Next Article in Journal
Adipic Acid Route: Oxidation of Cyclohexene vs. Cyclohexane
Next Article in Special Issue
New Bromo- and Iodo-Hydroxylactones with Two Methyl Groups Obtained by Biotransformation of Bicyclic Halolactones
Previous Article in Journal
Recent Advances in the Application of Metal–Organic Frameworks for Polymerization and Oligomerization Reactions
Previous Article in Special Issue
Biotechnological Approach for the Production of Enantiomeric Hydroxylactones Derived from Benzaldehyde and Evaluation of Their Cytotoxic Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tel-Cu-NPs Catalyst: Synthesis of Naphtho[2,3-g]phthalazine Derivatives as Potential Inhibiters of Tyrosinase Enzymes and Their Investigation in Kinetic, Molecular Docking, and Cytotoxicity Studies

1
Research Department of Chemistry, Nehru Memorial College (Affiliated to Bharathidasan University), Puthanampatti 621007, Tamil Nadu, India
2
Department of Zoology, College of Sciences, King Saud University (KSU), P.O. Box 2455, Riyadh 11451, Saudi Arabia
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(12), 1442; https://doi.org/10.3390/catal10121442
Submission received: 13 November 2020 / Revised: 3 December 2020 / Accepted: 4 December 2020 / Published: 9 December 2020
(This article belongs to the Special Issue Catalysis in the Synthesis of Biologically Active Compounds)

Abstract

:
Novel one-pot synthesis naphtho[2,3-g]phthalazine (1a1k) of Mannich base derivatives can be achieved via grindstone chemistry using a Tel-Cu-NPs (telmisartan-copper nanoparticles) catalyst. This method offers efficient mild reaction conditions and high yields. Tyrosinase inhibitory activity was evaluated for all synthesized compounds, along with analysis of kinetic behavior and molecular docking studies. The synthesized compound, 1c was (IC50 = 11.5 µM) more active than kojic acid (IC50 = 78.0 µM). Lineweaver Burk plots were used to analyze the kinetic behavior of the most active compound 1c, it was reversible and competitive behavior. Compound 1c and kojic acid occurred in the presence of 2-hydroxyketone, which has the same inhibitory mechanism. The molecular docking of compound 1c and the control kojic acid were docked against 2Y9X protein via the Schrodinger Suite. The compound 1c showed a respectable dock score (−5.6 kcal/mol) compared to kojic acid with a dock score of (−5.2 kcal/mol) in the 2Y9X protein. Cytotoxicity activity was also evaluated by using HepG2 (liver), MCF-7 (breast), and HeLa (cervical) cancer cell lines, and high activity for 1c (GI50 = 0.01, 0.03, and 0.04 µM, respectively) against all cell lines was found compared to standard and other compounds. Therefore, this study succeeded in testing a few promising molecules as potential antityrosinase agents.

1. Introduction

Out of several green chemistry methodologies, one simple and rapid methodology to prepare chemical and pharmaceutical compounds is via grindstone chemistry. This method was developed by Toda et al., who showed that many, but not all, reactions can be carried out by just grinding or triturating the solids together [1]. The Mannich reaction, which is a popular reaction in organic and medicinal chemistry domains [2], is a condensation reaction that forms an imminium-ion by enol-forming a carbonyl compound. The presently known Mannich type reactions face significant challenges. One of the prime challenges is the reaction time. Other challenges include harsh reaction conditions, toxicity, catalyst requirements, and the tedious methods required to separate and purify the final product(s). Some of the proposed methodologies to address these challenges include synthetic methodologies; for example, the use of Lewis acids or bases, ultrasound irradiation, or microwave, or the use of solubilizing agents or surfactant-type catalysts to address the insolubility of most organic compounds in water [3]. In addition, some of the known green trends in the Mannich reaction include ball-milling solvent-free [4], or with ionic liquid medium [5], nanoparticles [6], or enzymes in bio-catalytic conditions [7,8]. These green trends in the Mannich reaction also suffer with the challenges outlined above. Tyrosinase is a copper-containing mixed-function oxidase which is usually scattered in microorganisms, plants, and animals. Tyrosinase transforms into dopaquinone, which signifies the first step of melanin biosynthesis. o-quinones can polymerize naturally to melanin pigments [9]. o-diphenols are converted to o-quinones [10], which are involved in melanin synthesis in two separate reactions, and melanin is a biopolymer that is responsible for pigmentation [11]. However, the pigmentation can cause some dermatological disorders [12]. Tyrosinase is linked to neurodegeneration diseases associated with the creation of dopaquinone by oxidation of dopamine, which results in neuronal damage and cell death [13]. As a result, the medical, agricultural, and cosmetic industries are looking for effective tyrosinase inhibitors with minimal side effects. Derivatives of 9,10-anthraquinone have displayed various pharmacological effects, including laxative [14], anticancer [15], anti-inflammatory [16], antiarthritic [17], antifungal [18], antibacterial [19], antiviral [20], antiplatelet [21], and neuroprotective effects [22]. In addition, anthraquinone and its analogues have excellent properties, such as high excitation, emission, and absorption coefficients within the visible wavelength region, and low toxicity [23,24]. Some researchers have reported that benzo[g]phthalazines have the ability to form dinuclear complexes with Cu(II) when assayed with 1,4-bis(alkylamino)benzo[g]phthalazines. However, there has been no research focused on naphtho[g]phthalazine synthesis and biological screening [25]. This study focused on naphtho[2,3-g]phthalazine models covered by anthraquinone analogues in order to analyze the effect of antityrosinase activity.
Based on the above observations, the tyrosinase inhibitory activities of naphtho[2,3-g]phthalazine derivatives have not been previously studied. In this study, we designed new compounds bearing a carbonyl and hydroxyl group on the naphtho[2,3-g]phthalazine ring (Figure 1). Thus, we studied a new naphtho[2,3-g]phthalazine core belonging to the group of anthraquinone derivatives, and evaluated their anti-tyrosinase activity in vitro. Our kinetic studies also allowed us to discuss and compare molecular docking results compared with wet lab results.

2. Results and Discussion

2.1. Chemistry

2.1.1. Synthesis of Catalysis

Scheme 1 shows the route of preparation of telmisartan nanoparticles (Tel-Cu-NPs). First, the CuCl2·2H2O was mixed with telmisartan in ethanol with grinding for up to 10 min. A light blue solid material was obtained, and this solid material was mixed with 1 M NaOH and maintained at pH = 7. The obtained solid material was converted into a sky blue precipitate, followed by addition of water. The final solid material was filtered and purified with suitable techniques. The synthesized Cu-NPs were confirmed by XRD, SEM, and EDX analysis.

