Next Article in Journal
Thermal Properties of Illite-Zeolite Mixtures up to 1100 °C
Next Article in Special Issue
Differences in the Structure and Antimicrobial Activity of Hydrazones Derived from Methyl 4-Phenylpicolinimidate
Previous Article in Journal
Charge Accumulation in the Homo-Crosslinked-Polyethylene Bilayer
Previous Article in Special Issue
Attachment of Chiral Functional Groups to Modify the Activity of New GPx Mimetics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stereoselective Synthesis and Anticancer Activity of 2,6-Disubstituted trans-3-Methylidenetetrahydropyran-4-ones

by
Tomasz Bartosik
1,†,
Joanna Drogosz-Stachowicz
2,†,
Anna Janecka
2,
Jacek Kędzia
1,
Barbara Pacholczyk-Sienicka
1,
Jacek Szymański
3,
Katarzyna Gach-Janczak
2 and
Tomasz Janecki
1,*
1
Institute of Organic Chemistry, Lodz University of Technology, Żeromskiego 116, 90-924 Lodz, Poland
2
Department of Biomolecular Chemistry, Medical University of Łódź, Mazowiecka 6/8, 92-215 Lodz, Poland
3
Central Scientific Laboratory, Medical University of Łódź, Mazowiecka 6/8, 92-215 Lodz, Poland
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2022, 15(9), 3030; https://doi.org/10.3390/ma15093030
Submission received: 10 March 2022 / Revised: 14 April 2022 / Accepted: 18 April 2022 / Published: 21 April 2022

Abstract

:
In this report, we present efficient and stereoselective syntheses of 2,6-disubstituted trans-3-methylidenetetrahydropyran-4-ones and 2-(4-methoxyphenyl)-5-methylidenetetrahydropyran-4-one that significantly broaden the spectrum of the available methylidenetetrahydropyran-4-ones with various substitution patterns. Target compounds were obtained using Horner–Wadsworth–Emmons methodology for the introduction of methylidene group onto the pyranone ring. 3-Diethoxyphosphoryltetrahydropyran-4-ones, which were key intermediates in this synthesis, were prepared by fully or highly stereoselective addition of Gilman or Grignard reagents to 3-diethoxyphosphoryldihydropyran-4-ones. Addition occurred preferentially by axial attack of the Michael donors on the dihydropyranone ring. Relative configurations and conformations of the obtained adducts were assigned using a detailed analysis of the NMR spectra. The obtained methylidenepyran-4-ones were evaluated for cytotoxic activity against two cancer cell lines (HL-60 and MCF-7). 2,6-Disubstituted 3-methylidenetetrahydropyran-4-ones with isopropyl and phenyl substituents in position 2 were more cytotoxic than analogs with n-butyl substituent. Two of the most cytotoxic analogs were then selected for further investigation on the HL-60 cell line. Both analogs induced morphological changes characteristic of apoptosis in cancer cells, significantly inhibited proliferation and induced apoptotic cell death. Both compounds also generated DNA damage, and one of the analogs arrested the cell cycle of HL-60 cells in the G2/M phase. In addition, both analogs were able to inhibit the activity of topoisomerase IIα. Based on these findings, the investigated analogs may be further optimized for the development of new and effective topoisomerase II inhibitors.

Graphical Abstract

1. Introduction

The tetrahydropyran skeleton is present in a great number of natural products, which show various important biological properties, including anticancer activity. This activity depends on many structural and stereochemical features, and one of the most important is the substitution pattern of the tetrahydropyran ring. In this respect, many compounds containing a trans-2,6-disubstituted tetrahydropyrane moiety show significant cytotoxic activity. Representative examples are irciniastatins A and B 1 isolated from the Indo-Pacific marine sponge Ircinia ramosoa, which are powerful murine and human cancer cell growth inhibitors [1]; and aspergillide C 2 isolated from marine-derived fungus Aspergillus ostianus strain 01F313, which showed cytotoxic activity against mouse lymphocytic leukemia cells (L1210) [2] (Figure 1). Another important feature, which can strongly influence cytotoxic activity of tetrahydropyranes is the presence of exo-methylidene bond vicinal to the carbonyl group. This moiety is a strong Michael acceptor and can react with various bionucleophiles—in particular, with mercapto groups in cysteine residues of enzymes or other functional proteins. The exo-methylidene motif is characteristic of a large group of natural products, including sesquiterpene lactones (e.g., vernolepin 3) and is believed to be the reason for their strong and often very specific anticancer properties [3,4,5,6]. Several natural compounds containing the 2,6-disubstituted-3-methylidenetetrahydropyran-4-one skeleton are known. Representative examples are norperovskatone 4, recently isolated from Perovskia atriplicifolia, which possesses remarkable anti-hepatitis B virus (HBV) activity [7] or (+)-okilactomycin 5 [8] which exhibits in vitro cytotoxicity against a number of human cancer cell lines, including lymphoid leukemia L1210 and leukemia P388, and in vivo activity against Ehrlich ascites carcinoma.
Very recently, we reported on the synthesis and anticancer properties of a series of 2,2,6-trisubstituted 5-methylidenetetrahydropyran-4-ones (6) [9]. The analysis of the structure–activity relationship of the obtained compounds has revealed that cytotoxicity strongly depends on the nature of the substituent in position 6 (vicinal to the methylidene bond)—butyl and particularly iso-propyl groups being the best. The nature of substituents in position 2 is less important, although, in general, aromatic substituents are more advantageous for cancer growth inhibition than non-aromatic ones. One of the obtained compounds, 6-isopropyl-2,2-dimethyl-5-methylidenetetrahydropyran-4-one, showed very high cytotoxic activity against HL-60 human leukemia cells, induced apoptosis of these cells and caused the arrest of the cell cycle in the G2/M phase. Encouraged by these results, in this paper we present a simple, effective and stereoselective synthesis of 2,6-disubstituted trans-3-methylidenetetrahydropyran-4-ones 13a–o, which significantly broadens the spectrum of the available methylidenetetrahydropyran-4-ones with various substitution patterns. We applied synthetic methodology which proved its effectiveness in the synthesis of 5-methylidenetetrahydropyran-4-ones 6 and used Horner–Wadsworth–Emmons reagents for the olefination of formaldehyde to introduce an exo-methylidene group onto the tetrahydropyranone ring. In designing substituents for the new methylidenetetrahydropyran-4-one analogs, we turned special attention to these, which proved their effectiveness in enhancing the cytotoxicity of 2,2,6-trisubstituted 3-methylidenetetrahydropyra-4-ones 6, i.e., isopropyl and n-butyl groups. Additionally, we decided to introduce the ferrocenyl moiety onto the tetrahydropyran-4-one skeleton. The ferrocenyl group has proved itself to be an excellent choice to design new drugs [10,11,12], most probably because of its small size, aromaticity, hydrophobicity, low cytotoxicity against normal cells and redox behavior [13,14,15,16,17,18]. Several ferrocene derivatives have already been used as anticancer, antimalarial, antiviral and antibiotic agents [14,15,16,19,20].
The obtained methylidenepyran-4-ones 13a–o were evaluated for their cytotoxic activity against two cancer cell lines (HL-60 and MCF-7), and for comparison, against a normal cell line (HUVEC). Additionally, the influences of two selected analogs on proliferation, cell cycle distribution, apoptosis induction, DNA damage and topoisomerase IIα activity were also investigated.

