Abstract
LGALS3 promotes tumor progression in diverse cancers. However, the involvement of LGALS3 in glioblastoma has not yet been broadly illuminated.
Microarray was performed to detect the gene expression profiles of radioresistance in T98G cells and identified a universally upregulated gene, LGALS3. The impact of LGALS3 on the survival of glioblastoma cells facing ionizing irradiation or temozolomide was investigated by the Cell Counting Kit-8 (CCK-8). A total of 120 glioblastoma cases were collected to analyze the relationship between LGALS3 expression and patient prognosis. Another 961 patients with glioma and 1,351 healthy controls were recruited to study the association of SNPs across the LGALS3 gene with glioblastoma susceptibility. The functional SNP sites were also studied in cellular experiments.
An effective protection of LGALS3 from ionizing irradiation or temozolomide-induced cell death in T98G and U251 cells was found. In addition, high expression of LGALS3 could work as an independent risk factor for survival of patients with glioblastoma. Two SNP sites (rs4644 and rs4652) across the LGALS3 gene were associated with increased risk for glioblastoma, and the C allele of rs4652 and the A allele of rs4644 could enhance glioblastoma resistance to radio-chemotherapy, but not cell proliferation.
Our results suggest that LGALS3 is an important biomarker influencing glioblastoma risk and prognosis and a potential target for treating the malignancy, especially ones with resistance against the standard therapy.
LGALS3 promotes glioblastoma cells' resistance to ionizing irradiation and temozolomide and predicts poor prognosis. Targeting LGALS3 may limit the therapeutic resistance in glioblastoma and increase patient survival.
Introduction
Gliomas represent the most frequent primary tumors in the central nervous system with glioblastoma (WHO grade 4), which is the most common and malignant histologic subtype (1). Despite the use of multimodal treatments, consisting of maximum surgical resection followed by radiotherapy and temozolomide chemotherapy, patients with glioblastoma have to follow a fatal course with the median survival of approximately 15 months (1). Moreover, targeted therapies, such as temsirolimus and bevacizumab (2, 3), against molecules that drive the growth of the bulk of glioblastoma have so far also been unsuccessful in clinical trials. The failure of tumor control at the primary site is caused, in great part, by the infiltrative nature of glioblastoma and acquired resistance, which makes tumor cells survive initial therapies and contribute to inevitable tumor relapse (4). Therefore, it is imperative to better elucidate the complicated pathobiology of glioblastoma and understand the molecular details underlying the response of tumor cells to genotoxins, both of which will promote to design more effective therapeutic strategies.
Galectins are a family of proteins characterized by their ability to bind β-galactoside–containing glycoconjugates through a conserved carbohydrate-recognition domain (CRD) and involved in many physiologic events, such as cell proliferation and differentiation, apoptosis, and immune response (5, 6). Moreover, dysfunction of galectins has been found to promote cancer development and progression (7). LGALS3 (also known as galectin-3) is the only chimera-type galectin with a single CRD connected to a long and flexible N-terminal domain (8), and also is one of the most extensively studied galectins in cancer. For instance, LGALS3 expression in gastric cancer increases tumor cell motility and confers an aggressive phenotype (9, 10), and contributes to metastasis in lung cancer (11), whereas the removal of LGALS3 increases the susceptibility of diffuse large B-cell lymphoma cells to cell death (12), and LGALS3 deficiency reduces cell proliferation and migration in hepatocellular carcinoma (13). However, the loss of galectin-3 has been shown to be associated with the epithelial–mesenchymal transition, increased sphere-formation ability, and drug resistance in some breast cancers (14), which is in contradiction with previous results that LGALS3 could enhance the metastatic potential of breast cancer cells (15). All indicate the multifaceted functions of LGALS3 in tumor biology. In addition, the alterations of LGALS3 expression also appear heterogeneous in cancers that most tumors of the digestive tract and urinary system display upregulation of LGALS3, but tumors of the reproductive system present decreased LGALS3 expression (5). In glioma, most of the studies related to LGALS3 have shown its high levels in astrocytic tumors, whose pathologic grade was significantly associated with the expression level of LGALS3 (16–18). Moreover, LGALS3 is also being reported to influence the glioma cell migration (19) and promote cell survival under hypoxic and nutrient-deprived conditions (20). However, the involvement of LGALS3 in the chemo- and/or radioresistance in glioblastoma has not yet been demonstrated.
Genome-wide association studies (GWAS) offer a powerful means to search for genes that confer susceptibility to complex diseases (21). More recently, this tool has identified SNPs at several loci with increased risk for glioma, such as 17p13.1 (TP53), 5p15.33 (TERT), 7p11.2 (EGFR), 9p21.3 (CDKN2B-AS1), and 20q13.33 (RTEL1; ref. 22), which has advanced our understandings of etiology in the malignancy. Previously, two genetic variants at LGALS3 SNP sites (rs4644 and rs4652) were reported to contribute to the susceptibility of glioma, and the patients carrying AA genotype of LGALS3 + 292A>C (rs4652) had marked shorter overall survival period than those did not (23, 24). However, other SNPs located in the LGALS3 gene were not included in those studies and the effect of SNPs on glioblastoma progression had not yet been assessed.
