Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Andre Rosowsky is active.

Publication


Featured researches published by Andre Rosowsky.


Journal of Clinical Oncology | 1992

Cyclophosphamide pharmacokinetics: correlation with cardiac toxicity and tumor response.

Lois Ayash; Joel E. Wright; Olga Tretyakov; R Gonin; Anthony Elias; C Wheeler; Joseph Paul Eder; Andre Rosowsky; Karen H. Antman; Emil Frei

BACKGROUND Cyclophosphamide, which forms the nucleus for virtually all preparative regimens for autologous bone marrow transplantation (ABMT), is an alkylating agent of which cytotoxicity is not directly caused by the parent compound but by its biologically active metabolites. Its nonmyelosuppressive toxicity in the ABMT setting is cardiomyopathy. We attempted to determine any correlation between plasma levels of total cyclophosphamide and the subsequent development of cardiac dysfunction. PATIENTS AND METHODS Analyses of plasma levels and the derivation of plasma concentration-time curves (area under the curve [AUC]) were performed in 19 women with metastatic breast carcinoma, who received a continuous 96-hour infusion of cyclophosphamide, thiotepa, and carboplatin (CTCb) with ABMT. The assay for total cyclophosphamide measures the inactive parent compound; reliable assays of the active metabolites of cyclophosphamide are not yet available. RESULTS Six of 19 women developed moderate, but transient, congestive heart failure (CHF) as assessed by clinical and radiologic criteria. These patients had a significantly lower AUC of total cyclophosphamide (median, 2,888 mumol/L/h) than patients who did not develop CHF (median, 6,121 mumol/L/h) (P less than .002). Median duration of tumor response in these patients was also more durable; at least 22 months in patients with lower AUCs versus a median of 5.25 months in those with higher AUCs (P = .008). CONCLUSION These pharmacokinetic data support the premise that enhancement of cyclophosphamide activation may lead to both greater tumor cytotoxicity and increased but reversible end-organ toxicity. Early analysis of pharmacokinetic data may allow modulation of cyclophosphamide administration in an attempt to enhance therapeutic efficacy.


Annals of Internal Medicine | 1990

Continuous Infusion High-Dose Leucovorin with 5-Fluorouracil and Cisplatin for Untreated Stage IV Carcinoma of the Head and Neck

Arnon I. Dreyfuss; John R. Clark; Joel E. Wright; Charles M. Norris; Paul M. Busse; J. Lucarini; Barbara G. Fallon; D. Casey; Janet W. Andersen; Richard L. Klein; Andre Rosowsky; Daniel Miller; Emil Frei

STUDY OBJECTIVE To study the activity of continuous infusion cisplatin, 5-fluorouracil, and high-dose leucovorin (PFL) as induction chemotherapy in patients with previously untreated, advanced squamous cell carcinoma of the head and neck. DESIGN Nonrandomized, prospective trial. SETTING A comprehensive cancer center. PATIENTS Thirty-five patients (4 patients [11%], stage III; 31 patients [89%], stage IV [MO]), all evaluable for response and toxicity. INTERVENTIONS Two to three cycles of PFL before definitive, local-regional therapy (surgery and radiation therapy or radiation therapy alone). Chemotherapy included continuous intravenous infusion of cisplatin (25 mg/m2 body surface area daily, days 1 through 5); 5-fluorouracil (800 mg/m2 body surface area daily, days 2 through 6); and leucovorin (500 mg/m2 body surface area daily, days 1 through 6) administered once every 28 days. Pathologic response was evaluated by surgical resection or biopsy. Serum-reduced folates were measured before and 18 hours after the initiation of chemotherapy. RESULTS A clinical response to PFL was achieved in 28 of 35 (80%) patients: 23 (66%) patients had a complete response (90% CI, 50% to 79%) and 5 (14%) patients, a partial response. A complete response was confirmed pathologically in 14 of 19 (74%) patients. The most common toxicity was mucositis (grade 2 to 3; 94% of patients). Dose reduction for toxicity was necessary in 11 (31%) patients. There were no treatment-related deaths. Serum levels of leucovorin and (6S)5-methyltetrahydrofolate were measured in 7 patients. After 18 hours, the mean leucovorin level (+/- SD) was 34.3 +/- 1.5 mumol/L, of which only 8.0 +/- 0.5% was the active 6S isomer. The mean serum (6S)5-methyltetrahydrofolate was 9.2 +/- 0.6 mumol/L. CONCLUSIONS Continuous infusion cisplatin, 5-fluorouracil, and high-dose leucovorin is a new and highly active chemotherapy regimen that can achieve clinical and pathologically confirmed complete responses in a substantial proportion of patients with advanced, local-regional squamous cell carcinoma of the head and neck. Further studies are needed to confirm the activity of PFL and to determine its potential impact on local tumor control and disease-free and overall survival.


Biochemical Pharmacology | 2002

Multiple mechanisms of resistance to methotrexate and novel antifolates in human CCRF-CEM leukemia cells and their implications for folate homeostasis

Robert Mauritz; Godefridus J. Peters; David G. Priest; Yehuda G. Assaraf; Stavit Drori; Ietje Kathmann; Paul Noordhuis; Marlene A. Bunni; Andre Rosowsky; Gerrit Jansen

We determined the mechanisms of resistance of human CCRF-CEM leukemia cells to methotrexate (MTX) vs. those to six novel antifolates: the polyglutamatable thymidylate synthase (TS) inhibitors ZD1694, multitargeted antifolate, pemetrexed, ALIMTA (MTA) and GW1843U89, the non-polyglutamatable inhibitors of TS, ZD9331, and dihydrofolate reductase, PT523, as well as DDATHF, a polyglutamatable glycinamide ribonucleotide transformylase inhibitor. CEM cells were made resistant to these drugs by clinically relevant intermittent 24 hr exposures to 5-10 microM of MTX, ZD1694, GW1843U89, MTA and DDATHF, by intermittent 72 hr exposures to 5 microM of ZD9331 and by continuous exposure to stepwise increasing concentrations of ZD9331, GW1843U89 and PT523. Development of resistance required only 3 cycles of intermittent drug exposure to ZD1694 and MTA, but 5 cycles for MTX, DDATHF and GW1843U89 and 8 cycles for ZD9331. The predominant mechanism of resistance to ZD1694, MTA, MTX and DDATHF was impaired polyglutamylation due to approximately 10-fold decreased folylpolyglutamate synthetase activity. Resistance to intermittent exposures to GW1843U89 and ZD9331 was associated with a 2-fold decreased transport via the reduced folate carrier (RFC). The CEM cell lines resistant to intermittent exposures to MTX, ZD1694, MTA, DDATHF, GW1843U89 and ZD9331 displayed a depletion (up to 4-fold) of total intracellular reduced folate pools. Resistance to continuous exposure to ZD9331 was caused by a 14-fold increase in TS activity. CEM/GW70, selected by continuous exposure to GW1843U89 was 50-fold resistant to GW1843U89, whereas continuous exposure to PT523 generated CEM/PT523 cells that were highly resistant (1550-fold) to PT523. Both CEM/GW70 and CEM/PT523 displayed cross-resistance to several antifolates that depend on the RFC for cellular uptake, including MTX (95- and 530-fold). CEM/GW70 cells were characterized by a 12-fold decreased transport of [3H]MTX. Interestingly, however, CEM/GW70 cells displayed an enhanced transport of folic acid, consistent with the expression of a structurally altered RFC resulting in a 2.6-fold increase of intracellular folate pools. CEM/PT523 cells displayed a markedly impaired (100-fold) transport of [3H]MTX along with 12-fold decreased total folate pools. In conclusion, multifunctional mechanisms of resistance in CEM cells have a differential impact on cellular folate homeostasis: decreased polyglutamylation and transport defects lead to folate depletion, whereas a structurally altered RFC protein can provoke expanded intracellular folate pools.


Biochemical Pharmacology | 1979

Human thymidylate synthetase—III: Effects of methotrexate and folate analogs

Daniel W. Szeto; Cheng Yung-Chi; Andre Rosowsky; Yu Cheng-Sein; Edward J. Modest; James R. Piper; Carroll Temple; Robert D. Elliott; Jerry D. Rose; John A. Montgomery

Abstract The structure-activity relationship of human thymidylate synthetase (EC 2.1.1.45) was studied with two groups of folate analogs: (1) methotrexate (MTX) analogs modified at the glutamate residue and N 10 ; and (2) tetrahydrofolate (H 4 PteGlu) analogs modified at N 5 and N 10 . With respect to MTX analogs, it was found that: (1) substitution of the glutamate side chain by α-aminoadipic acid. α-aminopimelic acid or β-aminoglutaric acid slightly affects its K i ; (2) a free α-carboxyl group on the amino acid side chain of MTX, or any free carboxyl group in that vicinity plays an important role in the inhibitory potency of MTX analogs to the enzyme; (3)esterification or amidation of the α-carboxyl group of MTX decreases the inhibitory potency; and (4) free aspartyl or glutamyl conjugation through a peptide linkage to the γ-carboxyl group of the glutamate side chain decreases its K i to the enzyme by 5- and 8-fold respectively. Tetrahydrofolate analogs formed by inserting an ethylene, iminyl or a carbonyl bridge between the nitrogen at N 5 and N 10 or by substitution at the N 5 position were found to be poor inhibitors under our assay conditions.


Antimicrobial Agents and Chemotherapy | 2001

Dicyclic and Tricyclic Diaminopyrimidine Derivatives as Potent Inhibitors of Cryptosporidium parvum Dihydrofolate Reductase: Structure-Activity and Structure-Selectivity Correlations

Richard G. Nelson; Andre Rosowsky

ABSTRACT A structurally diverse library of 93 lipophilic di- and tricyclic diaminopyrimidine derivatives was tested for the ability to inhibit recombinant dihydrofolate reductase (DHFR) cloned from human and bovine isolates of Cryptosporidium parvum (J. R. Vásquez et al., Mol. Biochem. Parasitol. 79:153–165, 1996). In parallel, the library was also tested against human DHFR and, for comparison, the enzyme from Escherichia coli. Fifty percent inhibitory concentrations (IC50s) were determined by means of a standard spectrophotometric assay of DHFR activity with dihydrofolate and NADPH as the cosubstrates. Of the compounds tested, 25 had IC50s in the 1 to 10 μM range against one or bothC. parvum enzymes and thus were not substantially different from trimethoprim (IC50s, ca. 4 μM). Another 25 compounds had IC50s of <1.0 μM, and 9 of these had IC50s of <0.1 μM and thus were at least 40 times more potent than trimethoprim. The remaining 42 compounds were weak inhibitors (IC50s, >10 μM) and thus were not considered to be of interest as drugs useful against this organism. A good correlation was generally obtained between the results of the spectrophotometric enzyme inhibition assays and those obtained recently in a yeast complementation assay (V. H. Brophy et al., Antimicrob. Agents Chemother. 44:1019–1028, 2000; H. Lau et al., Antimicrob. Agents Chemother. 45:187–195, 2001). Although many of the compounds in the library were more potent than trimethoprim, none had the degree of selectivity of trimethoprim for C. parvum versus human DHFR. Collectively, the results of these assays comprise the largest available database of lipophilic antifolates as potential anticryptosporidial agents. The compounds in the library were also tested as inhibitors of the proliferation of intracellular C. parvum oocysts in canine kidney epithelial cells cultured in folate-free medium containing thymidine (10 μM) and hypoxanthine (100 μM). After 72 h of drug exposure, the number of parasites inside the cells was quantitated by indirect immunofluorescence microscopy. Sixteen compounds had IC50s of <3 μM, and five of these had IC50s of <0.3 μM and thus were comparable in potency to trimetrexate. The finding that submicromolar concentrations of several of the compounds in the library could inhibit in vitro growth of C. parvum in host cells in the presence of thymidine (dThd) and hypoxanthine (Hx) suggests that lipophilic DHFR inhibitors, in combination with leucovorin, may find use in the treatment of intractable C. parvum infections.


Antimicrobial Agents and Chemotherapy | 2000

Identification of Cryptosporidium parvum Dihydrofolate Reductase Inhibitors by Complementation in Saccharomyces cerevisiae

Victoria H. Brophy; John R. Vasquez; Richard G. Nelson; John R. Forney; Andre Rosowsky; Carol Hopkins Sibley

ABSTRACT There is a pressing need for drugs effective against the opportunistic protozoan pathogen Cryptosporidium parvum. Folate metabolic enzymes and enzymes of the thymidylate cycle, particularly dihydrofolate reductase (DHFR), have been widely exploited as chemotherapeutic targets. Although many DHFR inhibitors have been synthesized, only a few have been tested against C. parvum. To expedite and facilitate the discovery of effective anti-Cryptosporidium antifolates, we have developed a rapid and facile method to screen potential inhibitors of C. parvum DHFR using the model eukaryote, Saccharomyces cerevisiae. We expressed the DHFR genes of C. parvum, Plasmodium falciparum, Toxoplasma gondii, Pneumocystis carinii, and humans in the same DHFR-deficient yeast strain and observed that each heterologous enzyme complemented the yeast DHFR deficiency. In this work we describe our use of the complementation system to screen known DHFR inhibitors and our discovery of several compounds that inhibited the growth of yeast reliant on the C. parvum enzyme. These same compounds were also potent or selective inhibitors of the purified recombinantC. parvum DHFR enzyme. Six novel lipophilic DHFR inhibitors potently inhibited the growth of yeast expressing C. parvumDHFR. However, the inhibition was nonselective, as these compounds also strongly inhibited the growth of yeast dependent on the human enzyme. Conversely, the antibacterial DHFR inhibitor trimethoprim and two close structural analogs were highly selective, but weak, inhibitors of yeast complemented by the C. parvum enzyme. Future chemical refinement of the potent and selective lead compounds identified in this study may allow the design of an efficacious antifolate drug for the treatment of cryptosporidiosis.


Proteins | 2008

Structure‐based approach to pharmacophore identification, in silico screening, and three‐dimensional quantitative structure–activity relationship studies for inhibitors of Trypanosoma cruzi dihydrofolate reductase function

Norbert Schormann; Olga Senkovich; Kiera Walker; D.L. Wright; Amy C. Anderson; Andre Rosowsky; S. Ananthan; Bidhan A. Shinkre; Sadanandan E. Velu; Debasish Chattopadhyay

We have employed a structure‐based three‐dimensional quantitative structure–activity relationship (3D‐QSAR) approach to predict the biochemical activity for inhibitors of T. cruzi dihydrofolate reductase‐thymidylate synthase (DHFR‐TS). Crystal structures of complexes of the enzyme with eight different inhibitors of the DHFR activity together with the structure in the substrate‐free state (DHFR domain) were used to validate and refine docking poses of ligands that constitute likely active conformations. Structural information from these complexes formed the basis for the structure‐based alignment used as input for the QSAR study. Contrary to indirect ligand‐based approaches the strategy described here employs a direct receptor‐based approach. The goal is to generate a library of selective lead inhibitors for further development as antiparasitic agents. 3D‐QSAR models were obtained for T. cruzi DHFR‐TS (30 inhibitors in learning set) and human DHFR (36 inhibitors in learning set) that show a very good agreement between experimental and predicted enzyme inhibition data. For crossvalidation of the QSAR model(s), we have used the 10% leave‐one‐out method. The derived 3D‐QSAR models were tested against a few selected compounds (a small test set of six inhibitors for each enzyme) with known activity, which were not part of the learning set, and the quality of prediction of the initial 3D‐QSAR models demonstrated that such studies are feasible. Further refinement of the models through integration of additional activity data and optimization of reliable docking poses is expected to lead to an improved predictive ability. Proteins 2008.


Proteins | 2006

New insights into DHFR interactions: analysis of Pneumocystis carinii and mouse DHFR complexes with NADPH and two highly potent 5-(omega-carboxy(alkyloxy) trimethoprim derivatives reveals conformational correlations with activity and novel parallel ring stacking interactions.

Vivian Cody; Jim Pace; Kim Chisum; Andre Rosowsky

Structural data are reported for two highly potent antifolates, 2,4‐diamino‐5‐[3′,4′‐dimethoxy‐5′‐(5‐carboxy‐1‐pentynyl)]benzylpyrimidine (PY1011), with 5000‐fold selectivity for Pneumocystis carinii dihydrofolate reductase (pcDHFR), relative to rat liver DHFR, and 2,4‐diamino‐5‐[2‐methoxy‐5‐(4‐carboxybutyloxy)benzyl]pyrimidine (PY957), that has 80‐fold selectivity for pcDHFR. Crystal structures are reported for NADPH ternary complexes with PY957 and pcDHFR, refined to 2.2 Å resolution; with PY1011 and pcDHFR, refined to 2.0 Å resolution; and with PY1011 and mouse DHFR (mDHFR), refined to 2.2 Å resolution. These results reveal that the carboxylate of the ω‐carboxyalkyloxy side chain of these inhibitors form ionic interactions with the conserved Arg in the substrate binding pocket of DHFR. These data suggest that the enhanced inhibitory activity of PY1011 compared with PY957 is, in part, due to the favorable contacts with Phe69 of pcDHFR by the methylene carbons of the inhibitor side chain that are oriented by the triple bond of the 1‐pentynyl side chain. These contacts are not present in the PY957 pcDHFR complex, or in the PY1011 mDHFR complex. In the structure of mDHFR the site of Phe69 in pcDHFR is occupied by Asn64. These data also revealed a preference for an unusual parallel ring stacking interaction between Tyr35 of the active site helix and Phe199 of the C‐terminal β sheet in pcDHFR and by Tyr33 and Phe179 in mDHFR that is independent of bound ligand. A unique His174–His187 parallel ring stacking interaction was also observed only in the structure of pcDHFR. These ring stacking interactions are rarely found in any other protein families and may serve to enhance protein stability. Proteins 2006.


Journal of Medicinal Chemistry | 1998

Synthesis and Potent Antifolate Activity and Cytotoxicity of B-Ring Deaza Analogues of the Nonpolyglutamatable Dihydrofolate Reductase Inhibitor Nα-(4-Amino-4-deoxypteroyl)-Nδ-hemiphthaloyl-l-ornithine (PT523)

Andre Rosowsky; Joel E. Wright; Chitra M. Vaidya; Henry Bader; Ronald A. Forsch; Clara E. Mota; Jorge Pardo; Cindy S. Chen; Ying-Nan Chen

Six new B-ring analogues of the nonpolyglutamatable antifolate Nalpha-(4-amino-4-deoxypteroyl)-Ndelta-hemiphthaloy l-L-ornithine (PT523, 3) were synthesized with a view to determining the effect of modifications at the 5- and/or 8-position on dihydrofolate reductase (DHFR) binding and tumor cell growth inhibition. The 5- and 8-deaza analogues were prepared from methyl 2-L-amino-5-phthalimidopentanoate and 4-amino-4-deoxy-N10-formyl-5-deaza- and -8-deazapteroic acid, respectively. The 5,8-dideaza analogues were prepared from methyl 2-L-[(4-aminobenzoyl)amino]-5-phthalimidopentanoate and 2, 4-diaminoquinazoline-6-carbonitriles. The Ki for inhibition of human DHFR by the 5-deaza and 5-methyl-5-deaza analogues was about the same as that of 3 (0.35 pM), 11-fold lower than that of aminopterin (AMT, 1), and 15-fold lower than that of methotrexate (MTX, 2). However the Ki of the 8-deaza analogue was 27-fold lower than that of 1, and that of the 5,8-dideaza, 5-methyl-5,8-dideaza, and 5-chloro-5,8-dideaza analogues was approximately 50-fold lower. This trend was consistent with the published literature on the corresponding DHFR inhibitors with a glutamate side chain. In colony formation assays against the human head and neck squamous carcinoma cell line SCC25 after 72 h of treatment, the 5- and 8-deaza analogues were approximately as potent as 3, whereas the 5,8-dideaza analogue was 3 times more potent. 5-Methyl and 5-chloro substitution was also favorable, with the 5-methyl-5-deaza analogue being 2. 5-fold more potent than the 5-deaza analogue. However the effect of 5-methyl substitution was less pronounced in the 5,8-dideaza analogues than in the 5-deaza analogues. The 5-chloro-5,8-dideaza analogue of 3 was the most active member of the series, with an IC50 = 0.33 nM versus 1.8 nM for 3 and 15 nM for MTX. The 5-methyl-5-deaza analogue of 3 was also tested at the National Cancer Institute against a panel of 50 human tumor cell lines in culture and was consistently more potent than 3, with IC50 values in the low-nanomolar to subnanomolar range against most of the tumors. Leukemia and colorectal carcinoma cell lines were generally most sensitive, though good activity was also observed against CNS tumors and carcinomas of the breast and prostate. The results of this study demonstrate that B-ring analogues of 3 inhibit DHFR activity and tumor cell colony formation as well as, or better than, the parent compound. In view of the fact that 3 and its B-ring analogues cannot form polyglutamates, their high cytotoxicity relative to the corresponding B-ring analogues of AMT is noteworthy.


Antimicrobial Agents and Chemotherapy | 1995

Structure-activity and structure-selectivity studies on diaminoquinazolines and other inhibitors of Pneumocystis carinii and Toxoplasma gondii dihydrofolate reductase.

Andre Rosowsky; John B. Hynes; Sherry F. Queener

Twenty-eight 2,4-diaminopteridines with alkyl and aralkyl groups at the 6- and 7-positions, five 1,3-diamino-7,8,9,10-tetrahydropyrimido [4,5-c]isoquinolines with an alkyl, alkylthio, or aryl group at the 6-position, and nine 4,6-diamino-1,2-dihydro-s-triazines with one or two alkyl groups at the 2-position and a substituted phenyl or naphthyl group at the 1-position were evaluated as inhibitors of dihydrofolate reductase enzymes from Pneumocystis carinii, Toxoplasma gondii, and rat liver. Halogen substitution at the 5- or 6-position of 2,4-diaminoquinazoline favored selective binding to the P. carinii enzyme but not the T. gondii enzyme. For example, the 50% inhibitory concentrations of 2,4-diamino-6-chloroquinazoline as an inhibitor of P. carinii, T. gondii, and rat liver dihydrofolate reductase were 3.6, 14 and 29 microM, respectively, corresponding to 12-fold selectivity for the P. carinii enzyme but only marginal selectivity for the T. gondii enzyme. Greater than fivefold selectivity for P. carinii but not T. gondii dihydrofolate reductase was also observed for the 2,4-diaminoquinazolines with 5-methyl, 5-fluoro, 5- and 6-bromo, 6-chloro, and 5-chloro-6-bromo substitution. In contrast, alkyl and aralkyl substitution at the 6- and 7-positions of 2,4-diaminopteridines was found to be a favorable feature for selective inhibition of the T. gondii enzyme and, in two cases, for both enzymes. Nine of the fifty-one compounds tested against P. carinii dihydrofolate reductase and four of the thirty compounds tested against T. gondii dihydrofolate reductase displayed fivefold or greater selectivity for the microbial enzyme versus the rat liver enzyme. The most selective against both enzymes was 2,4-diamino-6,7-bis(cyclohexylmethyl) pteridine, with a selectivity ratio 2 orders of magnitude greater than the value reported for trimetrexate and piritrexim. Since substitution at the 7-position is generally considered to be detrimental to the binding of 2,4-diaminop-teridines and related compounds to mammalian dihydrofolate reductase, the selectivity observed in this study with the 6,7-bis(cyclohexylmethyl) analog may represent a useful approach to enhancing selective inhibition of the enzyme from nonmammalian species.

Collaboration


Dive into the Andre Rosowsky's collaboration.

Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

James H. Freisheim

University of Toledo Medical Center

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge