Identification of proteomic and metabolic signatures associated with chemoresistance of human epithelial ovarian cancer

  • Authors:
    • Wenjuan Wu
    • Qi Wang
    • Fuqiang Yin
    • Zhijun Yang
    • Wei Zhang
    • Hani Gabra
    • Li Li
  • View Affiliations

  • Published online on: August 10, 2016     https://doi.org/10.3892/ijo.2016.3652
  • Pages: 1651-1665
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Emerging drug resistance in epithelial ovarian cancer (EOC) thwarted progress in platinum‑based chemotherapy, resulting in increased mortality, morbidity and healthcare costs. The aim of this study was to detect the responses induced by chemotherapy at protein and metabolite levels, and to search for new plasma markers that can predict resistance to platinum‑based chemotherapy in EOC patients, leading to improved clinical response rates. Serum samples were collected and subjected to proteomic relative quantitation analysis and metabolomic analysis. Differentially expressed proteins and metabolites were subjected to bioinformatics and statistical analysis. Proteins that played a key role in platinum resistance were validated by western blotting and enzyme‑linked immunosorbent assay (ELISA). Metabolites that were the main contributors to the groups and closely with clinical characteristics were identified based on the database using nuclear magnetic resonance (NMR). In total, 248 proteins from two independent experiments were identified using isobaric tags for relative and absolute quantitation (iTRAQ)‑based quantitative proteomic approach. Among them, FN1, SERPINA1, GPX3 and ORM1 were chosen for western blotting and ELISA validation. Platinum resistance likely associated with differentially expressed proteins and FN1, SERPINA1 and ORM1 may play a positive role in chemotherapy. HPLC‑MS analysis of four groups revealed a total of 25,800 metabolic features, of which six compounds were chosen for candidate biomarkers and identified based on the database using NMR. The metabolic signatures of normal control (NC), platinum‑sensitive (PTS) and platinum‑resistant (PTR) groups were clearly separated from each other. Those findings may provide theoretical clues for the prediction of chemotherapeutic response and reverse of drug resistance, even lead to novel targets for therapeutic intervention.

Introduction

Ovarian cancer is the fifth leading cause of cancer deaths in women and has the highest overall mortality rate and poor 5-year survival. More than 90% of ovarian cancer cases are epithelial ovarian cancer (EOC), which represents a series of etiologically and molecularly distinct disease. Although early diagnosis and therapy are considered to be the most effective methods to improve the outcome of patients with any cancer, the majority of EOC patients often do not manifest clinical symptoms and receive medical intervention when their tumor cells have disseminated to the peritoneal cavity. Currently, the standard therapy for EOC is surgical resection followed by postoperative chemotherapy with carboplatin and paclitaxel (PTX). Despite initial responsiveness to cisplatin-based chemotherapy, surgical and chemotherapy is far from satisfactory and most patients eventually develop drug-resistant tumors and succumb to the recurrent disease. This is why the majority of EOC patients with advanced disease relapse within 5 years, and little progress has been made in improving overall survival rates. Previous studies have proposed numerous factors to influence drug-resistance, such as ATP-dependent efflux pumps, extracellular microenvironment, DNA repair mechanism, modification of the drug target, drug-induced cytotoxicity, disruptions in apoptotic signaling pathways and changes in the expression of protein associated with tumor resistance (1,2). Cisplatin has been used to treat various cancers primarily by causing DNA damage and has been accepted worldwide as a first-line anticancer drug for EOC chemotherapy. In this regard, to identify those patients who have potential recurrence and to overcome chemoresistance and therefore improving patient outcome are the serious challenges in the management of EOC patients. Nevertheless, none of the identified biomarkers for drug resistance have been proven acceptable for routine clinical use. Hence, identification of clinical reliable biomarkers has come to the forefront of investigation. Moreover, if a non-invasive but sensitive blood assay that can monitor responses to chemotherapy was available, it would be invaluable for guiding chemotherapy and greatly improving the overall survival rate of EOC patients.

Mass spectrometry (MS) is an important high-throughput, industrially stable, information-rich technique for profiling small molecular compounds and is widely used to assess potential diagnostic and prognostic biomarkers. We applied isobaric tags for relative and absolute quantitation (iTRAQ)-based quantitative proteomic approach (Fig. 1A) and HPLC-micrOTOF-Q II high-resolution mass spectrometer-based metabolic analysis to compare and identify proteins and metabolites with differential profile in normal control (NC) group, benign ovarian cyst (BOC), platinum-sensitive (PTS) and platinum-resistant (PTR) cohort of serum samples. The emergence of resistance to platinum-based therapy is the main clinical endpoint of this experiment. iTRAQ proteomic analysis that combines 2D-LC and MALDI-TOF-MS/MS is an established technique in which total proteins are enzymatically digested into a large array of small peptide fragments and then directly analyzed by liquid chromatography-mass spectrometry (LC-MS). A total of 64 proteins with different expression levels were identified. In the list of differentially identified proteins via this method, most of these proteins were in accordance with the previously published literatures and associated with cancers. In addition, to further explore the new biomarkers predicting the responses to cisplatin, four of these proteins (FN1, SERPINA1, ORM1 and GPX3) were confirmed in a large patient cohort using western blotting and commercial enzyme-linked immunosorbent assay (ELISA), respectively. Moreover, HPLC-micrOTOF-Q II MS coupled with multivariate analysis was utilized, good separations were obtained for PTR, PTS vs. health controls. Finally, six substances with low molecular weight were identified based on the database using nuclear magnetic resonance (NMR). Receiver operating characteristic (ROC) curve analysis was then used to elucidate the potential power of FN1, SERPINA1, ORM1 and six small molecular metabolites for discriminating between the PTS and PTR group. The findings of this study are expected to reveal new proteins and metabolites related to platinum resistance and to provide candidate biomarkers to predict clinical response to chemotherapy. However, we do need an effective serum marker to predict the patients who have no response to cisplatin chemotherapy and will progress or recur during or after chemotherapy that cannot be easily judged from ultrasound or CT scan. This prediction is fundamental since patients that are resistant might benefit from a different combinational chemotherapy.

Materials and methods

Human serum samples

After Institutional Review Board approval (Ethics Committees of the Affiliated Tumor Hospital of Guangxi Medical University, Nanning, China), we obtained specimens (Table I) between September 1998 and March 2013, including EOC and BOC specimens. EOC cases were assigned to the PTR and PTS group. The FIGO classification was used for clinical staging, and the Gynecologic Oncology Group criteria were used for histological grading. NC blood samples were voluntarily donated by healthy individuals. Patients were eligible to participate in this trial if they had a pathologically confirmed diagnosis of BOC or EOC. After giving informed consent, serum samples were collected by clean venipuncture, centrifuged for 10 min at 3,000 × g at 4°C and stored at −80°C until further analysis.

Table I

Clinical characteristics of 40 cases of samples used in screening, 129 cases in validating and 132 samples in comparing metabolomic profiles between NC, BOC, PTS and PTR groups.

Table I

Clinical characteristics of 40 cases of samples used in screening, 129 cases in validating and 132 samples in comparing metabolomic profiles between NC, BOC, PTS and PTR groups.

