Skip to main content
  • AACR Journals
    • Blood Cancer Discovery
    • Cancer Discovery
    • Cancer Epidemiology, Biomarkers & Prevention
    • Cancer Immunology Research
    • Cancer Prevention Research
    • Cancer Research
    • Clinical Cancer Research
    • Molecular Cancer Research
    • Molecular Cancer Therapeutics

AACR logo

  • Register
  • Log in
  • Log out
  • My Cart
Advertisement

Main menu

  • Home
  • About
    • The Journal
    • AACR Journals
    • Subscriptions
    • Permissions and Reprints
    • Reviewing
  • Articles
    • OnlineFirst
    • Current Issue
    • Past Issues
    • Meeting Abstracts
    • Collections
      • COVID-19 & Cancer Resource Center
      • Focus on Radiation Oncology
      • Novel Combinations
      • Reviews
      • Editors' Picks
      • "Best of" Collection
  • For Authors
    • Information for Authors
    • Author Services
    • Best of: Author Profiles
    • Submit
  • Alerts
    • Table of Contents
    • Editors' Picks
    • OnlineFirst
    • Citation
    • Author/Keyword
    • RSS Feeds
    • My Alert Summary & Preferences
  • News
    • Cancer Discovery News
  • COVID-19
  • Webinars
  • Search More

    Advanced Search

  • AACR Journals
    • Blood Cancer Discovery
    • Cancer Discovery
    • Cancer Epidemiology, Biomarkers & Prevention
    • Cancer Immunology Research
    • Cancer Prevention Research
    • Cancer Research
    • Clinical Cancer Research
    • Molecular Cancer Research
    • Molecular Cancer Therapeutics

User menu

  • Register
  • Log in
  • Log out
  • My Cart

Search

  • Advanced search
Molecular Cancer Therapeutics
Molecular Cancer Therapeutics
  • Home
  • About
    • The Journal
    • AACR Journals
    • Subscriptions
    • Permissions and Reprints
    • Reviewing
  • Articles
    • OnlineFirst
    • Current Issue
    • Past Issues
    • Meeting Abstracts
    • Collections
      • COVID-19 & Cancer Resource Center
      • Focus on Radiation Oncology
      • Novel Combinations
      • Reviews
      • Editors' Picks
      • "Best of" Collection
  • For Authors
    • Information for Authors
    • Author Services
    • Best of: Author Profiles
    • Submit
  • Alerts
    • Table of Contents
    • Editors' Picks
    • OnlineFirst
    • Citation
    • Author/Keyword
    • RSS Feeds
    • My Alert Summary & Preferences
  • News
    • Cancer Discovery News
  • COVID-19
  • Webinars
  • Search More

    Advanced Search

Article

Targeting Aurora2 Kinase in Oncogenesis: A Structural Bioinformatics Approach to Target Validation and Rational Drug Design1

Hariprasad Vankayalapati, David J. Bearss, José W. Saldanha, Rubén M. Muñoz, Sangeeta Rojanala, Daniel D. Von Hoff and Daruka Mahadevan
Hariprasad Vankayalapati
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
David J. Bearss
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
José W. Saldanha
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Rubén M. Muñoz
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Sangeeta Rojanala
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Daniel D. Von Hoff
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Daruka Mahadevan
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
DOI:  Published March 2003
  • Article
  • Figures & Data
  • Info & Metrics
  • PDF
Loading

Abstract

The aurora kinases are a novel oncogenic family of mitotic serine/threonine kinases (S/T kinases) that are overexpressed in a number of solid tumors, including pancreas and colorectal cancer. A PSI-BLAST search [National Center for Biotechnology Information (NCBI)] with the sequence of the S/T kinase domain of human aurora1 kinase [also known as AUR1, ARK2, AIk2, AIM-1, and STK12] and human aurora2 kinase (also known as AUR2, ARK1, AIK, BTAK, and STK15) showed a high sequence similarity to the three-dimensional structures of bovine cAMP-dependent kinase [Brookhaven Protein Data Bank code 1CDK], murine cAMP-dependent kinase (1APM), and Caenorhabditis elegans twitchin kinase (1KOA). When the aurora1 or aurora2 sequence was input into the tertiary structure prediction programs THREADER and 3D-PSSM (three-dimensional position-sensitive scoring matrix), the top structural matches were 1CDK, 1APM, and 1KOA, confirming that these domains are structurally conserved. The structural models of aurora1 and aurora2 were built using 1CDK as the template structure. Molecular dynamics and docking simulations, targeting the ATP binding site of aurora2 with adenylyl imidodiphosphate (AMP-PNP), staurosporine, and six small molecular S/T kinase inhibitors, identified active-site residues that interact with these inhibitors differentially. The docked structures of the aurora2-AMP-PNP and aurora2-staurosporine complexes indicated that the adenine ring of AMP-PNP and the indolocarbazole moiety of staurosporine have similar positions and orientations and provided the basis for the docking of the other S/T kinase inhibitors. Inhibitors with isoquinoline and quinazoline moieties were recognized by aurora2 in which H-89 and 6,7-dimethoxyquinazoline compounds exhibited high binding energies compared with that of staurosporine. The calculated binding energies for the docked small-molecule inhibitors were qualitatively consistent with the IC50 values generated using an in vitro kinase assay. The aurora2 structural model provides a rational basis for site-directed mutagenesis of the active site; design of novel H-89, staurosporine, and quinazoline analogues; and the screening of the available chemical database for the identification of other novel, small-molecular entities.

Introduction

In cell division, the M phase is composed of mitosis and cytokinesis. Progression through the M phase is dependent on several mitotic kinases, the activities of which are regulated by phosphorylation-dephosphorylation and proteolysis. Of the mitotic kinases, cyclin-dependent kinase 1 is the most prominent cell cycle regulator that orchestrates M-phase activities. However, a number of other mitotic protein kinases that participate in M phase have been identified; they include members of the polo, aurora, and NIMA (never in mitosis A) families and kinases implicated in mitotic checkpoints, mitotic exit, and cytokinesis (1). Aurora kinases are a family of oncogenic S/T kinases3 that localize to the mitotic apparatus (centrosome, poles of the bipolar spindle, or midbody) and regulate completion of centrosome separation, bipolar spindle assembly, and chromosome segregation. Three human homologues have been identified (aurora1, aurora2, and aurora3; Ref. 2), and they all share a highly conserved catalytic domain located in the COOH terminus, but their NH2 terminal extensions are of variable lengths with no sequence similarity (3). The human aurora kinases are expressed in proliferating cells and are also overexpressed in numerous tumor cell lines of the breast, ovary, prostate, pancreas, and colon (2, 4–6). It has been shown that aurora2 acts as an oncogene and is able to transform both Rat1 fibroblasts and mouse NIH3T3 cells in vitro, and aurora2 transforms NIH3T3 cells grown as tumors in nude mice (3, 4). It has been proposed (2) that overexpression of aurora2 may drive cells into aneuploidy, which in turn can accelerate the loss of tumor suppressor genes and/or amplification of oncogenes, leading to cellular transformation. This is likely to be achieved when those cells that overexpress aurora2 escape mitotic check points, which allows inappropriate activation of proto-oncogenes. Recently, we have shown up-regulation of aurora2 in a number of pancreatic cancer cell lines and have validated this as a target by demonstrating that antisense oligonucleotide treatment leads to a loss of aurora2 activity and increased apoptosis (7). Therefore, aurora2 is a potential target that requires further validation by the rational design of novel small-molecular inhibitors (8) and testing in pancreatic cancer cell lines and animal models for efficacy and potency.

To validate aurora2 as a drugable target in pancreatic cancer, we have undertaken a study to design specific inhibitors of this mitotic S/T kinase. Here we describe a structure-based design approach that uses three-dimensional modeling of aurora1 and aurora2 kinases. Homology modeling of PKs is well validated and documented (9–11). We have used a structural bioinformatic algorithm that uses PSI-BLAST (NCBI), THREADER, 3D-PSSM (three-dimensional position-sensitive scoring matrix), and SAP programs to determine the optimal template for homology modeling of aurora1 and aurora2. The crystal structure of the activated form of bovine cAMP-dependent PK was used as the best template for homology modeling using MD simulations in INSIGHT II. The modeled aurora2 structure was docked with known S/T kinase and aurora2 kinase inhibitors using the binary complex of cAMP-dependent PK-Mn2+-AMP-PNP (12, 13). The calculated binding energies from the docking analysis are in agreement with experimental IC50 values obtained from an in vitro kinase assay, which uses histone H1 phosphorylation to assess inhibitor activity.

