Skip to main content
Erschienen in: Arthritis Research & Therapy 5/2012

Open Access 01.10.2012 | Research article

Secretome analysis of chondroitin sulfate-treated chondrocytes reveals anti-angiogenic, anti-inflammatory and anti-catabolic properties

verfasst von: Valentina Calamia, Lucía Lourido, Patricia Fernández-Puente, Jesús Mateos, Beatriz Rocha, Eulalia Montell, Josep Vergés, Cristina Ruiz-Romero, Francisco J Blanco

Erschienen in: Arthritis Research & Therapy | Ausgabe 5/2012

Abstract

Introduction

Chondroitin sulfate (CS) is a symptomatic slow-acting drug for osteoarthritis (OA) widely used in the clinic. The aim of this work is to find proteins whose secretion from cartilage cells under proinflammatory stimuli (IL-1β) is regulated by CS, employing a novel quantitative proteomic approach.

Methods

Human articular chondrocytes released from three normal cartilages were grown in SILAC medium. When complete incorporation of the heavy isotope was achieved, chondrocytes were stimulated with IL-1β 5 ng/ml with or without CS pretreatment (200 µg/ml). Forty-eight hours later, chondrocyte secretomes were analyzed by nano-scale liquid chromatography-mass spectrometry. Real-time PCR, western blot and immunohistochemistry analyses were employed to confirm some of the results.

Results

We could identify 75 different proteins in the secretome of human articular chondrocytes. Eighteen of these were modulated by CS with statistical significance (six increased and 12 decreased). In normal chondrocytes stimulated with IL-1β, CS reduces inflammation directly by decreasing the presence of several complement components (CFAB, C1S, CO3, and C1R) and also indirectly by increasing proteins such as TNFα-induced protein (TSG6). TSG6 overexpression correlates with a decrease in pro-matrix metalloproteinase activation (observed in MMP1 and MMP3 levels). Finally, we observed a strong CS-dependent increase of an angiogenesis inhibitor, thrombospondin-1.

Conclusion

We have generated a quantitative profile of chondrocyte extracellular protein changes driven by CS in the presence of IL-1β. We have also provided novel evidences of its anti-angiogenic, anti-inflammatory, and anti-catabolic properties. Demonstration of the anti-angiogenic action of CS might provide a novel therapeutic approach for OA targeting.
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1186/​ar4040) contains supplementary material, which is available to authorized users.

Competing interests

The authors declare that they have no competing interests.

Authors' contributions

VC carried out the experimental work, analyzed the data and drafted the manuscript. LL and BR helped to collect and process protein samples, participated in western blot experiments and helped with statistical data analysis. PF-P and JM carried out the MS analysis and database search. EM and JV provided CS and helped with the study design. CR-R participated in the study design, interpretation of the data and manuscript preparation. FJB conceived and coordinated the project and revised the manuscript. All authors read and approved the final manuscript.
Abkürzungen
CS
chondroitin sulfate
Ct
threshold cycle
DMEM
Dulbecco's modified Eagle's medium
ECM
extracellular matrix
GS
glucosamine sulfate
HAC
human articular chondrocyte
IL
interleukin
LC
liquid chromatography
mAb
monoclonal antibody
MALDI
matrix-assisted laser desorption/ionization
MS
mass spectrometry
OA
osteoarthritis
PCR
polymerase chain reaction
SILAC
stable isotope labeling with amino acids in cell culture
TOF
time of flight
TFA
Trifluoroacetic acid
TNF
tumor necrosis factor
TSG6
TNFα-induced protein
TSP1
thrombospondin-1.

