Next Article in Journal
Parametric Analysis on Creep Deformation of Deep-Sea PMMA Observation Window
Next Article in Special Issue
Seaweed Calliblepharis jubata and Fucus vesiculosus Pigments: Anti-Dermatophytic Activity
Previous Article in Journal
Towards the Emergence of the Medical Metaverse: A Pilot Study on Shared Virtual Reality for Orthognathic–Surgical Planning
Previous Article in Special Issue
Inhalation with Vitamin D3 Metabolites—A Novel Strategy to Restore Vitamin D3 Deficiencies in Lung Tissue
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Flavone Hybrids and Derivatives as Bioactive Agents

by
László Hazai
,
Bernadett Zsoldos
,
Mónika Halmai
and
Péter Keglevich
*
Department of Organic Chemistry and Technology, Faculty of Chemical Technology and Biotechnology, Budapest University of Technology and Economics, Műegyetem rkp. 3, H-1111 Budapest, Hungary
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(3), 1039; https://doi.org/10.3390/app14031039
Submission received: 24 November 2023 / Revised: 20 January 2024 / Accepted: 23 January 2024 / Published: 25 January 2024

Abstract

:
Hybrid molecules can be defined as chemical entities with two or more structural domains, namely pharmacophores, having a specific biological effect. In many cases, when at least one of the components is biologically inactive, it is rather correct to call them “derivatives”, despite the fact that in the literature they are often mentioned also as hybrids. We have summarized such types of molecules, in which one of the components is mostly a real pharmacophore, i.e., flavone, which is one of the best-known natural bioactive substances. Structures, synthetic methods, medicinal indications, and more important activity data are presented.

1. Introduction

Flavones are among the widely known and biologically recognized natural compounds. Their characteristic basic structure is the phenylbenzopyrone skeleton. Depending on their structure, flavones have a different range of biological effects, such as antioxidant, antimicrobial, anticancer, antimalarial, anti-inflammatory, antiulcer, and anti-HIV effects [1,2].
Perhaps the best known of them are chrysin (1), kaempferol (2), and quercetin (3), which differ in the number and position of the hydroxyl groups (Figure 1). Chrysin (1) occurs in larger quantities in many plant species and in honey. It causes inhibition of cell proliferation through apoptosis and autophagy [3]. Kaempferol (2) and its derivatives are also found in plants, e.g., in beans, broccoli, strawberries, and tea fruits. Diets rich in fruits and vegetables can reduce the risk of coronary heart disease and cancer [4]. Quercetin (3) is found in onions, among many other plants. It causes the cancer-preventive effect through its antioxidant properties, by inhibition of various enzymes, and with modulation of signaling pathways [5].
Compounds containing two pharmacophores coupled covalently, i.e., hybrids, are described in most detail by Nepali et al. [6]. The types of hybrid molecules cover generally three possibilities (Figure 2). The two pharmacophores can be connected by a link-er chain (a), rigid or mobile, and can only be regularly substituted with one covalent bond, (b) or the two molecules can be simply merged (c) [7,8].
Hybrid molecules can also be classified according to the origin of the monomers. We distinguish between natural, synthetic, and mixed hybrids. The first two types consist of purely natural or synthetic units, while the third contains both.
There are several reviews in the literature about flavone hybrids, which primarily present a specific field of biological activity, e.g., anticancer agents [9] or esterase inhibitors [10]. On the other hand, further reviews summarize flavone hybrids formed with specific coupling components, such as 2-pyrrolidones (flavone alkaloids) and arylamines [11], with Vinca alkaloids [12] as well as 1,2,3-triazole pharmacophores [13].
Hundreds of molecules are known in the literature, in which the main component is flavone. Considering the nomenclature problems between “hybrids” and “derivatives”, we chose the solution of summarizing those compounds that were named hybrids by the cited authors themselves. Thus, in this review, flavone hybrids containing different pharmacophores are presented highlighting their biological effects.
As far as synthetic routes, flavones, like many other pharmacophores, e.g., amino acids (Section 2.3), carotenoids, and Vinca alkaloids (Section 2.4), are also natural substances, but the hybrids under discussion themselves are prepared synthetically. There are only a few hybrids of natural origin; these are mentioned in Section 2.7.
Classification of hybrids presented regarding the extraordinary diversity of the structure of the pharmacophores is presented from the point of view of importance and significance.

2. Flavone Hybrids with Different Pharmacophores

As can be seen, a lot of hybrids were synthesized, among which are the most important ones that were further investigated in different fields. These are considered the most typical structures, and examples of these characteristic hybrids are presented in Table 1 together with their biological activities.

2.1. Hybrids with 1,2,3-Triazoles and Their Derivatives

1,2,3-Triazole, as a well-known pharmacophore, and its derivatives have many positive properties, e.g., moderate dipole character, rigidity, stability shown in in vivo conditions, ability to form hydrogen bonds—which increases water solubility, as well as the possibility of binding to various biomolecules—led to the fact that its use in hybrids is very important. The ring itself does not occur in nature, although its compounds are known to have many bioactivities, such as anticancer, fungicidal, antibacterial, anti-HIV, antituberculosis, and antimalarial effects. The importance of the compounds containing the 1,2,3-triazole unit in medicinal chemistry is unquestionable, since—in contrast to other heterocycles—it is not protonated at a physiological pH due to its weak basicity [37,38].
One of the most classic flavone hybrid groups comprises compounds coupled with substituted 1,2,3-triazoles (Figure 3).
Thus, chrysin was reacted with propargyl bromide, and the propargyl derivative was subjected to a click reaction with various benzyl azide derivatives (prepared in situ) [14], resulting in the chrysin hybrids coupled with the pharmacophore in position 7. The antibacterial activity of the new compounds was investigated on 3-3 Gram-positive and Gram-negative pathogenic bacteria strains, respectively. Most of the derivatives showed moderate-to-excellent antibacterial effects.
Then, it was pointed out [15] that compound 5 (R=F, NO2) and its dimer derivative (6) (Figure 4) have moderate antiproliferative activity on several cell lines of different types of cancer. Among these, compound 6 (R=F) is the most important, with an IC50 = 0.73 µM activity on HeLa cells.
Chrysin triazole hybrids substituted on the triazole ring with an aryl group (7) (Figure 4) were synthesized with the general procedure, including the click reaction of the propargyl intermediate [16]. The compounds were investigated on several cancer cell lines and the best result was IC50 = 5.9 µM on the human gastric carcinoma cell line MGC-803. Replacing the 1,2,3-triazole ring to 1,2,4-triazole yielded no significant biological effect.
Antiproliferative and antimycobacterial flavone hybrids were synthesized containing two heterocyclic pharmacophores, isoxazole and benzimidazole [39]. The synthetic routes are presented in Figure 5. The merged isoxazole hybrid (9) was formed by the cyclization of the corresponding oxime prepared by the 4-formyl derivative (8), and the benzimidazole substituent (10) was built in with the usual procedure via reaction with o-phenylendiamine. Hybrid 11 was obtained in this case also via a click reaction.
Some of the compounds (11) showed moderate antiproliferative activity on cell line MCF-7 of breast cancer (IC50 = 14.5–19.1 µM), and hybrid 9 (R=CH3, R1=H) showed moderate antimycobacterial activity against a bovis strain.
The triazole ring is connected to the flavone via an amino group in the following hybrids (Figure 6). Amino derivative 12 was prepared via cyclization of the corresponding acylamino derivative and was then deacetylated. After an alkylating reaction with propargyl bromide, the resulting N-propargyl intermediate 13 was treated with azides and after the usual click reaction, the 14 hybrids were obtained [17]. The antiproliferative activity was tested on different human cancer cells, such as cervical, pancreatic breast, and neuroblastoma. The most promising result was against the MDA-MB-231 breast cancer cell line with GI50 ≤ 0.01 µM of compound 14 (R=4-nitrobenzyl).
The triazole ring was not only formed on the aromatic A ring of flavones, but also on the phenyl group (C ring) in position 2 (Figure 7).
Apigenin-7-methyl ether derivatives (15) were prepared employing a known procedure via a click reaction of the propargyl intermediate [40]. This type of compound especially 15 (R=2-F-phenyl) showed potent antitumor activity (IC50 = 10 µM) against ovarian carcinoma cell line SKOV-3. A high number of compounds 16 have been investigated, also using in silico methods [41]. Upon experimental tests, primarily the R=halogen-substituted compounds proved to have antibacterial and antifungal effects [18].
Finally, two more examples are presented (Figure 8). One of them belongs to the group of biologically exciting nucleoside analogs [42]. In these compounds, the triazole is linked to chrysin and the deoxythymine derivative together (17). The target compound was synthesized as of these types of hybrids, namely, a 1,3-dipolar cycloaddition reaction between the O-propargyl derivative of chrysin and 3′-azido-2′-deoxythymidine.
The second example is polyfluorinated flavones substituted on the phenyl ring by one or more 1,2,4-triazole rings (18). The rate of substitution depends on the reaction conditions and the excess of triazole. Weak fungistatic activity was observed among the prepared derivatives [43].