Materials and Methods

Sequence and Structure Analysis.

The aurora2 (ARK1; residues 98–403) and aurora1 (ARK2; residues 35–344) kinase domain sequences were used as probes to search a nonredundant database of sequences using PSI-BLAST (NCBI). Top-ranked sequences for which the three-dimensional structures of S/T kinase domains are available were the porcine heart cAMP-dependent PK (1CDK; Ref. 12), recombinant mouse cAMP-dependent PK (1APM; Ref. 14), and Caenorhabditis elegans twitchin kinase (1KOA; Ref. 15). The aurora1 and aurora2 domain sequences were analyzed using the programs THREADER (16) and 3D-PSSM, (17), which compare primary sequences with all of the known three-dimensional structures in the Brookhaven Protein Data Bank. The output is composed of the optimally aligned, lowest-energy, three-dimensional structures that are similar to the aurora kinases. The top-ranked structures were bovine cAMP-dependent kinase (1CDK), murine cAMP-dependent kinase (1APM) and twitchin kinase (1KOA) confirming the PSI-BLAST search. These three tertiary structures provide the three-dimensional templates for the homology modeling of aurora1 and aurora2. The crystal structure coordinates for the above S/T kinase domains were obtained from the Protein Data Bank (18). These domains were pairwise superimposed onto each other using the program SAP (19). The structural alignments from SAP were fine-tuned manually to better match residues within the regular secondary structural elements. The three manually aligned S/T kinase domain sequences with their respective secondary structures were viewed in Clustal X (20).

Homology Modeling and Refinement.

The modeling software used was INSIGHT II (version 2000, Accelrys Inc.) running on a Silicon Graphics Indigo2 workstation (Silicon Graphics Inc.) under the UNIX operating system. The crystal structure coordinates of the ternary complex of cAMP-dependent kinase (1CDK; Ref. 12) were retrieved from the Brookhaven Protein Data Bank. The catalytic domain served as the template for aurora1 and aurora2 modeling. The inhibitory peptide PKI (5–24) and the solvent molecules were deleted from the coordinate set before modeling. The two Mn2+ ions in the active-site pocket were retained and replaced by Mg2+ ions. The model building procedure for aurora1 and aurora2 entailed a deletion of residues 282GNLKD286 from the COOH-terminal end of cAMP-dependent PK and a deletion of the last 30 residues from the COOH terminus of cAMP-dependent PK (residues highlighted in blue in Fig. 1). Ser284 and Leu228 were inserted using the conformation of Lys204 of twitchin kinase (1KOA). All of the other side chains except for the identical residues were mutated to those of aurora1 using the Homology modeler and Biopolymer modules in Insight II.4 This model served as a template for the homology modeling (12, 13) of aurora2 kinase. The main differences between aurora1 and aurora2 are two insertions between Lys124-Lys125 and Ser388-Lys389. The peptide bonds between these residues were broken, and SGTPDIL and RVL residues were inserted in aurora1 consecutively without altering the geometry of the rest of the sequence. The conformation of the inserted fragments was obtained from the fragment database of low-energy conformations. Furthermore, all of the amino acid residues except for the identical residues were mutated to those of aurora2. The mutated side chains were manually positioned to minimize steric hindrance with adjacent residues. In general, insertions and deletions were found in loop regions connecting regular secondary structures and, hence, did not alter the hydrophobic core of the structure.

After the model building processes were complete, a series of minimizations were performed to relax the structure. Before model refinement, all of the close contacts caused by the mutation of side chains were fixed by manually rotating the χ1 and χ2 torsion angles and keeping the side chain torsion angles of the conserved residues in their original conformation. Hydrogen atoms were supplied by Biopolymer with CFF parameters (21), solvated by simulating with the distance dielectric constant, ϵ = 4rij (22, 23) and refined by the steepest descent and conjugate gradient minimization until convergence was reached, while restraining the position of the heavy atoms. Finally, the entire system was subjected to energy minimization using the 3000-step steepest descent followed by conjugate gradient minimization until an energy gradient of less than 0.001 kcal/mol/Å was achieved and from which, constraints were gradually removed.

Model Evaluation.

The final aurora2 model was examined using profile-3D (24). The profile-3D and three-dimensional-one-dimensional score plots of the model were positive over the entire length of protein in a moving-window scan to the template structure. Additionally, PROCHECK (25) was used to verify the correct geometry of the dihedral angles and the handedness of the model-built structure. For the PROCHECK statistics, an overall G factor of −0.31, hydrogen bond energy of 1.6, and only 0.29 bad contacts per 100 residues were observed, which is consistent with a good quality structure comparable with the crystal structure. The overall fold of the homology model (Fig. 3A) was found to be very similar to that of crystal structure template 1CDK, which was used for additional energy refinement, MD simulations, and docking studies of other S/T kinase inhibitors.

MD Simulations.

The energy-minimized aurora2 structure was used as starting model for MD simulations (26, 27), which were performed in the canonical ensemble (NVT) at 300°K using the CFF force field implemented in Discover_3 (version 2.9.5).5 Dynamics were equilibrated for 10 ps with time steps of 1 fs and continued for 10-ps simulations. The nonbonded cutoff distance of 8 Å and a distance-dependent dielectric constant (ϵ = 4rij) for water were used to simulate the aqueous media. All of the bonds to hydrogen were constrained. Dynamic trajectories were recorded every 0.5 ps for analysis. Individual simulations from the point of a stable trajectory generated time-averaged structures. The resulting low-energy structure was extracted, and energy was minimized to 0.001 kcal/mol/Å. To examine the conformational changes that occur during MD, the rms deviations were calculated from trajectories at 0.5-ps intervals and compared with the Cα backbone of cAMP-dependent PK (see Fig. 3B). The rms deviation for the two superimposed structures was 0.42 Å. Furthermore, the rms deviations were calculated for the protein backbone (0.37 Å) and the active-site pocket (0.41 Å) and were compared with crystal structure before the docking experiments. The resulting aurora2 structure served as the starting model for docking studies.

SA Docking.

To find sterically reasonable binding geometries and to explore the interactions of the proposed ligand recognition pocket, affinity docking with SA was performed in INSIGHT II (28, 29). The structures of ligands used for docking were from five crystal structure complexes of cAMP-dependent PK bound to AMP-PNP, (12), staurosporine (30, 31), H-89, H-7, H-8, (32, 33), and structures that were empirically built and energy minimized [KN-93 (34), ML-7 (35), and 6,7-dimethoxyquinazoline (8, 36); Fig. 2] in INSIGHT II. Partial charges were assigned to the ligands by the Gasteiger method defined within INSIGHT II.5 Systematic conformational searches were performed on each of the minimized ligands using 10-ps MD simulations at 300°K. For docking with AMP-PNP, the position of the ATP analogue was retained from its crystal structure with 1CDK, in which the adenine base served as a template for field-fit alignments with the indolocarbazole, fused biphenyl, and quinazoline moieties. The ATP analogue was then removed from the field-fit alignment, and each of the other ligands was docked into the active-site pocket with a position and orientation that were similar to those of AMP-PNP. The heavy atoms from the ATP analogue were used as sphere centers in the input to the docking procedure. To clarify the orientation of these inhibitors in the active-site pocket, the electrostatic potential at the van der Waals surface was determined using solvent surface calculations. Orientations with the lowest intermolecular potential energy were calculated. To explore the interaction of the ligands in the recognition pockets, SA docking of complex formations was carried out without constraints, to allow each of the protein-ligand complex systems to evolve freely. To explore the effects of solvation implicitly, a distance-dependent dielectric constant (ϵ = 4rij) was used (22, 23). The nonbonded cell multipole method was used for SA docking with input energy parameters. Docking simulations were performed at 500°K with 100 fs/stage (total of 50 stages), quenching the system to a final temperature of 300°K. The whole complex structure was energy minimized using 1000 steps. This provided 10 structures from SA docking, and their generated conformers were clustered according to rms deviation. The lowest global structure obtained was used for computing intermolecular binding energies (37). Furthermore, we have validated the robustness of the affinity-docking methodology (27, 38) by comparative docking of an ATP analogue, AMP-PNP, with 1CDK and aurora2 kinase. The binding mode of these complexes are shown in Fig. 4, A and B, respectively.

