Purpose:

Glioblastoma (GBM) is one of the most aggressive and lethal cancer types in humans. The standard treatment approach is surgery followed by chemoradiation. However, the molecular mechanisms of innate tumor radioresistance remain poorly understood.

Experimental Design:

We tested the expression of Smoothened (Smo) in primary and recurrent GBM tissues and cells. Then, we determined radiation effectiveness against primary and recurrent GBM cells. Lastly, the functional role of Smo in GBM radioresistance was further confirmed by in vitro and in vivo experiments.

Results:

We reported that Smo was significantly upregulated in recurrent GBM cell lines and tumor tissues following radiation treatment. Higher Smo expression indicated poor prognosis of GBM patients after radiation treatment. Smo had radioresistance effects in both GBM cells and human tumor xenografts. The mechanisms underlying these effects involved the attenuation of DNA damage repair caused by IR. Importantly, we found that the effect of Smo on radioresistance was mediated by Claspin polyubiquitination and proteasomal degradation, leading to the regulation of ATR–Chk1 signaling. Moreover, we found that Smo reduced Claspin polyubiquitination and proteasomal degradation by promoting USP3 transcription. Furthermore, we demonstrated that the Smo inhibitor GDC-0449 induced radiosensitivity to GBM.

Conclusions:

These data suggest that Smo confers radiation resistance in GBM by promoting USP3 transcription, leading to the activation of Claspin-dependent ATR–Chk1 signaling. These findings identify a potential mechanism of GBM resistance to radiation and suggest a potential therapeutic target for radiation resistance in GBM.

Translational Relevance

Glioblastomas (GBM) are extremely radioresistant, and the molecular mechanisms of innate tumor radioresistance remain poorly understood. Here, we report that Smo was significantly upregulated in recurrent GBM cells following radiation treatment. Smo had radioresistance effects in GBM. The mechanisms underlying these effects involved the attenuation of DNA damage repair caused by IR. Importantly, Smo confers radiation resistance in GBM by promoting USP3 transcription, leading to the activation of Claspin-dependent ATR–Chk1 signaling. Furthermore, we demonstrated that the Smo inhibitor GDC-0449 induced radiosensitivity to GBM. These findings identify a potential mechanism of GBM resistance to radiation and suggest a potential therapeutic target for radiation resistance in GBM.

Glioblastoma (GBM) is one of the most aggressive and lethal human cancer types and is a major health issue worldwide (1, 2). Despite standardized treatment that prolongs survival, the prognosis of GBM patients remains poor, and the median survival time is only approximately 12 months (3, 4). Previous studies have established that GBM typically recurs within the initial radiation treatment course, which indicates that GBM cells are extremely radioresistant (5, 6). Therefore, identifying novel mechanisms underlying radiation resistance in GBM is imperative.

The Hedgehog (Hh) pathway performs crucial functions in cell proliferation, apoptosis, and self-renewal (7, 8). Smoothened (Smo) is a component of the Hh pathway that triggers the downstream activation of Gli transcription factors (9, 10). Gli transcription factors consist of three proteins, Gli1, Gli2, and Gli3, which are responsible for regulating downstream target genes (9).

The ATM–Chk2 and ATR–Chk1 protein kinases make up two main signaling pathways activated by DNA damage (11–13). ATM–Chk2 and ATR–Chk1 protein kinase activation participates in numerous cellular events, including DNA damage recognition and processing, cell-cycle arrest, and apoptosis (12, 14). Once loaded to DNA, Rad17, and Claspin collaborate to fully activate ATR. Claspin also binds to Chk1 and is essential for the ATR phosphorylation of Chk1, which acts as both a sensor and a mediator of DNA replication (11, 12).

Ubiquitination is a vital posttranslational modification. However, deubiquitination, the reverse process mediated by deubiquitinating enzymes (DUB), is an important regulatory mechanism in controlling protein proteasomal degradation (15, 16). The deubiquitination enzyme ubiquitin-specific protease 3 (USP3) has been identified as a chromatin modifier that stabilizes genome integrity (17, 18).

In this study, we identified that Smo mediated radiation resistance in GBMs via promoting USP3 transcription, leading to the activation of Claspin-dependent ATR–Chk1 signaling. Thus, Smo may be a novel potential therapeutic target for GBMs.

Cell lines and reagents

Primary GBM cell lines derived from GBM surgical specimens were maintained in primary serum-free cultures grown on laminin. All cell lines were validated by short-tandem repeat DNA fingerprinting using an AmpFlSTR kit (Applied Biosystems), according to the manufacturer's instructions. During experiments, all GBM cell lines were cultured in DMEM supplemented with 10% FBS.

Patient tumors

Ten primary and the corresponding recurrent GBM patients who received radiotherapy (each pair was from the same patient) were all obtained from the Department of Neurosurgery, the First Affiliated Hospital of Nanjing Medical University. The histologic features of the tissues were independently examined independently by two neuropathologists according to the WHO criteria. The clinicopathologic and treatment characteristics of these patients are shown in Supplementary Table S1.

Real-time PCR assay

To quantify the mRNA expression of Smo, Claspin, and USP3, real-time PCR (qPCR) was carried out using SYBR Green (Applied Biosystems). The forward and reverse PCR primers for Smo were 5′-TGAAGGCTGCACGAATGAGG-3′ and 5′-CTTGGGGTTGTCTGTCCGAA-3′, the primers for Claspin were 5′-TGCCAAAGAAGCTACCGAGT-3′ and 5′-CCAGCATCTCATCCATAACACCT-3′, and the primers for USP3 were 5′-CAAGCTGGGACTGGTACAGAA-3′ and 5′-GCAGTGGTGCTTCCATTTACTT-3′. GAPDH was amplified as an internal control using the primers 5′-GAAGGTGAAGGTCGGAGTC-3′ and 5′-GAAGATGGTGATGGGATTTC-3′.

MGMT promoter methylation assay

Methylation specific polymerase chain reaction (MSP) was performed with primers specific for either methylated or the modified unmethylated DNA. The forward and reverse PCR primers for unmethylated reaction were 5′-TTTGTGTTTTGATGTTTGTAGGTTTTTGT-3′ and 5′-AACTCCACACTCTTCCAAAAACAAAACA-3′, and primers for the methylated reaction were 5′-TTTCGACGTTCGTAGGTTTTCGC-3′ and 5′-GCACTCTTCCGAAAACGAAACG-3′.

IHC assay

IHC assays were performed as previously described (19).

Colony formation assay

Radiation effectiveness was assessed by a colony formation assay. Cells were plated in 6-well plates, irradiated (0, 2, 4, 6, and 8 Gy), and returned to the incubator for 12 to 15 days. Then, cell colonies (>50 cells) were stained with crystal violet and counted. The surviving fraction was calculated as mean colony number/ (cells inoculated × plating efficiency). The data were analyzed using GraphPad Prism 5.0 software, and survival curves were plotted according to the single-hit multitarget model SF = 1 − (1 − e−D/D0)N, where SF is the surviving fraction and D is the radiation dose.

Immunostaining

Immunofluorescent staining was performed as described previously (20).

BrdUrd comet assay

Cells were plated in 6-well plates, irradiated (0 or 4 Gy), and returned to the incubator. DNA damage repair was assessed by single-cell gel electrophoresis assay under alkaline conditions as previously described (20).

Flow-cytometric assay

Cells were plated in 6-well plates, irradiated (0 or 4 Gy), and returned to the incubator for 12 hours. A flow cytometry assay was performed as described previously (21). The percentage of apoptotic cells was determined with Annexin V–PI/FITC staining. Cell-cycle progression was determined using propidium iodide staining.

Terminal deoxynucleotidyl transferase dUTP nick end labeling assay

A terminal deoxynucleotidyl transferase dUTP nick end labeling (TUNEL) apoptosis detection kit (Millipore) was used according to the manufacturer's instructions.

Western blot analysis

Antibodies against Smo, Gli1, Gli2, Rad17, Claspin, Plk1, ATR, p-ATR (Ser428), ATM, p-ATM (Ser1981), Chk1, p-Chk1 (Ser 317), p-Chk1 (Ser345), Chk2, p-Chk2 (T68; Abcam); HA (F-7), Myc (Santa Cruz Biotechnology); γ-H2AX, USP3, His, and β-Actin (Cell Signaling Technology) were used for Western blot analysis.

Coimmunoprecipitation

Coimmunoprecipitation (Co-IP) assays were performed as described previously (22).

Chromatin immunoprecipitation assay

A chromatin immunoprecipitation (Chip) assay was performed as described previously (23).

In vitro ubiquitylation of Claspin

The reaction was carried out at 37°C for 2 hours in 20 μL reaction buffer (20 mmol/L Tris-HCl, pH 7.2, 5 mmol/L MgCl2, 50 mmol/L NaCl, 1 mmol/L 2-mercaptoethanol, 10% glycerol) containing the following components: 50 mmol/L of ubiquitin WT, 100 nmol/L human E1, 2 μmol/L UbE2D3, 2 μmol/L HERC2, 2 mmol/L ATP, 5 μmol/L of GST-Claspin and His-Smo.

Subcutaneous and orthotopic xenografts

For subcutaneous xenograft studies, primary GBM2 cells (1 × 106) were inoculated into 6-week-old nude mice. Tumor volume was calculated using the formula: V = 0.5 × length × width2. After the tumor volume reached 50 mm3, the mice were irradiated with 0 or 2 Gy daily for 5 days. For orthotopic xenograft studies, GBM cells (5 × 105) were injected intracranially into the striatum of nude mice. The mice were subjected to IVIS scanning on the ninth day to confirm tumor uptake before treatment with radiation. All animal experiments were conducted with the approval of the Nanjing Medical University Institutional Committee for Animal Research and in conformity with national guidelines for the care and use of laboratory animals.

Statistical analyses

All results were presented as the mean ± SEM or the mean ± SD as indicated of at least three independent experiments. Data were subjected to Student t test for pairwise comparison or ANOVA for multivariate analysis. Differences were considered to be statistically significant at P < 0.05.

Study approval

GBM samples were collected with written informed consent from all patients. Human brain tumor specimens and the database were approved by the Institutional Review Board and the Ethics Committee of Nanjing Medical University.

The use of animals in this study was approved by Institutional Review Board and the Ethics Committee of Nanjing Medical University. The animals were administered treatment in accordance with the ethical standards formulated in the Helsinki Declaration and national guidelines and regulations.

Smo is markedly upregulated in recurrent GBMs after radiation treatment, and recurrent GBM cells show greater radioresistance

The mRNA expression levels of Smo were significantly higher in recurrent samples than in primary samples (Fig. 1A). Higher Smo expression indicated poor prognosis of GBM patients after radiation treatment (Supplementary Fig. S1A). Next, we analyzed Smo levels by IHC and Western blot, and Smo was markedly higher in the recurrent GBM samples than in the primary GBM samples (Fig. 1B; Supplementary Fig. S1B and S1C). In addition, Smo was significantly higher in the GBM cells than normal human astrocytes (NHA), especially higher in recurrent GBM cells (Supplementary Fig. S1D). To control for cell line heterogeneity, we established 4 primary GBM cell lines from 2 paired primary and recurrent GBM samples. We next determined GBM-associated molecular markers, including the methylation status of the MGMT promoter and the mutation status of IDH 1/2, EGFRvIII, and TP53 for the paired primary and recurrent GBM samples. All of the molecules and their mutations were consistent between the original tumor and cell line pairs (Supplementary Table S1). MGMT promoter methylation is reported to be a predictive biomarker of radiation response (24). Next, we evaluated MGMT promoter methylation and detected the protein expression of MGMT in the 2 paired primary and recurrent GBM cell lines. No differences were detected between these cell line pairs (Fig. 1C and D), which excluded the possibility that radiation resistance in recurrent GBM is due to MGMT promoter methylation. The radiation effectiveness against primary and recurrent GBM cells was assessed by a colony formation assay. Recurrent GBM cells showed greater resistance to radiation than primary GBM cells (Fig. 1E). Additionally, to verify cell survival after radiation, primary and recurrent GBM cells were differentially labeled with fluorescent dyes and mixed in defined ratios (primary GBM cells: red, recurrent GBM cells: green). In the presence of IR, the percentage of recurrent GBM cells after IR significantly increased (Fig. 1F; Supplementary Fig. S1E). We examined primary and recurrent GBM cells in response to IR-induced DNA damage by using γ-H2AX and the bromodeoxyuridine (BrdUrd) comet assay. We evaluated the presence of γ-H2AX foci in primary and recurrent GBM cells (Supplementary Fig. S1F). Although the numbers of γ-H2AX foci were similar in primary and in recurrent cells within 30 minutes after IR, the number of γ-H2AX foci in recurrent cells began to decrease within 2 hours after IR, whereas primary cells had a constantly high number of γ-H2AX foci, even after 12 hours. Primary and recurrent cells were equally susceptible to DNA damage by IR initially, but the percentage of cells with comet tails decreased more rapidly in recurrent cells than in matched primary cells (Supplementary Fig. S1G). These data confirm that recurrent GBM cells have greater radioresistance than primary GBM cells in vitro. In vivo, we used primary and recurrent GBM2 cells to generate orthotopic xenograft tumors in nude mice. From the 10th to the 15th day, the tumor-bearing mice were irradiated with 0 or 2 Gy daily for 5 days. Tumor progression was monitored using in vivo luminescence imaging. Xenografts carrying recurrent GBM2 cells displayed lager tumors (Supplementary Fig. S1H). Together, these results showed that Smo is markedly upregulated in recurrent GBM cell lines and tissues, and recurrent GBM cells show greater radioresistance.

Figure 1.

Smo is upregulated in recurrent GBM, and recurrent GBM cells show greater radioresistance. A, qPCR analysis of Smo expression in 10 paired primary and recurrent GBM patients who were insensitive to radiation therapy. **, P < 0.01; ***, P < 0.001. B, Representative images of IHC staining for Smo in two pairs of primary and recurrent GBM tissues. Scale bars, 100 μm. C, Analysis of methylation status of MGMT promoter primary and recurrent GBM cells (U, unmethylated; M, methylated). D, Two pairs of primary and recurrent GBM cells were subjected to Western blot analysis. E, Clonogenic survival curves for primary or recurrent GBM cells irradiated (0–8 Gy). The data are the mean of three independent experiments. F, Primary or recurrent GBM cells were labeled separately with cherry (red) and GFP (green) fluorescent dyes, with or without irradiation (4 Gy), and visualized by fluorescence microscopy at the indicated time points. Scale bars, 20 μm.

Figure 1.

Smo is upregulated in recurrent GBM, and recurrent GBM cells show greater radioresistance. A, qPCR analysis of Smo expression in 10 paired primary and recurrent GBM patients who were insensitive to radiation therapy. **, P < 0.01; ***, P < 0.001. B, Representative images of IHC staining for Smo in two pairs of primary and recurrent GBM tissues. Scale bars, 100 μm. C, Analysis of methylation status of MGMT promoter primary and recurrent GBM cells (U, unmethylated; M, methylated). D, Two pairs of primary and recurrent GBM cells were subjected to Western blot analysis. E, Clonogenic survival curves for primary or recurrent GBM cells irradiated (0–8 Gy). The data are the mean of three independent experiments. F, Primary or recurrent GBM cells were labeled separately with cherry (red) and GFP (green) fluorescent dyes, with or without irradiation (4 Gy), and visualized by fluorescence microscopy at the indicated time points. Scale bars, 20 μm.

Close modal

Smo overexpression confers radioresistance

To investigate whether Smo is associated with radioresistance, we selected primary GBM cells (the surviving fraction after 2 Gy irradiation was less than 45%) to overexpress Smo. As expected, Smo was overexpressed in the two primary GBM cell lines (Fig. 2A). Then, we determined the impact of Smo overexpression on primary GBM cells survival after IR. As shown in Fig. 2B, Smo overexpression made primary GBM cells more resistant to IR. Next, we assessed the apoptosis rate in Smo-overexpressing cells or vector control cells in response to IR. Smo-overexpressing cells had significantly fewer apoptotic cells (Annexin V–PI/FITC or TUNEL staining) and corresponding lower cleaved caspase-3 levels than control cells (Fig. 2CE; Supplementary Fig. S2A). These results demonstrate that Smo overexpression induces primary GBM cell resistance to IR and protects primary GBM cells from IR-induced apoptosis.

Figure 2.

Smo overexpression confers radiation resistance. A, Western blot analysis of Smo expression in primary GBM cells transfected with Smo or vector. β-Actin served as the loading control. B, Clonogenic survival curves for Smo-overexpressing or vector control primary GBM cells irradiated (0–8 Gy). The data represent the mean ± SEM from three independent experiments. C, Flow cytometry analysis of apoptosis in Smo-overexpressing or vector control GBM cells with or without irradiation (4 Gy, 12 hours later). The data represent the mean ± SEM from three independent experiments. **, P < 0.01. D, TUNEL analysis of apoptosis in Smo-overexpressing or vector control primary GBM cells with or without irradiation (4 Gy, 12 hours later). The data are the mean of three independent experiments. **, P < 0.01. E, Western blot analysis of cleaved caspase-3 expression in Smo-overexpressing or vector control primary GBM cells with or without irradiation (4 Gy, 12 hours later). β-Actin served as the loading control. F, Growth curve of Smo-overexpressing or vector control primary GBM1 cell–derived subcutaneous tumor xenografts after treatment with IR. *, P < 0.05; **, P < 0.01. G and H, Representative images of tumors originating from Smo-overexpressing or vector control primary GBM1 cells on the 35th day are shown (G); the tumor weights are the means of three independent experiments ± SEM (H). **, P < 0.01.

Figure 2.

