Next Article in Journal
Hydrophilic Dogwood Extracts as Materials for Reducing the Skin Irritation Potential of Body Wash Cosmetics
Previous Article in Journal
An Efficient Synthesis of Novel Bioactive Thiazolyl-Phthalazinediones under Ultrasound Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Neoflavonoids as Inhibitors of HIV-1 Replication by Targeting the Tat and NF-κB Pathways

1
Pharmaceutical Chemistry Area, Department of Pharmaceutical Sciences, University of Salamanca, Faculty of Pharmacy, CIETUS, IBSAL, Campus Miguel de Unamuno, 37007 Salamanca, Spain
2
Department of Cellular Biology, Physiology and Immunology, University of Córdoba, Faculty of Medicine Avda de Menendez Pidal s/n, 14004 Córdoba, Spain
3
CIFLORPAN, Center for Pharmacognostic Research on Panamanian Flora, College of Pharmacy, University of Panama, P.O. Box 0824-00172 Panama, Panama
4
National Centre of Microbiology, Institute Carlos III, Crt. Majadahonda a Pozuelo, 28220 Majadahonda, Madrid, Spain
5
Pharmacology Department, College of Pharmacy, Complutense University. Pz. Ramón Y Cajal s/n, 28040 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Molecules 2017, 22(2), 321; https://doi.org/10.3390/molecules22020321
Submission received: 12 January 2017 / Revised: 19 January 2017 / Accepted: 16 February 2017 / Published: 19 February 2017
(This article belongs to the Section Bioorganic Chemistry)

Abstract

:
Twenty-eight neoflavonoids have been prepared and evaluated in vitro against HIV-1. Antiviral activity was assessed on MT-2 cells infected with viral clones carrying the luciferase reporter gene. Inhibition of HIV transcription and Tat function were tested on cells stably transfected with the HIV-LTR and Tat protein. Seven 4-phenylchromen-2-one derivatives showed HIV transcriptional inhibitory activity but only the phenylchrome-2-one 10 inhibited NF-κB and displayed anti-Tat activity simultaneously. Compounds 10, 14, and 25, inhibited HIV replication in both targets at concentrations <25 μM. The assays of these synthetic 4-phenylchromen-2-ones may aid in the investigation of some aspects of the anti-HIV activity of such compounds and could serve as a scaffold for designing better anti-HIV compounds, which may lead to a potential anti-HIV therapeutic drug.

1. Introduction

UNAIDS report in 2016 indicated that tuberculosis remains the leading cause of death among people living with HIV, accounting for around one in three AIDS-related deaths. In 2014, the percentage of identified HIV-positive tuberculosis patients who started or continued on ART reached 77% [1]. This relevant fact should promote further studies to take advantage of this therapeutic ambivalence and to evaluate the possibility of using 4-phenylchromene-2-ones for treating patients suffering from both AIDS and tuberculosis.
Human immunodeficiency virus (HIV) is the cause of acquired immunodeficiency syndrome (AIDS) which is one of the leading causes of 2.9% of mortality in the world [2]. Although modern antiretroviral therapy (ART) using a combination of anti-HIV drugs has been highly effective in suppressing HIV load and decreasing mortality in AIDS patients, the emergence of drug resistances in HIV and the toxicity of the therapies currently in use have made the continued search for novel anti-HIV drugs necessary [3,4]. On the other hand, failures in efforts to develop an effective vaccine against HIV-1 infection [5] have emphasized the importance of ART in treating HIV-1-infected patients. Therefore, medicinal chemists are interested in the development of novel anti-HIV agents that might be particularly effective in controlling strains of HIV that are resistant to the current drugs [6].
The HIV viral cycle can be divided into early and late stages. Early stages comprise several steps, from viral attachment on the cell surface to integration in the host genome. Late stages include the processes of HIV mRNA synthesis, protein expression and morphogenesis. Once integrated, HIV can remain in a latent state in resting lymphocytes or undergo active replication. Transition from latency to HIV expression occurs mainly when cells are activated and requires the concerted action of cellular transcription factors and regulatory HIV proteins [7,8]. Among the transcription factors involved in LTR transactivation, the HIV proximal enhancer contains three binding sites for SP1 transcription factor and two binding sites for NF-κB. The NF-κB/Rel family of transcription factors represents a major inducible regulatory element involved in HIV transcription [9]. Located downstream of the basal promoter TAR sequence is the RNA target for the viral protein Tat, which acts in concert with other cellular factors [10], to generate full-length RNA transcripts [11]. Furthermore, NF-κB and Tat cooperate in driving HIV replication from the state of latency. Therefore, inhibition of the activity of these critical proteins should result in an effective blocking of viral replication [12,13,14].
The neoflavonoids with anti-HIV activity possessing 4-phenylcoumarin skeleton have been obtained mainly from Calophyllaceae family that includes the genera: Calophyllum [15,16,17,18,19,20,21]; Mammea [21,22,23,24,25,26]; Mesua [27,28,29,30,31,32,33]; Kielmeyera [34,35,36,37,38]; and Marila [39], from which several 4-phenyl chromen-2-one derivatives have been isolated.
The discovery, structural modification and structure-activity relationships studies of natural neoflavonoids with anti-HIV activity:(+)-Inophyllum B [40], (+)-Inophyllum C [19], and Inophyllum P [40], and a number of their synthetic derivatives have been successfully obtained in this work. These 4-phenylcoumarins have been proposed as suppressors of LTR-dependent transcription, but the mechanism of action has not been fully characterized [41]. In addition, isomesuol and mesuol inhibit TNF-α-induced HIV-1-LTR transcriptional activity by targeting the nuclear factor-κB (NF-κB) pathway. Mesuol inhibited the phosphorylation and the transcriptional activity of the NF-κB p65 subunit in TNFα-stimulated cells [42]. Isodispar B and Disparinol A are HIV transcription inhibitors, which inhibit both, NF-κB and Tat targets, affecting the HIV replication by synergistic effect [43]. Synthetic 4-phenylchromen-2-ones have been also reported to show antimicrobial [30,44,45], anti-mycobacterial [46] and anti-inflammatory activities [47].
In a previous paper we reported the anti-HIV activity of natural 4-phenylcoumarins isolated from Marila pluricostata. They were structurally related to Inophyllum coumarins series, but with one prenyl and other cyclized group across the hydroxyl group at position C-7 [43]. Furthermore, these compounds showed moderate anti-mycobacterial activity. This relevant fact induced us to prepare new similar, but simpler, derivatives with the idea in mind to obtain compounds with activity against both HIV and tuberculosis, and also to increase the structural diversity. With that, a better structure-activity relationship could be established. In this paper we reported the preparation and the anti-HIV activity of several neoflavone derivatives that showed anti-mycobacterial activity [48].

2. Results and Discussion

2.1. Chemistry

The 4-phenylchromen-2-ones or neoflavones 17 have been obtained by the Peckman condensation between ethyl benzoylacetate and different phenol derivatives that is, resorcinol for 1, phloroglucinol for 2, hydroxyhydroquinone for 3, 3,4-methylenedioxyphenol for 4, 3-methoxycatechol for 5, 2,3-dimethoxyphenol for 6 and β-naphthol for 7 in presence of concentred H2SO4 as condensing agent (Scheme 1). Some differences in the yield of the different compounds can be appreciated depending on the phenol derivative used (see the Experimental Section).
Phloroglucinol, as the starting material (Scheme 1), was treated with ethyl benzoylacetate by the Pechmann-Duisberg reaction giving 5,7-dihydroxy-4-phenylcoumarin (2). Friedel-Craft acylation or benzoylation of compound 2, by refluxing the phenol derivative with the corresponding acylchloride in a carbon disulfide/nitrobenzene mixture, and in the presence of aluminum trichloride, followed by Fries rearrangement provided mixtures of the 6-, 8-monoacylated and benzoylated neoflavones and 6,8-diacylated or dibenzoylated neoflavones [49,50,51,52]. Workup of the crude reaction products led to the isolation of neoflavones 821 whose spectroscopic properties were consistent with the structures shown in the Scheme 2. On the basis of detailed analysis of the H-H COSY (Correlation Spectroscopy), H-C HMQC (Heteronuclear Single Quantum Correlation), and H-C HMBC (Heteronuclear Multiple-Bond Correlation) 2D-NMR spectra, all of the compounds described were correctly characterized and their 13C-NMR data will be introduced in NAPROC-13 RMN spectroscopic database [53], for posterior online identification of natural products and their analogs and derivatives.
Neoflavones 2228 were obtained by the condensation of substituted cinnamic acids with the corresponding phenols, that is, phenol for 22, resorcinol for 23 and 24, 3,5-dichlorophenol for 25, 3,4-methylenedioxyphenol for 26 and 27, and 3-methoxybenzene-1,2-diol for 28. The reaction occurred in the presence of milder Friedel-Crafts catalyst BF3–Et2O and POCl3 in 54%–75% yields (Scheme 3). Previous attempts towards condensation of substituted cinnamic acids with corresponding phenols for the preparation of 4-phenylchromen-2-ones 2228 in the presence of concentrated HCl and HCl gas were unsuccessful. Additionally, some of these compounds were synthesized in low yields by microwave irradiation.
Some compounds prepared in this study have been described already in the literature (see the Experimental Section) and their structures were verified from an iterative search by 13C-NMR chemical shifts carried out within our NAPROC-13 RMN spectroscopic database [53], and later identified unambiguously. The rest of the prepared 4-phenylcoumarins were established on the basis of 1H- and 13C-NMR spectra. A combination of COSY, HMQC, and NOE experiments was utilized when necessary for a correct assignment of 1H and 13C chemical shifts.
The synthesized neoflavonoids 18 were evaluated in vitro against HIV-1 with the results shown in Table 1.

