Next Article in Journal
Triphenylamine-Containing Benzoic Acids: Synthesis, Liquid Crystalline and Redox Properties
Next Article in Special Issue
A Review of Biologically Active Oxime Ethers
Previous Article in Journal
Microanalysis Characterization and Immunomodulatory Effect for Selenium-Enriched Polysaccharide from Morchella esculenta (L.) Pers.
Previous Article in Special Issue
Synthesis of 4-Amino-N-[2 (diethylamino)Ethyl]Benzamide Tetraphenylborate Ion-Associate Complex: Characterization, Antibacterial and Computational Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Targeting Carbohydrate Mimetics of Tetrahydrofuran-Containing Acetogenins to Prostate Cancer

by
Patricia Gonzalez Periche
1,
Jacky Lin
1,
Naga V. S. D. K. Bhupathiraju
1,
Teja Kalidindi
2,
Delissa S. Johnson
2,
Nagavarakishore Pillarsetty
2 and
David R. Mootoo
1,*
1
Department of Chemistry, Hunter College and The Graduate Center, City University of New York, New York, NY 10065, USA
2
Department of Radiology, Memorial Sloan Kettering Cancer Center, New York, NY 10065, USA
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(7), 2884; https://doi.org/10.3390/molecules28072884
Submission received: 28 February 2023 / Revised: 20 March 2023 / Accepted: 21 March 2023 / Published: 23 March 2023
(This article belongs to the Special Issue Design, Synthesis and Biological Activity of Organic Compounds)

Abstract

:
The high potency of the tetrahydrofuran-containing acetogenins (THF-ACGs) against a broad range of human cancer cell lines has stimulated interest in structurally simpler mimetics. In this context, we have previously reported THF-ACG mimetics in which the THF and butenolide moieties of a mono-THF-ACG were replaced with carbohydrate and thiophene residues, respectively. In the present study, towards the targeting of these carbohydrate analogues to prostate cancer (PCa), we synthesized prodrugs in which a parent thiophene or butenolide congener was conjugated through a self-immolative linker to 2-[3-(1,3-dicarboxypropyl)ureido] pentanedioic acid (DUPA), a highly specific ligand for prostate-specific membrane antigen (PSMA), which is overexpressed on prostate tumors. Both prodrugs were found to be more active against receptor positive LNCaP than receptor-negative PC-3 cells, with 2.5 and 12 times greater selectivity for the more potent thiophene analog and the less active butenolide congener, respectively. This selectivity for LNCaP over PC-3 contrasted with the behavior of the parent drugs, which showed similar or significantly higher activity for PC-3 compared to LNCaP. These data support the notion that higher activity of these DUPA-derived prodrugs against LNCaP cells is connected to their binding to PSMA and suggest that the conjugation of PSMA ligands to this family of cytotoxic agents may be effective for targeting them to PCa.

Graphical Abstract

1. Introduction

Natural product-inspired small molecules remain a primary line of cancer therapeutics. In this context, tetrahydrofuran-containing acetogenins (THF-ACGs) have attracted attention for their high potencies against a broad range of human cancer cell lines [1,2,3,4]. Their unusual mechanism of action, which may in part be due to mitochondrial ATP disruption, may make them particularly effective against tumors that are significantly dependent on oxidative phosphorylation [5,6,7,8]. Accordingly, a variety of THF ACG analogues have been examined, with a major focus being on structures in which more complex segments of naturally occurring compounds have been replaced with simpler or easily accessed substitutes [3]. In this vein, we recently reported carbohydrate analogues of naturally occurring ACG 4-deoxyannomontacin (4-DAN) 1, wherein the THF and butenolide moieties were replaced with mannose and thiophene residues, respectively (Figure 1) [9]. These analogues 2 and 4 were prepared and screened as a mixture of C-10 epimers, as earlier studies on similar carbohydrate congeners suggested that the configuration at this central carbinol position should not impact on cytotoxicity, and for simplification of the synthesis [10]. Similar to 4-DAN, these mimetics showed cytotoxicity in the low-to-submicromolar range against two human prostate cancer (PCa) cell lines, PC-3 and LNCaP. Although their mode of action remains to be determined, 2 and 4 are attractive therapeutic leads given the prevalence of PCa [11,12], their relatively simple syntheses, and the possibility that their sugar motifs may confer favorable tumor selectivity and solubility properties.
Against this backdrop, we considered targeting PCa with prodrugs of 2 and 4 in which a tumor vector is conjugated via a self-immolative linker to the C-10 carbinol [13,14,15]. For this preliminary study, we employed an established methodology using 2-[3-(1,3-dicarboxypropyl)ureido] pentanedioic acid (DUPA) as a vector for the prostate-specific membrane antigen (PSMA) and a known disulfide-based linker. PSMA is a transmembrane glycoprotein that is overexpressed on a variety of primary and metastatic PCa and is a target for cancer diagnostics and therapeutics [16,17,18]. DUPA binds PSMA with high specificity and affinity and the receptor promotes endocytosis of bound ligands [19]. DUPA-linked conjugates of imaging and cytotoxic agents have shown significantly increased selectivity for PSMA-expressing tumors [20,21,22,23]. Drug release from prodrugs 3 and 5 is facilitated by interaction with thiols, which exist in higher concentration in cells than in plasma, thereby reducing the possibility of nonspecific uptake of the parent drugs 2 and 4, respectively [16,17,18,19,20,21,22,23,24]. Thus, the guiding hypothesis is that the DUPA ligand would elicit selective uptake of the prodrugs in PSMA-positive cells over PSMA-negative cells, thereby leading to increased activity against the former. Herein, we describe the synthesis and initial cytotoxicity data for 3 and 5.

2. Results and Discussion

2.1. Synthesis

The synthesis of the DUPA thiol segment started with 1-ethyl-3-(3-dimethylamino-propyl) carbodiimide (EDCI)-promoted coupling of the known DUPA carboxylic acid 6 and the HCl salt of amino tosylate 7 [20,21,22,23,25] (Scheme 1). This reaction delivered the chloride 8, which presumably resulted from the substitution of tosylate with in situ chloride. Treatment of 8 with potassium thioacetate yielded 9. Exposure of 9 to trifluoroacetic acid in the presence of triisopropylbutylsilane (TIBS) provided 10 [26].
For prodrug 3, the reaction of 11, an intermediate in the earlier synthesis of 2, and 12, the p-nitrophenyl carbonate of the 2-pyrido disulfide of 2-mercaptoethanol, followed by the removal of the acetal-protecting groups in the product, provided 13 [9,27,28] (Scheme 2). Finally, disulfide exchange on 13 with the in situ-generated thiol derivative of 10 delivered the prodrug conjugate 3. A similar sequence of reactions on 14, the precursor to 4, led to prodrug 5 [9]. Prodrugs 3 and 5 were characterized by NMR and HRMS (Supplementary Material).

2.2. Cytotoxicity Data

The cytotoxicity of 25 against the two human prostate tumor cell lines PC-3 and LNCaP was determined by the XTT cell viability assay after 48 h of treatment with the test compounds (Table 1 and Supplementary Material). LNCaP is known to overexpress the PSMA receptor, whereas PC-3 is receptor-negative [29]. The thiophene parent drug 2 showed low-micromolar–nanomolar-range activity, with eight times higher activity against PC-3 compared to LNCaP. The butenolide drug 4 was at least an order of magnitude less potent compared to 2, with very similar activity against both cell lines. In contrast, both prodrugs 3 and 5 were more selective for LNCaP than PC-3, albeit at lower potency compared to their respective drugs 2 and 4. Prodrug 3, the derivative of the thiophene drug 2 that was more selective for PC-3 over LNCaP cells, showed a reversal in selectivity, with approximately 2.6 times selectivity for LNCaP. Interestingly, the prodrug 5, the derivative of the less potent butenolide drug 4 that was similarly active against both cell lines, was 12 times more selective for LNCaP over PC-3. In comparison, doxorubicin, which is clinically used against PCa, was found to be 3.2 times more active against LNCaP cells [30].
That both prodrugs 3 and 5 are more selective against LNCaP cells compared to their parent drugs 2 and 4, respectively, which are more or similarly active against PC-3 cells, supports our hypothesis that the selectivity of these prodrugs is promoted by their binding to—and internalization by—PSMA on LNCaP cells. More detailed studies, for example, using cotreatment of cells with prodrug and a competing PSMA ligand, will be needed to confirm whether DUPA recognition is the basis for selectivity. Investigations with stable drug-DUPA conjugates that are not degradable intracellularly, as well as quantification of intracellular drugs and prodrug concentrations, will also be needed to interrogate whether the active drug entity is in fact the original drug. In the latter context, it could be argued that the activity of these prodrugs is primarily due to the DUPA moiety and not the original drug. However, this seems unlikely, given the fact that both 3 and 5 have the identical DUPA residue but are appreciably different in potency.
It is also noteworthy that the LNCaP selectivity observed for prodrug 5 is appreciably higher than for prodrug 3. This higher selectivity may be connected to the lower potency of 4, the parent drug in 5, compared to the drug entity in 3. A contributing factor may be the fact that 4 shows no selectivity for either cell line, whereas 2, the parent drug in 3, appears to be intrinsically selective for PC-3 compared to LNCaP cells. In the context of increasing selectivity of this class of DUPA conjugates to PSMA-expressing cells, known DUPA linkers that decrease overall prodrug hydrophobicity and increase PSMA affinity are envisaged [20,21,22]. Potency can be improved by increasing the activity of the drug entity, which may be possible by modifications in the lipid segments, as observed for the parent class of THF ACGs and analogues thereof [1,2,3,31,32].

3. Materials and Methods

Unless otherwise stated, all reactions were carried out under a nitrogen atmosphere in oven-dried glassware using standard syringe and septa technique. NMR spectra were obtained on 500 or 600 MHz instruments. Chemical shifts are relative to the deuterated solvent peak or the tetramethylsilane (TMS) peak at (δ 0.00) and are in parts per million (ppm). Assignments for selected nuclei were determined from 1H COSY and HSQC experiments. Thin-layer chromatography (TLC) was performed on 0.25 mm thick precoated silica gel or C18-W silica, HF254 aluminum sheets. Chromatograms were observed under UV (short- and long-wavelength) light, and/or were visualized by heating plates that were dipped in a solution of ammonium (VI) molybdate tetrahydrate (12.5 g) and cerium (IV) sulfate tetrahydrate (5.0 g) in 10% aqueous sulfuric acid (500 mL). Flash column chromatography (FCC) was performed using silica gel 60 A (230–400 mesh) or C18 (Carbon 17%) 60 A (230–400 mesh) and employed a stepwise solvent polarity gradient, correlated with TLC mobility. Hexanes used for FCC had a boiling point in the 40–60 °C range.

3.1. Chemical Synthesis

3.1.1. 6-aminohexyl 4-methylbenzenesulfonate Hydrochloride (7)

6-((tert-butoxycarbonyl)amino)hexyl 4-methylbenzenesulfonate (5.0 g, 0.013 mol) was stirred in a solution of 4M HCl in dioxane at rt for 16 h. At that time, the mixture was concentrated in vacuo. The residue was triturated with ethyl acetate to provide 7 (4.13 g, quantitative) as a white solid. 1H NMR (C5H5N, 500 MHz) δ 8.01 (d, J = 8.1 Hz, 2H), 7.33 (d, J = 8.1 Hz, 2H), 4.07 (t, J = 6.4 Hz, 2H), 3.18 (t, J = 7.5 Hz, 2H), 2.24 (s, 3H), 1.90 (m, 2H), 1.50 (m, 2H), 1.23 (m, 2H), 1.17 (m, 2H); 13C NMR (C5H5N, 125 MHz) δ 145.6, 134.2, 129.5, 128.8, 71.5, 40.0, 29.3, 28.4, 26.7, 25.5, 21.8. HRMS (ESI) calcd for C13H22ClNO3SNa (M + Na)+ 330.0901 found 330.0898.

3.1.2. Di-tert-butyl(((S)-1-(tert-butoxy)-5-((6-chlorohexyl)amino)-1,5-dioxopentan-2-yl)carbamoyl)-L-glutamate (8)

To a mixture of 6 (0.70 g, 1.4 mmol) and CH2Cl2 (70 mL) at 0 °C under N2, 7 (0.64 g, 2.1 mmol), EDCI (0.81 g, 4.24 mmol), Et3N (0.73 mL, 5.72 mmol), and DMAP (5 mg, 0.043 mmol) were added. The reaction mixture was then stirred at rt for 16 h, then washed successively with 1M HCl, saturated aqueous NaHCO3, and brine. The organic extract was dried (Na2SO4) and evaporated in vacuo. FCC of the residue afforded 8 (0.38 g, 44%) as an oil. Rf = 0.30 (30% EtOAc/hexanes). 1H NMR (CDCl3, 500 MHz) δ 6.61 (br s, 1H), 5.90 (m, 1H), 5.45 (m, 1H), 4.27 (m, 1H), 4.24 (m, 1H), 3.45 (m, 2H), 3.19 (m, 2H), 2.34 (m, 1H), 2.20 (m, 3H), 1.98 (m, 2H), 1.80 (m, 2H), 1.75 (m, 2H), 1.70-1.27 (m, 33H); 13C NMR (CDCl3, 125 MHz) δ 173.5, 172.6, 172.3, 172.1, 157.3, 82.2, 81.1, 80.8, 53.7, 53.2, 45.2, 39.6, 32.6, 32.1, 31.7, 29.5, 28.5, 28.3, 28.2, 26.7, 26.3. HRMS (ESI) calcd for C29H52ClN3O8Na (M + Na)+ 628.3335 found 628.3343.

3.1.3. Tri-tert-butyl (14S,18S)-2,11,16-trioxo-3-thia-10,15,17-triazaicosane-14,18,20-tricarboxylate (9)

KSAc (140 mg, 1.23 mmol) was added to a mixture of 8 (235 mg, 0.39 mmol) and 18-crown-6 ether (20 mg, 0.076 mmol) in anhydrous DMSO (2 mL). The mixture was stirred at rt for 4 h, after which the solvent was evaporated in vacuo. FCC of the residue afforded 9 (154 mg, 62%) as a white solid. Rf = 0.55 (50% EtOAc/hexanes). 1H NMR (CDCl3, 500 MHz) δ 6.48 (m, 1H), 5.50 (br s, 1H), 5.05 (m, 1H), 4.30 (m, 1H), 4.19 (m, 1H), 3.22 (m, 2H), 2.82 (t, J = 7.3 Hz, 2H), 2.40 (m, 1H), 2.29 (s, 3H), 2.25 (m, 3H), 2.03 (m, 2H), 1.84 (m, 2H), 1.58-1.25 (m, 35H); 13C NMR (CDCl3, 150 MHz) δ 196.5, 173.5, 172.6, 172.2, 172.1, 157.3, 82.2, 81.1, 80.8, 53.7, 53.2, 39.5, 32.1, 31.7, 30.9, 29.5, 29.4, 29.1, 28.5, 28.4 (2 signals), 28.3 (2 signals), 28.2, 26.4. HRMS (ESI) calcd for C31H55N3O9SNa (M + Na)+ 668.3551 found 668.3543.

3.1.4. (14. S,18S)-2,11,16-trioxo-3-thia-10,15,17-triazaicosane-14,18,20-tricarboxylic Acid (10)

TFA (0.25 mL) was added to a mixture of 9 (25 mg, 0.039 mmol), TIBS (0.02 mL), and CH2Cl2 (0.25 mL) at rt. The mixture was stirred for 1.5 h, then the volatiles were removed in vacuo. FCC of the residue (25% MeOH/CH2Cl2to HOAc:MeOH: CH2Cl2, 0.1:2:8) afforded 10 (20 mg, quant.) as a white solid. Rf = 0.30 (30% MeOH/ CH2Cl2). 1H NMR (MeOD, 400 MHz) δ 4.30 (dd, J = 5.1, 8.0 Hz, 1H), 4.20 (dd, 5.3, 8.4 Hz, 1H), 3.21 (m, 2H), 2.88 (t, J = 7.2 Hz, 2H), 2.38 (m, 4H), 2.32 (s, 3H), 2.16 (m, 1H), 2.04 (m, 1H), 1.91 (m, 2H), 1.58 (m, 4H), 1.39 (m, 4H); 13C NMR (CDCl3, 100 MHz) δ 197.8, 176.8, 174.7, 160.0, 54.7, 54.0, 40.4, 31.4, 30.8, 30.6, 30.3, 29.9, 29.6, 29.5, 29.0, 27.5. HRMS (ESI) calcd for C19H32N3O9S (M + H)+ 478.1854 found 478.1843.
2-(pyridin-2-yldisulfaneyl)ethyl(17-(thiophene-2-carboxamido)-1-(((2S,3S,4S,5S,6R)-3,4,5-trihydroxy-6-((undecyloxy)methyl)tetrahydro-2H-pyran-2-yl)oxy)heptadecan-8-yl)carbonate (13)
Compound 11 [9] (200 mg, 0.19 mmol) was dissolved in pyridine (2 mL) and 30% HF/pyridine (1 mL) was added. The reaction was stirred at 45 °C for 24 h at rt for additional 24 h, then diluted with saturated aqueous NaHCO3 and extracted with ethyl ether. FCC of the residue afforded the desilylated alcohol derivative (91 mg, 65%). Rf = 0.45 (40% EtOAc/hexanes). A mixture of this material (90 mg, 0.11 mmol), 12 (80 mg, 0.23 mmol), and 4-dimethylaminopyridine (DMAP) (30 mg, 0.25 mmol) in CH2Cl2 (1 mL) was stirred at rt for 16 h. The solvent was then evaporated, and the residue subject to FCC. The partially purified product (Rf = 0.45 (40% EtOAc/hexanes) was dissolved in CH2Cl2 (4 mL) and 5% AcCl in MeOH (0.6 mL) was added. The mixture was stirred at rt for 30 min, then evaporated under reduced pressure at rt. FCC of the residue afforded 13 (76 mg, 44% three steps) as a pale-yellow gum. Rf = 0.42 (8% MeOH/CH2Cl2). 1H NMR (CDCl3, 600 MHz) δ 8.44 (m, 1H), 7.67 (dt, J = 1.0, 8.1 Hz, 1H), 7.62 (dt, J = 1.8, 7.3 Hz, 1H), 7.46 (dd, J = 1.1, 3.8 Hz, 1H), 7.43 (dd, J = 1.1, 4.5 Hz, 1H), 7.08 (ddd, J = 1.0, 4.8, 7.2 Hz, 1H), 7.04 (dd, J = 3.7, 5.0 Hz 1H), 5.99 (br s, 1H), 4.80 (d, J = 1.0 Hz, 1H), 4.67 (m, 1H), 4.35 (t, J = 6.5 Hz, 2H), 3.89 (d, J = 1.7 Hz, 1H), 3.82 (dd, J = 2.0, 8.0 Hz, 1H), 3.76 (t, J = 8.9 Hz, 1H), 3.72-3.62 (m, 4H), 3.48 (m, 2H), 3.38 (m, 3H), 3.05 (t, J = 6.5 Hz, 2H), 1.57 (m, 10H), 1.27 (m, 36H), 0.86 (t, J = 7.1 Hz, 3H); 13C NMR (CDCl3, 600 MHz) δ 162.1, 159.7, 155.0, 149.9, 139.3, 137.3, 129.8, 128.0, 127.7, 121.1, 120.1 (2 signals), 99.7, 79.6, 72.3, 71.9, 71.8 (2 signals), 70.7 (2 signals), 69.4 (2 signals), 68.0, 67.9, 65.2, 40.2, 37.3, 34.1 (2 signals), 32.1, 29.8 (3 signals), 29.7, 29.6, 29.5 (3 signals), 29.4 (2 signals), 27.1, 26.2 (2 signals), 25.3, 22.9, 14.3. 1H NMR (CD3OD, 600 MHz) δ 8.42 (m, 1H), 7.87 (bdd, J = 0.6, 8.0 Hz, 1H), 7.81 (dt, J = 1.8, 8.1 Hz, 1H), 7.67 (dd, J = 1.0, 3.7 Hz, 1H), 7.64 (bdd, J = 0.5, 5.0 Hz, 1H), 7.24 (dd, J = 5.8, 7.2 Hz, 1H), 7.13 (dd, J = 3.8, 4.9 Hz, 1H), 4.72 (d, J = 1.2 Hz, 1H), 4.70 (m, 1H), 4.37 (t, J = 6.1 Hz, 2H), 3.79 (m, 1H), 3.78 (dd, J = 1.6, 9.2 Hz, 1H), 3.73 (m, 1H), 3.68 (m, 2H), 3.62 (dd, J = 3.8, 10.7 Hz, 1H), 3.55 (t, J = 9.6 Hz, 1H), 3.52 (t, J = 6.6 Hz, 2H), 3.43 (m, 1H), 3.35 (t, J = 7.2 Hz, 2H), 3.12 (t, J = 6.1 Hz, 2H), 1.60 (m, 10H), 1.35 (m, 36H), 0.91 (t, J = 7.1 Hz, 3H). 1H NMR (CD3OD, 150 MHz) δ 164.5, 161.2, 156.6, 150.6, 140.6, 139.3, 131.5, 129.5, 128.9, 122.6, 121.4, 101.7, 80.3, 73.5, 72.9, 72.8, 72.3, 72.0, 69.1, 68.7, 66.2, 41.1, 38.9, 35.3, 32.3, 31.0 (2 signals), 30.8, 30.7 (2 signals), 30.6, 30.5, 28.2, 27.5, 27.4, 26.4, 23.9, 14.6. HRMS (ESI) calcd for C47H79N2O10S3 (M + H)+ 927.4890 found 927.4891.

3.1.5. Mannose-thiophene-SS-DUPA (3)

Under a nitrogen atmosphere, 10 (31 mg, 0.06 mmol) was dissolved in MeOH (0.75 mL), and 1M NaOMe in MeOH (0.24 mL, 0.24 mmol) was added until pH 8–10. After 10 min, 5% AcCl in in MeOH (0.15 mL) was added until pH 6–7 to give a pale-brown suspension, which was added to a solution of 13 (38 mg, 0.04 mmol) in MeOH (0.5 mL) under nitrogen. The reaction mixture was stirred for 1h, then adjusted to pH 4 by addition of 5% AcCl in MeOH and evaporated at rt under reduced pressure. The residue was triturated with MeOH and the combined washings evaporated in vacuo. FCC of the residue on C18-SiO2 afforded unreacted 11 (15 mg). Rf = 0.05 (20% H2O/MeOH), and 3 (16 mg, 52% brsm) as a white solid: Rf = 0.32 (20% H2O/MeOH) 1H NMR (CD3OD, 600 MHz) δ 7.69 (dd, J = 1.1, 3.8 Hz, 1H, ArH), 7.64 (dd, J = 1.1, 5.0 Hz, 1H, ArH), 7.13 (dd, J = 3.8, 5.0 Hz, 1H, ArH), 4.74 (s, 1H, H-1′), 4.72 (m, 1H, H-10), 4.37 (m, 2H, H-a), 4.12 (dd, J = 6.2, 8.9 Hz, 2H, H-k, l), 3.82-3.54 (m, 7H, H-2′, 3′, 4′, 5′, 17, CH2-6′), 3.53 (t, J = 6.6 Hz, 2H, CH2-1″), 3.44 (m, 1H, H-17), 3.35 (t, J = 7.3 Hz, 2H, CH2-1), 3.20 (t, J = 6.6 Hz, 2H, H-h), 2.96 (t, J = 6.3 Hz, 2H, H-b), 2.75 (t, J = 7.3 Hz, 2H, H-c), 2.28 (m, 4H, CH2-i, n), 2.20-1.85 (m, 4H, CH2-j, m), 1.70 (m, 2H, CH2-d), 1.68-1.52 (m, 12H, CH2-2, 9, 11, 16, 2″, g), 1.49-1.25 (m, 40H), 0.92 (t, J = 7.1 Hz, 3H, CH3-11″); 13C NMR (CD3OD, 125 MHz) δ 182.1, 181.6, 180.0, 164.5, 160.0, 156.7, 140.6, 131.6, 129.5, 128.9, 101.7, 80.1, 73.5, 72.8 (2 signals), 72.3, 71.9, 69.1, 68.7, 66.7, 57.0, 56.0, 41.1, 40.5, 39.8, 35.4 (2 signals), 35.2, 33.3, 31.6, 31.0 (2 signals), 30.9, 30.8, 30.7, 30.6 (3 signals), 30.5, 30.2, 29.4, 28.2, 27.8, 27.5, 27.4, 26.4, 23.9, 14.7. HRMS (ESI) calcd for C59H103N4O18S3 (M + H)+ 1251.6424 found 1251.6420.

3.1.6. 1-((S)-5-methyl-2-oxo-2,5-dihydro-1H-1λ3-furan-3-yl)-15-(((2S,3S,4S,5S,6R)-3,4,5-trihydroxy-6-((undecyloxy)methyl)tetrahydro-2H-pyran-2-yl)oxy)pentadecan-8-yl (2-(pyridin-2-yldisulfaneyl)ethyl) Carbonate (15)

Compound 14 [9] (245 mg, 0.25 mmol) was dissolved in pyridine (1.5 mL), and 70% HF/pyridine (1 mL) was added. The reaction was stirred at 50 °C for a further 24 h, at which time additional 70% HF/pyridine (1 mL) was introduced and stirring continued at rt for 48 h. The reaction mixture was then diluted with saturated aqueous NaHCO3 and extracted with ethyl ether. FCC of the residue afforded unreacted 14 (30 mg) and the desilylated alcohol derivative (138 mg, 85% based on recovered starting material). Rf = 0.25 (5% acetone/CH2Cl2). 1H NMR (CDCl3, 500 MHz) δ 6.96 (s, 1H), 4.99 (s, 1H), 4.97 (m, 1H), 4.86 (apparent d, J = 6.2 Hz, 1H), 4.66 (apparent d, J = 6.3 Hz, 1H), 4.17 (m, 1H), 4.07 (d, J = 5.6 Hz, 1H), 3.68 (m, 4H), 3.56 (m, 2H), 3.47 (m, 1H), 3.38 (m, 5H), 2.41 (bs, 1H), 2.26 (m, 2H), 1.60-1.32 (m, 48H), 0.86 (t, J = 6.9 Hz, 3H); 13C NMR (CDCl3, 600 MHz) δ 174.1, 149.1, 134.5, 109.4, 97.1, 95.5, 78.7, 76.2, 73.4, 72.1, 72.0, 69.9, 68.4, 67.6, 56.2, 37.7 (2 signals), 32.1, 29.9, 29.8, 29.7 (2 signals), 29.6, 29.5 (2 signals), 29.3, 28.0, 27.6, 26.6, 26.3 (2 signals), 25.8, 25.3, 22.9, 19.4, 14.3.
A mixture of the product from the previous step (115 mg, 0.16 mmol), 12 (100 mg, 0.31 mmol), and DMAP (57 mg, 0.47 mmol) in CH2Cl2 (1 mL) was stirred at rt for 72 h. The solvent was then evaporated in vacuo. FCC of the residue provided the derived carbonate (150 mg, 93%). Rf = 0.55 (5% acetone/ CH2Cl2). 1H NMR (CDCl3, 600 MHz) δ 8.45 (d, J = 4.7 Hz, 1H), 7.63 (m, 2H), 7.07 (m, 1H), 6.95 (bs, 1H), 4.98 (s, 1H), 4.96 (m, 1H), 4.85 (apparent d, J = 6.2 Hz, 1H), 4.64 (m, 2H), 4.35 (t, J = 6.1 Hz, 1H), 4.10 (m, 1H), 4.05 (m, 1H), 3.56 (m, 3H), 3.48 (m, 1H), 3.45 (m, 1H), 3.37 (m, 5H), 3.04 (t, J = 6.1 Hz, 2H), 2.22 (m, 2H), 1.52 (m, 10H), 1.49 (s, 3H), 1.37 (d, J = 6.8 Hz, 3H), 1.29 (s, 3H), 1.27-1.21 (m, 30H), 0.84 (t, J = 6.9 Hz, 3H); 13C NMR (CDCl3, 500 MHz) δ 174.1, 159.7, 155.0, 149.9, 149.1, 137.3, 134.4, 121.1, 120.0, 109.4, 97.1, 96.5, 79.6, 78.6, 76.2, 73.3, 72.0, 69.8, 68.3, 67.6, 65.2, 56.2, 37.2, 34.2, 32.1, 29.9, 29.8, 29.7, 29.5, 29.4, 29.3 (2 signals), 28.0, 27.5, 26.6, 26.3, 25.3, 22.9, 19.4, 14.3.
A portion of the product from the previous step (75 mg, 0.074 mmol) was dissolved in CH2Cl2 (4 mL) and treated with 5% AcCl in MeOH (1.5 mL). The reaction was stirred for 7 h at rt, then adjusted to pH 6 with NaOMe/MeOH. The mixture was concentrated in vacuo. FCC of the residue (5% MeOH/CH2Cl2) afforded 15 (56 mg, 87%). Rf = 0.50 (5% MeOH/CH2Cl2). 1H NMR (CDCl3, 500 MHz) δ 8.51 (d, J = 4.5 Hz, 1H), 7.69 (m, 2H), 7.14 (m, 1H), 7.01 (s, 1H), 5.02 (m, 1H), 4.85 (s, 1H), 4.60 (m, 1H), 4.41 (t, J = 6.5 Hz, 2H), 3.94 (br s, 1H), 3.86 (dd, J = 3.3, 9.1 Hz, 1H), 3.82-3.45 (m, 5H), 3.54 (m, 2H), 3.43 (m, 1H), 3.10 (t, J = 6.5 Hz, 2H), 2.28 (t, J = 8.2 Hz, 2H), 1.59 (m, 12H), 1.40 (d, J = 6.8 Hz, 3H), 1.28 (m, 30H), 0.90 (t, J = 6.8 Hz, 3H); 13C NMR (CDCl3, 125 MHz) δ 174.2, 160.0, 155.0, 149.9, 137.3, 134.4, 121.1, 120.1, 99.7, 79.6, 72.3, 71.9, 71.8, 70.8, 70.6, 69.3 (2 signals), 68.0 (2 signals), 65.2, 37.2, 34.2, 32.1, 29.9, 29.8, 29.7, 29.6, 29.5 (2 signals), 29.4 (2 signals), 29.3 (2 signals), 27.5, 26.2 (3 signals), 25.3 (2 signals), 22.9, 19.4, 14.3. HRMS (ESI) calcd for C45H76NO11S2 (M + H)+ 870.4860 found 870.4850.

3.1.7. Mannose-butenolide-SS-DUPA (5)

Under a nitrogen atmosphere, 10 (80 mg, 0.16 mmol) was dissolved in MeOH (0.5 mL), and 1M NaOMe in MeOH (0.65 mL, 0.65 mmol) was added until pH 10-12. After 15 min, 5% AcCl in in MeOH (0.3 mL) was added until pH 6–7 to give a pale-brown suspension, which was added to a solution of 15 (54 mg, 0.064 mmol) in MeOH (0.5 mL) and THF (0.05 mL) under nitrogen. The reaction mixture was stirred for 30 min, then adjusted to pH 4–5 by addition of 5% AcCl in MeOH and evaporated at rt under reduced pressure. The residue was triturated with MeOH and the combined washings evaporated in vacuo. FCC of the residue on C18-SiO2 afforded 5 (24 mg, 31%) as a white amorphous solid. Rf = 0.45 (20% H2O/MeOH). 1H NMR (CD3OD, 600 MHz) δ 7.28 (s, 1H, H, =CH)), 5.08 (m, 1H, CH3CH-OC=O), 4.73 (partially buried s, 1H, H-1′), 4.71 (m, 1H, H-10), 4.38 (m, 2H, H-a), 4.12 (m, 2H, H-k, l), 3.81-3.54 (m, 7H, H-2′, 3′, 4′, 5′, 17, CH2-6′), 3.52 (t, J = 6.6 Hz, 2H, CH2-1″), 3.43 (m, 1H, H-17), 3.21 (t, J = 6.5 Hz, 2H, H-h), 2.97 (t, J = 6.3 Hz, 2H, H-b), 2.76 (t, J = 7.3 Hz, 2H, H-c), 2.26 (m, 6H, CH2- 3, i, n), 2.15–1.85 (m, 4H, CH2-j, m), 1.75 (m, 2H, CH2-d), 1.65 – 1.55 (m, 12H, CH2-4, 9, 11, 16, 2″, g), 1.50 – 1.30 (m, 39H), 0.92 (t, J = 6.8 Hz, 3H, CH3-11″); 13C NMR (CD3OD, 125 MHz) δ 181.9, 181.4, 179.9, 176.3, 175.7, 159.9, 156.6, 152.4, 134.5, 101.6, 80.0, 79.7, 73.4, 72.7 (2 signals), 72.2, 71.8, 68.9, 68.6, 66.6, 56.8, 55.9, 40.4, 39.7, 38.2, 35.3 (2 signals), 35.1, 33.2, 31.5, 30.9, (2 signals), 30.7, 30.6 (2 signals), 30.5, 30.4, 30.3, 30.1 (2 signals), 29.3, 28.5, 27.7, 27.4, 27.3, 26.3 (2 signals), 26.0, 23.8, 19.3, 14.5. HRMS (ESI) calcd for C57H100N3O19S2 (M + H)+ 1194.6387 found 1194.6299.

3.2. Cytotoxicity Measurements

3.2.1. Cell Lines and Cell Number Optimization

LNCaP (PSMA+, androgen-dependent) and PC-3 (PSMA-, androgen-independent) cells were purchased from the American Type Culture Collection (ATCC, Manassas, VA, USA), authenticated and routinely checked for mycoplasma contamination. LNCaP cells were cultured in Roswell Park Memorial Institute (RMPI) + 10% fetal calf serum (FCS) media and PC-3 cells were cultured in F12K + 10% FCS media. Media were supplemented with penicillin (100 I.U/mL) and streptomycin (100 µg/mL) to minimize bacterial contamination. The media were prepared by MSK Media preparation core. XTT reagent (2,3-bis-(2-methoxy-4-nitro-5-sulfophenyl)-2H-tetrazolium-5-carboxanilide) used for the study was purchased from ATCC (30–1011 K). Serial dilutions of cells in culture media were plated in triplicate from 100–100,000 in a 96-well plate. Cells were incubated at 37 °C in a CO2 incubator for 72 h. An activated solution of XTT reagent was made by adding 0.1 mL of activation reagent (sterile solution of N-methyl dibenzopyrazine methyl sulfate) to 5 mL of XTT reagent. 50 mL of activated XTT reagent was added per well. The plate was incubated for 2 h and absorbance from wells was measured at 475 nM. Background was measured at 660 nM. Background subtraction was made, and absorbance was plotted against cell number. The optimum cell number was determined based on the linear range of the XTT standard curve.

3.2.2. IC50 Determination

Serial dilutions of compounds to be tested were made in culture media from 1 nM–1 mM concentration. The determined cell number from the standard graph was plated in a 96-well plate in triplicate per each dilution of the compounds. The cells were allowed to attach overnight by incubation at 37 °C in a CO2 incubator. The media from the well were removed and media with serial dilutions of the compounds were added to the cells. The cells were incubated with the compound for 48 h at 37 °C in CO2 incubator. Activated XTT reagent was added to the wells and incubated for 2 h. Absorbance from the wells was measured at 475 nM and background at 660 nM. Background subtraction was made, and all the data were plotted in prism and IC50 values were determined.

4. Conclusions

Two prodrugs comprising carbohydrate mimetics of mono-THF-containing acetogenins conjugated via a self-immolative linker to a ligand for PSMA, were prepared. These conjugates were found to be approximately 3 and 12 times more active against LNCaP, a PSMA receptor-positive prostate tumor cell line, compared to PC-3, a PSMA receptor-negative prostate tumor cell line. This result contrasted with the activity of the nonconjugated cytotoxic agents, which showed similar or higher activity against PC-3 compared to LNCaP. These data suggest that targeting PSMA may be viable strategy for modulating the tumor selectivity of this family of cytotoxic agents and new directions for carbohydrate-based therapeutics. Studies on improving prodrug selectivity and potency using modified DUPA linkers and drug entities, as well as assays against a wider panel of human tumor and normal cells are directions for future work. Mechanistic and drug development investigations, including cytotoxicity assays in the presence of competing PSMA ligands, and determination of drug uptake, stability and cellular distribution will also be performed.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28072884/s1, NMR spectra for 26, 810, 13, 15, cytotoxicity data for 25.

Author Contributions

P.G.P. and J.L.: chemical synthesis; N.V.S.D.K.B., T.K. and D.S.J.: cytotoxicity experiments; N.P.: supervision of cytotoxicity experiments, manuscript review and editing; D.R.M.: conceptualization, supervision of chemical synthesis, manuscript preparation. All authors have read and agreed to the published version of the manuscript.

Funding

Chemical synthesis and the NEO-500 NMR spectrometer used to obtain data included in this publication were funded by the National Science Foundation grants CHE-1900509 and CHE-1828399, respectively. Funding support from MSK Core Center grant P30 CA008748 is acknowledged.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in on request from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alali, F.Q.; Liu, X.X.; McLaughlin, J.L. Annonaceous acetogenins: Recent progress. J. Nat. Prod. 1999, 62, 504–540. [Google Scholar] [CrossRef]
  2. McLaughlin, J.L. Paw paw and cancer: Annonaceous acetogenins from discovery to commercial products. J. Nat. Prod. 2008, 71, 1311–1321. [Google Scholar] [CrossRef] [PubMed]
  3. Kojima, N.; Tanaka, T. Medicinal chemistry of annonaceous acetogenins: Design, synthesis, and biological evaluation of novel analogues. Molecules 2009, 14, 3621–3661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Liaw, C.-C.; Wu, T.-Y.; Chang, F.-R.; Wu, Y.-C. Historic perspectives on Annonaceous acetogenins from the chemical bench to preclinical trials. Planta Med. 2010, 76, 1390–1404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Grba, D.N.; Blaza, J.N.; Bridges, H.R.; Agip, A.-N.A.; Yin, Z.; Murai, M.; Miyoshi, H.; Hirst, J. Cryo-electron microscopy reveals how acetogenins inhibit mitochondrial respiratory complex I. J. Biol. Chem. 2022, 298, 101602. [Google Scholar] [CrossRef]
  6. Xue, D.; Xu, Y.; Kyani, A.; Roy, J.; Dai, L.; Sun, D.; Neamati, N. Multiparameter optimization of oxidative phosphorylation inhibitors for the treatment of pancreatic cancer. J. Med. Chem. 2022, 65, 3404–3419. [Google Scholar] [CrossRef]
  7. Ahmad, F.; Cherukuri, M.K.; Choyke, P.L. Metabolic reprogramming in prostate cancer. Br. J. Cancer 2021, 125, 1185–1196. [Google Scholar] [CrossRef]
  8. Naguib, A.; Mathew, G.; Reczek, C.R.; Watrud, K.; Ambrico, A.; Herzka, T.; Salas, I.C.; Lee, M.F.; El-Amine, N.; Zheng, W.; et al. Mitochondrial complex I inhibitors expose a vulnerability for selective killing of Pten-mnull cells. Cell Rep. 2018, 23, 58–67. [Google Scholar] [CrossRef] [Green Version]
  9. Gonzalez Periche, P.; Ramdular, A.; Bhupathiraju, N.V.S.D.K.; Kalidindi, T.; Johnson, D.S.; Pillarsetty, N.; Mootoo, D.R. Synthesis of carbohydrate analogues of the THF-acetogenin 4-deoxyannomontacin and their cytotoxicity against human prostate cancer cell lines. Carbohydr. Res. 2022, 521, 108671. [Google Scholar] [CrossRef]
  10. Wang, F.; Kawamura, A.; Mootoo, D.R. Synthesis and antitumor activity of C-9 epimers of the tetrahydrofuran containing acetogenin 4-deoxyannoreticuin. Bioorg. Med. Chem. 2008, 16, 8413–8418. [Google Scholar] [CrossRef] [Green Version]
  11. Hoffman, R.M. Clinical practice. Screening for prostate cancer. N. Engl. J. Med. 2011, 365, 2013–2019. [Google Scholar] [CrossRef]
  12. Siegel, R.L.; Miller, K.D.; Wagle, N.S.; Jemal, A. Cancer Statistics 2023. CA Cancer J. Clin. 2023, 73, 17–48. [Google Scholar] [CrossRef]
  13. Shi, J.F.; Wu, P.; Cheng, X.L.; Wei, X.Y.; Jiang, Z.H. For studies on naturally occurring THF congeners conjugated to tumor vectors: Synthesis and cytotoxic property of annonaceous acetogenin glycoconjugates. Drug Design Dev. Therap. 2020, 14, 4993–5004. [Google Scholar] [CrossRef]
  14. Shi, J.F.; Wu, P.; Jiang, Z.H.; Wei, X.Y. Synthesis and tumor cell growth inhibitory activity of biotinylated annonaceous acetogenins. Eur. J. Med. Chem. 2014, 71, 219–228. [Google Scholar] [CrossRef]
  15. Queiroz, E.F.; Roblot, F.; Duret, P.; Figadère, B.; Gouyette, A.; Laprévote, O.; Serani, L.; Hocquemiller, R. Synthesis, spectroscopy and cytotoxicity of glycosylated acetogenin derivatives as promising molecules for cancer therapy. J. Med. Chem. 2000, 43, 1604–1610. [Google Scholar] [CrossRef]
  16. Uspenskaya, A.A.; Machulkin, A.E.; Mazhuga, A.G.; Beloglazkina, A.K. Conjugates of prostate-specific membrane antigen ligands with antitumor drugs. Pharm. Chem. J. 2019, 53, 288–297. [Google Scholar] [CrossRef]
  17. Afshar-Oromieh, A.; Babich, J.W.; Kratochwil, C.; Giesel, F.L.; Eisenhut, M.; Kopka, K.; Haberkorn, U. The rise of PSMA ligands for diagnosis and therapy of prostate cancer. J. Nucl Med. 2016, 57, 79S–89S. [Google Scholar] [CrossRef] [Green Version]
  18. Hawkey, N.M.; Sartor, A.O.; Morris, M.J.; Armstrong, A.J. Prostate-specific membrane antigen–targeted theranostics: Past, present, and future approaches. Clin. Adv. Hemat. Oncol. 2022, 20, 227–239. [Google Scholar]
  19. Pastorino, S.; Riondato, M.; Uccelli, L.; Giampiero, G.; Giovannini, E.; Duce, V.; Ciamiello, A. Toward the discovery and development of PSMA targeted inhibitors for nuclear medicine applications. Cur. Radiopharm. 2020, 13, 63–79. [Google Scholar] [CrossRef]
  20. Kularatne, S.A.; Wang, K.; Santhapuram, H.-K.R.; Low, P.S. Prostate-specific membrane antigen targeted imaging and therapy of prostate cancer using a PSMA inhibitor as a homing ligand. Mol. Pharm. 2009, 6, 780–789. [Google Scholar] [CrossRef]
  21. Kularatne, S.A.; Zhou, Z.; Yang, J.; Post, C.B.; Low, P.S. Design, synthesis, and preclinical evaluation of prostate-specific membrane antigen targeted 99mTc-radioimaging agents. Mol. Pharm. 2009, 6, 790–800. [Google Scholar] [CrossRef] [PubMed]
  22. Kularatne, S.A.; Venkatesh, C.; Santhapuram, H.-K.; Wang, K.; Vaitilingam, B.; Henne, W.A.; Low, P.S. Synthesis and biological analysis of prostate-specific membrane antigen-targeted anticancer prodrugs. J. Med. Chem. 2010, 53, 7767–7777. [Google Scholar] [CrossRef] [PubMed]
  23. Lv, Q.; Yang, J.; Zhang, R.; Yang, Z.; Yang, Z.; Wang, Y.; Xu, Y.; He, Z. Prostate-specific membrane antigen targeted therapy of prostate cancer using a DUPA−paclitaxel conjugate. Mol. Pharm. 2018, 15, 1842–1852. [Google Scholar] [CrossRef] [PubMed]
  24. Srinivasarao, M.; Low, P.S. Ligand targeted drug delivery. Chem. Rev. 2017, 117, 12133–12164. [Google Scholar] [CrossRef]
  25. Hansen, J.B.; Buchardt, O. A novel synthesis of tri-, di-, and mono-9-acridinyl derivatives of tetra-, tri-, and di-amines. J. Chem. Soc. Chem. Commun. 1983, 4, 162–164. [Google Scholar] [CrossRef]
  26. Leamon, C.P.; Parker, M.A.; Vlahov, I.R.; Xu, L.-C.; Reddy, J.A.; Vetzel, M.; Douglas, N. Synthesis and biological evaluation of EC20:  A new folate-derived, 99mTc-based radiopharmaceutical. Bioconj. Chem. 2002, 13, 1200–1210. [Google Scholar] [CrossRef]
  27. Satyam, A. Design and synthesis of releasable folate–drug conjugates using a novel heterobifunctional disulfide-containing linker. Bioorg. Med. Chem. Lett. 2008, 18, 3196–3199. [Google Scholar] [CrossRef]
  28. Dubikovskaya, E.A.; Thorne, S.H.; Pillow, T.H.; Contag, C.H.; Wender, P.A. Overcoming multidrug resistance of small-molecule therapeutics through conjugation with releasable octaarginine transporters. Proc. Natl. Acad. Sci. USA 2008, 105, 12128. [Google Scholar] [CrossRef] [Green Version]
  29. Xu, B.; Zhou, F.; Yan, M.-M.; Cai, D.-S.; Guo, W.-B.; Yang, Y.-Q.; Jia, X.-H.; Zhang, W.-X.; Li, T.; Ma, T.; et al. PSMA-oriented target delivery of novel anticancer prodrugs: Design, synthesis, and biological evaluations of oligopeptide-camptothecin conjugates. Int. J. Mol. Sci. 2018, 19, 3251. [Google Scholar] [CrossRef] [Green Version]
  30. Tacar, O.; Sriamornsak, P.; Dass, C.R. Doxorubicin: An update on anticancer molecular action, toxicity and novel drug delivery systems. J. Pharm. Pharmacol. 2013, 65, 157–170. [Google Scholar] [CrossRef]
  31. Abe, M.; Murai, M.; Ichimaru, N.; Kenmochi, A.; Yoshida, T.; Kubo, A.; Kimura, Y.; Moroda, A.; Akabe, H.; Nishioka, T.; et al. Dynamic function of the alkyl spacer of acetogenins in their inhibitory action with mitochondrial complex I (NADH-ubiquinone oxidoreductase). Biochemistry 2005, 44, 14898–14906. [Google Scholar] [CrossRef]
  32. Ohta, K.; Fushimi, T.; Okamura, M.; Akatsuka, A.; Dan, S.; Iwasaki, H.; Yamashita, M.; Kojima, N. Structure–antitumor activity relationship of hybrid acetogenins focusing on connecting groups between heterocycles and the linker moiety. RSC Adv. 2022, 12, 15728–15739. [Google Scholar] [CrossRef]
Figure 1. Natural and unnatural analogues of 4-DAN.
Figure 1. Natural and unnatural analogues of 4-DAN.
Molecules 28 02884 g001
Scheme 1. Synthesis of DUPA segment.
Scheme 1. Synthesis of DUPA segment.
Molecules 28 02884 sch001
Scheme 2. Synthesis of drug-DUPA conjugates.
Scheme 2. Synthesis of drug-DUPA conjugates.
Molecules 28 02884 sch002
Table 1. Cytotoxicity (IC50, mM) of 25.
Table 1. Cytotoxicity (IC50, mM) of 25.
Test CompoundLNCaPPC3Selectivity
LNCaP/PC-3
20.480.060.125
30.812.12.6
43.44.51.3
54.85812
Doxorubicin0.712.33.2
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Periche, P.G.; Lin, J.; Bhupathiraju, N.V.S.D.K.; Kalidindi, T.; Johnson, D.S.; Pillarsetty, N.; Mootoo, D.R. Targeting Carbohydrate Mimetics of Tetrahydrofuran-Containing Acetogenins to Prostate Cancer. Molecules 2023, 28, 2884. https://doi.org/10.3390/molecules28072884

AMA Style

Periche PG, Lin J, Bhupathiraju NVSDK, Kalidindi T, Johnson DS, Pillarsetty N, Mootoo DR. Targeting Carbohydrate Mimetics of Tetrahydrofuran-Containing Acetogenins to Prostate Cancer. Molecules. 2023; 28(7):2884. https://doi.org/10.3390/molecules28072884

Chicago/Turabian Style

Periche, Patricia Gonzalez, Jacky Lin, Naga V. S. D. K. Bhupathiraju, Teja Kalidindi, Delissa S. Johnson, Nagavarakishore Pillarsetty, and David R. Mootoo. 2023. "Targeting Carbohydrate Mimetics of Tetrahydrofuran-Containing Acetogenins to Prostate Cancer" Molecules 28, no. 7: 2884. https://doi.org/10.3390/molecules28072884

Article Metrics

Back to TopTop