MPHOSPH1 is a critical kinesin protein that functions in cytokinesis. Here, we show that MPHOSPH1 is overexpressed in hepatocellular carcinoma (HCC) cells, where it is essential for proliferation. Attenuating MPHOSPH1 expression with a tumor-selective shRNA-expressing adenovirus (Ad-shMPP1) was sufficient to arrest HCC cell proliferation in a manner associated with an accumulation of multinucleated polyploid cells, induction of postmitotic apoptosis, and increased sensitivity to taxol cytotoxicity. Mechanistic investigations showed that attenuation of MPHOSPH1 stabilized p53, blocked STAT3 phosphorylation, and prolonged mitotic arrest. In a mouse subcutaneous xenograft model of HCC, tumoral injection of Ad-shMPP1 inhibited MPHOSPH1 expression and tumor growth in a manner correlated with induction of apoptosis. Combining Ad-shMPP1 injection with taxol administration enhanced antitumor efficacy relative to taxol alone. Furthermore, Ad-shMPP1 tail vein injection suppressed formation of orthotopic liver nodules and prevented hepatic dysfunction. Taken together, our results identify MPHOSPH1 as an oncogenic driver and candidate therapeutic target in HCC. Cancer Res; 74(22); 6623–34. ©2014 AACR.

Kinesin superfamily (KIF) proteins are a group of proteins with a highly conserved motor domain, some of which move toward the plus end of microtubules in an ATP-dependent process reliant on their adenosine triphosphatase (ATPase) activity (1). KIF proteins participate in many essential cellular biologic activities including mitosis, meiosis, and the transport of macromolecules (2). Recently, there has been increasing evidence of aberrant expression of KIF proteins in a variety of cancers, suggesting the oncogenic potential of KIFs in addition to their normal cellular physiologic functions (3, 4). Among these KIFs, M-phase phosphoprotein 1 (MPHOSPH1, also referred to as KIF20B) was reported to be a plus-end-directed kinesin protein, playing a critical role in completion of cytokinesis in late telophase of mitosis (5). Recently, an oncogenic role for MPHOSPH1 in bladder and colorectal cancer cells was reported (4, 6).

Hepatocellular carcinoma (HCC) accounts for 90% of human primary liver cancers, which is the fifth most common cancer and the third most common cause of cancer-related death, particularly in East Asia. Advanced HCC often has a poor prognosis and only few chemotherapeutic drugs, such as sorafenib, demonstrate efficacy in increasing overall survival in advanced or metastatic HCC (7). However, due to the rapid development of drug resistance, the effective course of chemotherapy for HCC often lasts for only a few months (8). Resistance to chemotherapeutic agents, therefore, remains a major challenge to HCC therapy. Recently, KIF proteins have been shown to play critical roles in the development of resistance to antimitotic drugs, such as taxanes (7, 9). Combined chemotherapy targeting kinesins may thus represent a promising anticancer strategy, especially in drug-resistant solid tumors.

Another challenge in treating HCC is reduced hepatic function. Poor hepatic function is common in HCC, resulting from a variety of causes, such as hepatic ascites and malnutrition. In many cases, hepatic dysfunction results in a bad prognosis for patients with HCC. Therefore, maintaining hepatic function is critical for HCC therapy. For this reason, improving liver function of patients with HCC is usually initiated before carrying out further treatment, especially surgery. Consequently, due to the cytotoxicity of chemotherapy, an improved, combined strategy that prevents hepatic injury is urgently needed.

In the present study, we found that MPHOSPH1 expression is increased in HCC tissues compared with nontumor tissues, and that knockdown of endogenous MPHOSPH1 using an adenoviral vector inhibited HCC cell proliferation by triggering mitotic arrest and apoptosis. We also demonstrated that combining taxol treatment with MPHOSPH1 interference resulted in synergistic repression of HCC cell viability. In addition, we demonstrated that Ad-shMPP1 treatment not only suppressed HCC development, but also significantly reduced hepatic dysfunction caused by orthotopic HCC xenografts, suggesting that use of Ad-shMPP1 produced a dual effect; anti-HCC activity and protection of liver function in vivo.

Cell lines, animals, and tissue specimens

The human liver cell lines L02 and QSG7701 and the human HCC cell lines Hep3B, HepG2, and BEL-7404 were obtained from the Cell Bank of Type Culture Collection of the Chinese Academy of Sciences (Shanghai, China) or China Center for Type Culture Collection (Wuhan, China) within 6 months of the experiments carried out. All cell lines were authenticated by short tandem repeat profiling and were confirmed to be mycoplasma-negative before use. Cell lines were grown in DMEM (GIBCO BRL) supplemented with 10% FBS. The human samples including ten nonmalignant liver diseases (cirrhosis, choledocholithiasis, and cavernous hemangioma) and 19 HCC and four adjacent tissues were obtained from Wuhan Union Hospital, Tongji Medical College of Huazhong University of Science and Technology (Wuhan, China). The Institutional Review Board of Wuhan Union Hospital approved acquisition of tissue specimens and collection of human samples was in accordance with the established guidelines.

IHC assays

Clinical human tissues or the HCC xenograft samples were fixed in 4% formaldehyde, dehydrated with an ethanol gradient, and embedded in paraffin. Tissue sections were dewaxed and rehydrated according to a standard protocol. The sections were washed with PBS, treated with 3% H2O2, and blocked with a blocking solution. These steps were followed by overnight incubation with primary antibody at the proper dilution for IHC or hematoxylin and eosin (H&E) staining assays as previously described (10). To avoid experimental bias, the clinical human samples were scored from 0 to 3, based on the IHC staining of MPHOSPH1, by two people using a double-blind procedure.

Recombinant adenovirus

Recombinant adenoviruses (rec-Ads) were generated by respective homologous recombination between the shuttle plasmids and the packaging plasmid pBHGE3 in HEK293 cells. Generation, identification, production, purification, and titration of the recombinant adenovirus were performed as previously described (6).

RNA extraction and quantitative reverse transcriptase PCR

Total RNA was isolated from cells using TRIzol (Invitrogen). A total of 2 μg of RNA was used to synthesize the first single-strand cDNA using the RevertAid First Strand cDNA Synthesis Kit (Fermentas). The relative quantification of MPHOSPH1 cDNA and adenovirus gene E3 cDNA was performed using an ABI 7500 Fast System or a Bio-Rad CFX96 Real-Time PCR Detection System with Power SYBR Green PCR Master Mix (Applied Biosystems). GAPDH cDNA was used as an internal standard. The primers used are provided in Supplementary Materials.