2.1.2. Powder X-Ray Diffraction Studies

The structural analysis and phase crystallinity of the synthesized Cu-NPs were examined via the powder X-ray diffraction method. Figure 2 shows the Cu nanoparticle diffraction pattern at a 2θ value of 43.20°, 50.42°, and 74.15° with respect to the (111), (200), and (220) planes individually. A copper cubic lattice formation can be seen. JCPDS No. 040836 indicates that there was good agreement compared with the standard pattern for a pure face centered cubic of Cu-NPs. There were no impurity peaks. From the observed main diffracted peak, the average crystalline size could be calculated using the Scherer equation:
D ( h k l ) = k λ β cos θ ,
where D(hkl) is the size, k is the shape constant (0.89), λ is the incident X-ray (Cu kα source, λ = 0.15405 nm), β is the full width at half-maximum, and θ is the incident angle of the X-ray. The average crystallite size of the synthesized copper nanoparticles was 25.41 nm.

2.1.3. SEM and EDX Analysis

Figure 3 shows the morphology of the SEM images of the Cu-NPs obtained in water. A 200 nm particle size was found, conforming to the formation of Cu nanoparticles from telmisartan. Figure 4 shows the EDX of Cu-NPs that exhibited the elements Cu, C, and O in the nanoparticles, conformed by the formation of the nanoparticle.
Naphtho[2,3-g]phthalazine (1a1k) was synthesized as shown in Scheme 2. First, the 1,4-dihydroxy-anthraquinone of 1 with aldehyde (2.1 equiv.) in the presence of hydrazine hydrate (1.0 equiv.) and Tel-Cu-NPs catalysis in EtOH at RT gave naphtho[2,3-g]phthalazine with a 95% yield. Compounds of 1a1k were purified by column chromatography, and their structure was determined by spectroscopic analysis. Assignment of the selected characteristic IR bands of compounds 1a1k exposed the functional groups OH, NH, and C=O, which obtained strong absorption band ranges between 3406–3414 cm¹, 3280–3286 cm¹, and 1750–1754 cm−1, respectively. The 1H-NMR spectra show signals at δ 5.30–5.39 and 9.20–9.86 ppm corresponding to the OH and NH protons, respectively. The ¹³C NMR spectrum shows signals from δ 187.1–187.6, 151.5–153.9, and 50.7–60.8 ppm, that correspond to the C=O, -C-OH, and -C-N carbon atoms individually. In addition, mass spectra show that all compounds of molecular weight conformed to molecular ion peaks present in the mass spectral analysis.

2.1.4. Catalyst Recovery Studies

Figure 5 shows the recovery of the catalyst salvaged from at least 10 run times, with a slight loss in catalytic activity. The lessening of activity could be detected with the regenerated catalyst on salvaging due to the surface area of the catalyst during the reaction, or partial loss of the basic sites/regeneration. The application of the catalyst was inspected by optimizing the reaction conditions. A number of aldehydes were selected for the condensation reaction with the Cu-NPs (1 mole %) catalyst at room temperature in a solvent-free setting, and the yield is reported in Table 1. The Cu-catalyzed performance and reaction mechanism are shown in Scheme 3.

2.2. Biological Activity

2.2.1. Antityrosinase Activity

Inhibition of naphtho[2,3-g]phthalazine was tested using 3,4-dihydroxy-L-phenylalanine (L-DOPA) as the substrate. Kojic acid was used as one of the basic skin whitening elements, as a reference compound for this study. Table 2 indicates that values and effects of the naphtho[2,3-g]phthalazines (1a1k). Compound 1c, bearing a naphtho[2,3-g]phthalazine showed better activity, with an IC50 value of 11.5 µM, compared to kojic acid with an IC50 value of 78.0 µM.

2.2.2. Inhibitory Mechanism

2-Hydroxyketone is in compound 1a1k and kojic acid, and plays a major role in their mechanism. The compound 1c had the highest inhibition; the mechanism of inhibition is represented in Scheme 4 [26]. The kinetic behavior was studied using compound 1c at different concentrations with respect to L-DOPA, by mushroom tyrosinase. Figure 6 shows the lineweaver Burk plots, which indicated 1c is a competitive inhibitor. The result, due to the abscisse 1/[L-DOPA] is the reciprocal of the L-DOPA concentrations and reaction rate of tyrosinase activity.

2.2.3. Molecular Docking

The compound 1c and control Kojic acid were docked with 2Y9X protein via Schrodinger Suite. The compound 1c showed a dock score (−5.6 kcal/mol) higher than that of kojic acid, with a dock score of (−5.2 kcal/mol) in 2Y9X protein. The hydrogen bond stretches of compound 1c and kojic acid were strong hydrogen bonds, and 3.5Å in 2Y9X. The 1c forms three hydrogen bonds on the 2Y9X protein. The amino acid residues Arg268 (bond length: 5.85), Ser282 (bond length: 4.25), and Val283 (bond length: 3.62) were intricate in the hydrogen bonding interface. Residues of His61, His85, Phe90, His94, Phe192, Trp227, Val248, Gly249, Met257, Asn260, Thr261, His263, Phe264, Leu275, Pro277, Met280, Gly281, Pro284, His285, Ala286, Phe292, and His296 were involved in the hydrophobic interactions. The connections of 1c with the 2Y9X protein are shown in Figure 7. One hydrogen bond is formed from the control, kojic acid, in the 2Y9X protein receptor. The amino acid residue interaction with hydrogen bonding was involved in His263 (bond length: 1.79). Hydrophobic interactions were involved in the amino acid residues of Met280, Val283 and His296. The 2Y9X protein interactions with the control kojic acid are shown in Figure 8. The results of compound 1c show the better inhibition than the control kojic acid in mushroom tyrosinase protein 2Y9X. The tyrosinase enzyme inhibition values are listed in Table 3.

2.2.4. Cytotoxicity Activity

The 50% growth inhibition (GI50), tumor growth inhibition (TGI), and lethal concentration 50 (LC50) values were determined. Compound 1c was highly active against HepG2 cells (GI50:0.01 μM), MCF-7 (0.03 μM), HeLa (0.04 μM) when compared with doxorubicin and other compounds. All other compounds had moderate active in cytotoxicity screening. The cytotoxic screening results are presented in Table 4.

