CG&D
HOME HELP FEEDBACK SUBSCRIPTIONS ARCHIVE SEARCH TABLE OF CONTENTS
Cancer Research Clinical Cancer Research
Cancer Epidemiology Biomarkers & Prevention Molecular Cancer Therapeutics
Molecular Cancer Research Cell Growth & Differentiation

This Article
Right arrow Abstract Freely available
Right arrow Full Text (PDF)
Right arrow Alert me when this article is cited
Right arrow Alert me if a correction is posted
Services
Right arrow Similar articles in this journal
Right arrow Similar articles in PubMed
Right arrow Alert me to new issues of the journal
Right arrow Download to citation manager
Right arrow
Citing Articles
Right arrow Citing Articles via HighWire
Right arrow Citing Articles via Google Scholar
Google Scholar
Right arrow Articles by Rippmann, J. F.
Right arrow Articles by Schnapp, A.
Right arrow Search for Related Content
PubMed
Right arrow PubMed Citation
Right arrow Articles by Rippmann, J. F.
Right arrow Articles by Schnapp, A.
Cell Growth & Differentiation Vol. 11, 409-416, July 2000
© 2000 American Association for Cancer Research


Articles

Phosphorylation-dependent Proline Isomerization Catalyzed by Pin1 Is Essential for Tumor Cell Survival and Entry into Mitosis

Joerg F. Rippmann1, Silke Hobbie1, Christine Daiber, Bernd Guilliard, Margit Bauer, Joachim Birk, Herbert Nar, Pilar Garin-Chesa, Wolfgang J. Rettig and Andreas Schnapp2

Departments of Oncology Research [J. F. R., S. H., C. D., B. G., P. G-C., W. J. R., A. S.] and Medicinal Chemistry [M. B., J. B., H. N.], Boehringer Ingelheim Pharma KG, 88397 Biberach, Germany

Abstract

Pin1, a member of the parvulin family of peptidyl-prolyl cis-trans isomerases (PPIases) has been implicated in the G2-M transition of the mammalian cell cycle. Pin1 interacts with a series of mitotic phosphoproteins, including Polo-like kinase-1, Cdc25C, and Cdc27, and is thought to act as a phosphorylation-dependent PPIase for these target molecules. Pin1 recognizes phosphorylated serine-proline or threonine-proline peptide-bonds in test substrates up to 1300-fold better than in the respective unphosphorylated peptides. To test directly whether Pin1 regulates the G2-M transition and/or progression through mitosis by catalyzing phosphorylation-dependent prolyl isomerization of essential mitotic targets, we examined the consequences of Pin1 depletion, achieved by (a) overexpression of Pin1 antisense RNA, (b) overexpression of dominant-negative Pin1, and (c) by a known small-molecule Pin1-PPIase inhibitor, juglone. The results of all of the three lines of investigation show that the catalytic activity of Pin1 is essential for tumor cell survival and entry into mitosis.

Introduction

The cell cycle of eukaryotic cells is regulated by the coordinated phosphorylation and dephosphorylation of proteins and their ubiquitin-dependent proteolytic degradation. For example, entry into mitosis requires Cdc23 and Cyclin B, which assemble into a kinase complex responsible for the phosphorylation of important mitotic proteins at Ser/Pro and Thr/Pro motifs. Cyclin B itself must be degraded by the anaphase promoting complex/cyclosome complex to allow exit from mitosis. In Aspergillus nidulans, an additional kinase, NIMA, is required for proper progression through mitosis (1) . Cells lacking functional NIMA arrest in G2, whereas deregulated expression of NIMA induces premature mitosis (2) . A NIMA-like pathway has been proposed for vertebrate cells as well (3) . Recently, a protein interacting with NIMA (Pin1) has been identified by a yeast-two-hybrid screen, using NIMA as bait (4) . Pin1 is a small Mr 18,000 protein with two distinct functional domains, namely a NH2-terminal WW domain and a COOH-terminal PPIase domain). Pin1 was found to be a homologue to the Drosophila dodo protein (5) , and it complements the essential Saccharomyces cerevisiae homologue, Ess1/Ptf1 (4) . Deregulation of Pin1-levels in yeast, human HeLa cancer cells, and Xenopus laevis egg extracts has the reciprocal phenotype to NIMA manipulation (4, 5) , which suggests that Pin1 negatively regulates entry into mitosis and is essential for progression through mitosis. However, the deletion of the Pin1 homologue in Drosophila melanogaster (6) and in the mouse (7) did not result in readily observable phenotypes. The 39-amino-acid WW domain of Pin1 acts as a phosphoserine- and phosphothreonine-interaction module involved in the binding of at least 30 mitotic phosphoproteins, including Cdc25C, Plk1, Cdc27 (5, 8, 9) and phosphorylated {tau} protein (10) . Other functions for the Pin1-WW domain have been suggested, including regulation of subcellular localization, regulation of nuclear transport, transcription-promoting activity (11) , and pre-mRNA 3'-end formation (12) . The COOH-terminal domain of Pin1 is a phosphorylation-dependent PPIase that catalyzes the cis-to-trans isomerization of phosphoserine-proline and phosphothreonine-proline bonds with up to 1300-fold selectivity compared with unphosphorylated substrates (13) . Initial biochemical and crystal structure analysis revealed that Cys-113, His-59, His-157, and Ser-154 form a catalytic cascade with Cys-113 proposed to function as the general base in a nucleophilic reaction, and Lys-63, Arg-68, and Arg-69 bind the phosphate group of the substrate (13, 14) . Pin1 may, thus, act as a molecular switch by binding a subset of proteins phosphorylated at Ser/Thr-Pro motifs and regulating their biological activity through the isomerization of peptidyl-prolyl bonds (14) . However, proof that the PPIase activity of Pin1 is required for its function in vertebrate cells is still lacking. In this study, we addressed the importance of the Pin1 enzymatic activity in human tumor cell-lines by transient overexpression of Pin1 dominant negative proteins, by depletion of endogenous Pin1 through overexpression of Pin1-antisense RNA, and through experiments with the known Pin1-PPIase inhibitor, juglone.

Results

Catalytic Activity of Pin1 and Pin1 Mutant Proteins.
To study Pin1 at the biochemical and cellular level, wild-type Pin1 and a set of mutant proteins were characterized in vitro and in human tumor cell lines using transient transfection experiments. For biochemical characterization, Pin1, the prolyl-isomerase domain (PPIase), and a set of Pin1 mutants, including Pin1-C113A (C113A), Pin1-R68L (R68L), Pin1-R69L (R69L), and the double point-mutant Pin1-R68,69L (R68,69L) were expressed in Escherichia coli and purified to homogeneity. The rate constant of prolyl isomerization was determined by a protease-coupled isomerase assay (15) . When increasing amounts of the various proteins were incubated with the phosphorylated substrate, Ac-AApSPR-pNA, efficient isomerization was observed in the presence of full-length Pin1 and the PPIase domain (Fig. 1ACitation and Table 1Citation ). In contrast, Pin1-R68,69L exerted significant isomerase activity only at protein levels >1000 ng, whereas C113A was catalytically inactive in the protein range tested (Fig. 1ACitation ). Calculation of the specificity constant, kcat/Km, revealed that Pin1 catalyzed cis-to-trans isomerization approximately 400-fold more efficiently than R68,69L, 2-fold better than R68L, and 100-fold better than R69L, respectively (Table 1)Citation . When a nonphosphorylated substrate was used, Pin1 and R68,69L were equally active at high enzyme concentrations, whereas C113A again showed only little activity (Fig. 1BCitation ; Table 1Citation ). These data suggest that the side chain of C113 is required for the catalytic reaction per se, whereas the two basic residues R68 and R69 coordinate the binding of the phosphorylated substrate, with R69 being more important than R68.



View larger version (13K):
[in this window]
[in a new window]
 
Fig. 1. Catalytic activity of Pin1 and mutant proteins. The PPIase assay was performed in triplicate as described in "Materials and Methods" in the presence of increasing amounts of purified recombinant proteins by monitoring the release of 4-nitroaniline after isomerization of (A) phosphorylated or (B) unphosphorylated substrate. {blacksquare}, Pin1; , R68,69L; {triangleup}, C113A.

 

View this table:
[in this window]
[in a new window]
 
Table 1 Catalytic efficiency (kcat/Km) of Pin1 and Pin1 mutants [M-1s-1]a

 
The WW Domain of Pin1 Is Required for Localization in Nuclear Speckles.
The localization of Pin1 at nuclear substructures variously termed interchromatin granule clusters (IGCs) or speckles has been shown (4) . To identify the functional domain required for the observed localization, Pin1, the PPIase domain, the WW domain, and various Pin1 mutants were fused to GFP and transiently overexpressed in HeLa cells. Their subcellular localization was analyzed by microscopic inspection. As shown in Fig. 2ACitation , Pin1 and the WW domain localize almost exclusively in the cell nucleus and concentrate at discrete structures. C113A and R68,69L display a more diffuse staining pattern throughout the cell with a less distinctive nuclear staining. In contrast, the PPIase domain and GFP distribute throughout the cell and show no localization in nuclear speckles. Interestingly, the WW domain also colocalized with nuclear speckles identical to those detected by antibodies directed against the spliceosome assembly factor SC-35 (Fig. 2BCitation ) but not to promyelocytic leukemia oncogenic domains (PODs; data not shown). These data indicate that the Pin1-WW domain is sufficient to direct the localization of Pin1 into nuclear speckles.



View larger version (45K):
[in this window]
[in a new window]
 
Fig. 2. Intracellular localization of GFP-tagged Pin1 and GFP-tagged mutant proteins and their localization in nuclear speckles. A, HeLa cells transiently transfected with GFP, GFP-Pin1 (Pin1), GFP-Pin-C113A (C113A), GFP-Pin1-isomerase domain (PPIase), GFP-Pin1-WW domain (WW), and GFP-Pin1-R68,69L (R68,69L) were analyzed for the localization of the tagged proteins. B, colocalization of Pin1 and the Pin1-WW domain with the spliceosome assembly factor, SC-35. HeLa cells transiently transfected with GFP-Pin1 (Pin1), or GFP-WW domain (WW), were inspected for the intracellular localization of GFP, the SC-35 staining (Anti SC-35), and the superimposition of the individual images (Merge).

 
Depletion of Pin1 Induces Apoptosis in Eukaryotic Cells.
The importance of Pin1 for cell survival was addressed by overexpressing Pin1 and Pin1 mutants, with Pin1-antisense RNA, and by the use of the known Pin1 inhibitor, juglone (5-hydroxy-1,4-naphthoquinone; Ref. 16 ). In a first set of experiments, three different cell lines were transfected with expression vectors encoding GFP, a GFP-Pin1 fusion protein, or GFP and Pin1 antisense RNA. Two days after transfection cells were fixed and stained with DAPI, and the presence of apoptotic nuclei was scored by microscopic inspection. As shown in Fig. 3ACitation , apoptosis was slightly increased in cells transfected with GFP-Pin1 compared with GFP-only transfected cells. In contrast, Pin1 antisense overexpressing cells displayed a massive increase in apoptosis. A comparable strong induction of apoptosis was observed in HeLa cells overexpressing the PPIase domain, the catalytic inactive mutant C113A, and C113A lacking the WW domain (Fig. 3BCitation , C113A-{Delta}WW), which suggests that these molecules act as dominant-negative proteins. In contrast, cells transfected with the WW domain, R68,69L, and R68,69L lacking the WW domain (R68,69L-{Delta}WW) showed no significant increase in the number of apoptotic nuclei (Fig. 3BCitation ). Western blot analysis revealed a >100-fold overexpression of all GFP-Pin1 fusion proteins as compared with the endogenous levels of Pin1 (data not shown). Cultivation of HeLa cells in the presence of juglone induced apoptosis in a dose-dependent manner as detected by the appearance of a sub-G1 peak in flow cytometric DNA analysis, microscopic inspection, and TUNEL assays (Fig. 3CCitation and data not shown). The induction of apoptosis by juglone correlated with its IC50 value of 20 µM for Pin1 enzyme inhibition.