Cytotoxicity assay

HCC cells and normal cells were seeded in 24-well plates at a density of 20,000 cells per well. Twenty-four hours later, cells were infected with Ads at various multiplicity of infections (MOI). Four days after infection, the medium was removed, and the cells were exposed to 2% crystal violet in 20% methanol for 20 minutes. The plates were then washed with water and photographed.

Cell proliferation assays

Hep3B and HepG2 cells were seeded in 96-well plates at a density of 2,000 cells per well. For the combined treatment (Ads and taxol), cells were infected with Ads at an MOI of 2. Twenty-four hours after infection, taxol (Sigma-Aldrich) at a final concentration of 10 μg/mL was added to each well. Of note, 20 μL of 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT; 5 mg/mL) was added to each well at 48 hours later, and the plates were incubated at 37°C for 4 hours. The medium was removed and 150 μL of DMSO was added to each well. Absorbance at 490 nm was measured using a Microplate Reader (Thermo). Six replicate wells were counted for each assay.

In vitro combination studies

Synergism was determined using the software package Calcusyn (Biosoft). The combination index (CI) was calculated by the Chou–Talalay equation (11). A CI less than 0.9 was defined as synergism. The detailed description on the combination is provided in the Supplementary Materials.

Cell nuclei and cytoskeleton staining assay

Cells were seeded on glass cover slips in 6-well plates. Twenty-four hours later, cells were infected with Ads at an MOI of 2. Cells were then fixed with 0.5 mL of 4% paraformaldehyde for 10 minutes, washed twice with PBS, stained with Phalloidin Alexa Fluor 555 (Molecular Probes), washed twice with PBS, stained with DAPI (Sigma-Aldrich), and washed twice with PBS. Cell staining was analyzed using a BX51 upright microscope (Olympus) or a Leica TCS SP2 confocal microscope (Leica Microsystems).

DNA fragmentation assays

HepG2 cells were seeded in 10 cm culture dishes and infected with various Ads at MOI = 5. 72 hours postinfection, DNA was extracted and analyzed using agarose gel (12).

Flow-cytometric analyses

Cells fixed with 70% ethanol were incubated for 2 hours with an RNase A (40 μg/mL; Promega)/propidium iodide solution (50 μg/mL; Sigma-Aldrich). FACS was performed on a FACSCalibur flow cytometer (BD Biosciences).

SA-β-gal staining assay

β-galactosidase staining was performed using a senescence-associated β-Galactosidase Staining Kit (Beyotime). Cells were washed three times with PBS and fixed with 4% paraformaldehyde for 15 minutes at room temperature. Next, the cells were incubated overnight at 37°C in the dark with a solution containing 0.05 mg/mL X-gal.

Western blot assays

Cells were harvested and washed with PBS, and then lysed for analysis of proteins. Western blot analysis was performed as previously reported (6). Detailed antibody information is provided in the Supplementary Materials.

Animal experiments

Animal experiments were performed according to the Guide for the Care and Use of Laboratory Animals set forth by the Huazhong University of Science and Technology. Male BALB/c nude mice (4-week-old) were purchased from the Beijing HFK Bioscience Co. Ltd. Details of the experiments are provided in the Supplementary Materials.

Statistical analyses

All data were expressed as the mean ± SD and were analyzed using independent sample t tests and ANOVA using SPSS Base 10.0. Results were considered statistically significant when P < 0.05.

Overexpression of MPHOSPH1 in HCC tissues

MPHOSPH1 was previously classified as an oncogene based on studies of bladder cancer as well as in colorectal and lung cancer cell lines (4, 6). To determine whether MPHOSPH1 is upregulated in HCC, IHC detection of MPHOSPH1 was performed in samples from 19 patients with HCC (including four samples of normal adjacent tissues) and 10 nonmalignant tissues (Fig. 1A and Supplementary Table S1). Two people scored each IHC assay of all samples on a 0–3 point scale based on the MPHOSPH1 staining ranges and color degrees as shown in Fig. 1B and Supplementary Table S1 (Fig. 1B shows examples of MPHOSPH1 scoring 0–3). An IHC score of ≥2 is considered overexpression of MPHOSPH1. Approximately 84% (16 of 19) of HCC tissues demonstrated overexpressed MPHOSPH1, whereas, in contrast, only 14% (3 of 14) of nonmalignant liver tissues showed overexpression of MPHOSPH1 (Supplementary Table S1). Further statistical analysis showed a significant difference in the MPHOSPH1 expression level between HCC tissues and non-HCC tissues (Fig. 1C). In addition, although the HCC samples exhibited a range of Edmondson–Steiner grading (an index reflecting the differentiation degrees of HCC cells), we found no statistically significant correlation between the differentiation state of the HCC cells and the MPHOSPH1 expression level (Supplementary Table S1).

Figure 1.

Overexpression of MPHOSPH1 in HCC patient samples and human HCC cell lines. A, IHC was conducted on normal liver and HCC tissues for MPHOSPH1 protein expression. Magnification, ×20. B, intensity of MPHOSPH1 staining was scored on a scale of 0–3. C, MPHOSPH1 mRNA was measured by RT-PCR in three liver cancer cell lines and a normal liver cell line (***, P < 0.001).

Figure 1.

Overexpression of MPHOSPH1 in HCC patient samples and human HCC cell lines. A, IHC was conducted on normal liver and HCC tissues for MPHOSPH1 protein expression. Magnification, ×20. B, intensity of MPHOSPH1 staining was scored on a scale of 0–3. C, MPHOSPH1 mRNA was measured by RT-PCR in three liver cancer cell lines and a normal liver cell line (***, P < 0.001).

Close modal

The MPHOSPH1 mRNA levels in three HCC cell lines were also compared by qRT-PCR. The HCC cell lines HepG2, Hep3B, and BEL-7404 all showed elevated mRNA levels of MPHOSPH1 compared with that of the normal liver cell line L02 (Fig. 1D), suggesting that both MPHOSPH1 protein and mRNA levels may be used as potential biomarkers for HCC.

HCC-specific replication and MPHOSPH1 knockdown effects of the oncolytic recombinant adenovirus

We previously constructed a recombinant adenoviral vector based on the oncolytic Ad in which the viral E1B55KD gene was deleted (6). This Ad-vector has a tumor-specific survivin promoter controlling the expression of the adenoviral early gene E1A (Fig. 2A), thus restricting replication to tumor cells (6). To verify its ability to replicate in HCC cells, we infected HCC and normal liver cells with the Ad-vector at an MOI = 0.1. The ability to replicate was evaluated by assessing the adenoviral E3 gene mRNA level. Ad-vector replication in HepG2 cells was at least 100-fold higher than in the normal QSG7701 cells at 48 hours after infection (Fig. 2B), suggesting that it is a potent HCC-targeting oncolytic virus.

Figure 2.

HCC-specific replication and MPHOSPH1 knockdown effects of Ad-shMPP1 A, a schematic drawing of the Ad-shMPP1. Ψ, the encapsidation signal; ITR, the inverted terminal repeat. B, analysis of the replication ability of Ad-shMPP1 in the normal liver cell (QSG7701) and the HCC cell (HepG2) by RT-PCR. The adenovirus E3 region was amplified to estimate the viral DNA content 48 hours after infection (MOI = 0.1). C, quantification of the MPHOSPH1 transcript level by RT-PCR in HepG2 and Hep3B cells 48 hours after infection with Ad-vector and Ad-shMPP1 at MOI = 1. GAPDH was used as an internal control.

Figure 2.

HCC-specific replication and MPHOSPH1 knockdown effects of Ad-shMPP1 A, a schematic drawing of the Ad-shMPP1. Ψ, the encapsidation signal; ITR, the inverted terminal repeat. B, analysis of the replication ability of Ad-shMPP1 in the normal liver cell (QSG7701) and the HCC cell (HepG2) by RT-PCR. The adenovirus E3 region was amplified to estimate the viral DNA content 48 hours after infection (MOI = 0.1). C, quantification of the MPHOSPH1 transcript level by RT-PCR in HepG2 and Hep3B cells 48 hours after infection with Ad-vector and Ad-shMPP1 at MOI = 1. GAPDH was used as an internal control.

Close modal

To construct shRNA-expressing Ad, an MPHOSPH1-targeted shRNA (shMPP1) expression plasmid was used (6). We cloned the shMPP1 expression cassette into the oncolytic vector to generate Ad-shMPP1 (Fig. 2A). As expected, this tumor-specific oncolytic Ad-shMPP1 displayed potent interference ability against endogenous MPHOSPH1 expression in HCC cells. The relative mRNA level of MPHOSPH1 in HepG2 and Hep3B cells was significantly reduced following Ad-shMPP1 infection (95% and 60% downregulation at an MOI of 2 respectively; Fig. 2C).

Ad-shMPP1 inhibits HCC cell growth and potentiates the efficacy of taxol

To investigate the cytopathic effect (CPE) and the potential anti-HCC activity induced by Ad-shMPP1, different liver cancer cell lines and a normal liver cell line were infected with Ad-vector and Ad-shMPP1 at various MOIs. Ninety-six hours later, CPEs were determined by crystal violet staining. As shown in Fig. 3A, in normal liver L02 cells, infection with Ad-shMPP1 at an MOI of 5 did not cause a CPE, whereas in HepG2, Hep3B and BEL-7404 cells, treatment at the same MOI (in HepG2 cells, even at a lower MOI of 2, Fig. 3A) caused obvious CPE, demonstrating HCC cell-selective cytotoxicity and the relative safety of Ad-shMPP1.

Figure 3.

Antiproliferation effects of Ad-shMPP1 in HepG2 cells. A, the CPEs of Ads on HCC cells and normal liver cells at varied MOIs were determined by crystal violet staining 96 hours after infection. B, measurement of cell viability of HepG2 and Hep3B HCC cells 96 hours after infection with Ads or combined treatment with taxol using an MTT assay (**, P < 0.01; ***, P < 0.001). C, HepG2 and Hep3B cells were treated with Ad-shMPP1 or taxol or both in a series of concentrations and the combination index was calculated using CalcuSyn software. CI values less than 0.9 are considered synergism.

Figure 3.

Antiproliferation effects of Ad-shMPP1 in HepG2 cells. A, the CPEs of Ads on HCC cells and normal liver cells at varied MOIs were determined by crystal violet staining 96 hours after infection. B, measurement of cell viability of HepG2 and Hep3B HCC cells 96 hours after infection with Ads or combined treatment with taxol using an MTT assay (**, P < 0.01; ***, P < 0.001). C, HepG2 and Hep3B cells were treated with Ad-shMPP1 or taxol or both in a series of concentrations and the combination index was calculated using CalcuSyn software. CI values less than 0.9 are considered synergism.

Close modal

Some KIF proteins seem to be involved in the development of drug resistance by HCC to some microtubule-targeted agents, such as taxol (13). Thus, to further clarify whether Ad-shMPP1 could enhance the anti-HCC effects of taxol, 24 hours before taxol treatment, we infected HepG2 and Hep3B cells with Ad-shMPP1 at an MOI of 2, and 48 hours later, the cell viability was determined using an MTT assay. Ad-shMPP1 produced a significant cell growth arrest effect on HepG2 and Hep3B cells compared with Ad-vector control groups (Fig. 3B). More importantly, compared with the Ad or taxol mono-treatment group, the combined treatment amplified the anti-HCC effect (P < 0.01), indicating significantly enhanced cytotoxicity of taxol (Fig. 3B). To further investigate whether the combination of Ad-shMPP1 and taxol act synergistically, the CI value of Ad-shMPP1/taxol was calculated both in HepG2 and Hep3B cells using CalcuSyn software, yielding CI values of 0.538 and 0.705, respectively (Fig. 3C), suggesting that an increased antitumor effect of taxol was achieved by its combination with Ad-shMPP1 treatment.

Ad-shMPP1 knockdown induces apoptosis, polyploidy, and senescence in HCC cells

To investigate whether apoptosis is promoted during the course of HCC cell growth arrest induced by MPHOSPH1 downregulation, we infected HepG2 cells with Ad-vector and Ad-shMPP1 at MOI = 2, and performed Hoechst staining 48 hours after infection. Nuclear fragmentation and chromatin condensation, which are typical apoptotic features, were clearly observed in the Ad-shMPP1–infected cells (Fig. 4A). Apoptotic DNA fragmentation, a key feature of apoptosis, was also induced by Ad-shMPP1 infection in HepG2 cells (Fig. 4B).