Smo overexpression confers radiation resistance. A, Western blot analysis of Smo expression in primary GBM cells transfected with Smo or vector. β-Actin served as the loading control. B, Clonogenic survival curves for Smo-overexpressing or vector control primary GBM cells irradiated (0–8 Gy). The data represent the mean ± SEM from three independent experiments. C, Flow cytometry analysis of apoptosis in Smo-overexpressing or vector control GBM cells with or without irradiation (4 Gy, 12 hours later). The data represent the mean ± SEM from three independent experiments. **, P < 0.01. D, TUNEL analysis of apoptosis in Smo-overexpressing or vector control primary GBM cells with or without irradiation (4 Gy, 12 hours later). The data are the mean of three independent experiments. **, P < 0.01. E, Western blot analysis of cleaved caspase-3 expression in Smo-overexpressing or vector control primary GBM cells with or without irradiation (4 Gy, 12 hours later). β-Actin served as the loading control. F, Growth curve of Smo-overexpressing or vector control primary GBM1 cell–derived subcutaneous tumor xenografts after treatment with IR. *, P < 0.05; **, P < 0.01. G and H, Representative images of tumors originating from Smo-overexpressing or vector control primary GBM1 cells on the 35th day are shown (G); the tumor weights are the means of three independent experiments ± SEM (H). **, P < 0.01.

Close modal

We then used an in vivo approach to examine whether Smo conferred radioresistance. We subcutaneously injected primary GBM1 cells overexpressing Smo or a vector control into nude mice. After GBM engraftment was established, the mice were treated with IR at 0 or 2 Gy daily for 5 days. Primary GBM1 cells overexpressing Smo resulted in significantly higher tumor volumes and weights in irradiated mice than cells overexpressing the vector control (Fig. 2FH). Taken together, these results indicate that Smo overexpression contributes to the acquisition of radioresistance.

Smo inhibition sensitizes recurrent GBM cells to IR

Because Smo overexpression could confer IR resistance in GBM cells, we determined whether inhibiting Smo may modify GBM cell radiosensitivity. We infected two recurrent GBM cell lines (the surviving fraction after 2 Gy irradiation was over 65%) with three independent luciferase-encoding Smo shRNAs or control shRNA. As shown in Fig. 3A and Supplementary Fig. S3A, altered Smo expression in recurrent GBM cells was confirmed by Western blotting. We chose sh-Smo-1 for subsequent experiments due to its silencing efficiency. Smo-depleted recurrent GBM cells were more sensitive to radiation than control cells (Fig. 3B; Supplementary Fig. S3B).

Figure 3.

Smo silencing induces radiosensitivity. A, Western blot analysis of Smo in recurrent GBM2 cells transfected with three independent luciferase-encoding Smo shRNAs or control shRNA. β-Actin served as the loading control. B, Clonogenic survival curves for sh-Smo-1 or shCtrl recurrent GBM2 cells after IR (0–8 Gy). The data represent the mean ± SEM from three independent experiments. C, Immunodetection of γ-H2AX foci in sh-Smo-1 or shCtrl recurrent GBM2 cells with or without irradiation (4 Gy, 12 hours later). Quantitation of cells containing >10 γ-H2AX foci 0 to 12 hours later. The data represent the mean ± SEM from three independent experiments. Scale bars, 20 μm. D, Western blot analysis of γ-H2AX expression in sh-Smo-1 or shCtrl recurrent GBM2 cells with or without irradiation (4 Gy). β-Actin served as the loading control. E, Representative images (left) and quantification (right) of BrdUrd comet assays of recurrent GBM2 cells transduced with Smo shRNA or control at the indicated time after IR (4 Gy). The data are the means of three independent experiments ± SEM. Scale bars, 20 μm. F, Distribution of cell-cycle phases in sh-Smo-1 or shCtrl recurrent GBM cells with or without irradiation (4 Gy). The data are the means of three independent experiments ± SEM. **, P < 0.01. G, Flow cytometry analysis of apoptosis in sh-Smo-1 or shCtrl recurrent GBM cells with or without irradiation (4 Gy, 12 hours later). The data are the means of three independent experiments ± SEM. **, P < 0.01. H, Quantitation of TUNEL-positive cells 12 hours after IR (4 Gy, 12 hours later). The data are the means of three independent experiments ± SEM. **, P < 0.01. I, Western blot analysis of cleaved caspase-3 expression in sh-Smo-1 or shCtrl recurrent GBM cells with or without irradiation (4 Gy, 12 hours later). β-Actin served as the loading control. J,In vivo bioluminescent images of intracranial xenografts originating from Smo-depleted recurrent GBM2 or control cells with or without IR at 2 Gy daily for 5 days. K, Kaplan–Meier survival analysis of intracranial xenografts originating from Smo-depleted recurrent GBM2 or control cells with or without IR.

Figure 3.

Smo silencing induces radiosensitivity. A, Western blot analysis of Smo in recurrent GBM2 cells transfected with three independent luciferase-encoding Smo shRNAs or control shRNA. β-Actin served as the loading control. B, Clonogenic survival curves for sh-Smo-1 or shCtrl recurrent GBM2 cells after IR (0–8 Gy). The data represent the mean ± SEM from three independent experiments. C, Immunodetection of γ-H2AX foci in sh-Smo-1 or shCtrl recurrent GBM2 cells with or without irradiation (4 Gy, 12 hours later). Quantitation of cells containing >10 γ-H2AX foci 0 to 12 hours later. The data represent the mean ± SEM from three independent experiments. Scale bars, 20 μm. D, Western blot analysis of γ-H2AX expression in sh-Smo-1 or shCtrl recurrent GBM2 cells with or without irradiation (4 Gy). β-Actin served as the loading control. E, Representative images (left) and quantification (right) of BrdUrd comet assays of recurrent GBM2 cells transduced with Smo shRNA or control at the indicated time after IR (4 Gy). The data are the means of three independent experiments ± SEM. Scale bars, 20 μm. F, Distribution of cell-cycle phases in sh-Smo-1 or shCtrl recurrent GBM cells with or without irradiation (4 Gy). The data are the means of three independent experiments ± SEM. **, P < 0.01. G, Flow cytometry analysis of apoptosis in sh-Smo-1 or shCtrl recurrent GBM cells with or without irradiation (4 Gy, 12 hours later). The data are the means of three independent experiments ± SEM. **, P < 0.01. H, Quantitation of TUNEL-positive cells 12 hours after IR (4 Gy, 12 hours later). The data are the means of three independent experiments ± SEM. **, P < 0.01. I, Western blot analysis of cleaved caspase-3 expression in sh-Smo-1 or shCtrl recurrent GBM cells with or without irradiation (4 Gy, 12 hours later). β-Actin served as the loading control. J,In vivo bioluminescent images of intracranial xenografts originating from Smo-depleted recurrent GBM2 or control cells with or without IR at 2 Gy daily for 5 days. K, Kaplan–Meier survival analysis of intracranial xenografts originating from Smo-depleted recurrent GBM2 or control cells with or without IR.

Close modal

To investigate the impact of Smo inhibition on IR-induced DNA damage repair, we examined the profile of DNA damage at different time points via γ-H2AX. We evaluated the presence of γ-H2AX foci in sh-Smo-1 and shCtrl recurrent GBM cells (Fig. 3C; Supplementary Fig. S3C). Although the number of γ-H2AX foci was similar in sh-Smo-1 and shCtrl cells within 30 minutes after IR, the number of γ-H2AX foci in shCtrl cells began to decrease within 2 hours after IR, whereas sh-Smo-1 cells had a constantly high number of γ-H2AX foci, even after 12 hours. These results indicate that Smo inhibition delayed the repair of radiation-induced DNA damage. In addition, we determined the level of γ-H2AX by Western blotting to confirm this result (Fig. 3D; Supplementary Fig. S3D). Then, we compared the recovery of sh-Smo-1 and shCtrl cells in response to IR-induced DNA damage using a BrdUrd comet assay. sh-Smo-1 and shCtrl cells were equally susceptible to DNA damage to IR initially, but the percentage of cells with comet tails decreased more rapidly for shCtrl cells than for matched sh-Smo-1 cells (Fig. 3E; Supplementary Fig. S3E). These data indicate that sh-Smo-1 cells repaired the DNA damage less efficiently than shCtrl cells.

We next determined the effect of Smo knockdown on the cell cycle after IR in recurrent GBM cells (Fig. 3F; Supplementary Fig. S3F). Smo knockdown augmented IR-induced G2–M arrest in recurrent GBM cells. Tumor cells with impaired DNA damage repair ability are more prone to apoptosis. Therefore, we examined the capacity of Smo knockdown to induce apoptosis in response to radiation. As shown in Fig. 3G and Supplementary Fig. S3G, a remarkable increase in the number of apoptotic cells was observed in sh-Smo-1 cells compared with shCtrl cells. We also observed a significant increase in the number of apoptotic cells (determined by TUNEL staining) in sh-Smo-1 cells compared with shCtrl cells (Fig. 3H; Supplementary Fig. S3H and S3I). In accordance with these results, cleaved caspase-3 levels were higher in sh-Smo-1 cells than in shCtrl cells (Fig. 3I).

To test the in vitro results in an in vivo setting, we used recurrent GBM2 cells to generate orthotopic xenograft tumors in nude mice. From the 10th to the 15th day, the tumor-bearing mice were irradiated with 0 or 2 Gy daily for 5 days. Tumor progression was monitored using in vivo luminescence imaging. In stark contrast, xenografts carrying Smo-depleted recurrent GBM2 cells displayed significantly suppressed tumor growth (Fig. 3J; Supplementary Fig. S4A). The survival curves showed that IR-treated Smo-depleted xenografts exhibited significantly increased survival compared with IR-treated control xenografts (Fig. 3K). Furthermore, we examined the expression of γ-H2AX by IHC (Supplementary Fig. S4B) and examined apoptosis by TUNEL staining (Supplementary Fig. S4C). These data suggest that Smo inhibition sensitizes recurrent GBM cells to IR.

Smo mediates Claspin polyubiquitination and proteasomal degradation, leading to ATR–Chk1 signaling regulation

Recent studies have shown that the ATR/Chk1 and ATM/Chk2 pathways are two primary kinase signaling cascades activated in response to DNA damage (12). Activation of ATR and ATM results in the phosphorylation of a diverse array of downstream targets that participate in numerous cellular events, including DNA damage recognition and processing, cell-cycle arrest at different points, and apoptosis (14). Chk1 and Chk2 are selectively phosphorylated and activated by ATR and ATM, respectively, which triggers a wide range of distinct downstream responses (11). This prompted us to determine whether Smo knockdown enhanced the radiation effect through the ATR/Chk1 and/or ATM/Chk2 signaling pathways. Smo knockdown in recurrent GBM cells led to a significant decrease in the phosphorylation of Chk1 at Ser317 and Ser345 and ATR at Ser428 with a concomitant increase in γ-H2AX levels after IR (Fig. 4A). However, Smo knockdown in recurrent GBM cells did not affect the ATM/Chk2 pathway (Supplementary Fig. S5A). Claspin, Rad 17, and Plk1 are essential proteins for the ATR–Chk1-dependent activation of the DNA replication checkpoint response (13). First, we assessed the effect of Smo knockdown on the levels of Claspin, Rad17, and Plk1. As shown in Fig. 4B, Smo knockdown in recurrent GBM cells decreased the levels of total Claspin after IR, whereas the protein levels of Rad17 and Plk1 were not altered. Claspin is phosphorylated by ATR and binds directly to Chk1, which is important for Chk1 activation. Second, we determined how Smo inhibition affected Claspin expression in response to radiation-induced DNA damage. We performed qPCR to measure Claspin mRNA levels in sh-Smo-1 or shCtrl cells after radiation. However, no significant difference in Claspin mRNA levels was observed between sh-Smo-1 and shCtrl cells (Fig. 4C). Therefore, Claspin expression was not regulated at the mRNA level by Smo knockdown. Previous studies have reported that Claspin degradation occurs through ubiquitination and proteasomal degradation. Thus, we next treated sh-Smo-1 or shCtrl cells with the proteasome inhibitor MG132 and radiation. As shown in Fig. 4D, MG132 restored the levels of Claspin in sh-Smo-1 and shCtrl recurrent GBM cells. These results indicated that Smo could regulate Claspin protein stability via a proteasome-dependent pathway. It is known that the ubiquitin-proteasome system (UPS) promotes the proteasomal degradation of target proteins. Thus, we asked whether Smo mediated radiation-induced Claspin proteolysis via the UPS. Sh-Smo-1 and shCtrl recurrent GBM cells were transiently cotransfected with plasmids encoding Myc-tagged ubiquitin and HA-tagged Claspin and then irradiated. Sh-Smo-1 cells showed significantly higher radiation-induced Claspin protein levels with dramatically less polyubiquitination than shCtrl cells (Fig. 4E). Additionally, Smo overexpression in primary GBM cells led to a significant increase in the phosphorylation of Chk1 at Ser317 and Ser345 and ATR at Ser428, as well as a decrease in γ-H2AX levels after IR (Supplementary Fig. S5B). Smo cells showed significantly lower radiation-induced Claspin protein levels with dramatically less polyubiquitination than vector cells (Fig. 4F). To determine whether Claspin itself is ubiquitinated, the bacteria-expressed and purified GST-Claspin and His-Smo proteins were incubated with E1, UbE2D3 (E2), HERC2(E3), and ubiquitin for 2 hours. The mixtures were subjected to GST pulldown and Western blot with anti-Myc antibody. In vitro ubiquitylation assays showed that ubiquitination was present on Claspin itself, and Smo could not directly decrease the ubiquitin of Claspin (Supplementary Fig. S5C). To rule out shRNA off-target effects, we showed that Smo add-back reverses the following three phenotypes due to the sh-Smo-1 knockdown: increased IR sensitivity, persistent γH2AX foci, and Claspin destabilization (Supplementary Fig. S5D–S5G). Finally, to further confirm whether Smo mediated the Claspin-mediated ATR/Chk1 pathway, we determined whether Claspin could rescue Sh-Smo-1 cells from radiation-induced DNA damage. An Myc-tagged Claspin overexpression construct was expressed in sh-Smo-1 and shCtrl recurrent GBM cells, which were then irradiated. Exogenous expression of Claspin rescued the ATR/Chk1 pathway (Fig. 4G). Together, these data indicate that Smo regulates the ATR–Chk1 signaling pathway by mediating Claspin polyubiquitination and proteasomal degradation.

Figure 4.

Smo reduces Claspin polyubiquitination and proteasomal degradation, leading to the activation of ATR–Chk1 signaling. A, Western blot analysis of Smo, Gli1, Gli2, p-Chk1 (Ser317), p-Chk1 (Ser345), Chk1, p-ATR (Ser428), ATR, γ-H2AX, and β-Actin in sh-Smo-1 or shCtrl recurrent GBM cells with or without IR (12 hours later). B, Western blot analysis of Claspin, Rad17, Plk1, and β-Actin in sh-Smo-1 or shCtrl recurrent GBM cells 12 hours after IR. C, Real-time PCR analysis of Claspin in sh-Smo-1 or shCtrl recurrent GBM cells 12 hours after IR. D, Western blot analysis of Claspin and β-Actin in sh-Smo-1 or shCtrl recurrent GBM cells 12 hours after IR in the absence or presence of 20 mmol/L MG132. E, Myc-tagged ubiquitin and HA-tagged Claspin were cotransfected into sh-Smo-1 or shCtrl recurrent GBM cells 12 hours after IR. HA-Claspin was pulled down using an antibody against HA. Ubiquitin-conjugated HA-Claspin was subsequently detected by immunoblotting using an anti-Myc antibody. F, Myc-tagged ubiquitin and HA-tagged Claspin were cotransfected into Smo or vector primary GBM cells 12 hours after IR. HA-Claspin was pulled down using an antibody against HA. Ubiquitin-conjugated HA-Claspin was subsequently detected by immunoblotting using an anti-Myc antibody. G, Sh-Smo-1 or shCtrl recurrent GBM cells were transfected with a Myc-tagged Claspin construct or control vector 12 hours after IR. Western blot analysis of Claspin, p-Chk1 (Ser317), p-Chk1 (Ser345), Chk1, γ-H2AX, and β-Actin.

Figure 4.

Smo reduces Claspin polyubiquitination and proteasomal degradation, leading to the activation of ATR–Chk1 signaling. A, Western blot analysis of Smo, Gli1, Gli2, p-Chk1 (Ser317), p-Chk1 (Ser345), Chk1, p-ATR (Ser428), ATR, γ-H2AX, and β-Actin in sh-Smo-1 or shCtrl recurrent GBM cells with or without IR (12 hours later). B, Western blot analysis of Claspin, Rad17, Plk1, and β-Actin in sh-Smo-1 or shCtrl recurrent GBM cells 12 hours after IR. C, Real-time PCR analysis of Claspin in sh-Smo-1 or shCtrl recurrent GBM cells 12 hours after IR. D, Western blot analysis of Claspin and β-Actin in sh-Smo-1 or shCtrl recurrent GBM cells 12 hours after IR in the absence or presence of 20 mmol/L MG132. E, Myc-tagged ubiquitin and HA-tagged Claspin were cotransfected into sh-Smo-1 or shCtrl recurrent GBM cells 12 hours after IR. HA-Claspin was pulled down using an antibody against HA. Ubiquitin-conjugated HA-Claspin was subsequently detected by immunoblotting using an anti-Myc antibody. F, Myc-tagged ubiquitin and HA-tagged Claspin were cotransfected into Smo or vector primary GBM cells 12 hours after IR. HA-Claspin was pulled down using an antibody against HA. Ubiquitin-conjugated HA-Claspin was subsequently detected by immunoblotting using an anti-Myc antibody. G, Sh-Smo-1 or shCtrl recurrent GBM cells were transfected with a Myc-tagged Claspin construct or control vector 12 hours after IR. Western blot analysis of Claspin, p-Chk1 (Ser317), p-Chk1 (Ser345), Chk1, γ-H2AX, and β-Actin.