View larger version (29K):
[in this window]
[in a new window]
 
Fig. 3. Pin1 is essential for tumor cell survival. A, HeLa, 293, and SK-N-AS-cells, transiently transfected with GFP (GFP), GFP-Pin1 (Pin1), and GFP-Pin1 antisense (Pin1-AS) were analyzed for the frequency of apoptotic nuclei 48 h after transfection. B, HeLa cells, transiently transfected with GFP, GFP-tagged Pin1, or the indicated GFP-tagged mutant proteins, were analyzed for the frequency of apoptotic nuclei 72 h after transfection. A and B, average of three independent experiments is shown; bars, SE. C, juglone induces apoptosis in human tumor cells. Flow cytometric DNA analysis of HeLa cells grown for 72 h in the presence of the indicated concentration of juglone.

 
Depletion of Pin1 Blocks Entry into Mitosis and Induces Cell Death.
Because a function for Pin1 as an essential mitotic regulator has been suggested (4, 8) , we investigated the mitotic effects of Pin1 depletion in closer detail. Mitotic events were analyzed by microscopic inspection using DAPI staining and antibodies specific for {alpha}-tubulin, nuclear lamin, and phosphorylated histone H3 as markers. As shown in Fig. 5ACitation , histone H3 phosphorylation is detected at all mitotic stages, in agreement with previously published data (17, 18) . The same index of mitotic nuclei was detected in cells transfected with GFP-Pin1 and in GFP control cells at various times after transfection (Fig. 4ACitation and data not shown). Furthermore, Pin1 overexpression did not arrest cells at a certain mitotic phase. In contrast, the number of mitotic nuclei was reduced in cells overexpressing Pin1 antisense RNA (Fig. 4ACitation ). Mitotic nuclei were already absent 6 h after transfection (data not shown), which suggested that Pin1 depletion does not induce cell death in mitosis but, rather, impairs cells from entering mitosis.



View larger version (21K):
[in this window]
[in a new window]
 
Fig. 5. Pin1 depletion induces apoptosis in interphase. A, phosphorylation of histone H3 correlates with mitotic phases. Fixed HeLa cells were double-stained with DAPI and with an antibody recognizing phosphorylated histone H3. The nuclei of cells in prophase (P), metaphase (M), and telophase (T) are shown. B, mitotic cell death. HeLa cells were incubated for 48 h with 50 nM Taxol and double-stained as in A. C, Pin1 depletion induces apoptosis in interphase. Seventy-two h after transfection with GFP and Pin1 antisense, fixed cells were double-stained with DAPI and with an antibody recognizing phosphorylated histone H3. II and IV, same images as in I and III superimposed with GFP-emitted green fluorescence. Arrows, transfected apoptotic cells.

 


View larger version (15K):
[in this window]
[in a new window]
 
Fig. 4. Pin1 is required for entry into mitosis. A, Mitotic indices of cells transfected with GFP (GFP), GFP-Pin1 (Pin1), and GFP-Pin1 antisense (Pin1-AS). Forty-eight h after transfection, HeLa, 293, and SK-N-AS cells were stained with an antibody directed against phosphorylated histone H3, and the frequency of mitotic events was determined for the transfected cells. B, frequency of mitosis in HeLa cells transiently transfected with GFP, GFP-tagged Pin1, or the indicated mutant proteins. Seventy-two h after transfection, cells were stained with an antibody directed against phosphorylated histone H3, and the nuclei of transfected cells were scored for the frequency of mitosis. Average results from four independent experiments are shown. Bars, SE.

 
HeLa cells, overexpressing the Pin1 WW domain, R68,69L, or R68,69L-{Delta}WW showed mitotic indices comparable with the GFP control cells, whereas mitosis was significantly reduced in cells overexpressing the PPIase domain, C113A, or C113A-{Delta}WW (P < 0.05, according to Student’s test). Consistent with the experiments shown in Fig. 3CCitation , juglone-treated cells did not accumulate in mitosis. However, those nuclei that entered mitosis in the presence of juglone arrested in pro- and metaphase (data not shown).

Impaired Pin1 Function Induces Apoptosis in Interphase.
To study in closer detail at which stage of the cell cycle Pin1 function is essential for cell survival, apoptotic cells that were transfected with GFP and Pin1-antisense were inspected by immunofluorescence for the presence of phosphorylated histone H3. As shown in Fig. 5CCitation , phosphorylated histone H3 is not detected in apoptotic nuclei of Pin1-depleted cells, which suggests that these cells do not die in mitosis. In contrast, phosphorylated histone H3 is present in cells dying in metaphase after treatment with Taxol (Fig. 5BCitation ).

Discussion

Depletion of Pin1 by overexpression of Pin1 antisense RNA induces mitotic arrest and apoptosis in human tumor cells (4) . However, the Pin1 protein comprises distinct functional domains, including the WW and PPIase domains, and whether the enzymatic activity of the PPIase domain is required for this Pin1-mediated mechanism has not been previously defined. To address this question, GFP-tagged proteins—namely, Pin1, the isolated WW domain, the PPIase domain, and a set of Pin1-mutants—impaired either in catalysis (C113A, C113A-{Delta}WW), or substrate recognition (R68,69L, R68,69L-{Delta}WW), were overexpressed in human tumor cell lines.

Transient transfection of HeLa cells with GFP-tagged Pin1 and the WW domain resulted in a predominantly nuclear localization, with high-lighting of nuclear speckles (Fig. 2)Citation . In contrast, the PPIase domain showed no preferential nuclear staining but, instead, was distributed throughout the cell. The mutants C113A and R68,69L exerted a more diffuse but still speckled nuclear staining. These data suggest that the WW domain is essential and sufficient for the retention of Pin1 in nuclear speckles. However, additional interactions between cellular substrates and the active site of Pin1 may either influence the affinity toward proteins located in nuclear speckles or influence the nuclear import of Pin1. The presence of Pin1 in nuclear speckles may, therefore, reflect important WW domain as well as active-site interactions. Functionally, nuclear speckles could serve as storage sites for Pin1 as they do—it has been suggested—for splicing factors (19) .