Introduction

Osteoarthritis (OA) is one of the most prevalent chronic diseases affecting older people. Although its major feature is the progressive destruction of articular cartilage, it is now accepted that OA is a global disease of the joint, also involving the synovial membrane, subchondral bone and periarticular soft tissues [1]. Effective prevention of the structural damage must be a key objective of new therapeutic approaches to treat OA. However, drugs currently available are predominantly directed towards the symptomatic relief of pain and inflammation, doing little to reduce joint destruction [2].
Until now the pharmacological management of OA has been dominated by nonsteroidal anti-inflammatory drugs and analgesics (mainly paracetamol). However, the use of chondroitin sulfate (CS) by OA patients, alone or in combination with glucosamine sulfate (GS), has been rising globally over the last decade. Both molecules are well recognized as symptomatic slow-acting drugs for OA. Moreover, their application has an excellent safety profile, allowing long-term treatment [36]. Nevertheless, recent meta-analysis [7] and large-scale clinical trials [8] have demonstrated variable effects on OA symptoms, yielding conflicting results. For this reason, in 2010 we carried out the first pharmacoproteomic analysis of articular chondrocytes treated with exogenous CS and/or GS [9] with the aim of defining more clearly the effects of GS and CS on cartilage biology. In that work, we performed a classical proteomic approach by two-dimensional electrophoresis and mass spectrometry (MS) to describe the cellular proteome of normal human chondrocytes treated with both drugs, alone or in combination, in the presence of IL-1β, a proinflammatory cytokine that plays a pivotal role in the pathogenesis of OA [10]. A large number of target proteins of CS and GS were described, pointing out the wide range effects of these drugs on fundamental aspects of chondrocyte metabolism but also their alternative mechanisms of action in a system model of OA [9].
Once the utility of proteomics for analyzing the putative intracellular targets of CS and GS in cartilage cells was proved [9], we focused on the subset of chondrocyte extracellular proteins that are essential for cartilage extracellular matrix (ECM) synthesis and turnover processes. Furthermore, secreted proteins may end up in the bloodstream, and thereby may have potential use as non-invasive biomarkers [11]. For these reasons, the chondrocyte secretome has emerged as an attractive starting point for the discovery of new OA drug targets, for the monitoring of clinical trials or for the personalization and optimization of long-term therapies. We recently published the first quantitative study of the secretome of primary human articular chondrocytes (HACs) by chondrocyte metabolic labeling, using an in vitro model of inflammation by stimulation with IL-1β [12]. In the present work, we aimed to employ this model to generate a quantitative profile of chondrocyte extracellular protein changes driven by CS in the presence of the proinflammatory stimulus, which might provide novel molecular evidence for CS effects.

Materials and methods

Cartilage procurement and processing

Macroscopically normal human knee cartilage from three adult donors (70, 73 and 78 years old) with no history of joint disease was provided by the Tissue Bank and the Autopsy Service at CHU A Coruña for the proteomic analysis. The study was approved by the local ethics committee. Cartilage was processed as previously described [13].

Primary culture of chondrocytes

HACs were isolated as described previously [9, 13]. Briefly, cartilage surfaces were rinsed with saline buffer, and scalpels were used to cut parallel vertical sections 5 mm apart from the cartilage surface to the subchondral bone. These cartilage strips were dissected from the bone, and the tissue was incubated with trypsin at 37°C for 10 minutes and then digested with type IV clostridial collagenase. The release of chondrocytes from cartilage was achieved after 16 hours of digestion in an incubator at 37ºC, 5% carbon dioxide.

Chondrocyte metabolic labeling and differential treatment of SILAC cell populations