2. Materials and Methods

2.1. Chemistry

General information
NMR spectra were recorded on a Bruker DPX 250 or Bruker Avance II Plus instrument (Bruker, Billerica, MA, USA) at 250.13 MHz or 700 for 1H; 62.9 or 176 MHz for 13C; and 101.3 MHz for 31P NMR. Tetramethylsilane was used as an internal standard and 85% H3PO4 as an external standard. 31P NMR spectra were recorded using broadband proton decoupling. Melting points were determined in open capillaries and are uncorrected. Column chromatography was performed on silica gel 60 (230–400 mesh) (Aldrich, Sant Louis, MO, USA). Thin-layer chromatography was performed on the pre-coated TLC sheets of silica gel 60 F254 (Aldrich). The purity of the synthesized compounds was confirmed by the combustion elemental analyses (CH, elemental analyzer EuroVector 3018, Elementar Analysensysteme GmbH, Langenselbold, Germany). MS spectra of intermediates were recorded on Waters 2695-Waters ZQ 2000 LC/MS apparatus. EI mass spectra of final compounds were recorded on a GCMS-QP2010 ULTR A instrument (Shimadzu, Kioto, Japan). The mass spectra were obtained using the following operating conditions: electron energy of 70 eV and ion source temperature of 200 °C. Samples were introduced via a direct insertion probe heated from 30 to 300 °C. All reagents and starting materials were purchased from commercial vendors and used without further purification. Organic solvents were dried and distilled prior to use. Standard syringe techniques were used for transferring dry solvents.
General procedure for the synthesis of diethyl 4-alkyl(aryl)-4-hydroxy-2-oxobutylphosphonates 9a–e.
In a round-bottomed three-necked flask under an argon atmosphere, NaH (0.51 g, 17.00 mmol, 80% in mineral oil) was suspended in THF (48 mL). The suspension was stirred, and a solution of diethyl 2-oxopropylphosphonate 7 (3.00 g, 15.45 mmol) in THF (3 mL) was added dropwise. The reaction mixture was stirred for 30 min, cooled below −30 °C in dry ice–acetone bath and n-butyllithium (6.80 mL of 2.5 M solution in hexane, 17.00 mmol) was added dropwise. The reaction mixture was stirred for 30 min at this temperature, cooled to −78 °C in a dry ice–acetone bath and solution of aldehyde (18.54 mmol) in THF (6 mL) was added dropwise. Reaction mixture was stirred for 1.5 h at this temperature. After this time, the reaction was quenched by adding saturated solution of ammonium chloride (100 mL). The water layer was extracted with DCM (3 × 100 mL); the organic layers were combined, washed with brine and dried over MgSO4. The solvents were evaporated under reduced pressure, and the resulting crude product was purified by column chromatography (eluent ethyl acetate).
Characteristics of the compounds 9a-e are given in the Supporting Information.
General procedure for the synthesis of 6-alkyl(aryl)-3-diethoxyphosphoryldihydropyran-4-ones 11a–e.
The diethyl 4-alkyl(aryl)-4-hydroxy-2-oxobutylphosphonates (9a–d) (10.00 mmol) was dissolved in dry dichloromethane (100 mL). Next, dimethyl formamide dimethyl acetal (DMF–DMA) (30.00 mmol, 4.24 mL) was added dropwise and stirred for 2 h. After this time to reaction mixture boron trifluoride etherate (BF3 · Et2O) (25.00 mmol, 3.10 mL) was added dropwise. In the case of synthesis of 6-alkyl(aryl)-3-diethoxyphosphoryldihydropyran-4-ones 11e, the reaction was carried without boron trifluoride etherate and reduced amount of dimethyl formamide dimethyl acetal (DMF–DMA) (25.00 mmol, 3.53 mL). Reaction was controlled by 31P NMR. After completion of the reaction, to the reaction mixture was added saturated solution of sodium hydrocarbonate (100 mL). The water layer was extracted with ethyl acetate (3 × 100 mL), the organic layers were combined, washed with brine and dried over MgSO4. The solvents were evaporated under reduced pressure. Obtained crude product was used on next step without further purification.
General procedure for the synthesis of 2-alkyl(aryl)-6-alkyl(aryl)-3-diethoxyphosphoryltetrahydropyran-4-ones 12a–j.
To a suspension of copper iodide (I) (0.80 mmol, 305 mg) in THF (10.70 mL) was added dropwise n-BuLi 2.5 M in hexane solution (3.00 mmol, 1.25 mL) at 0 °C. The solution was stirred at this temperature for 20 min. After this time the mixture was cooled to −78 °C, chlorotrimethylsilane (5.00 mmol, 0.65 mL) was added dropwise and stirred for 10 min. Next, to obtained mixture was added a solution of 6-alkyl(aryl)-3-diethoxyphosphoryl- dihydropyran-4-one 11a-e (1.00 mmol) in dry THF (7.00 mL) and stirred at −78 °C for 4 h. After this time the reaction was quenched with saturated aqueous NH4Cl (7.50 mL) and stirred for 20 min. Next, the mixture was diluted with saturated aqueous NH4Cl (25.00 mL) and extracted with ethyl acetate (5 × 30 mL). The combined organic layers was washed with H2O (50.00 mL) and brine (50.00 mL), then dried over MgSO4 and concentrated under reduced pressure. Obtained crude product was purified by column chromatography using as eluent dichloromethane: acetone (20:1) mixture.
Characteristics of the compounds 12a–j are given in the Supporting information.
General procedure for the synthesis of 2-isopropyl-6-alkyl(aryl)-3-diethoxyphosphoryltetrahydropyran-4-ones 12k–o.
The solution of 6-alkyl(aryl)-3-diethoxyphosphoryldihydropyran-4-one 11a-e (1.0 mmol) in dry THF (10 mL) was cooled to 0 °C in ice-water bath and isopropylmagnesium chloride (3.0 mmol in THF) was added dropwise in argon atmosphere. After 2 h at this temperature, saturated solution of ammonium chloride (15 mL) was added. The water layer was washed with DCM (3 × 15 mL). Combined organic extracts were washed with brine (20 mL) and dried over MgSO4. The solvents were evaporated under reduced pressure and the resulting crude product was purified by column chromatography (eluent DCM : Acetone 20:1).
Characteristics of the compounds 12k–o are given in the Supporting information.
General procedure for the synthesis of 2-alkyl(aryl)-5-diethoxyphosphoryltetrahydropyran-4-ones 14a-d.
The solution of 6-alkyl(aryl)-3-diethoxyphosphoryldihydropyran-4-one 11a-d (1.0 mmol) in dry THF (10 mL) was cooled to −78 °C in dry ice–acetone bath and L-selectride® (1.1 mmol in THF) was added dropwise in argon atmosphere. After 1 h at this temperature, the reaction was carried out at 0 °C for 1 h. Reaction was completed by using saturated solution of ammonium chloride (15 mL). The water layer was washed with DCM (3 × 15 mL). Combined organic extracts were washed with brine (20 mL) and dried over MgSO4. The solvents were evaporated under reduced pressure and the resulting crude product was purified by column chromatography (eluent DCM : Acetone 20:1).
Characteristics of the compounds 14a-d are given in the Supporting information.
General procedure for the synthesis of trans-2,6,-disubstituted-3-methylidenetetrahydropyran-4-ones 13a–o and 15d.
To a vigorously stirred solution of 3-diethoxyphosphoryltetrahydropyran-4-ones 12a–o and 14d (0.2 mmol) in dry THF (2 mL), formaldehyde (36–38% solution in water, 0.167 mL, ca. 2.00 mmol) was added at 0 °C, followed by addition of K2CO3 (55.3 mg, 0.40 mmol) in water (0.56 mL). The resulting mixture was stirred vigorously at 0 °C for 2 h (for pyran-4-ones 12a–o and 14d) or at room temperature for 3 h (for pyran-4-ones 12k–o). Next Et2O (10 mL) was added and layers were separated. The water fraction was washed with Et2O (5 mL). Organic fractions were combined, washed with brine (10 mL) and dried over MgSO4. The solvents were evaporated under reduced pressure and the resulting crude product was purified by column chromatography (eluent dichloromethane: petroleum ether 5:1).
Trans-6-ethyl-2-isopropyl-3-methylidenetetrahydro-4H-pyran-4-one (13a) (16.0 mg, 44%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 0.88 (d, J = 6.6 Hz, 3H), 1.00 (t, J = 7.4 Hz, 3H), 1.02 (d, J = 6.6 Hz, 3H), 1.54 (dqd, J = 13.8, 7.5, 5.0 Hz, 1H), 1.63–1.69 (m, 1H), 1.92 (dhept, J = 9.4, 6.6 Hz, 1H), 2.31 (dd, J = 16.6, 10.2 Hz, 1H), 2.56 (dd, J = 16.6, 3.4 Hz, 1H), 3.85–3.96 (m, 1H), 4.04–4.14 (m, 1H), 5.13 (dd, J = 1.3, 1.3 Hz, 1H), 6.00 (dd, J = 1.3, 1.3 Hz, 1H). 13C NMR (176 MHz, chloroform-d) δ 10.02, 18.16, 19.30, 29.33, 29.62, 46.25, 70.36, 82.74, 121.08, 144.16, 198.69. EI-MS [M]+ = 182.0. Anal. Calcd for C11H18O2: C, 72.49; H, 9.95. Found: C, 72.28; H, 9.91.
Trans-2-butyl-6-ethyl-3-methylidenetetrahydro-4H-pyran-4-one (13b) (34.5 mg, 88%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 0.91 (t, J = 7.1 Hz, 3H), 0.97 (t, J = 7.1 Hz, 3H), 1.20–1.58 (m, 6H), 1.62 (dhept, J = 13.7, 7.4 Hz, 1H), 1.81 (dtd, J = 13.2, 9.7, 3.9 Hz, 1H), 2.30 (dd, J = 16.9, 10.5 Hz, 1H), 2.54 (dd, J = 16.9, 3.2 Hz, 1H), 3.88 (dddd, J = 10.6, 7.9, 4.9, 3.2 Hz, 1H), 4.59 (dddd, J = 10.0, 5.0, 1.5, 1.5 Hz, 1H), 5.17 (dd, J = 1.5, 1.5 Hz, 1H), 6.02 (d, J = 1.5, 1.5 Hz, 1H). 13C NMR (176 MHz, chloroform-d) δ 9.94, 14.13, 22.47, 27.68, 29.05, 34.11, 45.86, 69.66, 76.13, 120.26, 145.34, 197.83. EI-MS [M]+ = 196.0. Anal. Calcd for C12H20O2: C, 73.43; H, 10.27. Found: C, 73.50; H, 10.26.
Trans-6-ethyl-3-methylidene-2-phenyltetrahydro-4H-pyran-4-one (13c) (29.0 mg, 67%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 0.90 (t, J = 7.4 Hz, 3H), 1.49 (dqd, J = 13.8, 7.4, 5.0 Hz, 1H), 1.62 (dp, J = 13.8, 7.4 Hz, 1H), 2.39 (dd, J = 16.9, 10.6 Hz, 1H), 2.52 (dd, J = 16.9, 3.2 Hz, 1H), 3.67 (dddd, J = 10.6, 7.9, 5.0, 3.2 Hz, 1H), 5.26 (dd, J = 1.3, 1.3 Hz, 1H), 5.79 (s, 1H), 6.30 (dd, J = 1.3, 1.3 Hz, 1H), 7.27–7.38 (m, 5H). 13C NMR (176 MHz, chloroform-d) δ 9.70, 28.87, 45.95, 70.56, 78.27, 123.10, 127.93 (2 × C), 128.18, 128.62 (2 × C), 139.55, 142.36, 197.64. EI-MS [M]+ = 216.0. Anal. Calcd for C14H16O2: C, 77.75; H, 7.46. Found: C, 77.44; H, 7.48.