2.2. Evaluation of Antiviral Activity

The neoflavonoids synthesized (128) belong to three groups: simple 4-phenylchromen-2-ones, acyl-4-phenylchromen-2-ones and 3,4-dihydro-4-phenylchromen-2-ones, which have been evaluated in the anti HIV bioassay. The analysis of the results of inhibitory activity of NF-κB indicates that 4-phenylchromenones 9, 10, 13, 14, and 15 show a fair inhibitory activity at 50 μM. Regarding the specific HeLa-Tat-Luc assay results, compounds 9, 13, and 15 were nonspecific, whereas compounds 10 and 14 showed specificity. Furthermore compound 14 is only slightly toxic at 50 μM, but not toxic at 25 μM, and their activity as NF-κB and tat inhibitors is still strong (83.06% for NF-κB and 41.87% for Tat). Compound 14 turned out to be identical to the natural neoflavonoid Isodispar B previously isolated from Marila pluricostata [43]. Compound 10 NF-κB activity is also strong (70.53 at 25 μM) and it is nontoxic at 10 μM. Interestingly, the dichlorinated 3,4-dihydroflavonoid 25 showed specific anti-Tat activity, whereas all other 3,4-dihydroanalogs resulted inactive in this assay. It must be noted that simple structural differences within this series of 4-phenylchromen-2-ones and acyl 4-phenylchromen-2-ones, determine substantial changes in activity and selectivity. As an example, we could compare the specific NF-κB inhibitor 8-isovaleroyl-4-phenylchromen-2-one 14 and the specific Tat inhibitor 3,4-dihydro-5,7-dichloro-4-phenyl-chroman-2-one 25, also a 4-phenylchromen related. The drastic difference in bioactivity for these compounds should be due either to the presence of an isovaleroyl or heptanoyl group in C-6 or C-8, which potentiates the activity in both targets. However, regarding the comparison of compounds 7 and 14, it is worth noting that the presence of a benzene ring fused to the chromenone moiety, increases the anti-Tat activity, but the introduction 8-isovaleroyl or 6,8-diacetyl groups enhances the NF-κB inhibitory activity. Compound 10, when compared to Disparinol A, showed higher anti-HIV potency than this natural neoflavonoid. The acyl derivatives of 4-phenylchromen-2-ones are the most potent dual-target inhibitors.
The assays of these 4-phenyl-chromen-2-one derivatives may aid in the investigation of some aspects of the anti-HIV activity of this kind of compound that inhibited the transcription and could serve as a scaffold for designing better anti-HIV compounds, which may lead to a potential HIV therapeutic drug.

3. Experimental Section

3.1. General Information

All of the reagents for synthesis were commercially available and either used without further purification or purified by standard methods prior to use. Melting points were determined on a Büchi 510-K melting point apparatus (Büchi Labortechnik AG, Flawil, Switzerland) and are uncorrected. IR spectra were recorded (KBr 1%) in a Nicolet Impact 410 spectrophotometer. 1H-, 13C-NMR, COSY, HMQC, and HMBC were recorded on Brüker AC 200 (200 MHz) and Brüker DRX 400 (400 MHz) instruments. Chemical shifts (δ) are expressed in parts per million (ppm) relative to the residual solvent peak, and coupling constants are reported in Hertz (Hz). All signals assigned to hydroxyl groups were exchangeable with D2O. Reaction progress was monitored using analytical thin-layer chromatography (TLC) on precoated Merck silica gel Kiesegel 60 F254 plates, and the spots were detected under UV light (254 nm). The flash chromatography was conducted using silica gel 230–400 mesh. For EIMS and HRFABMS analysis, a VG-TS250 mass spectrometer (70 eV) was used. Elementary analyses were obtained with a LECO CHNS-932 and were within ±0.4% of the theoretical values.

3.2. General Procedures I for the Synthesis of Compounds 17

To a mixture of appropriate phenol (2 mmol) and ethyl benzoyl acetate (2 mmol), concentrated H2SO4, (1 mL) was added and stirred at room temperature for four days; after which the mixture was poured over crushed ice and extracted with AcOEt, (50 mL × 5). Evaporation gave a brown solid which, after chromatography (silica gel, hexane/AcOEt 10:1→1:1), afforded the corresponding 4-phenylcoumarin (17) as a white solid (yield, 25%–30%).
7-Hydroxy-4-phenyl-2H-chromen-2-one (1). Yield 85%; A white solid; m.p. 232–234 °C (MeOH). The spectral data (1H-NMR) were quite comparable with the data reported in [54].
5,7-Dihydroxy-4-phenyl-2H-chromen-2-one (2). Yield 70%; A white solid; m.p. 227–229 °C (MeOH). The spectral data (1H-NMR) were quite comparable with the data reported in [52].
6,7-Dihydroxy-4-phenyl-2H-chromen-2-one (3). Yield 68%; A white solid; m.p. 230–232 °C (CHCl3/MeOH); IR (KBr): ν = 3437, 3414, 1686, 1617, 1562 cm−1; 1H-NMR (MeOD) δ 7.52 (m, 2H), 7.52 (m, 3H), 6.86 (s, 1H), 6.83 (s, 1H), 6.13 (s, 1H); 13C-NMR (MeOD) δ 104.0, 111.3, 111.9, 112.3, 129.4, 129.4, 129.9, 129.9, 130.6, 137.2, 144.4, 150.5, 152.0, 158.4, 164.1. MS (EI) m/z: 254 (M+ C15H10O4, 8), 252 (52), 224 (100), 152 (80), 139 (13).
8-Phenyl-6H-[1,3]dioxolo[4,5-g]chromen-6-one (4). Prepared from benzo[d][1,3]dioxol-5-ol (2 mol) as described in the general procedure I. Yield 65%; A white solid; m.p. 190–192 °C (CDCl3/MeOH). IR (KBr): ν = 1712, 1627, 1563, 1503 cm−1; 1H-NMR (CDCl3) δ 7.50 (m, 3H), 7.41 (m, 2H), 6.86 (s, 1H), 6.82 (s, 1H), 6.22 (s, 1H), 6.04 (s, 2H); 13C-NMR (CDCl3) δ 98.5, 102.3, 104.3, 112.1, 112.8, 128.2, 128.2, 128.8, 128.8, 129.6, 135.6, 144.8, 151.1, 151.2, 155.8, 161.1. MS (EI) m/z: 266 (M+ C16H10O4, 66), 265 (88), 238 (100), 152 (22).
8-Hydroxy-7-methoxy-4-phenyl-2H-chromen-2-one (5). Prepared as described in general procedure I from 3-methoxybenzene-1,2-diol (2 mmol), yield 65%; A white solid; m.p. 220–222 °C (CDCl3/MeOH); IR (KBr): ν = 3342, 2937, 2838, 1697, 1562 cm‒1; 1H-NMR (MeOD) δ 7.47 (m, 2H), 7.47 (m, 3H), 6.95 (d, J = 9.1 Hz, 1H), 6.87 (d, J = 9.1 Hz, 1H), 6.15 (s, 1H), 3.94 (s, 3H); 13C-NMR (MeOD) δ 56.7, 108.7, 112.1, 114.3, 118.4, 129.2, 129.2, 129.5, 129.5, 130.4, 136.5, 143.8, 151.7, 158.2, 162.7. MS (EI) m/z: 268 (M+ C16H12O4, 60), 265 (88), 225 (100), 152 (19), 141 (65).
7,8-Dimethoxy-4-phenyl-2H-chromen-2-one (6). This was prepared from 2,3-dimethoxyphenol (2 mol) as described in the general procedure I, yield 70%; A white solid; m.p. 175–177 °C (CDCl3); IR (KBr): ν = 2968, 2933, 2844, 1718, 1602, 1557 cm−1; 1H-NMR (CDCl3) δ 7.51 (m, 3H), 7.43 (m, 2H), 7.18 (d, J = 9.2 Hz, 1H), 6.83 (d, J = 9.2 Hz, 1H), 6.22 (s, 1H), 4.02 (s, 3H), 3.95 (s, 3H); 13C-NMR (CDCl3) δ 56.4, 61.5, 108.1, 112.3, 113.8, 122.2, 128.4, 128.4, 128.8, 128.8, 129.6, 135.6, 136.5, 148.4, 155.5, 155.9, 160.6. MS (EI) m/z: 282 (M+ C17H14O4, 100), 267 (8), 254 (19), 239 (42), 152 (24), 139 (53).
4-Phenyl-2H-benzo[g]chromen-2-one (7). Prepared as described in general procedure II from naphthalen-2-ol (2 mmol), yield 60%; A white solid; m.p. 211–213 (CDCl3); IR (KBr): ν = 3054, 1722, 1633, 1593, 1553 cm‒1; 1H-NMR (CDCl3) δ 8.57 (m, 1H), 7.83 (m, 1H), 7.52–7,62 (m, 9H), 6.44 (s, 1H); 13C-NMR (CDCl3) δ 114.1, 114.5, 122.3, 122.7, 123.3, 123.9, 127.1, 127.6, 128.5, 128.5, 128.9, 128.9, 128.9, 129.6, 134.8, 135.6, 151.4, 156.5, 160.8. MS (EI) m/z: 272 (M+ C19H12O2, 52), 245 (18), 244 (100), 215 (64), 189 (10), 139 (7).