The isolated chondrocytes were recovered and plated at low density in SILAC DMEM-Flex (Invitrogen, Paisley, UK) deficient in arginine (R) and lysine (K) supplemented with 10% dialyzed fetal bovine serum (Gibco, Invitrogen), 4.5 g/l glucose (Sigma, St. Louis, MO, USA), 2 mM L-glutamine (Sigma), 100 units/ml penicillin and 100 µg/ml streptomycin. In the case of light media, standard L-lysine and L-arginine were used, while in the heavy media, isotope-labeled L-lysine (13C6) and isotope-labeled L-arginine (13C6,15N4) were used. For the initial cell expansion, 5×104 chondrocytes from each donor were seeded in two T-25 cell culture flasks (one grown in light medium and one in heavy medium). At confluence cells were recovered from each culture flask by trypsinization and seeded onto two six-multiwell plates (15×104 for well) for cell treatment. Chondrocytes were used at week 3 in primary culture, when 100% of labeling was reached. Verification of complete labeling was performed as previously described [12]. Briefly, a small aliquot of cells cultured in the heavy media were subjected to protein extraction. The extracts were then digested with trypsin and analyzed by nano-scale liquid chromatography (LC)-MS to determine the degree of incorporation by looking for the presence of light peptides. Verification of cell type was carried out by real-time PCR for the analysis of type II collagen mRNA expression under the conditions of study.
The chondroitin sulfate employed in this work is of bovine origin, with a CS content of 99.9% and a molecular weight of 15.12 kDa. Other characteristics (viscosity, sulfation sites, and so forth) have been previously detailed elsewhere [14]. Chondrocyte stimulation for the experiments was carried out following procedures previously described by our group, in which CS and IL-1β concentrations in the chondrocyte cultures were optimized for the proteomic studies [9, 12]. Briefly, cells were washed thoroughly to remove abundant serum proteins and were cultured in serum-free medium with or without chondroitin sulfate (200 μg/ml; Bioibérica, Barcelona, Spain). Two hours later, IL-1β was added to the culture media (5 ng/ml; Sigma). Finally, conditioned media were collected after 48 hours of culture. Cell viability was assessed by Trypan Blue dye exclusion.

Processing of conditioned media for analysis by LC-MS

Conditioned media obtained from three different donors were analyzed independently. In addition, the off-gel measurements were performed in duplicate to assess the technical reproducibility of the LC-MS set-up.
Conditioned media were collected, centrifuged and filtered using a 0.2 µm filter to ensure removal of any dead cells. Proteins in the individual medium were precipitated with 0.02% sodium deoxycholate for 10 minutes and then with 10% (v/v) trichloroacetic acid overnight at 4ºC. Precipitates were harvested by centrifugation at 13,000 rpm for 15 minutes at 4°C and then washed twice with ice-cold acetone. The protein pellets were dried in air and then resuspended in 6 M urea, 2 M thiourea and 25 mM ammonium bicarbonate. The protein content of the concentrated media was measured using the Bradford reagent from Sigma. Heavy and light samples were then mixed 1:1, and 4 μg of each mixed sample were in-solution reduced, alkylated and digested with trypsin. Digestion was performed overnight with 12.5 ng/l Sequencing Grade Modified Trypsin (Promega, Madison, WI, USA) at 37ºC. The mixtures were acidified with Trifluoroacetic acid (1% final concentration) to stop the enzymatic reaction. The resulted peptides were desalted and filtered through a C18 microcolumn (NuTip; Glygen, Columbia, MD, USA) and finally eluted from the C18 bed using 70% Acetonitrile/0.1% TFA. The organic component was removed by evaporating in a vacuum centrifuge and the peptides were resuspended in 2% Acetonitrile/0.1% TFA. Then 5 µl were injected into a reversed-phase column (Integrafit C18, Proteopep™ II; New Objective, Woburn, MA, USA) for nano-flow LC analysis, using a Tempo nanoLC (Eksigent, Dublin, CA, USA) equipped with a Sun Collect MALDI Spotter/Micro-Fraction Collector (SunChrom GmbH, Friedrichsdorf, Germany).