3. Experimental Section

3.1. Synthesis

FT IR (4000–400 cm−1) was recorded ED via Shimadzu 8201pc (Shimadzu, Tokyo, Japan). The ¹H and 13C NMR recording was via JEOL-300 Mhz (Jeol, Tokyo, Japan). The elementer analyzer model Varian EL III (Varian, Inc., Karlsruhe, Germany) was used for analysis of elemental presences. Thin-layer chromatography was used for purity checking.
Synthesis of Telmisartan Mediated Copper Nanoparticles (Tel-Cu-NPs)
The mixture of CuCl2· 2H2O (0.5 mmol, 5 mL), and telmisartan (1 mmol, 10 mL) was added to 10 mL of ethanol, then the reaction mixture was given 1 M NaOH with stirring and maintained at pH 7. The reaction mixture converted to a blue precipitate followed by the addition of 15 mL of water. The bile precipitate was obtained, then the precipitate was filtered and purified with suitable techniques.
General Preparation of Compounds (1a1k)
To a solution of 1,4-dihydroxy-anthraquinone (16.02 g, 0.10 mol) in EtOH (150 mL), aldehyde (0.30 mol), hydrazine hydrate (2.44 g, 0.02 mol), and catalyst Tel-Cu-NPs (1.0 mg) were mixed in a mortar at RT. Achievement of the product as checked by TLC. The insoluble crude product was filtered, then the final product was separated from the column chromatography using suitable eluting solvent. Similarly, compounds (1b1k) were synthesized by using the same method.
Catalyst recovery studies: In order to investigate the reusability of the catalyst, after completion of the reaction, copper nanoparticles were recovered by filtration of the mixture in a vacuum and thoroughly washed with ethylacetate. The final product was soluble in EtOAc solvent and we then separated the final product from the mixture. The copper nano catalyst was thoroughly washed with water, followed by filtering and then drying in an oven at 60 °C, and was then used for the next reaction without any reactivation [27].
Characterization of Target Compounds
5,12-dihydroxy-1,2,3,4-tetrahydronaphtho[2,3-g]phthalazine-6,11-dione (1a)
Yellow solid; yield 88%; mw: 296.28; mp: 126–127 °C; IR (KBr, cm−1): 3408, 3282, 2842, 1752, 1684, 804, 745; ¹H NMR (DMSO-d6) δ 9.20 (s, 2H), 8.29 (d, J = 11.6 Hz, 2H, Ar), 7.88 (2H, dd, J = 11.6 Hz, J = 11.2 Hz, Ar), 5.31 (s, 2H, OH), 3.91 (4H, s, CH2); 13C NMR (DMSO-d6) δ 187.1 (2C, C=O), 151.9 (2C, C-OH), 133.6, 132.1, 126.8 (6C, Ar), 132.0 (2C), 113.9 (2C), 50.7 (2C, C-N-N); EIMS(m/z): 296.30 [M]+; Anal. Calcd. For C16H12N2O4: C, 64.86; H, 4.08; N, 9.46; Found: C, 64.84; H, 4.10; N, 9.46;
5,12-dihydroxy-1,4-di(prop-1-en-1-yl)-1,2,3,4- tetrahydro naphtho[2,3-g]phthalazine-6,11-dione (1b)
Yellow solid; yield 85%; mw: 376.41; mp: 130–132 °C; IR (KBr): 3409, 3280, 2843, 1753, 1684, 806, 748 cm−1; ¹H NMR (300 MHz) δ 9.23 (s, 2H, NH), 8.27 (d, J = 11.4 Hz, 2H, Ar), 7.87 (dd, 2H, J = 11.6 Hz, J = 11.2 Hz, Ar), 6.08 (s, 2H), 5.41 (s, 2H, CH=CH), 5.39 (s, 2H, OH), 4.5 (s, 2H), 2.09 (s, 6H); 13C NMR (DMSO-d6) δ 187.3 (2C), 151.6 (2C, C-OH), 133.6, 132.1, 126.8 (6C, Ar), 127.1 (2C), 126.4 (2C), 126.2 (2C, C=C), 113.9 (2C), 60.4 (2C, C-N-N), 17.6 (2C, CH3); EI-MS m/z: 376.45 [M]+; Anal. Calcd. For C22H20N2O4: C, 70.20; H, 5.36; N, 7.44; Found: C, 70.22; H, 5.37; N, 7.43;
1,4-bis(2,6-dimethylhepta-1,5-dien-1-yl)-5,12-dihydroxy-1,2,3,4-tetrahydronaphtho[2,3-g]phthalazine-6,11-dione (1c)
Yellow solid, yield 84%; mw: 540.69; mp: 141–142 °C; IR (KBr): 3406, 3282, 2845, 1754, 1682, 804, 746 cm−1; ¹H NMR (300 MHz) δ 9.84 (s, 2H, NH), 8.26 (d, J = 11.2 Hz, 2H, Ar), 7.87 (dd, 2H, J = 11.4 Hz, J = 11.2 Hz, Ar), 5.80 (s, 2H, CH=C), 5.31 (s, 2H, OH), 5.20 (s, 2H, CH=C), 4.50 (s, 2H, CH), 2.0 (s, 8H, CH2), 1.82 (s, 12H, CH3), 1.70 (s, 6H); 13C NMR (75 MHz) δ 187.4 (2C), 151.5 (2C, C-OH), 135.7 (2C), 132.2 (2C), 133.6, 132.1, 126.8 (6C, Ar), 126.6 (2C), 123.5 (2C, C=C), 54.2 (2C, C-N-N), 113.9 (2C), 116.7 (2C, C=C), 39.7, 26.4 (4C, CH2), 24.6, 18.6, 16.4 (6C, CH3); EIMS (m/z) 540.72 [M]+; Anal. Calcd. For C34H40N2O4: C, 75.53; H, 7.46; N, 5.18; Found: C, 75.50; H, 7.48; N, 5.17;
5,12-dihydroxy-1,4-di(styryl)-1,2,3,4-tetrahydronaphtho[2,3-g]phthalazine-6,11-dione (1d)
Yellow solid, yield 83%; mw: 500.54; mp: 152–153 °C; IR (KBr, cm−1): 3408, 3280, 2843, 1751, 1683, 803, 742; ¹H NMR (DMSO-d6) δ 9.89 (s, 2H, NH), 8.26 (d, J = 11.0 Hz, 2H, Ar), 7.86 (dd, J = 11.1 Hz, J = 11.2 Hz, 2H, Ar), 7.24–7.40 (m, J = 11.6 Hz, 10H, Ph), 6.65 (s, 2H), 6.37 (s, 2H, CH=CH), 5.31(s, 2H, OH), 4.51 (2H, s, CH); 13C NMR (75 MHz, δ: 187.1 (2C), 151.9 (2C, C-OH), 126.6 (2C), 113.9 (2C), 129.0 (2C, C=C), 136.4, 128.6, 128.5, 127.9 (12C, Ph), 126.8, 132.8, 133.6 (6C, Ar), 123.3 (2C), 60.3 (2C, C-N-N); EIMS (m/z):500.58 [M]+; Anal. Calcd. For C32H24N2O4: C, 76.78; H, 4.83; N, 5.60; Found: C, 76.80; H, 4.80; N, 5.62
1,4-di(furan-2-yl)-5,12-dihydroxy-1,2,3,4-tetrahydronaphtho[2,3-g]phthalazine-6,11-dione (1e)