In Vitro Aurora Kinase Assay.

Aurora2 immunoprecipitation kinase assays were performed as described previously (6). Briefly, cells were first homogenized in lysis buffer [150 mm NaCl, 50 mm HEPES (pH 7.2), 1 mm EDTA, 1 mm EGTA, 1 mm DTT, 0.1% Tween 20, 0.1 mm phenylmethylsulfonylfluoride, 2.5 μg/ml leupeptin, 0.1 mm sodium orthovanadate (Sigma Chemicals, St. Louis, MO)] at 4°C. Lysates were then centrifuged at 10,000 × g for 10 min. Protein content was determined by the BCA protein assay (Pierce, Rockford, IL), and 200 μg was used for each sample. The supernatants were immunoprecipitated for 12 h at 4°C with protein A-agarose beads precoated with saturating amounts of aurora kinase-2 polyclonal antibody. Immunoprecipitated proteins on beads were washed twice with 1 ml of lysis buffer and twice with kinase buffer [50 mm HEPES (pH 7.0), 10 mm MgCl2, 5 mm MnCl2, and 1 mm DTT]. The beads were then resuspended in 40 μl of kinase buffer containing 10 μm ATP, 5 μCi of [γ-32P]ATP (6000 Ci/mmol; 1 Ci = 37 GBq; Amersham Corp., Arlington Heights, IL), and 4 μg of the histone H1 protein substrate. The samples were then incubated for 30 min at 30°C with occasional mixing. For dose-response studies to determine IC50, serial dilutions of the inhibitors were used (10–200 μm). The samples were then boiled in polyacrylamide gel sample buffer containing SDS and separated by electrophoresis. Phosphorylated proteins were quantified after exposure to autoradiographic film (Labscientific, Inc., Livingston, NJ) by densitometry using ImageQuant version 5.1 (Molecular Dynamics Computing Densitometer, Sunnyvale, CA). The IC50 values (Table 1) were calculated from the phosphorylation profile of the substrate histone H1 protein.

Results

Aurora Kinases Possess a Conserved yet Distinct S/T Kinase Catalytic Domain

The sequence identity and similarity between aurora1 and aurora2 kinases and cAMP-dependent PK are 30 and 60%, respectively, and that between aurora1 and aurora2 kinases and twitchin kinase is 17 and 54%, respectively. The aurora1 and aurora2 domain sequences were aligned with respect to the structural alignments obtained in Clustal X. Fig. 1 shows the secondary structural elements of the aligned kinase domains. The critical catalytic residues involved in the transfer of the γ-phosphoryl group of ATP to the substrates are highly or absolutely conserved between the aurora kinases, cAMP-dependent PK, and twitchin kinase, and these residues are highlighted in Fig. 1. In the prototype cAMP-dependent PK catalytic domain, they are 50GXGXXGX57, 72K, 91E, 121E, 166D, 168K, 170E, 171N, 184DFG, 197T (autophosphorylation site), 206APE, 220D, and 280R (12). The main differences between cAMP-dependent PK and the aurora kinases are: (a) the glycine-rich nucleotide binding motif 50GTGSFGRV57 is changed to 140GKGKFGNV147; (b) 206APE is changed to 297PPE299; (c) 284S and 228L in aurora kinases are inserted; (d) residues GNLKD are deleted from the COOH-terminal end of cAMP-dependent PK; and (e) the last 30 residues are deleted from the COOH terminus of cAMP-dependent PK. These results indicate that the aurora kinases possess a conserved S/T kinase catalytic domain similar to that of the cAMP-dependent PKs. However, the changed glycine-rich nucleotide-binding sequence motif, V123A213, E127T217, insertion of Ser284 in aurora2 and Leu228 in aurora1 (between Ser283 and Arg285) and A206P297 in the active site provide important structural differences between these two kinases that may explain their distinct substrate specificities.

Each S/T kinase structure is composed of a small NH2-terminal β-sheet domain and a large COOH-terminal α-helical domain, with the catalytic cleft between the two domains at which the substrate and ATP bind in the presence of Mg2+ ions (12). The crystal structure determined for bovine cAMP-dependent PK is phosphorylated in the activation loop and is in complex with AMP-PNP (Fig. 2) and the pseudo-substrate peptide PKI (5–24). This catalytically competent enzyme is in a state immediately before phosphoryl transfer, with the cleft between the NH2- and COOH-terminal domains in a closed conformation. The tertiary structure modeling of aurora1 and aurora2 on the catalytically competent cAMP-dependent PK structure has defined pertinent structural features in the aurora kinases and has guided our structure-based rational design process. Fig. 3A shows the model-built structure of aurora2. Secondary structural elements are indicated as red cylinders (α-helix), yellow arrows (β-sheet), green ribbons (coil), and blue ribbons (turns). The Cα backbone superimposition of the final minimized structure of aurora2 (blue) on cAMP-dependent PK (1CDK; red) is shown with a rms deviation of 0.42 Å (Fig. 3B). Because of the two COOH-terminal deletions in the aurora kinases compared with that of cAMP-dependent PK, the aurora kinases have evolved a more compact catalytic kinase core domain that may affect their substrate specificity and activity.

Docking of S/T Kinase Inhibitors Coupled to an in Vitro Kinase Assay Defines Distinct Chemical Moieties as Building Blocks for Potent Inhibitors of Aurora2

Aurora2-AMP-PNP Complex.

Affinity docking experiments indicate that AMP-PNP in complex with aurora2 exhibits similar conformational orientations to that of the crystal structure of cAMP-dependent PK bound to AMP-PNP (12). This comparative binding mode provides validation of affinity docking and a rationale for docking simulations of other S/T kinase inhibitors. Because of the change in sequence motif of the glycine-rich loop and residues interacting with the ATP analogue, AMP-PNP binds with an affinity in the ATP-binding pocket of aurora2 that is lower than that with 1CDK (Table 1). The crystal structure of the 1CDK-AMP-PNP complex (12) revealed that the high binding affinity of the ATP analogue in the nanomolar range is caused by a number of interactions, including polar, β,γ-phosphoryl coordination to Mn2+ ions, and hydrogen bonds with specific enzyme residues. From our docking and MD simulations, we observed a network of hydrogen bonding interactions with aurora2, which, in turn, stabilized the docked structure. The adenine base, its 6-amino group, and ring N1 atoms interact with Glu211 and Ala213, which have strong hydrogen bonding interactions: Glu211-C⋕O… H2N of adenine (1.97 Å, energy −2.34 kcal/mol) and Ala213-NH… N of ring N1 (2.26 Å, energy −2.65 kcal/mol), respectively. Weaker hydrogen bonding interactions for AMP-PNP with Glu121 (2.90 Å) and Val123 (3.20 Å) were observed in the 1CDK crystal structure complex. In comparative docking experiments, the docking of AMP-PNP with 1CDK revealed that the conformation and binding mode of AMP-PNP shows high similarity to that of the crystal structure (Fig. 4B). Another important finding from docking simulations suggests that the γ-phosphoryl group seems to play a significant role in hydrogen bonding interactions with the −NH atoms of Phe144 and Gly145 in aurora2. The 1CDK crystal structure complex shows that the 2′ hydroxyl group of the ribose moiety has potential hydrogen bonding interactions with Glu127 and Glu170. However, in the aurora2-AMP-PNP complex, the 2′OH group of the ribose moiety has a H-bonding interaction with Thr217 (Thr217-O… H of 2′OH-group, 2.40 Å). The conformation of AMP-PNP within the active site of aurora2 is maintained by seven hydrogen bonds formed arising from the ATP-binding site and the Gly/Lys pocket (Fig. 4). These hydrogen bonding interactions were computed from the final MD trajectories of the docked complex structure. The absence of the Glu260 hydrogen bonding interaction with the ribose moiety of aurora2 is attributable to the orientation of the β,γ-phosphoryl groups in the Gly/Lys pocket that prevent a strong electrostatic interaction with Lys162. The β- and γ-phosphoryl groups are also involved in close contacts with Lys143, Asn261, and Asp274 of the active site. The absence of critical H-bonding interactions with Lys162 and Glu260 presumably leads to an affinity binding of AMP-PNP (−69.3 kcal/mol) with the active-site residues of aurora2 that is lower than that with 1CDK (Table 1). These sequence and structural differences between cAMP-dependent PK and aurora2 kinase provide guidelines to better define the active-site pocket and to design novel and more specific aurora2 kinase inhibitors.