Figure 4.

MPHOSPH1 knockdown induces polyploidy and apoptosis in HCC cells. A, the nuclear morphology of HepG2 cells stained with Hoechst33258 dye were visualized 48 hours after infection. Arrows, apoptotic cells. Magnification, ×40. B, DNA fragmentation assays of HepG2 cells 48 hours after infection with Ads. DNA marker, DL2000. C, twenty-four hours after infection, the multinuclear HepG2 cells stained with DAPI and Phalloidin Alexa Fluor 555 were visualized by confocal microscopy. Scale bar, 10 μm. D, analysis of the polyploid HepG2 cells with PI staining by FACS 24 hours after infection. The numbers show the percentage of cells with >4N DNA content.

Figure 4.

MPHOSPH1 knockdown induces polyploidy and apoptosis in HCC cells. A, the nuclear morphology of HepG2 cells stained with Hoechst33258 dye were visualized 48 hours after infection. Arrows, apoptotic cells. Magnification, ×40. B, DNA fragmentation assays of HepG2 cells 48 hours after infection with Ads. DNA marker, DL2000. C, twenty-four hours after infection, the multinuclear HepG2 cells stained with DAPI and Phalloidin Alexa Fluor 555 were visualized by confocal microscopy. Scale bar, 10 μm. D, analysis of the polyploid HepG2 cells with PI staining by FACS 24 hours after infection. The numbers show the percentage of cells with >4N DNA content.

Close modal

Another obvious morphologic alteration following Ad-shMPP1 infection is the appearance of multinuclear cells due to the mitotic slippage from cell-cycle arrest without the formation of two daughter cells by cytokinesis discussed previously (6). In the current study, 24 hours after infection with Ad-shMPP1 at a relatively low MOI (MOI = 1), multinuclear cells were clearly observed by confocal microscopy (Fig. 4C). FACS analysis demonstrated accumulation of polyploid cells 24 hours after infection (>4N DNA content), which represents part of the multinucleate cells that were induced by Ad-shMPP1 (Fig. 4D). In agreement with our previous work, these results suggest that at an early stage of infection, prolonged mitotic arrest is induced because of failure of cytokinesis caused by reducing MPHOSPH1.

Moreover, 48 hours after infection with Ad-shMPP1 at an even lower MOI (MOI = 0.5), cellular senescence, rather than apoptosis, was accompanied by the growth arrest of HCC cells, as shown by altered cell shape and expression of pH-dependent β-galactosidase activity (Fig. 4E). Taken together, these results suggest that loss of MPHOSPH1 has a cytostatic effect on HCC cells probably through triggering mitotic arrest, and subsequent senescence or apoptosis depending on the MOI.

The p53 signal pathway inhibits proliferation of HCC cells after MPHOSPH1 knockdown

To study the mechanism underlying the MPHOSPH1 downregulation-induced HCC cell growth arrest, we analyzed the proteins involved in the process of apoptosis on Western blot analyses. The tumor suppressor p53, which is normal in HepG2 cells (14), was significantly elevated both in protein and mRNA levels when MPHOSPH1 was reduced (Fig. 5A and E). Phosphorylation of STAT3, downstream of p53, was suppressed in HepG2 cells after infection with Ad-shMPP1, whereas the base expression level of STAT3 was unaltered in comparison with the control (Fig. 5A–C), suggesting that STAT3 signal transduction was blocked in the process because STAT3 is activated by phosphorylation (15). Expression of the G1 phase cyclin-D1, a critical oncoprotein in several cancers as the driver for entering the cell cycle (16), was significantly reduced by the downregulation of MPHOSPH1 (Fig. 5A), suggesting a potential mechanism for mitotic arrest. Intriguingly, although infection with an Ad vector upregulates cyclin-B1 (Fig. 5A), the expression level of this M-phase cyclin was independent of MPHOSPH1 expression (Fig. 5A). In addition, Mad2, a critical spindle assembly checkpoint (SAC) promoting protein (17), was increased following Ad-shMPP1 infection (Fig. 5A), which suggested that mitotic arrest was induced by SAC. However, the expression of the mitotic inhibitor p27, which is also downstream of p53 and represses cell-cycle progression, was downregulated by MPHOSPH1 knockdown (Fig. 5A); this contradictory result indicates a complex molecular mechanism in the mitotic arrest process induced by reducing MPHOSPH1.

Figure 5.

Proteins involved in the inhibition of proliferation of HCC cells induced by MPHOSPH1 knockdown. A, Western blot analysis of proteins from HepG2 cells treated with Ads. B, Western blot analysis of proteins from HepG2 cells treated with Ads at different MOIs. C, densitometric quantitative analysis of western results. D, Western blot analysis of proteins from HepG2 cells with combined treatment of Ads and taxol. E, p53 mRNA was measured by RT-PCR in HepG2 cells treated with Ads. F, Western blot analysis of proteins from HepG2 cells transfected with shMPP1 expression plasmid at different times. G, measurement of cell viability of HepG2 cells 48 or 96 hours after transfection with shMPP1 expression plasmid or control plasmid using an MTT assay (*, P < 0.05; ***, P < 0.001).

Figure 5.

Proteins involved in the inhibition of proliferation of HCC cells induced by MPHOSPH1 knockdown. A, Western blot analysis of proteins from HepG2 cells treated with Ads. B, Western blot analysis of proteins from HepG2 cells treated with Ads at different MOIs. C, densitometric quantitative analysis of western results. D, Western blot analysis of proteins from HepG2 cells with combined treatment of Ads and taxol. E, p53 mRNA was measured by RT-PCR in HepG2 cells treated with Ads. F, Western blot analysis of proteins from HepG2 cells transfected with shMPP1 expression plasmid at different times. G, measurement of cell viability of HepG2 cells 48 or 96 hours after transfection with shMPP1 expression plasmid or control plasmid using an MTT assay (*, P < 0.05; ***, P < 0.001).