Yellow solid, yield 85%; mw: 428.39; mp: 150–151 °C; IR (KBr, cm−1): 3410, 3284, 2840, 1752, 1684, 805, 740; ¹H NMR (DMSO-d6) δ 9.86 (s, 2H, NH), 8.26 (d, J = 11.8 Hz, 2H, Ar), 7.86 (2H, dd, J = 11.9Hz, J = 11.2 Hz, Ar), 7.65 (d, 2H, J = 7.8 Hz, Furyl), 6.44 (dd, J = 7.8 Hz, J = 7.6 Hz, 2H, Furyl), 6.26 (d, 2H, J=, Furyl), 5.42 (s, 2H, CH), 5.30 (s, 2H, OH); 13C NMR (75 MHz) δ 187.2 (2C), 152.5, 106.7, 110.6, 142.1 (8C, Furyl), 151.6 (2C, C-OH), 133.6, 132.1, 126.8 (6C, Ar), 133.0 (2C), 113.9 (2C), 61.4 (2C, C-N-N); EIMS (m/z): 428.43 [M]+; Anal. Calcd. For C24H16N2O6: C, 67.29; H, 3.76; N, 6.54; Found: C, 67.30; H, 3.75; N, 6.55;
5,12-dihydroxy-1,4-diphenyl-1,2,3,4-tetrahydronaphtho[2,3-g]phthalazine-6,11-dione (1f)
Yellow solid, yield 82%; mw: 448.47; mp: 162–164 °C; IR (KBr, cm−1): 3412, 3286, 2842, 1754, 1686, 807, 742; ¹H NMR (DMSO-d6) δ 9.89 (s, 2H, NH), 8.22 (d, J = 11.8 Hz, 2H, Ar), 7.86 (dd,2H, J = 11.7 Hz, J = 11.6 Hz, Ar), 7.37 (d, 4H, J = 9.8 Hz, Ph), 7.33 (dd, 4H, J = 9.8 Hz, J = 9.2 Hz, Ph), 7.26 (dd, 2H, J = 9.8 Hz, J = 9.2 Hz, Ph), 5.29 (s, 2H, OH), 5.11 (2H, s, CH); 13C NMR (DMSO-d6) δ 187.6 (2C),153.9 (2C, C-OH), 142.7, 128.2, 129.2, 126.2 (12C, Ph), 134.9 (2C), 133.6, 132.1, 126.8 (6C, Ar), 114.1 (2C), 60.8 (2C, C-N-N); EIMS(m/z): 448.50 [M]+; Anal. Calcd. For C28H20N2O4: C, 74.99; H, 4.50; N, 6.25; Found: C, 75.01; H, 4.48; N, 6.26;
1,4-bis(4-chlorophenyl)-5,12-dihydroxy-1,2,3,4-tetrahydronaphtho[2,3-g]phthalazine-6,11-dione (1g)
Yellow solid, yield 84%; mw: 517.36; mp: 138–139 °C; IR (KBr, cm−1): 3413, 3285, 2840, 1752, 1684, 805, 740; ¹H NMR (DMSO-d6) δ 9.80 (s, 2H, NH), 8.25 (d, J = 11.7 Hz, 2H, Ar), 7.84 (2H, dd, J = 11.6 Hz, J = 11.2 Hz, 2H, Ar), 7.37 (d, J = 9.8 Hz, 4H, Ph), 7.17 (d, J = 9.8 Hz, J = 9.2 Hz, 4H, Ph), 5.32 (s, 2H, OH), 5.19 (s, 2H, CH); 13C NMR (DMSO-d6) δ: 187.1 (2C, C=O), 153.9 (2C, C-OH), 140.8, 131.8, 129.6, 129.3 (12C, Ph), 134.9 (2C), 133.6, 132.1, 126.8 (6C, Ar), 114.1 (2C), 60.6 (2C, C-N-N); EIMS (m/z): 517.40 [M]+; Anal. Calcd. For C28H18Cl2N2O4: C, 65.00; H, 3.51; N, 5.41; Found: C, 65.02; H, 3.53; N, 5.40;
5,12-dihydroxy-1,4-bis(4-hydroxyphenyl)-1,2,3,4-tetrahydronaphtho[2,3-g]phthalazine-6,11-dione (1h)
Yellow solid, yield 86%; mw: 480.47; mp: 158–159 °C; IR (KBr, cm−1): 3412, 3284, 2842, 1750, 1682, 804, 742; ¹H NMR (DMSO-d6) δ 9.80 (s, 2H, NH), 8.29 (d, J = 11.4 Hz, 2H, Ar), 7.85 (dd, J = 11.4 Hz, J = 11.2 Hz, 2H, Ar), 7.37 (d, J = 9.8 Hz, 4H, Ph), 7.17 (d, J = 9.8 Hz, 4H, Ph), 5.38(s, 4H, OH), 5.19 (s, 2H, CH); 13C NMR (DMSO-d6) δ 187.4 (2C), 153.2 (2C, C-OH), 156.0, 135.3, 116.4, 129.6, (12C, Ph), 134.9 (2C), 114.1 (2C), 133.6, 132.1, 126.8 (6C, Ar), 60.8 (2C, C-N-N); EIMS (m/z) 480.50 [M]+; Anal. Calcd. For C28H20N2O6: C, 69.99; H, 4.20; N, 5.83; Found: C, 69.97; H, 4.23; N, 5.85;
5,12-dihydroxy-1,4-bis(3-nitrophenyl)-1,2,3,4-tetrahydronaphtho[2,3-g]phthalazine-6,11-dione (1i)
Yellow solid, yield 85%; mw: 538.46; mp: 165–166 °C; IR (KBr, cm−1): 3410, 3280, 2840, 1752, 1680, 802,740; ¹H NMR (DMSO-d6) δ 9.76 (s, 2H, NH), 8.26 (d, J = 11.7 Hz, 2H, Ar), 7.86 (d, J = 9.8 Hz, 2H, Ph), 8.15 (s, 2H, Ph), 8.07 (d, J = 9.8 Hz, 2H, Ph), 7.76 (d, J = 9.8 Hz, 2H, Ph), 7.50 (dd, 2H, J = 9.8 Hz, J = 9.2 Hz, Ph), 5.36 (s, 2H, OH), 5.19 (s, 2H, CH); 13C NMR (DMSO-d6) δ 187.2 (2C), 153.5 (2C, C-OH), 143.6, 124.6, 148.4, 121.4, 130.1, 134.3 (12C, Ph), 134.9 (2C), 133.6, 132.1, 126.8 (6C, Ar), 114.1 (2C), 60.2 (2C, C-N-N); EIMS (m/z) 538.48 [M]+; Anal. Calcd. For C28H18N4O8: C, 62.46; H, 3.37; N, 10.40; Found: C, 62.47; H, 3.35; N, 10.38;
1,4-bis(4-(dimethylamino)phenyl)-5,12-dihydroxy-1,2,3,4-tetrahydronaphtho[2,3-g]phthalazine-6,11-dione (1j)
Yellow solid, yield 88%; mw: 534.61; mp: 169–171 °C; IR (KBr) 3412, 3282, 2842, 1754, 1682, 804, 742 cm−1; ¹H NMR (300 MHz) δ 9.78 (s, 2H, NH), 8.25 (d, J = 11.7 Hz, 2H, Ar), 7.86 (dd, 2H, J = 11.6 Hz, J = 11.2 Hz, Ar), 7.05 (d, J = 9.8 Hz, 4H, Ph), 6.60 (d, J = 9.8 Hz, 4H, Ph), 5.32 (s, 2H, OH), 5.19 (s, 2H), 3.01 (s, 12H, N(CH3)2); 13C NMR (75 MHz) δ 187.1 (2C), 153.9 (2C, C-OH), 148.6, 132.2, 113.4, 129.1 (12C, Ph), 134.9 (2C), 133.6, 132.1, 126.8 (6C, Ar), 114.1 (2C), 60.8 (2C, C-N-N), 41.3 (4C, N(CH3)2); EIMS (m/z) 534.65 [M]+; Anal. Calcd. For C32H30N4O4: C, 71.89; H, 5.66; N, 10.48; Found: C, 71.90; H, 5.68; N, 10.46; S, 8.96;
5,12-dihydroxy-1,4-bis(4-methoxyphenyl)-1,2,3,4-tetrahydronaphtho[2,3-g]phthalazine-6,11-dione (1k)
Yellow solid, yield 82%; mw: 508.52; mp: 137–139 °C; IR (KBr, cm−1): 3414, 3284, 2840, 1752, 1684, 802, 740; ¹H NMR (300 MHz) δ 9.86 (s, 2H, NH), 8.24 (d, J = 11.6 Hz, 2H, Ar), 7.89 (2H, dd, J = 11.6 Hz, J = 11.2 Hz, 2H, Ar), 7.12 (d, J = 9.8 Hz, 4H, Ph), 6.87 (d, J = 9.8 Hz, 4H, Ph), 5.31 (s, 2H, OH), 5.19 (s, 2H, CH), 3.77 (s, 6H, OCH3); 13C NMR (75 MHz) δ 187.4 (2C), 158.1, 135.0, 114.8, 129.2 (12C, Ph), 153.6 (2C, C-OH), 134.9 (2C), 133.6, 132.1, 126.8 (6C, Ar), 114.1 (2C), 60.3 (2C, C-N-N), 55.8 (2C, OCH3); EIMS(m/z) 508.54 [M]+; Anal. Calcd. For C30H24N2O6: C, 70.86; H, 4.76; N, 5.51; Found: C, 70.85; H, 4.77; N, 5.53;