Here, we applied microarrays to detect the expression profile of radioresistant glioblastoma cells and identified LGALS3 as a potential effector. We further demonstrated LGALS3-mediating glioblastoma resistance to radio- and chemotherapy and promoting tumor cell proliferation. Meanwhile, we confirmed that LGALS3 expression was upregulated in glioblastoma and associated with poor prognosis of patients, and illustrated that SNP sites (rs4644 and rs4652) were related to increased risk in glioblastoma, and the C allele of rs4652 and the A allele of rs4644 could promote glioblastoma resistance to radio-chemotherapy. These findings highlight the importance of LGALS3 in glioblastoma treatment resistance and its potential to be a therapeutic target for this tumor.
Materials and Methods
Study participants and clinical samples
The study protocol and acquisition of tissue specimens and blood samples were approved by both of the Ethical Review Committee of Second Military Medical University (Shanghai, China) and the Ethics Board of Fudan University (Shanghai, China). Written informed consent was supplied by all the participants. For the IHC-staining experiment, glioma specimens were obtained from the patients who underwent surgical treatment at Changzheng Hospital (Shanghai, China) from January 2000 to December 2010. Tumor histology of glioma was confirmed independently by two neuropathologists. Normal brain tissues were taken from severe trauma patients, on whom partial resection of the brain was required to reduce the increased intracranial pressure. For the GWAS, 961 newly diagnosed glioma patients and 1,351 healthy controls were consecutively enrolled between October 2004 and July 2009 from Huashan Hospital (Shanghai, China). Detailed recruitment criteria were defined previously (25). Each participant voluntarily donated peripheral whole blood sample of 3–5 mL for DNA extraction and genotyping.
Cell culture and lentivirus infection
Hek293T cells, the human glioblastoma cell lines (T98G and U251), and the human lung cancer cell line (A549) were obtained from the ATCC in 2013. These cell lines were last authenticated in the laboratory by short tandem repeat profiling based on eight markers in October 2017. Cells were cryopreserved and used within 6 months after thawing, not exceeding 10 passages. All cells were maintained in a humidified atmosphere of 5% CO2 at 37°C and cultured in DMEM supplemented with 10% FBS. The procedures of packaging and infecting with lentivirus were done according to our previous study (26).
DNA constructs and mutagenesis
High-fidelity enzyme KOD Plus Neo (Toyobo) was used for PCR amplification for the construction. With the primers (F, GCTCTAGAGCCACCATGGCAGACAATTTTTCGCTC; R, CGGGATCCTTAAGTAGATGACTTACAAGAGGGATAT), PCR-amplified human wild-type LGALS3 (+191C and +292A) was cloned into the XbaI-BamHI sites of the CD513B vector. Mutant LGALS3 (+191A and +292C) was made using the QuickChange site-directed mutagenesis Kit (Stratagene).
To efficiently knock down the LGALS3 expression, shRNA-4 and shRNA-5, generated with 5′-GGAAGAAAGACAGTCGGT-3′ and 5′-GCTCACTTGTTGCAGTACAAT-3′ targeting the LGALS3 transcript respectively, were selected (Fig. 1A). Corresponding sense and antisense oligonucleotides were synthesized, annealed, and cloned into the HpaI-XhoI sites of the PLL3.7 vector.
Impact of LGALS3 on cell survival under DNA damage stresses induced by radiation and temozolomide treatment. A, Detection of LGALS3 expression in T98G cells transfected with different vectors by Western blot analysis. B and C, When irradiated with 5 Gy or 10 Gy, T98G and U251 cells with overexpressed LGALS3 show higher survival rate, whereas the survival ability of LGALS3-depleted cells decreases significantly compared with the control. D and E, Treated with temozolomide (TMZ), both of the LGALS3-overexpressing T98G and U251 cells survive better than control cells at levels of 500, 1,000, 1,500, and 2,000 μmol/L, but an opposite result is observed in the cells with depleted LGALS3. F and G, LGALS3 significantly promotes T98G and U251 cell proliferation, which can be reversed by shRNAs targeting LGALS3. (★, P <0.05; CD513B and PLL3.7 are the controls).
Impact of LGALS3 on cell survival under DNA damage stresses induced by radiation and temozolomide treatment. A, Detection of LGALS3 expression in T98G cells transfected with different vectors by Western blot analysis. B and C, When irradiated with 5 Gy or 10 Gy, T98G and U251 cells with overexpressed LGALS3 show higher survival rate, whereas the survival ability of LGALS3-depleted cells decreases significantly compared with the control. D and E, Treated with temozolomide (TMZ), both of the LGALS3-overexpressing T98G and U251 cells survive better than control cells at levels of 500, 1,000, 1,500, and 2,000 μmol/L, but an opposite result is observed in the cells with depleted LGALS3. F and G, LGALS3 significantly promotes T98G and U251 cell proliferation, which can be reversed by shRNAs targeting LGALS3. (★, P <0.05; CD513B and PLL3.7 are the controls).
Ionizing irradiation and temozolomide treatment
Radioresistant T98G and A549 cells were established and maintained in our laboratory. Briefly, the radioresistant sublines (T98G-R and A549-R) were generated by repetitive pulse exposure of T98G glioblastoma cells and A549 lung cancer cells to ionizing irradiation using Gammacell-40137Cs γ (MDS Nordion) for 2 Gy (0.8 Gy/minute) each time. Every other day, the undead T98G-R and A549-R cells were passaged and irradiated. The cycle was repeated for 2 months. When exposed to a total radiation dose of 60 Gy, the live T98G-R cells were seeded at 6-cm dish for monoclonal growth, all of which, as well as the rest survival T98G-R cells and 549-R cells, would then be harvested for RNA extraction. Control T98G and A549 cells were given the same treatments without ionizing irradiation. For experiments verifying the function of LGALS3 in glioblastoma cell radioresistance, T98G and U251 received irradiation with 5 Gy or 10 Gy one time. To detect the effect of LGALS3 on temozolomide resistance, T98G and U251 cells were cultured in the presence of 0, 100, 500, 1,000, 1,500, or 2,000 μmol/L of temozolomide (provided by Merck Sharp & Dohme Ltd) and growing for 30 hours prior to assays.
RNA extraction and qRT-PCR
The total RNA was extracted from cell lines using TRIzol Reagent (Invitrogen) following the manufacturer's protocol. First-strand cDNA was synthesized using ReverTra Ace (Toyobo). The qRT-PCR was conducted by ABI PRISM 7900HT Sequence Detection system in the presence of SYBRGreen dye (Toyobo). Each sample was tested in triplicate. GAPDH was used as an internal control. All PCR primers used are as follow: LGALS3 forward primer, GGCCACTGATTGTGCCTTAT and reverse primer, GAAGCGTGGGTTAAAGTGGA; GAPDH forward primer, GCGAGATCCCTCCAAAATCAA and reverse primer, GTTCACACCCATGACGAACAT.
Western blot analysis
The cell samples were homogenized in lysis buffer containing a cocktail of 1% protease and 1% protease inhibitors (Sigma). Protein lysates were separated by 10% SDS-PAGE and transferred onto a PVDF membrane (Invitrogen), which were blocked in 5% milk for 1 hour and then treated with the appropriate antibody at 4°C overnight. The blots were visualized by an enhanced chemiluminescence detection system (Thermo Fisher Scientific). GAPDH was used as an internal control.
mRNA microarray
T98G and A549 cells treated with ionizing irradiation (60 Gy) and their respective control cells were harvested for RNA sample preparation. Then, the microarray hybridization was performed at KangChen Biotech Corporation by a HumanHT-12 v4 Expression BeadChip (Illumina). Data acquisition and analysis followed the standard Illumina Custom protocol. mRNAs with diffscore value >13 or <−13 were considered to be differentially expressed between radioresistance group and control.
IHC
The human specimens were fixed in formalin and then embedded in paraffin wax. The IHC assay was performed on these tissues to detect the LGALS3 expression and IDH1 mutation by methods described previously (26). LGALS3 antibody for IHC was obtained from Cell Signaling Technology (1:600, #87985). An anti-IDH1 R132H mAb (1:20, Dianova, DIA-H09) was used to detect the IDH1 status in glioma specimens that were recruited in the IHC assays. The intensity of positive-staining cytoplasm or nucleus of LGALS3 was scored on a scale of 0 to 3 (0: negative; 1: light brown; 2: medium brown; 3: dark brown), which was multiplied by the value of positive percentage in cytoplasm and nucleus, respectively. Then, the two scores of cellular cytoplasm and nucleus were added to estimate the expression level of LGALS3, which was divided into the high or low cut by the median total score 3. The status of IDH1 was directly interpreted as null or mutation, based on whether having a positive IHC staining.
The Cancer Genome Atlas data processing
The Cancer Genome Atlas (TCGA; www.cancergenome.nih.gov) provides multimodal data of more than 500 glioblastoma cases. Level 3 mRNA expression data, based on the Affymetrix microarrays (Human gene U133A) and Agilent microarrays (AgilentG4502A), clinical follow-up information, and molecular pathologic characteristics, were collected for further analysis. The expression of LGALS3 expression was classified as either high (expression value ≥11.595 or −0.70) or low (expression value <11.595 or −0.70), according to the data from the two types of microarrays, respectively.
Cell proliferation assay
Cells in the logarithmic phase of growth were seeded in 96-well plates (2,000 cells/well) in sextuple and allowed to grow for 1, 2, 3, 4, and 5 days. Cell proliferation assay was analyzed by the Cell Counting Kit-8 (CCK-8; Dojindo) following the manufacturer's instructions. The optical density was measured at 450 nm wavelength using a microplate reader (EL × 800, BIO-TEK). Data were obtained from three independent assays performed in triplicate.
Tag SNP and genotyping
Tag SNPs at the LGALS3 locus were selected from the genotype data for Han Chinese in Beijing, China (CHB) from the International HapMap Project (http://www.hapmap.ncbi.nlm.nih.gov) using the software Haploview V.4.0 (http://www.broad.mit.edu/mpg/haploview) to capture SNPs with a minimum minor allele frequency (MAF) of 0.05 in LGALS3 gene ± 2 kb with an r2 of 0.80 or greater.
Genomic DNA was extracted from venous blood using the Qiagen Blood Kit (Qiagen), and diluted to a final concentration of 15–20 ng/μL. Primers for amplification and extension reactions were designed using the Mass-ARRAY Assay Design Version 3.1 software (Sequenom). Genotyping was performed with the Mass-ARRAY iPLEX platform (Sequenom) using an allele-specific MALDI-TOF mass spectrometry assay (27). Then, genotyping quality was examined via a detailed quality control procedure that ensured a >99% successful call rate with duplicate calling of genotypes, internal positive control samples, and Hardy–Weinberg equilibrium (HWE) testing.
Statistical analysis
The significance of the differences between groups with continuous data was evaluated by the Student t test. Data were presented as the mean ± SD unless otherwise indicated. Differentially expressed genes (DEG) detected in microarray were calculated by the Illumina Custom algorism, as the Diffscore of each DEG was more than 13 or less than −13 between radioresistant T98G cells and controls with P < 0.05. Wilcoxon rank-sum test was used to compare the difference of counting data between two groups. Spearman analysis was used to examine the relationship between LGALS3 expression and glioma grade. Progression-free survival (PFS) and overall survival (OS) curves were plotted by the Kaplan–Meier method and compared by log-rank test. The identification of relevant prognostic factors was performed by univariate analysis, multivariate analysis, and stepwise backward Cox regression model. Variables with a value of P ≤ 0.2 on univariate analysis were added into the multivariate analysis. OS and PFS were calculated in days from the date of diagnosis to the time of death and to the time of tumor progression or recurrence, or death of the patient from glioblastoma, respectively. HWE in control subjects was assessed by the standard χ2 test (df = 1) for each SNP. Unconditional logistic regression was conducted to calculate ORs and 95% confidence intervals (95% CI) as estimates of relative risk for the single locus genotypes, adjusting for age and sex. SPSS software (SPSS) was used for all the statistical analyses. All statistical tests were two-tailed, the results were considered statistically significant when the P values were less than 0.05.
Results
Microarray identifies LGALS3 related to irradiation resistance in glioblastoma cells
To investigate the genes involved in resistance of radiotherapy of glioblastoma, T98G cell line was chosen to be irradiated with 60 Gy. After irradiation, viable T98G cells were monoclonally selected into eight positive groups for culture, which was named as T98-1a, -1n, -2d, -2h, -3c, -4b, -4d, and -4p, respectively. The rest and the unselected T98G cells were mixed cultured and named as T98G-4th. Irradiation of lung cancer cell A549 (A549R) with the same dose was performed as a parallel experiment. As indicated by microarray analysis, a total of 4307 DEGs (Supplementary Table S1) were found in all eight monoclonal groups comparing with the T98G control (Fig. 2A). Only 3 DEGs occurred in all eight groups, among which five and more groups contained 68 same DEGs.
Identification of LGALS3 from differentially expressed genes in radioresistant T98G cells. A, Different clones of radioresistant T98G and A549 cells show different numbers of differentially expressed genes. B and C, Kaplan–Meier analysis finds the prognostic value of LGALS3 using the data from both types of microarray in TCGA glioblastoma dataset. D, qPCR detection confirms the expression change of LGALS3 in different clones of radioresistant T98G and A549 cells.
Identification of LGALS3 from differentially expressed genes in radioresistant T98G cells. A, Different clones of radioresistant T98G and A549 cells show different numbers of differentially expressed genes. B and C, Kaplan–Meier analysis finds the prognostic value of LGALS3 using the data from both types of microarray in TCGA glioblastoma dataset. D, qPCR detection confirms the expression change of LGALS3 in different clones of radioresistant T98G and A549 cells.
Furthermore, TCGA dataset was introduced to help filter out both the radiation and patient prognosis–related DEGs in the 68 ones, of which S100A, BMP2, LGALS3 (Fig. 2B and C), and NNMT genes stand out by a preliminary screening analysis. Moreover, LGALS3 was then found included in both T98G-4th and A549R DEGs. qPCR further confirmed the consistency of LGALS3 expression detected in the microarray (Fig. 2D). These results revealed that glioblastoma cells might not be resistant to radiation through some common mechanisms, but also escape the radiation-induced death in some specific manners. LGALS3 was potentially involved in a common pathway related to irradiation resistance in glioblastoma cell, as well as other types of cancer cell.