3.3. General Procedures II: Synthesis of Compounds 822

Anhydrous aluminum trichloride (0.4 mmol) was added to a stirred suspension of compound 2 (0.1 mmol) in carbon disulfide (6 mL). Nitrobenzene (2 mL) was then added over 40 min, forming a homogeneous solution with evolution of HCl. The solution was heated under reflux for 30 min, appropriate acyl chloride (0.1 mmol) in nitrobenzene (1 mL) was added over 40 min before allowing it to cool with stirring. The mixture was poured onto ice/water and aqueous HCl and was extracted with ethyl acetate (25 mL, twice). Workup of the crude product by chromatography on silica gel led to the isolation of the different acyl derivatives products.
Following the general procedure II using allyl bromide and benzyl bromide the crude reaction product was chromatographed and eluted with hexane/EtOAc, to yield 8 and 21.
6-Allyl-5,7-dihydroxy-4-phenyl-2H-chromen-2-one (8). Yield 41%; A white solid; m.p. 196–198 °C (CHCl3/MeOH); IR (KBr): ν = 3217, 1686, 1626, 1591 cm‒1; 1H-NMR (CDCl3) δ 7.24 (m, 2H), 7.22 (m, 3H), 6.55 (s, 1H), 6.36 (s, 1H), 5.70 (m, 1H), 4.82 (t, J = 1.8 Hz, 1H) 4.76 (dd, J = 1.2 Hz, 1H), 3.16 (d, J = 6.2 Hz, 2H); 13C-NMR (CDCl3) δ 26.7, 95.8, 110.8, 110.9, 114.5, 114.6, 127.2, 128.9, 129.4, 135.9, 137.3, 154.3, 154.4, 155.3, 160.1, 161.8. (EI) m/z: 294 (M+ C18H14O4, 88), 281 (100), 268 (24), 253(46), 139 (15).
5,7-Dibenzyloxy-4-phenyl-2H-chromen-2-one (21). Yield 45%; A white solid; m.p. 175–177 °C (CH2Cl2). IR (KBr): ν = 3089, 3059, 3031, 2927, 2865, 1720, 1611, 1597, 1432, 1337, 1159, 1111, 1064, 726 cm‒1. 1H-NMR (CDCl3) δ 7.38 (m, 3H), 7.20 (m, 5H), 7.18 (m, 10 H), 5.99 (s, 1H), 6.61 (d, J = 2.4 Hz, 1H), 6.41 (d, J = 2.4 Hz, 1H), 5.10 (s, 1 H), 4.72 (s, 1H). 13C-NMR (CDCl3) δ 70.5, 70.8, 94.37, 94.87, 103.8, 113.10, 127.03, 127.11, 127.51, 127.62, 127.73, 127.99, 128.25, 128.40, 128.80, 135.08, 135.78, 139.71, 156.74, 157.21, 157.35, 160.74, 162.32. (EI) m/z: 434 (M+ C29H22O4, 17), 343 (12), 181 (14), 139 (6), 114 (6), 92 (52), 91 (100).
Following the general procedure II using acetyl chloride, the crude reaction product was chromatographed and eluted with hexane/EtOAc, to yield 12 and 17.
8-Acetyl-5,7-dihydroxy-4-phenyl-2H-chromen-2-one (12). Yield 25%; A white solid; m.p. 206–208 °C (Hex:AcOEt); IR (KBr): ν = 3224, 3083, 1689, 1627, 1592 cm‒1; 1H-NMR (CDCl3) δ 14.01 (s, 1H), 7.57 (m, 3H), 7.42 (m, 2H), 6.25 (s, 1H), 6.02 (s, 1H), 5.85 (s, 1H), 2.93 (s, 3H); 13C-NMR (CDCl3) δ 32.9, 99.3, 101.8, 103.7, 110.9, 126.8, 127.2, 127.9, 139.0, 157.9, 158.0, 159.9, 162.0, 167.8, 202.7. (EI) m/z, 272 (M+ C19H12O2, 52), 245 (18), 244 (100), 215(64), 189 (10). (EI) m/z: 296 (M+ C17H12O5, 18), 295 (100), 277 (22), 221(16), 165 (20), 139 (43).
6,8-Diacetyl-5,7-dihydroxy-4-phenyl-2H-chromen-2-one or (1,1’-(5,7-dihydroxy-2-oxo-4-phenyl-2H-chromene-6,8-diyl-diethanone) (17). Yield 52%; A white solid; m.p. 162–164 °C (Hex/AcOEt); IR (KBr): ν = 3467, 1733, 1717, 1593 cm‒1; 1H-NMR (CDCl3) δ 16.53 (s, 1H), 15.92 (s, 1H), 7.40 (m, 3H), 7.28 (m, 2H), 6.04 (s, 1H), 2.95 (s, 3H), 2.75 (s, 3H); 13C-NMR (CDCl3) δ 33.4, 33.5, 102.8, 106.3, 112.4, 126.9, 127.8, 128.5, 138.9, 156.5, 158.0, 161.5, 170.2, 172.2, 203.9, 205.5. (EI) m/z: 338 (M+ C19H14O6, 100), 323 (94), 313 (48), 295(90), 277 (26), 139 (29).
Following the general procedure II using propionyl chloride, the crude reaction product was chromatographed and eluted with hexane/EtOAc, to yield 13 and 18.
5,7-Dihydroxy-4-phenyl-8-propionyl-2H-chromen-2-one (13). Yield 28%; A white solid; m.p. 216–218 °C (Hex:AcOEt); IR (KBr): ν = 3218, 3067, 1693, 1616, 1592 cm−1; 1H-NMR (MeOD) δ 7.37 (m, 2H), 7.37 (m, 3H), 6.18 (s, 1H), 5.99 (s, 1H), 3.36 (c, J = 7.3 Hz, 2H), 1.26 (t, J = 7.3 Hz, 3H); 13C-NMR (MeOD) δ 8.6, 38.5, 100.2, 111.5, 127.6, 127.9, 128.6, 140.1, 158.3, 160.9, 162.6, 168.7, 206.7. (EI) m/z: 310 (M+ C18H14O5, 36), 282 (18), 281 (100), 252 (8), 171 (7), 139 (8).
5,7-Dihydroxy-4-phenyl-6,8-dipropionyl-2H-chromen-2-one (18). Yield 49%; A white solid; m.p. 152–154 °C (Hex:AcOEt); IR (KBr): ν = 3468, 3437, 3067, 2978, 2937, 2876 cm‒1; 1H-NMR (CDCl3) δ 16.64 (s, 1H), 15.97 (s, 1H), 7.40 (m, 3H), 7.28 (m, 2H), 6.03 (s, 1H), 3.41 (q, J = 7.3 Hz, 2H), 3.19 (q, J = 6.7 Hz, 2H), 1.28 (t, J = 7.3 Hz, 3H), 1.15 (t, J = 6.7 Hz, 3H); 13C-NMR (CDCl3) δ 8.2, 8.5, 38.2, 38.2, 101.4, 102.4, 106.1, 112.2, 126.9, 127.8, 128.5, 139.0, 156.7, 161.2, 169.9, 170.0, 172.0, 207.3, 208.7. (EI) m/z: 366 (M+ C21H18O6, 36), 338 (44), 337 (100), 319 (27), 308 (28), 281 (40), 178 (9), 138 (14).
Following the general procedure II using 3-methylbutanoyl chloride, the crude reaction product was chromatographed and eluted with hexane/EtOAc, to yield 9, 14, and 19.
5,7-Dihydroxy-6-(3-methylbutanoyl)-4-phenyl-2H-chromen-2-one (9). Yield 12%; A white solid; m.p. 206–208 °C (Hex:AcOEt); IR (KBr): ν = 3139, 3111, 2956, 2928, 2870, 1689, 1617, 1580 cm‒1; 1H-NMR (CDCl3) δ 15.79 (s, 1H), 11.39 (s, 1H), 7.52 (m, 3H), 7.39 (m, 2H), 6.63 (s, 1H), 5.98 (s, 1H), 2.90 (d, J = 6.3 Hz, 2H), 2.23 (m, 1H), 0.92 (d, J = 6.3 Hz, 3H), 0.92 (d, J = 6.3 Hz, 3H); 13C-NMR (CDCl3) δ 22.5, 22.5, 25.0, 53.0, 94.8, 101.4, 106.9, 111.3, 127.1, 127.4, 128.1, 139.0, 157.2, 159.6, 161.1, 163.7, 164.9, 207.3. (EI) m/z: 338 (M+ C20H18O5, 10), 282 (18), 281(100), 253 (3), 225(1), 171 (6), 139 (4).
5,7-Dihydroxy-8-(3-methylbutanoyl)-4-phenyl-2H-chromen-2-one (14). Yield 13%; A white solid; m.p. 210–212 °C (Hex:AcOEt); IR (KBr): ν = 3293, 2959, 2934, 2874, 2454, 1745, 1685, 1620, 1592 cm‒1; 1H-NMR (CDCl3) δ 14.14 (s, 1H), 7.55 (m, 3H), 7.43 (m, 2H), 6.25 (s, 1H), 6.01 (s, 1H), 6.00 (s, 1H), 3.17 (d, J = 6.3 Hz, 2H), 2.28 (m, 1H), 1.05 (d, J = 6.3 Hz, 3H), 1.05 (d, J = 6.3 Hz, 3H); 13C-NMR (CDCl3) δ 22.7, 22.7, 25.5, 53.6, 101.6, 101.8, 104.9, 112.0, 127.5, 129.6, 130.3, 136.5, 154.1, 155.7, 158.7, 159.7, 168.8, 205.9. (EI) m/z: 338 (M+ C20H18O5, 20), 323 (12), 281 (100), 254 (9), 171(6), 141 (6).
5,7-Dihydroxy-6,8-bis(3-methylbutanoyl)-4-phenyl-2H-chromen-2-one or 1,1′-(5,7-dihydroxy-2-oxo-4-phenyl-2H-chromene-6,8-diyl)bis(3-methylbutan-1-one) (19). Yield 50%; A white solid; m.p. 148–150 °C (Hex:AcOEt); IR (KBr): ν = 3469, 3435, 2959, 2931, 2871, 1756, 1618, 1597 cm‒1; 1H-NMR (CDCl3) δ 16.77 (s, 1H), 16.09 (s, 1H), 7.40 (m, 3H), 7.28 (m, 2H), 6.02 (s, 1H), 3.18 (d, J = 6.7 Hz, 2H), 3.02 (d, J = 6.6 Hz, 2H), 2.27 (m, 1H), 2.27 (m, 1H), 1.06 (d, J = 6.7 Hz, 3H), 1.06 (d, J = 6.7 Hz, 3H), 0.95 (d, J = 6.6 Hz, 3H), 0.95 (d, J = 6.6 Hz, 3H); 13C-NMR (CDCl3) δ 22.7, 22.7, 22.7, 22.7, 24.9, 25.8, 53.3, 53.3, 101.5, 102.6, 106.4, 112.2, 126.9, 127.8, 128.4, 139.1, 156.6, 158.0, 161.8, 170.3, 172.4, 206.6, 207.9. (EI) m/z: 422 (M+ C25H26O6, 34), 418 (25), 394 (36), 381 (43), 365 (100), 347 (36), 337(32), 281 (49), 171 (25), 139 (36).
Following the general procedure II using heptanoyl chloride, the crude reaction product was chromatographed and eluted with hexane/EtOAc, to yield 10, 15, and 20.
6-Heptanoyl-5,7-dihydroxy-4-phenyl-2H-chromen-2-one (10). Yield 12%; A white solid; m.p. 210–212 °C (Hex:AcOEt); IR (KBr): ν = 3596, 3254, 3058, 2954, 2928, 2856, 1774, 1688, 1717, 1593 cm‒1; 1H-NMR (CDCl3) δ 14.03 (s, 1H), 10.11 (s, 1H), 7.44 (m, 3H), 7.35 (m, 2H), 6.99 (s, 1H), 5.99 (s, 1H), 3.10 (t, J = 6.7 Hz, 2H), 1.64 (m, 4H), 1.24 (m, 4H), 0.87 (t, J = 6.7 Hz, 3H); 13C-NMR (CDCl3) δ 14.1, 22.6, 24.4, 29.0, 31.7, 44.7, 95.9, 101.9, 107.1, 111.6, 127.4, 128.4, 128.9, 138.3, 157.7, 159.3, 161.8, 164.0, 164.8, 207.8. (EI) m/z: 366 (M+ C22H22O5, 32), 351 (14), 310 (30), (309 (100), 281 (82), 253(11), 171 (18), 139 (20).
8-Heptanoyl-5,7-dihydroxy-4-phenyl-2H-chromen-2-one (15). Yield 27%; A white solid; m.p. 212–214 °C (Hex:AcOEt); IR (KBr): ν = 3246, 2951, 2928, 2868, 2844 cm−1, 1685, 1637, 1591; 1H-NMR (CDCl3) δ 14.11 (s, 1H), 7.54 (m, 3H), 7.43 (m, 2H), 6.25 (s, 1H), 6.11 (s, 1H), 6.00 (s, 1H), 3.30 (t, J = 6.7 Hz, 2H), 1.73 (m, 4H), 1.33 (m, 4H), 0.89 (t, J = 6.7 Hz, 3H); 13C-NMR (CDCl3) δ 14.1, 22.6, 24.6, 29.0, 31.8, 44.9, 101.5, 105.0, 111.9, 127.5, 129.7, 130.2, 136.4, 154.0, 158.3, 158.6, 159.6, 168.7, 206.2. (EI) m/z: 366 (M+ C22H22O5, 44), 351 (31), 309 (91), 281 (100), 267 (3), 253 (18), 171 (23), 139 (39).
6,8-Diheptanoyl-5,7-dihydroxy-4-phenyl-2H-chromen-2-one or 1,1’-(5,7-dihydroxy-2-oxo-4-phenyl-2H-chromene-6,8-diyl)diheptan-1-one (20). Yield 50%; A white solid; m.p. 164–166 °C (Hex:AcOEt); IR (KBr): ν = 3458, 34240, 2958, 2930, 2855, 1744, 1620, 1582 cm‒1; 1H-NMR (CDCl3) δ 16.69 (s, 1H), 16.02 (s, 1H), 7.40 (m, 3H), 7.28 (m, 2H), 6.02 (s, 1H), 3.33 (t, J = 7.3 Hz, 2H), 3.14 (t, J = 7.3 Hz, 2H), 1.76 (m, 4H), 1.64 (m, 4H), 1.32 (m, 4H), 1.32 (m, 4H), 0.90 (m, 3H), 0.90 (m, 3H); 13C-NMR (CDCl3) δ 14.0, 14.0, 22.6, 22.6, 24.3, 24.8, 29.0, 29.0, 31.6, 31.8, 44.8, 44.8, 101.4, 102.5, 106.2, 112.3, 127.0, 127.8, 128.5, 139.1, 156.6, 158.0, 161.2, 170.1, 172.2, 207.0, 208.4. (EI) m/z: 478 (M+ C29H34O6, 24), 421 (34), 393 (100), 323 (38), 293(52), 171 (20), 139 (14).
Following the general procedure II using benzoyl chloride, the crude reaction product was chromatographed and eluted with hexane/EtOAc, to 11 and 16.
6-Benzoyl-5,7-dihydroxy-4-phenyl-2H-chromen-2-one (11). Yield 18%; A white solid; m.p. 245–247 °C (AcOEt). The spectral data (IR, 1H-NMR and 13C-NMR) were quite comparable with the data reported in [46,48].
8-Benzoyl-5,7-dihydroxy-4-phenyl-2H-chromen-2-one (16). Yield 34%; A white solid; m.p. 253–256 °C (AcOEt). The spectral data (IR, 1H-NMR and 13C-NMR) were quite comparable with the data reported in [50,51].

