Abstract
The identification as cooperating targets of Proviral Integrations of Moloney virus in murine lymphomas suggested early on that PIM serine/threonine kinases play an important role in cancer biology. Whereas elevated levels of PIM1 and PIM2 were mostly found in hematologic malignancies and prostate cancer, increased PIM3 expression was observed in different solid tumors. PIM kinases are constitutively active and their activity supports in vitro and in vivo tumor cell growth and survival through modification of an increasing number of common as well as isoform-specific substrates including several cell cycle regulators and apoptosis mediators. PIM1 but not PIM2 seems also to mediate homing and migration of normal and malignant hematopoietic cells by regulating chemokine receptor surface expression. Knockdown experiments by RNA interference or dominant-negative acting mutants suggested that PIM kinases are important for maintenance of a transformed phenotype and therefore potential therapeutic targets. Determination of the protein structure facilitated identification of an increasing number of potent small molecule PIM kinase inhibitors with in vitro and in vivo anticancer activity. Ongoing efforts aim to identify isoform-specific PIM inhibitors that would not only help to dissect the kinase function but hopefully also provide targeted therapeutics. Here, we summarize the current knowledge about the role of PIM serine/threonine kinases for the pathogenesis and therapy of hematologic malignancies and solid cancers, and we highlight structural principles and recent progress on small molecule PIM kinase inhibitors that are on their way into first clinical trials.Identification of the PIM family of constitutively active proto-oncogenic serine/threonine kinases
Over two decades ago, cloning of retroviral integration sites in murine Moloney leukemia virus induced lymphomas has led to the identification of the PIM (Proviral Integration Moloney virus) gene locus.1 Over 50% of early T-cell lymphomas showed integrations near the PIM1 locus leading to deregulated expression of the PIM1 mRNA. The PIM1 gene locus was mapped to mouse chromosome 17, and to short arm of chromosome 6 (6p21) in the human genome. Further analysis revealed that the open reading frame of PIM1 encoded for a protein of 313aa extending over 6 exons, with highest homology to serine/threonine kinases.2 Predisposition to lymphomagenesis in PIM1 transgenic mice through cooperation with c-myc and N-myc demonstrated the proto-oncogenic activity of PIM1.3 Subsequent studies have characterized PIM1 as synergizing oncogene with over-expressed BCL2, GFI1, loss of FAS-L, or in collaboration of a leuke-mogenic fusion gene (E2A-PBX1).4 The PIM1 gene encodes for two isoforms of 34 and 44kD through the use of alternative initiation sites. Both isoforms contain the kinase domain and exhibited comparable in vitro kinase activity.5 PIM1 was found ubiquitously expressed and to function as a protein with a short half-life. Interestingly, the half-life of PIM1 (< 5 min) observed in normal peripheral leukocytes was significantly increased in K562, a Philadelphia chromosome-positive leukemia cell line derived from chronic myeloid leukemia in blast crisis.6 Abundant levels of PIM1 were found in hematopoietic cells. Moreover, sustained PIM1 expression was induced by cytokines that signal through structurally related receptors such as IL-3, GM-CSF, G-CSF or IL-6.7 Subsequently, several studies have documented that PIM1 is a major downstream target of the signal transducer and activator of transcription (STATs) induced by a large variety of additional receptors such as IL-2, IL-7, IL-9, IFNγ, EPO, FLT3 or TPO.7 PIM1 expression is not only regulated at the transcriptional, but also at the posttranscriptional, translational and posttranslational levels (Figure 1). Other studies have shown that PIM1 kinase is significantly protected from proteasomal degradation by heat shock proteins (Hsp70, Hsp90).8,9 Moreover, it has been proposed that micro-RNAs, miR-1 and miR-210, might be implicated in regulation of PIM1 expression.10,11
Germline inactivation of the PIM1 gene was associated with a mild phenotype as PIM1 deficient mice are ostensibly normal, healthy and fertile. However, subtle functional defects of the hematopoietic system have been identified: PIM1 mice showed erythrocytic microcytosis and PIM1 B cells and bone marrow-derived mast cells were impaired in interleukin-7 (IL-7) or IL-3 induced proliferation.12,13
Retroviral insertion site cloning in secondary transplants of Moloney murine virus induced lymphomas revealed PIM2 as a frequent but late event in tumorigenesis.14 Interestingly, proviral tagging in c-myc transgenic mice lacking PIM1 has led to compensatory activation of PIM2. The PIM2 gene located on chromosome Xp11 comprises 6 exons and is 53% identical to PIM1 at the amino acid level and shares preference and usage of non-AUG alternative initiation codons leading to 3 different isoforms. PIM2 is ubiquitously expressed with highest levels in brain and lymphoid cells, and like PIM1, PIM2 also potently synergizes in c-MYC induced lymphomagenesis.15
Through high throughput retroviral tagging in tumors of c-myc transgenic mice lacking PIM1 and PIM2, Mikkers and colleagues found selective activation of PIM3 suggesting that PIM3 can substitute for PIM1 and PIM2 in MuLV-induced lymphomagenesis.16 The PIM3 gene is located on chromosome 22q and encodes for a serine/threonine kinase with over 60% homology to PIM1 and PIM2, that is ubiquitously expressed with highest levels in kidney, breast and brain.17
PIM1, PIM2 and PIM3 compound knockout mice that survived the perinatal period displayed a profound reduction in body size suggesting that PIMs are important for body growth. Colony forming assays with bone marrow from PIM1PIM2PIM3 mice demonstrated that PIMs act redundantly in clonogenic growth in response to IL-3, IL-5, SCF and TPO. However, PIM1 seems to be the most crucial isoform for these responses. Despite these defects, it was possible to establish PIM compound knockout mice that were viable and fertile suggesting that the PIM family of serine/threonine kinases is important but dispensable for growth factor signaling.18
The oncogenic activity of PIM serine/threonine kinases is mediated by multiple cellular substrates
Expression of recombinant PIM1 protein demonstrated its activity as serine/threonine kinase. Interestingly, in addition to phosphorylation of cellular substrates, in vitro experiments also suggested autophosphorylation activity of PIM1.7,19 Examination of the substrate sequence specificity of PIM1 revealed strong preference for peptides containing (K/R)3-X-S/T-X (X=neither basic nor large hydrophobic residue)20 and positional peptide library screens identified a consensus sequence (ARKRRRHPS-GPPTA) that bound with low nM affinity to PIM kinases.21,22 Mass-spectrometric analysis of Xenopus laevis PIM1 identified potential autophosphorylation sites including serine 190 and threonine 205 both conserved between species.23 We have recently observed phosphorylation of PIM1 after heterologous expression in E.coli located on residue Ser261. However, this site does not seem to be phosphorylated in vitro, instead other putative phospho-rylation sites were found of which one would map to Ser8 located in the unstructured N-terminus of PIM1 (S Knapp, unpublished observation, 2009). Negative regulation of the stability of PIM kinases by the protein phosphatase 2A (PP2A) suggested that autophosphorylation and/or phosphorylation by so far unknown regulatory kinases might be important determinants of PIM function.24,25 Modification of PIM1 function through phosphorylation by an upstream kinase has also been proposed in interleukin-6 (IL-6)-treated prostate cancer cell lines. Phosphorylation of PIM1 on Tyr218 by the ETK tyrosine kinase was required for IL-6-induced activation of androgen receptor-mediated transcription.26,27 As PIM proteins harbor over 35 potential recognition sites for a large variety of kinases, the impact of PIM autophosphorylation or heterologous phosphorylation by upstream regulators for the oncogenic function needs further clarification (Figure 2).
Gain of function studies through overexpression of human or mouse PIM cDNAs in different cell types showed that PIM kinases act as survival factors by blocking apoptotic cell death through regulation of pro- and anti-apoptotic members of the BCL2 protein family.7,28 Several studies revealed a key role of PIM1 in cytokine-mediated survival signaling in murine mast cells and primary human basophils and eosinophils.13,29 The pro-survival activity of all three PIM kinases in several cell types might be (at least in part) explained by inactivation of the pro-apoptotic protein BAD through phosphorylation of the Ser112 gatekeeper residue and presumably also Ser136 and Ser155.30–33 Most recent work suggests that phosphorylation of the proline-rich Akt substrate 1 (PRAS1) might also have the potential to block apoptosis of murine factor-dependent hematopoietic progenitor cells (FDCP-1).34 In addition, PIM1 seem to impair the activity of the apoptosis signaling kinase 1 (ASK1), by direct phosphorylation, resulting in protection from H2O2-induced cell death of H1299 lung cancer cells.35
Intriguingly, overexpression of PIM1 also stimulated cell death signaling in Rat1 fibroblasts elicited by c-MYC most probably through interaction and modification of the Cdc25A cell cycle phosphatase.36 A series of in vitro experiments demonstrated that collaboration of PIM1 with c-myc is essential for STAT3-mediated cell cycle progression and survival of hematopoietic Ba/F3 cells.37 Recent in vitro work using human vascular endothelial cells suggested that functional cooperation of PIM kinases with c-myc might be based on PIM-mediated phosphorylation of his-tone H3. PIM1 seems to be recruited to the E-box elements of MYC leading to a MYC-MAX-PIM1 complex. This complex phosphorylates H3-S10 stimulating then the binding of RNA polymerase II that contributes to transcriptional activation of a subset of MYC target genes.38 However, it is currently not known which of the PIM1-co-regulated MYC target genes might be essential for transformation or whether H3-S10 can also be phosphorylated by PIM2 or PIM3. Interestingly, phosphorylation of H3 by PIM1 seems to provide essential docking sites for acetylation of Histone H4 at lysine 16 (H4K16ac) by the MOF his-tone acetyltransferase. The resulting nucleosomal mark then allows binding of the BRD4 bromodomain protein and the positive elongation factor b (p-TEFb) mediating transcriptional elongation.39
Murine hematopoietic cells (FDCP-1) were protected from irradiation or adriamycin-induced apoptosis by over-expression of PIM1. Interestingly, cellular protection was associated with nuclear localization of a large fraction of the short (p33) but not the long (p44) PIM1 isoform suggesting the existence of functionally important isoform-specific cellular substrates.40 A PIM1 consensus site (RKR-RQTSM) was found in the cell cycle regulator p21 (CDKN1A). PIM1 associated with and phosphorylated p21 on Thr145 resulting in stabilization and nuclear translocation. These observations made in various cell lines suggested that deregulated PIM1 activity might contribute to tumorigenesis at least in part by regulation of p21.41,42 PIM kinases seem also to regulate the p27 (CDKN1B) cyclin-dependent kinase inhibitor. All three PIM kinases bound and directly phosphorylated p27 at residues Thr157 and Thr198 that allows binding of p27 to 14-3-3 proteins resulting in its nuclear export and pro-teosome-dependent degradation. Through phosphorylation and inactivation of FoxO1a and FoxO3a, PIM kinases seem to directly repress p27 transcription as shown in solid cancer and leukemia cell lines.43 PIM kinases seem not only to interfere with G1-S (through modification of Cdc25A, p21 and p27) but also with the G2-S transition of the cell cycle by phosphorylating Cdc25C phosphatase and the Cdc25C-associated kinase (C-TAK1).44,45
Recognition motif-based searches as well as protein-protein interaction screens resulted in identification of several putative PIM substrates including SND1 (p100), PAP-1 (RP9), HP1γ (CBX3), SNX6, SOCS-1 and -3, RPS19, RUNX-1 and -3, ABCG2/BRC, API5, MYB, MYC, NFAT1, NUMA, PTPRO and p65/REL-A46–61 (Online Supplementary Table S1). Although most of these proposed substrates have not been validated as being in vivo executors of the proto-oncogenic function of the PIM kinases, some of them are of special interest.
The SOCS proteins are well-characterized regulators of the JAK/STAT signal transduction pathway. Two groups have reported that the stability of SOCS1 is regulated through interaction and phosphorylation by PIM serine/threonine kinases.50,51 Structure functional analysis showed that the N-terminal 79aa of SOCS1 are essential for interaction with PIM. However, the targeted S/T residues (>30 potential phospho-motifs are present but no apparent PIM recognition site can be found in this region) remain unknown. Malignant transformation by the v-Abl oncogene was associated with phosphorylation of SOCS1 on non-tyrosine residues and inhibition of SOCS1 mediated proteasomal targeting of activated JAK kinases. Interestingly, v-Abl was not able to efficiently transform primary bone marrow cells from PIM1PIM2 compound knockout mice but combined deficiency of PIM1/PIM2 and SOCS1 resulted in partial restoration of the v-Abl transformation efficiency. These observations suggested that modulation of SOCS1 by PIM kinases might play a key role in v-Abl mediated malignant transformation.62 Taken together, PIM kinases might participate on several levels in a complex feedback mechanism regulating the JAK2/STAT5 signaling pathway that plays an essential role in induction and maintenance of hematologic malignancies.63
Mutations in the gene encoding the ribosomal protein RPS19 account for about 25% of patients with Diamond Blackfan anemia (DBA). This is a congenital bone marrow failure syndrome that typically results in macrocytic anemia within the first year of life, associated with birth defects and increased incidence of cancer.64 Possible interaction and phosphorylation of RPS19 in the ribosome by PIM1 is interesting as PIM1 gene dosage seems to directly influence the size of the erythrocytes: PIM1 mice have a decreased erythrocyte size, whereas PIM1 transgenic mice exhibit an increase in the median cellular erythrocyte volume.12,52 Although the PIM1 targeted residues in RPS19 have not been mapped, the most common missense mutations in RPS19 are located in a serine/threonine-rich region predicted to be targeted by kinases.64 To address whether PIM1 could be a candidate gene for DBA forms without RPS19 mutations, Chiochetti and colleagues performed a mutational screen involving 99 DBA patients and found 2 missense mutations Pro311Thr (C/A) and Cys17Tyr (G/A).52 These mutations located at the very 5’ and 3’ end of PIM1 and their functional consequence remains to be studied.
Increasing evidence supports functional links between the integrity of ribosomal proteins and the p53 tumor suppressor. In mice, mutations in RPS19 and RPS20 caused dark skin and reduced erythrocyte count in mice through activation of p53.65 Whether PIM1 modulates p53 through interaction of RPS19 is currently not known. However, elevated PIM1 levels seemed to induce the p53 pathway in different cancer cell lines and murine embryonic fibroblasts (MEFs) and also correlated with increased Mdm2 levels in mantle cell lymphoma patients’ biopsies. PIM1 associated with and phosphorylated Mdm2 at Ser166 and Ser186 leading to stabilization of both proteins.66 More work is needed to validate the impact of PIM-mediated p53 regulation for induction and/or maintenance of malignant transformation.
PIM serine/threonine kinases in hematologic malignancies and solid cancers
Hematologic malignancies
Early studies demonstrated overexpression of PIM1 in a significant fraction (30%) of human myeloid and lymphoid leukemia (including 51 primary patients’ samples and 19 leukemia cell lines) in absence of any apparent gene rearrangements or amplifications.67 In cellular models of malignant myeloproliferative disorders, PIM1 and PIM2 were both found to be up-regulated and proposed to be a mediator of anti-apoptotic properties of oncogenic protein tyrosine kinases (PTKs) such as BCR/ABL, FLT3-ITD, or the JAK2V617F mutant, most probably mediated through aberrant JAK2/STAT5 activity.68–73 We and others have observed that overexpression of PIM1 was sufficient to induce IL-3 independence in murine hematopoietic Ba/F3 cells.74,75 Microarray experiments revealed upregulation of PIM1 expression in acute myeloid leukemia harboring alterations of the mixed-lineage leukemia (MLL) gene such as the MLL/ENL or MLL/AF9 fusion genes76 (J Schwaller, unpublished observation, 2009). Elevated PIM1 levels in acute myeloid leukemia are most likely the consequence of FLT3 activation (by overexpression or mutation) and/or of aberrant activation of HOXA9, a direct transcriptional regulator of PIM1 69,70,72,77 (Figure 3).
To address the role of PIM kinases for induction of PTK-mediated leukemic disorders, we have performed bone marrow reconstitution experiments using PIM knockout cells. Transplantation of wild-type or PIM2 bone marrow retrovirally expressing the FLT3-ITD mutant led to induction of typical lympho-myeloproliferative disease.78 In contrast, PIM1 bone marrow cells were not able to reconstitute lethally irradiated recipients and showed a significant defect for homing to the bone marrow and spleen. Grafting of hematopoietic stem cells is a complex process regulated by several signaling pathways of which the CXCL12/CXCR4 ligand/receptor system plays a predominant role.79,80 Interestingly, PIM1, but not PIM2 bone marrow cells expressed significantly lower amounts of surface CXCR4 and were impaired in migration towards a CXCL12 gradient. Blocking PIM1 activity by expression of a dominant-negative acting mutant, siRNAs or by a small molecule inhibitor resulted in impaired CXCR4 surface re-expression after ligand-induced receptor internalization. Site directed mutagenesis experiments and in vitro kinase assays suggested that PIM1 might regulate CXCR4 by direct phosphorylation of the S339 residue in the intracellular domain, known for proper receptor internalization and surface re-expression. Surface expression of CXCR4 is a known prognostic factor in acute myeloid leukemia.81 It is worth noting that a correlation between PIM1 overex-pression and surface CXCR4 expression was found in fresh blasts from acute myeloid leukemia patients. Treatment of the cells with a small molecule PIM inhibitor resulted in ex vivo downregulation of CXCR4 surface expression in 4 out of 6 patients tested. These observations suggested that PIM1 (but not PIM2) regulate homing and migration of leukemic cells through modification of surface CXCR4 expression.82
Several B-cell lymphoproliferative disorders have been associated with latent infections of Epstein-Barr virus or Kaposi sarcoma-associated herpesvirus (KSHV). Interestingly, Epstein-Barr virus infection of primary B lymphocytes has been associated with an increase of PIM mRNA expression, and over-expressed PIM kinases enhanced the activity of the viral transactivator EBNA2.83 Significantly elevated PIM expression levels were also found in malignant B cells that express the KSHV latency associated nuclear antigen (LANA). LANA has been shown to be a substrate of PIM1 that phosphorylates LANA within the N-terminal domain.84 In addition, a kinome wide expression library study identified activation of PIM1/PIM3 as a critical element for reactivation of a latent KSHV infection.85
B-cell non-Hodgkin’s lymphoma (NHLs) is characterized by chromosomal translocations leading to deregulation of several proto-oncogenes controlled by the immunoglobulin gene promoter and enhancer elements. Similar to the immunoglobulin variable region genes in normal B-cell development, aberrant somatic hypermutation of multiple loci, including the proto-oncogenes C-MYC, RhoH, PAX5 and PIM1, have been found in over 50% of diffuse large-cell lymphomas (DLBCL).86 Typically, these mutations are localized in the 5’ untranslated or coding region of the genes, are independent of chromosomal translocations and share features of typical variable-region associated somatic hypermutations. The lack of such mutations in normal germinal-center B cells suggests a direct role for the pathogenesis of malignant lymphomas; however, the molecular mechanisms are currently not understood. Strikingly, several somatic hypermutations affecting PIM1 have been found in cases of other subtypes of B-cell non-Hodgkin’s lymphoma including follicular cell lymphoma, AIDS-NHLs, and MALT lymphomas.87 Rather surprisingly, several PIM1 variants showed a significantly decreased in vitro kinase activity, suggesting a so far unknown kinase-independent oncogenic function of PIM1.88 Recent observations made in a cancer xenograft model, in which overexpression of a kinase-dead PIM1 (K67M) mutant resulted in the formation of larger tumors, supports the hypothesis of an oncogenic function of PIMs independent of catalytic activity.89 Gene expression profiling identified frequent upregulation of PIM1 expression in aggressive mantle B-cell lymphoma. As PIM1 expression levels seem to be a poor prognostic marker in intensively treated aggressive mantle cell lymphoma, further studies for its role as therapeutic target for this aggressive disease are warranted.90
Similar to PIM1, significant levels of PIM2 have been found in primary blasts from acute myeloid leukemia patients.69,91 Interestingly, recent work identified PIM2 (but not PIM1) as the main kinase that phosphorylates 4E-BP1 resulting in mTOR-independent translational control in acute myeloid leukemia cells. This study suggests that a potent PIM2 inhibitor might be able to block rapamycin-resistant translation of oncogenic proteins.91 PIM2 is also highly expressed in progenitor cells of the B-cell lineage and critically involved in signaling pathways regulating B-cell homeostasis.92 Moreover, PIM2 has been reported being over-expressed and associated with progression of several malignancies that originate from the B-cell lineage such as chronic lymphocytic leukemia (CLL), diffuse large B-cell lymphoma (DLBCL), mantle cell lymphoma (MCL) or myeloma.93,94 The ability of PIM2 to promote survival of lymphoid cells seems to be dependent on activation of nuclear factor (NF)-κB through the serine/threonine kinase Cot/Tpl2.95 As PIM2 is possibly a downstream target of NF-κB signaling, high levels of PIM2 might be the result of a feedback mechanism.96
Solid tumors
Biomarker delineation for prostate cancer by using gene expression profiling identified the PIM1 serine/threonine kinase being deregulated upon cancer progression. Further validation in over 700 clinical patients’ samples showed no or weak PIM1 expression in benign lesions, and moderate to strong PIM1 expression in over 50% of prostate cancer samples. PIM1 expression correlated significantly with a poor therapy outcome in prostate cancer.97 This study also revealed remarkably similar transcriptional co-regulation of PIM1 and c-myc, possibly mediating synergistic oncogenic effects. Subsequently, this hypothesis has been experimentally validated in vivo by transgenic mice that express human c-myc in the mouse prostate. Cross-species gene expression comparison revealed that “MYC-like” human cancers are characterized by significant upregulation of PIM1.98 Further studies found increased PIM1 expression in high-grade prostatic neoplasia. This finding suggests that PIM1 overexpression is an early event in prostate carcinogenesis.99 In vitro studies demonstrated enhanced tumor growth and protection from drug-induced apoptosis of prostate cancer cells upon overexpression of PIM kinases.100
Retroviral overexpression of PIM1 in immortalized, non-tumorigenic prostate or mammary epithelial cell lines or the LNCaP prostate carcinoma cell line has been shown to induce genomic instability characterized by a defect in the mitotic spindle checkpoint, abnormal mitotic spindles, centrosome amplification and chromosome missegregation resulting in poly- and aneuploidy.101,102 PIM1 induced chromosomal instability is not restricted to prostate cells but has also been observed in telomerase-immortalized human mammary epithelial cells and associated with dys-regulation of cyclin B1.102 However, further validation of these in vitro observations by a single team is needed. It would be interesting to see if expression levels of PIM1 correlate in vivo with the degree of genomic instability observed in human malignancies.
Less is known about the role of PIM1 in other solid cancers. While studying PIM1 expression during mammary development, Gapter and colleagues found elevated levels of PIM1 in most mammary carcinoma cell lines. In addition, progesterone increased PIM1 protein levels to some extent in non-tumorigenic mammary epithelia.103 Elevated PIM1 levels in prostate and breast cancer could be also the consequence of aberrant STAT5 activity that has been associated with disease progression in both tumor types.104 Peltola and colleagues found that elevated PIM1 expression might be predictive for radiation response in squamous cell carcinoma of the head and neck.105 In addition, increased PIM1 expression was proposed to be a prognostic marker for pancreatic ductal adenocarcinoma.106 Tumor-associated hypoxia seems to increase PIM1 expression and to support chemoresistance shown in several solid cancer cell lines.107 These observations suggest that targeting of PIM1 might be beneficial in combination with chemotherapeutics for the therapy of solid cancers.
Perineural invasion, a major mechanism that leads to the spread of prostate cancer cells, has been found to be associated with elevated PIM2 expression.108 Increased PIM2 levels in prostate cancer correlated with higher proliferation, a decreased rate of apoptosis and many established prognostic factors.109 In vitro studies using HepG2 cells suggested that PIM2 may act as pro-survival kinase in liver cancer.110
A search for target genes of common fusion proteins associated with human Ewing’s sarcoma revealed upregulation of PIM3. Overexpression of PIM3 in rodent fibroblasts (NIH3T3) showed a stronger transforming activity than the EWS-FLI fusion. Furthermore, co-expression of EWS-FLI with a dominant-negative acting PIM3 mutant prolonged survival of mice after subcutaneously injecting transduced NIH-3T3 cells suggesting that PIM3 might be important for transformation by EWS-fusion genes.111 PIM3 is also highly expressed in human hepato-cellular carcinoma but not in normal hepatocytes.112 Likewise, high PIM3 expression has also been observed in malignant lesions of the pancreas but not in normal pancreatic tissue.113 Furthermore, the same group found elevated PIM3 expression upon progression of gastric adeno-carcinoma.114 Interestingly, whereas growth and survival of certain solid tumor cells (e.g. prostate or liver cancer cell lines) has been severely affected by inactivation of a single PIM isoform, in leukemia cell lines knockdown of more than one PIM is required. The large number of expression and functional studies have suggested that inhibition of aberrant PIM activity by small molecules may open a new promising avenue for cancer therapy.
PIM kinases as targets for cancer therapy
Insights from the structure
The crystal structure of PIM1 has been reported by a number of laboratories.22,88,115,116 These structural studies revealed a classic bilobal protein kinase domain architecture and apart from the unique beta hairpin insert located N-terminal to helix αC, all conserved secondary structure elements of typical protein kinases were present (Figure 4). In protein kinases, the binding site for ATP is located in a deep cavity formed by the two kinase lobes and the connecting hinge region. However, ATP binding to this site differs in PIM kinases. The presence of a proline residue at position 123 does not allow formation of a second hydrogen bond of the hinge backbone to the adenine ring. In addition, the insertion of an additional residue in the kinase hinge results in structural changes in the hinge region due to the unique hinge sequence ERPXPX. The structure of PIM2 has been reported recently.117 However, so far no crystal structure has been reported for PIM3. Both PIM1 and PIM2 assume an active conformation in the reported crystal structures. The active state of kinases is characterized by the presence of the conserved lysine –glutamate salt bridge (Lys67 and Glu89 in PIM1), a closed lobe conformation and a well-structured activation segment. The activation segment is often unstructured in kinases that require phosphorylation for catalytic activity. Upon phosphorylation, the activation segment folds onto the lower lobe and structures the peptide-binding site leading to enzymatic activation. In contrast, PIM kinases are catalytically active in the absence of phosphorylation and the crystal structures of PIM provide an explanation for this constitutive activity. The unphosphorylated activation segment forms a large number of polar interactions with the lower kinase lobe stabilizing the observed active conformation. Nevertheless, PIM kinases do autophosphorylate, but the functional consequences of these post-translational modifications are currently unknown.22
Small molecule PIM kinase inhibitors
The unique hinge architecture of PIM kinases suggests that very selective inhibitors can be identified using this structural difference. Indeed, over 50 potential PIM inhibitors have become public but the kinome wide specificity of these inhibitors is largely unknown.
These inhibitors can be grouped into two main classes based on their binding mode (Figure 5). The first class of inhibitors represents typical ATP-mimetic compounds that form, similar to the adenine ring of the cofactor, a hydrogen bond with the hinge backbone oxygen of PIM1 (Glu121). These compounds comprise the broad-spectrum kinase inhibitor staurosporine and its analog K252, bisin-doyl maleinimides (BIM) and the related PKC inhibitor LY333531 (Figure 5, compound 1) as well as a number of extremely potent organometallic inhibitors with sub-nanomolar inhibitor potencies in vitro (compound 2).118–121 Also flavonoids form multiple polar interactions with the hinge backbone and are potent inhibitors of PIM kinases (compound 3).118,122
The second class of PIM inhibitors does not interact with the hinge region by forming classical hydrogen bonds and can therefore be considered as ATP competitive but not ATP mimetic inhibitors. This non-canonical binding mode has been first identified for pyrazolo[1,5-a]pyrimidines and the related imidazo[1,2-b]pyridazine (compound 4) and LY294002 (data not shown), initially described as a potent phosphatidylinositol 3-kinase inhibitor.116,118 One important aspect of this binding mode is the formation of polar interactions with the active site lysine (K67) and a conserved water molecule present in most co-crystal structures. These polar interactions anchor the inhibitor to the back of the ATP binding pocket. Typically inhibitor binding is additionally stabilized by a number of hydrophobic interactions. Imidazo[1,2-b]pyri-dazines have low nanomolar potency in vitro and display dose-dependently impaired survival of murine Ba/F3 cells that were made cytokine independent by overexpression of human PIMs.75 Optimization of a promising imida-zo[1,2-b]pyridazine lead compound (“sc-47”) resulted in the generation of a highly potent compound (SGI-1776) with in vitro activity against PIM1, PIM2 and PIM3 at nanomolar concentrations. SGI-1776 impaired the growth of human leukemic cell lines (KG1, MV4;11 and MOLM13) at a sub-micromolar concentration in vitro and had biological activity in MV4;11 xenografts in vivo.123 The most recent studies demonstrated that SGI-1776 induced apoptosis in chronic lymphocytic leukemia (CLL) cells as well as in prostate cancer cell lines.124,125 Encouraging experimental results initiated clinical trials to explore the safety of SGI-1776 for the treatment of refractory non-Hodgkin’s lymphoma and prostate cancer patients (ClinicalTrials.gov Identifier: NCT00848601).