LGALS3 promotes glioblastoma cell resistance to radiotherapy and temozolomide chemotherapy
To confirm the function of LGALS3 in glioblastoma radioresistance, we introduced exogenous LGALS3 expression or interference vectors in T98G and U251 cells to detect the cell survival rate after short-term exposure of ionizing irradiation. As analyzed by CCK-8, we found that both glioblastoma cells with LGALS3 overexpression showed higher survival rate in 5 Gy and 10 Gy groups than that of the control, however, the downregulation of LGALS3 was related to the decreased survival ability of T98G and U251 cells (Fig. 1B and C).
We next examined whether LGALS3 played an important role in resistance to temozolomide therapy. T98G and U251 cells transfected with LGALS3-overexpressing, -depleting, or control vector were treated with a broad range of temozolomide concentrations, and the cell survival was also determined by CCK-8. As shown in Fig. 1D and E, less glioblastoma cell death induced by temozolomide at levels of 500, 1,000, 1,500, and 2,000 μmol/L were observed in LGALS3-overexpressing groups. But, stronger cytotoxic effects on the cell survival were found in T98G and U251 cells with depleted LGALS3 compared with the control cells.
LGALS3 increases glioblastoma cell proliferation
To further address the biological importance of LGALS3 in glioblastoma, we detected whether other malignant phenotypes changed in T98G and U251 cells with stably increased or depleted LGALS3 expression. As determined by CCK-8 assay, both of the LGALS3-overexpressed T98G and U251 cells displayed higher proliferation rate than controls, whereas targeting LGALS3 by shRNA-4 and shRNA-5 could markedly inhibit cell proliferation (Fig. 1F and G). Collectively, these results suggest that LGALS3 may be a critical factor in regulating glioblastoma cell growth.
LGALS3 is upregulated in glioblastoma and serves as a prognostic factor
LGALS3 has been shown to be potentially associated with glioblastoma patient prognosis in TCGA database (Fig. 2; Supplementary Table S2). Thus, we then analyzed the LGALS3 expression and its correlation with patient survival in our cohort including 120 glioblastoma cases (Supplementary Table S3). As seen in Fig. 3, LGALS3 expression was noticeably increased in glioblastoma tissues when comparing with normal brains (3.10 ± 0.12 vs. 1.36 ± 0.19, P < 0.001), and positively correlated with pathologic grade of glioma (P < 0.001). To confirm the value of LGALS3 in the prognostic prediction of patients, univariate survival analysis (Supplementary Table S4) was first performed and found that patients with glioblastoma with low total LGALS3 expression score had a markedly longer OS and PFS than those in the high-score group in our cohort (OS, 10 vs. 18 months, P = 0.001; PFS, 7 vs. 15 months, P < 0.001; Fig. 3E and F). Furthermore, Cox proportional hazards regression model (Table 1) in the absence of MGMT methylation data (due to insufficient tissue for the analysis) identified LGALS3 expression as an independent prognostic factor of both OS (HR) = 1.946, P = 0.001) and PFS (HR = 2.294, P = 0.001) for patients with glioblastoma.
Expression and prognostic value of LGALS3 in glioblastoma. A, Representative staining image of LGALS3 in normal brain. B and C, Representative staining image of LGALS3 in glioblastoma (20×). D, The IHC score of LGALS3 expression in glioblastoma is higher than that in the normal brain, and it is associated with glioma pathologic grade. E and F, Kaplan–Meier analysis shows high expression of LGALS3 related to short PFS and OS of patients with glioblastoma.
Expression and prognostic value of LGALS3 in glioblastoma. A, Representative staining image of LGALS3 in normal brain. B and C, Representative staining image of LGALS3 in glioblastoma (20×). D, The IHC score of LGALS3 expression in glioblastoma is higher than that in the normal brain, and it is associated with glioma pathologic grade. E and F, Kaplan–Meier analysis shows high expression of LGALS3 related to short PFS and OS of patients with glioblastoma.
Multivariate analysis of factors associated with survival and progression of patients with glioblastoma
Survival . | Median survival (months, 95% CI) . | HR (95% CI) . | P . |
---|---|---|---|
OS | |||
LGALS3 (High vs. low) | 10 (8.454–11.546) vs. 18 (13.335–22.665) | 1.946 (1.304–2.904) | 0.001 |
Chemotherapy (yes vs. no) | 14 (15.303–24.688) vs. 9 (7.513–15.901) | 0.626 (0.406–0.967) | 0.035 |
PFS | |||
LGALS3 (High vs. low) | 7 (5.444–8.556) vs. 15 (9.010–20.99) | 2.294 (1.504–3.499) | 0.001 |
Radiotherapy (yes vs. no) | 11 (8.993–13.007) vs. 8 (6.156–9.844) | 0.600 (0.381–0.945) | 0.027 |
Gender (female vs. male) | 11 (6.664–15.336) vs. 9 (6.115–11.885) | 0.514 (0.325–0.812) | 0.004 |
Survival . | Median survival (months, 95% CI) . | HR (95% CI) . | P . |
---|---|---|---|
OS | |||
LGALS3 (High vs. low) | 10 (8.454–11.546) vs. 18 (13.335–22.665) | 1.946 (1.304–2.904) | 0.001 |
Chemotherapy (yes vs. no) | 14 (15.303–24.688) vs. 9 (7.513–15.901) | 0.626 (0.406–0.967) | 0.035 |
PFS | |||
LGALS3 (High vs. low) | 7 (5.444–8.556) vs. 15 (9.010–20.99) | 2.294 (1.504–3.499) | 0.001 |
Radiotherapy (yes vs. no) | 11 (8.993–13.007) vs. 8 (6.156–9.844) | 0.600 (0.381–0.945) | 0.027 |
Gender (female vs. male) | 11 (6.664–15.336) vs. 9 (6.115–11.885) | 0.514 (0.325–0.812) | 0.004 |
SNPs in LGALS3 are associated with glioblastoma risk
As known, LGALS3 facilitated glioblastoma development and progression, we next expanded our exploration of the association of LGALS3 gene with glioblastoma risk. A total of 961 glioma cases and 1,351 controls were recruited, and the characteristics of all participants were summarized in Supplementary Table S5. Five tag SNPs, rs4652, rs4644, rs7157768, rs8013027, and rs17128230, at LGALS3 locus, were discovered and the genotyping rate for all of them were over 99%. Among the healthy subjects, all genotypic frequencies were in HWE (P > 0.05). Notably, the allele frequencies of two SNPs (rs4644 and rs4652) were significantly different between the cases and the controls (P = 00.037 and 0.028, respectively; Supplementary Table S6). The genotype distribution of the five SNPs in the LGALS3 gene in case patients and control subjects and their associations with glioma risk are summarized in Supplementary Table S7. Logistic regression analyses revealed that CA+CC homozygotes of rs4652, after adjusting for age and sex, were significantly associated with an increased susceptibility to glioma compared with wild-type carriers in the dominant-effect model (adjusted OR = 1.19; 95% CI = 1.00–1.42). The homozygous variant genotype of rs4644 was individually associated with an increased risk of gliomas in the recessive-effect model (AA vs. AC+CC: adjusted OR = 2.10; 95% CI = 1.22–3.62). Furthermore, stratification analyses were performed using the histology subtypes (glioblastoma and other gliomas) for the two LGALS3 variants (rs4652 and rs4644). Analyses using codominant models showed both CC genotype of rs4652 and AA genotype of rs4644 associated with significantly increased risk of glioblastoma (OR = 1.44; 95% CI = 1.01–2.06 and OR = 2.81; 95% CI = 1.46–5.38, respectively; Table 2). Consistently, the AA genotype of rs4644 also displayed an increased association with risk in patients with glioblastoma in a recessive model (adjusted OR = 2.81; 95% CI = 1.47–5.36). However, no significant association was observed in the other glioma subgroup.
Analyses of the cumulative effect of adverse genotypes in rs4652 and rs4644 on glioblastoma risk
. | . | Glioblastoma . | |||
---|---|---|---|---|---|
Genomic model . | Genotype . | Control . | Case . | OR (95%CI) . | P . |
. | . | no. (%) . | no. (%) . | . | . |
rs4652 | |||||
Codominant | AA | 513 (38.3%) | 110 (34.0%) | 1.00 (reference) | |
CA | 634 (47.3%) | 154 (47.5%) | 1.14 (0.87–1.49) | 0.358 | |
