Next Article in Journal
A Preclinical Study of an 125I-Labeled PSMA Ligand for Prostate-Cancer Puncture
Next Article in Special Issue
New Antifungal Agents with Azole Moieties
Previous Article in Journal
Special Issue on Plant and Marine-Derived Natural Product Research in Drug Discovery: Strengths and Perspective
Previous Article in Special Issue
A Review of the Biological Activity of Amidrazone Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chalcone: A Promising Bioactive Scaffold in Medicinal Chemistry

by
Gayathri Rajendran
1,2,†,
Deepu Bhanu
1,2,†,
Baladhandapani Aruchamy
1,2,
Prasanna Ramani
1,2,*,
Nanjan Pandurangan
3,
Kondapa Naidu Bobba
4,
Eun Jung Oh
5,
Ho Yun Chung
5,6,
Prakash Gangadaran
6,7 and
Byeong-Cheol Ahn
6,7,*
1
Dhanvanthri Laboratory, Department of Sciences, Amrita School of Physical Sciences, Amrita Vishwa Vidyapeetham, Coimbatore 641112, India
2
Center of Excellence in Advanced Materials & Green Technologies (CoE–AMGT), Amrita School of Engineering, Amrita Vishwa Vidyapeetham, Coimbatore 641112, India
3
Department of Sciences, Amrita School of Arts and Sciences, Mysuru Campus, Amrita Vishwa Vidyapeetham, Mysuru 570026, India
4
Department of Radiology and Biomedical Imaging, University of California (San Francisco), San Francisco, CA 94143, USA
5
Department of Plastic and Reconstructive Surgery, CMRI, School of Medicine, Kyungpook National University, Kyungpook National University Hospital, Daegu 41944, Korea
6
BK21 FOUR KNU Convergence Educational Program of Biomedical Sciences for Creative Future Talents, Department of Biomedical Science, School of Medicine, Kyungpook National University, Daegu 41944, Korea
7
Department of Nuclear Medicine, School of Medicine, Kyungpook National University, Kyungpook National University Hospital, Daegu 41944, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally.
Pharmaceuticals 2022, 15(10), 1250; https://doi.org/10.3390/ph15101250
Submission received: 16 September 2022 / Revised: 26 September 2022 / Accepted: 1 October 2022 / Published: 11 October 2022
(This article belongs to the Special Issue Feature Reviews in Medicinal Chemistry)

Abstract

:
Chalcones are a class of privileged scaffolds with high medicinal significance due to the presence of an α,β-unsaturated ketone functionality. Numerous functional modifications of chalcones have been reported, along with their pharmacological behavior. The present review aims to summarize the structures from natural sources, synthesis methods, biological characteristics against infectious and non-infectious diseases, and uses of chalcones over the past decade, and their structure–activity relationship studies are detailed in depth. This critical review provides guidelines for the future design and synthesis of various chalcones. In addition, this could be highly supportive for medicinal chemists to develop more promising candidates for various infectious and non-infectious diseases.

1. Introduction

Molecular diversity is a crucial factor for any drug discovery program. The molecules of interest can be obtained through natural or chemical syntheses. Natural products remain a source of unexplored chemotypes (e.g., molecular scaffolds and pharmacophores) and offer many structural fragments for medicinal chemistry. Therefore, natural products are always considered to be invaluable sources of inspiration for drug discovery development. Recently, scientists have been attracted to smaller fragments or scaffolds (~300 Da) due to their unique properties, i.e., effective molecular interactions with the targeted biological receptors for pharmacological properties. These features are very crucial in the area of fragment-based drug design (FBDD). Such exceptional properties are found in flavonoids. Flavonoids are among the largest polyphenols that are widely distributed in fruits, vegetables, and the plant kingdom as secondary metabolites. Flavonoids are classified into chalcones, flavanones, flavones, isoflavones, aurones, neoflavones and biflavones. To date, more than 7500 molecularly diverse flavonoids have been reported. Among flavonoids, chalcones play a prominent role and have been found to be a precursor for the biosynthesis of flavonoids.
Chalcones have an α,β-unsaturated system with three carbon units as its structural basis [1]. They are known as 1,3-diphenylprop-2-en-1-one, which is implicated in their hydrophobic/hydrophilic nature that has sparked much research—especially in the field of infectious and non-infectious diseases [2]. Chalcones occur in the cis or trans form, of which the trans isomer is more thermodynamically stable than the cis form. Chalcones have been reported to have biological properties such as anticancer [3] antimicrobial [4], antiviral [5], antioxidant [6], anti-Alzheimer’s [7], antitumor [8], antidiabetic [9], anti-Parkinson’s [10], anti-inflammatory [11], and anti-nociceptive [12] effects (Figure 1). After a careful literature search, it has been found that in the past decade many scientific developments took place in chalcone chemistry, which is scattered in nature. Furthermore, heterocyclic rings are known to be bioactive, and incorporating heterocyclic moieties in the structural framework of chalcones will further increase their bioactivity. In recent years, numerous research projects have been completed, and scientists have synthesized various chalcone analogs. Chalcones bearing heterocyclic rings—such as quinoline [13], pyrrole [14], pyridine [15], indole [16], pyrazole [17], benzofuran [18], coumarin [19], isoxazole [20], benzimidazole [21], and azulene [22]—have been reported in the literature. There have also been reports of metallocene-based chalcone analogs such as ruthenocenyl and ferrocenyl chalcones in the literature [23]. Chalcones can also be used for other optical applications, such as fluorescent probes [24]. With this view, the effort was taken to summarize the structure, synthesis, properties, and applications of chalcones over the past decade, and their structure–activity relationship (SAR) studies are described in detail in this review.

2. Chalcones from Natural Sources

Chalcones can be biosynthesized by combining phenyl propane and acetate pathways through convergent syntheses. Coumaroyl-CoA is used for making phenylalanine in three enzymatic steps, and malonyl-CoA is synthesized by carboxylation of acetyl-CoA—a central intermediate in the Krebs tricarboxylic acid cycle. The chalcones are further classified based on their structure; for example, the compounds that usually occur in plants are commonly hydroxylated or methoxylated (Figure 2A); glycosylated and prenylated forms are found less frequently (Figure 2C), as are geranylated chromene-based derivatives (Figure 2B) and dihydroquinone-based chalcones (Figure 2D). Herbs that contain chalcones have been used as medicines in the past to treat a variety of illnesses. Chalcones can be found in large quantities in various plants, including Butea monosperma, Desmodium gangeticum, Humulus lupulus, Helichrysum rugulosum, Neoraputia magnifica, Angelica keiskei, Piper hispidum, Tarenna attenuata, and Calythropis aurea [12,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57].

3. Synthesis of Chalcones

Chalcones possess an α,β-unsaturated ketone with a diphenyl substitution, and their other functionality makes them unique as a scaffold with multifunctional biological properties. Procurement of chalcones from natural sources often entails logistical problems. Thus, they are chemically synthesized through the following methods:

3.1. Conventional Synthesis of Chalcones

Generally, chalcones are synthesized from an aldehyde and acetophenone through Claisen–Schmidt condensation in the presence of an acid or base. Numerous conditions have been established for synthesizing chalcones, as detailed below.
In the Claisen–Schmidt condensation reaction (Scheme 1), different bases—such as NaOH, KOH, LiOH, Ba(OH)2, and organic bases (i.e., piperidine and pyridine) have been used to facilitate the synthesis of chalcones. Likewise, acidic reagents such as HCl and AlCl3 have also been established. In both conditions, the yield of chalcone is comparatively less, with an average yield of 50–60% based on the substitutions
Later, to increase the yield, different catalysts were also tried [58,59]. For example, materials such as chitosan [60], Al–Mg hydrotalcite [61], and Cs-pollucite nanozeolite modified with organosilane [62] have been used for the reaction. The acid-based catalyst, i.e., BF3-Et2O, has many advantages over others, with the yield of the chalcone increasing drastically, which is also helpful for the ease of production [2]. Similarly, other materials—such as activated carbons [63], nanoporous AISBA-15 [64], cesium salt of 12-tungstophoric acid [65], ionic liquids [66], protonated aluminate mesoporous silica nanomaterials [67,68], (MWCNTs)-COOH-CeO2 hybrids [69], modified fluorapatite [68], Fe3O4-MOF core–shell magnetic microspheres [70], nanosized ZnWO4 [71], and graphene-supported ZnO nanoparticles [72]—have also been established. BF3-etherate can be used as a better catalyst for chalcone synthesis, as the products are obtained with higher yields—in the range of 75–96%. Using heterogeneous catalysts also yields chalcones in higher amounts, where the catalysts can be reused for many cycles of the reaction.

3.2. Greener Approaches for the Synthesis of Chalcones

3.2.1. Microwave-Assisted Method

Based on the disadvantages of the conventional methods, microwave-assisted synthesis has been developed for synthesizing chalcones. The main advantage of the microwave-assisted method is that the reaction is usually faster. For example, a few heteroaromatic chalcones can be yielded in 3–5 min without using solvents. Here, the catalysts (the conditions may be either acidic or basic) are combined with the starting materials to form a thick paste, from which the product can be isolated in a soluble form by the elution of catalysts. A series of chalcones have been synthesized using K2CO3 as a base along with microwave irradiation [73]. Furthermore, chalcones with pyrazoline moieties have been synthesized at higher yields under acidic conditions [74].

3.2.2. Ultrasound-Irradiated Synthesis

Similarly, another strategy—the ultrasound irradiation method—can also be established without solvents. The reaction rate is accelerated by ultrasonic irradiation, leading to the formation of the product in a much smaller amount of time and with a higher yield [75]. The completion of the reaction was shown in about 10 s during the synthesis of thiophene-based chalcones with cesium carbonate as a base [76]. Chalcones were also synthesized in the presence of a zeolite-based catalyst in solvent-free conditions, and the yield was greater than 95% [77].

3.2.3. Grinding Technique

This technique provides an alternate way to synthesize chalcones without solvents. The equivalent quantities of aromatic aldehydes, acetophenone, and the reagent are mixed and ground in a porcelain mortar for about 10 min. The desired product can be isolated after adding cold water through filtration [78]. A series of chalcones with different benzaldehydes and 2-acetyl-2-naphthol were synthesized using a grinding technique, and the products were obtained with a yield from 85 to 95% [79].

3.3. Coupling Reactions

Although chalcones are obtained with a good yield through the greener approaches, they do not generally take the form of multi-substituted chalcones. On the other hand, a coupling reaction could be a good platform for developing chalcone synthesis. Some essential coupling reactions for preparing chalcones are described below (Scheme 2).
Palladium provides an excellent opportunity for developing cross-coupling reactions for synthesizing chalcones; for example, PdCl2 is a suitable catalyst for preparing chalcones from potassium styryl trifluoroborates and benzoyl chlorides [80]. Chalcones have been synthesized via silver-catalyzed coupling of cinnamic acids with substituted α-keto acids [81]. In a similar vein, Sonogashira coupling has been utilized to prepare chalcones via the reaction of aryl halides bearing electron-withdrawing groups and prop-2-yl-1-ol [82]. Chalcone derivatives were synthesized using phenylacetylene and benzoyl chloride as starting materials; the reaction followed the same mechanism as Sonogashira coupling [83]. Subsequently, palladium acetate was effectively utilized for the generation of chalcones via the Suzuki reaction under basic conditions. Another method has been developed using palladium acetate through the oxidative carbonylation of boronic acid and styrene via a cross-coupling strategy [84].
Furthermore, the palladium dibutyl acetate derivative is utilized to develop chalcones through Stille coupling. The reaction involves C-C bond formation between organotin compounds and aryl halides. A new series of chalcones were synthesized by the reaction between organotin compounds and benzoyl chloride [85]. Three different types of coupling reactions have been reported in addition to the palladium coupling reactions. For example, a new series of chalcones were synthesized using gold nanoparticles via a coupling reaction between substituted methyl ketones and aryl alcohols (Scheme 3). The reported reaction is recyclable and gives a good yield of chalcones [86].
Furthermore, a cross-dehydrogenative coupling reaction was developed for chalcones through an enamine-based strategy using a persulfate catalyst. New chalcone analogs were synthesized in 2009 using CDC in the presence of ammonium persulfate as an oxidant instead of metal-based catalysts (Scheme 4) [87].
Further, Julia and Kocienski developed a coupling reaction involving C-C bond formation (Scheme 5). A series of E-enones were synthesized using this method, and the product formation was observed with higher yields [88]. The reaction involves β-ketosulfones and aldehyde in the presence of a base. The elimination of sulfur dioxide produces the chalcones.

3.4. Miscellaneous Reactions

Due to their lower thermodynamic stability, trans-chalcones are synthesized more frequently than cis-chalcones. As a result of this phenomenon, cis-chalcones have not been subjected to many applications compared to trans-chalcones. Cis-chalcones containing furanyl rings were synthesized using reductive (3 + 2) annulation of pyrylium salts with benzil in the presence of P(NMe2)3 (Scheme 6) [89,90].
Furthermore, a one-pot reaction has been reported in which the starting materials are directly added to it in the presence of an oxidant. No purification is performed after a single step; thus, one-pot synthesis can be expedited. Phenyl methanol was oxidized into benzaldehyde in the presence of chromium trioxide, which further underwent a condensation reaction with acetophenone to produce chalcones with a good yield (Scheme 7) [90].
Furthermore, phenyl cinnamate undergoes rearrangement to produce hydroxy-group-bearing chalcones (Scheme 8). The reaction is performed under a nitrogen atmosphere using benzene as the solvent [91].
Along the same line, the Wittig reaction (Scheme 9) has also been utilized to prepare chalcones. A series of chalcones were synthesized from aromatic aldehydes and arsonium salt in the presence of KF.2H2O via the Wittig reaction using the grinding method [92]. The reaction of α-bromoylides with aldehydes in the presence of NBS results in bromo-α,β-unsaturated ketone [93].

4. Chalcones against Infectious Diseases

Coronaviruses have affected millions of people across the world. Shigella bacterial infections have also been a significant threat in developing nations, and the mortality rate due to diseases such as TB and malaria is on the rise. Even though many medications are available to treat these diseases, the pathogens often resist them [94]. In recent years, there has been a significant advancement in the study of synthetic and medicinal organic chemistry. Chalcones are considered promising pharmaceuticals, and the following subsections detail their bioactivity against various infectious diseases.

4.1. Anti-Tubercular Activity

Tuberculosis (TB) is a contagious illness that primarily affects the lungs. It is ranked as the 7th leading global cause of rising mortality rates. Streptomycin, pyrazinamide, and rifampicin are just a few of the medications created to treat TB [95]. Many chalcone analogs have been tested for anti-tubercular activity, and there have been reports of chalcones with intense anti-tubercular activity [96,97,98] (Figure 3). Quinoline-based chalcone 51 showed anti-tubercular activity against the M. tuberculosis H37Rv strain, with better MIC values (10–80 µM) than the standard drug rifampicin [99]. Chalcones with extended phenyl skeletons (52 and 53) showed good activity against Mycobacterium tuberculosis, with MIC values almost equal to those of standard drugs such as rifampicin and streptomycin. The docking and QSAR study results of these conformationally restricted chalcones also matched the MABA assay results [100,101]. The spirochrome chalcones (54) were tested for inhibitory activity against M. tuberculosis. Moreover, their docking studies were performed with protein phosphotyrosinase phosphatase B. Docking studies indicated that one of the spirochrome chalcones was bound to the protein at the same pocket region as the natural ligand, and the MIC value indicated that it was more active than the standard drug used in the MABA assay [101].
Seventeen new C-dimethylated chalcones of class 55 were tested for their anti-tubercular activity, and some of them in that class showed higher activity than streptomycin (MIC = 10.75 µM) and ciprofloxacin (MIC = 9.43 µM). The docking results with the M. tuberculosis protein tyrosinase phosphatase were analyzed, and compounds with 1,2,3-trimethoxy and 1,2,3,5-tetramethoxy substitutions in the aldehyde ring were more active, which was consistent with results from the MABA assay [102,103]. A class of acetylenic chalcones (56) were synthesized and studied for their anti-tubercular activity, with good MIC values (20–100 µM) [103]. Twelve triazole-based chalcones were synthesized, and their activity was studied using the MABA assay and molecular docking (protein 4Y6U). The results indicated that structures (57) and scaffolds enhance the activity [96]. Thiazole-based chalcones (58) showed more promising activity than the standard drug pyrazinamide, with MIC values of 2.43 µM [104].

4.2. Antiviral Activity

Numerous viral infections, including HIV and coronaviruses, cause serious health problems in people all over the world. Numerous plants that naturally contain chalcones have been used to treat viral infections since time immemorial. Some of these are listed in Section 2. The antiviral activity of numerous recently synthesized chalcones has been investigated, and most inhibit the activity of viruses [105,106,107,108]. The structures of some active chalcones are listed in Figure 4. Compound 59 has an EC50 value of 51.65 µg/mL, and its docking studies show that it forms four hydrogen bonds in the docking sites of TMV-CP (tobacco mosaic virus coat protein), indicating that it shows good activity against TMV [109]. Chalcone analogs containing purine units were synthesized, and among them, 60—with a nitro substitution on the para position and difluoro substitution on positions 2 and 4—showed activity against TMV [110]. Chalcone analogs that inhibit the activity of the hepatitis C virus have also been reported (category 61) [111]. Some natural chalcone derivatives (62) were screened for their activity against SARS-CoV-2, with satisfactory EC50 values. The results matched the molecular docking studies with the protein RdRp, and the threshold values were calculated [112]. Chalcone derivatives with the purine derivative (63) showed a better EC50 value than the standard drug ribavirin in inhibiting the tobacco mosaic virus [5]. A 1,2,3-triazine-containing chalcone (64) was reported to be active against TMV [113]. Chalcone 65 showed activity against CMV (cucumber mosaic virus) [114]. Chalcones 66 and 67, containing 1,3,4-oxadiazole and 1,3,4-thiadiazole units, showed efficient activity against TMV and were more effective than ribavirin [115]. Chalcone 68 with an O-benzyl substitution was reported to have anti-HIV activity [108]. Influenza A virus activity was inhibited by compound 69, which was synthesized via Hoveyda–Grubbs cyclization [116]. Compound 70—a chalcone derivative bearing malonate and pyrimidine units—showed significant activity against CMV [117]. The 4-thioquinazoline chalcone analog 71 showed potent activity against TMV, with EC50 values ranging from 138.1 µM to 154.8 µM, making it much more effective than ribavirin [118].