The promising results generated on imidazo[1,2-b]pyri-dazines led to the development of a number of other bicyclic scaffolds with nitrogen atoms present at different positions in the aromatic ring system such as N-substituted-3-aryl-[1,2,4]triazolo[4,3-b]pyridazin-6-amine inhibitors and triazolo-pyridazines (compounds 5 and 6).126,127 Interestingly, Pierce et al. used a filter during the docking of compounds that searched for unusual aromatic CH – O bonds to the kinase hinge region instead of the typical ATP mimetic classical hydrogen bonds and identified a number of diverse inhibitors with Ki values between 0.091 and 4.5 μM. Docking studies also revealed indolyl-pyrrolones and pyridones (compound 7) as potent PIM inhibitors.128,129 Another interesting scaffold that has been reported to inhibit PIM kinases and for which activity in cell based assays could be demonstrated are benzyli-dene-thiazolidine-2,4-diones (compound 8).130
In addition, a number of inhibitors have been identified that interact via halogen atoms with the PIM hinge region. These inhibitors include a series of pyrazine compounds (compound 9) as well as a series of isoxazolo[3,4-b]quino-line-3,4(1H,9H)-diones (compound 6).131,132 Similar halogen hinge interactions have been reported for 4,5,6,7-tetrabro-mo-1H-benzimidazole, which also strongly inhibit CK2 and di-chloro-substituted β-carbolines (S Knapp, unpublished data, 2009 and pdb code: 3CXW).133, 134
An interesting aspect of many PIM inhibitors is the often observed selectivity of many inhibitors for PIM1 and PIM3 over PIM2, the latter usually being inhibited with much lower potency. Structural models are unable to explain this phenomenon that is likely to be related to differences in dynamic properties of the different PIM isoforms.118
Open questions and perspectives
Over 20 years have passed since the identification of PIM serine/threonine kinases as cooperating oncogenes. Several recent studies demonstrated that functional interference with PIM kinases impaired growth and survival of cancer cells. These findings initiated studies that aim to delineate the role of PIM kinases in cancer and its role as potential therapeutic targets. However, the functional redundancy of the three PIM kinases in cancer biology remains unclear. Do all three PIM kinases have proto-oncogenic activity in any cell type? Are PIMs important for maintenance of cancer-initiating cells (also referred to as cancer stem cells) escaping current chemotherapeutic strategies? A hallmark of normal and cancer-initiating cells is the capability of adapting to hypoxia.135 Expression of PIM kinases is up-regulated by hypoxia and mediates hypoxic-induced drug resistance in cancer cells.107 As PIM1 regulates the CXCL12/CXCR4 axis, inhibiting PIMs would not only affect survival but could also mobilize leukemic stem cells from their sanctuary niche.82 Nevertheless, it is not known whether long-term inhibition of one PIM isoform might lead to a compensatory upregulation of other PIM family members. A potential compensatory mechanism suggests that inhibitors of PIM kinases should block all PIM kinases for efficient cancer therapy. Despite the high homology in the ATP-binding pocket, several proposed small molecule PIM inhibitors preferentially inhibit the activity of PIM1 and PIM 3 rather than PIM2. A potential limiting factor for chemical targeting of PIM kinases is the presence of mutations currently known to be present in B-cell lymphomas and Diamond-Blackfan anemia. Interestingly, a point mutation in PIM1 has recently been identified in a case of acute myeloid leukemia.136 Massive parallel sequencing will help to identify genetic alterations of the kinome in all cancers. A recent study proposed that PIM kinases might be therapeutically targeted by monoclonal antibodies that recognized PIM1 located in the cytosol and nucleus as well as on the surface of human and murine cancer cells.137 Interestingly, treatment of prostate cancer cell lines with this PIM1-specific antibody resulted in disruption of PIM1/Hsp90 complexes, reduced BAD phosphorylation and activation of the mitochondrial apoptotic pathway. Although the exact molecular mechanisms of these effects of PIM antibodies remain to be elucidated, the substantial in vivo growth inhibition of prostate cancer xenografts and leukemia cells by application of the anti-PIM1 antibody not only confirmed PIM1 as therapeutic target but also suggested new avenues for therapeutic intervention.137,138
Heat shock protein (Hsp) mediates protection of PIMs from proteasomal degradation. This suggests a potential therapeutic interest of targeting aberrantly expressed PIM kinases by small molecule Hsp inhibitors such as 17-AAG/Geldanamycin-derivates.9,10 However, whether the reported anti-leukemic activity of such compounds is also mediated by destabilization and degradation of PIM kinases has not been experimentally demonstrated.139
Functional characterization of signaling pathways that result up- and downstream of PIM kinases need to be studied in more detail in order to delineate combined therapeutic approaches. A complicating factor is that kinases important for the evasion of apoptosis such as PIMs, AKT/PKB, or SGK1, regulate cell death through modification of common substrates that affect the apoptotic machinery (e.g. BAD, NF-κB, FKHRL) and cellular metabolism (e.g. 4E-BP1, TSC2-mTOR, GSK3β).140 PIM kinases have been shown to control mTOR inhibition (rapamycin) resistant proliferation and survival. Therefore, a combination of small molecule inhibitors targeting several survival kinases might be essential for a powerful cancer therapeutic platform.91,140 Oncogenic protein synthesis through eIF4E binding protein 1 (4E-BP1) seems to be mTOR-independent and controlled by PIM2. These observations suggest that the oncogenic eIF4F translation-initiating complex could be blocked with small molecule PIM2 inhibitors.91
Transformation by leukemogenic oncogenes such as constitutive active tyrosine kinases is often mediated by parallel activation of several signaling pathways like JAK-STAT and PI3K-AKT/PKB. Interestingly, both PIM and AKT/PKB kinases show a selectivity for arginine at the -5 and -3 positions within substrates.22 AKT/PKB and PIM kinases seem to play partly redundant roles in mediating growth and survival of hematopoietic cells most probably due to overlapping substrates like BAD, p21, p27, or Cot/Tpl-2.140,141 Using an unbiased phospho-proteomic approach, Choudhary and colleagues have recently not only identified 21 new potential PIM substrates, but also 8 novel proteins targeted by PIM and AKT/PKB.142 Whether PIM1 acts up- or downstream of AKT/PKB might be context and/or cell type specific.34,143 Nevertheless, the exact molecular relationship between PIMs and AKT/PKB remains to be elucidated.
Although transcriptional activation of PIM kinases is quite well understood, we know very little about upstream regulators that functionally modify PIM through posttranslational modifications. Identification and functional characterization of such pathways might also help understand how different isoforms of a PIM kinase (e.g. the long (p44) versus the short (p33) PIM1 isoform) meet their substrates that are located on the cell membrane, in the cytosol as well as in the nucleus. Interestingly, hypox-ia seems to induce nuclear localization of PIM1.144 Despite the many questions and obstacles which remain, we hope that collaborative research efforts will lead to the successful introduction of small molecule PIM inhibitors to the arsenal of targeted pharmacological weapons against human cancer.
Footnotes
- Funding: JS holds a research professorship from the Gertrude Von Meissner Foundation and is further supported by the Swiss Cancer League, the Swiss National Research Foundation, the Susy-Rueckert Foundation/Krebsliga beider Basel, Wilhelm-Sander Foundation (Munich) and the Novartis Foundation for Biomedical Research (Basel). SK is an investigator of the Structural Genomics Consortium (SGC), a registered charity (no. 1097737) funded by the Wellcome Trust, GlaxoSmithKline, Genome Canada, the Canadian Institutes of Health Research, the Ontario Innovation Trust, the Ontario Research and Development Challenge Fund, the Canadian Foundation for Innovation, VINNOVA, The Knut and Alice Wallenberg Foundation, The Swedish Foundation for Strategic Research, and Karolinska Institutet.
- The online version of this article has a Supplementary Appendix.
- Authorship and Disclosures JS initiated this work and wrote the report. SK provided the section on structural features of PIM kinase inhibitors. All authors read, gave comments, and approved the final version of the manuscript.
- The authors reported no potential conflicts of interest.
- Received October 28, 2009.
- Revision received December 1, 2009.
- Accepted December 21, 2009.
References
- Cuypers HT, Selten G, Quint W, Zijlstra M, Maandag ER, Boelens W. Murine leukemia virus-induced T-cell lymphoma-genesis: integration of proviruses in a distinct chromosomal region. Cell. 1984; 37(1):141-50. PubMedhttps://doi.org/10.1016/0092-8674(84)90309-XGoogle Scholar
- Selten G, Cuypers HT, Boelens W, Robanus-Maandag E, Verbeek J, Domen J. The primary structure of the putative oncogene pim-1 shows extensive homology with protein kinases. Cell. 1986; 46(4):603-11. PubMedhttps://doi.org/10.1016/0092-8674(86)90886-XGoogle Scholar
- van Lohuizen M, Verbeek S, Krimpenfort P, Domen J, Saris C, Radaszkiewicz T. Predisposition to lymphomagenesis in pim-1 transgenic mice: cooperation with c-myc and N-myc in murine leukemia virus-induced tumors. Cell. 1989; 56(4):673-82. PubMedhttps://doi.org/10.1016/0092-8674(89)90589-8Google Scholar
- Berns A, Breuer M, Verbeek S, van Lohuizen M. Transgenic mice as a means to study synergism between oncogenes. Int J Cancer Suppl. 1989; 4:22-5. PubMedGoogle Scholar
- Saris CJ, Domen J, Berns A. The pim-1 oncogene encodes two related protein-serine/threonine kinases by alternative initiation at AUG and CUG. Embo J. 1991; 10(3):655-64. PubMedGoogle Scholar
- Liang H, Hittelman W, Nagarajan L. Ubiquitous expression and cell cycle regulation of the protein kinase PIM-1. Arch Biochem Biophys. 1996; 330(2):259-65. PubMedhttps://doi.org/10.1006/abbi.1996.0251Google Scholar
- Wang Z, Bhattacharya N, Weaver M, Petersen K, Meyer M, Gapter L. Pim-1: a serine/threonine kinase with a role in cell survival, proliferation, differentiation and tumorigenesis. J Vet Sci. 2001; 2(3):167-79. PubMedGoogle Scholar
- Shay KP, Wang Z, Xing PX, McKenzie IF, Magnuson NS. Pim-1 kinase stability is regulated by heat shock proteins and the ubiq-uitin-proteasome pathway. Mol Cancer Res. 2005; 3(3):170-81. PubMedhttps://doi.org/10.1158/1541-7786.MCR-04-0192Google Scholar
- Mizuno K, Shirogane T, Shinohara A, Iwamatsu A, Hibi M, Hirano T. Regulation of Pim-1 by Hsp90. Biochem Biophys Res Commun. 2001; 281(3):663-9. PubMedhttps://doi.org/10.1006/bbrc.2001.4405Google Scholar
- Nasser MW, Datta J, Nuovo G, Kutay H, Motiwala T, Majumder S. Down-regulation of micro-RNA-1 (miR-1) in lung cancer. Suppression of tumorigenic property of lung cancer cells and their sensitization to doxorubicin-induced apoptosis by miR-1. J Biol Chem. 2008; 283(48):33394-405. PubMedhttps://doi.org/10.1074/jbc.M804788200Google Scholar
- Huang X, Ding L, Bennewith KL, Tong RT, Welford SM, Ang KK. Hypoxia-inducible mir-210 regulates normoxic gene expression involved in tumor initiation. Mol Cell. 2009; 35(6):856-67. PubMedhttps://doi.org/10.1016/j.molcel.2009.09.006Google Scholar
- Laird PW, van der Lugt NM, Clarke A, Domen J, Linders K, McWhir J. In vivo analysis of Pim-1 deficiency. Nucleic Acids Res. 1993; 21(20):4750-5. PubMedhttps://doi.org/10.1093/nar/21.20.4750Google Scholar
- Domen J, van der Lugt NM, Laird PW, Saris CJ, Clarke AR, Hooper ML. Impaired interleukin-3 response in Pim-1-deficient bone marrow-derived mast cells. Blood. 1993; 82(5):1445-52. PubMedGoogle Scholar
- Breuer ML, Cuypers HT, Berns A. Evidence for the involvement of pim-2, a new common proviral insertion site, in progression of lymphomas. Embo J. 1989; 8(3):743-8. PubMedGoogle Scholar
- Allen JD, Verhoeven E, Domen J, van der Valk M, Berns A. Pim-2 transgene induces lymphoid tumors, exhibiting potent synergy with c-myc. Oncogene. 1997; 15(10):1133-41. PubMedhttps://doi.org/10.1038/sj.onc.1201288Google Scholar
- Mikkers H, Allen J, Knipscheer P, Romeijn L, Hart A, Vink E. High-throughput retroviral tagging to identify components of specific signaling pathways in cancer. Nat Genet. 2002; 32(1):153-9. PubMedhttps://doi.org/10.1038/ng950Google Scholar
- Feldman JD, Vician L, Crispino M, Tocco G, Marcheselli VL, Bazan NG. KID-1, a protein kinase induced by depolarization in brain. J Biol Chem. 1998; 273(26):16535-43. PubMedhttps://doi.org/10.1074/jbc.273.26.16535Google Scholar
- Mikkers H, Nawijn M, Allen J, Brouwers C, Verhoeven E, Jonkers J. Mice deficient for all PIM kinases display reduced body size and impaired responses to haematopoietic growth factors. Mol Cell Biol. 2004; 24(13):6104-15. PubMedhttps://doi.org/10.1128/MCB.24.13.6104-6115.2004Google Scholar
- Hoover D, Friedmann M, Reeves R, Magnuson NS. Recombinant human pim-1 protein exhibits serine/threonine kinase activity. J Biol Chem. 1991; 266(21):14018-23. PubMedGoogle Scholar
- Friedmann M, Nissen MS, Hoover DS, Reeves R, Magnuson NS. Characterization of the proto-oncogene pim-1: kinase activity and substrate recognition sequence. Arch Biochem Biophys. 1992; 298(2):594-601. PubMedhttps://doi.org/10.1016/0003-9861(92)90454-5Google Scholar
- Palaty CK, Clark-Lewis I, Leung D, Pelech SL. Phosphorylation site substrate specificity determinants for the Pim-1 protoonco-gene-encoded protein kinase. Biochem Cell Biol. 1997; 75(2):153-62. PubMedhttps://doi.org/10.1139/bcb-75-2-153Google Scholar
- Bullock AN, Debreczeni J, Amos AL, Knapp S, Turk BE. Structure and substrate specificity of the Pim-1 kinase. J Biol Chem. 2005; 280(50):41675-82. PubMedhttps://doi.org/10.1074/jbc.M510711200Google Scholar
- Palaty CK, Kalmar G, Tai G, Oh S, Amankawa L, Affolter M. Identification of the autophosphorylation sites of the Xenopus laevis Pim-1 proto-oncogene-encoded protein kinase. J Biol Chem. 1997; 272(16):10514-21. PubMedhttps://doi.org/10.1074/jbc.272.16.10514Google Scholar
- Losman JA, Chen XP, Vuong BQ, Fay S, Rothman PB. Protein phosphatase 2A regulates the stability of Pim protein kinases. J Biol Chem. 2003; 278(7):4800-5. PubMedhttps://doi.org/10.1074/jbc.M208246200Google Scholar
- Ma J, Arnold HK, Lilly MB, Sears RC, Kraft AS. Negative regulation of Pim-1 protein kinase levels by the B56beta subunit of PP2A. Oncogene. 2007; 26(35):5145-53. PubMedhttps://doi.org/10.1038/sj.onc.1210323Google Scholar
- Kim O, Jiang T, Xie Y, Guo Z, Chen H, Qiu Y. Synergism of cytoplasmic kinases in IL6-induced ligand-independent activation of androgen receptor in prostate cancer cells. Oncogene. 2004; 23(10):1838-44. PubMedhttps://doi.org/10.1038/sj.onc.1207304Google Scholar
- Xie Y, Xu K, Dai B, Guo Z, Jiang T, Chen H. The 44 kDa Pim-1 kinase directly interacts with tyrosine kinase Etk/BMX and protects human prostate cancer cells from apoptosis induced by chemotherapeutic drugs. Oncogene. 2006; 25(1):70-8. PubMedGoogle Scholar
- Lilly M, Sandholm J, Cooper JJ, Koskinen PJ, Kraft A. The PIM-1 serine kinase prolongs survival and inhibits apoptosis-related mitochondrial dysfunction in part through a bcl-2-dependent pathway. Oncogene. 1999; 18(27):4022-31. PubMedhttps://doi.org/10.1038/sj.onc.1202741Google Scholar
- Didichenko SA, Spiegl N, Brunner T, Dahinden CA. IL-3 induces a Pim1-dependent antiapoptotic pathway in primary human basophils. Blood. 2008; 112(10):3949-58. PubMedhttps://doi.org/10.1182/blood-2008-04-149419Google Scholar
- Yan B, Zemskova M, Holder S, Chin V, Kraft A, Koskinen PJ. The PIM-2 kinase phosphorylates BAD on serine 112 and reverses BAD-induced cell death. J Biol Chem. 2003; 278(46):45358-67. PubMedhttps://doi.org/10.1074/jbc.M307933200Google Scholar
- Aho TL, Sandholm J, Peltola KJ, Mankonen HP, Lilly M, Koskinen PJ. Pim-1 kinase promotes inactivation of the pro-apoptotic Bad protein by phosphorylating it on the Ser112 gatekeeper site. FEBS Lett. 2004; 571(1–3):43-9. PubMedhttps://doi.org/10.1016/j.febslet.2004.06.050Google Scholar
- Fox CJ, Hammerman PS, Cinalli RM, Master SR, Chodosh LA, Thompson CB. The serine/threonine kinase Pim-2 is a tran-scriptionally regulated apoptotic inhibitor. Genes Dev. 2003; 17(15):1841-54. PubMedhttps://doi.org/10.1101/gad.1105003Google Scholar
- Macdonald A, Campbell DG, Toth R, McLauchlan H, Hastie CJ, Arthur JS. Pim kinases phosphorylate multiple sites on Bad and promote 14-3-3 binding and disso-ciation from Bcl-XL. BMC Cell Biol. 2006; 7:1. PubMedhttps://doi.org/10.1186/1471-2121-7-1Google Scholar
- Zhang F, Beharry ZM, Harris TE, Lilly MB, Smith CD, Mahajan S. PIM1 protein kinase regulates PRAS40 phosphorylation and mTOR activity in FDCP1 cells. Cancer Biol Ther. 2009; 8(9):846-53. PubMedhttps://doi.org/10.4161/cbt.8.9.8210Google Scholar
- Gu JJ, Wang Z, Reeves R, Magnuson NS. PIM1 phosphorylates and negatively regulates ASK1-mediated apoptosis. Oncogene. 2009; 28(48):4261-71. PubMedhttps://doi.org/10.1038/onc.2009.276Google Scholar
- Mochizuki T, Kitanaka C, Noguchi K, Muramatsu T, Asai A, Kuchino Y. Physical and functional interactions between Pim-1 kinase and Cdc25A phosphatase. Implications for the Pim-1-mediated activa-tion of the c-Myc signaling pathway. J Biol Chem. 1999; 274(26):18659-66. PubMedhttps://doi.org/10.1074/jbc.274.26.18659Google Scholar
- Shirogane T, Fukada T, Muller JM, Shima DT, Hibi M, Hirano T. Synergistic roles for Pim-1 and c-Myc in STAT3-mediated cell cycle progression and antiapoptosis. Immunity. 1999; 11(6):709-19. PubMedhttps://doi.org/10.1016/S1074-7613(00)80145-4Google Scholar
- Zippo A, De Robertis A, Serafini R, Oliviero S. PIM1-dependent phosphorylation of histone H3 at serine 10 is required for MYC-dependent transcriptional activation and oncogenic transformation. Nat Cell Biol. 2007; 9(8):932-44. PubMedhttps://doi.org/10.1038/ncb1618Google Scholar
- Zippo A, Serafini R, Rocchigiani M, Pennacchini S, Krepelova A, Oliviero S. Histone crosstalk between H3S10ph and H4K16ac generates a histone code that mediates transcription elongation. Cell. 2009; 138(6):1122-36. PubMedhttps://doi.org/10.1016/j.cell.2009.07.031Google Scholar
- Pircher TJ, Zhao S, Geiger JN, Joneja B, Wojchowski DM. Pim-1 kinase protects haematopoietic FDC cells from genotoxin-induced death. Oncogene. 2000; 19(32):3684-92. PubMedhttps://doi.org/10.1038/sj.onc.1203684Google Scholar
- Wang Z, Bhattacharya N, Mixter PF, Wei W, Sedivy J, Magnuson NS. Phosphorylation of the cell cycle inhibitor p21Cip1/WAF1 by Pim-1 kinase. Biochim Biophys Acta. 2002; 1593(1):45-55. PubMedGoogle Scholar
- Leverson JD, Koskinen PJ, Orrico FC, Rainio EM, Jalkanen KJ, Dash AB. Pim-1 kinase and p100 cooperate to enhance c-Myb activity. Mol Cell. 1998; 2(4):417-25. PubMedhttps://doi.org/10.1016/S1097-2765(00)80141-0Google Scholar
- Zhang Y, Wang Z, Magnuson NS. Pim-1 kinase-dependent phosphorylation of p21Cip1/WAF1 regulates its stability and cellular localization in H1299 cells. Mol Cancer Res. 2007; 5(9):909-22. PubMedhttps://doi.org/10.1158/1541-7786.MCR-06-0388Google Scholar
- Morishita D, Katayama R, Sekimizu K, Tsuruo T, Fujita N. Pim kinases promote cell cycle progression by phosphorylating and down-regulating p27Kip1 at the transcriptional and posttranscriptional levels. Cancer Res. 2008; 68(13):5076-85. PubMedhttps://doi.org/10.1158/0008-5472.CAN-08-0634Google Scholar
- Bachmann M, Hennemann H, Xing PX, Hoffmann I, Moroy T. The oncogenic serine/threonine kinase Pim-1 phosphorylates and inhibits the activity of Cdc25C-associated kinase 1 (C-TAK1): a novel role for Pim-1 at the G2/M cell cycle checkpoint. J Biol Chem. 2004; 279(46):48319-28. PubMedhttps://doi.org/10.1074/jbc.M404440200Google Scholar
- Bachmann M, Kosan C, Xing PX, Montenarh M, Hoffmann I, Moroy T. The oncogenic serine/threonine kinase Pim-1 directly phosphorylates and activates the G2/M specific phosphatase Cdc25C. Int J Biochem Cell Biol. 2006; 38(3):430-43. PubMedhttps://doi.org/10.1016/j.biocel.2005.10.010Google Scholar
- Maita H, Harada Y, Nagakubo D, Kitaura H, Ikeda M, Tamai K. PAP-1, a novel target protein of phosphorylation by pim-1 kinase. Eur J Biochem. 2000; 267(16):5168-78. PubMedGoogle Scholar
- Koike N, Maita H, Taira T, Ariga H, Iguchi-Ariga SM. Identification of heterochro-matin protein 1 (HP1) as a phosphorylation target by Pim-1 kinase and the effect of phosphorylation on the transcriptional repression function of HP1(1). FEBS Lett. 2000; 467(1):17-21. PubMedhttps://doi.org/10.1016/S0014-5793(00)01105-4Google Scholar
- Ishibashi Y, Maita H, Yano M, Koike N, Tamai K, Ariga H. Pim-1 translocates sorting nexin 6/TRAF4-associated factor 2 from cytoplasm to nucleus. FEBS Lett. 2001; 506(1):33-8. PubMedhttps://doi.org/10.1016/S0014-5793(01)02881-2Google Scholar
- Chen XP, Losman JA, Cowan S, Donahue E, Fay S, Vuong BQ. Pim serine/threo-nine kinases regulate the stability of Socs-1 protein. Proc Natl Acad Sci USA. 2002; 99(4):2175-80. PubMedhttps://doi.org/10.1073/pnas.042035699Google Scholar
- Peltola KJ, Paukku K, Aho TL, Ruuska M, Silvennoinen O, Koskinen PJ. Pim-1 kinase inhibits STAT5-dependent transcription via its interactions with SOCS1 and SOCS3. Blood. 2004; 103(10):3744-50. PubMedhttps://doi.org/10.1182/blood-2003-09-3126Google Scholar
- Chiocchetti A, Gibello L, Carando A, Aspesi A, Secco P, Garelli E. Interactions between RPS19, mutated in Diamond-Blackfan anemia, and the PIM-1 oncoprotein. Haematologica. 2005; 90(11):1453-62. PubMedGoogle Scholar
- Aho TL, Sandholm J, Peltola KJ, Ito Y, Koskinen PJ. Pim-1 kinase phosphorylates RUNX family transcription factors and enhances their activity. BMC Cell Biol. 2006; 7:21. PubMedhttps://doi.org/10.1186/1471-2121-7-21Google Scholar
- Xie Y, Xu K, Linn DE, Yang X, Guo Z, Shimelis H. The 44-kDa Pim-1 kinase phosphorylates BCRP/ABCG2 and thereby promotes its multimerization and drug-resistant activity in human prostate cancer cells. J Biol Chem. 2008; 283(6):3349-56. PubMedhttps://doi.org/10.1074/jbc.M707773200Google Scholar
- Peng C, Knebel A, Morrice NA, Li X, Barringer K, Li J. Pim kinase substrate identification and specificity. J Biochem. 2007; 141(3):353-62. PubMedhttps://doi.org/10.1093/jb/mvm040Google Scholar
- Winn LM, Lei W, Ness SA. Pim-1 phosphorylates the DNA binding domain of c-Myb. Cell Cycle. 2003; 2(3):258-62. PubMedGoogle Scholar
- Zhang Y, Wang Z, Li X, Magnuson NS. Pim kinase-dependent inhibition of c-Myc degradation. Oncogene. 2008; 27(35):4809-19. PubMedhttps://doi.org/10.1038/onc.2008.123Google Scholar
- Rainio EM, Sandholm J, Koskinen PJ. Cutting edge: Transcriptional activity of NFATc1 is enhanced by the Pim-1 kinase. J Immunol. 2002; 168(4):1524-7. PubMedhttps://doi.org/10.4049/jimmunol.168.4.1524Google Scholar
- Bhattacharya N, Wang Z, Davitt C, McKenzie IF, Xing PX, Magnuson NS. Pim-1 associates with protein complexes necessary for mitosis. Chromosoma. 2002; 111(2):80-95. PubMedhttps://doi.org/10.1007/s00412-002-0192-6Google Scholar
- Wang Z, Bhattacharya N, Meyer MK, Seimiya H, Tsuruo T, Tonani JA. Pim-1 negatively regulates the activity of PTP-U2S phosphatase and influences terminal differentiation and apoptosis of monoblas-toid leukemia cells. Arch Biochem Biophys. 2001; 390(1):9-18. PubMedhttps://doi.org/10.1006/abbi.2001.2370Google Scholar
- Nihira K, Ando Y, Yamaguchi T, Kagami Y, Miki Y, Yoshida K. Pim-1 controls NF-kappaB signalling by stabilizing RelA/p65. Cell Death Differ. 2010; 17(4):689-98. PubMedhttps://doi.org/10.1038/cdd.2009.174Google Scholar
- Chen JL, Limnander A, Rothman PB. Pim-1 and Pim-2 kinases are required for efficient pre-B-cell transformation by v-Abl onco-gene. Blood. 2008; 111(3):1677-85. PubMedhttps://doi.org/10.1182/blood-2007-04-083808Google Scholar
- Benekli M, Baumann H, Wetzler M. Targeting signal transducer and activator of transcription signaling pathway in leukemias. J Clin Oncol. 2009; 27(26):4422-32. PubMedhttps://doi.org/10.1200/JCO.2008.21.3264Google Scholar
- Flygare J, Karlsson S. Diamond-Blackfan anemia: erythropoiesis lost in translation. Blood. 2007; 109(8):3152-4. PubMedhttps://doi.org/10.1182/blood-2006-09-001222Google Scholar
- McGowan KA, Li JZ, Park CY, Beaudry V, Tabor HK, Sabnis AJ. Ribosomal mutations cause p53-mediated dark skin and pleiotropic effects. Nat Genet. 2008; 40(8):963-70. PubMedhttps://doi.org/10.1038/ng.188Google Scholar
- Hogan C, Hutchison C, Marcar L, Milne D, Saville M, Goodlad J. Elevated levels of oncogenic protein kinase Pim-1 induce the p53 pathway in cultured cells and correlate with increased Mdm2 in mantle cell lymphoma. J Biol Chem. 2008; 283(26):18012-23. PubMedhttps://doi.org/10.1074/jbc.M709695200Google Scholar
- Amson R, Sigaux F, Przedborski S, Flandrin G, Givol D, Telerman A. The human protooncogene product p33pim is expressed during fetal haematopoiesis and in diverse leukemias. Proc Natl Acad Sci USA. 1989; 86(22):8857-61. PubMedhttps://doi.org/10.1073/pnas.86.22.8857Google Scholar
- Nieborowska-Skorska M, Hoser G, Kossev P, Wasik MA, Skorski T. Complementary functions of the antiapoptotic protein A1 and serine/threonine kinase pim-1 in the BCR/ABL-mediated leukemogenesis. Blood. 2002; 99(12):4531-9. PubMedhttps://doi.org/10.1182/blood.V99.12.4531Google Scholar
- Mizuki M, Schwable J, Steur C, Choudhary C, Agrawal S, Sargin B. Suppression of myeloid transcription factors and induction of STAT response genes by AML-specific Flt3 mutations. Blood. 2003; 101(8):3164-73. https://doi.org/10.1182/blood-2002-06-1677Google Scholar
- Kim KT, Baird K, Ahn JY, Meltzer P, Lilly M, Levis M. Pim-1 is up-regulated by constitutively activated FLT3 and plays a role in FLT3-mediated cell survival. Blood. 2005; 105(4):1759-67. PubMedhttps://doi.org/10.1182/blood-2004-05-2006Google Scholar
- Adam M, Pogacic V, Bendit M, Chappuis R, Nawijn MC, Duyster J. Targeting PIM kinases impairs survival of haematopoietic cells transformed by kinase inhibitor-sensitive and kinase inhibitor-resistant forms of Fms-like tyrosine kinase 3 and BCR/ABL. Cancer Res. 2006; 66(7):3828-35. PubMedhttps://doi.org/10.1158/0008-5472.CAN-05-2309Google Scholar
- Agrawal S, Koschmieder S, Baumer N, Reddy NG, Berdel WE, Muller-Tidow C. Pim2 complements Flt3 wild-type receptor in haematopoietic progenitor cell transformation. Leukemia. 2008; 22(1):78-86. PubMedhttps://doi.org/10.1038/sj.leu.2404988Google Scholar
- Wernig G, Gonneville JR, Crowley BJ, Rodrigues MS, Reddy MM, Hudon HE. The Jak2V617F oncogene associated with myeloproliferative diseases requires a functional FERM domain for transformation and for expression of the Myc and Pim proto-oncogenes. Blood. 2008; 111(7):3751-9. PubMedhttps://doi.org/10.1182/blood-2007-07-102186Google Scholar
- Nosaka T, Kitamura T. Pim-1 expression is sufficient to induce cytokine independence in murine haematopoietic cells, but is dispensable for BCR-ABL-mediated transformation. Exp Haematol. 2002; 30(7):697-702. PubMedhttps://doi.org/10.1016/S0301-472X(02)00808-1Google Scholar
- Pogacic V, Bullock AN, Fedorov O, Filippakopoulos P, Gasser C, Biondi A. Structural analysis identifies imidazo[1,2-b]pyridazines as PIM kinase inhibitors with in vitro antileukemic activity. Cancer Res. 2007; 67(14):6916-24. PubMedhttps://doi.org/10.1158/0008-5472.CAN-07-0320Google Scholar
- Chen W, Kumar AR, Hudson WA, Li Q, Wu B, Staggs RA. Malignant transformation initiated by Mll-AF9: gene dosage and critical target cells. Cancer Cell. 2008; 13(5):432-40. PubMedhttps://doi.org/10.1016/j.ccr.2008.03.005Google Scholar
- Hu YL, Passegue E, Fong S, Largman C, Lawrence HJ. Evidence that the Pim1 kinase gene is a direct target of HOXA9. Blood. 2007; 109(11):4732-8. PubMedhttps://doi.org/10.1182/blood-2006-08-043356Google Scholar
- Grundler R, Thiede C, Miething C, Steudel C, Peschel C, Duyster J. Sensitivity toward tyrosine kinase inhibitors varies between different activating mutations of the FLT3 receptor. Blood. 2003; 102(2):646-51. PubMedhttps://doi.org/10.1182/blood-2002-11-3441Google Scholar
- Lapidot T, Dar A, Kollet O. How do stem cells find their way home?. Blood. 2005; 106(6):1901-10. PubMedhttps://doi.org/10.1182/blood-2005-04-1417Google Scholar
- Busillo JM, Benovic JL. Regulation of CXCR4 signaling. Biochim Biophys Acta. 2007; 1768(4):952-63. PubMedhttps://doi.org/10.1016/j.bbamem.2006.11.002Google Scholar
- Spoo AC, Lubbert M, Wierda WG, Burger JA. CXCR4 is a prognostic marker in acute myelogenous leukemia. Blood. 2004; 109(2):786-91. https://doi.org/10.1182/blood-2006-05-024844Google Scholar
- Grundler R, Brault L, Gasser C, Bullock AN, Dechow T, Woetzel S. Dissection of PIM serine/threonine kinases in FLT3-ITD-induced leukemogenesis reveals PIM1 as regulator of CXCL12-CXCR4-mediated homing and migration. J Exp Med. 2009; 206(9):1957-70. PubMedhttps://doi.org/10.1084/jem.20082074Google Scholar
- Rainio EM, Ahlfors H, Carter KL, Ruuska M, Matikainen S, Kieff E. Pim kinases are upregulated during Epstein-Barr virus infection and enhance EBNA2 activity. Virology. 2005; 333(2):201-6. PubMedhttps://doi.org/10.1016/j.virol.2005.01.001Google Scholar
- Bajaj BG, Verma SC, Lan K, Cotter MA, Woodman ZL, Robertson ES. KSHV encoded LANA upregulates Pim-1 and is a substrate for its kinase activity. Virology. 2006; 351(1):18-28. PubMedhttps://doi.org/10.1016/j.virol.2006.03.037Google Scholar
- Cheng F, Weidner-Glunde M, Varjosalo M, Rainio EM, Lehtonen A, Schulz TF. KSHV reactivation from latency requires Pim-1 and Pim-3 kinases to inactivate the latency-associated nuclear antigen LANA. PLoS Pathog. 2009; 5(3):e1000324. PubMedhttps://doi.org/10.1371/journal.ppat.1000324Google Scholar
- Pasqualucci L, Neumeister P, Goossens T, Nanjangud G, Chaganti RS, Kuppers R. Hypermutation of multiple proto-oncogenes in B-cell diffuse large-cell lymphomas. Nature. 2001; 412(6844):341-6. PubMedhttps://doi.org/10.1038/35085588Google Scholar
- Gaidano G, Pasqualucci L, Capello D, Berra E, Deambrogi C, Rossi D. Aberrant somatic hypermutation in multiple subtypes of AIDS-associated non-Hodgkin lymphoma. Blood. 2003; 102(5):1833-41. PubMedhttps://doi.org/10.1182/blood-2002-11-3606Google Scholar
- Kumar A, Mandiyan V, Suzuki Y, Zhang C, Rice J, Tsai J. Crystal structures of proto-oncogene kinase Pim1: a target of aberrant somatic hypermutations in diffuse large cell lymphoma. J Mol Biol. 2005; 348(1):183-93. PubMedhttps://doi.org/10.1016/j.jmb.2005.02.039Google Scholar
- Kim J, Roh M, Abdulkadir S.Paper presented at: ; 2008. Google Scholar
- Hsi ED, Jung SH, Lai R, Johnson JL, Cook JR, Jones D. Ki67 and PIM1 expression predict outcome in mantle cell lymphoma treated with high dose therapy, stem cell transplantation and rituximab: a Cancer and Leukemia Group B 59909 correlative science study. Leuk Lymphoma. 2008; 49(11):2081-90. PubMedhttps://doi.org/10.1080/10428190802419640Google Scholar
- Tamburini J, Green AS, Bardet V, Chapuis N, Park S, Willems L. Protein synthesis is resistant to rapamycin and constitutes a promising therapeutic target in acute myeloid leukemia. Blood. 2009; 114(8):1618-27. PubMedhttps://doi.org/10.1182/blood-2008-10-184515Google Scholar
- Woodland RT, Fox CJ, Schmidt MR, Hammerman PS, Opferman JT, Korsmeyer SJ. Multiple signaling pathways promote B lymphocyte stimulator dependent B-cell growth and survival. Blood. 2008; 111(2):750-60. PubMedhttps://doi.org/10.1182/blood-2007-03-077222Google Scholar
- Cohen AM, Grinblat B, Bessler H, Kristt D, Kremer A, Schwartz A. Increased expression of the hPim-2 gene in human chronic lymphocytic leukemia and non-Hodgkin lymphoma. Leuk Lymphoma. 2004; 45(5):951-5. PubMedhttps://doi.org/10.1080/10428190310001641251Google Scholar
- Huttmann A, Klein-Hitpass L, Thomale J, Deenen R, Carpinteiro A, Nuckel H. Gene expression signatures separate B-cell chronic lymphocytic leukaemia prognostic subgroups defined by ZAP-70 and CD38 expression status. Leukemia. 2006; 20(10):1774-82. PubMedhttps://doi.org/10.1038/sj.leu.2404363Google Scholar
- Hammerman PS, Fox CJ, Birnbaum MJ, Thompson CB. Pim and Akt oncogenes are independent regulators of haematopoietic cell growth and survival. Blood. 2005; 105(11):4477-83. PubMedhttps://doi.org/10.1182/blood-2004-09-3706Google Scholar
- Li J, Peet GW, Balzarano D, Li X, Massa P, Barton RW. Novel NEMO/IkappaB kinase and NF-kappa B target genes at the pre-B to immature B cell transition. J Biol Chem. 2001; 276(21):18579-90. PubMedhttps://doi.org/10.1074/jbc.M100846200Google Scholar
- Dhanasekaran SM, Barrette TR, Ghosh D, Shah R, Varambally S, Kurachi K. Delineation of prognostic biomarkers in prostate cancer. Nature. 2001; 412(6849):822-6. PubMedhttps://doi.org/10.1038/35090585Google Scholar
- Ellwood-Yen K, Graeber TG, Wongvipat J, Iruela-Arispe ML, Zhang J, Matusik R. Myc-driven murine prostate cancer shares molecular features with human prostate tumors. Cancer Cell. 2003; 4(3):223-38. PubMedhttps://doi.org/10.1016/S1535-6108(03)00197-1Google Scholar
- Cibull TL, Jones TD, Li L, Eble JN, Ann Baldridge L, Malott SR. Overexpression of Pim-1 during progression of prostatic adenocarcinoma. J Clin Pathol. 2006; 59(3):285-8. PubMedhttps://doi.org/10.1136/jcp.2005.027672Google Scholar
- Chen WW, Chan DC, Donald C, Lilly MB, Kraft AS. Pim family kinases enhance tumor growth of prostate cancer cells. Mol Cancer Res. 2005; 3(8):443-51. PubMedhttps://doi.org/10.1158/1541-7786.MCR-05-0007Google Scholar
- Roh M, Gary B, Song C, Said-Al-Naief N, Tousson A, Kraft A. Overexpression of the oncogenic kinase Pim-1 leads to genomic instability. Cancer Res. 2003; 63(23):8079-84. PubMedGoogle Scholar
- Roh M, Song C, Kim J, Abdulkadir SA. Chromosomal instability induced by Pim-1 is passage-dependent and associated with dysregulation of cyclin B1. J Biol Chem. 2005; 280(49):40568-77. PubMedhttps://doi.org/10.1074/jbc.M509369200Google Scholar
- Gapter LA, Magnuson NS, Ng KY, Hosick HL. Pim-1 kinase expression during murine mammary development. Biochem Biophys Res Commun. 2006; 345(3):989-97. PubMedhttps://doi.org/10.1016/j.bbrc.2006.04.110Google Scholar
- Tan SH, Nevalainen MT. Signal transducer and activator of transcription 5A/B in prostate and breast cancers. Endocr Relat Cancer. 2008; 15(2):367-90. PubMedhttps://doi.org/10.1677/ERC-08-0013Google Scholar
- Peltola K, Hollmen M, Maula SM, Rainio E, Ristamaki R, Luukkaa M. Pim-1 kinase expression predicts radiation response in squamocellular carcinoma of head and neck and is under the control of epidermal growth factor receptor. Neoplasia. 2009; 11(7):629-36. PubMedGoogle Scholar
- Reiser-Erkan C, Erkan M, Pan Z, Bekasi S, Giese NA, Streit S. Hypoxia-inducible proto-oncogene Pim-1 is a prognostic marker in pancreatic ductal adenocarcinoma. Cancer Biol Ther. 2008; 7(9):1352-9. PubMedhttps://doi.org/10.4161/cbt.7.9.6418Google Scholar
- Chen J, Kobayashi M, Darmanin S, Qiao Y, Gully C, Zhao R. Pim-1 plays a pivotal role in hypoxia-induced chemoresistance. Oncogene. 2009; 28(28):2581-92. PubMedhttps://doi.org/10.1038/onc.2009.124Google Scholar
- Ayala GE, Dai H, Ittmann M, Li R, Powell M, Frolov A. Growth and survival mechanisms associated with perineural invasion in prostate cancer. Cancer Res. 2004; 64(17):6082-90. PubMedhttps://doi.org/10.1158/0008-5472.CAN-04-0838Google Scholar
- Dai H, Li R, Wheeler T, Diaz de Vivar A, Frolov A, Tahir S. Pim-2 upregulation: biological implications associated with dis-ease progression and perinueral invasion in prostate cancer. Prostate. 2005; 65(3):276-86. PubMedhttps://doi.org/10.1002/pros.20294Google Scholar