Cells overexpressing the PPIase domain and the catalytic inactive mutants C113A and C113A-{Delta}WW showed a massive increase in the rate of dead cells comparable with those observed in cells overexpressing Pin1 antisense RNA. All of the hallmarks of apoptosis were observed, including chromatin condensation and fragmentation, membrane blebbing, and disassembly of nuclear lamins (20) . Presumably, the mutants C113A and C113A-{Delta}WW act as dominant negative versions of Pin1 by competing with endogenous Pin1 for the binding of physiological substrates but are unable to catalyze the essential conformational changes in substrate proteins required for cell survival. Our data suggest that the localization of Pin1 in nuclear speckles is not required for the induction of apoptosis because the overexpression of the PPIase domain and C113A-{Delta}WW, two mutants that do not bind to these nuclear substructures, still induces cell death. Transfection of Pin1, the WW domain, and the mutants R68,69L and R69,69L-{Delta}WW into human tumor cells did not result in an increase of apoptosis, despite the fact that these proteins also were at least 100-fold overexpressed compared with endogenous Pin1. These data show that high levels of Pin1 and of the WW domain are tolerated by cells and suggest that mutants like R68,69L and R69,69L-{Delta}WW do not compete with endogenous Pin1 for essential substrates, presumably because of their reduced affinity toward phosphorylated substrates.

An essential mitotic function for Pin1 has been proposed because depletion of Pin1 resulted in abnormal mitotic phenotypes and subsequent apoptosis (4) . We reinvestigated the effects of Pin1 depletion on mitosis using phosphorylation of histone H3 as a mitotic marker. Depletion of Pin1 by antisense RNA resulted in a complete absence of mitotic cells, in three of three human tumor cell lines tested. Similarly, mitosis was reduced in cells overexpressing the dominant negative proteins C113A and C113A-{Delta}WW, which suggests that the enzymatic activity of Pin1 is required for G2-M transition.

In contrast to previous observations (4) , we did not observe abnormal mitotic events in cells overexpressing Pin1 antisense RNA or the dominant negative mutants C113A and C113A-{Delta}WW, nor did we detect cells dying in mitosis. The morphological features like cell rounding, chromatin condensation, lamin disassembly, previously observed in Pin1depleted human tumor cells, actually may reflect apoptosis in interphase rather than mitotic cell death (20) .

This findings were underscored by the use of juglone, the only known inhibitor of the parvulin-like PPIases (16) . In agreement with the overexpression of C113A and C113A-{Delta}WW and with previous findings, juglone induced apoptosis (21) and prevented cells from entering into mitosis (22) . However, because juglone is a cysteine-reactive compound (16, 23) and has inhibitory activity on other proteins in vitro (23, 24) , these findings need to be reevaluated with specific Pin1 inhibitors.

In agreement with our findings, the requirement of Pin1 in an interphase checkpoint has recently been described (25) . In Xenopus leavis egg extracts Pin1 participates in the replication checkpoint dependent on its catalytic activity. This novel function of Pin1 may explain why, in the absence of DNA damage, no apparent phenotypes were observed in mice lacking Pin1. (6)

In conclusion, our data show that the WW domain of Pin1 is required for the localization of Pin1 in nuclear speckles and suggest that phosphorylation-dependent peptidyl-prolyl cis trans isomerization catalyzed by Pin1 is essential for cell survival and for entry into mitosis. Consequently, these Pin1-PPIase-dependent mechanisms may serve as novel targets for the treatment of diseases associated with impaired apoptosis.

Materials and Methods

Plasmids.
Pin1 was cloned by reverse transcription PCR from K-562 cells using primers 5'-CCGGATCCATGGCGGACGAGGAGAAGCTGC3' and 5'-CGAATTCAAGCTTCGAGGCCAGGCCTGGGCTCC-3'. Sitedirected mutations of Pin1 were introduced using PCR-based techniques. As template, the Pin1 cDNA cloned into pCR2.1 (Invitrogen, Carlsbad, CA) was used. For the generation of each mutant, two PCR reactions were carried out. For C113A, the primer 5'-CAGCGACGCCAGCTCAGCCAAGGC-3'; for R68L, the primer 5'-CAGTCACTGCGGCCCTCGTCCTGG-3'; and for R69L, the primer 5'-CAGTCACGGCTGCCCTCGTCCTGG-3' were used together with the T7 primer. For the second PCR, the corresponding complementary oligonucleotides and the M13reverse primer were used. For each construct, the two respective PCR products were annealed, a fill-in reaction was performed, and the resulting DNA fragment was used for a third PCR with T7 and M13reverse primers. The R68,69L mutation was generated by introducing the R69L mutation in a construct already containing the R68L mutation. The mutations were verified by automated sequencing. The WW domain (amino acids 1–41) was generated by PCR using primers 5'-CCGGATCCATGGCGGACGAGGAGAAGCTGC-3' and 5'-AGCTAGCTGTTGCCGCTGGGCCGC-3'. The isomerase domain (amino acids 45–163) was generated using primers 5'-CCGGATCCATATGGGCAAAAACGGGCAGGGGGAGC-3' and 5'-CGAATTCAAGCTTCGAGGCCAGGCCTGGGCTCC-3'. For expression of GST-Pin1 fusion proteins in bacteria, the cDNAs encoding Pin1, the WW domain, the PPIase domain and the various mutations were cloned into pGEX-4T-1 (Amersham Pharmacia Biotech, Freiburg, Germany). For mammalian expression, the respective cDNAs were cloned in frame to GFP into pEGFP-C1 (Clontech, Palo Alto, CA). The GFP-Pin antisense vector was obtained by cloning the Pin1 PCR fragment with EcoRI, BamHI into the EcoRI, BamHI sites of pEGFP-C1 (Clontech).