Clinical characteristicsNCBOCPTSPTRP-value
iTRAQ (n=40)
 No. of patients10101010-
 iTRAQ-labeled sample113 (run 1)114 (run 1)115 (run 1)116 (run 1)-
117 (run 2)118 (run 2)119 (run 2)121 (run 2)-
 Age (years) (mean ± SD)39.80±5.5344.1±18.1147.21±12.6444.85±16.17-
 Histological type
  Serous-6/104/106/10-
  Mucinous-2/102/101/10-
  Other-2/104/103/10-
 FIGO stage
  I–II--3/103/10-
  III–IV--7/107/10-
 Tumor grade
  Well-differentiated--1/102/10-
  Moderately-differentiated--1/101/10-
  Poorly-differentiated--8/107/10-
ELISA (n=129)
 No. of patients33-5244-
 Age (years) (mean ± SD)39.71±10.05-46.33±10.6547.08±10.85-
 Histological type
  Serous--26/5217/44-
  Mucinous--7/526/44-
  Other--19/5221/44-
 TNM stage
  I–II--19/522/44-
  III–IV--33/5242/44-
 Tumor grade
  Well-differentiated--14/527/44-
  Moderately-differentiated--9/5210/44-
  Poorly-differentiated--29/5227/44-
FN1 (mean ± SD)69.14±13.29-62.41±12.7871.08±13.190.004
ORM1 (mean ± SD)157.43±18.26-173.64±22.69221.12±34.600.000
SERPINA1 (mean ± SD)756.19±244.39-685.69±204.59816.26±245.530.021
Metabolomics (n=132)
 No. of patients4194537-
 Age (years) (mean ± SD)39.61±9.2542.56±15.4446.56±10.0347.41±12.46-
 Histological type
  Serous--23/4511/37-
  Mucinous--6/454/37-
  Other--14/4519/37-
  NA--2/453/37-
 FIGO stage
  I–II--13/453/37-
  III–IV--31/4532/37-
  NA--1/452/37-
 Tumor grade
  Well-differentiated--6/453/37-
  Moderately-differentiated--5/454/37-
  Poorly-differentiated--27/4520/37-
  NA--7/4510/37-
 Primary therapy outcome
  Success--22/4513/37-
  CR+PR--21/4523/37-
  SD+PD--2/451/37-
  NA

[i] NC, normal control; BOC, benign ovarian cyst; PTS, platinum-sensitive; PTR, platinum-resistant; iTRAQ, isobaric tags for relative and absolute quantitation; ELISA, enzyme-linked immunosorbent assay; NA, not available.

Chemicals and reagents

The iTRAQ™ Reagent kit and mass calibration standards were purchased from Applied Biosystems (Bedford, MA, USA). Sequencing grade trypsin was obtained from Promega (Madison, WI, USA). Amicon Ultra-15 Certifugal Filter Units (3 kDa) were purchased from EMD Millipore (Billerica, MA, USA). All the solvents and chemicals used in this experiment were of LC-MS or analytical grade. HPLC grade water and acetonitrile (ACN) were purchased from Merck KGaA (Darmstadt, Germany). BCA assay kit was purchased from Pierce Biotechnology, Inc. (Rockford, IL, USA). Methyl methanethiosulfonate (MMTS) and methanol were obtained from Thermo Fisher Scientific (Rockford, IL, USA). Triethylammonium bicarbonate (TEAB), trifluoroacetic acid (TFA), formic acid and α-cyano-4-hydroxycinnamic acid (CHCA) were all obtained from Sigma (St. Louis, MO, USA).

Proteomic analysis
Depletion of high abundant proteins

Pooled serum samples were depleted of the 14 most highly abundant proteins using antibody-based depletion with Human 14 Multiple Affinity Removal System (MARS Hu-14; Agilent Technologies, Inc., Palo Alto, CA, USA), according to the manufacturer’s instructions. Crude serum samples were thawed on ice. Equal amounts of blood (20 μl) from 10 individuals in each group were pooled. Thereafter, MARS Hu-14 column was used to deplete specific 14 high-abundant proteins ~94% of total protein mass from human serum (Fig. 1B). The total protein concentrations of the depleted sera were determined using the BCA Protein Assay Reagent Kit (Pierce Biotechnology, Inc.) (Fig. 1C).

iTRAQ labeling

Prior to iTRAQ analysis, aliquots of 100 μg protein from each of the four sample pools were reduced using dissolution buffer (0.5 M TEAB) to a volume of 20 μl. To each of the four pools, 1 μl denaturant (2% SDS) and 2 μl reducing reagent [50 mM tris(2-carboxyethyl)phosphine] were added. Each pool was incubated at 60°C for 1 h. Cysteines were stopped by adding 1 μl cystine-blocking reagent (200 mM MMTS in isopropanol), and samples were incubated for additional 10 min at room temperature. The samples were digested with Sequencing Grade Modified Trypsin (Promega) at a protein-to-trypsin ratio of 30:1, at 37°C overnight. After that, peptides from each of the four depleted serum pools were labeled with 8-plex iTRAQ reagents (AB SCIEX, Foster City, CA, USA) according to the manufacturer’s instructions. The labels were applied in the following order: NC pool (113 Da), BOC pool (114 Da), PTS pool (115 Da); PTR pool (116 Da), so as to run the same sample in duplicate in each run. The four labeled samples were then evaporated to a volume of roughly 30 μl using a SpeedVac Concentrator and combined as a mixture followed by cleaning up by a strong cation exchange (SCX) column.

2D-LC-ESI-MS/MS

The combined peptide sample was subjected to SCX chromatography employing a PolySulfoethyl A column (2.1×200 nm; PolyLC, Inc., Columbia, MD, USA), on a high-pressure LC-pump (1200 series; Agilent Technologies, Inc.). The mixed sample was diluted in 10 mM KH2PO4 (pH 3.0), 25% v/v ACN (mobile phase A). Peptides were eluted with a linear gradient of 0–500 mM KCl (mobile phase B: 25% v/v ACN, 10 mM KH2PO4, 500 mM KCl, pH 3.0) for 115 min at a flow rate of 0.2 ml/min. Fractions were collected at 2-min intervals, and 16 fractions were collected from 24.5 to 98.5 min. Each SCX fraction was desalted using C18 Spin Columns (The Nest Group, Inc., Southborough, MA, USA) following the manufacturer’s instructions and then vacuum centrifuged to dryness. The peptide fractions were separated on a nano-reverse-phase LC system (Tempo™ LC MALDI Spotting System; Applied Biosystems), using a Magic C18AQ column (150 mm × 200 μm, 3 μm, 200 Å; Michrom Bioresources, Inc., Auburn, CA, USA), at a flow rate of 2 μl/min. A binary gradient with buffer A (98% H2O, 2% ACN, and 0.1% TFA) and buffer B (2% H2O, 98% ACN, and 0.1% TFA) was employed as the mobile phase. The peptide solutions were first loaded for 20 min using buffer A only on the pre-column, and the separation occurred over a period of 110 min. The elution from the column was mixed in 1:1 ratio with 5 mg/ml CHCA with a flow rate of 2 μl/min, and spotted onto the MALDI plates in a 44×28 spot array format. MS and MS/MS analysis was performed on a TOF-TOF 5800 MALDI platform (Applied Biosystems). MS spectra were recorded in the positive-ion reflector mode covering 700/800–4000 mass-to-charge ratio (m/z) acquiring 1,500 laser shots per spectrum (30 subspectra of 50 shots). After screening of all LC-MALDI sample positions the fragmentation of automatically selected precursors was performed at a collision energy of 2 kV with collision-induced dissociation gas (air). Up to 20 of the most intense ion signals per spot position, characterized by an S/N >45, were selected as precursors for MS/MS acquisition.

Database searches and criteria

Peptide matching, protein identification, and relative protein quantification for the iTRAQ experiment were performed with ProteinPilot v4.0 software (Applied Biosystems) in which the paragon search algorithm was applied. MS/MS spectra were searched against the UniProt/Swiss-Prot database for species of Homo sapiens. The database was searched using the following parameters: trypsin was used as the digestion agent, MMTS as a fixed modification of cysteine, thorough as search effort, and biological modification as the ID focus. Identifications of proteins were only accepted with a ‘local false discovery rate (FDR)’ estimation of ≤5% and an unused ProtScore ≥1.3 (>95% CI). In addition, proteins were considered for further statistical analysis when meet the following standards: one or more unique peptides with 95% confidence had to be identified; proteins were considered up- or downregulated when their fold changes were >1.3 or <0.77. The results obtained from ProteinPilot were exported to Microsoft Excel for manual interpretation. The protein lists from the two iTRAQ experiments (run 1 and 2; Table II) were merged with ratios calculated to the reference pool.

Table II

Differentially expressed proteins identified by iTRAQ between PTR group compared to PTS group.

Table II

Differentially expressed proteins identified by iTRAQ between PTR group compared to PTS group.

Unused protein scoreSequence coverage (%)Accession no.NameGene symbolSpeciesiTRAQ ratioExpression pattern

Run 1Run 2
13.0147.80P02763α-1-acid glycoprotein 1ORM1aHuman3.322.34Up
4.5641.80P19652α-1-acid glycoprotein 2ORM2aHuman-2.33Up
8.5246.50D1MGQ2α-2 globin chainHBA2Human1.581.56Up
23.7472.70P02647Apolipoprotein A-I APOA1aHuman2.001.72Up
3.5268.00P02652Apolipoprotein A-IIAPOA2Human1.33-Up
2.9923.60Q13790Apolipoprotein FAPOFHuman1.36-Up
2.0030.00P08519 Apolipoprotein(a)LPAHuman-1.57Up
19.9135.80P22792Carboxypeptidase N subunit 2CPN2Human-1.45Up
8.2730.10P06276 CholinesteraseBCHEaHuman-1.31Up
1.6025.20P31146 Coronin-1ACORO1AHuman1.81-Up
1.8635.30Q5VVP7C-reactive protein, pentraxin-relatedCRPHuman3.822.46Up
17.6355.50E9KL23Epididymis secretory sperm binding protein Li 44a SERPINA1aHuman6.023.80Up
4.2344.00P02675Fibrinogen β chainFGBaHuman2.42-Up
111.1354.60P02751 FibronectinFN1Human2.52-Up
16.4355.20O75636 Ficolin-3FCN3Human-1.37Up
63.0979.10P00738 HaptoglobinHPaHuman2.712.59Up
18.6584.40D9YZU5Hemoglobin, βHBBaHuman1.761.46Up
1.4614.30P01880Ig δ chain C regionIGHDHuman-2.54Up
2.0339.30P01860Ig γ-3 chain C regionIGHG3Human1.361.47Up
3.8951.90P0CG05Ig λ-2 chain C regionsIGLC2Human-1.72Up
13.8841.20P01871Ig μ chain C regionIGHMHuman-1.37Up
224.1370.60P02751–8Isoform 8 of fibronectinFN1aHuman-2.35Up
1.3543.70Q9BWP8–9Isoform 9 of collectin-11COLEC11Human1.34-Up
25.7170.90P02750Leucine-rich α-2-glycoproteinLRG1Human-1.55Up
2.4319.00Q6Q3G8 Lysosomal-associated membrane protein 2, isoform CRA_b LAMP2aHuman1.81-Up
2.236.70Q6UXB8Peptidase inhibitor 16PI16Human-1.30Up
2.5647.40P06702Protein S100A9 S100A9aHuman1.36-Up
94.9182.10P02787 SerotransferrinTFaHuman1.631.58Up
75.0673.60P02768Serum albuminALBaHuman-1.50Up
3.2770.00E9PR14Serum amyloid A proteinSAA2Human4.572.63Up
2.0417.80E7ES66Uncharacterized proteinGP1BAHuman1.50-Up
5.5137.80B4E1D3Uncharacterized proteinFGBHuman-2.46Up
2.0120.40C9JC84Uncharacterized proteinFGGHuman-2.60Up
8.9921.90P04275von Willebrand factorVWFaHuman-1.33Up
499.3871.20P04114Apolipoprotein B-100APOBaHuman-0.74Down
3.8261.50B2R526Apolipoprotein C-IAPOC1Human0.63-Down
6.0065.40P02655Apolipoprotein C-IIAPOC2Human0.620.68Down
7.1861.50B0YIW2Apolipoprotein C-III variant 1 APOC3aHuman0.600.58Down
23.9577.30P02649Apolipoprotein EAPOEaHuman-0.60Down
8.9643.20Q96KN2β-Ala-His dipeptidaseCNDP1Human-0.71Down
2.0130.60P49913Cathelicidin antimicrobial peptideCAMPHuman0.71-Down
4.0021.70P00488Coagulation factor XIII A chain F13A1aHuman0.750.67Down
2.6021.90Q12860 Contactin-1 CNTN1aHuman0.72-Down
2.1022.60P22352Glutathione peroxidase 3GPX3aHuman-0.76Down
5.5243.20Q1KLZ0HCG15971, isoform CRA_a PS1TP5BP1aHuman0.620.39Down
50.4370.10P05546Heparin cofactor II SERPIND1Human-0.76Down
3.7831.50P26927Hepatocyte growth factor-like proteinMST1Human0.75-Down
110.7952.90P19823Inter-α-trypsin inhibitor heavy chain H2 ITIH2aHuman-0.77Down
1.6614.60Q9NPH3Interleukin-1 receptor accessory proteinIL1RAPHuman0.52-Down
2.0036.20O14791–2Isoform 2 of apolipoprotein L1APOL1Human-0.76Down
4.0019.20P16070–5Isoform 5 of CD44 antigenCD44aHuman0.74-Down
204.2580.80B7ZKJ8ITIH4 proteinITIH4Human-0.59Down
44.3554.80P01042 Kininogen-1KNG1aHuman0.77-Down
8.2943.20Q5SQS3Mannan-binding lectinMBL2aHuman0.690.59Down
1.3311.80P04180 Phosphatidylcholine-sterol acyltransferaseLCATHuman0.64-Down
2.0013.60Q53Y44 ProfilinPFN1Human-0.68Down
99.8376.50P00734 ProthrombinF2aHuman-0.65Down
21.2060.10P02743Serum amyloid P componentAPCSHuman-0.58Down
9.4327.90P27169Serum paraoxonase/arylesterase 1PON1aHuman0.66-Down
2.0017.70O00391Sulfhydryl oxidase 1 QSOX1aHuman-0.37Down
2.0023.10D3DUS9Triosephosphate isomeraseTPI1aHuman-0.70Down
3.0551.60P67936Tropomyosin α-4 chainTPM4Human-0.68Down
6.3126.60A6NHF2Uncharacterized proteinBTDHuman0.41-Down
2.0021.50E7EX29Uncharacterized protein YWHAZaHuman0.75-Down

a The peptides identified with 95% confidence.

{ label (or @symbol) needed for fn[@id='tfn3-ijo-49-04-1651'] } iTRAQ, isobaric tags for relative and absolute quantitation; PTR, platinum-resistant; PTS, platinum- sensitive.

Western blotting

To confirm the identity of the proteins discovered by iTRAQ, western blotting was performed (Fig. 3A–E). Briefly, equal volumes of non-depleted serum from NC, PTS and PTR individuals (n=9 per group) were electrophoretically separated by 10% SDS-PAGE and transferred to a 0.45-mm polyvinylidene fluoride membranes (EMD Millipore) using a Bio-Rad wet transfer apparatus. Anti-human ORM1 (2 μg/ml), FN1 (1:1,000 dilution), SERPINA1 (1:1,000) and GPX3 (1:500) antibodies were from R&D Systems, Inc. (cat. no. MAB3694), Sigma (cat. no. F3648), OriGene Technologies, Inc., (cat. no. TA500376) and Abcam (cat. no. ab27325). Secondary antibodies were DyLight 680 anti-mouse (cat. no. 072-06-18-06, 1:5,000 dilution; KPL, Inc.) and IRDye 680RD donkey anti-rabbit (cat. no. 926-32223, 1:5,000 dilution; LI-COR Biosciences). Membranes were blocked with 5% skim milk in phosphate-buffered saline (PBS) with 0.1% Tween-20 for 2 h at room temperature. The concentration of primary and secondary antibodies was consistent as recommended in the instructions. Then membranes were incubated with primary antibodies overnight at 4°C, followed by fluorescent secondary antibodies (1:5,000) for 1 h at ambient temperature. After washing three times in PBST, proteins were detected with Odyssey infrared imaging system (LI-COR Biosciences, Lincoln, NE, USA) following the manufacturer’s instructions.

ELISA

Based on the iTRAQ and western blotting findings above we selected four targets, SERPINA1, ORM1, GPX3 and FN1, the protein markers potentially associated with PTR, for the validation using ELISA method. ORM1, FN1, SERPINA1 ELISA kits were obtained from and utilized according to the manufacture’s instructions. ORM1 and FN1 serum samples were diluted 100-fold, and SERPINA1 serum samples were diluted 50-fold. All samples and standards were tested in triplicate. Absorbance was determined using Power Scan 4 multiplex microplate reader (DS Pharma Biomedical Co. Ltd., Osaka, Japan) and analysis of results was conducted by SPSS 16.0 software.

Statistical analysis

Statistical analyses were performed with SPSS 16.0 software (SPSS, Inc., Chicago, IL, USA). A comparative analysis of multiple groups was analyzed by one-way ANOVA or Kruskal-Wallis test and multiple comparisons were performed with the least significant difference test. Results are presented as means ± SD. ROC curves were used to determine the diagnostic value of the markers. P<0.05 was considered statistically significant.

Metabolical analysis
Serum sample preparation

Prior to serum preparation, samples were thawed on ice for 1 h. A total of 200 μl of serum (stored at −80°C) was resuspended with 800 μl cold CAN (stored at −20°C) mixed thoroughly, and precipitated on ice for 2 h. The samples were centrifuged at 4°C and 12,000 × g for 15 min. The supernatant was then transferred to a new tube. Subsequently, serums were lyophilized for 24 h using a freeze dryer (Beijing Songyuan Huaxing Biotechnology Co., Ltd., Beijing, China). Methanol (200 μl) was added to lyophilized samples, vortexed, sonicated for 5 min, and centrifuged (12,000 × g, 4°C, 5 min). The supernatant (150 μl) was collected for further analysis.

Metabolic signature via LC-MS

LC-MS analysis was performed using Agilent HPLC (1290 series) fitted with a Zorbax Rx-C8 column (5 μm, 150×2.1 mm; Agilent Technologies, Inc.) and coupled to a Bruker Daltonics’ micrOTOF-Q II high-resolution mass spectrometer. The flow rate was 0.25 ml/min, injection volume 5 μl and column temperature 30°C. The mobile phase was consisted of solvent A (0.1% formic acid, 99.9% water) and solvent B (0.1% formic acid, 99.9% ACN). HPLC conditions were 15% solvent B changing linearly to 40% solvent B over 5 min, to 80% solvent B over 10 min, 80% solvent B over 5 min, to 90% solvent B over 5 min, and then 90% solvent B over 15 min. Finally, mobile phase constituents reverted to starting conditions for 5 min re-equilibration. Total analysis time was 45 min. Mass spectral analysis was operated on a micrOTOF-Q II high-resolution mass spectrometer (Bruker Daltonics) linked to an Agilent HPLC (1290 series) by HyStar software (Bruker Daltonics). Electrospray ionization (ESI) (positive ion mode) was used to identify the molecular ion mass [M+H]. Source parameters are: ESI capillary voltage, 4,500 V; nebulizing gas pressure, 1 bar; drying gas flow, 6 l/min; and drying gas temperature, 220°C. Data were acquired in a mass range of 50–1,500 m/z.

Data analysis

Following LC-MS, raw MS data were converted into a matrix that is compatible with multivariate statistical analysis and interpretation by using an in-house set of tools, such as the Compass software package (Bruker Daltonics). Signals obtained from each sample in the chromatogram were segmented into a series of regions characterized by retention time and m/z using the Compass software, furthermore, the theoretical m/z values were compared with the experimental values from MS signals. Based on the exact m/z, elemental formulas were generated using the DataAnalysis software (Bruker Daltonics). C, H, N, O, P and S were the elements of the formulas. The lists of generated formulas were searched against the METLIN database (https://metlin.scripps.edu/) to identify compounds. Principal component analysis (PCA) was then operated utilizing Profile Analysis software (Compass software package; Bruker Daltonics).

Results

Proteomic analysis
Serum proteomic data analysis

To enhance the detection of the lower abundance proteins, most of the 14 abundant proteins were removed in equal volumes from each sample. Technical replicate samples were used to increase the reliability of the iTRAQ technique for relative quantitation. The relative expression levels, statistical parameters and the peptide information of identified proteins for each pool were obtained from two (replicate) peptide spectra data as described above. Subsequently, all the identified proteins were filtered with manually selected filter exclusion parameters. Thus, in the first iTRAQ data set (run 1), identification of 197 proteins was made. Similarly, 184 proteins were discovered in the second iTRAQ data set (run 2). The proteins identified from the two iTRAQ data sets were subsequently combined, and a total of 248 unique proteins were identified and quantified. Proteins were considered up- or downregulated when their ratios were >1.3 or <0.77 (Fig. 2C). Therefore 64 proteins were screened out as candidate biomarkers in one or two separate experiments as differentially expressed proteins between PTS and PTR sets: 33 of which were increased (PTR/PTS >1.3) and 31 were decreased (PTR/PTS <0.77). Candidate biomarkers selected by these criteria are summarized in Table II. For better understanding of the data structure in our experiment, a clustering algorithm for grouping proteins was required (Fig. 2B), and k-means clustering approach was then performed to group the data based on the degree of similarity between the PTS and PTR group. The different trends of identified proteins during PTS and PTR can be grouped into six subsets with a similar expression pattern (Fig. 2E). In the first subset, most of the 20 proteins were upregulated in a stepwise way in NC, BOC, PTS and PTR groups. Most are extracellular proteins involved in cell adhesion, cell communication and immune system process. The trend of these proteins differentially expressed in PTS and PTR groups were of interest as these could provide leads for potentially useful biomarkers of platinum status. In the second set, all the 10 proteins were specifically increased in PTS group and are therefore interesting candidates for EOC diagnosis or prognostic studies.

Gene ontology analysis

Data analysis of 64 unique proteins identified by two iTRAQ experiments was performed using the Blast2GO database (http://www.blast2go.com/b2ghome) and Web Gene Ontology Annotation Plot (WEGO: http://wego.genomics.org.cn/cgi-bin/wego/index.pl) to class each protein into its respective cellular components, molecular function and biological process (Fig. 2A). For the 64 differential proteins, the subcellular distributions were enriched mainly in extracellular region (90.6 and 78.1%) (the two numbers represent the upregulated and downregulated proteins proportion of the total, respectively), which imply that most of these proteins are secretary proteins. According to GO molecular function analysis, the top three common functional annotations were binding (87.5 and 93.8%), catalytic (28.1 and 56.3%), and enzyme regulator (25 and 31.3%). Most of the differential proteins were involved in biological regulation (90.6 and 84.4%), response to stimulus (81.3 and 87.5%), and pigmentation (84.4 and 71.9%). To clearly show the expression trend of differential proteins during cancer progress, k-means clustering method was used to classify the 64 protein. The results are shown in Fig. 2E. For further text mining, PANTHER Classification System (http://www.pantherdb.org/) was used to carry out the GO analysis. During tumor progression, 20 proteins were upregulated gradually in cluster A and most of these members participated in immune system process, cell adhesion, and cell communication, which would make sense in connection with drug resistance.

Pathway and biological interaction network analyses

Enrichment analysis of associated diseases and drugs was performed for the differentially expressed proteins that met our thresholds (fold, rank) using the IPAD browser (http://bioinfo.hsc.unt.edu/IPAD/) tools. Results of associated diseases and drug analysis showed that 30 proteins (14 upregulated, 16 downregulated) among them were significantly associated with EOC (indicated as pentagrams in Table II) and GPX3 was associated with cisplatin. In addition, DAVID Bioinformatics Resources 6.7 (http://david.abcc.ncifcrf.gov/home.jsp) was used to investigate possible interactions between the 30 proteins associated with EOC, which revealed that the differential proteins were significantly enriched in complement and coagulation cascades and ECM-receptor interaction. To model the signaling network potentially affected in the context of platinum status, the 17 focus proteins with fold changes between PTS and PTR group >1.5 were then subjected to network analysis using STRING software (http://string-db.org/). The network analysis identified ALB, APOA1, SERPINA1, FN1, ORM1 and TF as the major molecules affected in PTR patients (Fig. 2D). Candidate proteins with the most extreme deviation from the NC, BOC, PTS groups, including SERPINA1 (3.8-fold increased), ORM1 (3.32-fold increased), FN1 (2.35-fold increased) and cisplatin-associated GPX3 (1.35-fold decreased), were commonly identified in two iTRAQ experiments, and chosen for further analyses.

Co-occurrence analysis with COREMINE

COREMINE was used to performe co-occurrence analysis based on literature. The 64 differentially expressed proteins and the following list of keywords were used to interrogate the tools: drug resistance, neoplasm; drug resistance; drug resistance, multiple. In order to restrict the number of proteins potentially associated with drug resistance or MDR, p<0.01 was considered statistically significant. The cumulative frequency top 50 protein lists out of connected proteins, which showed a p<0.01 and the 64 differentially expressed proteins were compared to look for the degree of overlap. Finally, proteomic and co-occurrence analysis shared the following 11 proteins: ALB, CRP, FN1, S100A8, TF, VWF, APOC2, APOE, CAT, CD44, F2.

Western blotting

A total of 27 serum samples, composing 9 from NC group, 9 from PTS group, and 9 from PTR group, were subjected to western blotting against SERPINA1, ORM1, GPX3 and FN1. These proteins were selected for western blotting primarily the following factors: big fold changes of differential expression, correlation with cancer/drug resistance from a literature-based text mining, the expression trend in four pools and the availability of commercial antibodies. Our results indicated that three of the four candidates have similar trends with the proteomic results (SERPINA1, ORM1, FN1) in the serum of PTR cases, compared to PTS cases, which implied the credibility of proteomic analysis (Fig. 3A–D). One-way ANONA was applied to calculate means ± SD from each group along with p-values.

Clinical relevance of SERPINA1, ORM1, GPX3 and FN1

In the initial experiment 10 NC, 10 PTS and 10 PTR samples were used to validate the expression levels of SERPINA1, ORM1, GPX3 and FN1. The results illustrated that statistical significant difference between PTS and PTR was seen for SERPINA1, ORM1 and FN1, but not GPX3 (data not shown). However, the expression of GPX3, which was observed to be downregulated (FC=1.35) in proteomic analysis, was not significantly different (p>0.05) by ELISA. Consequently, we carried a full validation study for SERPINA1, ORM1 and FN1, using the entire 129 samples collected (data are shown in Table I). Consistent with the iTRAQ results in the previous experiment, relative quantitation of SERPINA1, ORM1 and FN1 (Fig. 3F–I) between PTS 52 samples and PTR 44 samples were all found to be significantly upregulated (p<0.05). Likewise, to further evaluate the diagnostic significance of these three proteins, a ROC curve analysis was constructed for each protein by plotting sensitivity vs. specificity. The overall predictive accuracy of each protein was reflected by the area under the ROC curve (AUC), a commonly used indicator for estimating the diagnostic efficacy of a potential biomarker. FN1 and SERPINA1 with ROC areas of 0.679 and 0.666, respectively, suggest that their use as a biomarker may not be reliable. Unlike the FN1 and SERPINA1, the AUC for ORM1 was 0.91 and its sensitivity and specificity for predicting PTR was 71 and 97.4%, respectively, which could clearly separate the PTS patients from the PTR individuals. These results highlight a potential role for ORM1 in the response to platinum therapy.

Metabolic analysis

A total of 25,800 metabolic features was observed in our study. Data of identified compounds were subjected to t-test analysis to identify significant metabolic patters and variations. Compounds having p<0.01 and fold-change >2 were considered as statistically significant. PCA and multivariate statistics were then applied to identify key PTR-associated metabolic perturbations in PTR compared to PTS. Unsupervised PCA of the resultant data showed clear metabolic separation of PTR from PTS along the first principal component, and clear distinctions of EOC (PTR and PTS) from healthy individuals along the second principal component (Fig. 4A). The BOC sera were not obviously grouped because of their limited sample size. PCA loading plots (Fig. 4B) provided six metabolite features contributing to the separation of groups along PC1 and PC2. Six known compounds were identified using NMR based on database (Table III). The levels of the six potential biomarkers in blood from PTR, PTS and NC group were determined by LC-MS/MS. Compared to PTS, PTR exhibit a specific metabolic trait characterized by decreased levels of calycanthidine and increased levels of 1-monopalmitin, ricinoleic acid methl ester, polyoxyethylene (600)mono-ricinoleate/glycidyl stearate. Furthermore, the concentration of dodemorph was higher and of C16 sphinganine was lower in the EOC compared to NC (Table III). ROC curve was used to calculate sensitivity and specificity of the four biomarkers for PTR compared with PTS. The AUC for 1-monopalmitin, ricinoleic acid methyl ester, polyoxyethylene (600)mono-ricinoleate and calycanthidine was 0.892, 0.900, 0.883 and 0.109, respectively, and their sensitivity and specificity for predicting PTR were 83.8 and 75%; 81.1 and 86.4%; 83.8 and 75%; 90.9 and 73%, respectively, which could clearly separate the PTS patients from the PTR individuals. The combinational four biomarkers achieved an AUC value (AUC=0.925) while the statistical analysis provided 86.5% sensitivity and 81.8% specificity for the prediction of PTR (Fig. 5).

Table III

Relevant analytical data for the metabolites identified in PTR, PTS, BOC and NC groups.

Table III

Relevant analytical data for the metabolites identified in PTR, PTS, BOC and NC groups.

Retention time (min)Adductm/zError (mDa)σ valueMolecular formulaTrendCAS/PubChemPossible metabolite
19.9[M+H]282.2807−1.50.0064 C18H35NOUp (EOC/NC)1593-77-7 Dodemorph
9.4[M+H]274.2777−3.60.0329 C16H35NO2Down (EOC/NC)4266342C16 sphinganine
16.9[M+H]313.2775−3.90.0334 C19H36O3Up (PTR/PTS)141-24-2Ricinoleic acid methyl ester
16.8[M+H]331.2871−2.80.006 C19H38O3Up (PTR/PTS)542-44-9 1-Monopalmitin
19.2[M+H]341.3071−2.10.0067 C21H40O3Up (PTR/PTS)977137-78-2Polyoxyethylene (600)mono-ricinoleate
19.2[M+H]341.3071−2.10.0067 C21H40O3Up (PTR/PTS)7460-84-6Glycidyl stearate
19.9[M+H]361.2419−4.60.0235 C23H28N4Down (PTR/PTS)5516-85-8 Calycanthidine

[i] PTR, platinum-resistant; PTS, platinum-sensitive; BOC, benign ovarian cyst; NC, normal control; m/z, mass-to-charge ratio; EOC, epithelial ovarian cancer.

Discussion

Cisplatin is one of chemotherapeutical agents commonly used to treat EOC, which causes DNA damage via forming inter- and/or intrastrand DNA adduct lesions and eventually cytotoxicity. However, the benefits of chemotherapy can be attenuated because of the emergence of platinum resistance. To eradicate the mechanisms of platinum resistance in EOC is a difficult task. The recent development of proteomic approaches applied to investigate drug-resistance mechanisms has greatly helped in addressing these issues. Comparative proteomic approach is a powerful tool, which might help to guide future research and cross validation of various proteomic profiling with a high throughput. In our experiment, a panel of 64 different proteins that have altered expression in PTR patients were compared to the parental PTS group using a shotgun quantitative proteomics approach, and four of these proteins were confirmed with western blotting and ELISA. The results of serum FN1, SERPINA1, GPX3 and ORM1 from 2D-LC-MS/MS analysis were validated in a 139 cohort using a different methodology. Western blotting and ELISA confirmed that the serum level of FN1, SERPINA1 and ORM1 was upregulated in PTR group, which indicated that the FN1, SERPINA1 and ORM1 serum levels might be a tool for screening and diagnosis of PTR. However, it should be noted that although the change in direction (up- or down-regulated) of GPX3 detected by western blotting between PTS and PTR group was consistent with iTRAQ, the changes measured by ELISA assay in 43 patients was not statistically significant (p>0.05). The differences in fold change determined by iTRAQ, western blotting and ELISA can be attributed to methodological factors such as the use of isobaric tags and/or differences inherent in the technical method. ROC curve analysis was applied to find the cut-off value of serum FN1, SERPINA1 and ORM1 to discriminate between PTS and PTR group. The sensitivity and specificity were calculated. ROC curves show, ORM1 with 71% sensitivity and 97.4% specificity could give a higher accuracy (Fig. 3). However, FN1 and SERPINA1 were not reliable for clinical diagnosis because of low sensitivity and specificity. Our comprehensive study of proteomics led to the possibility that monitoring the level of serum ORM1 could be clinically useful for the screening and diagnosis of PTR patients.

α-1-antitrypsin (SERPINA1) is an inhibitor of serine proteases principally secreted by hepatocytes, but also by monocytes, neutrophils, macrophages and alveolar epithelial cells, and plays a critical role in modulating host immunity, inhibiting T lymphocyte-mediated antitumor function and thereby accelerated tumor proliferation, and metastasis (35). Moreover, there have been numerous studies documenting a link between SERPINA1 and various cancers, although for most the mechanism for the linkage is unclear. Giving the expression levels of SERPINA1 in rat bladder tumor tissues were 2.5-fold higher than those in normal bladder tissues using two-dimensional difference gel electrophoresis (2D-DIGE) (6). Similar conclusions were also obtained in pancreatic tumors, hepatocellular carcinoma, non-small cell lung cancer, gastric cancer in gastric juice, prostate cancer patients and malignancy in insulinomas. Much attention has been focused on the role of SERPINA1 as a tumor suppressor, but no report has shown directly the relation between SERPINA1 and chemotherapy drugs. In our study, the contribution of SERPINA1 to drug resistance was implicated in human serum samples. SERPINA1 shows one of the largest fold increases (3.8-fold increased) in protein expression level in PTR cohort compared with PTS cohort (p<0.05). Nevertheless, our data are in good agreement with prior studies elicited above, but disagree with the results of Normandin et al (7).

Fibronectin is a glycoprotein that is involved in cell adhesion, signal transduction and migration processes including embryogenesis, wound healing, blood coagulation, host defense, and metastasis, especially possibly suppression of apoptosis (810). There have been many reports on the relation between FN1 and human tumors. Similar thesis reported that fibronectin was involved in Ras, Erk, Akt and ECM pathways and mediate various signals such as cancer cell adhesion, growth migration and invasion (11,12). Akiyama et al showed that FN1 played a causal role in tumor neovascularization and metastasis (13). In addition, a recent study found that FN1 is one of the key genes in regulating SOX2 cell migration, invasion, colony formation and drug resistance in ovarian cancer cells (1416). Qian et al also indicated that FN1 is targeted by let-7g to promote mammary carcinoma cell migration and invasion via p44/42 MAPK and MMPs (17). FN1 was also suggested as a marker for renal cell carcinoma aggressiveness (18,19). Moreover, FN1 was shown to be a direct target gene for miR-1 and miR-200. While miR-1 may play a role as a tumor suppressor gene in laryngeal carcinoma. Similarly, miR-200 is crucial for the maintenance of epithelial identity, behavior, and sensitivity to chemotherapy in ovarian cancer cell line (20,21), which confirmed our previous observation by miRNA microarrays with samples obtained from the same patients as this study (22). All these findings suggest that a functional relation is present between FN1 and platinum response, which supports our data in EOC.

As a member of guutathione peroxidases, GPX3 is located in 5q23 and has critical roles in the detoxification of hydrogen peroxide and other oxygen-free radicals. Previous studies have demonstrated that GPX3 had a broader downregulated pattern in a variety of cancers, such as ovarian, cervical, thyroid, head and neck, lung, colorectal, gastric, gallbladder, breast, and esophageal cancers than in healthy controls. These reports suggest that GPX3 contains a tumor-suppressor function. The mechanisms involved in mediating the GPX3 tumor-suppressor function are mainly due to promoter hypermethylation (23), the downregulation of c-Met expression (24,25), and the role of antioxidant enzymes which are involved in reactive oxygen species (ROS) metabolism. As a messenger molecule, ROS might increase cancer cell proliferation, genetic mutations, instability, and thereby invasion and angiogenesis (26). In addition, ROS also mediates the induction of tumor cell death via many chemotherapeutic agents such as platinum (27). Although the researchers failed to measure the serum concentration of GPX3, this statement is supported by our results. However, GPX3 is identified to be highly expressed in clear cell adenocarcinoma compared to control tissues at a DNA, mRNA and protein level on cell lines and clinical samples of ovarian clear cell adenocarcinoma (28,29). Although the molecular biological mechanism is not clarified, these results might indicate that GPX3 activity is tumor-specific. In our present study, GPX3 was shown to be downregulated in PTR group compared with PTS group, which confirmed the previous results, but the exact mechanism, in response to anticancer drugs remains to be further understood.

The α-1-acid glycoprotein primarily synthesized by the liver is an acute-phase reactant with immunomodulatory and immunosuppressive properties (30) and its serum levels are increased by inflammation, stress, and chronic disease such as cancer (31). Two main biological functions were involved in α-1-acid glycoprotein, binding and transporting of endogenous substances or drugs, and a strong immunomodulatory function. Previous investigations in patients with carcinoma of the breast, lung, ovary and endometrium have suggested that serum ORM1 concentrations were increased two times higher than that in healthy individuals, and ORM1 might act as blocking agent protecting tumor cells against immunological attack, thereby contributing to the ‘immune escape’ of the tumor (32,33). ORM1 can also interfere with cytokine function by inducing the secretion of soluble TNFα receptor and IL-1, -6 and -12 receptor antagonist (30,34). Although the mechanisms by which ORM1 mediates its functions are not fully understood, ORM1 has been shown to bind to the chemokine receptor CCR5 in macrophages, the asialoglycoprotein receptor in hepatocytes, the surface lectin-like receptor Siglec-5 in neutrophils and can also modulate TNFα-induced phosphorylation of p38 MAPK, MEK1/2, c-Jun N-terminal kinase which is required for angiogenesis in macrophages (3539), but not VEGF-induced signaling. In addition, ORM1 has been shown to enhance endothelial cell migration and capillary tube formation in vitro (40). Moreover, several reports suggested that the serum levels of α-1-acid glycoprotein influenced the pharmacokinetics (PK)/pharmacodynamics (PD) of chemotherapy drugs such as docetaxel, PTX and imatinib (4144). As these reports remarked, α-1-acid glycoprotein may function as a carrier of PTX from the serum into the liver via the α-1-acid glycoprotein receptors, and this might result in the enhancement of the PTX metabolism. Although ORM1 has been reported to be associated with cancers or metabolisms of chemotherapy drugs according to previous reports, no studies have underlined the importance of ORM1 in cisplatin resistance in PTR patients, and this is the first time that ORM1 was identified as an important biomarker of response to cisplatin-based chemotherapy. The mechanism of this phenomenon may be attributed to the PK/PD changes of cisplatin, however, further studies will be required to fully understand ORM1 functional roles in drug resistance.

In the present study, we undertook a non-destructive metabolomic technique (HPLC-micrOTOF-Q II MS/MS) to investigate the metabolic traits. All the six metabolites in our experiment were identified as fatty acid or derivatives. Profiling of metabolomics elucidated changes in the levels of fatty acid metabolism, which confirmed our previous observations by proteome approach and conclusions of many addressed articles on chemotherapeutic resistance and metabolism, and served as an insightful reference to the mechanism research of drug resistance. Fatty acid synthese (FASN) providing proliferating cancer cell lipids for membrane biogenesis was assumed to have metabolic characteristics of cancel cells (45). Expression level of FASN is significantly upregulated in kinds of neoplasm and correlates with poor prognosis, but in a health individual is very low even undetectable, suggesting that FASN serves as a metabolic oncogene (46). Fatty acids were used by proliferating tumor cells for membrane assembly, lipid modifications of proteins, and as an efficient source of energy, all are required to sustain neoplasm growth and survival (47). Furthermore, it is shown that FASN is overexpressed in drug-resistant breast neoplasm cell line (MCF7/AdVp3000), and that reducing the expression of FASN increased the drug sensitivity in MCF7 and MDA-MB-468 (breast cancer cell lines) (48). Analogously, FASN was reported to be associated with acquired trastuzumab/docetaxel/5-fluorouracil resistance in breast cancer or radiation and gemcitabine in pancreatic neoplasm. FASN also played an active role in chemotherapy resistance of HER-2/neu-induced breast neoplasm. FASN not only played a key role in acquired resistant phenotype but also in inherent resistant phenotype in hepatocellular carcinoma (49). Roodhart et al identified two platinum-induced polyunsaturated fatty acids which induce resistance to chemotherapeutic drugs. When the central enzymes associated with the production of polyunsaturated fatty acids were blocked, the mesenchymal stem cells induced resistance which was prevented (50). All the above further confirmed the metabolism abnormality of fatty acid is induced by PTR.

In conclusion, we identified a panel of new ovarian epithelial cancer serum protein biomarkers, which have an indicator value for platinum status and allow patients who have a high chance of being resistant to cisplatin-based chemotherapy to receive an alternative therapy. Although thousands of metabolites were identified, links were weak and annotated only a small proportion of the total analytes. In further studies, the role of these differentially proteins or compounds in cisplatin resistance needs to be validated on a large scale to evaluate the clinical benefit of using these candidate biomarkers for diagnosis or prognosis analyses. The contribution of the identified biomarkers in cisplatin resistance should also be explored to help understand and design chemosensitizing agents. In addition, our study demonstrated that metabolomics and proteomics could validate one another partially and their combination could better elucidate the mechanism of drug resistance and provide candidate molecular targets for personalizing therapeutic interventions and treatment efficacy monitoring.

Acknowledgements

This study was supported by the National Natural Science Foundation of China (grant no. 81572579), Guangxi Scientific Research and Technological Development Program Topics (no. 14124004) and the Specialized Research Fund for the Doctoral Program of Higher Education (nos. 20124503110003 and 2013.01–2016.10).

References

1 

Leonessa F and Clarke R: ATP binding cassette transporters and drug resistance in breast cancer. Endocr Relat Cancer. 10:43–73. 2003. View Article : Google Scholar : PubMed/NCBI

2 

Gottesman MM, Fojo T and Bates SE: Multidrug resistance in cancer: Role of ATP-dependent transporters. Nat Rev Cancer. 2:48–58. 2002. View Article : Google Scholar : PubMed/NCBI

3 

Zhang D, Wu M, Nelson DE, Pasula R and Martin WJ II: Alpha-1-antitrypsin expression in the lung is increased by airway delivery of gene-transfected macrophages. Gene Ther. 10:2148–2152. 2003. View Article : Google Scholar : PubMed/NCBI

4 

Hansson M, Jönsson S, Persson AM, Calafat J, Tapper H and Olsson I: Targeting proteins to secretory lysosomes of natural killer cells as a principle for immunoregulation. Mol Immunol. 40:363–372. 2003. View Article : Google Scholar : PubMed/NCBI

5 

Cekmen M, Evereklioglu C, Er H, Inalöz HS, Doganay S, Türköz Y and Ozerol IH: Vascular endothelial growth factor levels are increased and associated with disease activity in patients with Behçet’s syndrome. Int J Dermatol. 42:870–875. 2003. View Article : Google Scholar : PubMed/NCBI

6 

Lu Y, Liu P, Wen W, Grubbs CJ, Townsend RR, Malone JP, Lubet RA and You M: Cross-species comparison of orthologous gene expression in human bladder cancer and carcinogen-induced rodent models. Am J Transl Res. 3:8–27. 2010.PubMed/NCBI

7 

Normandin K, Péant B, Le Page C, de Ladurantaye M, Ouellet V, Tonin PN, Provencher DM and Mes-Masson AM: Protease inhibitor SERPINA1 expression in epithelial ovarian cancer. Clin Exp Metastasis. 27:55–69. 2010. View Article : Google Scholar : PubMed/NCBI

8 

de la Fuente MT, Casanova B, Garcia-Gila M, Silva A and Garcia-Pardo A: Fibronectin interaction with alpha4beta1 integrin prevents apoptosis in B cell chronic lymphocytic leukemia: Correlation with Bcl-2 and Bax. Leukemia. 13:266–274. 1999. View Article : Google Scholar : PubMed/NCBI

9 

Schwarzbauer JE and DeSimone DW: Fibronectins, their fibrillogenesis, and in vivo functions. Cold Spring Harb Perspect Biol. 3:32011. View Article : Google Scholar

10 

Tapper J, Kettunen E, El-Rifai W, Seppälä M, Andersson LC and Knuutila S: Changes in gene expression during progression of ovarian carcinoma. Cancer Genet Cytogenet. 128:1–6. 2001. View Article : Google Scholar : PubMed/NCBI

11 

Mutlu P, Ural AU and Gündüz U: Differential gene expression analysis related to extracellular matrix components in drug-resistant RPMI-8226 cell line. Biomed Pharmacother. 66:228–231. 2012. View Article : Google Scholar : PubMed/NCBI

12 

Ahmed N, Riley C, Rice G and Quinn M: Role of integrin receptors for fibronectin, collagen and laminin in the regulation of ovarian carcinoma functions in response to a matrix microenvironment. Clin Exp Metastasis. 22:391–402. 2005. View Article : Google Scholar : PubMed/NCBI

13 

Akiyama SK, Olden K and Yamada KM: Fibronectin and integrins in invasion and metastasis. Cancer Metastasis Rev. 14:173–189. 1995. View Article : Google Scholar : PubMed/NCBI

14 

Shibata K, Kikkawa F, Nawa A, Suganuma N and Hamaguchi M: Fibronectin secretion from human peritoneal tissue induces Mr 92,000 type IV collagenase expression and invasion in ovarian cancer cell lines. Cancer Res. 57:5416–5420. 1997.PubMed/NCBI

15 

Lou X, Han X, Jin C, Tian W, Yu W, Ding D, Cheng L, Huang B, Jiang H and Lin B: SOX2 targets fibronectin 1 to promote cell migration and invasion in ovarian cancer: New molecular leads for therapeutic intervention. OMICS. 17:510–518. 2013. View Article : Google Scholar : PubMed/NCBI

16 

Jinawath N, Vasoontara C, Jinawath A, Fang X, Zhao K, Yap KL, Guo T, Lee CS, Wang W, Balgley BM, et al: Oncoproteomic analysis reveals co-upregulation of RELA and STAT5 in carboplatin resistant ovarian carcinoma. PLoS One. 5:e111982010. View Article : Google Scholar : PubMed/NCBI

17 

Qian P, Zuo Z, Wu Z, Meng X, Li G, Wu Z, Zhang W, Tan S, Pandey V, Yao Y, et al: Pivotal role of reduced let-7g expression in breast cancer invasion and metastasis. Cancer Res. 71:6463–6474. 2011. View Article : Google Scholar : PubMed/NCBI

18 

Waalkes S, Atschekzei F, Kramer MW, Hennenlotter J, Vetter G, Becker JU, Stenzl A, Merseburger AS, Schrader AJ, Kuczyk MA, et al: Fibronectin 1 mRNA expression correlates with advanced disease in renal cancer. BMC Cancer. 10:5032010. View Article : Google Scholar : PubMed/NCBI

19 

Yokomizo A, Takakura M, Kanai Y, et al: Use of quantitative shotgun proteomics to identify fibronectin 1 as a potential plasma biomarker for clear cell carcinoma of the kidney. Cancer biomarkers: Section. Dis Markers. 10:175–183. 2011.

20 

He X, Wang Y, Zhang W, Li H, Luo R, Zhou Y, Li C, Liao M, Huang H, Lv X, et al: Screening differential expression of serum proteins in AFP-negative HBV-related hepatocellular carcinoma using iTRAQ -MALDI-MS/MS. Neoplasma. 61:17–26. 2014. View Article : Google Scholar

21 

Cochrane DR, Spoelstra NS, Howe EN, Nordeen SK and Richer JK: MicroRNA-200c mitigates invasiveness and restores sensitivity to microtubule-targeting chemotherapeutic agents. Mol Cancer Ther. 8:1055–1066. 2009. View Article : Google Scholar : PubMed/NCBI

22 

Liu L, Zou J, Wang Q, Yin FQ, Zhang W and Li L: Novel microRNAs expression of patients with chemotherapy drug-resistant and chemotherapy-sensitive epithelial ovarian cancer. Tumour Biol. 35:7713–7717. 2014. View Article : Google Scholar : PubMed/NCBI

23 

Zhang X, Yang JJ, Kim YS, Kim KY, Ahn WS and Yang S: An 8-gene signature, including methylated and down-regulated glutathione peroxidase 3, of gastric cancer. Int J Oncol. 36:405–414. 2010.PubMed/NCBI

24 

Bottaro DP, Rubin JS, Faletto DL, Chan AM, Kmiecik TE, Vande Woude GF and Aaronson SA: Identification of the hepatocyte growth factor receptor as the c-met proto-oncogene product. Science. 251:802–804. 1991. View Article : Google Scholar : PubMed/NCBI

25 

Park M, Dean M, Kaul K, Braun MJ, Gonda MA and Vande Woude G: Sequence of MET protooncogene cDNA has features characteristic of the tyrosine kinase family of growth-factor receptors. Proc Natl Acad Sci USA. 84:6379–6383. 1987. View Article : Google Scholar : PubMed/NCBI

26 

Azad MB, Chen Y and Gibson SB: Regulation of autophagy by reactive oxygen species (ROS): Implications for cancer progression and treatment. Antioxid Redox Signal. 11:777–790. 2009. View Article : Google Scholar

27 

Davis W Jr, Ronai Z and Tew KD: Cellular thiols and reactive oxygen species in drug-induced apoptosis. J Pharmacol Exp Ther. 296:1–6. 2001.

28 

Lee HJ, Do JH, Bae S, Yang S, Zhang X, Lee A, Choi YJ, Park DC and Ahn WS: Immunohistochemical evidence for the over-expression of Glutathione peroxidase 3 in clear cell type ovarian adenocarcinoma. Med Oncol. 28(Suppl 1): S522–S527. 2011. View Article : Google Scholar

29 

Saga Y, Ohwada M, Suzuki M, Konno R, Kigawa J, Ueno S and Mano H: Glutathione peroxidase 3 is a candidate mechanism of anticancer drug resistance of ovarian clear cell adenocarcinoma. Oncol Rep. 20:1299–1303. 2008.PubMed/NCBI

30 

Hochepied T, Berger FG, Baumann H and Libert C: Alpha(1)-acid glycoprotein: An acute phase protein with inflammatory and immunomodulating properties. Cytokine Growth Factor Rev. 14:25–34. 2003. View Article : Google Scholar

31 

Fournier T, Medjoubi-N N and Porquet D: Alpha-1-acid glycoprotein. Biochim Biophys Acta. 1482:157–171. 2000. View Article : Google Scholar : PubMed/NCBI

32 

Fan C, Stendahl U, Stjernberg N and Beckman L: Association between orosomucoid types and cancer. Oncology. 52:498–500. 1995. View Article : Google Scholar : PubMed/NCBI

33 

Duché JC, Urien S, Simon N, Malaurie E, Monnet I and Barré J: Expression of the genetic variants of human alpha-1-acid glycoprotein in cancer. Clin Biochem. 33:197–202. 2000. View Article : Google Scholar : PubMed/NCBI

34 

Tilg H, Vannier E, Vachino G, Dinarello CA and Mier JW: Antiinflammatory properties of hepatic acute phase proteins: Preferential induction of interleukin 1 (IL-1) receptor antagonist over IL-1 beta synthesis by human peripheral blood mononuclear cells. J Exp Med. 178:1629–1636. 1993. View Article : Google Scholar : PubMed/NCBI

35 

Matsumoto K, Nishi K, Kikuchi M, Watanabe H, Nakajou K, Komori H, Kadowaki D, Suenaga A, Maruyama T and Otagiri M: Receptor-mediated uptake of human alpha1-acid glycoprotein into liver parenchymal cells in mice. Drug Metab Pharmacokinet. 25:101–107. 2010. View Article : Google Scholar : PubMed/NCBI

36 

Lee YS, Choi JW, Hwang I, Lee JW, Lee JH, Kim AY, Huh JY, Koh YJ, Koh GY, Son HJ, et al: Adipocytokine orosomucoid integrates inflammatory and metabolic signals to preserve energy homeostasis by resolving immoderate inflammation. J Biol Chem. 285:22174–22185. 2010. View Article : Google Scholar : PubMed/NCBI

37 

Gunnarsson P, Levander L, Påhlsson P and Grenegård M: The acute-phase protein alpha 1-acid glycoprotein (AGP) induces rises in cytosolic Ca2+ in neutrophil granulocytes via sialic acid binding immunoglobulin-like lectins (siglecs). FASEB J. 21:4059–4069. 2007. View Article : Google Scholar : PubMed/NCBI

38 

Atemezem A, Mbemba E, Vassy R, Slimani H, Saffar L and Gattegno L: Human alpha1-acid glycoprotein binds to CCR5 expressed on the plasma membrane of human primary macrophages. Biochem J. 356:121–128. 2001. View Article : Google Scholar : PubMed/NCBI

39 

Ligresti G, Aplin AC, Dunn BE, Morishita A and Nicosia RF: The acute phase reactant orosomucoid-1 is a bimodal regulator of angiogenesis with time- and context-dependent inhibitory and stimulatory properties. PLoS One. 7:e413872012. View Article : Google Scholar : PubMed/NCBI

40 

Irmak S, Oliveira-Ferrer L, Singer BB, Ergün S and Tilki D: Pro-angiogenic properties of orosomucoid (ORM). Exp Cell Res. 315:3201–3209. 2009. View Article : Google Scholar : PubMed/NCBI

41 

Goh BC, Lee SC, Wang LZ, Fan L, Guo JY, Lamba J, Schuetz E, Lim R, Lim HL, Ong AB, et al: Explaining interindividual variability of docetaxel pharmacokinetics and pharmacodynamics in Asians through phenotyping and genotyping strategies. J Clin Oncol. 20:3683–3690. 2002. View Article : Google Scholar : PubMed/NCBI

42 

Delbaldo C, Chatelut E, Ré M, Deroussent A, Séronie-Vivien S, Jambu A, Berthaud P, Le Cesne A, Blay JY and Vassal G: Pharmacokinetic-pharmacodynamic relationships of imatinib and its main metabolite in patients with advanced gastrointestinal stromal tumors. Clin Cancer Res. 12:6073–6078. 2006. View Article : Google Scholar : PubMed/NCBI

43 

Bruno R, Olivares R, Berille J, Chaikin P, Vivier N, Hammershaimb L, Rhodes GR and Rigas JR: Alpha-1-acid glycoprotein as an independent predictor for treatment effects and a prognostic factor of survival in patients with non-small cell lung cancer treated with docetaxel. Clin Cancer Res. 9:1077–1082. 2003.PubMed/NCBI

44 

Katori N, Sai K, Saito Y, Fukushima-Uesaka H, Kurose K, Yomota C, Kawanishi T, Nishimaki-Mogami T, Naito M, Sawada J, et al: Genetic variations of orosomucoid genes associated with serum alpha-1-acid glycoprotein level and the pharmacokinetics of paclitaxel in Japanese cancer patients. J Pharm Sci. 100:4546–4559. 2011. View Article : Google Scholar : PubMed/NCBI

45 

Pandey PR, Liu W, Xing F, Fukuda K and Watabe K: Anti-cancer drugs targeting fatty acid synthase (FAS). Recent Patents Anticancer Drug Discov. 7:185–197. 2012. View Article : Google Scholar

46 

Flavin R, Peluso S, Nguyen PL and Loda M: Fatty acid synthase as a potential therapeutic target in cancer. Future Oncol. 6:551–562. 2010. View Article : Google Scholar : PubMed/NCBI

47 

Chen Y, Ma Z, Li A, Li H, Wang B, Zhong J, Min L and Dai L: Metabolomic profiling of human serum in lung cancer patients using liquid chromatography/hybrid quadrupole time-of-flight mass spectrometry and gas chromatography/mass spectrometry. J Cancer Res Clin Oncol. 141:705–718. 2015. View Article : Google Scholar

48 

Liu H, Liu Y and Zhang JT: A new mechanism of drug resistance in breast cancer cells: Fatty acid synthase overexpression-mediated palmitate overproduction. Mol Cancer Ther. 7:263–270. 2008. View Article : Google Scholar : PubMed/NCBI

49 

Meena AS, Sharma A, Kumari R, Mohammad N, Singh SV and Bhat MK: Inherent and acquired resistance to paclitaxel in hepatocellular carcinoma: Molecular events involved. PLoS One. 8:e615242013. View Article : Google Scholar : PubMed/NCBI

50 

Roodhart JM, Daenen LG, Stigter EC, Prins HJ, Gerrits J, Houthuijzen JM, Gerritsen MG, Schipper HS, Backer MJ, van Amersfoort M, et al: Mesenchymal stem cells induce resistance to chemotherapy through the release of platinum-induced fatty acids. Cancer Cell. 20:370–383. 2011. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

October-2016
Volume 49 Issue 4

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
Wu W, Wang Q, Yin F, Yang Z, Zhang W, Gabra H and Li L: Identification of proteomic and metabolic signatures associated with chemoresistance of human epithelial ovarian cancer. Int J Oncol 49: 1651-1665, 2016
APA
Wu, W., Wang, Q., Yin, F., Yang, Z., Zhang, W., Gabra, H., & Li, L. (2016). Identification of proteomic and metabolic signatures associated with chemoresistance of human epithelial ovarian cancer. International Journal of Oncology, 49, 1651-1665. https://doi.org/10.3892/ijo.2016.3652
MLA
Wu, W., Wang, Q., Yin, F., Yang, Z., Zhang, W., Gabra, H., Li, L."Identification of proteomic and metabolic signatures associated with chemoresistance of human epithelial ovarian cancer". International Journal of Oncology 49.4 (2016): 1651-1665.
Chicago
Wu, W., Wang, Q., Yin, F., Yang, Z., Zhang, W., Gabra, H., Li, L."Identification of proteomic and metabolic signatures associated with chemoresistance of human epithelial ovarian cancer". International Journal of Oncology 49, no. 4 (2016): 1651-1665. https://doi.org/10.3892/ijo.2016.3652