Nano-scale LC-MALDI-MS analysis

LC eluate was deposited onto an Opti-TOF LC MALDI target plate (1,534-spot format; ABSciex, Framingham, MA, USA) with a speed of one spot per 15 seconds. Before spotting, the LC microfractions were mixed with MALDI matrix (3 mg/ml α-cyano-4-hydroxycinnamic acid in 70% Acetonitrile and 0.1% TFA containing 10 fmol/μl angiotensin as internal standard). Peptide-containing LC spots were analyzed in a 4800 MALDI-TOF/TOF instrument (ABSciex) with a 200 Hz repetition rate (Nd:YAG laser). MS full-scan spectra were acquired from 800 to 4,000 m/z. A total of 1,500 laser shots were accumulated for each time-of-flight MS spectrum at an optimized fixed laser setting. Tandem MS mode was operated with 1 kV collision energy with CID gas (air) over a range of 60 to -20 m/z of the precursor mass value. The precursor mass window was 300 ppm (full width at half-maximum) in relative mode. A minimum of 800 and a maximum of 1,500 laser shots were accumulated with laser stop conditions set at 10 product ion peaks of signal-to-noise ratio >100 at an optimized, fixed laser setting with metastable suppressor option on. Data-dependent tandem MS settings included acquisition of up to 20 of the most intense ion signals per spot. If two or more consecutive spots in an LC run with precursor m/z were within 200 ppm tolerance, the spot with the maximum signal-to-noise ratio was subjected to tandem MS analysis.

Data analysis

Peptide and protein identification and comparative quantification were performed using the Protein Pilot software vs 3.0 (ABSciex) with Paragon Algorithm. MS/MS data was searched against the UniProt/Swiss-Prot database of protein sequences (August 2010; Swiss-Prot, Geneva, Switzerland), using the following parameters: sample type set as SILAC (Lis+6, Arg+10), cysteine alkylation with Iodoacetamide, urea denaturation, one missed cleavage allowed in trypsin digestion and focus in biological modifications. Only proteins with a threshold >95% confidence (>1.3 unused score) were considered for protein identification. Data were normalized for mixing error by bias corrections.

Real-time PCR assays

Total RNA was isolated from chondrocytes (5×105 per well) using Trizol Reagent (Invitrogen, Carlsbad, CA, USA), following the manufacturer's instructions. cDNA was synthesized from 1 μg total RNA, using the Transcriptor First Strand cDNA Synthesis Kit (Roche Applied Science, Mannheim, Germany) in accordance with the manufacturer's instructions, and was analyzed by quantitative real-time PCR. The quantitative real-time PCR assay was performed in the LightCycler 480 instrument (Roche Applied Science) using 96-well plates. Primers for thrombospondin-1 (TSP1), TNFα-induced protein (TSG6) and the housekeeping genes, HPRT1 and RPLP0, were designed using the Universal Probe Library tool from the Roche website [15]. Primer sequences were as follows: TSP1 forward, 5'-GCTGCACTGAGTGTCACTGTC-3'; TSP1 reverse, 5'-TCAGGAACTGTGGCATTGG-3'; TSG6 forward, 5'-GCTAGAGGCAGCCAGAAAAA-3'; TSG6 reverse, 5'-ATCCAACTCTGCCCTTAGCC-3'; HPRT1 forward, 5'-TGACCTTGATTTATTTTGCATACC-3'; HPRT1 reverse, 5'-CGAGCAAGACGTTCAGTCCT-3'; RPLP0 forward, 5'-TCTACAACCCTGAAGTGCTTGAT-3', PRPL0 reverse 5'-CAATCTGCAGACAGACACTGG-3'. The results were analyzed using the LightCycler 480 software release 1.5.0 (Roche Applied Science), which automatically recorded the threshold cycle (Ct). An untreated cell sample (basal) was used as the calibrator; the fold-change for this sample was 1.0. Target gene Ct values were normalized against HPRT1 and RPLP0. Data were analyzed using the 2-ΔΔCt method and expressed as the fold-change of the test sample compared with the basal condition [16].