2.2. Hybrids with Triphenylphosphine

One of the ways to attack cancer cells is to introduce the drug into the mitochondria. For this, four main strategies were developed: the use of lipophilic cations, the use of cell-penetrating peptides, the application of nanoparticles, and physical penetration. Among them, the most commonly used method is the use of lipophilic cations [44].
Triphenylphosphonium (TPP+)-based cations have a hydrophobic surface and can easily diffuse through the inner membrane. In addition, it was shown that TPP+ cations can localize in the mitochondrial matrix, at a concentration 1000 times higher than in the cytoplasm [45]. Since cancer cells have an increased mitochondrial and plasma membrane potential compared to healthy cells, the TPP+ cation can cause selective cell death [46].
As proof of this, quercetin and triphenylphosphine hybrids were synthesized using a 4C linker chain (Figure 9). During the course of the synthesis, the specific substituted flavone (19) with protecting groups was treated with the corresponding dihalogen butane, followed by a deacetylation reaction, producing the molecule coupled with the linker (20). After an activation step with NaI, a reaction with triphenylphosphine afforded the mitochondria-targeted derivative 21 [19].
The correlation between oxidative behavior and cytotoxic effect was established by examining Jurkat cells [20].
Further derivatives were synthesized (Figure 10) and also presented their mitochondrial accumulation.
The side chain containing the triphenylphosphonium salt was also built in two different places of the polyphenol, in positions 3 and 5, respectively [47].
Moreover, compound 21 was conjugated with mitochondria-targeted self-assembled nanoparticles, and in vitro cytotoxicity to HeLa, HepG2, MCF-7, and A549 cells were investigated and showed that the nanoparticles were more effective therapeutic agents compared to quercetin [48].

2.3. Flavone Hybrids with Amino Acids

Most often, the amino acid side chain was formed in position 7 of flavones, and chrysin proved to be the most used flavone among them. Thus, several amino acids were coupled to chrysin for metabolic studies, e.g., glycine, alanine, valine, leucine, methionine, etc. [49].
Two methods were used known from peptide chemistry. In the first one, after hydrolysis of 24, hybrids were obtained. In compound 26, the amino acid was coupled to the flavone via a short linker (Figure 11).
The hybrids of 27 (Figure 12) were prepared employing similar procedures; however, at the coupling reaction with amino acids, the well-known N-hydroxybenzotriazole–EDCI system was used [21].
The most potent antiproliferative activity was obtained in the case of compound 27 (R=butane-2-yl, R1=Me, n = 1; isoleucine ester hybrid; IC50 = 3.78 µM against human gastric carcinoma MGC-803 cell lines).
Compounds with a similar structure were also prepared with a longer linker (Figure 12). Amongst these, 27 (R=2-Me-propyl, R1=H, n = 4; leucine hybrid) proved to be an effective anticancer molecule. On the breast cancer MCF-7 cell line, IC50 = 16.6 µM was obtained; moreover, the compound displayed moderate inhibition of EGFR [22].
Further derivatives of flavones combined with amino acids are compounds where the short linker is substituted with a long carbon chain alkyl group (Figure 12). Compound 28, where the amino acid component is leucine in methyl ester form (R1=Me) showed good activities on breast cancer cells MCF-7 cell line (IC50 = 32.7 µM) and on MDA-MB-231 (IC50 = 8.2 µM) [50]. In addition, compound 28 had outstanding selectivity between other cancer cells and human vascular endothelial cells.
Amino acids containing apigenin (29) and quercetin derivatives (30) were also synthesized using protected amino acids. From the apigenin hybrids, the Lys-coupled molecule proved to be the best on the biological tests, having a sedative effect on mouse brain tissues (single dose 0.4 mg/g) passing through the blood–brain barrier [51]. In the case of the quercetin–glutamic acid conjugate (IC50 = 0.14 µM on drug-resistant human MES-SA uterine sarcoma cell line), a potent Pgp-based MDR-modulating effect was found [23].
Flavone–amino acid hybrids (32) were prepared from 7-bromo derivative (31) employing the Buchwald–Hartwig reaction conditions (Figure 13). Not only were single amino acid-containing derivatives (32a) synthesized, but also dipeptide hybrids (32b).
The products were investigated on several cancer types [24], and hybrids coupled with single-amino acid esters hydrolyzed to the corresponding carboxylic acids proved to be the most effective, especially 32a (A=Val, R=H). However, it should be noted that the authors did not use a positive control in the course of the screening, which is a non-recommended practice.

2.4. Flavones Coupled with Representative Natural Components

Flavones containing a carotenoid-like side chain were synthesized from the corresponding C26-aldehyde and the phosphorous ester-substituted flavone using the Wittig–Horner–Emmons reaction (Figure 14). Antioxidant characteristics of products 33 (R=H, OH) were investigated [52].
Furthermore, the UVA photoprotective properties of compound 33 (R=OH) were also tested [53].
Important groups of natural compounds are the flavone alkaloids and Vinca alkaloids [11]. Flavone hybrids of vindoline (Figure 14) coupled in position 10 of vindoline (34, 35) were synthesized with short or longer linkers. The connection of vindoline was also built to position 17 using a 4C linker (36) via known procedures. From the new derivatives, compound 34 showed important antiproliferative activity in several cell types, and an excellent GI50 value (1.1 µM) was obtained on the melanoma LOX IMVI cell line [12].

2.5. Cyclic Saturated Amines in Flavone Derivatives

Chrysin analogs (37) with antibacterial activities were synthesized using the maximum number of 6C long spacers (Figure 15).
Morpholine, piperidine, and N-methylpiperazine were the coupling components and the key intermediate bromoalkyl chain was prepared from chrysin with dibromoalkanes [25]. The sulfonylpiperazine hybrids (37d) proved to be antioxidant and cytotoxic agents [26]. The halogen analogs (37d, R1=Cl, diCl, diF) showed anticancer effects against the SK-OVS (human ovarian cancer), HeLa (human cervical cancer), and HT-29 (human colon adenocarcinoma) cell lines; the methoxy and dimethoxy derivatives proved to be effective against the A-549 (human non-small cell lung carcinoma) and HT-29 cell lines (minimum value of IC50 = 4.67 µg/mL). At the same time, it should be mentioned that no positive control was used in the cytotoxicity assays, only literature data for gefitinib were obtained.
Several flavone hybrids substituted in position 8 with piperazine (38) and piperidine (39) were prepared (Figure 16). The key step in the synthesis was a typical Mannich reaction between flavone 40 and the piperazine derivative in the presence of paraformaldehyde, resulting in the desired product 38. Compound 38 (R=thiophen-2-ylmethyl) showed an excellent IC50 for PARP-1 (14.7 nM) and in PARP-2 (0.9 µM) inhibition; thus, this hybrid could be considered a lead molecule in cancer treatment [27]. Moreover, the above-mentioned compound exerted a selective cytotoxic effect on breast cancer gene 1 (BRCA1)-deficient cells (SK-OV-3 cells) via inhibiting PARP-1. Poly(ADP-ribose) polymerase-1 (PARP-1) is a potential target for the discovery of anticancer drugs.
The nipecotic acid–flavone hybrids (39) were characterized as anti-Alzheimer agents [54] with a similar synthetic route.

2.6. Heterocyclic Pharmacophores

Flavone hybrids were synthesized with several different heterocycles (Figure 17). In connection with this type of hybrid, only a single paper was published; there was no more special continuation of research on the corresponding structures. Thus, they are briefly presented, highlighting the most effective derivative.
Maleinimide derivatives (41), similar to some flavone alkaloids, can be classified here with antiinflammatory activity [55]. Some imidazopyrimidine hybrids (42) showed antiplasmodial activity [28].
Compound 43 containing a thiazole ring displayed antifungal activity against C. albicans with an MIC of 12.5 µg/mL [29]. Furopyrazol-substituted hybrids (44) with an antiproliferative effect were prepared in large quantities with different substituents in the aromatic ring and in the 2-phenyl group. On cell lines HEP-2 (human laryngeal carcinoma), A549 (human lung adenocarcinoma), and HeLa, IC50 = 2.5–10.0 µM values were obtained [30]. Furthermore, toxicity studies revealed that compound 44 specifically targets cancer cell lines.
1,3-Oxazines coupled in two different positions to the flavone skeleton (45a,b) exhibited excellent activities against both hormone-dependent and hormone-independent human breast cancer cells and against HeLa cells (IC50 = 14–18 µM) [56].
Flavone–triazine hybrids (46) possessed good anticancer activity on human breast cancer cells (MCF-7) [57], and acridine derivatives (47, 48) coupled via shorter and longer linkers [31,32] proved to be MAO inhibitors and anti-Alzheimer agents, respectively. From a pharmacological point of view, the hybrid containing a benzodiazepine ring (49) is a very exciting structure combination. In the case of n = 3, the compound showed significant DNA binding activity, as well as in vitro cytotoxicity [58].

2.7. Flavone–Flavone and Flavone–Coumarin Hybrids

Different analogs of amentoflavone (50, R=OH, R1=4-OH, AMF) were synthesized (Figure 18) via a convergent total synthetic route [27].
The new compounds, especially 50 (R=H, R1=4-F), proved to be PARP-1 and PARP-2 inhibitors, with IC50 = 179.2 nM and 8.5 µM activities, respectively. Nevertheless, AMF (a natural compound), isolated from Selaginella moellendorffii Hieron, has a selective PARP-1-inhibitory effect. A further natural compound is the 51 flavone–coumarin hybrid, which was isolated from a Yemenien plant Gnidia socotrana. The connection formed between positions 8 and 6″; moreover, another hybrid was also isolated with 6–8″ coupling, which is regarded as an atropisomer [59]. Several derivatives of this isolated hybrid (51) were synthesized; thus, the anti-diabetic compounds 52 were prepared via Suzuki coupling in the key step [60].
Among the flavone–coumarin hybrids, the substituted compounds 53 had a common aromatic A ring (AA′). According to this, they are merged hybrids [61]. These derivatives were synthesized from the flavone ring provided with the corresponding reactive substituents.

2.8. Aryl Derivatives of Flavones

The aryl-substituted flavones are summarized in Figure 19. Compound 54 was prepared via Suzuki cross-coupling between the corresponding 3-bromoflavone and 3,4,5-trimethoxybenzeneboronic acid under usual reaction conditions and using the required protecting groups. Derivative 55 was obtained via Knoevenagel condensation of the 2,3-dihydroflavone (i.e., the appropriate flavanone) with 3,4,5-trimethoxybenzaldehyde in the presence of a piperidine base. Compounds 54 and 55, together with some similar derivatives, presented moderate tubulin polymerization-inhibitory activity [62].
Aryl derivatives coupled with the 2-phenyl group via an amide bond (56) were synthesized using linear synthetic methods. The amide group was formed by the reactions of the corresponding carboxylic acids and amines [63,64,65]. Compounds 56a,b,c,d showed potent anticancer activity in several cancer cells, and proved to be potential lead molecules for further investigations [66]. It is important to emphasize that compound 56b,d exhibited IC50 values above 100 µM in two non-tumor cell lines (L132 (human lung epithelial cell) and IOSE-80PC (human ovarian epithelial cell)) using MTT cells viability assay. This suggests that these derivatives are selective for cancer cells.
Further effects were also obtained for 56e,f,g (where g is the same as d), showing a promising possibility for the development of anti-inflammatory agents on the base of biological investigations of proinflammatory mediators [33]. However, it should be noted that no positive control was used during the determination of the IC50 values in connection with PGE2 production inhibition.
In the course of changing the methoxy substituents to methyl or halogen, especially chloro atom, the hit compounds obtained (56h,i) proved to have a potential inhibitory effect against the STE20/GCK-IV kinase family [63].
Molecule 57 was selected from the many similar compounds [64] as the most potent derivative.
Arylamino flavone 55 exhibited nanomolar antiproliferative activity in the MCF7 cell line (breast cancer, GI50 = 30 nM) and the HCT-15 cell line (colon cancer, GI50 = 60 nM).

2.9. Miscellaneous

Recently, a review was published presenting only very few hybrids, focusing more on flavone derivatives and especially the detailed synthetic procedures together with biological activities [65].
Finally, without any detail, some flavone hybrids are presented that cannot be classified into any of the above groups and for which no further investigation data were found. Although in some cases, their biological effects are important, these compounds did not form a large family of further derivatives.
The following figure (Figure 20) shows some typical hybrid skeletons which are also worth mentioning among the miscellaneous structures.
Flavones with a thiosemicarbazone side chain showed antimicrobial activity [67]. In addition, flavones bearing a hydroxamic acid across an alkyl chain linker showed anti-tumor activity against breast cancer [68]. Stilbene coupled to flavones in position 7 (58) exhibits potential antimalarial effects (IC50 = 5.07 μM) [69]. Flavone–cyanoacetamide adducts [70] possess strong AChE-binding affinity, allowing for the possibility of innovative drug development for AD. Flavones, thioflavones, and selenoflavones containing a naphthyl group in position 2 showed MDR-reversing activity, along with with antimicrobial and cytotoxic effects [71]. Some antiproliferative naphthoflavones were also synthesized [34,72]. In fact, these compounds are flavones condensed with a benzene ring in positions 5 and 6, e.g., compound 59 as a xanthine oxidase inhibitor (IC50 = 0.62 μM) [34]. Naphthoquinone, as a substituent, is also attached to a flavone [73]. These types of compounds were prepared via a rhodium(III)-catalyzed direct oxidative cross-coupling reaction. In the anticancer flavone–estradiol merged hybrids (60), there is a common aromatic A-ring of the flavone and the steroid in the molecule [35]. Desmoschinensisflavones A and B, having a flavonebenzylbenzoate hybrid structural framework, were isolated from Desmos chinensis Lour [74]. Some interesting flavone esters are also known; quercetin was esterified with aspirin in position 7 [75] and showed moderate anticancer activities, while chrysin was coupled with linolenic and linoleic acid on 7-OH [76], showing a competitive inhibitory effect on mushroom tyrosinase activity. Kaempferol connected with carbohydrate derivatives is known to have antibacterial effects (MIC = 0.05 µg/mL) [36]. Further chrysin derivatives were synthetized containing N,N-dimethylethylenediamine, including its carbamate (61) [77], which showed important antiproliferative activity, especially in the case of melanoma. Moreover, new hybrids were prepared by forming a 7-(3-amino-2-hydroxy-propyloxy) side chain possessing anticancer activity against prostate cancer cells [78].

