Skip to main content
ACS Medicinal Chemistry Letters logoLink to ACS Medicinal Chemistry Letters
. 2010 Aug 4;1(8):400–405. doi: 10.1021/ml100110x

Concise Syntheses of the Abyssinones and Discovery of New Inhibitors of Prostate Cancer and MMP-2 Expression

Rebecca L Farmer †,§, Margaret M Biddle , Antoinette E Nibbs , Xiaoke Huang §, Raymond C Bergan §,, Karl A Scheidt †,‡,*
PMCID: PMC2992379  NIHMSID: NIHMS227003  PMID: 21116437

Abstract

graphic file with name ml-2010-00110x_0009.jpg

The first asymmetric syntheses of four members of the abyssinone class of natural products (I, II, III, and IV 4′-OMe) via quinine- or quinidine-derived thiourea-catalyzed intramolecular conjugate additions of β-keto ester alkylidenes are reported. This concise strategy delivers all four natural products and their corresponding antipodes. A preliminary evaluation of all of these small molecules against a metastatic human prostate cancer cell line has identified that these compounds selectively and differentially inhibit cell growth and downregulate the expression of matrix metalloproteinase-2 (MMP-2) at nontoxic concentrations.

Keywords: Abyssinones, prostate cancer, MMP-2 expression, hydrogen-bonding catalysis


The abyssinones are a family of chiral, enantiomerically enriched flavanones that exhibit a diverse range of promising biological activities (see Figure 1).13 Although extracts containing these small molecules have been used as traditional remedies, any investigation of their specific anticancer properties necessitates a general synthetic approach that can access them in enantiomerically enriched form. Our program geared toward the synthesis of pyran-containing natural products prompted our interest in these compounds.4,5 Despite their therapeutic promise, the abyssinones have not been (a) synthesized with control over the absolute stereochemistry and (b) evaluated for their ability to inhibit cancer progression.68 The unadorned structure of flavanones belies the challenge in executing a strategy that installs and maintains the configuration at the C2 position.9 This stereocenter in the abyssinones is quite sensitive—mildly basic conditions promote reversible ring-opening to achiral 2′-hydroxychalcones.10 In addition, flavanones containing electron-donating substituents in the C4′ position are particularly susceptible to racemization.9,11 A limited number of approaches for the stereoselective synthesis of flavanones have been developed, but these approaches are not general and would not provide the abyssinone core in a concise manner.1216

Figure 1.

Figure 1

Retrosynthetic analysis of abyssinone natural products.

Our synthetic plan is outlined in Figure 1. The key step in the strategy is the application of our asymmetric thiourea-catalyzed cyclization, which can provide controlled access to either stereoisomer of these natural products.9,11 A Knoevenagel condensation between an appropriately protected β-keto ester and different aldehydes corresponding to each abyssinone would provide the requisite alkylidenes for our thiourea-catalyzed cyclization. Our greatest concern was maintaining the integrity of the newly formed C2 stereocenter, which would depend heavily on the identification of mild decarboxylation/deprotection conditions. We anticipated that we could maximize efficiency if the decarboxylation and unmasking of the C7 phenol were performed in a single flask as the last step of the synthesis.

The first challenge encountered in the total synthesis of the abyssinones involved accessing the alkylidene substrates for the thiourea cyclization (see Figure 2 for a representative example with abyssinone I). Using the requisite aldehyde 5, prepared from 4-hydroxybenzaldehyde,17 we attempted the condensation with the protected β-keto ester 7 under standard Knoevenagel conditions (piperidinium acetate, Dean−Stark apparatus). These reactions produced the desired E alkylidene 8 in >95:5 E:Z for all aldehydes employed in the Knoevenagel condensation. (We were not able to generate and isolate in quantity the Z alkylidene species with which to conduct cyclization studies.) Unfortunately, the reaction conditions also led to the formation of significant amounts of racemic cyclization adducts (±9). After we surveyed various conditions, we discovered that bis-morpholine aminal 6, used directly without purification, underwent smooth Knoevenagel condensation with β-ketoester 7 (2 equiv of glacial acetic acid at 22 °C in toluene) to deliver alkylidene 8 with minimal levels of racemic cyclized compounds. As predicted, the resulting alkylidene 8 underwent the desired intramolecular conjugate addition when exposed to 10 mol % quinine-derived C6′ thiourea catalyst I.18 Multiple iterations of the cyclization of alkylidene 8 consistently afforded carboxyflavanone 9 in good yield (78−80%) and with excellent enantioselectivity (90−91% ee). Importantly, the alkylidene precursors underwent smooth cyclization with excellent control over the newly formed C2 stereocenter using our chiral thiourea catalysis conditions.