3.2. Biological Activity

3.2.1. Anti-Tyrosinase Activity

The mushroom tyrosinase (powder, ≥1000 unit/mg solid, EC 1.14.18.1) inhibitory activities were measured spectrophotometrically, as defined previously with some modifications [28]. For preparation of the test solution, 12.428 U of mushroom tyrosinase was added to L-DOPA 1.5 mM and 0.1 mM sodium phosphate buffer (pH 6.5), and incubated at 30 °C for 2 min. A total of 3.0 mL of test sample was monitored at 475 nm by a Perkin Elmer Lambda 35 Spectrophotometer (Perkin Elmer Corp, Waltham, MA, USA). Standard kojic acid was used in this screening, and 50% inhibition (IC50) was calculated and analyzed in three independent experiments.
Tyrosinase inhibitory activity (%) = [(A − B) − (C − D)] / (A − B) × 100
where A = absorbance of incubation (after) of the blank solution; B = absorbance incubation (before) of the blank solution; C = absorbance of incubation (after) the blank solution; D = absorbance of incubation (before) the blank solution.

3.2.2. Molecular Docking

Agaricus bisporus-mushroom tyrosinase (PDBID: 2Y9X) was downloaded from http://www.rcsb.org [29,30].

Grid Generation and Molecular Docking

The tyrosinase structure was arranged via “Protein Preparation Wizard” by the Maestro interface in the Schrodinger Suite. Originally, bond orders were allotted and the target protein was added. Later, the protein was minimized to attain the congregated RMSD-root mean square deviation of 0.30Å via the OPLS_2005 force field. The active position (binding pocket) of the target protein was demarcated by literature survey and the Protein Data Bank [31,32,33]. Compound 3c and control kojic acid were drawn by using 2D sketcher in the Schrodinger Suite for docking studies. The studies were achieved by the target protein with a Glide docking procedure to forecast the conformational positions and docking scores (binding energies) of the ligands inside the active area of the protein. All over the docking mock-ups, both partial and suppleness of the active site remains was achieved by Glide/SP/XP and induced fit docking (IFD) methods. The inhibitor devouring minimum binding energy value that was best scoring, and the outcomes, were examined via the Discovery Studio 2019 software package.

3.2.3. Cytotoxic Activity

The compounds (1a1k) were evaluated for cytotoxic activity using the method in our literature report [34]. The details of the experiment are presented in the Supplementary Materials. The cytotoxic values were confirmed based on at least three independent evaluations.

4. Conclusions

Naphtho[2,3-g]phthalazine (1a1k) Mannich base derivatives were synthesized via the grindstone method using the Tel-Cu-NPs catalyst. We synthesized eleven naphtho[2,3-g]phthalazine derivatives (1a1k) and evaluated their anti-tyrosinase activity. Compound 1c significantly inhibited tyrosinase (IC50 = 11.5 µg/mL). Based on molecular docking, compound 1c had a good dock score (−5.6 kcal/mol) compared to kojic acid (−5.2 kcal/mol) in 2Y9X protein. Cytotoxicity was screened for in all compounds; 1c was highly active (GI50 = 0.01, 0.03, and 0.04 µM) against HepG2 (liver), MCF-7 (breast), and HeLa (cervical) cancer cell lines, compared to doxorubicin and other compounds. Therefore, these results suggest that naphtho[2,3-g]phthalazine derivatives could serve as a new class of anti-tyrosinase agents.

Supplementary Materials

The Supplementary Materials are available online at https://www.mdpi.com/2073-4344/10/12/1442/s1. Characterization data including 1H and 13C NMR spectra for all compounds.

Author Contributions

A.D. and S.A. methodology preparation for biological activity; K.S. software and docking result preparation and analysis chemical data analysis, A.I. investigation of all part of this manuscript chemistry and Biology. The manuscript was written writing—original draft preparation through the contributions of all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Researchers Supporting Project number (RSP-2020/27), King Saud University, Riyadh, Saudi Arabia.