3. Scope and Limitations

The potential disadvantages of hybrids are well known, i.e., high molar mass, risk of off-target toxicity, new and unexpected target recognition, and increased risk of poor solubility [79]. These compounds usually do not comply with Lipinski’s and Veber’s rules [80]. However, it is very important that some of the derivatives of flavonoids with these activities were revealed to be more potent than parent compounds or even commercially available drugs, using in vitro assays. Therefore, future exploitation of the mechanism of action underlying most potent compounds may contribute to the discovery of innovative chemotherapeutic agents.
Nevertheless, a hybridization protocol has emerged as a powerful asset for the development of more potent and less toxic drugs that exhibit increased specificity compared to conventional mono-therapy. There is substantial potential in refining and exploring novel designs, utilizing linker structures, and meticulously choosing partners and functional entities. These endeavors offer the prospect of discovering new drug candidates that could outperform existing anti-infective and anticancer medications in terms of safety and efficacy. Notably, numerous hybrid molecules have advanced through both early and late phases of clinical development, highlighting the promising role of hybridization as an innovative approach to drug discovery.

4. Conclusions

In this paper, a complete overview of the hybrids of flavones was outlined. It should be mentioned that in many cases, instead of the term “hybrid”, “derivative” seemed to be better. The characteristic families of different structures were presented, with the most typical synthetic routes; however, our most important goal was to show the biological effects. Studying all the results, it can be seen that among the effects, anticancer activity dominates in most cases. Moreover, among the compounds prepared are the most relevant structures having mitochondria-targeted characteristics and possessing MDR-modulating effects. From the evaluation of the results, it is clear how important it is that at least one of the components is a natural compound in hybrids.
Interestingly, most flavone hybrids show antiproliferative effects. Considering that cancer is still an unsolved medicinal problem all over the world, it is very important to enhance the potential role of hybrids in the treatment of cancer diseases.
Moreover, it can also be seen that the research on hybrids represents one of the possible ways of evaluating new anticancer agents.