Figure 2.

Figure 2

Initial route to abyssinone I (1).

Initially, the decarboxylation and deprotection of 9 were accomplished in a single flask utilizing MgBr2·OEt2, but the low yields for this reaction sequence (less than 25%) led us to investigate higher yielding reactions conditions. Unfortunately, various O-aryl-protected t-butyl carboxyflavanones could not be converted into 14 without significant epimerization or functionalization of the abyssinones containing prenyl side chains. We anticipated that exchanging the t-butyl ester for an allyl ester and replacing the methoxymethyl group with an allyl group might allow for mild, single-flask decarboxylation/deprotection employing palladium catalysis.19 The requisite allyl-protected alkylidenes 1013 were synthesized directly from the corresponding aldehydes using our mild bis-aminal approach. The key asymmetric cyclizations of 1013 were catalyzed by exposure to 10 mol % of either the quinine- or quinidine-derived thiourea I or II(20) at −25 °C in toluene (Table 1). The use of morpholine in the presence of 5 mol % Pd(PPh3)4 promoted the deprotection and decarboxylation at 22 °C to afford the natural product abyssinones I, II, III, and IV 4′-OMe ether cleanly and in high yields.21 The levels of enantioenrichment for each compound from the asymmetric conjugate addition (i.e., 1013 to 1417) were excellent. The quinidine-derived thiourea catalyst II provided each of the abyssinones (after deprotection/decarboxylation) with the naturally occurring configuration at C2, while employing the pseudoenantiomeric thiourea I generated the unnatural (R)-abyssinones with comparable levels of stereoselectivity. These results are especially satisfying since (1) flavanones containing alkoxy or hydroxy substituents in the C4′ position are susceptible to epimerization due to stabilized benzylic cation formation and (2) flavanones can undergo reversible ring opening to form 2′-hydroxychalcones upon exposure to mild bases. The overall yields for this process (conjugate addition, allyl deprotection, and decarboxylation) provided an efficient method for the construction of these flavanones and allowed for the evaluation of our synthetic method in terms of more complex synthetic efforts.

Table 1. Catalytic Asymmetric Synthesis of Abyssinones.

graphic file with name ml-2010-00110x_0008.jpg

graphic file with name ml-2010-00110x_0011.jpg
a

Enantiomeric excess of abyssinones (14) after palladium(0) deprotection determined by HPLC analysis (Chiralcel OD-H or AD-H).

b

Isolated yield from 1013.

We next sought to determine if the enantiomerically enriched abyssinones and their corresponding enantiomers would elicit important and stereodependent biological activity. Members of the broad flavonoid family (over 5000 natural products) have exhibited a wide variety of anticancer effects, acting as antioxidants, angiogenesis inhibitors, and potent cytotoxic agents.22 In addition, several flavonoids have been shown to inhibit the activity and downregulate the expression of pro-metastatic enzymes, such as the matrix metalloproteinases (MMPs), in a variety of tumor cell lines.2326 However, these studies have focused primarily on achiral isoflavones, and none of these investigations to date have been conducted on the abyssinone family of flavanone natural products.

Our studies focused on prostate cancer (PCa), which is the second most common cause of cancer-related death in U.S. men. Mortality from PCa is caused primarily by the development of metastatic disease, in which PCa cells move from the prostate gland to distant sites in the body and continue their unchecked growth.27 Proteases such as MMPs increase cell invasion, and thus, their synthesis by cancer cells facilitates movement and metastatic behavior.28 MMP-2 has been shown to be a particularly important target for human PCa, since its increased expression in tissue portends metastasis.29 Decreased MMP-2 expression also leads to decreased invasion of human PCa cells in vitro and to decreased metastasis of human PCa cells in a murine model of metastasis.25,30 Consequently, the MMPs, and MMP-2 in particular, represent an important and extensively studied therapeutic target. However, while some first-generation MMP inhibitors have been effective in preclinical models, there has been little success in subsequent clinical trials, predominantly due to severe systemic toxicity.31,32

An alternative promising strategy for counteracting the pro-invasive effect of MMP-2 is controlling the amount of this enzyme produced by tumor cells (i.e., upregulation and downregulation).30 We have demonstrated that PCa cells treated with the isoflavone genistein exhibit decreased levels of both MMP-2 gene transcript and protein, which leads to an overall reduction in invasive potential.25 Furthermore, studies performed on additional cancer cell lines have demonstrated that siRNA knockdown of MMP-2 leads to a reduction of both invasion and tumor-induced angiogenesis.33 Developing a method to downregulate MMP expression levels also has the advantage of being potentially less toxic than traditional MMP inhibitors. Thus, we sought to evaluate the ability of the abyssinones generated in our laboratory to inhibit PCa cell growth and downregulate the expression of MMP-2, since these types of interventions could both attenuate/prevent metastasis and increase survival rates for PCa. We were also eager to evaluate the effect of stereochemistry on the ability of the abyssinones to inhibit cell proliferation and downregulate MMP-2 expression.

Our studies began by evaluating the impact of enantioenriched and racemic abyssinones (I, II, III, and IV 4′-OMe) on metastatic PCa (PC3-M) cell growth. Given that enantioenriched abyssinones have never been synthesized, we were particularly interested to determine if the enantiomers of each compound demonstrated differential cytotoxicity when compared to each other and also to the racemic mixture.34 Furthermore, these cytotoxicity studies were instrumental for defining nontoxic levels of 14, which guided the MMP-2 transcript expression evaluations (vide infra). Metastatic variant human PC3-M cells were treated for 3 days with 0−50 μM racemic, (R)- or (S)-14 (12 compounds total) under conditions of exponential cell growth, and then, MTT assays were performed.35 The CC50 values were determined for each compound and are shown in Figure 3. As indicated by Figure 3, (S)-abyssinones III (3) and IV 4′-OMe (4) both demonstrated statistically significant levels of inhibition of PC3-M cell proliferation when compared to the (R)-enantiomers and the racemic mixtures. In fact, both (S)-abyssinones III and IV 4′-OMe caused an approximate 2-fold decrease in the proliferation of PC3-M cells, as they possessed CC50 values of 17 and 12 μM, respectively, as compared to the CC50 values of the (R)-enantiomers and racemates at 42 [(R)-3], 57 [(±)-3], 46 [(R)-4], and 33 μM [(±)-4]. These results clearly indicate the key role of stereochemistry in the biological activity of these compounds and validate our enantioselective synthetic approach for these particular natural products. Further evidence for the importance of the enantioselective synthesis of the abyssinones is also provided by the growth inhibition data for the racemic compounds (Figure 3). Racemic abyssinone II was actually significantly less active than both of the enantioenriched compounds. This particularly interesting result indicates that there may be some functional antagonism between the enantiomers of abyssinone II that lead to a significant reduction of activity in the racemic mixture. Finally, we also performed some preliminary experiments to try to determine if the abyssinones were inhibiting cell growth via a cytotoxic (i.e., cell death) or cytostatic (i.e., cell cycle arrest) mechanism. Interestingly, when the absorbance values at 540 nm for the treated cells at 50 μM (t = 72 h) were compared with those for untreated cells at the start of the experiment (t = 0 h), (S)-abyssinones III and IV 4′-OMe both appeared to reduce the number of cells to a level at or below that of the start of the experiment. This result suggested that these compounds were acting through a cytotoxic mechanism, which was verified by cell imaging that indicated gross cell death at 72 h (see the Supporting Information). However, all of the other compounds, including (R)-abyssinones III and IV 4′-OMe, demonstrated absorbance values that were greater than that of the t = 0 measurement, indicating that they were most likely acting through a cytostatic or combination (cytostatic and cytotoxic) mechanism. Therefore, all of these combined findings highlight the importance of our hydrogen-bonding catalysis approach, since this information could not have been obtained without sufficient amounts of the enantioenriched compounds for analysis.

Figure 3.

Figure 3

CC50 values for each enantiomer and the racemic mixtures of abyssinones 14 against proliferation of PC3-M cells. For experimental details, see the Supporting Information.

We next chose to evaluate the impact of (R)- and (S)-14 (eight compounds total) on the levels of MMP-2 transcript, given that several members of the flavonoid family of natural products have been shown to downregulate the expression of this important pro-metastatic enzyme. Our aim was not only to determine the cytotoxicity of these compounds (i.e., their chemotherapeutic potential) but also to evaluate their ability to target MMP-2 synthesis and to act as chemopreventative agents capable of blocking progression to a more aggressive disease state. Cells were treated with each compound at 3 μM for 3 days, since this dose was not associated with cytotoxicity by the MTT assay. Importantly, we chose this nontoxic dose for our analysis to eliminate any confounding nonspecific effects due to cell toxicity. After abyssinone exposure, MMP-2 transcript levels were measured by isolating RNA and performing reverse transcription and quantitative real time polymerase chain reactions (qRT-PCR). MMP-2 transcript levels were measured instead of MMP-2 protein levels because MMP-2 expression is tightly regulated at the transcriptional level36 and because flavonoids have been shown to inhibit MMP-2 transcript expression. Furthermore, decreased levels of MMP-2 mRNA expression have also been correlated with a lower pathological stage of PCa and, by extension, an improved clinical outcome.29,37 The measurement of transcript levels provides a direct measure of the ability of the abyssinones to target cell-based processes responsible for regulating this clinically relevant pro-metastatic enzyme. Once the MMP-2 transcript levels were measured by qRT-PCR, they were normalized to that of the internal control gene GAPDH (glyceraldehyde-3-phosphate dehydrogenase), which was measured in both the treated and the untreated cells.38 GAPDH is expressed at constant levels across a wide array of biological systems and is widely used for control purposes. We found that the GAPDH expression levels were the same for both the treated and the untreated cells, which indicated that the effects on MMP-2 expression were not a result of nonspecific downregulation of transcription.

With both the MMP-2 studies and the cell growth studies, the abyssinone enantiomers exhibited statistically significant differential biological activity (Figure 4). In particular, for both abyssinones III and IV 4′-OMe, the (R)-enantiomers suppressed MMP-2 expression to 65−78% of untreated control cells, respectively, at a nontoxic concentration (3 μM). Interestingly, the activity against MMP-2 expression for the unnatural (R)-enantiomers was significantly greater than that for the corresponding natural (S)-enantiomers. While there are limited examples of natural products whose unnatural enantiomers demonstrate comparable biological potency (e.g., fredericamycin A39), we are not aware of compounds other than ent-roseophillin40 and (R)-abyssinones III and IV 4′-OMe that are more active than their natural enantiomers. These examples clearly demonstrate that the construction of unnatural antipodes by stereoselective synthesis provides new avenues for potential therapeutic development. The in vitro assays described above interrogate distinctly different biological functions and highlight the importance of successfully installing the stereochemical elements during the syntheses of the abyssinones. The cytotoxicity and MMP-2 studies conducted using each enantiomer show different response profiles across the abyssinones (14), with promising abilities to differentially target cell growth and metastatic potential.

Figure 4.

Figure 4

MMP-2 expression levels of abyssinone-treated PC3-M cells at 3 μM. For experimental details, see the Supporting Information.

In summary, the first asymmetric syntheses of abyssinones I−III and IV 4′-OMe have been accomplished using a chiral thiourea-catalyzed intramolecular conjugate addition. A tandem allyl deprotection/allyl ester decarboxylation generates the final products in a single flask while maintaining the stereochemical integrity at the C2 center. Our synthetic route delivers the natural products and the corresponding antipodes with excellent levels of enantioselectivity, thereby facilitating the evaluation of individual stereoisomers and fueling the discovery of their differential bioactivity. The key observations that the abyssinones specifically and differentially curtail cell growth and suppress MMP-2 expression in whole cells at nontoxic concentrations underscore the utility of our successful catalytic asymmetric synthesis of these natural products. These combined experiments fully integrate asymmetric catalysis, target synthesis application, and chemical biology discovery with medicinal relevance. Continued investigations of the abyssinones and related compounds to understand their biological activity against various cancer models are ongoing in our laboratories.

Supporting Information Available

General information, general procedures, comparative analysis, selected NMR spectra, HPLC traces, and general biological assay procedures and additional data. This material is available free of charge via the Internet at http://pubs.acs.org.

Support for this work has been provided to K.A.S. by the American Cancer Society (Scholar Grant 09-016-01 CDD), AstraZeneca, Amgen, GlaxoSmithKline, the H Foundation, the Northwestern Medical Foundation (Dixon Translational Award), and the Robert H. Lurie Cancer Center (NU) and to R.C.B. by a Veterans Administration Merit Award. This work was funded in part by the Chicago Biomedical Consortium with support from The Searle Funds at The Chicago Community Trust. M.M.B. acknowledges the NIH for a postdoctoral training fellowship (T32 AG00260) directed by the Center for Molecular Innovation and Drug Discovery (NU). A.E.N. and R.L.F. thank the NIH for predoctoral fellowships (F31CA132617 and F31CA138097, respectively) and the Malkin Scholars Program for cancer research grants.

Funding Statement

National Institutes of Health, United States

Supplementary Material

ml100110x_si_001.pdf (1.7MB, pdf)

References

  1. Kamat V. S.; Chuo F. Y.; Kubo I.; Nakanishi K. Anti-Microbial Agents from an East African Medicinal Plant Erythrina abyssinica. Heterocycles 1981, 15, 1163–1170. [Google Scholar]
  2. Lee D.; Bhat K. P. L.; Fong H. H. S.; Farnsworth N. R.; Pezzuto J. M.; Kinghorn A. D. Aromatase Inhibitors from Broussonetia papyrifera. J. Nat. Prod. 2001, 64, 1286–1293. [DOI] [PubMed] [Google Scholar]
  3. Na M.; Jang J.; Njamen D.; Mbafor J. T.; Fomum Z. T.; Kim B. Y.; Oh W. K.; Ahn J. S. Protein Tyrosine Phosphatase-1B Inhibitory Activity of Isoprenylated Flavonoids Isolated from Erythrina mildbraedii. J. Nat. Prod. 2006, 69, 1572–1576. [DOI] [PubMed] [Google Scholar]
  4. Custar D. W.; Zabawa T. P.; Scheidt K. A. Total Synthesis and Structural Revision of the Marine Macrolide Neopeltolide. J. Am. Chem. Soc. 2008, 130, 804–805. [DOI] [PubMed] [Google Scholar]
  5. Morris W. J.; Custar D. W.; Scheidt K. A. Stereoselective Synthesis of Tetrahydropyran-4-ones from Dioxinones Catalyzed by Scandium(III) Triflate. Org. Lett. 2005, 7, 1113–1116. [DOI] [PubMed] [Google Scholar]
  6. Rao G. V.; Swamy B. N.; Chandregowda V.; Reddy G. C. Synthesis of (±)-Abyssinone I and Related Compounds: Their Anti-Oxidant and Cytotoxic Activities. Eur. J. Med. Chem. 2009, 44, 2239–2245. [DOI] [PubMed] [Google Scholar]
  7. Maiti A.; Cuendet M.; Croy V. L.; Endringer D. C.; Pezzuto J. M.; Cushman M. Synthesis and Biological Evaluation of (±)-Abyssinone II and Its Analogues as Aromatase Inhibitors for Chemoprevention of Breast Cancer. J. Med. Chem. 2007, 50, 2799–2806. [DOI] [PubMed] [Google Scholar]
  8. Moriarty R. M.; Grubjesic S.; Surve B. C.; Chandersekera S. N.; Prakash O.; Naithani R. Synthesis of Abyssinone II and Related Compounds as Potential Chemopreventive Agents. Eur. J. Med. Chem. 2006, 41, 263–267. [DOI] [PubMed] [Google Scholar]
  9. Biddle M. M.; Lin M.; Scheidt K. A. Catalytic Enantioselective Synthesis of Flavanones and Chromanones. J. Am. Chem. Soc. 2007, 129, 3830–3831. [DOI] [PubMed] [Google Scholar]
  10. Nising C. F.; Bräse S. The Oxa-Michael Reaction: From Recent Developments to Applications in Natural Product Synthesis. Chem. Soc. Rev. 2008, 37, 1218–1228. [DOI] [PubMed] [Google Scholar]
  11. Wang L. J.; Liu X. H.; Dong Z. H.; Fu X.; Feng X. M. Asymmetric Intramolecular Oxa-Michael Addition of Activated α,β-Unsaturated Ketones Catalyzed by a Chiral N,N′-Dioxide Nickel(II) Complex: Highly Enantioselective Synthesis of Flavanones. Angew. Chem., Int. Ed. 2008, 47, 8670–8673. [DOI] [PubMed] [Google Scholar]
  12. Brown M. K.; Degrado S. J.; Hoveyda A. H. Highly Enantioselective Cu-Catalyzed Conjugate Additions of Dialkylzinc Reagents to Unsaturated Furanones and Pyranones: Preparation of Air-Stable and Catalytically Active Cu-Peptide Complexes. Angew. Chem., Int. Ed. 2005, 44, 5306–5310. [DOI] [PubMed] [Google Scholar]
  13. Hodgetts K. J. Inter- and Intramolecular Mitsunobu Reaction Based Approaches to 2-Substituted Chromans and Chroman-4-ones. Tetrahedron 2005, 61, 6860–6870. [Google Scholar]
  14. Hodgetts K. J.; Maragkou K. I.; Wallace T. W.; Wootton R. C. R. Conjugate Addition to 3-Arylsulfinylchromones as a Synthetic Route to Homochiral 2-Substituted Chromanones: Scope and Limitations. Tetrahedron 2001, 57, 6793–6804. [Google Scholar]
  15. Kawasaki M.; Kakuda H.; Goto M.; Kawabata S.; Kometani T. Asymmetric Synthesis of 2-Substituted Chroman-4-ones Using Lipase-Catalyzed Kinetic Resolutions. Tetrahedron: Asymmetry 2003, 14, 1529–1534. [Google Scholar]
  16. Noda Y.; Watanabe M. Synthesis of Both Enantiomers of Flavanone and 2-Methylchromanone. Helv. Chim. Acta 2002, 85, 3473–3477. [Google Scholar]
  17. See the Supporting Information for synthetic details.
  18. Marcelli T.; van der Haas R. N. S.; van Maarseveen J. H.; Hiemstra H. Asymmetric Organocatalytic Henry Reaction. Angew. Chem., Int. Ed. 2006, 45, 929–931. [DOI] [PubMed] [Google Scholar]
  19. Tsuji J.; Nisar M.; Shimizu I. Facile Palladium-Catalyzed Decarboxylation Reaction of Allylic β-Keto Esters. J. Org. Chem. 1985, 50, 3416–3417. [Google Scholar]
  20. Dickmeiss G.; De Sio V.; Udmark J.; Poulsen T. B.; Marcos V.; Jorgensen K. A. Organocatalytic Asymmetric Desymmetrization-Fragmentation of Cyclic Ketones. Angew. Chem., Int. Ed. 2009, 48, 6650–6653. [DOI] [PubMed] [Google Scholar]
  21. All spectroscopic data (1H, 13C, IR, MS, and [α]) of the synthetic abyssinones matched the values reported for the natural products; see the Supporting Information.
  22. Andersen O. M.; Markham K. R.. Flavanoids: Chemistry, Biochemistry and Applications; CRC Press: Boca Raton, 2006. [Google Scholar]
  23. Puli S.; Lai J. C. K.; Bhushan A. Inhibition of Matrix Degrading Enzymes and Invasion in Human Glioblastoma (U87MG) Cells by Isoflavones. J. Neuro-Oncol. 2006, 79, 135–142. [DOI] [PubMed] [Google Scholar]
  24. Sartor L.; Pezzato E.; Dell’Aica I.; Caniato R.; Biggin S.; Garbisa S. Inhibition of Matrix-Proteases by Polyphenols: Chemical Insights for Anti-Inflammatory and Anti-Invasion Drug Design. Biochem. Pharmacol. 2002, 64, 229–237. [DOI] [PubMed] [Google Scholar]
  25. Huang X. K.; Chen S.; Xu L.; Liu Y. Q.; Deb D. K.; Platanias L. C.; Bergan R. C. Genistein Inhibits p38 Map Kinase Activation, Matrix Metalloproteinase Type 2, and Cell Invasion in Human Prostate Epithelial Cells. Cancer Res. 2005, 65, 3470–3478. [DOI] [PubMed] [Google Scholar]
  26. Xu L.; Bergan R. C. Genistein Inhibits Matrix Metalloproteinase Type 2 Activation and Prostate Cancer Cell Invasion by Blocking the Transforming Growth Factor β-Mediated Activation of Mitogen-Activated Protein Kinase-Activated Protein Kinase 2−27-kDa Heat Shock Protein Pathway. Mol. Pharmacol. 2006, 70, 869–877. [DOI] [PubMed] [Google Scholar]
  27. Jemal A.; Siegel R.; Ward E.; Hao Y.; Xu J.; Murray T.; Thun M. J. Cancer Statistics, 2008. CA Cancer J. Clin. 2008, 58, 71–96. [DOI] [PubMed] [Google Scholar]
  28. Woodhouse E. C.; Chuaqui R. F.; Liotta L. A. General Mechanisms of Metastasis. Cancer 1997, 80, 1529–1537. [DOI] [PubMed] [Google Scholar]
  29. Stearns M.; Stearns M. E. Evidence for Increased Activated Metalloproteinase 2 (MMP-2a) Expression Associated with Human Prostate Cancer Progression. Oncol. Res. 1996, 8, 69–75. [PubMed] [Google Scholar]
  30. Lakshman M.; Xu L.; Ananthanarayanan V.; Cooper J.; Takimoto C. H.; Helenowski I.; Pelling J. C.; Bergan R. C. Dietary Genistein Inhibits Metastasis of Human Prostate Cancer in Mice. Cancer Res. 2008, 68, 2024–2032. [DOI] [PubMed] [Google Scholar]
  31. Overall C. M.; Lopez-Otin C. Strategies for MMP Inhibition in Cancer: Innovations for the Post-Trial Era. Nat. Rev. Cancer 2002, 2, 657–672. [DOI] [PubMed] [Google Scholar]
  32. Fisher J. F.; Mobashery S. Recent Advances in MMP Inhibitor Design. Cancer Metastasis Rev. 2006, 25, 115–136. [DOI] [PubMed] [Google Scholar]
  33. For an example, see Chetty C.; Bhoopathi P.; Joseph P.; Chittivelu S.; Rao J. S.; Lakka S. Adenovirus-Mediated Small Interfering RNA Against Matrix Metalloproteinase-2 Suppresses Tumor Growth and Lung Metastasis in Mice. Mol. Cancer Ther. 2006, 5, 2289–2299. [DOI] [PubMed] [Google Scholar]
  34. When (R)-abyssinone III is exposed to cell culture media and is heated to 37 °C, there is no racemization at 9 and 24 h as observed by HPLC (97:3 hexanes:IPA on AD-H Chiralcel column). However, (S)-abyssinone II does undergo loss of enantiomeric excess over a period of 24 h under the same conditions, presumably due to the free 4′-OH phenol. Abyssinones I and IV 4′-OMe have similar structures to abyssinones III and do not have free phenols like abyssinone II. As such, these compounds are also expected to be stable to epimerization under the biological testing conditions.
  35. Kyle E.; Neckers L.; Takimoto C.; Curt G.; Bergan R. Genistein-Induced Apoptosis of Prostate Cancer Cells is Preceded by a Specific Decrease in Focal Adhesion Kinase Activity. Mol. Pharmacol. 1997, 51, 193–200. [DOI] [PubMed] [Google Scholar]
  36. Vihinen P.; Kahari V. M. Matrix Metalloproteinases in Cancer: Prognostic Markers and Therapeutic Targets. Int. J. Cancer 2002, 99, 157–166. [DOI] [PubMed] [Google Scholar]
  37. Kuniyasu H.; Troncoso P.; Johnston D.; Bucana C. D.; Tahara E.; Fidler I. J.; Pettaway C. A. Relative Expression of Type IV Collagenase, E-cadherin, and Vascular Endothelial Growth Factor/Vascular Permeability Factor in Prostatectomy Specimens Distinguishes Organ-Confined from Pathologically Advanced Prostate Cancers. Clin. Cancer Res. 2000, 6, 2295–2308. [PubMed] [Google Scholar]
  38. Ding Y.; Xu L.; Chen S.; Jovanovic B. D.; Helenowski I. B.; Kelly D. L.; Catalona W. J.; Yang X. J.; Pins M.; Ananthanarayanan V.; Bergan R. C. Characterization of a Method for Profiling Gene Expression in Cells Recovered from Intact Human Prostate Tissue Using RNA Linear Amplification. Prostate Cancer Prostatic Dis. 2006, 9, 379–391. [DOI] [PubMed] [Google Scholar]
  39. Boger D. L.; Huter O.; Mbiya K.; Zhang M. S. Total Synthesis of Natural and ent-Fredericamycin A. J. Am. Chem. Soc. 1995, 117, 11839–11849. [Google Scholar]
  40. Boger D. L.; Hong J. Y. Asymmetric Total Synthesis of ent-(−)-Roseophilin: Assignment of Absolute Configuration. J. Am. Chem. Soc. 2001, 123, 8515–8519. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ml100110x_si_001.pdf (1.7MB, pdf)

Articles from ACS Medicinal Chemistry Letters are provided here courtesy of American Chemical Society

RESOURCES