Next Article in Journal
Personalized Treatment for Infantile Ascending Hereditary Spastic Paralysis Based on In Silico Strategies
Next Article in Special Issue
Behavioral Despair Is Blocked by the Flavonoid Chrysin (5,7-Dihydroxyflavone) in a Rat Model of Surgical Menopause
Previous Article in Journal
Self-Organization of Porphyrin–POM Dyads: Nonplanar Diacids and Oxoanions in Low-Dimensional H-Bonding Networks
Previous Article in Special Issue
Plumbagin Suppresses Breast Cancer Progression by Downregulating HIF-1α Expression via a PI3K/Akt/mTOR Independent Pathway under Hypoxic Condition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pharmacotherapeutics Applications and Chemistry of Chalcone Derivatives

1
PAPRSB Institute of Health Sciences, Universiti Brunei Darussalam, Gadong BE1410, Brunei
2
Faculty of Data Science and Information Technology, INTI International University, Nilai 71800, Malaysia
3
Faculty of Engineering, Universiti Teknologi Malaysia, Skudai 81300, Malaysia
4
Centre for Sustainability of Ecosystem and Earth Resources (Pusat ALAM), Universiti Malaysia Pahang, Kuantan 26300, Malaysia
5
Department of Pharmacy Practice, Faculty of Pharmacy, Universitas Airlangga, Surabaya 60115, Indonesia
6
Faculty of Pharmacy, Quest International University, Ipoh 30250, Malaysia
7
Faculty of Integrated Technologies, Universiti Brunei Darussalam, Gadong BE1410, Brunei
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(20), 7062; https://doi.org/10.3390/molecules27207062
Submission received: 22 September 2022 / Revised: 16 October 2022 / Accepted: 17 October 2022 / Published: 19 October 2022

Abstract

:
Chalcones have been well examined in the extant literature and demonstrated antibacterial, antifungal, anti-inflammatory, and anticancer properties. A detailed evaluation of the purported health benefits of chalcone and its derivatives, including molecular mechanisms of pharmacological activities, can be further explored. Therefore, this review aimed to describe the main characteristics of chalcone and its derivatives, including their method synthesis and pharmacotherapeutics applications with molecular mechanisms. The presence of the reactive α,β-unsaturated system in the chalcone’s rings showed different potential pharmacological properties, including inhibitory activity on enzymes, anticancer, anti-inflammatory, antibacterial, antifungal, antimalarial, antiprotozoal, and anti-filarial activity. Changing the structure by adding substituent groups to the aromatic ring can increase potency, reduce toxicity, and broaden pharmacological action. This report also summarized the potential health benefits of chalcone derivatives, particularly antimicrobial activity. We found that several chalcone compounds can inhibit diverse targets of antibiotic-resistance development pathways; therefore, they overcome resistance, and bacteria become susceptible to antibacterial compounds. A few chalcone compounds were more active than conventional antibiotics, like vancomycin and tetracycline. On another note, a series of pyran-fused chalcones and trichalcones can block the NF-B signaling complement system implicated in inflammation, and several compounds demonstrated more potent lipoxygenase inhibition than NSAIDs, such as indomethacin. This report integrated discussion from the domains of medicinal chemistry, organic synthesis, and diverse pharmacological applications, particularly for the development of new anti-infective agents that could be a useful reference for pharmaceutical scientists.

1. Introduction

Chalcone is a collective group of ketones (flavonoids) that has a three-carbon α,β-unsaturated carbonyl group attached to two aromatic rings, namely rings A and B (Figure 1). The numbering system of chalcone shown in Figure 1 is followed throughout this article. Other chemical names of chalcone include benzyl acetophenone or benzylideneacetophenone. They are produced by certain plant species such as Angelica, Glycyrrhiza, Humulus, and Scutellaria as precursors to the biosynthesis of flavonoids and isoflavonoids and intermediates to the synthesis of heterocyclic compounds with biologically interesting properties such as pyrazolines, isoxazoles, cyanopyridines and pyrimidines.
Chalcone is part of plants’ most prominent class of secondary metabolites. This is used in plant defense mechanisms to combat reactive oxygen species for the plant to survive and prevent molecular damage as well as damage caused by microorganisms, insects, and animals [1]. Chalcone can also be chemically synthesized in the laboratory using the Claisen Schmidt or aldol condensation reaction [2]. Chalcone has been reported to exert multiple beneficial properties, such as anti-inflammatory, antibacterial, antifungal, antidiabetic properties, and anticancer activities. It also improves vision, memory, joint and muscle discomfort, liver and kidney function, sleep, prevents cancer, strengthens the immune system, and beautifies skin and hair. [3,4,5].
Chalcone and its derivatives have shown an inhibitory effect against methicillin-resistant Staphylococcus aureus (MRSA) [6] due to the presence of −OH groups in the B ring and the lipophilicity of the A ring (Figure 1) [7,8]. A combination of Chalcones and antibiotics (i.e., oxacillin) has also shown a synergistic effect in treating MRSA infections [9,10,11,12].
Additionally, some chalcone derivatives exhibited antifungal activities, particularly against Microsporum gypseum [13,14,15,16]. They inhibit the β (1, 3)-glucan and chitin synthases responsible for the formation and normal functioning of the fungal cell wall [17,18]. Some chalcone compounds have shown superior antifungal effects compared to ketoconazole, a broad-spectrum oral antifungal agent [13]. Chalcone compounds are also effective inhibitors of inflammatory enzymes, such as cyclooxygenase (COX), lipooxygenase (LOX), interleukins (IL), and prostaglandins (PGs) [19]. The most researched novel series of chalcones revealed their potent inhibitory effects on nitric oxide (NO) formation and the release of glucuronidase and lysozyme, which are responsible for inflammatory responses [20]. Heterocyclic rings and methoxy substitutions on the attached rings of chalcones contribute to the anticancer properties of such chalcone compounds [21,22,23]. The investigated chalcone compounds: isoliquiritigenin, flavokawain, and xanthohumol, showed cytotoxic and apoptosis induction that promotes antitumor activities [23]. However, the mechanisms of the compound are unclear, causing uncertainty about chalcones’ astounding abilities. It is a compound that has attracted the attention of many scientists due to its plethora of therapeutic efficacy.
Numerous studies have been conducted to explore the pharmacological properties of chalcone and its derivatives in recent years. Even though there are a few reviews available, such as antiviral [24,25], preclinical studies [4], synthesis chalcone [26], and antidiabetic [27], comprehensive reviews focusing on the overall purported health benefits, including molecular mechanisms of pharmacological actions of chalcone and its derivatives evaluating the recent progress of therapeutical applications is insufficient. Therefore, herein, we reviewed the research findings to discuss the synthesis of chalcone and its derivatives and assess their pharmacotherapeutic efficacy, focusing on the recent experimental studies. This report integrated discussion from the domains of medicinal chemistry, organic synthesis, and diverse pharmacological applications, particularly for the development of new anti-infective agents that could be a useful reference for pharmaceutical scientists.

2. Chemistry of Chalcone and Its Derivatives

Chalcone and its derivatives have been long used in various traditional medicine systems, including homeopathic and Chinese medicine. They are traditionally prepared by the reaction of benzaldehydes and active methylene ketones under homogeneous conditions using the Claisen-Schmidt condensation and a more recent invention known as the aldol condensation [28]. However, recent discoveries of methods for producing chalcones provide different advantages depending on the type of catalyst, solvent, base, and reaction conditions [29].

2.1. Claisen Schmidt Condensation

The Claisen Schmidt condensation reaction involves an aldehyde with the carbonyl group without hydrogen atoms in the α-position, and a ketone, using a heterogeneous acid catalyst to produce the desired α,β-unsaturated ketone. This is one of the methods of synthesizing chalcone in the laboratory (Figure 2) due to the equimolar quantities of acetophenone and benzaldehyde. Claisen Schmidt condensation uses an aqueous-alcoholic alkali with a (concentration of 10 to 60%) to catalyze the reaction between acetophenone and benzaldehyde by dehydration [30]. The reaction can take place either for 12–15 h at a temperature of 50 °C or for one week at room temperature (20–25 °C).
The heterogeneous acid catalyst is useful in producing chalcones because of the increased purity of end-products, decreased amounts of undesired products, reduced reaction time, and cost-effective procedure. Ionic liquids (ILs) are prepared using this condensation reaction but use a multi-sulfonic acid group ion liquid as the catalyst. ILs have gained interest due to their high catalytic activity, small catalyst usage, easy filtration, recyclable catalyst, and constant catalytic activity (Figure 2). However, the product yield will be decreased due to the reaction conditions, which promotes the Cannizzaro reaction. This redox reaction produces primary alcohol and carboxylic acid from two aldehyde molecules.

2.2. Aldol Condensation

Aldol condensation is another synthetic method commonly used after the Claisen-Schmidt condensation. The aldol condensation reaction (Figure 3), or the solid-state reaction, replaces aldehydes with benzylidene-diacetate and uses heat (200–350 °C) and a base such as potassium hydroxide as a catalyst for the reaction between the two compounds. It uses calcium, barium or strontium hydroxides or carbonates as catalysts in a liquid mixture containing water with a low boiling point that can perform distillation at a constant temperature. Other synthetic reactions can increase the reaction rate using microwave radiation without solvents; it also provides fluorescence emission profiles that can be used as biological markers.
Reacting a ketone with an aldehyde (with the carbonyl group-containing no hydrogen atoms in the α position with an acidic heterogeneous catalyst-activated carbon) will reduce the reaction time, cost, and impurities in the final products. Another technique involves the phase transfer method for synthesizing heterocyclic ring-containing chalcones, which incorporates a third ring into the chalcone’s skeleton chain. The starting constituents used are 1-phenyl-3-aryl-4-formyl pyrazole and acetophenone, with the catalyst being tetrabutylammonium bromide in the presence of an inorganic alkaline solution; the reaction is performed under microwave radiation (Figure 3).

2.3. Synthesisand Chemistry

Another method uses plants such as Sedum jinianum, S. plumbizincicola, S. alfredi and Potentilla griffithii with a high concentration of metals such as zinc, copper, cadmium, and magnesium, and at least one of the metal elements is used as a catalyst. They are first heated and treated with acid, then filtered and purified to be attached to a carrier, becoming a metal catalyst in synthesizing chalcones. Although this method promotes less pollution, it is unconventional to extract metal from these plants to be used as mere catalysts.
A different method uses a fluorine-containing biphasic catalyst as a result of reacting 4-dimethylamino pyridine with fluorinated alkyl iodide to react benzaldehyde with acetophenone at 50 to 100 °C for 1 to 3 h. The reacted compounds were cooled, filtered, distilled, and recrystallized with more than 99% purity. This invention requires an easy processing method, nature-friendly, low synthesis cost, and ease of re-obtaining the fluoric catalyst.
There are methods to synthesize hydroxyflavones that use soluble resin, specifically polyethylene glycol (PEG) that reacted with benzyloxy-2-hydroxy-acetophenone to be used in a reaction with benzaldehyde with a base as a catalyst. The advantages of this method are high accessibility to the reactant compounds and high purity percentage.
Another popular method is the one-pot synthesis. In this process, a direct reaction in one step is utilized to prepare inorganic components, while the organic component operates as a surface capping material or template. This method shortens the time to separate and purify the products and increases the product yield. Another technique of one-pot synthesis is slowly adding primary alcohol with chromium oxide (CrO3) to produce furochalcones, or the addition of cheap catalysts, which are copper salt, 2,2′-bipyridine, and 2,2,6,6-Tetramethylpiperidinyloxy (TEMPO) kept at a temperature of −10 to 100 °C for 10 to 96 h to produce a milder reaction.
A different synthesis method to produce the α,β-unsaturated carbonyl system that does not require condensation reaction offers more direct response with fewer vigor conditions, a cheaper cost of reactants, a more straightforward operating system, and a higher yield of products. The raw materials are the halogenated aromatic hydrocarbons with a ketone in the carbon skeleton and aromatic alkynes, using an alkali, a phosphine ligand, and palladium as catalysts, at a temperature of 60 to 150 °C [31,32].