3.4. General Procedures III: Synthesis of Compounds 2228

To a mixture of POCl3 (10 mmol) and BF3–Et2O (20 mmol) at 0 °C, appropriated cinnamic acid (5 mmol) was added and the reaction mixture stirred for 15 min at 0 °C. Phenol (5 mmol) was added to the above reaction mixture in small portions and stirring continued at 25–28 °C for 4–12 h. The reaction mixture was poured on to ice-water; sodium acetate (1 g) was added and the mixture was warmed on a water bath for 2 min. It was cooled, extracted with ethyl acetate (2 × 150 mL), washed with water (150 mL), dried, and the solvent removed under reduced pressure to obtain the crude product, which was purified by column chromatography using acetone–chloroform as eluent to afford pure 4-phenyldihydro-coumarins 2228 in 60%–75% yields.
4-(4-Methoxyphenyl)chroman-2-one (22). This was prepared from p-methoxycinnamic acid and phenol using the general procedure III. Yield 62%; A white solid; m.p. 160–162 °C (CH2Cl2/MeOH); IR (KBr): ν = 2943, 2927, 2907, 2833, 1832, 1717, 1701, 1608 cm‒1; 1H-NMR (CDCl3) δ 7.31 (br d, J = 8.1 Hz, 1H), 7.12 (d, J = 8.2 Hz, 2H), 7.07 (br d, J = 8.2 Hz, 1H), 6.97 (m, 2H), 6.87 (d, J = 8.2 Hz, 3H), 4.30 (t, J = 7.3 Hz, 1H), 3.79 (s, 3H), 3.08 (dd, J = 7.3 Hz, 15.7 Hz, 1H), 2.96 (dd, J = 7.3, 15.7 Hz, 2H); 13C-NMR (CDCl3) δ 37.2, 40.0, 55.3, 114.5 (2C), 117.1, 124.7, 126.3, 128.3, 128.6 (3C), 132.2, 151.7, 159.0, 167.8. (EI) m/z: 254 (M+ C16H14O3, 65), 226 (14), 212 (15), 211 (72), 197 (22), 182 (15), 181 (100), 168 (13), 139 (12).
4-(3,4-Dimethoxyphenyl)-7-hydroxychroman-2-one (23). This was prepared from 3,4-dimethoxycinnamic acid and resorcinol as described in the general procedure III. Yield 67%; A white solid; m.p. 167–169 °C (CHCl3/MeOH); IR (KBr): ν = 3434, 2960, 2936, 1762, 1624, 1595 cm‒1; 1H-NMR (CDCl3) δ 6.83 (d, J = 7.7 Hz, 1H), 6.68 (dd, J = 7.7, 2.4 Hz, 1H), 6.68 (d, J = 2.4 Hz, 1H), 6.67 (d, J = 7.7 Hz, 1H), 6.66 (s, 1H), 6.59 (dd, J = 8.2, 2.4 Hz, 1H), 6.46 (br s, 1H), 4.21 (t, J =7.3 Hz, 1H), 3.85 (s, 3H), 3.81 (s, 3H), 3.06 (dd, J = 7.3 Hz, 14.9 Hz, 1H), 2.95 (dd, J = 7.3, 14.9 Hz, 1H); 13C-NMR (CDCl3) δ 37.5, 39.6, 55.9, 55.9, 104.2, 110.6, 111.6, 112.1, 117.5, 119.8, 129.1, 133.2, 148.3, 149.3, 152.1, 156.5, 168.7. (EI) m/z: 300 (M+ C17H16O5, 100), 269 (14), 257 (36), 243 (26), 227 (81), 190 (14), 139 (8).
4-(4-Hydroxyphenyl)-7-methoxychroman-2-one (24). This was prepared from p-coumaric acid and resorcinol as described in the general procedure III. Yield 68%; A white solid; m.p. 169–171 °C (CHCl3/MeOH); IR (KBr): ν = 3436, 2938, 2904, 2840, 1762, 1615 cm‒1; 1H-NMR (CDCl3) δ 6.98 (d, J = 8.3 Hz, 2H), 6.89 (d, J = 8.3 Hz, 1H), 6.77 (d, J = 8.3 Hz, 2H), 6.66 (d, J = 2.4 Hz, 1H), 6.64 (dd, J = 8.3, 2.4 Hz, 1H), 5.54 (br s, 1H), 4.22 (t, J = 6.8 Hz, 1H), 3.80 (s, 3H), 3.05 (dd, J = 6.8, 15.1Hz, 1H), 2.93 (dd, J = 6.8, 15.1 Hz, 1H); 13C-NMR (CDCl3) δ 37.4, 39.3, 55.6, 102.4, 110.7, 115.6, 115.6, 118.0, 128.7, 128.7, 128.8, 132.6, 152.3, 155.0, 159.9, 168.1. (EI) m/z: 270 (M+ C16H14O4, 38), 242 (17), 228 (10), 227 (100), 211 (40), 184 (15), 128 (18).
5,7-Dichloro-4-phenylchroman-2-one (25). This was prepared from cinnamic acid and 3,5-dichorophenol as described in the general procedure III. Yield 70%; A white solid; m.p. 230–232 °C (CHCl3); IR (KBr): ν = 3077, 1713, 1599, 1567 cm‒1; 1H-NMR (CDCl3) δ 7.27 (m, 3H), 7.22 (d, J = 1.9 Hz, 1H), 7.11 (d, J = 1.9 Hz, 1H), 7.06 (m, 2H), 4.64 (dd, J = 5.4, 3.4 Hz, 1H), 3.10 (dd, J = 3.4, 12.2 Hz, 1H), 3.04 (dd, J = 5.4, 12.2 Hz, 1H); 13C-NMR (CDCl3) δ 36.7, 38.5, 116.5, 122.2, 125.6, 126.7, 126.7, 127.9, 129.2, 129.2, 134.5, 138.9, 152.9, 165.7. (EI) m/z: 293 (M+ C15H10O2Cl2, 36), 291 (59), 276 (23), 274 (36), 256 (13), 251 (64), 249 (100), 215 (19), 152 (33).
8-(4-Hydroxy-3,5-dimethoxyphenyl)-7,8-dihydro-[1,3]dioxolo[4,5-g]chromen-6-one (26). This was prepared from 4-hydroxy-3,5-dimethoxycinnamic acid and sesamol as described in the general procedure III. Yield 69%; A white solid; m.p. 162–164 °C (eter); IR (KBr): ν = 3458, 3023, 3007, 2956, 2930, 2913, 2836, 1742, 1627, 1609 cm‒1; 1H-NMR (CDCl3) δ 6.65 (s, 1H), 6.41 (s, 1H), 6.36 (s, 2H), 5.96 (s, 2H), 5.52 (br s, 1H), 4.13 (t, J =7.3 Hz, 1H), 3.84 (s, 3H), 3.84 (s, 3H), 3.09 (dd, J = 7.3, 15.8 Hz, 1H), 2.92 (dd, J = 7.3, 15.8 Hz, 1H); 13C-NMR (CDCl3) δ 37.2, 40.8, 56.4, 56.4, 99.1, 101.7, 104.2, 104.2, 107.2, 118.2, 131.5, 134.2, 144.5, 146.1, 147.5, 147.5, 147.5, 167.8. (EI) m/z: 344 (M+ C18H16O7, 100), 326 (13), 295 (14), 271 (82), 256 (21), 167 (37), 133 (12).
8-(3,4,5-Trimethoxyphenyl)-7,8-dihydro-[1,3]dioxolo[4,5-g]chromen-6-one (27). This was prepared from 3,4,5-trimethoxycinnamic acid and sesamol as described in the general procedure III. Yield 61%; A white solid; m.p. 162–164 °C (diethyl ether); IR (KBr): ν = 3023, 3007, 2956, 2930, 2913, 2836, 1742, 1627, 1609, 1584 cm‒1; 1H-NMR (CDCl3) δ 6.66 (s, 1H), 6.43 (s, 1H), 6.35 (s, 2H), 5.96 (s, 2H), 4.15 (t, J = 6.8 Hz, 1H), 3.83 (s, 3H), 3.81 (s, 6H), 3.05 (dd, J = 6.8, 16.1 Hz, 1H), 2.92 (dd, J = 6.8, 16.1 Hz, 1H); 13C-NMR (CDCl3) δ 37.1, 41.0, 56.1, 56.1, 60.8, 99.2, 101.8, 104.4, 104.5, 107.2, 117.8, 136.1, 136.1, 137.5, 144.5, 146.1, 147.6, 153.7, 167.6. (EI) m/z: 358 (M+ C19H18O7, 100), 325 (29), 315 (21), 285 (83), 241 (27), 215 (35), 181 (81), 133 (38).
8-Hydroxy-7-methoxy-4-(3,4,5-trimethoxyphenyl)chroman-2-one (28). This was prepared from 3,4,5-trimethoxycinnamic acid and 3-methoxybenzene-1,2-diol as described in the general procedure III. Yield 60%; A white solid; m.p. 164–166 °C (CHCl3/MeOH); IR (KBr): ν = 3309, 3001, 2951, 2841, 1759, 1629, 1588 cm‒1; 1H-NMR (CDCl3) δ 6.64 (d, J = 8.5 Hz, 1H), 6.48 (d, J = 8.5 Hz, 1H), 6.37 (s, 2H), 5.73 (s, 1H), 4.24 (t, J = 7.3 Hz, 1H), 3.90 (s, 3H), 3.83 (s, 3H), 3.80 (s, 6H), 3.09 (dd, J = 7.3, 14.1 Hz, 1H), 2.97 (dd, J = 7.3, 14.1 Hz, 1H); 13C-NMR (CDCl3) δ 37.2, 40.8, 56.4, 56.4, 56.4, 60.8, 104.6, 104.6, 107.0, 117.9, 119.2, 133.8, 136.2, 137.4, 139.4, 147.2, 153.6, 153.6, 166.8. (EI) m/z: 360 (M+ C19H20O7, 100), 345 (22), 327 (35), 317 (19), 287 (99), 272 (24), 217 (17), 181 (31), 173 (22).

