Next Article in Journal
Reactive Paper Spray Ionization Mass Spectrometry for Rapid Detection of Estrogens in Cosmetics
Previous Article in Journal
Exploring the Anti-Cancer Effects of Fish Bone Fermented Using Monascus purpureus: Induction of Apoptosis and Autophagy in Human Colorectal Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Gypsogenin Battling for a Front Position in the Pentacyclic Triterpenes Game of Thrones on Anti-Cancer Therapy: A Critical Review—Dedicated to the Memory of Professor Hanaa M. Rady

by
Mohamed O. Radwan
1,2,*,
Howaida I. Abd-Alla
2,
Azhaar T. Alsaggaf
3,
Hatem El-Mezayen
4,
Mohammed A. S. Abourehab
5,6,
Mohamed E. El-Beeh
7,
Hiroshi Tateishi
1,
Masami Otsuka
1,8 and
Mikako Fujita
1,*
1
Medicinal and Biological Chemistry Science Farm Joint Research Laboratory, Faculty of Life Sciences, Kumamoto University, Kumamoto 862-0973, Japan
2
Chemistry of Natural Compounds Department, National Research Centre, Giza 12622, Egypt
3
Department of Chemistry, Taibah University, Madinah 42353, Saudi Arabia
4
Biochemistry Department, Helwan University, Cairo 11795, Egypt
5
Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Minia University, Minia 61519, Egypt
6
Department of Pharmaceutics, Faculty of Pharmacy, Umm Al-Qura University, Makkah 21955, Saudi Arabia
7
Biology Department, Al-Jumum University College, Umm Al-Qura University, Makkah 21955, Saudi Arabia
8
Department of Drug Discovery, Science Farm Ltd., Kumamoto 862-0976, Japan
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(15), 5677; https://doi.org/10.3390/molecules28155677
Submission received: 3 July 2023 / Revised: 19 July 2023 / Accepted: 25 July 2023 / Published: 27 July 2023

Abstract

:
In the last decade, gypsogenin has attracted widespread attention from medicinal chemists by virtue of its prominent anti-cancer potential. Despite its late identification, gypsogenin has proved itself as a new anti-proliferative player battling for a frontline position among other classic pentacyclic triterpenes such as oleanolic acid, glycyrrhetinic acid, ursolic acid, betulinic acid, and celastrol. Herein, we present the most important reactions of gypsogenin via modification of its four functional groups. Furthermore, we demonstrate insights into the anti-cancer activity of gypsogenin and its semisynthetic derivatives and go further by introducing our perspective to judiciously guide the prospective rational design. The present article opens a new venue for a better exploitation of gypsogenin chemical entity as a lead compound in cancer chemotherapy. To the best of our knowledge, this is the first review article exploring the anti-cancer activity of gypsogenin derivatives.

Graphical Abstract

1. Introduction

Cancer is the second major global cause of mortality preceded with cardiovascular diseases [1,2,3,4]. Having said that, cancer cases are soaring at an alarming worldwide rate. Surprisingly, in some countries, cancer has exceeded cardiovascular disorders as a leading mortality cause [5]. This horrifying fact has been discussed in a cohort study that pointed out a transition in the main causes of deaths among youth in some countries [6]. Medicinal chemists are continuously urged to innovate new chemical entities to surmount resistance, reduce side effects, and enhance the efficacy of commercial drugs in the hard-fought battle against cancer [7,8,9,10,11,12]. Many natural products have provided skeletons and structural references for the invention of modern drugs [13,14,15,16]. Found in higher plants, pentacyclic triterpenes (PTs) are bio-nutrient phytochemicals endowed with a diverse range of bioactivities such as hepatoprotective [17,18,19], anti-inflammatory [20,21,22], anti-hypertensive [19,23,24,25], anti-atherosclerotic [23,24], anti-viral [26,27,28], anti-fibrosis [29,30,31], and anti-ulcer effects [20,32,33]. In particular, PTs have ubiquitous applications in terms of anti-cancer drug discovery [34,35,36,37,38,39,40,41,42].
The literature is loaded with plenty of success stories linking PTs derivatives with a prominent role in the prevention of cancer initiation, promotion, angiogenesis, and progression through disrupting different intermittent mechanisms and pathways. The number of scientific publications and citations linking PTs and cancer has been soaring over the past twenty years, according to the Web of Science database (Figure 1). PTs are generally non-cytotoxic, albeit minor derivatizations can lead to dramatic changes in activity.
PTs comprise four main chemical skeletons, namely oleanane, ursane, lupane, and friedelane. Oleanolic acid, from the oleanane type, is one of the most extensively studied PTs in terms of medicinal chemistry. Oleanolic acid suppresses proliferation of hepatocellular carcinoma [42,43], human bladder cancer [44], breast cancer [45,46], lung carcer [47], and colon cancer [48,49]. To possess such diverse activities, oleanolic acid modulates multiple cell-signaling pathways [50]. Two oleanolic acid derivatives, CDDO and CDDO-Me, have already entered clinical trials for the treatment of solid tumors and lymphoma, allowing oleanolic acid to top the throne of pentacyclic triterpenes in terms of chemotherapy [51,52]. Glycyrrhetinic acid is another representative of oleanane-type triterpenoids with ubiquitous anti-cancer activities [53,54,55,56,57,58]. Ursolic acid [48,59,60,61,62], betulinic acid [63,64,65,66,67,68], and celastrol [69,70,71,72], representing ursane, lupane, and friedelane type triterpenoids, respectively, were reported to possess multifaceted anti-cancer properties.
Gypsogenin (3-hydroxy-23-oxoolean-12-en-28-oic acid), a less-explored PT, extracted from Gypsophila oldhamiana in a saponin form linked with sugar moieties. It is generated as pure sapogenin via acid hydrolysis [73]. It has an oleanane-type skeleton and possesses four active sites, C-3 hydroxyl, ring C double bond, C-23 aldehyde group and C-28 carboxylic acid, which are amenable to a wide range of chemical transformations (Figure 2). The hydroxyl, alkene, and carboxyl groups exist in most PTs. Nevertheless, the aldehyde group is unique, as other classic triterpenes lack such a group, which represents a structural alert for most medicinal chemists due to its high reactivity [74].
Previously, aldehydes used to have an unfavorable reputation due to their toxicity and metabolic instability. Nonetheless, in modern chemical biology, they have been applied as covalent probes to target lysine residues in proteins by forming a covalent imine adduct. In this regard, roblitinib development as exquisitely selective inhibition of FGFR4 signaling was based on the presence of an aldehyde group. The latter is responsible for creating a reversible-covalent bond with the target while avoiding the safety concerns of irreversible covalent inhibitors [75]. Taken together, the aldehyde group will play an important role in drug discovery in the 21st century to find ligands for traditionally undruggable targets [74,76]. This may give gypsogenin and advantage over other PTs.
Recently, gypsogenin proved itself as an outstanding entity that can enter the competition between PTs for a frontline position as a lead anti-cancer agent. Most previous reports linked gypsogenin to anti-cancer effects. It is unlikely that other bioactivity will be found for gypsogenin and its derivatives; one example is the observed strong inhibition of acetylcholinesterase, which provides a basis for potential Alzheimer’s therapy involving natural products [77]. Stunningly, the first carboxamide series of gypsogenin came out in 2018, which points out the shortage of structure–activity relationship (SAR) studies on this precious PT [73]. Moreover, no gypsogenin derivatives with modified ring C were synthesized before 2023.
Several PTs exhibit limited water solubility and low bioavailability, which can be addressed by derivatization [78]. Derivatization not only optimizes triterpenes’ pharmacokinetics, but also their pharmacodynamics. Herein, we summarized the chemical modifications of gypsogenin four functional groups and focused on the anti-cancer effect of gypsogenin and its semi-derivatives. We generated SAR for gypsogenin and its derivatives against leukemia, breast cancer, and lung cancer. We present our recommendations for prospective work and the missing information that should be addressed. Our study represents a cornerstone reference for any future research linking gypsogenin and cancer. We believe that future extensive SAR studies of gypsogenin will advance it to a frontline position in the pentacyclic triterpenes Game of Thrones on anti-cancer therapy.

2. Methodology

This review article is the first to discuss gypsogenin and its derivative from a medicinal chemistry perspective. We used the keywords gypsogenin derivative and anti-cancer for our search in PubMed and Web of Science. This disclosed approximately 60 articles and patents, of which 27 were considered for this review. As this study focuses on medicinal chemistry aspects, we excluded the anti-cancer activity of the naturally found gypsogenin saponins and considered the semi-synthetic derivatives of gypsogenin for this review.