CC | 193 (14.4%) | 60 (18.5%) | 1.44 (1.01–2.06) | 0.044 | |
Dominant | AA/CA+CC | 827 (61.7%) | 214 (66.0%) | 1.21 (0.94–1.56) | 0.148 |
rs4644 | |||||
Codominant | CC | 931 (68.9%) | 217 (67.0%) | 1.00 (reference) | |
AC | 396 (29.3%) | 91 (28.1%) | 1.00 (0.76–1.31) | 0.984 | |
AA | 24 (1.8%) | 16 (4.9%) | 2.81 (1.46–5.38) | 0.002 | |
Recessive | CC+AC/AA | 1327 (98.2%) | 308 (95.1%) | 2.81 (1.47–5.36) | 0.002 |
. | . | Glioblastoma . | |||
---|---|---|---|---|---|
Genomic model . | Genotype . | Control . | Case . | OR (95%CI) . | P . |
. | . | no. (%) . | no. (%) . | . | . |
rs4652 | |||||
Codominant | AA | 513 (38.3%) | 110 (34.0%) | 1.00 (reference) | |
CA | 634 (47.3%) | 154 (47.5%) | 1.14 (0.87–1.49) | 0.358 | |
CC | 193 (14.4%) | 60 (18.5%) | 1.44 (1.01–2.06) | 0.044 | |
Dominant | AA/CA+CC | 827 (61.7%) | 214 (66.0%) | 1.21 (0.94–1.56) | 0.148 |
rs4644 | |||||
Codominant | CC | 931 (68.9%) | 217 (67.0%) | 1.00 (reference) | |
AC | 396 (29.3%) | 91 (28.1%) | 1.00 (0.76–1.31) | 0.984 | |
AA | 24 (1.8%) | 16 (4.9%) | 2.81 (1.46–5.38) | 0.002 | |
Recessive | CC+AC/AA | 1327 (98.2%) | 308 (95.1%) | 2.81 (1.47–5.36) | 0.002 |
The LGALS3 with +191A and +292C enhances glioblastoma resistance to chemo- and radiotherapy
We then detected the impact of +191A and +292C haplotype on glioblastoma progression and treatment resistance. As shown in Fig. 4A and D, both T98G and U251 cells with overexpression of +191A and +292C LGALS3 displayed similar abilities of proliferation as those of LGALS3 wild-type (+191C and +292A) cells. However, LGALS3 with +191A and +292C haplotype increased the survival rate of T98G and U251 cells under different DNA damage stresses (Fig. 4B, C, E, and F). Collectively, these findings indicated that SNPs rs4644 and rs4652 were involved in the regulation of LGALS3 function in glioblastoma.
Effects of the C allele of rs4652 and the A allele of rs4644 on glioblastoma cell proliferation and resistance against radiochemotherapy. A and D, No difference of cell proliferation abilities is found between LGALS3 +191 A - +292 C and +191 C - +292 A groups in T98G and U251 cells. B and E, +191A and +292C haplotype promote cell survival under conditions with the dose of 5 Gy and 10 Gy irradiation. C and F, +191A and +292C haplotype also make the cell survival rate increased compared with the wild-type at levels of 1,000, 1,500, and 2,000 μmol/L temozolomide. (★, P <0.05).
Effects of the C allele of rs4652 and the A allele of rs4644 on glioblastoma cell proliferation and resistance against radiochemotherapy. A and D, No difference of cell proliferation abilities is found between LGALS3 +191 A - +292 C and +191 C - +292 A groups in T98G and U251 cells. B and E, +191A and +292C haplotype promote cell survival under conditions with the dose of 5 Gy and 10 Gy irradiation. C and F, +191A and +292C haplotype also make the cell survival rate increased compared with the wild-type at levels of 1,000, 1,500, and 2,000 μmol/L temozolomide. (★, P <0.05).
Discussion
In this study, we have addressed the oncologic functions of LGALS3 in treatment resistance and glioblastoma progression, LGALS3 gene expression in predicting patient prognosis, and the role of LGALS3 gene susceptibility in tumorigenesis. Our results revealed that LGALS3 promotes glioblastoma cell resistance to irradiation and temozolomide chemotherapy, improved the ability of proliferation, and acted as an independent risk factor for OS. And we also identified two SNPs (rs4652 and rs4644) in LGALS3 associated with glioblastoma risk and found that the C allele of rs4652 and the A allele of rs4644 contributed to more resistance to chemo-radiotherapy. These data suggest that LGALS3 may be a driver of glioblastoma development and recurrence, and a potential target for cancer treatment.
Radiotherapy is a mainstay treatment for many types of cancer including glioblastoma, and the temozolomide agent has been approved for treating this tumor for over one decade (28). The goal of chemo-radiotherapy is particularly to achieve direct or indirect DNA damage, which results in the termination of cancer cell division and proliferation and ultimately triggers cell apoptosis (29, 30). However, DNA repair mechanisms removing DNA lesions induced by ionizing irradiation or temozolomide and the activation of pathways blocking apoptosis are found to facilitate cancer cell survival and, in great part, to drive glioblastoma resistance towards the standardized therapeutic regimen (31, 32). Previous studies have shown that LGALS3 contains an Asp-Trp-Gly-Arg (NWGR) motif in the C-terminal domain, which is highly conserved in the BH1 domain of Bcl-2 gene family including Bcl-2, Bcl-xL, and Bax, and critical for the antiapoptotic function of Bcl-2 (33). Subsequently, LGALS3 has been demonstrated as a vital antiapoptotic molecule, which can act like Bcl-2, and to promote treatment resistance in multitype cancers (33). Besides, LGALS3 is also found to interact with Bcl-2 in a lactose-inhibitable manner (34), whereas Bcl-2 is linked to chemoresistance and aggressive recurrence in glioblastoma (35). Another Bcl-2 family member, Bcl-xL/Bcl-2–associated death promoter (Bad), accounting for the intrinsic pathway of apoptosis, can be downregulated by LGALS3 after treatment with chemotherapeutic drug (36), and the expression of LGALS3 also stimulates the phosphorylation of Bad (ph-Bad), which is in a highly expressed and inactive state in glioblastoma (37). In addition, as responding to cancer-killing effects induced by agents, mitochondria will release cytochrome c to upregulate Bax delivering the death signal to form Bax/Bak homodimers or antagonize Bcl-2 antiapoptosis effect through forming Bax/Bcl-2 heterodimers at outer membrane of mitochondria, which results in disruption of mitochondrial membrane potential and activation of caspases in glioblastoma (38). However, LGALS3 can inhibit mitochondrial depolarization and damage, and then contribute to the prevention of cytochrome c release, fewer caspases activity, suppression of apoptosis, and anticancer drug resistance (36, 39). Thus, it is indicated that LGALS3 may function through a cell death inhibition pathway that may involve Bcl-2 family. Moreover, cancer stem cell in glioblastoma has been demonstrated as a core contributor in chemo- and radioresistance through preferential activation of the DNA damage checkpoint response and extensive DNA repair aberrations (40, 41). Recently, studies in other cancers showed that LGALS3 expression could identify a subset of cancer stem cells with high levels of stem cell characteristics, including chemotherapy resistance (42). It is speculated that the increased level of LGALS3 in glioblastoma stem cell (43) may drive the therapeutic resistance by maintaining the stemness of cancer stem cells in the malignancy, which needs further investigations.
In addition, the typical oncogenic pathway RAS signaling, which drives tumorigenesis and progression in glioblastoma (44), was also reported in the downstream of LGALS3, and the overexpression of LGALS3 activated RAS and induced cancer cell proliferation (45). Taken together, both mentioned may explain why the high expression level of LGALS3 is related to short survival time for patients with glioblastoma.
Our study also underscored the functional single nucleotide variants of LGALS3 in glioblastoma progression. Previous findings revealed that rs4652 and rs4644 could contribute to glioma risk and the SNPs at rs4652 (+292 A>C) were associated with tumor grade and prognosis of patients with glioma (23, 24).Though lacking sufficient tissue for enough DNA to validate the relationship between the two SNPs and clinical and pathologic characteristics of the patients in our cohort, we first determined the positive impact of the C allele of rs4652 combined with the A allele of rs4644 on resistance to ionizing irradiation and chemotherapeutic drug in glioblastoma. As reported previously, amino acids Ala62-Tyr63 of the LGALS3 harbor the cleavage site for matrix metalloproteinases (MMP) and rs4644 rendering residue 64 of LGALS3 changing from histidine to proline may alter the pattern of MMPs cleavage, which may result in elevated LGALS3 level (46, 47). However, different from the enhanced phenotypes of treatment resistance, the difference of cell proliferation abilities between +191 A - +292 C and +191 C - +292 A groups was not observed in T98G and U251 cells, which indicated diverse signaling pathways underlying LGALS3 in glioblastoma progression. In addition, it is also possible that SNPs change the configuration of LGALS3, which facilitates the interaction of LGALS3 with MUC1 on the cell surface to promote EGFR dimerization and activation, and activated EGFR has been found to promote glioblastoma cell therapeutic resistance (48–50). Besides, the two SNPs (rs4652 and rs4644) were identified to contribute to glioma risk, which needs additional studies to confirm the association between gliomagenesis and these SNPS, the genes, and the pathways underlying LGALS3. Especially, LGALS3 pathway–based approaches may have a potential in the identification of novel genes conferring tumor susceptibility.
In conclusion, this study demonstrates the oncogenic role of LGALS3, along with its functional SNPs, in glioblastoma chemo- and radioresistance. Further understanding of the mechanism involved by LGALS3 may accelerate to develop a useful therapeutic intervention for this malignancy.
Disclosure of Potential Conflicts of Interest
No potential conflicts of interest were disclosed.
Authors' Contributions
Conception and design: H. Wang, Y. Yan, H. Chen, J. Chen
Development of methodology: H. Wang, L. Hu, Y. Yan, J. Chen
Acquisition of data (provided animals, acquired and managed patients, provided facilities, etc.): H. Wang, Q. Huang
Analysis and interpretation of data (e.g., statistical analysis, biostatistics, computational analysis): H. Wang, X. Song, Q. Huang, D. Yun
Writing, review, and/or revision of the manuscript: H. Wang, X. Song, T. Xu, D. Yun
Administrative, technical, or material support (i.e., reporting or organizing data, constructing databases): H. Wang, T. Xu, Y. Wang, H. Chen, D. Lu, J. Chen
Study supervision: D. Lu, J. Chen
Acknowledgments
This project was supported by the National Natural Science Foundation of China (grant numbers 81572501 and 81272781). We thank Dr. Mei Jiang for her help in English language editing.
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.
References
Supplementary data
Table S1. DEGs in each group.