3.5. Antiviral Activity Assays

The anti-HIV activity of these neoflavonoids on Tat and NF-κB functions has been evaluated. To this aim, we have used two stably transfected cell lines. The previously described 5.1 cell line [55] is a Jurkat-derived clone stably transfected with a plasmid containing the luciferase gene under the control of HIV-LTR. In this cell clone, activation with TNFα induces NF-κB activation and subsequent HIV-1 expression. We have also analysed the anti-HIV activity in HeLa-Tat-Luc cells, in which the HIV-1 LTR is directly activated by the HIV-1 Tat protein. A compound was considered active in one assay if it inhibited the target function by more than 50% (NF-κB) or 30% (Tat) at either 25 or 50 μM concentration. The active compounds were submitted for further evaluation through a HeLa-Tet-ON assay, as previously described [56]. In the Hela-Tet-ON cells the luciferase expression is under control of an artificial promoter that can be activated by tetracycline. Therefore, compounds that inhibit tetracycline-induced luciferase activity were considered non-specific for luciferase-based anti-HIV assays.
Cell viability was evaluated in non-infected treated cultures following the same protocol as in the recombinant virus assay and measuring cell toxicity with a classical MTT assay. IC50 were calculated using GraphPad Prism software (non-linear regression, log (inhibitor) vs. response).

4. Conclusions

A series of twenty-eight neoflavonoids have been synthesized and evaluated against HIV-1 in vitro. Ten 4-phenylchromen-2-one derivatives displayed HIV specific transcriptional inhibition and five displayed nonspecific mechanisms of action. The heptanoylchromen-one 10 was the more potent Tat antagonist, while compound 14 showed high inhibition of the NF-κB pathway. A preliminary SAR analysis established that the presence of the acyl group is essential for the anti HIV in both targets.

Supplementary Materials

The following are available online at https://www.mdpi.com/1420-3049/22/2/321/s1, Figure S1: Spectroscopic data of neoflavonoid derivatives 310, 1215, 1728.

Acknowledgments

This research is part of the following projects funded by the Spanish Ministry of Economy and Competitiveness and Instituto de Salud Carlos III (PI16/CIII/034); the Spanish AIDS Research Network (RD16CIII/0002/0001) that is included in the Spanish I+D+I Plan and is co-financed by ISCIII-Subdirección General de Evaluacion and European Funding for Regional Development (FEDER), Acknowledgements are also due to the University of Panama, and to the National Secretariat of Science, Technology and Innovation (SENACYT) of Panama for SNI distinguished scientist stimulus award to MPG.