Acknowledgments

This work was funded by Researchers Supporting Project number (RSP-2020/27), King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Toda, F.; Tanaka, K.; Sekikawa, A. Host–guest complex formation by a solid–solid reaction. J. Chem. Soc. Chem. Commun. 1987, 4, 279. [Google Scholar] [CrossRef]
  2. Mannich, C. Eine Synthese von β-ketonbasen. Arch. Pharm. 1917, 255, 261–276. [Google Scholar] [CrossRef] [Green Version]
  3. Manabe, K.; Kobayashi, S. Mannich-Type Reactions of Aldehydes, Amines, and Ketones in a Colloidal Dispersion System Created by a Brønsted Acid−Surfactant-Combined Catalyst in Water. Org. Lett. 1999, 1, 1965–1967. [Google Scholar] [CrossRef]
  4. Jing-Bo, Y.; Gang, P.; Zhi-Jiang, J.; Zi-Kun, H.; Wei-Ke, S. Mechanochemical Oxidative Mannich Reaction: Evaluation of Chemical and Mechanical Parameters for the Mild and Chemoselective Coupling of N-tert-butoxycarbonyltetrahydroquinolines and Ketones. Eur. J. Org. Chem. 2016, 32, 5340–5344. [Google Scholar]
  5. Khanna, G.; Aggarwal, K.; Khuran, J.L. An efficient and confluent approach for the synthesis of novel 3,4-dihydro-2H-naphtho[2,3-e][1,3]oxazine-5,10-dione derivatives by a three component reaction in ionic liquid. RSC Adv. 2015, 5, 46448–46454. [Google Scholar] [CrossRef]
  6. Ghomi, J.S.; Zahedi, S. Novel ionic liquid supported on Fe3O4 nanoparticles and its application as a catalyst in Mannich reaction under ultrasonic irradiation. Ultrason Sonochem 2017, 34, 916–923. [Google Scholar] [CrossRef]
  7. Ling-Ling, W.; Yang, X.; Da-Cheng, Y.; Zhi, G.; Yan-Hong, H. Biocatalytic asymmetric Mannich reaction of ketimines using wheat germ lipase, Catal. Sci. Technol. 2016, 6, 3963–3970. [Google Scholar]
  8. Mostafa, A.A.F.; SathishKumar, C.; Al-Askar, A.M.; Sayed, S.R.; SurendraKumar, R.; Idhayadhulla, A. Synthesis of novel benzopyran-connected pyrimidine and pyrazole derivatives via a green method using Cu(II)-tyrosinase enzyme catalyst as potential larvicidal, antifeedant activities. RSC Adv. 2019, 9, 25533–25543. [Google Scholar] [CrossRef] [Green Version]
  9. Whitaker, J.R. Polyphenol Oxidase. In Food Enzymes; Springer: Boston, MA, USA, 1995. [Google Scholar] [CrossRef]
  10. Fenoll, L.G.; Penalver, M.J.; RodríguezLopez, J.N.; Varon, R.; GarcíaCanovas, F.; Tudela, J. Tyrosinase kinetics: Discrimination between two models to explain the oxidation mechanism of monophenol and diphenol substrates. Int. J. Biochem. Cell. Biol. 2004, 36, 235–246. [Google Scholar] [CrossRef]
  11. Costin, G.E.; Hearing, V.J. Human skin pigmentation: Melanocytes modulate skin color in response to stress. FASEB J. 2007, 21, 976–994. [Google Scholar] [CrossRef]
  12. Vashi, N.A.; Kundu, R.V. Facial hyperpigmentation: Causes and treatment. Br. J. Dermatol. 2013, 169, 41–56. [Google Scholar] [CrossRef] [PubMed]
  13. Hasegawa, T. Tyrosinase-Expressing Neuronal Cell Line as in vitro Model of Parkinson’s disease. Int. J. Mol. Sci. 2010, 11, 1082–1089. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Van Gorkom, B.A.; De Vries, E.G.; Karrenbeld, A.; Leibeuker, J.H. Review article: Anthranoid laxatives and their potential carcinogenic effects, Aliment. Pharmacol. Ther. 1999, 13, 443–452. [Google Scholar] [CrossRef] [PubMed]
  15. Shrestha, J.P.; Fosso, M.Y.; Bearss, J.; Chang, C.W.T. Synthesis and anticancer structure activity relationship investigation of cationic anthraquinone analogs. Eur. J. Med. Chem. 2014, 77, 96–102. [Google Scholar] [CrossRef]
  16. Khan, N.; Karodi, R.; Siddiqui, A.; Thube, S.; Rub, R. Development of anti-acne gel formulation of anthraquinones rich fraction from Rubia cordifolia (Rubiaceae). Int. J. Appl. Res. Nat. Prod. 2011, 4, 28–36. [Google Scholar]
  17. Davis, R.; Agnew, P.; Shapiro, E. Antiarthritic activity of anthraquinones found in aloe for podiatric medicine. J. Am. Podiatr. Med. Assoc. 1986, 76, 61–66. [Google Scholar] [CrossRef]
  18. Wuthi-udomlert, M.; Kupittayanant, P.; Gritsanapan, W. In vitro evaluation of antifungal activity of anthraquinone derivatives of Senna alata. J. Health Res. 2010, 24, 117–122. [Google Scholar]
  19. Fosso, M.Y.; Chan, K.Y.; Gregory, R.; Chang, C.W.T. Library Synthesis and Antibacterial Investigation of Cationic Anthraquinone Analogs. ACS Comb. Sci. 2012, 14, 231–235. [Google Scholar] [CrossRef]
  20. Barnard, D.L.; Fairbairn, D.W.; O’Neill, K.L.; Gage, T.L.; Sidwell, R.W. Anti-human cytomegalovirus activity and toxicity of sulfonated anthraquinones and anthraquinone derivatives. Antiviral. Res. 1995, 28, 317–329. [Google Scholar] [CrossRef]
  21. Seo, E.J.; Ngoc, T.M.; Lee, S.M.; Kim, Y.S.; Jung, Y.S. Chrysophanol-8-O-glucoside, an Anthraquinone Derivative in Rhubarb, Has Antiplatelet and Anticoagulant Activities. J. Pharmacol. Sci. 2012, 118, 245–254. [Google Scholar] [CrossRef] [Green Version]