2.4. Approach to Design of Chalcone Derivatives from the Natural Sources

Different approaches have been used to design chalcone derivatives, as outlined below:

2.4.1. Isoliquiritigenin

It can be extracted from the plant Radix Glycyrrhizae. Isoliquiritigenin is used in cosmetics due to its beneficial effects on the skin, including treatment of skin conditions such as acne, eczema, and irritation, as well as desirable effects, such as skin whitening, anti-aging, and eye drop preparations. There have been claims that isoliquiritigenin can activate the GABAA receptor, bind to the γ-subunit and act as a positive allosteric modulator that elicits similar effects as a benzodiazepine. It also prevents and treats cardio-cerebrovascular diseases [33] (Figure 4).

2.4.2. Licochalcone A

It is present in high concentrations in the plant Glycyrrhiza inflate [34]. Licochalcone A and isoliquiritigenin compounds are used in the cosmetic industry; they are both used in acne treatment and skin whitening. Licochalcone A is also used to prepare skin toner and hair cosmetics. The compound in essential oils produces bath salts to clean pores in the skin, control sebum production and retain skin moisture. Licochalcone A is also used for the treatment, improvement, and prevention of adenosine 5′ monophosphate-activated protein kinase (AMPK)-related diseases, an enzyme involved in metabolism, specifically for lipids [34]. Influenza virus infection-related diseases were treated for other conditions in which licochalcone A was used [35,36,37] (Figure 4).

2.4.3. Xanthoangelol

It is a major component of the plant Angelica keiskei. Xanthoangelol is an isoprenyl-chalcone compound with antioxidant properties that are used to prevent diseases involving lipid metabolism or inflammation [38] (Figure 4).

2.4.4. Isobavachalcone

It can be isolated from Psoralea corylifolia or Piper longum fruits. It belongs to the same chalcone family as xanthoangelol but has different properties [39]. It inhibits melanin formation causing skin whitening [40]. Isobavachalcone is also used to reduce nerve inflammation and inhibit cholesterol absorption [41]. Other uses of isobavachalcone prevent and control diseases [42] (Figure 4).

2.4.5. Xanthohumol

It is present in the Humulus lupulus and belongs to the prenylated chalcone family. Its antioxidant property is a ‘broad spectrum’ cancer chemo-preventive agent. Along with hydroxytyrosol, it is used as a nasal spray for viral infection, allergic reactions, or vasomotor rhinitis of the nasal mucosa [43]. Xanthohumol has demonstrated the inhibitory activity of the enzyme α-glucosidase, an enzyme responsible for carbohydrate metabolism. Therefore, it is used in metabolic diseases such as diabetes and other diseases such as AIDS, osteoporosis, and malignant tumors [44,45] (Figure 4).

2.4.6. Nardoaristolone A

It can be extracted from Nardostachys chinensis and classified as terpenoid chalcones [46]. Like any other chalcones, it has been used to treat different skin cancers. It has systemic effects such as increasing the red blood cell count and aids in small bowel movements. Nardoaristolone A has been used in medications for tuberculosis and endometrium cancers [1,45] (Figure 4).

2.5. Role of Chalcone Moiety in Synthesis of Derivatives

The chalcone moiety can be used to produce other chalcone derivatives such as cyanopyridines, pyrazolines, isoxazoles, and pyrimidines with different heterocyclic ring systems. These derivatives containing hydroxyl, ether, acid, or amino groups have diverse functionality and can be used to produce more complex chalcones. Aminochalcones, a chalcone moiety, can produce benzothiazole chalcones and other derivatives through alkylation, hydrolysis, and esterification or amide formation. Chalcone derivatives can be produced through the Phase Transfer Catalysis through alkylation, using oxygen or sulfur.
Chalcone derivatives can also be produced using the α,β-unsaturated system by substitution of the functional group at the two positions, and dihydrochalcones can be produced by reducing the double bond in the saturated system; these derivatives will be used in the synthesis of pyrazoles and flavonoids as well as other heterocyclic compounds.
A chalcone derivative which is 3-phenyl-1-(4-methyl) phenyl-2-bromo-propylene-1-one, can be produced by reacting ρ-toualdehyde with acetophenone using an inorganic base as a catalyst in an aldol condensation reaction, and the products react with a halogen in an addition reaction followed by an elimination reaction with an inorganic base. Another method also used a heterogeneous catalyst (hexagonal boron nitride h-BN) that is hydrogenated and has a Frustrated Lewis Pair (FLP)-type electronic structure that produces 100% yield.
Another method to obtain dihydrochalcones, specifically phloretin, used bacterial or plant chalcone isomerases and enolate reductase as a catalyst. An electrolytic hydrogenation method using hydrogen, which becomes active by electrolysis of water, is needed, followed by an addition reaction with a ketone without a catalyst to obtain neohesperidin dihydrochalcones. Dihydrochalcones and quinazolinyl derivatives can be produced by substituting carboxylic or nitro groups at the β carbon of the carbon skeleton of chalcones, respectively [47] (Figure 5).

2.5.1. Semi Synthetic Derivatives of Chalcones

Chalcone isocordoin and its semisynthetic derivatives were tested for Anti-inflammatory and anti-hypersensitive effects in mice (Figure 6).

2.5.2. Characterization of Chalcones

The structure of the synthesized chalcones can be characterized by IR, NMR and mass spectroscopy.

2.5.3. UV Spectrum

The UV spectrum of chalcones consists of two essential absorption band: band I and relatively a minor band, band II. In chalcones, band I usually appears in 340–390 nm, although a minor inflection or peak often occurs at 300–320 nm. Band II appears in 220–270 nm.

2.5.4. IR Spectrum

In the IR spectra of chalcones asymmetric and symmetric stretching vibrations of the aromatic C–H bonds are seen at 3120–3080 cm−1 and 3060–3040 cm−1 ranges with two low intensity bands. C–H stretching band of the =C–H group is observed at 3030–3010 cm−1. The bands at 1610–1570 cm−1 are assigned to the vibrations of the aromatic ring. The inplane deformation of the =C–H bond is appeared as broad weak band at 1460–1430 cm−1. The carbonyl stretching vibrations for the enones (=C–C=O) can be found between 1650 and 1685 cm−1.

2.5.5. NMR Spectrum

The 1H-NMR spectrum of double bonds of chalcones were seen at 5.4 and 6.1 ppm. The aromatic regions were observed at 6.9–8.1 ppm.
In 13C-NMR spectrum of chalcones, the carbonyl carbon usually appears between δ 186.6 and 196.8. The α- and β- carbon atoms with respect to the carbonyl group give characteristic signals between δ 116.1–128.1 and δ 136.9–145.4 respectively.

2.5.6. Mass Spectrum

Basic fragmentation pathways of chalcones are obtained by loss of the phenyl group from the A or B ring, and loss of CO.

3. Basic Fragmentation Pathways of Chalcones Are Obtained by Loss of the Phenyl Group from the A or B Ring, and Loss of CO. Pharmacotherapeutic Activities

Chalcone and its derivatives have shown diverse pharmacological activities. A summary of different pharmacological properties with their salient mechanisms of action is shown in Figure 7. The pharmacophore responsible for various activities changes depending on the activity. The details of these functions have been discussed comprehensively in the following sections.

3.1. Anti-Bacterial Agent

Chalcone is an antibacterial agent with moderate to high activity due to the presence of the reactive αβ-unsaturated system. Their flexibility to change their structure by incorporating different types of substituent groups into the aromatic ring can potentially achieve a higher potency, lower toxicity, and a wider spectrum of antibacterial activity [54].
Many chalcone derivatives showed potential antibacterial activities against different pathogenic strains, including antibiotic-resistant bacteria. A summary list is shown in Table 1. Several chalcones were significantly potential than the standard antibiotics (Table 1). For example, compounds 14, 68, 1113, 1519, 2129 (Table 1) have shown strong antibacterial activity against Gram-positive bacteria Staphylococcus aureus and Enterococcus faecalis, and against Gram-negative bacteria Escherichia coli and Salmonella enterica [55]. Gram-positive bacteria were more susceptible to cationic molecules than Gram-negative bacteria. For example, compounds 14 showed the highest activity (MIC ranged 1–2 µg/mL) against Gram-positive bacteria, whereas MIC ranged 2–8 1–2 µg/mL for Gram-negative bacteria. The hydrophobicity of the alkyl chain was responsible for varying strength of antibacterial potentiality. Compounds 7 with medium hydrophobicity exhibited the highest activity against the tested bacteria; however, compounds 610 with different alkyl chain lengths showed different antibacterial sensitivity. The results reported in Table 1 demonstrated that increasing alkyl chain length (i.e., compounds 810) causes decreasing antibacterial activity. Since the long hydrophobic chain has an aggregation tendency, this might cause this decreasing antibacterial activity [56]. Another study reported that 31 compounds, among them, compounds 47, 50, and 51 (Table 1), were more active against the tested bacteria, B. cereus, E. coli, P. aeruginosa, and S. aureus [57]. Another study tested both sulfones and bisulfones chalcones (11 compounds) for their antibacterial activity against Gram-positive strains B. subtilis and S. aureus and Gram-negative strains P. aeruginosa and S. typhimurium [58]. In this study, in comparison to standard antibiotics Ampicillin and Kanamycin, compound 61 was slightly better against B. subtilis and compounds 65, 66, and 67 were significantly potential against S. typhimurium. In another study using monomeric chalcone compounds, there was higher antibacterial activity in Gram-positive bacteria than the Gram-negative bacteria [59]. A screening method of chalcone derivatives against S. aureus, B. subtilis, E. coli, and P. aeruginosa showed increased lipophilic area, and the smaller molecular size of chalcones increased their antimicrobial activity [60].
Chu et al. [55] reported the antibacterial activity of several cationic Chalcones against antibiotic resistance strains, including nine clinical isolates of MRSA, 12 clinical isolated of KPC-2-producing Klebsiella pneumoniae (KPC), and 12 clinical isolated of New Delhi metallo-β-lactamase 1 (NDM-1)-producing Enterobacteriaceae. A few selected molecules (compounds 14, 7, 12, 1619, 26, and 28) were tested against MRSA and found MIC values of 0.25–32 μg/mL (Table 1). Among these, compound 7 showed the lowest MIC 0.25 μg/mL, and exhibited the highest activity against clinical isolates 6 of MRSA. The MICs of the other compounds were primarily 0.5–8 μg/mL [55].
For KPC, compounds 17, 26, and 28 exhibited the best antibacterial activity (MIC 2 μg/mL), which was equivalent to vancomycin. However, compound 3 showed average activity MIC ranging from 4–8 μg/mL. The MIC for NDM ranged mainly from 1–16 μg/mL; however, among them, compound 28 exhibited the highest activity (MIC: 2–8 μg/mL) [55]. Overall, compound 28 was highly effective against KPC and NDM, whereas compound 7 was for MRSA. Furthermore, the antibacterial activity of the tested molecules against NDM was better than the KPC.

3.2. Antibacterial Mechanisms

Chalcone and its derivatives have the potential to target diverse receptors. This diverse inhibition process includes (i) efflux pump inhibitor (EPI), (ii) type II fatty acid biosynthetic pathway (FAS-II), (iii) DNA replication, (iv) filamentous temperature-sensitive mutant Z (FtsZ), (v) virulence factor, and (vi) protein tyrosine phosphatases (PTPs). A schematic representation of the antibacterial mechanisms of chalcone and its derivatives is shown in Figure 8. The details mechanisms can be found elsewhere (see review [5]).
Chalcones were the potential to exhibit cytoplasmic membrane depolarization and inner membrane permeabilization. Compounds 14, 7, 11, 12, 1619 were used to evaluate the ability of the cytoplasmic membrane depolarization and the inner membrane permeabilization against S. aureus and E. coli to understand their mode of action [55]. The highest cytoplasmic membrane depolarization activity of S. aureus was exhibited by compounds 3, 7, and 18, whereas compounds 11, 1, and 7 showed for E. coli. Both S. aureus and E. coli cytoplasmic membrane depolarization was at least level by the other compounds.
The ability of bacterial efflux pumps to rapidly export antibacterial drugs makes them sustainable against drugs. Even though the origin of antibiotic resistance is numerous and complex, efflux pumps are one of the critical class of resistance determinants [62,63,64,65,66,67,68,69]. About 117 Chalcone molecule was tested against the NorA efflux pump of S. aureus and found at least 20 effective compounds that were able to inhibit this efflux pump. However, five compounds were active, and among them, two compounds, 4-phenoxy-4′-dimethylaminoethoxychalcone and 4-dimethylamino-4′- dimethylaminoethoxychalcone were highly active, which were equipotent to reserpine with IC50-values of 9.0 and 7.7 lM, respectively. Additionally, three compounds synergistically increased the effect of ciprofloxacin on S. aureus, of which 4-phenoxy-4′-dimethylaminoethoxychalcone exhibited a fourfold higher activity at 3.13 μg/mL, being twice as potent as reserpine [70]. Additionally, the ability to inhibition of efflux pump was concentration-dependent. Higher molecule concentration increases many-fold inhibition of efflux pump.
Chalcone derivatives, particularly trans-3-(1H-indol-3-yl)-1-(4′-benzyloxyphenyl)-2-propen-1-one, 1-(4″-biphenyl)-3- (3′4′-dihydroxyphenyl)-2-propen-1-one, 1-(4″-hydroxy-3″- methylphenyl)-3-(4′-hydroxyphenyl)-2-propen-1-one, 3-(4′–chlorophenyl)-1-(4″-hydroxyphenyl)-2-propen-1-one and LTG-oxime showed effect on clinical isolates of MRSA by modulating the bacterial efflux pump. These compounds exhibit synergistic interaction with antibiotics– norfloxacin under both in vitro and in vivo conditions [71]. In short, chalcone and its derivatives exhibited antibacterial actions via various targets, which mainly include efflux pump inhibitory (EPI), interfering DNA replication, and filamentous temperature-sensitive mutant Z (FtsZ).
Condensed pyrimidine derivatives, ring-fused chalcones, and flavanones were also reported to have antibacterial activity. In treating certain oral infections caused by bacteria, chalcones were included along with the antibiotics, which were found to enhance the effects of the antibiotics. Rhodanine derivatives have strong antimicrobial activity against methicillin-resistant and non-methicillin-resistant S. aureus strains. Quinoxalines derivatives inhibited bacterial activity against the gram-positive bacteria, Bacillus subtilis, and Staphylococcus aureus [72].

3.3. Anti-Fungal Agent

Chalcones act on glutathione, cysteine molecules, and proteins involved in the cell separation of yeast cells. A hybrid molecule of pharmacophore of fluconazole and chalcone was tested for their antifungal activity and showed inhibitory activity against Candida albicans. Chalcone with unsubstituted thiophene ring B and thiomethyl substitution at the p-position of ring A (compared to o- and m-) produced credible antifungal activity on fluconazole-sensitive and fluconazole-resistant strains of yeast [73].
The highest antifungal efficacy against standardized strains of Trichophyton rubrum was demonstrated by the chalcone 3-(2-chlorophenyl)-1-(2′-hydroxy-4′,6′-dimethoxyphenyl)prop-2-en-1-one (MIC = 12.5 μg/mL), which also inhibited all ten clinical isolates of T. rubrum (IC50 12.5 and IC90 25 μg/mL). The Neurospora crassa assay revealed a blotchy look in the inhibitory halo formed by this chalcone, strongly indicating that it may work by inhibiting the fungal cell wall [74]. This study also demonstrated that noticeable hyphal curling was observed in the hazy zone at a magnification of 400 or less; therefore, this chalcone appears to be a hyphal deformity inducer.
The highest antifungal activity (MIC: 1.95 μg/mL) was shown by the compounds 62, 63, 68, and 69 against C. albicans (Table 2), followed by the compounds 66, 67 (MIC: 3.90 μg/mL), and compound 65 has shown the same activity (MIC: 15.62 μg/mL) as Amphotericin-B. The remaining compounds have shown lower activity (MIC ≥ 125–62.50 μg/mL) against C. albicans [58]. In comparison to C. albicans, all compounds were less active against A. niger. In this case, compounds 66 and 67 have shown comparatively better antifungal activity (MIC: 7.81 μg/mL) against A. niger. However, as compared to Amphotericin-B, compounds 61, 65 (MIC: 15.62 μg/mL), 68 (MIC: 31.25 g/mL), and 63, 64 (MIC: 62.50 μg/mL) have shown good antifungal activity against A. niger (Table 2). Compounds 66 and 67 have demonstrated equal antifungal activity with Nystatin against A. niger, possibly due to the presence of a high electron releasing group (Tri-OMe) or without substitutions on the aromatic ring. Similarly, compounds 62, 63, and 69 were most active against C. albicans as compared to standard drugs Amphotericin-B and Nystatin may be due to the presence of the electron-withdrawing group.
Other observations included significant antifungal activity by fluoro-substitution at the p- position on ring A and chalcone compounds with smaller halogen sizes [73]. Several studies on the antifungal activity of chalcones have results that contradict one another. However, they have a common finding: placing the hydroxyl group at the m-position in ring A produces a significant antifungal effect [54].

3.4. Anti-Malarial Agent

Numerous studies have been conducted to evaluate the antimalarial activity of chalcone and its derivatives for decades to develop novel, safe, less toxic, and highly active antimalarials [75,76,77,78,79,80,81,82,83,84,85,86,87,88]. Among the many chalcones-chloroquinoline hybrid compounds, 3-(4-[1-(7-Chloro-quinolin-4-yl)-1H-[1,2,3]triazol-4-yl-methoxy]-3-methoxy-phenyl)-1-(2,4-dimethoxy-phenyl)-propenone was the most active against different isolates, chloroquine-sensitive (CQS) strain D10 (IC50 0.04 µM) and chloroquine-resistant (CQR) strains Dd2 (IC50 0.07 µM) and W2 (IC50 0.09 µM) of Plasmodium falciparum [81].
A series of chalcone derivatives were investigated to determine their antimalarial activity on P. falciparum cysteine protease; the results showed varying antimalarial activity depending on the steric and hydrophobic properties, molar refractivity, and molecular length against chloroquine-resistant P. falciparum, and the molecular weight against mefloquine-resistant strains [89]. In another study, phenylurenyl chalcone derivatives were synthesized and tested as inhibitors against a chloroquine-resistant strain of P. falciparum to evaluate the activity of the cysteine protease falcipain-2, globin hydrolysis, β-hematin formation, and murine P. berghei malaria [75]. Among the tested Chalcones, the most active antimalarial compound was 1-[3′-N-(N′-phenylurenyl) phenyl]-3(3,4,5-trimethoxyphenyl)-2-propen-1-one (IC50 of 1.76 μM).
A series of novel keto-enamine Chalcone-chloroquine hybrids were tested against chloroquine-sensitive P. falciparum 3D7 strain and found some compounds that exhibited comparable antimalarial activity [76]. Later, highly potent antimalarial compounds were evaluated for their in vivo efficacy using Swiss mice infected by chloroquine-resistant (N-67) strain of P. yoelii and demonstrated that compounI(E)-2-tert-Butyl-6-((2-(7-chloroquinolin-4-ylamino)ethylamino)methylenI4-((E)-3-(4-fluorophenyl)-3-oxoprop-1-enyl)cyclohexa-2,4-dienone, and (E)-4-((E)-3-(4-Bromophenyl)-3-oxoprop-1-enyl)-2-tert-butyl-6-((2-(7-chloroquinolin-4-ylamino)ethylamino)methylene)cyclohexa-2,4-dienone each suppressed 99.9% parasitemia on day 4, possibly by the inhibition of hemozoin formation. This mode of action may be the primary mechanism of action of these compounds against malaria parasites.
A study reported that the C2–C3 double bond is responsible for the high inhibitory activity of chalcones because it links the A and B rings as well as stabilizes the molecular conformation, which allows the drug molecule to bind to the active site more efficiently; there is decreased inhibitory activity by steric interactions due to the substitutions on the bridge portion of chalcone derivatives; increased inhibitory activity was seen in electron-donating substitution on ring A and chloro- or fluoro- substitution on ring B; an increase in antimalarial activity in placing quinolinyl group to ring B [90]. A small hydrophobic nitrogen heterocyclic group at ring B and a small lipophilic functional group at ring A can increase antimalarial activity [90].

3.5. Anti-Protozoal and Anti-Filarial Agent

Even though chalcone and its derivatives have diverse biological applications; however, limited applications have been found for antiprotozoal activity [91,92,93]. Since structural modifications of the main pharmacophore of chalcone is the synthesis process of new derivatives, expecting high biological activity of novel compounds is normal. However, the antiparasitic potencies of the α,β-double bond modified chalcones only differ marginally from the potencies of the parent chalcones [93]. This may be limited the further evaluation of chalcone for antiprotozoal activity.
Twenty Chalcone compounds were isolated from plants and tested against extracellular promastigotes of Leishmania donovani, L. infantum, L. enrietii and L. major, and against intracellular amastigote L. donovani residing within murine macrophages [93]. Most of the compounds were active (EC50 0.07–2.01 µg/mL) against the extracellular Leishmania (L. donovani). A few chalcones, namely 2′,4′-dihydroxy-4-methoxychalcone, 2′-hydroxy-3,4-dimethoxychalcone and 2-hydroxy-4,4′-dimethoxychalcone (EC50 0.39–0.41 µg/mL) were also significantly inhibited the intracellular survival of L. donovani parasites. However, all the compounds showed cytotoxicity (EC50 0.19–2.06 µg/mL) while tested on mammalian macrophages derived from murine bone marrow [92]. Hayat et al. [91] screened fifteen Chalcone molecules, of which only four compounds were found to be more active against the Entamoeba histolytica, and one compound was moderately active compared to reference drug metronidazole (IC50: 1.46 μM). In addition, all the tested compounds were non-toxic against the human breast cancer MCF-7 cell line (IC50: 1.56–50 μM).
TIe chalcIne of interest in having anti-leishmanial activities is licochalcone A; it modifies the structure and function of the Leishm’nia braziliensis’s mitochondria as well as the activity of mitochondrial dehydrogenases. In addition, the aforementioned chalcone prevents the promastigotes and amastigotes, which are produced by the parasite, from developing. There is also a decrease in parasite load inside the liver and spleen [94]. Another protozoon called Trypanosoma cruzi was used to study the effects of chalcone. The chalcones with no substitution groups exhibit anti-trypanosomal activity [95].
Chalcones were tested on Setaria cervi extracted from female adults for anti-filarial effects by inhibiting glutathione-S-transferase (GST) enzyme. There was a significant inhibitory activity of GST, which led to irreversible inhibition of the viability and motility of the parasites, as well as reduced glutathione levels. Results indicated that lipophilic groups of the ring in chalcone containing oxygen or nitrogen were behind the anti-filarial effects, where methoxy groups exhibited the most significant activity [96].

3.6. Anti-Inflammatory Agent

The main component of chalcone that contributes to its anti-inflammatory property is the α,β-unsaturated carbon skeletal system. Polymethoxychalcones showed potent inhibitory effects on inflammation, mainly through the production of high IL-1 β-induced PGE2, prostaglandin inhibitors, inflammatory cytokines, and normal human epidermal keratinocyte inhibitors. A series of pyran-fused chalcones and trichalcones can inhibit the NF-κB signaling complement system involved in inflammation. At the same time, some compounds showed better inhibitory activity against lipoxygenase than the standard drug indomethacin. Inflammatory cytokines, particularly TNF-α, IL-1β, IL-6, and IL-8 production, were inhibited by chalcones with heterocyclic systems in their moiety. Chalcones converted to Mannich bases also inhibit the release of inflammatory mediators from mast cells, macrophages, neutrophils, and microglial cells. Other chalcone derivatives, such as dihydroxychalcones, also exhibit anti-inflammatory activity [46].
In inflammatory diseases, nardoaristolone A and isobava chalcone significantly inhibited pain-related acute and chronic inflammation and treated rheumatoid arthritis. Other chalcone compounds can also be used to suppress secondary tumor formation and treat rheumatoid arthritis and osteoarthritis. SGCH 19 (1-(7-chloro-3-methyl-1-phenyl-2-naphthyl)-3-(2-furyl)-2-propen-1-one), SGCH 20 (1-(7-chloro-3-methyl-1-phenyl-2-naphthyl)-3-(2-thienyl)-2-propen-1-one), and chalcone compounds with fluoride or chloride groups have a high potency as an anti-inflammatory agent, and may be more potent than the reference drugs, indomethacin and ibuprofen. It was concluded that electron-withdrawing groups (EWG) are responsible for the increased anti-inflammatory activity of chalcones [97].
The anti-inflammatory activity also depends on the position of the substituted group on the phenyl ring A and increases in the order: −3, 4-(OH)2, −4-OH, −3,4-OCH2O, −3-OCH3–4-OH, −4-OCH3, −4-N(CH3)2. Chalcones with fluoride or chloride groups also show potent anti-inflammatory activity, with 4-Cl and 3-NO2 being more powerful than 2, 4-Cl, and 2-NO2, respectively [2].
Three enzymes were tested with the inhibitory activity of chalcone derivatives on these enzymes, namely the mammalian alpha-amylase, cyclooxygenase (COX), and monoamine oxidases (MAOs) [98,99,100]. In a study involving porcine pancreatic alpha-amylase, trans-chalcones act as a competitive inhibitor by interacting with Trp59 and Tyr62 [101].

3.7. Anti-Cancer Agent

In recent years, numerous studies evaluated and revealed the anticancer potential of chalcone and chalcone-based different types of derivatives through in vitro, in vivo as well as molecular docking studies [102,103,104,105,106,107,108,109,110,111]. Both in vitro and in vivo methods were used to explore the anti-melanoma effects of flavokawain B on human melanoma cells and the processes of cell death that were mediated by the generation of reactive oxygen species (ROS) [108]. The findings show that flavokawain B decreased the viability of human melanoma cells as well as the expression of the B-Raf proto-oncogene, serine/threonine kinase (BRAF), and extracellular signal-regulated kinase. The execution of apoptosis was aided by Caspase-3 activation, the PARP cleavage pathway, and Bcl2-associated X (Bax)/B-cell lymphoma 2 (Bcl-2) dysregulation. Additionally, flavokawain B also inhibited tumor growth in nude mice with xenografts [108]. This study concluded that flavokawain B could potentially be used for managing human melanoma cancer by executing and producing ROS-modulated apoptotic and autophagic cell death.
Chouiter et al. [105] synthesized novel chalcone derivatives and tested for their anticancer activity using human lung (A549) and stomach (AGS) cancer cell lines and evaluated them in the noncancer human lung fibroblast (MRC-5) cell lines. All the cell lines tested showed no toxicity for 2-pyrazolines, although the AGS cell line was hazardous for chalcones containing a benzimidazole moiety. Mechanistic studies revealed that these substances cause loss of cell viability and mitochondrial membrane potential while inducing morphological characteristics consistent with regulated cell death by caspase activation, particularly by caspase-3. Boronic chalcones were investigated for anticancer activity, and results indicated the chalcones, particularly AM114, produced a cytotoxic effect in cancer cells by inhibiting the chymotrypsin-like activity of the 20S proteasome in these cells, which in turn accumulated p53 and p21 proteins, creating a favorable condition for cell apoptosis [112]. In other studies, the double bond was replaced with a thiophene ring, which showed the antiproliferative activity of cancer cells and the growth inhibition activity of cancer cell lines at nanomolar to low micromolar concentrations. Some studies have shown inhibition of tubulin polymerization by binding [3H] colchicine to tubulin and arresting the G2/M phase of K562 cells [113,114]. 2′-hydroxychalcone was studied for their antitumor effects on HepG2 hepatocellular carcinoma cells, and results showed they initiate cell apoptosis and inhibit cell proliferation. This was due to the ‘ring A,’ which does not contain methoxy groups. Other chalcones promoted cell apoptosis and cytotoxicity in prostate cancer cells through tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) [114].
In addition, studies conducted on the cytotoxic activity of chalcones to breast cancer cell lines MCF-7 and T47D showed significant cytotoxicity, especially the derivatives containing nitro‘groups, such as ‘N-4-hydroxy-3-(3-(2/3/4-nitrophenyl)-acryloyl)’phenyl acetamide’. Another study on quinolinyl chalcone derivatives discovered the potent anti-cancer activity of up to 103% by SGCH 3 (3-(4-chlorophenyl)-1-(3-methyl-1-phenyl-2-naphthyl)-2-propen-1-one) [97]. The coumarinyl-chalcone derivatives (3 h and 3 m) exhibit cytotoxicity towards human cell lines of the lungs, breast, and blood cells, albeit less potent than the standard drug, Imatinib [113,115].

4. Conclusions

This review summarized the recent antimicrobial and other pharmacological activities of chalcone compounds with their mode of action. Moreover, we have discussed different methods of chalcone derivatives synthesis and characterization. Notably, chalcone derivatives demonstrated potential antimicrobial activities, including antibiotic-resistant strains like MRSA, KPC, and NDM. Chalcone derivatives potentially inhibit the different targets pathway of microbial resistance. Several chalcone compounds demonstrated optimum characteristics of an effective antimicrobial agent. However, their potential must be warranted using preclinical and clinical studies. Chalcone derivatives can also be a lead compound with anti-inflammatory and anticancer properties, but further study is required to corroborate its clinical activities. Since structure-activity has a potential relationship, even though some modifications revealed negligible improvement, further study is needed for more extensive evaluation, including molecular mechanisms of actions. Even though a plethora of health benefits are evident from experimental studies, evaluating the pharmacological activities of chalcone and its derivatives is still required, particularly in preclinical and clinical studies. As chalcone and its derivatives can be synthesized in the laboratory and chemical structure can be modified, more effective action against a specific disease pathway can be achieved.

Author Contributions

Conceptualization, L.C.M. and J.S.D.; methodology, L.C.M., A.H. and V.K.; formal analysis, L.C.M., S.M. and J.S.D.; resources, J.S.D., K.W.G., N.K. and H.P.G.; data curation, V.K., K.W.G. and M.J.L.; Supervision, L.C.M., N.K., A.H. and H.P.G.; Visualization, M.S.H. and S.M.; Funding acquisition, J.S.D., A.H., H.Y. and K.W.G.; writing original draft preparation, S.M., J.S.D., V.K. and L.C.M.; writing, review and editing, M.J.L., A.H., S.K.S.D., H.Y. and M.S.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data have been included in the manuscript.

Acknowledgments

We thank Afza Firzanah for the preliminary data collation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Matos, M.J.; Vazquez-Rodriguez, S.; Uriarte, E.; Santana, L. Potential pharmacological uses of chalcones: A patent review (from June 2011–2014). Expert Opin. Ther. Pat. 2015, 25, 351–366. [Google Scholar] [CrossRef] [PubMed]
  2. Chavan, B.B.; Gadekar, A.S.; Mehta, P.P.; Vawhal, P.K.; Kolsure, A.K.; Chabukswar, A.R. Synthesis and Medicinal Significance of Chalcones- A Review. Asian J. Biomed. Pharm. Sci. 2016, 6, 56. [Google Scholar]
  3. Bandgar, B.P.; Gawande, S.S. Synthesis and biological screening of a combinatorial library of beta-chlorovinyl chalcones as anticancer, anti-inflammatory and antimicrobial agents. Bioorg. Med. Chem. 2010, 18, 2060–2065. [Google Scholar] [CrossRef] [PubMed]
  4. Salehi, B.; Quispe, C.; Chamkhi, I.; El Omari, N.; Balahbib, A.; Sharifi-Rad, J.; Bouyahya, A.; Akram, M.; Iqbal, M.; Docea, A.O.; et al. Pharmacological Properties of Chalcones: A Review of Preclinical Including Molecular Mechanisms and Clinical Evidence. Front. Pharmacol. 2021, 11, 592654. [Google Scholar] [CrossRef] [PubMed]
  5. Dan, W.; Dai, J. Recent developments of chalcones as potential antibacterial agents in medicinal chemistry. Eur. J. Med. Chem. 2020, 187, 111980. [Google Scholar] [CrossRef] [PubMed]
  6. Bandgar, B.P.; Patil, S.A.; Gacche, R.N.; Korbad, B.L.; Hote, B.S.; Kinkar, S.N.; Jalde, S.S. Synthesis and biological evaluation of nitrogen-containing chalcones as possible anti-inflammatory and antioxidant agents. Bioorganic Med. Chem. Lett. 2010, 20, 730–733. [Google Scholar] [CrossRef]
  7. Alcaráz, L.E.; Blanco, S.E.; Puig, O.N.; Tomás, F.; Ferretti, F.H. Antibacterial activity of flavonoids against methicillin-resistant Staphylococcus aureus strains. J. Biol. 2000, 205, 231–240. [Google Scholar] [CrossRef] [PubMed]
  8. Nowakowska, Z. A review of anti-infective and anti-inflammatory chalcones. Eur. J. Med. Chem. 2007, 42, 125–137. [Google Scholar] [CrossRef] [PubMed]
  9. Talia, J.M.; Debattista, N.B.; Pappano, N.B. New antimicrobial combinations: Substituted chalcones- oxacillin against methicillin resistant Staphylococcus aureus. Braz. J. Microbiol. 2011, 42, 470–475. [Google Scholar] [CrossRef] [Green Version]
  10. Williamson, E.M. Synergy and other interactions in phytomedicines. Phytomedicine 2001, 8, 401–409. [Google Scholar] [CrossRef]
  11. Hemaiswarya, S.; Kruthiventi, A.K.; Doble, M. Synergism between natural products and antibiotics against infectious diseases. Phytomedicine 2008, 15, 639–652. [Google Scholar] [CrossRef]
  12. 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]
  13. Gupta, D.; Jain, D.K. Chalcone derivatives as potential antifungal agents: Synthesis, and antifungal activity. J. Adv. Pharm. Technol. Res. 2015, 6, 114–117. [Google Scholar] [CrossRef]
  14. López, S.N.; Castelli, M.V.; Zacchino, S.A.; Domínguez, J.N.; Lobo, G.; Charris-Charris, J.; Cortés, J.C.; Ribas, J.C.; Devia, C.; Rodríguez, A.M.; et al. In vitro antifungal evaluation and structure-activity relationships of a new series of chalcone derivatives and synthetic analogues, with inhibitory properties against polymers of the fungal cell wall. Bioorg. Med. Chem. 2001, 9, 1999–2013. [Google Scholar] [CrossRef]
  15. Tsuchiya, H.; Sato, M.; Akagiri, M.; Takagi, N.; Tanaka, T.; Iinuma, M. Anti-Candida activity of synthetic hydroxychalcones. Pharmazie 1994, 49, 756–758. [Google Scholar] [PubMed]
  16. Sato, M.; Tsuchiya, H.; Akagiri, M.; Fujiwara, S.; Fujii, T.; Takagi, N.; Matsuura, N.; Iinuma, M. Growth inhibitory properties of chalcones to Candida. Lett. Appl. Microbiol. 1994, 18, 53–55. [Google Scholar] [CrossRef]
  17. Reddy, N.P.; Aparoy, P.; Reddy, T.C.; Achari, C.; Sridhar, P.R.; Reddanna, P. Design, synthesis, and biological evaluation of prenylated chalcones as 5-LOX inhibitors. Bioorg. Med. Chem. 2010, 18, 5807–5815. [Google Scholar] [CrossRef]
  18. Lahtchev, K.L.; Batovska, D.I.; Parushev, S.P.; Ubiyvovk, V.M.; Sibirny, A.A. Antifungal activity of chalcones: A mechanistic study using various yeast strains. Eur. J. Med. Chem. 2008, 43, 2220–2228. [Google Scholar] [CrossRef]
  19. Mahapatra, D.K.; Bharti, S.K.; Asati, V. Chalcone Derivatives: Anti-inflammatory Potential and Molecular Targets Perspectives. Curr. Top. Med. Chem. 2017, 17, 3146–3169. [Google Scholar] [CrossRef] [PubMed]
  20. Hsieh, H.K.; Tsao, L.T.; Wang, J.P.; Lin, C.N. Synthesis and anti-inflammatory effect of chalcones. J. Pharm. Pharm. 2000, 52, 163–171. [Google Scholar] [CrossRef]
  21. Karthikeyan, C.; Moorthy, N.S.; Ramasamy, S.; Vanam, U.; Manivannan, E.; Karunagaran, D.; Trivedi, P. Advances in chalcones with anticancer activities. Recent Pat. Anticancer Drug Discov. 2015, 10, 97–115. [Google Scholar] [CrossRef]
  22. Sharma, V.; Kumar, V.; Kumar, P. Heterocyclic chalcone analogues as potential anticancer agents. Anticancer Agents Med. Chem. 2013, 13, 422–432. [Google Scholar]
  23. León-González, A.J.; Acero, N.; Muñoz-Mingarro, D.; Navarro, I.; Martín-Cordero, C. Chalcones as Promising Lead Compounds on Cancer Therapy. Curr. Med. Chem. 2015, 22, 3407–3425. [Google Scholar] [CrossRef] [PubMed]
  24. 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]
  25. Marinov, R.; Markova, N.; Krumova, S.; Yotovska, K.; Zaharieva, M.M.; Genova-Kalou, P. Antiviral properties of chalcones and their synthetic derivatives: A mini review. Pharmacia 2020, 67, 325. [Google Scholar] [CrossRef]
  26. Jasim, H.A.; Nahar, L.; Jasim, M.A.; Moore, S.A.; Ritchie, K.J.; Sarker, S.D. Chalcones: Synthetic chemistry follows where nature leads. Biomolecules 2021, 11, 1203. [Google Scholar] [CrossRef] [PubMed]
  27. 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]
  28. Sun, Y.-F.; Cui, Y.-P. The synthesis, characterization and properties of coumarin-based chromophores containing a chalcone moiety. Dyes Pigments 2008, 78, 65–76. [Google Scholar] [CrossRef]
  29. Gomes, M.N.; Muratov, E.N.; Pereira, M.; Peixoto, J.C.; Rosseto, L.P.; Cravo, P.V.L.; Andrade, C.H.; Neves, B.J. Chalcone Derivatives: Promising Starting Points for Drug Design. Molecules 2017, 22, 1210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Dong, F.; Jian, C.; Zhenghao, F.; Kai, G.; Zuliang, L. Synthesis of chalcones via Claisen–Schmidt condensation reaction catalyzed by acyclic acidic ionic liquids. Catal. Commun. 2008, 9, 1924–1927. [Google Scholar] [CrossRef]
  31. Wadey, B.L. Plasticizers. In Encyclopedia of Physical Science and Technology, 3rd ed.; Meyers, R.A., Ed.; Academic Press: New York, NY, USA, 2003; pp. 441–456. [Google Scholar]
  32. Larsen, R.D. 3.08-1,3-Selenazoles. In Comprehensive Heterocyclic Chemistry II; Katritzky, A.R., Rees, C.W., Scriven, E.F.V., Eds.; Pergamon: Oxford, UK, 1996; pp. 493–510. [Google Scholar]
  33. Zhan, C.; Yang, J. Protective effects of isoliquiritigenin in transient middle cerebral artery occlusion-induced focal cerebral ischemia in rats. Pharmacol. Res. 2006, 53, 303–309. [Google Scholar] [CrossRef]
  34. Furusawa, J.; Funakoshi-Tago, M.; Mashino, T.; Tago, K.; Inoue, H.; Sonoda, Y.; Kasahara, T. Glycyrrhiza inflata-derived chalcones, Licochalcone A, Licochalcone B and Licochalcone D, inhibit phosphorylation of NF-kappaB p65 in LPS signaling pathway. Int. Immunopharmacol. 2009, 9, 499–507. [Google Scholar] [CrossRef]
  35. Chen, M.; Christensen, S.B.; Blom, J.; Lemmich, E.; Nadelmann, L.; Fich, K.; Theander, T.G.; Kharazmi, A. Licochalcone A, a novel antiparasitic agent with potent activity against human pathogenic protozoan species of Leishmania. Antimicrob. Agents Chemother. 1993, 37, 2550–2556. [Google Scholar] [CrossRef] [Green Version]
  36. Chen, M.; Christensen, S.B.; Theander, T.G.; Kharazmi, A. Antileishmanial activity of licochalcone A in mice infected with Leishmania major and in hamsters infected with Leishmania donovani. Antimicrob. Agents Chemother. 1994, 38, 1339–1344. [Google Scholar] [CrossRef] [Green Version]
  37. Zhai, L.; Blom, J.; Chen, M.; Christensen, S.B.; Kharazmi, A. The antileishmanial agent licochalcone A interferes with the function of parasite mitochondria. Antimicrob. Agents Chemother. 1995, 39, 2742–2748. [Google Scholar] [CrossRef] [Green Version]
  38. Ogawa, H.; Okada, Y.; Kamisako, T.; Baba, K. Beneficial effect of xanthoangelol, a chalcone compound from Angelica keiskei, on lipid metabolism in stroke-prone spontaneously hypertensive rats. Clin. Exp. Pharmacol. Physiol. 2007, 34, 238–243. [Google Scholar] [CrossRef] [PubMed]
  39. Li, Y.; Qin, X.; Li, P.; Zhang, H.; Lin, T.; Miao, Z.; Ma, S. Isobavachalcone isolated from Psoralea corylifolia inhibits cell proliferation and induces apoptosis via inhibiting the AKT/GSK-3β/β-catenin pathway in colorectal cancer cells. Drug Des. Dev. Ther. 2019, 13, 1449–1460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Martinson, K.; Stueven, N.; Monte, A.; Huang, C. A Novel Stilbene-Like Compound That Reduces Melanin through Inhibiting Melanocyte Differentiation and Proliferation without Inhibiting Tyrosinase. Cosmetics 2018, 5, 45. [Google Scholar] [CrossRef] [Green Version]
  41. Lee, H.; Li, H.; Kweon, M.; Choi, Y.; Kim, M.J.; Ryu, J.H. Isobavachalcone from Angelica keiskei Inhibits Adipogenesis and Prevents Lipid Accumulation. Int. J. Mol. Sci. 2018, 19, 1693. [Google Scholar] [CrossRef] [Green Version]
  42. Liu, X.; Li, W.; Hu, B.; Wang, M.; Wang, J.; Guan, L. Identification of isobavachalcone as a potential drug for rice blast disease caused by the fungus Magnaporthe grisea. J. Biomol. Struct. Dyn. 2019, 37, 3399–3409. [Google Scholar] [CrossRef]
  43. Liu, M.; Yin, H.; Liu, G.; Dong, J.; Qian, Z.; Miao, J. Xanthohumol, a Prenylated Chalcone from Beer Hops, Acts as an α-Glucosidase Inhibitor in Vitro. J. Agric. Food Chem. 2014, 62, 5548–5554. [Google Scholar] [CrossRef]
  44. Proença, C.; Freitas, M.; Ribeiro, D.; Oliveira, E.F.T.; Sousa, J.L.C.; Tomé, S.M.; Ramos, M.J.; Silva, A.M.S.; Fernandes, P.A.; Fernandes, E. α-Glucosidase inhibition by flavonoids: An in vitro and in silico structure-activity relationship study. J Enzym. Inhib Med Chem 2017, 32, 1216–1228. [Google Scholar] [CrossRef]
  45. Rahman, M. Chalcone: A Valuable Insight into the Recent Advances and Potential Pharmacological Activities. Chem. Sci. 2011, 2011. [Google Scholar] [CrossRef]
  46. 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]
  47. El-Meligie, S.; Taher, A.T.; Kamal, A.M.; Youssef, A. Design, synthesis and cytotoxic activity of certain novel chalcone analogous compounds. Eur. J. Med. Chem. 2017, 126, 52–60. [Google Scholar] [CrossRef]
  48. Bukhari, S.N.; Jasamai, M.; Jantan, I. Synthesis and biological evaluation of chalcone derivatives (mini review). Mini Rev. Med. Chem. 2012, 12, 1394–1403. [Google Scholar]
  49. Cheng, P.; Yang, L.; Huang, X.; Wang, X.; Gong, M. Chalcone hybrids and their antimalarial activity. Arch. Der Pharm. 2020, 353, e1900350. [Google Scholar]
  50. Ouyang, Y.; Li, J.; Chen, X.; Fu, X.; Sun, S.; Wu, Q. Chalcone Derivatives: Role in Anticancer Therapy. Biomolecules 2021, 11, 894. [Google Scholar] [CrossRef]
  51. 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]
  52. Rudrapal, M.; Khan, J.; Dukhyil, A.A.B.; Alarousy, R.; Attah, E.I.; Sharma, T.; Khairnar, S.J.; Bendale, A.R. Chalcone Scaffolds, Bioprecursors of Flavonoids: Chemistry, Bioactivities, and Pharmacokinetics. Molecules 2021, 26, 7177. [Google Scholar] [CrossRef]
  53. Sahu, N.K.; Balbhadra, S.S.; Choudhary, J.; Kohli, D.V. Exploring pharmacological significance of chalcone scaffold: A review. Curr. Med. Chem. 2012, 19, 209–225. [Google Scholar] [CrossRef]
  54. 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]
  55. 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]
  56. Sarig, H.; Rotem, S.; Ziserman, L.; Danino, D.; Mor, A. Impact of self-assembly properties on antibacterial activity of short acyl-lysine oligomers. Antimicrob. Agents Chemother. 2008, 52, 4308–4314. [Google Scholar] [CrossRef]
  57. Ávila, H.P.; Smânia, E.d.F.A.; Monache, F.D.; Smânia, A. Structure–activity relationship of antibacterial chalcones. Bioorganic Med. Chem. 2008, 16, 9790–9794. [Google Scholar] [CrossRef]
  58. Konduru, N.K.; Dey, S.; Sajid, M.; Owais, M.; Ahmed, N. Synthesis and antibacterial and antifungal evaluation of some chalcone based sulfones and bisulfones. Eur. J. Med. Chem. 2013, 59, 23–30. [Google Scholar] [CrossRef]
  59. 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]
  60. Tang, Y.L.; Li, Y.K.; Li, M.X.; Gao, H.; Yang, X.B.; Mao, Z.W. Synthesis of New Piperazine Substituted Chalcone Sulphonamides as Antibacterial Agents. Curr. Org. Synth. 2020, 17, 136–143. [Google Scholar] [CrossRef]
  61. Soto, S.M. Role of efflux pumps in the antibiotic resistance of bacteria embedded in a biofilm. Virulence 2013, 4, 223–229. [Google Scholar] [CrossRef] [Green Version]
  62. Du, D.; Wang-Kan, X.; Neuberger, A.; van Veen, H.W.; Pos, K.M.; Piddock, L.J.V.; Luisi, B.F. Multidrug efflux pumps: Structure, function and regulation. Nat. Rev. Microbiol. 2018, 16, 523–539. [Google Scholar] [CrossRef]
  63. Sun, J.; Deng, Z.; Yan, A. Bacterial multidrug efflux pumps: Mechanisms, physiology and pharmacological exploitations. Biochem. Biophys. Res. Commun. 2014, 453, 254–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Nikaido, H. Antibiotic Resistance Caused by Gram-Negative Multidrug Efflux Pumps. Clin. Infect. Dis. 1998, 27, S32–S41. [Google Scholar] [CrossRef] [PubMed]
  65. Blair, J.M.A.; Richmond, G.E.; Piddock, L.J.V. Multidrug efflux pumps in Gram-negative bacteria and their role in antibiotic resistance. Future Microbiol. 2014, 9, 1165–1177. [Google Scholar] [CrossRef] [PubMed]
  66. Alcalde-Rico, M.; Hernando-Amado, S.; Blanco, P.; Martínez, J.L. Multidrug Efflux Pumps at the Crossroad between Antibiotic Resistance and Bacterial Virulence. Front. Microbiol. 2016, 7, 1483. [Google Scholar] [CrossRef] [PubMed]
  67. Amaral, L.; Martins, A.; Spengler, G.; Molnar, J. Efflux pumps of Gram-negative bacteria: What they do, how they do it, with what and how to deal with them. Front. Pharmacol. 2014, 4, 168. [Google Scholar] [CrossRef] [Green Version]
  68. Sharma, A.; Gupta, V.K.; Pathania, R. Efflux pump inhibitors for bacterial pathogens: From bench to bedside. Indian J. Med. Res. 2019, 149, 129–145. [Google Scholar]
  69. Holler, J.G.; Slotved, H.-C.; Mølgaard, P.; Olsen, C.E.; Christensen, S.B. Chalcone inhibitors of the NorA efflux pump in Staphylococcus aureus whole cells and enriched everted membrane vesicles. Bioorganic Med. Chem. 2012, 20, 4514–4521. [Google Scholar] [CrossRef]
  70. Gaur, R.; Gupta, V.K.; Pal, A.; Darokar, M.P.; Bhakuni, R.S.; Kumar, B. In vitro and in vivo synergistic interaction of substituted chalcone derivatives with norfloxacin against methicillin resistant Staphylococcus aureus. Rsc Adv. 2015, 5, 5830–5845. [Google Scholar] [CrossRef]
  71. Vijay, K.; Karakavalasa, P.; Vasanthi, R. Synthesis, characterization and pharmacological evaluation of some novel quinoxaline derived chalcones. Der Pharma Chem. 2013, 5, 301–307. [Google Scholar]
  72. Xu, M.; Wu, P.; Shen, F.; Ji, J.; Rakesh, K.P. Chalcone derivatives and their antibacterial activities: Current development. Bioorg. Chem. 2019, 91, 103133. [Google Scholar] [CrossRef]
  73. Kucerova-Chlupacova, M.; Kunes, J.; Buchta, V.; Vejsova, M.; Opletalova, V. Novel pyrazine analogs of chalcones: Synthesis and evaluation of their antifungal and antimycobacterial activity. Molecules 2015, 20, 1104–1117. [Google Scholar] [CrossRef] [Green Version]
  74. Boeck, P.; Leal, P.C.; Yunes, R.A.; Filho, V.C.; López, S.; Sortino, M.; Escalante, A.; Furlán, R.L.E.; Zacchino, S. Antifungal Activity and Studies on Mode of Action of Novel Xanthoxyline-Derived Chalcones. Arch. Der Pharm. 2005, 338, 87–95. [Google Scholar] [CrossRef] [PubMed]
  75. Domínguez, J.N.; León, C.; Rodrigues, J.; Gamboa de Domínguez, N.; Gut, J.; Rosenthal, P.J. Synthesis and evaluation of new antimalarial phenylurenyl chalcone derivatives. J. Med. Chem. 2005, 48, 3654–3658. [Google Scholar] [CrossRef]
  76. Sashidhara, K.V.; Kumar, M.; Modukuri, R.K.; Srivastava, R.K.; Soni, A.; Srivastava, K.; Singh, S.V.; Saxena, J.K.; Gauniyal, H.M.; Puri, S.K. Antiplasmodial activity of novel keto-enamine chalcone-chloroquine based hybrid pharmacophores. Bioorganic Med. Chem. 2012, 20, 2971–2981. [Google Scholar] [CrossRef] [PubMed]
  77. 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] [PubMed]
  78. Domínguez, J.N.; León, C.; Rodrigues, J.; de Domínguez, N.G.; Gut, J.; Rosenthal, P.J. Synthesis and antimalarial activity of sulfonamide chalcone derivatives. Il Farm. 2005, 60, 307–311. [Google Scholar] [CrossRef] [PubMed]
  79. Tomar, V.; Bhattacharjee, G.; Rajakumar, S.; Srivastava, K.; Puri, S.K. Synthesis of new chalcone derivatives containing acridinyl moiety with potential antimalarial activity. Eur. J. Med. Chem. 2010, 45, 745–751. [Google Scholar] [CrossRef]
  80. 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]
  81. 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. Bioorganic Med. Chem. 2010, 18, 8243–8256. [Google Scholar] [CrossRef] [PubMed]
  82. Pingaew, R.; Saekee, A.; Mandi, P.; Nantasenamat, C.; Prachayasittikul, S.; Ruchirawat, S.; Prachayasittikul, V. Synthesis, biological evaluation and molecular docking of novel chalcone–coumarin hybrids as anticancer and antimalarial agents. Eur. J. Med. Chem. 2014, 85, 65–76. [Google Scholar] [CrossRef] [PubMed]
  83. Li, R.; Kenyon, G.L.; Cohen, F.E.; Chen, X.; Gong, B.; Dominguez, J.N.; Davidson, E.; Kurzban, G.; Miller, R.E.; Nuzum, E.O. In vitro antimalarial activity of chalcones and their derivatives. J. Med. Chem. 1995, 38, 5031–5037. [Google Scholar] [CrossRef]
  84. Sinha, S.; Medhi, B.; Sehgal, R. Chalcones as an emerging lead molecule for antimalarial therapy: A review. J. Mod. Med. Chem 2013, 1, 64–77. [Google Scholar]
  85. Liu, M.; Wilairat, P.; Croft, S.L.; Tan, A.L.-C.; Go, M.-L. Structure–activity relationships of antileishmanial and antimalarial chalcones. Bioorganic Med. Chem. 2003, 11, 2729–2738. [Google Scholar] [CrossRef]
  86. Tadigoppula, N.; Korthikunta, V.; Gupta, S.; Kancharla, P.; Khaliq, T.; Soni, A.; Srivastava, R.K.; Srivastava, K.; Puri, S.K.; Raju, K.S.R. Synthesis and insight into the structure–activity relationships of chalcones as antimalarial agents. J. Med. Chem. 2013, 56, 31–45. [Google Scholar] [CrossRef]
  87. 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. Bioorganic Med. Chem. Lett. 2010, 20, 942–944. [Google Scholar] [CrossRef]
  88. Domínguez, J.N.; Charris, J.E.; Lobo, G.; de Domínguez, N.G.; Moreno, M.a.M.; Riggione, F.; Sanchez, E.; Olson, J.; Rosenthal, P.J. Synthesis of quinolinyl chalcones and evaluation of their antimalarial activity. Eur. J. Med. Chem. 2001, 36, 555–560. [Google Scholar] [CrossRef]
  89. Insuasty, B.; Montoya, A.; Becerra, D.; Quiroga, J.; Abonia, R.; Robledo, S.; Vélez, I.D.; Upegui, Y.; Nogueras, M.; Cobo, J. Synthesis of novel analogs of 2-pyrazoline obtained from [(7-chloroquinolin-4-yl)amino]chalcones and hydrazine as potential antitumor and antimalarial agents. Eur. J. Med. Chem. 2013, 67, 252–262. [Google Scholar] [CrossRef] [PubMed]