Aurora2-Staurosporine Complex.

In the case of affinity SA docking simulations on staurosporine (Fig. 2), the position and orientation within the active-site pocket of aurora2 were identical to those of the cAMP-dependent PK-staurosporine complex (1STC; Ref. 30). In particular, the hydrogen bonding interactions computed from affinity were identical to those in the crystal structure of the cAMP-dependent PK-staurosporine complex (Fig. 5). The amino acid residues that recognize and interact with staurosporine are Gly140, Val147, Glu181, Glu211, Ala213, Thr217, Asn261, and Asp274. A strong intermolecular hydrogen bond to Ala213 −NH with the lactam carbonyl group (2.03 Å) of staurosporine was observed in all of the structures obtained from MD simulations. However, Glu211-C⋕O… HN of the lactam amide hydrogen bonding interaction, present in the cAMP-dependent PK-staurosporine complex, was not observed in the initial docked aurora2-staurosporine complex. This is most likely attributable to the indolocarbazole and pyranosyl moieties of staurosporine, which induced conformational changes in the neighboring active site residues of aurora2. The staurosporine-induced conformational changes were investigated by calculating the rms between the initial and final complex structures obtained from MD simulations. The 0.23-Å difference observed in the active-site pocket of aurora2 on staurosporine binding is shown in Table 2. An analysis of the low-energy trajectory from the final run of simulations reveals that the amide proton forms an H-bond interaction with Glu211 (2.72 Å). It is interesting to note that the indolocarbazole moiety exhibits a more stable orientation in all of the structures generated during MD simulations. The phenyl ring syn to the methoxy group is positioned in the ATP-binding pocket and is in close contact with Glu181 and Asp274 residues and are within 2.50–3.00 Å, whereas the phenyl ring anti to methoxy group is oriented into the Gly-Lys pocket. The pyranosyl methylamino group is positioned within H-bonding distance to Thr217 (1.97 Å) and Glu260 (3.32 Å). On the basis of these observations and experimental data (Table 1), staurosporine (a nanomolar PK inhibitor) binds to aurora2 with a high binding energy of −91.6 kcal/mol (see Table 1).

Aurora2-H-series Complexes.

We have analyzed the mode of binding of three isoquinoline H-series PK inhibitors, H-89, H-8, and H-7 (Fig. 2), with aurora2 kinase (32, 33). The conformations of these isoquinolines were obtained from the crystal structure complexes of cAMP-dependent PKs and were subjected to affinity docking experiments without altering their crystallographic geometries. AMP-PNP served as a template for SA docking, and an analysis of dynamic trajectories revealed that the isoquinoline moiety occupied the ATP binding site and is positioned similarly to that of the adenine base of AMP-PNP. The N-alkyl chain of H-8 and H-89 extended into the β,γ-phosphoryl site. In all of the H-series isoquinolines, we observed that the nitrogen ring interacts with Ala213, which is located close to Glu211 at the hinge region and is in the purine binding site of AMP-PNP. Although the H-89 isoquinoline structure is different from that of the AMP-PNP, it has similar structural elements, including an isoquinoline nitrogen atom at position 13, two −NH groups at positions 4 and 17, and a sulfonyl group at position 1. These play an important role in the recognition of the active-site pocket of aurora2 and represent another class of S/T kinase inhibitors with an IC50 of 107 μm (Table 1). A number of SA docking simulations with different initial configurations identified a binding mode with a binding energy (−103.5 kcal/mol) substantially higher than that observed for AMP-PNP, staurosporine, and other isoquinoline structures. In this model, the N13 atom is positioned such that it H-bonds with the Ala213 −NH group (−N… HN-Ala213, 2.80 Å), which mimics the 1CDK Val123 amide H-bonding interaction (Fig. 6). Unlike the β,γ-phosphoryl group, the bromophenyl moiety is located deep in the Gly/Lys pocket and exhibits close hydrophobic interactions (3–3.50 Å) with Phe144 and Gly145. The carbonyl group of Asn261 in aurora2 forms a strong hydrogen bonding interaction with the N17 amide proton, and Glu260 is also in close contact with the same amide proton. The isoquinoline ring and sulfonamide moiety make a number of close contacts with Gly140, Gly145, and Val147. Moreover, the styrene moiety is also in close contact with Lys141 and Gly142. These interactions determine a stable binding mode observed in several dynamic trajectories analyzed within ±10 kcal/mol of that of the low-energy global structure generated during MD simulations.

For the H-8 and H-7 compounds (data not shown), the docking is based on the aurora2-H-89 complex. The N13 isoquinoline ring atom involved in H-bonding interactions with the Ala213 amide group is within a distance of 1.95–2.45 Å. The replacement of the bromophenyl styrene moiety of H-89 with the N-alkyl side chain and piprazine group in the case of H-8 and H-7 leads to less potent inhibitors of aurora2. Although the isoquinoline and sulfonyl moieties exhibit structural similarity, the H-series of compounds lacks a number of van der Waals and hydrophobic contacts with the enzyme. Such interactions play an important role for high binding affinity, as observed in the crystal structures of cAMP-dependent PK in complex with inhibitors. As shown in Fig. 6, the simulations demonstrate those structural changes that alter the shape of the sulfonyl substituents, leading to significant changes in binding energies (Table 1). Thus, the −NH of the N-methyl group of H-8, oriented to the Asp274 residue and its methyl function, are positioned within 3.5–4 Å. The sulfonyl substituent had only one contact with Vall47. Whereas the piprazine moiety is positioned within the ribose binding site of AMP-PNP, its −NH group is 3.5 Å away from Lys141. These results suggest that although H-8 and H-7 compounds bind to aurora2, the higher affinity of H-89 can be explained by its strong hydrogen bonding and hydrophobic interactions with both the ATP-binding site and the residues within the Gly/Lys pocket.

Aurora2-6,7-Dimethoxyquinazoline Complex.

Although quinazoline-containing compounds have been reported to be very potent and selective EGFR receptor tyrosine kinase inhibitors (39, 40), recent data suggest that these compounds also bind to S/T kinases (8). The 6,7-dimethoxyquinazoline which binds to aurora2 with an IC50 value of 0.00785 μm (8), represents another class of high-affinity aurora2 inhibitors. The aurora2-H-89 complex provided the basis for docking the quinazoline ring and, subsequently, the H-89 molecule was removed. The entire aurora2-6,7-dimethoxyquinazoline complex was optimized by several cycles of SA docking simulations and minimization using the CFF force field. From SA docking simulations, it was observed that the quinazoline moiety was positioned in an orientation that was similar to that of the adenine base of AMP-PNP and was anchored deep in the ATP binding site by forming three hydrogen bonds (Fig. 7). The N1 atom of quinazoline and N3 atom of the pyrimidine rings are involved in potential hydrogen bonding interactions with Ala213 (Ala213-NH… N1, 2.78 Å, energy −1.61 kcal/mol) and Asp274 (Asp274-NH… N3, 2.05 Å, energy −1.29 kcal/mol). Additionally, the −C⋕O of Asp274 is involved in hydrogen bonding interactions with 2-(N-benzoylamino) −NH group (Fig. 7). The oxygen atom of the 7-methoxy substituent is also positioned within hydrogen bonding distance with the backbone amide of Ala213. Glu211 has strong hydrogen bonding interactions with AMP-PNP and staurosporine, but no such an interaction was observed with the quinazoline compound. This is attributable to the lack of a donating functional group and strong steric interaction of methoxy group with Glu211. The 2-(N-benzoylamino) group had a similar position and orientation to that of the γ-phosphoryl group of AMP-PNP or the bromophenyl group of H-89 and interacts with Phe144 and Gly145, which in turn interacts with Lys143 located 3.50 Å from the carbonyl group. In the final model, we observed that the 5-aminopyrimidine moiety and the corresponding H-89 sulfonyl moiety positioned similarly with in the active site. As shown by the AMP-PNP model, the similar manner in which H-89 and quinazoline (Fig. 8) bind indicates that the polar, nonpolar, and hydrogen bonding interactions are essential for the high activity observed for these compounds. Such a large number of polar, nonpolar, and hydrogen bonding contacts with the active-site residues are observed for 6,7-dimethoxyquinazoline, which exhibits a high binding energy (−104.9 kcal/mol) compared with that of H-89 (Table 1). These results suggest that the rigid substituents of the β- and γ-phosphoryl positions are well tolerated.

