Open Access

Expression and phosphorylation of stathmin correlate with cell migration in esophageal squamous cell carcinoma

  • Authors:
    • Fei Liu
    • Yu-Lin Sun
    • Yang Xu
    • Fang Liu
    • Li-Shun Wang
    • Xiao-Hang Zhao
  • View Affiliations

  • Published online on: November 28, 2012     https://doi.org/10.3892/or.2012.2157
  • Pages: 419-424
  • Copyright: © Liu et al. This is an open access article distributed under the terms of Creative Commons Attribution License [CC BY_NC 3.0].

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


Abstract

Microtubules play extensive roles in cellular processes, including cell motility. Stathmin is an important protein which destabilizes microtubules. The essential function of stathmin is closely associated with its phosphorylation status. Stathmin is overexpressed in many human cancers and has a significant relationship with clinical characteristics such as grade, tumor size and prognosis. We demonstrated that stathmin was overexpressed in ESCC tissues using both 2-DE and immunohistochemistry analysis. In addition, overexpression of stathmin was significantly correlated with histological grade in ESCC. However, no correlation was found with age, gender and lymph node metastasis. Knockdown of stathmin with siRNA impaired cell migration in KYSE30 and KYSE410 cells. When EC0156 cells were treated with paclitaxel, stathmin was stably phosphorylated and migration was impaired. These observations suggest that stathmin may have a more important function in ESCC development and migration. The present study provides further understanding of the importance of stathmin in ESCC therapy or diagnosis.

Introduction

Esophageal cancer (EC) is the eighth most common cancer worldwide (1), while esophageal squamous cell carcinoma (ESCC) is the predominant histological subtype in Asia, especially in China. It is characterized by high incidence and mortality rate (1,2). We found that stathmin is a differentially expressed protein between cancer and adjacent normal tissues in ESCC using proteomic technology.

Stathmin is an important protein which destabilizes microtubules (3,4). Microtubules are essential for many cellular processes, including mitosis, intracellular transport, supportment of cell shape and cell motility, and stathmin playes an important role in the regulation of microtubule which was involved in the construction and function of the mitotic spindle (5). Phosphorylation of stathmin led to a loss of the microtubule-destabilizing activity (68). Inhibition of stathmin phosphorylation produced strong mitotic phenotypes characterized by disassembly and disorganization of mitotic spindles and abnormal chromosome distributions (6). Stathmin phosphorylation gradient was necessary for correct spindle formation. Gradients of diffusible morphogens are known to be crucial for the supracellular self-organization of tissues and organisms (9).

Many studies had reported that stathmin was overexpressed across a broad range of human cancers, including acute leukemia, lymphoma, neuroblastoma and ovarian, prostate, breast and lung cancer (1012).

The upregulation of stathmin in ESCC was reported and associated with differentiation degree, lymph node metastasis, invasive depth and TNM stage (13,14). Wang et al demonstrated that knockdown of stathmin by antisense oligonucleotide can inhibit the proliferation of ECa109 cells (15). The expression and exact biological function of stathmin in ESCC, especially motility, remained largely unclear.

Materials and methods

Cell culture

ESCC cell lines EC0156 was established by our laboratory (16). KYSE30, KYSE140, KYSE150, KYSE170, KYSE180, KYSE410 and KYSE510 were donated by Dr Y. Shimada. EC0156 was cultured in Dulbecco’s modified Eagle’s medium (HyClone, UT, USA) supplemented with 10% fetal calf serum, 100 U/ml penicillin and 100 μg/ml streptomycin (Gibco, NY, USA). The other cell lines were cultured in RPMI-1640 medium supplemented with 10% heat-inactivated fetal bovine serum. All the cells were incubated at 37°C in a humidified atmosphere of 5% CO2.

Tissue specimens

From January 1999 to 2002, 8 pairs of specimens for two dimensional electrophoresis were obtained from surgically resected ESCC tissues in Cancer Hospital of Chinese Academy of Medical Sciences (CAMS). Another 50 tissues specimes for immnohistochemistry were also obtained from surgically resected esophageal carcinoma in Cancer Hospital of CAMS from January 1999 to 2009. Tissue specimens (n=93) for immunohistochemistry were purchased as microarray (Outdo Biotech Co., Shanghai, China). All specimens were frozen immediately, stored at liquid nitrogen. The median age of the patients of the 143 ESCC tissues for immunochemistry was 60 years (range, 29–84 years).

Protein extraction and quantification

Approximately 1×107 cells were grown to 80% confluence and washed six times in 1X phosphate-buffered saline (PBS). Soluble proteins were extracted with lysis (50 mM Tris-HCl pH 7.4, 150 mM NaCl, 1% Triton X-100, 0.1% SDS) with protease inhibitor cocktail (2.5 μM AEBSF, 0.04 μg/ml aprotinin, 0.04 μg/ml leupeptin, 1 mM EDTA ) by super-sonication and followed by centrifugation at 12,000 g for 15 min. Protein concentration was measured by the Bradford method.

Two dimensional electrophoresis

The soluble proteins from individual tissue specimen and the pooled tissue samples were separated by two dimensional electrophoresis (2-DE). Commercial IPG strips (pH 3–10NL, 18 cm; Amersham Biosciences, Uppsala, Sweden), were rehydrated overnight with 450 μl solution containing 8 M urea, 2% w/v CHAPS, 20 mM DTT, 0.5% v/v IPG buffer, 0.002% bromophenol blue and 1000 μg protein. Electrofocusing was carried out for 60 kVh at 20°C following the manufacturer’s instruction. Prior to the second dimension, the IPG strips were equilibrated for 30 min with 50 mM Tris-HCl pH 8.8, 6 M urea, 30% v/v glycerol, 2% w/v SDS, and a trace of bromophenol blue followed by reduction with 1% of DTT and alkylation with 2.5% of iodoacetamide. The IPG strips were placed into 12% SDS-polyacrylamide gels (26×20 cm) and were further electrophoresed by an EttanII-DE system (Amersham Biosciences) with a programmable power control, 0.5 h at 0.5 W per gel, then at 15 W per gel until the dye front reached the gel bottom. The separated proteins were visualized by Coomassie Brilliant Blue staining.

Protein identification

Briefly, images of the stained gels were acquired with an Image Scanner (Amersham Biosciences) using transmissive light. The gel images were first analyzed by eyes and subsequently were analyzed by ImageMaster 2D Elite 4.01 (Amersham Biosciences). Protein spots with signals differential intensity reaching 2.0 in 2D gels were excised, then were digested in modified trypsin solution with a final substrate-to-trypsin ratio of 40:1 (W:W) (in 25 mM ammonium bicarbonate). The digested peptides from 2-DE gel spots were analyzed by MALDI-TOF-MS using an Ettan-MALDI-TOF system (Amersham Biosciences). Monoisotopic peptide masses obtained from MALDI-TOF were used to search the NCBInr protein database using Mascot algorithm.

Immunohistochemistry

For immunohistochemical staining multiple tissue arrays (MTA) of formalin-fixed and ESCC and their matched adjacent normal tissues were incubated with stathmin mAb (3352, Cell Signaling Technology, UK) or control IgG. After washing with 1X PBS, slides were reacted with the biotin-labeled second antibody and then visualized using an ultrasensitive streptavidin-peroxidases system (Maxim Biotech, Fuzhou, China). Semi-quantitative criteria of the stathmin immunoreaction were modified according to previous publications (17,18). Immunostaining was scored as follows: 0, negative; 1, weak; 2, moderate; and 3, strong staining. The percentage of stathmin staining area was graded as 0, no positive staining; 1, <10%; 2, 10–50%; or 3, 50–100%. The staining index was calculated as the multiples of staining intensity and staining area, as described (18,19) The staining index <1 was considered as negative, while 1–4 as weak and >4 as strong.

Western blot analysis

Western blotting was performed as previously described (20). In briefly, protein extracted from cell lines and tissues specimens were separated using 12% SDS-PAGE, then transferred to polyvinylidene fluoride (PVDF) membranes (Millipore, Bedford, MA). Membranes were blocked by 10% skim milk in 1X PBS. The membranes were incubated with the primary antibodies against stathmin (ab52630, Abcam, UK) or β-actin (Cat. No. A-5316, Sigma, MO) in suitable dilutions. Secondary antibodies were anti-rabbit IgG and anti-mouse IgG, respectively. Signals were detected by chemiluminescence using the ECL kit.

siRNA transfection

The double-strand small interfering RNAs (siRNAs) were synthesized in duplex and purified forms using Genechem Co. (Shanghai, China). siRNAs targeting stathmin [5′-GAAACGAGAGCACGAGAAAtt-3′ (forward) and 5′-UUUCUCGUGCUCUCGUUUCtt-3′ (reverse)] and non-specific scramble siRNA oligonucletide sequences [5′-UUC UCCGAACGUGUCACGUtt-3′ (forward) and 5′-ACGUGAC ACGUUCGGAGAAtt-3′ (reverse)] were transfected separately into KYSE30 and KYSE410 cells using Lipofectamine 2000 (Invitrogen, CA, USA) according to manufacturer′s protocol. After 72 h, proteins were extracted.

In vitro wound-healing assay

KYSE30, KYSE410 or EC0156 cells were incubated overnight yielding confluent monolayer for wounding. Wounds were scratched using a pipette tip, photographs were taken immediately (time 0 h) and 24 h after wounding. The distance migrated by the cell monolayer to close the wounded area during this time period was measured. Photos were taken by Leica DMCI microscope (Leica, German) at ×100 magnification.

Statistical analysis

Data were analyzed by SPSS 16.0 (SPSS Inc., Chicago, IL, USA). χ2 test was used to analyze the relationship between stathmin expression with clinicopathologic characteristics. P-values <0.05 were considered as statistically significant.

Results

The identification of stathmin in ESCC tissues

Firstly, we analyzed differientially expressed proteins using proteomic method between tumor and corresponding normal tissues in ESCC. Results showed stathmin was detected in all 2-DE gels of the 8 pairs of ESCC tissues (Fig. 1), as an obvious differentially expressed spot.

Stathmin is overexpressed in ESCC tissues

Subsequently, we employed immunohistochemistry to analyze the expression of stathmin in 143 ESCC tissue microarray using the anti-stathmin antibody. Strong staining was seen in ESCC tissues, however, the normal tissues were weaker or negatively stained (Fig. 2). The strong staining of stathmin in cancer was 70.63% (101/143), while normal in 27.97% (40/143). In addition, the weak staining of stathmin in cancer was 20.28% (29/143), while normal in 15.38% (22/143). There was a significantly different staining mode between cancer and normal (P<0.05). Statistical analysis showed that overexpression of stathmin was significantly correlated with histological grade (P<0.05) (Table I). However, no correlation was found between stathmin expression and age, gender and lymph node metastasis (P>0.05) (Table I).

Table I

The correlation between stathmin expression and clinicopathological characteristics in ESCC specimens.

Table I

The correlation between stathmin expression and clinicopathological characteristics in ESCC specimens.

Stathmin

CharacteristicsAll cases (n) (%)Negative (%)Weak (%)Strong (%)P-value
Age>0.05
 ≥6080 (55.94)7 (8.75)14 (17.50)59 (73.75)
 <6063 (44.06)6 (9.52)15 (23.81)42 (66.67)
Gender>0.05
 Male108 (75.52)11 (10.19)24 (22.22)73 (67.59)
 Female35 (24.48)2 (5.71)5 (14.29)28 (80.00)
Tumor grade0.019
 Well differentiated36 (25.17)8 (22.22)8 (22.22)20 (55.56)
 Moderately differentiated80 (55.94)5 (6.25)16 (20.00)59 (73.75)
 Poor differentiated27 (18.89)0 (0.00)5 (18.52)22 (81.48)
Lymph node metastasisa>0.05
 Present38 (40.43)3 (7.90)3 (7.90)32 (84.20)
 Not present56 (59.57)2 (3.57)7 (12.50)47 (83.93)

a Only 94 cases have lymph node information.

siRNA-mediated reduction in stathmin expression resulted in impaired cell migration

We chose two ESCC cell lines (KYSE30 and KYSE410) as a model. SiRNA was employed to knockdown stathmin, and western blotting to detect the effect of siRNA. The results showed that the expression of stathmin was reduced after transfection of siRNA oligonucleotide for 72 h in KYSE30 and KYSE410 (Fig. 3A). Subsequently, wound-healing assay showed that the speed of wound recovery of siRNA stathmin was much slower than the scramble in both KYSE30 and KYSE410 (Fig. 3B). The results revealed that cell migration was impaired when deficient of stathmin.

The phosphorylation of stathmin reduced the motility of EC0156

Considering phosphorylation had been closely associtated with the function of stathmin, we wondered whether stathmin phosphorylation has influence on ESCC cell lines. We selected EC0156 for the next study.

EC0156 was treated with paclitaxel, a compound extracted from taxus plants, at gradient dosage from 0 to 1.6 μg/ml. The results revealed that the 19 kDa band of stathmin was decreased after paclitaxel treatment, whereas a new band at about 21 kDa gradiently increased in a dose-dependent manner. We confirmed the new band was phosphorylated stathmin at Ser-16. The tendency increased dramatically between 0.1 and 1.6 μg/ml, reaching a peak at a dosage of 0.8 μg/ml. Conversely, KYSE30 and EC0156 without treatment had no changes (Fig. 4A). Besides, rounded cells were observed in most of the EC0156 through microscopic analysis (Fig. 4B). The number of cells appered slightly reduced. Wound-healing assay showed that the speed of wound recovery of EC0156 cells treated with pacitaxel was much slower than the control, suggesting that cell migration was impaired after stabilized phophorylation of stathmin (Fig. 4C).

Discussion

Previously, we analyzed differientially expressed proteins using proteomic methods between tumor and nomal tissues in ESCC. We found an obviously overexpressed spot in all tumor 2-DE gels, which was identified as stathmin. Stathmin is ubiquitous, highly conserved 19 kDa cytosolic phosphoprotein that regulates microtubule dynamics (21).

We employed immunohistochemistry to detect the expression of stathmin in ESCC tissue microarray. Here we demonstrated strong expression in 77.14% (108/140) of ESCC tissues. In addition, overexpression of stathmin was significantly correlated with histological grade (P<0.05). However, no correlation was found with age, gender and lymph node metastasis (P>0.05). Wang et al revealed that the positive rate of stathmin in 75 ESCC samples was 81.3% and the relative contents of stathmin were significantly correlated with the differentiation degree, lymph node metastasis, invasive depth and TNM stage of ESCC (13). In addition, the basic tendency of stronger stathmin staining in ESCC was consistent with our previous study, which only used 13 ESCC samples (22). So far, many studies have demonstrated stathmin to be overexpressed in many human cancers, including mesothelioma tumor (11), malignant pheochromocytomas (23,24), cervical carcinoma (25), primary nasopharyngeal carcinoma (26), gastric cancer (27), hepatocellular carcinoma (28,29), medulloblastoma (30), endometrial cancer (18) and urothelial carcinoma (31). Above all, overexpression of stathmin was significantly correlated with clinical stage, tumor grade and lymph node metastasis in cancers, including cervical carcinoma, nasopharyngeal carcinoma, gastric cancer, hepatocellular carcinoma and endometrial cancer (18,2530). Moreover, stathmin may be regarded as a survival prognosis factor in ovarian cancer (32), endometrial cancer (18) and urothelial carcinoma (31). Our previous work, although the ESCC cases were only 13 for immunohistochemistry to analyze the expression of stathmin, we still found the expression of stathmin was overexpressed in ESCC tissue, but there was no correlation with tumor grades (22). In the present study, we expanded the ESCC case numbers for immunohistochemitry. Finally, we found overexpression of stathmin was significantly correlated with histological grade. However, the relationship between stathmin expression and lymph node metastasis did not replicate. There may be two reasons. The limited cases with lymph node metastasis in tissue microarray may lead to information being lost in our study. Another factor worth considered is the tiny difference of the pathologic criteria. In addition, the case number of the present study was much greater than the previous reports in ESCC.

We demonstrated that the speed of wound recovery in stathmin knockdown cells was much slower than the scramble in both KYSE30 and KYSE410 cells, suggesting cell migration was impaired when deficient of stathmin. Similar report exists in gastic cancer, in which cell invasion was significantly reduced by stathmin siRNA in the Matrigel invasion assay (27). Stathmin depletion with siRNA caused significant inhibition of lamellipodia formation, which directed by Pak1-WAVE2-kinesin complex (33). The lamellipodia formation may promote cell migration and invasion. Thus, stathmin depletion might inactivate lamellipodia formation leading to reduction of cell migration.

We demonstrated that the band at 19 kDa of stathmin was decreased when EC0156 was treated by paclitaxel. Meantime, a new band at about 21 kDa appeared in a dose-dependent manner. Subsequently, the new band was confirmed as phosphorylated stathmin at Ser-16. The basic function of stathmin was closely associated with its phosphorylated state. The stathmin-tubulin interaction was dependent on phosphorylation of stathmin (34). Rapidly switching phosphorylation of stathmin regulated microtubule assembly (21). Phosphorylation at Ser-16 and Ser-63 strongly reduced stathmin-tubulin complex formation (35). The effects of stathmin on dynamic instability were strongly attenuated by phosphorylation at Ser-16-and Ser-63 (6). Many protein kinases, including CDC2 (36), MAP (37), Auro B (38) and BGLF4 (39), can phosphorylate stathmin. Stathmin is known to undergo phosphorylation at Ser16 upon cell stimulation, such as paclitaxel at low concentration (40). Paclitaxel is an anticancer drug which interferes with microtubules (41). In the present study, we found paclitaxel may induce stathmin phosphorylation at Ser-16 and influenced the function of stathmin.

Through microscopic analysis, rounded cells were observed in most of EC0156 after the treatment with paclitaxel. The total number EC0156 appeared slightly reduced. Besides, the wound recovery speed of EC0156 treated with paclitaxel was diminished compared to the control, suggesting that cell migration may be impaired after stathmin phophorylation. It is possibly that stabilized phosphorylation of stathmin may disrupt microtubule dynamics and finally impaired the motility of EC0156, which were supported by other reports. Migration can be viewed as a repeated sequences of events that include formation of pseudopodia protrusions, attachment and translocation of the cell body in the direction of the new adhesion sites. The microtubule dynamics instability is important to generate an asymmetrical microtubules array (42). Stronger or long time stabilized phosphorylation of stathmin may impair the microtuble dynamics. Cernuda-Morollon et al found that T-cell receptor (TCR)-induced stathmin phosphorylation prevented T cell polarization on ICAM-1 by increasing Rac1 activity and reduced the migration of T cell and changed microtubule dynamics leading to loss of migratory polarity (43). Di Paolo et al revealed that phosphorylation at three sites (Ser-16, Ser-25, Ser-63) completely inhibited the tubulin binding capacity of stathmin (35). This phenomenon was also observated in cancer cells. Belletti et al revealed that Ser-16 phosphorylation of stathmin enhanced sarcoma cell adhesion and inhibited sarcoma cell motility by mutant methods (44). The present study revealed that paclitaxel may act as a stimulus factor to induce the phosphorylaiton of stathmin leading to impairment of microtubule dynamics. Our observations provide new evidence for understanding the interaction between stathmin phosphorylation and microtubules.

Our observation at cell level was not in accordance with immunohistochemistry analysis. Except for information lost or cancer diversity, different microenvironment between cultured cells and tissues may also contribute. In summary, stathmin was overexpressed in ESCC tissues and overexpression of stathmin was significantly correlated with histological grade. In addition, deficiency or stabilized phosphorylation of stathmin both impaired the migration of ESCC cells. The present study might highlight the potential of stathmin in the therapy and diagnosis of ESCC.

Acknowledgements

We thank Dr Rong Wang in Mount Sinai School of Medicine and Dr Si-qi Liu in Beijing Genomics Institute for critical suggestion. This study was supported by National High-Tech R&D Program (No. 2012AA020206, 2012AA02A503), SKPBR (No. 2011CB910703) and NSFC (No. 91029725, 81071789, 81071811) of China.

References

1 

Jemal A, Siegel R, Ward E, Hao Y, Xu J and Thun MJ: Cancer statistics, 2009. CA Cancer J Clin. 59:225–249. 2009. View Article : Google Scholar

2 

Lin DC, Du XL and Wang MR: Protein alterations in ESCC and clinical implications: a review. Dis Esophagus. 22:9–20. 2009. View Article : Google Scholar : PubMed/NCBI

3 

Steinmetz MO: Structure and thermodynamics of the tubulin-stathmin interaction. J Struct Biol. 158:137–147. 2007. View Article : Google Scholar : PubMed/NCBI

4 

Belmont LD and Mitchison TJ: Identification of a protein that interacts with tubulin dimers and increases the catastrophe rate of microtubules. Cell. 84:623–631. 1996. View Article : Google Scholar : PubMed/NCBI

5 

Cassimeris L: The oncoprotein 18/stathmin family of microtubule destabilizers. Curr Opin Cell Biol. 14:18–24. 2002. View Article : Google Scholar : PubMed/NCBI

6 

Manna T, Thrower DA, Honnappa S, Steinmetz MO and Wilson L: Regulation of microtubule dynamic instability in vitro by differentially phosphorylated stathmin. J Biol Chem. 284:15640–15649. 2009. View Article : Google Scholar : PubMed/NCBI

7 

Marklund U, Larsson N, Gradin HM, Brattsand G and Gullberg M: Oncoprotein 18 is a phosphorylation-responsive regulator of microtubule dynamics. EMBO J. 15:5290–5298. 1996.PubMed/NCBI

8 

Amayed P, Pantaloni D and Carlier MF: The effect of stathmin phosphorylation on microtubule assembly depends on tubulin critical concentration. J Biol Chem. 277:22718–22724. 2002. View Article : Google Scholar : PubMed/NCBI

9 

Niethammer P, Bastiaens P and Karsenti E: Stathmin-tubulin interaction gradients in motile and mitotic cells. Science. 303:1862–1866. 2004. View Article : Google Scholar : PubMed/NCBI

10 

Liu Z, Lu H, Shi H, et al: PUMA overexpression induces reactive oxygen species generation and proteasome-mediated stathmin degradation in colorectal cancer cells. Cancer Res. 65:1647–1654. 2005. View Article : Google Scholar

11 

Kim JY, Harvard C, You L, et al: Stathmin is overexpressed in malignant mesothelioma. Anticancer Res. 27:39–44. 2007.PubMed/NCBI

12 

Nakashima D, Uzawa K, Kasamatsu A, et al: Protein expression profiling identifies maspin and stathmin as potential biomarkers of adenoid cystic carcinoma of the salivary glands. Int J Cancer. 118:704–713. 2006. View Article : Google Scholar : PubMed/NCBI

13 

Wang F, Wang LX, He W, Zhu LN, Zhao PR and Fan QX: Expression of stathmin in esophageal squamous cell carcinoma and its biological significance. Nan Fang Yi Ke Da Xue Xue Bao. 30:1552–1557. 2010.(In Chinese).

14 

Wang X, Fan QX, Zhao PR and Wang F: Expression and significance of stathmin in esophageal squamous cell carcinoma. J Basic Clin Oncol. 120:372–374. 2007.

15 

Wang F, Wang LX, Fan QX and Zhao PR: Inhibitory effects of ASODN of stathmin gene on cultured Eca109 cell lines. J Zhengzhou University (Medical Sciences). 43:414–418. 2008.

16 

Wang Q, Xu Y, Zhao X, et al: A facile one-step in situ functionalization of quantum dots with preserved photoluminescence for bioconjugation. J Am Chem Soc. 129:6380–6381. 2007. View Article : Google Scholar : PubMed/NCBI

17 

Kouzu Y, Uzawa K, Koike H, et al: Overexpression of stathmin in oral squamous-cell carcinoma: correlation with tumour progression and poor prognosis. Br J Cancer. 94:717–723. 2006.PubMed/NCBI

18 

Trovik J, Wik E, Stefansson IM, et al: Stathmin overexpression identifies high risk patients and lymph node metastasis in endometrial cancer. Clin Cancer Res. 17:3368–3377. 2011. View Article : Google Scholar : PubMed/NCBI

19 

Salvesen HB, Das S and Akslen LA: Loss of nuclear p16 protein expression is not associated with promoter methylation but defines a subgroup of aggressive endometrial carcinomas with poor prognosis. Clin Cancer Res. 6:153–159. 2000.PubMed/NCBI

20 

Sun Y, Mi W, Cai J, et al: Quantitative proteomic signature of liver cancer cells: tissue transglutaminase 2 could be a novel protein candidate of human hepatocellular carcinoma. J Proteome Res. 7:3847–3859. 2008. View Article : Google Scholar

21 

Rana S, Maples PB, Senzer N and Nemunaitis J: Stathmin 1: a novel therapeutic target for anticancer activity. Expert Rev Anticancer Ther. 8:1461–1470. 2008. View Article : Google Scholar : PubMed/NCBI

22 

Liu F, Liu F, Sun YL and Zhao XH: Significance of STMN1 expression in esophageal squamous cell carcinoma. World Chin J Digestol. 18:1306–1312. 2010.

23 

Sadow PM, Rumilla KM, Erickson LA and Lloyd RV: Stathmin expression in pheochromocytomas, paragangliomas, and in other endocrine tumors. Endocr Pathol. 19:97–103. 2008. View Article : Google Scholar : PubMed/NCBI

24 

Bjorklund P, Cupisti K, Fryknas M, et al: Stathmin as a marker for malignancy in pheochromocytomas. Exp Clin Endocrinol Diabetes. 118:27–30. 2010. View Article : Google Scholar : PubMed/NCBI

25 

Xi W, Rui W, Fang L, Ke D, Ping G and Hui-Zhong Z: Expression of stathmin/op18 as a significant prognostic factor for cervical carcinoma patients. J Cancer Res Clin Oncol. 135:837–846. 2009. View Article : Google Scholar : PubMed/NCBI

26 

Cheng AL, Huang WG, Chen ZC, et al: Identification of novel nasopharyngeal carcinoma biomarkers by laser capture microdissection and proteomic analysis. Clin Cancer Res. 14:435–445. 2008. View Article : Google Scholar : PubMed/NCBI

27 

Jeon TY, Han ME, Lee YW, et al: Overexpression of stathmin1 in the diffuse type of gastric cancer and its roles in proliferation and migration of gastric cancer cells. Br J Cancer. 102:710–718. 2010. View Article : Google Scholar : PubMed/NCBI

28 

Yuan RH, Jeng YM, Chen HL, et al: Stathmin overexpression cooperates with p53 mutation and osteopontin overexpression, and is associated with tumour progression, early recurrence, and poor prognosis in hepatocellular carcinoma. J Pathol. 209:549–558. 2006. View Article : Google Scholar

29 

Gan L, Guo K, Li Y, et al: Up-regulated expression of stathmin may be associated with hepatocarcinogenesis. Oncol Rep. 23:1037–1043. 2010.PubMed/NCBI

30 

Kuo MF, Wang HS, Kuo QT, et al: High expression of stathmin protein predicts a fulminant course in medulloblastoma. J Neurosurg Pediatr. 4:74–80. 2009. View Article : Google Scholar : PubMed/NCBI

31 

Lin WC, Chen SC, Hu FC, et al: Expression of stathmin in localized upper urinary tract urothelial carcinoma: correlations with prognosis. Urology. 74:1264–1269. 2009. View Article : Google Scholar : PubMed/NCBI

32 

Su D, Smith SM, Preti M, et al: Stathmin and tubulin expression and survival of ovarian cancer patients receiving platinum treatment with and without paclitaxel. Cancer. 115:2453–2463. 2009. View Article : Google Scholar : PubMed/NCBI

33 

Takahashi K and Suzuki K: Membrane transport of WAVE2 and lamellipodia formation require Pak1 that mediates phosphorylation and recruitment of stathmin/Op18 to Pak1-WAVE2-kinesin complex. Cell Signal. 21:695–703. 2009. View Article : Google Scholar : PubMed/NCBI

34 

Ravelli RB, Gigant B, Curmi PA, et al: Insight into tubulin regulation from a complex with colchicine and a stathmin-like domain. Nature. 428:198–202. 2004. View Article : Google Scholar : PubMed/NCBI

35 

Di Paolo G, Antonsson B, Kassel D, Riederer BM and Grenningloh G: Phosphorylation regulates the microtubule-destabilizing activity of stathmin and its interaction with tubulin. FEBS Lett. 416:149–152. 1997.PubMed/NCBI

36 

Moreno FJ and Avila J: Phosphorylation of stathmin modulates its function as a microtubule depolymerizing factor. Mol Cell Biochem. 183:201–209. 1998. View Article : Google Scholar : PubMed/NCBI

37 

Antonsson B, Kassel DB, Ruchti E and Grenningloh G: Differences in phosphorylation of human and chicken stathmin by MAP kinase. J Cell Biochem. 80:346–352. 2001. View Article : Google Scholar : PubMed/NCBI

38 

Gadea BB and Ruderman JV: Aurora B is required for mitotic chromatin-induced phosphorylation of Op18/Stathmin. Proc Natl Acad Sci USA. 103:4493–4498. 2006. View Article : Google Scholar : PubMed/NCBI

39 

Chen PW, Lin SJ, Tsai SC, et al: Regulation of microtubule dynamics through phosphorylation on stathmin by Epstein-Barr virus kinase BGLF4. J Biol Chem. 285:10053–10063. 2010. View Article : Google Scholar : PubMed/NCBI

40 

Andersen SS: Spindle assembly and the art of regulating microtubule dynamics by MAPs and Stathmin/Op18. Trends Cell Biol. 10:261–267. 2000. View Article : Google Scholar : PubMed/NCBI

41 

Mistry SJ, Bank A and Atweh GF: Synergistic antiangiogenic effects of stathmin inhibition and taxol exposure. Mol Cancer Res. 5:773–782. 2007. View Article : Google Scholar : PubMed/NCBI

42 

Honore S, Pasquier E and Braguer D: Understanding microtubule dynamics for improved cancer therapy. Cell Mol Life Sci. 62:3039–3056. 2005. View Article : Google Scholar : PubMed/NCBI

43 

Cernuda-Morollon E, Millan J, Shipman M, Marelli-Berg FM and Ridley AJ: Rac activation by the T-cell receptor inhibits T cell migration. PLoS One. 5:e123932010. View Article : Google Scholar : PubMed/NCBI

44 

Belletti B, Nicoloso MS, Schiappacassi M, et al: Stathmin activity influences sarcoma cell shape, motility, and metastatic potential. Mol Biol Cell. 19:2003–2013. 2008. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

February 2013
Volume 29 Issue 2

Print ISSN: 1021-335X
Online ISSN:1791-2431

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Liu F, Sun Y, Xu Y, Liu F, Wang L and Zhao X: Expression and phosphorylation of stathmin correlate with cell migration in esophageal squamous cell carcinoma. Oncol Rep 29: 419-424, 2013
APA
Liu, F., Sun, Y., Xu, Y., Liu, F., Wang, L., & Zhao, X. (2013). Expression and phosphorylation of stathmin correlate with cell migration in esophageal squamous cell carcinoma. Oncology Reports, 29, 419-424. https://doi.org/10.3892/or.2012.2157
MLA
Liu, F., Sun, Y., Xu, Y., Liu, F., Wang, L., Zhao, X."Expression and phosphorylation of stathmin correlate with cell migration in esophageal squamous cell carcinoma". Oncology Reports 29.2 (2013): 419-424.
Chicago
Liu, F., Sun, Y., Xu, Y., Liu, F., Wang, L., Zhao, X."Expression and phosphorylation of stathmin correlate with cell migration in esophageal squamous cell carcinoma". Oncology Reports 29, no. 2 (2013): 419-424. https://doi.org/10.3892/or.2012.2157