4.3. Antimalarial Activity

Malaria is caused by a parasite called Plasmodium falciparum, which is transmitted by mosquitoes. Antimalarial medications such as chloroquine and artemisinin are readily available. However, modern P. falciparum strains are resistant to available medications. Many chalcones that occur naturally have been used to treat malaria [119,120]. Additionally, a large number of synthetically developed chalcones have been tested for their efficacy against various parasite strains. Compound 72 (Figure 5), with a pyrimidine moiety, showed antimalarial activity, with an IC50 value of 21.4 µM [121]. A sulfonyl-based 4-methoxychalcone (73) could be used as an efficient oral drug against malaria, with an IC50 value of 2.06 µM; furthermore, the docking study indicated a comparatively high binding score of −7.3 Kcal/mol [122]. As expected, the imidazole group contributes to the activity, as indicated in compound 74, which showed potent antimalarial activity, with an IC50 value of 1.1 µM [123].
Since the quinolone group has already explored for its antiviral activity, the chalcone with the quinolone scaffold (75) showed an IC50 value of 0.031 µM and high binding energy with the heme protein [124]. Similarly, another derivative (76) showed IC50 values in the range of 44.06–59.37 µM and potency against the CQ-3D7 strain [125]. An unusual cyrhetrenyl chalcone (77) a derivative containing a ferrocene moiety—showed effective inhibition against strains CQ-3D7 and CQ-W2 [126]. Compound 78—a chalcone extended with a pyrrolidine unit—was effective against the 3D7 strain. The docking results proved that the molecule shows H-bonding and aromatic interactions with the proteins 1J2I and 1J3K [127]. An azide–alkyne coupling can produce a triazole-bearing chalcone from the combination of quinolone and natural moieties, as in the case of 79, which was found to be active against three different P. falciparum strains: CQ-D10, CQ-Dd2, and CQ-W2 [128]. A normal methoxylated chalcone possesses isopropyl ether linkage, as indicated in compound 80, and can act as an excellent antimalarial agent with inhibitory activity against two parasitic strains [120,121,129]. Furthermore, chromene-based chalcones 81–84 have inhibitory activity against strains CQ-Dd2, 3D7, and KI [130,131]. Quinolone-based chalcone 85, synthesized by MW irradiation, was proven to have inhibitory activity against P. falciparum [132]. Chalcone 86, which contains an acridine unit in its structural framework, possesses high efficacy against malaria [133]. Similarly, prenylated chalcone 87 inhibits the activity of the CQ-3D7 strain of P. falciparum [134]. A 4-aminoquinolinyl chalcone amide in conjunction with the furano derivative (88) was active against two strains, with IC50 values ranging from 0.004 µM to 0.07 µM [135]. Furthermore, 89 is an indole-moiety-containing chalcone that is active against malaria, with an IC50 value of 2.5 µM [136].

4.4. Antibacterial Activity

One of the leading health problems affecting human populations is bacterial infection. Pathogens becoming more resistant to antibiotics have caused the effectiveness of current antibiotics to decline. Both natural and artificial chalcones have demonstrated activity against some Gram-positive and Gram-negative strains [137,138]. Biofouling is one of the significant problems in marine-based product development. The chalcones shown in Figure 6 are active against different bacterial strains. In one of the studies, the organism V. natriegens (MD6) and other marine strains (P. fluorescens and B. flexus) were taken for antibacterial studies. The compounds 90, 91 (0.31–0.61 µM), and 92 (0.0.25–0.195 µM) were reported as having suitable microbial activities against all three organisms and could be used as anti-biofouling agents against marine microorganisms. [139]. Of these compounds, the thio-based chalcone (90) was more effective (0.058–0.22 µM) in inhibiting biofouling strains through cell membrane disruption. Chalcones 93, 94, and 95 were proven to have good activity, with MIC values in the range of 0.78–1.56 µM [140]. On the other hand, polyhydroxylated chalcones bearing hydroxy groups—9698, 100 (62.5–250 µM), and 102 (6.25–25 µM)—were active against methicillin-resistant Staphylococcus aureus [141]. Another polyhydroxylated chalcone (101) showed enhanced inhibitory activity against Gram-positive organisms (0.1–0.717 µM) such as Bacillus subtilis and Staphylococcus aureus, as well as the Gram-negative bacteria Pseudomonas aeruginosa, Escherichia coli, Salmonella typhi, and Proteus vulgaris [142]. The natural chalcone 102 was reported to be active against two Gram-positive bacteria [143].
The extensive analysis of bacterial studies proves that heterocyclic chalcones are superior to homocyclic chalcones. The trans-chalcone containing a difurano ring (103) (28 µM) potentially better inhibits the activity of S. aureus compared to the standard drug amoxicillin [144]. The thiazole-containing chalcone 104 (1.4 µM) is similar to vancomycin against S. aureus [145]. Three nitrogen-heterocycle-bearing chalcones—105, 106, and 107 (20 µM)—act against the Gram-negative bacterium S. dysenteriae and two Gram-positive bacteria (S. Typhi and S aureus) [146]. Compounds 108, 109, and 110 are metal complexes formed by chalcones that show inhibitory activity against E. coli, S. aureus, and K. pneumonia (2–52 µM) [147]. Compound 111 is a bis-chalcone-bearing thiophene unit that shows an MIC value (10–12.5 µM) better than that of the standard [148]. Chalcones 112 (128 µM), 113 (128 µM), and 114 (32 µM) are active against S. aureus [149]. Compound 115, with an MIC value of 2 µM, has an inhibitory activity similar to that of norfloxacin [150]. A cationic chalcone (116) inhibited the activity of S. aureus and MRSA strains, with MIC values ranging from 0.25 to 1 µM [151]. A β-carboline-linked chalcone (117) inhibited the activity of S. aureus, with an MIC of 440 µM and a ZOI of 15 mm [152,153]. A chalcone bearing a rhodamine-3-acetic acid unit (118) showed similar inhibitory activity (2 µM) to norfloxacin against S. aureus [153]. Compound 119 can be used as an efficient antibacterial agent with inhibitory activity against four different bacterial strains: Pseudomonas aeruginosa (44.06 mM), Escherichia coli (55.83 mM), Staphylococcus aureus (44.59 mM), and Bacillus subtilis (79.76 mM) [154].

5. Chalcones for Non-Infectious Diseases

Contact with an infected person does not result in the transmission of non-infectious diseases. They develop as a result of people’s changing lifestyles. These illnesses are challenging to diagnose, and if they are not found at an early stage there are no highly effective medicines available to treat them. Chalcones exhibit activity against cancer, diabetes, and other non-contagious illnesses. In this section, a number of the data from the literature are thoroughly discussed.

5.1. Anti-Alzheimer’s Activity

Alzheimer’s disease is a neurological condition that causes memory loss in people—especially those over 65 years old. It also results in other behavioral problems and a decline in language and orientation abilities. The factors vital to its cause are oxidative stress, acetylcholinesterase (AChE), and Aβ1-42 (amyloid beta) aggregation [155,156,157,158,159,160]. Compound 119a (Figure 7) showed promising activity, inhibiting the aggregation of Aβ1-42 [161,162]. The docking studies on newly synthesized chalcone-O-alkylamine analogs were performed with acetylcholinesterase, butyrylcholinesterase (BuChE), and Aβ1-42. It was found that compound 120 was bound to the same pocket region as the natural ligand. Moreover, the in vitro analysis showed that it could cross the BBB and inhibit MAO-B and BACE-1 in the treatment of neurodegenerative disorders [162]. A naturally derived chalcone—L licochalcone B (compound 121)—exhibits potent anti-Alzheimer’s activity [7]. Butein (compound 122) [163] and phloridzin (compound 123) [164] are some chalcones or derivatives occurring in nature, showing anti-Alzheimer’s activity as they induce oxidative stress.

5.2. Anticancer Activity

Cancer is one of the most serious illnesses and the leading cause of death worldwide. It is introduced by uncontrolled gene mutation or unregulated cell division in the body. Although drugs with minimal side effects have not yet been created, many chemotherapeutic drugs have been developed in recent years. Chalcones are pharmacologically significant molecules [165,166]. Numerous recently created chalcones exhibit activity against several cancer cell lines, such as MCF-7, A549, and PC3 [166,167,168,169].

5.2.1. Anti-Breast-Cancer Activity

Large numbers of chalcones (Figure 8 and Figure 9; Table 1) have been studied for their anti-breast-cancer activities, mainly in MCF-7 cells. Based on their structures, they can be classified into two broad areas: homocyclic and heterocyclic chalcone derivatives. Furthermore, homocyclic chalcones can be divided into three classes: (i) hydroxylated, (ii) methoxylated, and (iii) chalcones with extended functionality. It has been found that chalcones with extended functionality show higher IC50 values when compared to other categories, of which the homocyclic chalcone 134 with electron-withdrawing groups such as nitro and trifluoro extensions showed the best IC50 value (0.03 µM).
Similarly, heterocyclic chalcone derivatives are also categorized based on their functional derivatives. For example, a chalcone with a furano derivative (139) (IC50 = 0.00035 µM) showed excellent inhibitory activity in the nanomolar range compared with other counterparts (i.e., 138, 140, and 141). Furthermore, one representation from each of the categories—such as imidazole 143, IC50 = 0.56 µM), thiazole (146, IC50 = 0.18 µM)), pyrimidine (150, IC50 = 0.14 µM), and oxazoline (154, IC50 = 0.35 µM) derivatives—showed inhibitory activity in the range of nanomolar concentrations. However, the indole/quinolone/pyridine classes showed comparatively less activity compared to derivatives 139, 143, 146, 150, and 154. Of the listed compounds, heterocyclic derivatives showed comparatively higher anti-breast-cancer activity than homocyclic derivatives. The observation identified that furano derivative 139 with electron-withdrawing ability was the best inhibitor of breast cancer cell lines.

5.2.2. Anti-Lung-Cancer Activity

Furthermore, chalcones have been extended for their anti-lung-cancer activities. In the area, chalcones have been categorized in two ways: (i) chalcones with extended functionality, and (ii) heterocyclic chalcones. In the homocyclic category, compounds 160, 163, 165, and 166 show inhibitory activities in the nanomolar range. In the heterocyclic category, unlike other inhibitory activities, only a few compounds (168 and 169) showed the best inhibitory activities in the nanomolar range. Amongst the compounds reported, the chalcones with extended heterocyclic functionality showed the best inhibitory activity, with IC50 values of 0.33–2.07 nM (Figure 10) in the A549 and HLF cell lines.

5.2.3. Chalcones with Broad-Spectrum Anticancer Activities

Some of the chalcones in Table 2 indicate a broad spectrum of anticancer activities on different cancer cell lines. The homocyclic chalcones—such as 174, 179, 182, 185, and 187—have broad-spectrum anticancer activities with moderate IC50 values. Furthermore, heterocyclic chalcones (173, 175, 176, 181, and 184) showed lower to larger IC50 values, and chalcones with extended hetero functionality (177, 178, 179, 182, 183, and 186) were found to be the best in the broad-spectrum cancer cell lines, as indicated in the table. Moreover, chalcones with nitrogen- and sulfur-containing heterocycles and with methoxy substitutions have been reported to be active against microbes, leukemia, prostate cancer, and colon cancer [218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235].

5.3. Antidiabetic Activity

Diabetes is a metabolic disorder characterized by high blood glucose levels. The inhibition of crucial enzymes—such as α-glucosidase, α-amylase, tyrosinase phosphatase, AMP-kinase, and aldose reductase—can cure hyperglycemia [9,239,240,241]. Compounds such as bischalcone 188 inhibit the activity of α-glucosidase, with an IC50 value of 23.7 µM [242]. Compound 189 (0.92 µM) is a trihydroxy chalcone that is effective against diabetes as it inhibits AMP-kinase activity [243]. A chalcone with an aryloxy propylamine unit (190) has been reported as an effective antidiabetic agent, inhibiting type 2 diabetes in rats [244]. Compound 191 shows good inhibitory activity against PTP1B [245]. Compound 192, with the presence of methoxy and halogen groups, exhibits activity against diabetes [238]. Chalcone 193 (0.5–2.5 µM) was active against type 2 diabetes even though its docking results were not high as expected [239]. Compounds 194 and 195 are aminochalcones, while 196 is a naturally derived hydroxychalcone showing inhibitory activity against the enzymes peptidase, α-glucosidase, and PPAR, which decrease the blood glucose levels [9]. A triazine-bearing chalcone (197) showed potent antidiabetic activity [246]. Compounds 198–202 (Figure 11) exhibit higher antidiabetic activity than the standards rosiglitazone and pioglitazone (250–300 mg/dL) [247]. Chalcone 203 shows inhibitory activity against α-glucosidase, with an IC50 value of 67.77 µM [248].

5.4. Anti-Parkinson’s Activity

Parkinson’s disease is a neurological disorder involving the loss of dopamine neurons in the nigrostriatal pathway and decreased levels of neurotransmitters. Coordination, mental health, and mobility are all impacted by this. Chalcones have been found to inhibit MAO-B receptors in the literature (Figure 12) [10,249]. Compound 204 is an ethoxylated chalcone that can bind reversibly with the MAO-B pocket site, with an IC50 value of 0.53 µM [10]. The trifluoromethyl-substituted chalcone 205 is a suitable hMAO-B inhibitor [250]. Compounds 206 and 207 are active against Parkinson’s disease, with IC50 values of 0.29 µM and 0.087 µM [251]. Chalcone 208 can act as a good Parkinson’s drug, with an IC50 value of 0.0044 µM [252]. Compounds 209 and 210 (1–108 µM) are potent MAO-B inhibitors [253]. Compound 211 (1.20 µM) is an indole-based chalcone that can easily cross the BBB, and the results have been analyzed via PAMPA assay [254]. An imidazole-containing chalcone, 212 (20–100 µM), acts as both an MAO-A and MAO-B inhibitor [255]. Chalcones bearing oxime ethers—213 (76.8 µM) and 214 (81.8 µM)—and the thienyl chalcone 215 are active against diabetes, and their docking studies indicate that they bind efficiently at the pocket regions of hMAO-B [256].

6. SAR Studies

The extensive search and analysis showed that chalcones possess outstanding biological properties with respect to infectious and non-infectious diseases. The first part of the study indicated that 60 new chalcones have been isolated from different families—mainly from Fabaceae, Euphorbiaceae, Cyperaceae, Piperaceae, Lecythidaceae, Annonaceae, Myrtaceae, Umbelliferae, Lauraceae, Rutaceae, Rubiaceae, Caprifoliaceae, and Zingiberaceae. The structures of the reported chalcones in the plants are hydroxylated, methoxylated, a combination of hydroxylated and methoxylated, prenylated, non-prenylated, glycosylated, cinnamoylated, dihydroformylated, and so on. Scarcity and chemotaxonomy lead to the synthesis of chalcones via chemical methods. With this in mind, many synthesis methods have been discovered and listed. However, extensive protection and de-protection strategies are necessary to prepare functional-based chalcones such as hydroxyl-, amino-, and thio-based compounds. Both synthetic (compounds 124 and 125) and natural chalcones (compounds 1–14) bearing hydroxyl functional groups demonstrate higher activity in breast cancer cell lines [257,258]. Similarly, halogenated synthetic chalcones and methoxylated chalcones derived from both natural and synthetic sources have proven to be highly active against liver, prostrate, and lung cancer cell lines [259,260,261,262]. In addition, trihydroxylated and methoxylated chalcones showed higher activity than di- and mono-substituted ones [263,264,265].
From this report, it is very much clear that natural analogs of homocyclic chalcones with extended hetero functionality and heterocyclic chalcones possess excellent inhibitory activity against infectious and non-infectious diseases (Figure 13). In a similar vein, naturally occurring chalcones are poorly bioavailable molecules due to their decreased absorption in the intestine. However, the new chalcone categories are yet to be analyzed in terms of this particular property of interest. Similarly, the reported chalcones can be taken to the next level of studies, including animal studies, clinical trials, and much more.

7. Conclusions and Perspectives

Chalcones are well-known chemical compounds with high pharmacological importance. They can be of natural or synthetic origin. Many natural chalcones have been used as medicines since ancient times. Since the 1980s, research has been progressing in this particular area. Various routes have been developed for the synthesis of chalcones. The principal method used to synthesize chalcones is Claisen–Schmidt condensation. Other synthetic methodologies include coupling reactions in the presence of organometallic catalysts and rearrangement reactions. Light-enhanced synthesis methodologies have also been reported in the literature. Many studies have been conducted by varying the structural framework of chalcones by incorporating different heterocyclic moieties, and most showed different bioactivities. Metal-incorporated chalcones and their complexes have also been reported. Most chalcones are fluorescent and can be employed as fluorescent probes. Compared to synthetic chalcones, the extraction of natural chalcones often takes a long time and produces low yields. Chalcones obtained from natural sources are homocyclic, demonstrating lower bioactivity when compared to compounds synthesized with heterocyclic units. Many organic reactions have been reported in the literature to obtain a wide variety of substituted chalcones that are employed as anticancer, antidiabetic, antioxidant, antibacterial, anthelmintic, antiulcer, antiviral, insecticidal, and antiprotozoal agents. Therefore, the ability to structurally modify chalcones through synthesis has better advantages than natural chalcones, such as higher yields, easy handling, cost-effectiveness, and much more. The literature indicates that chalcones are most active against cancerous disorders—mostly against breast cancer cell lines, and also against lung cancer. Furthermore, SAR studies indicate that they have good ADMET properties, making them efficient as drug molecules. From this point of view, it is clear that these molecules’ research interest and importance will further increase in the coming years.

Author Contributions

G.R. and D.B. contributed equally to the collection of the literature and preparing the initial draft. B.A., P.R., N.P., K.N.B., E.J.O., H.Y.C., P.G. and B.-C.A. contributed to the editing and subsequent revisions. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF), funded by the Ministry of Education (NRF-2022R1I1A1A01068652, NRF-2022R1I1A3069687, and NRF-2022R1I1A1A01069734). This work was also supported by a grant from the National Research Foundation of Korea, funded by the Korean government (MSIT) (NRF-2020R1A2C2009496 and NRF-2022R1A2C2005057).

Data Availability Statement

Not applicable.

Acknowledgments

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Syam, S.; Abdelwahab, S.I.; Al-Mamary, M.A.; Mohan, S. Synthesis of Chalcones with Anticancer Activities. Molecules 2012, 17, 6179–6195. [Google Scholar] [CrossRef] [PubMed]
  2. Narender, T.; Papi Reddy, K. A Simple and Highly Efficient Method for the Synthesis of Chalcones by Using Borontrifluoride-Etherate. Tetrahedron Lett. 2007, 48, 3177–3180. [Google Scholar] [CrossRef]
  3. Constantinescu, T.; Lungu, C.N.; Jazvinš’cak, M.; Jembrek, J. Anticancer Activity of Natural and Synthetic Chalcones. Int. J. Mol. Sci. 2021, 22, 11306. [Google Scholar] [CrossRef] [PubMed]
  4. Mustafa, M.; Mostafa, Y.A. A Facile Synthesis, Drug-Likeness, and in Silico Molecular Docking of Certain New Azidosulfonamide–Chalcones and Their in Vitro Antimicrobial Activity. Mon. Für Chem. Chem. Mon. 2020, 151, 417–427. [Google Scholar] [CrossRef]
  5. Fu, Y.; Liu, D.; Zeng, H.; Ren, X.; Song, B.; Hu, D.; Gan, X. New chalcone derivatives: Synthesis, antiviral activity and mechanism of action. RSC Adv. 2020, 10, 24483–24490. [Google Scholar] [CrossRef]
  6. Osipova, V.P.; Polovinkina, M.A.; Telekova, L.R.; Velikorodov, A.V.; Stepkina, N.N.; Berberova, N.T. Synthesis and Antioxidant Activity of New Hydroxy Derivatives of Chalcones. Russ. Chem. Bull. 2020, 69, 504–509. [Google Scholar] [CrossRef]
  7. Cao, Y.; Xu, W.; Huang, Y.; Zeng, X. Licochalcone B, a Chalcone Derivative from Glycyrrhiza Inflata, as a Multifunctional Agent for the Treatment of Alzheimer’s Disease. Nat. Prod. Res. 2020, 34, 736–739. [Google Scholar] [CrossRef]
  8. Zhou, W.; Zhang, W.; Peng, Y.; Jiang, Z.H.; Zhang, L.; Du, Z. Design, Synthesis and Anti-Tumor Activity of Novel Benzimidazole-Chalcone Hybrids as Non-Intercalative Topoisomerase II Catalytic Inhibitors. Molecules 2020, 25, 3180. [Google Scholar] [CrossRef]
  9. Rammohan, A.; Bhaskar, B.V.; Venkateswarlu, N.; Gu, W.; Zyryanov, G.V. Design, Synthesis, Docking and Biological Evaluation of Chalcones as Promising Antidiabetic Agents. Bioorg. Chem. 2020, 95, 103527. [Google Scholar] [CrossRef]
  10. Maliyakkal, N.; Saleem, U.; Anwar, F.; Shah, M.A.; Ahmad, B.; Umer, F.; Almoyad, M.A.A.; Parambi, D.G.T.; Beeran, A.A.; Nath, L.R.; et al. Ameliorative Effect of Ethoxylated Chalcone-Based MAO-B Inhibitor on Behavioural Predictors of Haloperidol-Induced Parkinsonism in Mice: Evidence of Its Antioxidative Role against Parkinson’s Diseases. Environ. Sci. Pollut. Res. 2022, 29, 7271–7282. [Google Scholar] [CrossRef]
  11. Hsieh, H.-K.; Tsao, L.-T.; Wang, J.-P.; Lin, C.-N. Synthesis and Anti-Inflammatory Effect of Chalcones. J. Pharm. Pharmacol. 2000, 52, 163–171. [Google Scholar] [CrossRef] [PubMed]
  12. de Campos-Buzzi, F.; de Campos, J.P.; Tonini, P.P.; Corrêa, R.; Yunes, R.A.; Boeck, P.; Cechinel-Filho, V. Antinociceptive Effects of Synthetic Chalcones Obtained from Xanthoxyline. Arch. der Pharm. 2006, 339, 361–365. [Google Scholar] [CrossRef] [PubMed]
  13. Abdullah, M.I.; Mahmood, A.; Madni, M.; Masood, S.; Kashif, M. Synthesis, Characterization, Theoretical, Anti-Bacterial and Molecular Docking Studies of Quinoline Based Chalcones as a DNA Gyrase Inhibitor. Bioorg. Chem. 2014, 54, 31–37. [Google Scholar] [CrossRef] [PubMed]
  14. Özdemir, A.; Altintop, M.D.; Sever, B.; Gençer, H.K.; Kapkaç, H.A.; Atli, Ö.; Baysal, M. A New Series of Pyrrole-Based Chalcones: Synthesis and Evaluation of Antimicrobial Activity, Cytotoxicity, and Genotoxicity. Molecules 2017, 22, 2112. [Google Scholar] [CrossRef] [Green Version]
  15. Mojarrab, M.; Soltani, R.; Aliabadi, A. Pyridine Based Chalcones: Synthesis and Evaluation of Antioxidant Activity of 1-Phenyl-3-(Pyridin-2-Yl) Prop-2-En-1-One Derivatives. Jundishapur J. Nat. Pharm. Prod. 2013, 8, 125. [Google Scholar] [CrossRef] [Green Version]
  16. Özdemir, A.; Altintop, M.D.; Turan-Zitouni, G.; Çiftçi, G.A.; Ertorun, Ö.; Alataş, Ö.; Kaplancikli, Z.A. Synthesis and Evaluation of New Indole-Based Chalcones as Potential Antiinflammatory Agents. Eur. J. Med. Chem. 2015, 89, 304–309. [Google Scholar] [CrossRef]
  17. Chavan, H.V.; Adsul, L.K.; Kotmale, A.S.; Dhakane, V.D.; Thakare, V.N.; Bandgar, B.P. Design, Synthesis, Characterization and in Vitro and in Vivo Anti-Inflammatory Evaluation of Novel Pyrazole-Based Chalcones. J. Enzym. Inhib. Med. Chem. 2015, 30, 22–31. [Google Scholar] [CrossRef] [Green Version]
  18. Naik, N.; Kumar, V.H.; Dias, S.M.; Swami, R.J. Novel 4-methoxy-2-acetyl benzofuran based chalcones: A new perceptivity into their antioxidant potentials. Int. J. Pharm. Pharm. Sci. 2013, 5, 242–247. [Google Scholar]
  19. Rajeshirke, M.; Sreenath, M.C.; Chitrambalam, S.; Joe, I.H.; Sekar, N. Enhancement of NLO Properties in OBO Fluorophores Derived from Carbazole-Coumarin Chalcones Containing Carboxylic Acid at the N-Alykl Terminal End. J. Phys. Chem. C 2018, 122, 14313–14325. [Google Scholar] [CrossRef]
  20. Shaik, A.; Bhandare, R.R.; Palleapati, K.; Nissankararao, S.; Kancharlapalli, V.; Shaik, S. Antimicrobial, Antioxidant, and Anticancer Activities of Some Novel Isoxazole Ring Containing Chalcone and Dihydropyrazole Derivatives. Molecules 2020, 25, 1047. [Google Scholar] [CrossRef] [Green Version]
  21. Parikh, K.; Joshi, D. Antibacterial and Antifungal Screening of Newly Synthesized Benzimidazole-Clubbed Chalcone Derivatives. Med. Chem. Res. 2013, 22, 3688–3697. [Google Scholar] [CrossRef]
  22. Bala, D.; Jinga, L.I.; Popa, M.; Hanganu, A.; Voicescu, M.; Bleotu, C.; Tarko, L.; Nica, S. Design, Synthesis, and Biological Evaluation of New Azulene-Containing Chalcones. Materials 2022, 15, 1629. [Google Scholar] [CrossRef]
  23. Singh, A.; Rani, A.; Gut, J.; Rosenthal, P.J.; Kumar, V. Piperazine-linked 4-aminoquinoline-chalcone/ferrocenyl-chalcone conjugates: Synthesis and antiplasmodial evaluation. Chem. Biol. Drug Des. 2017, 90, 590–595. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, H.; Guo, C.; Guo, S.; Fan, J.; Wang, L.; Shi, D. Chalcone-Analogue Fluorescent Probes for Detecting Thiophenols in Seawater Samples. Talanta 2019, 201, 301–308. [Google Scholar] [CrossRef]
  25. Malek, S.N.A.; Phang, C.W.; Ibrahim, H.; Wahab, N.A.; Sim, K.S. Phytochemical and Cytotoxic Investigations of Alpinia Mutica Rhizomes. Molecules 2011, 16, 583–589. [Google Scholar] [CrossRef] [Green Version]
  26. Yang, X.W.; Wang, J.S.; Wang, Y.H.; Xiao, H.T.; Hu, X.J.; Mu, S.Z.; Yan-Lin, M.; Lin, H.; He, H.P.; Li, N.; et al. Tarennane and Tarennone, Two Novel Chalcone Constituents from Tarenna Attenuata. Planta Med. 2007, 73, 496–498. [Google Scholar] [CrossRef]
  27. Liu, Y.; Zhang, X.; Kelsang, N.; Tu, G.; Kong, D.; Lu, J.; Zhang, Y.; Liang, H.; Tu, P.; Zhang, Q. Structurally Diverse Cytotoxic Dimeric Chalcones from Oxytropis Chiliophylla. J. Nat. Prod. 2018, 81, 307–315. [Google Scholar] [CrossRef] [PubMed]
  28. Funakoshi-Tago, M.; Tanabe, S.; Tago, K.; Itoh, H.; Mashino, T.; Sonoda, Y.; Kasahara, T. Licochalcone A Potently Inhibits Tumor Necrosis Factor α-Induced Nuclear Factor-ΚB Activation through the Direct Inhibition of IκB Kinase Complex Activation. Mol Pharm. 2009, 76, 745–753. [Google Scholar] [CrossRef]
  29. Tomazela, D.M.; Pupo, M.T.; Passador, E.A.P.; da Silva, M.F.D.G.F.; Vieira, P.C.; Fernandes, J.B.; Rodrigues Fo, E.; Oliva, G.; Pirani, J.R. Pyrano Chalcones and a Flavone from Neoraputia Magnifica and Their Trypanosoma Cruzi Glycosomal Glyceraldehyde-3-Phosphate Dehydrogenase-Inhibitory Activities. Phytochemistry 2000, 55, 643–651. [Google Scholar] [CrossRef]
  30. Akihisa, T.; Tokuda, H.; Hasegawa, D.; Ukiya, M.; Kimura, Y.; Enjo, F.; Suzuki, T.; Nishino, H. Chalcones and Other Compounds from the Exudates of Angelica Keiskei and Their Cancer Chemopreventive Effects. J. Nat. Prod. 2006, 69, 38–42. [Google Scholar] [CrossRef]
  31. Cui, Y.; Ao, M.; Li, W.; Hu, J.; Yu, L. Anti-Inflammatory Activity of Licochalcone a Isolated from Glycyrrhiza Inflata. Z. Für Nat. Sect. C J. Biosci. 2008, 63, 361–365. [Google Scholar] [CrossRef] [PubMed]
  32. Pascoal, A.C.R.F.; Ehrenfried, C.A.; Lopez, B.G.C.; de Araujo, T.M.; Pascoal, V.D.B.; Gilioli, R.; Anheê, G.F.; Ruiz, A.L.T.G.; de Carvalho, J.E.; Stefanello, M.A.; et al. Antiproliferative Activity and Induction of Apoptosis in PC-3 Cells by the Chalcone Cardamonin from Campomanesia Adamantium (Myrtaceae) in a Bioactivity-Guided Study. Molecules 2014, 19, 1843–1855. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Tuntipaleepun, M.; Chakthong, S.; Ponglimanont, C.; Plodpai, P.; Voravuthikunchai, S.P. Antifungal and Cytotoxic Substances from the Stem Barks of Desmos Chinensis. Chin. Chem. Lett. 2012, 23, 587–590. [Google Scholar] [CrossRef]
  34. Liu, M.L.; Duan, Y.H.; Hou, Y.L.; Li, C.; Gao, H.; Dai, Y.; Yao, X.S. Nardoaristolones A and B, Two Terpenoids with Unusual Skeletons from Nardostachys Chinensis Batal. Org. Lett. 2013, 15, 1000–1003. [Google Scholar] [CrossRef] [PubMed]
  35. Yang, E.B.; Zhang, K.; Cheng, L.Y.; Mack, P. Butein, a Specific Protein Tyrosine Kinase Inhibitor. Biochem. Biophys. Res. Commun. 1998, 245, 435–438. [Google Scholar] [CrossRef]
  36. Ren, Y.; Yuan, C.; Qian, Y.; Chai, H.B.; Chen, X.; Goetz, M.; Kinghorn, A.D. Constituents of an Extract of Cryptocarya Rubra Housed in a Repository with Cytotoxic and Glucose Transport Inhibitory Effects. J. Nat. Prod. 2014, 77, 550–556. [Google Scholar] [CrossRef] [Green Version]
  37. Rees, K.A.; Bermudez, C.; Edwards, D.J.; Elliott, A.G.; Ripen, J.E.; Seta, C.; Huang, J.X.; Cooper, M.A.; Fraser, J.A.; Yeo, T.C.; et al. Flemingin-Type Prenylated Chalcones from the Sarawak Rainforest Plant Desmodium Congestum. J. Nat. Prod. 2015, 78, 2141–2144. [Google Scholar] [CrossRef]
  38. Inamori, Y.; Baba, K.; Tsujibo, H.; Taniguch, M.; Nakata, K.; Kozawa, M. Antibacterial Activity of Two Chalcones, Xanthoangelol and 4-Hydroxyderricin, Isolated from the Root of Angelica Keiskei KOIDZUMI. Chem. Pharm. Bull. 1991, 39, 1604–1605. [Google Scholar] [CrossRef] [Green Version]
  39. Costa, G.M.; Endo, E.H.; Cortez, D.A.G.; Nakamura, T.U.; Nakamura, C.V.; Dias Filho, B.P. Antimicrobial Effects of Piper Hispidum Extract, Fractions and Chalcones against Candida Albicans and Staphylococcus Aureus. J. Mycol. Médicale 2016, 26, 217–226. [Google Scholar] [CrossRef]
  40. Wang, Z.; Wang, N.; Han, S.; Wang, D.; Mo, S.; Yu, L.; Huang, H.; Tsui, K.; Shen, J.; Chen, J. Dietary Compound Isoliquiritigenin Inhibits Breast Cancer Neoangiogenesis via VEGF/VEGFR-2 Signaling Pathway. PLoS ONE 2013, 8, e68566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Agarkar, S.A.; Kulkarni, R.R.; Dhas, V.V.; Chinchansure, A.A.; Hazra, P.; Joshi, S.P.; Ogale, S.B. Isobutrin from Butea Monosperma (Flame of the Forest): A Promising New Natural Sensitizer Belonging to Chalcone Class. ACS Appl. Mater. Interfaces 2011, 3, 2440–2444. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, D.; Liang, J.; Zhang, J.; Wang, Y.; Chai, X. Natural Chalcones in Chinese Materia Medica: Licorice. Evidence-Based Complement. Altern. Med. 2020, 2020, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Malik, H.S.; Bilal, A.; Ullah, R.; Iqbal, M.; Khan, S.; Ahmed, I.; Krohn, K.; Saleem, R.S.Z.; Hussain, H.; Faisal, A. Natural and Semisynthetic Chalcones as Dual FLT3 and Microtubule Polymerization Inhibitors. J. Nat. Prod. 2020, 83, 3111–3121. [Google Scholar] [CrossRef] [PubMed]
  44. Aoki, N.; Muko, M.; Ohta, E.; Ohta, S. C-Geranylated Chalcones from the Stems of Angelica Keiskei with Superoxide-Scavenging Activity. J. Nat. Prod. 2008, 71, 1308–1310. [Google Scholar] [CrossRef] [PubMed]
  45. Kil, Y.S.; Choi, S.K.; Lee, Y.S.; Jafari, M.; Seo, E.K. Chalcones from Angelica Keiskei: Evaluation of Their Heat Shock Protein Inducing Activities. J. Nat. Prod. 2015, 78, 2481–2487. [Google Scholar] [CrossRef]
  46. Daikonya, A.; Katsuki, S.; Kitanaka, S. Antiallergic Agents from Natural Sources 9. Inhibition of Nitric Oxide Production by Novel Chalcone Derivatives from Mallotus Philippinensis (Euphorbiaceae). Chem. Pharm. Bull. 2004, 52, 1326–1329. [Google Scholar] [CrossRef] [Green Version]
  47. Fu, L.C.; Huang, X.A.; Lai, Z.Y.; Hu, Y.J.; Liu, H.J.; Cai, X.L. A New 3-Benzylchroman Derivative from Sappan Lignum (Caesalpinia Sappan). Molecules 2008, 13, 1923–1930. [Google Scholar] [CrossRef] [Green Version]
  48. Phrutivorapongkul, A.; Lipipun, V.; Ruangrungsi, N.; Kirtikara, K.; Nishikawa, K.; Maruyama, S.; Watanabe, T.; Ishikawa, T. Studies on the Chemical Constituents of Stem Bark of Millettia Leucantha: Isolation of New Chalcones with Cytotoxic, Anti-Herpes Simplex Virus and Anti-Inflammatory Activities. Chem. Pharm. Bull. 2003, 51, 187–190. [Google Scholar] [CrossRef] [Green Version]
  49. Svetaz, L.; Tapia, A.; López, S.N.; Furlán, R.L.E.; Petenatti, E.; Pioli, R.; Schmeda-Hirschmann, G.; Zacchino, S.A. Antifungal Chalcones and New Caffeic Acids Esters from Zuccagnia Punctata Acting against Soybean Infecting Fungi. J. Agric. Food Chem. 2004, 52, 3297–3300. [Google Scholar] [CrossRef]
  50. Wang, J.P.; Tsao, L.T.; Raung, S.L.; Lin, C.N. Investigation of the Inhibitory Effect of Broussochalcone A on Respiratory Burst in Neutrophils. Eur. J. Pharmacol. 1997, 320, 201–208. [Google Scholar] [CrossRef]
  51. Ohnogi, H.; Kudo, Y.; Tahara, K.; Sugiyama, K.; Enoki, T.; Hayami, S.; Sagawa, H.; Tanimura, Y.; Aoi, W.; Naito, Y.; et al. Six New Chalcones from Angelica Keiskei Inducing Adiponectin Production in 3T3-L1 Adipocytes. Biosci. Biotechnol. Biochem. 2012, 76, 961–966. [Google Scholar] [CrossRef] [PubMed]
  52. Haraguchi, H.; Inoue, J.; Tamura, Y.; Mizutani, K. Antioxidative components of Psoralea corylifolia (Leguminosae). Phytother. Res. Int. J. Devoted Pharmacol. Toxicol. Eval. Nat. Prod. Deriv. 2002, 16, 539–544. [Google Scholar] [CrossRef] [PubMed]
  53. Kulkarni, R.R.; Tupe, S.G.; Gample, S.P.; Chandgude, M.G.; Sarkar, D.; Deshpande, M.V.; Joshi, S.P. Antifungal Dimeric Chalcone Derivative Kamalachalcone e from Mallotus Philippinensis. Nat. Prod. Res. 2014, 28, 245–250. [Google Scholar] [CrossRef] [PubMed]
  54. Strathmann, J.; Gerhauser, C. Anti-Proliferative and Apoptosis-Inducing Properties of Xanthohumol, a Prenylated Chalcone from Hops (Humulus Lupulus L.). In Natural Compounds as Inducers of Cell Death; Springer: Dordrecht, The Netherlands, 2012; Volume 1, pp. 69–93. [Google Scholar] [CrossRef]
  55. McRae, J.M.; Yang, Q.; Crawford, R.J.; Palombo, E.A. Acylated Flavonoid Tetraglycoside from Planchonia Careya Leaves. Phytochem. Lett. 2008, 1, 99–102. [Google Scholar] [CrossRef]
  56. Chung, M.-I.; Weng, J.-R.; Lai, M.-H.; Yen, M.-H.; Lin, C.-N. A New Chalcone, Xanthones, and a Xanthonolignoid from Hypericum geminiflorum. J. Nat. Prod. 1999, 62, 1033–1035. [Google Scholar] [CrossRef]
  57. Maxwell, C.A.; Hartwig, U.A.; Joseph, C.M.; Phillips, D.A. A Chalcone and Two Related Flavonoids Released from Alfalfa Roots Induce Nod Genes of Rhizobium Meliloti. Plant Physiol 1989, 91, 842–847. [Google Scholar] [CrossRef] [Green Version]
  58. Nair, A.D.; Athira, C.K.; Manikandan, P.; Ramani, P. One-Pot Synthesis of Modified 4-Aryl-4H-Chromenes and Their Preliminary Anti-Cancer Studies. J. Indian Chem. Soc. 2019, 96, 19–22. [Google Scholar]
  59. Pandurangan, N.; Bose, C.; Banerji, A. Synthesis and Antioxygenic Activities of Seabuckthorn Flavone-3-Ols and Analogs. Bioorg. Med. Chem. Lett. 2011, 21, 5328–5330. [Google Scholar] [CrossRef]
  60. Hernawan; Purwono, B.; Triyono; Hanafi, M. The Use of Chitosan as a Solid Base Catalyst for the Chalcones Synthesis. IOP Conf. Ser. Earth Environ. Sci. 2020, 462, 2055. [Google Scholar] [CrossRef]
  61. Climent, M.J.; Corma, A.; Iborra, S.; Velty, A. Activated Hydrotalcites as Catalysts for the Synthesis of Chalcones of Pharmaceutical Interest. J. Catal. 2004, 221, 474–482. [Google Scholar] [CrossRef]
  62. Mohammad, S.A.G.; Khoerunnisa, F.; Rigolet, S.; Jean Daou, T.; Ling, T.C.; Ng, E.P. Hierarchical Cs–Pollucite Nanozeolite Modified with Novel Organosilane as an Excellent Solid Base Catalyst for Claisen–Schmidt Condensation of Benzaldehyde and Acetophenone. Processes 2020, 8, 96. [Google Scholar] [CrossRef]
  63. Winter, C.; Caetano, J.N.; Araújo, A.B.C.; Chaves, A.R.; Ostroski, I.C.; Vaz, B.G.; Pérez, C.N.; Alonso, C.G. Activated Carbons for Chalcone Production: Claisen-Schmidt Condensation Reaction. Chem. Eng. J. 2016, 303, 604–610. [Google Scholar] [CrossRef]
  64. Elamathi, P.; Chandrasekar, G.; Balamurali, M.M. Nanoporous AlSBA-15 Catalysed Claisen–Schmidt Condensation for the Synthesis of Novel and Biologically Active Chalcones. J. Porous Mater. 2020, 27, 817–829. [Google Scholar] [CrossRef]
  65. Rafiee, E.; Rahimi, F. A Green Approach to the Synthesis of Chalcones via Claisen-Schmidt Condensation Reaction Using Cesium Salts of 12-Tungstophosphoric Acid as a Reusable Nanocatalyst. Mon. Für Chem. Chem. Mon. 2013, 144, 361–367. [Google Scholar] [CrossRef]
  66. Das, S.; Porashar, B.; Saikia, S.; Borah, R. Brönsted Acidic Ionic Liquids Catalysed Sequential Michael-Like Addition of Indole with Chalcones via Claisen-Schmidt Condensation. ChemistrySelect 2020, 5, 3041–3047. [Google Scholar] [CrossRef]
  67. Sazegar, M.R.; Mahmoudian, S.; Mahmoudi, A.; Triwahyono, S.; Jalil, A.A.; Mukti, R.R.; Nazirah Kamarudin, N.H.; Ghoreishi, M.K. Catalyzed Claisen–Schmidt Reaction by Protonated Aluminate Mesoporous Silica Nanomaterial Focused on the (E)-Chalcone Synthesis as a Biologically Active Compound. RSC Adv. 2016, 6, 11023–11031. [Google Scholar] [CrossRef]
  68. Jioui, I.; Dânoun, K.; Solhy, A.; Jouiad, M.; Zahouily, M.; Essaid, B.; Len, C.; Fihri, A. Modified Fluorapatite as Highly Efficient Catalyst for the Synthesis of Chalcones via Claisen–Schmidt Condensation Reaction. J. Ind. Eng. Chem. 2016, 39, 218–225. [Google Scholar] [CrossRef]
  69. Heidarzadeh, T.; Nami, N.; Zareyee, D. Preparation of (MWCNTs)-COOH/CeO2Hybrid as an Efficient Catalyst for Claisen-Schmidt Condensation. J. Appl. Chem. Res. 2021, 15, 44–57. [Google Scholar]
  70. Ke, F.; Qiu, L.G.; Zhu, J. Fe3O4@MOF Core-Shell Magnetic Microspheres as Excellent Catalysts for the Claisen-Schmidt Condensation Reaction. Nanoscale 2014, 6, 1596–1601. [Google Scholar] [CrossRef]
  71. Paul, A.; Devi, M.; Dhar, S.S. Incorporation of Nanosized ZnWO4 and Fe3O4 on Graphitic Carbon Nitride to Fabricate a Novel, Highly Active Magnetically Recoverable Catalyst in Claisen–Schmidt Condensation. J. Phys. Chem. Solids 2020, 136, 109117. [Google Scholar] [CrossRef]
  72. Li, Z.; Zhao, H.; Han, H.; Liu, Y.; Song, J.; Guo, W.; Chu, W.; Sun, Z. Graphene-Supported ZnO Nanoparticles: An Efficient Heterogeneous Catalyst for the Claisen-Schmidt Condensation Reaction without Additional Base. Tetrahedron Lett. 2017, 58, 3984–3988. [Google Scholar] [CrossRef]
  73. Srivastava, Y.K. Ecofriendly Microwave Assisted Synthesis of Some Chalcones. Rasayan J. Chem. 2008, 1, 884–886. [Google Scholar]
  74. Farooq, S.; Ngaini, Z.; Mortadza, N.A. Microwave-Assisted Synthesis and Molecular Docking Study of Heteroaromatic Chalcone Derivatives as Potential Antibacterial Agents. Bull. Korean Chem. Soc. 2020, 41, 918–924. [Google Scholar] [CrossRef]
  75. Calvino, V.; Picallo, M.; López-Peinado, A.J.; Martín-Aranda, R.M.; Durán-Valle, C.J. Ultrasound Accelerated Claisen–Schmidt Condensation: A Green Route to Chalcones. Appl. Surf. Sci. 2006, 252, 6071–6074. [Google Scholar] [CrossRef]
  76. Perozo-Rondón, E.; Martín-Aranda, R.M.; Casal, B.; Durán-Valle, C.J.; Lau, W.N.; Zhang, X.F.; Yeung, K.L. Sonocatalysis in Solvent Free Conditions: An Efficient Eco-Friendly Methodology to Prepare Chalcones Using a New Type of Amino Grafted Zeolites. Catal. Today 2006, 114, 183–187. [Google Scholar] [CrossRef]
  77. Homerin, G.; Nica, A.S.; Farce, A.; Dubois, J.; Ghinet, A. Ultrasounds-Mediated 10-Seconds Synthesis of Chalcones as Potential Farnesyltransferase Inhibitors. Bioorg. Med. Chem. Lett. 2020, 30, 127149. [Google Scholar] [CrossRef]
  78. Kumar, S.; Lamba, M.S.; Makrandi, J.K. An Efficient Green Procedure for the Synthesis of Chalcones Using C-200 as Solid Support under Grinding Conditions. Green Chem. Lett. Rev. 2008, 1, 123–125. [Google Scholar] [CrossRef]
  79. Yeshwant, S.M.; Nanded, M.; Nalwar, Y.; Zangade, S.; Mokle, S.; Vibhute, A.; Vibhute, Y. An Efficient and Operationally Simple Synthesis of Some New Chalcones by Using Grinding Technique. Chem. Sci. J. 2009, 2011, 13. [Google Scholar]
  80. Al-Masum, M.; Ng, E.; Wai, M.C. Palladium-Catalyzed Direct Cross-Coupling of Potassium Styryltrifluoroborates and Benzoyl Chlorides—A One Step Method for Chalcone Synthesis. Tetrahedron Lett. 2011, 52, 1008–1010. [Google Scholar] [CrossRef]
  81. Diana, E.J.; Kanchana, U.S.; Mathew, T.V.; Anilkumar, G. Recent Developments in the Metal Catalysed Cross-Coupling Reactions for the Synthesis of the Enone System of Chalcones. Appl. Organomet. Chem. 2020, 34, e5987. [Google Scholar] [CrossRef]
  82. Braun, R.U.; Ansorge, M.; Müller, T.J.J. Coupling-Isomerization Synthesis of Chalcones. Chemistry 2006, 12, 9081–9094. [Google Scholar] [CrossRef]
  83. Hsieh, C.T.; Ötvös, S.B.; Wu, Y.C.; Mándity, I.M.; Chang, F.R.; Fülöp, F. Highly Selective Continuous-Flow Synthesis of Potentially Bioactive Deuterated Chalcone Derivatives. ChemPlusChem 2015, 80, 859–864. [Google Scholar] [CrossRef] [PubMed]
  84. Rao, M.L.N.; Venkatesh, V.; Jadhav, D.N. A Palladium Catalyzed Atom-Efficient Cross-Coupling Reactivity of Triarylbismuths with α,β-Unsaturated Acyl Chlorides. J. Organomet. Chem. 2008, 693, 2494–2498. [Google Scholar] [CrossRef]
  85. Yamakawa, T.; Kinoshita, H.; Miura, K. Synthetic Utility of Tribenzyltin Hydride and Its Derivatives as Easily Accessible, Removable, and Decomposable Organotin Reagents. J. Organomet. Chem. 2013, 724, 129–134. [Google Scholar] [CrossRef]
  86. Kim, S.; Bae, S.W.; Lee, J.S.; Park, J. Recyclable Gold Nanoparticle Catalyst for the Aerobic Alcohol Oxidation and C–C Bond Forming Reaction between Primary Alcohols and Ketones under Ambient Conditions. Tetrahedron 2009, 65, 1461–1466. [Google Scholar] [CrossRef]
  87. Li, C.J. Cross-Dehydrogenative Coupling (CDC): Exploring C-C Bond Formations beyond Functional Group Transformations. Acc. Chem. Res. 2009, 42, 335–344. [Google Scholar] [CrossRef] [PubMed]
  88. Kumar, A.; Sharma, S.; Tripathi, V.D.; Srivastava, S. Synthesis of Chalcones and Flavanones Using Julia–Kocienski Olefination. Tetrahedron 2010, 66, 9445–9449. [Google Scholar] [CrossRef]
  89. Tan, P.; Wang, S.R. Reductive (3 + 2) Annulation of Benzils with Pyrylium Salts: Stereoselective Access to Furyl Analogues of Cis-Chalcones. Org. Lett. 2019, 21, 6029–6033. [Google Scholar] [CrossRef] [PubMed]
  90. Mahapatra, D.K.; Bharti, S.K.; Asati, V. Anti-Cancer Chalcones: Structural and Molecular Target Perspectives. Eur. J. Med. Chem. 2015, 98, 69–114. [Google Scholar] [CrossRef] [PubMed]
  91. Ramakrishnan, V.T.; Kagan, J. The Photochemical Synthesis of 2′-Hydroxychalcones from Phenyl Cinnamates. J. Org. Chem. 1970, 35, 2901–2904. [Google Scholar] [CrossRef]
  92. Wang, C.Q. Solvent-Free Stereoselective Synthesis of Chalcones via Wittig Reaction of Arsonium Ylide by Grinding. In Advanced Materials Research; Trans Tech Publications Ltd.: Wollerau, Switzerland, 2014; Volume 864, pp. 2132–2135. [Google Scholar]
  93. Huang, Z.; Wang, L.; Huang, X. Stereoselective Synthesis of α-Bromo-α,β-Unsaturated Ketones via Wittig Reaction. Synth. Commun. 2006, 33, 757–762. [Google Scholar] [CrossRef]
  94. Baker, R.E.; Mahmud, A.S.; Miller, I.F.; Rajeev, M.; Rasambainarivo, F.; Rice, B.L.; Takahashi, S.; Tatem, A.J.; Wagner, C.E.; Wang, L.F.; et al. Infectious Disease in an Era of Global Change. Nat. Rev. Microbiol. 2021, 20, 193–205. [Google Scholar] [CrossRef] [PubMed]
  95. Sahoo, B.M.; Banik, B.K.; Mahato, A.K.; Shanthi, C.N.; Mohantad, B.C. Microwave-assisted synthesis of antitubercular agents: A novel approach. In Green Approaches in Medicinal Chemistry for Sustainable Drug Design; Elsevier: Amsterdam, The Netherlands, 2020; pp. 779–818. [Google Scholar] [CrossRef]
  96. Kaur, H.; Singh, R.; Kant, R. Synthesis, Molecular Docking, and Evaluation of Triazole and Chalcone Conjugate as Antitubercular Agent. Res. Sq. 2021. [Google Scholar] [CrossRef]
  97. Bhoot, D.; Khunt, R.C.; Parekh, H.H. Synthesis and Biological Evaluation of Chalcones and Acetyl Pyrazoline Derivatives Comprising Furan Nucleus as an Antitubercular Agents. Med. Chem. Res. 2012, 21, 3233–3239. [Google Scholar] [CrossRef]
  98. Muradás, T.C.; Abbadi, B.L.; Villela, A.D.; Macchi, F.S.; Bergo, P.F.; de Freitas, T.F.; Sperotto, N.D.M.; Timmers, L.F.S.M.; de Souza, O.N.; Picada, J.N.; et al. Pre-Clinical Evaluation of Quinoxaline-Derived Chalcones in Tuberculosis. PLoS ONE 2018, 13, e0202568. [Google Scholar] [CrossRef] [Green Version]
  99. Rao, N.S.; Shaik, A.B.; Routhu, S.R.; Hussaini, S.M.A.; Sunkari, S.; Rao, A.V.S.; Reddy, A.M.; Alarifi, A.; Kamal, A. New Quinoline Linked Chalcone and Pyrazoline Conjugates: Molecular Properties Prediction, Antimicrobial and Antitubercular Activities. ChemistrySelect 2017, 2, 2989–2996. [Google Scholar] [CrossRef]
  100. Yadav, D.K.; Ahmad, I.; Shukla, A.; Khan, F.; Negi, A.S.; Gupta, A. QSAR and Docking Studies on Chalcone Derivatives for Antitubercular Activity against M.Tuberculosis H37Rv. J. Chemom. 2014, 28, 499–507. [Google Scholar] [CrossRef]
  101. Mujahid, M.; Yogeeswari, P.; Sriram, D.; Basavanag, U.M.V.; Díaz-Cervantes, E.; Córdoba-Bahena, L.; Robles, J.; Gonnade, R.G.; Karthikeyan, M.; Vyas, R.; et al. Spirochromone-chalcone conjugates as antitubercular agents: Synthesis, bio evaluation and molecular modeling studies. RSC Adv. 2015, 5, 106448–106460. [Google Scholar] [CrossRef]
  102. Anandam, R.; Jadav, S.S.; Ala, V.B.; Ahsan, M.J.; Bollikolla, H.B. Synthesis of New C-Dimethylated Chalcones as Potent Antitubercular Agents. Med. Chem. Res. 2018, 27, 1690–1704. [Google Scholar] [CrossRef]
  103. Hans, R.H.; Guantai, E.M.; Lategan, C.; Smith, P.J.; Wan, B.; Franzblau, S.G.; Gut, J.; Rosenthal, P.J.; Chibale, K. Synthesis, Antimalarial and Antitubercular Activity of Acetylenic Chalcones. Bioorg. Med. Chem. Lett. 2010, 20, 942–944. [Google Scholar] [CrossRef]
  104. Kasetti, A.B.; Singhvi, I.; Nagasuri, R.; Bhandare, R.R.; Shaik, A.B. Thiazole–Chalcone Hybrids as Prospective Antitubercular and Antiproliferative Agents: Design, Synthesis, Biological, Molecular Docking Studies and In Silico ADME Evaluation. Molecules 2021, 26, 2847. [Google Scholar] [CrossRef]
  105. Trivedi, J.C.; Bariwal, J.B.; Upadhyay, K.D.; Naliapara, Y.T.; Joshi, S.K.; Pannecouque, C.C.; de Clercq, E.; Shah, A.K. Improved and Rapid Synthesis of New Coumarinyl Chalcone Derivatives and Their Antiviral Activity. Tetrahedron Lett. 2007, 48, 8472–8474. [Google Scholar] [CrossRef]
  106. Elkhalifa, D.; Al-Hashimi, I.; Al Moustafa, A.E.; Khalil, A. A comprehensive review on the antiviral activities of chalcones. J. Drug Target. 2021, 29, 403–419. [Google Scholar] [CrossRef] [PubMed]
  107. Onyilagha, J.C.; Malhotra, B.; Elder, M.; French, C.J.; Towers, G.N. Comparative studies of inhibitory activities of chalcones on tomato ringspot virus (ToRSV). Can. J. Plant Pathol. 1997, 19, 133–137. [Google Scholar] [CrossRef]
  108. Cole, A.L.; Hossain, S.; Cole, A.M.; Phanstiel, O. Synthesis and Bioevaluation of Substituted Chalcones, Coumaranones and Other Flavonoids as Anti-HIV Agents. Bioorg. Med. Chem. 2016, 24, 2768–2776. [Google Scholar] [CrossRef] [PubMed]
  109. Zhou, D.; Xie, D.; He, F.; Song, B.; Hu, D. Antiviral Properties and Interaction of Novel Chalcone Derivatives Containing a Purine and Benzenesulfonamide Moiety. Bioorg. Med. Chem. Lett. 2018, 28, 2091–2097. [Google Scholar] [CrossRef] [PubMed]
  110. Gan, X.; Wang, Y.; Hu, D.; Song, B. Design, Synthesis, and Antiviral Activity of Novel Chalcone Derivatives Containing a Purine Moiety. Chin. J. Chem. 2017, 35, 665–672. [Google Scholar] [CrossRef]
  111. Mateeva, N.; Eyunni, S.V.K.; Redda, K.K.; Ononuju, U.; Hansberry, T.D.; Aikens, C.; Nag, A. Functional Evaluation of Synthetic Flavonoids and Chalcones for Potential Antiviral and Anticancer Properties. Bioorg. Med. Chem. Lett. 2017, 27, 2350–2356. [Google Scholar] [CrossRef]
  112. Duran, N.; Polat, M.F.; Aktas, D.A.; Alagoz, M.A.; Ay, E.; Cimen, F.; Tek, E.; Anil, B.; Burmaoglu, S.; Algul, O. New Chalcone Derivatives as Effective against SARS-CoV-2 Agent. Int. J. Clin. Pract. 2021, 75, e14846. [Google Scholar] [CrossRef]
  113. Tang, X.; Su, S.; Chen, M.; He, J.; Xia, R.; Guo, T.; Chen, Y.; Zhang, C.; Wang, J.; Xue, W. Novel chalcone derivatives containing a 1, 2, 4-triazine moiety: Design, synthesis, antibacterial and antiviral activities. RSC Adv. 2019, 9, 6011–6020. [Google Scholar] [CrossRef] [Green Version]
  114. Wang, Y.J.; Zhou, D.G.; He, F.C.; Chen, J.X.; Chen, Y.Z.; Gan, X.H.; Hu, D.Y.; Song, B.A. Synthesis and Antiviral Bioactivity of Novel Chalcone Derivatives Containing Purine Moiety. Chin. Chem. Lett. 2018, 29, 127–130. [Google Scholar] [CrossRef]
  115. Gan, X.; Hu, D.; Chen, Z.; Wang, Y.; Song, B. Synthesis and Antiviral Evaluation of Novel 1,3,4-Oxadiazole/Thiadiazole-Chalcone Conjugates. Bioorg. Med. Chem. Lett. 2017, 27, 4298–4301. [Google Scholar] [CrossRef] [PubMed]
  116. Bizzarri, B.M.; Fanelli, A.; Piccinino, D.; De Angelis, M.; Dolfa, C.; Palamara, A.T.; Nencioni, L.; Zippilli, C.; Crucianelli, M.; Saladino, R.; et al. Synthesis of stilbene and chalcone inhibitors of influenza A virus by SBA-15 supported Hoveyda-Grubbs metathesis. Catalysts 2019, 9, 983. [Google Scholar] [CrossRef]
  117. Chen, Z.; Li, P.; Hu, D.; Dong, L.; Pan, J.; Luo, L.; Zhang, W.; Xue, W.; Jin, L.; Song, B. Synthesis, Antiviral Activity, and 3D-QSAR Study of Novel Chalcone Derivatives Containing Malonate and Pyridine Moieties. Arab. J. Chem. 2019, 12, 2685–2696. [Google Scholar] [CrossRef] [Green Version]
  118. Wan, Z.; Hu, D.; Li, P.; Xie, D.; Gan, X. Synthesis, antiviral bioactivity of novel 4-thioquinazoline derivatives containing chalcone moiety. Molecules 2015, 20, 11861–11874. [Google Scholar] [CrossRef] [Green Version]
  119. Awasthi, S.K.; Mishra, N.; Kumar, B.; Sharma, M.; Bhattacharya, A.; Mishra, L.C.; Bhasin, V.K. Potent Antimalarial Activity of Newly Synthesized Substituted Chalcone Analogs in Vitro. Med. Chem. Res. 2009, 18, 407–420. [Google Scholar] [CrossRef]
  120. Qin, H.L.; Zhang, Z.W.; Lekkala, R.; Alsulami, H.; Rakesh, K.P. Chalcone Hybrids as Privileged Scaffolds in Antimalarial Drug Discovery: A Key Review. Eur. J. Med. Chem. 2020, 193, 112215. [Google Scholar] [CrossRef]
  121. Insuasty, B.; Ramírez, J.; Becerra, D.; Echeverry, C.; Quiroga, J.; Abonia, R.; Robledo, S.M.; Vélez, I.D.; Upegui, Y.; Muñoz, J.A.; et al. An Efficient Synthesis of New Caffeine-Based Chalcones, Pyrazolines and Pyrazolo[3,4-b][1,4]Diazepines as Potential Antimalarial, Antitrypanosomal and Antileishmanial Agents. Eur. J. Med. Chem. 2015, 93, 401–413. [Google Scholar] [CrossRef]
  122. De Oliveira, M.E.; Cenzi, G.; Nunes, R.R.; Andrighetti, C.R.; Valadão, D.M.D.S.; Dos Reis, C.; Simões, C.M.O.; Nunes, R.J.; Junior, M.C.; Taranto, A.G.; et al. Antimalarial activity of 4-metoxychalcones: Docking studies as falcipain/plasmepsin inhibitors, ADMET and lipophilic efficiency analysis to identify a putative oral lead candidate. Molecules 2013, 18, 15276–15287. [Google Scholar] [CrossRef]
  123. Yadav, N.; Dixit, S.K.; Bhattacharya, A.; Mishra, L.C.; Sharma, M.; Awasthi, S.K.; Bhasin, V.K. Antimalarial Activity of Newly Synthesized Chalcone Derivatives In Vitro. Chem. Biol. Drug Des. 2012, 80, 340–347. [Google Scholar] [CrossRef]
  124. Hameed, A.; Masood, S.; Hameed, A.; Ahmed, E.; Sharif, A.; Abdullah, M.I. Anti-Malarial, Cytotoxicity and Molecular Docking Studies of Quinolinyl Chalcones as Potential Anti-Malarial Agent. J. Comput. Mol. Des. 2019, 33, 677–688. [Google Scholar] [CrossRef] [PubMed]
  125. Gayam, V.; Ravi, S. Cinnamoylated Chloroquine Analogues: A New Structural Class of Antimalarial Agents. Eur. J. Med. Chem. 2017, 135, 382–391. [Google Scholar] [CrossRef] [PubMed]
  126. Arancibia, R.; Biot, C.; Delaney, G.; Roussel, P.; Pascual, A.; Pradines, B.; Klahn, A.H. Cyrhetrenyl Chalcones: Synthesis, Characterization and Antimalarial Evaluation. J. Organomet. Chem. 2013, 723, 143–148. [Google Scholar] [CrossRef]
  127. Syahri, J.; Nasution, H.; Nurohmah, B.A.; Purwono, B.; Yuanita, E.; Zakaria, N.H.; Hassan, N.I. Design, synthesis and biological evaluation of aminoalkylated chalcones as antimalarial agent. Sains Malays. 2020, 49, 2667–2677. [Google Scholar] [CrossRef]
  128. Guantai, E.M.; Ncokazi, K.; Egan, T.J.; Gut, J.; Rosenthal, P.J.; Smith, P.J.; Chibale, K. Design, Synthesis and in Vitro Antimalarial Evaluation of Triazole-Linked Chalcone and Dienone Hybrid Compounds. Bioorg. Med. Chem. 2010, 18, 8243–8256. [Google Scholar] [CrossRef]
  129. Gutteridge, C.E.; Thota, D.S.; Curtis, S.M.; Kozar, M.P.; Li, Q.; Xie, L.; Zhang, J.; Melendez, V.; Asher, C.O.; Luong, T.T.; et al. In Vitro Biotransformation, in Vivo Efficacy and Pharmacokinetics of Antimalarial Chalcones. Pharmacology 2011, 87, 96–104. [Google Scholar] [CrossRef]
  130. Kumar, R.; Mohanakrishnan, D.; Sharma, A.; Kaushik, N.K.; Kalia, K.; Sinha, A.K.; Sahal, D. Reinvestigation of Structure–Activity Relationship of Methoxylated Chalcones as Antimalarials: Synthesis and Evaluation of 2,4,5-Trimethoxy Substituted Patterns as Lead Candidates Derived from Abundantly Available Natural β-Asarone. Eur. J. Med. Chem. 2010, 45, 5292–5301. [Google Scholar] [CrossRef]
  131. Tadigoppula, N.; Korthikunta, V.; Gupta, S.; Kancharla, P.; Khaliq, T.; Soni, A.; Srivastava, R.K.; Srivastava, K.; Puri, S.K.; Raju, K.S.R.; et al. Synthesis and Insight into the Structure-Activity Relationships of Chalcones as Antimalarial Agents. J. Med. Chem. 2013, 56, 31–45. [Google Scholar] [CrossRef]
  132. Sarveswari, S.; Vijayakumar, V.; Siva, R.; Priya, R. Synthesis of 4-Hydroxy-2(1h)-Quinolone Derived Chalcones, Pyrazolines and Their Antimicrobial, in Silico Antimalarial Evaluations. Appl. Biochem. Biotechnol. 2015, 175, 43–64. [Google Scholar] [CrossRef]
  133. Tomar, V.; Bhattacharjee, G.; Kamaluddin; Rajakumar, S.; Srivastava, K.; Puri, S. Synthesis of new chalcone derivatives containing acridinyl moiety with potential antimalarial activity. Eur. J. Med. Chem. 2010, 45, 745–751. [Google Scholar] [CrossRef]
  134. Syahri, J.; Rullah, K.; Armunanto, R.; Yuanita, E.; Nurohmah, B.A.; Aluwi, M.F.F.M.; Kok, L.; Wai, B.P. Synthesis, biological evaluation, QSAR analysis, and molecular docking of chalcone derivatives for antimalarial activity. Parasite 2016, 4, 8. [Google Scholar] [CrossRef]
  135. Smit, F.J.; N’Da, D.D. Synthesis, in vitro antimalarial activity and cytotoxicity of novel 4-aminoquinolinyl-chalcone amides. Bioorg. Med. Chem. 2014, 22, 1128–1138. [Google Scholar] [CrossRef] [PubMed]
  136. Jyoti; Gaur, R.; Kumar, Y.; Cheema, H.S.; Kapkoti, D.S.; Darokar, M.P.; Khan, F.; Bhakuni, R.S. Synthesis, Molecular Modelling Studies of Indolyl Chalcone Derivatives and Their Antimalarial Activity Evaluation. Nat. Prod. Res. 2021, 35, 3261–3268. [Google Scholar] [CrossRef] [PubMed]
  137. Mouzié, C.M.; Guefack, M.-G.F.; Kianfé, B.Y.; Serondo, H.U.; Ponou, B.K.; Siwe-Noundou, X.; Teponno, R.B.; Krause, R.W.M.; Kuete, V.; Tapondjou, L.A. A New Chalcone and Antimicrobial Chemical Constituents of Dracaena Stedneuri. Pharmaceuticals 2022, 15, 725. [Google Scholar] [CrossRef]
  138. Lagu, S.B.; Yejella, R.P.; Bhandare, R.R.; Shaik, A.B. Design, Synthesis, and Antibacterial and Antifungal Activities of Novel Trifluoromethyl and Trifluoromethoxy Substituted Chalcone Derivatives. Pharmaceuticals 2020, 13, 375. [Google Scholar] [CrossRef]
  139. Sivakumar, P.M.; Prabhawathi, V.; Doble, M. Antibacterial Activity and QSAR of Chalcones against Biofilm-Producing Bacteria Isolated from Marine Waters. SAR QSAR Environ. Res. 2010, 21, 247–263. [Google Scholar] [CrossRef]
  140. Satokata, A.A.C.; Souza, J.H.; Silva, L.L.O.; Santiago, M.B.; Ramos, S.B.; de Assis, L.R.; Theodoro, R.D.S.; e Oliveira, L.R.; Regasini, L.O.; Martins, C.H.G. Chalcones with potential antibacterial and antibiofilm activities against periodontopathogenic bacteria. Anaerobe 2022, 76, 102588. [Google Scholar] [CrossRef]
  141. Moreira Osório, T.; Delle Monache, F.; Domeneghini Chiaradia, L.; Mascarello, A.; Regina Stumpf, T.; Roberto Zanetti, C.; Bardini Silveira, D.; Regina Monte Barardi, C.; de Fatima Albino Smânia, E.; Viancelli, A.; et al. Antibacterial Activity of Chalcones, Hydrazones and Oxadiazoles against Methicillin-Resistant Staphylococcus Aureus. Bioorg. Med. Chem. Lett. 2012, 22, 225–230. [Google Scholar] [CrossRef]
  142. Sivakumar, P.M.; Ganesan, S.; Veluchamy, P.; Doble, M. Novel Chalcones and 1,3,5-Triphenyl-2-Pyrazoline Derivatives as Antibacterial Agents. Chem. Biol. Drug Des. 2010, 76, 407–411. [Google Scholar] [CrossRef]
  143. Li, Y.; Sun, B.; Zhai, J.; Fu, L.; Zhang, S.; Zhang, J.; Liu, H.; Xie, W.; Deng, H.; Chen, Z.; et al. Synthesis and Antibacterial Activity of Four Natural Chalcones and Their Derivatives. Tetrahedron Lett. 2019, 60, 151165. [Google Scholar] [CrossRef]
  144. Abdula, A.; Abdula, A.M. Synthesis, Characterization and Antibacterial Activity of (E)-Chalcone Derivatives. Eur. J. Chem. 2013, 4, 207–210. [Google Scholar] [CrossRef] [Green Version]
  145. Sashidhara, K.V.; Rao, K.B.; Kushwaha, P.; Modukuri, R.K.; Singh, P.; Soni, I.; Shukla, P.K.; Chopra, S.; Pasupuleti, M. Novel Chalcone-Thiazole Hybrids as Potent Inhibitors of Drug Resistant Staphylococcus Aureus. ACS Med. Chem. Lett. 2015, 6, 809–813. [Google Scholar] [CrossRef] [Green Version]
  146. Venkatesan, P.; Sumathi, S. Piperidine Mediated Synthesis of N-heterocyclic Chalcones and Their Antibacterial Activity. J. Heterocycl. Chem. 2009, 47, 81–84. [Google Scholar] [CrossRef]
  147. Dkhar, L.; Banothu, V.; Pinder, E.; Phillips, R.M.; Kaminsky, W.; Kollipara, M.R. Ru, Rh and Ir Metal Complexes of Pyridyl Chalcone Derivatives: Their Potent Antibacterial Activity, Comparable Cytotoxicity Potency and Selectivity to Cisplatin. Polyhedron 2020, 185, 114606. [Google Scholar] [CrossRef]
  148. Asiri, A.M.; Khan, S.A. Synthesis and anti-bacterial activities of a bis-chalcone derived from thiophene and its bis-cyclized products. Molecules 2011, 16, 523–531. [Google Scholar] [CrossRef] [Green Version]
  149. Tran, T.-D.; Nguyen, T.-T.; Do, T.-H.; Huynh, T.-N.; Tran, C.-D.; Thai, K.-M. Synthesis and antibacterial activity of some heterocyclic chalcone analogues alone and in combination with antibiotics. Molecules 2012, 17, 6684–6696. [Google Scholar] [CrossRef] [PubMed]
  150. Zheng, C.J.; Jiang, S.M.; Chen, Z.H.; Ye, B.J.; Piao, H.R. Synthesis and Anti-Bacterial Activity of Some Heterocyclic Chalcone Derivatives Bearing Thiofuran, Furan, and Quinoline Moieties. Arch. der Pharm. 2011, 344, 689–695. [Google Scholar] [CrossRef]
  151. Chu, W.C.; Bai, P.Y.; Yang, Z.Q.; Cui, D.Y.; Hua, Y.G.; Yang, Y.; Yang, Q.Q.; Zhang, E.; Qin, S. Synthesis and Antibacterial Evaluation of Novel Cationic Chalcone Derivatives Possessing Broad Spectrum Antibacterial Activity. Eur. J. Med. Chem. 2018, 143, 905–921. [Google Scholar] [CrossRef] [PubMed]
  152. Venkataramana Reddy, P.O.; Hridhay, M.; Nikhil, K.; Khan, S.; Jha, P.N.; Shah, K.; Kumar, D. Synthesis and Investigations into the Anticancer and Antibacterial Activity Studies of β-Carboline Chalcones and Their Bromide Salts. Bioorg. Med. Chem. Lett. 2018, 28, 1278–1282. [Google Scholar] [CrossRef] [PubMed]
  153. Chen, Z.H.; Zheng, C.J.; Sun, L.P.; Piao, H.R. Synthesis of New Chalcone Derivatives Containing a Rhodanine-3-Acetic Acid Moiety with Potential Anti-Bacterial Activity. Eur. J. Med. Chem. 2010, 45, 5739–5743. [Google Scholar] [CrossRef] [PubMed]
  154. Santosh, R.; Selvam, M.K.; Kanekar, S.U.; Nagaraja, G.K. Synthesis, Characterization, Antibacterial and Antioxidant Studies of Some Heterocyclic Compounds from Triazole-Linked Chalcone Derivatives. ChemistrySelect 2018, 3, 6338–6343. [Google Scholar] [CrossRef]
  155. Rani, A.; Singh, A.; Kaur, J.; Singh, G.; Bhatti, R.; Gumede, N.; Kisten, P.; Singh, P.; Kumar, V. 1H-1, 2, 3-triazole grafted tacrine-chalcone conjugates as potential cholinesterase inhibitors with the evaluation of their behavioral tests and oxidative stress in mice brain cells. Bioorg. Chem. 2021, 114, 105053. [Google Scholar] [CrossRef] [PubMed]
  156. Mathew, B.; Haridas, A.; Uçar, G.; Baysal, I.; Joy, M.; Mathew, G.E.; Lakshmanan, B.; Jayaprakash, V. Synthesis, Biochemistry, and Computational Studies of Brominated Thienyl Chalcones: A New Class of Reversible MAO-B Inhibitors. ChemMedChem 2016, 11, 1161–1171. [Google Scholar] [CrossRef] [PubMed]
  157. Rampa, A.; Montanari, S.; Pruccoli, L.; Bartolini, M.; Falchi, F.; Feoli, A.; Cavalli, A.; Belluti, F.; Gobbi, S.; Tarozzi, A.; et al. Chalcone-Based Carbamates for Alzheimer’s Disease Treatment. Futur. Med. Chem. 2017, 9, 749–764. [Google Scholar] [CrossRef]
  158. Thapa, P.; Upadhyay, S.P.; Suo, W.Z.; Singh, V.; Gurung, P.; Lee, E.S.; Sharma, R.; Sharma, M. Chalcone and Its Analogs: Therapeutic and Diagnostic Applications in Alzheimer’s Disease. Bioorg. Chem. 2021, 108, 104681. [Google Scholar] [CrossRef]
  159. Zhang, X.; Rakesh, K.P.; Bukhari, S.N.A.; Balakrishna, M.; Manukumar, H.M.; Qin, H.L. Multi-Targetable Chalcone Analogs to Treat Deadly Alzheimer’s Disease: Current View and Upcoming Advice. Bioorg. Chem. 2018, 80, 86–93. [Google Scholar] [CrossRef]
  160. Cong, L.; Dong, X.; Wang, Y.; Deng, Y.; Li, B.; Dai, R. On the Role of Synthesized Hydroxylated Chalcones as Dual Functional Amyloid-β Aggregation and Ferroptosis Inhibitors for Potential Treatment of Alzheimer’s Disease. Eur. J. Med. Chem. 2019, 166, 11–21. [Google Scholar] [CrossRef]
  161. Wang, X.Q.; Zhou, L.Y.; Tan, R.X.; Liang, G.P.; Fang, S.X.; Li, W.; Xie, M.; Wen, Y.H.; Wu, J.Q.; Chen, Y.P. Design, Synthesis, and Evaluation of Chalcone Derivatives as Multifunctional Agents against Alzheimer’s Disease. Chem. Biodivers. 2021, 18, e2100341. [Google Scholar] [CrossRef]
  162. Bai, P.; Wang, K.; Zhang, P.; Shi, J.; Cheng, X.; Zhang, Q.; Zheng, C.; Cheng, Y.; Yang, J.; Lu, X.; et al. Development of Chalcone-O-Alkylamine Derivatives as Multifunctional Agents against Alzheimer’s Disease. Eur. J. Med. Chem. 2019, 183, 111737. [Google Scholar] [CrossRef]
  163. Lee, D.S.; Jeong, G.S. Butein Provides Neuroprotective and Anti-neuroinflammatory Effects through Nrf2/ARE-dependent Haem Oxygenase 1 Expression by Activating the PI3K/Akt Pathway. Br. J. Pharmacol. 2016, 173, 2894. [Google Scholar] [CrossRef] [Green Version]
  164. Bhullar, K.S.; Rupasinghe, H.P.V. Polyphenols: Multipotent Therapeutic Agents in Neurodegenerative Diseases. Oxid. Med. Cell. Longev. 2013, 2013, 891748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Singh, M.; Kaur, M.; Silakari, O. Flavones: An Important Scaffold for Medicinal Chemistry. Eur. J. Med. Chem. 2014, 84, 206–239. [Google Scholar] [CrossRef] [PubMed]
  166. Quattrocchio, F.; Baudry, A.; Lepiniec, L.; Grotewold, E. The regulation of flavonoid biosynthesis. In The Science of Flavonoids; Springer: New York, NY, USA, 2006; pp. 97–122. [Google Scholar] [CrossRef]
  167. Nijveldt, R.J.; Van Nood, E.L.S.; Van Hoorn, D.E.; Boelens, P.G.; Van Norren, K.; Van Leeuwen, P.A. Flavonoids: A review of probable mechanisms of action and potential applications. Am. J. Clin. Nutr. 2001, 74, 418–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Thomas, N.; Zachariah, S.M.; Ramani, P. 4-Aryl-4H-Chromene-3-Carbonitrile Derivates: Synthesis and Preliminary Anti-Breast Cancer Studies. J. Heterocycl. Chem. 2016, 53, 1778–1782. [Google Scholar] [CrossRef]
  169. Doroghazi, J.R.; Albright, J.C.; Goering, A.W.; Ju, K.S.; Haines, R.R.; Tchalukov, K.A.; Labeda, D.P.; Kelleher, N.L.; Metcalf, W.W. A Roadmap for Natural Product Discovery Based on Large-Scale Genomics and Metabolomics. Nat. Chem. Biol. 2014, 10, 963–968. [Google Scholar] [CrossRef]
  170. Kim, S.H.; Lee, E.; Baek, K.H.; Kwon, H.B.; Woo, H.; Lee, E.S.; Kwon, Y.; Na, Y. Chalcones, Inhibitors for Topoisomerase I and Cathepsin B and L, as Potential Anti-Cancer Agents. Bioorg. Med. Chem. Lett. 2013, 23, 3320–3324. [Google Scholar] [CrossRef]
  171. Wang, G.; Liu, W.; Gong, Z.; Huang, Y.; Li, Y.; Peng, Z. Design, Synthesis, Biological Evaluation and Molecular Docking Studies of New Chalcone Derivatives Containing Diaryl Ether Moiety as Potential Anticancer Agents and Tubulin Polymerization Inhibitors. Bioorg. Chem. 2020, 95, 103565. [Google Scholar] [CrossRef]
  172. Wang, G.; Liu, W.; Gong, Z.; Huang, Y.; Li, Y.; Peng, Z. Synthesis, Biological Evaluation, and Molecular Modelling of New Naphthalene-Chalcone Derivatives as Potential Anticancer Agents on MCF-7 Breast Cancer Cells by Targeting Tubulin Colchicine Binding Site. J. Enzym. Inhib. Med. Chem. 2020, 35, 139–144. [Google Scholar] [CrossRef] [PubMed]
  173. Mokale, S.N.; Dube, P.N.; Bhavale, S.A.; Sayed, I.; Begum, A.; Nevase, M.C.; Shelke, V.R.; Mujaheed, A. Synthesis, in-Vitro Screening, and Docking Analysis of Novel Pyrrolidine and Piperidine-Substituted Ethoxy Chalcone as Anticancer Agents. Med. Chem. Res. 2015, 24, 1842–1856. [Google Scholar] [CrossRef]
  174. Khanapure, S.; Jagadale, M.; Bansode, P.; Choudhari, P.; Rashinkar, G. Anticancer Activity of Ruthenocenyl Chalcones and Their Molecular Docking Studies. J. Mol. Struct. 2018, 1173, 142–147. [Google Scholar] [CrossRef]
  175. Zhang, H.; Liu, J.J.; Sun, J.; Yang, X.H.; Zhao, T.T.; Lu, X.; Gong, H.B.; Zhu, H.L. Design, Synthesis and Biological Evaluation of Novel Chalcone Derivatives as Antitubulin Agents. Bioorg. Med. Chem. 2012, 20, 3212–3218. [Google Scholar] [CrossRef] [PubMed]
  176. Reddy, P.B.; Reddy, M.B.M.; Reddy, R.; Chhajed, S.; Gupta, P.P. Molecular Docking, PKPD, and Assessment of Toxicity of Few Chalcone Analogues as EGFR Inhibitor in Search of Anticancer Agents. Struct. Chem. 2020, 31, 2249–2255. [Google Scholar] [CrossRef]
  177. Sharma, A.; Chakravarti, B.; Gupt, M.P.; Siddiqui, J.A.; Konwar, R.; Tripathi, R.P. Synthesis and Anti Breast Cancer Activity of Biphenyl Based Chalcones. Bioorg. Med. Chem. 2010, 18, 4711–4720. [Google Scholar] [CrossRef]
  178. Lu, C.-F.; Wang, S.-H.; Pang, X.-J.; Zhu, T.; Li, H.-L.; Li, Q.-R.; Li, Q.-Y.; Gu, Y.-F.; Mu, Z.-Y.; Jin, M.-J.; et al. Synthesis and biological evaluation of amino chalcone derivatives as antiproliferative agents. Molecules 2020, 25, 5530. [Google Scholar] [CrossRef] [PubMed]
  179. Kumar, S.K.; Hager, E.; Pettit, C.; Gurulingappa, H.; Davidson, N.E.; Khan, S.R. Design, synthesis, and evaluation of novel boronic-chalcone derivatives as antitumor agents. J. Med.Chem. 2003, 46, 2813–2815. [Google Scholar] [CrossRef] [PubMed]
  180. Ngameni, B.; Cedric, K.; Mbaveng, A.T.; Erdoğan, M.; Simo, I.; Kuete, V.; Daştan, A. Design, Synthesis, Characterization, and Anticancer Activity of a Novel Series of O-Substituted Chalcone Derivatives. Bioorg. Med. Chem. Lett. 2021, 35, 127827. [Google Scholar] [CrossRef]
  181. Kumar, D.; Raj, K.K.; Malhotra, S.V.; Rawat, D.S. Synthesis and anticancer activity evaluation of resveratrol–chalcone conjugates. MedChemComm 2014, 5, 528–535. [Google Scholar] [CrossRef]
  182. Saxena, H.O.; Faridi, U.; Kumar, J.K.; Luqman, S.; Darokar, M.P.; Shanker, K.; Chanotiya, C.S.; Gupta, M.M.; Negi, A.S. Synthesis of Chalcone Derivatives on Steroidal Framework and Their Anticancer Activities. Steroids 2007, 72, 892–900. [Google Scholar] [CrossRef]
  183. Shenvi, S.; Kumar, K.; Hatti, K.S.; Rijesh, K.; Diwakar, L.; Reddy, G.C. Synthesis, Anticancer and Antioxidant Activities of 2,4,5-Trimethoxy Chalcones and Analogues from Asaronaldehyde: Structure–Activity Relationship. Eur. J. Med. Chem. 2013, 62, 435–442. [Google Scholar] [CrossRef]
  184. Hsieh, C.Y.; Ko, P.W.; Chang, Y.J.; Kapoor, M.; Liang, Y.C.; Chu, H.L.; Lin, H.H.; Horng, J.C.; Hsu, M.H. Design and synthesis of benzimidazole-chalcone derivatives as potential anticancer agents. Molecules 2019, 24, 3259. [Google Scholar] [CrossRef] [Green Version]
  185. Kamal, A.; Reddy, J.S.; Ramaiah, M.J.; Dastagiri, D.; Bharathi, E.V.; Sagar, M.V.P.; Pushpavalli, S.N.C.V.L.; Ray, P.; Pal-Bhadra, M. Design, synthesis and biological evaluation of imidazopyridine/pyrimidine-chalcone derivatives as potential anticancer agents. MedChemComm 2010, 1, 355–360. [Google Scholar] [CrossRef]
  186. Farghaly, T.A.; Masaret, G.S.; Muhammad, Z.A.; Harras, M.F. Discovery of Thiazole-Based-Chalcones and 4-Hetarylthiazoles as Potent Anticancer Agents: Synthesis, Docking Study and Anticancer Activity. Bioorg. Chem. 2020, 98, 103761. [Google Scholar] [CrossRef] [PubMed]
  187. Qin, H.L.; Shang, Z.P.; Jantan, I.; Tan, O.U.; Hussain, M.A.; Sher, M.; Bukhari, S.N.A. Molecular Docking Studies and Biological Evaluation of Chalcone Based Pyrazolines as Tyrosinase Inhibitors and Potential Anticancer Agents. RSC Adv. 2015, 5, 46330–46338. [Google Scholar] [CrossRef]
  188. Sankappa Rai, U.; Isloor, A.M.; Shetty, P.; Pai, K.S.R.; Fun, H.K. Synthesis and in Vitro Biological Evaluation of New Pyrazole Chalcones and Heterocyclic Diamides as Potential Anticancer Agents. Arab. J. Chem. 2015, 8, 317–321. [Google Scholar] [CrossRef]
  189. Madhavi, S.; Sreenivasulu, R.; Yazala, J.P.; Raju, R.R. Synthesis of Chalcone Incorporated Quinazoline Derivatives as Anticancer Agents. Saudi Pharm. J. 2017, 25, 275–279. [Google Scholar] [CrossRef] [Green Version]
  190. Mirzaei, S.; Hadizadeh, F.; Eisvand, F.; Mosaffa, F.; Ghodsi, R. Synthesis, Structure-Activity Relationship and Molecular Docking Studies of Novel Quinoline-Chalcone Hybrids as Potential Anticancer Agents and Tubulin Inhibitors. J. Mol. Struct. 2020, 1202, 127310. [Google Scholar] [CrossRef]
  191. Kamal, A.; Srinivasulu, V.; Nayak, V.L.; Sathish, M.; Shankaraiah, N.; Bagul, C.; Reddy, N.V.S.; Rangaraj, N.; Nagesh, N. Design and Synthesis of C3-Pyrazole/Chalcone-Linked Beta-Carboline Hybrids: Antitopoisomerase I, DNA-Interactive, and Apoptosis-Inducing Anticancer Agents. ChemMedChem 2014, 9, 2084–2098. [Google Scholar] [CrossRef]
  192. Alswah, M.; Bayoumi, A.H.; Elgamal, K.; Elmorsy, A.; Ihmaid, S.; Ahmed, H.E. Design, synthesis and cytotoxic evaluation of novel chalcone derivatives bearing triazolo [4,3-a]-quinoxaline moieties as potent anticancer agents with dual EGFR kinase and tubulin polymerization inhibitory effects. Molecules 2017, 23, 48. [Google Scholar] [CrossRef] [Green Version]
  193. Suma, V.R.; Sreenivasulu, R.; Rao, M.V.B.; Subramanyam, M.; Ahsan, M.J.; Alluri, R.; Rao, K.R.M. Design, Synthesis, and Biological Evaluation of Chalcone-Linked Thiazole-Imidazopyridine Derivatives as Anticancer Agents. Med. Chem. Res. 2020, 29, 1643–1654. [Google Scholar] [CrossRef]
  194. Ammar, Y.A.; Fayed, E.A.; Bayoumi, A.H.; Ezz, R.R.; Alsaid, M.S.; Soliman, A.M.; Ghorab, M.M. New Chalcones Bearing Isatin Scaffold: Synthesis, Molecular Modeling and Biological Evaluation as Anticancer Agents. Res. Chem. Intermed. 2017, 43, 6765–6786. [Google Scholar] [CrossRef]
  195. Bagul, C.; Rao, G.K.; Makani, V.K.K.; Tamboli, J.R.; Pal-Bhadra, M.; Kamal, A. Synthesis and biological evaluation of chalcone-linked pyrazolo [1,5-a] pyrimidines as potential anticancer agents. MedChemComm 2017, 8, 1810–1816. [Google Scholar] [CrossRef] [PubMed]
  196. Nelson, G.; Alam, M.A.; Atkinson, T.; Gurrapu, S.; Sravan Kumar, J.; Bicknese, C.; Johnson, J.L.; Williams, M. Synthesis and Evaluation of P-N,N-Dialkyl Substituted Chalcones as Anti-Cancer Agents. Med. Chem. Res. 2013, 22, 4610–4614. [Google Scholar] [CrossRef]
  197. Al Zahrani, N.A.; El-Shishtawy, R.M.; Elaasser, M.M.; Asiri, A.M. Synthesis of novel chalcone-based phenothiazine derivatives as antioxidant and anticancer agents. Molecules 2020, 25, 4566. [Google Scholar] [CrossRef] [PubMed]
  198. Peerzada, M.N.; Khan, P.; Ahmad, K.; Hassan, M.I.; Azam, A. Synthesis, Characterization and Biological Evaluation of Tertiary Sulfonamide Derivatives of Pyridyl-Indole Based Heteroaryl Chalcone as Potential Carbonic Anhydrase IX Inhibitors and Anticancer Agents. Eur. J. Med. Chem. 2018, 155, 13–23. [Google Scholar] [CrossRef]
  199. Mohamed, M.F.; Mohamed, M.S.; Shouman, S.A.; Fathi, M.M.; Abdelhamid, I.A. Synthesis and Biological Evaluation of a Novel Series of Chalcones Incorporated Pyrazole Moiety as Anticancer and Antimicrobial Agents. Appl. Biochem. Biotechnol. 2012, 168, 1153–1162. [Google Scholar] [CrossRef]
  200. Pragathi, Y.J.; Veronica, D.; Anitha, K.; Rao, M.V.B.; Raju, R.R. Synthesis and Biological Evaluation of Chalcone Derivatives of 1,2,4-Thiadiazol-Benzo[d]Imidazol-2-Yl)Quinolin-2(1H)-One as Anticancer Agents. Chem. Data Collect. 2020, 30, 100556. [Google Scholar] [CrossRef]
  201. Chauhan, S.S.; Singh, A.K.; Meena, S.; Lohani, M.; Singh, A.; Arya, R.K.; Cheruvu, S.H.; Sarkar, J.; Gayen, J.R.; Datta, D.; et al. Synthesis of Novel β-Carboline Based Chalcones with High Cytotoxic Activity against Breast Cancer Cells. Bioorg. Med. Chem. Lett. 2014, 24, 2820–2824. [Google Scholar] [CrossRef]
  202. Gurrapu, N.; Praveen Kumar, E.; Kolluri, P.K.; Putta, S.; Sivan, S.K.; Subhashini, N.J.P. Synthesis, Biological Evaluation and Molecular Docking Studies of Novel 1,2,3-Triazole Tethered Chalcone Hybrids as Potential Anticancer Agents. J. Mol. Struct. 2020, 1217, 128356. [Google Scholar] [CrossRef]
  203. Coşkun, D.; Tekin, S.; Sandal, S.; Coşkun, M.F. Synthesis, Characterization, and Anticancer Activity of New Benzofuran Substituted Chalcones. J. Chem. 2016, 2016, 7678486. [Google Scholar] [CrossRef] [Green Version]
  204. Djemoui, A.; Naouri, A.; Ouahrani, M.R.; Djemoui, D.; Lahcene, S.; Lahrech, M.B.; Boukenna, L.; Albuquerque, H.M.T.; Saher, L.; Rocha, D.H.A.; et al. A Step-by-Step Synthesis of Triazole-Benzimidazole-Chalcone Hybrids: Anticancer Activity in Human Cells+. J. Mol. Struct. 2020, 1204, 127487. [Google Scholar] [CrossRef]
  205. Dong, J.; Huang, G.; Zhang, Q.; Wang, Z.; Cui, J.; Wu, Y.; Meng, Q.; Li, S. Development of benzochalcone derivatives as selective CYP1B1 inhibitors and anticancer agents. MedChemComm 2019, 10, 1606–1614. [Google Scholar] [CrossRef] [PubMed]
  206. Jayashree, B.S.; Patel, H.H.; Mathew, N.S.; Nayak, Y. Synthesis of Newer Piperidinyl Chalcones and Their Anticancer Activity in Human Cancer Cell Lines. Res. Chem. Intermed. 2016, 42, 3673–3688. [Google Scholar] [CrossRef]
  207. Khan, N.S.; Khan, P.; Ansari, M.F.; Srivastava, S.; Hasan, G.M.; Husain, M.; Hassan, M.I. Thienopyrimidine-Chalcone Hybrid Molecules Inhibit Fas-Activated Serine/Threonine Kinase: An Approach to Ameliorate Antiproliferation in Human Breast Cancer Cells. Mol. Pharm. 2018, 15, 4173–4189. [Google Scholar] [CrossRef] [PubMed]
  208. Mphahlele, M.J.; Maluleka, M.M.; Parbhoo, N.; Malindisa, S.T. Synthesis, Evaluation for Cytotoxicity and Molecular Docking Studies of Benzo[c]Furan-Chalcones for Potential to Inhibit Tubulin Polymerization and/or EGFR-Tyrosine Kinase Phosphorylation. Int. J. Mol. Sci. 2018, 19, 2552. [Google Scholar] [CrossRef] [Green Version]
  209. Tatsuzaki, J.; Bastow, K.F.; Nakagawa-Goto, K.; Nakamura, S.; Itokawa, H.; Lee, K.H. Dehydrozingerone, Chalcone, and Isoeugenol Analogues as in Vitro Anticancer Agents. J. Nat. Prod. 2006, 69, 1445–1449. [Google Scholar] [CrossRef] [Green Version]
  210. Wan, M.; Xu, L.; Hua, L.; Li, A.; Li, S.; Lu, W.; Pang, Y.; Cao, C.; Liu, X.; Jiao, P. Synthesis and Evaluation of Novel Isoxazolyl Chalcones as Potential Anticancer Agents. Bioorg. Chem. 2014, 54, 38–43. [Google Scholar] [CrossRef]
  211. Kumar, D.; Maruthi Kumar, N.; Tantak, M.P.; Ogura, M.; Kusaka, E.; Ito, T. Synthesis and Identification of α-Cyano Bis(Indolyl)Chalcones as Novel Anticancer Agents. Bioorg. Med. Chem. Lett. 2014, 24, 5170–5174. [Google Scholar] [CrossRef]
  212. Rahimzadeh Oskuei, S.; Mirzaei, S.; Reza Jafari-Nik, M.; Hadizadeh, F.; Eisvand, F.; Mosaffa, F.; Ghodsi, R. Design, Synthesis and Biological Evaluation of Novel Imidazole-Chalcone Derivatives as Potential Anticancer Agents and Tubulin Polymerization Inhibitors. Bioorg. Chem. 2021, 112, 104904. [Google Scholar] [CrossRef]
  213. Bandgar, B.P.; Gawande, S.S. Synthesis and Biological Screening of a Combinatorial Library of β-Chlorovinyl Chalcones as Anticancer, Anti-Inflammatory and Antimicrobial Agents. Bioorg. Med. Chem. 2010, 18, 2060–2065. [Google Scholar] [CrossRef]
  214. Kamal, A.; Mallareddy, A.; Suresh, P.; Shaik, T.B.; Lakshma Nayak, V.; Kishor, C.; Shetti, R.V.C.R.N.C.; Sankara Rao, N.; Tamboli, J.R.; Ramakrishna, S.; et al. Synthesis of Chalcone-Amidobenzothiazole Conjugates as Antimitotic and Apoptotic Inducing Agents. Bioorg. Med. Chem. 2012, 20, 3480–3492. [Google Scholar] [CrossRef]
  215. Zhao, L.; Mao, L.; Hong, G.; Yang, X.; Liu, T. Design, Synthesis and Anticancer Activity of Matrine–1H-1,2,3-Triazole–Chalcone Conjugates. Bioorg. Med. Chem. Lett. 2015, 25, 2540–2544. [Google Scholar] [CrossRef] [PubMed]
  216. Wang, Y.; Hedblom, A.; Koerner, S.K.; Li, M.; Jernigan, F.E.; Wegiel, B.; Sun, L. Novel Synthetic Chalcones Induce Apoptosis in the A549 Non-Small Cell Lung Cancer Cells Harboring a KRAS Mutation. Bioorg. Med. Chem. Lett. 2016, 26, 5703–5706. [Google Scholar] [CrossRef] [PubMed]
  217. Zhu, C.; Wang, R.; Zheng, W.; Chen, D.; Yue, X.; Cao, Y.; Qin, W.; Sun, H.; Wang, Y.; Liu, Z.; et al. Synthesis and Evaluation of Anticancer Activity of BOC26P, an Ortho-Aryl Chalcone Sodium Phosphate as Water-Soluble Prodrugs in Vitro and in Vivo. Biomed. Pharmacother. 2017, 96, 551–562. [Google Scholar] [CrossRef] [PubMed]
  218. Park, S.; Kim, E.H.; Kim, J.; Kim, S.H.; Kim, I. Biological Evaluation of Indolizine-Chalcone Hybrids as New Anticancer Agents. Eur. J. Med. Chem. 2018, 144, 435–443. [Google Scholar] [CrossRef]
  219. Jain, U.K.; Bhatia, R.K.; Rao, A.R.; Singh, R.; Saxena, A.K.; Sehar, I. Design and Development of Halogenated Chalcone Derivatives as Potential Anticancer Agents. Trop. J. Pharm. Res. 2014, 13, 73–80. [Google Scholar] [CrossRef]
  220. de Vasconcelos, A.; Campos, V.F.; Nedel, F.; Seixas, F.K.; Dellagostin, O.A.; Smith, K.R.; de Pereira, C.M.P.; Stefanello, F.M.; Collares, T.; Barschak, A.G. Cytotoxic and Apoptotic Effects of Chalcone Derivatives of 2-Acetyl Thiophene on Human Colon Adenocarcinoma Cells. Cell Biochem. Funct. 2013, 31, 289–297. [Google Scholar] [CrossRef]
  221. Gupta, S.; Maurya, P.; Upadhyay, A.; Kushwaha, P.; Krishna, S.; Siddiqi, M.I.; Sashidhara, K.V.; Banerjee, D. Synthesis and Bio-Evaluation of Indole-Chalcone Based Benzopyrans as Promising Antiligase and Antiproliferative Agents. Eur. J. Med. Chem. 2018, 143, 1981–1996. [Google Scholar] [CrossRef]
  222. Wu, J.-Z.; Cheng, C.-C.; Shen, L.-L.; Wang, Z.-K.; Wu, S.-B.; Li, W.-L.; Chen, S.-H.; Zhou, R.-P.; Qiu, P.-H. Synthetic chalcones with potent antioxidant ability on H2O2-induced apoptosis in PC12 cells. Int. J. Mol.Sci. 2014, 15, 18525–18539. [Google Scholar] [CrossRef] [Green Version]
  223. Singh, A.; Fong, G.; Liu, J.; Wu, Y.-H.; Chang, K.; Park, W.; Kim, J.; Tam, C.; Cheng, L.W.; Land, K.M.; et al. Synthesis and preliminary antimicrobial analysis of isatin–ferrocene and isatin–ferrocenyl chalcone conjugates. ACS Omega 2018, 3, 5808–5813. [Google Scholar] [CrossRef]
  224. Kurt, B.Z.; Ozten Kandas, N.; Dag, A.; Sonmez, F.; Kucukislamoglu, M. Synthesis and Biological Evaluation of Novel Coumarin-Chalcone Derivatives Containing Urea Moiety as Potential Anticancer Agents. Arab. J. Chem. 2020, 13, 1120–1129. [Google Scholar] [CrossRef]
  225. Xu, F.; Li, W.; Shuai, W.; Yang, L.; Bi, Y.; Ma, C.; Yao, H.; Xu, S.; Zhu, Z.; Xu, J. Design, Synthesis and Biological Evaluation of Pyridine-Chalcone Derivatives as Novel Microtubule-Destabilizing Agents. Eur. J. Med. Chem. 2019, 173, 1–14. [Google Scholar] [CrossRef] [PubMed]
  226. Stanojković, T.; Marković, V.; Matić, I.Z.; Mladenović, M.P.; Petrović, N.; Krivokuća, A.; Petković, M.; Joksović, M.D. Highly Selective Anthraquinone-Chalcone Hybrids as Potential Antileukemia Agents. Bioorg. Med. Chem. Lett. 2018, 28, 2593–2598. [Google Scholar] [CrossRef] [PubMed]
  227. Reddy, M.V.R.; Pallela, V.R.; Cosenza, S.C.; Mallireddigari, M.R.; Patti, R.; Bonagura, M.; Truongcao, M.; Akula, B.; Jatiani, S.S.; Reddy, E.P. Design, Synthesis and Evaluation of (E)-α-Benzylthio Chalcones as Novel Inhibitors of BCR-ABL Kinase. Bioorg. Med. Chem. 2010, 18, 2317–2326. [Google Scholar] [CrossRef]
  228. Ferrer, R.; Lobo, G.; Gamboa, N.; Rodrigues, J.; Abramjuk, C.; Jung, K.; Lein, M.; Charris, J.E. Synthesis of [(7-Chloroquinolin-4-Yl)Amino]Chalcones: Potential Antimalarial and Anticancer Agents. Sci. Pharm. 2009, 77, 725–742. [Google Scholar] [CrossRef] [Green Version]
  229. Kumar, D.; Kumar, N.M.; Akamatsu, K.; Kusaka, E.; Harada, H.; Ito, T. Synthesis and Biological Evaluation of Indolyl Chalcones as Antitumor Agents. Bioorg. Med. Chem. Lett. 2010, 20, 3916–3919. [Google Scholar] [CrossRef]
  230. Szliszka, E.; Czuba, Z.P.; Mazur, B.; Sedek, L.; Paradysz, A.; Krol, W. Chalcones enhance TRAIL-induced apoptosis in prostate cancer cells. Int. J. Mol. Sci. 2009, 11, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Kamal, A.; Reddy, V.S.; Santosh, K.; Bharath Kumar, G.; Shaik, A.B.; Mahesh, R.; Chourasiya, S.S.; Sayeed, I.B.; Kotamraju, S. Synthesis of Imidazo[2,1-b][1,3,4]Thiadiazole–Chalcones as Apoptosis Inducing Anticancer Agents. Medchemcomm 2014, 5, 1718–1723. [Google Scholar] [CrossRef]
  232. Wu, C.M.; Lin, K.W.; Teng, C.H.; Huang, A.M.; Chen, Y.C.; Yen, M.H.; Wu, W.B.; Pu, Y.S.; Lin, C.N. Chalcone derivatives inhibit human platelet aggregation and inhibit growth in human bladder cancer cells. Biol. Pharm. Bull. 2014, 37, 1191–1198. [Google Scholar] [CrossRef] [Green Version]
  233. Gargantilla, M.; López-Fernández, J.; Camarasa, M.J.; Persoons, L.; Daelemans, D.; Priego, E.M.; Pérez-Pérez, M.J. Inhibition of XPO-1 Mediated Nuclear Export through the Michael-Acceptor Character of Chalcones. Pharmaceuticals 2021, 14, 1131. [Google Scholar] [CrossRef]
  234. Muchtaridi, M.; Syahidah, H.N.; Subarnas, A.; Yusuf, M.; Bryant, S.D.; Langer, T. Molecular Docking and 3D-Pharmacophore Modeling to Study the Interactions of Chalcone Derivatives with Estrogen Receptor Alpha. Pharmaceuticals 2017, 10, 81. [Google Scholar] [CrossRef] [Green Version]
  235. Alam, M.J.; Alam, O.; Perwez, A.; Rizvi, M.A.; Naim, M.J.; Naidu, V.G.M.; Imran, M.; Ghoneim, M.M.; Alshehri, S.; Shakeel, F. Design, Synthesis, Molecular Docking, and Biological Evaluation of Pyrazole Hybrid Chalcone Conjugates as Potential Anticancer Agents and Tubulin Polymerization Inhibitors. Pharmaceuticals 2022, 15, 280. [Google Scholar] [CrossRef] [PubMed]
  236. Chen, J.; Yan, J.; Hu, J.; Pang, Y.; Huang, L.; Li, X. Synthesis, Biological Evaluation and Mechanism Study of Chalcone Analogues as Novel Anti-Cancer Agents. RSC Adv. 2015, 5, 68128–68135. [Google Scholar] [CrossRef]
  237. Bandgar, B.P.; Gawande, S.S.; Bodade, R.G.; Totre, J.V.; Khobragade, C.N. Synthesis and Biological Evaluation of Simple Methoxylated Chalcones as Anticancer, Anti-Inflammatory and Antioxidant Agents. Bioorg. Med. Chem. 2010, 18, 1364–1370. [Google Scholar] [CrossRef]
  238. Venkatarao, V.; Kumar, L.; Jha, A.; Sridhar, G. Synthesis and Biological Evaluation of Chalcone Fused Quinoline Derivatives as Anticancer Agents. Chem. Data Collect. 2019, 22, 100236. [Google Scholar] [CrossRef]
  239. Konidala, S.K.; Kotra, V.; Danduga, R.C.S.R.; Kola, P.K. Coumarin-Chalcone Hybrids Targeting Insulin Receptor: Design, Synthesis, Anti-Diabetic Activity, and Molecular Docking. Bioorg. Chem. 2020, 104, 104207. [Google Scholar] [CrossRef]
  240. Adelusi, T.I.; Du, L.; Chowdhury, A.; Xiaoke, G.; Lu, Q.; Yin, X. Signaling Pathways and Proteins Targeted by Antidiabetic Chalcones. Life Sci. 2021, 284, 118982. [Google Scholar] [CrossRef] [PubMed]
  241. Kaushal, R.; Kaur, M. Bio-Medical Potential of Chalcone Derivatives and Their Metal Complexes as Antidiabetic Agents: A Review. J. Coord. Chem. 2021, 74, 725–742. [Google Scholar] [CrossRef]
  242. Cai, C.Y.; Rao, L.; Rao, Y.; Guo, J.X.; Xiao, Z.Z.; Cao, J.Y.; Huang, Z.S.; Wang, B. Analogues of Xanthones—Chalcones and Bis-Chalcones as α-Glucosidase Inhibitors and Anti-Diabetes Candidates. Eur. J. Med. Chem. 2017, 130, 51–59. [Google Scholar] [CrossRef]
  243. Shin, J.; Jang, M.G.; Park, J.C.; Koo, Y.D.; Lee, J.Y.; Park, K.S.; Chung, S.S.; Park, K. Antidiabetic Effects of Trihydroxychalcone Derivatives via Activation of AMP-Activated Protein Kinase. J. Ind. Eng. Chem. 2018, 60, 177–184. [Google Scholar] [CrossRef]
  244. Shukla, P.; Satyanarayana, M.; Verma, P.C.; Tiwari, J.; Dwivedi, A.P.; Srivastava, R.; Rehuja, N.; Srivastava, S.P.; Gautam, S.; Tamrakar, A.K.; et al. Chalcone-based aryloxypropanolamine as a potential antidiabetic and antidyslipidaemic agent. Curr. Sci. 2017, 112, 1675–1689. [Google Scholar] [CrossRef]
  245. Mahapatra, D.K.; Asati, V.; Bharti, S.K. Chalcones and Their Therapeutic Targets for the Management of Diabetes: Structural and Pharmacological Perspectives. Eur. J. Med. Chem. 2015, 92, 839–865. [Google Scholar] [CrossRef] [PubMed]
  246. Shinde, R.S.; Salunke, S.D. Facile synthesis of some triazine based chalcones as potential antioxidant and anti-diabetic agents. J. Chem. Pharm. 2015, 7, 114–120. [Google Scholar] [CrossRef]