Western blot analysis

Western blot analyses were performed utilizing standard procedures. Briefly, 20 μg secreted proteins and 50 μg intracellular proteins were loaded and resolved using 10% SDS-PAGE. The separated proteins were then transferred to polyvinylidene fluoride membranes (Immobilon P; Millipore Co., Bedford, MA, USA) by electroblotting and probed with specific antibodies against TSP1 (Santa Cruz Biotechnology, Santa Cruz, CA), TSG6 (Abnova, Taipei, Taiwan), MMP1 and MMP3 (Abcam, Cambridge, UK). Immunoreactive bands were detected and housekeeping control GAPDH (Sigma). Immunoreactive bands were detected by chemiluminescence using corresponding horseradish peroxidase-conjugated secondary antibodies and enhanced chemiluminescence detection reagents (GE Healthcare, Uppsala, Sweden), and then digitized using the LAS 3000 image analyzer (Fujifilm, Tokyo, Japan). For secretome samples, equivalent loadings were verified by Ponceau Red (Sigma) staining after transference (data not shown). Quantitative changes in band intensities were evaluated using ImageQuant 5.2 software (GE Healthcare).

Immunohistochemical analysis

Cartilage shavings from three healthy donors different from those selected for the proteomics strategy were cut into 6 mm discs using a sterile biopsy punch, and one disc/well was placed into 96-well plates containing 200 µl serum-free DMEM. Plates were incubated for 48 hours with 200 µg/ml CS in presence of IL-1β (5 ng/ml). Frozen samples were then cut at 4 μm with a cryostat (Leica Microsystems, Barcelona, Spain) for immunohistochemical analysis. Sections were incubated with primary antibody to detect the presence of TSP1 (1:50; Santa Cruz Biotechnology, Heidelberg, Germany). The peroxidase/DAB ChemMate™ DAKO EnVision™ detection kit (Dako, Barcelona, Spain) was used to determine antigen-antibody interactions. Negative staining controls were achieved by omitting the primary mAb. Samples were visualized using an optical microscope.

Statistical analysis

Each experiment was repeated at least three times. The statistical significance of the differences between mean values was determined using a two-tailed t-test, considering P ≤ 0.05 significant. In the proteomic analysis, normalization tools and the statistical package from Protein Pilot software were employed (ABSciex). We considered statistically significant only those changes with P ≤ 0.05 and a ratio ≥1.2 (or ≤0.83). Where appropriate, results are expressed as the mean ± standard error.

Results and discussion

Most CS exists as the sugar chains of aggrecan in the cartilage, and its high water-retaining capacity ensures proper cartilage hydration [17]. However, several data in the literature reveal that the mechanism of action of CS is not limited to the fact that it is part of the aggrecan; in vivo studies in animal models and in vitro studies with human and animal articular cells suggest that the effects of CS result from a combination of numerous factors [18]. We have performed a gel-free quantitative proteomics experiment for the secretome analysis (cell-conditioned media) of HACs treated with bovine CS (95% purity) in the presence of IL-1β. Although HAC supernatants lack the complexity of the intact cartilage ECM, chondrocyte secretome may represent an attractive subproteome for understanding the chondroprotective action of CS [19, 20].