Author Contributions

Conceptualization, L.H.; writing—original draft preparation, L.H.; writing—review and editing, P.K., B.Z. and M.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been connected to project no. RRF-2.3.1-21-2022-00015 with support provided by the European Union (Széchenyi Plan Plus, National Laboratory Program, PharmaLab).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Beecher, G.R. Overview of dietary flavonoids: Nomenclature, occurrence and intake. J. Nutr. 2003, 133, 3248–3254. [Google Scholar] [CrossRef] [PubMed]
  2. Ren, W.; Qiao, Z.; Wang, H.; Zhu, L.; Zhang, L. Flavonoids: Promising anticancer agents. Med. Res. Rev. 2003, 23, 519–534. [Google Scholar] [CrossRef] [PubMed]
  3. Mehdi, S.H.; Nafees, S.; Zafaryab, M.; Khan, M.A.; Rizvi, M.M.A. Chrysin: A Promising Anticancer Agent its Current Trends and Future Perspectives. Eur. J. Exp. Biol. 2018, 8, 3–16. [Google Scholar] [CrossRef]
  4. Mei, Q.; Wang, C.; Yuan, W.; Zhang, G. Selective methylation of kaempferol via benzylation and deacetylation of kaempferol acetates. Beilstein J. Org. Chem. 2015, 11, 188–193. [Google Scholar] [CrossRef]
  5. Murakami, A.; Ashida, H.; Terao, J. Multitargeted cancer prevention by quercetin. Cancer Lett. 2008, 269, 315–325. [Google Scholar] [CrossRef] [PubMed]
  6. Nepali, K.; Sharma, S.; Sharma, M.; Bedi, P.M.S.; Dhar, K.L. Rational approaches, design strategies, structure activity relationship and mechanistic insights for anticancer hybrids. Eur. J. Med. Chem. 2014, 77, 422–487. [Google Scholar] [CrossRef] [PubMed]
  7. Meunier, B. Hybrid molecules with a dual mode of action: Dream or reality? Acc. Chem. Res. 2008, 41, 69–77. [Google Scholar] [CrossRef]
  8. Pawelczyk, A.; Sowa-Kasprzak, K.; Olender, D.; Zaprutko, L. Molecular consortia—Various structural and synthetic concepts for more effective therapeutics synthesis. Int. J. Mol. Sci. 2018, 19, 1104. [Google Scholar] [CrossRef]
  9. Spatafora, C.; Tringali, C. Natural-derived Polyphenols as Potential Anticancer Agents. Anti-Cancer Agents Med. Chem. 2012, 12, 902–918. [Google Scholar] [CrossRef]
  10. Singh, H.; Singh, J.V.; Kaur, N.; Sanduja, M.; Singh, G.; Singh, B.; Preet, M.; Sharma, S. Rational Approaches, Design Strategies, Structure Activity Relationship and Mechanistic Insights for Esterase Inhibitors. Mini-Rev. Med. Chem. 2018, 18, 837–894. [Google Scholar] [CrossRef]
  11. Mayer, S.; Keglevich, A.; Für Sepsey, C.; Bölcskei, H.; Ilkei, V.; Keglevich, P.; Hazai, L. Results in Chemistry of Natural Organic Compounds. Synthesis of New Anticancer Vinca Alkaloids and Flavone Alkaloids. Chemistry 2020, 2, 46. [Google Scholar] [CrossRef]
  12. Mayer, S.; Nagy, N.; Keglevich, P.; Szigetvári, Á.; Dékány, M.; Szántay, C., Jr.; Hazai, L. Synthesis of Novel Vindoline-Chrysin Hybrids. Chem. Biodivers. 2021, 18, e202100725. [Google Scholar] [CrossRef]
  13. Pereira, D.; Pinto, M.; Correia-da-Silva, M.; Cidade, H. Recent Advances in Bioactive Flavonoid Hybrids Linked by 1,2,3-Triazole Ring Obtained by Click Chemistry. Molecules 2021, 27, 230. [Google Scholar] [CrossRef]
  14. Li, X.; Cai, Y.; Yang, F.; Meng, Q. Synthesis and molecular docking studies of chrysin derivatives as antibacterial agents. Med. Chem. Res. 2017, 26, 2225–2234. [Google Scholar] [CrossRef]
  15. Németh-Rieder, A.; Keglevich, P.; Hunyadi, A.; Latif, A.D.; Zupkó, I.; Hazai, L. Synthesis and In Vitro Anticancer Evaluation of Flavone-1,2,3-Triazole Hybrids. Molecules 2023, 28, 626. [Google Scholar] [CrossRef] [PubMed]
  16. Luan, T.; Quan, Z.; Fang, Y.; Yang, H. Design, Synthesis and Antiproliferative Activity of Chrysin Derivatives Bearing Triazole Moieties. Chin. J. Org. Chem. 2020, 40, 440–446. [Google Scholar] [CrossRef]
  17. Sowjanya, T.; Rao, Y.J.; Murthy, N.Y.S. Synthesis and Antiproliferative Activity of New 1,2,3-Triazole/Flavone Hybrid Heterocycles Against Human Cancer Cell Lines. Russ. J. Gen. Chem. 2017, 87, 1864–1871. [Google Scholar] [CrossRef]
  18. Kant, R.; Kumar, D.; Agarwal, D.; Gupta, R.D.; Tilak, R.; Awasthi, S.K.; Agarwal, A. Synthesis of newer 1,2,3-triazole linked chalcone and flavone hybrid compounds and evaluation of their antimicrobial and cytotoxic activities. Eur. J. Med. Chem. 2016, 113, 34–49. [Google Scholar] [CrossRef] [PubMed]
  19. Mattarei, A.; Biasutto, L.; Rastrelli, F.; Garbisa, S.; Marotta, E.; Zoratti, M.; Paradisi, C. Regioselective O-derivatization of quercetin via ester intermediates. An improved synthesis of Rhamnetin and development of a new mitochondriatropic derivative. Molecules 2010, 15, 4722–4736. [Google Scholar] [CrossRef] [PubMed]
  20. Mattarei, A.; Sassi, N.; Durante, C.; Biasutto, L.; Sandoná, G.; Marotta, E.; Garbisa, S.; Gennaro, A.; Paradisi, C.; Zoratti, M. Redox properties and cytotoxicity of synthetic isomeric mitochondriatropic derivatives of the natural polyphenol quercetin. Eur. J. Org. Chem. 2011, 2011, 5577–5586. [Google Scholar] [CrossRef]
  21. Song, X.; Liu, Y.; He, J.; Zheng, X.; Lei, X.; Jiang, G.; Zjan, L. Synthesis of novel amino acid derivatives containing chrysin as anti-tumor agents against human gastric carcinoma MGC-803 cells. Med. Chem. Res. 2015, 24, 1789–1798. [Google Scholar] [CrossRef]
  22. Liu, Y.-M.; Li, Y.; Liu, R.-F.; Xiao, J.; Zhou, B.-N.; Zhang, Q.-Z.; Song, J.-X. Synthesis, characterization and preliminary biological evaluation of chrysin amino acid derivatives that induce apoptosis and EGFR downregulation. J. Asian Nat. Prod. Res. 2021, 23, 39–54. [Google Scholar] [CrossRef]
  23. Kim, M.K.; Choo, H.; Chong, Y. Water-Soluble and Cleavable Quercetin−Amino Acid Conjugates as Safe Modulators for P-Glycoprotein-Based Multidrug Resistance. J. Med. Chem. 2014, 57, 7216–7233. [Google Scholar] [CrossRef]
  24. Pajtás, D.; Kónya, K.; Kiss-Szikszai, A.; Dzubák, B.; Pethő, Z.; Varga, Z.; Panyi, G.; Patonay, T. Optimization of the Synthesis of Flavone−Amino Acid and Flavone−Dipeptide Hybrids via Buchwald−Hartwig Reaction. J. Org. Chem. 2017, 82, 4578–4587. [Google Scholar] [CrossRef]
  25. Babu, K.S.; Babu, T.H.; Srinivas, P.V.; Kishore, K.H.; Murthy, U.S.N.; Rao, J.M. Synthesis and biological evaluation of novel C(7) modified chrysin analogues as antibacterial agents. Bioorg. Med. Chem. Lett. 2006, 16, 221–224. [Google Scholar] [CrossRef]
  26. Patel, R.V.; Mistry, B.M.; Syed, R.; Parekh, N.M.; Shin, H.-S. Sulfonylpiperazines based on a flavone as antioxidant and cytotoxic agents. Arch. Pharm. Chem. Life Sci. 2019, 352, e1900051. [Google Scholar] [CrossRef]
  27. Long, H.; Hu, X.; Wang, B.; Wang, Q.; Wang, R.; Liu, S.; Xiong, F.; Jiang, Z.; Zhang, X.-Q.; Ye, W.-C.; et al. Discovery of Novel Apigenin−Piperazine Hybrids as Potent and Selective Poly (ADP-Ribose) Polymerase-1 (PARP-1) Inhibitors for the Treatment of Cancer. J. Med. Chem. 2021, 64, 12089–12108. [Google Scholar] [CrossRef]
  28. Raiguru, B.P.; Panda, J.; Mohapatra, S.; Nayak, S. Palladium-catalyzed facile synthesis of imidazo [1,2-a]pyridine-flavone hybrids and evaluation of their antiplasmodial activity. J. Mol. Struct. 2023, 1294, 136282. [Google Scholar] [CrossRef]
  29. Sivakumar, K.K.; Kumar, A.N.; Munirathinam, M.; Ponnilavarasan, I.; Preedeep, A.; Priya, A.S.M. Design, conventional and microwave assisted synthesis of molecular hybrids of thiazol-2-amine containing flavone/chalcone moieties possessing antimicrobial properties. Eur. J. Biomed. Pharm. Sci. 2018, 5, 246–256. [Google Scholar]
  30. Tangeti, V.S.; Vasundhara, D.; Satyanarayana, K.V.V.V.; Kumar, K.S.P. Synthesis and Antiproliferative Activity of Some Dihydro-1H-furo [2,3-c]pyrazole-Flavone Hybrids. Asian J. Chem. 2017, 29, 1525–1532. [Google Scholar] [CrossRef]
  31. Remya, R.S.; Ramalakshmi, N.; Nalini, C.N.; Amuthalakshmi, S. Design, synthesis, and in vitro evaluation of novel acridine derivatives as monoamine oxidase inhibitors. Rasayan J. Chem. 2022, 15, 2318–2325. [Google Scholar] [CrossRef]
  32. Liao, S.; Deng, H.; Huang, S.; Yang, J.; Wang, S.; Yin, B.; Zheng, T.; Zhang, D.; Liu, J.; Gao, G.; et al. Design, synthesis and evaluation of novel 5,6,7-trimethoxyflavone–6-chlorotacrine hybrids as potential multifunctional agents for the treatment of Alzheimer’s disease. Bioorg. Med. Chem. Lett. 2015, 25, 1541–1545. [Google Scholar] [CrossRef]
  33. Hassan, A.H.E.; Yoo, S.Y.; Lee, K.W.; Yoon, Y.M.; Ryu, H.W.; Jeong, Y.; Shin, J.-S.; Kang, S.-Y.; Kim, S.-Y.; Lee, H.-H.; et al. Repurposing mosloflavone/5,6,7-trimethoxyflavone-resveratrol hybrids: Discovery of novel p38-α MAPK inhibitors as potent interceptors of macrophage-dependent production of proinflammatory mediators. Eur. J. Med. Chem. 2019, 180, 253–267. [Google Scholar] [CrossRef]
  34. Singh, H.; Sharma, S.; Ojha, R.; Gupta, M.K.; Nepali, K.; Bedi, P.M.S. Synthesis and evaluation of naphthoflavones as a new class of non purine xanthine oxidase inhibitors. Bioorg. Med. Chem. Lett. 2014, 24, 4192–4197. [Google Scholar] [CrossRef]
  35. Molnár, B.; Gopisetty, M.K.; Nagy, F.I.; Adamecz, D.I.; Kása, Z.; Kiricsi, M.; Frank, É. Efficient access to domain-integrated estradiol-flavone hybrids via the corresponding chalcones and their in vitro anticancer potential. Steroids 2022, 187, 109099. [Google Scholar] [CrossRef]
  36. Lv, X.H.; Liu, H.; Ren, Z.L.; Wang, W.; Tang, F.; Cao, H.Q. Design, synthesis and biological evaluation of novel flavone Mannich base derivatives as potential antibacterial agents. Mol. Divers. 2019, 23, 299–306. [Google Scholar] [CrossRef]
  37. Agalave, S.G.; Maujan, S.R.; Pore, V.S. Click Chemistry: 1,2,3-Triazoles as Pharmacophores. Chem. Asian J. 2011, 6, 2696–2718. [Google Scholar] [CrossRef]
  38. Kolb, H.C.; Sharpless, K.B. The growing impact of click chemistry on drug discovery. Drug Discov. Today 2003, 8, 1128–1137. [Google Scholar] [CrossRef]
  39. Rao, Y.J.; Sowjanya, T.; Thirupathi, G.; Murthy, N.Y.S.; Kotapalli, S.S. Synthesis and biological evaluation of novel flavone/triazole/benzimidazole hybrids and flavone/isoxazole-annulated heterocycles as antiproliferative and antimycobacterial agents. Mol. Divers. 2018, 22, 803–814. [Google Scholar] [CrossRef]
  40. Qi, Y.; Ding, Z.; Yao, Y.; Ma, D.; Ren, F.; Yang, H.; Chen, A. Novel triazole analogs of apigenin-7-methyl ether exhibit potent antitumor activity against ovarian carcinoma cells via the induction of mitochondrial-mediated apoptosis. Exp. Ther. Med. 2019, 17, 1670–1676. [Google Scholar] [CrossRef]
  41. Thillainayagam, M.; Malathi, K.; Ramaiah, S. In-Silico molecular docking and simulation studies on novel chalcone and flavone hybrid derivatives with 1, 2, 3—Triazole linkage as vital inhibitors of Plasmodium falciparum Dihydroorotate Dehydrogenase. J. Biomol. Struct. Dyn. 2018, 36, 3993–4009. [Google Scholar] [CrossRef] [PubMed]
  42. Yuan, J.-W.; Qu, L.-B.; Chen, X.-L.; Qu, Z.-B.; Liu, X.-Q.; Ke, D.-D. An Efficient Synthesis of Mono and Bis-1,2,3-triazole AZT Derivatives via Copper(I)-catalyzed Cycloaddition. J. Chin. Chem. Soc. 2011, 58, 24–30. [Google Scholar] [CrossRef]
  43. Panova, N.A.; Shcherbakov, K.V.; Zhilina, E.F.; Burgart, Y.V.; Saloutin, V.I. Synthesis of Mono- and Polyazole Hybrids Based on Polyfluoroflavones. Molecules 2023, 28, 869. [Google Scholar] [CrossRef]
  44. Wang, J.; Li, J.; Xiao, Y.; Fu, B.; Qin, Z. TPP-based mitocans: A potent strategy for anticancer drug design. RSC Med. Chem. 2020, 11, 858–875. [Google Scholar] [CrossRef]
  45. Gardner, Z.S.; Schumacher, T.J.; Ronayne, C.T.; Kumpati, G.P.; Williams, M.J.; Yoshimura, A.; Mereddy, V.R. Synthesis and biological evaluation of novel 2-alkoxycarbonylallylester phosphonium derivatives as potential anticancer agents. Bioorg. Med. Chem. Lett. 2021, 45, 128136. [Google Scholar] [CrossRef]
  46. Tsepaeva, O.V.; Nemtarev, A.V.; Abdullin, T.I.; Grigor’eva, L.R.; Kuznetsova, E.V.; Akhmadishina, R.A.; Mironov, V.F. Design, Synthesis, and Cancer Cell Growth Inhibitory Activity of Triphenylphosphonium Derivatives of the Triterpenoid Betulin. J. Nat. Prod. 2017, 80, 2232–2239. [Google Scholar] [CrossRef]
  47. Biasutto, L.; Mattarei, A.; Paradisi, C. Synthesis and testing of novel isomeric mitochondriatropic derivatives of resveratrol and quercetin. In Mitochondrial Medicine, Volume II, Manipulating Mitochondrial Function; Methods in Molecular Biology, Volume 1265; Weissig, W., Edeas, M., Eds.; Springer: New York, NY, USA, 2015; Chapter 13; pp. 161–179. [Google Scholar] [CrossRef]
  48. Xing, L.; Lyu, J.-Y.; Yang, Y.; Cui, P.-F.; Gu, L.-Q.; Qiao, J.-B.; He, Y.-J.; Zhang, T.-Q.; Sun, M.; Lu, J.-J.; et al. pH-Responsive de-PEGylated nanoparticles based on triphenylphosphine-quercetin self-assemblies for mitochondria-targeted cancer therapy. Chem. Commun. 2017, 53, 8790–8793. [Google Scholar] [CrossRef]
  49. Veselovskaya, M.V.; Garazd, M.M.; Ogorodniichuk, A.S.; Garazd, Y.L.; Khilya, V.P. Synthesis of amino-acid derivatives of chrysin. Chem. Nat. Compd. 2008, 44, 704–711. [Google Scholar] [CrossRef]
  50. Li, Y.; Zhang, Q.; Hr, J.; Yu, W.; Xiao, J.; Guo, Y.; Zhu, X.; Liu, Y. Synthesis and biological evaluation of amino acid derivatives containing chrysin that induce apoptosis. Nat. Prod. Res. 2019, 35, 529–538. [Google Scholar] [CrossRef] [PubMed]
  51. Wong, T.-Y.; Tsai, M.-S.; Hsu, L.-C.; Lin, S.-W.; Liang, P.-H. Traversal of the Blood−Brain Barrier by Cleavable l-Lysine Conjugates of Apigenin. J. Agric. Food Chem. 2018, 66, 8124–8131. [Google Scholar] [CrossRef] [PubMed]
  52. Beutner, S.; Frixel, S.; Ernst, H.; Hoffmann, T.; Hernandez-Blanco, I.; Hundsdoerfer, C.; Kiesendahl, K.; Kock, S.; Martin, H.-D.; Mayer, B.; et al. Carotenylflavonoids, a novel group of potent, dual-functional antioxidants. Arkivoc 2007, 8, 279–295. [Google Scholar] [CrossRef]
  53. Hundsdörfer, C.; Stahl, W.; Müller, T.J.J.; De Spirt, S. UVA Photoprotective Properties of an Artificial Carotenylflavonoid Hybrid Molecule. Chem. Res. Toxicol. 2012, 25, 1692–1698. [Google Scholar] [CrossRef] [PubMed]
  54. DhunMati, K.; Nalin, C.N.; Ramalakshmi, N.; Niraimathi, V.; Amuthalakshmi, S. Synthesis, Molecular Docking and Invitro Evaluation of Nipecotic Acid-Flavone Hybrids as Anti Alzheimer Agents—A Multi Target Directed Ligand Approach. J. Pharm. Neg. Res. 2022, 13, 229–238. [Google Scholar]
  55. Zhang, Y.; Xu, Z.; Zhan, L.; Gao, Y.; Zheng, B.; Zhou, Y.; Sheng, Y.; Liang, G.; Song, Z. Design, synthesis and biological evaluation of novel chromone-maleimide hybrids as potent anti-inflammatory agents against LPS-induced acute lung injury. Bioorg. Chem. 2022, 128, 160049. [Google Scholar] [CrossRef] [PubMed]
  56. Kumar, S.; Verma, N.; Kumar, N.; Patel, A.; Roy, P.; Pruthi, V.; Ahmed, N. Design, Synthesis, Molecular Docking and Biological Studies of Novel Phytoestrogen-Tanaproget Hybrids. Synth. Commun. 2016, 46, 460–474. [Google Scholar] [CrossRef]
  57. Abu-Aisheh, M.N.; Mustafa, M.S.; Mubarak, M.S.; El-Abadelah, M.M.; Voelter, W. Synthesis of Some 1-(Flavone-7-yl)-4,5-dihydro-1,2,4-triazin-6(1H)-ones and Related Congeners. Lett. Org. Chem. 2012, 9, 465–473. [Google Scholar] [CrossRef]
  58. Kamal, A.; Ramu, R.; Khanna, G.B.R.; Saxena, A.K.; ShanMugavel, M.; Pandita, R.M. Synthesis and evaluation of new pyrrolo [2,1-c][1,4]benzodiazepine hybrids linked to a flavone moiety. Arkivoc 2005, 2005, 83–91. [Google Scholar] [CrossRef]
  59. Franke, K.; Porzel, A.; Schmidt, J. Flavone-coumarin hybrids from Gnidia socotrana. Phytochemistry 2002, 61, 873–878. [Google Scholar] [CrossRef]
  60. Pan, G.; Zhao, L.; Xiao, N.; Yang, K.; Ma, Y.; Zhao, X.; Fan, Z.; Zhang, Y.; Yao, Q.; Yu, P. Total synthesis of 8-(6″-umbelliferyl)-apigenin and its analogs as anti-diabetic reagents. Eur. J. Med. Chem. 2016, 122, 674–683. [Google Scholar] [CrossRef]
  61. Rao, H.S.P.; Tangeti, V.S. Synthesis of 3-Aroylcoumarin-Flavone Hybrids. Lett. Org. Chem. 2012, 9, 218–220. [Google Scholar] [CrossRef]
  62. Quintin, J.; Roullier, C.; Thoret, S.; Lewin, G. Synthesis and anti-tubulin evaluation of chromone-based analogues of combretastatins. Tetrahedron 2006, 62, 4038–4051. [Google Scholar] [CrossRef]
  63. Hassan, A.H.E.; Lee, K.-T.; Lee, Y.S. Flavone-based arylamides as potential anticancers: Design, synthesis and in vitro cell-based/cell-free evaluations. Eur. J. Med. Chem. 2020, 187, 111965. [Google Scholar] [CrossRef] [PubMed]
  64. Mayer, S.; Keglevich, P.; Ábrányi-Balogh, P.; Szigetvári, Á.; Dékány, M.; Szántay, C., Jr.; Hazai, L. Synthesis and in vitro evaluation of novel chrysin and 7-aminochrysin derivatives. Molecules 2020, 25, 888. [Google Scholar] [CrossRef] [PubMed]
  65. Leonte, D.; Ungureanu, D.; Zaharia, V. Flavones and Related Compounds: Synthesis and Biological Activity. Molecules 2023, 28, 6528. [Google Scholar] [CrossRef] [PubMed]
  66. Hassan, A.H.E.; Choi, E.; Yoon, Y.M.; Lee, K.W.; Yoo, S.Y.; Cho, M.C.; Yang, J.S.; Kim, H.I.; Hong, J.Y.; Shin, J.-S.; et al. Natural products hybrids: 3,5,4′-Trimethoxystilbene-5,6,7-trimethoxyflavone chimeric analogs as potential cytotoxic agents against diverse human cancer cells. Eur. J. Med. Chem. 2019, 161, 559–580. [Google Scholar] [CrossRef] [PubMed]
  67. Singh, G.; Malhotra, R. Synthesis and antimicrobial activity of thiosemicarbazide induced hydrazone of 4-oxo-4H-chromene-3-carbaldehyde. AIP Conf. Proc. 2017, 1860, 020064. [Google Scholar] [CrossRef]
  68. Wei, M.; Xie, M.; Zhang, Z.; Wei, Y.; Zhang, J.; Pan, H.; Li, B.; Wang, J.; Song, Y.; Chong, C.; et al. Design and synthesis of novel flavone-based histone deacetylase inhibitors antagonizing activation of STAT3 in breast cancer. Eur. J. Med. Chem. 2020, 206, 112677. [Google Scholar] [CrossRef] [PubMed]
  69. Raiguru, B.P.; Mohapatra, S.; Nayak, S.; Sahal, D.; Yadav, M.; Acharya, B.N. Flavone-stilbene hybrids: Synthesis and evaluation as potential antimalarial agents. Eur. J. Med. Chem. Rep. 2022, 4, 100029. [Google Scholar] [CrossRef]
  70. Basha, S.J.; Mohan, P.; Yeggoni, D.P.; Babu, Z.R.; Kumar, P.B.; Rao, A.D.; Subramaynam, R.; Damu, A.G. New Flavone-Cyanoacetamide Hybrids with a Combination of Cholinergic, Antioxidant, Modulation of β-Amyloid Aggregation, and Neuroprotection Properties as Innovative Multifunctional Therapeutic Candidates for Alzheimer’s Disease and Unraveling Their Mechanism of Action with Acetylcholinesterase. Mol. Pharm. 2018, 15, 2206–2223. [Google Scholar] [CrossRef]
  71. Marc, M.A.; Kincses, A.; Rácz, B.; Nasim, M.J.; Sarfraz, M.; Lázaro-Milla, C.; Dominguez-Álvarez, E.; Jacob, C.; Spengler, G.; Almendros, P. Antimicrobial, Anticancer and Multidrug-Resistant Reversing Activity of Novel Oxygen-, Sulfur- and Selenoflavones and Bioisosteric Analogues. Pharmaceuticals 2020, 13, 453. [Google Scholar] [CrossRef]
  72. Kumar, D.; Singh, O.; Nepali, K.; Bedi, P.M.S.; Qayum, A.; Singh, S.; Jain, S.K. Naphthoflavones as antiproliferative agents: Design, synthesis and biological evaluation. Anti-Cancer Agents Med. Chem. 2016, 16, 881–890. [Google Scholar] [CrossRef]
  73. Samanta, R.; Narayan, R.; Antonchick, A.P. Rhodium(III)-Catalyzed Direct Oxidative Cross Coupling at the C5 Position of Chromones with Alkenes. Org. Lett. 2012, 23, 6108–6111. [Google Scholar] [CrossRef]
  74. Polbuppha, I.; Suthiphasipl, V.; Maneerat, T.; Charoensup, R.; Limtharakul, T.; Cheenpracha, S.; Pyne, S.G.; Laphookhieo, S. Desmoschinensisflavones A and B, two rare flavones having a hybrid benzyl benzoate esterflavone structural framework from Desmos chinensis Lour. RSC Adv. 2020, 10, 45076–45080. [Google Scholar] [CrossRef]
  75. Lu, C.; Huang, F.; Li, Z.; Ma, J.; Li, H.; Fang, L. Synthesis and Bioactivity of Quercetin Aspirinates. Bull. Korean Chem. Soc. 2014, 35, 518–520. [Google Scholar] [CrossRef]
  76. Jamali, Z.; Behbahani, G.R.R.; Zare, K.; Gheibi, N. Effect of chrysin omega—3 and 6 fatty acid esters on mushroom tyrosinase activity, stability, and structure. J. Food Biochem. 2018, 43, e12728. [Google Scholar] [CrossRef]
  77. Mayer, S.; Herr, D.M.; Nagy, N.; Donkó-Tóth, V.; Keglevich, P.; Weber, M.; Dékány, M.; Hazai, L. Synthesis and In Vitro Anticancer Evaluation of Chrysin Containing Hybrids and Other Chrysin Derivatives. Period. Polytech. Chem. Eng. 2023, 67, 316–336. [Google Scholar] [CrossRef]
  78. Jeon, K.-H.; Park, S.; Shin, J.-H.; Jung, A.-R.; Hwang, S.-Y.; Seo, S.H.; Jo, H.; Na, Y.; Kwon, Y. Synthesis and evaluation of 7-(3-aminopropyloxy)-substituted flavone analogue as a topoisomerase II catalytic inhibitor and its sensitizing effect to enzalutamide in castration-resistant prostate cancer cells. Eur. J. Med. Chem. 2023, 246, 114999. [Google Scholar] [CrossRef] [PubMed]
  79. Serafin, P.; Kleczkowska, P. Bombesins: A New Frontier in Hybrid Compound Development. Pharmaceutics 2023, 15, 2597. [Google Scholar] [CrossRef] [PubMed]
  80. Singh, A.K.; Kumar, A.; Singh, H.; Sonawane, P.; Paliwal, H.; Thareja, S.; Pathak, P.; Grishina, M.; Jaremko, M.; Emwas, A.-H.; et al. Concept of Hybrid Drugs and Recent Advancements in Anticancer Hybrids. Pharmaceuticals 2022, 15, 1071. [Google Scholar] [CrossRef]