  247. Hsieh, C.T.; Hsieh, T.J.; El-Shazly, M.; Chuang, D.W.; Tsai, Y.H.; Yen, C.T.; Wu, S.F.; Wu, Y.C.; Chang, F.R. Synthesis of Chalcone Derivatives as Potential Anti-Diabetic Agents. Bioorg. Med. Chem. Lett. 2012, 22, 3912–3915. [Google Scholar] [CrossRef] [PubMed]
  248. Chinthala, Y.; Thakur, S.; Tirunagari, S.; Chinde, S.; Domatti, A.K.; Arigari, N.K.; Srinivas, K.V.N.S.; Alam, S.; Jonnala, K.K.; Khan, F.; et al. Synthesis, Docking and ADMET Studies of Novel Chalcone Triazoles for Anti-Cancer and Anti-Diabetic Activity. Eur. J. Med. Chem. 2015, 93, 564–573. [Google Scholar] [CrossRef] [PubMed]
  249. Tarozzi, A.; Królicka, E.; Kie, K.; Ła, D. Chalcones as Potential Ligands for the Treatment of Parkinson’s Disease. Pharmaceuticals 2022, 15, 847. [Google Scholar] [CrossRef]
  250. Hammuda, A.; Shalaby, R.; Rovida, S.; Edmondson, D.E.; Binda, C.; Khalil, A. Design and Synthesis of Novel Chalcones as Potent Selective Monoamine Oxidase-B Inhibitors. Eur. J. Med. Chem. 2016, 114, 162–169. [Google Scholar] [CrossRef]
  251. Mathew, B.; Parambi, D.G.T.; Mathew, G.E.; Uddin, M.S.; Inasu, S.T.; Kim, H.; Marathakam, A.; Unnikrishnan, M.K.; Carradori, S. Emerging Therapeutic Potentials of Dual-Acting MAO and AChE Inhibitors in Alzheimer’s and Parkinson’s Diseases. Arch. Pharm. Chem. Life Sci. 2019, 352, 1900177. [Google Scholar] [CrossRef]
  252. Hitge, R.; Smit, S.; Petzer, A.; Petzer, J.P. Evaluation of Nitrocatechol Chalcone and Pyrazoline Derivatives as Inhibitors of Catechol-O-Methyltransferase and Monoamine Oxidase. Bioorg. Med. Chem. Lett. 2020, 30, 127188. [Google Scholar] [CrossRef]
  253. Parambi, D.G.T.; Saleem, U.; Shah, M.A.; Anwar, F.; Ahmad, B.; Manzar, A.; Itzaz, A.; Harilal, S.; Uddin, M.S.; Kim, H.; et al. Exploring the Therapeutic Potentials of Highly Selective Oxygenated Chalcone Based MAO-B Inhibitors in a Haloperidol-Induced Murine Model of Parkinson’s Disease. Neurochem. Res. 2020, 45, 2786–2799. [Google Scholar] [CrossRef]
  254. Sasidharan, R.; Manju, S.L.; Uçar, G.; Baysal, I.; Mathew, B. Identification of Indole-Based Chalcones: Discovery of a Potent, Selective, and Reversible Class of MAO-B Inhibitors. Arch. Pharm. 2016, 349, 627–637. [Google Scholar] [CrossRef]
  255. Sasidharan, R.; Baek, S.C.; Sreedharannair Leelabaiamma, M.; Kim, H.; Bijo, M. Imidazole Bearing Chalcones as a New Class of Monoamine Oxidase Inhibitors. Biomed. Pharmacother. 2018, 106, 8–13. [Google Scholar] [CrossRef] [PubMed]
  256. Oh, J.M.; Rangarajan, T.M.; Chaudhary, R.; Singh, R.P.; Singh, M.; Singh, R.P.; Tondo, A.R.; Gambacorta, N.; Nicolotti, O.; Mathew, B.; et al. Novel class of chalcone oxime ethers as potent monoamine oxidase-B and acetylcholinesterase inhibitors. Molecules 2020, 25, 2356. [Google Scholar] [CrossRef] [PubMed]
  257. Padhye, S.; Ahmad, A.; Oswal, N.; Dandawate, P.; Rub, R.A.; Deshpande, J.; Swamy, K.V.; Sarkar, F.H. Fluorinated 2′-Hydroxychalcones as Garcinol Analogs with Enhanced Antioxidant and Anticancer Activities. Bioorg. Med. Chem. Lett. 2010, 20, 5818–5821. [Google Scholar] [CrossRef]
  258. Kim, S.Y.; Lee, I.S.; Moon, A. 2-Hydroxychalcone and Xanthohumol Inhibit Invasion of Triple Negative Breast Cancer Cells. Chem. Interact. 2013, 203, 565–572. [Google Scholar] [CrossRef] [PubMed]
  259. Gul, H.I.; Yamali, C.; Gunesacar, G.; Sakagami, H.; Okudaira, N.; Uesawa, Y.; Kagaya, H. Cytotoxicity, apoptosis, and QSAR studies of phenothiazine derived methoxylated chalcones as anticancer drug candidates. Med. Chem. Res. 2018, 27, 2366–2378. [Google Scholar] [CrossRef]
  260. Prabhakar, V.; Balasubramanian, R.; Sathe, P.; Krishna, C.M.; Juvekar, A. In Vitro Anticancer Activity of Monosubstituted Chalcone Derivatives. Int. J. Tumor Ther. 2014, 3, 1–9. [Google Scholar] [CrossRef]
  261. Schmitt, F.; Draut, H.; Biersack, B.; Schobert, R. Halogenated naphthochalcones and structurally related naphthopyra-zolines with antitumor activity. Bioorg. Med. Chem. Lett. 2016, 26, 5168–5171. [Google Scholar] [CrossRef]
  262. Mai, C.W.; Yaeghoobi, M.; Abd-Rahman, N.; Kang, Y.B.; Pichika, M.R. Chalcones with Electron-Withdrawing and Electron-Donating Substituents: Anticancer Activity against TRAIL Resistant Cancer Cells, Structure–Activity Relationship Analysis and Regulation of Apoptotic Proteins. Eur. J. Med. Chem. 2014, 77, 378–387. [Google Scholar] [CrossRef]
  263. Echeverria, C.; Santibañez, J.F.; Donoso-Tauda, O.; Escobar, C.A.; Ramirez-Tagle, R. Structural Antitumoral Activity Relationships of Synthetic Chalcones. Int. J. Mol. Sci. 2009, 10, 221–231. [Google Scholar] [CrossRef]
  264. Zhuang, C.; Zhang, W.; Sheng, C.; Zhang, W.; Xing, C.; Miao, Z. Chalcone: A Privileged Structure in Medicinal Chemistry. Chem. Rev. 2017, 117, 7762–7810. [Google Scholar] [CrossRef]
  265. Rammohan, A.; Reddy, J.S.; Sravya, G.; Rao, C.N.; Zyryanov, G.V. Chalcone Synthesis, Properties and Medicinal Applications: A Review. Environ. Chem. Lett. 2020, 18, 433–458. [Google Scholar] [CrossRef]
Figure 1. Structure of chalcone and its bioactivities (e.g., anticancer, antibacterial, antiviral, anti-tubercular).
Figure 1. Structure of chalcone and its bioactivities (e.g., anticancer, antibacterial, antiviral, anti-tubercular).
Pharmaceuticals 15 01250 g001
Figure 2. Naturally occurring chalcones: (A) hydroxylated and methoxylated chalcones; (B) chromene-based chalcones; (C) prenylated chalcones; (D) quinone-, cinnamoyl-, and dihydro-bearing chalcones.
Figure 2. Naturally occurring chalcones: (A) hydroxylated and methoxylated chalcones; (B) chromene-based chalcones; (C) prenylated chalcones; (D) quinone-, cinnamoyl-, and dihydro-bearing chalcones.
Pharmaceuticals 15 01250 g002aPharmaceuticals 15 01250 g002b
Scheme 1. Chalcone synthesis via Claisen–Schmidt condensation.
Scheme 1. Chalcone synthesis via Claisen–Schmidt condensation.
Pharmaceuticals 15 01250 sch001
Scheme 2. Various methods to synthesize chalcones.
Scheme 2. Various methods to synthesize chalcones.
Pharmaceuticals 15 01250 sch002
Scheme 3. Synthesis of chalcones by aerobic oxidative cross-coupling reaction.
Scheme 3. Synthesis of chalcones by aerobic oxidative cross-coupling reaction.
Pharmaceuticals 15 01250 sch003
Scheme 4. Synthesis of chalcones by cross-dehydrogenative coupling.
Scheme 4. Synthesis of chalcones by cross-dehydrogenative coupling.
Pharmaceuticals 15 01250 sch004
Scheme 5. Synthesis of substituted chalcones through Julia–Kocienski olefination.
Scheme 5. Synthesis of substituted chalcones through Julia–Kocienski olefination.
Pharmaceuticals 15 01250 sch005
Scheme 6. Synthesis of cis-chalcones by reductive (3+2) annulation.
Scheme 6. Synthesis of cis-chalcones by reductive (3+2) annulation.
Pharmaceuticals 15 01250 sch006
Scheme 7. One-pot synthesis of chalcones.
Scheme 7. One-pot synthesis of chalcones.
Pharmaceuticals 15 01250 sch007
Scheme 8. Synthesis of chalcones via photo-Fries rearrangement.
Scheme 8. Synthesis of chalcones via photo-Fries rearrangement.
Pharmaceuticals 15 01250 sch008
Scheme 9. Wittig reaction for the synthesis of chalcones.
Scheme 9. Wittig reaction for the synthesis of chalcones.
Pharmaceuticals 15 01250 sch009
Figure 3. Structures of some chalcones showing anti-tubercular activity.
Figure 3. Structures of some chalcones showing anti-tubercular activity.
Pharmaceuticals 15 01250 g003
Figure 4. Chalcones with antiviral activity.
Figure 4. Chalcones with antiviral activity.
Pharmaceuticals 15 01250 g004
Figure 5. Structures of chalcones with antimalarial activity.
Figure 5. Structures of chalcones with antimalarial activity.
Pharmaceuticals 15 01250 g005
Figure 6. Chalcones with anti-bacterial activity: (A) homocyclic chalcones; (B) heterocyclic chalcones.
Figure 6. Chalcones with anti-bacterial activity: (A) homocyclic chalcones; (B) heterocyclic chalcones.
Pharmaceuticals 15 01250 g006
Figure 7. Chalcones for Alzheimer’s disease.
Figure 7. Chalcones for Alzheimer’s disease.
Pharmaceuticals 15 01250 g007
Figure 8. Homocyclic chalcones with anti-breast-cancer activity [170,171,172,173,174,175,176,177,178,179,180,181,182,183].
Figure 8. Homocyclic chalcones with anti-breast-cancer activity [170,171,172,173,174,175,176,177,178,179,180,181,182,183].
Pharmaceuticals 15 01250 g008
Figure 9. Heterocyclic chalcones with anti-breast-cancer activity [23,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208].
Figure 9. Heterocyclic chalcones with anti-breast-cancer activity [23,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208].
Pharmaceuticals 15 01250 g009aPharmaceuticals 15 01250 g009b
Figure 10. (A,B) Chalcones with anti-lung-cancer activity and their IC50 values [187,193,200,209,210,211,212,213,214,215,216,217].
Figure 10. (A,B) Chalcones with anti-lung-cancer activity and their IC50 values [187,193,200,209,210,211,212,213,214,215,216,217].
Pharmaceuticals 15 01250 g010aPharmaceuticals 15 01250 g010b
Figure 11. Structures of chalcones with antidiabetic activity.
Figure 11. Structures of chalcones with antidiabetic activity.
Pharmaceuticals 15 01250 g011
Figure 12. Chalcones showing anti-Parkinson’s activity.
Figure 12. Chalcones showing anti-Parkinson’s activity.
Pharmaceuticals 15 01250 g012
Figure 13. Flowchart depicting the structure–activity relationships of natural and synthetic chalcones.
Figure 13. Flowchart depicting the structure–activity relationships of natural and synthetic chalcones.
Pharmaceuticals 15 01250 g013
Table 1. Chalcones with anti-breast-cancer activities.
Table 1. Chalcones with anti-breast-cancer activities.
Homocyclic chalcone derivatives (breast cancer cell lines; IC50 values in µM)
Hydroxylated chalconesMethoxylated chalconesChalcones with extended functionality
1241.371287.31340.03
125-1292.541353.5
1264.413030.55136-
12791.413123.45137<10
--1322.2--
--1331.2--
Heterocyclic chalcone derivatives (breast cancer cell lines; IC50 values in µM)
Furano-based chalconesImidazole-based chalconesThiazole-based chalcones
1382.21428.911451.97
1390.00035–0.591430.561460.18
1401.451445.8914712
1411.8----
Pyrimidine-based chalconesIndole-based chalconesOxazoline/pyrazole/quinoline/pyridine-based chalcones
1487.415131.661540.35
1496.521522.251553.9
1500.14153121562.32
----1571.8
Table 2. Chalcones with diverse anticancer activities.
Table 2. Chalcones with diverse anticancer activities.
Compound NumberStructure of ChalconesTypes of Cancer
173Pharmaceuticals 15 01250 i001Cervical cancer (0.027 ± 0.01 µM), prostate cancer (0.031 ± 0.05 µM), leukemia (0.031 ± 0.12 µM), lung cancer (0.026 ± 0.03 µM) [236]
174Pharmaceuticals 15 01250 i002Lung cancer, colon cancer, renal adenocarcinoma, pancreatic carcinoma [237]
175Pharmaceuticals 15 01250 i003Lung cancer (1.39–3.17 µM), breast cancer (1.97–4.14 µM), hepatocarcinoma (1.56–3.79 µM) [186]
176Pharmaceuticals 15 01250 i004Breast cancer (2.2 ± 0.3 µM), prostate cancer(0.9 ± 0.5 µM), lung cancer (1.10 ± 0.5 µM), pancreatic carcinoma (1.2 ± 0.2 µM) [187]
177Pharmaceuticals 15 01250 i005Lung cancer (0.10–2.90 µM), breast cancer (0.14–0.17 µM), colon adenocarcinomas (0.13–2.89 µM) [189]
178Pharmaceuticals 15 01250 i006Lung cancer (0.66 ± 0.071 µM), breast cancer (0.18 ± 0.094 µM), prostate carcinoma (1.03 ± 0.45 µM) [193]
179Pharmaceuticals 15 01250 i007Breast cancer (3.44 ± 0.19 µM), liver carcinoma (4.64 ± 0.23 µM), adenocarcinoma (6.31 ± 0.27 μM) [171]
180Pharmaceuticals 15 01250 i008Colon cancer (11.78 µM), breast cancer (31.66 µM), liver cancer (13.95 µM) [194]
181Pharmaceuticals 15 01250 i009Breast cancer (3.9–4.1 µM), liver cancer (3.8–5.0 µM), colorectal cancer (3.3 µM) [199]
182Pharmaceuticals 15 01250 i010Breast cancer (2.54 µM), colorectal cancer (1.83 µM), gastric carcinoma (1.52 µM) [178]
183Pharmaceuticals 15 01250 i011Breast cancer (0.012 ± 0.007 µM), lung cancer (0.074 ± 0.004 µM), colon cancer (0.074 ± 0.004 μM), ovarian cancer (0.083 ± 0.002 μM) [200]
184Pharmaceuticals 15 01250 i012Breast cancer (0.33–0.89 µM), melanoma (0.11–1.28 µM), lung cancer (0.34–7.56 µM) [238]
185Pharmaceuticals 15 01250 i013Ovarian carcinoma (6.66 µM), breast cancer (30.55 µM), lung cancer (36.35 µM) [181]
186Pharmaceuticals 15 01250 i014Colon carcinoma (33.31 µM), cervical carcinoma (21.80 µM), breast cancer (23.45 µM), lung cancer (4.28 µM) [206]
187Pharmaceuticals 15 01250 i015Breast cancer, synovial carcinoma, cervical carcinoma (2.2–4.5 µM) [183]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rajendran, G.; Bhanu, D.; Aruchamy, B.; Ramani, P.; Pandurangan, N.; Bobba, K.N.; Oh, E.J.; Chung, H.Y.; Gangadaran, P.; Ahn, B.-C. Chalcone: A Promising Bioactive Scaffold in Medicinal Chemistry. Pharmaceuticals 2022, 15, 1250. https://doi.org/10.3390/ph15101250

AMA Style

Rajendran G, Bhanu D, Aruchamy B, Ramani P, Pandurangan N, Bobba KN, Oh EJ, Chung HY, Gangadaran P, Ahn B-C. Chalcone: A Promising Bioactive Scaffold in Medicinal Chemistry. Pharmaceuticals. 2022; 15(10):1250. https://doi.org/10.3390/ph15101250

Chicago/Turabian Style

Rajendran, Gayathri, Deepu Bhanu, Baladhandapani Aruchamy, Prasanna Ramani, Nanjan Pandurangan, Kondapa Naidu Bobba, Eun Jung Oh, Ho Yun Chung, Prakash Gangadaran, and Byeong-Cheol Ahn. 2022. "Chalcone: A Promising Bioactive Scaffold in Medicinal Chemistry" Pharmaceuticals 15, no. 10: 1250. https://doi.org/10.3390/ph15101250

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