Trans-2,6-diisopropyl-3-methylidenetetrahydro-4H-pyran-4-one (13d) (19.6 mg, 50%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 0.87 (d, J = 6.7 Hz, 3H), 0.92 (d, J = 6.8 Hz, 3H), 0.99–1.04 (m, 6H), 1.75 (dhept, J = 13.5, 6.8 Hz, 1H), 1.91 (dhept, J = 9.6, 6.7 Hz, 1H), 2.35 (dd, J = 16.5, 10.6 Hz, 1H), 2.55 (dd, J = 16.5, 3.3 Hz, 1H), 3.68 (ddd, J = 10.6, 6.8, 3.3 Hz, 1H), 4.08 (ddd, J = 9.6, 1.3, 1.3 Hz, 1H), 5.12 (dd, J = 1.3, 1.3 Hz, 1H), 5.98 (dd, J = 1.3, 1.3 Hz, 1H). 13C NMR (176 MHz, chloroform-d) δ 18.28, 18.33, 18.56, 19.35, 29.34, 33.57, 44.03, 73.78, 83.03, 120.89, 144.18, 199.12. EI-MS [M]+ = 196.0. Anal. Calcd for C12H20O2: C, 73.43; H, 10.27. Found: C, 73.31; H, 10.24.
Trans-2-butyl-6-isopropyl-3-methylidenetetrahydro-4H-pyran-4-one (13e) (34.9 mg, 83%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 0.89–0.93 (m, 6H), 0.99 (d, J = 6.7 Hz, 3H), 1.27–1.50 (m, 5H), 1.73 (hept, J = 6.7 Hz, 1H), 1.77–1.87 (m, 1H), 2.35 (dd, J = 16.8, 11.0 Hz, 1H), 2.53 (dd, J = 16.8, 3.0 Hz, 1H), 3.64 (ddd, J = 11.0, 6.7, 3.0 Hz, 1H), 4.60 (dddd, J = 10.0, 4.7, 1.5, 1.5 Hz, 1H), 4.98–5.39 (m, 1H), 6.02 (dd, J = 1.5, 1.5 Hz, 1H). 13C NMR (176 MHz, chloroform-d) δ 14.14, 18.28, 18.48, 22.48, 27.74, 33.24, 34.03, 43.66, 73.06, 76.34, 120.18, 145.35, 198.25. EI-MS [M]+ = 210.0. Anal. Calcd for C13H22O2: C, 74.24; H, 10.54. Found: C, 74.38; H, 10.50.
Trans-6-isopropyl-3-methylidene-2-phenyltetrahydro-4H-pyran-4-one (13f) (24.0 mg, 52%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 0.84 (d, J = 6.8 Hz, 3H), 0.92 (d, J = 6.8 Hz, 3H), 1.67–1.79 (m, 1H), 2.43 (dd, J = 16.9, 10.9 Hz, 1H), 2.51 (dd, J = 16.9, 3.2 Hz, 1H), 3.42 (ddd, J = 10.9, 6.6, 3.2 Hz, 1H), 5.27 (dd, J = 1.3, 1.3 Hz, 1H), 5.81 (s, 1H), 6.31 (dd, J = 1.3, 1.3 Hz, 1H), 7.28–7.38 (m, 5H). 13C NMR (176 MHz, chloroform-d) δ 18.10, 18.25, 33.07, 43.75, 73.83, 78.36, 122.97, 128.06 (2 × C), 128.18, 128.58 (2 × C), 139.49, 142.24, 198.08. EI-MS [M]+ = 230.0. Anal. Calcd for C15H18O2: C, 78.23; H, 7.88. Found: C, 78.39; H, 7.87.
Trans-2-isopropyl-3-methylidene-6-phenyltetrahydro-4H-pyran-4-one (13g) (22.1 mg, 48%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 0.94 (d, J = 6.6 Hz, 3H), 1.04 (d, J = 6.6 Hz, 3H), 1.99–2.10 (m, 1H), 2.76 (dd, J = 16.7, 9.8 Hz, 1H), 2.85 (dd, J = 16.7, 3.9 Hz, 1H), 4.22 (ddd, J = 9.0, 1.3, 1.3 Hz, 1H), 5.10 (dd, J = 9.8, 3.9 Hz, 1H), 5.20 (dd, J = 1.3, 1.3 Hz, 1H), 6.08 (dd, J = 1.3, 1.3 Hz, 1H), 7.28–7.34 (m, 1H), 7.35–7.41 (m, 4H). 13C NMR (176 MHz, chloroform-d) δ 17.97, 19.16, 29.70, 47.04, 70.92, 82.59, 121.33, 126.05 (2 × C), 128.03, 128.67 (2 × C), 140.65, 143.69, 197.78. EI-MS [M]+ = 230.0. Anal. Calcd for C15H18O2: C, 78.23; H, 7.88. Found: C, 78.21; H, 7.90.
Trans-2-buthyl-3-methylidene-6-phenyltetrahydro-4H-pyran-4-one (13h) (37.6 mg, 77%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 0.91 (t, J = 7.2 Hz, 3H), 1.27–1.68 (m, 5H), 1.93 (dtd, J = 14.4, 9.5, 4.9 Hz, 1H), 2.76 (dd, J = 16.9, 10.0 Hz, 1H), 2.84 (dd, J = 16.9, 3.7 Hz, 1H), 4.66–4.72 (m, 1H), 5.10 (dd, J = 10.0, 3.7 Hz, 1H), 5.26 (dd, J = 1.3, 1.3 Hz, 1H), 6.11 (dd, J = 1.3, 1.3 Hz, 1H), 7.29–7.34 (m, 1H), 7.35–7.41 (m, 4H). 13C NMR (176 MHz, chloroform-d) δ 14.12, 22.52, 27.71, 34.04, 46.91, 70.46, 76.40, 120.66, 126.18 (2 × C), 128.14, 128.80 (2 × C), 140.79, 145.00, 197.23. EI-MS [M]+ = 244.0. Anal. Calcd for C16H20O2: C, 78.65; H, 8.25. Found: C, 78.33; H, 8.28.
Trans-3-methylidene-2,6-diphenyltetrahydro-4H-pyran-4-one (13i) (31.7 mg, 60%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 2.85 (d, J = 7.0 Hz, 2H), 4.90 (dd, J = 7.0 Hz, 1H), 5.32 (s, 1H), 5.88 (s, 1H), 6.38 (s, 1H), 7.28–7.42 (m, 10H). 13C NMR (176 MHz, chloroform-d) δ 47.04, 71.29, 78.66, 123.59, 126.21 (2 × C), 128.06 (2 × C), 128.22, 128.48, 128.81 (2 × C), 128.84 (2 × C), 139.15, 140.53, 142.38, 197.05. EI-MS [M]+ = 264.0. Anal. Calcd for C18H16O2: C, 81.79; H, 6.10. Found: C, 81.92; H, 6.12.
Trans-2-isopropyl-6-(4-methoxyphenyl)-3-methylidenetetrahydro-4H-pyran-4-one (13j) (21.3 mg, 41%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 0.94 (d, J = 6.6 Hz, 3H), 1.03 (d, J = 6.6 Hz, 3H), 2.04 (dhept, J = 8.9, 6.6 Hz, 1H), 2.77 (dd, J = 16.7, 9.1 Hz, 1H), 2.81 (dd, J = 16.7, 4.3 Hz, 1H), 3.81 (s, 3H), 4.18 (ddd, J = 8.9, 1.3, 1.3 Hz, 1H), 5.05 (dd, J = 9.1, 4.3 Hz, 1H), 5.18 (dd, J = 1.3, 1.3 Hz, 1H), 6.07 (d, J = 1.3, 1.3 Hz, 1H), 6.87–6.96 (m, 2H), 7.27–7.35 (m, 2H).13C NMR (176 MHz, chloroform-d) δ 18.06, 19.31, 29.91, 47.05, 55.46, 70.80, 82.46, 114.20 (2 × C), 121.34, 127.64 (2 × C), 132.82, 143.94, 159.55, 198.15. EI-MS [M]+ = 260.0. Anal. Calcd for C16H20O3: C, 73.82; H, 7.74. Found: C, 73.60; H, 7.75.
Trans-2-butyl-6-(4-methoxyphenyl)-3-methylidenetetrahydro-4H-pyran-4-one (13k) (45.0 mg, 82%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 0.91 (t, J = 7.2 Hz, 3H), 1.29–1.51 (m, 4H), 1.59 (ddt, J = 14.1, 10.6, 5.3 Hz, 1H), 1.91 (dtd, J = 14.3, 9.6, 4.8 Hz, 1H), 2.77 (dd, J = 16.9, 9.3 Hz, 1H), 2.81 (dd, J = 16.9, 4.2 Hz, 1H), 3.81 (s, 3H), 4.64 (dddd, J = 9.4, 5.2, 1.5, 1.5 Hz, 1H), 5.05 (dd, J = 9.3, 4.2 Hz, 1H), 5.24 (dd, J = 1.3 Hz, 1.3 Hz, 1H), 6.10 (dd, J = 1.3 Hz, 1.3 Hz, 1H), 6.89–6.92 (m, 2H), 7.28–7.31 (m, 2H). 13C NMR (176 MHz, chloroform-d) δ 14.12, 22.52, 27.70, 34.05, 46.75, 55.43, 70.22, 76.13, 114.17 (2 × C), 120.53, 127.63 (2 × C), 132.79, 145.10, 159.52, 197.45. EI-MS [M]+ = 274.0. Anal. Calcd for C17H22O3: C, 74.42; H, 8.08. Found: C, 74.11; H, 8.11.
Trans-6-(4-methoxyphenyl)-3-methylidene-2-phenyltetrahydro-4H-pyran-4-one (13l) (29.4 mg, 50%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 2.82 (dd, J = 16.9, 4.1 Hz, 1H), 2.86 (dd, J = 17.0, 9.7 Hz, 1H), 3.80 (s, 3H), 4.86 (dd, J = 9.7, 4.0 Hz, 1H), 5.29 (dd, J = 1.3, 1.3 Hz, 1H), 5.82 (dd, J = 1.3, 1.3 Hz, 1H), 6.36 (dd, J = 1.3, 1.3 Hz, 1H), 6.83–6.96 (m, 2H), 7.22–7.25 (m, 2H), 7.31–7.36 (m, 1H), 7.37–7.40 (m, 4H). 13C NMR (176 MHz, chloroform-d) δ 46.86, 55.46, 71.07, 78.48, 114.19 (2 × C), 123.51, 127.70 (2 × C), 128.06 (2 × C), 128.43, 128.82 (2 × C), 132.49, 139.30, 142.53, 159.59, 197.27. EI-MS [M]+ = 294.0. Anal. Calcd for C19H18O3: C, 77.53; H, 6.16. Found: C, 77.61; H, 6.15.
Trans-6-ferrocenyl-2-isopropyl-3-methylidenetetrahydro-4H-pyran-4-one (13m) (21.6 mg, 32%) Orange solid mp 136–138 °C. 1H NMR (700 MHz, chloroform-d) δ 0.92 (d, J = 6.7 Hz, 3H), 1.01 (d, J = 6.7 Hz, 3H), 1.99 (dp, J = 7.4, 6.6 Hz, 1H), 2.85 (dd, J = 16.9, 7.6 Hz, 1H), 2.95 (dd, J = 16.9, 4.8 Hz, 1H), 4.08–4.10 (m, 1H), 4.11 (dtd, J = 2.6, 1.3, 0.4 Hz, 1H), 4.15 (s, 5H), 4.16–4.18 (m, 1H), 4.21 (td, J = 2.5, 1.3 Hz, 1H), 4.28 (dt, J = 2.6, 1.3 Hz, 1H), 5.01 (dd, J = 7.6, 4.8 Hz, 1H), 5.17 (dd, J = 1.4, 1.4 Hz, 1H), 6.11 (dd, J = 1.4, 1.4 Hz, 1H). 13C NMR (176 MHz, chloroform-d) δ 17.41, 19.38, 30.33, 45.07, 66.65, 68.27, 68.40, 68.54, 68.87 (5 × C), 68.97, 80.51, 87.14, 121.17, 144.18, 198.26. EI-MS [M]+ = 338.0. Anal. Calcd for C19H22FeO2: C, 67.47; H, 6.56. Found: C, 67.66; H, 6.53.
Trans-2-buthyl-6-ferrocenyl-3-methylidenetetrahydro-4H-pyran-4-one (13n) (42.3 mg, 60%) Orange oil. 1H NMR (700 MHz, chloroform-d) δ 0.91 (t, J = 7.2 Hz, 3H), 1.19–1.51 (m, 4H), 1.60 (dddd, J = 14.0, 10.3, 5.5, 4.6 Hz, 1H), 1.79 (dtd, J = 14.1, 9.5, 4.6 Hz, 1H), 2.85 (dd, J = 17.0, 7.4 Hz, 1H), 2.94 (dd, J = 17.1, 4.9 Hz, 1H), 4.10 (dtd, J = 2.5, 1.3, 0.4 Hz, 1H), 4.15 (s, 5H), 4.16–4.18 (m, 1H), 4.21 (td, J = 2.5, 1.2 Hz, 1H), 4.27 (dt, J = 2.6, 1.3 Hz, 1H), 4.41–4.49 (m, 1H), 5.01 (dd, J = 7.4, 4.9 Hz, 1H), 5.21 (dd, J = 1.7, 1.1 Hz, 1H), 6.10 (dd, J = 1.8, 1.1 Hz, 1H). 13C NMR (176 MHz, chloroform-d) δ 14.17, 22.55, 27.52, 33.90, 44.96, 66.67, 68.26, 68.29, 68.40, 68.85 (5 × C), 68.94, 74.25, 87.11, 120.28, 145.40, 197.59. EI-MS [M]+ = 352.0. Anal. Calcd for C20H24FeO2: C, 68.19; H, 6.87. Found: C, 67.98; H, 6.90.
Trans-6-ferrocenyl-3-methylidene-2-phenyltetrahydro-4H-pyran-4-one (13o) (38.7 mg, 52%) Orange oil. 1H NMR (700 MHz, chloroform-d) δ 2.92 (dd, J = 17.1, 7.8 Hz, 1H), 2.97 (dd, J = 17.1, 4.8 Hz, 1H), 4.10 (s, 5H), 4.12–4.15 (m, 1H), 4.19 (td, J = 2.4, 1.3 Hz, 2H), 4.20–4.24 (m, 1H), 4.88 (dd, J = 7.7, 4.8 Hz, 1H), 5.14 (dd, J = 1.4, 1.4 Hz, 1H), 5.61 (dd, J = 1.6, 1.6 Hz, 1H), 6.32 (dd, J = 1.4, 1.4 Hz, 1H), 7.29–7.46 (m, 5H). 13C NMR (176 MHz, chloroform-d) δ 45.55, 66.51, 68.20, 68.35, 68.88 (6 × C), 69.14, 77.41, 87.00, 123.60, 128.14 (2 × C), 128.43, 128.74 (2 × C), 139.65, 143.27, 197.19. EI-MS [M]+ = 372.0. Anal. Calcd for C22H20FeO2: C, 70.99; H, 5.42. Found: C, 71.21; H, 5.41.
2-(4-Methoxyphenyl)-5-methylidenetetrahydro-4H-pyran-4-one (15d) (19.6 mg, 45%). Colorless oil. 1H NMR (700 MHz, chloroform-d) δ 2.71 (dd, J = 17.4, 11.1 Hz, 1H), 2.81 (dd, J = 17.4, 3.2 Hz, 1H), 3.81 (s, 3H), 4.57 (ddd, J = 14.5, 2.2, 2.2 Hz, 1H), 4.79 (ddd, J = 14.5, 1.2, 1.2 Hz, 1H), 4.79 (dd, J = 11.1, 3.2 Hz, 1H), 5.32 (ddd, J = 2.2, 1.2, 1.2 Hz, 1H), 6.15 (ddd, J = 2.2, 1.2, 1.2 Hz, 1H), 6.86–6.95 (m, 2H), 7.27–7.32 (m, 2H). 13C NMR (176 MHz, chloroform-d) δ 47.50, 55.46, 69.99, 77.67, 114.22 (2 × C), 120.55, 127.27 (2 × C), 132.70, 140.84, 159.63, 196.43. EI-MS [M]+ = 218.0. Anal. Calcd for C13H14O3: C, 71.54; H, 6.47. Found: C, 71.43; H, 6.49.