Protein Expression and Purification.
Bacterial cells (XL1-blue; Stratagene, Heidelberg, Germany) that harbored pGEX-4T-1-derived expression plasmids were grown in Luria Bertani media. Protein expression was induced with 1 mM IPTG in midexponential growth phase. Four h after the addition of IPTG, cells were harvested by centrifugation (10,000 x g, 10 min, 4°C) and the pellet was resuspended in lysis buffer [20 mM disodium phosphate (pH 7.0), 1 mg/ml lysozyme, 10 mM EDTA, 1% (v/v) Triton X-100, 10 µg/ml DNase, protease-inhibitors (Roche Diagnostics, Mannheim, Germany)] at one-tenth of the culture volume. Cells were lysed by three cycles of freeze-thawing, and unsoluble protein was separated by centrifugation (10,000 x g, 10 min, 4°C). The supernatant was incubated with one-fifth volume of preswollen glutathione Sepharose 4B (Amersham Pharmacia Biotech, Freiburg, Germany) for 45 min at 4°C. Beads were separated from the supernatant and washed by repeated centrifugation (300 x g, 5 min, 4°C) and resuspension in PBS. Pin1 or mutant proteins were released by incubation with bovine {alpha}-thrombin (1 µg/ml; Sigma, Deisenhofen, Germany) for 3 h at 30°C. Beads were separated from the supernatant by centrifugation (300 x g, 10 min, 4°C). Bovine {alpha}-thrombin was removed by incubation of the supernatant with benzamidine-Sepharose (Amersham Pharmacia Biotech, Freiburg, Germany) at 4°C for 30 min and subsequent centrifugation (300 x g, 10 min, 4°C). After addition of 1 mM DTT and 5 µg/ml Pefabloc (Serva, Heidelberg, Germany) the supernatant was filtered (0.22 µm microcentrifuge filter units; Millipore, Bedford) and concentrated (Centricon 10; Amicon; Bedford, MA). Aliquots (20 µg/ml) were frozen in liquid nitrogen and stored at -80°C.

Generation of Polyclonal Antisera against Pin1.
Two female rabbits were immunized with 12.5 µg of purified GST-Pin1 per animal, together with Friend’s adjuvans (Life Technologies, Inc., Karlsruhe, Germany). Prior to the s.c. (6 sides, 150 µl/side) and muscle (4 sides, 25 µl/side) administration, the protein was adsorbed to 2% AluGelS (Serva, Heidelberg, Germany). Immunization was repeated three times with 25 µg of GST-Pin1, and animals were bled 2 weeks after the final immunization.

Western Blot Analysis.
HeLa cell extracts were prepared by detaching the adherent cells with trypsin/EDTA (Life Technologies, Inc., Karlsruhe, Germany), triplicate washing with PBS, and lysis of the cell pellet by 4x SDS-loading buffer (RotiLoad; Roth, Karlsruhe, Germany) at 95°C for 5 min. An aliquot corresponding to 2 x 105 cells was loaded on a 15% SDS-polyacrylamide gel. The proteins were separated by electrophoresis and electroblotted onto nitrocellulose (Schleicher & Schuell, Dassel, Germany). Endogenous or overexpressed Pin1 mutant proteins were detected by incubation of the membrane for 2 h at room temperature with rabbit anti-GST-Pin1 serum 1:250 in M-TBS and subsequent incubation for 1 h with horseradish peroxidase-linked donkey antirabbit serum (Amersham, Braunschweig, Germany; 1:1000 in 5% M-TBS). After each step, membrane was washed three times with 0.05% Tween 20 in TBS. Luminescent signals were generated with SuperSignal West Pico substrate (Pierce, IL) and recorded by autoradiography.

Determination of PPIase Activity.
The PPIase activity was measured as described (15) . Briefly, purified proteins (1–6000 ng) were preincubated with 72 µM substrate (Ac-AApSPR-pNA or Ac-AASPR-pNA; Interactiva, Ulm, Germany) in a reaction volume of 60 µl in the presence of 35 mM HEPES (pH 7.8), 0.4 mg/ml BSA, and 2 mM DTT for 30 min at 4°C. Proteolysis of substrate was initiated by the addition of 70 µl trypsin (86 µg/ml) in 35 mM HEPES (pH 7.8). The release of p-nitroaniline was followed with a microplate reader (MR5/7000/DIAS; Dynex Technologies, Denkendorf, Germany) at 390 nm every 9 s. Resulting plots were fitted to a first-order reaction and from the rate constant kobs., the catalytic efficiency kcat/Km was calculated as described previously (15) .

Cell Culture and Transfection of Mammalian Cells.
HeLa cells (ATCC, CCL-2) were grown in RPMI 1640 (Bio Whittaker, de PetitRechain, Belgium), supplemented with 10% FCS; the human kidney cell line 293 (ATCC, CRL-1573) and the neuroblastoma cell line SK-N-AS (ATCC, CRL-2137) were grown in DMEM F-12 (Bio Whittaker) supplemented with 10% FCS. Twenty-four h before transfection, 2 x 105 cells were seeded onto poly-D-lysine-coated chamber slides and were incubated at 37°C in 5% CO2. Transfection was performed using 1 µg of plasmid DNA and 5 µl of Fugene reagent (Roche Diagnostics, Mannheim, Germany) according to the manufacturer’s protocol.

Immunofluorescence Staining.
The following primary antibodies were used: antiphospho-histone-H3 rabbit serum (1:500 dilution; Upstate Biotechnology, Lake Placid, NY); mouse monoclonal IgG recognizing spliceosome assembly factor SC-35 (1:200 dilution, clone SC-35; Sigma, St. Louis, MO); mouse monoclonal IgG PG-M3, directed against promyelocytic leukemia (1:200 dilution; Santa Cruz Biotechnology, Santa Cruz, CA); mouse monoclonal antilamin solution (undiluted; Progen, Heidelberg, Germany); and mouse monoclonal anti-{alpha}-tubulin IgG (1:500 dilution; Sigma). Briefly, cells, fixed for 20 min with 2% paraformaldehyde, were treated with 0.25% Triton X-100 for 10 min and with 2% BSA/PBS for 30 min and were incubated for 1 h with 200 µl of the respective antibody-solution diluted in 2% BSA/PBS. For fluorescent labeling, cells were further incubated with secondary rhodamine-labeled goat antirabbit IgG, or rhodamine-labeled goat antimouse IgG (1:1000 dilution; Jackson Immuno Research Laboratories, West Grove, PA) in 2% BSA/PBS for 30 min. Cellular DNA was stained with 2.5 µg/ml DAPI in PBS for 5 min. All of the steps were performed at room temperature, and cells were washed two times with PBS after each step. Finally, cells were mounted in 80% glycerol/PBS and stored at 4°C. For documentation and microscopic analysis, a fluorescence microscope (Axioplan 2; Zeiss, Jena, Germany) and a laser scanning microscope (DM IRB/E; Leica, Bensheim, Germany) were used.

