Open Access

Establishment and characterization of human osteosarcoma cells resistant to pyropheophorbide-α methyl ester-mediated photodynamic therapy

  • Authors:
    • Yong Tao
    • Yunsheng Ou
    • Hang Yin
    • Yanyang Chen
    • Shenxi Zhong
    • Yongjian Gao
    • Zenghui Zhao
    • Bin He
    • Qiu Huang
    • Qianxing Deng
  • View Affiliations

  • Published online on: September 27, 2017     https://doi.org/10.3892/ijo.2017.4136
  • Pages: 1427-1438
  • Copyright: © Tao et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

The present study was performed to establish and characterize new human osteosarcoma cell lines resistant to pyropheophorbide-α methyl ester‑mediated photodynamic therapy (MPPa-PDT). MPPa-PDT-resistant cells are isolated from the human osteosarcoma MG63 and HOS cell lines and two resistant populations were finally acquired, including MG63/PDT and HOS/PDT. Cell Counting Kit-8 (CCK-8) assay was used to determine the MPPa-PDT, cisplatin (CDDP) resistance and proliferation of MG63, MG63/PDT, HOS and HOS/PDT cells. The intracellular ROS were analyzed using DCFH-DA staining. The colony formation, invasion and migration of parental and resistant cells were compared. FCM was employed to examine the cell cycle distribution, the apoptosis rate and the proportion of CD133+ cells. The fluorescence intensity of intracellular MPPa was observed by fluorescence microscopy and quantified using microplate reader. The protein levels were assessed by western blotting (WB). Compared with two parental cells, MG63/PDT and HOS/PDT were 1.67- and 1.61-fold resistant to MPPa-PDT, respectively, and also exhibited the resistance to CDDP. FCM assays confirmed that both MG63/PDT and HOS/PDT cells treated with MPPa-PDT displayed a significantly lower apoptosis rate in comparison with their corresponding parental cells. The expression of apoptosis-related proteins (i.e. cleaved-caspase 3 and cleaved‑PARP), intracellular ROS and the antioxidant proteins (HO-1 and SOD1) in MG63/PDT and HOS/PDT cells was also lower than that in parental cells. Both MG63/PDT and HOS/PDT cells exhibited changes in proliferation, photosensitizer absorption, colony formation, invasion, migration and the cell cycle distribution as compared to MG63 and HOS cells, respectively. Compared to MG63 and HOS cells, both resistant cell lines had a higher expression of CD133, survivin, Bcl-xL, Bcl-2, MRP1, MDR1 and ABCG2, but a lower expression of Bax. The present study successfully established two resistant human osteosarcoma cell lines which are valuable to explore the resistance-related mechanisms and the approaches to overcome resistance.

Introduction

Osteosarcoma is well known as the most common primary malignant bone tumor in clinic. It is strongly invasive and has poor prognosis, and mainly occurs in children and adolescents (1-3). Photodynamic therapy (PDT), a new promising approach for tumor therapy, is featured by excellent selectivity, less side reactions, and visible light of specific wavelength to excite photosensitizer enriched in tumor tissue, which can subsequently result in reactive oxygen, mainly singlet oxygen to cause tumor cell death (4,5). In addition, PDT has already been used in the treatment of skin and esophageal cancer, and other tumors (6-8). MPPa, a second generation photosensitizer derived from chlorophyll, has many advantages including stability, single component, rapid absorption and metabolism and strong photosensitivity (9). It has been reported that MPPa-PDT can kill tumor cells such as nasopharyngeal carcinoma, prostate and breast cancer cells (10-12).

However, photodynamic therapy may lead to the resistance of tumor cells (13). Numerous mechanisms are reported to mediate the resistance of tumor cells (14). Furthermore, the surviving tumor cells play an important role in the recurrence and deterioration of tumor, and result in poor prognosis (15,16). It has been reported that the increased expression of HO1, SOD1 and other antioxidant proteins can protect tumor cells from ROS damage (17). ATP-binding cassette (ABC) transporters (e.g. ABCG2, MRP1 and MDR1) were found to mediate tumor cell resistance, and alleviate cell damage through pumping out intracellular toxins (18). Furthermore, the overexpression of anti-apoptotic protein P-survivin was responsible for chemoresistance and radio-resistance of tumor cells (19). The Bcl-2 protein family can regulate the permeability of mitochondrial membrane, and its expression can decrease the sensitivity of tumor cells to antitumor therapy (20). The present study aimed to establish human osteosarcoma cell lines resistant to MPPa-PDT. The resistant cell lines can also be employed to investigate the resistance mechanism of human osteosarcoma cells to MPPa-PDT and explore the approach to overcome MPPa-PDT resistance.

Materials and methods

Reagents and instruments

MPPa and 2′7′-dichlorofluorescin diacetate (DCFH-DA) were purchased from Sigma-Aldrich (St. Louis, MO, USA). BCA and trypsin were purchased from Beyotime Biotech (Shanghai, China). Cell viability and cytotoxicity test kits [Cell Counting Kit-8 (CCK-8)] were obtained from Dojindo Molecular Technologies (Kumamoto, Japan). Extracellular matrix gel was obtained from BD Biosciences (Franklin Lakes, NJ, USA). Annexin V-propidium iodide (PI) double-staining test kit was purchased from KeyGen Biotech (Nanjing, China). LED equipment was purchased from Chongqing Jingyu Laser Technology Co. Ltd. (Chongqing, China).

Antibodies

Primary antibodies were: β-actin (1:1,000), cleaved caspase-3 (1:1,000), SOD1 (1:1,000), cleaved PARP (1:1,000), Bcl-2 (1:1,000), Bcl-xL (1:1,000), Bax (1:1,000), CD133 (1:1,000), MDR1 (1:500), MRP1 (1:500) and P-survivin (1:1,000; all from Cell Signaling Technology, Inc. Danvers, MA, USA), HO1 (1:500; Proteintech Group, Inc., Wuhan, China) and CD133-APC (1:20; BD Biosciences). Secondary antibodies were: HRP monoclonal antibody anti-IgG of mouse and HRP monoclonal antibody anti-IgG of rabbit (Cell Signaling Technology, Inc.).

Cell line and culture

MG63 and HOS cells were obtained from the Chinese Academy of Sciences (Shanghai, China), and cultured in Dulbecco's modified Eagle's medium (DMEM) supplemented with 10% fetal bovine serum (FBS) (both from HyClone, Beijing, China), 100 µg/ml penicillin and 100 µg/ml streptomycin (Beyotime Biotech) at 37°C in a humidified atmosphere containing 5% CO2.

Resistance induction to MPPa-PDT

MG63 and HOS cells were cultured in the dark and incubated with different MPPa concentrations (Fig. 1A and B) for 20 h, and then washed twice with phosphate-buffered saline (PBS). The culture medium was replaced, and the cells were exposed to red light (630 nm, 40 mW/cm2) in the continuous output mode. Treatment conditions that caused survival rates of 40–60% were chosen. The dead cells were wiped off and the surviving cells were cultured in fresh complete medium continuously for 24 h. Twenty hours later, the cells were harvested and replaced. They were subjected to a new PDT with intermittently increased doses of MPPa. The final populations were obtained following a total of 10 cycles of PDT, and they were MG63/PDT and HOS/PDT.

MPPa-PDT sensitivity assay

CCK-8 was used to investigate the sensitivity of cell lines MG63, MG63/PDT, HOS and HOS/PDT to MPPa-PDT. Cells were plated in 96-well plates at a density of 5×103 cells/well with 3 duplications. After a 24-h incubation, the culture medium was replaced with the fresh medium containing different concentrations of MPPa (0, 0.25, 0.5, 0.75 and 1.0 µM for MG63 and MG63/PDT; 0, 0.1, 0.15, 0.2 and 0.25 µM for HOS and HOS/PDT cells). The cells were then cultured for 20 h. The culture medium was replaced, and the cells were exposed to red light (630 nm, 40 mW/cm2). Twenty-four hours later, the cells were incubated for 1 h with 10 µl CCK-8 in each well. A microplate reader was employed to detect the absorption values of CCK-8 at 450 nm. The cell viability was calculated according to the following formulation:

Cell viability(%)=Average OD in experiment group/average OD in control group×100% Resistance indices(RIs)=IC50values for resistant cells/IC50values for parental cells

Based on the results of the cell viability test, we chose an MPPa concentration of 0.45 µM for MG63 and MG63/PDT cells and 0.15 µM for HOS and HOS/PDT cells with a light energy density of 4.8 J/cm2 as the treatment conditions.

Detection of apoptosis rate by Annexin V-PI double staining and FCM

Cells were seeded in 6-well plates at a density of 1×105 cells/well. All cells were harvested after corresponding treatments, and assessed by FCM after Annexin V-PI double staining (KeyGen Biotech).

Cell proliferation assay by CCK-8

Cells were seeded in 96-well plates at a density of 5×103 cells/well. When all cells attached, the cell viability was determined using the CCK-8 assay at 0, 12, 24 and 48 h.

Cell cycle analysis by FCM

Cells were seeded in 6-well plates at a density of 1×105 cells/well. After incubation for ~36 h, when fusion was up to ~60–70%, all groups were collected at the same time and washed twice, and fixed by suspending in 70% ethanol at 4°C for 24 h, and then subjected to FCM.

Assessment of intracellular ROS level by DCFH-DA staining

Cells were inoculated in 6- and 24-well plates, at a density of 1×105 and 5×104 cells/well, respectively. Following the corresponding treatments, the cells were further incubated for 2 h. Then, DCFH-DA (10 µM) was added at 37°C for 20 min. Finally, cells in 24-well plates were observed by fluorescence microscope (FM) after being washed 3 times, and the cells in 6-well plates were determined by FCM after being trypsinized and collected.

Assessment of CD133 by FCM

Cells were seeded in 6-well plates at a density of 1×105 cells/well. All cells were trypsinized, centrifuged and resuspended in PBS. CD133-APC (1:20) was added to corresponding groups and incubated in the dark for 10 min at 4°C. After being washed, the cells were resuspended in PBS and analyzed by FCM.

Measurement of intracellular MPPa

Cells were inoculated in 6- and 96-well plates at a density of 1×105 and 5×103 cells/well, respectively. After attachment, the cells were incubated with different MPPa concentrations (2, 4 and 8 µM) for 20 h. Then, an inverted FM was used to observe the fluorescence of cells with 4 µM MPPa in 6-well plates, and a microplate reader was adopted to examine the fluorescence value of cells in 96-well plates (λexc 525 nm; λem 680 nm).

Colony formation assay

MG63 and MG63/PDT were seeded into 6-well plates at a density of 200 cells/well. HOS and HOS/PDT were seeded in 6-well plates at a density of 500 cells/well. The culture medium was replaced every 3 days. Two weeks later, colonies were fixed with polyformaldehyde (4%) for 20 min and visualized by crystal violet solution.

Invasion and migration assays

The ECM (1:8, 100 µl) was added to the top chamber and placed in an incubator for 2 h. When the ECM was dried, 4×104 MG63 and MG63/PDT or 8×104 HOS and HOS/PDT cells in serum-free medium were added to the top chamber. In the lower chamber, complete medium was added. After incubation at 37°C in 5% CO2 for 48 h, the cells on the upper surface of the membrane were removed using a cotton swab. The cells on the lower surface of the membrane were fixed with polyformaldehyde (4%) for 15 min and stained by crystal violet solution (0.1%). Cells were observed by inverted phase contrast microscope. Five randomized fields at a magnification of ×40 were selected. For the migration assay, a protocol similar to the invasive assay was performed, but without the ECM layer in the chamber, and the cells added as well as the incubation time was half of that in invasion assay, respectively.

Western blot analysis

Cells were rinsed with PBS and lysed by RIPA buffer containing a phosphatase and protease inhibitor cocktail. Protein concentration was assessed by BCA. Protein samples (40 µg) were electrophoresed and blotted on polyvinylidene fluoride membranes, which were blocked in Tris-buffered saline containing Tween-20 and 5% non-fat milk for 1 h at room temperature, and incubated with the corresponding primary antibodies overnight at 4°C. After being rinsed, the membranes were subjected to the horseradish peroxidase-conjugated secondary antibody for 1 h and developed by electrochemiluminescence. Quantity One software was used to detect the gray values of some western bands. The relative expression of target proteins was displayed using the ratio of target protein/β-actin. Three independent experiments were performed.

Statistical analysis

Data are expressed as the mean ± SD and analyzed by SPSS (SPSS, Inc., Chicago, IL, USA). Differences between groups were determined using the one-way or two-way ANOVA test for intergroup and independent-sample t-test for two groups. At P<0.05, the difference was considered significant.

Results

Establishment of the resistant cell lines

Parental cells, MG63 and HOS cells, were treated by PDT with increased concentration of MPPa for 10 cycles (Fig. 1A and B). The 10th generation of resistant cells obtained were named MG63/PDT and HOS/PDT, respectively. We employed MG63, HOS and MG63/PDT, HOS/PDT as experimental objects for the following studies. In order to verify the tolerance of resistant cells compared with their corresponding parental cells, the cell viability and apoptosis rate were assessed after PDT. A CCK-8 assay demonstrated that the concentration at 50% inhibition (IC50) for MG63/PDT (0.704±0.016 µM) was 1.67-fold more resistant than that for MG63 (0.421±0.028 µM; P<0.001), and the concentration at 50% inhibition (IC50) for HOS/PDT (0.226±0.008 µM) was 1.61-fold higher as compared to that for HOS (0.140±0.004 µM) (P<0.001; Fig. 1C and D). FCM indicated that the apoptosis rate of MG63/PDT (20.04±2.16%) was markedly lower than that of MG63 (40.58±2.34%) (P<0.001; Fig. 1E), when they were subjected to PDT with 0.45 µM MPPa. The apoptosis rate of HOS/PDT (17.80±1.26%) subjected to PDT with 0.15 µM MPPa was also significantly lower than that of HOS (37.22±0.81%) (P<0.001; Fig. 1F).

Measurement of intracellular ROS by FCM and FM

FCM and FM demonstrated that there was no significant difference of intracellular ROS levels between MG63 and MG63/PDT cells without PDT treatment (P=0.267). However, the ROS level in MG63 cells was markedly higher than that in MG63/PDT after PDT treatment (P=0.001) (Fig. 2A, B and E). Similar results were observed between HOS and HOS/PDT cells (Fig. 2C, D and F).

Cell proliferation test

After 12 h of culture, the MG63/PDT cells began to grow slower compared with the MG63 cells and the difference of growth speed was significant after 24 h (P=0.037; Fig. 3A). However, HOS/PDT began to grow slower than the parental HOS cells after 24 h, and the difference of growth speed was significant after 48 h (P=0.005; Fig. 3B). The proliferation curves of two types of resistant cells were more flat compared with that of the relatively primitive cells.

Examination of cell cycle distribution by FCM

FCM revealed that the proportion of cells in the G0/G1, S and G2/M phase were 19.56, 69.54 and 10.9% for MG63; 46.35, 47.12 and 6.53% for MG63/PDT (Fig. 4A); 24.17, 53.95 and 21.88% for HOS; and 39.67, 44.07 and 16.27% for HOS/PDT (Fig. 4C), respectively. The MG63/PDT and HOS/PDT cells exhibited shorter S and G2/M phases with a concomitantly longer G0/G1 phase in cycle distribution than their comparative cells (P<0.05; Fig. 4B and D).

Invasion assay

Transwell assays with a layer of Matrigel on the top inserts were employed to examine the invasive capability. The results revealed that the number of cells penetrated both the Matrigel and membrane was 232±22.7 for MG63/PDT cells, 42±11 for MG63 cells (Fig. 5A and B), 86.33±14.64 for HOS/PDT cells and 18.33±3.51 for HOS cells (Fig. 5C and D), demonstrating that MG63/PDT and HOS/PDT had better ability of invasiveness compared with their corresponding parental cells (P<0.05).

Migration assay

Transwell assay was performed to detect migration ability. The results indicated that the number of cells penetrating the membrane was 195±12.12 for MG63/PDT cells, 13.67±5.68 for MG63 cells (Fig. 6A and B), 687.33±31.72 for HOS/PDT cells and 263.67±20.84 for HOS cells (Fig. 6C and D), revealing that the migration ability of MG63/PDT and HOS/PDT cells were increased compared with parental MG63 and HOS cells, respectively (P<0.05).

Colony formation assay

Both MG63/PDT and HOS/PDT cells had more potential in forming colonies. The number of MG63/PDT cell colonies (36.3±5.0) was significantly higher compared to MG63 cells (21.3±5.7, P=0.027; Fig. 7A and B). The number of HOS/PDT cell colonies (22.3±4.5) was also markedly higher than HOS cells (6.3±1.5, P=0.004; Fig. 7C and D).

Determination of osteosarcoma stem marker CD133 by FCM

FCM revealed that the proportion of CD133+ cells in MG63/PDT, MG63, HOS/PDT and HOS cells was 11.49±0.71, 0.43±0.09, 5.76±0.52 and 0.50±0.06%, respectively. MG63/PDT and HOS/PDT exhibited more CD133+ cells than their corresponding original cells (Fig. 8; P<0.05).

Measurement of intracellular MPPa

The fluorescence intensity of MPPa cells was observed by FM. The results revealed that the fluorescence intensity of MPPa in MG63 cells was stronger than that of MG63/PDT (Fig. 9A), and the fluorescence intensity of MPPa in HOS cells was stronger than that of HOS/PDT cells (Fig. 9C). Microplate reader was employed to quantify fluorescence intensity of MPPa in the cells, revealing that MPPa content in MG63 and HOS cells was significantly higher than that of MG63/PDT and HOS/PDT cells, respectively (P<0.05; Fig. 9B and D).

Cytotoxicity by CDDP

CCK-8 was adopted to assess the cyto-toxicity of CDDP on PDT resistant and parental cells. Survival rates of MG63/PDT and HOS/PDT were significantly higher than those of the corresponding original MG63 and HOS cells (P<0.05; Fig. 10).

Western blot assays

WB was used to detect the expression of different proteins. The results demonstrated that there was no significant difference of the expression of antioxidant-related protein HO-1, SOD1 and pro-apoptotic proteins cleaved-caspase 3, cleaved-PARP between MG63 and MG63/PDT cells (P>0.05). After treatment with MPPa-PDT for 12 h, the expression of HO-1, SOD1, cleaved-caspase 3 and cleaved-PARP in MG63 and MG63/PDT cells was increased, but their expression in MG63/PDT cells was lower than those in MG63 cells (P<0.05; Fig. 11A and B). The expression of HO-1, SOD1, cleaved-caspase 3 and cleaved-PARP in HOS and HOS/PDT cells displayed similar results (Fig. 11C and D).

To further study the characteristics of the resistance and the mechanisms, the expression of drug resistance-related proteins (e.g. ABCG2, MRP1, MDR, apoptosis-related proteins, Bcl-2, Bcl-xL, P-survivin and Bax) and cancer stemness markers CD133 in MG63, MG63/PDT, HOS and HOS/PDT were detected. The results indicated that the levels of ABCG2, MRP1, MDR and the anti-apoptotic proteins Bcl-2, Bcl-xL, P-survivin in MG63/PDT and HOS/PDT cells were all more highly expressed than those in MG63 and HOS cells (P<0.05; Fig. 12). However, the pro-apoptotic protein Bax in MG63/PDT and HOS/PDT cells was significantly downregulated compared with that in their parental cells (P<0.05; Fig. 12). Furthermore, the expression for the cancer stem marker CD133 in MG63/PDT cells and HOS/PDT cells was markedly increased relative to MG63 and HOS cells (P<0.05; Fig. 12).

Discussion

The 5-year survival rate of osteosarcoma patients has reached 60-70% with the continuous improvement of treatment, but the treatment is still not effective for the patients with recurrence, metastasis and chemoresistance. PDT has emerged as an important approach to treat tumors, with some advantages including strong targeting, non-invasive and repeatable treatment (4). Our previous study found that MPPa-mediated photodynamic therapy could significantly kill osteosarcoma MG63 cells, suggesting that it may be used in the clinical treatment of osteosarcoma patients (3). It is well known that resistance is very common in chemotherapy, radiotherapy and other cancer therapy (21,22). This may be related to the existence of heterogeneous cells resistant to the therapy or the induction of resistance in tumor cells (23). Some surviving cells after the treatment become more aggravated, more prone to invasion and metastasis (13,16). However, PDT still encountered the problem of resistance due to the expression of resistance-related proteins.

Milla et al successfully isolated squamous carcinoma cells (SCCs) resistant to PDT by repeated methyl d-aminolevulinic acid (Me-ALA-PDT) treatment of LD90 doses for tumor cells (24). The present study selected LD90 doses of MPPa-PDT for human osteosarcoma cell lines MG63 and HOS to establish new human osteosarcoma cell lines. However, after 3 days of treatment, all the cells died and failed to form resistance. This may be related to mismatch speed of resistance-related molecule expression. Thus, we chose a relatively mild treatment condition of IC40-IC60. The MG63 and HOS cells were subjected to 10 cycles of PDT by gradually increasing the dose of MPPa, and finally MPPa-PDT-resistant cells were obtained, named MG63/PDT and HOS/PDT, respectively. In order to verify the resistance of newly constructed osteosarcoma cell lines MG63/PDT and HOS/PDT to MPPa-PDT, we examined the expression of cleaved-caspase 3 and cleaved-PARP, apoptosis, cell viability in MG63, MG63/PDT, HOS and HOS/PDT cells after MPPa-PDT treatment. The results revealed that MG63/PDT and HOS/PDT cells were more resistant to MPPa-PDT compared to their corresponding parental cells. There may be some mechanisms that protected them from the damage of MPPa-PDT in osteosarcoma cells.

ROS is the main mechanism by which PDT kills osteosar-coma cells (3,25). In the present study, ROS in resistant cells MG63/PDT and HOS/PDT and parental cells MG63 and HOS, was analyzed by FCM and FM. The results demonstrated that there was no difference in the ROS level between resistant and parental cells in the absence of treatment. However, after treatment with PDT, the amount of ROS in resistant cells was significantly lower than that in parental cells, suggesting that the resistant cells changed some signal molecules to decrease the production of ROS. The amount of ROS induced by PDT depends on the type and the dose of the photosensitizer, irradiation time and the ability of cells to antioxidative stress. HO-1 not only degrades heme, but also promotes antioxidation, anti-inflammation and anti-apoptosis (26,27). Ciesla et al found that upregulation of HO-1 expression in rhabdomyosarcoma could reduce intracellular ROS content and promote cell survival (28). Lv et al reported that inhibition of HO-1 could increase the sensitivity of laryngeal carcinoma to CDDP. Early studies also found that HO-1 expression could decrease the damage of photodynamic therapy to tumors (29). SOD1 is an important antioxidant enzyme in cells, and is capable of decomposing superoxide, and free cells of ROS damage. Soares et al reported that inhibition of SOD1 increased the sensitivity of tumor cells to photodynamic therapy (30,31). In the present study, HO-1 and SOD1 expression were examined after MPPa-PDT treatment by same MPPa and light dose. However, the results were contrary to our expectation. The expression of HO-1 and SOD1 in resistant cells was significantly lower than those in parental cells, though both of them were induced by MPPa-PDT. In addition, there was no significant difference in the expression of HO-1 and SOD1 between resistant and parental cells without MPPa-PDT treatment. The results indicated that there may be another pathway in resistant cells that induces the resistance to MPPa-PDT. Higher expression of antioxidant machinery of cells definitely should result in low ROS levels in response to a particular treatment. Primitively, we hypothesized that MPPa-PDT-resistant osteosarcoma cells may produce more antioxidant proteins than original cells in order to clean out ROS. However, the results demonstrated that MPPa-PDT-resistant osteosarcoma cells had less antioxidant proteins than original osteosarcoma cells. Tian et al found that the inhibition of antioxidants may increase ROS and the damage of MPPa-PDT on tumor cells (32). There may be another reason for the high expression of antioxidants in parental cells. In one treatment, the expression of antioxidant proteins was adjusted in osteosarcoma cells according to the amount of ROS and formed tolerance to treatment, however this warrants further exploration. Concomitantly, the expression of other anti-oxidative stress kinases in resistant cells was also worthy of further study.

Our previous study revealed that the cytotoxicity of MPPa-PDT on osteosarcoma cells was in a dose-dependent manner (3). Milla et al found that the amount of Me-ALA in resistant SCC was less than that in parental SCC cells after only Me-ALA treatment for 4 h (24). We hypothesized that certain intracellular content after MPPa treatment resulted in the difference of ROS between resistant and parental cells. By detecting the fluorescence intensity of intracellular MPPa, we found that the content of MPPa in the resistant cells was significantly less than that in parental cells after the identical MPPa treatment, suggesting that the resistant cells could decrease the production of ROS induced by PDT owing to the less content of intracellular photosensitizer compared with that in parental cells. The ABC transporter family proteins with ATP enzyme activity (i.e. ABCG2, MDR1 and MRP1) can diminish intracellular drug concentration by accelerating drug efflux. Ishikawa et al found that inhibition of ABCG2 increased the uptake of photosensitizer PpIX by tumor cells (33). Liu et al reported that inhibition of MDR1 or MRP1 could increase the sensitivity of osteosarcoma cells to chemotherapy (34,35). In the present study, WB results implied that the expression of ABCG2, MDR1 and MRP1 in the resistant cells was signifi-cantly unregulated compared to parental cells. The decrease of MPPa content in the tolerant cells may be linked to the high expression of ABC transporter protein. However, the definite relationship between their high expression and the resistance to MPPa-PDT of MG63/PDT and HOS/PDT cells warrants further studies.

In addition to the upregulation of the ABC transporter, the anti-apoptotic and pro-apoptotic proteins may also be involved in the formation of cell resistance. The BCL-2 protein family is widely involved in the process of cell apoptosis. The pro-apoptotic protein Bax can increase the permeability of the mitochondrial membrane to facilitate the release of cytochrome c from the mitochondria into the cytoplasm. On the contrary, anti-apoptotic protein Bcl-2 and Bax-xL can stabilize the mitochondrial membrane by inhibiting Bax (20). Previous studies have found that MPPa-PDT can induce apoptosis of osteosarcoma MG63 cells by downregulation of Bcl-2 and the promotion of the expression of Bax (3). Survivin is a member of the inhibitor of apoptosis protein, which can bind caspase 3, an apoptosis executor, thereby inhibiting its activity (19). Ferrario et al found that a reduction of survivin can increase the cytotoxicity of PDT on breast cancer cells (36). The expression of P-survivin is unregulated in PDT-resistant squamous carcinoma cells (24). Our results revealed that the expression of anti-apoptotic protein Bcl-2, Bcl-xL and P-survivin was significantly increased in the resistant cells MG63/PDT and HOS/PDT compared with the parental cells MG63 and HOS, and the expression of pro-apoptotic protein Bax was notably decreased. The results demonstrated that upregulation of anti-apoptotic proteins and downregulation of pro-apoptotic proteins may be one of the mechanisms responsible for PDT-resistance of MG63/PDT and HOS/PDT cells.

Under normal circumstances, the resistant cells may accelerate proliferation to overcome the cytotoxicity of drugs (37). In our experiments, we found that the proliferation ability of the resistant cells MG63/PDT and HOS/PDT was weaker than that of their primary MG63 and HOS cells, respectively. The results are similar to those found in the CDDP resistant osteosarcoma cells SOSP-9607/CDDP (38). The slow proliferation of the resistant cells may be caused by the increased quiescent cells (16). In line with this, our results also demonstrated that there were more cells in the G0/G1 phase in resistant cells MG63/PDT and HOS/PDT compared with their corresponding parental cells. PDT can cause DNA damage, which can activate or inhibit some signal molecules to block the cell cycle in the G1 phase and promote DNA repair (39,40). The results revealed that more cells in the G0/G1 phase of the resistant cells MG63/PDT and HOS/PDT could enhance their DNA repair ability, which also may be one of the mechanisms responsible for the PDT-resistance.

Casas et al found that the ability of invasion and migration in PDT-resistant cells was significantly weaker than that of their primary cells (41). However, Milla et al found that the migration ability of PDT-resistant SCC cells was stronger than that of their original cells (24). Han et al found that the invasion ability of CDDP-resistant osteosarcoma cells was also significantly enhanced compared to their primary cells (37). Our results revealed that the invasion, migration and clone formation capacity of resistant cells MG63/PDT and HOS/PDT were also significantly stronger than those of parental cells. The invasion, migration and clone formation may vary depending on the treatment methods, and the type of cells and photosensitizers. The results revealed that the recurrence of osteosarcoma after PDT may become more troublesome.

Cancer stem cells (CSCs) have the capacity of self-renewal and differentiation, and are accountable for proliferation, metastasis, drug resistance and recurrence of tumors (42). CD133 is currently recognized as a tumor stem cell marker, and many researchers have chosen CD133 as a marker for osteosarcoma stem cells to screen osteosarcoma stem cells (43-47). ABCG2 is not only associated with drug resistance, but also as a surface marker of CSC (48,49). It was reported that osteosarcoma-cancer-stem cells exhibited higher expression of ABCG2 and CD133 (50). In the present study, ABCG2 was upregulated in resistant cells. The present study also revealed that MG63/PDT and HOS/PDT cells presented higher expression of CD133 and possessed more CD133+ cells compared to their parental cells, demonstrating that the stemness of osteosarcoma cells was enhanced during resistant cell construction by MPPa-PDT, and osteosarcoma stem cells may be involved in PDT-resistance of MG63/PDT and HOS/PDT cells. However, the proportion of CD44+ cells, another cancer stem cell maker (51), between parental and resistant cells did not exhibit any significant difference and the slow proliferation and less cells in the G2/M phase of MG63/PDT and HOS/PDT cells were inconsistent with the appearance of cancer stem cells. It has been reported that PDT can conquer the resistance of tumor cells to chemotherapeutic drugs (52). CCK-8 assay was used to detect the viability of the cells treated with different doses of CDDP. The results revealed that the PDT-resistant osteosarcoma cells also appeared to be more resistant to CDDP, suggesting that MG63/PDT and HOS/PDT cells were not only resistant to PDT, but also to chemotherapy.

In conclusion, we successfully constructed two new PDT-resistant osteosarcoma cell lines MG63/PDT and HOS/PDT which featured by strong invasiveness, migration and stemness, and high expression of anti-apoptotic proteins as well as the ABC transporter family proteins, and low expression of pro-apoptotic protein and CDDP tolerance. The newly constructed PDT-resistant cell lines will be beneficial in the exploration of the biological characteristics of recurrent osteosarcoma, the methods of conquering PDT-resistance and to clarify possibly related mechanisms.

Acknowledgments

The present study was supported by the National Natural Science Foundation of China (81572634), and the Graduate Scientific Innovation Project of Chongqing Education Committee (CYS15141). The authors wish to thank Mr. Yong Zhu, of the Department of Orthopedics, The First Affiliated Hospital of Chongqing Medical University (Chongqing, China), for their advice and supervision with regards to the statistical analysis and modification of the manuscript.

References

1 

Wang W, Yang J, Wang Y, Wang D, Han G, Jia J, Xu M and Bi W: Survival and prognostic factors in Chinese patients with osteosarcoma: 13-year experience in 365 patients treated at a single institution. Pathol Res Pract. 213:119–125. 2017. View Article : Google Scholar : PubMed/NCBI

2 

Whelan J, McTiernan A, Cooper N, Wong YK, Francis M, Vernon S and Strauss SJ: Incidence and survival of malignant bone sarcomas in England 1979–2007. Int J Cancer. 131:E508–E517. 2012. View Article : Google Scholar

3 

Huang Q, Ou YS, Tao Y, Yin H and Tu PH: Apoptosis and autophagy induced by pyropheophorbide-α methyl ester-mediated photodynamic therapy in human osteosarcoma MG-63 cells. Apoptosis. 21:749–760. 2016. View Article : Google Scholar : PubMed/NCBI

4 

Agostinis P, Berg K, Cengel KA, Foster TH, Girotti AW, Gollnick SO, Hahn SM, Hamblin MR, Juzeniene A, Kessel D, et al: Photodynamic therapy of cancer: An update. CA Cancer J Clin. 61:250–281. 2011. View Article : Google Scholar : PubMed/NCBI

5 

Khaled YS, Wright KE, Melcher A and Jayne D: Anti-cancer effects of oncolytic viral therapy combined with photodynamic therapy in human pancreatic cancer cell lines. Spring Meeting for Clinician Scientists in Training 2015. 26–Feb. 2015, https://doi.org/10.1016/S0140-6736(15)60371-3.

6 

Cheng Y, Chang Y, Feng Y, Liu N, Sun X, Feng Y, Li X and Zhang H: Simulated sunlight-mediated photodynamic therapy for melanoma skin cancer by titanium-dioxide-nanoparticle-gold-nanocluster-graphene heterogeneous nanocomposites. Small. 13:16039352017. View Article : Google Scholar

7 

Kuzyniak W, Schmidt J, Glac W, Berkholz J, Steinemann G, Hoffmann B, Ermilov EA, Gürek AG, Ahsen V, Nitzsche B, et al: Novel zinc phthalocyanine as a promising photosensitizer for photodynamic treatment of esophageal cancer. Int J Oncol. 50:953–963. 2017. View Article : Google Scholar : PubMed/NCBI

8 

Yano T, Kasai H, Horimatsu T, Yoshimura K, Teramukai S, Morita S, Tada H, Yamamoto Y, Kataoka H, Kakushima N, et al: A multicenter phase II study of salvage photodynamic therapy using talaporfin sodium (ME2906) and a diode laser (PNL6405EPG) for local failure after chemoradiotherapy or radiotherapy for esophageal cancer. Oncotarget. 8:22135–22144. 2017.PubMed/NCBI

9 

Luo T, Wilson BC and Lu QB: Evaluation of one- and two-photon activated photodynamic therapy with pyropheophorbide-a methyl ester in human cervical, lung and ovarian cancer cells. J Photochem Photobiol B. 132:102–110. 2014. View Article : Google Scholar : PubMed/NCBI

10 

Li KM, Sun X, Koon HK, Leung WN, Fung MC, Wong RN, Lung ML, Chang CK and Mak NK: Apoptosis and expression of cytokines triggered by pyropheophorbide-a methyl ester-mediated photodynamic therapy in nasopharyngeal carcinoma cells. Photodiagn Photodyn Ther. 3:247–258. 2006. View Article : Google Scholar

11 

Tian Y, Leung W, Yue K and Mak N: Cell death induced by MPPa-PDT in prostate carcinoma in vitro and in vivo. Biochem Biophys Res Commun. 348:413–420. 2006. View Article : Google Scholar : PubMed/NCBI

12 

Tian YY, Hu XY, Leung WN, Yuan HQ, Zhang LY, Cui FA and Tian X: Investigation of photodynamic effect caused by MPPa-PDT on breast cancer Investigation of photodynamic effect caused by MPPa-PDT. Laser Phys Lett. 9:754–758. 2012. View Article : Google Scholar

13 

Gilaberte Y, Milla L, Salazar N, Vera-Alvarez J, Kourani O, Damian A, Rivarola V, Roca MJ, Espada J, González S, et al: Cellular intrinsic factors involved in the resistance of squamous cell carcinoma to photodynamic therapy. J Invest Dermatol. 134:2428–2437. 2014. View Article : Google Scholar : PubMed/NCBI

14 

Zamarrón A, Lucena SR, Salazar N, Sanz-Rodríguez F, Jaén P, Gilaberte Y, González S and Juarranz Á: Isolation and charac-terization of PDT-resistant cancer cells. Photochem Photobiol Sci. 14:1378–1389. 2015. View Article : Google Scholar

15 

Anderson SJ, Wapnir I, Dignam JJ, Fisher B, Mamounas EP, Jeong JH, Geyer CE Jr, Wickerham DL, Costantino JP and Wolmark N: Prognosis after ipsilateral breast tumor recurrence and locoregional recurrences in patients treated by breast-conserving therapy in five National Surgical Adjuvant Breast and Bowel Project protocols of node-negative breast cancer. J Clin Oncol. 27:2466–2473. 2009. View Article : Google Scholar : PubMed/NCBI

16 

Wang C, Guo LB, Ma JY, Li YM and Liu HM: Establishment and characterization of a paclitaxel-resistant human esophageal carcinoma cell line. Int J Oncol. 43:1607–1617. 2013. View Article : Google Scholar : PubMed/NCBI

17 

Jeon M, Rahman N and Kim YS: Cytoprotective effect of Makgeolli lees on paraquat induced oxidative stress in A549 cells via activation of NRF2 and antioxidant genes. J Microbiol Biotechnol. 26:277–286. 2016. View Article : Google Scholar

18 

Dean M: ABC transporters, drug resistance, and cancer stem cells. J Mammary Gland Biol Neoplasia. 14:3–9. 2009. View Article : Google Scholar : PubMed/NCBI

19 

Zhang B, Pan JS, Liu JY, Han SP, Hu G and Wang B: Effects of chemotherapy and/or radiotherapy on survivin expression in ovarian cancer. Methods Find Exp Clin Pharmacol. 28:619–625. 2006. View Article : Google Scholar

20 

Czabotar PE, Lessene G, Strasser A and Adams JM: Control of apoptosis by the BCL-2 protein family: Implications for physiology and therapy. Nat Rev Mol Cell Biol. 15:49–63. 2014. View Article : Google Scholar

21 

Restifo NP, Smyth MJ and Snyder A: Acquired resistance to immunotherapy and future challenges. Nat Rev Cancer. 16:121–126. 2016. View Article : Google Scholar : PubMed/NCBI

22 

Wicki A, Mandalà M, Massi D, Taverna D, Tang H, Hemmings BA and Xue G: Acquired resistance to clinical cancer therapy: A Twist in physiological signaling. Physiol Rev. 96:805–829. 2016. View Article : Google Scholar : PubMed/NCBI

23 

Solyanik GI: Multifactorial nature of tumor drug resistance. Exp Oncol. 32:181–185. 2010.

24 

Milla LN, Cogno IS, Rodríguez ME, Sanz-Rodríguez F, Zamarrón A, Gilaberte Y, Carrasco E, Rivarola VA and Juarranz A: Isolation and characterization of squamous carcinoma cells resistant to photodynamic therapy. J Cell Biochem. 112:2266–2278. 2011. View Article : Google Scholar : PubMed/NCBI

25 

Tu P, Huang Q, Ou Y, Du X, Li K, Tao Y and Yin H: Aloe-emodin-mediated photodynamic therapy induces autophagy and apoptosis in human osteosarcoma cell line MG-63 through the ROS/JNK signaling pathway. Oncol Rep. 35:3209–3215. 2016. View Article : Google Scholar : PubMed/NCBI

26 

Chau LY: Heme oxygenase-1: Emerging target of cancer therapy. J Biomed Sci. 22:222015. View Article : Google Scholar : PubMed/NCBI

27 

Furfaro AL, Traverso N, Domenicotti C, Piras S, Moretta L, Marinari UM, Pronzato MA and Nitti M: The Nrf2/HO-1 axis in cancer cell growth and chemoresistance. Oxid Med Cell Longev. 2016:19581742016. View Article : Google Scholar

28 

Ciesla M, Marona P, Kozakowska M, Jez M, Seczynska M, Loboda A, Bukowska-Strakova K, Szade A, Walawender M, Kusior M, et al: Heme oxygenase-1 controls an HDAC4-miR-206 pathway of oxidative stress in rhabdomyosarcoma. Cancer Res. 76:5707–5718. 2016. View Article : Google Scholar : PubMed/NCBI

29 

Lv X, Song DM, Niu YH and Wang BS: Inhibition of heme oxygenase-1 enhances the chemosensitivity of laryngeal squamous cell cancer Hep-2 cells to cisplatin. Apoptosis. 21:489–501. 2016. View Article : Google Scholar : PubMed/NCBI

30 

Soares HT, Campos JR, Gomes-da-Silva LC, Schaberle FA, Dąbrowski JM and Arnaut LG: Pro-oxidant and antioxidant effects in Photodynamic Therapy: Cells recognize that not all exogenous ROS are alike. ChemBioChem. 17:836–842. 2016. View Article : Google Scholar : PubMed/NCBI

31 

Wright KE, MacRobert AJ and Phillips JB: Inhibition of specific cellular antioxidant pathways increases the sensitivity of neurons to meta-tetrahydroxyphenyl chlorin-mediated photodynamic therapy in a 3D co-culture model. Photochem Photobiol. 88:1539–1545. 2012. View Article : Google Scholar : PubMed/NCBI

32 

Tian S, Yong M, Zhu J, Zhang L, Pan L, Chen Q, Li KT, Kong YH, Jiang Y, Yu TH, et al: Enhancement of the effect of methyl pyropheophorbide-a-mediated photodynamic therapy was achieved by increasing ROS via inhibition of Nrf2-HO-1 or Nrf2-ABCG2 signaling. Anticancer Agents Med Chem. 17:12017. View Article : Google Scholar

33 

Ishikawa T, Nakagawa H, Hagiya Y, Nonoguchi N, Miyatake S and Kuroiwa T: Key role of human ABC transporter ABCG2 in photodynamic therapy and photodynamic diagnosis. Adv Pharmacol Sci. 2010:5873062010.PubMed/NCBI

34 

Li C, Guo D, Tang B, Zhang Y, Zhang K and Nie L: Notch1 is associated with the multidrug resistance of hypoxic osteosarcoma by regulating MRP1 gene expression. Neoplasma. 63:734–742. 2016. View Article : Google Scholar : PubMed/NCBI

35 

Liu T, Li Z, Zhang Q, De Amorim Bernstein K, Lozano-Calderon S, Choy E, Hornicek FJ and Duan Z: Targeting ABCB1 (MDR1) in multi-drug resistant osteosarcoma cells using the CRISPR-Cas9 system to reverse drug resistance. Oncotarget. 7:83502–83513. 2016.PubMed/NCBI

36 

Ferrario A, Rucker N, Wong S, Luna M and Gomer CJ: Survivin, a member of the inhibitor of apoptosis family, is induced by photodynamic therapy and is a target for improving treatment response. Cancer Res. 67:4989–4995. 2007. View Article : Google Scholar : PubMed/NCBI

