopyrone
CassiasideAnti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) Butyrylcholinesterase inhibitory activity
(c) β-secretase inhibitory activity
-0–100 µg/mL(a) IC50 = 18.1 µg/mL
(b) IC50 = 177 µg/mL
(c) IC50 = 1.85 µg/mL
[10]
Antidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(a) IC50 = 48.55 µg/mL
(b) IC50 = 129.2 µg/mL
[9]
Hepatoprotective activityIn vitroHepatoprotective efficacy against t-BHP-induced cell death in HepG2 cells 25 µM(a) protect cells damage
(b) increased GSH level and reduce ROS level
[37]
Parkinson’s disease activityIn vitro(a) MAO-A inhibitory activity
(b) MAO-B inhibitory activity
-400 µM(a) IC50 = 11.26 µM
(b) IC50 ≥ 400 µM
[19]
Isotoralactone; ToralactoneAntimicrobial activityIn vitroAntibacterial activity against
(a) Staphylococcus aureus 209P and
(b) Escherichia coli NIHJ
-0–100 µg/mLMIC (a) 2–3 µg/Ml
(b) 5.5–12 µg/mL
[46]
Cassiaside B2, Cassiaside C2Antiallergic activityIn vitroInhibition of histamine release in rat peritoneal mast cells-100 µMCassiaside B2 inhibit 17.2%; Cassiaside C2
Inhibit 53.9%
[6]
Toralactone GentiobiosideAntidiabetic activityIn vitro(a) PTP 1B inhibitory activity
(b) α-glucosidase inhibitory activity
-(a) 0–100 µg/mL
(b) 0–400 µg/mL
(a) IC50 = 81.1µg/mL
(b) IC50 = 37.60 µg/mL
[9]
Anti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) Butyrylcholinesterase inhibitory activity
(c) β-secretase inhibitory activity
-0–100 µg/mL(a) IC50 = 91.3 µg/mL
(b) IC50 = 117 µg/mL
(c) IC50 = 69.0 µg/mL
[10]
Hepatoprotective activityIn vitroHepatoprotective efficacy against t-BHP-induced cell death in HepG2 cells-20 µMIncreased in Nrf2/ARE-luciferase activity, and upregulated NQO1, GLC, HO-1 levels[12]
rubrofusarin, Rubrofusarin 6-O-β-d-glucopyranoside, Rubrofusarin 6-O-β-d-gentiobioside, Nor-rubrofusarin 6-O-β-d-glucosideAnti-Alzheimer’s activityIn vitro(a) Acetylcholinesterase inhibitory activity
(b) β-secretase inhibitory activity
-(a) 0–100 µM
(b) 0–750 µM
(a)15.95–148 µM
(b) 14.0–190 µM
[55]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ali, M.Y.; Park, S.; Chang, M. Phytochemistry, Ethnopharmacological Uses, Biological Activities, and Therapeutic Applications of Cassia obtusifolia L.: A Comprehensive Review. Molecules 2021, 26, 6252. https://doi.org/10.3390/molecules26206252

AMA Style

Ali MY, Park S, Chang M. Phytochemistry, Ethnopharmacological Uses, Biological Activities, and Therapeutic Applications of Cassia obtusifolia L.: A Comprehensive Review. Molecules. 2021; 26(20):6252. https://doi.org/10.3390/molecules26206252

Chicago/Turabian Style

Ali, Md Yousof, Seongkyu Park, and Munseog Chang. 2021. "Phytochemistry, Ethnopharmacological Uses, Biological Activities, and Therapeutic Applications of Cassia obtusifolia L.: A Comprehensive Review" Molecules 26, no. 20: 6252. https://doi.org/10.3390/molecules26206252

Article Metrics

Back to TopTop