Aurora2 in Complex with KN-93 and ML-7.

The binding modes of KN-93 and ML-7 (data not shown) were also explored using the above models. ML-7 was positioned and oriented similarly to H-7. The 4-methoxyphenyl sulfonyl amide moiety of KN-93 was found to occupy the ATP binding site of aurora2. The interaction of the chlorophenylstyrene group with Asp274 and Asn261 and its binding mode is dissimilar to that observed for the inhibitor class of H-89 and H-8 bound to aurora2. Interestingly, KN-93 did not exhibit any hydrogen bonding interactions. Specifically, the presence of the methyl (amino) methyl and hydroxyethyl groups had significant steric clashes with Gly140, Val147, and Asp274, presumably because of its rigid conformation.

Discussion

The development of potent and selective nucleotide analogue-based inhibitors that interact with individual PKs at their ATP binding site are being evaluated in various stages of clinical trials (41). Given the proof of concept that the inhibition of dysregulated PKs in human cancers is potentially achievable and could provide clinical efficacy, we have targeted human aurora2 kinase in pancreatic cancer. An iterative structure-based small-molecule drug design approach in combination with an in vitro kinase assay has been developed to test the potency and efficacy of novel aurora2 kinase inhibitors.

The tertiary structure models of aurora1 and aurora2 were built using the catalytic domain of PKA complexed with a 20-amino acid substrate analogue inhibitor and Mn2+-ATP (12). The cAMP-dependent PK structure revealed a two-domain PK with a deep cleft between the domains. Mn2+-ATP and a portion of the inhibitor peptide occupy the active site cleft. The NH2-terminal smaller domain is associated with nucleotide binding and its largely antiparallel β-sheet structure constitutes a nucleotide binding motif. The COOH-terminal larger domain is predominantly α-helical with a single β-sheet at the domain interface. This domain is primarily involved in peptide binding and catalysis. Most of the invariant amino acids in this conserved catalytic core are clustered at the sites of nucleotide binding and catalysis. The tertiary structure of the aurora2 model built on 1CDK has significant identity and similarity indicating a conserved catalytic core. The main differences in the active site of aurora2 and cAMP-dependent PK are the changed glycine-rich nucleotide-binding motif 50GKGKFGNV57, which should be sufficient to provide distinct substrate specificities and deletion of the last 30 residues from the COOH terminus of cAMP-dependent PK (Fig. 1) and to provide for a more compact aurora2 catalytic kinase core domain that may be relevant to its functional localization to mitotic spindles.

A number of crystal structures of cAMP-dependent PK in complex with ATP analogues and pseudo-substrates (42, 43) have provided details as to the mechanisms of inactivation by modulation of one or more of the four conserved structural elements; these are: (a) inhibition of the ATP binding pocket; (b) distortion of the glycine-rich loop; (c) alteration of the position of α-helix C; and (d) alteration of the conformation of the phosphorylation-dependent activation segment. The crystal structure determined for bovine cAMP-dependent PK is phosphorylated in the activation loop and is in complex with AMP-PNP and the pseudo-substrate peptide PKI (5–24). This catalytically competent enzyme is in a state immediately before phosphoryl transfer, with the cleft between the NH2- and COOH-terminal domains in a closed conformation. The tertiary structure modeling of aurora1 and aurora2 on the catalytically competent cAMP-dependent PK structure has defined pertinent structural features in the aurora kinases that has guided the rational design of novel chemical entities.

To gain insight into the binding mode of nucleoside analogue-based inhibitors, we have docked known S/T kinase inhibitors. The potency of these docked inhibitors has been evaluated in an in vitro kinase assay. The complex structure of AMP-PNP with cAMP-dependent PK and aurora2 from SA docking simulations reveals a network of hydrogen bonding interactions. The sequence differences in the active-site pocket between the two kinases explains the lower binding energy of the AMP-PNP-aurora2 complex. This is attributable to the weak hydrogen bonding interactions of the adenine base with Glu211, Ala213, and Glu127Thr217 mutation in aurora2, the absence of a hydrogen bonding interaction of the 2′OH group of ribose with Lys162 and Glu260 and the orientation of the β,γ-phosphoryl groups in the Gly/Lys pocket (Fig. 4). However, in the case of staurosporine, the position and orientation in the active site was identical to that of the cAMP-dependent PK-staurosporine complex (1STC; Ref. 30). In particular, the hydrogen bonding interactions computed from affinity were identical with the mostly conserved active-site residues (Fig. 5) and accommodated staurosporine by minor conformational changes during MD simulations (Table 2).

In the case of the H-series of compounds (H-89, H-8, and H-7) the isoquinoline moiety occupied the ATP binding site and is positioned similarly to that of the adenine base of AMP-PNP. The N-alkyl chain of H-89 extended into the β,γ-phosphoryl site and the isoquinoline nitrogen ring interacts with Ala213. Although the H-89 isoquinoline structure is different from that of the AMP-PNP, it has similar structural elements including an isoquinoline nitrogen atom at position 13 which is involved in a hydrogen bonding interaction with Ala213 (Val123 in 1CDK). The bromophenyl moiety is located deep in the Gly/Lys pocket and exhibits a hydrophobic interaction with Phe144 (Fig. 6). Furthermore, the strong hydrogen bonding and hydrophobic interactions with both the ATP and Gly/Lys pocket supports the calculated binding energy (Table 1). The aurora2-H-89 complex was used to model 6,7-dimethoxyquinazoline, and the quinazoline moiety is positioned similar to that of the adenine base of AMP-PNP and anchored deep in the ATP binding site (Fig. 7). The 2-(N-benzoylamino) group had similar positions and orientations to those of the γ-phosphoryl group of AMP-PNP and interacts with Gly145, which in turn is in close proximity to Phe144. The N1 atom of quinazoline and N3 atom of pyrimidine ring exhibits a strong hydrogen bonding interaction with Ala123 and Asp274. In quinazoline, the 2-substituted amide hydrogen is in close contact with Gly145 and its 5-substituted amide hydrogen has strong interactions with Gly140 and Val147. Furthermore, the aminopyrimidine and benzoly carbonyl groups are positioned close to Asp274. These interactions explain its high binding energy (−104.9 kcal/mol) when compared with that of H-89. Compounds KN-93 and ML-7 (data not shown) essentially had poor interactions with aurora2, mainly because of their rigid conformations and lack of hydrogen bonding interactions. The 4-chlorophenyl-2-propenyl-(methylamino)methylphenyl substituent of KN-93 at the 2-position is close to the hydrophilic pocket and appears to be sterically crowded. The 4-position projects toward the Gly/Lys pocket in the region occupied by the bromostyrene moiety of H-89. Taken together, these results suggest that SA docking complexes of aurora2 with H-89 and 6,7-dimethoxyquinazolines furnish a number of guidelines for analogue design.