  22. Jackson, T.C.; Verrier, J.D.; Kochanek, P.M. Anthraquinone-2-sulfonic acid (AQ2S) is a novel neurotherapeutic agent. Cell Death Dis. 2013, 4, 451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Bhasin, A.K.; Chauhan, P.; Chaudhary, S. A novel sulfur-incorporated naphthoquinone as a selective “turn-on” fluorescence chemical sensor for rapid detection of Ba2+ ion in semi-aqueous medium. Sens. Actuators B Chemical. 2019, 294, 116–122. [Google Scholar] [CrossRef]
  24. Kumar, P.; Ghosh, A.; Jose, D.A. A simple colorimetric sensor for the detection of moisture in organic solvents and building materials: Applications in rewritable paper and fingerprint imaging. Analyst 2019, 144, 594–601. [Google Scholar] [CrossRef] [PubMed]
  25. Rodríguez-Ciria, M.; Sanz, A.M.; Yunta, M.J.R.; Gómez-Contreras, F.; Navarro, P.; Sánchez-Moreno, M.; Boutaleb-Charki, S.; Osuna, A.; Castiñeiras, A.; Pardo, M.; et al. 1,4-Bis (alkyla mino) benzo[g]phthalazines able to form dinuclear complexes of Cu(II) which as free ligands behave as SOD inhibitors and show efficient in vitro activity against Trypanosoma cruzi. Bioorg. Med. Chem. 2007, 15, 2081–2091. [Google Scholar]
  26. Battaini, G.; Monzani, E.; Casella, L.; Santagostini, L.; Pagliarin, R. Inhibition of the catecholase activity of biomimetic dinuclear copper complexes by kojic acid. J. Biol. Inorg. Chem. 2000, 5, 262–268. [Google Scholar] [CrossRef]
  27. Naeimi, H.; Brojerdi, S.S. Facile and Efficient One-Pot Synthesis of Anthraquinones from Benzene Derivatives Catalyzed by Silica Sulfuric Acid. Polycycl. Aromat. Compd. 2014, 34, 504–517. [Google Scholar] [CrossRef]
  28. Xia, L.L.; Idhayadhulla, A.; Lee, Y.R.; Wee, Y.J.; Kim, S.H. Anti-tyrosinase, antioxidant, and antibacterial activities of novel 5-hydroxy-4-acetyl-2,3-dihydronaphtho[1,2-b]furans. Eur. J. Med. Chem. 2014, 86, 605–612. [Google Scholar]
  29. Friesner, R.A.; Murphy, R.B.; Repasky, M.P.; Frye, L.L.; Greenwood, J.R.; Halgren, T.A.; Sanschagrin, P.C.; Mainz, D.T. Extra Precision Glide:  Docking and Scoring Incorporating a Model of Hydrophobic Enclosure for Protein−Ligand Complexes. J. Med. Chem. 2006, 49, 6177–6196. [Google Scholar] [CrossRef] [Green Version]
  30. Farid, R.; Day, T.; Friesner, R.A.; Pearlstein, R.A. New insights about HERG blockade obtained from protein modeling, potential energy mapping, and docking studies. Bioorg. Med. Chem. 2006, 14, 3160–3173. [Google Scholar] [CrossRef]
  31. Saeed, A.; Mahesar, P.A.; Channar, P.A.; Abbas, Q.; Larik, F.A.; Hassan, M.; Raza, H.; Seo, S.Y. Synthesis, molecular docking studies of coumarinyl-pyrazolinyl substituted thiazoles as non-competitive inhibitors of mushroom tyrosinase. Bioorg. Chem. 2017, 74, 187–196. [Google Scholar]
  32. Larik, F.A.; Saeed, A.; Channar, P.A.; Muqadar, U.; Abbas, Q.; Hassan, M.; Seo, S.Y.; Bolte, M. Design, synthesis, kinetic mechanism and molecular docking studies of novel 1-pentanoyl-3-arylthioureas as inhibitors of mushroom tyrosinase and free radical scavengers. Eur. J. Med. Chem. 2017, 141, 273–281. [Google Scholar] [CrossRef] [PubMed]
  33. Hassan, M.; Ashraf, Z.; Abbas, Q.; Raza, H.; Seo, S.Y. Exploration of Novel Human Tyrosinase Inhibitors by Molecular Modeling, Docking and Simulation Studies. Interdiscip. Sci. 2018, 10, 68–80. [Google Scholar] [CrossRef] [PubMed]
  34. Kumar, R.S.; Moydeen, M.; Al-Deyab, S.S.; Aseer, M.; Idhayadhulla, A. Synthesis of new morpholine-connected pyrazolidine derivatives and their antimicrobial, antioxidant, and cytotoxic activities. Bioorg. Med. Chem. Lett. 2017, 27, 66–71. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Target molecules of tyrosinase enzyme inhibitors.
Figure 1. Target molecules of tyrosinase enzyme inhibitors.
Catalysts 10 01442 g001
Scheme 1. Synthesis of Tel-Cu-NPs (telmisartan-copper nanoparticles).
Scheme 1. Synthesis of Tel-Cu-NPs (telmisartan-copper nanoparticles).
Catalysts 10 01442 sch001
Figure 2. XRD Studies of telmisartan mediated copper nanoparticle.
Figure 2. XRD Studies of telmisartan mediated copper nanoparticle.
Catalysts 10 01442 g002
Figure 3. Telmisartan Cu-NPs of SEM images.
Figure 3. Telmisartan Cu-NPs of SEM images.
Catalysts 10 01442 g003
Figure 4. Telmisartan-copper nanoparticle of EDX.
Figure 4. Telmisartan-copper nanoparticle of EDX.
Catalysts 10 01442 g004
Scheme 2. Synthesis of target compounds (1a1k).
Scheme 2. Synthesis of target compounds (1a1k).
Catalysts 10 01442 sch002
Scheme 3. Mechanism of reaction compound (1a1k).
Scheme 3. Mechanism of reaction compound (1a1k).
Catalysts 10 01442 sch003
Figure 5. Recyclability of telmisartan mediated copper nanoparticle.
Figure 5. Recyclability of telmisartan mediated copper nanoparticle.
Catalysts 10 01442 g005
Scheme 4. Tyrosinase with reversible competitive binding of compound 1c (binuclear active site).