37 

Otto T and Sicinski P: Cell cycle proteins as promising targets in cancer therapy. Nat Rev Cancer. 17:93–115. 2017. View Article : Google Scholar : PubMed/NCBI

38 

Han T, Zhu X, Wang J, Zhao H, Ma Q, Zhao J, Qiu X and Fan Q: Establishment and characterization of a cisplatin-resistant human osteosarcoma cell line. Oncol Rep. 32:1133–1139. 2014.PubMed/NCBI

39 

Baldea I, Olteanu DE, Bolfa P, Tabaran F, Ion RM and Filip GA: Melanogenesis and DNA damage following photodynamic therapy in melanoma with two meso-substituted porphyrins. J Photochem Photobiol B. 161:402–410. 2016. View Article : Google Scholar : PubMed/NCBI

40 

Borgstahl GEO, Brader K, Mosel A, Liu S, Kremmer E, Goettsch KA, Kolar C, Nasheuer HP and Oakley GG: Interplay of DNA damage and cell cycle signaling at the level of human replication protein A. DNA Repair. 21:12–23. 2014. View Article : Google Scholar : PubMed/NCBI

41 

Casas A, Di Venosa G, Vanzulli S, Perotti C, Mamome L, Rodriguez L, Simian M, Juarranz A, Pontiggia O, Hasan T, et al: Decreased metastatic phenotype in cells resistant to aminolevulinic acid-photodynamic therapy. Cancer Lett. 271:342–351. 2008. View Article : Google Scholar : PubMed/NCBI

42 

Stuckey DW and Shah K: Stem cell-based therapies for cancer treatment: Separating hope from hype. Nat Rev Cancer. 14:683–691. 2014. View Article : Google Scholar : PubMed/NCBI

43 

Tirino V, Desiderio V, d'Aquino R, De Francesco F, Pirozzi G, Graziano A, Galderisi U, Cavaliere C, De Rosa A, Papaccio G, et al: Detection and characterization of CD133+ cancer stem cells in human solid tumours. PLoS One. 3:e34692008. View Article : Google Scholar

44 

Veselska R, Hermanova M, Loja T, Chlapek P, Zambo I, Vesely K, Zitterbart K and Sterba J: Nestin expression in osteosarcomas and derivation of nestin/CD133 positive osteosarcoma cell lines. BMC Cancer. 8:3002008. View Article : Google Scholar : PubMed/NCBI

45 

He A, Qi W, Huang Y, Feng T, Chen J, Sun Y, Shen Z and Yao Y: CD133 expression predicts lung metastasis and poor prognosis in osteosarcoma patients: A clinical and experimental study. Exp Ther Med. 4:435–441. 2012. View Article : Google Scholar : PubMed/NCBI

46 

Li J, Zhong XY, Li ZY, Cai JF, Zou L, Li JM, Yang T and Liu W: CD133 expression in osteosarcoma and derivation of CD133+ cells. Mol Med Rep. 7:577–584. 2013. View Article : Google Scholar

47 

Ricci-Vitiani L, Lombardi DG, Pilozzi E, Biffoni M, Todaro M, Peschle C and De Maria R: Identification and expansion of human colon-cancer-initiating cells. Nature. 445:111–115. 2007. View Article : Google Scholar

48 

Robey RW, Polgar O, Deeken J, To KW and Bates SE: ABCG2: Determining its relevance in clinical drug resistance. Cancer Metastasis Rev. 26:39–57. 2007. View Article : Google Scholar : PubMed/NCBI

49 

Ding XW, Wu JH and Jiang CP: ABCG2: A potential marker of stem cells and novel target in stem cell and cancer therapy. Life Sci. 86:631–637. 2010. View Article : Google Scholar : PubMed/NCBI

50 

Di Fiore R, Santulli A, Ferrante RD, Giuliano M, De Blasio A, Messina C, Pirozzi G, Tirino V, Tesoriere G and Vento R: Identification and expansion of human osteosarcoma-cancer-stem cells by long-term 3-aminobenzamide treatment. J Cell Physiol. 219:301–313. 2009. View Article : Google Scholar : PubMed/NCBI

51 

Jaggupilli A and Elkord E: Significance of CD44 and CD24 as cancer stem cell markers: An enduring ambiguity. Clin Dev Immunol. 2012:7080362012. View Article : Google Scholar : PubMed/NCBI

52 

Spring BQ, Rizvi I, Xu N and Hasan T: The role of photodynamic therapy in overcoming cancer drug resistance. Photochem Photobiol Sci. 14:1476–1491. 2015. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

November-2017
Volume 51 Issue 5

Print ISSN: 1019-6439
Online ISSN:1791-2423

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Tao Y, Ou Y, Yin H, Chen Y, Zhong S, Gao Y, Zhao Z, He B, Huang Q, Deng Q, Deng Q, et al: Establishment and characterization of human osteosarcoma cells resistant to pyropheophorbide-α methyl ester-mediated photodynamic therapy. Int J Oncol 51: 1427-1438, 2017
APA
Tao, Y., Ou, Y., Yin, H., Chen, Y., Zhong, S., Gao, Y. ... Deng, Q. (2017). Establishment and characterization of human osteosarcoma cells resistant to pyropheophorbide-α methyl ester-mediated photodynamic therapy. International Journal of Oncology, 51, 1427-1438. https://doi.org/10.3892/ijo.2017.4136
MLA
Tao, Y., Ou, Y., Yin, H., Chen, Y., Zhong, S., Gao, Y., Zhao, Z., He, B., Huang, Q., Deng, Q."Establishment and characterization of human osteosarcoma cells resistant to pyropheophorbide-α methyl ester-mediated photodynamic therapy". International Journal of Oncology 51.5 (2017): 1427-1438.
Chicago
Tao, Y., Ou, Y., Yin, H., Chen, Y., Zhong, S., Gao, Y., Zhao, Z., He, B., Huang, Q., Deng, Q."Establishment and characterization of human osteosarcoma cells resistant to pyropheophorbide-α methyl ester-mediated photodynamic therapy". International Journal of Oncology 51, no. 5 (2017): 1427-1438. https://doi.org/10.3892/ijo.2017.4136