js-k, a glutathione/glutathione s-transferase-activated nitric oxide donor of the diazeniumdiolate...

Upload: kawtherahmed

Post on 03-Apr-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/28/2019 JS-K, a Glutathione/Glutathione S-Transferase-activated Nitric Oxide Donor of the Diazeniumdiolate Class with Pote

    1/10

    2003;2:409-417.Mol Cancer TherPaul J. Shami, Joseph E. Saavedra, Lai Y. Wang, et al.

    1Antineoplastic ActivityPotentNitric Oxide Donor of the Diazeniumdiolate Class with

    Transferase-activatedS-JS-K, a Glutathione/Glutathione

    Updated version

    http://mct.aacrjournals.org/content/2/4/409Access the most recent version of this article at:

    Cited Articles

    http://mct.aacrjournals.org/content/2/4/409.full.html#ref-list-1This article cites by 26 articles, 2 of which you can access for free at:

    Citing articles

    http://mct.aacrjournals.org/content/2/4/409.full.html#related-urlsThis article has been cited by 9 HighWire-hosted articles. Access the articles at:

    E-mail alerts related to this article or journal.Sign up to receive free email-alerts

    SubscriptionsReprints and

    [email protected] atTo order reprints of this article or to subscribe to the journal, contact the AACR Publications

    Permissions

    [email protected] at

    To request permission to re-use all or part of this article, contact the AACR Publications

    Research.on March 22, 2013. 2003 American Association for Cancermct.aacrjournals.orgDownloaded from

    http://mct.aacrjournals.org/content/2/4/409http://mct.aacrjournals.org/content/2/4/409http://mct.aacrjournals.org/content/2/4/409.full.html#ref-list-1http://mct.aacrjournals.org/content/2/4/409.full.html#ref-list-1http://mct.aacrjournals.org/content/2/4/409.full.html#related-urlshttp://mct.aacrjournals.org/content/2/4/409.full.html#related-urlshttp://mct.aacrjournals.org/cgi/alertshttp://mct.aacrjournals.org/cgi/alertsmailto:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]://mct.aacrjournals.org/http://mct.aacrjournals.org/http://mct.aacrjournals.org/mailto:[email protected]:[email protected]://mct.aacrjournals.org/cgi/alertshttp://mct.aacrjournals.org/content/2/4/409.full.html#related-urlshttp://mct.aacrjournals.org/content/2/4/409.full.html#ref-list-1http://mct.aacrjournals.org/content/2/4/409
  • 7/28/2019 JS-K, a Glutathione/Glutathione S-Transferase-activated Nitric Oxide Donor of the Diazeniumdiolate Class with Pote

    2/10

    JS-K, a Glutathione/Glutathione S-Transferase-activatedNitric Oxide Donor of the Diazeniumdiolate Class withPotent Antineoplastic Activity1

    Paul J. Shami,2 Joseph E. Saavedra, Lai Y. Wang,Challice L. Bonifant, Bhalchandra A. Diwan,Shivendra V. Singh, Yijun Gu, Stephen D. Fox,Gregory S. Buzard, Michael L. Citro,David J. Waterhouse, Keith M. Davies, Xinhua Ji,and Larry K. Keefer

    Division of Medical Oncology, Department of Internal Medicine,University of Utah and Salt Lake City Veterans Administration MedicalCenters, Salt Lake City, Utah 84148 [P. J. S., L. Y. W.]; Basic ResearchProgram [J. E. S., B. A. D., G. S. B., M. L. C.] and Analytical ChemistryLaboratory [S. D. F.], Science Applications International Corporation-Frederick, National Cancer Institute at Frederick, Frederick, Maryland21702; Department of Pharmacology, University of Pittsburgh CancerInstitute, Pittsburgh, Pennsylvania 15213 [S. V. S., Y. G.];Macromolecular Crystallography Laboratory, National Cancer Instituteat Frederick, Frederick, Maryland 21702 [Y. G., X. J.]; Department ofChemistry, George Mason University, Fairfax, Virginia 22030 [K. M. D.];and Laboratory of Comparative Carcinogenesis, National CancerInstitute at Frederick, Frederick, Maryland 21702 [C. L. B., D. J. W.,L. K. K.]

    Abstract

    We have previously shown that nitric oxide (NO)

    inhibits growth and induces differentiation and

    apoptosis in acute myeloid leukemia cells, with the HL-

    60 human myeloid leukemia line being particularly

    sensitive to NO-mediated cytolysis. With the goal

    of identifying a prodrug that can target NO to theleukemia cells without inducing NO-mediated systemic

    hypotension, we have screened a series of O2-aryl

    diazeniumdiolates designed to be stable at

    physiological pH but to release NO upon reaction

    with glutathione. O2-(2,4-Dinitrophenyl) 1-[(4-

    ethoxycarbonyl)piperazin-1-yl]diazen-1-ium-1,2-diolate

    (JS-K) proved to be the most active antiproliferative

    agent among those tested in HL-60 cells, with an IC50of 0.20.5 M. After 5 days of exposure to 0.5 M JS-K,

    HL-60 cells had differentiated and acquired some of

    the phenotypic features of normal monocytes. One- to

    2-day treatment with JS-K at concentrations of 0.51

    M resulted in apoptosis induction in a concentration-and caspase-dependent manner. JS-K also inhibited

    the growth of solid tumor cell lines but to a lesser

    extent than HL-60 cells. JS-K was administered i.v. to

    nonobese diabetic-severe combined immune deficient

    mice at doses of up to 4 mol/kg without inducing

    significant hypotension. The growth of s.c. implanted

    HL-60 cells was reduced by50% when the mice

    received i.v. injections three times/week with 4 mol/

    kg boluses of JS-K. Histological examination of tumor

    explants from JS-K-treated animals revealed extensive

    necrosis. Similar results were seen with s.c. human

    prostate cancer (PPC-1) xenografts. Our data indicate

    that JS-K is a promising lead compound for the

    possible development of a novel class of antineoplastic

    agents.

    Introduction

    AML3 is a life-threatening disease with an annual incidence

    of 2.25 per 100,000 (1). Despite therapy with different classes

    of chemotherapeutic agents, including anthracyclines, 1--

    D-arabinofuranosylcytosine, etoposide, all-trans-retinoic

    acid, as well as high-dose therapy with stem cell rescue, the

    overall 5-year disease-free survival remains 30%. This il-

    lustrates the need for agents with novel mechanisms of

    action.This study is based on the premise that the unusual sen-

    sitivity of AML cells to the cytotoxic effects of NO can be

    exploited for improved therapy of this malignancy. We pre-

    viously showed that NO-releasing drugs of the diazenium-

    diolate class induce apoptosis in the HL-60 human myeloid

    leukemia cell line at relatively low concentrations (for exam-

    ple, an IC50 of 0.006 mM for a cell-permeant prodrug that is

    activated for NO release by intracellular esterases; Ref. 2).

    This is in marked contrast to, for example, normal vascular

    smooth muscle cells, which experience cytostasis without

    toxicity in the continuous presence of 0.20.5 mM (Z)-1-[2-

    (2-Aminoethyl)-N-(2-ammonioethyl)amino]diazen-1-ium-1,2-

    diolate (half-life for NO release of 20 h) for 6 days (3), and to

    primary hepatocytes, which can actually be protected from

    apoptosis by exposure to a diazeniumdiolate that is acti-

    vated for NO release by cytochrome P450 (4).

    Received 11/25/02; revised 1/30/03; accepted 2/12/03.The costs of publication of this article were defrayed in part by thepayment of page charges. This article must therefore be hereby markedadvertisement in accordance with 18 U.S.C. Section 1734 solely to indi-cate this fact.1 This work is supported by a Translational Research Award from theLeukemia and Lymphoma Society (to P. J. S.). This project was alsosupported, in part, through National Cancer Institute Contract No. NO1-CO-12400 (J. E. S., B. A. D., S. D. F., G. S. B., M. L. C.), and NationalInstitute of Environmental Health Sciences Grant ES09140 (to S. V. S.).2 To whom requests for reprints should be addressed, at SLC VA MedicalCenter, Box 151M, 500 Foothill Boulevard, Salt Lake City, UT 84148.Phone: (801) 582-1565; Fax: (801) 583-9624; E-mail: [email protected].

    utah.edu.

    3 The abbreviations used are: AML, acute myeloid leukemia; BSO, buthi-onine sulfoximine; CDNB, 1-chloro-2,4-dinitrobenzene; DNP-SG, S-(2,4-dinitrophenyl)glutathione; GSH, glutathione; GST, glutathione S-transfer-ase; GSTCD, 1-( S-glutathionyl)-2,4,6-trinitrocyclohexadienate anion;HPLC, high performance liquid chromatography; JS-K, O2-(2,4-dinitro-phenyl) 1-[(4-ethoxycarbonyl)piperazin-1-yl]diazen-1-ium-1,2-diolate; MS,mass spectrometry; MTT, 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetra-zolium bromide; NAC, N-acetyl-L-cysteine; NO, nitric oxide; NOD-SCID,nonobese diabetic-severe combined immune deficient; NSE, nonspecificesterase; xGSTY1-1, class Y (where Y is: A, ; M, ; P, ) GST of subunit

    type 1 from x (where x is: h, human; r, rat).

    409Vol. 2, 409417, April 2003 Molecular Cancer Therapeutics

    Research.on March 22, 2013. 2003 American Association for Cancermct.aacrjournals.orgDownloaded from

    http://mct.aacrjournals.org/http://mct.aacrjournals.org/http://mct.aacrjournals.org/
  • 7/28/2019 JS-K, a Glutathione/Glutathione S-Transferase-activated Nitric Oxide Donor of the Diazeniumdiolate Class with Pote

    3/10

    One might, therefore, expect NO-based therapies to ex-

    hibit some selectivity for the NO-sensitive leukemia cells

    relative to normal tissues. With this in mind, we have

    screened a library of arylated diazeniumdiolates that were

    designed to be activated for NO release by reaction with

    cellular thiols such as GSH (5) with or without catalysis by

    GST, the major isoforms (, , ) of which are expressed in

    75, 55, and 95% of AML cases, respectively (6). We report

    here on the in vitro and in vivo antineoplastic activity of the

    most potent growth inhibitor of this family tested to date,JS-K (structure in Fig. 1).

    Materials and Methods

    Chemicals. JS-K (7) and 4-carbethoxy-PIPERAZI/NO (Ref.

    8; structures shown in Fig. 1) were synthesized as described

    previously. S-(2,4-Dinitrophenyl)glutathione was prepared by

    the method of Mancini et al. (9). The pan-caspase inhibitor

    Z-VAD-FMK was from Biomol (Plymouth Meeting, PA).

    Daunorubicin and etoposide were from Dr. Grayden Harker

    (University of Utah). HPLC grade solvents were purchased

    from VWR Scientific Co. (South Plainfield, NJ). All other

    chemicals were from Sigma (St. Louis, MO) unless otherwise

    noted.

    Chromatography and MS. HPLC separations were car-

    ried out with a Phenomenex Luna 5 C18(2) column (Tor-

    rance, CA) using a gradient of acetonitrile:water (each con-

    taining 0.1% formic acid) at a flow rate of 1 ml/min

    (percentage of acetonitrile: 25% for 0 5 min followed by a

    linear program to 70% at 15 min).HPLC-MS studies were performed on an Agilent Capillary

    Series 1100 LC/MSD Ion Trap mass spectrometer with elec-

    trospray ionization in the positive ion mode. Separations

    were effected as above, except that the flow rate was 15

    l/min, and the gradient was 10% acetonitrile for 0 5 min

    followed by a linear program to 70% at 15 min.

    Molecular Modeling. CDNB is a model substrate used

    extensively for monitoring activity of GST (10). It has been

    shown that GSTs with a higher specific activity toward CDNB

    stabilize the Meisenheimer complex (Fig. 1) at the transition

    state better than the isoforms with poor catalytic activity for

    CDNB-GSH conjugation (5). Crystal structures of a transition

    state analogue, GSTCD

    , in complex with rGSTM1-1, a class rat GST isoform (11), and with hGSTP1-1 (isoleucine-

    104, alanine-113 variant), a class human GST isoform (12),

    have been reported and provided the foundation for model-

    ing the transition state of GST-catalyzed GSH conjugation of

    JS-K. The initial model of the Meisenheimer complex of GSH

    and JS-K for hGSTM1-1 and that for hGSTP1-1 built based

    on the crystal structures of GSTCD in rGSTM1-1 (11) and

    that in hGSTP1-1 (12), respectively. The models were subject

    to geometry optimization using the conjugate gradient

    method of Powell (13) and docked into the active sites of

    ligand-free hGSTM1-1 and hGSTP1-1 built based on the

    rGSTM1-1 (11) and hGSTP1-1 (12) structures, respectively.

    The geometry of the protein-Meisenheimer complexes wasthen optimized, and the energy was minimized (13). Both

    complexes were built in dimeric form considering the fact

    that the biologically active forms of GSTs are dimeric pro-

    teins and that the glutathionyl moiety of GSH interacts with

    the side chains from both subunits (14 17). The Engh and

    Huber (18) geometric parameters were used as the basis of

    the force field. No crystal structure of class GST-bound

    GSTCD is available. The Meisenheimer complex for class

    GST was therefore built by modifying those for rGSTM1-1

    and hGSTP1-1 based on the structures of hGSTA1-1-bound

    S-benzyl-GSH (19) and the GSH conjugate of ethacrynic acid

    (20). After energy minimization, the Meisenheimer complex

    for hGSTA1-1 was docked in the active center of the enzyme,

    and the protein-Meisenheimer complex was subject to en-

    ergy minimization as described above. The molecular mod-

    eling studies were carried out with program suites O (21) and

    X-PLOR (22).

    Determination of Specific Activity of Human GSTs to-

    ward JS-K. Purified preparations of recombinant hG-

    STA1-1, hGSTM1-1, and hGSTP1-1 were obtained from

    Panvera (Madison, WI). The activity of human GST toward

    CDNB was determined, as described by Habig et al. (10),

    before activity measurements with JS-K to ensure that the

    enzyme preparations were catalytically active. For activity

    measurement toward JS-K, the reaction mixture in a final

    volume of 1 ml contained 100 mM potassium phosphate

    buffer (pH 6.5), 1 mM GSH, 0.045 mM JS-K, and an appro-

    Fig. 1. Structures of JS-K, our control arylating agent CDNB, and theircommon GSH conjugate (DNP-SG). The structures of the Meisenheimercomplexespresumed to be intermediates in the two conjugation reactionsare also shown. 4-Carbethoxy-PIPERAZI/NO (the coproduct of the GSH/

    JS-K reaction) generates NO spontaneously at physiological pH asshown; its half-life for NO release at 37C in 0.1 M phosphate buffer (pH7.4) was found to be 6.0 min (k 1.9 103 s1).

    410 Antitumor Effects of JS-K, a Nitric Oxide Prodrug

    Research.on March 22, 2013. 2003 American Association for Cancermct.aacrjournals.orgDownloaded from

    http://mct.aacrjournals.org/http://mct.aacrjournals.org/http://mct.aacrjournals.org/
  • 7/28/2019 JS-K, a Glutathione/Glutathione S-Transferase-activated Nitric Oxide Donor of the Diazeniumdiolate Class with Pote

    4/10

    priate amount of human GST isoenzyme protein. The reac-

    tion was started by the addition of JS-K, and the rate of

    reaction was monitored by measuring decrease in absorb-

    ance of JS-K at 298 nm because of its reaction with GSH.

    The specific activity toward JS-K was calculated using an

    extinction coefficient of 18 mM1

    cm1

    at 298 nm.Measurement of NO Release. Chemiluminescence de-

    tection and quantification of NO evolving from the reactions

    of JS-K were conducted using an NO-specific Thermal En-

    ergy Analyzer (Model 502A; Thermedics, Analytical Instru-

    ment Division, Waltham, MA) essentially as described previ-

    ously (23). Briefly, pH 7.4 phosphate buffer containing 1 mM

    GSH was sparged with inert gas until a steady detector

    response was established. Where indicated, GSTs were

    added to a final concentration of 1.67 g of enzyme/ml. The

    NO release profile was followed at 37C for 45 min after

    injecting JS-K at a final concentration of 133 nM to start the

    reaction. The resulting curve was integrated to quantify the

    amount of NO released/mol of compound.Cell Lines and Culture Conditions. HL-60, DLD1, and

    U937 cells were from American Type Culture Collection (Ma-

    nassas, VA). Meth A cells were from Dr. Wolfram Samlowski

    (University of Utah). The PPC-1 cell line was provided by Dr.

    Graeme Bolger (University of Alabama). For the cell growth

    and apoptosis experiments, cells were cultured at a density

    of 150,000 cells/ml in RPMI 1640 with 10% fetal bovine

    serum at 37C in a 5% CO2-humidified atmosphere. Agents

    were added at the indicated concentrations 24 h after culture

    initiation. At the indicated time intervals, cells were harvested

    and washed twice in PBS before processing for analysis of

    growth, differentiation, and apoptosis.

    Cell Growth, Differentiation, and Apoptosis Assays.

    The number of viable cells was determined using the MTT

    assay according to the manufacturers protocol (Promega,

    Madison, WI) or using a Coulter counter. Differentiation was

    evaluated using Wright and NSE staining of cells collected on

    microscope slides by cytospin as described previously (24).

    Apoptosis was assayed by flow cytometry and by determin-

    ing DNA fragmentation using agarose gel electrophoresis as

    described previously (25). For the flow cytometry assay, we

    used the propidium iodide staining method of Nicoletti et al.

    (26).

    In Vivo Studies of JS-K. NOD/SCID mice were bred and

    maintained at the Huntsman Cancer Institute at the Univer-

    sity of Utah. Experiments were performed on male or female

    mice 6 8 weeks of age at the Animal Care Facility of the Salt

    Lake City Veterans Administration Medical Center after ap-

    proval by the Institutional Animal Care and Use Committees.

    We measured systolic blood pressure on unanesthetized

    NOD/SCID mice using an occluding tail cuff and a pulse

    transducer connected to a blood pressure transducer/mon-

    itor from World Precision Instruments (Sarasota, FL). Signals

    from the blood pressure monitor and pulse transducer were

    transmitted to a MacLab2 data acquisition device (pur-

    chased from Stoelting, Wood Dale, IL) that feeds directly into

    a Macintosh computer. The recorded data were analyzed

    using the Chart data analysis software purchased from

    Stoelting. Measurements were done in triplicate at each time

    point.

    To study the in vivo antineoplastic potency of JS-K, NOD/

    SCID mice received injections in the flanks s.c. with HL-60 or

    PPC-1 cells (2.5 106 cells/flank). When s.c. tumors were

    palpable, treatment with JS-K or an equal volume of vehicle

    (20% DMSO in PBS) was started using the indicated doses

    and route. Tumor size was measured daily or every other dayusing a Vernier caliper. Tumor volume was calculated using

    the formula: width length [(width length)/2] 0.5236.

    Fifteen to 20 days after tumor cell implantation, animals were

    sacrificed by CO2

    inhalation, and tumors were collected for

    histochemical analysis.

    Histological Analysis of Tumors. At the completion of the

    experiments, animals were sacrificed, and s.c. tumors were

    dissected out, fixed in 10% formaldehyde, and imbedded in

    paraffin. Four-m sections were cut and stained with H&E.

    Calculations and Statistical Analysis. Results are ex-

    pressed as averages of multiple experiments with SE. SE

    was calculated as the SD of different measurements divided

    by the square root of the number of measurements. Differ-ences were considered statistically significant if the P was

    0.05 as calculated using the t test.

    Results

    Reactivity of JS-K with GSH. The diazeniumdiolate ion has

    been judged to resemble chloride as a leaving group in SNAr

    reactions (7). Because CDNB (structure shown in Fig. 1) is

    known to react with GSH, we anticipated that JS-K would be

    similarly converted to DNP-SG (structure also shown in Fig.

    1). This was confirmed by HPLC-MS; as shown in Fig. 2A, an

    85% conversion of JS-K to DNP-SG occurred within a 30-

    min incubation period at 37C in pH 7.4 phosphate buffer.

    Similar results were seen in RPMI 1640 cell culture. Pseudo-first-order kinetic plots for the reaction of GSH with JS-K in

    0.1 M phosphate buffer (pH 7.4) were obtained with GSH (15

    mM) in large excess of the substrate. Excellent first-order

    behavior was observed over several half-lives and measured

    first-order rate constants showed a linear dependence on

    (GSH; Fig. 2C). The slope and y intercept of the line yielded

    values for the second-order [k2 (1.02 0.04) M1 s1] and

    first-order [k1 (4 12) 10

    5 s1] rate constants for the

    reactions of JS-K with GSH and water, respectively, at 37C

    in 0.1 M phosphate buffer, pH 7.4. The UV spectral changes

    accompanying the reaction in a second cell culture medium

    (DMEM) are shown in Fig. 2B.

    Hydrolysis of JS-K. JS-K proved resistant to simple hy-

    drolysis under these conditions, as reflected in the near-zero

    y intercept of Fig. 2C. The value of k1

    obtained from the

    intercept is in statistical agreement with the rate constant for

    JS-K hydrolysis (1 106 s1) measured separately in the

    absence of GSH. The small amount of the hydrolysis product

    (2,4-dinitrophenol) seen in the chromatogram of Fig. 2A was

    apparently formed in the DMSO stock solution, which in this

    case had been stored in the refrigerator with intermittent use

    during several weeks. Hydrolysis was much more facile at pH

    12, which was expected for a compound type designed to be

    activated for NO release by nucleophilic attack.

    Solubility Limits. JS-K showed an interesting but poten-

    tially nettlesome tendency to remain supersaturated in aque-

    ous solutions, only to separate from the aqueous phase at a

    411Molecular Cancer Therapeutics

    Research.on March 22, 2013. 2003 American Association for Cancermct.aacrjournals.orgDownloaded from

    http://mct.aacrjournals.org/http://mct.aacrjournals.org/http://mct.aacrjournals.org/
  • 7/28/2019 JS-K, a Glutathione/Glutathione S-Transferase-activated Nitric Oxide Donor of the Diazeniumdiolate Class with Pote

    5/10

    time course that was difficult to predict or control. For ex-

    ample, our first attempts to follow the hydrolysis of 50 M

    JS-K spectrophotometrically sometimes showed little

    change, as expected from Fig. 2C, but at other times re-

    vealed a variable and sometimes rapid rate of absorbance

    loss without production of new maxima. The problem couldbe overcome by cleaning the cuvette with nitric acid before

    filling it with buffer then adding the DMSO/JS-K stock solu-

    tion, in which case absorbance changed little with time.

    However, rinsing this solution out with distilled water fol-

    lowed by refilling the cuvette with buffer and adding the

    stock solution often led to shrinking absorbance with con-

    comitant appearance of cloudiness in the cuvette; dissolu-

    tion of the insoluble material, after isolating it by centrifuga-

    tion, showed that the resulting solid (dissolved in acetonitrile)

    was essentially pure JS-K in an amount equivalent to the

    quantity lost from the supernatant. Quantification of the JS-K

    in the buffer by HPLC and UV spectrophotometry after all

    spectral changes had stopped indicated that JS-Ks solubil-ity limit in 0.1 M phosphate buffer at 37C and pH 7.4 con-

    taining 1% DMSO was 10 M. The relative insolubility of

    JS-K in aqueous media should be taken into account during

    any experimental work involving dilution of organic JS-K

    solutions with aqueous media because the observed tend-

    ency toward supersaturation can lead to JS-K concentra-

    tions that are much lower or higher than expected and thus

    to erroneous results.

    Catalysis of the GSH/JS-K Reaction and NO Release

    by GST. The reaction of GSH with CDNB (Fig. 1) is catalyzed

    by several classes of human GSTs, and, thus, this electro-

    philic substrate is often used for quantifying their activity.

    Given the similarity of diazeniumdiolate ions to chloride as a

    leaving group in SNAr reactions (7), we expected JS-K also to

    undergo GST-catalyzed conjugation with GSH. To gain in-

    sights into the effect the obvious steric differences between

    chloride and the diazeniumdiolate ion shown in Fig. 1 may

    imply, we modeled the accommodation of the Meisenheimer

    complex of JS-K in the active sites of the three major classes

    of human GSTs, i.e., hGSTA1-1, hGSTM1-1, and hGSTP1-1.

    As illustrated in Fig. 3, AC, both hGSTA1-1 and hGSTM1-1

    classes of GSTs accommodate the Meisenheimer complex

    very well, but hGSTP1-1 appears to have serious steric con-

    flicts with the diazeniumdiolate moiety of the transition state

    complex. On the basis of molecular modeling, we predicted

    that hGSTA1-1 and hGSTM1-1 should be more effective

    than hGSTP1-1 for catalyzing the GSH conjugation of JS-K.These predictions were confirmed by determining the activ-

    ities of recombinant hGSTA1-1, hGSTM1-1, and hGSTP1-1

    preparations toward JS-K. The data are summarized in Table 1.

    Specific activities of GSTs toward CDNB were determined be-

    fore activity measurement with JS-K to ensure that the enzyme

    preparations were catalytically active. The specific activities of

    the GSTs toward CDNB were comparable with the values pub-

    lished in the literature (Ref. 27; Table 1). In agreement with our

    max 354 nm, rather than DNP-SG, max 340 nm. C, increases in JS-Kconsumption rate at an initial concentration of 50 M in 0.1 M phosphate

    (pH 7.4) at 37C as a function of increasing GSH concentration.

    Fig. 2. Reactions of JS-K with thiols. A, HPLC traces of a reactionmixture containing 50 M JS-K and 500 M GSH at 37C in 0.1 M phos-phate (pH 7.4) at an observing wavelength of 300 nm immediately aftermixing (top) and 30 min later (bottom). The structures of JS-K and its GSHconjugate DNP-SG (extinction coefficients at 300 nm of 16 and 3.7 m M1

    cm1, respectively) are shown in Fig. 1; 2,4-DNP is the JS-K hydrolysisproduct, 2,4-dinitrophenol, and DMSO is the cosolvent. B, UV spectralchanges during 30 min at 37C of a 50 M JS-K solution in serum-freeDMEMin the absence (left) and presence (right) of 1 mM GSH. Note that in this

    cystine-containing medium, the product is S-(2,4-dinitrophenyl)cysteine,

    412 Antitumor Effects of JS-K, a Nitric Oxide Prodrug

    Research.on March 22, 2013. 2003 American Association for Cancermct.aacrjournals.orgDownloaded from

    http://mct.aacrjournals.org/http://mct.aacrjournals.org/http://mct.aacrjournals.org/
  • 7/28/2019 JS-K, a Glutathione/Glutathione S-Transferase-activated Nitric Oxide Donor of the Diazeniumdiolate Class with Pote

    6/10

    prediction, hGSTP1-1 was much less active than hGSTA1-1 or

    hGSTM1-1 for GSH conjugation of JS-K.

    That these reactions led to NO generation as predicted

    was demonstrated by purging gases from the solution as

    they formed into an NO-specific chemiluminescence detec-

    tor. The results are shown in Fig. 3D. Consistent with JS-Ks

    resistance to hydrolysis but its reactivity toward GSH as

    noted above, no NO was detected until 1 mM GSH was

    added to the pH 7.4 phosphate/0.67 M JS-K solution. In the

    absence of enzyme, NO release began immediately upon

    adding GSH, increasing in rate until plateauing at 8 min, and

    integrating to a total of 1.1 mol of NO/mol of JS-K within 43

    min of mixing. The hGSTP1-1 isoform catalyzed this reac-

    tion, but only weakly; that this was not attributable to a

    deleterious effect of JS-K exposure on the enzymes activity

    was demonstrated in its unfettered ability, in the presence of

    up to 80 M JS-K, to catalyze CDNBs conjugation with GSH.

    The hGSTA1-1 and hGSTM1-1 isoforms proved much supe-rior to hGSTP1-1 as catalysts for JS-K conjugation. The data

    of Fig. 3D on NO release are thus consistent with the

    conclusion from the JS-K consumption studies of Table 1

    that JS-K is metabolized much better by hGSTA1-1 and

    hGSTM1-1 than by hGSTP1-1.

    Growth Inhibitory Properties of JS-K. Fig. 4A shows a

    comparison between the growth inhibitory ability of JS-K and

    those of the chemotherapeutic agents daunorubicin and eto-

    poside in our HL-60 assay system. The IC50

    s of JS-K, dauno-

    rubicin, and etoposide were 0.5, 0.01, and 0.3 M, respec-

    tively. CDNB, a compound with the same aryl ring as JS-K

    that does not release NO (Fig. 1), inhibited the in vitro growth

    of HL-60 cells but at much higher concentrations, with anIC

    50estimated at 6.7 M (data not shown). JS-K also inhib-

    ited the growth of U937 (monocytic leukemia) cells with an

    IC50 of 0.3 M (data not shown). Solid tumor cell growth was

    also inhibited by JS-K, although to a lesser extent than

    leukemia cells (Fig. 5); the IC50

    s for the three lines we tested,

    PPC-1, DLD-1, and Meth A, were an order of magnitude

    greater than those for the two leukemia lines.

    To determine whether modulation of the GST pathway

    affects JS-Ks antineoplastic properties, we performed ex-

    periments using NAC or BSO. NAC increases intracellular

    GSH levels, whereas BSO inhibits its synthesis (28). Treat-

    ment of HL-60 cells with NAC (0.3 0.5 mM) or BSO (0.2 0.3

    mM) did not significantly affect cell growth. Pretreatment of

    Fig. 3. Catalysis of JS-Ks NO release by GST. AC illustrate the accommodationof the Meisenheimer complex of JS-K in the active center of (A) hGSTA1-1, (B)hGSTM1-1, and (C) hGSTP1-1 as predicted by the molecular modeling studies. Theactive centers of the GSTs are shown as surface representations and the Meisen-heimer complexes as ball-and-stick models with atomic color scheme (carbon, gray;nitrogen, blue; oxygen, red; and sulfur, yellow). The illustration was prepared usingGrasp (29) and Raster3D (30). D, chemiluminescence traces showing the time courseof NO release from 0.67 M JS-K at 37C in 0.1 M phosphate (pH 7.4) containing 1.0mM GSH both alone and with 1.67 g/ml hGSTA1-1, hGSTM11, or hGSTP11.

    Table 1 Specific activities (mol min1 mg1) of GST-catalyzed

    glutathionylation of CDNB and JS-K in 0.10 M phosphate buffer at pH6.5 and 25Ca

    Enzyme CDNB JS-K

    hGSTA1-1 34 3.7

    hGSTM1-1 200 14.9

    hGSTP1-1 (I104, A113) 95 0.15

    a The concentrations of CDNB and JS-K were 1.0 and 0.045 mM, respec-tively. The GSH concentration was 1.0 mM for both substrates. The proteinconcentrations were 0.025, 0.01, and 0.215 mg/ml for hGSTA1-1,hGSTM1-1, and hGSTP1-1, respectively.

    413Molecular Cancer Therapeutics

    Research.on March 22, 2013. 2003 American Association for Cancermct.aacrjournals.orgDownloaded from

    http://mct.aacrjournals.org/http://mct.aacrjournals.org/http://mct.aacrjournals.org/
  • 7/28/2019 JS-K, a Glutathione/Glutathione S-Transferase-activated Nitric Oxide Donor of the Diazeniumdiolate Class with Pote

    7/10

    the cells for 2 6 h with NAC prevented the JS-K-induced

    growth inhibition. Pretreatment of HL-60 cells with BSO for

    2 6 h did not prevent the JS-K-induced growth inhibition,

    whereas pretreatment of the cells with BSO for 24 h en-

    hanced it (data not shown).

    Induction of Leukemia Cell Apoptosis by JS-K. Be-

    cause we had previously shown that spontaneous and

    esterase-activated NO generators induce apoptosis in leu-

    kemia cells (2, 25), we sought to determine whether JS-K had

    a similar effect. Three days after addition of JS-K at concen-

    trations of 0.5 and 1 M, the percentage of apoptotic HL-60

    cells increased from 7 to 27 and 43%, respectively (Fig. 4B).

    The flow cytometry experiments were confirmed with DNA

    laddering assays (Fig. 4C). The pan-caspase inhibitor C-

    VAD-FMK prevented the JS-K-induced growth inhibitory and

    apoptotic effects (Fig. 4D).

    Effect of JS-K on Leukemia Cell Differentiation. We

    have previously shown that NO induces HL-60 cells to dif-

    ferentiate along the monocytic phenotype (24). We therefore

    determined whether JS-K induced differentiation as well.

    HL-60 cells were treated with JS-K at a concentration of 0.5

    M f or 35 days. Wright stain revealed morphological

    changes consistent with a monocytic phenotype, namely

    development of folded nuclei, large cytoplasms, and cyto-

    plasmic vacuoles (Fig. 6). NSE (an enzyme specific to the

    monocytic lineage) staining showed that JS-K increased the

    percentage of HL-60 cells expressing NSE from 1 to 40%

    (Fig. 6).

    Fig. 4. Antiproliferative and proapoptotic effects of JS-K on HL-60 leukemia cells in vitro. A, comparison of the antiproliferative effects of JS-K with thoseof other chemotherapeutic agents. Cells were cultured with JS-K, daunorubicin, or etoposide for 3 days at the indicated concentrations. Cell viability wasdetermined using the MTT assay (averages and SE of three separate experiments). B, apoptosis induction by JS-K. Cells were cultured with JS-K at theindicated concentrations. At 72 h, the percentage of apoptotic cells was determined by flow cytometry (see text; averages and SE of three separateexperiments). C, DNA laddering in HL-60 cells resulting from 3-day exposure to 1 M JS-K (gel representative of three different experiments; MWM,molecular weight marker). D, reversal by the caspase inhibitor C-VAD-FMK of JS-Ks cytostatic and proapoptotic effects on HL-60 cells. Cells were culturedwith JS-K (0.75 M), C-VAD-FMK (50 M), or the combination for 3 days. Cell viability and apoptosis were determined using the MTT assay and flowcytometry, respectively (averages and SE of three separate experiments).

    414 Antitumor Effects of JS-K, a Nitric Oxide Prodrug

    Research.on March 22, 2013. 2003 American Association for Cancermct.aacrjournals.orgDownloaded from

    http://mct.aacrjournals.org/http://mct.aacrjournals.org/http://mct.aacrjournals.org/
  • 7/28/2019 JS-K, a Glutathione/Glutathione S-Transferase-activated Nitric Oxide Donor of the Diazeniumdiolate Class with Pote

    8/10

    In Vivo Effects of JS-K. Before determining in vivo an-

    tineoplastic activity of JS-K, we next sought to identify a

    dose that would not induce significant hypotension in NOD/

    SCID mice. Using i.v. administration, we were able to esca-

    late the dose of JS-K up to 4 mol/kg without observing

    significant hypotension, whereas doses of 5 mol/kg or

    higher had a notable hypotensive effect (data not shown).

    Using estimates of the mouse blood volume, 4 mol/kg of

    JS-K would be expected to yield peak blood levels of 17

    M, which is far above its in vitro IC50

    .

    NOD/SCID mice were then implanted s.c. in the flanks with2.5 106 HL-60 cells/flank. When the tumors were palpable,

    treatment was started with JS-K administered i.v. via the tail

    vein three times/week at a dose of 4 mol/kg. Control mice

    received an equal volume of vehicle through the same route

    and according to the same schedule. JS-K treatment in-

    duced a significant inhibition of in vivo leukemia cell growth

    (Fig. 7A). Sixteen days after starting therapy, the average

    tumor volumes in control and JS-K-treated mice were 8.34

    0.72 and 3.13 1.14 cm3 (P 0.039), respectively, reflecting

    a 50% reduction in tumor volume in treated mice. Histo-

    logical analysis of HL-60 cell tumors obtained from vehicle-

    treated mice revealed a uniform population of densely

    packed myeloblasts. The cells were highly invasive, pene-

    trating the surrounding tissues, and showing high mitotic

    activity. Necrosis was minimal. On the other hand, histolog-

    ical analysis revealed extensive (50%) cell necrosis in

    HL-60 cell tumors obtained from the JS-K-treated mice as

    compared with 10% in controls (Fig. 7B).

    To determine whether JS-K inhibits thein vivo growth of solid

    tumor cells, NOD/SCID mice were implanted with 2.5 106

    PPC-1 (prostate carcinoma) cells and treated with 4 mol/kg

    JS-K or an equal volume of vehicle i.v. three times/week. Similar

    to the observation with HL-60 cells, JS-K treatment inhibited

    the growth of PPC-1 cells in vivo. Nineteen days after start of

    therapy, s.c. tumor implant volumes were 0.368 0.082 and

    0.107 0.053 cm3 (P 0.0073) in vehicle and JS-K-treated

    animals, respectively (Fig. 7C). Similar to HL-60 cells, PPC-1

    cells were highly aggressive and invaded the surrounding tis-

    sues. Histological analysis revealed extensive tumor necrosis in

    implants obtained from JS-K-treated animals (Fig. 7D).

    Discussion

    The results suggest that JS-K might serve as a promising

    lead compound in the search for new classes of antineoplas-

    tic agents with novel mechanisms of action. In a test of its

    antiproliferative effect on the HL-60 human myeloid leukemia

    cell line, its consistently submicromolar IC50 (three separatedeterminations of 0.2, 0.5, and 0.5 M) compared favorably

    with those of the clinical antileukemics daunorubicin (0.01

    M) and etoposide (0.3 M). Chemical characterization of

    JS-K showed that it resists hydrolysis in the absence of

    strong nucleophiles but generates copious NO upon reaction

    with GSH. The experiments additionally confirmed predic-

    tions based on molecular models of transition state geometry

    that two of three major human isoforms of GST could

    strongly catalyze NO release from JS-K.

    Consistent with our hypothesis that such an NO prodrug

    might direct its toxic action selectively to the NO-sensitive

    malignant cells while sparing less sensitive normal tissues,

    JS-K injected i.v. three times/week inhibited the growth of

    s.c. tumors formed from HL-60 xenografts in NOD-SCID

    mice by approximately one-half. This suggests that JS-K

    may serve as a useful lead for developing therapies against

    leukemia.

    As far as mechanism of action is concerned, the in vitro

    experiments showed all of the characteristics of previous

    studies with structurally different NO donors (including in-

    duction of both apoptosis and differentiation in HL-60 cells)

    and were thus consistent with NO-induced cytolysis. How-

    ever, it is important to recognize that other pathways may

    also be operative. For example, arylation of thiol groups in

    critical protein residues could be contributing to the toxicity;

    however, this may be a minor effect, as CDNB (which trans-

    fers the same 2,4-dinitrophenyl group to cellular nucleophiles

    Fig. 6. Effect of JS-K on differentiation of leukemia cells. HL-60 cellswere cultured with (right) and without (left) 0.5 M JS-K for a period of 5days. Differentiation was assessed morphologically (Wright stain, top) and

    by determining NSE (bottom) expression (photomicrographs of one ex-periment representative of three).

    Fig. 5. Inhibitory effect of JS-K on three different solid tumor cell lines.PPC-1 (prostate), DLD-1 (colon), and Meth A (mammary) cells were cul-tured with JS-K at the indicated concentrations for 3 days. Cell viabilitywas determined using the MTT assay (averages and SE of three separate

    experiments).

    415Molecular Cancer Therapeutics

    Research.on March 22, 2013. 2003 American Association for Cancermct.aacrjournals.orgDownloaded from

    http://mct.aacrjournals.org/http://mct.aacrjournals.org/http://mct.aacrjournals.org/
  • 7/28/2019 JS-K, a Glutathione/Glutathione S-Transferase-activated Nitric Oxide Donor of the Diazeniumdiolate Class with Pote

    9/10

    as JS-K but without generating any NO) was an order of

    magnitude less potent in the HL-60 screen. Whether there is

    synergy between the growth inhibitory effects of NO and

    arylation remains to be determined. Transcarbamoylation is

    also a possible route to toxicity. The finding that N-acetyl-L-

    cysteine prevented the growth inhibitory effects of JS-K

    while BSO enhanced them suggests that one or more reac-

    tive nitrogen/oxygen species (NO and/or the product of NO s

    reaction with other reactive oxygen species) play a key role

    in effecting the antineoplastic activity of JS-K.

    As to the speci fic molecular targets that may be in-

    volved, exposure to JS-K can in principle render any crit-

    ical thiol group nucleophilically unreactive by S-nitrosation

    or S-(2,4-dinitrophenyl)ation. It is likely that the cysteine-

    containing caspases are directly or indirectly among the

    mechanistically important targets of JS-Ks action be-

    cause the nonspecific caspase inhibitor C-VAD-FMK re-

    duced JS-Ks cytostatic and apoptotic effects on HL-60

    cells. Other pathways are possible, though, and a fuller

    explanation of the drugs antineoplastic activity must

    await the outcome of future hypothesis testing.

    Whatever the mechanism(s) that may be involved, the

    present results support the choice of JS-K as a worthy

    lead compound for additional drug discovery efforts. The

    significant in vivo activity demonstrated in Fig. 7 was seen

    in first generation experiments that could hardly be viewed

    as optimized; the dose regimen was chosen based on

    preliminary studies of bolus sizes that could be adminis-

    tered without inducing systemic hypotension or other

    manifestations of toxicity. It is entirely possible that a

    continuous administration schedule would be much more

    effective. Structural modification aimed at improving the

    problematic solubility of this compound could beneficially

    affect its absorption, distribution, and transport proper-

    ties. With substantial efforts underway in other laborato-

    ries4 as well as our own aimed at further characterizing

    and refining the chemotherapeutic potential of this inter-

    esting lead, we are hopeful that our ultimate goal of intro-

    4 L. Jia; K. Tew and V. Findlay; J. Liu and M. Waalkes, personal commu-

    nications.

    Fig. 7. Effects of i.v. JS-K on growth of both human leukemia and solid tumor xenografts in NOD-SCID mice. A, growth rate of tumors produced when2.5 106 HL-60 cells were implanted s.c. in NOD/SCID mice [3 animals each for JS-K and vehicle treatments]. When tumors became palpable, JS-K wasadministered i.v. at a dose of 4 mol/kg three times/week. Tumor volumes were measured at regular intervals. At the time of experiment termination, tumor

    implants were dissected out and analyzed histologically. Differences between vehicle and JS-K values from days 9 16 were significant at the P 0.05 level.B, JS-K induced extensive necrosis in these tumors (photomicrograph from 1 animal representative of 3; H&E stain, 100 magnification). C, growth rateof tumors produced when 2.5 106 human prostate carcinoma (PPC-1) cells were implanted s.c. in NOD/SCID mice (5 animals each for JS-K and vehicletreatments) and followed as in A above. Differences between vehicle and JS-K values were significant at the P 0.05 level for all time points after day 1.D, JS-K induced extensive necrosis in these tumors (photomicrograph from 1 animal representative of 5; H&E stain,100 magnification).

    416 Antitumor Effects of JS-K, a Nitric Oxide Prodrug

    Research.on March 22, 2013. 2003 American Association for Cancermct.aacrjournals.orgDownloaded from

    http://mct.aacrjournals.org/http://mct.aacrjournals.org/http://mct.aacrjournals.org/
  • 7/28/2019 JS-K, a Glutathione/Glutathione S-Transferase-activated Nitric Oxide Donor of the Diazeniumdiolate Class with Pote

    10/10

    ducing new and improved treatments for AML and other

    malignant diseases will be significantly advanced.

    References

    1. Greer, J. P., Baer, and Kinney, M. C. In: G. R. Lee, J. Foerster, J.

    Lukens, F. Paraskevas, J. P. Greer, and G. M. Rodgers (eds.), Wintrobes

    Clinical Hematology, Ed. 10, pp. 22722319. Baltimore: Williams and

    Wilkins, 1999.

    2. Saavedra, J. E., Shami, P. J., Wang, L. Y., Davies, K. M., Booth, M. N.,

    Citro, M. L., and Keefer, L. K. Esterase-sensitive nitric oxide donors of the

    diazeniumdiolate family. In vitro antileukemic activity. J. Med. Chem., 43:

    261269, 2000.

    3. Mooradian, D. L., Hutsell, T. C., and Keefer, L. K. Nitric oxide (NO)

    donor molecules: effect of NO release rate on vascular smoothmusclecell

    proliferation in vitro. J. Cardiovasc. Pharmacol., 25: 674 678, 1995.

    4. Saavedra, J. E., Billiar, T. R., Williams, D. L., Kim, Y-M., Watkins, S. C.,

    and Keefer, L. K. Targeting nitric oxide (NO) delivery in vivo. Design of a

    liver-selective NO donor prodrug that blocks tumor necrosis factor--

    induced apoptosis and toxicity in the liver. J. Med. Chem., 40: 19471954,

    1997.

    5. Bico, P., Chen, C. Y., Jones, M., Erhardt, J., and Dirr, H. Class piglutathione S-transferase: Meisenheimer complex formation. Biochem.

    Mol. Biol. Int., 33: 887 892, 1994.

    6. Sargent,J. M., Williamson, C., Hall, A.G., Elgie, A.W., and Taylor, C. G.

    In: G. J. L. Kaspers, R. Pieters, and A. J. P. Veerman (eds.), Drug Resist-

    ance in Leukemia and Lymphoma, III, pp. 205 209. New York: Kluwer

    Academic/Plenum Publishers, 1999.

    7. Saavedra, J. E., Srinivasan, A., Bonifant, C. L., Chu, J., Shanklin, A. P.,

    Flippen-Anderson, J. L., Rice, W. G., Turpin, J. A., Davies, K. M., and

    Keefer, L. K. The secondary amine/nitric oxide complex ion R2N[N(O)NO]-

    as nucleophile and leaving group in SNAr reactions. J. Org. Chem., 66:

    3090 3098, 2001.

    8. Saavedra, J. E., Booth, M. N., Hrabie, J. A., Davies, K. M., and Keefer,

    L. K. Piperazine as a linker for incorporating the nitric-oxide releasing

    diazeniumdiolate group into other biomedically relevant functional mole-

    cules. J. Org. Chem., 64: 5124 5131, 1999.

    9. Mancini, I., Guella, G., Chiasera, G., and Pietra, F. Glutathione S-

    transferase catalysed dehalogenation of haloaromatic compounds which

    lack nitro groups near the reaction centre. Tetrahedron Lett., 39: 1611

    1614, 1998.

    10. Habig, W. H., Pabst, M. J., and Jakoby, W. B. Glutathione S-trans-

    ferases. The first enzymatic step in mercapturic acid formation. J. Biol.

    Chem., 249: 7130 7139, 1974.

    11. Ji, X., Armstrong, R. N., and Gilliland, G. L. Snapshots along the

    reaction coordinate of an SNAr reaction catalyzed by glutathione trans-

    ferase. Biochemistry, 32: 12949 12954, 1993.

    12. Prade, L., Huber, R., Manoharan, T. H., Fahl, W. E., and Reuter, W.

    Structures of class pi glutathione S-transferase from human placenta in

    complex with substrate, transition-state analogue and inhibitor. Structure

    (Lond.), 5: 12871295, 1997.

    13. Powell, M. J. D. Restart procedures for the conjugate gradient

    method. Math. Prog., 12: 241254, 1977.

    14. Armstrong, R. N. Glutathione S-transferases: structure and mecha-

    nism of an archetypical detoxication enzyme. Adv. Enzymol. Relat. Areas

    Mol. Biol., 69: 1 44, 1994.

    15. Wilce, M. C. J., and Parker, M. W. Structure and function of glutathi-

    one S-transferases. Biochim. Biophys. Acta, 1205: 118, 1994.

    16. Dirr, H., Reinemer, P., and Huber, R. X-ray crystal structures of

    cytosolic glutathione S-transferases. Implications for protein architecture,

    substrate recognition and catalytic function. Eur. J. Biochem., 220: 645

    661, 1994.

    17. Armstrong, R. N. Structure, catalytic mechanism, and evolution of the

    glutathione transferases. Chem. Res. Toxicol., 10: 218, 1997.

    18. Engh, R. A., and Huber, R. Accurate bond and angle parameters for

    X-ray protein structure refinement. Acta Crystallogr., 47: 392 400, 1991.

    19. Sinning, I., Kleywegt, G. J., Cowan, S. W., Reinemer, P., Dirr, H. W.,

    Huber, R., Gilliland, G. L., Armstrong, R. N., Ji, X., Board, P. G., Olin, B.,

    Mannervik, B., and Jones, T. A. Structure determination and refinement of

    human alpha class glutathione transferase A1-1, and a comparison with

    the mu and pi class enzymes. J. Mol. Biol., 232: 192212, 1993.

    20. Cameron, A. D., Sinning, I., LHermite, G., Olin, B., Board, P. G.,

    Mannervik, B., and Jones, T. A. Structural analysis of human alpha-class

    glutathione transferase A1-1 in the apo-form and in complexes with

    ethacrynic acid and its glutathione conjugate. Structure (Lond.), 3: 717727, 1995.

    21. Jones, T. A., and Kjeldgaard, M. Electron-density map interpretation.

    Methods Enzymol., 277: 173208, 1997.

    22. Brunger, A. T., and Rice, L. M. Crystallographic refinement by simu-

    lated annealing: methods and applications. Methods Enzymol., 277: 243

    269, 1997.

    23. Keefer, L. K., Nims, R. W., Davies, K. M., and Wink, D. A. NONOates

    (1-substituted diazen-1-ium-1,2-diolates) as nitric oxide donors: conven-

    ient nitric oxide dosage forms. Methods Enzymol., 268: 281293, 1996.

    24. Magrinat, G., Mason, S. N., Shami, P. J., and Weinberg, J. B. Nitric

    oxide modulation of human leukemia cell differentiation and gene expres-

    sion. Blood, 80: 1880 1884, 1992.

    25. Shami, P. J., Sauls, D. L., and Weinberg, J. B. Schedule and concen-

    tration-dependent induction of apoptosis in leukemia cells by nitric oxide.

    Leukemia (Baltimore), 12: 14611466, 1998.

    26. Nicoletti, I., Migliorati, G., Pagliacci, M. C., Grignani, F., and Riccardi,

    C. A rapid and simple method for measuring thymocyte apoptosis by

    propidium iodide staining and flow cytometry. J. Immunol. Methods, 139:

    271279, 1991.

    27. Hayes, J. D., and Pulford, D. J. The glutathione S-transferase super-

    gene family: regulation of GST and the contribution of the isoenzymes to

    cancer chemoprotection and drug resistance. Crit. Rev. Biochem. Mol.

    Biol., 30: 445 600, 1995.

    28. Anderson, M. E. Glutathione: an overview of biosynthesis and mod-

    ulation. Chem. Biol. Interact., 111112: 114, 1998.

    29. Nicholls, A., Sharp, K. A., and Honig, B. Protein folding and associ-

    ation: insights from the interfacial and thermodynamic properties of hy-

    drocarbons. Proteins Struct. Funct. Genet., 11: 281296, 1991.

    30. Merritt, E. A., and Bacon, D. J. Raster3D: photorealistic molecular

    graphics. Methods Enzymol., 277: 505524, 1997.

    417Molecular Cancer Therapeutics