Close modal

Smo stabilizes Claspin by promoting USP3 transcription

Deubiquitinating enzymes (DUB) are a group of proteases that regulate ubiquitin-dependent pathways by cleaving ubiquitin-protein bonds (25). To identify potential DUBs responsible for Claspin degradation, we used a DUB siRNA library designed for effective and convenient siRNA to knock down the entire DUB family. DUB library screening revealed that siRNA-mediated inhibition of all DUB genes decreased Claspin protein levels. Among the 98 known human DUBs (Supplementary Table S2), the knockdown of only 5 (USP3, USP9X, USP9Y, USP20, and USP50) types of DUBs decreased Claspin protein levels by more than half (Fig. 5A; Supplementary Fig. S6A). A recent study showed that the Hedgehog (Hh) signaling pathway could promote DUB transcription (23). We asked whether Smo stabilizes Claspin by promoting DUB transcription. We detected the mRNA levels of USP3, USP9X, USP9Y, USP20, and USP50 in sh-Smo-1, shCtrl, Smo, and vector control HEK293 cells. Previous studies reported that USP7, USP20, and USP29 could stabilize Claspin level (26–28). We also detected the mRNA levels of USP7 and USP29 in sh-Smo-1, shCtrl, Smo, and vector control HEK293 cells. The results showed that Smo knockdown decreased the mRNA levels of only USP3, and Smo overexpression increased the mRNA levels of USP3 (Fig. 5B; Supplementary Fig. S7A and S7B). In addition, Smo affected the mRNA levels of USP3 through Gli 2 (Fig. 5C), but not Gli 1 and Gli 3 (Supplementary Fig. S7C and S7D). Transcription factors may bind to specific regions of promoters to regulate gene expression, and we found one potential Gli 2-binding site (Fig. 5D). Reporter gene assays showed that the USP3 promoter could promote reporter gene expression more effectively than the pGL3-Basic control vector (Fig. 5E). Then, a chromatin immunoprecipitation assay was performed to clarify the association of Gli 2 and the USP3 promoter region in recurrent GBM cells (Fig. 5F).

Figure 5.

Smo stabilizes Claspin through USP3. A, DUB siRNA library screening revealed that siRNA-mediated inhibition of multiple DUB genes decreased Claspin protein levels. **, P < 0.01. B, Smo silencing decreased the mRNA expression of USP3, and Smo upregulation increased the mRNA expression of USP3. The data are the means of three independent experiments ± SEM. C, Smo mediated USP3 mRNA expression through Gli2. **, P < 0.01. D, Diagram showing a potential Gli2-binding site in the promoter region of USP3. E, HEK293T cells were transfected with pGL3-Basic or pGL3-USP3-Pro together with an increasing amount of HA-Gli2. The pRL-TK plasmid was used as an internal control for cell number and transfection efficiency. Reporter gene expression was measured by dual-luciferase assays. The data are the means of three independent experiments ± SEM. **, P < 0.01. F, Chromatin immunoprecipitation assays detected the binding of Gli2 with a USP3 promoter in recurrent GBM cells. G, HEK293T cells were transfected with increasing amounts of wild-type USP3 or USP3-C168S, and Claspin expression was detected by Western blotting. H, Recurrent GBM cells were transfected with plasmids expressing wild-type USP3 or USP3-C168S, and Claspin expression was detected by Western blotting. I and J, Recurrent GBM cells were transfected with two independent luciferase-encoding USP3 shRNAs and then treated with DMSO or 20 μmol/L MG132 for 6 hours. Cell lysates were detected by Western blotting using a Claspin antibody. K and L, Sh-USP3 or shCtrl recurrent GBM cells were treated with 50 μg/mL cycloheximide. Whole-cell lysates were harvested at the indicated times. **, P < 0.01.

Figure 5.

Smo stabilizes Claspin through USP3. A, DUB siRNA library screening revealed that siRNA-mediated inhibition of multiple DUB genes decreased Claspin protein levels. **, P < 0.01. B, Smo silencing decreased the mRNA expression of USP3, and Smo upregulation increased the mRNA expression of USP3. The data are the means of three independent experiments ± SEM. C, Smo mediated USP3 mRNA expression through Gli2. **, P < 0.01. D, Diagram showing a potential Gli2-binding site in the promoter region of USP3. E, HEK293T cells were transfected with pGL3-Basic or pGL3-USP3-Pro together with an increasing amount of HA-Gli2. The pRL-TK plasmid was used as an internal control for cell number and transfection efficiency. Reporter gene expression was measured by dual-luciferase assays. The data are the means of three independent experiments ± SEM. **, P < 0.01. F, Chromatin immunoprecipitation assays detected the binding of Gli2 with a USP3 promoter in recurrent GBM cells. G, HEK293T cells were transfected with increasing amounts of wild-type USP3 or USP3-C168S, and Claspin expression was detected by Western blotting. H, Recurrent GBM cells were transfected with plasmids expressing wild-type USP3 or USP3-C168S, and Claspin expression was detected by Western blotting. I and J, Recurrent GBM cells were transfected with two independent luciferase-encoding USP3 shRNAs and then treated with DMSO or 20 μmol/L MG132 for 6 hours. Cell lysates were detected by Western blotting using a Claspin antibody. K and L, Sh-USP3 or shCtrl recurrent GBM cells were treated with 50 μg/mL cycloheximide. Whole-cell lysates were harvested at the indicated times. **, P < 0.01.

Close modal

Then, we detected the expression of USP3 and Claspin in primary and recurrent GBM tissues and cells. We found that mRNA and protein levels of USP3 and protein levels of Claspin in recurrent GBM were higher than that in primary GBM. But the mRNA expression of Claspin was not changed (Supplementary Fig. S8A-S8C). We next examined whether USP3 could stabilize Claspin. We found that USP3-WT overexpression increased the protein levels of Claspin, but the catalytically inactive C168S mutant of USP3 failed to induce Claspin levels (Fig. 5G and H). USP3 knockdown in recurrent GBM cells decreased Claspin expression, and this effect was partially reversed by MG132 (Fig. 5I and J). Moreover, USP3 depletion promoted the degradation of Claspin in recurrent GBM cells treated with cycloheximide (Fig. 5K and L). Lastly, USP3 overexpression reverse sh-Smo–induced decreasing in Claspin, CHK1 phosphorylation, and γ-H2AX after IR (Supplementary Fig. S8D). So, Smo stabilizes Claspin via USP3.

USP3 interacts with Claspin

We then examined the interaction between USP3 and Claspin. Co-IP experiments in HEK293T and recurrent GBM cells showed that USP3 and Claspin interact with each other (Fig. 6A and B). Confocal images showed colocalization of USP3 (green) and Claspin (red) in recurrent GBM cells (Fig. 6C). USP3 contains two conserved protein domains: a zinc-finger domain (ZnF) and a catalytic domain (UCH). To identify which USP3 domain mediated the interaction with Claspin, we generated two deletion mutants containing the ZnF and UCH domains of USP3 (Fig. 6D). We found that the UCH domain mediated the interaction with Claspin. In addition, we generated six truncated fragments spanning full-length Claspin as described in a previous study (28), and our domain mapping experiments revealed that the N-terminus of Claspin (1–330) mediated the interaction with USP3 (Fig. 6E). USP3 (UCH) showed an inhibitory effect on Claspin ISRE-luc activation, but USP3 (ZnF) did not have any inhibitory effect on Claspin-induced ISRE-luc activation (Supplementary Fig. S9A).

Figure 6.

USP3 interacts with Claspin. A and B, The interaction between USP3 and Claspin was confirmed by coimmunoprecipitation in HEK293T and recurrent GBM cells. C, Confocal images showing colocalization of USP3 (green) and Claspin (red) in recurrent GBM cells. Nuclei were counterstained with DAPI (blue). Scale bars, 10 μm. D, Constructs of full-length USP3 [USP3 (FL)] or USP3 containing only the ZnF [USP3 (ZnF)] or UCH domain [USP3 (UCH)]. The interaction between USP3 (FL), USP3 (ZnF), and USP3 (UCH) with Claspin was confirmed by coimmunoprecipitation in HEK293T cells. E, Schematic structure of Claspin fragments. The interaction between Claspin (FL), Claspin (F1), Claspin (F2), Claspin (F3), Claspin (F4), and Claspin (F5) and USP3 was confirmed by coimmunoprecipitation in HEK293T cells.

Figure 6.

USP3 interacts with Claspin. A and B, The interaction between USP3 and Claspin was confirmed by coimmunoprecipitation in HEK293T and recurrent GBM cells. C, Confocal images showing colocalization of USP3 (green) and Claspin (red) in recurrent GBM cells. Nuclei were counterstained with DAPI (blue). Scale bars, 10 μm. D, Constructs of full-length USP3 [USP3 (FL)] or USP3 containing only the ZnF [USP3 (ZnF)] or UCH domain [USP3 (UCH)]. The interaction between USP3 (FL), USP3 (ZnF), and USP3 (UCH) with Claspin was confirmed by coimmunoprecipitation in HEK293T cells. E, Schematic structure of Claspin fragments. The interaction between Claspin (FL), Claspin (F1), Claspin (F2), Claspin (F3), Claspin (F4), and Claspin (F5) and USP3 was confirmed by coimmunoprecipitation in HEK293T cells.

Close modal

USP3 deubiquitinates Claspin

To determine whether USP3 deubiquitinates Claspin, HEK293 cells were cotransfected with HA-Claspin and Myc-Ub. Ubiquitinated Claspin was incubated with Flag-USP3 (WT or C168S) or control. As shown in Fig. 7A, USP3-WT, but not USP3-C168S, significantly reduced the ubiquitination of Claspin. In recurrent GBM cells, USP3 knockdown induced the ubiquitination of Claspin (Fig. 7B). Two major forms of polyubiquitin chains are known to be formed through distinct types of linkages (Lys48- or Lys63-linked chains). Lys48-linked ubiquitin chain consistently guides proteolysis by the proteasome (37), but the lysine 63-linked ubiquitination of the substrate protein indicates processing by the endosomal–lysosomal-dependent degradation pathway. Accordingly, we wondered which type of polyubiquitin modifications on Claspin protein was affected by USP3. USP3 significantly disassembled Lys48-linked polyubiquitylation of Claspin, but had no significant effect on the Lys63-linked polyubiquitylation of Claspin (Fig. 7C). Furthermore, we expressed an Lys48-resistant (Lys48R) form of ubiquitin in USP3 inhibition recurrent GBM cells and found that enforced expression of Lys48R ubiquitin reduced USP3 inhibition–induced Claspin downregulation (Fig. 7D). Together, these results demonstrate that USP3 deubiquitinates Claspin.

Figure 7.

USP3 deubiquitinates Claspin. A, HEK293T cells were transfected with HA-Claspin, Myc-Ub, and either wild-type Flag-USP3 (WT) or USP3-C163S. Cells were treated with 20 μmol/L MG132 for 6 hours. B, ShCtrl and sh-USP3 recurrent GBM cells were transfected with Myc-Ub, and cell lysates were subjected to IP with Claspin antibody, followed by IB with antibodies against Myc and Claspin. Cells were treated with 20 μmol/L MG132 for 6 hours. C, Recurrent GBM cells were cotransfected with HA-Claspin, Flag-USP3, and the indicated Myc-Ub Lys0, Lys48-only, or Lys63-only plasmids, and then the Claspin ubiquitylation linkage was analyzed. D, ShCtrl and sh-USP3 recurrent GBM cells transfected with Ub WT or Ub Lys48R were cultured for 72 hours. The resulting products were analyzed by Western blotting using the indicated antibodies. The data are the means of three independent experiments ± SEM. *, P < 0.05. E, Schematic of Smoothened promotes USP3 mediate Claspin deubiquitination, leading to the activation of Claspin-dependent ATR–Chk1 signaling.

Figure 7.

USP3 deubiquitinates Claspin. A, HEK293T cells were transfected with HA-Claspin, Myc-Ub, and either wild-type Flag-USP3 (WT) or USP3-C163S. Cells were treated with 20 μmol/L MG132 for 6 hours. B, ShCtrl and sh-USP3 recurrent GBM cells were transfected with Myc-Ub, and cell lysates were subjected to IP with Claspin antibody, followed by IB with antibodies against Myc and Claspin. Cells were treated with 20 μmol/L MG132 for 6 hours. C, Recurrent GBM cells were cotransfected with HA-Claspin, Flag-USP3, and the indicated Myc-Ub Lys0, Lys48-only, or Lys63-only plasmids, and then the Claspin ubiquitylation linkage was analyzed. D, ShCtrl and sh-USP3 recurrent GBM cells transfected with Ub WT or Ub Lys48R were cultured for 72 hours. The resulting products were analyzed by Western blotting using the indicated antibodies. The data are the means of three independent experiments ± SEM. *, P < 0.05. E, Schematic of Smoothened promotes USP3 mediate Claspin deubiquitination, leading to the activation of Claspin-dependent ATR–Chk1 signaling.

Close modal

GDC-0449 inhibition of Smo induces radiosensitivity to GBM

To leverage our results for clinical application, we used a small-molecule inhibitor of Smo, GDC-0449. GDC-0449 treatment induced recurrent GBM cells sensitive to IR but not to NHAs (Supplementary Fig. S10A). GDC-0449 treatment reduced the protein levels of Smo and Claspin, and induced he protein levels of γ-H2AX after IR in recurrent GBM cells (Supplementary Fig. S10B). GDC-0449 treatment, like Smo knockdown, promoted Claspin ubiquitination and degradation (Supplementary Fig. S10C). In vivo, GDC-0449 with IR treatment recurrent GBM2 cells displayed significantly suppressed tumor growth compared with IR treatment only (Supplementary Fig. S10D).

Radiotherapy is the most widely used modality in the treatment of GBM (29, 30). Nonetheless, GBM radioresistance is the main cause of local failure after radiotherapy (5, 6). Several studies indicated the influence of Hedgehog signaling to promote more aggressive disease and tumor resistance to therapeutics (31, 32). Smo plays a critical role in Hedgehog signaling, which activates Gli transcription factors. Additionally, we previously proved that Smo was a poor prognosis factor and a potential therapeutic target in glioma (33). In this study, we investigated the specific role of Smo in the modulation of GBM radioresistance. First, we detected paired primary and recurrent GBM patients who were treated with radiation and observed that Smo mRNA and protein expression were higher in recurrent GBMs than in primary GBMs. By correlating of Smo expression with survival for GBM patients after radiation treatment, we found that higher Smo expression indicated poor prognosis of GBM patients after radiation treatment. Additionally, recurrent GBM cells show greater resistance to radiation than primary GBM cells. Smo overexpression in primary GBM cells promoted acquired radiation resistance in vitro and in vivo. However, Smo inhibition sensitized recurrent GBM cells to radiation in vitro. These sensitizing effects are the result of the inhibition of DNA damage repair, leading to G2 arrest and apoptosis. These data are further supported by patient-specific orthotopic GBM xenograft models.

ATR has been shown to be responsible for promoting G2 arrest and Claspin degradation turns off the G2 checkpoint (34). Of note, our data suggested G2 arrest after IR in Smo depletion recurrent GBM cells. Considering that Smo, a cell-surface modulator, initiates an upstream signaling but might not solely depend on the Clasping–ATR axis here, we speculated the previously reported Gli1 inhibition (35, 36) may play a role in promoting G2 arrest. Indeed, Smo inhibition decreased the expression of Gli family including both Gli1 and Gli2 in our cell models. When the expression of Gli1 was further regulated in recurrent GBM cells to evaluate the cell cycle, we found a significant G2 arrest after IR in Gli1 depletion recurrent GBM cells. Conversely, G2 arrest restored when we added back Gil1 (Supplementary Fig. S11A and S11B). Therefore, we have identified a mechanism by which the increase G2 arrest was mainly regulated by Gil1. Thus, our findings uncover the importance of Smo in the radiation resistance of GBM and the applicability of Smo as a potential prognostic and predictive marker for radiation sensitivity in GBM.

DNA damage responses are controlled by two main signaling pathways in vertebrates, the ATM–Chk2 and ATR–Chk1 protein kinase pathways (37). ATM and ATR are two serine/threonine kinase members of a phosphoinositide 3-kinase–like family (38, 39). Chk1 and Chk2, which are selectively phosphorylated and activated by ATR and ATM, respectively, trigger a wide range of distinct downstream responses (40). Activated ATR–Chk1 and ATM–Chk2 regulate DNA damage repair and induce apoptosis (12). The present study showed that Smo knockdown could disrupt the activation of ATR/Chk1 but not ATM–Chk2 signaling in response to radiation-induced DNA damage, leading to inhibited DNA damage repair and increased apoptosis. Claspin is a critical protein for the ATR-dependent phosphorylation of Chk1. Previous studies have shown that Claspin is associated with DNA damage in human cells (41, 42). Claspin initiates DNA replication in human cells, while Claspin downregulation upregulates DNA damage (43). In this study, Smo mediated Claspin degradation via UPS-mediated proteolysis. We demonstrated that Smo knockdown could promote radiation-induced Claspin polyubiquitination and proteasomal degradation, resulting in Chk1 inactivation and increased radiosensitivity to GBM. However, Smo overexpression could decrease radiation-induced Claspin polyubiquitination and proteasomal degradation. Our findings demonstrate the connection between Smo and the regulation of Claspin proteolysis.