- Gong J, Wang J, Ren K, Liu C, Li B, Shi Y. Serine/threonine kinase Pim-2 promotes liver tumorigenesis induction through mediating survival and preventing apoptosis of liver cell. J Surg Res. 2009; 153(1):17-22. PubMedhttps://doi.org/10.1016/j.jss.2008.03.033Google Scholar
- Deneen B, Welford SM, Ho T, Hernandez F, Kurland I, Denny CT. PIM3 proto-oncogene kinase is a common transcriptional target of divergent EWS/ETS oncoproteins. Mol Cell Biol. 2003; 23(11):3897-908. PubMedhttps://doi.org/10.1128/MCB.23.11.3897-3908.2003Google Scholar
- Fujii C, Nakamoto Y, Lu P, Tsuneyama K, Popivanova BK, Kaneko S. Aberrant expression of serine/threonine kinase Pim-3 in hepatocellular carcinoma development and its role in the proliferation of human hepatoma cell lines. Int J Cancer. 2005; 114(2):209-18. PubMedhttps://doi.org/10.1002/ijc.20719Google Scholar
- Li YY, Popivanova BK, Nagai Y, Ishikura H, Fujii C, Mukaida N. Pim-3, a proto-oncogene with serine/threonine kinase activity, is aberrantly expressed in human pancreatic cancer and phosphorylates bad to block bad-mediated apoptosis in human pancreatic cancer cell lines. Cancer Res. 2006; 66(13):6741-7. PubMedhttps://doi.org/10.1158/0008-5472.CAN-05-4272Google Scholar
- Zheng HC, Tsuneyama K, Takahashi H, Miwa S, Sugiyama T, Popivanova BK. Aberrant Pim-3 expression is involved in gastric adenoma-adenocarcinoma sequence and cancer progression. J Cancer Res Clin Oncol. 2008; 134(4):481-8. PubMedhttps://doi.org/10.1007/s00432-007-0310-1Google Scholar
- Jacobs MD, Black J, Futer O, Swenson L, Hare B, Fleming M. Pim-1 ligand-bound structures reveal the mechanism of serine/threonine kinase inhibition by LY294002. J Biol Chem. 2005; 280(14):13728-34. PubMedhttps://doi.org/10.1074/jbc.M413155200Google Scholar
- Qian KC, Wang L, Hickey ER, Studts J, Barringer K, Peng C. Structural basis of constitutive activity and a unique nucleotide binding mode of human Pim-1 kinase. J Biol Chem. 2005; 280(7):6130-7. PubMedhttps://doi.org/10.1074/jbc.M409123200Google Scholar
- Bullock AN, Russo S, Amos A, Pagano N, Bregman H, Debreczeni JE. Crystal structure of the PIM2 kinase in complex with an organoruthenium inhibitor. PLoS One. 2009; 4(10):e7112. PubMedhttps://doi.org/10.1371/journal.pone.0007112Google Scholar
- Bullock AN, Debreczeni JE, Fedorov OY, Nelson A, Marsden BD, Knapp S. Structural basis of inhibitor specificity of the human protooncogene proviral insertion site in moloney murine leukemia virus (PIM-1) kinase. J Med Chem. 2005; 48(24):7604-14. PubMedhttps://doi.org/10.1021/jm0504858Google Scholar
- Fedorov O, Marsden B, Pogacic V, Rellos P, Muller S, Bullock AN. A systematic interaction map of validated kinase inhibitors with Ser/Thr kinases. Proc Natl Acad Sci USA. 2007; 104(51):20523-8. PubMedhttps://doi.org/10.1073/pnas.0708800104Google Scholar
- Debreczeni JE, Bullock AN, Atilla GE, Williams DS, Bregman H, Knapp S. Ruthenium half-sandwich complexes bound to protein kinase Pim-1. Angew Chem Int Ed Engl. 2006; 45(10):1580-5. PubMedhttps://doi.org/10.1002/anie.200503468Google Scholar
- Bregman H, Meggers E. Ruthenium halfsandwich complexes as protein kinase inhibitors: an N-succinimidyl ester for rapid derivatizations of the cyclopentadienyl moiety. Org Lett. 2006; 8(24):5465-8. PubMedhttps://doi.org/10.1021/ol0620646Google Scholar
- Holder S, Zemskova M, Zhang C, Tabrizizad M, Bremer R, Neidigh JW. Characterization of a potent and selective small-molecule inhibitor of the PIM1 kinase. Mol Cancer Ther. 2007; 6(1):163-72. PubMedhttps://doi.org/10.1158/1535-7163.MCT-06-0397Google Scholar
- Gourley E, Liu XH, Lamb J, Grand C, Vankayalapati H, Warner S, Bearss D.Paper presented at: ; 2008. Google Scholar
- Chen LS, Redkar S, Bearss D, Wierda WG, Gandhi V. Pim kinase inhibitor, SGI-1776, induces apoptosis in chronic lymphocytic leukemia cells. Blood. 2009; 114(19):4150-7. PubMedhttps://doi.org/10.1182/blood-2009-03-212852Google Scholar
- Mumenthaler SM, Ng PY, Hodge A, Bearss D, Berk G, Kanekal S. Pharmacologic inhibition of Pim kinases alters prostate cancer cell growth and resensitizes chemoresistant cells to taxanes. Mol Cancer Ther. 2009; 8(10):2882-93. PubMedhttps://doi.org/10.1158/1535-7163.MCT-09-0293Google Scholar
- Pierce AC, Jacobs M, Stuver-Moody C. Docking study yields four novel inhibitors of the protooncogene Pim-1 kinase. J Med Chem. 2008; 51(6):1972-5. PubMedhttps://doi.org/10.1021/jm701248tGoogle Scholar
- Grey R, Pierce AC, Bemis GW, Jacobs MD, Moody CS, Jajoo R. Structure-based design of 3-aryl-6-amino-triazolo[4,3-b]pyridazine inhibitors of Pim-1 kinase. Bioorg Med Chem Lett. 2009; 19(11):3019-22. PubMedhttps://doi.org/10.1016/j.bmcl.2009.04.061Google Scholar
- Cheney IW, Yan S, Appleby T, Walker H, Vo T, Yao N. Identification and structure-activity relationships of substituted pyridones as inhibitors of Pim-1 kinase. Bioorg Med Chem Lett. 2007; 17(6):1679-83. PubMedhttps://doi.org/10.1016/j.bmcl.2006.12.086Google Scholar
- Olla S, Manetti F, Crespan E, Maga G, Angelucci A, Schenone S. Indolyl-pyrrolone as a new scaffold for Pim1 inhibitors. Bioorg Med Chem Lett. 2009; 19(5):1512-6. PubMedhttps://doi.org/10.1016/j.bmcl.2009.01.005Google Scholar
- Beharry Z, Zemskova M, Mahajan S, Zhang F, Ma J, Xia Z. Novel benzyli-dene-thiazolidine-2,4-diones inhibit Pim protein kinase activity and induce cell cycle arrest in leukemia and prostate cancer cells. Mol Cancer Ther. 2009; 8(6):1473-83. PubMedhttps://doi.org/10.1158/1535-7163.MCT-08-1037Google Scholar
- Tong Y, Stewart KD, Thomas S, Przytulinska M, Johnson EF, Klinghofer V. Isoxazolo[3,4-b]quinoline-3,4(1H,9H)-diones as unique, potent and selective inhibitors for Pim-1 and Pim-2 kinases: chemistry, biological activities, and molecular modeling. Bioorg Med Chem Lett. 2008; 18(19):5206-8. PubMedhttps://doi.org/10.1016/j.bmcl.2008.08.079Google Scholar
- Qian K, Wang L, Cywin CL, Farmer BT, Hickey E, Homon C. Hit to lead account of the discovery of a new class of inhibitors of Pim kinases and crystallo-graphic studies revealing an unusual kinase binding mode. J Med Chem. 2009; 52(7):1814-27. PubMedhttps://doi.org/10.1021/jm801242yGoogle Scholar
- Cozza G, Mazzorana M, Papinutto E, Bain J, Elliott M, di Maira G. Quinalizarin as a potent, selective and cell-permeable inhibitor of protein kinase CK2. Biochem J. 2009; 421(3):387-95. PubMedhttps://doi.org/10.1042/BJ20090069Google Scholar
- Pagano MA, Bain J, Kazimierczuk Z, Sarno S, Ruzzene M, Di Maira G. The selectivity of inhibitors of protein kinase CK2: an update. Biochem J. 2008; 415(3):353-65. PubMedhttps://doi.org/10.1042/BJ20080309Google Scholar
- Olivotto M, Dello Sbarba P. Environmental restrictions within tumor ecosystems select for a convergent, hypoxia-resistant pheno-type of cancer stem cells. Cell Cycle. 2008; 7(2):176-87. PubMedhttps://doi.org/10.4161/cc.7.2.5315Google Scholar
- Akagi T, Shih LY, Ogawa S, Gerss J, Moore SR, Schreck R. Single nucleotide polymorphism genomic arrays analysis of t(8;21) acute myeloid leukemia cells. Haematologica. 2009; 94(9):1301-6. PubMedhttps://doi.org/10.3324/haematol.2009.005744Google Scholar
- Hu XF, Li J, Vandervalk S, Wang Z, Magnuson NS, Xing PX. PIM-1-specific mAb suppresses human and mouse tumor growth by decreasing PIM-1 levels, reducing Akt phosphorylation, and activating apoptosis. J Clin Invest. 2009; 119(2):362-75. PubMedGoogle Scholar
- Li J, Hu XF, Loveland BE, Xing PX. Pim-1 expression and monoclonal antibody targeting in human leukemia cell lines. Exp Hematol. 2009; 37(11):1284-94. PubMedhttps://doi.org/10.1016/j.exphem.2009.08.002Google Scholar
- Banerji U, Judson I, Workman P. The clinical applications of heat shock protein inhibitors in cancer - present and future. Curr Cancer Drug Targets. 2003; 3(5):385-90. PubMedhttps://doi.org/10.2174/1568009033481813Google Scholar
- Amaravadi R, Thompson CB. The survival kinases Akt and Pim as potential pharmacological targets. J Clin Invest. 2005; 115(10):2618-24. PubMedhttps://doi.org/10.1172/JCI26273Google Scholar
- Fox CJ, Hammerman PS, Thompson CB. The Pim kinases control rapamycin-resistant T cell survival and activation. J Exp Med. 2005; 201(2):259-66. PubMedhttps://doi.org/10.1084/jem.20042020Google Scholar
- Choudhary C, Olsen JV, Brandts C, Cox J, Reddy PN, Bohmer FD. Mislocalized activation of oncogenic RTKs switches downstream signaling outcomes. Mol Cell. 2009; 36(2):326-39. PubMedhttps://doi.org/10.1016/j.molcel.2009.09.019Google Scholar
- Muraski JA, Rota M, Misao Y, Fransioli J, Cottage C, Gude N. Pim-1 regulates cardiomyocyte survival downstream of Akt. Nat Med. 2007; 13(12):1467-75. PubMedhttps://doi.org/10.1038/nm1671Google Scholar
- Chen J, Kobayashi M, Darmanin S, Qiao Y, Gully C, Zhao R. Hypoxia-mediated up-regulation of Pim-1 contributes to solid tumor formation. Am J Pathol. 2009; 175(1):400-11. PubMedhttps://doi.org/10.2353/ajpath.2009.080972Google Scholar