A novel hPirh2 splicing variant without ubiquitin protein ligase activity interacts with p53 and is down-regulated in hepatocellular carcinoma

A novel hPirh2 splicing variant without ubiquitin protein ligase activity interacts with p53 and is down-regulated in hepatocellular carcinoma

FEBS Letters 584 (2010) 2772–2778 journal homepage: www.FEBSLetters.org A novel hPirh2 splicing variant without ubiquitin protein ligase activity in...

2MB Sizes 0 Downloads 112 Views

FEBS Letters 584 (2010) 2772–2778

journal homepage: www.FEBSLetters.org

A novel hPirh2 splicing variant without ubiquitin protein ligase activity interacts with p53 and is down-regulated in hepatocellular carcinoma Gang Wu a,1, Meiqian Sun a,1, Liping Zhang a, Juan Zhou a, Yi Wang b, Keke Huo a,* a b

State Key Laboratory of Genetic Engineering, Institute of Genetics, School of Life Sciences, Fudan University, Shanghai 200433, China Department of Pathology, Shanghai Tenth People’s Hospital, Tongji University, Shanghai 200072, China

a r t i c l e

i n f o

Article history: Received 6 November 2009 Revised 13 April 2010 Accepted 27 April 2010 Available online 7 May 2010 Edited by Veli-Pekka Lehto Keywords: Pirh2 Splicing variant Hepatocellular carcinoma p53 Ubiquitin ligase

a b s t r a c t A novel splice variant of hPirh2, named hPirh2b, was isolated from human fetal liver cDNA library. hPirh2b has a 38-nucleotide deletion and encodes a 188-amino acid protein with a truncated RINGH2 domain. It shows no ubiquitin protein ligase activity. A low level of expression of hPirh2 was found both at transcriptional and translational level in human hepatocellular carcinoma (HCC) when compared to non-cancerous tissue. Statistical analysis showed that the low expression is associated with lack of differentiation of HCC. In direct binding studies hPirh2b bound p53 indicating that RING-H2 domain is not needed for this interaction. Structured summary: MINT-7889990: NTKLBP1 (uniprotkb: Q5T7V8) physically interacts (MI:0915) with Pirh2b (uniprotkb:Q2KN33) by two-hybrid (MI:0018) MINT-7890105: hPirh2b (uniprotkb:Q2KN33) physically interacts (MI:0915) with p53 (uniprotkb:P04637) by anti tag coimmunoprecipitation (MI:0007) Ó 2010 Federation of European Biochemical Societies. Published by Elsevier B.V. All rights reserved.

1. Introduction Ubiquitin-mediated protein degradation plays an important role in controlling the levels of various cellular proteins and therefore impacts on diverse cellular processes such as cell cycle progression, receptor modulation, transcription regulation and signal transduction. The key components that dictate its substrate specificities are E3 ubiquitin ligases. Most E3 ligases are differentially expressed in cancers. Targeting E3 ubiquitin ligases for cancer therapy has gained increasing attention, largely because of their ability to trigger the degradation of oncogenes or tumour suppressors [1–4]. So E3 ligases represent potential therapeutic targets. Pirh2 is a newly identified RING finger protein with E3 ligase activity. Its notable feature is a capacity to negatively regulate p53 through physical interaction and ubiquitin-mediated proteolysis independent of Mdm2 [5]. It was initially cloned as androgen receptor N-terminal interacting protein [6]. In northern blot analysis in various adult mouse tissues, it occurs as two transcripts of Abbreviations: hPirh2, human p53-induced RING-H2 protein; HCC, human hepatocellular carcinoma; ARNIP, androgen receptor N-terminal interacting protein; TMA, tissue microarray * Corresponding author. Fax: +86 21 55664526. E-mail address: [email protected] (K. Huo). 1 The authors contributed equally to this work.