Scheme 4. Tyrosinase with reversible competitive binding of compound 1c (binuclear active site).
Catalysts 10 01442 sch004
Figure 6. Inhibition of compound 1c—Lineweaver–Burk plot.
Figure 6. Inhibition of compound 1c—Lineweaver–Burk plot.
Catalysts 10 01442 g006
Figure 7. Molecular docking of compound 1c with binding site of 2Y9X protein.
Figure 7. Molecular docking of compound 1c with binding site of 2Y9X protein.
Catalysts 10 01442 g007
Figure 8. Molecular docking of kojic acid binding site of 2Y9X protein.
Figure 8. Molecular docking of kojic acid binding site of 2Y9X protein.
Catalysts 10 01442 g008
Table 1. Catalyst recyclability.
Table 1. Catalyst recyclability.
EntryCatalyst Use Yield (%)
11st 86
22nd 85
33rd 84
44th 84
55th 83
66th 81
77th 84
88th 82
99th 81
1010th81
Table 2. The compounds (1a1k)—Tyrosinase screening.
Table 2. The compounds (1a1k)—Tyrosinase screening.
CompoundConcentration (µg/mL) aIC50
µg/mL (µM)
2550100
1a11.02 ± 0.1522.81 ± 0.1946.74 ± 0.98>100
1b03.05 ± 0.2817.84 ± 0. 1236.52 ± 0.98>100
1c58.75 ± 0.3472.84 ± 0. 6588.52 ± 0.5706.26 (11.5)
1d18.75 ± 0.2237.84 ± 0.4368.52 ± 0.1871.46 (142.7)
1e13.75 ± 0.5624.84 ± 0.2355.52 ± 0.0691.51 (213.6)
1f20.75 ± 0.6653.84 ± 0.1978.52 ± 0.6356.24 (125.4)
1g14.75 ± 0.5431.84 ± 0.1362.52 ± 0.1780.68 (155.9)
1h19.75 ± 0.4723.84 ± 0.2443.52 ± 0.05>100
1i16.75 ± 0.3821.84 ± 0.2940.52 ± 0.29>100
1j19.75 ± 0.2334.84 ± 0.3454.52 ± 0.8888.59 (165.7)
1k21.75 ± 0.1741.84 ± 0.1067.52 ± 0.1569.18 (136.0)
Kojic acid55.61 ± 0.8768.10 ± 0.1184.12 ± 0.1811.09 (78.0)
a Data represent the mean ± standard error of the mean values of three separate experiments.
Table 3. Molecular docking interactions of compound 1c and control kojic acid against mushroom Table 2. Y9X.
Table 3. Molecular docking interactions of compound 1c and control kojic acid against mushroom Table 2. Y9X.
LigandDock ScoreH. BondGlide E. ModelNo of InteractionsInteracting ResiduesBond Length
1c−5.6−2.4−36.8653Arg268, Ser282, and Val2835.85, 4.25, and 3.62
Kojic acid−5.2−2.2−17.0301His263 1.79
Table 4. Cytotoxicity activity of compounds (2a2k).
Table 4. Cytotoxicity activity of compounds (2a2k).
Comp.
No.
HepG2MCF-7HeLa
GI50 (µM)TGI (µM)LC50 (µM)GI50 (µM)TGI (µM)LC50 (µM)GI50 (µM)TGI (µM)LC50 (µM)
1a 3.60 ± 0.18 8.90 ± 0.7415.40 ± 0.12 0.71 ± 0.0814.60 ± 0.7022.60 ± 0.32 3.50 ± 0.1907.80 ± 0.1614.80 ± 0.10
1b 0.05 ± 0.02 0.19 ± 0.05 0.93 ± 0.02 0.15 ± 0.02 0.38 ± 0.340.89 ± 0.08 0.21 ± 0.17 0.59 ± 0.1001.58 ± 0.18
1c 0.01 ± 0.02 0.15 ± 0.31 0.61 ± 0.11 0.03 ± 0.00 0.29 ± 0.410.98 ± 0.09 0.04 ± 0.06 0.58 ± 0.19 0.98 ± 0.09
1d 0.16 ± 0.07 0.49 ± 0.28 1.20 ± 0.01 9.80 ± 0.9518.20 ± 0.7436.20 ± 0.63 4.30 ± 0.2009.40 ± 0.1215.10 ± 0.10
1e 4.70 ± 0.1010.30 ± 0.2428.20 ± 0.40 10.0 ± 0.3420.10 ± 0.6442.00 ± 0.4110.10 ± 0.2421.30 ± 0.1436.70 ± 1.23
1f 6.20 ± 0.9413.20 ± 0.1022.30 ± 0.52 09.5 ± 0.1515.60 ± 0.2432.90 ± 0.35 5.70 ± 0.0319.20 ± 0.1926.50 ± 0.95
1g 0.13 ± 0.05 0.28 ± 0.12 0.56 ± 0.02 0.48 ± 0.08 0.95 ± 0.041.59 ± 0.74 0.06 ± 0.01 0.16 ± 0.06 0.30 ± 0.05
1h 6.10 ± 0.1212.80 ± 0.1326.90 ± 0.2211.60 ± 0.9126.70 ± 0.2849.50 ± 0.7413.30 ± 0.1429.40 ± 0.1852.00 ± 0.04
1i12.30 ± 0.1826.80 ± 0.2545.30 ± 0.19 1.28 ± 0.10 2.91 ± 0.326.49 ± 0.74 4.10 ± 0.1809.20 ± 0.0615.80 ± 0.62
1j 6.30 ± 0.4213.10 ± 0.6526.90 ± 0.72 2.80 ± 0.65 4.90 ± 0.341.20 ± 0.08 6.10 ± 0.0713.90 ± 0.1030.10 ± 0.18
1k15.50 ± 0.0230.10 ± 0.3187.20 ± 0.1117.10 ± 0.7426.10 ± 0.4149.30 ± 0.0919.60 ± 0.0631.90 ± 0.1987.60 ± 0.19
Std. 0.01 ± 0.00 0.13 ± 0.01 0.58 ± 0.02 0.02 ± 0.000.21 ± 0.060.74 ± 0.09 0.05 ± 0.01 0.41 ± 0.10 0.88 ± 0.05
Value expressed are means ± SD error of the mean values of three separate experiments. Standard: Doxorubicin.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Selvaraj, K.; Daoud, A.; Alarifi, S.; Idhayadhulla, A. Tel-Cu-NPs Catalyst: Synthesis of Naphtho[2,3-g]phthalazine Derivatives as Potential Inhibiters of Tyrosinase Enzymes and Their Investigation in Kinetic, Molecular Docking, and Cytotoxicity Studies. Catalysts 2020, 10, 1442. https://doi.org/10.3390/catal10121442

AMA Style

Selvaraj K, Daoud A, Alarifi S, Idhayadhulla A. Tel-Cu-NPs Catalyst: Synthesis of Naphtho[2,3-g]phthalazine Derivatives as Potential Inhibiters of Tyrosinase Enzymes and Their Investigation in Kinetic, Molecular Docking, and Cytotoxicity Studies. Catalysts. 2020; 10(12):1442. https://doi.org/10.3390/catal10121442

Chicago/Turabian Style

Selvaraj, Keerthana, Ali Daoud, Saud Alarifi, and Akbar Idhayadhulla. 2020. "Tel-Cu-NPs Catalyst: Synthesis of Naphtho[2,3-g]phthalazine Derivatives as Potential Inhibiters of Tyrosinase Enzymes and Their Investigation in Kinetic, Molecular Docking, and Cytotoxicity Studies" Catalysts 10, no. 12: 1442. https://doi.org/10.3390/catal10121442

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