Smo is an important component of the Hh pathway, which triggers the activation of Gli transcription factors and then regulates downstream target genes (9). Protein ubiquitination is tightly controlled by ubiquitin ligases or their removal through deubiquitinating enzymes (44, 45). Deubiquitinases sever ubiquitin from substrates and stop ubiquitin-dependent signaling, and they are now known to be essential and specific components of the ubiquitin–proteasome pathway (46, 47). In the present study, we first determined the potential DUBs that could mediate Claspin degradation, and we selected 5 DUBs from the entire DUB family. A recent study reported that Gli 1 could promote USP48 transcription (23). Similar to Gli1, Gli2 is a key transcriptional target of Hh signaling. Indeed, the present study showed that Smo induced only USP3 expression through transactivation by Gli2. USP3 is a member of the DUB family that has been reported to regulate various cellular events through deubiquitinating various target proteins, such as RIG-I and p53 (17, 48). Here, we identify that USP3 removes ubiquitin conjugates from Claspin, which is turn suppresses Claspin's proteasome degradation. Ectopic expression of USP3 attenuated Claspin polyubiquitination and proteasomal degradation. Conversely, USP3 knockdown induces Claspin polyubiquitination and proteasomal degradation. More importantly, USP3 contains two domains: a ZnF domain and a UCH domain. The results show that only the UCH domain of USP3 can interact with Claspin but not the ZnF domain of USP3. Moreover, the N-terminus of Claspin (1–330) mediates the interaction with USP3. Thus, our findings show that Smo promotes radiation resistance in GBM through USP3-induced Claspin-dependent ATR–Chk1 signaling (Fig. 7E). GDC-0449 is an inhibitor of Smo and has entered clinical trial for glioma. In this study, we used GDC-0449 to assess the impact of pharmacologic inhibition of Smo in GBM. We found that GDC-0449 could promote Claspin ubiquitination and degradation, which lead to recurrent GBM cells sensitive to IR.

In summary, our study revealed that Smo is critical in modulating radiation resistance in GBMs. Smo regulates USP3 expression through transactivation by Gli2, leading to the activation of Claspin-dependent ATR–Chk1 signaling. Our results reveal a novel mechanism underlying acquired radiation resistance in GBM and have important implications in the development of treatment strategies for radiation-resistant GBMs.

Statistical analysis

All experiments were performed in triplicate. The data were subjected to Student t test for pairwise comparison or ANOVA for multivariate analysis. Differences were considered to be statistically significant at P < 0.05.

No potential conflicts of interest were disclosed.

All the data set and materials generated and/or analyzed during the current study are available.

Conception and design: Y. Tu, Y. Chen, J. Ji

Development of methodology: Y. Tu, Z. Chen, G. Sun, H. Chao, L. Fan, Y. You, J. Ji

Acquisition of data (provided animals, acquired and managed patients, provided facilities, etc.): Y. Tu, Z. Chen, C. Li, Y. You, J. Ji

Analysis and interpretation of data (e.g., statistical analysis, biostatistics, computational analysis): Y. Tu, Z. Chen, Z. Bao, Y. You, J. Ji

Writing, review, and/or revision of the manuscript: Y. Tu, H. Chao, Y. Qu, Y. Chen, J. Ji

Administrative, technical, or material support (i.e., reporting or organizing data, constructing databases): Y. Tu, J. Ji

Study supervision: P. Zhao, J. Ji

This study was supported by grants from the National Natural Science Foundation of China (81972153, 81772602, 91742105, and 91942309), Jiangsu Province's Natural Science Foundation (BK20160047), Jiangsu Province's Key Discipline of Medicine (XK201117), Jiangsu Province Medical Key Talent (ZDRCB2016002), Jiangsu Provincial Key Research Development Program of China (BE2018750), and the Priority Academic Program Development of Jiangsu Higher Education Institution.