For the docked structures, the calculated binding energies from the highest to the lowest are 6,7-dimethoxyquinazoline > H-89 > staurosporine > H-8 > AMP-PNP > KN-93 > ML-7 > H-7. For the in vitro kinase assay the IC50 values from the highest to the lowest are 6,7-dimethoxyquinazoline > staurosporine > H-89 > ML-7 > KN-93 (Table 1). These results qualitatively validate the modeling and docking studies described for this series of S/T kinase inhibitors. Molecular modeling was used to evaluate and select potential substituents for the isoquinoline and quinazoline templates. A de novo small fragment and available chemical database library using LUDI4 (44) and FlexX (45) virtual docking computational procedures were performed using SYBYL (Tripos, St. Louis, MO) to generate novel compounds specific to aurora2. The high-score compounds from FlexX having a quinazoline moiety were selected, and are currently being synthesized and screened, for aurora2 kinase inhibitory activity.

Recently, the crystal structure of the human aurora 2 kinase (46) was published (Brookhaven Protein Data Bank entry 1MUO on hold). The structure determination was based on molecular replacement using an aurora2 kinase homology model, which was based on 1CDK. This essentially validates the model for docking studies undertaken in this report. The main conclusions from the crystal structure were that the homology model is consistent with an overall kinase fold and active site except for a lack of electron density of the activation loop, which includes a conserved tryptophan. Although the coordinates of the crystal structure of aurora2 are not available, the binding mode of the docked AMP-PNP in the ATP binding site are similar in 1CDK and the modeled aurora2.

In conclusion, we have developed a structural bioinformatic approach that has facilitated the tertiary structure modeling of aurora2, studied the binding mode of ATP analogue and inhibitors, and evaluated these compounds using an in vitro kinase assay. A structural analysis of the docked ATP analogue and S/T kinase inhibitors has provided guidelines for the rational design of novel aurora2 kinase inhibitors. Furthermore, using the available chemical database and FlexX, we have identified a series of novel compounds that are currently been synthesized for in vitro and in vivo testing.

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

Structure-based sequence alignment in Clustal X of the catalytic PK domains of aurora2 (ARK1), aurora1 (ARK2), bovine cAMP-dependent PK (1CDK), murine cAMP-dependent PK (1APM), and C. elegans twitchin kinase (1KOA). The α-helices are α1 to α11 (black bars), β-strands are β1 to β11 (gray bars), identical residues (shaded and *), highly conserved residues (:), and similar residues (.). Magenta, the active site residues; green, inserted residues; red, deleted residues; blue, N- extensions of aurora2 and -1 not included in the modeling and the COOH-terminal residues of 1CDK and 1APM not included in the modeling. Threonine197 (1CDK) is equivalent to Thr288 (aurora2 and 1, yellow) and is the autophosphorylation site.

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

The structures of the ATP analogue and S/T kinase inhibitors evaluated for inhibition of aurora2 kinase activity. These are AMP-PNP, staurosporine, H-89, H-8, H-7, KN-93, ML-7, and 6,7-dimethoxyquinazoline.

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

A, the homology model of aurora2 kinase. Secondary structural elements are labeled as α-helix (red cylinder), β-sheet (yellow arrow), coil (green ribbon), and turns (blue ribbons). B, the Cα backbone superimposition of the final minimized structure of aurora2 (blue) on cAMP-dependent PK (1CDK; red) with an rms deviation of 0.23 Å.

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

A, binding mode of AMP-PNP in the ATP binding site of aurora2 kinase. The purine base occupies the adenine binding site and the β- and γ-phosphoryl groups are buried in the Gly/Lys pocket. Black dashed lines, the hydrogen bonds; distances are in Å. B, binding mode of AMP-PNP in the ATP binding site of 1CDK. The potential hydrogen bonding interactions and the active-site residues involved in binding are shown.

Fig. 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 5.

The binding mode of aurora2-staurosporine (color-by-atom type) complex. The active site residues of the protein involved in interactions are shown; black dashed lines, the hydrogen bonds.

Fig. 6.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 6.

The stereo-view of the binding mode of the aurora2-H89 complex. The isoquinoline ring is positioned deep in the ATP binding site and interacts with A213 and N261. The bromophenyl moiety extends into the Gly/Lys pocket. Dotted lines, the two hydrogen bonds; labels, the active site residues.

Fig. 7.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 7.

The binding mode of the aurora2-6,7-dimethoxyquinazoline complex. The active site backbone of the protein is shown as a blue ribbon, the quinazoline ring (color-by-atom) anchored deep in the purine binding site and is involved in hydrogen bonding interactions with E211 and D274 (black dashed lines; see “Results” for explanation). The active site residues involved in the interactions are labeled and clipped for a better view of the binding site.

Fig. 8.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 8.

The superposed structures of staurosporine (ash), 6,7-dimethoxyquinazoline (blue), H-89 (red), and AMP-PNP (magenta) in the ATP-binding pocket of aurora2. The enzyme active site is clipped and residues shown in color-by-atom type.

View this table:
  • View inline
  • View popup
Table 1

The measured (IC50) and calculated energies (kcal/mol) of the ATP analogue and inhibitors tested for aurora2 kinase binding

View this table:
  • View inline
  • View popup
Table 2

The calculated rms deviation (Å) for the superimposed aurora2 and aurora2-staurosporine complex

The rms deviations were calculated for the active-site residues and the Cα backbone.

Acknowledgments

We thank Drs. Laurence H. Hurley and Emmanuelle J. Meuillet-May for helpful discussions and Dr. David M. Bishop for preparing, proofreading, and editing the manuscript and figures.

Footnotes

  • The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked advertisement in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

  • ↵1 Supported by NIH Grants CA95031 and CA88310 from the National Cancer Institute.

  • ↵3 The abbreviations used are: S/T kinase, serine/threonine kinase; AMP-PNP, adenylyl imidodiphosphate; NCBI, National Center for Biotechnology Information; MD, molecular dynamic(s); PK, protein kinase; rms, root mean square; SA, simulated annealing.

  • 4 INSIGHT II 2000. Molecular modeling software, Accelrys Inc., San Diego, CA.

  • ↵5 Discover_3 (Version 2.9.5). Molecular mechanics force fields, INSIGHT II, 2000. Molecular modeling software, Accelrys Inc., San Diego, CA.

    • Accepted December 20, 2002.
    • Received August 23, 2002.
    • Revision received December 4, 2002.
  • Molecular Cancer Therapeutics

References

  1. ↵
    Nigg, E. A. Mitotic kinases as regulators of cell division and its checkpoints.Nat. Rev. Mol. Cell. Biol. , 2:21 –32,2001 .
    OpenUrlCrossRefPubMed
  2. ↵
    Bischoff, J. R., and Plowman, G. D. The Aurora/Ipl1p kinase family: regulators of chromosome segregation and cytokinesis.Trends Cell Biol. , 9:454 –459,1999 .
    OpenUrlCrossRefPubMed
  3. ↵
    Bischoff, J. R., Anderson, L., Zhu, Y., Mossie, K., Ng, L., Souza, B., Schryver, B., Flanagan, P., Clairvoyant, F., Ginther, C., Chan, C. S. M., Novotny, M., Slamon, D. J., and Plowman, G. D. A homologue of Drosophila aurora kinase is oncogenic and amplified in human colorectal cancers.EMBO J. , 17:3052 –3065,1998 .
    OpenUrlAbstract
  4. ↵
    Giet, R., and Prignet, C. Aurora/Ipl1p-related kinases, a new oncogenic family of mitotic serine-threonine kinases.J. Cell Sci. , 112:3591 –3601,1999 .
    OpenUrlAbstract/FREE Full Text
  5. ↵
    Tatsuka, M., Katayama, H., Ota, T., Tanaka, T., Odashima, S., Suzuki, F., and Terada, Y. Multinuclearity and increased ploidy caused by overexpression of the aurora- and Ipl1-like midbody-associated protein mitotic kinase in human cancer cells.Cancer Res. , 58:4811 –4816,1998 .
    OpenUrlAbstract/FREE Full Text
  6. ↵
    Zhou, H., Kuang, J., Zhong, L., Kuo, W. L., Gray, J. W., Sahin, A., Brinkley, B. R., and Sen, S. Tumour amplified kinase STK15/BTAK induces centrosome amplification, aneuploidy and transformation.Nat. Genet. , 20:189 –193,1998 .
    OpenUrlCrossRefPubMed
  7. ↵
    Han, H., Bearss, D. J., Browne, L. W., Calaluce, R., Nagle, R. B., and Von Hoff, D. D. Identification of differentially expressed genes in pancreatic cancer cells using a cDNA microarray.Cancer Res. , 62:2890 –2896,2002 .
    OpenUrlAbstract/FREE Full Text
  8. ↵
    Austen, M., John, K., Henri, F., and George, B.Preparation of quinazolines as aurora 2 kinase inhibitors. PCT Int. Appl, WO0121596 United Kingdom:2001 .
  9. ↵
    Sun, L., Tran, N., Liang, C., Tang, F., Rice, A., Schreck, R., Waltz, K., Shawver, L. K., McMahon, G., and Tang, C. Design, synthesis, and evaluations of substituted 3-[(3- or 4-carboxyethylpyrrol-2-yl)methylidenyl]indolin-2-ones as inhibitors of VEGF, FGF, and PDGF receptor tyrosine kinases.J. Med. Chem. , 42:5120 –5130,1999 .
    OpenUrlCrossRefPubMed
  10. Krystal, G. W., Honsawek, S., Kiewlich, D., Liang, C., SVasile, S., Sun, L., McMahon, G., and Lipson, K. E. Indolinone tyrosine kinase inhibitors block kit activation and growth of small cell lung cancer cells.Cancer Res. , 61:3660 –3668,2001 .
    OpenUrlAbstract/FREE Full Text
  11. ↵
    Knighton, D. R., Cadena, D. L., Zheng, J., Eyck, L. F. T., Taylor, S. S., Sowadski, J. M., and Gill, G. N. Structural features that specify tyrosine kinase activity deduced from homology modeling of the epidermal growth factor receptor.Proc. Natl. Sci. USA, USA. , 90:5001 –5005,1993 .
    OpenUrlAbstract/FREE Full Text
  12. ↵
    Bossemeyer, D., Engh, R. A., Kinzel, V., Ponstingl, H., and Huber, R. Phosphotransferase and substrate binding mechanism of the cAMP-dependent protein kinase catalytic subunit from porcine heart as deduced from the 2.0 Å structure of the complex with Mn2+ adenylyl imidodiphosphate and inhibitor peptide PKI(5–24).EMBO J. , 12:849 –859,1993 .
    OpenUrlPubMed
  13. ↵
    Zheng, J., Trafny, E., Knighton, D., Xuong, N-H., Taylor, S., Ten Eyck, L., and Sowadski, J. 2.2 Å refined crystal structure of the catalytic subunit of cAMP-dependent protein kinase complexed with MnATP and a peptide inhibitor.Acta Crystallogr. D Biol. Crystallogr. , 49:362 –365,1993 .
  14. ↵
    Knighton, D. R., Zheng, J. H., Ten Eyck, L. F., Ashford, V. A., Xuong, N. H., Taylor, S. S., and Sowadski, J. M. Crystal structure of the catalytic subunit of cyclic adenosine monophosphate-dependent protein kinase.Science (Wash. DC) , 253:407 –414,1991 .
    OpenUrlAbstract/FREE Full Text
  15. ↵
    Kobe, B., Heierhorst, J., Feil, S. C., Parker, M. W., Benian, G. M., Weiss, K. R., and Kemp, B. E. Giant protein kinases: domain interactions and structural basis of autoregulation.EMBO J. , 15:6810 –6821,1996 .
    OpenUrlPubMed
  16. ↵
    Jones, D. T. GenTHREADER: an efficient and reliable protein fold recognition method for genomic sequences.J. Mol. Biol. , 287:797 –815,1999 .
    OpenUrlCrossRefPubMed
  17. ↵
    Kelley, L. A., MacCallum, R. M., and Sternberg, M. J. E. Enhanced genome annotation using structural profiles in the program 3D-PSSM.J. Mol. Biol. , 299:499 –520,2000 .
    OpenUrlPubMed
  18. ↵
    Sussman, J. L., Lin, D., Jiang, J., N. Manning, N. O., Prilusky, J., Ritter, O., and Abola, E. E. Protein data bank (PDB): database of three-dimensional structural information of biological macromolecules.Acta Crystallogr. D Biol. Crystallogr. , 54:1078 –1084,1998 .
    OpenUrlCrossRefPubMed
  19. ↵
    Taylor, W. R., and Orengo, C. A. Protein structure alignment.J. Mol. Biol. , 208:1 –22,1989 .
    OpenUrlCrossRefPubMed
  20. ↵
    Thompson, J. D., Gibson, T. J., Plewniak, F., Jeanmougin, F., and Higgins, D. G. The CLUSTAL X windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools.Nucleic Acid Res. , 24:4876 –4882,1997 .
  21. ↵
    Maple, J. R., Hwang, M-J., Jalkanen, K. J., Stockfisch, T. P., and Hagler, A. T. Derivation of class II force fields: V. Quantum force field for amides, peptides, and related compounds.J. Comp. Chem. , 19:430 –458,1998 .
    OpenUrlCrossRef
  22. ↵
    Garemyr, R., and Elofsson, A. Study of the electrostatics treatment in molecular dynamics simulations.Proteins Struct. Funct. Genet. , 37:417 –428,1999 .
    OpenUrlCrossRefPubMed
  23. ↵
    Orozco, M., Laughton, C. A., Herzyk, P., and Neidle, S. Molecular-mechanics modelling of drug-DNA structure: the effects of differing dielectric treatment on helix parameters and comparison with a fully solvated model.J. Biomol. Struct. Dyn. , 8:359 –373,1990 .
    OpenUrlPubMed
  24. ↵
    Luthy, R., Bowie, J. U., and Eisenberg, D. Assessment of protein models with three-dimensional profiles.Nature (Lond.) , 356:83 –85,1992 .
    OpenUrlCrossRefPubMed
  25. ↵
    Laskowski, R. A., MacArthur, M. W., Moss, D. S., and Thornton, J. M. PROCHECK: a program to check the stereochemical quality of protein structures.J. Appl. Cryst. , 26:283 –291,1993 .
  26. ↵
    Marhefka, C. A., Moore, I. I., B. M., Bishop, T. C., Kirkovsky, L., Mukherjee, A., Dalton, J. T., and Miller, D. D. Homology modeling using multiple molecular dynamics simulations and docking studies of the human androgen receptor ligand binding domain bound to testosterone and nonsteroidal ligands.J. Med. Chem. , 44:1729 –1740,2001 .
    OpenUrlCrossRefPubMed
  27. ↵
    Kiyama, R., Tamura, Y., Watanabe, F., Tsuzuki, H., Ohtani, M., and Yodo, M. Homology modeling of gelatinase catalytic domains and docking simulations of novel sulfonamide inhibitors.J. Med. Chem. , 42:1723 –1738,1999 .
    OpenUrlCrossRefPubMed
  28. ↵
    Luty, B. A., Wasserman, Z. R., Stouten, P. F. W., Hodge, C. N., Zacharias, M., and McCammon, J. A. A molecular mechanics/grid method for evaluation of ligand-receptor interactions.J. Comp. Chem. , 16:454 –464,1995 .
    OpenUrlCrossRef
  29. ↵
    Stouten, P. F. W., Frömmel, C., Nakamura, H., aand Sander, C. An effective solvation term based on atomic occupancies for use in protein simulations.Mol. Simulation , 10:97 –120,1993 .
  30. ↵
    Prade, L., Engh, R. A., Girod, A., Kinzel, V., Huber, R., and Bossemeyer, D. Staurosporine-induced conformational changes of cAMP-dependent protein kinase catalytic subunit explain inhibitory potential.Structure (Lond.) , 5:1627 –1637,1997 .
    OpenUrlCrossRefPubMed
  31. ↵
    Link, J. T., Raghavan, S., Gallant, M., Danishefski, S. J., Chou, T. C., and Ballas, L. M. Staurosporine and ent-staurosporine: the first total synthesis, prospects for a regioselective approach, and activity profile.J. Am. Chem. Soc. , 118:2825 –2842,1996 .
    OpenUrlCrossRef
  32. ↵
    Engh, R. A., Gorod, A., Kinzel, V., Huber, R., and Bossemeyer, D. Crystal structures of catalytic subunit of cAMP-dependent protein kinase in complex with isoquinolinesulfonyl protein kinase inhibitors H7, H8, and H89.J. Biol. Chem. , 271:26157 –26164,1996 .
    OpenUrlAbstract/FREE Full Text
  33. ↵
    Xu, R-M., Carmel, G., Kuret, J., and Cheng, X. Structural basis for selectivity of the isoquinoline sulfonamide family of protein kinase inhibitors.Proc. Natl. Acad. Sci. USA , 93:6308 –6313,1996 .
    OpenUrlAbstract/FREE Full Text
  34. ↵
    Anderson, M. E., Braun, A. P., Wu, Y., Lu, T., Wu, Y., Schulman, H., and Sung, R. J. KN-93, an inhibitor of multifunctional Ca++/calmodulin-dependent protein kinase, decreases early after-depolarizations in rabbit heart.J. Pharmacol. Exp. Ther. , 287:996 –1006,1998 .
    OpenUrlAbstract/FREE Full Text
  35. ↵
    Matsuoka, I., Nakahata, N., and Nakanishi, H. Selective inhibition of collagen-induced arachidonic acid liberation by 1-(5-iodonaphthalene-1-sulfonyl)-1H-hexahydro-1,4-diazepine hydrochloride (ML-7), a myosin light chain kinase inhibitor, in washed rabbit platelets.Biochem. Pharmacol. , 54:1019 –1026,1997 .
    OpenUrlCrossRefPubMed
  36. ↵
    Austen, M., and John, K.Preparation of quinazolines as aurora 2 kinase inhibitors. PCT Int. Appl. WO 0121597 (UK).,208 pp,2001 .
  37. ↵
    Åqvist, J., Medina, C., and Samuelsson, J-E. A new method for predicting binding affinity in computer-aided drug design.Protein Eng. , 7:385 –391,1994 .
    OpenUrlAbstract/FREE Full Text
  38. ↵
    Riccio, A., Tamburrini, M., Giardina, B., and Prisco, G. Molecular dynamics analysis of a second phosphate site in the hemoglobins of the seabird, south polar skua. Is there a site-site migratory mechanism along the central cavity.Biophys J. , 81:1938 –1946,2001 .
    OpenUrlPubMed
  39. ↵
    Bridges, A. J., Zhou, H., Cody, D. R., Rewcastle, G. W., McMicheal, A., Hollis Showalter, H. D., Fry, D. W., Kraker, A. J., and Denny, W. A. Tyrosine kinase inhibitors. 8. An unusually steep structure–activity relationship for analogues of 4-(3-bromoanilino)-6, 7-dimethoxyquinazoline (PD 153035), a potent inhibitor of the epidermal growth factor receptor.J. Med. Chem. , 39:267 –276,1996 .
    OpenUrlCrossRefPubMed
  40. ↵
    Bridges, A. J. Chemical inhibitors of protein kinases.Chem. Rev. , 101:2541 –2572,2001 .
    OpenUrlCrossRefPubMed
  41. ↵
    Scapin, G. Structural biology in drug design: selective protein kinase inhibitors.Drug Discovery Today , 7:601 –611,2002 .
    OpenUrlCrossRefPubMed
  42. ↵
    Narayana, N., Cox, S., Xuong, N-H., Ten Eyck, L. F., and Taylor, S. S. A binary complex of the catalytic subunit of cAMP-dependent protein kinase and adenosine further defines conformational flexibility.Structure (Lond.) , 5:921 –935,1997 .
    OpenUrlCrossRefPubMed
  43. ↵
    Narayana, N., Diller, T. C., Koide, K., Bunnage, M. E., Nicalaou, K. C., Brunton, L. L., Xuong, N-H., Ten Eyck, L. F., and Taylor, S. S. Crystal structure of the potent natural product inhibitor Balanol in complex with the catalytic subunit of cAMP-dependent protein kinase.Biochemistry , 38:2367 –2376,1999 .
    OpenUrlCrossRefPubMed
  44. ↵
    Bohm, H. J. The computer program LUDI: a new method for the de novo design of enzyme inhibitors.J. Comput-Aided Mol. Des. , 6:61 –78,1992 .
  45. ↵
    Rarey, M., Kramer, B., Lenguer, T., and Klebe, G. A. A fast flexible docking method using an incremental construction algorithm.J. Mol. Biol. , 261:470 –489,1996 .
    OpenUrlCrossRefPubMed
  46. ↵
    Cheetham, G. M. T., Knegtel, R. M. A., Coll, J. T., Renwick, S. B., Swenson, L., Weber, P., Lippke, J. A., and Austen, D. A. Crystal structure of aurora-2, an oncogenic serine/threonine kinase.J. Biol. Chem. , 277:42419 –42422,2002 .
    OpenUrlAbstract/FREE Full Text
View Abstract
PreviousNext
Back to top
Molecular Cancer Therapeutics: 2 (3)
March 2003
Volume 2, Issue 3
  • Table of Contents
  • About the Cover

Sign up for alerts

View this article with LENS

Open full page PDF
Article Alerts
Sign In to Email Alerts with your Email Address
Email Article

Thank you for sharing this Molecular Cancer Therapeutics article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
Targeting Aurora2 Kinase in Oncogenesis: A Structural Bioinformatics Approach to Target Validation and Rational Drug Design1
(Your Name) has forwarded a page to you from Molecular Cancer Therapeutics
(Your Name) thought you would be interested in this article in Molecular Cancer Therapeutics.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Citation Tools
Targeting Aurora2 Kinase in Oncogenesis: A Structural Bioinformatics Approach to Target Validation and Rational Drug Design1
Hariprasad Vankayalapati, David J. Bearss, José W. Saldanha, Rubén M. Muñoz, Sangeeta Rojanala, Daniel D. Von Hoff and Daruka Mahadevan
Mol Cancer Ther March 1 2003 (2) (3) 283-294;

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Share
Targeting Aurora2 Kinase in Oncogenesis: A Structural Bioinformatics Approach to Target Validation and Rational Drug Design1
Hariprasad Vankayalapati, David J. Bearss, José W. Saldanha, Rubén M. Muñoz, Sangeeta Rojanala, Daniel D. Von Hoff and Daruka Mahadevan
Mol Cancer Ther March 1 2003 (2) (3) 283-294;
del.icio.us logo Digg logo Reddit logo Twitter logo CiteULike logo Facebook logo Google logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Jump to section

  • Article
    • Abstract
    • Introduction
    • Materials and Methods
    • Results
    • Discussion
    • Acknowledgments
    • Footnotes
    • References
  • Figures & Data
  • Info & Metrics
  • PDF
Advertisement

Related Articles

Cited By...

More in this TOC Section

  • Prediction of individual response to platinum/paclitaxel combination using novel marker genes in ovarian cancers
  • Low doses of cisplatin or gemcitabine plus Photofrin/photodynamic therapy: Disjointed cell cycle phase-related activity accounts for synergistic outcome in metastatic non–small cell lung cancer cells (H1299)
  • Semisynthetic homoharringtonine induces apoptosis via inhibition of protein synthesis and triggers rapid myeloid cell leukemia-1 down-regulation in myeloid leukemia cells
Show more Article
  • Home
  • Alerts
  • Feedback
  • Privacy Policy
Facebook  Twitter  LinkedIn  YouTube  RSS

Articles

  • Online First
  • Current Issue
  • Past Issues
  • Meeting Abstracts

Info for

  • Authors
  • Subscribers
  • Advertisers
  • Librarians

About MCT

  • About the Journal
  • Editorial Board
  • Permissions
  • Submit a Manuscript
AACR logo

Copyright © 2021 by the American Association for Cancer Research.

Molecular Cancer Therapeutics
eISSN: 1538-8514
ISSN: 1535-7163

Advertisement