1.7 and 1.6 kb [5]. Human p53-induced RING-H2 protein (hPirh2) consists of 261 amino acids and has a RING-H2 (C3H2C3) finger domain. It is defined by a consensus sequence with six cysteines and two histidines that coordinate two zinc ions accounts for E3 ligase activity of Pirh2 in vitro and in vivo [6]. Here, we describe the identification and characterization of a novel splicing variant of hPirh2 from human fetal liver cDNA library. We named this novel splicing variant as hPirh2b. The expression pattern of hPirh2 in various human normal tissues, human hepatocellular carcinoma (HCC) and non-cancerous tissues was studied. We also show that hPirh2b, which lacks RING-H2 domain, has no ubiquitin ligase activity. It can interact with p53 in vitro and in vivo. These results will help us to understand the potential role of Pirh2 in the development of HCC. 2. Materials and methods 2.1. Yeast two-hybrid screening The coding sequence of hNTKLBP1 (human N-terminal kinaselike-binding protein 1, NM_152281.2) was cloned into pGBKT7 as a bait and the recombined plasmid was transfected into the yeast strain AH109 followed by the protocol of pACT2 human fetal liver cDNA library (CLONTECH MATCHMAKER Two-Hybrid system 3).

0014-5793/$36.00 Ó 2010 Federation of European Biochemical Societies. Published by Elsevier B.V. All rights reserved. doi:10.1016/j.febslet.2010.04.075

G. Wu et al. / FEBS Letters 584 (2010) 2772–2778

Transformants were selected according to the manufacturer’s instruction. Prey plasmids were isolated and sequenced, followed by homology search using the BLAST algorithm on NCBI web site (http://www.ncbi.nlm.nih.gov). 2.2. Northern blot analysis Northern blot analysis was carried out using Human 8-lane MTN Blot membrane (CLONTECH). The hPirh2 cDNA probe was labeled with [a-32P]dCTP using Random Primer DNA Labeling Kit (Takara). Prehybridization was done at 68 °C for 3 h with herring sperm DNA. Hybridization was at 68 °C overnight followed by washing and autoradiography at 70 °C. 2.3. Quantitative real-time reverse transcription PCR Pieces of HCC tissues were collected from tumors of patients undergoing surgery at Zhong Shan Hospital, Shanghai. Total RNA from normal and tumor tissues was extracted by TRIzol (GibcoBRL) and then treated with amplification grade DNase I (Invitrogen) before cDNA synthesis. Reverse transcription was carried out with Reverse Transcriptase Kit (Takara). qPCR was performed in triplicate using Sybr Green Mastermix on the ABI Prism 7900 Sequence Detection System (Applied Biosystems). GAPDH was used as an endogenous control. qPCR conditions were 95 °C for 15 min, 40 cycles of 95 °C for 15 s and 60 °C for 1 min, followed by the condition for melt curve analysis to confirm the specificity of the amplification product: 95 °C for 15 s, 60 °C for 15 s and 95 °C for 15 s. PCR primer sequences are as follows: hPirh2, 50 -TATGACCAGGTATTGG 50 -CAGTGAAATTCTACGTCCTCCAG-30 (antiAGACAGC-30 (sense), sense); hPirh2b, 50 -GAAATGTTGAAAGAATATGACCAGG-30 (sense), 50 50 -GAACAGTGGATCGTCCATTACAGT-30 (antisense); GAPDH, 0 0 GCACCGTCAAGGCTGAGAAC-3 (sense), 5 -ATGGTGGTGAAGACGCC AGT-30 (antisense). The primers of hPirh2b were designed to anneal the splice junction of exon7 and exon8 of hPirh2 uniquely, ensuring specific amplification of hPirh2b. 2.4. Tissue microarray-based immunohistochemistry The expression of hPirh2 protein in HCC was analyzed using tissue microarray (TMA), which contained 190 pairs of HCC tissue and surrounding liver tissue. Six normal liver tissue samples from hepatic haemangioma were selected as normal controls. All liver tissues were verified by pathological examination. Immunohistochemistry analysis was performed using the Elivision plus twostep System (Dako, Carpinteria, CA) following manufacturer’s instructions. Slides were incubated with primary 1:25 anti-hPirh2 rabbit pAb antibody (Merck). Diaminobenzidine and hematoxylin were used as chromogen and counterstain, respectively. Dark brown granules were taken as positive reaction. 2.5. Glutathione S-transferase (GST) pull-down assay hPirh2 and hPirh2b were cloned into pGEX4T-1 vector (Amersham Pharmacia Biotech) to generate constructs encoding GSThPirh2 and GST-hPirh2b fusion proteins. p53 (72–390aa) was cloned into pGBKT7 which contains a c-Myc epitope tag. Mycp53 fusion protein was produced by the TNT-coupled reticulocyte lysate system (Amersham Pharmacia Biotech) following the manufacturer’s instruction and detected by Western blot with anti-Myc antibody (Invitrogen). For interaction studies, GST-hPirh2 and GST-hPirh2b were allowed bind to glutathione-Sepharose 4B beads which were then incubated with 5 ll Myc-p53 protein in 200 ll NETN buffer (50 mM Tris–HCl, pH 7.5, 200 mM NaCl, 2 mM EDTA, 0.1% NP-40, 1 mM PMSF) for

2773

3 h at 4 °C, respectively. The beads were washed three times by H buffer (20 mM HEPES, 50 mM KCl, 20% glycerol, 0.1% Nonidet P-40, and 0.007% b-mercaptoethanol). Then the pellets were boiled in 20 ll loading buffer, subjected to 15% SDS–PAGE, transferred to PVDF membrane and immunoblotted with antiMyc antibody. 2.6. Co-immunoprecipitation assay in vivo hPirh2b was cloned into pCMV-Myc (Clontech) to produce Myc tagged hPirh2b. Wild-type p53 expression vector (pcDNA3.0-p53) was generously provided by Dr. Long Yu. 2.6  106 HEK293T cells were maintained in DMEM supplemented with 10% bovine calf serum (GIBCO), grown on 60 mm dishes before the day of transfection. 5 lg total DNA of either Myc-hPirh2b and pcDNA3.0-p53 or pCMV-Myc and pcDNA3.0-p53 (in 1:1 molar ratio) was transfected with Lipofectamine 2000 reagent (Invitrogen) following the operation protocol. 48 h later, cells were washed twice with ice-cold PBS and lysed with 1 ml lysis buffer (20 mM Tris, pH 7.5, 150 mM NaCl, 1 mM Na2EDTA, 1 mM EGTA, 10% glycerol, 2.5 mM Na4P2O7, 1% Triton, 1 mM Na3VO4, 1 mM b-glycerophosphate, 10 lg/ml aprotinin, 10 lg/ml leupeptin, and 1 mM PMSF). Lysate supernatant was incubated with EZview™ Red Anti-c-Myc Affinity Gel (Sigma) and co-immunoprecipiation was done according to the operation manual. Samples were resolved by 12% SDS– PAGE and subjected to Western blotting using anti-p53 antibody (Calbiochem). 2.7. Ubiquitination assay in vitro The ubiquitination assay was performed according to a previously published report [7] with modifications. GST-hPirh2 or GST-hPirh2b immobilized to the glutathione-Sepharose 4B beads was combined with rabbit E1 (40 ng, Calbiochem), UbcH5b (100 ng, Calbiochem), His-tagged ubiquitin (2 lg, Calbiochem) in a ubiquitination buffer (50 mM Tris–HCl (pH 7.4), 2 mM ATP, 5 mM MgCl2, 2 mM DTT, 30 mM creatine phosphate, 0.05 mg/ml creatine phosphokinase and BL-21 cell lysates), and the mixture was incubated at 30 °C for 1.5–2 h. The reaction was terminated by adding loading buffer and processed for 12% SDS–PAGE. Western blot experiments were done using the anti-His antibody (Invitrogen). 2.8. Subcellular localization Hela cells were cultured in DMEM, supplemented with 10% fetal bovine serum (Life Technologies) under the condition of 5% CO2 at 37 °C. hPirh2 and hPirh2b were cloned into the pEGFP-C1 plasmid (Clontech) to generate the expression constructs pEGFP-C1-hPirh2 and pEGFP-C1-hPirh2b in which GFP is tagged to the N-terminus of hPirh2 and hPirh2b. All transfections were performed using Lipofectamine 2000 reagent. After 48 hrs, fluorescent image analysis was performed using confocal laser scanning microscopy (Zeiss LSM 510 META, Axiovert 200, Germany). 2.9. Statistical analysis t-test was used to analyze the relationship between the expression of hPirh2 as determined by qPCR and the differentiation of HCC. For the result of TMA, chi-square test was applied to assess the differences of hPirh2 protein expression between HCC and the surrounding liver tissue. A probability level of 0.05 was chosen for statistical significance. Statistical analysis was performed with the SPSS12.0 software package (Chicago, IL, USA).

2774

G. Wu et al. / FEBS Letters 584 (2010) 2772–2778

3. Results 3.1. Isolation of hPirh2b from fetal liver cDNA library and sequence characterization of hpirh2b Apart from hPirh2 identified as the interaction protein of hNTKLBP1 by yeast two-hybrid in [8], a cDNA of 1320 nt was selected for closer scrutiny. Sequencing revealed an open reading frame of 564 nucleotides corresponding to the nt 27–590 in the original contig (GenBank accession number AY888047) (Fig. 1). From the nucleotide sequence of 564 nt ORF, a protein sequence of 188 amino was deduced, with a calculated molecular mass of 20.7 kDa. Comparison with hPirh2 shows 96% identity. SMART analysis (http://smart.embl-heidelberg.de/) revealed a Znf_CHY domain. Furthermore, when compared with hPirh2, a deletion of 38 nucleotides was found in hPirh2b, corresponding to the RING finger domain in Pirh2.

Search against the genome database revealed that hPirh2b was represented by a genomic contig (NT_086651) in chromosome 4q21. The coding region of hPirh2b can be divided into eight exons through alignment of the hPirh2b cDNA with this genomic contig (Fig. 2). All the putative exon–intron junctions conform to the GT/AG rule (Table 1). Compared with another splicing variant of hPirh2 (accession number NM_015436), hPirh2b uses a different 50 splice site in exon 8. This results in a 38-nucleotide deletion in 50 terminus of exon8 and creates a premature termination codon in exon 8. Thus, the exons 6 and 7 and parts of exons 5 and 8, encoding RING-H2 domain in Pirh2 are spliced out in hPirh2b. 3.2. Expression of hPirh2 in normal and HCC/non-cancerous tissue The expression of hPirh2 in normal tissue investigated by Northern blot (Fig. 3A). Highest level of expression was found in heart and skeletal muscles, followed by liver and pancreas. The expres-

Fig. 1. The nucleotide and deduced amino acid sequences of hPirh2b. Start and stop codons are marked in bold italics type. The polyadenylation signal is indicated by black box. Two dark shadings highlight a confidently predicted Znf_CHY domain (aa20–94) and a truncated RING finger domain (aa145–178).

2775

G. Wu et al. / FEBS Letters 584 (2010) 2772–2778

Fig. 2. The schematic representation of two splicing variants of hPirh2 and comparison of the RING-H2 domains. (A) hPirh2 and hPirh2b have 9 and 8 coding exons, respectively. (B) The conserved cysteine and histidine residues in RING-H2 domain (C3H2C3) are indicated by black box. In the truncated RING-H2 domain of hPirh2b, the seventh and eighth cysteines are lost.

Table 1 Exon–intron boundary sequence of hPirh2b. Exon

cDNA position

Number

Length (bp)

Position at NT_086651

Gene

1 2 3 4 5 6 7 8

90 120 116 79 45 59 27 28

23511578–23511667 23506538–23506657 23506181–23506296 23491719–23491797 23489507–23489551 23489342–23489400 23489229–23489255 23488254–23488281

27-116 117-236 237-352 353-431 432-476 477-535 536-562 563-590

Splicing acceptor

– tgggtgttagGCACCTTGCT tcttttttagGCCCAACAGA gttttgctagGATTGGTCCA ttttaaacagTGTATTGAAA tcatttacagGACATTCACA tttctttcagAACGTGTTAT tctgctttagATATGACCAG

Splicing donor

TCTCCTAAAGgtgacgcctt AATTCAACATgtaagatttt GAATTTGTAGgtactgtagg AAGACACAAGgtatataatt ATGTTTGGAGgtaggcttaa TTTTACATAGgtaaagtaac TGTTGAAAGAgtgagtgttt –

Intron Number

Length (bp)

1 2 3 4 5 6 7

– 4920 241 14383 2167 106 86 947

Intron and exon nucleotide sequence are shown in lowercase and uppercase letters, respectively. Bold italics lettering indicates donor and acceptor splice site. Not including 50 or 30 UTRs or stop codon.

sion of hPirh2 in HCC/non-cancerous tissues was also detected but only at a low level. By qPCR analysis of 79 pairs of HCC/non-cancerous tissues, we found that the level of expression of hPirh2 was lower in 66% cases (52) when compared with the corresponding normal liver. hPirh2b was found at lower level of expression in 58% of cases when compared with the normal tissue (Fig. 3B). Statistically, a significant correlation could be seen between a low level of expression of hPirh2 (when compared with normal tissue) and the lack of differentiation of HCC (P < 0.05 t-test). However, there was no association between the differential expression of hPirh2b and the differentiation grade of HCC. Also TMA-based immunohistochemistry revealed a low level of expression of hPirh2 in HCC when compared with normal tissue; 39% positive (74 out of 190) in HCCs; 73% positive1 (39 out of 190 tissues) in normal tissue; P = 1.839E-11) (Fig. 3C and D). 3.3. hPirh2b can interact with p53 in vitro and in vivo GST pull-down assay was applied to tell whether or not hPirh2b can interact with p53. GST-hPirh2 and GST-hPirh2b fusion proteins were expressed in E. coli strain BL-21. The calculated molecular weights of the expressed fusion proteins are 58 and 49 kDa, corre-

spondingly, as determined by 12% SDS–PAGE (Fig. 4A). Immunoblotting with ant-Myc antibodies showed that the myc-p53 fusion protein was bound to GST-hPirh2 and GST-hPirh2b fusion proteins but not GST alone (Fig. 4B). To further confirm the interaction between hPirh2b and p53, co-immunoprecipitation experiment was carried out. We found that p53 was present in the Myc-hPirh2b immunoprecipitates but not in the pCMV-Myc immunoprecipitates (Fig. 4C). 3.4. hPirh2b has no ubiquitin protein ligase activity An in vitro ubiquitination assay system was used to determine whether hPirh2b can mediate protein ubiquitination. Affinity purified GST-hPirh2, GST-hPirh2b and GST were added to bacterial extracts containing E1, E2 (UbcH5b) and His-tagged ubiquitin, respectively. After the ubiquitination reaction, the samples were subjected to SDS–PAGE and immunoblotted with His antibody to detect ubiquitinated products. Positive immunoreaction was seen only in lane 1 indicating that GST-hPirh2 fusion protein but not GST-hPirh2b nor GST can catalyze the autoubiquitination activity (Fig. 5). Thus, it can be concluded that hPirh2b is devoid of ligase activity. This is most likely due to its lack of the RING-H2 domain.

2776

G. Wu et al. / FEBS Letters 584 (2010) 2772–2778

Fig. 3. (A) Expression patterns of hPirh2 in various human normal tissues. Three hPirh2 transcripts of about 1.6, 1.7 and 4.4 kb were detected. (B) qPCR results of hPirh2 and hPirh2b were normalized and converted into ratio against the expression level of GAPDH. (C and D) Normal liver tissue and HCC was immunostained with anti-Pirh2 antibody respectively. Positive staining is indicated as brown color.

Fig. 4. Interaction of hPirh2 and hPirh2b with p53 in vitro and in vivo. (A) GST-hPirh2, GST-hPirh2b and GST were expressed and purified respectively. (B) p53 binds to hPirh2 and hPirh2b, but not to GST alone. The translation product of p53 in vitro was used as a control. (C) hPirh2b interacts with p53 in vivo. Myc-hPirh2b and p53 were cotransfected into HEK293T cells. Myc tagged vector and p53 were co-transfected as control. Lysate was immunoprecipitated with anti-Myc antibody and analyzed by immunoblotting with anti-p53 antibody.

3.5. Subcellular localization of hPirh2 isoforms In an attempt to identify whether the truncated RING-H2 domain will influence the subcellular localization of hPirh2, we produced expression constructs in which GFP is tagged to the N-terminus of hPirh2 and hPirh2b. GFP-hPirh2 exhibited a

diffuse nuclear and cytoplasmic distribution which is in agreement with previous reports (data not shown) [9–11]. However, there is no significant difference about the subcellular localization of hPirh2b from hPirh2. hPirh2b homogeneously distributed throughout the cytoplasm and the nuclei in Hela cells (Fig. 6).

G. Wu et al. / FEBS Letters 584 (2010) 2772–2778

Fig. 5. hPirh2b has no ubiquitin protein ligase activity. GST-hPirh2, GST-hPirh2b and GST were evaluated for E3 activity in the presence of E1, E2, and His-tagged Ub as indicated. No ubiquitinated products were detected in the presence of GST or GST-hPirh2b.

4. Discussion Here we report cloning and characterization of a novel splice variant of hPirh2. Recently, three other isoforms of hPirh2 were identified in [11,12]. Common to these variant forms is that they show alternative splicing in the RING-H2 domain [11]. Alternative pre-mRNA splicing is a sophisticated and ubiquitous nuclear process which plays an essential role in the regulation of many biologic processes. Increasing amounts of evidence strongly suggest that the alternative pre-mRNA splicing of some E3 ligases occurs frequently in human cancer cells [13,14]. Pirh2 was originally identified as a p53-inducible and p53interacting protein which promotes p53 ubiquitination and degradation [5]. It had been found to be expressed higher in lung cancer, prostatic carcinoma, head and neck cancer [10,15,16]. During the preparation of this manuscript, Wang et al. found that the increased expression levels of Pirh2 were correlated with a poor prognosis in HCC patients [17]. That suggested Pirh2 could be as a potential drug target and a prognostic biomarker of cancers. However, further studies are still needed to explore the potential role which hPirh2b plays in the genesis and progression of HCC.

2777

Given Pirh2 has been shown to be one of the interaction protein of p53 [5], we next sought to investigate the relationship between hPirh2b and p53. Although there is a RING domain deletion in hPirh2b, the interaction between hPirh2b and p53 have been found by GST pull-down assay and co-immunoprecipitation, which suggested that the intact RING-H2 domain is not necessary for Pirh2–p53 interaction. RING-H2 is commonly found in ubiquitin ligases. Thus, it was pertinent to test whether Pirh2b, analogous to Pirh2, has ubiquitin ligase activity. Ubiquitin assay failed to show any ligase activity pointing out to the central role of RING-H2 in this activity. Interestingly, a RING-less mutant of mPirh2 and another two hPirh2 RING-H2 domain variants retain the ability to repress p53-mediated transcriptional activity [5,11], a functional feature of intact Pirh2. Thus, Pirh2 is able to affect the transactivation function of p53 by using some other mechanism than ubiquitination followed by degradation. The present results suggest that hPirh2b utilizes similar, ubiquitin-independent mechanisms in its functional interaction with p53. Its differential expression in various tissues and tumors may be a reflection of its differential role in p53-mediated tumorigenesis. In some studies, presence or absence of RING domain has been found to be associated with altered subcellular distribution of some ubiquitin protein ligases with putative pathogenetic significance [18]. The nuclear and nucleolar subcellular localization play an essential role in the function and stability of Pirh2 [9]. No significant difference was found in the subcellular localizations of hPirh2b from hPirh2. Thus, presence or absence of RING-H2 does not seem to be associated with the altered subcellular distribution at least at the light microscopic level. In previous reports, mPirh2 transcripts of 1.6 and 1.7 kb were found in various adult mouse tissues [5,6]. In addition, a minor 1.3 kb transcript was only detected in rat PC-12 (rat adrenal pheochromocytoma) cells and mouse TM4 (mouse Sertoli cells) cells [6]. These results suggest that Pirh2 isoforms are expressed differentially among tissues. To date, more than 40 different splicing variants of Mdm2 transcripts have been identified both in tumors and normal tissues. Most of them are associated with tumor progression and prognosis [19–21]. Due to the similar type of functional partnership of Mdm2 and Pirh2 in relation to p53, it can be surmised that the tissue variable Pirh2 isoforms could provide tissue specific regulatory pathways underlying some specific tumorigenic pathways. Acknowledgments This work was supported by the Chinese 863 program (Grant No. 2006AA02A310), CNHLPP program (Grant No. 2004BA711A19) and Shanghai Science and Technology Developing Program (Grant No. 03DZ14024) to K. Huo.

Fig. 6. Subcellular localization of GFP-tagged hPirh2b. hPirh2b homogeneously distributed throughout the cytoplasm and the nuclei of Hela cells. Nuclei were stained by DAPI.

2778

G. Wu et al. / FEBS Letters 584 (2010) 2772–2778

References [1] Sun, Y. (2003) Targeting E3 ubiquitin ligases for cancer therapy. Cancer Biol. Ther. 2, 623–629. [2] Robinson, P.A. and Ardley, H.C. (2004) Ubiquitin-protein ligases – novel therapeutic targets? Curr. Protein Pept. Sci. 5, 163–176. [3] Hoeller, D., Hecker, C.M. and Dikic, I. (2006) Ubiquitin and ubiquitin-like proteins in cancer pathogenesis. Nat. Rev. Cancer 6, 776–788. [4] Bernassola, F., Karin, M., Ciechanover, A. and Melino, G. (2008) The HECT family of E3 ubiquitin ligases: multiple players in cancer development. Cancer Cell 14, 10–21. [5] Leng, R.P. et al. (2003) Pirh2, a p53-induced ubiquitin-protein ligase, promotes p53 degradation. Cell 112, 779–791. [6] Beitel, L.K., Elhaji, Y.A., Lumbroso, R., Wing, S.S., Panet-Raymond, V., Gottlieb, B., Pinsky, L. and Trifiro, M.A. (2002) Cloning and characterization of an androgen receptor N-terminal-interacting protein with ubiquitin-protein ligase activity. J. Mol. Endocrinol. 29, 41–60. [7] Lorick, K.L., Jensen, J.P., Fang, S., Ong, A.M., Hatakeyama, S. and Weissman, A.M. (1999) RING fingers mediate ubiquitin-conjugating enzyme (E2)-dependent ubiquitination. Proc. Natl. Acad. Sci. USA 96, 11364–11369. [8] Zhang, L., Li, J., Wang, C., Ma, Y. and Huo, K. (2005) A new human gene hNTKLBP1 interacts with hPirh2. Biochem. Biophys. Res. Commun. 330, 293–297. [9] Logan, I.R., Sapountzi, V., Gaughan, L., Neal, D.E. and Robson, C.N. (2004) Control of human PIRH2 protein stability: involvement of TIP60 and the proteosome. J. Biol. Chem. 279, 11696–11704. [10] Logan, I.R., Gaughan, L., McCracken, S.R., Sapountzi, V., Leung, H.Y. and Robson, C.N. (2006) Human PIRH2 enhances androgen receptor signaling through inhibition of histone deacetylase 1 and is overexpressed in prostate cancer. Mol. Cell. Biol. 26, 6502–6510. [11] Corcoran, C.A., Montalbano, J., Sun, H., He, Q., Huang, Y. and Sheikh, M.S. (2009) Identification and characterization of two novel isoforms of Pirh2

[12] [13] [14]

[15]

[16]

[17]

[18]

[19]

[20] [21]

ubiquitin ligase that negatively regulate p53 independent of RING finger domains. J. Biol. Chem. 284, 21955–21970. Shi, J., Huang, Y. and Sheikh, M.S. (2010) Identification of Pirh2D, an additional novel isoform of Pirh2 ubiquitin ligase. Mol. Cell Pharmacol. 2, 21–23. Venables, J.P. (2004) Aberrant and alternative splicing in cancer. Cancer Res. 64, 7647–7654. Wang, Z., Lo, H.S., Yang, H., Gere, S., Hu, Y., Buetow, K.H. and Lee, M.P. (2003) Computational analysis and experimental validation of tumor-associated alternative RNA splicing in human cancer. Cancer Res. 63, 655–657. Duan, W., Gao, L., Druhan, L.J., Zhu, W.G., Morrison, C., Otterson, G.A. and Villalona-Calero, M.A. (2004) Expression of Pirh2, a newly identified ubiquitin protein ligase, in lung cancer. J. Natl. Cancer Inst. 96, 1718–1721. Shimada, M. et al. (2009) High expression of Pirh2, an E3 ligase for p27, is associated with low expression of p27 and poor prognosis in head and neck cancers. Cancer Sci. 100, 866–872. Wang, X.M., Yang, L.Y., Guo, L., Fan, C. and Wu, F. (2009) P53-induced RING-H2 protein, a novel marker for poor survival in hepatocellular carcinoma after hepatic resection. Cancer 115, 4554–4563. Cookson, M.R., Lockhart, P.J., McLendon, C., O’Farrell, C., Schlossmacher, M. and Farrer, M.J. (2003) RING finger 1 mutations in Parkin produce altered localization of the protein. Hum. Mol. Genet. 12, 2957–2965. Lukas, J., Gao, D.Q., Keshmeshian, M., Wen, W.H., Tsao-Wei, D., Rosenberg, S. and Press, M.F. (2001) Alternative and aberrant messenger RNA splicing of the mdm2 oncogene in invasive breast cancer. Cancer Res. 61, 3212– 3219. Bartel, F., Taubert, H. and Harris, L.C. (2002) Alternative and aberrant splicing of MDM2 mRNA in human cancer. Cancer Cell 2, 9–15. Sigalas, I., Calvert, A.H., Anderson, J.J., Neal, D.E. and Lunec, J. (1996) Alternatively spliced mdm2 transcripts with loss of p53 binding domain sequences: transforming ability and frequent detection in human cancer. Nat. Med. 2, 912–917.