Close modal

In addition, p-STAT1, which is induced by IFNα and is a part of the ISGF3 transcription factor complex (18), was significantly upregulated after Ad-vector and Ad-shMPP1 treatment (Fig. 5B and C), suggesting that an IFN effect is triggered by Ad infection, which may explain why the Ad-vector also induces some cytotoxicity (Fig. 3A and B) because the IFN effect is capable of enhancing the antitumor effect of the oncolytic Ad (19). Interestingly, Ad-shMPP1 slightly downregulated the total expression level of STAT1 compared with Ad-vector (Fig. 5B and C), which shows a dual-effect of Ad-shMPP1 on the cells. First, it reduces the total level of STAT1, presumably due to reducing MPHOSPH1 expression; and second, as an oncolytic Ad, it significantly induces the phosphorylation of STAT1. This result indicates that infection with Ad-shMPP1 initiates a complicated process.

Other proteins involved in the combined treatment with Ad-shMPP1 and taxol were also analyzed by Western blotting. KIF inhibitors have been reported to induce the antiapoptotic Hsp70 in myeloma cells (20), potentially revealing an underlying mechanism by which cancer cells resist the KIF-targeted drugs through the antiapoptotic pathway. In the present study, we did not detect upregulation of Hsp70 following infection of HCC cells with Ad-shMPP1 (Fig. 5D), which suggests that Hsp70 may be not involved in the taxol resistance of HCC cells. In addition, cyclin-D1 was downregulated, not only in Ad-shMPP1–infected HepG2 cells, but also after the combined treatment with taxol to an even greater extent (Fig. 5D), suggesting a critical role for mitotic arrest in the affect of the Ad-shMPP1/taxol combined treatment. In addition, the expression of CDKI (cyclin-dependent kinase inhibitor) p15, which is also an antimitotic protein, was not affected by MPHOSPH1 knockdown (Fig. 5D), suggesting that the CDKI pathway may not be involved.

To minimize the interference of the Ad system (Fig. 5A and B), we also used a plasmid containing the shMPP1 cassette (p-shMPP1) to downregulate MPHOSPH1 in HCC cells (6). MPHOSPH1 was significantly downregulated by p-shMPP1 transfection in HepG2 cells (Fig. 5F), and, consistent with the Ad results (Fig. 5A), MPHOSPH1 knockdown significantly induced p53 expression while reducing the expression of p27, p-STAT3, and cyclin-D1 (Fig. 5F). The MTT assay also demonstrated that the downregulation of MPHOSPH1 by p-shMPP1 significantly attenuated HCC cell proliferation (Fig. 5G), in agreement with the MTT results from the Ad infection (Fig. 3B).

Knockdown of endogenous MPHOSPH1 inhibits HCC xenografts growth in vivo

To evaluate the efficacy of an MPHOSPH1-targeted strategy against HCC, a subcutaneous HepG2 cell xenograft model was used. Intratumoral injections of 2 × 109 plaque-forming unit (pfu) Ads or taxol were carried out after the mean tumor volume reached 80 to 120 mm3. As shown in Fig. 6A, at the end of a 6-week experiment, the Ad-shMPP1–treated animals demonstrated significant suppression of tumor growth with an average tumor volume of about 250 mm3, whereas the average tumor volumes of PBS or Ad-vector–injected animals were 960 and 770 mm3, respectively, indicating approximately 74% suppression of the xenograft tumor growth by Ad-shMPP1 (P < 0.05). The efficacy of combined treatment was also investigated. The combined treatment with Ad-shMPP1 and taxol demonstrated enhanced anti-HCC activity, with the mean tumor volumes of 180 mm3, whereas the groups of taxol and taxol/Ad-vector were 790 and 700 mm3 respectively, in contrast (P < 0.05; Fig. 6B), indicating the potent anti-HCC activity of the MPHOSPH1 knockdown in combination with taxol treatment.

Figure 6.

In vivo knockdown of endogenous MPHOSPH1 inhibits HCC xenograft growth. A and B, measurement of the tumor volume of different groups treated with Ads or taxol every 4 days (*, P < 0.05). C, IHC and H&E staining analysis of tumor sections derived from tumors. Magnification, ×20. D, AST and ALB values of the orthotopic HCC xenografts implanted mice 21 days after injection of Ads or PBS (*, P < 0.05; #, 0.05 < P < 0.07). E, analysis of the relative content of rec-Ad in mouse livers by RT-PCR after intravenous injection. The adenovirus E3 region was amplified to estimate the viral DNA content using GAPDH as an internal control.

Figure 6.

In vivo knockdown of endogenous MPHOSPH1 inhibits HCC xenograft growth. A and B, measurement of the tumor volume of different groups treated with Ads or taxol every 4 days (*, P < 0.05). C, IHC and H&E staining analysis of tumor sections derived from tumors. Magnification, ×20. D, AST and ALB values of the orthotopic HCC xenografts implanted mice 21 days after injection of Ads or PBS (*, P < 0.05; #, 0.05 < P < 0.07). E, analysis of the relative content of rec-Ad in mouse livers by RT-PCR after intravenous injection. The adenovirus E3 region was amplified to estimate the viral DNA content using GAPDH as an internal control.

Close modal

To study whether MPHOSPH1 knockdown also triggers apoptosis in HCC cells in vivo, pathologic examination was performed 7 days after the injection of Ads. IHC indicated significant reduction of the endogenous MPHOSPH1 in HCC cells induced by Ad-shMMP1 (Fig. 6C), demonstrating its ability to downregulate MPHOSPH1 in vivo. Furthermore, H&E staining showed large areas of nuclear fragmentation in HCC cells from Ad-shMPP1–treated xenografts, while few apoptotic cells were detected in the control groups (Fig. 6C). All these results suggest that Ad-shMPP1 may exert anti-HCC effects in vivo by interfering with the expression of endogenous MPHOSPH1 and inducing apoptosis in HCC tissues.

Furthermore, an orthotopic xenograft model was also used to better evaluate the anti-HCC efficacy of Ad-shMPP1 in vivo through intravenous injection. At 7 days after intrahepatic implantation of HepG2 cells, a total of 2 × 109 pfu of Ads were injected through the tail vein. Three weeks after the injection, only 33.3% (2 of 6) of the Ad-shMPP1–treated mice developed liver tumor nodules, and in the other four mice receiving Ad-shMPP1 (66.6%, 4 of 6), no macroscopic or microscopic tumors were found in their livers (Supplementary Table S2 and Supplementary Fig. S1). In contrast, all animals (6 of 6) in the control groups (PBS and Ad-vector groups) developed visible liver tumor nodules (Supplementary Table S2 and Supplementary Fig. S1).

Ad-shMPP1 also prevented the hepatic damage normally caused by liver cancer. Compared with the control groups, the aspartate transaminase (AST) value, which is often elevated in patients with HCC due to hepatic injury, was significantly lower after intravenous injection of Ad-shMPP1 in the in situ HCC xenograft implanted mice (Fig. 6D and Supplementary Table S2). Another hepatic marker, albumin (ALB), which reflects the protein synthesis ability of liver and is commonly impaired by hepatic ascites or malnutrition, was significantly improved by Ad-shMPP1 treatment of HCC animals (P < 0.05 vs. PBS group and P < 0.07 vs. Ad-vector group, Fig. 6D and Supplementary Table S2). The alanine transaminase (ALT) and alkaline phosphatase values were not statistically different between groups (Supplementary Table S2), although a trend toward improvement was nevertheless observed for ALT (P = 0.1107, Supplementary Table S2). To assess how long the Ad replicated after injection, qRT-PCR was used to evaluate the Ad content in animal's liver by determining the relative mRNA levels of the adenoviral E3 gene. We found that the oncolytic Ad persists in liver tissues for at least 8 days after intravenous injection, and that by 12 days postinjection, the Ads were nearly completely eliminated (Fig. 6E).

Recently, kinesins have attracted significant attention as potential new targets for cancer therapy in the clinic (13, 21). Some agents that target mitotic kinesin proteins, such as Eg5, have demonstrated potent antitumor effects in preclinical models (22). However, the results from current phase II clinical trials for these agents are disappointing. For instance, ispinesib (SB-715992, Cytokinetics and GlaxoSmithKline), the first Eg5 inhibitor to enter clinical trials, has been closed for further trials as a monotherapy due to its lack of clinical efficacy for a number of solid tumors (23–25). An explanation for the clinical failure of mitosis-specific suppressors is the much smaller proportion of mitotic cells in human tumors compared with the animal xenograft models (26). Because of the insufficient antitumor effect of monotherapy, several ongoing clinical studies have incorporated KIF inhibitors with chemotherapeutic agents including carboplatin, capecitabine, and docetaxel to explore potential synergistic effects (27). In addition, drug resistance continues to be a major challenge in HCC chemotherapy (8). Notably, some KIF proteins, such as KIFC3 and MCAK, play a critical role in the development of drug resistance in cancer cells (9, 28, 29). Therefore, a combined strategy that brings synergistic efficacy and reduces development of drug resistance is urgently required for combating malignant HCC.

In searching for novel HCC targets that potentially increase chemotherapy efficacy, we focused on the kinesin protein MPHOSPH1, first identified through its in vivo interaction with the mitotic peptidyl-prolylisomerase PIN1 (30). In the current study, we investigated the efficacy of the combined treatment of targeting MPHOSPH1 in combination with taxol in HCC xenograft models, and demonstrated that reducing endogenous MPHOSPH1 significantly enhanced the cytotoxic effect of taxol against HCC, which suggests that MPHOSPH1 is a potential target for HCC therapy. Because of the intrinsic cytotoxicity of the Ad-vector, however, the observed combined effect (Fig. 3C) may simply indicate the synergistic interaction of Ad-shMPP1 as a complex with taxol, which would not be interpreted as the interaction of taxol and the shRNA of MPHOSPH1.

We recently reported that depletion of MPHOSPH1 in colorectal cancer cells resulted in mitotic arrest due to failure of cytokinesis and subsequent apoptosis (6). Intriguingly, our previous study indicated that the observed multinucleate (polyploidy) cells eventually underwent apoptosis (6). In the present study, we also found that significant mitotic arrest was induced by downregulation of endogenous MPHOSPH1 in HCC cells, leading to polyploidy (the multinucleated cells in Fig. 4C and D), which suggested that the mitotic slippage was induced presumably by the prolonged mitotic arrest (6). Intriguingly, mitotic slippage that causes polyploid cells is considered to be a protective mechanism that cancer cells use against mitotic inhibition (31). However, the infected cells with polyploidy observed here eventually underwent apoptosis (Fig. 4A and B), suggesting that the mitotic slippage process triggered postmitotic cell death instead.

SNP is a useful tool to study the potential pattern of gene expression. By applying “SNPinfo” tool to predict the SNPs in the MPHOSPH1 promoter region (32), we found three SNPs with Minor Allele Frequency > 0.05 (rs10881625, rs9325443, and rs9325444) to be located at the transcription factor binding sites, which may regulate gene expression by impacting transcription factors binding. However, this is only a prediction and experimental verification is still needed.

Moreover, p53, which promotes postmitotic apoptosis due to its accumulation following mitotic arrest (33), was significantly elevated in Ad-shMPP1–treated HepG2 cells (Fig. 5A). Expression of the oncogene cyclin-D1 as well as phosphorylation of STAT3, which were reported to be regulated by p53 (34, 35), were also markedly inhibited by MPHOSPH1 knockdown, suggesting that STAT3 activation was blocked through the p53 pathway. A probable mechanism was proposed previously by Blagosklonny (33), suggesting that during the mitotic arrest when the nuclear envelope is dissolved and chromosomes are condensed, transcription is absent, while the p53 mRNA is long-lived and accumulates (Fig. 5A and E). However, p53 cannot induce its targets because transcription is blocked. When the cell escapes from the mitotic arrest (mitotic slippage), transcription resumes, and the accumulated p53 suddenly induces its proapoptotic targets such as PUMA, Bax, and CD95, which can trigger apoptosis (Fig. 4A and B; refs. 6, 33). This hypothesis is consistent with the present study and our previous work (6).

Interestingly, in addition to HCC with wild-type p53, we also demonstrated in Hep3B cells, which are p53-null, sensitivity to MPHOSPH1 interference, although to a lesser extent (Fig. 3). In a recent study, Forys and colleagues reported that loss of p53 leads to an increase in ARF protein levels, which function to limit the proliferation and tumorigenicity of p53-null cells, suggesting a complementary effect of ARF when p53 is deficient (36). It will be of great future interest to further investigate whether ARF or other tumor suppressors are involved in the p53-null cells to better understand the alternative regulation mechanism of MPHOSPH1 knockdown.

A carcinogenic link to MPHOSPH1 was previously found in bladder cancer (4), and we recently reported its upregulation in colorectal cancer cells (6). Here, we demonstrated that infection with Ad-shMPP1 not only suppressed HCC development, but also protected hepatic function from HCC impairment (Fig. 6D). Significantly, improved AST and ALB were observed following Ad-shMPP1 treatment. Because the estimation of the hepatic function reserve is critical in the prognosis of patients with primary liver cancer, improving poor hepatic function of patients is critical for facilitating proper treatment and reducing the risk of surgery. Another potential benefit of Ad-shMPP1 treatment is the IFN effect, which can be induced by Ad infection (Fig. 5B), and might further enhance the antitumor effects of the oncolytic Ad (19), although the IFN itself may also impair the replication of oncolytic Ad (37). In summary, the observed HCC inhibition and the improved hepatic function following treatment with Ad-shMPP1 may present an outstanding advance in clinical outcomes of HCC treatment.

No potential conflicts of interest were disclosed.

Conception and design: X. Liu, Y. Zhou, X. Liu, K. Huang

Development of methodology: X. Liu, Y. Zhou, K. Huang

Acquisition of data (provided animals, acquired and managed patients, provided facilities, etc.): X. Liu, A. Peng, L. Huang, K. Ji, L. Zheng

Analysis and interpretation of data (e.g., statistical analysis, biostatistics, computational analysis): X. Liu, A. Peng, K. Huang

Writing, review, and/or revision of the manuscript: X. Liu, H. Gong, R. B. Petersen, L. Zheng, K. Huang

Administrative, technical, or material support (i.e., reporting or organizing data, constructing databases): X. Liu, A. Peng

Study supervision: X. Liu, K. Huang

This work was supported by the Natural Science Foundation of China (Nos. 81202557, 31271370, 81100687, 81172971, 81222043, and 31471208), the Program for New Century Excellent Talents in University (NECT11-0170), the Municipal Key Technology Program of Wuhan (Wuhan Bureau of Science and Technology, No. 201260523174), the Health Bureau of Wuhan (WX12B06), and the Natural Science Foundation of Hubei Province (2013CFB359).

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.
Diefenbach
RJ
,
Mackay
JP
,
Armati
PJ
,
Cunningham
AL
. 
The C-terminal region of the stalk domain of ubiquitous human kinesin heavy chain contains the binding site for kinesin light chain
.
Biochemistry
1998
;
37
:
16663
70
.
2.
Goldstein
LS
,
Philp
AV
. 
The road less traveled: emerging principles of kinesin motor utilization
.
Annu Rev Cell Dev Biol
1999
;
15
:
141
83
.
3.
Corson
TW
,
Zhu
CQ
,
Lau
SK
,
Shepherd
FA
,
Tsao
MS
,
Gallie
BL
. 
KIF14 messenger RNA expression is independently prognostic for outcome in lung cancer
.
Clin Cancer Res
2007
;
13
:
3229
34
.
4.
Kanehira
M
,
Katagiri
T
,
Shimo
A
,
Takata
R
,
Shuin
T
,
Miki
T
, et al
Oncogenic role of MPHOSPH1, a cancer-testis antigen specific to human bladder cancer
.
Cancer Res
2007
;
67
:
3276
85
.
5.
Abaza
A
,
Soleilhac
JM
,
Westendorf
J
,
Piel
M
,
Crevel
I
,
Roux
A
, et al
M phase phosphoprotein 1 is a human plus-end-directed kinesin-related protein required for cytokinesis
.
J Biol Chem
2003
;
278
:
27844
52
.
6.
Liu
XR
,
Cai
Y
,
Cao
X
,
Wei
RC
,
Li
HL
,
Zhou
XM
, et al
A new oncolytic adenoviral vector carrying dual tumour suppressor genes shows potent anti-tumour effect
.
J Cell Mol Med
2012
;
16
:
1298
309
.
7.
Xie
B
,
Wang
DH
,
Spechler
SJ
. 
Sorafenib for treatment of hepatocellular carcinoma: a systematic review
.
Dig Dis Sci
2012
;
57
:
1122
9
.
8.
Kim
HY
,
Park
JW
. 
Molecularly targeted therapies for hepatocellular carcinoma: sorafenib as a stepping stone
.
Dig Dis
2011
;
29
:
303
9
.
9.
De
S
,
Cipriano
R
,
Jackson
MW
,
Stark
GR
. 
Overexpression of kinesins mediates docetaxel resistance in breast cancer cells
.
Cancer Res
2009
;
69
:
8035
42
.
10.
Liu
X
,
Cao
X
,
Wei
R
,
Cai
Y
,
Li
H
,
Gui
J
, et al
Gene-viro-therapy targeting liver cancer by a dual-regulated oncolytic adenoviral vector harboring IL-24 and TRAIL
.
Cancer Gene Ther
2012
;
19
:
49
57
.
11.
Chou
TC
,
Talalay
P
. 
Quantitative analysis of dose-effect relationships: the combined effects of multiple drugs or enzyme inhibitors
.
Adv Enzyme Regul
1984
;
22
:
27
55
.
12.
Herrmann
M
,
Lorenz
HM
,
Voll
R
,
Grunke
M
,
Woith
W
,
Kalden
JR
. 
A rapid and simple method for the isolation of apoptotic DNA fragments
.
Nucleic Acids Res
1994
;
22
:
5506
7
.
13.
Liu
X
,
Gong
H
,
Huang
K
. 
Oncogenic role of kinesin proteins and targeting kinesin therapy
.
Cancer Sci
2013
;
104
:
651
6
.
14.
Lee
TK
,
Lau
TC
,
Ng
IO
. 
Doxorubicin-induced apoptosis and chemosensitivity in hepatoma cell lines. Cancer Chemother Pharmacol
. 
2002
;
49
:
78
86
.
15.
Darnell
JE
 Jr
,
Kerr
IM
,
Stark
GR
. 
Jak-STAT pathways and transcriptional activation in response to IFNs and other extracellular signaling proteins
.
Science
1994
;
264
:
1415
21
.
16.
Musgrove
EA
,
Caldon
CE
,
Barraclough
J
,
Stone
A
,
Sutherland
RL
. 
Cyclin D as a therapeutic target in cancer
.
Nat Rev Cancer
2011
;
11
:
558
72
.
17.
De Antoni
A
,
Pearson
CG
,
Cimini
D
,
Canman
JC
,
Sala
V
,
Nezi
L
, et al
The Mad1/Mad2 complex as a template for Mad2 activation in the spindle assembly checkpoint
.
Curr Biol
2005
;
15
:
214
25
.
18.
Choi
JC
,
Holtz
R
,
Petroff
MG
,
Alfaidy
N
,
Murphy
SP
. 
Dampening of IFN-gamma-inducible gene expression in human choriocarcinoma cells is due to phosphatase-mediated inhibition of the JAK/STAT-1 pathway
.
J Immunol
2007
;
178
:
1598
607
.
19.
Su
C
,
Peng
L
,
Sham
J
,
Wang
X
,
Zhang
Q
,
Chua
D
, et al
Immune gene-viral therapy with triplex efficacy mediated by oncolytic adenovirus carrying an interferon-gamma gene yields efficient antitumor activity in immunodeficient and immunocompetent mice
.
Mol Ther
2006
;
13
:
918
27
.
20.
Liu
M
,
Aneja
R
,
Liu
C
,
Sun
L
,
Gao
J
,
Wang
H
, et al
Inhibition of the mitotic kinesin Eg5 up-regulates Hsp70 through the phosphatidylinositol 3-kinase/Akt pathway in multiple myeloma cells
.
J Biol Chem
2006
;
281
:
18090
7
.
21.
Rath
O
,
Kozielski
F
. 
Kinesins and cancer
.
Nat Rev Cancer
2012
;
12
:
527
39
.
22.
Sakowicz
R
,
Finer
JT
,
Beraud
C
,
Crompton
A
,
Lewis
E
,
Fritsch
A
, et al
Antitumor activity of a kinesin inhibitor
.
Cancer Res
2004
;
64
:
3276
80
.
23.
Knox
JJ
,
Gill
S
,
Synold
TW
,
Biagi
JJ
,
Major
P
,
Feld
R
, et al
A phase II and pharmacokinetic study of SB-715992, in patients with metastatic hepatocellular carcinoma: a study of the National Cancer Institute of Canada Clinical Trials Group (NCIC CTG IND.168)
.
Invest New Drugs
2008
;
26
:
265
72
.
24.
Tang
PA
,
Siu
LL
,
Chen
EX
,
Hotte
SJ
,
Chia
S
,
Schwarz
JK
, et al
Phase II study of ispinesib in recurrent or metastatic squamous cell carcinoma of the head and neck
.
Invest New Drugs
2008
;
26
:
257
64
.
25.
Lee
CW
,
Belanger
K
,
Rao
SC
,
Petrella
TM
,
Tozer
RG
,
Wood
L
, et al
A phase II study of ispinesib (SB-715992) in patients with metastatic or recurrent malignant melanoma: a National Cancer Institute of Canada Clinical Trials Group trial
.
Invest New Drugs
2008
;
26
:
249
55
.
26.
Komlodi-Pasztor
E
,
Sackett
DL
,
Fojo
AT
. 
Inhibitors targeting mitosis: tales of how great drugs against a promising target were brought down by a flawed rationale
.
Clin Cancer Res
2012
;
18
:
51
63
.
27.
Blagden
SP
,
Molife
LR
,
Seebaran
A
,
Payne
M
,
Reid
AH
,
Protheroe
AS
, et al
A phase I trial of ispinesib, a kinesin spindle protein inhibitor, with docetaxel in patients with advanced solid tumours
.
Br J Cancer
2008
;
98
:
894
9
.
28.
Ganguly
A
,
Yang
H
,
Cabral
F
. 
Overexpression of mitotic centromere-associated Kinesin stimulates microtubule detachment and confers resistance to paclitaxel
.
Mol Cancer Ther
2011
;
10
:
929
37
.
29.
Tan
MH
,
De
S
,
Bebek
G
,
Orloff
MS
,
Wesolowski
R
,
Downs-Kelly
E
, et al
Specific kinesin expression profiles associated with taxane resistance in basal-like breast cancer
.
Breast Cancer Res Treat
2012
;
131
:
849
58
.
30.
Kamimoto
T
,
Zama
T
,
Aoki
R
,
Muro
Y
,
Hagiwara
M
. 
Identification of a novel kinesin-related protein, KRMP1, as a target for mitotic peptidyl-prolyl isomerase Pin1
.
J Biol Chem
2001
;
276
:
37520
8
.
31.
Huang
HC
,
Shi
J
,
Orth
JD
,
Mitchison
TJ
. 
Evidence that mitotic exit is a better cancer therapeutic target than spindle assembly
.
Cancer Cell
2009
;
16
:
347
58
.
32.
Xu
Z
,
Taylor
JA
. 
SNPinfo: integrating GWAS and candidate gene information into functional SNP selection for genetic association studies
.
Nucleic Acids Res
2009
;
37
:
W600
5
.
33.
Blagosklonny
MV
. 
Mitotic arrest and cell fate: why and how mitotic inhibition of transcription drives mutually exclusive events
.
Cell Cycle
2007
;
6
:
70
4
.
34.
Liebermann
DA
,
Hoffman
B
,
Vesely
D
. 
p53 induced growth arrest versus apoptosis and its modulation by survival cytokines
.
Cell Cycle
2007
;
6
:
166
70
.
35.
Lin
J
,
Jin
X
,
Rothman
K
,
Lin
HJ
,
Tang
H
,
Burke
W
. 
Modulation of signal transducer and activator of transcription 3 activities by p53 tumor suppressor in breast cancer cells
.
Cancer Res
2002
;
62
:
376
80
.
36.
Forys
JT
,
Kuzmicki
CE
,
Saporita
AJ
,
Winkeler
CL
,
Maggi
LB
 Jr
,
Weber
JD
. 
ARF and p53 coordinate tumor suppression of an oncogenic IFN-beta-STAT1-ISG15 signaling axis
.
Cell Rep
2014
;
7
:
514
26
.
37.
Feigenblum
D
,
Walker
R
,
Schneider
RJ
. 
Adenovirus induction of an interferon-regulatory factor during entry into the late phase of infection
.
J Virol
1998
;
72
:
9257
66
.