Author Contributions

The authors of these research have participated as follows: “José Luis López-Pérez, Esther Del Olmo and Arturo San Feliciano” conceived and designed structures and synthesis project; “Dionisio A. Olmedo” performed all the synthetic experiments; “José Luis López-Pérez and Dionisio A. Olmedo” interpreted the results, discussed the experimental data and prepared the manuscript; “Rocío Sancho, Eduardo Muñoz, Luis M. Bedoya and José Alcamí” conducted the biological assay and provided the experimental procedure and results; “Dionisio A. Olmedo, José Luis López-Pérez, Arturo San Feliciano, Luis M. Bedoya and Mahabir P. Gupta contributed in overall redaction and revision of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. UNAIDS. The Joint United Nations Programme on HIV/AIDS. Report on the World Statistic the People Living with HIV. 2015. Available online: http://aidsinfo.unaids.org/ (accessed on 21 August 2016).
  2. Piot, P.; Bartos, M.; Ghys, P.D.; Walker, N.; Schwartlander, B. The global impact of HIV/AIDS. Nature 2001, 410, 968–973. [Google Scholar] [CrossRef]
  3. McNicholl, I.R.; McNicholl, J.J. On the horizon: Promising investigational antiretroviral agents. Curr. Pharm. Des. 2006, 12, 1091–1103. [Google Scholar] [CrossRef] [PubMed]
  4. Wainberg, M.A.; Jeang, K.T. 25 years of HIV-1 research—Progress and perspectives. BMC Med. 2008, 6. [Google Scholar] [CrossRef] [PubMed]
  5. Barouch, D.H. Challenges in the development of an HIV-1 vaccine. Nature 2008, 455, 613–619. [Google Scholar] [CrossRef] [PubMed]
  6. Adamson, C.S.; Freed, E.O. Recent progress in antiretrovirals—Lessons from resistance. Drug Discov. Today 2008, 13, 424–432. [Google Scholar] [CrossRef] [PubMed]
  7. Rabson, A.B.; Lin, H.C. NF-κB and HIV: Linking viral and immune activation. Adv. Pharmacol. 2000, 48, 161–207. [Google Scholar] [PubMed]
  8. Gatignol, A.; Duarte, M.; Daviet, L.; Chang, Y.N.; Jeang, K.T. Sequential steps in Tat trans-activation of HIV-1 mediated through cellular DNA, RNA, and protein binding factors. Gene Expr. 1996, 5, 217–228. [Google Scholar] [PubMed]
  9. Chen, B.K.; Feinberg, M.B.; Baltimore, D. The κB sites in the human immunodeficiency virus type 1 long terminal repeat enhance virus replication yet are not absolutely required for viral growth. J. Virol. 1997, 71, 5495–5504. [Google Scholar]
  10. Greene, W.C. The molecular biology of human immunodeficiency virus type 1 infection. N. Engl. J. Med. 1991, 324, 308–317. [Google Scholar] [PubMed]
  11. Alcamí, J.; Lain de Lera, T.; Folgueira, L.; Pedraza, M.A.; Jacque, J.M.; Bachelerie, F.; Noriega, A.R.; Hay, R.T.; Harrich, D.; Gaynor, R.B. Absolute dependence on kappa B responsive elements for initiation and Tat-mediated amplification of HIV transcription in blood CD4 T lymphocytes. EMBO J. 1995, 14, 1552–1560. [Google Scholar] [PubMed]
  12. Stevenson, M. Tat’s seductive side. Nat. Med. 2003, 9, 163–164. [Google Scholar] [CrossRef] [PubMed]
  13. Yeni, P.G.; Hammer, S.M.; Carpenter, C.C.; Cooper, D.A.; Fischl, M.A.; Gatell, J.M.; Gazzard, B.G.; Hirsch, M.S.; Jacobsen, D.M.; Katzenstein, D.A.; et al. Antiretroviral treatment for adult HIV infection in 2002: Updated recommendations of the International AIDS Society-USA Panel. JAMA 2002, 288, 222–235. [Google Scholar] [CrossRef] [PubMed]
  14. Karin, M.; Yamamoto, Y.; Wang, Q.M. The IKK NF-κB system: A treasure trove for drug development. Nat. Rev. Drug Discov. 2004, 3, 17–26. [Google Scholar] [CrossRef] [PubMed]
  15. Ferchichi, L.; Derbre, S.; Mahmood, K.; Toure, K.; Guilet, D.; Litaudon, M.; Awang, K.; Hadi, A.H.A.; Le Ray, A.M.; Richomme, P. Bioguided fractionation and isolation of natural inhibitors of advanced glycation end-products (AGEs) from Calophyllum flavoramulum. Phytochemistry 2012, 78, 98–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Brenzan, M.A.; Nakamura, C.V.; Dias Filho, B.P.; Ueda-Nakamura, T.; Young, M.C.M.; García Cortez, D.A. Antileishmanial activity of crude extract and coumarin from Calophyllum brasiliense leaves against Leishmania amazonensis. Parasitol. Res. 2007, 101, 715–722. [Google Scholar] [CrossRef] [PubMed]
  17. Guilet, D.; Helesbeux, J.-J.; Seraphin, D.; Sevenet, T.; Richomme, P.; Bruneton, J. Novel cytotoxic 4-phenylfuranocoumarins from Calophyllum dispar. J. Nat. Prod. 2001, 64, 563–568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Guilet, D.; Seraphin, D.; Rondeau, D.; Richomme, P.; Bruneton, J. Cytotoxic coumarins from Calophyllum dispar. Phytochemistry 2001, 58, 571–575. [Google Scholar] [CrossRef]
  19. Kashman, Y.; Gustafson, K.R.; Fuller, R.W.; Cardellina, J.H.; McMahon, J.B.; Currens, M.J.; Buckheit, R.W.; Hughes, S.H.; Cragg, G.M.; Boyd, M.R. HIV inhibitory natural products. Part 7. The calanolides, a novel HIV-inhibitory class of coumarin derivatives from the tropical rainforest tree, Calophyllum lanigerum. J. Med. Chem. 1992, 35, 2735–2743. [Google Scholar] [CrossRef] [PubMed]
  20. Patil, A.D.; Freyer, A.J.; Eggleston, D.S.; Haltiwanger, R.C.; Bean, M.F.; Taylor, P.B.; Caranfa, M.J.; Breen, A.L.; Bartus, H.R. The inophyllums, novel inhibitors of HIV-1 reverse transcriptase isolated from the Malaysian tree, Calophyllum inophyllum Linn. J. Med. Chem. 1993, 36, 4131–4138. [Google Scholar] [CrossRef] [PubMed]
  21. Games, D.E. Identification of 4-phenyl and 4-alkylcoumarins in Mammea americana L., Mammea africana G. Don and Calophyllum inophyllum by gas chromatography. Mass Spectrometry. Tetrahedron Lett. 1972, 13, 3187–3190. [Google Scholar] [CrossRef]
  22. Reutrakul, V.; Leewanich, P.; Tuchinda, P.; Pohmakotr, M.; Jaipetch, T.; Sophasan, S.; Santisuk, T. Cytotoxic coumarins from Mammea harmandii. Planta Med. 2003, 69, 1048–1051. [Google Scholar] [PubMed]
  23. Prachyawarakorn, V.; Mahidol, C.; Ruchirawat, S. NMR study of seven coumarins from Mammea siamensis. Pharm. Biol. 2000, 38, 58–62. [Google Scholar] [CrossRef] [PubMed]
  24. Crombie, L.; Jones, R.C.F.; Palmer, C.J. Synthesis of the mammea coumarins. Part 1. The coumarins of the mammea A, B, and C series. J. Chem. Soc. Perkin Trans. 1 1987, 317–331. [Google Scholar] [CrossRef]
  25. Carpenter, I.; McGarry, E.J.; Scheinmann, F. Extractives from Guttiferae. Part XXI: The isolation and structure of nine coumarins from the bark of Mammea africana G. Don. J. Chem. Soc. C. 1971, 3783–3790. [Google Scholar] [CrossRef]
  26. Crombie, L.; Games, D.E.; McCormick, A. Extractives of Mammea americana L. Part II. The 4-phenylcoumarins isolation and structure of Mammea A/AA, A/A cyclo D, A/BA, A/AB, and A/BB. J. Chem. Soc. C. 1967, 2255–2260. [Google Scholar] [CrossRef]
  27. Rouger, C.; Derbré, S.; Charreau, B.; Pabois, A.; Cauchy, T.; Litaudon, M.; Awang, K.; Richomme, P. Lepidotol A from Mesua lepidota Inhibits Inflammatory and Immune Mediators in Human Endothelial Cells. J. Nat. Prod. 2015, 78, 2187–2197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Cheng Lian, G.; Sin The, S.; Hui Mah, S.; Rahmani, M.; Taufiq-Yap, Y.H.; Awang, K. A Novel Cyclodione Coumarin from the Stem Bark of Mesua beccariana. Molecules 2011, 16, 7249–7255. [Google Scholar]
  29. Awang, K.; Chan, G.; Litaudon, M.; Ismail, N.H.; Martin, M.-T.; Gueritte, F.O. 4-Phenylcoumarins from Mesua elegans with acetylcholinesterase inhibitory activity. Bioorg. Med. Chem. 2010, 18, 7873–7877. [Google Scholar] [CrossRef] [PubMed]