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.
Diplas
BH
,
He
X
,
Brosnan-Cashman
JA
,
Liu
H
,
Chen
LH
,
Wang
Z
, et al
The genomic landscape of TERT promoter wildtype-IDH wildtype glioblastoma
.
Nat Commun
2018
;
9
:
2087
.
2.
Pavlyukov
MS
,
Yu
H
,
Bastola
S
,
Minata
M
,
Shender
VO
,
Lee
Y
, et al
Apoptotic cell-derived extracellular vesicles promote malignancy of glioblastoma via intercellular transfer of splicing factors
.
Cancer Cell
2018
;
34
:
119
35
.
3.
Benitez
JA
,
Ma
J
,
D'Antonio
M
,
Boyer
A
,
Camargo
MF
,
Zanca
C
, et al
Publisher correction: PTEN regulates glioblastoma oncogenesis through chromatin-associated complexes of DAXX and histone H3.3
.
Nat Commun
2018
;
9
:
16217
.
4.
Sanai
N
,
Li
J
,
Boerner
J
,
Stark
K
,
Wu
J
,
Kim
S
, et al
Phase 0 trial of AZD1775 in first-recurrence glioblastoma patients
.
Clin Cancer Res
2018
;
24
:
3820
8
.
5.
Wahba
A
,
Rath
BH
,
O'Neill
JW
,
Camphausen
K
,
Tofilon
PJ
. 
The XPO1 inhibitor selinexor inhibits translation and enhances the radiosensitivity of glioblastoma cells grown in vitro and in vivo
.
Mol Cancer Ther
2018
;
17
:
1717
26
.
6.
De Bacco
F
,
D'Ambrosio
A
,
Casanova
E
,
Orzan
F
,
Neggia
R
,
Albano
R
, et al
MET inhibition overcomes radiation resistance of glioblastoma stem-like cells
.
EMBO Mol Med
2016
;
8
:
550
68
.
7.
Qi
X
,
Schmiege
P
,
Coutavas
E
,
Wang
J
,
Li
X
. 
Structures of human patched and its complex with native palmitoylated sonic hedgehog
.
Nature
2018
;
560
:
128
32
.
8.
Bal
E
,
Park
HS
,
Belaid-Choucair
Z
,
Kayserili
H
,
Naville
M
,
Madrange
M
, et al
Mutations in ACTRT1 and its enhancer RNA elements lead to aberrant activation of Hedgehog signaling in inherited and sporadic basal cell carcinomas
.
Nat Med
2017
;
23
:
1226
33
.
9.
Raleigh
DR
,
Reiter
JF
. 
Misactivation of hedgehog signaling causes inherited and sporadic cancers
.
J Clin Invest
2019
;
129
:
465
75
.
10.
Grausam
KB
,
Dooyema
SDR
,
Bihannic
L
,
Premathilake
H
,
Morrissy
AS
,
Forget
A
, et al
ATOH1 promotes leptomeningeal dissemination and metastasis of sonic hedgehog subgroup medulloblastomas
.
Cancer Res
2017
;
77
:
3766
77
.
11.
Rodriguez-Bravo
V
,
Guaita-Esteruelas
S
,
Florensa
R
,
Bachs
O
,
Agell
N
. 
Chk1- and claspin-dependent but ATR/ATM- and Rad17-independent DNA replication checkpoint response in HeLa cells
.
Cancer Res
2006
;
66
:
8672
9
.
12.
Smith
J
,
Tho
LM
,
Xu
N
,
Gillespie
DA
. 
The ATM-Chk2 and ATR-Chk1 pathways in DNA damage signaling and cancer
.
Adv Cancer Res
2010
;
108
:
73
112
.
13.
Peng
C
,
Chen
Z
,
Wang
S
,
Wang
HW
,
Qiu
W
,
Zhao
L
, et al
The error-prone DNA polymerase kappa promotes temozolomide resistance in glioblastoma through Rad17-dependent activation of ATR-Chk1 signaling
.
Cancer Res
2016
;
76
:
2340
53
.
14.
Cimprich
KA
,
Cortez
D
. 
ATR: an essential regulator of genome integrity
.
Nat Rev Mol Cell Biol
2008
;
9
:
616
27
.
15.
Jin
X
,
Yan
Y
,
Wang
D
,
Ding
D
,
Ma
T
,
Ye
Z
, et al
DUB3 promotes BET inhibitor resistance and cancer progression by deubiquitinating BRD4
.
Mol Cell
2018
;
71
:
592
605
.
16.
Rawat
R
,
Starczynowski
DT
,
Ntziachristos
P
. 
Nuclear deubiquitination in the spotlight: the multifaceted nature of USP7 biology in disease
.
Curr Opin Cell Biol
2019
;
58
:
85
94
.
17.
Cui
J
,
Song
Y
,
Li
Y
,
Zhu
Q
,
Tan
P
,
Qin
Y
, et al
USP3 inhibits type I interferon signaling by deubiquitinating RIG-I-like receptors
.
Cell Res
2014
;
24
:
400
16
.
18.
Cheng
YC
,
Shieh
SY
. 
Deubiquitinating enzyme USP3 controls CHK1 chromatin association and activation
.
Proc Natl Acad Sci U S A
2018
;
115
:
5546
51
.
19.
Chen
Z
,
Chen
X
,
Chen
P
,
Yu
S
,
Nie
F
,
Lu
B
, et al
Long non-coding RNA SNHG20 promotes non-small cell lung cancer cell proliferation and migration by epigenetically silencing of P21 expression
.
Cell Death Dis
2017
;
8
:
e3092
.
20.
Bao
S
,
Wu
Q
,
McLendon
RE
,
Hao
Y
,
Shi
Q
,
Hjelmeland
AB
, et al
Glioma stem cells promote radioresistance by preferential activation of the DNA damage response
.
Nature
2006
;
444
:
756
60
.
21.
Liu
X
,
Chong
Y
,
Tu
Y
,
Liu
N
,
Yue
C
,
Qi
Z
, et al
CRM1/XPO1 is associated with clinical outcome in glioma and represents a therapeutic target by perturbing multiple core pathways
.
J Hematol Oncol
2016
;
9
:
108
.
22.
Zheng
H
,
Shen
M
,
Zha
YL
,
Li
W
,
Wei
Y
,
Blanco
MA
, et al
PKD1 phosphorylation-dependent degradation of SNAIL by SCF-FBXO11 regulates epithelial-mesenchymal transition and metastasis
.
Cancer Cell
2014
;
26
:
358
73
.
23.
Zhou
A
,
Lin
K
,
Zhang
S
,
Ma
L
,
Xue
J
,
Morris
SA
, et al
Gli1-induced deubiquitinase USP48 aids glioblastoma tumorigenesis by stabilizing Gli1
.
EMBO Rep
2017
;
18
:
1318
30
.
24.
Rivera
AL
,
Pelloski
CE
,
Gilbert
MR
,
Colman
H
,
De La Cruz
C
,
Sulman
EP
, et al
MGMT promoter methylation is predictive of response to radiotherapy and prognostic in the absence of adjuvant alkylating chemotherapy for glioblastoma
.
Neuro Oncol
2010
;
12
:
116
21
.
25.
Gersch
M
,
Wagstaff
JL
,
Toms
AV
,
Graves
B
,
Freund
SMV
,
Komander
D
. 
Distinct USP25 and USP28 oligomerization states regulate deubiquitinating activity
.
Mol Cell
2019
;
74
:
436
51
.
26.
Faustrup
H
,
Bekker-Jensen
S
,
Bartek
J
,
Lukas
J
,
Mailand
N
. 
USP7 counteracts SCFbetaTrCP- but not APCCdh1-mediated proteolysis of Claspin
.
J Cell Biol
2009
;
184
:
13
9
.
27.
Martin
Y
,
Cabrera
E
,
Amoedo
H
,
Hernandez-Perez
S
,
Dominguez-Kelly
R
,
Freire
R
. 
USP29 controls the stability of checkpoint adaptor Claspin by deubiquitination
.
Oncogene
2015
;
34
:
1058
63
.
28.
Zhu
M
,
Zhao
H
,
Liao
J
,
Xu
X
. 
HERC2/USP20 coordinates CHK1 activation by modulating CLASPIN stability
.
Nucleic Acids Res
2014
;
42
:
13074
81
.
29.
Ma
J
,
Benitez
JA
,
Li
J
,
Miki
S
,
Ponte de Albuquerque
C
,
Galatro
T
, et al
Inhibition of nuclear PTEN tyrosine phosphorylation enhances glioma radiation sensitivity through attenuated DNA repair
.
Cancer Cell
2019
;
35
:
504
18
.
30.
Zhang
L
,
Cheng
F
,
Wei
Y
,
Zhang
L
,
Guo
D
,
Wang
B
, et al
Inhibition of TAZ contributes radiation-induced senescence and growth arrest in glioma cells
.
Oncogene
2019
;
38
:
2788
99
.
31.
Kudo
K
,
Gavin
E
,
Das
S
,
Amable
L
,
Shevde
LA
,
Reed
E
. 
Inhibition of Gli1 results in altered c-Jun activation, inhibition of cisplatin-induced upregulation of ERCC1, XPD and XRCC1, and inhibition of platinum-DNA adduct repair
.
Oncogene
2012
;
31
:
4718
24
.
32.
Mazumdar
T
,
DeVecchio
J
,
Shi
T
,
Jones
J
,
Agyeman
A
,
Houghton
JA
. 
Hedgehog signaling drives cellular survival in human colon carcinoma cells
.
Cancer Res
2011
;
71
:
1092
102
.
33.
Tu
Y
,
Niu
M
,
Xie
P
,
Yue
C
,
Liu
N
,
Qi
Z
, et al
Smoothened is a poor prognosis factor and a potential therapeutic target in glioma
.
Sci Rep
2017
;
7
:
42630
.
34.
Gewurz
BE
,
Harper
JW
. 
DNA-damage control: Claspin destruction turns off the checkpoint
.
Curr Biol
2006
;
16
:
R932
4
.
35.
Zinke
J
,
Schneider
FT
,
Harter
PN
,
Thom
S
,
Ziegler
N
,
Toftgard
R
, et al
beta-Catenin-Gli1 interaction regulates proliferation and tumor growth in medulloblastoma
.
Mol Cancer
2015
;
14
:
17
.
36.
Sun
Y
,
Guo
W
,
Ren
T
,
Liang
W
,
Zhou
W
,
Lu
Q
, et al
Gli1 inhibition suppressed cell growth and cell cycle progression and induced apoptosis as well as autophagy depending on ERK1/2 activity in human chondrosarcoma cells
.
Cell Death Dis
2014
;
5
:
e979
.
37.
Wang
M
,
Guo
L
,
Wu
Q
,
Zeng
T
,
Lin
Q
,
Qiao
Y
, et al
ATR/Chk1/Smurf1 pathway determines cell fate after DNA damage by controlling RhoB abundance
.
Nat Commun
2014
;
5
:
4901
.
38.
Cheng
A
,
Zhao
T
,
Tse
KH
,
Chow
HM
,
Cui
Y
,
Jiang
L
, et al
ATM and ATR play complementary roles in the behavior of excitatory and inhibitory vesicle populations
.
Proc Natl Acad Sci U S A
2018
;
115
:
E292
E301
.
39.
Munk
S
,
Sigurethsson
JO
,
Xiao
Z
,
Batth
TS
,
Franciosa
G
,
von Stechow
L
, et al
Proteomics reveals global regulation of protein SUMOylation by ATM and ATR kinases during replication stress
.
Cell Rep
2017
;
21
:
546
58
.
40.
Wang
WJ
,
Wu
SP
,
Liu
JB
,
Shi
YS
,
Huang
X
,
Zhang
QB
, et al
MYC regulation of CHK1 and CHK2 promotes radioresistance in a stem cell-like population of nasopharyngeal carcinoma cells
.
Cancer Res
2013
;
73
:
1219
31
.
41.
Cabrera
E
,
Hernandez-Perez
S
,
Koundrioukoff
S
,
Debatisse
M
,
Kim
D
,
Smolka
MB
, et al
PERK inhibits DNA replication during the unfolded protein response via claspin and Chk1
.
Oncogene
2017
;
36
:
678
86
.
42.
McGarry
E
,
Gaboriau
D
,
Rainey
MD
,
Restuccia
U
,
Bachi
A
,
Santocanale
C
. 
The deubiquitinase USP9X maintains DNA replication fork stability and DNA damage checkpoint responses by regulating CLASPIN during S-phase
.
Cancer Res
2016
;
76
:
2384
93
.
43.
Bianco
JN
,
Bergoglio
V
,
Lin
YL
,
Pillaire
MJ
,
Schmitz
AL
,
Gilhodes
J
, et al
Overexpression of claspin and timeless protects cancer cells from replication stress in a checkpoint-independent manner
.
Nat Commun
2019
;
10
:
910
.
44.
Gong
P
,
Davidson
GA
,
Gui
W
,
Yang
K
,
Bozza
WP
,
Zhuang
Z
. 
Activity-based ubiquitin-protein probes reveal target protein specificity of deubiquitinating enzymes
.
Chem Sci
2018
;
9
:
7859
65
.
45.
Jeong
JS
,
Jung
C
,
Seo
JS
,
Kim
JK
,
Chua
NH
. 
The Deubiquitinating enzymes UBP12 and UBP13 positively regulate MYC2 levels in jasmonate responses
.
Plant Cell
2017
;
29
:
1406
24
.
46.
Hutchins
AP
,
Liu
S
,
Diez
D
,
Miranda-Saavedra
D
. 
The repertoires of ubiquitinating and deubiquitinating enzymes in eukaryotic genomes
.
Mol Biol Evol
2013
;
30
:
1172
87
.
47.
Reyes-Turcu
FE
,
Ventii
KH
,
Wilkinson
KD
. 
Regulation and cellular roles of ubiquitin-specific deubiquitinating enzymes
.
Annu Rev Biochem
2009
;
78
:
363
97
.
48.
Fu
S
,
Shao
S
,
Wang
L
,
Liu
H
,
Hou
H
,
Wang
Y
, et al
USP3 stabilizes p53 protein through its deubiquitinase activity
.
Biochem Biophys Res Commun
2017
;
492
:
178
83
.