Apoptosis.
The index of apoptotic cells was generally determined after DAPI staining. Nuclei with highly condensed and fragmented DNA were scored positive. This phenotype was confirmed by positive TUNEL staining of the cells, which was performed according to the manufacturer’s protocol (ApopTag fluorescein direct in situ apoptosis detection kit; Intergen Company, Oxford, United Kingdom).

Fluorescence-activated Cell Sorting.
Ten thousand cells per analysis were used. Cells were incubated for 72 h with the indicated concentrations of juglone, fixed with 2% paraformaldehyde for 20 min at room temperature, and permeabilized with 0.25% Triton X-100 in PBS by incubation for 5 min on ice. Cells were pelleted by centrifugation (1000 x g, 5 min, 4°C), resuspended in propidium iodine staining buffer (0.1% RNase, 10 µg/ml propidium iodine in PBS) and incubated for 20 min at room temperature. The DNA content was analyzed using a FACS Calibur (Becton Dickinson, Heidelberg, Germany).

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 These two authors contributed equally to this work. Back

2 To whom requests for reprints should be addressed, at Department of Oncology Research, Boehringer Ingelheim Pharma KG, Birkendorfer Strasse 65, 88397 Biberach an der Riss, Germany. Phone: 49-7351-54-5240; Fax: 49-7351-54-5146; E-mail: andreas.schnapp{at}bc.boehringer-ingelheim.com Back

3 The abbreviations used are: Cdc, cell division cycle; Ac, acetyl; DAPI, 4',6'-diamino-2-phenylindol; GFP, green fluorescent protein; GST, glutathione S-transferase; IPTG, isopropyl-1-thio-ß-D-galactoside; NIMA, never in mitosis A; pNA, para-nitroaniline; PPIase, peptidyl-prolyl cis-trans isomerase; TUNEL, terminal deoxynucleotidyl transferase-mediated nick end labeling; M-TBS, 5% dried milk, 25 mM Tris (pH 7.4), 150 mM NaC. Back

Received for publication 12/15/99. Revision received 3/23/00. Accepted for publication 5/16/00.

References

  1. Osmani S. A., May G. S., Morris N. R. Regulation of the mRNA levels of nimA, a gene required for the G2-M transition in Aspergillus nidulans. J. Cell Biol, 104: 1495-1504, 1987.[Abstract/Free Full Text]
  2. Osmani S. A., Pu R. T., Morris N. R. Mitotic induction and maintenance by overexpression of a G2-specific gene that encodes a potential protein kinase. Cell, 53: 237-244, 1988.[Medline]
  3. Lu K. P., Hunter T. Evidence for a NIMA-like mitotic pathway in vertebrate cells. Cell, 81: 413-424, 1995.[Medline]
  4. Lu K. P., Hanes S. D., Hunter T. A human peptidyl-prolyl isomerase essential for regulation of mitosis. Nature (Lond.), 380: 544-547, 1996.[Medline]
  5. Crenshaw D. G., Yang J., Means A. R., Kornbluth S. The mitotic peptidyl-prolyl isomerase, Pin1, interacts with Cdc25 and Plx1. EMBO, 17: 1315-1327, 1998.[Abstract]
  6. Fujimori F., Takahashi K., Uchida C., Uchida T. Mice lacking Pin1 develop normally, but are defective in entering cell cycle from G0 arrest. Biochem. Biophys. Res. Commun, 265: 658-663, 1999.[Medline]
  7. Maleszka R., Hanes S. D., Hackett R. L., deCouet H. G., Miklos G. L. The Drosophila melanogaster dodo (dod) gene, conserved in humans, is functionally interchangeable with the ESS1 cell division gene of Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA, 93: 447-451, 1996.[Abstract/Free Full Text]
  8. Shen M., Stukenberg P. T., Kirschner M. W., Lu K. P. The essential mitotic peptidyl-prolyl isomerase Pin1 binds and regulates mitosis-specific phosphoproteins. Genes Dev, 12: 706-720, 1998.[Abstract/Free Full Text]
  9. Lu P-J., Zhou X. Z., Shen M., Lu K. P. Function of WW domains as phosphoserine- or phosphothreonine-binding modules. Science (Washington DC), 283: 1325-1328, 1999.[Abstract/Free Full Text]
  10. Lu P-J., Wulf G., Zhou X. Z., Davies P., Lu K. P. The prolyl isomerase Pin1 restores the function of Alzheimer-associated phosphorylated {tau} protein. Nature (Lond.), 399: 784-788, 1999.[Medline]
  11. Komuro A., Saeki M., Kato S. Npw38, a novel nuclear protein possessing a WW domain capable of activating basal transcription. Nucleic Acids Res, 27: 1957-1965, 1999.[Abstract/Free Full Text]
  12. Morris D. P., Phatnani H. P., Greenleaf A. L. Phospho-carboxyl-terminal domain binding and the role of a prolyl isomerase in pre-mRNA 3'-end formation. J. Biol. Chem, 274: 31583-31587, 1999.[Abstract/Free Full Text]
  13. Yaffe M. B., Schutkowski M., Shen M., Zhou X. Z., Stukenberg P. T., Rahfeld J-U., Xu J., Kuang J., Kirschner M. W., Fischer G., Cantley L. C., Lu K. P. Sequence-specific and phosphorylation-dependent proline isomerization: a potential mitotic regulatory mechanism. Science (Washington DC), 278: 1957-1960, 1997.[Abstract/Free Full Text]
  14. Ranganathan R., Lu K. P., Hunter T., Noel J. P. Structural and functional analysis of the mitotic rotamase Pin1 suggests substrate recognition is phosphorylation dependent. Cell, 89: 876-886, 1997.
  15. Fischer G., Bang H., Mech C. Determination of enzymatic catalysis for the cis-trans-isomerization of peptide binding in proline-containing peptides (in German). Biomed. Biochim. Acta, 43: 1101-1111, 1984.[Medline]
  16. Henning L., Christner C., Kipping M., Schelbert B., Rücknagel K. P., Grabley S., Küllertz G., Fischer G. Selective inactivation of parvulin-like peptidyl-prolyl cis/trans isomerase by juglone. Biochemistry, 37: 5953-5960, 1998.[Medline]
  17. Wei Y., Yu L., Bowen J., Gorovsky M. A., Allis C. D. Phosphorylation of histone H3 is required for proper chromosome condensation and segregation. Cell, 97: 99-109, 1999.[Medline]
  18. Goto H., Tomono Y., Ajiro K., Kasoko H., Fujita M., Sakurai M., Okawa K., Iwamatsu A., Okigaki T., Takahashi T., Inagaki M. Identification of a novel phosphorylation site on histone H3 coupled with mitotic chromosome condensation. J. Biol. Chem, 274: 25543-25549, 1999.[Abstract/Free Full Text]
  19. Sleeman J. E., Lamond A. I. Nuclear organization of pre-mRNA splicing factors. Curr. Opin. Cell Biol, 11: 372-377, 1999.[Medline]
  20. King L. K., Cidlowski J. A. Cell cycle and apoptosis: common pathways to life and death. J. Cell. Biochem, 58: 175-180, 1995.[Medline]
  21. Okada T. A., Roberts E., Brodie A. F. Mitotic abnormalities produced by juglone in Ehrlich ascites tumor cells. Proc. Soc. Exp. Biol. Med, 126: 583-588, 1967.[Abstract/Free Full Text]
  22. Stockwell B. R., Haggarty S. J., Schreiber S. L. High-throughput screening of small molecules in miniaturized mammalian cell-based assays involving post-translational modifications. Chem. Biol, 6: 71-83, 1999.[Medline]
  23. Frew, T., Powis, G., Berggren, M., Gallegos, A., Abraham, R. T., Ashendel, C. L., Zalkow, L. H., Hudson, C., Gruszecka-Kowalik, E., Burgess, E. M., Benedetti-Doctorovich, V., Kerrigan, J. E., Lambropoulos, J., Merriman, R., and Bonjouklian, R. Novel quinone antiproliferative inhibitors of phosphatidylinositol-3-kinase. Anticancer Drug Des. 10: 347–359, 1995.
  24. Duhaiman A. S. Characterization of {zeta}-crystallin inhibition by juglone. Biochem. Biophys. Res. Commun, 218: 648-652, 1996.[Medline]
  25. Winkler K. E., Swenson K. I., Kornbluth S., Means A. R. Requirements of the prolyl isomerase Pin1 for the replication checkpoint. Science (Washington DC), 287: 1644-1647, 2000.[Abstract/Free Full Text]