  30. Verotta, L.; Lovaglio, E.; Vidari, G.; Finzi, P.V.; Neri, M.G.; Raimondi, A.; Parapini, S.; Taramelli, D.; Riva, A.; Bombardelli, E. 4-Alkyl- and 4-phenylcoumarins from Mesua ferrea as promising multidrug resistant antibacterials. Phytochemistry 2004, 65, 2867–2879. [Google Scholar] [CrossRef] [PubMed]
  31. Morel, C.; Dartiguelongue, C.; Youhana, T.; Oger, J.M.; Seraphin, D.; Duval, O.; Richomme, P.; Bruneton, J. New coumarins from Mesua racemosa: Isolation and synthesis. Heterocycles 1999, 51, 2183–2191. [Google Scholar]
  32. Morel, C.; Guilet, D.; Oger, J.M.; Seraphin, D.; Sevenet, T.; Wiart, C.; Hadi, A.H.A.; Richomme, P.; Bruneton, J. 6-Acylcoumarins from Mesua racemosa. Phytochemistry 1999, 50, 1243–1247. [Google Scholar] [CrossRef]
  33. Bala, K.R.; Seshadri, T.R. Isolation and synthesis of some coumarin components of Mesua ferrea seed oil. Phytochemistry 1971, 10, 1131–1134. [Google Scholar] [CrossRef]
  34. Cruz, F.G.; Moreira, L.d.M.; Santos, N.A.S.; Guedes, M.L.S. Additional Coumarins from Kielmeyera reticulate. J. Braz. Chem. Soc. 2002, 13, 704–707. [Google Scholar] [CrossRef]
  35. Cruz, F.G.; da Silva-Neto, J.T.; Guedes, M.L.S. Xanthones and Coumarins from Kielmeyera lathrophyton. J. Braz. Chem. Soc. 2001, 12, 117–122. [Google Scholar] [CrossRef]
  36. Gramacho, R.d.S.; Nagem, T.J.; de Oliveira, T.T.; de Queiroz, M.E.L.R.; Neves, A.A.; Saddi, N. Phenylcoumarins from Kielmeyera elata. Phytochemistry 1999, 51, 579–581. [Google Scholar] [CrossRef]
  37. Cruz, F.G.; Santos, N.A.S.; David, J.M.; Guedes, M.L.S.; Chávez, J.P. Coumarins from Kielmeyera argentea. Phytochemistry 1998, 48, 703–706. [Google Scholar] [CrossRef]
  38. Cruz, F.G.; Moreira, L.M.; David, J.M.; Guedes, M.L.S.; Chávez, J.P. Coumarins from Kielmeyera reticulata. Phytochemistry 1998, 47, 1363–1366. [Google Scholar]
  39. López-Pérez, J.L.; Olmedo, D.A.; del Olmo, E.; Vásquez, Y.; Solís, P.N.; Gupta, M.P.; San Feliciano, A. Cytotoxic 4-phenylcoumarins from the leaves of Marila pluricostata. J. Nat. Prod. 2005, 68, 369–373. [Google Scholar] [CrossRef] [PubMed]
  40. Ishikawa, T. Chemistry of Anti HIV-1 Active Calophyllum Coumarins. J. Synth. Org. Chem. Jpn. 1998, 56, 116–124. [Google Scholar] [CrossRef]
  41. Chiang, C.C.; Mouscadet, J.F.; Tsai, H.J.; Liu, C.T.; Hsu, L.Y. Synthesis and HIV-1 integrase inhibition of novel bis- or tetra-coumarin analogues. Chem. Pharm. Bull. 2007, 55, 1740–1743. [Google Scholar] [CrossRef] [PubMed]
  42. Márquez, N.; Sancho, R.; Bedoya, L.M.; Alcamí, J.; López-Pérez, J.L.; San Feliciano, A.; Fiebich, B.L.; Muñoz, E. Mesuol, a natural occurring 4-phenylcoumarin, inhibits HIV-1 replication by targeting the NF-κB pathway. Antivir. Res. 2005, 66, 137–145. [Google Scholar] [CrossRef] [PubMed]
  43. Bedoya, L.M.; Beltrán, M.; Sancho, R.; Olmedo, D.A.; Sánchez-Palomino, S.; Olmo, E.; López-Pérez, J.L.; Muñoz, E.; San Feliciano, A.; Alcamí, J. 4-Phenylcoumarins as HIV transcription inhibitors. Bioorg. Med. Chem. Lett. 2005, 15, 4447–4450. [Google Scholar] [CrossRef] [PubMed]
  44. Krishna, C.; Bhargavi, M.V.; Rao, C.P.; Krupadanama, D. Synthesis and antimicrobial assessment of novel coumarins featuring 1,2,4-oxadiazole. Med. Chem. Res. 2015, 24, 3743–3751. [Google Scholar] [CrossRef]
  45. Chin, Y.P.; Huang, W.J.; Hsu, F.L.; Lin, Y.L.; Lin, M.H. Synthesis and evaluation of antibacterial activities of 5,7-Dihydroxycoumarin derivatives. Arch. Pharm. 2011, 344, 386–393. [Google Scholar] [CrossRef] [PubMed]
  46. Hwang, C.H.; Jaki, B.U.; Klein, L.L.; Lankin, D.C.; McAlpine, J.B.; Napolitano, J.G.; Fryling, N.A.; Franzblau, S.G.; Cho, S.H.; Stamets, P.E.; et al. Chlorinated Coumarins from the Polypore Mushroom Fomitopsis officinalis and Their Activity against Mycobacterium tuberculosis. J. Nat. Prod. 2013, 76, 1916–1922. [Google Scholar] [CrossRef] [PubMed]
  47. Lin, C.M.; Huang, S.T.; Lee, F.W.; Kuo, H.S.; Lin, M.H. 6-Acyl-4-aryl/alkyl-5,7-dihydroxycoumarins as anti-inflammatory agents. Bioorg. Med. Chem. 2006, 14, 4402–4409. [Google Scholar] [CrossRef] [PubMed]
  48. Del Olmo, E. Final Report Project X.11. Iberoamerican Program of Science and Technology for Development: Madrid, Spain, 2004; (unpublished results). [Google Scholar]
  49. Cao, S.G.; Wu, X.H.; Sim, K.Y.; Tan, B.H.K.; Vittal, J.J.; Pereira, J.T.; Goh, S.H. Minor coumarins from Calophyllum teysmannii var. inophylloide and synthesis of cytotoxic calanone derivatives. Helv. Chim. Acta 1998, 81, 1404–1416. [Google Scholar] [CrossRef]
  50. Cao, S.G.; Sim, K.Y.; Goh, S.H. Three new coumarins from Calophyllum teysmannii var. inophylloide (Guttiferae). Heterocycles 1997, 45, 2045–2052. [Google Scholar]
  51. Kulkarni, M.V.; Kulkarni, G.M.; Lin, C.H.; Sun, C.M. Recent advances in coumarins and 1-azacoumarins as versatile biodynamic agents. Curr. Med. Chem. 2006, 13, 2795–2818. [Google Scholar] [CrossRef] [PubMed]
  52. Palmer, C.J.; Josephs, J.L. Synthesis of the Calophyllum coumarins. Part 2. J. Chem. Soc. Perkin Trans 1 1995, 3135–3152. [Google Scholar] [CrossRef]
  53. NAPROC-13 RMN Spectroscopic Database USAL. Available online: http://c13.usal.es (accessed on 17 February 2017).
  54. Gaurav Taneja, A.G.; Raghuvanshi, A.; Kant, R.; Maulik, P.R. Diversity-oriented general protocol for the synthesis of privileged oxygen scaffolds: Pyrones, coumarins, benzocoumarins and naphthocoumarins. Org. Biomol. Chem. 2013, 11, 5239–5253. [Google Scholar]
  55. Kamat, S.P.; D’Souza, A.M.; Paknikar, S.K.; Beauchamp, P.S. A convenient one-pot synthesis of 4-methyl-3-phenyl-, 3-aryl- and 3-aryl-4-phenylcoumarins. J. Chem. Res. Synop. 2002, 242–246. [Google Scholar] [CrossRef]
  56. Sancho, R.; Medarde, M.; Sánchez-Palomino, S.; Madrigal, B.M.; Alcamí, J.; Muñoz, E.; San Feliciano, A. Anti-HIV activity of some synthetic lignanolides and intermediates. Bioorg. Med. Chem. Lett. 2004, 14, 4483–4486. [Google Scholar] [CrossRef] [PubMed]
  • Sample Availability: Samples of the compounds are available from José Luis López-Pérez, E-Mail: [email protected].
Scheme 1. Preparation of 4-phenylcoumarins 17.
Scheme 1. Preparation of 4-phenylcoumarins 17.
Molecules 22 00321 sch001
Scheme 2. Preparation of neoflavones 821.
Scheme 2. Preparation of neoflavones 821.
Molecules 22 00321 sch002
Scheme 3. Preparation of neoflavones 2228.
Scheme 3. Preparation of neoflavones 2228.
Molecules 22 00321 sch003
Table 1. Anti-HIV Activity of neoflavonoids.
Table 1. Anti-HIV Activity of neoflavonoids.
CompoundNF-κB (5.1 LTR)Hela-Tat-lucSpecificity (HeLa-Tet-On-Luc)Toxicity MT2 (%)
25 µM50 µM25 µM50 µM50 µM50 µM
1NT−11.28NT27.88NT18.40
2NT−4.10NT−7.43NT8.78
3NT34.27NT6.59NT2.22
47.3021.665.5834.64S3.55
522.9823.83−2.6344.60S11.30
6NT19.12NT16.25NT2.88
7NT51.71NT94.69U4.00
8NT9.25NT21.74NT8.33
9NT68.74NT80.84UNT
1070.5368.19NT83.32S<10
11NT37.81NT26.63SNT
12NT20.40NT5.81SNT
13NT67.29NT66.72UNT
1483.0686.6041.8769.32S17.02
15NT79.41NT80.37UNT
16NT−17.05NT20.87SNT
17NT83.70NT44.93UNT
18NT35.05NT30.20SNT
1959.8666.04NT12.99SNT
20NT10.70NT22.30SNT
22NT15.20NT6.06NT1.61
23NT11.94NT−34.54NT2.55
2436.9543.21NT−18.03NT2.00
2535.2053.9957.4672.27S3.50
26NT13.37NT−28.08NT2.59
27NT20.48NT15.08NT4.73
28NT5.67NT−65.28NT6.13
Mesuol71.0077.90NT71.30S>4 μM
S = Specific activity; U = Unspecific mode of action; NT = Not tested.

Share and Cite

MDPI and ACS Style

Olmedo, D.A.; López-Pérez, J.L.; Del Olmo, E.; Bedoya, L.M.; Sancho, R.; Alcamí, J.; Muñoz, E.; Feliciano, A.S.; Gupta, M.P. Neoflavonoids as Inhibitors of HIV-1 Replication by Targeting the Tat and NF-κB Pathways. Molecules 2017, 22, 321. https://doi.org/10.3390/molecules22020321

AMA Style

Olmedo DA, López-Pérez JL, Del Olmo E, Bedoya LM, Sancho R, Alcamí J, Muñoz E, Feliciano AS, Gupta MP. Neoflavonoids as Inhibitors of HIV-1 Replication by Targeting the Tat and NF-κB Pathways. Molecules. 2017; 22(2):321. https://doi.org/10.3390/molecules22020321

Chicago/Turabian Style

Olmedo, Dionisio A., José Luis López-Pérez, Esther Del Olmo, Luis M. Bedoya, Rocío Sancho, José Alcamí, Eduardo Muñoz, Arturo San Feliciano, and Mahabir P. Gupta. 2017. "Neoflavonoids as Inhibitors of HIV-1 Replication by Targeting the Tat and NF-κB Pathways" Molecules 22, no. 2: 321. https://doi.org/10.3390/molecules22020321

Article Metrics

Back to TopTop