2.2. Biology

Metabolic activity by MTT assay
The MTT (3-(4,5-dimethyldiazol-2-yl)-2,5-diphenyl tetrazolium bromide) assay was performed to assess the cytotoxicity of the investigated compounds. Briefly, HL-60, MCF-7 and HUVEC cells were seeded in 24-well plates at density of 8 × 104/mL. After 24 h incubation, cells were treated with various concentrations of the compounds and incubated for another 24 h or 48 h. Then, the cells were incubated with MTT solution (100 µL; 5 mg/mL in PBS) for 1.5 h and centrifuged. The formazan product was dissolved in DMSO, and the absorption was measured using FlexStation 3 Multi-Mode Microplate Reader (Molecular Devices, LLC, San Jose, CA, USA) at 560 nm. Assays were performed twice in triplicate for each analog.
Cell morphology
HL-60 cells were treated with compound 13d or 13g at IC50 and 2IC50 for 24 h. Then, the cells were examined for morphological changes using a light microscope at 40x magnification and photographed.
Apoptosis assay
Apoptotic cells were detected using FITC Annexin V Apoptosis Detection Kit I (BD Biosciences, San Jose, CA, USA). HL-60 cells (2.0 × 105 cells/mL) seeded in 6-well plates and cultured overnight were treated with various concentrations of 13d or 13g for 24 h. Then, the cells were washed with PBS and resuspended in the binding buffer. After staining with FITC Annexin V and propidium iodide (15 min, room temperature) cells were analyzed by flow cytometry using CytoFlex and Kaluza Analysis Software v2 (Beckman Coulter Inc., Brea, CA, USA).
Analysis of apoptosis, DNA damage and cell cycle
Apoptosis, DNA damage, proliferation and cell cycle kinetics were investigated using the Apoptosis, DNA Damage and Cell Proliferation Kit (BD Biosciences, San Jose, CA, USA) according to the manufacturer’s protocol. Briefly, HL-60 cells (2.0 × 105 cells/mL) were seeded in 6-well plates and left to grow for 24 h. Then, the cells were treated with 13d or 13g at IC50 and 2IC50 concentrations. After 24 h incubation, the cells were treated with BrdU solution for 8 h. Then, the cells were collected by centrifugation, fixed and permeabilized. To expose the incorporated BrdU, the cells were incubated with DNase (300 µg/mL in PBS) for 1 h at 37 °C. Afterwards, the cells were stained with fluorescent antibodies, including Anti-BrdU, Anti-Cleaved PARP (Asp214) and Anti-H2AX (pS139) for 20 min at room temperature. Total DNA for cell cycle analysis was stained using DAPI solution (1 µg/mL in staining buffer). Finally, the cells were analyzed by flow cytometry using CytoFLEX (Beckman Coulter, Inc., Brea, CA, USA) and the data analysis was performed with Kaluza Analysis Software v2 (Beckam Coulter, Inc., Brea, CA. USA).
Topoisomerase IIα activity
Topoisomerase IIα activity inhibition was assessed using the Human Topoisomerase II Decatenation Assay (Inspiralis, Norwich, UK). Briefly, kDNA was incubated with ATP, assay buffer (50 mM Tris-HCl pH 7.5, 125 mM NaCl, 10 mM MgCl2, 5 mM DTT, 100 µg/mL albumin), human topoisomerase II enzyme and various concentrations of the investigated compounds for 30 min at 37 °C. The reaction was stopped by addition of STEB (40% (w/v) sucrose, 100 mM Tris-HCl pH 8, 1 mM EDTA, 0.5 mg/mL bromophenol blue) followed by DNA extraction with chloroform/iso-amyl alcohol mixture (24:1). The samples were loaded onto a 1% agarose gel and run in a TBE buffer at 85 V for 65 min. Gels were stained with ethidium bromide (1 µg/mL) for 15 min, destained in water and visualized with a transilluminator (ChemiDoc MP Imaging System, Bio-Rad, Hercules, CA, USA).
Statistical analysis
All data are expressed as mean ±SEM. Statistical analyses were carried out using Prism 6.0 (GraphPad Software Inc., San Diego, CA, USA). Statistical significance was assessed using Student’s t-test (for comparison of two groups) or one-way ANOVA followed by a post-hoc multiple comparison Student–Newman–Keuls test (for comparison of three or more groups). *p < 0.05, **p < 0.01 and ***p < 0.001 were considered significant.