3. Gypsogenin Extraction and Chemical Transformation

The difficulty of isolation of gypsogenin from plants and the high price of commercially available gypsogenin limited extensive SAR studies. One extraction example showed that starting with 20 kg of air-dried roots of Gypsophila oldhamiana yields as little as 1.3 g of pure gypsogenin. The procedures were initiated via water extraction of the water-soluble saponins before drying under a vacuum. The mixture was subjected to acid hydrolysis using 10% HCl for 72 h before neutralization with NaOH and extraction with ethyl acetate. After evaporation, the mixture was applied to column chromatography using 10:1 hexane–ethyl acetate eluent to give rise to gypsogenin as a white solid [73,79,80]. Gypsogenin can also be found in other species of Gypsophila, such as bermejoi, simonii [81], paniculate, and arrostii [82]. Additionally, it is available in plants belonging to the Caryophyllaceae family, such as Agrostemma githago (Lychnis githago) [83,84], Melandrium firmum [85], and different Stellaria species [86,87]. Furthermore, plants that belong to the Amaranthaceae family, e.g., Beta vulgaris L [88] and Chenopodium quino [89], contain gypsogenin. Greatrex et al. synthesized gypsogenin from the alcoholic PT analogue, hederagenin, via oxidation [90].
As we mentioned above, gypsogenin has four functional groups that can be feasibly modified to enhance its pharmacodynamic and pharmacokinetic profile. The 3-OH group was acetylated using the conventional method used for other PTs—reflux with acetic anhydride in dry pyridine—as described by Emirdag et al. [91]. The addition of dimethyl amino pyridine (DMAP) as a catalyst was used elsewhere to improve yield [77,92]. The 3-OH group was recently oxidized, forming the 3-keto analogue. This was achieved by mixing gypsogenin with Dess–Martin periodinane in dichloromethane at 0 °C for 15 min [92]. The authors also reported 3-OH etherification using different alkyl bromides in the presence of potassium iodide and potassium carbonate in dimethyl formamide (DMF) at 60 °C [92]. Dehydration of gypsogenin by thionyl chloride in (DMF) eliminates the 3-OH group and produces its unsaturated 2,3 dehydro- analogue [92].
Gypsogenic acid (Figure 2), the dicarboxylic acid analogue of gypsogenin, can be isolated from Gypsophila oldhamiana roots, especially if a portion of gypsogenin is transformed into gypsogenic acid during the hydrolysis step. In addition, its 3-acetyl analogue was synthesized through oxidation of 3-acetyl gypsogenin (1) by sodium hypochlorite and hydrogen peroxide in the presence of sodium dihydrogen phosphate at room temperature [93]. A similar oxidation process could be achieved via vigorous stirring with potassium permanganate in ethanol water mixture at room temperature [93].
The 4-aldehyde group of gypsogenin is versatile and has been reacted in different ways. Its oximation by using hydroxylamine hydrochloride in pyridine at 105 °C afforded compound 2 in a good yield (Figure 3) [73,91]. It was also reacted with thiosemicarbazide in a 1:1 MeOH: water mixture under reflux forming a thiosemicarbazone analogue [91]. Another amination of gypsogenin’s 4-aldehyde was performed in acetic acid using phenyl hydrazine or 2,4-dinitrophenylhydrazine solvent at room temperature; the latter resulted in the formation of Schiff base 5 [73].
We have performed reductive amination of gypsogenin’s 4-aldehyde group using different amines and sodium triactoxyborohydride for in situ reduction of the formed Schiff base in dichloroethane solvent at room temperature (compounds 12, 13, 14, 15, and 17) [94,95]. The yield of this reaction was generally poor due to the low solubility of gypsogenin in dichloroethane. That is why another group performed this reaction in methanol while using sodium borohydride as a reducing agent to obtain compound 16 [92].
The third functional group of gypsogenin is 28-COOH, which is widely found in PTs. A feasible esterification process involves activation by potassium carbonate in DMF at room temperature, followed by addition of appropriate alkyl bromide. This was applied for synthesis of 6 [95], 8, and 9 [96] in good yields. Hybrids of gypsogenin and chalcones achieved via ester bond were disclosed in a recent patent [79]. Esterification of 3-acetyl gypsogenin with different substituted chalcones was achieved using N, N′-Dicyclohexylcarbodiimide (DCC) in the presence of 4-dimethylaminopyridine compounds 10 and 11 (Figure 3).
Different amides of 3-acetyl gypsogenin were produced via activation of the carboxyl group with oxalyl chloride, followed by addition of the appropriate amine in the presence of triethyl amine as a catalyst in dichloromethane [73,92]. This general method was applied for the synthesis of the amides shown in Figure 3, such as compounds 18, 19 [92], and 20 [93] in good yields. Bisamidation was performed for 3-acetyl gypsogenic acid, adopting the same procedures to obtain derivatives such as 22 through reaction with two different amines for each carboxyl group [93]. Some reported bisamides were synthesized by reacting dichloride of gypsogenic acid with the two molar equivalents of the same appropriate amine [77,93].
Facile oxidation approaches of ring C were recently conducted using different conditions resulting in different products. Stirring of gypsogenin with hydrogen peroxide and formic acid in dichloromethane at room temperature afforded the epoxide congener (24). On the other hand, oxidation of gypsogenin using selenium dioxide in acetic acid under reflux gave rise to the 11-keto derivative (25) [92] (Figure 4). The produced enone system of ring C imitates that naturally found in glycyrrhetinic acid. The molecular formula and molecular weight of the compounds in Figure 2, Figure 3 and Figure 4 were summarized in Table S1 (Supporting data).

4. Anti-Cancer Effect of Gypsogenin, Gypsogenic Acid, and Their Semisynthetic Derivatives

4.1. Anti-Leukemic Activity

Gypsogenic acid did not show observable activity against chronic myeloid leukemia (K562) and acute myeloid leukemia (HL-60), where its IC50 exceeded 100 µM for both cells [97]. Another study was in accordance with this, recording the activity of gypsogenic acid IC50 against K562 as 227.6 µM; however, HL-60 was more sensitive (IC50 61.1 µM) [98]. The latter value shows a discrepancy with the previous report by Lee’s group [97]. Gypsogenic acid demonstrated low activity against other lymphoid leukemias SKW-3 and BV-173 (IC50 79.1 and 41.4 µM, respectively) [98].
Later on, we found that gypsogenin highly outperforms gypsogenic acid with IC50 12.7 µM against K562, highlighting the crucial role of the 4-aldehyde group [96]. Simultaneously, Emirdag et al. revealed that gypsogenin has anti-proliferative effect on HL-60 (IC50 10.4 µM) by inducing apoptosis [82,91]. 3-acetyl gypsogenin, 1, has almost the same effect of gypsogenin on HL-60 (IC50 10.77 µM). Gypsogenin activity is increased by oximation of its aldehyde group (compound 2 IC50 3.9 µM). Mutually, the 3-acetylated oxime analogue 3 surpassed the activity of 1 (IC50 5.9 µM) [91]. Gypsogenin benzyl ester 6 has IC50 8.1 µM; however, its acetylation product 7 has IC50 6.7 µM [91].
By virtue of its notable apoptotic effect, 6 was further benchmarked for its effect on K562 cell line, where it showed moderate activity (IC50 9.3 µM) [99]. However, this study represented a turning point for a better understanding of gypsogenin’s molecular target. Compound 6 inhibited ABL1 tyrosine kinase with IC50 8.71 µM. This is assumed to be the main target for its cytotoxic effect on K562. It is needless to say that the presence of other off targets cannot be excluded. Concomitantly, 6 inhibited other kinases such as C-terminal Src kinase (CSK) and Lyn kinase isoform B; LYN B (IC50 1.5 µM and 2.9 µM, respectively) [99]. It is clear that oximation of 6 is detrimental for its activity on both K562 and HL-60, as the respective IC50 value of 4 is 21.3 µM and 10.6 µM [99].
Ciftci et al. moved forward with a structure–activity relationship study of 6 and succeeded in enhancing its activity [96]. As mentioned above, the free aldehyde group is crucial for activity against leukemia. Therefore, Ciftci et al. came up with substituted congeners of 6, keeping a free 4-aldehyde group [96]. Compounds 8 and 9 have IC50 4.7 and 3.1 µM, respectively, against K562 cells. Additionally, IC50 of 8 and 9 for ABL1 tyrosine kinase was 7.1 µM and 6.1 µM, respectively. Both compounds have induced an explicit apoptosis effect, especially 8, whose apoptosis induction was clearer than imatinib, a gold standard ABL1 kinase inhibitor for CML therapy. Concomitantly, 8 suppressed the downstream signaling of extracellular signal-regulated kinase (ERK) phosphorylation [96]. In a similar vein, both compounds exhibited moderate activity on MT-2 and Jurkat cells. Interestingly, the IC50 of 9 for MT-2 and Jurkat was 7.2 µM and 4.8 µM, respectively. The authors evaluated both compounds to determine their cytotoxic effect on peripheral blood mononuclear cells (PBMC) and calculated the selectivity index as the ratio of IC50 between PBMC and K562 cells. The higher selectivity index value of compound 8, 11.0, than compound 9, 8.0, reflects the favorable safety profile of compound 8.
A recent report by Ulusoy et al. showed that reductive amination of the 4-aldehyde group with different aromatic and alicyclic amines leads to either reduction or complete abrogation of anti-K562 activity [95]. The hit compound in this study, 13, had IC50 11.3 µM which is even less active than the parent compound, gypsogenin [95]. Furthermore, 13 inhibited ABL1 kinase in a moderate fashion (IC50 value of 13.0 µM). This is further evidence of the crucial role of the 4-aldehyde group for anti-K562 activity (Figure 5). In addition, 13 had less effect on MT-2 and Jurkat than 8 and 9. Compound 13 had moderate effect on a panel of kinases at 30 µM of drug concentration, especially for BRK, BTK, LYN B, and SRC. Compound 14 with the more hydrophobic 4-isopropyl substitution exhibited less activity (IC50 23.8 µM), whereas the presence of a bulky N-piperazinyl benzyl moiety abolished activity as shown for 12 (IC50 > 100 µM). The activity was also abolished in the presence of an electron-donating substitution, as was the case for 15 (IC50 > 100 µM).
So far, there has been no report linking gypsogenin or gypsogenic acid carboxamides and leukemia. This is the same case for modified ring C derivatives and gypsogenin–chalcone hybrids. In a word, gypsogenin benzyl esters have been the most active derivatives against K562 and HL-60 leukemias until now. The SAR pertaining to activity against K562 and HL-60 is afforded in Figure 5.

4.2. Anti-Breast Cancer Activity

Gypsogenin has moderate cytotoxic activity for MCF-7 (IC50 9.0 µM); however, its benzyl ester derivative 6 has IC50 5.1 µM [91]. Surprisingly, substituted benzyl esters such as 8 and 9 showed less activity than gypsogenin with respective IC50 51.58 µM and 15.3 µM. Notably, the 3-acetyl analogues 1 and 7 possess less activity (IC50 20.5 µM and 65.1 µM, respectively). However, oximation of gypsogenin and 6 slightly improves their cytotoxic effect, as shown for 2 and 4. The exact mechanism of action is yet to be elucidated [91]. Notably, compound 1 has low IC50 value of 5.4 µM against triple-negative breast cancer cell (TNBC) line (MDA-MB-231). In this regard, two gypsogenin–chalcone hybrids demonstrated moderate effect, too, namely, 10 and 11 with respective IC50 11.0 µM and 7.9 µM [79]. This can be a clue for targeting TNBC, which is an aggressive form of breast cancer that does not respond to hormonal therapy [100].
Wu et al. found that gypsogenic acid has a weak antiproliferative effect on MCF-7 (IC50 26.8 µM), which also highlights the role of the 4-aldehyde group. The authors highly enhanced gypsogenin and gypsogenic acid activity through mono-and bisamidation [93]. Gypsogenin carboxamide with imidazole, compound 20, has IC50 3.7 µM, which is similar to the gypsogenic acid mono-amide of only C28 with pyrazole, compound 23, whose IC50 is 3.8 µM. Gypsogenic acid bisamide of both C23 and C28, compound, 22 demonstrated pronounced activity (IC50 4.1 µM). The favorable safety profile of those carboxamides is shown by measuring their activity on human umbilical vein endothelial cells (HUVEC cells). It was determined that 22 possesses the highest selectivity index (24.0) among the mentioned active compounds.
Further evidence of the efficiency of gypsogenin amides was disclosed this year by Sun et al. [92]. Two amides, 18 and 19, possess IC50 5.7 µM and 13.8 µM, respectively, towards MCF-7. They also synthesized compound 16 via reductive amination reaction using methylamine; its IC50 is 11.3 µM, which is greater than that of gypsogenin (IC50 9.0 µM). The selectivity index of 16, 18, and 19 exceeds 30 when related to their effect on HUVEC.
Ring C-oxidized gypsogenin derivatives have recently been developed (Figure 4) [92]. The epoxide derivative (24) has IC50 26.6 µM on MCF-7. In parallel with this, the 11-keto derivative (25) has similar activity (IC50 25.3 µM), implying that oxidation of ring C reduces MCF-7 sensitivity. Conclusively, gypsogenin carboxamides are the best cytotoxic entities against MCF-7 compared to other derivatives (Figure 6).

4.3. Anti-Lung Cancer Activity

Gypsogenin can inhibit the growth and metastasis of Lewis lung cancer through inhibition of tumor angiogenesis and induction of apoptosis [101]. Different molecular targets were implicated in this mechanism. Gypsogenin downregulated mutant P53 and vascular endothelial growth factor (VEGF). It reduces the expression of Bcl-2 protein and raises Bax expression, promoting tumor apoptosis. The anti-proliferative effect of gypsogenin, (1), and 3-acetyl gypsogenic acid against A549 lung cancer cells is moderate (IC50 19.6, 30.8, and 23.7 µM, respectively) [73,93]. Oximation of gypsogenin and 1 maintains the activity without significant change [73]. 2,4-dinitrophenyl)hydrazono derivative of gypsogenin (5) demonstrated a strong cytotoxic effect on A549 cells (IC50 3.1 µM) [73,80]. In accordance, the amino product (16) exhibited stronger cytotoxic effect (IC50 1.5 µM) [92].
The two carboxamides 20 and 23 showed a bit higher activity than compound 5 (IC50 2.5 and 2.8 µM, respectively) [93]. Both compounds destroyed the cell membrane and increased its permeability, leading to the outflow of intracellular nucleic acid, but they weakly induced apoptosis and arrested A549 cell cycle of [93]. Another anti-lung cancer hit is the gypsogenic acid bisamidation product of (22), whose IC50 value is 2.0 µM. However, it is noteworthy that mono-amidation products 20 and 23 surpass its activity but with a lower selectivity index for HUVEC.
Concomitantly, compounds 18 and 19 showed a sub-micromolar effect on A549 (IC50 0.5 µM and 0.9 µM, respectively) and induced both apoptosis through damaging the cell membrane and arresting the cell cycle. Combining in silico and in vitro tools defined VEGF1 as a gypsogenin target [92]. Remarkably, compound 18 showed a higher binding affinity to VEGF1 than the parent compound, which is in accordance with the cytotoxicity results. Gypsogenin esters showed disappointing results, such as those found for 8, whose IC50 exceeds 100 µM and 9 which is less active than the parent compound (IC50 24.5 µM). On the contrary, esterification with chalcone moieties elevated A549 sensitivity; the IC50 of 10 and 11 is 4.9 µM, and 1.3 µM, respectively [79]. This result denotes the role of chalcone moiety in conferring gypsogenin with high activity.
The epoxide analogue (24) has almost the same activity as the parent compound (IC50.18.7 µM), whereas the 11-keto derivative (25) has slightly better activity (IC50.13.5 µM) [92]. In conclusion, gypsogenin carboxamides and chalcone hybrids are the most promising anti-proliferative entities against A549 (Figure 7).

4.4. Other Anti-Cancer Activities

A batch of gypsogenin derivatives demonstrated other notable anti-cancer effects. In this regard, we will focus mainly on compounds with at least single-digit micromolar IC50 values. Gypsogenin and its 3-acetyl form (1) possess remarkable cytotoxic activity against HeLa (cervical cancer) [79]. Compound 2 has notable anti-proliferative activity against SaoS-2 cells (osteosarcoma) and HeLa cells. Its 3-acetylated derivative (3) also has a similar effect on SaoS-2 but not on HeLa. It is noteworthy that gypsogenin has IC50 7.8 against SaoS-2 which is better than 1, 2, and 3. On the other hand, 3 is distinguished by its prominent activity against HT-29 cells (colorectal adenocarcinoma) [91] (Table 1).
Another study showed that gypsogenin can suppress gastric cancer cells NCI-N87 proliferation by targeting VEGF and MM-9 and promoting the expression of caspase-3 and Bax proteins [102]. Compounds 5 and 21 were reported mainly for targeting colon cancer cells (LOVO) through strong induction of apoptosis and dose-dependent S-phase arrest in cells. Both compounds exhibited moderate effect on SKOV3 (ovarian cancer) and HepG2 cells (Hepatocellular carcinoma) [73]. The amino compound 16 also exhibited notable activity against LOVO. Compounds 2 and 3 showed no or moderate activity towards LOVO [92]. The most active compound against LOVO cells is compound 8 with submicromolar cytotoxicity, implying that gypsogenin carboxamides usually outperform other derivatives [92] (Table 1).
Three amides were reported by Wu et al., 20, 22, and 23 with outstanding activities against HepG2, TE-1 (esophageal cancer), and MC3-8 (colon cancer) cells [93]. Gypsogenin–chalcone hybrids 10 and 11 showed outstanding activity against HeLa and pancreatic cancer cells (PANC-1). Gypsogenin 28-COOH ester 9 showed better activity in HeLa cells than 8 [96]. Ciftci et al. revealed new derivatives that suppress glioma proliferation through EGFR inhibition. The amino derivative compound 17 has the strongest effect against EGFR and glioma cells U251, T98G, and U87 (Table 1). Consequently, the titled compound clearly induced apoptosis of U251 cells in a comparable fashion to cisplatin. This study revealed that gypsogenin benzyl esters were less effective than 17 on glioma cells [94] (Table 1). Furthermore, at 30 µM concentration, compound 17 showed moderate inhibition for a panel of other kinases, including ABL1 tyrosine kinase.

5. Conclusions and Future Directions

Befitting its anti-cancer promise, we presented a critical review of gypsogenin and its derivatives. Gypsogenin possesses a versatile and unique aldehyde group that can be utilized to create covalent interactions with undruggable targets. We dissected how gypsogenin was employed for semi-synthesis by reacting its four functional groups, then we demonstrated the bioactivity of the most important derivatives in the literature. So far, gypsogenin carboxamides have demonstrated high cytotoxic activity against breast and lung cancer. The bisamides of gypsogenic acid possess prominent activity as well; however, their anti-leukemic activity is yet to be explored. Gypsogenin benzyl esters showed pronounced activity against CML. Ring C-modified gypsogenin derivatives are weak antiproliferative agents against lung and breast cancer, but they have not been tested for their anti-leukemic effect. Gypsogenin and its derivatives were reported to target kinases such as ABL1 and VEGF. The selectivity index of some active compounds is high, reflecting their potential high safety. Further medicinal chemistry studies on gypsogenin are urgently needed to afford more active hits and elucidate their other plausible molecular targets.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28155677/s1, Table S1: Molecular formula and molecular weight of compounds in Figure 2, Figure 3, and Figure 4.

Author Contributions

Conceptualization, M.O.R.; methodology, M.O.R.; software, M.O.R.; investigation, H.I.A.-A.; resources, A.T.A.; data curation H.I.A.-A.; writing—original draft preparation, M.O.R.; writing—review and editing, H.I.A.-A., A.T.A., H.E.-M. and M.F.; visualization, A.T.A., M.A.S.A. and M.E.E.-B.; supervision, H.I.A.-A., H.T., M.O. and M.F.; project administration, M.A.S.A., M.O. and M.F.; funding acquisition, M.A.S.A. and M.E.E.-B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Deanship of scientific research at Umm Al-Qura University by grant code (23UQU4340520DSR009).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the Deanship of scientific research at Umm Al-Qura University for supporting this work via grant code (23UQU4340520DSR009).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ferlay, J.; Soerjomataram, I.; Dikshit, R.; Eser, S.; Mathers, C.; Rebelo, M.; Parkin, D.M.; Forman, D.; Bray, F. Cancer Incidence and Mortality Worldwide: Sources, Methods and Major Patterns in GLOBOCAN 2012. Int. J. Cancer 2015, 136, E359–E386. [Google Scholar] [CrossRef] [PubMed]
  2. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global Cancer Statistics 2018: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA. Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [Green Version]
  3. Santucci, C.; Carioli, G.; Bertuccio, P.; Malvezzi, M.; Pastorino, U.; Boffetta, P.; Negri, E.; Bosetti, C.; La Vecchia, C. Progress in Cancer Mortality, Incidence, and Survival: A Global Overview. Eur. J. Cancer Prev. Off. J. Eur. Cancer Prev. Organ. ECP 2020, 29, 367–381. [Google Scholar] [CrossRef]
  4. Ferlay, J.; Colombet, M.; Soerjomataram, I.; Dyba, T.; Randi, G.; Bettio, M.; Gavin, A.; Visser, O.; Bray, F. Cancer Incidence and Mortality Patterns in Europe: Estimates for 40 Countries and 25 Major Cancers in 2018. Eur. J. Cancer Oxf. Engl. 1990 2018, 103, 356–387. [Google Scholar] [CrossRef]
  5. Ciftci, H.I.; Can, M.; Ellakwa, D.E.; Suner, S.C.; Ibrahim, M.A.; Oral, A.; Sekeroglu, N.; Özalp, B.; Otsuka, M.; Fujita, M.; et al. Anticancer Activity of Turkish Marine Extracts: A Purple Sponge Extract Induces Apoptosis with Multitarget Kinase Inhibition Activity. Investig. New Drugs 2020, 38, 1326–1333. [Google Scholar] [CrossRef]
  6. Dagenais, G.R.; Leong, D.P.; Rangarajan, S.; Lanas, F.; Lopez-Jaramillo, P.; Gupta, R.; Diaz, R.; Avezum, A.; Oliveira, G.B.F.; Wielgosz, A.; et al. Variations in Common Diseases, Hospital Admissions, and Deaths in Middle-Aged Adults in 21 Countries from Five Continents (PURE): A Prospective Cohort Study. Lancet 2019, 395, 785–794. [Google Scholar] [CrossRef] [PubMed]
  7. Tan, S.; Li, D.; Zhu, X. Cancer Immunotherapy: Pros, Cons and Beyond. Biomed. Pharmacother. Biomed. Pharmacother. 2020, 124, 109821. [Google Scholar] [CrossRef]
  8. Liu, Z.; Ren, Y.; Weng, S.; Xu, H.; Li, L.; Han, X. A New Trend in Cancer Treatment: The Combination of Epigenetics and Immunotherapy. Front. Immunol. 2022, 13, 809761. [Google Scholar] [CrossRef] [PubMed]
  9. Pereira, B.; Billaud, M.; Almeida, R. RNA-Binding Proteins in Cancer: Old Players and New Actors. Trends Cancer 2017, 3, 506–528. [Google Scholar] [CrossRef]
  10. Bhinder, B.; Gilvary, C.; Madhukar, N.S.; Elemento, O. Artificial Intelligence in Cancer Research and Precision Medicine. Cancer Discov. 2021, 11, 900–915. [Google Scholar] [CrossRef]
  11. Radwan, M.O.; Toma, T.; Arakaki, Y.; Kamo, M.; Inoue, N.; Koga, R.; Otsuka, M.; Tateishi, H.; Fujita, M. New Insight into the Bioactivity of Substituted Benzimidazole Derivatives: Repurposing from Anti-HIV Activity to Cell Migration Inhibition Targeting HnRNP M. Bioorg. Med. Chem. 2023, 86, 117294. [Google Scholar] [CrossRef] [PubMed]
  12. Rupaimoole, R.; Slack, F.J. MicroRNA Therapeutics: Towards a New Era for the Management of Cancer and Other Diseases. Nat. Rev. Drug Discov. 2017, 16, 203–222. [Google Scholar] [CrossRef] [PubMed]
  13. Alqathama, A.; Yonbawi, A.R.; Shao, L.; Bader, A.; Abdalla, A.N.; Gibbons, S.; Prieto, J.M. The in Vitro Cytotoxicity against Human Melanoma Cells, Tyrosinase Inhibition and Antioxidant Activity of Grewia Tenax Leaves Extracts. Boletin Latinoam. Caribe Plantas Med. Aromat. 2022, 22, 268–276. [Google Scholar] [CrossRef]
  14. Bader, A.; Abdalla, A.N.; Obaid, N.A.; Youssef, L.; Naffadi, H.M.; Elzubier, M.E.; Almaimani, R.A.; Flamini, G.; Pieracci, Y.; El-Readi, M.Z. In Vitro Anticancer and Antibacterial Activities of the Essential Oil of Forsskal’s Basil Growing in Extreme Environmental Conditions. Life 2023, 13, 651. [Google Scholar] [CrossRef]
  15. Abo-Elghiet, F.; Ibrahim, M.H.; El Hassab, M.A.; Bader, A.; Abdallah, Q.M.A.; Temraz, A. LC/MS Analysis of Viscum Cruciatum Sieber Ex Boiss. Extract with Anti-Proliferative Activity against MCF-7 Cell Line via G0/G1 Cell Cycle Arrest: An in-Silico and in-Vitro Study. J. Ethnopharmacol. 2022, 295, 115439. [Google Scholar] [CrossRef]
  16. Bader, A.; Bkhaitan, M.M.; Abdalla, A.N.; Abdallah, Q.M.A.; Ali, H.I.; Sabbah, D.A.; Albadawi, G.; Abushaikha, G.M. Design and Synthesis of 4-O-Podophyllotoxin Sulfamate Derivatives as Potential Cytotoxic Agents. Evid. Based Complement. Alternat. Med. 2021, 2021, e6672807. [Google Scholar] [CrossRef]
  17. Gutiérrez-Rebolledo, G.A.; Siordia-Reyes, A.G.; Meckes-Fischer, M.; Jiménez-Arellanes, A. Hepatoprotective Properties of Oleanolic and Ursolic Acids in Antitubercular Drug-Induced Liver Damage. Asian Pac. J. Trop. Med. 2016, 9, 644–651. [Google Scholar] [CrossRef] [Green Version]
  18. Xu, G.B.; Xiao, Y.H.; Zhang, Q.Y.; Zhou, M.; Liao, S.G. Hepatoprotective Natural Triterpenoids. Eur. J. Med. Chem. 2018, 145, 691–716. [Google Scholar] [CrossRef]
  19. Ayeleso, T.B.; Matumba, M.G.; Mukwevho, E. Oleanolic Acid and Its Derivatives: Biological Activities and Therapeutic Potential in Chronic Diseases. Molecules 2017, 22, 1915. [Google Scholar] [CrossRef] [Green Version]
  20. Aly, A.M.; Al-Alousi, L.; Salem, H.A. Licorice: A Possible Anti-Inflammatory and Anti-Ulcer Drug. AAPS PharmSciTech 2005, 6, E74–E82. [Google Scholar] [CrossRef] [Green Version]
  21. Tsai, S.J.; Yin, M.C. Antioxidative and Anti-Inflammatory Protection of Oleanolic Acid and Ursolic Acid in PC12 Cells. J. Food Sci. 2008, 73, H174–H178. [Google Scholar] [CrossRef]
  22. Radwan, M.O.; Ismail, M.A.H.; El-Mekkawy, S.; Ismail, N.S.M.; Hanna, A.G. Synthesis and Biological Activity of New 18β-Glycyrrhetinic Acid Derivatives. Arab. J. Chem. 2016, 9, 390–399. [Google Scholar] [CrossRef] [Green Version]
  23. Somova, L.I.; Shode, F.O.; Ramnanan, P.; Nadar, A. Antihypertensive, Antiatherosclerotic and Antioxidant Activity of Triterpenoids Isolated from Olea Europaea, Subspecies Africana Leaves. J. Ethnopharmacol. 2003, 84, 299–305. [Google Scholar] [CrossRef]
  24. Somova, L.O.; Nadar, A.; Rammanan, P.; Shode, F.O. Cardiovascular, Antihyperlipidemic and Antioxidant Effects of Oleanolic and Ursolic Acids in Experimental Hypertension. Phytomedicine 2003, 10, 115–121. [Google Scholar] [CrossRef]
  25. Zhang, S.; Liu, Y.; Wang, X.; Tian, Z.; Qi, D.; Li, Y.; Jiang, H. Antihypertensive Activity of Oleanolic Acid Is Mediated via Downregulation of Secretory Phospholipase A2 and Fatty Acid Synthase in Spontaneously Hypertensive Rats. Int. J. Mol. Med. 2020, 46, 2019–2034. [Google Scholar] [CrossRef] [PubMed]
  26. Pompei, R.; Laconi, S.; Ingianni, A. Antiviral Properties of Glycyrrhizic Acid and Its Semisynthetic Derivatives. Mini-Rev. Med. Chem. 2012, 9, 996–1001. [Google Scholar] [CrossRef] [PubMed]
  27. Sun, Z.-G.; Zhao, T.-T.; Lu, N.; Yang, Y.-A.; Zhu, H.-L. Research Progress of Glycyrrhizic Acid on Antiviral Activity. Mini-Rev. Med. Chem. 2019, 19, 826–832. [Google Scholar] [CrossRef] [PubMed]
  28. Tohmé, M.J.; Giménez, M.C.; Peralta, A.; Colombo, M.I.; Delgui, L.R. Ursolic Acid: A Novel Antiviral Compound Inhibiting Rotavirus Infection in Vitro. Int. J. Antimicrob. Agents 2019, 54, 601–609. [Google Scholar] [CrossRef] [PubMed]
  29. Yang, Y.; Huang, Y.; Huang, C.; Lv, X.; Liu, L.; Wang, Y.; Li, J. Antifibrosis Effects of Triterpene Acids of Eriobotrya Japonica (Thunb.) Lindl. Leaf in a Rat Model of Bleomycin-Induced Pulmonary Fibrosis. J. Pharm. Pharmacol. 2012, 64, 1751–1760. [Google Scholar] [CrossRef] [PubMed]
  30. Lee, M.K.; Lee, K.Y.; Jeon, H.Y.; Sung, S.H.; Kim, Y.C. Antifibrotic Activity of Triterpenoids from the Aerial Parts of Euscaphis Japonica on Hepatic Stellate Cells. J. Enzyme Inhib. Med. Chem. 2009, 24, 1276–1279. [Google Scholar] [CrossRef] [PubMed]
  31. Xiang, H.; Han, Y.; Zhang, Y.; Yan, W.; Xu, B.; Chu, F.; Xie, T.; Jia, M.; Yan, M.; Zhao, R.; et al. A New Oleanolic Acid Derivative against CCl4-Induced Hepatic Fibrosis in Rats. Int. J. Mol. Sci. 2017, 18, 553. [Google Scholar] [CrossRef] [Green Version]
  32. Farina, C.; Pinza, M.; Pifferi, G. Synthesis and Anti-Ulcer Activity of New Derivatives of Glycyrrhetic, Oleanolic and Ursolic Acids. Il Farm. 1998, 53, 22–32. [Google Scholar] [CrossRef] [PubMed]
  33. Somensi, L.B.; Costa, P.; Boeing, T.; Mariano, L.N.B.; Longo, B.; Magalhães, C.G.; Duarte, L.P.; Maciel e Silva, A.T.; de Souza, P.; de Andrade, S.F.; et al. Gastroprotective Properties of Lupeol-Derived Ester: Pre-Clinical Evidences of Lupeol-Stearate as a Potent Antiulcer Agent. Chem. Biol. Interact. 2020, 321, 108964. [Google Scholar] [CrossRef] [PubMed]
  34. Chudzik, M.; Korzonek-Szlacheta, I.; Król, W. Triterpenes as Potentially Cytotoxic Compounds. Molecules 2015, 20, 1610–1625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Tang, Z.-Y.; Li, Y.; Tang, Y.-T.; Ma, X.-D.; Tang, Z.-Y. Anticancer Activity of Oleanolic Acid and Its Derivatives: Recent Advances in Evidence, Target Profiling and Mechanisms of Action. Biomed. Pharmacother. 2022, 145, 112397. [Google Scholar] [CrossRef]
  36. Salvador, J.A.R.; Leal, A.S.; Valdeira, A.S.; Gonçalves, B.M.F.; Alho, D.P.S.; Figueiredo, S.A.C.; Silvestre, S.M.; Mendes, V.I.S. Oleanane-, Ursane-, and Quinone Methide Friedelane-Type Triterpenoid Derivatives: Recent Advances in Cancer Treatment. Eur. J. Med. Chem. 2017, 142, 95–130. [Google Scholar] [CrossRef]
  37. Laszczyk, M.N. Pentacyclic Triterpenes of the Lupane, Oleanane and Ursane Group as Tools in Cancer Therapy. Planta Med. 2009, 75, 1549–1560. [Google Scholar] [CrossRef]
  38. Ghante, M.H.; Jamkhande, P.G. Role of Pentacyclic Triterpenoids in Chemoprevention and Anticancer Treatment: An Overview on Targets and Underling Mechanisms. J. Pharmacopunct. 2019, 22, 55–67. [Google Scholar] [CrossRef]
  39. Shaheen, U.; Ragab, E.A.; Abdalla, A.N.; Bader, A. Triterpenoidal Saponins from the Fruits of Gleditsia Caspica with Proapoptotic Properties. Phytochemistry 2018, 145, 168–178. [Google Scholar] [CrossRef]
  40. Kumar, A.; Gupta, K.B.; Dhiman, M.; Arora, S.; Jaitak, V. New Pentacyclic Triterpene from Potentilla Atrosanguinea Lodd. as Anticancer Agent for Breast Cancer Targeting Estrogen Receptor-α. Nat. Prod. Res. 2022, 36, 4352–4357. [Google Scholar] [CrossRef]
  41. Ghosh, A.; Panda, C.K. Role of Pentacyclic Triterpenoid Acids in the Treatment of Bladder Cancer. Mini Rev. Med. Chem. 2022, 22, 1331–1340. [Google Scholar] [CrossRef] [PubMed]
  42. Liese, J.; Abhari, B.A.; Fulda, S. Smac Mimetic and Oleanolic Acid Synergize to Induce Cell Death in Human Hepatocellular Carcinoma Cells. Cancer Lett. 2015, 365, 47–56. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, X.; Bai, H.; Zhang, X.; Liu, J.; Cao, P.; Liao, N.; Zhang, W.; Wang, Z.; Hai, C. Inhibitory Effect of Oleanolic Acid on Hepatocellular Carcinoma via ERK-P53-Mediated Cell Cycle Arrest and Mitochondrial-Dependent Apoptosis. Carcinogenesis 2013, 34, 1323–1330. [Google Scholar] [CrossRef]
  44. Mu, D.-W.; Guo, H.-Q.; Zhou, G.-B.; Li, J.-Y.; Su, B. Oleanolic Acid Suppresses the Proliferation of Human Bladder Cancer by Akt/MTOR/S6K and ERK1/2 Signaling. Int. J. Clin. Exp. Pathol. 2015, 8, 13864–13870. [Google Scholar]
  45. Amara, S.; Zheng, M.; Tiriveedhi, V. Oleanolic Acid Inhibits High Salt-Induced Exaggeration of Warburg-like Metabolism in Breast Cancer Cells. Cell Biochem. Biophys. 2016, 74, 427–434. [Google Scholar] [CrossRef] [Green Version]
  46. Chakravarti, B.; Maurya, R.; Siddiqui, J.A.; Bid, H.K.; Rajendran, S.M.; Yadav, P.P.; Konwar, R. In Vitro Anti-Breast Cancer Activity of Ethanolic Extract of Wrightia Tomentosa: Role of pro-Apoptotic Effects of Oleanolic Acid and Urosolic Acid. J. Ethnopharmacol. 2012, 142, 72–79. [Google Scholar] [CrossRef] [PubMed]
  47. Zhao, X.; Liu, M.; Li, D. Oleanolic Acid Suppresses the Proliferation of Lung Carcinoma Cells by MiR-122/Cyclin G1/MEF2D Axis. Mol. Cell. Biochem. 2015, 400, 1–7. [Google Scholar] [CrossRef]
  48. Furtado, R.A.; Rodrigues, E.P.; Araújo, F.R.R.; Oliveira, W.L.; Furtado, M.A.; Castro, M.B.; Cunha, W.R.; Tavares, D.C. Ursolic Acid and Oleanolic Acid Suppress Preneoplastic Lesions Induced by 1,2-Dimethylhydrazine in Rat Colon. Toxicol. Pathol. 2008, 36, 576–580. [Google Scholar] [CrossRef]
  49. Janakiram, N.B.; Indranie, C.; Malisetty, S.V.; Jagan, P.; Steele, V.E.; Rao, C.V. Chemoprevention of Colon Carcinogenesis by Oleanolic Acid and Its Analog in Male F344 Rats and Modulation of COX-2 and Apoptosis in Human Colon HT-29 Cancer Cells. Pharm. Res. 2008, 25, 2151–2157. [Google Scholar] [CrossRef]
  50. Žiberna, L.; Šamec, D.; Mocan, A.; Nabavi, S.F.; Bishayee, A.; Farooqi, A.A.; Sureda, A.; Nabavi, S.M. Oleanolic Acid Alters Multiple Cell Signaling Pathways: Implication in Cancer Prevention and Therapy. Int. J. Mol. Sci. 2017, 18, 643. [Google Scholar] [CrossRef] [Green Version]
  51. Yadav, V.R.; Prasad, S.; Sung, B.; Kannappan, R.; Aggarwal, B.B. Targeting Inflammatory Pathways by Triterpenoids for Prevention and Treatment of Cancer. Toxins 2010, 2, 2428–2466. [Google Scholar] [CrossRef] [Green Version]
  52. Borella, R.; Forti, L.; Gibellini, L.; De Gaetano, A.; De Biasi, S.; Nasi, M.; Cossarizza, A.; Pinti, M. Synthesis and Anticancer Activity of CDDO and CDDO-Me, Two Derivatives of Natural Triterpenoids. Molecules 2019, 24, 4097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Hsu, Y.-C.; Hsieh, W.-C.; Chen, S.-H.; Li, Y.-Z.; Liao, H.-F.; Lin, M.-Y.; Sheu, S.-M. 18β-Glycyrrhetinic Acid Modulated Autophagy Is Cytotoxic to Breast Cancer Cells. Int. J. Med. Sci. 2023, 20, 444–454. [Google Scholar] [CrossRef]
  54. Chen, J.; Zhang, Z.-Q.; Song, J.; Liu, Q.-M.; Wang, C.; Huang, Z.; Chu, L.; Liang, H.-F.; Zhang, B.-X.; Chen, X.-P. 18β-Glycyrrhetinic-Acid-Mediated Unfolded Protein Response Induces Autophagy and Apoptosis in Hepatocellular Carcinoma. Sci. Rep. 2018, 8, 9365. [Google Scholar] [CrossRef] [Green Version]
  55. Sun, Y.; Dai, C.; Yin, M.; Lu, J.; Hu, H.; Chen, D. Hepatocellular Carcinoma-Targeted Effect of Configurations and Groups of Glycyrrhetinic Acid by Evaluation of Its Derivative-Modified Liposomes. Int. J. Nanomed. 2018, 13, 1621–1632. [Google Scholar] [CrossRef] [Green Version]
  56. Lee, C.S.; Kim, Y.J.; Lee, M.S.; Han, E.S.; Lee, S.J. 18beta-Glycyrrhetinic Acid Induces Apoptotic Cell Death in SiHa Cells and Exhibits a Synergistic Effect against Antibiotic Anti-Cancer Drug Toxicity. Life Sci. 2008, 83, 481–489. [Google Scholar] [CrossRef] [PubMed]
  57. Yamaguchi, H.; Noshita, T.; Yu, T.; Kidachi, Y.; Kamiie, K.; Umetsu, H.; Ryoyama, K. Novel Effects of Glycyrrhetinic Acid on the Central Nervous System Tumorigenic Progenitor Cells: Induction of Actin Disruption and Tumor Cell-Selective Toxicity. Eur. J. Med. Chem. 2010, 45, 2943–2948. [Google Scholar] [CrossRef] [PubMed]
  58. Roohbakhsh, A.; Iranshahy, M.; Iranshahi, M. Glycyrrhetinic Acid and Its Derivatives: Anti-Cancer and Cancer Chemopreventive Properties, Mechanisms of Action and Structure- Cytotoxic Activity Relationship. Curr. Med. Chem. 2016, 23, 498–517. [Google Scholar] [CrossRef] [PubMed]
  59. Zafar, S.; Khan, K.; Hafeez, A.; Irfan, M.; Armaghan, M.; ur Rahman, A.; Gürer, E.S.; Sharifi-Rad, J.; Butnariu, M.; Bagiu, I.-C.; et al. Ursolic Acid: A Natural Modulator of Signaling Networks in Different Cancers. Cancer Cell Int. 2022, 22, 399. [Google Scholar] [CrossRef]
  60. Raphael, T.J.; Kuttan, G. Effect of Naturally Occurring Triterpenoids Ursolic Acid and Glycyrrhizic Acid on the Cell-Mediated Immune Responses of Metastatic Tumor-Bearing Animals. Immunopharmacol. Immunotoxicol. 2008, 30, 243–255. [Google Scholar] [CrossRef]
  61. Kim, D.-K.; Baek, J.H.; Kang, C.-M.; Yoo, M.-A.; Sung, J.-W.; Kim, D.-K.; Chung, H.-Y.; Kim, N.D.; Choi, Y.H.; Lee, S.-H.; et al. Apoptotic Activity of Ursolic Acid May Correlate with the Inhibition of Initiation of DNA Replication. Int. J. Cancer 2000, 87, 629–636. [Google Scholar] [CrossRef] [PubMed]
  62. Liu, X.-S.; Jiang, J. Induction of Apoptosis and Regulation of the MAPK Pathway by Ursolic Acid in Human Leukemia K562 Cells. Planta Med. 2007, 73, 1192–1194. [Google Scholar] [CrossRef] [PubMed]
  63. Pisha, E.; Chai, H.; Lee, I.S.; Chagwedera, T.E.; Farnsworth, N.R.; Cordell, G.A.; Beecher, C.W.; Fong, H.H.; Kinghorn, A.D.; Brown, D.M. Discovery of Betulinic Acid as a Selective Inhibitor of Human Melanoma That Functions by Induction of Apoptosis. Nat. Med. 1995, 1, 1046–1051. [Google Scholar] [CrossRef] [PubMed]
  64. Hordyjewska, A.; Ostapiuk, A.; Horecka, A.; Kurzepa, J. Betulin and Betulinic Acid: Triterpenoids Derivatives with a Powerful Biological Potential. Phytochem. Rev. 2019, 18, 929–951. [Google Scholar] [CrossRef] [Green Version]
  65. Fulda, S. Betulinic Acid: A Natural Product with Anticancer Activity. Mol. Nutr. Food Res. 2009, 53, 140–146. [Google Scholar] [CrossRef]
  66. Selzer, E.; Pimentel, E.; Wacheck, V.; Schlegel, W.; Pehamberger, H.; Jansen, B.; Kodym, R. Effects of Betulinic Acid Alone and in Combination with Irradiation in Human Melanoma Cells. J. Investig. Dermatol. 2000, 114, 935–940. [Google Scholar] [CrossRef] [Green Version]
  67. Fulda, S.; Debatin, K.-M. Sensitization for Anticancer Drug-Induced Apoptosis by Betulinic Acid. Neoplasia 2005, 7, 162–170. [Google Scholar] [CrossRef] [Green Version]
  68. Li, X.; Jiang, W.; Li, W.; Dong, S.; Du, Y.; Zhang, H.; Zhou, W. Betulinic Acid-Mediating MiRNA-365 Inhibited the Progression of Pancreatic Cancer. Oncol. Res. 2023, 31, 505–514. [Google Scholar] [CrossRef]
  69. Lim, H.Y.; Ong, P.S.; Wang, L.; Goel, A.; Ding, L.; Li-Ann Wong, A.; Ho, P.C.; Sethi, G.; Xiang, X.; Goh, B.C. Celastrol in Cancer Therapy: Recent Developments, Challenges and Prospects. Cancer Lett. 2021, 521, 252–267. [Google Scholar] [CrossRef]
  70. Yang, H.; Chen, D.; Cui, Q.C.; Yuan, X.; Dou, Q.P. Celastrol, a Triterpene Extracted from the Chinese “Thunder of God Vine,” Is a Potent Proteasome Inhibitor and Suppresses Human Prostate Cancer Growth in Nude Mice. Cancer Res. 2006, 66, 4758–4765. [Google Scholar] [CrossRef] [Green Version]
  71. Nagase, M.; Oto, J.; Sugiyama, S.; Yube, K.; Takaishi, Y.; Sakato, N. Apoptosis Induction in HL-60 Cells and Inhibition of Topoisomerase II by Triterpene Celastrol. Biosci. Biotechnol. Biochem. 2003, 67, 1883–1887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Kannaiyan, R.; Manu, K.A.; Chen, L.; Li, F.; Rajendran, P.; Subramaniam, A.; Lam, P.; Kumar, A.P.; Sethi, G. Celastrol Inhibits Tumor Cell Proliferation and Promotes Apoptosis through the Activation of C-Jun N-Terminal Kinase and Suppression of PI3 K/Akt Signaling Pathways. Apoptosis Int. J. Program. Cell Death 2011, 16, 1028–1041. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, H.; Mu, Y.; Wang, F.; Song, L.; Sun, J.; Liu, Y.; Sun, J. Synthesis of Gypsogenin Derivatives with Capabilities to Arrest Cell Cycle and Induce Apoptosis in Human Cancer Cells. R. Soc. Open Sci. 2018, 5, 171510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Gampe, C.; Verma, V.A. Curse or Cure? A Perspective on the Developability of Aldehydes as Active Pharmaceutical Ingredients. J. Med. Chem. 2020, 63, 14357–14381. [Google Scholar] [CrossRef] [PubMed]
  75. Weiss, A.; Adler, F.; Buhles, A.; Stamm, C.; Fairhurst, R.A.; Kiffe, M.; Sterker, D.; Centeleghe, M.; Wartmann, M.; Kinyamu-Akunda, J.; et al. FGF401, A First-In-Class Highly Selective and Potent FGFR4 Inhibitor for the Treatment of FGF19-Driven Hepatocellular Cancer. Mol. Cancer Ther. 2019, 18, 2194–2206. [Google Scholar] [CrossRef]
  76. Covalent Docking of Large Libraries for the Discovery of Chemical Probes|Nature Chemical Biology. Available online: https://www.nature.com/articles/nchembio.1666 (accessed on 29 June 2023).
  77. Heller, L.; Schwarz, S.; Weber, B.A.; Csuk, R. Gypsogenin Derivatives: An Unexpected Class of Inhibitors of Cholinesterases. Arch. Pharm. 2014, 347, 707–716. [Google Scholar] [CrossRef]
  78. Furtado, N.A.J.C.; Pirson, L.; Edelberg, H.; Miranda, L.M.; Loira-Pastoriza, C.; Preat, V.; Larondelle, Y.; André, C.M. Pentacyclic Triterpene Bioavailability: An Overview of In Vitro and In Vivo Studies. Molecules 2017, 22, 400. [Google Scholar] [CrossRef] [Green Version]
  79. Gypsogenin Türevi Ile Kalkon Bileşiklerinin Yeni Yari Sentezi ve bu Türevlerinin Insan Kanser Hücre Hatlari Üzerindeki Çalişmalari—TR201922043A2|PatentGuru. Available online: https://www.patentguru.com/TR201922043A2 (accessed on 11 July 2023).
  80. Sun, J.; Zhang, H.; Mou, Y.; Sun, J.; Wang, F.; Wu, Z.; Wang, Y.; Song, L. Gypsogenin Derivatives. CN107236017A. 10 October 2017. Available online: https://patents.google.com/patent/CN107236017A/en?oq=CN+107236017A (accessed on 11 July 2023).
  81. Acebes, B.; Díaz-Lanza, A.M.; Bernabé, M. A Saponin from the Roots of Gypsophila bermejoi. Phytochemistry 1998, 49, 2077–2079. [Google Scholar] [CrossRef]
  82. Emirdağ-Öztürk, S.; Babahan, İ.; Özmen, A. Synthesis, Characterization and In Vitro Anti-Neoplastic Activity of Gypsogenin Derivatives. Bioorganic Chem. 2014, 53, 15–23. [Google Scholar] [CrossRef]
  83. Sikriwal, D.; Ghosh, P.; Batra, J.K. Ribosome Inactivating Protein Saporin Induces Apoptosis through Mitochondrial Cascade, Independent of Translation Inhibition. Int. J. Biochem. Cell Biol. 2008, 40, 2880–2888. [Google Scholar] [CrossRef]
  84. Kucukkurt, I.; Ince, S.; Enginar, H.; Eryavuz, A.; Fidan, A.F.; Kargioglu, M. Protective effects of Agrostemma githago L. and Saponaria officinalis L. extracts against ionizing radiation-induced oxidative damage in rats. Rev. Med. Vet. 2011, 162, 289–296. [Google Scholar]
  85. Chandra, S.; Rawat, D.S. Medicinal Plants of the Family Caryophyllaceae: A Review of Ethno-Medicinal Uses and Pharmacological Properties. Integr. Med. Res. 2015, 4, 123–131. [Google Scholar] [CrossRef] [Green Version]
  86. Oladeji, O.S.; Oyebamiji, A.K. Stellaria media (L.) Vill.—A Plant with Immense Therapeutic Potentials: Phytochemistry and Pharmacology. Heliyon 2020, 6, e04150. [Google Scholar] [CrossRef]
  87. Li, Y.; Wang, J.; Li, L.; Song, W.; Li, M.; Hua, X.; Wang, Y.; Yuan, J.; Xue, Z. Natural Products of Pentacyclic Triterpenoids: From Discovery to Heterologous Biosynthesis. Nat. Prod. Rep. 2022. [Google Scholar] [CrossRef]
  88. Mikołajczyk-Bator, K.; Błaszczyk, A.; Czyżniejewski, M.; Kachlicki, P. Characterisation and Identification of Triterpene Saponins in the Roots of Red Beets (Beta vulgaris L.) Using Two HPLC-MS Systems. Food Chem. 2016, 192, 979–990. [Google Scholar] [CrossRef] [PubMed]
  89. Burnouf-Radosevich, M.; Delfel, N.E.; England, R. Gas Chromatography-Mass Spectrometry of Oleanane- and Ursane-Type Triterpenes—Application to Chenopodium Quinoa Triterpenes. Phytochemistry 1985, 24, 2063–2066. [Google Scholar] [CrossRef]
  90. Anderson, O.; Beckett, J.; Briggs, C.C.; Natrass, L.A.; Cranston, C.F.; Wilkinson, E.J.; Owen, J.H.; Williams, R.M.; Loukaidis, A.; Bouillon, M.E.; et al. An Investigation of the Antileishmanial Properties of Semi-Synthetic Saponins. RSC Med. Chem. 2020, 11, 833–842. [Google Scholar] [CrossRef] [PubMed]
  91. Emirdağ-Öztürk, S.; Karayıldırım, T.; Çapcı-Karagöz, A.; Alankuş-Çalışkan, Ö.; Özmen, A.; Poyrazoğlu-Çoban, E. Synthesis, Antimicrobial and Cytotoxic Activities, and Structure-Activity Relationships of Gypsogenin Derivatives against Human Cancer Cells. Eur. J. Med. Chem. 2014, 82, 565–573. [Google Scholar] [CrossRef]
  92. Sun, K.-P.; Zhao, T.-T.; Liu, L.; Mu, X.-D.; Sun, J.-Y. Anticancer Structure-Activity Relationships and Potential Target Exploration of the Natural Product Gypsogenin. ChemistrySelect 2023, 8, e202300072. [Google Scholar] [CrossRef]
  93. Wu, G.; Chu, H.; Wang, J.; Mu, Y.; Sun, J. Synthesis of Gypsogenin and Gypsogenic Acid Derivatives with Antitumor Activity by Damaging Cell Membranes. New J. Chem. 2019, 43, 18898–18914. [Google Scholar] [CrossRef]
  94. Ciftci, H.I.; Radwan, M.O.; Sever, B.; Hamdy, A.K.; Emirdağ, S.; Ulusoy, N.G.; Sozer, E.; Can, M.; Yayli, N.; Araki, N.; et al. EGFR-Targeted Pentacyclic Triterpene Analogues for Glioma Therapy. Int. J. Mol. Sci. 2021, 22, 10945. [Google Scholar] [CrossRef]
  95. Ulusoy, N.G.; Emirdağ, S.; Sözer, E.; Radwan, M.O.; Çiftçi, H.; Aksel, M.; Bölükbaşı, S.Ş.; Özmen, A.; Yaylı, N.; Karayıldırım, T.; et al. Design, Semi-Synthesis and Examination of New Gypsogenin Derivatives against Leukemia via Abl Tyrosine Kinase Inhibition and Apoptosis Induction. Int. J. Biol. Macromol. 2022, 222, 1487–1499. [Google Scholar] [CrossRef] [PubMed]
  96. Ciftci, H.I.; Radwan, M.O.; Ozturk, S.E.; Ulusoy, N.G.; Sozer, E.; Ellakwa, D.E.; Ocak, Z.; Can, M.; Ali, T.F.S.; Abd-Alla, H.I.; et al. Design, Synthesis and Biological Evaluation of Pentacyclic Triterpene Derivatives: Optimization of Anti-ABL Kinase Activity. Molecules 2019, 24, 3535. [Google Scholar] [CrossRef] [Green Version]
  97. Lee, I.; Yoo, J.K.; Na, M.; Min, B.S.; Lee, J.; Yun, B.S.; Jin, W.; Kim, H.; Youn, U.; Chen, Q.C.; et al. Cytotoxicity of Triterpenes Isolated from Aceriphyllum Rossii. Chem. Pharm. Bull. 2007, 55, 1376–1378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Krasteva, I.; Yotova, M.; Yosifov, D.; Benbassat, N.; Jenett-Siems, K.; Konstantinov, S. Cytotoxicity of Gypsogenic Acid Isolated from Gypsophila trichotoma. Pharmacogn. Mag. 2014, 10 (Suppl. S2), S430–S433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Ciftci, H.I.; Ozturk, S.E.; Ali, T.F.S.; Radwan, M.O.; Tateishi, H.; Koga, R.; Ocak, Z.; Can, M.; Otsuka, M.; Fujita, M. The First Pentacyclic Triterpenoid Gypsogenin Derivative Exhibiting Anti-ABL1 Kinase and Anti-Chronic Myelogenous Leukemia Activities. Biol. Pharm. Bull. 2018, 41, 570–574. [Google Scholar] [CrossRef] [Green Version]
  100. Almansour, N.M. Triple-Negative Breast Cancer: A Brief Review About Epidemiology, Risk Factors, Signaling Pathways, Treatment and Role of Artificial Intelligence. Front. Mol. Biosci. 2022, 9, 836417. [Google Scholar]
  101. Tian, G.; Zhou, L.; Zhong, Y.; Xu, W.; Bai, H.; Liu, L.; Cui, S. Experimental Studies of the Therapeutic Effect of Gypsophila oldhamiana Gypsogenin on Lewis Lung Cancer in Mice. Chin. J. Clin. Oncol. 2008, 5, 206–210. [Google Scholar] [CrossRef]
  102. Liu, Y.; Li, X.; Jiang, S.; Ge, Q. Inhibitory effect of Gypsophila oldhamiana gypsogenin on NCI-N87 gastric cancer cell line. Eur. J. Inflamm. 2018, 16. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Number of citations and scientific publications containing research linking triterpenes with anti-cancer activity over the period 2000–2023. Data were obtained from the Web of Science database by searching for the keywords triterpene cancer.
Figure 1. Number of citations and scientific publications containing research linking triterpenes with anti-cancer activity over the period 2000–2023. Data were obtained from the Web of Science database by searching for the keywords triterpene cancer.
Molecules 28 05677 g001
Figure 2. Structure of gypsogenin, gypsogenic acid and 3-acetyl gypsogenin (1) highlighting the four functional groups.
Figure 2. Structure of gypsogenin, gypsogenic acid and 3-acetyl gypsogenin (1) highlighting the four functional groups.
Molecules 28 05677 g002
Figure 3. Structure of gypsogenin and gypsogenic acid bioactive derivatives through reaction with 3-OH, C-23-CHO or -COOH, and C28-COOH.
Figure 3. Structure of gypsogenin and gypsogenic acid bioactive derivatives through reaction with 3-OH, C-23-CHO or -COOH, and C28-COOH.
Molecules 28 05677 g003
Figure 4. Gypsogenin derivatives with modified ring C.
Figure 4. Gypsogenin derivatives with modified ring C.
Molecules 28 05677 g004
Figure 5. Summary of gypsogenin derivatives SAR pertaining to cytotoxicity against K562 and HL-60 cells.
Figure 5. Summary of gypsogenin derivatives SAR pertaining to cytotoxicity against K562 and HL-60 cells.
Molecules 28 05677 g005
Figure 6. Summary of gypsogenin derivatives SAR pertaining to cytotoxicity against breast cancer cells.
Figure 6. Summary of gypsogenin derivatives SAR pertaining to cytotoxicity against breast cancer cells.
Molecules 28 05677 g006
Figure 7. Summary of gypsogenin derivatives SAR pertaining to cytotoxicity against lung cancer cells.
Figure 7. Summary of gypsogenin derivatives SAR pertaining to cytotoxicity against lung cancer cells.
Molecules 28 05677 g007
Table 1. Gypsogenin derivatives with different cytotoxic activities.
Table 1. Gypsogenin derivatives with different cytotoxic activities.
CompoundCell Line and IC50 Value (µM)
HT-29 [92]Saos-2 [82]HeLa [91]
Gypsogenin10.47.822.4
111.18.235.0
210.87.98.7
36.78.9>100
LOVO [93]
165.8
187.2
190.8
24>30
2517.8
LOVO [74]HePG2 [73]SKOV3 [73]
52.910.09.7
213.512.513.1
HepG2 [94]TE-1 [93]MC3-8 [93]
204.04.72.9
223.65.44.8
232.24.22.6
HeLa [80,97]PANC-1 [80]
Gypsogenin9.413.5
13.35.0
835.2-
95.6-
109.58.7
1110.27.9
U251T98GU87
17 [95]5.88.117.0
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

Radwan, M.O.; Abd-Alla, H.I.; Alsaggaf, A.T.; El-Mezayen, H.; Abourehab, M.A.S.; El-Beeh, M.E.; Tateishi, H.; Otsuka, M.; Fujita, M. Gypsogenin Battling for a Front Position in the Pentacyclic Triterpenes Game of Thrones on Anti-Cancer Therapy: A Critical Review—Dedicated to the Memory of Professor Hanaa M. Rady. Molecules 2023, 28, 5677. https://doi.org/10.3390/molecules28155677

AMA Style

Radwan MO, Abd-Alla HI, Alsaggaf AT, El-Mezayen H, Abourehab MAS, El-Beeh ME, Tateishi H, Otsuka M, Fujita M. Gypsogenin Battling for a Front Position in the Pentacyclic Triterpenes Game of Thrones on Anti-Cancer Therapy: A Critical Review—Dedicated to the Memory of Professor Hanaa M. Rady. Molecules. 2023; 28(15):5677. https://doi.org/10.3390/molecules28155677

Chicago/Turabian Style

Radwan, Mohamed O., Howaida I. Abd-Alla, Azhaar T. Alsaggaf, Hatem El-Mezayen, Mohammed A. S. Abourehab, Mohamed E. El-Beeh, Hiroshi Tateishi, Masami Otsuka, and Mikako Fujita. 2023. "Gypsogenin Battling for a Front Position in the Pentacyclic Triterpenes Game of Thrones on Anti-Cancer Therapy: A Critical Review—Dedicated to the Memory of Professor Hanaa M. Rady" Molecules 28, no. 15: 5677. https://doi.org/10.3390/molecules28155677

Article Metrics

Back to TopTop