Figure 1. Structure of phenylbenzopyrone skeleton (marked in red) and the best-known flavones.
Figure 1. Structure of phenylbenzopyrone skeleton (marked in red) and the best-known flavones.
Applsci 14 01039 g001
Figure 2. Different types of connections of pharmacophores in hybrids ((a): hybrids having linker, (b): fused hybrids, (c): merged hybrids).
Figure 2. Different types of connections of pharmacophores in hybrids ((a): hybrids having linker, (b): fused hybrids, (c): merged hybrids).
Applsci 14 01039 g002
Figure 3. Synthesis of chrysin hybrids with benzyl-substituted 1,2,3-triazoles.
Figure 3. Synthesis of chrysin hybrids with benzyl-substituted 1,2,3-triazoles.
Applsci 14 01039 g003
Figure 4. Bis-substituted and aryltriazolyl chrysin hybrids.
Figure 4. Bis-substituted and aryltriazolyl chrysin hybrids.
Applsci 14 01039 g004
Figure 5. 7-Hydroxyflavones coupled with heterocycles.
Figure 5. 7-Hydroxyflavones coupled with heterocycles.
Applsci 14 01039 g005
Figure 6. Synthesis of aminoflavone hybrids with substituted 1,2,3-triazoles.
Figure 6. Synthesis of aminoflavone hybrids with substituted 1,2,3-triazoles.
Applsci 14 01039 g006
Figure 7. 1,2,3-Triazole substituted flavones on the 2-phenyl ring.
Figure 7. 1,2,3-Triazole substituted flavones on the 2-phenyl ring.
Applsci 14 01039 g007
Figure 8. Two special examples of triazole-containing hybrids.
Figure 8. Two special examples of triazole-containing hybrids.
Applsci 14 01039 g008
Figure 9. Synthesis of 7-O-(4-triphenylphosphoniumbutyl) quercetin iodide (21).
Figure 9. Synthesis of 7-O-(4-triphenylphosphoniumbutyl) quercetin iodide (21).
Applsci 14 01039 g009
Figure 10. Triphenylphosphonium-containing side chains in positions 3 and 5 of quercetin.
Figure 10. Triphenylphosphonium-containing side chains in positions 3 and 5 of quercetin.
Applsci 14 01039 g010
Figure 11. Chrysin derivatives containing amino acids.
Figure 11. Chrysin derivatives containing amino acids.
Applsci 14 01039 g011
Figure 12. Chrysin (27,28)–, apigenin (29)– and quercetin (30)–amino acid hybrids.
Figure 12. Chrysin (27,28)–, apigenin (29)– and quercetin (30)–amino acid hybrids.
Applsci 14 01039 g012
Figure 13. Synthesis of flavone–amino acid hybrids using the Buchwald–Hartwig reaction.
Figure 13. Synthesis of flavone–amino acid hybrids using the Buchwald–Hartwig reaction.
Applsci 14 01039 g013
Figure 14. Flavone hybrids containing β-carotene (marked in blue) and vindoline (marked in red).
Figure 14. Flavone hybrids containing β-carotene (marked in blue) and vindoline (marked in red).
Applsci 14 01039 g014
Figure 15. Flavone–cyclic amine hybrids in position 7.
Figure 15. Flavone–cyclic amine hybrids in position 7.
Applsci 14 01039 g015
Figure 16. Piperazine- and piperidine-substituted flavones.
Figure 16. Piperazine- and piperidine-substituted flavones.
Applsci 14 01039 g016
Figure 17. Flavones coupled with different heterocycles.
Figure 17. Flavones coupled with different heterocycles.
Applsci 14 01039 g017
Figure 18. Flavone–flavone and flavone–coumarin hybrids.
Figure 18. Flavone–flavone and flavone–coumarin hybrids.
Applsci 14 01039 g018
Figure 19. Aryl derivatives of flavones coupled directly or with a short linker.
Figure 19. Aryl derivatives of flavones coupled directly or with a short linker.
Applsci 14 01039 g019
Figure 20. Selected structures with characteristic coupling components.
Figure 20. Selected structures with characteristic coupling components.
Applsci 14 01039 g020
Table 1. The most characteristic hybrid skeletons of flavones and their biological activities.
Table 1. The most characteristic hybrid skeletons of flavones and their biological activities.
FlavonePharmacophoreBiological EffectActivity *Control Activity SectionReference
chrysin1,2,3-triazolesantibacterialMIC = 6.25 µg/mL
(E. coli)
penicillin,
MIC = 50 µg/mL
2.1.[14]
chrysin1,2,3-triazolesantiproliferativeIC50 = 0.733 µM
(HeLa)
cisplatin,
IC50 = 12.2 µM
2.1.[15,16]
6-aminoflavone1,2,3-triazolesantiproliferativeGI50 < 0.01 µM
(MDA-MB-231)
paclitaxel,
GI50 = 0.091 µM
2.1.[17]
luteolin1,2,3-triazolesantiplasmodialIC50 = 3.85 µM
(P. falciparum 3D7)
artemisinin,
IC50 = 1.12 µM
2.1.[18]
quercetintriphenylphosphineantioxidantEC50 = 6.3 µM
(DPPH assay)
quercetin,
EC50 = 6.0 µM
2.2.[19,20]
chrysinamino acidsantiproliferativeIC50 = 3.78 µM
(MGC-803)
cisplatin,
IC50 = 4.40 µM
2.3.[21,22]
quercetinamino acidsMDR modulatorIC50 = 0.14 µM
(MES-SA/Dx5)
doxorubicin,
IC50 = 8.20 µM
2.3.[23]
2-phenylchromen-
4-one
amino acidsantiproliferativeIC50 = 9.2 µM
(CCRF-CEM)
n.d.2.3.[24]
chrysinvindolineantiproliferativeGI50 = 1.1 µM
(LOX IMVI)
n.d.2.4.[12]
chrysinmorpholineantibacterialMIC = 6.25 µg/mL
(B. sphaericus)
streptomycin,
MIC = 12.5 µg/mL
2.5.[25]
chrysinpiperazinesantitumorIC50 = 4.67 µM
(HeLa)
gefitinib,
IC50 = 17.9 µM
2.5.[26]
apigeninpiperazinesantitumorIC50 = 0.0147 µM
(PARP-1 inhibition)
olaparib,
IC50 = 0.0051 µM
2.5.[27]
2-phenylchromen-
4-one
imidazo[1,2-a]
pyridine
antiplasmodialIC50 = 1.98 µM
(P. falciparum K1)
chloroquine,
IC50 = 0.25 µM
2.6.[28]
2-phenylchromen-
4-one
2-amino-thiazoleantibacterialMIC = 12.5 µg/mL
(B. lintus)
ciprofloxacin,
MIC < 1.5 µg/mL
2.6.[29]
7-hydroxyflavonedihydro-1H-furo[2,3-c]pyrazoleantitumorIC50 = 2.5 µM
(Hep2)
doxorubicin,
IC50 = 10 µM
2.6.[30]
7-hydroxyflavoneacridinemonoamine
oxidase inhibitor
IC50 = 3.24 µM
(MAO-B inhibition)
tacrine,
IC50 = 7.85 µM
2.6.[31]
5,6,7-trimethoxy-
flavone
6-chlorotacrineanti-Alzheimer’s
disease
IC50 = 0.0128 µM
(AChE inhibition)
6-chlorotacrine,
IC50 = 0.0785 µM
2.6.[32]
chrysinapigeninantiproliferativeIC50 = 0.179 µM
(PARP-1 inhibition)
olaparib,
IC50 = 0.0051 µM
2.7.[27]
mosloflavoneresveratrolanti-inflammatoryIC50 = 0.31 µM
(PGE2 inhibition)
n.d.2.8.[33]
2-phenylchromen-
4-one
naphthopyranxanthine
oxidase inhibitor
IC50 = 0.62 µM
(XO inhibition)
apigenin,
IC50 = 1.11 µM
2.9.[34]
2-phenylchromen-
4-one
estradiolantiproliferativeIC50 = 3.9 µM
(HeLa)
cisplatin,
IC50 > 330 µM
2.9.[35]
kaempferolcarbohydrateantibacterialMIC = 0.05 µg/mL
(S. gallinarum)
ciprofloxacin,
MIC = 0.25 µg/mL
2.9.[36]
* The activities presented are the value of the most effective derivative. The type of microorganism or cell line tested is shown in parentheses.
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

Hazai, L.; Zsoldos, B.; Halmai, M.; Keglevich, P. Flavone Hybrids and Derivatives as Bioactive Agents. Appl. Sci. 2024, 14, 1039. https://doi.org/10.3390/app14031039

AMA Style

Hazai L, Zsoldos B, Halmai M, Keglevich P. Flavone Hybrids and Derivatives as Bioactive Agents. Applied Sciences. 2024; 14(3):1039. https://doi.org/10.3390/app14031039

Chicago/Turabian Style

Hazai, László, Bernadett Zsoldos, Mónika Halmai, and Péter Keglevich. 2024. "Flavone Hybrids and Derivatives as Bioactive Agents" Applied Sciences 14, no. 3: 1039. https://doi.org/10.3390/app14031039

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop