U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Probe Reports from the NIH Molecular Libraries Program [Internet]. Bethesda (MD): National Center for Biotechnology Information (US); 2010-.

Cover of Probe Reports from the NIH Molecular Libraries Program

Probe Reports from the NIH Molecular Libraries Program [Internet].

Show details

ML265: A potent PKM2 activator induces tetramerization and reduces tumor formation and size in a mouse xenograft model

, , , , , , , , , , , , , , , , , , , , and .

Author Information and Affiliations

Received: ; Last Update: May 8, 2013.

Cancer cells have altered metabolic processes compared to normal differentiated cells and the expression of the M2 isozyme of pyruvate kinase (PKM2) plays an important role in this aberrant metabolism. The M1 isoform is a highly active enzyme typically expressed in muscle and brain tissue, the alternatively spliced M2 variant is considerably less active and expressed in many tumors studied to date. This report describes the use of the PKM2 activator, ML265, and details some of the biophysical, ex vivo and in vivo activity of this compound. ML265 induces the more active tetrameric state of PKM2 and the X-ray co-crystal structure shows that the activator binds at the dimer-dimer interface between two subunits of PKM2. This compound was tested in a H1299 mouse xenograft model and showed significant reduction in tumor size, weight, and occurrence with no apparent toxicity over the 7-week experiment.

Screening Center Name & PI: NIH Chemical Genomics Center, Christopher P. Austin

Chemistry Center Name & PI: NIH Chemical Genomics Center, Christopher P. Austin

Assay Submitter & Institution: Matthew G. Vander Heiden, Koch Institute for Integrative Cancer Research at Massachusetts Institute of Technology and Dana Farber Cancer Institute, Harvard Medical School

PubChem Summary Bioassay Identifier (AID): 602359

Probe Structure & Characteristics

Image ml265fu1
CID/ML#Target NameIC50/EC50 (nM) [SID, AID]Anti-target Name(s)IC50/EC50 (μM) [SID, AID]Fold SelectiveSecondary Assay(s) Name: IC50/EC50 (nM) [SID, AID]
CID 44246499/ML265hPyk-M292 [SID 85267885, AID 602384]hPyk-M1, -R, and -LInactive against all 3 isoforms [SID 85267885, AID 602383, AID 602381, AID 602379]>100 fold over all 3 isoformsLDH assay290
[SID 85267885, AID 602377]

Recommendations for Scientific Use of the Probe

ML265 is a potent activator of PKM2 in both biochemical (AC50 = 92 nM) and cell-based assays with high selectivity over PKM1, PKR and PKL. The compound was found to induce tetramerization based upon size exclusion chromatography, sucrose gradient ultracentrifugation and immunoprecipitation experiments. A high resolution X-ray structure revealed that ML265 binds at the dimer-dimer interface of the PKM2 homotetramer. Importantly, this compound is capable of activating PKM2 in cell lysate of pervanadate treated cells, which is a condition known to inhibit PKM2 activity through accumulation of phosphotyrosine peptides. ML265 significantly increased the doubling time of H1299 cells under hypoxic conditions, but interestingly showed no effect under normoxia. PKM2 expression has also been shown to effect PGAM1 phosphorylation and ML265 was able to reduce this phosphorylation event. After we determined that ML265 had an acceptable in vitro microsomal stability profile, it was advanced into in vivo PK experiments and ultimately showed a reduction in tumor size, weight, and occurrence in a H1299 mouse xenograft model.

1. Introduction

Cancer cells exhibit altered metabolism and utilization of extracellular nutrients, thereby providing a unique strategy to effect tumor growth17. Tumors are known to import significantly higher amounts of glucose compared to normal tissue and utilize glucose’s carbons as cellular building blocks for proliferation3. Associated with this enhanced glucose uptake, the expression of the M2 isoform of pyruvate kinase (PKM2) is a factor contributing to biosynthesis and tumor growth8. ShRNA replacement of PKM2 with the M1 splice variant (PKM1) does not support rapid proliferation and increases cellular doubling time8.

Catalyzing the final step in glycolysis, pyruvate kinase transfers the phosphate group from phosphoenolpyruvate (PEP) to adenosine diphosphate (ADP) to yield adenosine triphosphate (ATP) and pyruvate. Two genes are responsible for encoding the four pyruvate kinase isozymes; encoded by the Pkrl gene, PKL is mainly expressed in the liver and PKR is expressed in red blood cells. Most tissues express either the PKM1 or PKM2 isoform, which are both encoded by the Pkm gene. PKM1 is found in the majority of normal tissues, but the PKM2 isoform is expressed in highly proliferative cells including all cancer cell lines and tumors studied to date9. The sole difference between the PKM1 and PKM2 isoforms is a 56 amino acid stretch of which only 21 differ between the two isozymes1012. This distinction, which arises from alternative splicing by mutual exclusion of a single exon, causes substantially different catalytic and regulatory properties between the two isoforms. The encoding of the 21 unique amino acids in PKM1 result in an enzyme that is not allosterically regulated and exhibits high constitutive enzymatic activity13. In the less active PKM2 isoform, these 21 amino acids create an allosteric site where the upstream glycolytic intermediate, fructose-1,6-bisphosphate (1,6-FBP) binds and activates the enzyme in a feed-forward mechanism14. Evidence suggests that 1,6-FBP binding induces homotetramer formation, which comprises the most active form of the enzyme15,16.

Of the four pyruvate kinase isoforms, PKM2 is the only isozyme for which peptides and proteins with phosphorylated tyrosine residues inhibit its activity by causing the release of FBP1719. Additionally, PKM2, but not PKM1, is inhibited under elevated intracellular reactive oxygen species (ROS) via direct oxidation of cysteine 35820. It has also been shown that there is a correlation between PKM2 expression and enhanced phosphorylation of H11 on phosphoglycerate mutase 1 (PGAM1) by PEP21. This seems to be a mechanism for cells to generate pyruvate without elevating ATP levels, thus allowing glycolysis to proceed at a high rate22. In H1299, A549 and SN12C cell lines, shRNA replacement of PKM2 with PKM1 reversed many characteristics of the Warburg effect including reduced lactate production and enhanced oxygen consumption8. Additionally, PKM1 replacement of PKM2 resulted in a decrease of PEP phosphorylation on H11 of PGAM121. PKM2 expression appears to be critical for tumor formation as all tumors that formed in a mouse xenograft models initiated with PKM1 expressing H1299 cells had re-expressed PKM28. Though a number of non-metabolic roles for PKM2 have recently emerged9, one hypothesis is that proliferation is supported by PKM2 expression due to its low activity, and that pharmacological activation of this enzyme to the levels of PKM1 could have anti-proliferative affects. In an effort to test this theory, we have identified a number of small molecules capable of potent and selective activation of PKM2 in vitro2325. Here we show that one of our PKM2 activators, ML265, can increase pyruvate kinase activity in cells to comparable levels of PKM1 expressing cells and has appropriate pharmacokinetic properties enabling an in vivo mouse xenograft experiment.

Prior Art

We have previously described 3 unique chemotypes capable of potent in vitro activation of PKM22325 including a bis-sulfonamide series24 (Figure 1), a thieno-pyrrole-pyridazinone series23 and a tetrahydroquinoline-6-sulfonamide series25. Nice SAR and numerous analogs with potencies <100 nM were described for all series. At the time the studies described herein were initiated, the tetrahydroquinoline-6-sulfonamide series had not been fully developed and only the bis-sulfonamide and thieno-pyrrole-pyridazinone series were compared. The thieno-pyrrole-pyridazinone series was chosen for the detailed in vitro experiments and in vivo work due to its superior in vitro ADME profile compared to ML203 (data not shown).

Figure 1. Previous PKM2 activator ML probes.

Figure 1

Previous PKM2 activator ML probes.

2. Materials and Methods

General Chemistry Methods

All air or moisture sensitive reactions were performed under positive pressure of nitrogen with oven-dried glassware. Anhydrous solvents such as dichloromethane (DCM), N,N-dimethylforamide (DMF), acetonitrile, methanol and triethylamine, trifluoroacetic acid (TFA) were purchased from Sigma-Aldrich. Preparative purification was performed on a Waters semi-preparative HPLC system. The column used was a Phenomenex Luna C18 (5 micron, 30 × 75 mm) at a flow rate of 45 mL/min. The mobile phase consisted of acetonitrile and water (each containing 0.1% trifluoroacetic acid). A gradient of 10% to 50% acetonitrile over 8 minutes was used during the purification. Fraction collection was triggered by UV detection (220 nM). Analytical analysis was performed on an Agilent LC/MS (Agilent Technologies, Santa Clara, CA). Method 1: A 7 minute gradient of 4% to 100% Acetonitrile (containing 0.025% trifluoroacetic acid) in water (containing 0.05% trifluoroacetic acid) was used with an 8 minute run time at a flow rate of 1 mL/min. A Phenomenex Luna C18 column (3 micron, 3 × 75 mm) was used at a temperature of 50 °C. Method 2: A 3 minute gradient of 4% to 100% Acetonitrile (containing 0.025% trifluoroacetic acid) in water (containing 0.05% trifluoroacetic acid) was used with a 4.5 minute run time at a flow rate of 1 mL/min. A Phenomenex Gemini Phenyl column (3 micron, 3 × 100 mm) was used at a temperature of 50 °C. Purity determination was performed using an Agilent Diode Array Detector for both Method 1 and Method 2. Mass determination was performed using an Agilent 6130 mass spectrometer with electrospray ionization in the positive mode. 1H NMR spectra were recorded on Varian 400 MHz spectrometers. Chemical Shifts are reported in ppm with undeuterated solvent (DMSO-h6 at 2.49 ppm) as internal standard for DMSO-d6 solutions. All of the analogs tested in the biological assays have a purity greater than 95% based on both analytical methods. High resolution mass spectrometry was recorded on Agilent 6210 Time-of-Flight LC/MS system. Confirmation of molecular formulae was accomplished using electrospray ionization in the positive mode with the Agilent Masshunter software (version B.02).

2.1. Assays

Luminescent pyruvate kinase activation assay

The primary pyruvate kinase activation assay was performed in 1536-well white microtiter plates using 4 μL/well assay volume with final concentrations of 0.1 nM human PKM2, 0.4 nM PKM1, 3 nM PKL, or 0.6 nM PKR, 0.5 mM PEP, and 0.1 mM ADP in assay buffer containing 50 mM Imidazole pH 7.2, 50 mM KCl, 7 mM MgCl2, 0.01% Tween, and 0.05% BSA. Compounds were dissolved in DMSO and 23 nL were transferred with a 1536-well pintool. The enzymatic reaction was allowed to proceed for 60 minutes, and then ATP was detected using a bioluminescent detection reagent containing D-luciferin and firefly luciferase (KinaseGlo Plus, Promega). Luminescence was then measured on the Perkin Elmer ViewLux using an exposure of 1 sec (with 1× binning). Kinase activity levels were graphed using GraphPad Prism to determine initial velocities (V0), which were then plotted for each set of conditions.

Confirmatory LDH kinetic assay

All compounds were also tested in a kinetic mode by coupling the generation of pyruvate by pyruvate kinase to the depletion of NADH through lactate dehydrogenase. For PKM2, 3 μL of substrate mix (final concentration, 50 mM Tris-Cl pH 8.0, 200 mM KCl, 15 mM MgCl2, 0.1 mM PEP, 4.0 mM ADP, and 0.2 mM NADH) were dispensed into Kalypsys black-solid 1,536 well plates using the Aurora Discovery BioRAPTR Flying Reagent Dispenser (FRD; Beckton-Dickenson, Franklin Lakes, NJ) and 23 nL of compounds were delivered via a Kalypsys pin tool; then, 1 μL of enzyme mix (final concentrations, 10 nM PKM2 and 1 μM of LDH) was added. Plates were immediately placed in ViewLux (Perkin Elmer) and NADH fluorescence was determined at 30 second exposure intervals for between 3 and 6 minutes. Data were normalized to the uninhibited and EC100 activation using known activators, such as fructose-1,6-bis-phosphate. The data has been deposited in PubChem (AID 602377).

PKM2 Km determination

M2 substrate Km values were determined using a modified time course version of the primary luminescence assay. The assay was performed in 1536-well white microtiter plates using 4 μL/well assay volume with final concentrations of 5 nM human PKM2, and 4 mM of either PEP or ADP in assay buffer containing 50 mM Tris pH 7.5, 100 mM KCl and 5 mM MgCl2. The non-titrated substrate in each reaction was included at non-limiting concentrations (4 mM), and the opposing substrate was diluted 1:2 in assay buffer to give an 8-pt titration (final concentration range 6.0 mM - 0.02 mM). Activator (either 1,6-FBP, ML265, or DMSO control) was added to the PKM2 solution at a final concentration of 10 μM, and the resulting solution was then added to the substrate solution to start the reaction. The PKM2 reactions were stopped (and ATP levels quantified), at 0, 5 and 10 min timepoints by addition of KinaseGlo Plus bioluminescent detection reagent. Luminescence was then measured on the Perkin Elmer ViewLux using an exposure of 1 sec (with 1× binning). M2 activity levels at each timepoint were graphed using GraphPad Prism to determine initial velocities (V0), which were then plotted for each set of conditions. Curves were fit and Km values calculated using an allosteric sigmoidal nonlinear regression algorithm in Prism.

Cell lines

293T (human embryonic kidney) and A549 (human non-small cell lung carcinoma) cells were obtained from ATCC and cultured in DMEM (Mediatech) supplemented with 10% fetal calf serum (FCS), 2 mM glutamine, 100 U/mL penicillin and 100 μg/mL streptomycin. H1299 (human non-small cell lung carcinoma), T.T (human medullary thyroid carcinoma), SN12C (human renal cell carcinoma), and SKMel28 (human melanoma) were obtained from ATCC and cultured in RPMI (Mediatech) supplemented as above. All cells were cultured in a humidified incubator at 37 °C/5% CO2 unless otherwise stated. Glucose concentration in the media was 25 mM (4.5 g/L) unless otherwise stated. Hypoxia treatments were performed using an InVivo2 400 humidified workstation (Ruskinn, Pencoed, UK). For all hypoxia treatments (and corresponding normoxic control cultures), the media were supplemented with 20 mM HEPES buffer. Cells expressing specific Flag-tagged isoforms of mouse pyruvate kinase M, or mutants thereof, in the absence of endogenous PKM2 were derived as described8.

Western Blot

Cells or tissues were lysed in RIPA buffer with 2 mM DTT and protease inhibitors and clarified by centrifugation at 21,000 × g. The protein content of supernatants was quantified by Bradford assay, and analyzed by SDS-PAGE followed by Western blot using standard protocols and the following primary antibodies: anti-pyruvate kinase (Abcam, ab6191), anti-PKM1 (Sigma, SAB4200094), anti-PKM2 (Cell Signaling, 4053), anti-FLAG (Sigma, F3165), and anti-actin (Abcam, ab1801). Iso-electric focusing/SDS-PAGE two-dimensional Western blot analysis was performed as described previously20 after incubation of cells with DMSO or drug as indicated for one hour.

Immunoprecipitation

Cells attached to culture dishes were quickly washed once with a large volume (20–30 mL) of ice-cold PBS, snap-frozen in a liquid nitrogen bath and stored at −80 °C until further processing. Cells were lysed in 700 μL PK lysis buffer (50 mM Tris-HCl pH 7.5, 1 mM EDTA, 150 mM NaCl, 1% Igepal-630) supplemented freshly prior to use with protease inhibitors [10 μg/mL phenymethylsulfonyl fluoride, 4 μg/mL aprotinin, 4 μg/mL leupeptin, and 4 μg/mL pepstatin (pH 7.4)] and 1 mM dithiothreitol (DTT). Lysates were centrifuged (20,000xg, 10 min, at 4 °C), supernatants were transferred to fresh Eppendorf tubes containing 20 μL of 50% Flag-agarose (Sigma-A2220) bead slurry in PK lysis buffer and incubated rotating at 4 °C for 1 hr. Under these conditions, lysates were immunodepleted of detectable Flag-tagged proteins. Immunoprecipitates were washed 4 times with PK lysis buffer (1 mL = 100 bead-volumes per wash) then eluted from beads with 3xFlag peptide (150 μg/mL final concentration, Sigma, F4799, dissolved in 50 mM Tris-HCl pH 7.4, 150 mM NaCl) for 30 min rotating at 4 °C. Following a brief centrifugation of the beads, eluates were transferred to fresh Eppendorf tubes, supplemented with SDS-PAGE loading buffer and analyzed by SDS-PAGE.

Xenograft Experiments

H1299 parental and H1299 cells with constitutive expression of a mouse PKM1 cDNA (H1299-PKM1 cells) were propagated in RPMI supplemented with 10% fetal bovine serum, penicillin/streptomycin, 2 mM glutamine, and hygromycin for transgene selection. Cells were harvested, resuspended in sterile PBS, and 5×106 cells were injected subcutaneously into nu/nu mice. Tumor growth was monitored via caliper measurement, the mice were sacrificed and tumors harvested after the time indicated. Tumors were weighed, divided, and either flash-frozen in liquid nitrogen or fixed in formalin for later analysis.

PKM2 activity assay

Pyruvate kinase activity was measured by monitoring pyruvate-dependent conversion of NADH to NAD+ by lactate dehydrogenase (LDH). Briefly, for cell line experiments, the medium was replaced with fresh medium 1 hr prior to the start of treatment with DMSO or activator. Also, where indicated, 100 μM pervanadate was added 10 min. prior to cell lysis. Cells were lysed on ice with RIPA buffer containing 2 mM DTT and protease inhibitors and clarified by centrifugation at 21,000 × g. 5 μL of the supernatant was used to assess pyruvate kinase activity26. Pyruvate kinase activity was subsequently normalized for total protein content. For the experiment in Figure 4, the amino acid sequences of the peptides were: GGAVDDDYAQFANGG (M2tide) and GGAVDDDpYAQFANGG (P-M2tide)17.

Figure 4. ML265 activates PKM2 in the presence of phosphorylated peptides.

Figure 4

ML265 activates PKM2 in the presence of phosphorylated peptides. a) Size exclusion chromatography of recombinant PKM2 after incubation with an M2 binding peptide motif (M2tide) or phosphorylated version of the same peptide motif (pM2tide). b) ML265’s (more...)

Sucrose-gradient ultracentrifugation

Approximately 2.3 mL of a 10–40% sucrose gradient was layered in ultracentrifuge tubes and left to equilibrate overnight. Recombinant PKM2 was incubated with drug or 1 mM FBP for 30 minutes on ice before layering on top of the gradient. The gradients were then spun at 55,000 rpm for 10 hours using a Beckman TLS-55 rotor and fractions collected. The resulting fractions analyzed by SDS-PAGE and Coomassie blue staining. Intensity of staining was quantified using IR fluorescence on a LiCOR instrument.

Size exclusion chromatography

Recombinant protein or ~2 mg of cellular protein from hypotonically lysed cells was separated on a HiPrep 16/60 Sephacryl S-200 HR column (GE) in 50 mM sodium phosphate, 150 mM sodium chloride, pH 7.2. Fractions were analyzed by UV absorbance or SDS-PAGE with Western blot as indicated. For the experiment in Figure 4 the peptides used are the same as described under PKM2 activity assay.

Crystallography methods

For each complex crystal 360 0.5° oscillation images were collected in a continuous sweep. Data were reduced with HKL-200027 (DASA-58) or HKL-300028 (TEPP-46). Crystals belonged to space group P21. Structures were solved by direct replacement with the isomorphous Protein Data Bank (PDB)29 entry 3GQY. Activator geometry restraints were obtained at the PRODRG30 server. Iterations of model rebuilding, refinement and geometry validation were performed with COOT31, REFMAC32, and MOLPROBITY33, respectively. The refined models were deposited34 in the PDB.

Cloning, Expression, and Purification of Full Length PKM2

The cDNA sequence of full-length human PKM2 was amplified by PCR and cloned into the expression plasmid pET28a-LIC, and the construct was verified by DNA sequencing. The resultant plasmid was transformed into Escherichia coli strain BL21(DE3) and the cells were grown in Terrific Broth at 37 °C with 50 μg/mL kanamycin. When the OD 600 reached 1.0, the cells were induced by the addition of 0.5 mM isopropyl β-D-thiogalactoside for overnight at 18 °C and the recombinant PKM2 was expressed as a histag fusion protein containing 19 N-terminal amino acid residues (MGSSHHHHHHSSGLVPRGS). Cells were harvested by centrifugation, frozen in liquid nitrogen, and stored at −80 °C. The thawed cell pellet was resuspended in binding buffer (10 mM HEPES, pH 7.5; 300 mM KCl; 5 mM imidazole, 5 mM MgCl2; 5% glycerol; 2.5 mM TCEP), supplemented with protease inhibitor cocktail (Sigma) and 0.5% CHAPS, and then lysed by sonication. The lysate was clarified by centrifugation, and the supernatant was loaded onto DE52 column (Whatman) pre-equilibrated with binding buffer. The collected flow-through was loaded again onto Ni-NTA gravity flow column (Qiagen) equilibrated with binding buffer, and the bound histagged PKM2 protein was then eluted with elution buffer (10 mM HEPES, pH 7.5; 300 mM KCl; 5 mM imidazole, 5 mM MgCl2; 5% glycerol; 2.5 mM TCEP). The fractions eluted from Ni-NTA column were pooled, concentrated, and loaded onto a HiLoad 26/60 Superdex 200 column (Amersham Biosciences) pre-equilibrated with 10 mM HEPES, pH 7.5; 100 mM KCl; 5 mM MgCl2; 5% glycerol; 2.5 mM TCEP. The fractions containing PKM2 protein were collected, concentrated to 20 mg/mL using Amicon Ultra-15 centrifugal filter device (Millipore), and stored at −80°C until use. The final protein purity was confirmed by SDS-PAGE.

Protein Crystallization and structure determination

For co-crystallization, the purified PKM2 was incubated at room temperature for overnight in the presence of 5 mM activators (TEPP-46 or DASA-58) and then set up crystallization trays using the sitting-drop vapor diffusion method with droplets of protein solution (0.5 μL) and reservoir solution (0.5 μL). The best diffracting crystals were obtained within 2 days from a reservoir solution containing 25% P3350, 0.2 M NH4OAc and 0.1 M Bis-Tris pH 6.5. Crystals were cryo-protected with 50% Paratone-N, 50% mineral oil and were frozen directly in liquid nitrogen for x-ray data collection. Data collection was carried out at the Advanced Photon Source beamline 23ID-B.

Cell doubling time

20,000 cells were seeded in 12-well plates at day -1 in media containing 5.6 mM D-glucose and incubated at the indicated O2 concentrations. In all cases media contained 10% dialyzed FCS and 20 mM HEPES pH 8.0. At day 0 the media were replaced with fresh media that had been equilibrated since the time of cell seeding at 37 °C under the corresponding oxygen concentrations. At days 0, 2, 4 and 6 cells were fixed with 10% formalin and at the end of the experiment stained with 0.1% w/v crystal violet in 20% methanol, shaking for 15 min at room temperature, and washed with water twice for 10 min each; the plates were then dried on air. Cell-bound crystal violet was solubilized in 1 mL 10% v/v acetic acid and, because the amount of dye bound to cells is proportional to the number of cells, accumulation of cell mass was assessed by measurement of crystal violet absorbance at 595 nm in a spectrophotometer. Doubling times were calculated using exponential regression (http://www.doubling-time.com/compute.php).

LC/MS/MS Tandem mass spectrometry

For all mass spectrometry (MS) experiments, pyruvate kinase immunoprecipitates were separated using SDS-PAGE, the gel was stained with Coomassie blue, destained, and the pyruvate kinase band was excised. Samples were washed with 50% acetonitrile, subjected to reduction with 10 mM dithiothreitol (DTT) for 30 minutes, alkylation with 55 mM iodoacetamide with 45 minutes, and in-gel digestion with TPCK modified trypsin (Promega) or chymotrypsin (Princeton Scientific) overnight at pH 8.3, followed by reversed-phase microcapillary/tandem mass spectrometry (LC/MS/MS). LC/MS/MS was performed using an EASY-nLC splitless nanoflow HPLC (Thermo Fisher Scientific) with a self-packed 75 μm id × 15 cm C18 Picofrit column (New Objective) coupled to a LTQ-Orbitrap XL mass spectrometer (Thermo Scientific) in data-dependent positive ion mode at 300 nL/min with one full MS-FT scan followed by six MS/MS-IT scans via collision induced dissociation (CID). MS/MS spectra were searched against the concatenated target and decoy (reversed) Swiss-Prot protein database using Sequest (Proteomics Browser Software (PBS), Thermo Fisher Scientific) with differential modifications for Met oxidation (+15.99), deamidation of Asn and Gln (+0.984) and fixed Cys alkylation (+57.02). Peptide sequences were identified if they initially passed the following Sequest scoring thresholds against the target database: 1+ ions, Xcorr ≥ 2.0 Sf ≥ 0.4, P ≥ 5; 2+ ions, Xcorr ≥ 2.0, Sf ≥ 0.4, P ≥ 5; 3+ ions, Xcorr ≥ 2.60, Sf ≥ 0.4, P ≥ 5 against the target protein database. Passing MS/MS spectra were manually inspected to be sure that all b- and y- fragment ions aligned with the assigned sequence. False discovery rates (FDR) of peptide hits were estimated below 1.5% based on reversed database hits.

In vivo PK methods

CompoundML265
FormulationIV: 40% w/v beta hydroxy cylcodextrin in water
IP: 40% w/v beta hydroxy cylcodextrin in water
PO: 0.5% w/v carboxy methyl cellulose (400–1000 cps) with 0.1 % v/v Tween 80
DoseIV: 1 mg/kg, PO: 10 mg/kg; IP: 10 and 50 mg/kg
Test SystemMale Balb/c mice
Fasting ConditionsFasted from 2 h pre-dose to 2 h post-dose
Study DesignFor intravenous administration, the solution of the ML265 was administered as bolus at a dose level of 1mg/kg via tail vein at a slow and steady rate. The dosing volume was 5 mL/kg. For intraperitoneal administration, the solution formulation was administered at a dose level of 10 mg/kg and 50 mg/kg through the peritoneum. The dosing volume for i.p. route was 10 mL/kg. For oral dosing, the test compound was administered as a suspension at a dose of 10 mg/kg, by stomach intubation using a 16-gauge oral feeding needle. The dosing volume for oral administration was 10 mL/kg. Blood samples (0.06 mL) were collected from retro-orbital sinus of set of three mouse each at 0, 0.08 (iv & ip), 0.25, 0.5, 1, 2, 4, 6 (only for po), 8 and 24 h in labeled micro centrifuge tube containing K2EDTA as anticoagulant.
The plasma samples were stored at −70 ºC until bioanalysis.
AnalysisPlasma samples were analyzed for ML265 by LC-MS/MS method. LLOQ = 0.636 ng/mL

2.2. Probe Characterization

*Purity >95% as determined by LC/MS and 1H NMR analyses.

*Purity >95% as determined by LC/MS and 1H NMR analyses

2-Methylsulfinyl-4-methyl-6-[(3-aminophenyl)methyl]-4H-thieno[3,2-b]pyrrole[3,2-d]pyridazinone Retention Time: (Method 1) = 3.304 min and (Method 2) = 2.306 min; 1H NMR (400 MHz, DMSO-d6) δ ppm 8.67 (1H, s), 8.02 (1H, s), 7.20 (1H, dd, J= 8.8, 7.6 Hz), 6.90 (1H, d, J= 7.6 Hz), 6.84-6.78 (2H, m), 5.30 (2H, d, J= 2.8 Hz), 4.26 (3H, s), 3.00 (3H, s); HRMS (ESI) m/z (M+H)+ calcd. for C17H17N4O2S2, 373.0793; found 373.0796.

Table 1Extended probe submitted to the MLSMR

Structures of each analog are depicted below the table.

Internal IDMLS IDSIDCIDML #TypeSource
NCGC00186528MLS0031785408526788544246499ML265Extended ProbeNCGC
NCGC00186527MLS0031785358526788444246498NAAnalogNCGC
NCGC00182625MLS0031785369937659746912192NAAnalogNCGC
NCGC00182627MLS0031785379937659846912193NAAnalogNCGC
NCGC00182628MLS0031785385764626525243750NAAnalogNCGC
NCGC00182629MLS0031785399937659946912194NAAnalogNCGC
Image ml265fu3
Figure 2. Stability of ML265 in ADP-Glo assay buffer (Panel A) and DPBS (pH 7.4) buffer (Panel B) at room temperature over 48 hr.

Figure 2Stability of ML265 in ADP-Glo assay buffer (Panel A) and DPBS (pH 7.4) buffer (Panel B) at room temperature over 48 hr

2.3. Probe Preparation

Preparation of (ML265)

Scheme 1. Preparation of: 2-methylsulfinyl-4-methyl-6-[(3-aminophenyl)methyl]-4H-thieno[3,2-b]pyrrole[3,2-d]pyridazinone (ML265).

Scheme 1Preparation of: 2-methylsulfinyl-4-methyl-6-[(3-aminophenyl)methyl]-4H-thieno[3,2-b]pyrrole[3,2-d]pyridazinone (ML265)

Step 1: To a solution of 2-bromo-4-methyl-4H-thieno[3,2-b]pyrrole[3,2-d]pyridazinone (9.5 g, 33.4 mmol) in DMF (120 mL) were added 3-nitrobenzylbromide (14.5 g, 66.9 mmol) and potassium carbonate (10.2 g, 73.6 mmol) and the mixture was heated at 100 °C overnight. After cooling to room temperature, water (100 mL) was added and the mixture was stirred for 10 min to dissolve the inorganic salts. The majority of the desired product was the solid which was filtered, first washed with water (3×50 ml) then EtOAc (3×50 ml). Part of product was still dissolved in the filtrate. The organic layer of the filtrate was separated and washed with brine, dried over Na2SO4. After the removal of organic solvent in vacuo, the residue was purified by silica gel chromatography (EtOAc/hexane= 1/4 to 1/1 gradient) to recover the desired product. The total amount of 2-bromo-4-methyl-6-[(3-nitrophenyl)methyl]-4H-thieno[3,2-b]pyrrole[3,2-d]pyridazinone was 10.5 g (75%).

Step 2: To a solution of 2-bromo-4-methyl-6-[(3-nitrophenyl)methyl]-4H-thieno[3,2-b]-pyrrole[3,2-d]pyridazinone (5.0 g, 11.9 mmol) in EtOH (40 mL) was added SnCl2.2H2O (21.5 g, 95.2 mmol). Concentrated HCl (40 mL) was added dropwise with ice/water cooling. The mixture was heated at 35 °C overnight. The pH was adjusted with 1N NaOH to pH = 9 and filtered through a pad of Celite. Excess EtOAc was used to wash the solid and the filtrate was washed with brine, dried over Na2SO4. After the removal of EtOAc in vacuo, the crude residue was purified by silica gel chromatography (MeOH/DCM= 1/100 to 1/20 gradient) to give 2-bromo-4-methyl-6-[(3-aminophenyl)methyl]-4H-thieno[3,2-b]pyrrole[3,2-d]pyridazinone (3.8 g, 82%) as a light yellow solid.

Step 3: To a solution of 2-bromo-4-methyl-6-[(3-aminophenyl)methyl]-4H-thieno[3,2-b]pyrrole[3,2-d]pyridazinone (200 mg, 0.51 mmol) in DMF (4 mL) in a sealed tube were added copper (I) bromide (74 mg, 0.51 mmol) and sodium thiomethoxide (108 mg, 1.53 mmol) and the sealed tube was purged with N2 for 1 min then sealed. The mixture was heated at 100 °C overnight. After cooling to room temperature, excess EtOAc was added and the precipitate was filtered through a pad of Celite. The filtrate was washed with brine and the organic layer was separated, dried over Na2SO4. After the removal of EtOAc, the crude residue was purified through silica gel chromatography (DCM/MeOH= 1/100 to 1/20) to give 2-methylthio-4-methyl-6-[(3-aminophenyl)methyl]-4H-thieno[3,2-b]pyrrole[3,2-d]pyridazinone (119 mg, 65%) as a light yellow solid.

Step 4: To a solution of 2-methylthio-4-methyl-6-[(3-aminophenyl)methyl]-4H-thieno[3,2-b]pyrrole[3,2-d]pyridazinone (100 mg, 0.28 mmol) in DCM (2 mL) cooled to −78 °C was added a m-CPBA (77wt%, 69 mg, 0.31 mmol) in DCM (1 ml) solution dropwise within 2 min. The mixture was stirred at this temperature for 1 h. Excess DCM (ca. 10 ml) was added and the reaction mixture was allowed to warm to room temperature. The mixture was washed with 10% aqueous NaHCO3 solution, brine and the organic layer was dried over Na2SO4. After the removal of DCM, the crude residue was purified through silica gel chromatography (MeOH/DCM= 1/100 to 1/10) to give 2-methylsulfinyl-4-methyl-6-[(3-aminophenyl)methyl]-4H-thieno[3,2-b]pyrrole-[3,2-d]pyridazinone (83 mg, 80%) as a light yellow solid.

3. Results

3.1. Summary of Screening Results

The tetrameric state is the most active form of pyruvate kinase, which is induced by the endogenous activator 1,6-FBP for PKL, PKR and PKM2. Typically, this feed forward mechanism allows cells to respond to high glycolytic flux and process the resulting PEP to pyruvate and ATP in an appreciably exothermic reaction (ΔG°’ = −7.5 kcal/mol). Cancer cells express PKM2 and somewhat paradoxically exhibit high rates of aerobic glycolysis while having remarkably low pyruvate kinase activity when compared to PKM1 expressing cells. In line with our previously reported activators, ML265 potently activates PKM2 in vitro with an AC50 = 92 nM and shows a high degree of selectivity over the other 3 pyruvate kinase isoforms (Figure 3).

Figure 3. ML265 shows selective activation of PKM2 over PKM1, PKR and PKL in the luminescent pyruvate kinase activity assays using KinaseGlo.

Figure 3

ML265 shows selective activation of PKM2 over PKM1, PKR and PKL in the luminescent pyruvate kinase activity assays using KinaseGlo.

Cancer cells are known to exist in highly phosphorylated states and certain peptide motifs with phosphorylated tyrosines are known PKM2 inhibitors17,18. Binding of these phosphorylated peptides results in 1,6-FBP extrusion from the allosteric site. Use of a known phosphorylated peptide motif that binds to PKM2 (pM2tide) causes a shift in abundance of the more active tetrameric state to the less active monomeric state compared to the same, non-phosphorylated peptide (M2tide) (Figure 4a). ML265 was able to maintain potent activation of PKM2 in the presence of both M2tide and pM2tide (Figure 4b). Treating cells with pervanadate, a known phosphatase inhibitor, causes increased amounts of phosphorylated proteins and decreased PKM2 activity17. To test whether ML265 could activate PKM2 in this cellular context, we determined pyruvate kinase activity in cell lysate of both pervanadate and non-pervanadate treated cells (Figure 4c). Remarkably, at 1 μM, ML265 was able to activate PKM2 regardless of pervanadate treatment. Cantley and co-workers have shown that RNAi mediated replacement of PKM2 with PKM1 results in higher pyruvate kinase activity and reduces cellular proliferation in a hypoxic environment8. To study ML265’s activity on cellular proliferation, the doubling time of H1299 cells were studied over a 6 day period. The activator had no effect at normoxia, but did show an increase in doubling time under hypoxia at 30 μM (Figure 5).

Figure 5. ML265’s affect on doubling time of H1299 cells under normoxia and hypoxia.

Figure 5

ML265’s affect on doubling time of H1299 cells under normoxia and hypoxia.

The fact that cells rapidly utilize aerobic glycolysis with expression of PKM2, the less active enzyme compared to PKM1, led researchers to examine the fate of the enzyme’s high energy substrate, PEP21. They found that in PKM2 expressing cells, PEP activates PGAM1 by readily donating a phosphate group to H11 on PGAM1. One hypothesis is that this reaction allows cells to keep the glycolytic pathway active while not producing ATP. Increases in the ATP/AMP ratio is known to inhibit glycolysis through control of phospofructokinase (PFK). RNAi mediated knockdown of PKM2 and expression of PKM1 led to a significant decrease in phosphorylation on H11 of PGAM121. To assess whether ML265 could reduce this phosphorylation event, H1299 cells were treated with compound for one hour, lysed and analyzed by 2D iso-electric focusing (Figure 6). Compared to DMSO control, a significant reduction in the ratio of phosphorylated to non-phosphorylated PGAM1 was observed.

Figure 6. H1299 cells incubated with 50 μM ML265 for 1 hour, lysed followed by 2-D electrophoresis, Western blot and PGAM1 antibody staining.

Figure 6

H1299 cells incubated with 50 μM ML265 for 1 hour, lysed followed by 2-D electrophoresis, Western blot and PGAM1 antibody staining.

3.2. Dose Response Curves for Probe

We had previously shown that a structural analog of ML265 activates PKM2 via an increase in PEP affinity23. In the absence of activator, PKM2 has a relatively low affinity for PEP (KM ~ 1.5 mM), but treatment with either 1,6-FBP or ML265 resulted in a significant shift in KM to ~0.08 and ~0.09 mM respectively (Figure 7a). In accordance with our previous studies, this effect was not seen with respect to ADP affinity (Figure 7b).

Figure 7. PKM2 activation (scaled from initial velocities) as a function of a) PEP and b) ADP concentration in the presence of 10μM 1,6-FBP (gray circles), ML265 (black circles) or DMSO control (white circles).

Figure 7

PKM2 activation (scaled from initial velocities) as a function of a) PEP and b) ADP concentration in the presence of 10μM 1,6-FBP (gray circles), ML265 (black circles) or DMSO control (white circles).

The homotetramer of PKM2 is the most active form and its endogenous activator, 1,6-FBP is thought to induce this subunit association. To investigate this hypothesis, we used size exclusion chromatography to separate the monomers from tetramers and then determined the effect of 1,6-FBP on PKM2 activity (Figure 8a). As expected, the untreated tetramer had considerably higher kinase activity than the untreated monomer. Addition of 100 μM 1,6-FBP had little effect on the tetramer activity, but resulted in ~70% increase in activity for the monomer (Figure 8b). In an effort to understand whether our activators could induce the more active tetrameric state, we incubated ML265 with recombinant PKM2 and then used sucrose gradient ultracentrifugation followed by SDS-PAGE to separate the protein. As can be seen in Figure 8c, ML265 caused a significant shift to tetrameric species, similar to PKM1 and consistent with the affect seen with 1,6-FBP on PKM2.

Figure 8. a) Size exclusion chromatography was used to separate monomeric and tetrameric PKM2 species and the fractions were b) assayed for pyruvate kinase activity with and without 1,6-FBP (100 μM).

Figure 8

a) Size exclusion chromatography was used to separate monomeric and tetrameric PKM2 species and the fractions were b) assayed for pyruvate kinase activity with and without 1,6-FBP (100 μM). c) Sucrose gradient ultracentrifugation profiles of recombinant (more...)

In an effort to further elucidate this action, we were able to generate a high resolution X-ray structure of ML265 bound to PKM2 (Figure 9). The structure showed that two activators and four 1,6-FBP molecules bind per tetramer. The two equivalents of ML265 bind at the dimer-dimer interface and are completely buried within this interfacial pocket. ML265 is accommodated through van der Waals interactions and water mediated hydrogen bonds to the pocket lining residues (Figure 9b).

Figure 9. a) X-ray structure of the PKM2 tetramer with two equivalents of ML265 and four equivalents of 1,6-FBP bound.

Figure 9

a) X-ray structure of the PKM2 tetramer with two equivalents of ML265 and four equivalents of 1,6-FBP bound. ML265 binds at the dimer-dimer interface, termed A-A’. b) The sulfoxide and carbonyl group of ML265 form water mediated hydrogen bond (more...)

3.3. In vivo Activity

ShRNA replacement of PKM2 with PKM1 in H1299 cells impairs tumor formation in mouse xenograft models8. To determine whether our prior PKM2 activator probe (ML202) or ML265 was more suitable for a comparable xenograft model, we first assessed aqueous solubility and microsomal stability (Table 2). Both activators were comparable with no discernable difference in solubility as well as stability in mouse, rat and human liver microsomes. At this point both compounds were advanced to in vivo PK experiments to determine the more suitable candidate for the efficacy model. As can be seen in Figures 10, ML265 gave superior plasma concentrations that persisted at higher levels over the 24 hour study. ML265 also displayed good oral bioavailability, low clearance, a long half-life and good volume of distrubtion (Table 3). These characteristics were deemed appropriate for use in the mouse xenograft efficacy model.

Table 2. In vitro ADME profile for ML202 and ML265.

Table 2

In vitro ADME profile for ML202 and ML265.

Figure 10. Plasma concentration-time profiles of a) ML202 and b) ML265 in male Balb/c mice following intravenous, intraperitoneal and oral administration.

Figure 10

Plasma concentration-time profiles of a) ML202 and b) ML265 in male Balb/c mice following intravenous, intraperitoneal and oral administration.

Table 3. PK parameters of ML265 in male Balb/c mice.

Table 3

PK parameters of ML265 in male Balb/c mice.

To determine the proper dosing regimen, a multi-day drug exposure simulation was performed using the single dose PK data (Figure 11a). Using 90 nM (AC50 from biochemical assay) as the minimum plasma concentration for the duration of the experiment and 15.4% as the free-fraction in mouse plasma, 50 mg/kg BID was chosen as the regimen to keep plasma levels of drug well above 90 nM. ML265 was then dosed orally at 150 mg/kg in a A549 mouse xenograft model and maximal PKM2 activation was achieved (Figure 11b). We also wanted to determine the tolerability of bi-daily dosing in a 5-day study looking at percent change in body weight. To this end, ML265 was well tolerated at the desired 50 mg/kg BID regimen (Figure 11c).

Figure 11. a) Simulation of three BID dosages of ML265 derived from single dose PK experiment.

Figure 11

a) Simulation of three BID dosages of ML265 derived from single dose PK experiment. b) Pharmacodynamic study on tumor lysates after two consecutive days of 150 mg/kg BID dosing of ML265 in an A549 xenograft model. c) Monitoring bodyweight of mice given (more...)

ML265 was advanced into a 7-week H1299 mouse xenograft study using immunocompromised (nu/nu) mice. The mice were randomly divided into two cohorts with one group receiving the established 50 mg/kg BID of ML265, the other cohort received only vehicle. Blood counts, serum chemistries and histological examination of various tissues were followed to determine if there were overt toxicities associated with ML265 treatment (Table 4).

Table 4. Serum chemistries and blood counts of mice treated with ML265.

Table 4

Serum chemistries and blood counts of mice treated with ML265.

Gratifyingly, no significant aberrations were observed in these toxicity measures when compared to the vehicle treated mice. ML265 was able to delay tumor formation compared to vehicle-treated mice and the umors that did form were significantly smaller (Figure 12). Further examination of the tumors revealed that ML265 was detectable signifying that the tumor was being exposed to the activator.

Figure 12. H1299 cells were injected subcutaneously into nu/nu mice which were subsequently randomly divided into two cohorts, one given vehicle and the other 50 mg/kg BID ML265 throughout the duration of the experiment.

Figure 12

H1299 cells were injected subcutaneously into nu/nu mice which were subsequently randomly divided into two cohorts, one given vehicle and the other 50 mg/kg BID ML265 throughout the duration of the experiment. a) Injection sites were monitored for tumor (more...)

4. Discussion

PKM2 has emerged as an important enzyme in the metabolic reprogramming of cancer cells and its expression profile make it a valuable target for further study. Notably, its splice variant, PKM1 has entirely different regulatory properties affecting its activity; whereas PKM1 is constitutively active, PKM2 requires an upstream glycolytic intermediate, 1,6-FBP, for allosteric activation. RNAi replacement of PKM2 with PKM1 results in higher oxygen consumption, lower lactate production, reduced proliferation and significantly reduced tumor size and weight in mouse xenograft models8. In an effort to study the hypothesis that pharmacological activation of PKM2 could mimic these RNAi effects, we have developed novel PKM2 activators2325. These compounds are remarkably selective over PKM1, PKR and PKL and potently activate PKM2 with low nanomolar AC50’s. The tetrameric structure of pyruvate kinase is known to be the most active form of the enzyme and ML265 seems to activate PKM2 via induction of this complex. Sucrose gradient ultracentrifugation of both PKM1 and PKM2 indeed showed that PKM1 exists predominantly in the tetrameric form, whereas PKM2 exists with a higher proportion of the monomeric form. Treatment of PKM2 with either 1,6-FBP or ML265 causes a marked shift from monomeric to tetrameric species. Co-crystallization studies of PKM2 with ML265 indicated that two small molecules bind per tetramer, each at the dimer-dimer interface. The compound makes water-mediated and direct hydrogen bonds to the amino acid residues of both monomers in each dimer. Somewhat surprisingly, four equivalents of 1,6-FBP still bound per tetramer, highlighting the distinct location that the endogenous activator binds compared to the synthetic small molecules.

Increased amounts of phosphorylated proteins in cancer cells arise through elevated levels of kinase signaling. Experimentally, exogenous production of highly phosphorylated proteins can be induced by pervanadate treatment via inhibition of phosphatase activity. This pervanadate treatment inhibits pyruvate kinase activity in PKM2 expressing cells, but not PKM1 expressing cells. Experiments support that the inhibition arises from binding of phosphorylated peptides to PKM2 causing release of 1,6-FBP, and dissociation of the enzyme into its less active monomeric form. In vitro experiments showed that non-phosphorylated counterparts of these peptide motifs do not inhibit the enzyme17. ML265 is able to potently activate PKM2 in the presence of both the non-phosphorylated and phosphorylated peptide. Additionally, activation of pyruvate kinase activity in PKM2 expressing cells to similar levels as PKM1 expressing cells was observed in both pervanadate treated and untreated cells. ML265 was able to significantly reduce cellular proliferation in a hypoxic environment, further demonstrating the PKM2 acitvator’s ability to function in conditions typically seen in tumors.

Though a number of recent reports have revealed both non-enzymatic and non-metabolic consequences of PKM2 expression3538, RNAi experiments and pharmacological activation support the importance of PKM2 enzymatic activity in tumor growth. After showing good microsomal stability, aqueous solubility and Caco2 permeability, ML265 was progressed to in vivo PK where it showed good bioavailability, exposure, a long half-life and low clearance. The 7-week mouse xenograft model showed that the activation of PKM2 with ML265 was able to significantly reduce tumor size and occurrence without showing signs of acute toxicity. We believe this small molecule will help unravel many of the roles and functional consequences of PKM2 expression, and help elucidate aspects of the metabolic reprogramming occurring in tumors.

5. References

1.
Tennant DA, Duran RV, Gottlieb E. Targeting metabolic transformation for cancer therapy. Nat Rev Cancer. 2010;10:267–77. [PubMed: 20300106]
2.
Vander Heiden MG. Targeting cancer metabolism: a therapeutic window opens. Nat Rev Drug Discov. 2011;10:671–84. [PubMed: 21878982]
3.
Vander Heiden MG, Cantley LC, Thompson CB. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science. 2009;324:1029–33. [PMC free article: PMC2849637] [PubMed: 19460998]
4.
Cairns RA, Harris IS, Mak TW. Regulation of cancer cell metabolism. Nat Rev Cancer. 2011;11:85–95. [PubMed: 21258394]
5.
Levine AJ, Puzio-Kuter AM. The control of the metabolic switch in cancers by oncogenes and tumor suppressor genes. Science. 2010;330:1340–4. [PubMed: 21127244]
6.
Trachootham D, Alexandre J, Huang P. Targeting cancer cells by ROS-mediated mechanisms: a radical therapeutic approach? Nat Rev Drug Discov. 2009;8:579–91. [PubMed: 19478820]
7.
Weissleder R. Molecular imaging in cancer. Science. 2006;312:1168–71. [PubMed: 16728630]
8.
Christofk HR, et al. The M2 splice isoform of pyruvate kinase is important for cancer metabolism and tumour growth. Nature. 2008;452:230–3. [PubMed: 18337823]
9.
Mazurek S. Pyruvate kinase type M2: a key regulator of the metabolic budget system in tumor cells. Int J Biochem Cell Biol. 2011;43:969–80. [PubMed: 20156581]
10.
Noguchi T, Inoue H, Tanaka T. The M1- and M2-type isozymes of rat pyruvate kinase are produced from the same gene by alternative RNA splicing. J Biol Chem. 1986;261:13807–12. [PubMed: 3020052]
11.
Clower CV, et al. The alternative splicing repressors hnRNP A1/A2 and PTB influence pyruvate kinase isoform expression and cell metabolism. Proc Natl Acad Sci U S A. 2010;107:1894–9. [PMC free article: PMC2838216] [PubMed: 20133837]
12.
Yamada K, Noguchi T. Regulation of pyruvate kinase M gene expression. Biochem Biophys Res Commun. 1999;256:257–62. [PubMed: 10079172]
13.
Ikeda Y, Tanaka T, Noguchi T. Conversion of non-allosteric pyruvate kinase isozyme into an allosteric enzyme by a single amino acid substitution. J Biol Chem. 1997;272:20495–501. [PubMed: 9252361]
14.
Ikeda Y, Noguchi T. Allosteric regulation of pyruvate kinase M2 isozyme involves a cysteine residue in the intersubunit contact. J Biol Chem. 1998;273:12227–33. [PubMed: 9575171]
15.
Ashizawa K, Willingham MC, Liang CM, Cheng SY. In vivo regulation of monomer-tetramer conversion of pyruvate kinase subtype M2 by glucose is mediated via fructose 1,6-bisphosphate. J Biol Chem. 1991;266:16842–6. [PubMed: 1885610]
16.
Ashizawa K, McPhie P, Lin KH, Cheng SY. An in vitro novel mechanism of regulating the activity of pyruvate kinase M2 by thyroid hormone and fructose 1, 6-bisphosphate. Biochemistry. 1991;30:7105–11. [PubMed: 1854723]
17.
Christofk HR, Vander Heiden MG, Wu N, Asara JM, Cantley LC. Pyruvate kinase M2 is a phosphotyrosine-binding protein. Nature. 2008;452:181–6. [PubMed: 18337815]
18.
Ikeda Y, Taniguchi N, Noguchi T. Dominant negative role of the glutamic acid residue conserved in the pyruvate kinase M(1) isozyme in the heterotropic allosteric effect involving fructose-1,6-bisphosphate. J Biol Chem. 2000;275:9150–6. [PubMed: 10734049]
19.
Hitosugi T, et al. Tyrosine phosphorylation inhibits PKM2 to promote the Warburg effect and tumor growth. Sci Signal. 2009;2:ra73. [PMC free article: PMC2812789] [PubMed: 19920251]
20.
Anastasiou D, et al. Inhibition of PKM2 by reactive oxygen species contributes to cellular antioxidant responses. Science. 2011;334:1278. [PMC free article: PMC3471535] [PubMed: 22052977]
21.
Vander Heiden MG, et al. Evidence for an alternative glycolytic pathway in rapidly proliferating cells. Science. 2010;329:1492–9. [PMC free article: PMC3030121] [PubMed: 20847263]
22.
Locasale JW, Vander Heiden MG, Cantley LC. Rewiring of glycolysis in cancer cell metabolism. Cell Cycle. 2010;9:4253. [PubMed: 21045562]
23.
Jiang JK, et al. Evaluation of thieno[3,2-b]pyrrole[3,2-d]pyridazinones as activators of the tumor cell specific M2 isoform of pyruvate kinase. Bioorg Med Chem Lett. 2010;20:3387–93. [PMC free article: PMC2874658] [PubMed: 20451379]
24.
Boxer MB, et al. Evaluation of substituted N,N′-diarylsulfonamides as activators of the tumor cell specific M2 isoform of pyruvate kinase. J Med Chem. 2010;53:1048–55. [PMC free article: PMC2818804] [PubMed: 20017496]
25.
Walsh MJ, et al. 2-Oxo-N-aryl-1,2,3,4-tetrahydroquinoline-6-sulfonamides as activators of the tumor cell specific M2 isoform of pyruvate kinase. Bioorg. Med. Chem. Lett. 2011;21:6322. [PMC free article: PMC3224553] [PubMed: 21958545]
26.
Vander Heiden MG, et al. Identification of small molecule inhibitors of pyruvate kinase M2. Biochem Pharmacol. 2009;79:1118–24. [PMC free article: PMC2823991] [PubMed: 20005212]
27.
Otwinowski Z, Minor W. Methods in Enzymology. Vol. 276. (Academic Press; 1997. Processing of X-ray diffraction data collected in oscillation mode; pp. 307–326. [PubMed: 27754618]
28.
Minor W, Cymborowski M, Otwinowski Z, Chruszcz M. HKL-3000: the integration of data reduction and structure solution--from diffraction images to an initial model in minutes. Acta Crystallogr D Biol Crystallogr. 2006;62:859–66. [PubMed: 16855301]
29.
Berman HM, et al. The Protein Data Bank. Nucleic Acids Res. 2000;28:235–42. [PMC free article: PMC102472] [PubMed: 10592235]
30.
Schuttelkopf AW, van Aalten DM. PRODRG: a tool for high-throughput crystallography of protein-ligand complexes. Acta Crystallogr D Biol Crystallogr. 2004;60:1355–63. [PubMed: 15272157]
31.
Emsley P, Lohkamp B, Scott WG, Cowtan K. Features and development of Coot. Acta Crystallogr D Biol Crystallogr. 2010;66:486–501. [PMC free article: PMC2852313] [PubMed: 20383002]
32.
Murshudov GN, Vagin AA, Dodson EJ. Refinement of macromolecular structures by the maximum-likelihood method. Acta Crystallogr D Biol Crystallogr. 1997;53:240–55. [PubMed: 15299926]
33.
Davis IW, Murray LW, Richardson JS, Richardson DC. MOLPROBITY: structure validation and all-atom contact analysis for nucleic acids and their complexes. Nucleic Acids Res. 2004;32:W615–9. [PMC free article: PMC441536] [PubMed: 15215462]
34.
Yang H, et al. Automated and accurate deposition of structures solved by X-ray diffraction to the Protein Data Bank. Acta Crystallogr D Biol Crystallogr. 2004;60:1833–9. [PubMed: 15388930]
35.
Luo W, et al. Pyruvate kinase M2 is a PHD3-stimulated coactivator for hypoxia-inducible factor 1. Cell. 2011;145:732–44. [PMC free article: PMC3130564] [PubMed: 21620138]
36.
Hoshino A, Hirst JA, Fujii H. Regulation of cell proliferation by interleukin-3-induced nuclear translocation of pyruvate kinase. J Biol Chem. 2007;282:17706–11. [PubMed: 17446165]
37.
Stetak A, et al. Nuclear translocation of the tumor marker pyruvate kinase M2 induces programmed cell death. Cancer Res. 2007;67:1602–8. [PubMed: 17308100]
38.
Yang W, et al. Nuclear PKM2 regulates β-catenin transactivation upon EGFR activation. Nature. 2011;478:118–122. [PMC free article: PMC3235705] [PubMed: 22056988]

Views

  • PubReader
  • Print View
  • Cite this Page
  • PDF version of this page (827K)

Related information

Similar articles in PubMed

See reviews...See all...

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...