3. Results and Discussion

3.1. Chemistry

Addition of diethyl 2-oxopropylphosphonate 7 to aliphatic or aromatic aldehydes was performed using conditions described by Wada et al. [21]. Treatment of 7 [22] with 1.1 equivalents of sodium hydride at room temperature and next 1.1 equivalents of n-butyllithium solution (2,5 M in hexane) at the temperature range from –30 to –40 °C gave dianion 8, which was reacted with 1.2 equivalents of aldehyde (Scheme 1). After reaction, work-up and purification by column chromatography produced 4-hydroxy-2-oxoalkylphosphonates 9a-e in good yields (Table 1). Conversion of 9a-e into 3-diethoxyphosphoryldihydropyran-4-ones 11a-e was realized performing the reaction with N,N-dimethylformamide dimethyl acetal (DMF-DMA, 3 equivalents) followed by treatment of the intermediate dimethylaminovinylphosphonates 10a-e with 2.5 equivalents of BF3 · Et2O for 24 h, at room temperature. Surprisingly, cyclization of 9e with ferrocenyl substituent, which was obtained by reaction of dianion 8 with ferrocenecarboxaldehyde, was inefficient in these conditions. Analysis of the 31P NMR spectra of the crude reaction mixture showed a number of signals which indicated degradation of the obtained compound. However, when the reaction was performed with three equivalents of DMF-DMA for 3 h at room temperature, without addition of BF3 · Et2O, the expected 3-diethoxyphosphoryl-6-ferrocenyldihydropyran-4-one 11e was obtained. Attempts to purify 11a-e by column chromatography failed due to their decomposition. However, inspection of 1H, 31P and 13C NMR spectra of these compounds showed that their purity can be estimated as 80–95%. Therefore, crude dihydropyran-4-ones 11a-e were used in the next step.
With 3-diethoxyphosphoryldihydropyran-4-ones 11a-e in hand, we decided to introduce various substituents next to the phosphoryl group by performing conjugate addition to these compounds (Scheme 2). The reactions proceeded smoothly with Gilman reagents, Ph2CuLi and n-Bu2CuLi in the presence of TMSCl at −78 °C in THF, to give the expected 2,6-disubsttituted-3-diethoxyphosphoryltetrahydropyran-4-ones 12a–j as mixtures of two diastereoisomers, r-2-trans-3-trans-6-12a–j and r-2-cis-3-trans-6-12a–j, and enol forms, trans-enol-12a–j. Yields and isomer ratios are given in Table 2. It is worth noticing that in all cases, yields are significantly higher for the phenyl substituent than for the n-butyl one. Evidently, the phenyl Gilman reagent is a better nucleophile in these reactions than the n-butyl one.
Structures of the obtained 3-diethoxyphosphoryl-2,6-disubstitutedtetrahydropyran-4-ones 12a–j were established using 1H, 13C and 31P NMR spectra, and mass spectrometry. Analysis of NMR spectra allowed us to assign the relative configurations and preferred conformations of the obtained isomers. Figure 2 shows characteristic coupling constants for r-2-trans-3-trans-6- and r-2-cis-3-trans-6-2-n-butyl-6-ethyl-3-diethoxyphosphoryltetrahydropyran-4-one 12a taken from the NMR spectra of the mixture of both diastereoisomers. Large coupling constants JH5–H6 (10.9 Hz and 8.7 Hz) and small coupling constants JH5′–H6 (2.9 Hz and 4.0 Hz) indicate diaxial and axial-equatorial relationships of these protons, respectively, and prove the equatorial position of the ethyl group in both diastereoisomers. On the other hand, a large coupling constant JC1′(n-Bu)-P = 13.6 Hz in r-2-trans-3-trans-6-12a proves the diaxial relation of n-butyl and phosphoryl groups, whereas JC1′(n-Bu)-P ~ 0 Hz in r-2-cis-3-trans-6-12a indicates the axial-equatorial position of these groups. Signals of the enol form, trans-enol-12a, were also present, including a characteristic doublet of hydroxyl group with chemical shifts 11–12 ppm and coupling constant JH–P ~1 Hz. This coupling constant shows the existence of resonance assisted hydrogen bond (RAHB) reported recently for the related organophosphorus compounds [23]. As similar sets of coupling constants were observed for 2,6-disubstituted 3-diethoxyphosphoryltetrahydropyran-4-ones 12b–j, we believe that these assignments are valid for the whole series.
It is worth stressing that we did not observe the formation of the second pair of diastereoisomers, i.e., r-2-trans-3-cis-6- and r-2-cis-3-cis-6-12a–j or the corresponding cis-enol forms. This observation shows that the addition of Gilman reagents to 3-diethoxyphosphoryldihydropyran-4-ones 11a–e is fully diastereoselective, and as expected, proceeds exclusively via axial attack of the Gilmann reagent on the dihydropyranone ring.
Unfortunately, the introduction of an isopropyl group in position 2 by addition of the appropriate Gilman reagent, i-Pr2CuLi, to 3-diethoxyphosphoryldihydropyran-4-ones 11a–e, was inefficient. However, when we performed this addition using Grignard reagent, i-PrMgCl, the reaction proceeded smoothly and gave expected 6-substituted-3-diethoxyphosphoryl-2-isopropyltetrahydropyran-4-ones 12k–o in good yields (Scheme 3, Table 3). The only drawback was poor stereoselectivity of these reactions. In each case, a complex mixture of four diastereoisomers and two enol forms were formed. Although the axial attack was still favored, the substantial amount of r-2-trans-3-cis-6-12k–o along with their epimers, r-2-cis-3-cis-6-12k–o, and cis-enols-12k–o, were formed.
Figure 3 shows all six possible isomers of 3-diethoxyphosphoryl-2,6-diisopropylpyran-4-one 12l and characteristic coupling constants which could be read from the NMR spectra. Despite many signals present in these spectra, we were able to assign signals and consequently relative configurations and conformations for five isomers in the mixture. The sixth isomer, r-2-cis-3-trans-6-12l, was present in too small an amount (1% according to 31P NMR spectrum, Table 3) to find its characteristic signals in 1H or 13C NMR spectra. Small values of coupling constants JC1′-P for both r-2-trans-3-cis-6 and r-2-cis-3-cis-6-12l (3.5 Hz and 3.9 Hz, respectively) indicate a cis relationship between phosphoryl and isopropyl groups. In turn, small coupling constant JH2-P in r-2-trans-3-cis-6-12l (11.8 Hz) and big coupling constant JH2-P in r-2-cis-3-cis-6-12l (39.3 Hz) prove axial/equatorial and diaxial relationships between the H2 proton and phosphoryl group in these diastereoisomers, respectively. Similarities in the NMR spectra of 12l and remaining 6-substituted 2-isopropyl-3-diethoxyphosphoryltetrahydropyran-4-ones 12k–o allow us to propose the same stereochemical relationships in all pyranones 12k–o.
Finally, the 2,6-disubstituted 3-diethoxyphosphoryltetrahydropyran-4-ones 12a–o were converted to 3-methylidenetetrahydropyran-4-ones 13a–o by performing Horner–Wadsworth–Emmons olefination. The reaction was carried out with 2 equivalents of K2CO3 and 10 equivalents of formaline at 0 °C in THF as a solvent (Scheme 4). 2-n-Butyl- and 2-phenyl-3-methylidenetetrahydropyran-4-ones 13a–j were formed as single trans-diastereoisomers, whereas 2-isopropyl-3-methylidenetetrahydropyran-4-ones 12k–o were formed as mixtures of trans- and cis-diastereoisomers. However, diastereoisomers cis-16k–o were unstable and decomposed during column chromatography. Therefore, only trans-2,6-disubstituted 3-methylidenetetrahydropyran-4-ones 13a–o were purified by column chromatography and obtained in moderate to very good yields (Table 4). Interestingly, all 2-n-butyl-3-methylidenetetrahydropyran-4-ones 13b,e,h,k,n were obtained in significantly higher yields than 3-methylidenetetrahydropyran-4-ones with an isopropyl or phenyl group in position 2. It is very likely that greater steric hindrance of isopropyl or phenyl groups, in comparison to n-butyl, impedes HWE olefination of formaldehyde.
To broaden the substitution pattern of methylidenetetrahydropyran-4-ones, we subjected 3-diethoxyphosphoryldihydropyran-4-ones 11a–d to a double bond reduction. A solution of lithium tri-sec-butylborohydride (L-Selectride®, 1M in THF) was used as a reducing agent, and the reaction was carried out in an inert atmosphere in the presence of 1.1 equivalents of reductant, with THF as a solvent at temperature range from −78 to 0 °C (Scheme 5). After reaction work-up and purification by column chromatography, corresponding 2-substituted 5-diethoxyphosphoryltetrahydropyran-4-ones 14a–d were obtained with very good yields as mixtures of trans- and cis-diastereoisomers and enol form. (Table 5).
Relative configurations of the obtained compounds were assigned by the analysis of 1H and 13C NMR spectra, which showed that in all cases trans form dominated and the R1 group was in equatorial position. Characteristic coupling constants for 14c are shown in Figure 4. Coupling constants of H-2 proton with vicinal protons H-3 and H-3′ (JH2-H3 = 10.8 or 11.6 Hz and JH2-H3′ = 2.9 or 3.4 Hz) indicate the equatorial positions of ethyl substituents in both diastereoisomers. On the other hand, the coupling constant of the H-5 proton with proton H-6 in trans-diastereoisomer (JH5-H6 = 10.8 Hz) indicate equatorial positions of the diethoxyphosphoryl group. In cis-diastereoisomer, the same coupling constant was significantly smaller (JH5-H6 = 4.2 Hz). Moreover, the coupling constant of H-6 proton with diethoxyphosphoryl group was very high (JH6-P = 34.7 Hz), which indicates the axial position of diethoxyphosphoryl group. Similar sets of coupling constants were observed in other 2-substituted 5-diethoxyphosphoryltetrahydropyran-4-ones 14.
Next, we made an attempt to transform the obtained 2-substituted-5-diethoxyphosphoryltetrahydropyran-4-ones 14a-d into target 2-substituted-5-methylidenetetrahydropyran-4-ones 15a-d, using Horner–Wadsworth–Emmons olefination. The reaction was carried out with 2 equivalents of K2CO3, 10 equivalents of formalin, THF as a solvent and room temperature for 3 h. Unfortunately, the obtained 3-methylidenetetrahydropyran-4-ones 15a-c,e, turned out to be unstable at room temperature and decomposed during column chromatography. Only 2-(4-methoxyphenyl)-5-methylidenetetrahydropyran-4-one 15d proved to be stable for a prolonged period of time (at least 7 days at room temperature) and after purification by column chromatography was obtained in 45% yield (Scheme 6).

3.2. Biology

3.2.1. In Vitro Cytotoxicity

The cytotoxic activity of the obtained methylidenetetrahydropyran-4-ones was investigated in two cancer cell lines, human promyelocytic leukemia (HL-60) and breast cancer adenocarcinoma (MCF-7), and in healthy umbilical vein endothelial cells (HUVEC), for comparison. An anticancer drug, carboplatin, was used as a reference compound. Metabolic activity of the cells was assessed after 48 h incubation with various concentrations of the analogs (Table 6). Purity of all final compounds was determined on the basis of their 1H NMR spectra [24] (signal-to-noise ratio and presence of unidentified signals) and was at least 98%, in most cases even over 99%.
Generally, the tested compounds were more cytotoxic for leukemic than for breast cancer cells. Analysis of the structure–activity relationship revealed that the cytotoxicity of methylidenetetrahydropyran-4-ones 13a–o strongly depends on the nature of the substituent in position 2 (R2), and to a much lesser extent, on the structure of the substituent in position 6 (R1). This observation can be explained by the fact that R2 substituent is much closer to the methylidene group, which is believed to be the reaction site for various bionucleophiles, as mentioned in the Introduction. The best substituent in position 6 seems to be the isopropyl group; however, the differences in cytotoxic activity among other substituents in this position are not very significant. On the other hand, the presence and nature of the substituent next to the methylidene moiety is crucial for the activity. Compound 15d, with no substituent in this position, is practically inactive. Evidently, the isopropyl group is the best substituent for position 2 in both tested cell lines. Almost all methylidenetetrahydropyran-4-ones 13 (with the exception of 13a) bearing isopropyl in this position are highly cytotoxic, with IC50 values below 10 μM. The n-butyl or phenyl groups contribute much less to the cytotoxic activity of the analogs.
The most cytotoxic analogs were 13g, 13j and 13m, with IC50 values below 4 µM against the HL-60 cell line. However, these compounds showed low cancer/healthy cell IC50 ratios, and therefore, poor selectivity for HL-60 cells compared to normal HUVEC cells. Analog 13d showed high cytotoxicity in HL-60 cells (IC50 = 7.65 ± 0.05 µM) and was the most selective one, exhibiting over 3-fold higher inhibitory activity in HL-60 than in HUVEC cells.
Analogs 13d and 13g were selected for further assessment of their anticancer properties. The cytotoxicity of these two compounds in HL-60 cells was evaluated after 24 h incubation. The IC50 values were 15.65 ± 0.61 µM (Figure 5B) and 6.49 ± 0.31 µM (Figure 6B), respectively, and they were used in the subsequent assays.

3.2.2. Antiproliferative Activity

To examine the influences of the investigated compounds on HL-60 cell proliferation, cells were exposed to a thymidine analog, bromodeoxyuridine (BrdU), that incorporates into newly synthesized DNA during proliferation. The results indicate that both analogs significantly inhibited cell proliferation. Treatment with 13d at IC50 and 2IC50 concentrations decreased the BrdU-incorporating cell populations to 31.5 ± 1.65% and 26.4 ± 6.9%, respectively (Figure 5F), whereas 13g at IC50 and 2IC50 concentrations decreased the populations to 35.3 ± 1.23% and 31.9 ± 6.02%, respectively (Figure 6F).
The cell cycle has a critical role in regulation of cell growth, proliferation and division of cells after DNA damage. Therefore, it often serves as a target in cancer therapy development [25]. The cell cycle distribution was investigated using DAPI DNA staining and flow cytometry analysis. 13d at 2IC50 concentration significantly raised the number of cells in the G2/M phase (24.1 ± 1.40% of cells compared to control 14.9 ± 1.37%), whereas the population of cells in G0/G1 phase was diminished (Figure 5D,E). 13g caused a slight increase in the G2/M cell population; however, the change was not statistically significant (Figure 6D,E).

3.2.3. Cell Morphology Changes

HL-60 cells grow in suspension and usually have a roughly spherical shape. Treatment with 13d or 13g for 24 h induced several morphological changes in the cells. Some became elongated and asymmetrical; others exhibited characteristics typical for apoptotic cells, including shrinkage, membrane blebbing and formation of apoptotic bodies (Figure 5A and Figure 6A).

3.2.4. Induction of Apoptosis

One of the typical features of apoptotic cell death is translocation of phosphatidylserine from the inside surface to the extracellular surface of the cell membrane [26], which can be detected with Annexin V [27]. Therefore, to examine the ability of the investigated compounds to induce apoptosis in HL-60 cells, we performed Annexin V and propidium iodide (PI) staining (Figure 5C and Figure 6C). In the control group, 90.3 ± 1.15% of HL-60 cells were healthy, 3.3 ± 0.36% were in early stage of apoptosis (Annexin V-positive and PI-negative) and 5.5 ± 1.02% were in late-stage apoptosis (Annaxin V-positive and PI-positive). 13d did not induce a significant change in the number of apoptotic cells at IC50; however, at 2IC50, the analog caused increases in both early and late apoptotic cell numbers to 13.0 ± 2.64% and 24.6 ± 2.34%, respectively. Similarly, 13g was not effective at IC50, but at 2IC50, early-stage apoptosis was detected in 8.7 ± 0.57% of cells, and 39.5 ± 1.16% of cells were in the late stage apoptosis.
Another hallmark of apoptosis is cleavage of poly(ADP-ribose) polymerase 1 (PARP1) into 24 kDa and 89 kDa fragments executed by caspases 3 and 7. PARP 1 fragmentation causes its inactivation, which in turn leads to inhibition of DNA repair and promotes caspase-mediated fragmentation of DNA during programmed cell death [28]. Therefore, detection of 89 kDa fragments of cleaved PARP with Anti-Cleaved PARP (Asp214) antibodies and flow cytometry was used to investigate apoptosis. Incubation of HL-60 cells with 13d only at 2IC50 concentration led to significant increase in population of cells with cleaved PARP (49.0 ± 1.61%), in comparison to the control (0.7 ± 0.13%) (Figure 5G). Treatment with 13g also caused a significant PARP cleavage in 13.5 ± 3.6% of the cell population only at 2IC50 concentration (Figure 6G).

3.2.5. Induction of DNA Damage

DNA damage can be the result of either single- or double-strand breaks (DSBs). DSB are considered the most dangerous, as it is assumed that in metazoan, just one DSB has the potential to trigger cell death [29]. Cancer cells are often more susceptible to DSBs than healthy cells. Therefore, drugs able to induce DSB formation are widely used in anticancer therapies [30]. One of the earliest signals for DSB detection include phosphorylation of the histone H2AX on the Ser139, generating γH2AX [31]. In our study, induction of DSB was investigated by flow cytometry of HL-60 cells stained with anti-γH2AX fluorescent antibody. Treatment of HL-60 cells with 13d at 2IC50 led to significant γH2AX generation in 38.4 ± 5.89% of cells in comparison to the control (1.1 ± 0.17%) [Figure 5H]. 13g was markedly less genotoxic, causing at 2IC50, H2AX phosphorylation in 9.6 ± 2.94% of the cell population (Figure 6H).

3.2.6. Topoisomerase IIα Activity

Some of the most effective DNA-damaging anticancer drugs include inhibitors of topoisomerases [32]. DNA topoisomerases are enzymes capable of changing DNA topology by generating transient breaks to mitigate strain. There are two types of topoisomerases, namely, type I enzymes cleaving one of the DNA strands and type II enzymes creating DSB [33]. Type II topoisomerases separate into subfamilies α and β. Topoisomerase IIβ is evenly distributed in all cells, both normal and cancerous, whereas topoisomerase IIα is overexpressed in rapidly proliferating cells, what makes it an optimal target for anticancer therapy [34]. Therefore, the inhibitory activity of the investigated analogs on topoisomerase IIα was also explored.
The ability of the analogs to inhibit topoisomerase II activity was investigated by measuring decatenation of kinetoplast DNA (kDNA) by human topoisomerase IIα. Topoisomerase II has the ability to decatenate interlinked dsDNA minicircles. When the enzyme is inactive, the kDNA molecules remain in the well after gel electrophoresis, whereas when topoisomerase II is active, the minicircles are released and migrate into the gel. Both compounds inhibited topoisomerase II activity at 0.5 mM and 1 mM concentrations. 13g showed stronger inhibitory activity towards topoisomerase IIα than 13d (Figure 7).

4. Conclusions

In the present investigation, we showed that Horner–Wadsworth–Emmons methodology is a very useful tool for the efficient synthesis of 2,6-disubstituted 3-methylidenetetrahydropyran-4-ones 13a–o and 2-(4-methoxyphenyl)-5-methylidenetetrahydropyran-4-one 15d. Addition of Gilman or Grignard reagents to intermediate diethoxyphosphoryltetrahydropyran-4-ones 13a-e was fully or highly stereoselective, respectively, and finally gave access to diastereomerically pure trans-2,6-disubstituted 3-methylidenetetrahydropyran-4-ones 13a–o with various substitution patterns. It is worth stressing that, in spite of very complex diastereoismeric mixtures of the obtained adducts, careful NMR analysis allowed us to establish their relative configurations and conformations. The in vitro assessment of the obtained analogs using the conventional MTT assay showed that in HL-60 and MCF-7 cells, analogs with isopropyl and phenyl substituents in position R2 were more cytotoxic than those with n-butyl substituents. Among the novel compounds, 13g was the most cytotoxic and 13d the most selective, as compared with normal HUVEC cells. Therefore, the anticancer potential of 13d and 13g in HL-60 cells was further investigated. Our findings showed that both analogs induced morphological changes characteristic of apoptosis in cancer cells, significantly inhibited proliferation and induced apoptotic cell death. Moreover, both compounds generated DNA damage, and analog 13d arrested the cell cycle of HL-60 cells in the G2/M phase. It has been reported that topoisomerase II inhibitory activity is associated with generation of DNA damage and induction of cell cycle arrest at the G2/M phase [35]. Both 13d and 13g were able to inhibit the activity of topoisomerase IIα; however, 13g was more potent. Based on these findings, the investigated analogs may be further optimized for the development of new and effective topoisomerase II inhibitors.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma15093030/s1. General procedures and characterization data for diethyl 4-hydroxy-2-oxoalkylphosphonates 9ae, 3-diethoxyphosphoryltetrahydropyran-4-ones 12ao and 2-alkyl(aryl)-5-diethoxyphosphoryltetrahydropyran-4-ones 14ad. Copies of 1H, 13C and 31P NMR spectra of all obtained compounds.

Author Contributions

Conceptualization, T.J. and A.J.; methodology, T.J., A.J., J.K., J.S. and K.G.-J.; software, T.B. and J.D.-S.; validation, J.K. and K.G.-J.; formal analysis, T.B., J.D.-S. and B.P.-S.; investigation, T.B. and J.D.-S.; resources, T.B., J.D.-S.; data curation, T.B. and J.D.-S.; writing—original draft preparation, T.B. and J.D.-S.; writing—review and editing, T.J., A.J.; visualization, T.B., J.D.-S.; supervision, A.J. and T.J.; project administration, J.K. and K.G.-J.; funding acquisition, T.J., A.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the grants from the Medical University of Lodz, number 503/1-156-02/503-11-001-19-00, and from Lodz University of Technology, number 501-3-32-1-2.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Pettit, G.R.; Xu, J.-P.; Chapuis, J.-C.; Pettit, R.K.; Tackett, L.P.; Doubek, D.L.; Hooper, J.N.A.; Schmidt, J.M. Antineoplastic Agents. 520. Isolation and Structure of Irciniastatins A and B from the Indo-Pacific Marine Sponge Ircinia ramosa. J. Med. Chem. 2004, 47, 1149–1152. [Google Scholar] [CrossRef] [PubMed]
  2. Kito, K.; Ookura, R.; Yoshida, S.; Namikoshi, M.; Ooi, T.; Kusumi, T. New Cytotoxic 14-Membered Macrolides from Marine-Derived Fungus Aspergillus ostianus. Org. Lett. 2008, 10, 225–228. [Google Scholar] [CrossRef] [PubMed]
  3. Janecki, T. Targets in Heterocyclic Systems; Attanasi, O.A., Spinelli, D., Eds.; Italian Society of Chemistry: Rome, 2006; Volume 10, pp. 301–320. [Google Scholar]
  4. Feray, L.; Bertrand, M.P. Targets in Heterocyclic Systems; Attanasi, O.A., Spinelli, D., Eds.; Italian Society of Chemistry: Rome, 2008; Volume 12, pp. 31–58. [Google Scholar]
  5. Kitson, R.R.A.; Millemaggi, A.; Taylor, R.J.K. The Renaissance of α-Methylene-γ-butyrolactones: New Synthetic Approaches. Angew. Chem. Int. Ed. 2009, 48, 9426–9451. [Google Scholar] [CrossRef] [PubMed]
  6. Elford, T.G.; Hall, D.G. Advances in 2-(Alkoxycarbonyl)allylboration of Carbonyl Compounds and Other Direct Methods for the Preparation of α-Exo-Alkylidene γ-Lactones. Synthesis 2010, 2010, 893–907. [Google Scholar] [CrossRef]
  7. Jiang, Z.-Y.; Zhou, J.; Huang, C.-G.; Hu, Q.-F.; Huang, X.-Z.; Wang, W.; Zhang, L.-Z.; Li, G.-P.; Xia, F.-T. Two novel antiviral terpenoids from the cultured Perovskia atriplicifolia. Tetrahedron 2015, 71, 3844–3849. [Google Scholar] [CrossRef]
  8. Imai, H.; Suzuki, K.-I.; Morioka, M.; Numasaki, Y.; Kadota, S.; Nagai, K.; Sato, T.; Iwanami, M.; Saito, T. Okilactomycin, a novel antibiotic produced by a Streptomyces species. J. Antibiot. 1987, 40, 1475–1482. [Google Scholar] [CrossRef] [Green Version]
  9. Bartosik, T.; Kędzia, J.; Drogosz-Stachowicz, J.; Janecka, A.; Krajewska, U.; Mirowski, M.; Janecki, T. Synthesis of 2,2,6-Trisubstituted 5-Methylidenetetrahydropyran-4-ones with Anticancer Activity. Molecules 2020, 25, 611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Fouda, M.F.R.; Abd-Elzaher, M.M.; Abdelsamaia, R.A.; Labib, A.A. On the medicinal chemistry of ferrocene. Appl. Organomet. Chem. 2007, 21, 613–625. [Google Scholar] [CrossRef]
  11. Ornelas, C. Application of ferrocene and its derivatives in cancer research. New J. Chem. 2011, 35, 1973–1985. [Google Scholar] [CrossRef]
  12. Gasser, G.; Metzler-Nolte, N. The potential of organometallic complexes in medicinal chemistry. Curr. Opin. Chem. Biol. 2012, 16, 84–91. [Google Scholar] [CrossRef] [Green Version]
  13. Lal, B.; Badshah, A.; Altaf, A.A.; Khan, N.; Ullah, S. Miscellaneous applications of ferrocene-based peptides/amides. Appl. Organomet. Chem. 2011, 25, 843–855. [Google Scholar] [CrossRef]
  14. Hillard, E.; Vessieres, A.; Thouin, L.; Jaouen, G.; Amatore, C. Ferrocene-Mediated Proton-Coupled Electron Transfer in a Series of Ferrocifen-Type Breast-Cancer Drug Candidates. Angew. Chem. Int. Ed. 2006, 45, 285–290. [Google Scholar] [CrossRef] [PubMed]
  15. James, P.; Neudorfl, J.; Eissmann, M.; Jesse, P.; Prokop, A.; Schmalz, H.G. Enantioselective Synthesis of Ferrocenyl Nucleoside Analogues with Apoptosis-Inducing Activity. Org. Lett. 2006, 8, 2763–2766. [Google Scholar] [CrossRef] [PubMed]
  16. Jaouen, G.; Top, S.; Vessieres, A.; Leclercq, G.; Mc, J.M. The First Organometallic Selective Estrogen Receptor Modulators (SERMs) and Their Relevance to Breast Cancer. Curr. Med. Chem. 2004, 11, 2505–2517. [Google Scholar] [CrossRef] [PubMed]
  17. Forgues, S.F.; Nicot, B.D. Ferrocene and ferrocenyl derivatives in luminescent systems. J. Photochem. Photobiol. A Chem. 2000, 132, 137–159. [Google Scholar] [CrossRef]
  18. Nguyen, P.; Elipe, P.G.; Manners, I. Organometallic Polymers with Transition Metals in the Main Chain. Chem. Rev. 1999, 99, 1515–1548. [Google Scholar] [CrossRef]
  19. Biot, C.; Glorian, G.; Maciejewski, L.A.; Brocard, J.S.; Domarle, O.; Blampain, G.; Millet, P.; Georges, A.J.; Abessolo, H.; Dive, D.; et al. Synthesis and Antimalarial Activity in Vitro and in Vivo of a New Ferrocene-Chloroquine Analogue. J. Med. Chem. 1997, 40, 3715–3718. [Google Scholar] [CrossRef]
  20. Nguyen, A.; Vessieres, A.; Hillard, E.A.; Top, S.; Pigeon, P.; Jaouen, J. Ferrocifens and Ferrocifenols as New Potential Weapons against Breast Cancer. Chimia 2007, 61, 716–724. [Google Scholar] [CrossRef] [Green Version]
  21. Wada, E.; Kanemasa, S.; Tsuge, O. New Synthesis of 2-Oxo-3-alkenylphosphonates and Hetero Diels–Alder Reactions with Vinyl Ethers Leading to 5-Substituted 2-Phosphinyl-2-cyclohexen-1-ones. Bull. Chem. Soc. Jpn. 1989, 62, 860–868. [Google Scholar] [CrossRef]
  22. Diethyl 2-oxopropylphosphonate 7 was purchased from Merck®.
  23. Koszuk, J.; Bartosik, T.; Wojciechowski, J.; Wolf, W.M.; Janecka, A.; Drogosz, J.; Długosz, A.; Krajewska, U.; Mirowski, M.; Janecki, T. Synthesis of 3-Methylidene-1-tosyl-2,3-dihydroquinolin-4-(1H)-ones as Potent Cytotoxic Agents. Chem. Biodivers. 2018, 15, e1800242. [Google Scholar] [CrossRef]
  24. Hou, T.; Raftery, D. Analysis and Purification Methods in Combinatorial Chemistry; Yan, B., Ed.; John Wiley & Sons: Hoboken, NJ, USA, 2004; Volume 163, pp. 137–173. [Google Scholar]
  25. Schwartz, G.K.; Shah, M.A. Targeting the cell cycle: A new approach to cancer therapy. J. Clin. Oncol. 2005, 23, 9408–9421. [Google Scholar] [CrossRef] [PubMed]
  26. Tyurina, Y.Y.; Shvedova, A.A.; Kawai, K.; Tyurin, V.A.; Kommineni, C.; Quinn, P.J.; Kagan, V.E. Phospholipid signaling in apoptosis: Peroxidation and externalization of phosphatidylserine. Toxicology 2000, 148, 93–101. [Google Scholar] [CrossRef]
  27. Van Engeland, M.; Nieland, L.J.; Ramaekers, F.C.; Schutte, B.; Reutelingsperger, C.P. Annexin V-affinity assay: A review on an apoptosis detection system based on phosphatidylserine exposure. Cytom. J. Int. Soc. Anal. Cytol. 1998, 31, 1–9. [Google Scholar] [CrossRef]
  28. Mashimo, M.; Onishi, M.; Uno, A.; Tanimichi, A.; Nobeyama, A.; Mori, M.; Fujii, T. The 89-kDa PARP1 cleavage fragment serves as a cytoplasmic PAR carrier to induce AIF-mediated apoptosis. J. Biol. Chem. 2021, 296, 100046. [Google Scholar] [CrossRef]
  29. Jackson, S.P. Sensing and repairing DNA double-strand breaks. Carcinogenesis 2002, 23, 687–696. [Google Scholar] [CrossRef]
  30. Aparicio, T.; Baer, R.; Gautier, J. DNA double-strand break repair pathway choice and cancer. DNA Repair 2014, 19, 169–175. [Google Scholar] [CrossRef] [Green Version]
  31. Stope, M.B. Phosphorylation of histone H2A. X as a DNA-associated biomarker. World Acad. Sci. J. 2021, 3, 31. [Google Scholar] [CrossRef]
  32. O’Connor, M.J. Targeting the DNA damage response in cancer. Mol. Cell 2015, 60, 547–560. [Google Scholar] [CrossRef] [Green Version]
  33. Dehshahri, A.; Ashrafizadeh, M.; Afshar, E.G.; Pardakhty, A.; Mandegary, A.; Mohammadinejad, R.; Sethi, G. Topoisomerase inhibitors: Pharmacology and emerging nanoscale delivery systems. Pharmacol. Res. 2020, 151, 104551. [Google Scholar] [CrossRef]
  34. Marinello, J.; Delcuratolo, M.; Capranico, G. Anthracyclines as topoisomerase II poisons: From early studies to new perspectives. Int. J. Mol. Sci. 2018, 19, 3480. [Google Scholar] [CrossRef] [Green Version]
  35. Larsen, A.K.; Escargeuil, A.E.; Skladanowski, A. From DNA damage to G~ 2 arrest: The many roles of topoisomerase II. Prog. Cell Cycle Res. 2003, 5, 295–300. [Google Scholar] [PubMed]
Figure 1. Compounds containing tetrahydropyrane ring with anti-HBV or cytotoxic activity.
Figure 1. Compounds containing tetrahydropyrane ring with anti-HBV or cytotoxic activity.
Materials 15 03030 g001
Scheme 1. Synthesis of diethyl 4-hydroxy-2-oxoalkylphosphonates 9a-e and 3-diethoxyphosphoryldihydropyran-4-ones 11a-e.
Scheme 1. Synthesis of diethyl 4-hydroxy-2-oxoalkylphosphonates 9a-e and 3-diethoxyphosphoryldihydropyran-4-ones 11a-e.
Materials 15 03030 sch001
Scheme 2. Synthesis of 6-substituted 2-n-butyl- or 2-phenyl-3-diethoxyphosphoryltetrahydropyran-4-ones 12a–j using Gillman reagents, as Michael donors.
Scheme 2. Synthesis of 6-substituted 2-n-butyl- or 2-phenyl-3-diethoxyphosphoryltetrahydropyran-4-ones 12a–j using Gillman reagents, as Michael donors.
Materials 15 03030 sch002
Figure 2. Characteristic coupling constants of all isomers of 2-n-butyl-6-ethyl-3-diethoxyphosphoryltetrahydropyran-4-one 12a.
Figure 2. Characteristic coupling constants of all isomers of 2-n-butyl-6-ethyl-3-diethoxyphosphoryltetrahydropyran-4-one 12a.
Materials 15 03030 g002
Scheme 3. Synthesis of 6-substituted 3-diethoxyphosphoryl-2-isopropyltetrahydropyran-4-ones 12k–o using a Grignard reagent, i-PrMgCl, as a Michael donor.
Scheme 3. Synthesis of 6-substituted 3-diethoxyphosphoryl-2-isopropyltetrahydropyran-4-ones 12k–o using a Grignard reagent, i-PrMgCl, as a Michael donor.
Materials 15 03030 sch003
Figure 3. Characteristic coupling constants of the isomers of 3-diethoxyphosphoryl-2,6-diisopropylpyran-4-one 12l.
Figure 3. Characteristic coupling constants of the isomers of 3-diethoxyphosphoryl-2,6-diisopropylpyran-4-one 12l.
Materials 15 03030 g003
Scheme 4. Synthesis of trans-2,6-disubstituted 3-methylidenetetrahydropyran-4-ones trans-13a–o.
Scheme 4. Synthesis of trans-2,6-disubstituted 3-methylidenetetrahydropyran-4-ones trans-13a–o.
Materials 15 03030 sch004
Scheme 5. Synthesis of 2-substituted 3-diethoxyphosphoryltetrahydropyran-4-ones 14a-d.
Scheme 5. Synthesis of 2-substituted 3-diethoxyphosphoryltetrahydropyran-4-ones 14a-d.
Materials 15 03030 sch005
Figure 4. Characteristic coupling constants of all isomers of 5-diethoxyphosphoryl-2-phenyltetrahydropyran-4-one 14c.
Figure 4. Characteristic coupling constants of all isomers of 5-diethoxyphosphoryl-2-phenyltetrahydropyran-4-one 14c.
Materials 15 03030 g004
Scheme 6. Synthesis of 2-(4-methoxyphenyl)-5-methylidenetetrahydropyran-4-one 15d.
Scheme 6. Synthesis of 2-(4-methoxyphenyl)-5-methylidenetetrahydropyran-4-one 15d.
Materials 15 03030 sch006
Figure 5. Effect of 13d on HL-60 cells after 24h incubation. (A) Morphological changes of HL-60 cells treated with 13d at IC50 and 2IC50 concentrations observed under a light microscope. (B) Metabolic activity of HL-60 cells treated with increasing concentrations of 13d assessed by MTT assay. (C) HL-60 cells treated with 13d at IC50 and 2IC50 concentrations stained with Annexin V and PI were analyzed by flow cytometry. (DH) Cells treated with 13d at IC50 and 2 IC50 concentrations were stained with Anti-BrdU, Anti-Cleaved-PARP (Asp214) and Anti-H2AX (pS139) antibodies and DAPI solution and analyzed by flow cytometry. The error bars represent means of three replicates ± SEM. ns—statistically non-significant; * p < 0.05, ** p < 0.01, *** p < 0.001 were considered significant in comparison to untreated control.
Figure 5. Effect of 13d on HL-60 cells after 24h incubation. (A) Morphological changes of HL-60 cells treated with 13d at IC50 and 2IC50 concentrations observed under a light microscope. (B) Metabolic activity of HL-60 cells treated with increasing concentrations of 13d assessed by MTT assay. (C) HL-60 cells treated with 13d at IC50 and 2IC50 concentrations stained with Annexin V and PI were analyzed by flow cytometry. (DH) Cells treated with 13d at IC50 and 2 IC50 concentrations were stained with Anti-BrdU, Anti-Cleaved-PARP (Asp214) and Anti-H2AX (pS139) antibodies and DAPI solution and analyzed by flow cytometry. The error bars represent means of three replicates ± SEM. ns—statistically non-significant; * p < 0.05, ** p < 0.01, *** p < 0.001 were considered significant in comparison to untreated control.
Materials 15 03030 g005
Figure 6. Effect of 13g on HL-60 cells after 24h incubation. (A)—Morphological changes of HL-60 cells treated with 13g at IC50 and 2IC50 concentrations observed under a light microscope. (B)—Metabolic activity of HL-60 cells treated with increasing concentrations of 13d assessed by MTT assay. (C)—HL-60 cells treated with 13g at IC50 and 2 IC50 concentrations stained with Annexin V and PI were analyzed by flow cytometry. (DH)—Cells treated with 13g at IC50 and 2 IC50 concentrations were stained with Anti-BrdU, Anti-Cleaved-PARP (Asp214) and Anti-H2AX (pS139) antibodies and DAPI solution and analyzed by flow cytometry. The error bars represent means of three replicates ± SEM. ns—statistically non-significant; * p < 0.05, ** p < 0.01, *** p < 0.001 were considered significant in comparison to untreated control.
Figure 6. Effect of 13g on HL-60 cells after 24h incubation. (A)—Morphological changes of HL-60 cells treated with 13g at IC50 and 2IC50 concentrations observed under a light microscope. (B)—Metabolic activity of HL-60 cells treated with increasing concentrations of 13d assessed by MTT assay. (C)—HL-60 cells treated with 13g at IC50 and 2 IC50 concentrations stained with Annexin V and PI were analyzed by flow cytometry. (DH)—Cells treated with 13g at IC50 and 2 IC50 concentrations were stained with Anti-BrdU, Anti-Cleaved-PARP (Asp214) and Anti-H2AX (pS139) antibodies and DAPI solution and analyzed by flow cytometry. The error bars represent means of three replicates ± SEM. ns—statistically non-significant; * p < 0.05, ** p < 0.01, *** p < 0.001 were considered significant in comparison to untreated control.
Materials 15 03030 g006
Figure 7. Inhibition of topoisomerase IIα by 13d and 13g at 0.5 and 1 mM concentrations.
Figure 7. Inhibition of topoisomerase IIα by 13d and 13g at 0.5 and 1 mM concentrations.
Materials 15 03030 g007
Table 1. Yields of diethyl 4-hydroxy-2-oxoalkylphosphonates 9a-e.
Table 1. Yields of diethyl 4-hydroxy-2-oxoalkylphosphonates 9a-e.
CompoundR1Yield [%] a
9
aEt71
bi-Pr73
cPh76 b
dp-MeOPh84 b
e Materials 15 03030 i00180
a Yields of pure, isolated products based on 7. b These compounds are not new and were reported previously [21].
Table 2. Yields and isomers ratio of 6-substituted 2-n-butyl- or 2-phenyl-3-diethoxyphosphoryltetrahydropyran-4-ones 12a–j.
Table 2. Yields and isomers ratio of 6-substituted 2-n-butyl- or 2-phenyl-3-diethoxyphosphoryltetrahydropyran-4-ones 12a–j.
12R1R2r-2-trans-3-trans-6 atrans-enol ar-2-cis-3-
trans-6 a
Yield b [%]
aEtn-Bu8111865
bPh6729484
ci-Prn-Bu77121176
dPh7227197
ePhn-Bu5246288
fPh6236289
gp-MeOPhn-Bu8711273
hPh7227181
i Materials 15 03030 i002n-Bu74141279
jPh5542363
a Ratios of trans/enol/cis isomers were taken from 31P NMR spectra of the crude products. b Yields of pure isolated products.
Table 3. Yields and ratios of isomers of 6-substituted 3-diethoxyphosphoryl-2-isopropyltetrahydropyran-4-ones 12k–o.
Table 3. Yields and ratios of isomers of 6-substituted 3-diethoxyphosphoryl-2-isopropyltetrahydropyran-4-ones 12k–o.
12R1R1r-2-trans-3-trans-6 atrans-enol ar-2-cis-3-
trans-6 a
r-2-trans-3-
cis-6 a
cis-enol ar-2-cis-3-
cis-6 a
Yield b [%]
kEt46112732263
li-Pr56112921168
mPh57512311387
np-MeOPh44214031067
o Materials 15 03030 i003815141859
a Ratios of trans/enol/cis isomers were taken from 31P NMR spectra of the crude products. b Yields of pure isolated products.
Table 4. Yields of trans-2,6-disubstituted 3-methylidenetetrahydropyran-4-ones trans-13a–o obtained.
Table 4. Yields of trans-2,6-disubstituted 3-methylidenetetrahydropyran-4-ones trans-13a–o obtained.
13R1R2Yield a [%]
aEti-Pr44
bn-Bu88
cPh67
di-Pri-Pr50
en-Bu83
fPh52
gPhi-Pr48
hn-Bu77
iPh60
jp-MeOC6H4i-Pr41
kn-Bu82
lPh50
m Materials 15 03030 i004i-Pr32
nn-Bu60
oPh52
a Yields of pure, isolated products.
Table 5. Yields and isomer ratios of 2-substituted 3-diethoxyphosphoryltetrahydropyran-4-ones 14a-d.
Table 5. Yields and isomer ratios of 2-substituted 3-diethoxyphosphoryltetrahydropyran-4-ones 14a-d.
14R1transaenol acisaYield b [%]
aEt52222671
bi-Pr55113476
cPh52133574
dp-MeOC6H462271162
a Ratios of trans/enol/cis isomers were taken from 31P NMR spectra of the crude products. b Yields of pure isolated products.
Table 6. In vitro cytotoxic activity of methylidenetetrahydropyran-4-ones 13a–o and 15d tested on two cancer cell lines and one normal cell line.
Table 6. In vitro cytotoxic activity of methylidenetetrahydropyran-4-ones 13a–o and 15d tested on two cancer cell lines and one normal cell line.
AnalogR1R2IC50 [μM] a
HL-60MCF-7HUVEC
Et
13ai-Pr21.30 ± 1.0020.36 ± 0.35
13bn-Bu39.15 ± 0.15106.5 ± 0.50
13cPh11.85 ± 0.1536.00 ± 0.41
i-Pr
13di-Pr7.65 ± 0.0514.00 ± 1.0024.7 ± 0.65
13en-Bu78.75 ± 5.25>150
13fPh8.45 ± 0.0569.65 ± 2.1611.94 ± 2.32
Ph
13gi-Pr3.69 ± 0.088.63 ± 0.073.84 ± 0.01
13hn-Bu134.00 ± 1.00>150
13iPh36.00 ± 1.41>200
p-MeOC6H4
13ji-Pr3.75 ± 0.079.11 ± 0.024.25 ± 0.05
13kn-Bu>100>100
13lPh25.01 ± 0.94203.00 ± 56.00
Materials 15 03030 i005
13mi-Pr3.87 ± 0.165.73 ± 0.36
13nn-Bu62.60 ± 0.61>100
13oPh>100>100
15dp-MeOC6H4H163.00 ± 7.07>200
Carboplatin 2.53 ± 0.203.76 ± 0.355.35 ± 0.05
a Compound concentration required to inhibit tumor cell proliferation by 50%. Data are expressed as mean ± SEM from the concentration–response curves of at least three experiments.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bartosik, T.; Drogosz-Stachowicz, J.; Janecka, A.; Kędzia, J.; Pacholczyk-Sienicka, B.; Szymański, J.; Gach-Janczak, K.; Janecki, T. Stereoselective Synthesis and Anticancer Activity of 2,6-Disubstituted trans-3-Methylidenetetrahydropyran-4-ones. Materials 2022, 15, 3030. https://doi.org/10.3390/ma15093030

AMA Style

Bartosik T, Drogosz-Stachowicz J, Janecka A, Kędzia J, Pacholczyk-Sienicka B, Szymański J, Gach-Janczak K, Janecki T. Stereoselective Synthesis and Anticancer Activity of 2,6-Disubstituted trans-3-Methylidenetetrahydropyran-4-ones. Materials. 2022; 15(9):3030. https://doi.org/10.3390/ma15093030

Chicago/Turabian Style

Bartosik, Tomasz, Joanna Drogosz-Stachowicz, Anna Janecka, Jacek Kędzia, Barbara Pacholczyk-Sienicka, Jacek Szymański, Katarzyna Gach-Janczak, and Tomasz Janecki. 2022. "Stereoselective Synthesis and Anticancer Activity of 2,6-Disubstituted trans-3-Methylidenetetrahydropyran-4-ones" Materials 15, no. 9: 3030. https://doi.org/10.3390/ma15093030

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