  90. Kalaria, P.N.; Karad, S.C.; Raval, D.K. A review on diverse heterocyclic compounds as the privileged scaffolds in antimalarial drug discovery. Eur. J. Med. Chem. 2018, 158, 917–936. [Google Scholar] [CrossRef] [PubMed]
  91. Hayat, F.; Moseley, E.; Salahuddin, A.; Van Zyl, R.L.; Azam, A. Antiprotozoal activity of chloroquinoline based chalcones. Eur. J. Med. Chem. 2011, 46, 1897–1905. [Google Scholar] [CrossRef]
  92. Kayser, O.; Kiderlen, A.F. In vitro leishmanicidal activity of naturally occurring chalcones. Phytother. Res. 2001, 15, 148–152. [Google Scholar] [CrossRef]
  93. Nielsen, S.F.; Kharazmi, A.; Christensen, S.B. Modifications of the α, β-double bond in chalcones only marginally affect the antiprotozoal activities. Bioorganic Med. Chem. 1998, 6, 937–945. [Google Scholar] [CrossRef]
  94. de Mello, M.V.P.; Abrahim-Vieira, B.A.; Domingos, T.F.S.; de Jesus, J.B.; de Sousa, A.C.C.; Rodrigues, C.R.; Souza, A.M.T. A comprehensive review of chalcone derivatives as antileishmanial agents. Eur. J. Med. Chem. 2018, 150, 920–929. [Google Scholar] [CrossRef] [PubMed]
  95. de Brito, D.H.A.; Almeida-Neto, F.W.Q.; Ribeiro, L.R.; Magalhães, E.P.; de Menezes, R.R.P.P.B.; Sampaio, T.L.; Martins, A.M.C.; Bandeira, P.N.; Marinho, M.M.; Marinho, E.S.; et al. Synthesis, structural and spectroscopic analysis, and antiproliferative activity of chalcone derivate (E)-1-(4-aminophenyl)-3-(benzo[b]thiophen-2-yl)prop-2-en-1-one in Trypanosoma cruzi. J. Mol. Struct. 2022, 1253, 132197. [Google Scholar] [CrossRef]
  96. Sangshetti, J.N.; Shinde, D.B.; Kulkarni, A.; Arote, R. Two decades of antifilarial drug discovery: A review. RSC Adv. 2017, 7, 20628–20666. [Google Scholar] [CrossRef] [Green Version]
  97. Kotra, V.; Ganapaty, S.; Adapa, S.R. Synthesis of a new series of quinolinyl chalcones as anticancer and anti-inflammatory agents. Indian J. Chem. 2010, 49B, 1109–1116. [Google Scholar]
  98. Ali, M.; Khan, M.; Zaman, K.; Wadood, A.; Iqbal, M.; Alam, A.; Shah, S.; Ashfaq Ur, R.; Yousaf, M.; Rafique, R.; et al. Chalcones: As Potent α-amylase Enzyme Inhibitors; Synthesis, In Vitro, and In Silico Studies. Med. Chem. 2021, 17, 903–912. [Google Scholar] [CrossRef]
  99. Herencia, F.; Ferrándiz, M.L.; Ubeda, A.; Domínguez, J.N.; Charris, J.E.; Lobo, G.M.; Alcaraz, M.J. Synthesis and anti-inflammatory activity of chalcone derivatives. Bioorg. Med. Chem. Lett 1998, 8, 1169–1174. [Google Scholar] [CrossRef]
  100. Chimenti, F.; Fioravanti, R.; Bolasco, A.; Chimenti, P.; Secci, D.; Rossi, F.; Yáñez, M.; Orallo, F.; Ortuso, F.; Alcaro, S. Chalcones: A Valid Scaffold for Monoamine Oxidases Inhibitors. J. Med. Chem. 2009, 52, 2818–2824. [Google Scholar] [CrossRef] [PubMed]
  101. Najafian, M.; Ebrahim-Habibi, A.; Hezareh, N.; Yaghmaei, P.; Parivar, K.; Larijani, B. Trans-chalcone: A novel small molecule inhibitor of mammalian alpha-amylase. Mol. Biol. Rep. 2011, 38, 1617–1620. [Google Scholar] [CrossRef]
  102. Abou-Zied, H.A.; Youssif, B.G.M.; Mohamed, M.F.A.; Hayallah, A.M.; Abdel-Aziz, M. EGFR inhibitors and apoptotic inducers: Design, synthesis, anticancer activity and docking studies of novel xanthine derivatives carrying chalcone moiety as hybrid molecules. Bioorganic Chem. 2019, 89, 102997. [Google Scholar] [CrossRef]
  103. Burmaoglu, S.; Ozcan, S.; Balcioglu, S.; Gencel, M.; Noma, S.A.A.; Essiz, S.; Ates, B.; Algul, O. Synthesis, biological evaluation and molecular docking studies of bis-chalcone derivatives as xanthine oxidase inhibitors and anticancer agents. Bioorganic Chem. 2019, 91, 103149. [Google Scholar] [CrossRef] [PubMed]
  104. Castaño, L.F.; Cuartas, V.; Bernal, A.; Insuasty, A.; Guzman, J.; Vidal, O.; Rubio, V.; Puerto, G.; Lukáč, P.; Vimberg, V. New chalcone-sulfonamide hybrids exhibiting anticancer and antituberculosis activity. Eur. J. Med. Chem. 2019, 176, 50–60. [Google Scholar] [CrossRef]
  105. Chouiter, M.I.; Boulebd, H.; Pereira, D.M.; Valentão, P.; Andrade, P.B.; Belfaitah, A.; Silva, A.M.S. New chalcone-type compounds and 2-pyrazoline derivatives: Synthesis and caspase-dependent anticancer activity. Future Med. Chem. 2020, 12, 493–509. [Google Scholar] [CrossRef]
  106. 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. A step-by-step synthesis of triazole-benzimidazole-chalcone hybrids: Anticancer activity in human cells+. J. Mol. Struct. 2020, 1204, 127487. [Google Scholar] [CrossRef]
  107. Gao, F.; Huang, G.; Xiao, J. Chalcone hybrids as potential anticancer agents: Current development, mechanism of action, and structure-activity relationship. Med. Res. Rev. 2020, 40, 2049–2084. [Google Scholar] [CrossRef]
  108. Hseu, Y.-C.; Chiang, Y.-C.; Vudhya Gowrisankar, Y.; Lin, K.-Y.; Huang, S.-T.; Shrestha, S.; Chang, G.-R.; Yang, H.-L. The in vitro and in vivo anticancer properties of chalcone flavokawain b through induction of ros-mediated apoptotic and autophagic cell death in human melanoma cells. Cancers 2020, 12, 2936. [Google Scholar] [CrossRef]
  109. 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]
  110. 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]
  111. 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. Bioorganic Chem. 2020, 95, 103565. [Google Scholar] [CrossRef]
  112. Yang, J.-L.; Ma, Y.-H.; Li, Y.-H.; Zhang, Y.-P.; Tian, H.-C.; Huang, Y.-C.; Li, Y.; Chen, W.; Yang, L.-J. Design, synthesis, and anticancer activity of novel trimethoxyphenyl-derived chalcone-benzimidazolium salts. ACS Omega 2019, 4, 20381–20393. [Google Scholar] [CrossRef]
  113. Yan, W.; Yang, T.; Yang, J.; Wang, T.; Yu, Y.; Wang, Y.; Chen, Q.; Bai, P.; Li, D.; Ye, H.; et al. SKLB060 Reversibly Binds to Colchicine Site of Tubulin and Possesses Efficacy in Multidrug-Resistant Cell Lines. Cell. Physiol. Biochem. Int. J. Exp. Cell. Physiol. Biochem. Pharmacol. 2018, 47, 489–504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Kłósek, M.; Mertas, A.; Król, W.; Jaworska, D.; Szymszal, J.; Szliszka, E. Tumor Necrosis Factor-Related Apoptosis-Inducing Ligand-Induced Apoptosis in Prostate Cancer Cells after Treatment with Xanthohumol-A Natural Compound Present in Humulus lupulus L. Int. J. Mol. Sci. 2016, 17, 837. [Google Scholar] [CrossRef] [Green Version]
  115. Sathish Kumar, K.; Kotra, V.; Phani Kumar, K.; Praveena Devi, C.B.; Anusha, N.; Hari Babu, B.; Adil, S.F.; Shaik, M.R.; Khan, M.; Al-Warthan, A.; et al. ZnCl2 catalyzed new coumarinyl-chalcones as cytotoxic agents. Saudi J. Biol. Sci. 2021, 28, 386–394. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Core structure of chalcone molecule.
Figure 1. Core structure of chalcone molecule.
Molecules 27 07062 g001
Figure 2. Claisen-Schmidt condensation reaction.
Figure 2. Claisen-Schmidt condensation reaction.
Molecules 27 07062 g002
Figure 3. Heterocyclic ring-containing chalcone using a phase transfer catalyst.
Figure 3. Heterocyclic ring-containing chalcone using a phase transfer catalyst.
Molecules 27 07062 g003
Figure 4. The derivatives of Chalcones from natural sources.
Figure 4. The derivatives of Chalcones from natural sources.
Molecules 27 07062 g004
Figure 5. Chemical reactions from Chalcones derivatives.
Figure 5. Chemical reactions from Chalcones derivatives.
Molecules 27 07062 g005
Figure 6. Semisynthetic derivatives of chalcones.
Figure 6. Semisynthetic derivatives of chalcones.
Molecules 27 07062 g006
Figure 7. Pharmacological properties of chalcone and its derivatives (Information was sourced from references [19,48,49,50,51,52,53].
Figure 7. Pharmacological properties of chalcone and its derivatives (Information was sourced from references [19,48,49,50,51,52,53].
Molecules 27 07062 g007
Figure 8. Antibacterial mechanisms of chalcone and its derivatives. Antibiotic resistance bacteria inhibit the function of antibiotic or antimicrobial agents by activating the efflux pump, modifying the active site of the enzyme, and degrading the enzyme. However, chalcone and its derivatives inhibit or block the functions of resistance plasmids through different mechanisms. Additionally, they inhibit DNA gyrase, DNA replication, and FtsZ. Therefore, bacteria cannot cell divide, and ultimately the bacteria die (Information was sourced from reference [61].
Figure 8. Antibacterial mechanisms of chalcone and its derivatives. Antibiotic resistance bacteria inhibit the function of antibiotic or antimicrobial agents by activating the efflux pump, modifying the active site of the enzyme, and degrading the enzyme. However, chalcone and its derivatives inhibit or block the functions of resistance plasmids through different mechanisms. Additionally, they inhibit DNA gyrase, DNA replication, and FtsZ. Therefore, bacteria cannot cell divide, and ultimately the bacteria die (Information was sourced from reference [61].
Molecules 27 07062 g008
Table 1. Antibacterial activity of the different types of chalcone and its derivatives [57,59,61,62].
Table 1. Antibacterial activity of the different types of chalcone and its derivatives [57,59,61,62].
No.Materials TestedAntimicrobial AssayTest Microorganism (MIC, µg/mL)AntibioticAntimicrobial Effect
S. aureus ATCC 29213E. faecalis ATCC 29212E. coli
ATCC 25922
S. enterica ATCC 1307
1(E)-N-(2-((4-cinnamoylphenyl)amino)-2-oxoethyl)-N,N-dimethyloctan-1-aminium chlorideBMD1234VAN, MEMStrong
2(E)-N-(2-((4-(3-(3-chlorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N dimethyloctan-1-aminium chlorideBMD2248VAN, MEMStrong
3(E)-N-(2-((4-(3-(3-chlorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyloctan-1-aminium chlorideBMD2144VAN, MEMStrong
4(E)-N-(2-((4-(3-(4-fluorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyloctan-1-aminium chlorideBMD2224VAN, MEMStrong
5(E)-N-(2-((4-(3-(3-fluorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyloctan-1-aminium chlorideBMD3264>128>128VAN, MEMNot good
6(E)-N-(2-((4-(3-(2-fluorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethylbutan-1-aminium chlorideBMD161664128VAN, MEMGood except S. enterica
7(E)-N-(2-((4-(3-(2-fluorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyloctan-1-aminium chlorideBMD0.5124VAN, MEMStrong
8(E)-N-(2-((4-(3-(2-fluorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyldodecan-1-aminium chlorideBMD481632VAN, MEMGood
9(E)-N-(2-((4-(3-(2-fluorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyltetradecan-1-aminium chlorideBMD8>128>128>128VAN, MEMEffective against S. aureus only
10(E)-N-(2-((4-(3-(2-fluorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyloctadecan-1-aminium chlorideBMD>128>128>128>128VAN, MEMNo activity
11(E)-N-(2-((4-(3-(2,3-difluorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyloctan-1-aminium chlorideBMD0.5228VAN, MEMStrong
12(E)-N-(2-((4-(3-(2,4-difluorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyloctan-1-aminium chlorideBMD2244VAN, MEMStrong
13(E)-N-(2-((4-(3-(2,6-difluorophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyloctan-1-aminium chlorideBMD3232128128VAN, MEMFair
14(E)-N-(2-((4-(3-(2-ethoxy-5-nitrophenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyloctan-1-aminium chlorideBMD82>128>128VAN, MEMStrong against S. aureus and E. faecalis
15(E)-N-(2-((4-(3-(4-(tert-butyl)phenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N, N-dimethyloctan-1-aminium chlorideBMD24816VAN, MEMStrong
16(E)-N,N-dimethyl-N-(2-oxo-2((4(3(2(trifluoromethyl)phenyl)acryloyl)phenyl)amino)ethyl)octan-1-aminium chlorideBMD241616VAN, MEMStrong
17(E)-N,N-dimethyl-N-(2-oxo-2-((4-(3-(p-tolyl)acryloyl)phenyl)amino)ethyl)octan-1-aminium chlorideBMD1148VAN, MEMStrong
18(E)-N-(2-((4-(3-(4-methoxyphenyl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyloctan-1-aminium chlorideBMD1244VAN, MEMStrong
19(E)-N,N-dimethyl-N-(2-((4-(3-(naphthalen-2-yl)acryloyl)phenyl)amino)-2-oxoctan-1-aminium chlorideBMD2288VAN, MEMStrong
20(E)-N,N-dimethyl-N-(2-oxo-2-((4-(3-(pyridin4yl)acryloyl)phenyl)amino)ethyl)butan-1-aminium chlorideBMD128128>128>128VAN, MEMNot good
21(E)-N,N-dimethyl-N-(2-oxo-2-((4-(3-(pyridin-4yl)acryloyl)phenyl)amino)ethyl)octan-1-aminium chlorideBMD481632VAN, MEMGood
22(E)-N,N-dimethyl-N-(2-oxo-2-((4-(3-(pyridin-3-yl)acryloyl)phenyl)amino)ethyl)octan-1-aminium chlorideBMD481632VAN, MEMGood
23(E)-N,N-dimethyl-N-(2-oxo-2-((4-(3-(pyridinedin-2yl)acryloyl)phenyl)amino)ethyl)octan-1-aminium chlorideBMD481632VAN, MEMGood
24(E)-N-(2-((4-(3-(6-bromopyridin-2-yl)acryloyl)phenyl)amino)-2-oxoethyl)-N, N-dimethyloctan-1-aminium chlorideBMD481632VAN, MEMGood
25(E)-((2-((4-(3-(3-fluoropyridin-2-yl)acryloyl)phenyl)amino)-2-oxoethyl)(methyl)(octyl)-l4-azanyl)methylium chlorideBMD28816VAN, MEMStrong
26(E)-N-(2-((4-(3-(furan-2-yl)acryloyl)phenyl)amino)-2-oxoethyl)-N,N-dimethyloctan-1-aminium chlorideBMD14416VAN, MEMStrong
27(E)-N,N-dimethyl-N-(2-oxo-2-((4-(3-(thiophen-2-yl)acryloyl)phenyl)amino)ethyl)butan-1-aminium chlorideBMD0.5148VAN, MEMStrong
28(E)-N, N-dimethyl-N-(2-oxo-2-((4-(3-(thiophen-2-yl)acryloyl)phenyl)amino)ethyl)octan-1-aminium chlorideBMD323264128VAN, MEMFair
29(E)-N,N-dimethyl-N-(2-oxo-2-((3-(3-(thiophen2-yl)acryloyl)phenyl)amino)ethyl) octan-1-aminium chlorideBMD241616VAN, MEMStrong
S. aureus ATCC 25923B. cereus ATCC 11778E. coli
ATCC 25922
P. aeruginosa ATCC 27853
302,2′,4,4′,5,5′-hexamethoxychalconeBMD>2000>2000>20002000TETNot active
312′-hydroxy-4,4’,5′-trimethoxychalconeBMD2000100020001000TETNot active
323,4-methylenedloxy-2′-3′,4′,6′-tetramethoxychalconeBMD>2000200020002000TETNot active
334,4′-dimethoxychalconeBMD>2000>2000>20002000TETNot active
343’,4′-dimethoxychalconeBMD1000100020001000TETNot active
352-hydroxy’-3’,4′-dimethoxychalconeBMD1000100010001000TETNot active
362′-acetoxy-3′-4′,4’,6′-tetramethoxychalconeBMD2000200010001000TETNot active
372,3′,4,4′,5-pentamethoxychalconeBMD2000200010002000TETNot active
382,2′,4′,5-tetramethoxychalconeBMD100025020002000TETActive against B. cereus
392,3,3′,4,4′,6-hexamethoxychalconeBMD2000200020002000TETNot active
40CordoinBMD2000200020002000TETNot active
414-hydroxycordoinBMD2000100020002000TETNot active
42IsocordoinBMD2000200020002000TETNot active
434-hydroxyisocordoinBMD31.231.210001000TETStrong for S. aureus and B. cereus
44DerricinBMD2000200020002000TETNot active
452-hydroxyderricinBMD2000200020002000TETNot active
463-hydroxyderricinBMD2000100020002000TETNot active
474-hydroxyderricinBMD7.83.920002000TETStrong
484-methoxyderricinBMD2000200020002000TETNot active
492′,4,4′-trihydroxychalconeBMD62.562.520002000TETStrong
502’, 4,4’-trihydroxy-3-preny’-3’geranylchalconeBMD31.215.610001000TETStrong
512’, 4,4’-trihydroxy’-3’geranylchalconeBMD31.215.610001000TETStrong for S. aureus and B. cereus
524-hydroxyisolonchocarpinBMD1000100020002000TETNot active
53LonchocarpinBMD2000200010002000TETNot active
544-hydroxylonchocarpinBMD2000200020002000TETNot active
554-hydroxyisolonchocarpinBMD2000100020002000TETNot active
56IsolonchocarpinBMD2000200010001000TETNot active
574-hydroxy-4′-methoxychalconeBMD50050020001000TETNot active
582-hydroxydihydrochalconeBMD2000200020002000TETNot active
592’,4,5′-trihydroxy-3,4-methylene-dioxy-dihydrochalconeBMD2000200020002000TETNot active
602’,4,5′-trihydroxy-dihydrochalconeBMD250250>20002000TETActive against S. aureus and B. cereus
S. aureusB. subtilisS. typhimuriumP. aeruginosa
611,3-Bis(4-chlorophenyl)-3-(phenylsulfonyl)propan-1-oneBMD 250 15.62 125 125 AMP, KMNStrong
621-Phenyl-3-(4-chlorophenyl)-3-(phenylsulfonyl)propane-1-oneBMD 125 62.5 125 62.5 AMP, KMNStrong
631-(4-Chlorophenyl)-3-(3-nitrophenyl)-3-phenylsulfonylprop-ane-1-oneBMD 125 62.5 62.5 62.5 AMP, KMNStrong
641-(4-Bromophenyl)-3-phenyl-3-(phenylsulfonyl) propane-1-oneBMD 125 62.5 250 62.5 AMP, KMNStrong
651-(4-Bromophenyl)-3-(3,4-dimethoxyphenyl)-3-(phenylsulfonyl)propane-1-oneBMD 62.5 62.5 1.95 62.5 AMP, KMNStrong
661-(4-Bromophenyl)-3-(3,4,5-trimethoxyphenyl)-3-(phenylsulfonyl) propane-1-oneBMD 125 31.25 1.95 125 AMP, KMNStrong
671-Phenyl-3-phenyl- 3-phenylsulfonylpropane-1-oneBMD 62.5 31.25 1.95 125 AMP, KMNStrong
681,5-Di(4-methylphenyl)-1,5-bis(phenylsulfonyl)pentan-3-oneBMD 62.5 62.5 31.25 62.5 AMP, KMNStrong
691,5-Di(4-chlorophenyl)-1,5-bis(phenylsulfonyl)pentan-3-oneBMD 250 31.25 15.62 250 AMP, KMNStrong
701,5-Di(phenyl)-1,5-bis(phenylsulfonyl)pentan-3-oneBMD 62.5 125 250 62.5 AMP, KMNStrong
711,5-Di(4-methoxyphenyl)-1,5-bis(phenylsulfonyl)pentan-3-oneBMD 250 31.25 62.5 125 AMP, KMNStrong
MIC: Minimum inhibitory concentration; BMD: Broth microdilution; VAN: Vancomycin (MIC: 2 µg/mL for S. aureus); MEM: Meropenem (MIC: <0.125 µg/mL for E. coli); TET: Tetracycline (MIC: 1 µg/mL for S. aureus, 0.25 µg/mL for B. cereus, 2 µg/mL cereus, 32 µg/mL for P. aeruginosa); AMP: Ampicillin (MIC: 500 µg/mL for S. aureus, 7.81 µg/mL for B. subtilis, 15.62 µg/mL for S. typhimurium, 250 µg/mL for P. aeruginosa); KMN: Kanamycin (MIC: 500 µg/mL for S. aureus, 250 µg/mL for B. subtilis, 125 µg/mL for S. typhimurium, 125 µg/mL for P. aeruginosa).
Table 2. Antifungal activity of chalcone and its derivatives [66].
Table 2. Antifungal activity of chalcone and its derivatives [66].
No.Materials Tested
Drug
Test AssayAntifungal Strains (MIC μg/mL)Standard AntifungalAntifungal Effect
A. nigerC. albicans
611,3-Bis(4-chlorophenyl)-3-(phenylsulfonyl)propan-1-oneBMD15.6231.25AMP-B, NSNVery strong
621-Phenyl-3-(4-chlorophenyl)-3-(phenylsulfonyl)propane-1-oneBMD1251.95AMP-B, NSNVery strong
631-(4-Chlorophenyl)-3-(3-nitrophenyl)-3-phenylsulfonylprop-ane-1-oneBMD62.51.95AMP-B, NSNVery strong
641-(4-Bromophenyl)-3-phenyl-3-(phenylsulfonyl) propane-1-oneBMD62.5125AMP-B, NSNVery strong
651-(4-Bromophenyl)-3-(3,4-dimethoxyphenyl)-3-(phenylsulfonyl)propane-1-oneBMD15.6215.62AMP-B, NSNVery strong
661-(4-Bromophenyl)-3-(3,4,5-trimethoxyphenyl)-3-(phenylsulfonyl) propane-1-oneBMD7.813.9AMP-B, NSNVery strong
671-Phenyl-3-phenyl- 3-phenylsulfonylpropane-1-oneBMD7.813.9AMP-B, NSNVery strong
681,5-Di(4-methylphenyl)-1,5-bis(phenylsulfonyl)pentan-3-oneBMD31.251.95AMP-B, NSNVery strong
691,5-Di(4-chlorophenyl)-1,5-bis(phenylsulfonyl)pentan-3-oneBMD1251.95AMP-B, NSNVery strong
701,5-Di(phenyl)-1,5-bis(phenylsulfonyl)pentan-3-oneBMD125250AMP-B, NSNStrong
711,5-Di(4-methoxyphenyl)-1,5-bis(phenylsulfonyl)pentan-3-oneBMD125125AMP-B, NSNStrong
BMD: Broth microdilution; AMP: Amphotericin-B (MIC: 500 μg/mL for A. niger, 15.62 μg/mL for C. albicans), NSN: Nystatin (MIC: 7.81 μg/mL for A. niger and C. albicans).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dhaliwal, J.S.; Moshawih, S.; Goh, K.W.; Loy, M.J.; Hossain, M.S.; Hermansyah, A.; Kotra, V.; Kifli, N.; Goh, H.P.; Dhaliwal, S.K.S.; et al. Pharmacotherapeutics Applications and Chemistry of Chalcone Derivatives. Molecules 2022, 27, 7062. https://doi.org/10.3390/molecules27207062

AMA Style

Dhaliwal JS, Moshawih S, Goh KW, Loy MJ, Hossain MS, Hermansyah A, Kotra V, Kifli N, Goh HP, Dhaliwal SKS, et al. Pharmacotherapeutics Applications and Chemistry of Chalcone Derivatives. Molecules. 2022; 27(20):7062. https://doi.org/10.3390/molecules27207062

Chicago/Turabian Style

Dhaliwal, Jagjit Singh, Said Moshawih, Khang Wen Goh, Mei Jun Loy, Md. Sanower Hossain, Andi Hermansyah, Vijay Kotra, Nurolaini Kifli, Hui Poh Goh, Sachinjeet Kaur Sodhi Dhaliwal, and et al. 2022. "Pharmacotherapeutics Applications and Chemistry of Chalcone Derivatives" Molecules 27, no. 20: 7062. https://doi.org/10.3390/molecules27207062

Article Metrics

Back to TopTop