Next Article in Journal
The Fabrication of High-Hardness and Transparent PMMA-Based Composites by an Interface Engineering Strategy
Next Article in Special Issue
A Triazaspirane Derivative Inhibits Migration and Invasion in PC3 Prostate Cancer Cells
Previous Article in Journal
High Yield Synthesis of Curcumin and Symmetric Curcuminoids: A “Click” and “Unclick” Chemistry Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chemistry and Biological Activities of Naturally Occurring and Structurally Modified Podophyllotoxins

1
Guangxi Key Laboratory of Medicinal Resources Protection and Genetic Improvement, Guangxi Botanical Garden of Medicinal Plants, Nanning 530023, China
2
Guangdong Province Key Laboratory of Pharmacodynamic Constituents of TCM and New Drugs Research, International Cooperative Laboratory of Traditional Chinese Medicine Modernization and Innovative Drug Development of Ministry of Education (MOE) of China, College of Pharmacy, Jinan University, Guangzhou 510632, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(1), 302; https://doi.org/10.3390/molecules28010302
Submission received: 26 November 2022 / Revised: 17 December 2022 / Accepted: 25 December 2022 / Published: 30 December 2022
(This article belongs to the Special Issue Synthesis and Application of Anticancer Inhibitors)

Abstract

:
Plants containing podophyllotoxin and its analogues have been used as folk medicines for centuries. The characteristic chemical structures and strong biological activities of this class of compounds attracted attention worldwide. Currently, more than ninety natural podophyllotoxins were isolated, and structure modifications of these molecules were performed to afford a variety of derivatives, which offered optimized anti-tumor activity. This review summarized up to date reports on natural occurring podophyllotoxins and their sources, structural modification and biological activities. Special attention was paid to both structural modification and optimized antitumor activity. It was noteworthy that etoposide, a derivative of podophyllotoxin, could prevent cytokine storm caused by the recent SARS-CoV-2 viral infection.

1. Introduction

Podophyllotoxin and related derivatives (briefly called podophyllotoxins) are widely distributed in plant kingdom, which had long been used in folk medicines for the treatment of snake bites [1,2], cancer, astriction, etc.; [3]. Podophyllotoxin (1) was firstly isolated from Podophyllum peltatum by Podwyssotzki in 1884 as crystals, and its structure was elucidated until 1930. Henceforth, compounds with such skeletons were increasingly discovered, such as epipodophyllotoxin (2), 4′-demethylpicropodophyllotoxin (7), deoxylpodophyllotoxin (14), etc.; [4]. Podophyllotoxins were a group of natural occurring aryltetralin lignans with characteristic four conjugated rings system and a free-rotating tri-substituted benzene ring. This group of secondary metabolites possessed strong bioactivities, such as antivirus, antitumor and anti-inflammatory, etc., [4,5,6,7,8,9,10]. Among these bioactivities, antitumor was the most attractive function [6,11,12]. Some semi-synthesized podophyllotoxins, such as etoposide and teniposide, have been successfully used in clinics [7,13,14,15,16]. However, drug resistance and various adverse drug reactions, including anemia, hair loss, severe gastrointestinal disturbances [17,18], hepatoxicity [19], immunosuppression [20] and neurologic symptoms [21] etc.; limited their clinical usages. So, continued efforts were still imperatively laid on the structural modification and pharmacological evaluation of podophyllotoxin derivatives to look for more potent agents [3]. In addition, extensive structure-activity relationship (SAR) studies have demonstrated that even a small alteration in the structures will cause significant change of their biological activities or even the molecular targets [22,23,24,25,26,27,28].
This review focused on up-to-date studies on the natural podophyllotoxins and their natural sources, structural modification on rings A, B, C, D and E, biological activities including antitumor, antiviral anti-inflammation, miscellaneous effects and toxicity. In addition, total chemical synthesis, biosynthesis and ADME were also included in this review.

2. Natural Occurring Podophyllotoxins

Podophyllotoxins were not only found in typical species such as Podophyllum peltatum L. and Podophyllum emodi Wall. (syn. P. hexandrum Royle), but also could be found in other genera, e.g., Sinopodophyllum, Diphylleia and Dysosma (Berberidaceae), Polygala (Polygalaceae), Anthriscus (Apiaceae), Linun (Linaceae), Hyptis (Verbenaceae), Juniperus, Callitris and Thujopsis (Cupressaceae), Haplophyllum (Rutaceae), Commiphora (Burseraceae) and Hernandia (Hernandiaceae) [29,30]. Hitherto, more than ninety podophyllotoxins were identified. These molecules can be classified into aglycones and their glycosides. Some of these aglycones are seco-podophyllotoxins produced by cleavages of A, C or D ring. The glycoside derivatives usually contained a sugar chain consisted of one to several monosaccarides attached at C-4 (glucose or apiose).

2.1. Podophyllotoxins

The alycones included the isomers of podophyllotoxin (25), and pharmacological study revealed that the 1,2-cis and 2,3-trans configurations were of crucial importance for the biological activities since the C-2 diasteromer of podophyllotoxin (3) was nearly inactive [31]. In contrast, the configuration of C-4 was less important because the semisynthesis of clinical used drugs (etoposide and teniposide) used 1 or 2 as the starting materials without considering the stereochemistry at C-4 [32]. Other analogs included the 4′-demethyl, 4-oxidized, and dehydroxy derivatives of podophyllotoxin and its isomers (619). Among these derivatives, deoxypodophyllotoxin (14) was the most extensively distributed constituent and it was reported to show various activities, such as reducing pigmentation [33], antiasthmatic [34], antitumor [22], etc. Compounds 2024 were three acidic podophyllotoxin derivatives and compound 24, the 4-acetyl substituented product of 4′-demethyl-podophyllotoxin, was reported to show more potent cytotoxic activity than etoposide (97) [35]. Angeloyl podophyllotoxin (26) was firstly isolated as a natural product from Anthriscus sylvestris by an activity-guided isolation method [36], and it was reported to activate the caspase-3 in human promyeloid leukemic HL-60 cells [37]. Compounds 2632 were featured with a methoxy substituent at C-5. 5-methoxypodophyllotoxin (26) and 5-methoxy-4-epipodophyllotoxin (27) isolated from the bark of Libocedrus chealieri showed strong cytotoxicity on KB cells with IC50 values at nanomolar concentrations. The mechanism was related to their effect on tubulin assembly [38]. 5-Methoxypodophyllotoxin -7-O-n-hexanoate (31) has been detected by HPLC-MS in several Linum species, but was first isolated from seeds of Linum flavum as the only identified aryltetralin lignans in this plant. Hernandin (33), the 6-methoxy derivative of deoxypodophyllotoxin (14) was firstly isolated from the seeds of Hernandia ovigera L. and its structure was elucidated by both spectroscopic and X-ray crystallographic mehods [39]. α-Peltatin (34) and β-peltatin (35) are two main components isolated from the dried rhizomes and roots of Podophyllum peltatum L. [40,41]. The 5-methyl ether of β-peltatin (32) was found to exist in genera Bursera, Jeniperus and Thujopsis, which also showed potent cytotoxic activity. Diphyllin (46) was a common cytotoxic agent extensively found in many species of genera Diphylleia, whose structure was featured with a tetradehydrogenated C-ring, a methylenedioxy at C-3′ and C-4′ and two methoxy at C-6 and C-7 [42,43]. Justicidin A-B (4748) isolated in genera Justicia were two ligands with the same structural features as diphyllin (46). Justicidin C (49) was an isomer of Justicidin B (48) with a C-11 carbonyl in contrast that at C-12 in 48. Haploymyrtin (50) [44] was 7-demethoxy derivative of diphyllin from Haplophyllum myrtifolium, whose total synthesis attracted intensive attentions [45,46,47]. Compound 58 with a butenyl ether group at C-7 instead of a hydroxy in 50 was isolated from the same plant [48]. Isodiphyllin (36) with the exchanged substitution of B and E rings in diphyllin (46) was isolated from Dysosma versipellis. Compound 37 from Haplophyllum cappadocicum was a demethoxy derivative of podophyllotoxin. Clilinaphthalide A and B (4142) were two ligands bearing the same structural features as 36 but with a methoxy group at C-4. Compounds 3840 were 5′-demethoxy anologues of podophyllotoxin. Compounds 4345 were ring A opened derivatives. Polygamain (51) was featured with methylenedioxy groups at both B and E rings, and was identified as a cytoxic agent from Haplophyllum ptilostylum. Compound 52 was the 4α isomer of 51, which was obtained from Bursera simaruba. Erlangerin A-D (5356) were four ligands isolated from Commiphora erlangeriana. Compounds 53 and 54 belonged to the polygamatin-type; while 55 and 56 were related to podophyllotoxin (1). Compound 57 was a polygamatin-type ligand from Justicia heterocarpa whose structure was confirmed by X-ray diffraction analysis. The chemical structures of natural podophyllotoxins aglycones are shown in Figure 1.

2.2. Seco-Podophyllotoxins

Some seco-podophyllotoxins existed in plants (Figure 2). Compounds 5966 were C ring cleavaged derivatives. Yatein (59) was a common constituent which has been found in species Juniperus chinensis, Bursera simaruba and Hernandia nymphaefolia, et al. [49,50,51]. Compounds 60 and 61 were the demethyl or methoxy derivatives of yatein (59). These two compounds had been isolated from Juniperus sabina and Hernandia peltata. Nemerosin (62) was isolated from Anthriscus sylvestris through a bioassay-guided isolation method, but was reported to be less active than podophyllotoxin-type compounds [52]. Bursehernin (63) was another common seco-podophyllotoxin existing extensively in genera Hernandia [53,54]. Compounds 6771 were D ring opened derivatives.

2.3. Podophyllotoxin Glycosides

The sugar units usually attached at C-4 consisted of one to several monosaccarides, glucose or apiose (Figure 3). Compounds 7285 were monoglycosides. Among these compounds, 7277 and 79 were podophyllotoxin-type glucosides, 8081 were apioside of diphyllin-type anologues, and 78 and 82 were featured with acetyl substitution attached at the sugar unit. In these structures, the sugar units located at C-4. Compounds 8386 were also monoglycosides, but the sugar unit attached at C-4′ or C-5 position. Compounds 8793 were diglucosides with the sugar side chain linked at C-4. Bispicropodophyllin glucoside (94) was an unique dimeric lignan from Withania coagulans, in which the C ring of both units were opened to form two esters linking the two units [55]. Ciliatoside A (95) and B (96) were lignan glycosides possessing potent anti-inflammatory effect from Justicia ciliate, with a three and four sugar side chain, respectively [56].
The main podophyllotoxins and their natural sources were summarized (Table 1 and Figure 1, Figure 2 and Figure 3).

3. Structural Modification in Podophyllotoxins

Most of the natural occurring podophyllotoxins were limited in applications either by their insufficient resources or prohibitive toxicity. In the mid of nineteenth century, investigations on the synthesis or semisynthesis of podophyllotoxins were undertaken to construct new molecules with optimized antineoplastic activity and less toxicity [128], which led to the generation of two widely used anticancer drugs, etoposide and teniposide [129]. Compared to the parent compounds, etoposide and teniposide showed moderate toxicity, improved therapeutic index (TI) and acceptable efficiency in the treatment of many cancers, especially small cell lung carcinoma and testicular cancer [130]. Nevertheless, limitations such as poor solubility and growing drug resistance still existed during their applications [14,131]. So, podophyllotoxin and its derivatives were still hotspots of modifications for novel anticancer agents. Many previous reviews summarized the synthesis or semisynthesis of podophyllotoxin derivatives including simple esterification, demethylation, oxidation, etc. Recently, many researchers were interested in introducing hereronuclears into podophyllotoxins according to the bioisostere theory, as well as the synthesis of spin-labeled derivatives or conjugates with anticancer drugs, e.g., 5-fluorouracil (5-Fu) [27,132].

3.1. Introducing Heteronuclears into Podophyllotoxins

Bioisosterism was a rational strategy in molecular modifications [133]. Recently, substitution of carbon atoms with heteronuclears was carried out to synthesize podophyllotoxin analogues. Pharmacological studies revealed that some nitrogen-containing derivatives, such as GL-331 (97) and TOP-53 (98), exhibited more potent cytotoxic activities than their parent compounds (Figure 4). Additionally, their abilities to reverse multidrug resistance and inhibit P-glycoprotein induced drug efflux were improved.

3.1.1. Ring A

Ring A was reported to be important for the cytotoxicity [3]. Cleavage or changes of A-ring (Figure 5), such as replacing it with a pyridazine ring (99), will decrease the cytotoxicity, as well as TOPO-II inhibitory activity [134]. But many A-ring modifications were still performed to improve their inhibitory activity on reverse transcriptase (RT) in HIV, and minimize the side-effects. A series of A-ring opened compounds 100103 were synthesized, and were tested to possess potent anti-HIV activities with the average EC50 less than 0.001 μg/mL and the therapeutic index (TI) value more than 120 (against HIV) [135].

3.1.2. Ring B

Modification in ring B was relatively rare. Introducing a hydroxyl group into ring B (104) could remarkably improve the TOPO-II inhibitory activity of epipodophyllotoxin [3]. It was reported that the alkoxy-substituted benzene ring was replaced by a pyrazole moiety (105110). The antiproliferative properties of these heterocyclic compounds were comparable with the currently used anticancer drug etoposide [136].

3.1.3. Ring C

A variety of ring C modified podophyllotoxins have been synthesized, and their diverse biological profiles were attractive. TOPO-II is the major target of podophyllotoxins in cancer therapy, since C-4 position in ring C was identified to be the TOPO-II binding site [3]. It was reported that a bulky at C-4 could enhance their cytotoxic activities [137]. Some researches indicated that replacing the saccharide chain with a non-saccharide moiety can remarkably reverse the drug resistance of etoposide [138].
Interestingly, many modifications were focused on transforming the C-4 of podophyllotoxins into an amino group (Figure 6), and pharmacological evaluations showed that some derivatives exhibited superior activity, particularly against the drug-resistant cell lines [139,140]. Introducing amino group into these molecules also made further modifications possible. A series of saturated aliphatic amide derivatives (111117) were synthesized by linking 4β-amino-4-deoxypodophyllotoxin with succinic acid [141], among which, compounds 113117 showed more potent antitumor activity than etoposide. More inspiringly, they could reverse MDR against K562/AO2 in vitro. Moreover, some N-substituted-5-methoxy derivatives (118123) were synthesized and were tested to show comparative activity against HeLa cancer cells with etoposide [142]. In addition, some 4-β-anilino amides exhibited more potent selectivity against several cancer cell lines. For examples, compound 124 (IC50 1.11 μM against A549; 3.23 μM against MCF-7), 127 (IC50 0.71 μM against A549; 0.92 μM against MCF-7) [140], 128 (ED50 2.4 μM against A549, ED50 4.5 μM against MCF-7) and 129 (ED50 0.7 μM against KB; 3.5 μM against KB-7d) [139]. Compounds 137155 were a series of sulfonamides [138,143], and some of these compounds were 2–10 times more potent than etoposide. Interestingly, compounds 154 and 155 showed selectivity against MDR-MCF7 cell line; while morpholino- and the piperazino-containing sulfonamides derivatives 152 and 153 exhibited selectivity against P388 leukemia and A549 lung carcinoma cell lines. Besides, 4-β-anilino-podophyllotoxins (156) [144], as well as derivatives with expanding conjugated system, i.e., 4β-N-polyaromatic podophyllotoxins (157166), were synthesized [145]. All of them exhibited significant in vitro anticancer activity and the mechanism were investigated to involve the inhibition of DNA topo-II. Compounds 160161 and 165166 were more potent than compounds 162–164 (with C-4′ hydroxyl group in E-ring), indicating that the aromatic group at C-4 and a methoxy at C-4′ might play a vital role in their cytotoxic activity. Several 4β-amino hetereoaromatic ring derivatives (167174) [146] also exhibited promising anticancer activity against colon cancer cell lines, and compound 174 even showed selectivity against CNS malignant cells.
4β-hydroxyl group was the key position of the structural modification on podophyllotoxins. These semisynthetic derivatives showed distinct TOPO-II inhibitory activities. Most of them exceeded their parent compounds, indicating the side chain at C-4 also played a key role in their bioactivity besides the skeleton. Morever, some compounds did not even obey the established structure-active relationship. For example, compounds 142, 144 and 150 without a bulky side chain at C-4 also exhibit potent TOPO-II inhibitory activity; while compound 147, a sulfamide with a long aliphatic side chain showed no TOPO-II binding affinity. Among some polyaromatic substituted 4β-amino podophyllotoxins, 4′-methoxyl derivatives are more cytotoxic than the 4′-hydroxyl compounds. The above structure-activity relationships indicated the existence of some new binding sites for podophyllotoxins on DNA TOPO-II. Furthermore, some compounds are specific for certain cancer cell lines, e.g., colon and prostate, revealing the involvement of some other novel mechanisms.
Structural modification was also performed to introduce other elements into podophyllotoxins, such as Se or metals. Compounds 175 and 176 with 4β-Se showed enhancement of cell death in a time- and dose-dependent manners, and the mechanism involved the translocation of Bax, the activation of the mitochondrial pathway and apoptosis through the release of proapoptotic factors [147]. Forming complex was an alternative method to introduce metal ions into podophyllotoxins. Hydrazide-podophyllic metal complexes could interact with DNA in different ways. The complexes of Ni and Co-HDPP interacted with DNA mainly by insertion; while the interaction of Zn-HDPP with DNA by partial insertion [148].
Instead of substituting the C-4 hydroxyl group with an amino group, nitrogen atom could also be inserted into podophylltoxin skeleton as a part of the ring C. These derivatives could be subdivided into 2-aza-podophyllotoxins and 4-aza-podophyllotoxins. Compound 177, one of the 2-aza-podophyllotoxin was found to exhibit significant activity against several human cancer cell lines [149], but the mechanism was still unclear. 2-aza-podophyllotoxins could inhibit TOPO-II in malignant cells. Different from some natural occurring or semi-synthesized podophyllotoxins, an oxidized E-ring would be an essential motif of 2-aza-podophyllotoxins analogues (178) [150]. 4-aza-4-deoxypodophyllotoxin showed potent cytotoxicity against P388 leukemia cells [151]. Another group of dehydro-podophyllotoxins were synthesized. A series of 4-aza-2,3-dehydro-4-deoxypodophyllotoxins (180188) [152] showed two fold potent cytotoxicity against P-388 leukemia cells than podophyllotoxin. However, a planar 4-aza-C-ring (179) is not favorable with IC50 > 20 μM. The author also proposed an in silico model to predict IC50 of different compounds. Some A-ring removed or replaced 4-aza-podophyllotoxins compounds 193196 exhibited strong anticancer activity. Their structures are showed in Figure 6. But the exact mechanisms are still under investigation [153]. What is more, SAR of these 4-aza-podophyllotoxins was different from that of some natural or semisynthetic derivatives. It was reported that transfused D-ring was an essential motif for binding microtublin or TOPO-II, and a dioxymethene A-ring has a positive impact on its cytoxic effect. But pharmaceutical results of these 4-aza derivatives revealed that some 4-aza 2,3-dehyro-podophyllotoxins also exhibited promising cytotoxicity. Other 4-aza derivatives like compound 197 and 198, with different linker between ring C and E ring were synthesized, which possessed inhibitory activity on tubulin polymerization, as well as promising antitumor activities [154].
Forming lactone or lactam motif in ring C led to the synthesis of compounds 199205, which possessed moderate cytotocixities in several cancer cell lines excepted that the C-lactone derivatives showed potency on colon cancer cell line [155].

3.1.4. Ring D

Although ring D was generally supposed to be an essential part for the activities of podophyllotoxins, a series of ring D opened deoxypodophyllotoxin (206211) showed selective cytotoxicities against the HL-60 cell line (Figure 7) [156]. Oxidization followed by further modification at C-9 led to the synthesis of some carboxylic acid derivatives as esters, amides, nitriles and anhydrides (212225). Their cytotoxicities were at micromolar range, although less potent than the parent compounds [157]. Besides, reaction of podophyllic aldehyde with aliphatic, aromatic, and heteroaromatic amines led to the synthesis of a series of imines (227236), and biological evaluations indicated that they could induce microtubule depolymerization, and cells arrested at the G2/M phase [158]. As a continuation of the above research work, the same group further synthetized several series of nonlactonic podophyllic aldehyde analogues (237268), featured with combinations of aldehyde, imine, amine, ester, and amide functionalities at C-9 and C-9′ of the cyclolignan skeleton. Among these compounds, 249253 with an aldehyde or imine at C-9 and an ester at C-9′ were the most potent with IC50 values in the nanomolar range, and some of them showed several times more potent cytotoxicity against HT-29 and A-549 carcinoma than MB-231 melanoma cells. The mechanisms of these structures were found to be involved with two different mechanisms, i.e., cell death induction by cell cycle arrest and the microtubule-disrupting capacity [159].
Besides those ring D opened derivatives, compound 267 with a 1,5-disubstituted triazole ring instead of the lactone motif was synthetized, which showed moderate cytotoxicity with similar mechanism to podophyllotoxin [160]. Another derivative (268) with a substituted cyclosulfite ring exhibited significantly cytotoxicity [161].

3.1.5. Ring E

Several 4′-ester derivatives of GL-331 (269271) (Figure 8), which were 4β-amino derivative of epipodophyllotoxin under Phase II clinical evaluation [162] were synthetized and showed inhibitory activity on KB and resistant KB-7d tumor cells. The molecular target was confirmed to be DNA topo II. These findings challenged the long-standing premise that a free 4′-hydroxy group was essential for the topo II inhibition [4,163]. Subsequently, the same research group introduced some solubility enhancing moieties to the 4′-hydroxyl position and synthesized eight novel 4′-ester 4β-arylamino analogues (272279) with improved activity profiles and water-solubility compared with etoposide. Based on the above results, the authors proposed a SAR of these analogues: the pendent E ring and the variable 4β-substitution were respectively defined as the enzyme and DNA interacting domains, and the latter was critical to DNA cleavage specificity and drug-resistance [162]. Other E ring modification led to the synthesis of a N-alkyl-4-amino-1,2-dihydroquinoline-lactone (280) whose pendent E ring could not rotate freely, and its bioactivity was still under testing [164].
Bioisosteric replacement of the phenolic ring with nitrogen-containing heterocycles, such as pyrazoles and triazoles could overcome the reduced drug bioavailability caused by oxidation and glucuronidation of phenolic hydroxyl groups [165,166] Based on compound 180, a dihydropyridopyrazole analogue of podophyllotoxin, a series of E ring modified derivatives (281304) were synthetized, as substituted E ring with aliphatic, aromatic or heteroaromatic groups. Among these derivatives, those with bromine at meta-position of the aromatic ring E (285, 291294) showed potent cytoxicities, but the mechanism still needed further investigation [167].

3.2. Spin Labeled Podophyllotoxins

Stable nitroxyl radicals could be used to improve anti-cancer profiles of drugs [168]. Tian’s group initiated their synthesis work of spin-labled podophyllotoxins in early 1980th [169]. Subsequently, a series of nitroxyl spin-labeled ester derivatives were synthesized, and the modification positions varied from 4-hydroxyl group, 4-amino group, 4′-hydroxyl to the carboxyl group in the open lactone ring (Figure 9) [170,171,172]. Introduction of nitroxyl radical moieties into 4β-amino-4′-demethylepipodophyllotoxin (305312) greatly enhanced the antioxidative effect, antitumor and anti-drug resistance activities [173,174]. Besides, a series of 4′-spin-labeled compounds 313320 were designed and synthesized, and pharmacological experiments showed that most of these molecules exhibited more potent cytotoxicities against HL-60, RPMI-8226 and A549 than the parent compounds. In addition, the synthesized derivatives showed either similar or better antioxidative activities than etoposide [175].
It was well known that cancer formation and development was closely linked to inflammation [176,177,178,179]. In addition, after an inflammatory stimulus, reactive oxygen species (ROS) produced, which could cause cell or DNA damage and eventually mediate carcinogenesis [180]. These cytotoxic podophyllotoxins combined with an antioxidative property was able to reduce tissue damage induced by ROS and prevent tumorigenesis.

3.3. Conjugates of Podophyllotoxins

Anticancer drugs were usually joined together for the synergistic treatment of cancer since few tumors are sensitive enough to be cured by single drugs. The anticancer drugs could be connected directly or by means of a linker [181]. The combination of podophyllotoxins and other anticancer drugs led to the synthesis of a series of conjugates.
The connection of podophyllotoxins with other anticancer drugs by various linkers resulted in the construct of many combined agents (Figure 10). Compounds 321329 were conjugates consisted of podophyllotoxin and antimetabolite 5-FU using different spacers. Among them, 4β-N-substituted-phenylalanine 5-Fu pentyl ester-4′-demethylepipodophyllotoxin (321329) was tested to be the most potent cytotoxic activity against HL-60 and A-549 cell, which was stable in plasma [182]. Besides, another series of derivatives (330339) were synthesized via combining demethyepipodophyllotoxin and 5-FU through a peptide bond derived from natural L-amino acids. These compounds displayed more potent anticancer activity in vitro than etoposide, and showed synergistic effects [183].
A series of thiocolchicine podophyllotoxin derivatives (340343) connected by the disulfide bond were constructed based on a combinatorial chemistry method. The biological evaluation demonstrated that divalent compounds were not merely the sum of the single compound’s activities, thus reflecting a different interaction with the biological target [184]. Inspired by the pharmaceutical results mentioned above, the same research group synthesized hybrids of naturally occurring antimitotic compounds. One of these molecules, the hybrid of vinorelbine and podophyllotoxin (344) linked by succinic anhydride showed good cytotoxicity but with a low efficacy for the inhibition of tubulin assembly, suggesting a different biological target [185]. Furthermore, another group of condensed dimeric compounds 345349 of thiocolchicine and/or podophyllotoxin with six different dicarboxylic acids were synthesized. Among them, three compounds showed a significant inhibitory activity on the polymerization of tubulin in vitro and causing obvious disruption to the microtubule network in vivo, indicating the spacer unit played an important role on their biological activity [186].
Another example was the hybrid of etoposide and amsacrine, both of which are inhibitors of TOPO-II. The pharmaceutical results indicated that the linkers were highly important for their biological profiles. Compound 351 was more potent than both etoposide and amsacrine according to its DNA cleavage assay, whereas 350 without an ethylene spacer was less potent. Nevertheless, 350 targeted on tubulin polymerization other than its effect on topoisomerase II suggesting the etoposide-amsacrine hybrids might lead to the discovery of dual inhibitors targeting both topoisomerase II and tubulin [187]. Another example was the combination of podophyllotoxin and indibulin, which was also a potent microtubulin inhibitor. Further modification of this conjugate led to the synthesis of a series of 4α-O- and 4β-N-indol-3-yl-glyoxyl-substituted derivatives (352361) of podophyllotoxin [188]. Among them, 354 was tested to be more potent than etoposide. Moreover, YB-1EPN (356) and L1EPO (362) were investigated to have the activities to overcome P-glycoprotein-mediated multidrug resistance in the KBV200 and K562/A02 cell lines, respectively [189,190].
Besides the hybrids of podophyllotoxins with anticancer agents to improve their pharmacological profiles, similar conjugates were also synthesized to optimized their antiviral activities. Conjugates containing stavudine which was a nucleoside reverse inhibitor and podophyllotoxin analogues (363367) showed increasing bioactivities. Subsequent SAR research showed 7β-amide, cyano group and an opened A-ring or 4′-demethylation are favorable for the anti-HIV activity [191].

4. Biological Activities of Natural Occurring Podophyllotoxins

Podophyllotoxins were a group of highly bioactive compounds. Historically, podophyllotoxin and its analogues were extracted from plants and directly used as a mixture mainly for external applications [3]. Later, scientists found these compounds could be used in viral infections, such as HPV and HIV diseases [75,192,193]. With the development of pharmacological investigations, the neutrophil activation [194], abnormal vascular vessels destroying [195], radioprotection [196,197,198], antioxidation [199,200], skin pigmentation reduction [33], anti-inflammation, anti-hyperplasia [201] and allergic reaction regulation [202] were extensively studied. Besides, these compounds were found to affect sodium and calcium concentrations in neuron [203]. Besides, podophyllotoxins showed insecticidal activities, for example, podophyllotoxin analogs showed antifeedant activity [204,205]. Similar to the anti-tumor activity, the transfused lactone ring was essential [206].

4.1. Antitumor Activity

Extensive pharmacological tests showed that although these compounds shared very similar skeleton, their targets were varied. Podophyllotoxin (1) bound to the β-subset of microtublin at the colchicine site and potently inhibits the microtubule assembly [207], resulting in G2/M arrest [208]. Deoxypodophyllotoxin (14) induces G2/M cell-cycle arrest followed by apoptosis through multiple cellular processes, involving the activation of ATM, upregulation of p53 and Bax, then activation of caspase-3 and -7 [22,59] and the Cdk1/cyclinB1 complex through Cdc25C [209]. 4′-Demethylepipodophyllotoxin and its derivates, such as episode bound to topoisomerase-Iiα [210] in cancer cells, stabilized a cleavable complex between DNA and topoisomerase II, consequently resulted in DNA strand breaks and led to cytotoxic effect [211]. Picropodophyllin (3) specific inhibited the IGF-1R kinase activity [28,212,213,214]. However, some of compounds could also interfere with signal transduction (e.g., podophyllotoxin (1) [59,207]) in cells, and their exact mechanisms were still in debate [215].

4.2. Antiviral Activities

Natural products were one of the most important sources of antiviral agents and lead compounds [216]. Podophyllotoxin (1) solution and cream could be clinically used in HPV infection patients [10,217], the mechanism involved directly binding a hinge domain E2 in the HPV virus and inhibited the E2/E7 interaction [218]. Some structurally modified podophyllotoxins were found to be effective against HIV.
The pandemic coronavirus disease 2019, caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), was mainly transmitted via the inhalation system and characterized by fever, cough and difficulty in breathing. Some natural products were found to exhibit useful effects against the COVID-19 [219]. Especially, etoposide, a derivative of podophyllotoxin, could prevent cytokine storm caused by SARS-CoV-2 viral infection [220,221]. Furthermore, molecular docking based on RMSD and RMSF data supported the use of etoposide as an inhibitor of COVID-19 [222].

4.3. Anti-Inflammation Activities

Deoxypodophyllotoxin (14) could interfere with many inflammation processes and exhibited potent anti-inflammation activity in pharmacological researches. In inflammation initiating phase, deoxypodophyllotoxin (14) could abolished LPS-induced iNOS expression by inhibiting NF-kappa B [223]; Deoxypodophyllotoxin (14) could decrease the mRNA levels of Th2 cytokines [34]; and it could inhibit TNF-alpha-induced ICAM-1 expression through nuclear factor-kappa B (NF-kappa B) in a dose-dependent manner [224]. It could also inhibit inflammatory cell migration and MMP-2/9 activities, and the MMP-9 transcription [225]. Besides, podophyllotoxins (1) showed antioxidative effect. It would help to clean the reactive oxygen species (ROS), and decrease the inflammation induced tissue damage.

4.4. Miscellaneous Activities

Podophyllotoxin was used as medical cream and applied to genital warts and molluscum contagiosum [226]. Podophyllotoxin (1) exposure could affect mouse oocyte maturation by disturbing microtubule dynamics and meiotic spindle formation [3]. Acetylpodophyllotoxin (51) displayed direct antigiardial killing activity on Caco-2 cells [227]. In addition, some podophyllotoxin derivatives exhibited insecticidal activity with final mortality rates of 70%. Especially, a chlorine or bromine atom introduced at the C2′ or C2′ and C6′ positions on the E ring of podophyllotoxin could increase the insecticidal activity [228,229].

4.5. Toxicity and Protection

However, podophyllotoxin was well known for its potent cytotoxic properties because of its poor selectivity against tumor cells and narrow therapeutic window. A young woman presented with podophyllin intoxication following topical application of podophyllin resin to genital condylomata acuminata. The disorder was marked by hallucinatory psychosis, bone marrow depression, and mild hepatic dysfunction [230]. A 22 years old man developed a severe sensorimotor neuropathy following ingestion of podophyllin, which had been prescribed for genital condylomata. The initial toxic symptoms were vomiting and diarrhea, followed by peripheral neuropathy. The neuropathy was still present after 18 months [231]. It was noteworthy that most podophyllotoxin intoxication usually results from the accidental ingestion or topical application of podophyllum resin [232]. In addition, etoposide was reported to show immunosuppression, which deserved attention in chemoimmunotherapy [233].
In order to alleviate the toxicity of podophyllotoxin, scientists used some polyphenols e.g., curcumin [234], quercetin [235] and kaempferol [235] to prevent the toxic effect. The protective mechanisms were due to the antioxidant activity of those polyphenols against the oxidative stress induced by podophyllotoxin and the competitive binding of polyphenols against podophyllotoxin in the same colchicines-binding sites.

5. Total Chemical Synthesis, Biosynthesis and ADME

The total chemical synthesis protocol of podophyllotoxin was introduced in 1996 [236]; however it is time-consuming with low yield. Another efficient and stereoselective strategy for the total synthesis of podophyllotoxin included 12 steps with 29% overall yield [237]. Further investigation showed that this approach can be simplified to an eight step approach with an equal overall yield [238]. The main steps of these syntheses are shown in Figure 11. Later, Ting et al. reported a short total synthesis of podophyllotoxin which could be finished in five steps with 41% overall yield [239]. Xiao et al. reported a nickel-catalyzed approach for the construction of diastereodivergent cores embedded in podophyllum lignans [240]. Besides, an enantioselective total synthesis of (−)-podophyllotoxin was accomplished by organocatalytic Heck cyclization [241]. To date, several elegant strategies have been developed for the synthesis of podophyllotoxin; however, more concise with high yield total synthesis had so far remained an unmet challenge.
In order to produce more podophyllotoxins, many experiments focused on biosynthesis of podophyllotoxins in cultures of plant cell lines [242,243,244,245] and endophytic fungus [246]. The biosynthesis of podophyllotoxin was considered to be an attractive alternative because of the much simpler and greener steps and relatively higher yield. The current biosynthesis pathway of podophyllotoxins in plants involved the process of L-phenylalanine/L-tyrosin→coniferyl alcohol→pinoresinol→(-)-secoisolariciresinol→(-)-matairesinol→(-)-pluviatolide→podophyllotoxin→glycosylation modification of podophyllotoxin [247]. Furthermore, chemoenzymatic synthesis had led to the asymmetric configuration of podophyllotoxin. For example, milligram-level synthesis of (−)-deoxypodophyllotoxin has been achieved in tobacco. At the same time, part of the biosynthetic pathway of podophyllotoxin had been expressed in Escherichia coli and Saccharomyces cerevisiae, and different podophyllotoxin intermediates have been obtained. However, limitation still existed. For example, enzymes were characterized by their high selectivity, and thus, the substrates were limited, and not all desired podophyllotoxin-type products can be produced using this method.
In addition, microbial transformation of natural products is an important approach to synthesize derivatives with improved pharmacological properties. Many podophyllotoxin derivative with higher activity and water-solubility were produced via biotransformation by microorganisms, such as Penicillium purpurogenum [248], Pseudomonas aeruginosa [249], Cunninghamella echinulata [250] and Bacillus fusiformis [251]. Microbial transformations can not only obtain new derivatives, but also provide a natural enzyme library with various catalytic types, which has gradually become a choice for biosynthesis because of the high stereoselectivity and regioselectivity, mild reaction conditions and simple operation steps.
The ADME processes of podophyllotoxins in animals were not clearly evaluated, especially for some new derivates. Experiments using enzyme to predict the metabolic pathway were performed. The results showed that CYP3A4, the main human metabolizing enzyme, had the ability to transform deoxypodophyllotoxin into epipodophyllotoxin [252,253]; while CYP1A2 and CYP2C9 could not accomplish this biotransformation. Furthermorde, etoposie and related semi-synthesized podophyllotoxins could be degraded (3-O-demethylation) [254].

6. Conclusions and Remarks

As described above, podophyllotoxins are widely distributed in nature. Slight structurally modified podophyllotoxins showed different bioactivities from the parent compounds. Their cytotoxicity, safety, pharmacological activity against MDR cell line or selectivity against certain cancer cell lines varied with the structural changes. Limitations on the podophyllotoxin studies existed in several aspects.
Firstly, the mechanism of some compounds was still unknown. So, their targets and pharmacophore of these molecules were still uncertain. Meanwhile, this information was vital to understand the mechanism and further development of these compounds. Secondly, previously established SAR was facing challenge. Some compounds without defined essential motif still showed remarkable cytotoxicity. This could be the result of modification changing the 3-D structures of these molecules. So, previously established SAR seemed to be no longer as comprehensive as before, especially when it was used to predict the ester or ether of these compounds. Thirdly, for many compounds, cytotoxity as the unique activity of this kind of compounds was only tested in limited cell lines, so the cytotoxity cannot be predicted on other cancer cells. Fourthly, for most compounds, the ADME parameters and in vivo activity were not studied. However, cancer was a very complex malignant disease. Different kind of cells were anchored in cancer tissues. So, inhibiting the quick dividing cancer cells did not mean to the cure of this disease [255,256,257,258,259,260]. Finally, the microenvironment [261,262,263] also played very important role in the generation, development and metastasis of a tumor; however, microenvironment was less studied in podophyllotoxins. In order to find a potent drug with high therapeutics, more experiments should be conducted.
In summary, podophyllotoxins were very promising compounds because of their unique chemical structures and diverse bioactivities. Structure modifications make them more suitable for clinical use. A slight change in these chemical structures lead to a remarkable change in their activity. So, the establishment of a comprehensive SARs, which was more suitable for the natural and modified podophyllotoxins, was necessary.

Author Contributions

Conceptualization and project administration, R.J.; validation, L.R. and Z.S.; investigation, L.J. and F.C.; data curation, L.J. and L.R.; writing—original draft preparation, L.J. and L.R.; writing—review and editing, R.J.; funding acquisition, R.J., L.R. and Z.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Guangxi Natural Science Foundation Program (grant number 2021GXNSFBA220068), Scientific Research Funding Project of Guangxi Botanical Garden of Medicinal Plants (grant number GYYJ 202001), National Natural Science Foundation of China (grant number 81872760 and 82204975), Natural Science Foundation of Guangdong province (grant number 2021A1515011251), Guangdong scientific scheme (grant number 2021A0505030032), and Independent Research Project of Guangxi Medicinal Plant Conservation Talent Center (grant number GXYYXGD202203).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the public research platform in the College of Pharmacy, Jinan University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shah, Z.; Gohar, U.F.; Jamshed, I.; Mushtaq, A.; Mukhtar, H.; Zia-Ui-Haq, M.; Toma, S.I.; Manea, R.; Moga, M.; Popovici, B. Podophyllotoxin, history, recent advances and future prospects. Biomolecules 2021, 11, 603. [Google Scholar] [CrossRef]
  2. Chou, S.L.; Chou, M.Y.; Kao, W.F.; Yen, D.H.T.; Huang, C.I.; Lee, C.H. Cessation of nail growth following Bajiaolian intoxication. Clin. Toxicol. 2008, 46, 159–163. [Google Scholar] [CrossRef]
  3. Hu, L.L.; Zhou, X.; Zhang, H.L.; Wu, L.L.; Tang, L.S.; Chen, L.L.; Duan, J.L. Exposure to podophyllotoxin inhibits oocyte meiosis by disturbing meiotic spindle formation. Sci. Rep. 2018, 8, 10145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Liu, Y.Q.; Yang, L.; Tian, X. Podophyllotoxin, current perspectives. Curr. Bioact. Compd. 2007, 3, 37–66. [Google Scholar]
  5. Cheng, Y.C.; Ying, C.X.; Leung, C.H.; Li, Y. New targets and inhibitors of HBV replication to combat drug resistance. J. Clin. Virol. 2005, 34, S147–S150. [Google Scholar] [CrossRef]
  6. Talib, W.H.; Daoud, S.; Mahmod, A.I.; Hamed, R.A.; Awajan, D.; Abuarab, S.F.; Odeh, L.H.; Khater, S.; Al Kury, L.T. Plants as a source of anticancer agents, from bench to bedside. Molecules 2022, 27, 4818. [Google Scholar] [CrossRef] [PubMed]
  7. Yu, X.; Che, Z.; Xu, H. Recent advances in the chemistry and biology of podophyllotoxins. Chem. Eur. J. 2017, 23, 4467–4526. [Google Scholar] [CrossRef]
  8. Murray, M.L.; Meadows, J.; Doré, C.J.; Copas, A.J.; Haddow, L.J.; Lacey, C.; Jit, M.; Soldan, K.; Bennett, K.; Tetlow, M.; et al. Human papillomavirus infection, protocol for a randomised controlled trial of imiquimod cream (5%) versus podophyllotoxin cream (0.15%), in combination with quadrivalent human papillomavirus or control vaccination in the treatment and prevention of recurrence of anogenital warts (HIPvac trial). BMC Med. Res. Methodol. 2018, 18, 125. [Google Scholar]
  9. Sosa, S.; Balick, M.J.; Arvigo, R.; Esposito, R.G.; Pizza, C.; Altinier, G.; Tubaro, A. Screening of the topical anti-inflammatory activity of some Central American plants. J. Ethnopharmacol. 2002, 81, 211–215. [Google Scholar] [CrossRef]
  10. Fan, H.Y.; Zhu, Z.L.; Xian, H.C.; Wang, H.F.; Chen, B.J.; Tang, Y.J.; Tang, Y.L.; Liang, X.H. Insight into the molecular mechanism of podophyllotoxin derivatives as anticancer drugs. Front. Cell Dev. Biol. 2021, 9, 709075. [Google Scholar] [CrossRef]
  11. Gordaliza, M. Natural products as leads to anticancer drugs. Clin. Transl. Oncol. 2007, 9, 767–776. [Google Scholar] [CrossRef] [PubMed]
  12. Goldsmith, M.A.; Carter, S.K. 4′-Demethyl-epipodophyllotoxin-[beta]-d-thenylidene glucoside (VM-26)-A brief review. Eur. J. Cancer 1973, 9, 477–482. [Google Scholar] [CrossRef] [PubMed]
  13. Yusenko, M.; Jakobs, A.; Klempnauer, K.H. A novel cell-based screening assay for small-molecule MYB inhibitors identifies podophyllotoxins teniposide and etoposide as inhibitors of MYB activity. Sci. Rep. 2018, 8, 13159. [Google Scholar] [CrossRef] [Green Version]
  14. Tachibana, Y.; Zhu, X.K.; Krishnan, P.; Lee, K.H.; Bastow, K.F. Characterization of human lung cancer cells resistant to 4 ‘-O-demethyl-4 beta-(2″-nitro-4″-fluoroanilino)-4-desoxypodophyllotoxin, a unique compound in the epipodophyllotoxin antitumor class. Anti-Cancer Drugs 2000, 11, 19–28. [Google Scholar] [CrossRef] [PubMed]
  15. Hong, W.G.; Cho, J.H.; Hwang, S.G.; Lee, E.; Lee, J.; Kim, J.I.; Um, H.D.; Park, J.K. Chemosensitizing effect of podophyllotoxin acetate on topoisomerase inhibitors leads to synergistic enhancement of lung cancer cell apoptosis. Int. J. Oncol. 2016, 48, 2265–2276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Infante Lara, L.; Fenner, S.; Ratcliffe, S.; Isidro-Llobet, A.; Hann, M.; Bax, B.; Osheroff, N. Coupling the core of the anticancer drug etoposide to an oligonucleotide induces topoisomerase II-mediated cleavage at specific DNA sequences. Nucleic Acids Res. 2018, 46, 2218–2233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Srivastava, V.; Negi, A.S.; Kumar, J.K.; Gupta, M.M.; Khanuja, S.P.S. Plant-based anticancer molecules, A chemical and biological profile of some important leads. Bioorg. Med. Chem. 2005, 13, 5892–5908. [Google Scholar] [CrossRef]
  18. Lee, N.J.; Jeong, I.C.; Cho, M.Y.; Jeon, C.W.; Yun, B.C.; Kim, Y.O.; Kim, S.H.; Chung, I. Synthesis and in vitro antitumor activity of phthalimide polymers containing podophyllotoxin. Eur. Polym. J. 2006, 42, 3352–3359. [Google Scholar] [CrossRef]
  19. Sun, D.; Gao, X.; Wang, Q.; Krausz, K.W.; Fang, Z.; Zhang, Y.; Xie, C.; Gonzalez, F.J. Metabolic map of the antiviral drug podophyllotoxin provides insights into hepatotoxicity. Xenobiotica 2021, 51, 1047–1059. [Google Scholar] [CrossRef]
  20. Gordaliza, M.; Castro, M.A.; del Corral, J.M.; López-Vázquez, M.; Feliciano, A.S.; Faircloth, G.T. In vivo immunosuppressive activity of some cyclolignans. Bioorg. Med. Chem. Lett. 1997, 7, 2781–2786. [Google Scholar] [CrossRef]
  21. Chou, S.L.; Chou, M.Y.; Kao, W.F.; Yen, D.H.T.; Yen, L.Y.; Huang, C.I.; Lee, C.H. Bajiaolian poisoning–a poisoning with high misdiagnostic rate. Am. J. Emerg. Med. 2010, 28, 85–89. [Google Scholar] [CrossRef] [PubMed]
  22. Shin, S.Y.; Yong, Y.; Kim, C.G.; Lee, Y.H.; Lim, Y. Deoxypodophyllotoxin induces G2/M cell cycle arrest and apoptosis in HeLa cells. Cancer Lett. 2009, 287, 231–239. [Google Scholar] [CrossRef] [PubMed]
  23. Yong, Y.; Shin, S.Y.; Lee, Y.H.; Lim, Y. Antitumor activity of deoxypodophyllotoxin isolated from Anthriscus sylvestris, Induction of G2/M cell cycle arrest and caspase-dependent apoptosis. Bioorg. Med. Chem. Lett. 2009, 19, 4367–4371. [Google Scholar] [CrossRef] [PubMed]
  24. Botta, B.; Delle Monache, G.; Misiti, D.; Vitali, A.; Zappia, G. Aryltetralin lignans, chemistry, pharmacology and biotransformations. Curr. Med. Chem. 2001, 8, 1363–1381. [Google Scholar] [CrossRef] [PubMed]
  25. Lizama, C.; Ludwig, A.; Moreno, R.D. Etoposide induces apoptosis and upregulation of TACE/ADAM17 and ADAM10 in an in vitro male germ cell line model. Biochim. Biophys. Acta (BBA) 2011, 1813, 120–128. [Google Scholar] [CrossRef] [Green Version]
  26. Liu, Y.Q.; Yang, L.; Tian, X. Design, synthesis, and biological evaluation of novel pyridine acid esters of podophyllotoxin and esters of 4′-demethylepipodophyllotoxin. Med. Chem. Res. 2008, 16, 319–330. [Google Scholar] [CrossRef]
  27. Girnita, A.; Girnita, L.; Del Prete, F.; Bartolazzi, A.; Larsson, O.; Axelson, M. Cyclolignans as inhibitors of the insulin-like growth factor-1 receptor and malignant cell growth. Cancer Res. 2004, 64, 236–242. [Google Scholar] [CrossRef] [Green Version]
  28. Ohshima-Hosoyama, S.; Hosoyama, T.; Nelon, L.D.; Keller, C. IGF-1 receptor inhibition by picropodophyllin in medulloblastoma. Biochem. Biophys. Res. Commun. 2010, 399, 727–732. [Google Scholar] [CrossRef]
  29. Xu, H.; Lv, M.; Tian, X. A Review on hemisynthesis, biosynthesis, biological activities, mode of action, and structure-activity relationship of podophyllotoxins: 2003–2007. Curr. Med. Chem. 2009, 16, 327–349. [Google Scholar] [CrossRef]
  30. Gordaliza, M.; Garcia, P.A.; del Corral, J.M.M.; Castro, M.A.; Gomez-Zurita, M.A. Podophyllotoxin, distribution, sources, applications and new cytotoxic derivatives. Toxicon 2004, 44, 441–459. [Google Scholar] [CrossRef]
  31. Canel, C.; Moraes, R.M.; Dayan, F.E.; Ferreira, D. Podophyllotoxin. Phytochemistry 2000, 54, 115–120. [Google Scholar] [CrossRef] [PubMed]
  32. Stahelin, H.F.; von Wartburg, A. The chemical and biological route from podophyllotoxin glucoside to etoposide, ninth Cain memorial Award lecture. Cancer Res. 1991, 51, 5–15. [Google Scholar] [PubMed]
  33. Olaru, O.T.; Niţulescu, G.M.; Orțan, A.; Dinu-Pîrvu, C.E. Ethnomedicinal, phytochemical and pharmacological profile of Anthriscus sylvestris as an alternative source for anticancer lignans. Molecules 2015, 20, 15003–15022. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Lin, C.X.; Lee, E.; Jin, M.H.; Yook, J.; Quan, Z.; Ha, K.; Moon, T.C.; Kim, M.J.; Kim, K.J.; Lee, S.H.; et al. Deoxypodophyllotoxin (DPT) inhibits eosinophil recruitment into the airway and Th2 cytokine expression in an OVA-induced lung inflammation. Planta Med. 2006, 72, 786–791. [Google Scholar] [CrossRef] [PubMed]
  35. Sun, Y.J.; Li, Z.L.; Chen, H.; Liu, X.Q.; Zhou, W.; Hua, H.M. Three new cytotoxic aryltetralin lignans from Sinopodophyllum emodi. Bioorg. Med. Chem. Lett. 2011, 21, 3794–3797. [Google Scholar] [CrossRef]
  36. Lim, Y.H.; Leem, M.J.; Shin, D.H.; Chang, H.B.; Hong, S.W.; Moon, E.Y.; Lee, D.K.; Yoon, S.J.; Woo, W.S. Cytotoxic constituents from the roots of Anthriscus sylvestris. Arch. Pharm. Res. 1999, 22, 208–212. [Google Scholar] [CrossRef]
  37. Jeong, G.S.; Kwon, O.K.; Park, B.Y.; Oh, S.R.; Ahn, K.S.; Chang, M.J.; Oh, W.K.; Kim, J.C.; Min, B.S.; Kim, Y.C.; et al. Lignans and coumarins from the roots of Anthriscus sylvestris and their increase of caspase-3 activity in HL-60 cells. Biol. Pharm. Bull. 2007, 30, 1340–1343. [Google Scholar] [CrossRef] [Green Version]
  38. Zhang, Y.J.; Litaudon, M.; Bousserouel, H.; Martin, M.T.; Thoison, O.; Leonce, S.; Dumontet, V.; Sevenet, T.; Gueritte, F. Sesquiterpenoids and cytotoxic lignans from the bark of Libocedrus chevalieri. J. Nat. Prod. 2007, 70, 1368–1370. [Google Scholar] [CrossRef]
  39. Yamaguchi, H.; Arimoto, M.; Tanoguchi, M.; Ishida, T.; Inoue, M. Studies on the constituents of the seeds of Hernandia ovigera L. III. Structures of two new lignans. Chem. Pharm. Bull. 1982, 30, 3212–3218. [Google Scholar] [CrossRef] [Green Version]
  40. Hartwell, J.L.; Detty, W.E. Beta-Peltatin, a new component of podophyllin. J. Am. Chem. Soc. 1948, 70, 2833. [Google Scholar] [CrossRef]
  41. Hartwell, J.L. Alpha-Peltatin, a new compound isolated from podophyllum peltatum. J. Am. Chem. Soc. 1947, 69, 2918. [Google Scholar] [CrossRef] [PubMed]
  42. Zhao, C.; Zhang, N.; He, W.; Li, R.; Shi, D.; Pang, L.; Dong, N.; Xu, H.; Ji, H. Simultaneous determination of three major lignans in rat plasma by LC-MS/MS and its application to a pharmacokinetic study after oral administration of Diphylleia sinensis extract. Biomed. Chromatogr. 2014, 28, 463–467. [Google Scholar] [CrossRef] [PubMed]
  43. Broomhead, A.J.; Dewick, P.M. Tumor-inhibitory aryltetralin lignans in Podophyllum versipelle, Diphylleia cymosa and Diphylleia grayi. Phytochemistry 1990, 29, 3831–3837. [Google Scholar] [CrossRef]
  44. Evcim, U.; Gozler, B.; Freyer, A.J.; Shamma, M. Haplomyrtin and (-)-haplomyrfolin, two lignans from Haplophyllum myrtifolium. Phytochemistry 1986, 25, 1949–1951. [Google Scholar] [CrossRef]
  45. Schaaf, G.M.; Feld, W.A. Optimizing the synthesis of haplomyrtin. In Abstracts of Papers of the American Chemical Society; Amer Chemical Society: Washington, DC, USA, 1999; Volume 217. [Google Scholar]
  46. Barrow, W.S. The Synthesis of Haplomyrtin Utilizing the Triisopropylsilyl Protecting Group. Master’s Thesis, Wright State University, Dayton, OH, USA, 2013. [Google Scholar]
  47. Hunter, N.E. Towards the Total Synthesis of Haplomyrtin. Master’s Thesis, Wright State University, Dayton, OH, USA, 2010. [Google Scholar]
  48. Sağlam, H.; Gözler, T.; Gözler, B. A new prenylated arylnaphthalene lignan from Haplophyllum myrtifolium. Fitoterapia 2003, 74, 564–569. [Google Scholar] [CrossRef] [PubMed]
  49. Tanoguchi, M.; Arimoto, M.; Saika, H.; Yamaguchi, H. Studies on the constituents of the seeds of Hernandia ovigera L. VI. Isolation and structural determination of three lignans. Chem. Pharm. Bull. 1987, 35, 4162–4168. [Google Scholar] [CrossRef] [Green Version]
  50. Donoso-Fierro, C.; Tiezzi, A.; Ovidi, E.; Ceccarelli, D.; Triggiani, D.; Mastrogiovanni, F.; Taddei, A.R.; Pérez, C.; Becerra, J.; Silva, M.; et al. Antiproliferative activity of yatein isolated from Austrocedrus chilensis against murine myeloma cells, cytological studies and chemical investigations. Pharm. Biol. 2015, 53, 378–385. [Google Scholar] [CrossRef]
  51. Miyata, M.; Itoh, K.; Tachibana, S. Extractives of Juniperus chinensis L. I, isolation of podophyllotoxin and yatein from the leaves of J. chinensis. J. Wood Sci. 1998, 44, 397–400. [Google Scholar] [CrossRef]
  52. Wickramaratne, D.B.; Mar, W.; Chai, H.; Castillo, J.J.; Farnsworth, N.R.; Soejarto, D.D.; Cordell, G.A.; Pezzuto, J.M.; Kinghorn, A.D. Cytotoxic constituents of Bursera permollis. Planta Med. 1995, 61, 80–81. [Google Scholar] [CrossRef]
  53. Ito, C.; Matsui, T.; Wu, T.S.; Furukawa, H. Isolation of 6,7-demethylenedesoxy-podophyllotoxin from Hernandia ovigera. Chem. Pharm. Bull. 1992, 40, 1318–1321. [Google Scholar] [CrossRef] [Green Version]
  54. Pettit, G.R.; Meng, Y.; Gearing, R.P.; Herald, D.L.; Pettit, R.K.; Doubek, D.L.; Chapuis, J.C.; Tackett, L.P. Antineoplastic Agents. 522. Hernandia peltata (Malaysia) and Hernandia nymphaeifolia (Republic of Maldives). J. Nat. Prod. 2004, 67, 214–220. [Google Scholar] [CrossRef] [PubMed]
  55. Nur-e-Alam, M.; Yousaf, M.; Qureshi, S.; Baig, I.; Nasim, S.; Attaur, R.; Choudhary, M.I. A novel dimeric podophyllotoxin-type lignan and a new withanolide from Withania coagulans. Helv. Chim. Acta 2003, 86, 607–614. [Google Scholar] [CrossRef]
  56. Day, S.H.; Chiu, N.Y.; Tsao, L.T.; Wang, J.P.; Lin, C.N. New lignan glycosides with potent antiinflammatory effect, isolated from Justicia ciliata. J. Nat. Prod. 2000, 63, 1560–1562. [Google Scholar] [CrossRef] [PubMed]
  57. Lee, S.K.; Kim, Y.; Jin, C.; Lee, S.H.; Kang, M.J.; Jeong, T.C.; Jeong, S.Y.; Kim, D.H.; Yoo, H.H. Inhibitory effects of deoxypodophyllotoxin from Anthriscus sylvestris on human CYP2C9 and CYP3A4. Planta Med. 2010, 76, 701–704. [Google Scholar] [CrossRef]
  58. Ikeda, R.; Nagao, T.; Okabe, H.; Nakano, Y.; Matsunaga, H.; Katano, M.; Mori, M. Antiproliferative constituents in Umbelliferae plants. III. Constituents in the root and the ground part of Anthriscus sylvestris Hoffm. Chem. Pharm. Bull. 1998, 46, 871–874. [Google Scholar] [CrossRef] [Green Version]
  59. Dall’Acqua, S.; Giorgetti, M.; Cervellati, R.; Innocenti, G. Deoxypodophyllotoxin content and antioxidant activity of aerial parts of Anthriscus sylvestris Hoffm. Zeitschrift für Naturforschung C 2006, 61, 658–662. [Google Scholar] [CrossRef]
  60. Sánchez-Monroy, M.B.; León-Rivera, I.; Llanos-Romero, R.E.; García-Bores, A.M.; Guevara-Fefer, P. Cytotoxic activity and triterpenes content of nine Mexican species of Bursera. Nat. Prod. Res. 2021, 35, 4881–4885. [Google Scholar] [CrossRef]
  61. Jutiviboonsuk, A.; Zhang, H.; Tan Ghee, T.; Ma, C.; Van Hung, N.; Manh Cuong, N.; Bunyapraphatsara, N.; Soejarto, D.D.; Fong Harry, H.S. Bioactive constituents from roots of Bursera tonkinensis. Phytochemistry 2005, 66, 2745–2751. [Google Scholar] [CrossRef]
  62. Peraza-Sanchez, S.R.; Pena-Rodriguez, L.M. Isolation of picropolygamain from the resin of Bursera simaruba. J. Nat. Prod. 1992, 55, 1768–1771. [Google Scholar] [CrossRef]
  63. Antúnez-Mojica, M.; Rojas-Sepúlveda, A.M.; Mendieta-Serrano, M.A.; Gonzalez-Maya, L.; Marquina, S.; Salas-Vidal, E.; Alvarez, L. Lignans from Bursera fagaroides affect in vivo cell behavior by disturbing the tubulin cytoskeleton in zebrafish embryos. Molecules 2018, 24, 8. [Google Scholar] [CrossRef] [Green Version]
  64. Dehelean, C.A.; Marcovici, I.; Soica, C.; Mioc, M.; Coricovac, D.; Iurciuc, S.; Cretu, O.M.; Pinzaru, I. Plant-derived anticancer compounds as new perspectives in drug discovery and alternative therapy. Molecules 2021, 26, 1109. [Google Scholar] [CrossRef] [PubMed]
  65. Aynehchi, Y. Deoxypodophyllotoxin, the cytotoxic principle of Callitris columellaris. J. Pharm. Sci. 1971, 60, 121–122. [Google Scholar] [CrossRef] [PubMed]
  66. Van Uden, W.; Pras, N.; Maingré, T.M. The accumulation of podophyllotoxin-β-D-glucoside by cell suspension cultures derived from the conifer Callitris drummondii. Plant Cell Rep. 1990, 9, 257–260. [Google Scholar] [CrossRef] [PubMed]
  67. Mohammadhosseini, M.; Venditti, A.; Frezza, C.; Serafini, M.; Bianco, A.; Mahdavi, B. The Genus Haplophyllum Juss.: Phytochemistry and Bioactivities—A Review. Molecules 2021, 26, 4664. [Google Scholar] [CrossRef]
  68. Gozler, B.; Arar, G.; Gozler, T.; Hesse, M. Isodaurinol, an arylnaphthalene lignan from Haplophyllum cappadocicum. Phytochemistry 1992, 31, 2473–2475. [Google Scholar]
  69. Gozler, B.; Gozler, T.; Saglam, H.; Hesse, M. Minor lignans from Haplophyllum cappadocicum. Phytochemistry 1996, 42, 689–693. [Google Scholar] [CrossRef]
  70. Dekebo, A.; Lang, M.; Polborn, K.; Dagne, E.; Steglich, W. Four lignans from Commiphora erlangeriana. J. Nat. Prod. 2002, 65, 1252–1257. [Google Scholar] [CrossRef]
  71. Habtemariam, S. Cytotoxic and cytostatic activity of erlangerins from Commiphora erlangeriana. Toxicon 2003, 41, 723–727. [Google Scholar] [CrossRef]
  72. Ma, C.; Yang, J.S.; Luo, S.R. Study on lignans from Diphylleia sinensis. Acta Pharm. Sin. 1993, 28, 690–694. [Google Scholar]
  73. Yuan, Y.; Wang, Y.; Huang, M.; Xu, R.; Zeng, H.; Nie, C.; Kong, J. Development and characterization of molecularly imprinted polymers for the selective enrichment of podophyllotoxin from traditional Chinese medicines. Anal. Chim. Acta 2011, 695, 63–72. [Google Scholar] [CrossRef]
  74. Xu, X.; Gao, X.; Jin, L.; Bhadury, P.S.; Yuan, K.; Hu, D.; Song, B.; Yang, S. Antiproliferation and cell apoptosis inducing bioactivities of constituents from Dysosma versipellis in PC3 and Bcap-37 cell lines. Cell Div. 2011, 6, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Jiang, R.W.; Zhou, J.R.; Hon, P.M.; Li, S.L.; Zhou, Y.; Li, L.L.; Ye, W.C.; Xu, H.X.; Shaw, P.C.; But, P.P.H. Lignans from Dysosma versipellis with Inhibitory Effects on Prostate Cancer Cell Lines. J. Nat. Prod. 2007, 70, 283–286. [Google Scholar] [CrossRef] [PubMed]
  76. Feng, H.; Chen, G.; Zhang, Y.; Guo, M. Potential multifunctional bioactive compounds from Dysosma versipellis explored by bioaffinity ultrafiltration-HPLC/MS with Topo I, Topo II, COX-2 and ACE2. J. Inflamm. Res. 2022, 15, 4677–4692. [Google Scholar] [CrossRef] [PubMed]
  77. Karuppaiya, P.; Tsay, H.S. Therapeutic values, chemical constituents and toxicity of Taiwanese Dysosma pleiantha—A review. Toxicol. Lett. 2015, 236, 90–97. [Google Scholar] [CrossRef]
  78. Uden, W. The production of podophyllotoxin and related cytotoxic lignans by plant cell cultures. Plant Cell Tissue Organ. Cult. 1992, 20, 81–87. [Google Scholar] [CrossRef] [Green Version]
  79. Hu, S.; Zhou, Q.; Wu, W.R.; Duan, Y.X.; Gao, Z.Y.; Li, Y.W.; Lu, Q. Anticancer effect of deoxypodophyllotoxin induces apoptosis of human prostate cancer cells. Oncol. Lett. 2016, 12, 2918–2923. [Google Scholar] [CrossRef] [Green Version]
  80. Santos, E.O.; Lima, L.S.; David, J.M.; Martins, L.C.; Guedes, M.L.S.; David, J.P. Podophyllotoxin and other aryltetralin lignans from Eriope latifolia and Eriope blanchetii. Nat. Prod. Res. 2011, 25, 1450–1453. [Google Scholar] [CrossRef]
  81. David, J.P.; Silva, E.F.; Moura, D.L.; Guedes, M.L.; Assunção, R.J.; David, J.M. Lignans and triterpenes from cytotoxic extract of Eriope blanchetii. Quim. Nova 2001, 24, 730–733. [Google Scholar] [CrossRef] [Green Version]
  82. Raffauf, R.F.; Kelley, C.J.; Ahmad, Y.; Le Quesne, P.W. Alpha- and beta-peltatin from Eriope macrostachya. J. Nat. Prod. 1987, 50, 772–773. [Google Scholar] [CrossRef]
  83. Lim, S.; Grassi, J.; Akhmedjanova, V.; Debiton, E.; Balansard, G.; Beliveau, R.; Barthomeuf, C. Reversal of P-glycoprotein-mediated drug efflux by eudesmin from Haplophyllum perforatum and cytotoxicity pattern versus diphyllin, podophyllotoxin and etoposide. Planta Med. 2007, 73, 1563–1567. [Google Scholar] [CrossRef]
  84. Bessonova, I.A. Components of Haplophyllum bucharicum. Chem. Nat. Compd. 2000, 36, 323–324. [Google Scholar] [CrossRef]
  85. Štefánik, M.; Bhosale, D.S.; Haviernik, J.; Straková, P.; Fojtíková, M.; Dufková, L.; Huvarová, I.; Salát, J.; Bartáček, J.; Svoboda, J.; et al. Diphyllin Shows a Broad-Spectrum Antiviral Activity against Multiple Medically Important Enveloped RNA and DNA Viruses. Viruses 2022, 14, 354. [Google Scholar] [CrossRef] [PubMed]
  86. Al-Abed, Y.; Abu-Zarga, M.; Sabri, S.; Atta Ur, R.; Voelter, W. A arylnaphthalene lignan from Haplophyllum buxbaumii. Phytochemistry 1998, 49, 1779–1781. [Google Scholar] [CrossRef] [PubMed]
  87. Al-Abed, Y.; Sabri, S.; Zarga, M.A.; Shah, Z.; Attaur, R. Chemical constituents of the flora of Jordan, part V-B. Three new arylnaphthalene lignan glucosides from Haplophyllum buxbaumii. J. Nat. Prod. 1990, 53, 1152–1161. [Google Scholar] [CrossRef]
  88. Nakul, G.S.; Zarga, M.H.A.; Sabri, S.S.; Al-Eisawi, D.M. Chemical constituents of the flora of Jordan. Part III. Mono-O-acetyl diphyllin apioside, a new arylnaphthalene lignan from Haplophyllum buxbaumii. J. Nat. Prod. 1987, 50, 748–750. [Google Scholar] [CrossRef]
  89. Aimaiti, S.; Saito, Y.; Fukuyoshi, S.; Goto, M.; Miyake, K.; Newman, D.J.; O’Keefe, B.R.; Lee, K.H.; Nakagawa-Goto, K. Isolation, Structure Elucidation, and Antiproliferative Activity of Butanolides and Lignan Glycosides from the Fruit of Hernandia nymphaeifolia. Molecules 2019, 24, 4005. [Google Scholar] [CrossRef] [Green Version]
  90. Udino, L.; Abaul, J.; Bourgeois, P.; Gorrichon, L.; Duran, H.; Zedde, C. Lignans from the seeds of Hernandia sonora. Planta Med. 1999, 65, 279–281. [Google Scholar] [CrossRef]
  91. Ito, C.; Itoigawa, M.; Ogata, M.; Mou, X.Y.; Tokuda, H.; Nishino, H.; Furukawa, H. Lignans as anti-tumor-promoter from the seeds of Hernandia ovigera. Planta Med. 2001, 67, 166–168. [Google Scholar] [CrossRef]
  92. Gu, J.Q.; Park, E.J.; Totura, S.; Riswan, S.; Fong, H.H.; Pezzuto, J.M.; Kinghorn, A.D. Constituents of the twigs of Hernandia ovigera that inhibit the transformation of JB6 murine epidermal cells. J. Nat. Prod. 2002, 65, 1065–1068. [Google Scholar] [CrossRef]
  93. Liu, H.P. Studies on the Chemical Constituentis of Rhamnus Hainanensis and Hernandia ovigera L. Master’s Thesis, Qingdao University of Science and Technology, Shangdou, China, 2013. [Google Scholar]
  94. Fragoso-Serrano, M.; Pereda-Miranda, R. Dereplication of podophyllotoxin and related cytotoxic lignans in Hyptis verticillata by ultra-high-performance liquid chromatography tandem mass spectrometry. Phytochem. Anal. 2020, 31, 81–87. [Google Scholar] [CrossRef]
  95. Picking, D.; Delgoda, R.; Boulogne, I.; Mitchell, S. Hyptis verticillata Jacq, a review of its traditional uses, phytochemistry, pharmacology and toxicology. J. Ethnopharmacol. 2013, 147, 16–41. [Google Scholar] [CrossRef] [PubMed]
  96. Bridi, H.; de Carvalho Meirelles, G.; von Poser, G.L. Subtribe Hyptidinae (Lamiaceae): A promising source of bioactive metabolites. J. Ethnopharmacol. 2021, 264, 113225. [Google Scholar] [CrossRef] [PubMed]
  97. Lu, Y.H.; Wei, B.L.; Ko, H.H.; Lin, C.N. DNA strand-scission by phloroglucinols and lignans from heartwood of Garcinia subelliptica Merr. and Justicia plants. Phytochemistry 2008, 69, 225–233. [Google Scholar] [CrossRef] [PubMed]
  98. Li, S.J.; Liang, Z.Z.; Li, J.J.; Zhang, X.; Zheng, R.H.; Zhao, C.Q. Update on naturally occurring novel arylnaphthalenes from plants. Phytochem. Rev. 2020, 19, 337–403. [Google Scholar] [CrossRef]
  99. Al-Juaid, S.S.; Abdel-Mogib, M. A novel podophyllotoxin lignan from Justicia heterocarpa. Chem. Pharm. Bull. 2004, 52, 507–509. [Google Scholar] [CrossRef] [Green Version]
  100. Suryanarayana, A.A.; Shankar, S.; Manoharan, J.P.; Vidyalakshmi, S. In-Silico analysis and molecular docking studies of phytoconstituents of Justicia adhatoda as potential inhibitors of SARS-CoV2 target proteins. In Proceedings of the First International Conference on Combinatorial and Optimization, ICCAP 2021, Chennai, India, 7–8 December 2021. [Google Scholar]
  101. Weng, J.R.; Ko, H.H.; Yeh, T.L.; Lin, H.C.; Lin, C.N. Two new arylnaphthalide lignans and antiplatelet constituents from Justicia procumbens. Arch. Pharm. 2004, 337, 207–212. [Google Scholar] [CrossRef]
  102. Day, S.H.; Lin, Y.C.; Tsai, M.L.; Tsao, L.T.; Ko, H.H.; Chung, M.I.; Lee, J.C.; Wang, J.P.; Won, S.J.; Lin, C.N. Potent cytotoxic lignans from Justicia procumbens and their effects on nitric oxide and tumor necrosis factor-alpha production in mouse macrophages. J. Nat. Prod. 2002, 65, 379–381. [Google Scholar] [CrossRef]
  103. Zhao, Y.; Ku, C.F.; Xu, X.Y.; Tsang, N.Y.; Zhu, Y.; Zhao, C.L.; Liu, K.L.; Li, C.C.; Rong, L.; Zhang, H.J. Stable axially chiral isomers of arylnaphthalene lignan glycosides with antiviral potential discovered from Justicia procumbens. J. Org. Chem. 2021, 86, 5568–5583. [Google Scholar] [CrossRef]
  104. Xu, X.Y.; Wang, D.Y.; Ku, C.F.; Zhao, Y.; Cheng, H.; Liu, K.L.; Rong, L.J.; Zhang, H.J. Anti-HIV lignans from Justicia procumbens. Chin. J. Nat. Med. 2019, 17, 945–952. [Google Scholar] [CrossRef]
  105. Liu, B.; Zhang, T.; Xie, Z.T.; Hong, Z.C.; Lu, Y.; Long, Y.M.; Ji, C.Z.; Liu, Y.T.; Yang, Y.F.; Wu, H.Z. Effective components and mechanism analysis of anti-platelet aggregation effect of Justicia procumbens L. J. Ethnopharmacol. 2022, 294, 115392. [Google Scholar] [CrossRef]
  106. Hemmati, S.; Seradj, H. Justicidin B, a promising bioactive lignan. Molecules 2016, 21, 820. [Google Scholar] [CrossRef]
  107. Premjet, D.; Tachibana, S. Production of podophyllotoxin by immobilized cell cultures of Juniperus chinensis. Pakistan J. Biol. Sci. 2004, 7, 1130–1134. [Google Scholar]
  108. Xu, S.; Li, X.; Liu, S.; Tian, P.; Li, D. Juniperus sabina L. as a source of podophyllotoxins, extraction optimization and Anticholinesterase activities. Int. J. Mol. Sci. 2022, 23, 10205. [Google Scholar] [CrossRef] [PubMed]
  109. Xie, S.; Li, G.; Qu, L.; Zhong, R.; Chen, P.; Lu, Z.; Zhou, J.; Guo, X.; Li, Z.; Ma, A.; et al. Podophyllotoxin Extracted from Juniperus sabina Fruit Inhibits Rat Sperm Maturation and Fertility by Promoting Epididymal Epithelial Cell Apoptosis. Evid. Based Complement Alternat. Med. 2017, 2017, 6958982. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Guerrero, E.; Abad, A.; Montenegro, G.; Del Olmo, E.; López-Pérez, J.L.; San Feliciano, A. Analgesic and anti-inflammatory activity of podophyllotoxin derivatives. Pharm. Biol. 2013, 51, 566–572. [Google Scholar] [CrossRef] [PubMed]
  111. Kašparová, M.; Jand, M.; Tůmová, L.; Spilková, J. Production of podophyllotoxin by plant tissue cultures of Juniperus virginiana. Nat. Prod. Commun. 2017, 12, 101–103. [Google Scholar] [CrossRef] [Green Version]
  112. Maqbool, M.; Cushman, K.E.; Gerard, P.D.; Bedir, E.; Lata, H.; Moraes, R.M. Podophyllotoxin content in leaves of Eastern red cedar (Juniperus virginiana). Acta Hortic. 2004, 629, 87–92. [Google Scholar] [CrossRef]
  113. Belma, K. Arytetralin lignans from Linum catharticum L. Biochem. Syst. Ecol. 1998, 26, 795–796. [Google Scholar] [CrossRef]
  114. Klaes, M.; Ellendorff, T.; Schmidt, T.J. 6-Methoxypodophyllotoxin-7-O-n-Hexanoate, a new aryltetralin lignan ester from seeds of Linum flavum. Planta Med. 2010, 76, 719. [Google Scholar] [CrossRef]
  115. Mikame, K.; Sakakibara, N.; Umezawa, T.; Shimada, M. Lignans of Linum flavum var. compactum. J. Wood Sci. 2002, 48, 440–445. [Google Scholar] [CrossRef]
  116. Renouard, S.; Corbin, C.; Drouet, S.; Medvedec, B.; Doussot, J.; Colas, C.; Maunit, B.; Bhambra, A.S.; Gontier, E.; Jullian, N.; et al. Investigation of Linum flavum (L.) Hairy root cultures for the production of anticancer aryltetralin lignans. Int. J. Mol. Sci. 2018, 19, 990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Mohagheghzadeh, A.; Gholami, A.; Soltani, M.; Hemmati, S.; Alfermann, A.W. Linum mucronatum, Organ to organ lignan variations. Z. Nat. C 2005, 60, 508–510. [Google Scholar] [CrossRef] [PubMed]
  118. Mohagheghzadeh, A.; Hemmati, S.; Mehregan, I.; Alfermann, A.W. Linum persicum, lignans and placement in Linaceae. Phytochem. Rev. 2004, 2, 363–369. [Google Scholar] [CrossRef]
  119. Javidnia, K.; Miri, R.; Rezai, H.; Jafari, A.; Azarmehr, A.; Amirghofran, Z. Biological Activity and Aryltetraline Lignans of Linum persicum. Pharm. Biol. 2005, 43, 547–550. [Google Scholar] [CrossRef]
  120. Vasilev, N.P.; Ionkova, I.I. Isolation and structure elucidation of aryltetralin lignans from Linum tauricum ssp. bulgaricum. Pharmacogn. Mag. 2006, 2, 169–174. [Google Scholar]
  121. Anand, U.; Biswas, P.; Kumar, V.; Ray, D.; Ray, P.; Loake, V.I.P.; Kandimalla, R.; Chaudhary, A.; Singh, B.; Routhu, N.K.; et al. Podophyllum hexandrum and its active constituents, Novel radioprotectants. Biomed. Pharmacother. 2022, 146, 112555. [Google Scholar] [CrossRef]
  122. Zilla, M.K.; Nayak, D.; Amin, H.; Nalli, Y.; Rah, B.; Chakraborty, S.; Kitchlu, S.; Coswami, A.; Ali, A. 4′-Demethyl-deoxypodophyllotoxin glucoside isolated from Podophyllum hexandrum exhibits potential anticancer activities by altering Chk-2 signaling pathway in MCF-7 breast cancer cells. Chem.Biol. Interact. 2014, 224, 100–107. [Google Scholar] [CrossRef]
  123. Jackson, D.E.; Dewick, P.M. Aryltetralin lignans from Podophyllum hexandrum and Podophyllum peltatum. Phytochemistry 1984, 23, 1147–1152. [Google Scholar] [CrossRef]
  124. Hoffmann, J.J.; Wiedhopf, R.M.; Cole, J.R. Cytotoxic and tumor inhibitory agent from Polygala macradenia gray (Polygalaceae): 4′-demethyldeoxypodophyllotoxin. J. Pharm. Sci. 1977, 66, 586–587. [Google Scholar] [CrossRef]
  125. Zhao, C.; Nagatsu, A.; Hatano, K.; Shirai, N.; Kato, S.; Ogihara, Y. New lignan glycosides from Chinese medicinal plant, Sinopodophillum emodi. Chem. Pharm. Bull. 2003, 51, 255–261. [Google Scholar] [CrossRef] [Green Version]
  126. Zhao, C.; Cao, W.; Nagatsu, A.; Ogihara, Y. Three new glycosides from Sinopodophyllum emodi (WALL.) YING. Chem. Pharm. Bull. 2001, 49, 1474–1476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Yokosuka, A.; Yamada, C.; Saito, M.; Yokogawa, S.; Mimaki, Y. Chemical constituents of the leaves of Thujopsis dolabrata and their cytotoxicity. Chem. Pharm. Bull. 2022, 70, 720–725. [Google Scholar] [CrossRef] [PubMed]
  128. Hande, K.R. Etoposide, four decades of development of a topoisomerase II inhibitor. Eur. J. Cancer 1998, 34, 1514–1521. [Google Scholar] [CrossRef] [PubMed]
  129. Bender, R.P.; Jablonksy, M.J.; Shadid, M.; Romaine, I.; Dunlap, N.; Anklin, C.; Graves, D.E.; Osheroff, N. Substituents on etoposide that interact with human topoisomerase II alpha in the binary enzyme-drug complex, Contributions to etoposide binding and activity. Biochemistry 2008, 47, 4501–4509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Lichota, A.; Gwozdzinski, K. Anticancer activity of natural compounds from plant and marine environment. Int. J. Mol. Sci. 2018, 19, 3533. [Google Scholar] [CrossRef] [Green Version]
  131. Jin, F.; Zhao, L.; Guo, Y.J.; Zhao, W.J.; Zhang, H.; Wang, H.T.; Shao, T.; Zhang, S.L.; Wei, Y.J.; Feng, J.; et al. Influence of Etoposide on anti-apoptotic and multidrug resistance-associated protein genes in CD133 positive U251 glioblastoma stem-like cells. Brain Res. 2010, 1336, 103–111. [Google Scholar] [CrossRef]
  132. Vasilcanu, D.; Weng, W.H.; Girnita, A.; Lui, W.O.; Vasilcanu, R.; Axelson, M.; Larsson, O.; Larsson, C.; Girnita, L. The insulin-like growth factor-1 receptor inhibitor PPP produces only very limited resistance in tumor cells exposed to long-term selection. Oncogene 2006, 25, 3186–3195. [Google Scholar] [CrossRef] [Green Version]
  133. Lima, L.M.; Barreiro, E.J. Bioisosterism, A useful strategy for molecular modification and drug design. Curr. Med. Chem. 2005, 12, 23–49. [Google Scholar] [CrossRef]
  134. Bertounesque, E.; Meresse, P.; Monneret, C.; Florent, J.C. Synthetic approach to condensed heterocyclic analogues from etoposide revisited. Synthesis of A-ring pyridazine etoposide. Tetrahedron Lett. 2007, 48, 5781–5784. [Google Scholar] [CrossRef]
  135. Zhu, X.K.; Guan, J.; Xiao, Z.; Cosentino, L.M.; Lee, K.H. Anti-AIDS agents. Part 61, Anti-HIV activity of new podophyllotoxin derivatives. Bioorg. Med. Chem. 2004, 12, 4267–4273. [Google Scholar] [CrossRef]
  136. Magedov, I.V.; Manpadi, M.; Van Slambrouck, S.; Steelant, W.F.A.; Rozhkova, E.; Przheval’skii, N.M.; Rogelj, S.; Kornienko, A. Discovery and investigation of antiproliferative and apoptosis-inducing properties of new heterocyclic podophyllotoxin analogues accessible by a one-step multicomponent synthesis. J. Med. Chem. 2007, 50, 5183–5192. [Google Scholar] [CrossRef] [PubMed]
  137. Zhao, M.; Feng, M.; Bai, S.F.; Zhang, Y.; Chao Bi, W.; Chen, H. Synthesis and antitumor activity of novel podophyllotoxin derivatives. Chin. Chem. Lett. 2009, 20, 901–904. [Google Scholar] [CrossRef]
  138. Kamal, A.; Ashwini Kumar, B.; Arifuddin, M.; Dastidar, S.G. Synthesis of 4[beta]-amido and 4[beta]-sulphonamido analogues of podophyllotoxin as potential antitumour agents. Bioorg. Med. Chem. 2003, 11, 5135–5142. [Google Scholar] [CrossRef]
  139. Xiao, Z.; Bastow, K.F.; Vance, J.R.; Lee, K.H. Antitumor agents. Part 227, Studies on novel 4′-O-demethyl-epipodophyllotoxins as antitumor agents targeting topoisomerase II. Bioorg. Med. Chem. 2004, 12, 3339–3344. [Google Scholar]
  140. Xi, W.L.; Cai, Q.; Tang, Y.B.; Sun, H.; Xiao, Z.Y. Design and synthesis of novel cytotoxic podophyllotoxin derivatives. Chin. Chem. Lett. 2010, 21, 1153–1156. [Google Scholar] [CrossRef]
  141. Zhao, M.; Zhang, Y.; Yang, Z.; Cao, B.; Zheng, Y.; Chen, H. Synthesis of podophyllotoxin derivatives and their antitumor activity in vitro. Chin. J. Med. Chem. 2009, 19, 85–88. [Google Scholar]
  142. Lu, Y.; Zuo, S.; Shi, S.; Li, X.; Zhang, Y.; Lu, J.; Chen, H. Synthesis and antitumor activities of podophyllotoxin derivatives. Chin. J. Med. Chem. 2010, 20, 90–95. [Google Scholar]
  143. Guianvarc’h, D.; Duca, M.; Boukarim, C.; Kraus-Berthier, L.; Léonce, S.; Pierré, A.; Pfeiffer, B.; Renard, P.; Arimondo, P.B.; Monneret, C.; et al. Synthesis and biological activity of sulfonamide derivatives of epipodophyllotoxin. J. Med. Chem. 2004, 47, 2365–2374. [Google Scholar] [CrossRef] [PubMed]
  144. Hu, C.Q.; Xu, D.Q.; Du, W.T.; Qian, S.J.; Wang, L.; Lou, J.S.; He, Q.J.; Yang, B.; Hu, Y.Z. Novel 4 beta-anilino-podophyllotoxin derivatives, design synthesis and biological evaluation as potent DNA-topoisomerase II poisons and anti-MDR agents. Mol. Biosyst. 2010, 6, 410–420. [Google Scholar] [CrossRef]
  145. Kamal, A.; Kumar, B.A.; Suresh, P.; Agrawal, S.K.; Chashoo, G.; Singh, S.K.; Saxena, A.K. Synthesis of 4[beta]-N-polyaromatic substituted podophyllotoxins, DNA topoisomerase inhibition, anticancer and apoptosis-inducing activities. Bioorg. Med. Chem. 2010, 18, 8493–8500. [Google Scholar] [CrossRef]
  146. Bhat Bilal, A.; Reddy, P.B.; Agrawal Satyam, K.; Saxena, A.K.; Kumar, H.M.S.; Qazi, G.N. Studies on novel 4beta-[(4-substituted)-1,2,3-triazol-1-yl] podophyllotoxins as potential anticancer agents. Eur. J. Med. Chem. 2008, 43, 2067–2072. [Google Scholar] [CrossRef] [PubMed]
  147. Miao, R.; Han, Y.; An, L.; Yang, J.; Wang, Q. Seleno-podophyllotoxin derivatives induce hepatoma SMMC-7721 cell apoptosis through Bax pathway. Cell Bio. Int. 2008, 32, 217–223. [Google Scholar] [CrossRef] [PubMed]
  148. Wang, P.H.; Zhang, Q.; Yuan, W.B.; Wang, L.F.; Song, Y.M.; Liu, H.H. Studies on the mechanism of the interaction between hydrazide-podophyllic Ni(II), Co(II), Zn(II) metal complexes and DNA. Spectrosc. Spectr. Anal. 2006, 26, 1298–1302. [Google Scholar]
  149. Usami, Y.; Arimoto, M.; Kobayashi, K.; Honjou, M.; Yamanaka, M.; Miyao, M.; Ichikawa, H.; Bastow, K.F.; Lee, K.H. Synthesis of aryltetralin type 2-azalignans using schollkopf’s bislactim-ether methodology. Heterocycles 2009, 78, 2041–2052. [Google Scholar] [CrossRef]
  150. Iida, A.; Kano, M.; Kubota, Y.; Koga, K.; Tomioka, K. Podophyllotoxin aza-analogue, a novel DNA topoisomerase II inhibitor. Chem. Pharm. Bull. 2000, 48, 486–489. [Google Scholar] [CrossRef] [Green Version]
  151. Hitotsuyanagi, Y.; Kobayashi, M.; Morita, H.; Itokawa, H.; Takeya, K. Synthesis of (-)-4-aza-4-deoxypodophyllotoxin from (-)-podophyllotoxin. Tetrahedron Lett. 1999, 40, 9107–9110. [Google Scholar] [CrossRef]
  152. Hitotsuyanagi, Y.; Fukuyo, M.; Tsuda, K.; Kobayashi, M.; Ozeki, A.; Itokawa, H.; Takeya, K. 4-aza-2,3-dehydro-4-deoxypodophyllotoxins, Simple aza-podophyllotoxin analogues possessing potent cytotoxicity. Bioorg. Med. Chem. Lett. 2000, 10, 315–317. [Google Scholar] [CrossRef]
  153. Shi, C.L.; Wang, J.X.; Chen, H.; Shi, D.Q. Regioselective synthesis and in vitro anticancer activity of 4-aza-podophyllotoxin derivatives catalyzed by l-proline. J. Comb. Chem. 2010, 12, 430–434. [Google Scholar] [CrossRef]
  154. Labruere, R.; Gautier, B.; Testud, M.; Seguin, J.; Lenoir, C.; Desbene-Finck, S.; Helissey, P.; Garbay, C.; Chabot, G.G.; Vidal, M.; et al. Design, synthesis, and biological evaluation of the first podophyllotoxin analogues as potential vascular-disrupting agents. Chem. Med. Chem. 2010, 5, 2016–2025. [Google Scholar] [CrossRef]
  155. Singh, P.; Faridi, U.; Srivastava, S.; Kumar, J.K.; Darokar, M.P.; Luqman, S.; Shanker, K.; Chanotiya, C.S.; Gupta, A.; Gupta, M.M.; et al. Design and synthesis of C-ring lactone- and lactam-based podophyllotoxin analogs as anticancer agents. Chem. Pharm. Bull. 2010, 58, 242–246. [Google Scholar] [CrossRef]
  156. Zhao, Y.; Feng, J.H.; Ding, H.X.; Xiong, Y.; Cheng, C.H.K.; Hao, X.J.; Zhang, Y.M.; Pan, Y.J.; Gueritte, F.; Wu, X.M.; et al. Synthesis and cytotoxicity of racemic isodeoxypodophyllotoxin analogues with isoprene-derived side chains. J. Nat. Prod. 2006, 69, 1145–1152. [Google Scholar] [CrossRef] [PubMed]
  157. Castro, M.A.; Miguel del Corral, J.M.; Gordaliza, M.; Garcia, P.A.; Gomez-Zurita, M.A.; San Feliciano, A. Synthesis and cytotoxic evaluation of C-9 oxidized podophyllotoxin derivatives. Bioorg. Med. Chem. 2007, 15, 1670–1678. [Google Scholar] [CrossRef]
  158. Castro, M.A.; Miguel del Corral, J.M.; Gordaliza, M.; García, P.A.; Gómez-Zurita, M.A.; García-Grávalos, M.D.; de la Iglesia-Vicente, J.; Gajate, C.; An, F.; Mollinedo, F.; et al. Synthesis and biological evaluation of new selective cytotoxic cyclolignans derived from podophyllotoxin. J. Med. Chem. 2004, 47, 1214–1222. [Google Scholar] [CrossRef] [PubMed]
  159. Castro, M.A.; Miguel del Corral, J.M.; Garcia, P.A.; Rojo, M.V.; de la Iglesia-Vicente, J.; Mollinedo, F.; Cuevas, C.; San Feliciano, A. Synthesis and Biological Evaluation of New Podophyllic Aldehyde Derivatives with Cytotoxic and Apoptosis-Inducing Activities. J. Med. Chem. 2010, 53, 983–993. [Google Scholar] [CrossRef] [PubMed]
  160. Imperio, D.; Pirali, T.; Galli, U.; Pagliai, F.; Cafici, L.; Canonico, P.L.; Sorba, G.; Genazzani, A.A.; Tron, G.C. Replacement of the lactone moiety on podophyllotoxin and steganacin analogues with a 1,5-disubstituted 1,2,3-triazole via ruthenium-catalyzed click chemistry. Bioorg. Med. Chem. 2007, 15, 6748–6757. [Google Scholar] [CrossRef]
  161. Xiao, Z.; Han, S.; Bastow, K.F.; Lee, K.H. Antitumor agents. Part 232, Synthesis of cyclosulfite podophyllotoxin analogues as novel prototype antitumor agents. Bioorg. Med. Chem. Lett. 2004, 14, 1581–1584. [Google Scholar] [CrossRef] [PubMed]
  162. Xiao, Z.; Vance, J.R.; Bastow, K.F.; Brossi, A.; Wang, H.K.; Lee, K.H. Antitumor agents. Part 235, Novel 4′-ester etoposide analogues as potent DNA topoisomerase II inhibitors with improved therapeutic potential. Bioorg. Med. Chem. 2004, 12, 3363–3369. [Google Scholar] [CrossRef]
  163. Wu, C.C.; Li, T.K.; Farh, L.; Lin, L.Y.; Lin, T.S.; Yu, Y.J.; Yen, T.J.; Chiang, C.W.; Chan, N.L. Structural basis of type II topoisomerase inhibition by the anticancer drug etoposide. Science 2011, 333, 459–462. [Google Scholar] [CrossRef] [Green Version]
  164. Labruere, R.; Helissey, P.; Desbene-Finck, S.; Giorgi-Renault, S. Design and effective synthesis of the first 4-aza-2,3-didehydropodophyllotoxin rigid aminologue, a N-methyl-4-[(3,4,5- trimethoxyphenyl) amino]]-1,2- dihydroquinoline-lactone. J. Org. Chem. 2008, 73, 3642–3645. [Google Scholar] [CrossRef]
  165. Wright, J.L.; Gregory, T.F.; Kesten, S.R.; Boxer, P.A.; Serpa, K.A.; Meltzer, L.T.; Wise, L.D.; Espitia, S.A.; Konkoy, C.S.; Whittemore, E.R.; et al. Subtype-Selective N-Methyl-D-Aspartate Receptor Antagonists, Synthesis and Biological Evaluation of 1-(Heteroarylalkynyl)-4-benzylpiperidines. J. Med. Chem. 2000, 43, 3408–3419. [Google Scholar] [CrossRef]
  166. Wilkening, R.R.; Ratcliffe, R.W.; Fried, A.K.; Meng, D.; Sun, W.; Colwell, L.; Lambert, S.; Greenlee, M.; Nilsson, S.; Thorsell, A.; et al. Estrogen receptor beta-subtype selective tetrahydrofluorenones, use of a fused pyrazole as a phenol bioisostere. Bioorg. Med. Chem. Lett. 2006, 16, 3896–3901. [Google Scholar] [CrossRef] [PubMed]
  167. Magedov, I.V.; Manpadi, M.; Rozhkova, E.; Przheval’skii, N.M.; Rogelj, S.; Shors, S.T.; Steelant, W.F.A.; slambrouck, S.V.; Kornienko, A. Structural simplification of bioactive natural products with multicomponent synthesis: Dihydropyridopyrazole analogues of podophyllotoxin. Bioorg. Med. Chem. Lett. 2007, 17, 1381–1385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Herr, K.; Fleckenstein, M.; Brodrecht, M.; Höfler, M.V.; Heise, H.; Aussenac, F.; Gutmann, T.; Reggelin, M.; Buntkowsky, G. A novel strategy for site selective spin-labeling to investigate bioactive entities by DNP and EPR spectroscopy. Sci. Rep. 2021, 11, 13714. [Google Scholar] [CrossRef] [PubMed]
  169. Yang, L.; Nan, X.; Li, W.Q.; Wang, M.J.; Zhao, X.B.; Liu, Y.Q.; Zhang, Z.J.; Lee, K.H. Synthesis of novel spin-labeled podophyllotoxin derivatives as potential antineoplastic agents, Part XXV. Med. Chem. Res. 2014, 23, 4926–4931. [Google Scholar] [CrossRef]
  170. Jin, Y.; Chen, S.W.; Tian, X. Synthesis and biological evaluation of new spin-labeled derivatives of podophyllotoxin. Bioorg. Med. Chem. 2006, 14, 3062–3068. [Google Scholar] [CrossRef]
  171. Zhang, J.Q.; Zhang, Z.W.; Hui, L.; Chen, S.W.; Tian, X. Novel semisynthetic spin-labeled derivatives of podophyllotoxin with cytotoxic and antioxidative activity. Bioorg. Med. Chem. Lett. 2009, 20, 983–986. [Google Scholar] [CrossRef]
  172. Zhang, J.Q.; Zhang, Z.W.; Hui, L.; Tian, X. Design, synthesis and biological evaluation of novel spin-labeled derivatives of podophyllotoxin. Nat. Prod. Commun. 2010, 5, 241–244. [Google Scholar] [CrossRef]
  173. Kou, L.; Wang, M.J.; Wang, L.T.; Zhao, X.B.; Nan, X.; Yang, L.; Liu, Y.Q.; Morris-Natschke, S.L.; Lee, K.H. Toward synthesis of third-generation spin-labeled podophyllotoxin derivatives using isocyanide multicomponent reactions. Eur. J. Med. Chem. 2014, 75, 282–288. [Google Scholar] [CrossRef] [Green Version]
  174. Zhang, X.Y.; Li, W.G.; Wu, Y.J.; Tian, X. Antioxidative and antitumor activity of derivatives of 4-beta -amino-4′-demethylepipodophyllotoxin and their structure-activity relationship. Pharmazie 2007, 62, 432–438. [Google Scholar]
  175. Zhang, Z.W.; Zhang, J.Q.; Hui, L.; Chen, S.W.; Tian, X. First synthesis and biological evaluation of novel spin-labeled derivatives of deoxypodophyllotoxin. Eur. J. Med. Chem. 2010, 45, 1673–1677. [Google Scholar] [CrossRef]
  176. Khandrika, L.; Kumar, B.; Koul, S.; Maroni, P.; Koul, H.K. Oxidative stress in prostate cancer. Cancer Lett. 2009, 282, 125–136. [Google Scholar] [CrossRef]
  177. Khandia, R.; Munjal, A. Interplay between inflammation and cancer. Adv. Protein Chem. Struct. Biol. 2020, 119, 199–245. [Google Scholar] [PubMed]
  178. Elinav, E.; Nowarski, R.; Thaiss, C.A.; Hu, B.; Jin, C.; Flavell, R.A. Inflammation-induced cancer, crosstalk between tumours, immune cells and microorganisms. Nat. Rev. Cancer 2013, 13, 759–771. [Google Scholar] [CrossRef] [PubMed]
  179. Baniyash, M. Chronic inflammation, immunosuppression and cancer, new insights and outlook. Semin. Cancer Biol. 2006, 16, 80–88. [Google Scholar] [CrossRef] [PubMed]
  180. Minelli, A.; Bellezza, I.; Conte, C.; Culig, Z. Oxidative stress-related aging, A role for prostate cancer? Biochim. Biophys. Acta (BBA) 2009, 1795, 83–91. [Google Scholar] [CrossRef] [PubMed]
  181. Peters, G.J.; van der Wilt, C.L.; van Moorsel, C.J.; Kroep, J.R.; Bergman, A.M.; Ackland, S.P. Basis for effective combination cancer chemotherapy with antimetabolites. Pharmacol. Ther. 2000, 87, 227–253. [Google Scholar] [CrossRef]
  182. Chen, S.W.; Tian, X.; Tu, Y.Q. Synthesis and cytotoxic activity of novel derivatives of 4′-demethylepipodophyllotoxin. Bioorg. Med. Chem. Lett. 2004, 14, 5063–5066. [Google Scholar] [CrossRef]
  183. Zhang, F.M.; Yao, X.J.; Tian, X.; Tu, Y.Q. Synthesis and biological evaluation of new 4beta-5-Fu-substituted 4′-demethylepipodophyllotoxin derivatives. Molecules 2006, 11, 849–857. [Google Scholar] [CrossRef] [Green Version]
  184. Danieli, B.; Giardini, A.; Lesma, G.; Passarella, D.; Peretto, B.; Sacchetti, A.; Silvani, A.; Pratesi, G.; Zunino, F. Thiocolchicine-podophyllotoxin conjugates, Dynamic libraries based on disulfide exchange reaction. J. Org. Chem. 2006, 71, 2848–2853. [Google Scholar] [CrossRef]
  185. Passarella, D.; Giardini, A.; Peretto, B.; Fontana, G.; Sacchetti, A.; Silvani, A.; Ronchi, C.; Cappelletti, G.; Cartelli, D.; Borlak, J.; et al. Inhibitors of tubulin polymerization, Synthesis and biological evaluation of hybrids of vindoline, anhydrovinblastine and vinorelbine with thiocolchicine, podophyllotoxin and baccatin III. Bioorgan. Med. Chem. 2008, 16, 6269–6285. [Google Scholar] [CrossRef]
  186. Passarella, D.; Peretto, B.; Yepes, R.B.Y.; Cappelletti, G.; Cartelli, D.; Ronchi, C.; Snaith, J.; Fontana, G.; Danieli, B.; Borlak, J. Synthesis and biological evaluation of novel thiocolchicine-podophyllotoxin conjugates. Eur. J. Med. Chem. 2010, 45, 219–226. [Google Scholar] [CrossRef]
  187. Arimondo, P.; Boukarim, C.; Bailly, C.; Dauzonne, D.; Monneret, C. Design of two etoposide-amsacrine conjugates, topoisomerase II and tubuline polymerization inhibition and relation to cytotoxicity. Anti-Cancer Drug Des. 2000, 15, 413–421. [Google Scholar]
  188. Yu, P.F.; Chen, H.; Wang, J.; He, C.X.; Cao, B.; Li, M.; Yang, N.; Lei, Z.Y.; Cheng, M.S. Design, synthesis and cytotoxicity of novel podophyllotoxin derivatives. Chem. Pharm. Bull. 2008, 56, 831–834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Chen, H.; Bi, W.C.; Cao, B.; Yang, Z.X.; Chen, S.W.; Shang, H.; Yu, P.F.; Yang, J. A novel podophyllotoxin derivative (YB-1EPN) induces apoptosis and down-regulates express of P-glycoprotein in multidrug resistance cell line KBV200. Eur. J. Pharmacol. 2010, 627, 69–74. [Google Scholar] [CrossRef]
  190. Chen, H.; Wang, J.; Zhang, J.Z.; Wang, Y.Z.; Cao, B.; Bai, S.F.; Yu, P.F.; Bi, W.C.; Xie, W.L. L1EPO, a Novel Podophyllotoxin Derivative Overcomes P-Glycoprotein-Mediated Multidrug Resistance in K562/A02 Cell Line. Biol. Pharm. Bull. 2009, 32, 609–613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Chen, S.W.; Wang, Y.H.; Jin, Y.; Tian, X.; Zheng, Y.T.; Luo, D.Q.; Tu, Y.Q. Synthesis and anti-HIV-1 activities of novel podophyllotoxin derivatives. Bioorg. Med. Chem. Lett. 2007, 17, 2091–2095. [Google Scholar] [CrossRef]
  192. Li, J.H.; Chen, Y.; Wang, J.S.; Kong, W.; Jin, Y.H. Etoposide Induces Mitochondria-Associated Apoptotic Cell Death in Human Gastric Carcinoma Cells. Chem. Res. Chin. Univ. 2008, 24, 597–602. [Google Scholar] [CrossRef]
  193. Zhang, J.; Zhou, C.; Liu, S.; Chen, H.; Yang, C. Effect of podophyllotoxin on human gastric cancer cell line SGC 7901. Zhongnan Daxue Xuebao Yixueban 2008, 33, 718–722. [Google Scholar]
  194. Chirumbolo, S.; Conforti, A.; Lussignoli, S.; Metelmann, H.; Bellavite, P. Effects of Podophyllum peltatum compounds in various preparations and dilutions on human neutrophil functions in vitro. Br. Homeopath. J. 1997, 86, 16–26. [Google Scholar] [CrossRef]
  195. Zhang, W.J.; Wu, L.Q.; Liu, H.L.; Ye, L.Y.; Xin, Y.L.; Grau, G.E.; Lou, J.N. Abnormal blood vessels formed by human liver cavernous hemangioma endothelial cells in nude mice are suitable for drug evaluation. Microvasc. Res. 2009, 78, 379–385. [Google Scholar] [CrossRef]
  196. Lata, M.; Prasad, J.; Singh, S.; Kumar, R.; Singh, L.; Chaudhary, P.; Arora, R.; Chawla, R.; Tyagi, S.; Soni, N.L.; et al. Whole body protection against lethal ionizing radiation in mice by REC-2001, A semi-purified fraction of Podophyllum hexandrum. Phytomedicine 2009, 16, 47–55. [Google Scholar] [CrossRef]
  197. Bala, M.; Goel, H.C. Radioprotective effect of podophyllotoxin in Saccharomyces cerevisiae. J. Environ. Pathol. Toxicol. Oncol. 2004, 23, 139–144. [Google Scholar] [CrossRef] [PubMed]
  198. Arora, R.; Chawla, R.; Dhaker, A.S.; Adhikari, M.; Sharma, J.; Singh, S.; Gupta, D.; Kumar, R.; Sharma, A.; Sharma, R.K.; et al. Podophyllum hexandrum as a potential botanical supplement for the medical management of nuclear and radiological emergencies (NREs) and free radical-mediated ailments, leads from in vitro/in vivo radioprotective efficacy evaluation. J. Diet. Suppl. 2010, 7, 31–50. [Google Scholar] [CrossRef]
  199. Chawla, R.; Arora, R.; Kumar, R.; Sharma, A.; Prasad, J.; Singh, S.; Sagar, R.; Chaudhary, P.; Shukla, S.; Kaur, G.; et al. Antioxidant activity of fractionated extracts of rhizomes of high-altitude Podophyllum hexandrum, Role in radiation protection. Mol. Cell. Biochem. 2005, 273, 193–208. [Google Scholar] [CrossRef]
  200. Singh Pankaj, K.; Kumar, R.; Sharma, A.; Arora, R.; Chawla, R.; Jain Swatantra, K.; Sharma Rakesh, K. Podophyllum hexandrum fraction (REC-2006) shows higher radioprotective efficacy in the p53-carrying hepatoma cell line, a role of cell cycle regulatory proteins. Integr. Cancer Ther. 2009, 8, 261–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Razuvaev, A.; Henderson, B.; Girnita, L.; Larsson, O.; Axelson, M.; Hedin, U.; Roy, J. The cyclolignan picropodophyllin attenuates intimal hyperplasia after rat carotid balloon injury by blocking insulin-like growth factor-1 receptor signaling. J. Vasc. Surg. 2007, 46, 108–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Lin, C.X.; Son, M.J.; Ju, H.K.; Moon, T.C.; Lee, E.; Kim, S.H.; Kim, M.J.; Son, J.K.; Lee, S.H.; Chang, H.W. Deoxypodophyllotoxin, a naturally occurring lignan, inhibits the passive cutaneous anaphylaxis reaction. Planta Med. 2004, 70, 474–476. [Google Scholar]
  203. Xu, P.; Sun, Q.; Wang, X.J.; Zhang, S.G.; An, S.S.; Cheng, J.; Gao, R.; Xiao, H. Pharmacological effect of deoxypodophyllotoxin, A medicinal agent of plant origin, on mammalian neurons. Neurotoxicology 2010, 31, 680–686. [Google Scholar] [CrossRef]
  204. Zhang, Y.; Lv, M.; Xu, H. Insecticidal activity of twin compounds from podophyllotoxin and cytisine. Bioorg. Med. Chem. Lett. 2021, 43, 128104. [Google Scholar] [CrossRef]
  205. Linder, S.; Shoshan, M.C.; Gupta, R.S. Picropodophyllotoxin or podophyllotoxin does not induce cell death via insulin-like growth factor-I receptor. Cancer Res. 2007, 67, 2899. [Google Scholar] [CrossRef] [Green Version]
  206. Gao, R.; Gao, C.; Tian, X.; Yu, X.; Di, X.; Xiao, H.; Zhang, X. Insecticidal activity of deoxypodophyllotoxin, isolated from Juniperus sabina, and related lignans against larvae of Pieris rapae. Pest Manag. Sci. 2004, 60, 1131–1136. [Google Scholar] [CrossRef] [PubMed]
  207. Chen, Y.Q.; Xie, X. Podophyllotoxin induces CREB phosphorylation and CRE-driven gene expression via PKA but not MAPKs. Mol. Cells 2010, 29, 41–50. [Google Scholar] [CrossRef] [PubMed]
  208. Tseng, C.J.; Wang, Y.J.; Liang, Y.C.; Jeng, J.H.; Lee, W.S.; Lin, J.K.; Chen, C.H.; Liu, I.C.; Ho, Y.S. Microtubule damaging agents induce apoptosis in HL 60 cells and G2/M cell cycle arrest in HT 29 cells. Toxicology 2002, 175, 123–142. [Google Scholar] [CrossRef] [PubMed]
  209. Wang, B.; Chen, L.; Zhen, H.; Zhou, L.; Shi, P.; Huang, Z. Proteomic changes induced by podophyllotoxin in human cervical carcinoma HeLa cells. Am. J. Chin. Med. 2013, 41, 163–175. [Google Scholar] [CrossRef]
  210. Naik, P.K.; Dubey, A.; Soni, K.; Kumar, R.; Singh, H. The binding modes and binding affinities of epipodophyllotoxin derivatives with human topoisomerase II[alpha]. J. Mol. Graph. Model 2010, 29, 546–564. [Google Scholar] [CrossRef]
  211. Asami, Y.; Jia, D.W.; Tatebayashi, K.; Yamagata, K.; Tanokura, M.; Ikeda, H. Effect of the DNA topoisomerase II inhibitor VP-16 on illegitimate recombination in yeast chromosomes. Gene 2002, 291, 251–257. [Google Scholar] [CrossRef]
  212. Karasic, T.B.; Hei, T.K.; Ivanov, V.N. Disruption of IGF-1R signaling increases TRAIL-induced apoptosis, A new potential therapy for the treatment of melanoma. Exp. Cell Res. 2010, 316, 1994–2007. [Google Scholar] [CrossRef] [Green Version]
  213. Colon, E.; Zaman, F.; Axelson, M.; Larsson, O.; Carlsson-Skwirut, C.; Svechnikov Konstantin, V.; Soder, O. Insulin-like growth factor-I is an important antiapoptotic factor for rat leydig cells during postnatal development. Endocrinology 2007, 148, 128–139. [Google Scholar] [CrossRef]
  214. Duan, Z.; Choy, E.; Harmon, D.; Yang, C.; Ryu, K.; Schwab, J.; Mankin, H.; Hornicek, F.J. Insulin-like growth factor-I receptor tyrosine kinase inhibitor cyclolignan picropodophyllin inhibits proliferation and induces apoptosis in multidrug resistant osteosarcoma cell lines. Mol. Cancer Ther. 2009, 8, 2122–2130. [Google Scholar] [CrossRef] [Green Version]
  215. Tanaka, T.; Halicka, H.D.; Traganos, F.; Seiter, K.; Darzynkiewicz, Z. Induction of ATM activation, histone H2AX phosphorylation and apoptosis by etoposide-Relation to cell cycle phase. Cell Cycle 2007, 6, 371–376. [Google Scholar] [CrossRef]
  216. Xu, X.Y.; Wang, D.Y.; Li, Y.P.; Deyrup, S.T.; Zhang, H.J. Plant-derived lignans as potential antiviral agents: A systematic review. Phytochem. Rev. 2022, 21, 239–289. [Google Scholar] [CrossRef] [PubMed]
  217. Della Fera, A.N.; Warburton, A.; Coursey, T.L.; Khurana, S.; McBride, A.A. Persistent Human Papillomavirus Infection. Viruses 2021, 13, 321. [Google Scholar] [CrossRef] [PubMed]
  218. Saitoh, T.; Kuramochi, K.; Imai, T.; Takata, K.I.; Takehara, M.; Kobayashi, S.; Sakaguchi, K.; Sugawara, F. Podophyllotoxin directly binds a hinge domain in E2 of HPV and inhibits an E2/E7 interaction in vitro. Bioorg. Med. Chem. 2008, 16, 5815–5825. [Google Scholar] [CrossRef] [PubMed]
  219. Saleem, A.; Akhtar, M.F.; Haris, M.; Abdel-Daim, M.M. Recent updates on immunological, pharmacological, and alternative approaches to combat COVID-19. Inflammopharm 2021, 29, 1331–1346. [Google Scholar] [CrossRef] [PubMed]
  220. Ahamad, S.; Branch, S.; Harrelson, S.; Hussain, M.K.; Saquib, M.; Khan, S. Primed for global coronavirus pandemic: Emerging research and clinical outcome. Eur. J. Med. Chem. 2021, 209, 112862. [Google Scholar] [CrossRef]
  221. Patel, M.; Dominguez, E.; Sacher, D.; Desai, P.; Chandar, A.; Bromberg, M.; Caricchio, R.; Criner, G.J.; Temple University COVID-19 Research Group. Etoposide as salvage therapy for cytokine storm due to coronavirus disease 2019. Chest 2021, 159, e7–e11. [Google Scholar] [CrossRef]
  222. Rashid, H.U.; Ahmad, N.; Abdalla, M.; Khan, K.; Martines, M.A.U.; Shabana, S. Molecular docking and dynamic simulations of Cefixime, Etoposide and Nebrodenside A against the pathogenic proteins of SARS-CoV-2. J. Mol. Struct. 2022, 1247, 131296. [Google Scholar] [CrossRef]
  223. Jin, M.H.; Moon, T.C.; Quan, Z.J.; Lee, E.; Kim, Y.K.; Yang, J.H.; Suh, S.J.; Jeong, T.C.; Lee, S.H.; Kim, C.H.; et al. The naturally occurring flavolignan, deoxypodophyllotoxin, inhibits lipopolysaccharide-induced iNOS expression through the NF-kappa B activation in RAW264.7 macrophage cells. Biol. Pharm. Bull. 2008, 31, 1312–1315. [Google Scholar] [CrossRef] [Green Version]
  224. Jin, M.; Lee, E.; Yang, J.H.; Lu, Y.; Kang, S.; Chang, Y.C.; Lee, S.H.; Suh, S.J.; Kim, C.H.; Chang, H.W. Deoxypodophyllotoxin Inhibits the Expression of Intercellular Adhesion Molecule-1 Induced by Tumor Necrosis Factor-alpha in Murine Lung Epithelial Cells. Biol. Pharm. Bull. 2010, 33, 1–5. [Google Scholar] [CrossRef] [Green Version]
  225. Suh, S.J.; Kim, J.R.; Jin, U.H.; Choi, H.S.; Chang, Y.C.; Lee, Y.C.; Kim, S.H.; Lee, I.S.; Moon, T.C.; Chang, H.W.; et al. Deoxypodophyllotoxin, flavolignan, from Anthriscus sylvestris Hoffm. inhibits migration and MMP-9 via MAPK pathways in TNF-alpha-induced HASMC. Vasc. Pharmacol. 2009, 51, 13–20. [Google Scholar] [CrossRef]
  226. Petersen, C.S.; Agner, T.; Ottevanger, V.; Larsen, J.; Ravnborg, L. A single-blind study of podophyllotoxin cream 0.5% and podophyllotoxin solution 0.5% in male patients with genital warts. Genitourin Med. 1995, 71, 391–392. [Google Scholar] [CrossRef] [PubMed]
  227. Gutiérrez-Gutiérrez, F.; Puebla-Pérez, A.M.; González-Pozos, S.; Hernández-Hernández, J.M.; Pérez-Rangel, A.; Alvarez, L.P.; Tapia-Pastrana, G.; Castillo-Romero, A. Antigiardial Activity of Podophyllotoxin-Type Lignans from Bursera fagaroides var. fagaroides. Molecules 2017, 22, 799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Che, Z.; Yu, X.; Zhi, X.; Fan, L.; Yao, X.; Xu, H. Synthesis of novel 4α-(acyloxy)-2′(2′,6′)-(di)halogenopodophyllotoxin derivatives as insecticidal agents. J. Agric. Food Chem. 2013, 61, 8148–8155. [Google Scholar] [CrossRef] [PubMed]
  229. Wang, R.; Zhi, X.; Li, J.; Xu, H. Synthesis of Novel Oxime Sulfonate Derivatives of 2′(2′,6′)-(Di)chloropicropodophyllotoxins as Insecticidal Agents. J. Agric. Food Chem. 2015, 63, 6668–6674. [Google Scholar] [CrossRef]
  230. Filley, C.M.; Graff-Richard, N.R.; Lacy, J.R.; Heitner, M.A.; Earnest, M.P. Neurologic manifestations of podophyllin toxicity. Neurology 1982, 32, 308–311. [Google Scholar] [CrossRef] [PubMed]
  231. O’Mahony, S.; Keohane, C.; Jacobs, J.; O’Riordain, D.; Whelton, M. Neuropathy due to podophyllin intoxication. J. Neurol. 1990, 237, 110–112. [Google Scholar] [CrossRef] [PubMed]
  232. Kao, W.F.; Hung, D.Z.; Tsai, W.J.; Lin, K.P.; Deng, J.F. Podophyllotoxin intoxication: Toxic effect of Bajiaolian in herbal therapeutics. Hum. Exp. Toxicol. 1992, 11, 480–487. [Google Scholar] [CrossRef]
  233. Yang, W.; Zhang, Y.; Xia, D.; Xu, X. The Immunosuppression of Etoposide Deserves Attention in Chemoimmunotherapy. Am. J. Clin. Oncol. 2021, 44, 224–225. [Google Scholar] [CrossRef]
  234. Li, J.; Dai, C.X.; Sun, H.; Jin, L.; Guo, C.Y.; Cao, W.; Wu, J.; Tian, H.Y.; Luo, C.; Ye, W.C.; et al. Protective effects and mechanisms of curcumin on podophyllotoxin toxicity in vitro and in vivo. Toxicol. Appl. Pharmacol. 2012, 265, 190–199. [Google Scholar] [CrossRef]
  235. Li, J.; Sun, H.; Jin, L.; Cao, W.; Zhang, J.; Guo, C.Y.; Ding, K.; Luo, C.; Ye, W.C.; Jiang, R.W. Alleviation of podophyllotoxin toxicity using coexisting flavonoids from Dysosma versipellis. PLoS ONE 2013, 8, e72099. [Google Scholar] [CrossRef] [Green Version]
  236. Li, J.; Zhang, X.; Renata, H. Asymmetric Chemoenzymatic Synthesis of (-)-Podophyllotoxin and Related Aryltetralin Lignans. Angew. Chem. Int. Ed. Engl. 2019, 58, 11657–11660. [Google Scholar] [CrossRef] [PubMed]
  237. Wu, Y.; Zhang, H.; Zhao, Y.; Zhao, J.; Chen, J.; Li, L. A New and Efficient Strategy for the Synthesis of Podophyllotoxin and Its Analogues. Org. Lett. 2007, 9, 1199–1202. [Google Scholar] [CrossRef] [PubMed]
  238. Wu, Y.; Zhao, J.; Chen, J.; Pan, C.; Li, L.; Zhang, H. Enantioselective sequential conjugate addition-allylation reactions, a concise total synthesis of (+)-podophyllotoxin. Org. Lett. 2009, 11, 597–600. [Google Scholar] [CrossRef] [PubMed]
  239. Ting, C.P.; Maimone, T.J. C-H Bond Arylation in the Synthesis of Aryltetralin Lignans: A Short Total Synthesis of Podophyllotoxin. Angew. Chem. Int. Ed. Engl. 2014, 53, 3115–3119. [Google Scholar] [CrossRef] [PubMed]
  240. Xiao, J.; Cong, X.W.; Yang, G.Z.; Wang, Y.W.; Peng, Y. Divergent Asymmetric Syntheses of Podophyllotoxin and Related Family Members via Stereoselective Reductive Ni-Catalysis. Org. Lett. 2018, 20, 1651–1654. [Google Scholar] [CrossRef]
  241. Hajra, S.; Garai, S.; Hazra, S. Catalytic Enantioselective Synthesis of (-)-Podophyllotoxin. Org. Lett. 2017, 19, 6530–6533. [Google Scholar] [CrossRef]
  242. Anrini, M.; Jha, S. Characterization of Podophyllotoxin Yielding Cell Lines of Podophyllum hexandrum. Caryologia 2009, 62, 220–235. [Google Scholar] [CrossRef] [Green Version]
  243. Shen, S.; Tong, Y.; Luo, Y.; Huang, L.; Gao, W. Biosynthesis, total synthesis, and pharmacological activities of aryltetralin-type lignan podophyllotoxin and its derivatives. Nat. Prod. Rep. 2022, 39, 1856–1875. [Google Scholar] [CrossRef]
  244. Sedaghat, S.; Ezatzadeh, E.; Alfermann, A.W. Podophyllotoxin from suspension culture of Linum album. Nat. Prod. Res. 2008, 22, 984–989. [Google Scholar] [CrossRef]
  245. Baldi, A.; Srivastava, A.K.; Bisaria, V.S. Effect of Aeration on Production of Anticancer Lignans by Cell Suspension Cultures of Linum album. Appl. Biochem. Biotech. 2008, 151, 547–555. [Google Scholar] [CrossRef]
  246. Kusari, S.; Lamshoft, M.; Spiteller, M. Aspergillus fumigatus Fresenius, an endophytic fungus from Juniperus communis L. Horstmann as a novel source of the anticancer pro-drug deoxypodophyllotoxin. J. Appl. Microbiol. 2009, 107, 1019–1030. [Google Scholar] [CrossRef] [PubMed]
  247. Meng, Z.; Yao, T.; Zhao, W.; Li, H.; Tang, Y.J. Research progress in biosynthesis of podophyllotoxin and its derivatives. Chin. J. Biotechnol. 2021, 37, 2026–2038. [Google Scholar]
  248. Jia, K.Z.; Zhan, X.; Li, H.M.; Li, S.; Shen, Y.; Qingsheng, Q.; Zhang, Y.; Li, Y.Z.; Tang, Y.J. A novel podophyllotoxin derivative with higher anti-tumor activity produced via 4′-demethylepipodophyllotoxin biotransformation by Penicillium purpurogenum. Process Biochem. 2020, 96, 220–227. [Google Scholar] [CrossRef]
  249. Li, Y.; Li, Y.Y.; Mi, Z.Y.; Li, D.S.; Tang, Y.J. Novel biotransformation process of podophyllotoxin to produce podophyllic acid and picropodophyllotoxin by Pseudomonas aeruginosa CCTCC AB93066, part II: Process optimization. Bioresour. Technol. 2009, 100, 2271–2277. [Google Scholar] [CrossRef] [PubMed]
  250. Messiano, G.B.; Wijeratne, E.M.; Lopes, L.M.; Gunatilaka, A.A. Microbial transformations of aryltetralone and aryltetralin lignans by Cunninghamella echinulata and Beauveria bassiana. J. Nat. Prod. 2010, 73, 1933–1937. [Google Scholar] [CrossRef] [PubMed]
  251. Tang, Y.J.; Xu, X.L.; Zhong, J.J. A novel biotransformation process of 4′-demethylepipodophyllotoxin to 4′-demethylepipodophyllic acid by Bacillus fusiformis CICC 20463, Part II: Process optimization. Bioprocess. Biosyst. Eng. 2010, 33, 237–246. [Google Scholar] [CrossRef] [PubMed]
  252. Vasilev, N.P.; Julsing, M.K.; Koulman, A.; Clarkson, C.; Woerdenbag, H.J.; Ionkova, I.; Bos, R.; Jaroszewski, J.W.; Kayser, O.; Quax, W.J. Bioconversion of deoxypodophyllotoxin into epipodophyllotoxin in E. coli using human cytochrome P450 3A4. J. Biotechnol. 2006, 126, 383–393. [Google Scholar] [CrossRef] [PubMed]
  253. Julsing, M.K.; Vasilev, N.P.; Schneidman-Duhovny, D.; Muntendam, R.; Woerdenbag, H.J.; Quax, W.J.; Wolfson, H.J.; Ionkova, I.; Kayser, O. Metabolic stereoselectivity of cytochrome P450 3A4 towards deoxypodophyllotoxin, In silico predictions and experimental validation. Eur. J. Med. Chem. 2008, 43, 1171–1179. [Google Scholar] [CrossRef]
  254. Kitamura, H.; Okudela, K.; Yazawa, T.; Sato, H.; Shimoyamada, H. Cancer stem cell, Implications in cancer biology and therapy with special reference to lung cancer. Lung Cancer 2009, 66, 275–281. [Google Scholar] [CrossRef]
  255. Cheng, J.X.; Liu, B.L.; Zhang, X. How powerful is CD133 as a cancer stem cell marker in brain tumors? Cancer Treat. Rev. 2009, 35, 403–408. [Google Scholar] [CrossRef]
  256. Walcher, L.; Kistenmacher, A.K.; Suo, H.; Kitte, R.; Dluczek, S.; Strauß, A.; Blaudszun, A.R.; Yevsa, T.; Fricke, S.; Kossatz-Boehlert, U. Cancer Stem Cells-Origins and Biomarkers, Perspectives for Targeted Personalized Therapies. Front. Immunol. 2020, 11, 1280. [Google Scholar] [CrossRef] [PubMed]
  257. Bjerkvig, R.; Johansson, M.; Miletic, H.; Niclou, S.P. Cancer stem cells and angiogenesis. Semin. Cancer Biol. 2009, 19, 279–284. [Google Scholar] [CrossRef] [PubMed]
  258. Nassar, D.; Blanpain, C. Cancer Stem Cells, Basic Concepts and Therapeutic Implications. Annu. Rev. Pathol. 2016, 11, 47–76. [Google Scholar] [CrossRef] [PubMed]
  259. Salnikov, A.V.; Kusumawidjaja, G.; Rausch, V.; Bruns, H.; Gross, W.; Khamidjanov, A.; Ryschich, E.; Gebhard, M.M.; Moldenhauer, G.; Büchler, M.W.; et al. Cancer stem cell marker expression in hepatocellular carcinoma and liver metastases is not sufficient as single prognostic parameter. Cancer Lett. 2009, 275, 185–193. [Google Scholar] [CrossRef] [PubMed]
  260. Cao, H.; Xu, W.; Qian, H.; Zhu, W.; Yan, Y.; Zhou, H.; Zhang, X.; Xu, X.; Li, J.; Chen, Z.; et al. Mesenchymal stem cell-like cells derived from human gastric cancer tissues. Cancer Lett. 2009, 274, 61–71. [Google Scholar] [CrossRef]
  261. Allavena, P.; Sica, A.; Solinas, G.; Porta, C.; Mantovani, A. The inflammatory micro-environment in tumor progression. The role of tumor-associated macrophages. Crit. Rev. Oncol. Hematol. 2008, 66, 1–9. [Google Scholar] [CrossRef]
  262. Apte, R.N.; Krelin, Y.; Song, X.; Dotan, S.; Recih, E.; Elkabets, M.; Carmi, Y.; Dvorkin, T.; White, R.M.; Gayvoronsky, L.; et al. Effects of micro-environment- and malignant cell-derived interleukin-1 in carcinogenesis, tumour invasiveness and tumour-host interactions. Eur. J. Cancer 2006, 42, 751–759. [Google Scholar] [CrossRef]
  263. Liu, J.; Gao, M.; Yang, Z.; Zhao, Y.; Guo, K.; Sun, B.; Gao, Z.; Wang, L. Macrophages and Metabolic Reprograming in the Tumor Microenvironment. Front. Oncol. 2022, 12, 795159. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of natural podophyllotoxins aglycones.
Figure 1. Chemical structures of natural podophyllotoxins aglycones.
Molecules 28 00302 g001
Figure 2. Chemical structures of natural seco-podophyllotoxins.
Figure 2. Chemical structures of natural seco-podophyllotoxins.
Molecules 28 00302 g002
Figure 3. Chemical structures of natural podophyllotoxin glycosides.
Figure 3. Chemical structures of natural podophyllotoxin glycosides.
Molecules 28 00302 g003
Figure 4. Chemical structures of GL-331 and TOP-53.
Figure 4. Chemical structures of GL-331 and TOP-53.
Molecules 28 00302 g004
Figure 5. A-ring and B-ring modified podophyllotoxins.
Figure 5. A-ring and B-ring modified podophyllotoxins.
Molecules 28 00302 g005
Figure 6. C-ring modified podophyllotoxins.
Figure 6. C-ring modified podophyllotoxins.
Molecules 28 00302 g006aMolecules 28 00302 g006b
Figure 7. D-ring modified podophyllotoxins.
Figure 7. D-ring modified podophyllotoxins.
Molecules 28 00302 g007
Figure 8. E-ring modified podophyllotoxins.
Figure 8. E-ring modified podophyllotoxins.
Molecules 28 00302 g008
Figure 9. Spin labeled podophyllotoxins.
Figure 9. Spin labeled podophyllotoxins.
Molecules 28 00302 g009
Figure 10. Conjugates containing podophyllotoxins.
Figure 10. Conjugates containing podophyllotoxins.
Molecules 28 00302 g010
Figure 11. Total chemical synthesis of podophyllotoxin.
Figure 11. Total chemical synthesis of podophyllotoxin.
Molecules 28 00302 g011
Table 1. Podophyllotoxins and their plant resources.
Table 1. Podophyllotoxins and their plant resources.
PlantCompoundsBiological ActivitiesReferences
Anthriscus neglectaDeoxypodophyllotoxin 14Antitumor (induction of apoptosis) and inhbiton of CYP2C9 and CYP3A4 enzymes[57]
Anthriscus sylvestrisPicropodophyllin 3Antitumor (induction of HL-60 apoptosis)[33,37,58,59]
Deoxypodophyllotoxin 14Antitumor (inhibtions on MK-1, HeLa and B16F10 cells; induction of HL-60 apoptosis)
Deoxypicropodophyllin 15Antitumor (induction of HL-60 apoptosis)
Angeloyl podophyllotoxin 25Antitumor (induction of HL-60 apoptosis)
Nemerosin 62Antitumor (inhibition against MK-1, HeLa, and B16F10 cells)
Bursera morelensisDeoxypodophyllotoxin 14Antitumor (inhibition agains HCT-15 and SK-LU1 cells) [60]
5′-demethoxydeoxypodophyllotoxin 40Antitumor (inhibition agains HCT-15 and SK-LU1 cells)
Bursera tonkinensis4′-demethyldesoxypodophyllotoxin 17Antitumor (inhibition agains KB, Col2 and LNCaP cells)[61]
Bursehernin 63-
Isolariciresinol 70-
5-methoxy-Isolariciresinol 71-
4-demethyldesoxypodophyllotoxin-4-O-β-D-glucoside 83-
Bursera simarubaPicropolygamain 52-[62]
Bursera fagaroidesAcetyl podophyllotoxin 20Antitumor (disturbing tubulin)[63]
Peltatin-A-methyl ether 32Antitumor (disturbing tubulin)
5′-desmethoxy-β-Peltatin-A-methyl ether 38Antitumor (disturbing tubulin)
Callitris columellarisDeoxypodophyllotoxin 14Antitumor (induction of apoptosis)[64,65]
Callitris drummondiPodophyllotoxin 1-[66]
Haplophyllum cappadocicum4-deoxyisodiphyllin 37-[67,68,69]
Tuberculatin 80-
Matairesinol 64-
Justicidin A 47-
Justicidin B 48-
Diphyllin 46-
Haplomyrtoside 81-
Haplomyrtin 50-
1β-Polygamain 51-
Majidine 91-
Commiphora erlangerianaErlangerin A 53-[70,71]
Erlangerin B 54-
Erlangerin C 55Cytotoxicity in RAW 264.7 cells; antitumor (inhibtion of HeLa, EAhy926 and L929 cells)
Erlangerin D 56Cytotoxicity in RAW 264.7 cells; antitumor (inhibtion of HeLa, EAhy926 and L929 cells)
Podophyllotoxin 1Cytotoxicity in RAW 264.7 cells; antitumor (inhibtion of HeLa, EAhy926 and L929 cells)
Diphylleia sinensisDeoxypodophyllotoxin 14-[42,72,73]
Isopicropodophyllone 10-
Diphyllin 46-
Picropodophyllin 3-
Podophyllotoxone 8-
Justicidin A 47-
4′-demethylpodophyllotoxin 6-
Picropodophyllin glucoside 74-
4′-demethylpodophyllotoxin 6-
Dysosma versipellisPodophyllotoxin 1Antitumor (inhibition of LNCaP, PC-3, A549 and HT-29 cells) [74,75,76]
4′-demethyldeoxypodophyllotoxin 17Antitumor (inhibition against PC3 and LNcap-37 cells)
Dehydropodophyllotoxin 12-
Diphyllin 46-
Podophyllotoxone 8-
4′-demethyldehydropodophyllotoxin 13-
Isopicropodophyllone 10-
Isodiphyllin 36-
Picropodophyllotoxin-4-O-β-D-glucopyranosyl-(1→6)-β-D-glucopyranoside 90-
L-picropodophyllotoxin-4-O-β-D-glucopyranoside 77-
4′-demethyl podophyllotoxone 11-
Podophyllotoxone 8Antitumor (inhibition against PC3 and LNcap-37 cells)
α-peltatin 34-
β-peltatin 35-
Deoxypodophyllotoxin 14Antitumor (inhibition of LNCaP and PC-3 cells)
Podophyllotoxin-4-O-β-D-glucoside 72-
4-demethylpodophyllotoxin-7¢-O-β-D-glucopyranoside 76-
α-peltatin-5-O-β-D-glucopyranoside 84-
β-peltatin-5-O-β-D-glucopyranoside 85-
4′-demethylpodophyllotoxin 6Antitumor (inhibition of LNCaP and PC-3 cells)
Dysosma pleianthaDeoxypodophyllotoxin 14Antiviral, anticancer[77]
Podophyllotoxone 8
4′-demethylpodophyllotoxin 6
4′-demethyldesoxypodophyllotoxin 17
4′-demethyl podophyllotoxone 3
Podophyllotoxin 1
4′-demethylpodophyllotoxin 6
Diphylleia cymosa4′-demethyldesoxypodophyllotoxin 17-[43]
Diphyllin 46-
4′-demethylpodophyllotoxin 6-
Podophyllotoxin 1-
β-peltatin 35-
Diphylleia grayiPicropodophyllin 3-[43,78,79]
Deoxypodophyllotoxin 14Antitumor (inhibition of the prostate cancer cells through Akt/p53/Bax/PTEN pathway)
Diphyllin 46-
Podophyllotoxin 1Antitumor targeting the mitosis
β-apopicropodophyllin 18-
Eriope blanchetiiβ-peltatin 35-[80,81]
α-peltatin 34-
Yatein 59-
Podophyllotoxin 1-
Eriope macrostachyaβ-peltatin 35-[82]
α-peltatin 34-
Haplophyllum perforatumDiphyllin 46Antitumor (inhibition against PC3, DLD1, A549, MDCK, MDCK-MDR1 cells)[83]
Haplophyllum myrtifolium7-O-(3-methyl-2-butenyl)isodaurinol 58-[44,48]
Haplomyrtin 50-
(-)-haplomyrfolin 65-
1β-Polygamain 51-
Haplophyllum bucharicumJusticidin B 48-[84,85]
Diphyllin 46-
Haplophyllum buxbaumiiJusticidin B 48-[86,87,88]
Diphyllin 46-
(-)-Tuberculatin 80-
Mono-O-acetyldiphyllin apioside 82-
Ciliatoside A 95-
Ciliatoside B 96-
Majidine 91-
Hernandia peltataDeoxypodophyllotoxin 14-[54]
Deoxypicropodophyllin 15-
5′-methoxyyatein 61-
Bursehernin 63-
Hernandia nymphaeifoliaDeoxypodophyllotoxin 14-[54,89]
Deoxypicropodophyllin 15-
Bursehernin 63-
Yatein 59-
5′-methoxyYatein 61-
Hernandia sonoraPodophyllotoxin 1-[90]
Picropodophyllin 3-
Deoxypodophyllotoxin 14-
Hernandin 33-
Podophyllotoxin acetate 51-
5-Methoxypodophyllotoxin 26-
5-methoxypodophyllotoxin acetate 30-
(3S,4R)-3-[(S)-(acetyloxy)(3,4,5-trimethoxyphenyl)-
methyl]dihydro-4-[(7-methoxy-1,3-benzodioxol-5-yl)-
methyl]-2(3H)-Furanone 66-
Hernandia ovigera6,7-demethylenedesoxypodophyllotoxin 45Antivirus (inhibition against EBV early antigen activation)[39,49,53,91,92,93]
Deoxypicropodophyllin 15-
Podophyllotoxin 1-
Bursehernin 63Antivirus (inhibition against EBV early antigen activation)
Hernandin 33-
Dehydrodeoxypodophyllotoxin 19Antivirus (inhibition against EBV early antigen activation)
Yatein 59-
Dehydropodophyllotoxin 12Antivirus (inhibition against EBV early antigen activation)
Deoxypodophyllotoxin 14Antivirus (inhibition against EBV early antigen activation)
5-methoxy-desoxypodophyllotoxin 29-
Hyptis verticillata4′-Demethylpicropodophyllotoxin 7-[94,95,96]
4′-demethyldeoxypodophyllotoxin 17Mitosis disturbance and antifungus
β-peltatin 35Mitosis disturbance and antivirus
Dehydropodophyllotoxin 12Mitosis disturbance and antibacteria
Dehydrodeoxypodophyllotoxin 19-
Yatein 59Mitosis disturbance and antifungus
Isodeoxypodophyllotoxin 16Mitosis disturbance
Deoxypicropodophyllin 15Mitosis disturbance
β-apopicropodophyllin 18Mitosis disturbance and antifungus
Justicia ciliataJusticidin A 47 [56,97,98]
Justicidin B 49-
Cilinaphthalide A 41-
Cilinaphthalide B 42Antiplatelet
Ciliatoside A 95DNA damage and anti-inflammation (inhibited the accumulation of NO in RAW 264.7)
Ciliatoside B 96Anti-inflammation (inhibition of NO in RAW 264.7)
Diphyllin 46-
Justicia heterocarpaFuro[3′,4′:6,7]naphtho[2,3-d]-1,3-dioxol-6(5aH)-one, 5,8,8a,9-tetrahydro-5a-hydroxy-5-(6-methoxy-1,3-benzodioxol-5-yl)-, (5S,5aS,8aS)- 57-[99]
Justicia adhatodaDiphyllin 46Antitumor, antivirus (SARS-CoV2)[100]
Justicidin B 49Antiinflammatory, antiplatelet aggregation,
cytotoxicity, antiviral (SARS-CoV2), fungicidal
Justicidin A 47Antivirus (SARS-CoV2)
Podophyllotoxin 1Antitumor, antivirus (SARS-CoV2)
Justicia procumbensTuberculation 80Antitumor (breakage of plasmid), enhancement of TNF-alpha generation and antiplatelet [97,101,102,103,104,105,106]
Justicidin A 47Antiplatelet; cytotoxicity and enhancement of TNF-alpha generation
procumbenoside A 92Antitumor (breakage of plasmid) and antivirus (HIV-1)
procumbenoside B 93-
Ciliatoside A 95Antitumor (breakage of plasmid)
Ciliatoside B 96-
Justicidin C 49-
Justicidin B 48Antitumor, antiplatelet, anti-inflammation, antifungus, antivirus and antibacteria
Diphyllin 46Cytotoxic and antivirus (HIV-1)
Mono-O-acetyldiphyllin apioside 82-
Isodiphyllin 36-
Juniperus chinensisPodophyllotoxin 1-[51,107]
Yatein 59-
Juniperus sabinaepipicropodophyllotoxin 4-[108,109]
4-acetyl epipodophyllotoxin 22-
4-acetyl epipicropodophyllotoxin 23-
4-acetyl junaphtoic acid 67
Junaphtoic acid 68-
Podophyllotoxin 1Anticholinesterase, antifertility effect (inducing epididymal epithelial cell apoptosis)
Deoxypodophyllotoxin 14Anticholinesterase
3-O-demethylYatein 60-
Juniperus thuriferaPodophyllotoxone 8Analgesic and anti-inflammation[110]
Deoxypodophyllotoxin 14Analgesic and anti-inflammation
Juniperus virginianaPodophyllotoxin 1-[111,112]
Libocedrus chevalieri5-methoxy-4-epipodophyllotoxin 27Antitumor (inhibition against leukemia L1210 cells)[38]
5-Methoxypodophyllotoxin 26Antitumor (inhibition against cancer cells)
5-methoxypodophyllotoxin-4-O-β-D-glucoside 79-
Podophyllotoxin-4-O-β-D-glucoside 72-
Linum catharticumPodophyllotoxin 1-[113]
β-peltatin 35-
Podophyllotoxin-4-O-β-D-glucoside 72-
5-Methoxypodophyllotoxin 26-
5-methoxy podophyllotoxin acetate 30-
5-methoxypodophyllotoxin-4-O-β-D-glucoside 79-
Linum flavum5-Methoxypodophyllotoxin 26-[114,115,116]
5-methoxypodophyllotoxin-7-O-n-hexanoate 31-
β-peltatin 35-
α-peltatin 34-
Podophyllotoxin 1-
5-methoxypodophyllotoxin-4-O-β-D-glucoside 79-
α-peltatin-5-O-β-D-glucopyranoside 84-
β-peltatin-5-O-β-D-glucopyranoside 85-
Linum mucronatum5-Methoxypodophyllotoxin 26-[117]
β-peltatin 35-
5′-demethoxy-methoxypodophyllotoxin 39-
Podophyllotoxin 1-
Yatein 59-
Linum persicum5-Methoxypodophyllotoxin 26-[118,119]
5-methoxy podophyllotoxin acetate 30-
Podophyllotoxin 1-
β-peltatin 35-
α-peltatin 34-
Linum tauricum4′-demethyl-6-methoxypodophyllotoxin 28-[120]
Podophyllotoxin 1-
4′-Demethylpodophyllotoxin 6-
Podophyllum hexandrumPodophyllotoxin 1Antitumor (inhibition against MCF-7 cells)[121,122]
Isopicropodophyllotoxin 5-
Sinolignan A 78-
Sinolignan B 90-
Deoxypodophyllotoxin 14
Isopicropodophyllone 10-
Picropodophyllone 9-
Podophyllotoxone 8-
Picropodophyllin 3-
Deoxypicropodophyllin 15-
Dehydropodophyllotoxin 12-
Isopicropodophyllone 10-
4′-demethyl-picropodophyllotoxin 7-
3′,4′-demethylene-podophyllotoxin 43-
3′,4′-demethylene-4-demethyl-podophyllotoxin 44-
4′-demethyl-deoxypodophyllotoxin 17Antitumor (inhibition against MCF-7 cells)
4′-demethyl-podophyllotoxin 6Antitumor (inhibition against MCF-7 cells)
4′-demethyl-dehydropodophyllotoxin 13-
4-demethylpodophyllotoxin-7¢-O-β-D-glucopyranoside 76-
Podophyllotoxin-4-O-β-D-glucopyranoside 72-
4-Demethyl-deoxypodophyllotoxin-4-O-β-D-glucopyranoside 83Antitumor (inhibition against MCF-7 cells)
Picropodophyllotoxin-7′-O-β-D-glucopyranosyl-(1→6)-β-D-glucopyranoside 90-
Isopodophyllotoxin-7′-O-β-D-glucopyranosyl-(1→6)-β-D-glucopyranoside 88-
Me epipodophyllate 7′-O-β-D-glucopyranosyl-(1→6)-β-D-glucopyranoside 87-
Diphyllin 46-
Podophyllum peltatumEpipodophyllotoxin 2-[40,41,123]
Isopicropodophyllone 10-
β-peltatin 35-
α-peltatin 34-
Podophyllotoxone 8-
4′-demethyl podophyllotoxone 11-
4′-demethyldeoxypodophyllotoxin 17-
4′-demethylpodophyllotoxin 6-
Podophyllotoxin 1Antioxidant
Deoxypodophyllotoxin 14Antioxidant
4-O-β-D-glucopyranoside epipodophyllotoxin 73-
Polygala macradenia4′-demethyldeoxypodophyllotoxin 17Antitumor (inhibition against P-388 lymphocytic leukemia and human epidermoid carcinoma)[124]
Sinopodophyllum emodiPodophyllotoxin 1-[35,125,126]
isopicropodophyllone 10
Dehydropodophyllotoxin 12-
Deoxypodophyllotoxin 14-
Picropodophyllin acetate 21Antitumor (inhibition against HeLa and KB cells)
4′-acetyl-4′-demethyl-podophyllotoxin 24Antitumor (inhibition against HeLa and KB cells)
4-O-β-D-glucopyranoside 4′-demethylpicropodophyllotoxin 75Antitumor (inhibition against HeLa and KB cells)
4-O-β-D-glucopyranosyl-(1
→6)-β-D-glucopyranoside of
picropodophyllotoxin 89
-
4-O-β-D-glucopyranoside 4′-demethylepipodophyllotoxin 86-
Thujopsis dolabrataDeoxypodophyllotoxin 14Cytotoxic (inhibition against HL-60 and Caki-1 cells)[127]
Desoxypodophillic acid 69-
Desoxypicopodophyllin 15-
β-peltatin 35Cytotoxic (inhibition against HL-60 and Caki-1 cells)
β-peltatin-A-methylether 32-
Withania coagulansbispicropodophyllin glucoside 94-[55]
Note: “-” means bioactivity was not reported in the references.
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

Jin, L.; Song, Z.; Cai, F.; Ruan, L.; Jiang, R. Chemistry and Biological Activities of Naturally Occurring and Structurally Modified Podophyllotoxins. Molecules 2023, 28, 302. https://doi.org/10.3390/molecules28010302

AMA Style

Jin L, Song Z, Cai F, Ruan L, Jiang R. Chemistry and Biological Activities of Naturally Occurring and Structurally Modified Podophyllotoxins. Molecules. 2023; 28(1):302. https://doi.org/10.3390/molecules28010302

Chicago/Turabian Style

Jin, Lu, Zhijun Song, Fang Cai, Lijun Ruan, and Renwang Jiang. 2023. "Chemistry and Biological Activities of Naturally Occurring and Structurally Modified Podophyllotoxins" Molecules 28, no. 1: 302. https://doi.org/10.3390/molecules28010302

Article Metrics

Back to TopTop