This article has been cited by other articles:


Home page
Mol. Cell. Biol.Home page
A. S. Farrell, C. Pelz, X. Wang, C. J. Daniel, Z. Wang, Y. Su, M. Janghorban, X. Zhang, C. Morgan, S. Impey, et al.
Pin1 Regulates the Dynamics of c-Myc DNA Binding To Facilitate Target Gene Regulation and Oncogenesis
Mol. Cell. Biol., August 1, 2013; 33(15): 2930 - 2949.
[Abstract] [Full Text] [PDF]


Home page
Biochm. Soc. Trans.Home page
G. Gudavicius, H. Soufari, S. Upadhyay, C. D. Mackereth, and C. J. Nelson
Resolving the functions of peptidylprolyl isomerases: insights from the mutagenesis of the nuclear FKBP25 enzyme
Biochm. Soc. Trans., June 1, 2013; 41(3): 761 - 768.
[Abstract] [Full Text] [PDF]


Home page
J. Virol.Home page
Y.-S. Lim, H. T. L. Tran, S.-J. Park, S.-A. Yim, and S. B. Hwang
Peptidyl-Prolyl Isomerase Pin1 Is a Cellular Factor Required for Hepatitis C Virus Propagation
J. Virol., September 1, 2011; 85(17): 8777 - 8788.
[Abstract] [Full Text] [PDF]


Home page
Cancer Res.Home page
A. van der Horst and K. K. Khanna
The Peptidyl-Prolyl Isomerase Pin1 Regulates Cytokinesis through Cep55
Cancer Res., August 15, 2009; 69(16): 6651 - 6659.
[Abstract] [Full Text] [PDF]


Home page
J. Cell Sci.Home page
J. Kim, T. G. Choi, Y. Ding, Y. Kim, K. S. Ha, K. H. Lee, I. Kang, J. Ha, R. J. Kaufman, J. Lee, et al.
Overexpressed cyclophilin B suppresses apoptosis associated with ROS and Ca2+ homeostasis after ER stress
J. Cell Sci., November 1, 2008; 121(21): 3636 - 3648.
[Abstract] [Full Text] [PDF]


Home page
J Biol ChemHome page
C. Fila, C. Metz, and P. van der Sluijs
Juglone Inactivates Cysteine-rich Proteins Required for Progression through Mitosis
J. Biol. Chem., August 1, 2008; 283(31): 21714 - 21724.
[Abstract] [Full Text] [PDF]


Home page
Mol. Cell. Biol.Home page
F. Suizu, A. Ryo, G. Wulf, J. Lim, and K. P. Lu
Pin1 Regulates Centrosome Duplication, and Its Overexpression Induces Centrosome Amplification, Chromosome Instability, and Oncogenesis
Mol. Cell. Biol., February 15, 2006; 26(4): 1463 - 1479.
[Abstract] [Full Text] [PDF]


Home page
Mol. Cell. Biol.Home page
S.-Y. Chen, G. Wulf, X. Z. Zhou, M. A. Rubin, K. P. Lu, and S. P. Balk
Activation of {beta}-Catenin Signaling in Prostate Cancer by Peptidyl-Prolyl Isomerase Pin1-Mediated Abrogation of the Androgen Receptor-{beta}-Catenin Interaction
Mol. Cell. Biol., February 1, 2006; 26(3): 929 - 939.
[Abstract] [Full Text] [PDF]


Home page
J Biol ChemHome page
F. Eckerdt, J. Yuan, K. Saxena, B. Martin, S. Kappel, C. Lindenau, A. Kramer, S. Naumann, S. Daum, G. Fischer, et al.
Polo-like Kinase 1-mediated Phosphorylation Stabilizes Pin1 by Inhibiting Its Ubiquitination in Human Cells
J. Biol. Chem., November 4, 2005; 280(44): 36575 - 36583.
[Abstract] [Full Text] [PDF]


Home page
J Biol ChemHome page
J. W. Hong, M. S. Ryu, and I. K. Lim
Phosphorylation of Serine 147 of tis21/BTG2/pc3 by p-Erk1/2 Induces Pin-1 Binding in Cytoplasm and Cell Death
J. Biol. Chem., June 3, 2005; 280(22): 21256 - 21263.
[Abstract] [Full Text] [PDF]


Home page
J Biol ChemHome page
J. D. Joseph, S. N. Daigle, and A. R. Means
PINA Is Essential for Growth and Positively Influences NIMA Function in Aspergillus nidulans
J. Biol. Chem., July 30, 2004; 279(31): 32373 - 32384.
[Abstract] [Full Text] [PDF]


Home page
J Biol ChemHome page
E. Bayer, S. Goettsch, J. W. Mueller, B. Griewel, E. Guiberman, L. M. Mayr, and P. Bayer
Structural Analysis of the Mitotic Regulator hPin1 in Solution: INSIGHTS INTO DOMAIN ARCHITECTURE AND SUBSTRATE BINDING
J. Biol. Chem., July 3, 2003; 278(28): 26183 - 26193.
[Abstract] [Full Text] [PDF]


Home page
J. Cell Sci.Home page
A. Ryo, Y.-C. Liou, K. P. Lu, and G. Wulf
Prolyl isomerase Pin1: a catalyst for oncogenesis and a potential therapeutic target in cancer
J. Cell Sci., March 1, 2003; 116(5): 773 - 783.
[Abstract] [Full Text] [PDF]


Home page
EMBO Rep.Home page
P. E. Shaw
Peptidyl-prolyl isomerases: a new twist to transcription
EMBO Rep., June 1, 2002; 3(6): 521 - 526.
[Abstract] [Full Text] [PDF]


Home page
J Biol ChemHome page
P.-J. Lu, X. Z. Zhou, Y.-C. Liou, J. P. Noel, and K. P. Lu
Critical Role of WW Domain Phosphorylation in Regulating Phosphoserine Binding Activity and Pin1 Function
J. Biol. Chem., January 25, 2002; 277(4): 2381 - 2384.
[Abstract] [Full Text] [PDF]


Home page
Plant Physiol.Home page
J.-L. Magnard, M. Yang, Y.-C. S. Chen, M. Leary, and S. McCormick
The Arabidopsis Gene Tardy Asynchronous Meiosis Is Required for the Normal Pace and Synchrony of Cell Division during Male Meiosis
Plant Physiology, November 1, 2001; 127(3): 1157 - 1166.
[Abstract] [Full Text] [PDF]


Home page
J. Cell Sci.Home page
H.-k. Huang, S. L. Forsburg, U. P. John, M. J. O'Connell, and T. Hunter
Isolation and characterization of the Pin1/Ess1p homologue in Schizosaccharomyces pombe
J. Cell Sci., October 15, 2001; 114(20): 3779 - 3788.
[Abstract] [Full Text] [PDF]


Home page
EMBO J.Home page
G. M. Wulf, A. Ryo, G. G. Wulf, S. W. Lee, T. Niu, V. Petkova, and K. P. Lu
Pin1 is overexpressed in breast cancer and cooperates with Ras signaling in increasing the transcriptional activity of c-Jun towards cyclin D1
EMBO J., July 2, 2001; 20(13): 3459 - 3472.
[Abstract] [Full Text] [PDF]


Home page
Cell Growth Differ.Home page
T. Waerner, P. Gardellin, K. Pfizenmaier, A. Weith, and N. Kraut
Human RERE Is Localized to Nuclear Promyelocytic Leukemia Oncogenic Domains and Enhances Apoptosis
Cell Growth Differ., April 1, 2001; 12(4): 201 - 210.
[Abstract] [Full Text]


Home page
J Biol ChemHome page
I. Landrieu, B. Odaert, J.-M. Wieruszeski, H. Drobecq, P. Rousselot-Pailley, D. Inze, and G. Lippens
p13SUC1 and the WW Domain of PIN1 Bind to the Same Phosphothreonine-Proline Epitope
J. Biol. Chem., January 12, 2001; 276(2): 1434 - 1438.
[Abstract] [Full Text] [PDF]


Home page
J Biol ChemHome page
P.-J. Lu, X. Z. Zhou, Y.-C. Liou, J. P. Noel, and K. P. Lu
Critical Role of WW Domain Phosphorylation in Regulating Phosphoserine Binding Activity and Pin1 Function
J. Biol. Chem., January 25, 2002; 277(4): 2381 - 2384.
[Abstract] [Full Text] [PDF]


Home page
J Biol ChemHome page
I. Landrieu, B. Odaert, J.-M. Wieruszeski, H. Drobecq, P. Rousselot-Pailley, D. Inze, and G. Lippens
p13SUC1 and the WW Domain of PIN1 Bind to the Same Phosphothreonine-Proline Epitope
J. Biol. Chem., January 12, 2001; 276(2): 1434 - 1438.
[Abstract] [Full Text] [PDF]


This Article
Right arrow Abstract Freely available
Right arrow Full Text (PDF)
Right arrow Alert me when this article is cited
Right arrow Alert me if a correction is posted
Services
Right arrow Similar articles in this journal
Right arrow Similar articles in PubMed
Right arrow Alert me to new issues of the journal
Right arrow Download to citation manager
Right arrow
Citing Articles
Right arrow Citing Articles via HighWire
Right arrow Citing Articles via Google Scholar
Google Scholar
Right arrow Articles by Rippmann, J. F.
Right arrow Articles by Schnapp, A.
Right arrow Search for Related Content
PubMed
Right arrow PubMed Citation
Right arrow Articles by Rippmann, J. F.
Right arrow Articles by Schnapp, A.


HOME HELP FEEDBACK SUBSCRIPTIONS ARCHIVE SEARCH TABLE OF CONTENTS
Cancer Research Clinical Cancer Research
Cancer Epidemiology Biomarkers & Prevention Molecular Cancer Therapeutics
Molecular Cancer Research Cell Growth & Differentiation