Secretome profiling of IL-1β and CS-treated HACs

Given the key role of chondrocytes in ECM synthesis and turnover, and also the importance of these mechanisms for tissue maintenance (which are disturbed in OA and other joint diseases), we examined the effect of CS in the subset of proteins secreted by chondrocytes (secretome) in an inflammatory environment (IL-1β). Inflammatory molecules, such as proinflammatory cytokines, are critical mediators of the disturbed metabolism and enhance the catabolism of joint tissue involved in OA pathophysiology [21]. For this purpose, supernatants from IL-1β-stimulated chondrocytes, with or without CS treatment, were collected after 48 hours of incubation and were analyzed. Owing to the low complexity of the secretome samples, we carried out a monodimensional approach: we combined equal amounts of proteins from the experimental conditions to be compared (treated or untreated with CS, both in the presence of the cytokine), and then these samples were digested in solution with trypsin. The correspondent tryptic peptides were separated by LC and the peptides were subsequently eluted and subjected to mass spectrometry analysis (MALDI-MS/MS).
This procedure resulted in the identification of 75 proteins present in the culture media of IL-1β-treated cells with statistical confidence (73 with Protein Pilot score ≥2). Some of them had not been previously reported to be secreted by chondrocytes [12], but they were found in serum [22] and/or synovial fluid [23] of OA patients and thus possess putative biomarker value. A complete list of these proteins is shown in Table 1. The majority of the identified secreted proteins were cartilage ECM proteins, or proteins with well-established matrix functions. Furthermore, several mediators of the inflammatory response were detected. The molecular function of the identified proteins was categorized by GeneOntology and is shown in Figure 1. The most abundant proteins identified in the samples (in terms of Protein Pilot hits, see Table 1) included well-known cartilage-related proteins, namely fibronectin (FN1) and chitinase-3-like protein 1 (CHI3L1), as well as ECM degradative enzymes, such as stromelysin-1 (MMP3) and interstitial collagenase (MMP1).
Table 1
Proteins identified in the secretome of IL-1β-stimulated chondrocytes with or without CS treatment
Symbol
Score
Peptides (95%)
Covariance (%)
Accession numbera
Name
Function
OA serum [22]
OA synovial fluid [23]
MMP2
20.24
17
36.4
[P08253]
72 kDa type IV collagenase
Angiogenesis/collagen degradation
 
x
A1AT
2
2
6
[P01009]
Alpha-1-antitrypsin
Inhibitor of serine proteases
  
ANXA2
2
1
36
[P07355]
Annexin A2
Heat-stress response
  
B2MG
4
3
26.9
[P61769]
Beta-2-microglobulin
Immunity
x
 
PGS1
4.03
3
13.9
[P21810]
Biglycan
Collagen fiber assembly/ECM component
  
CATB
3.06
2
15.3
[P07858]
Cathepsin B
Thiol protease
  
CCL2
6
4
38.4
[P13500]
C-C motif chemokine 2b
Inflammatory response
  
CCL8
4
2
38.4
[P80075]
C-C motif chemokine 8b
Inflammatory response
  
CH3L1
63.26
96
82
[P36222]
Chitinase-3-like protein 1
ECM component
 
x
CH3L2
2
2
17.7
[Q15782]
Chitinase-3-like protein 2b
ECM component
  
CLUS
10.09
9
43.2
[P10909]
Clusterin
Immunity
x
x
CCD80
2
1
23.3
[Q76M96]
Coiled-coil domain-containing protein 80
Cell adhesion and matrix assembly
  
CO3A1
2.02
1
14.3
[P02461]
Collagen alpha-1(III) chain
ECM component
  
CO6A1
4.16
4
28.5
[P12109]
Collagen alpha-1(VI) chain
ECM component
  
COCA1
17.35
10
21.1
[Q99715]
Collagen alpha-1(XII) chain
ECM component
  
CO1A2
27.65
25
55.2
[P08123]
Collagen alpha-2(I) chain
ECM component
  
C1R
8.13
7
31.5
[P00736]
Complement C1r subcomponent
Immunity
x
x
C1S
10.6
9
27.3
[P09871]
Complement C1s subcomponent
Immunity
x
x
CO3
13.4
7
26.4
[P01024]
Complement C3
Immunity
x
 
CFAB
8.29
6
17.2
[P00751]
Complement factor B
Immunity
x
x
CXCL3
4.01
4
69.2
[P19876]
C-X-C motif chemokine 3b
inflammatory response
  
CXCL5
2
1
20.2
[P42830]
C-X-C motif chemokine 5b
Inflammatory response
  
CXCL6
4.13
5
38.6
[P80162]
C-X-C motif chemokine 6b
Inflammatory response
  
CYTC
4
3
67.1
[P01034]
Cystatin-C
Inhibitor of cysteine proteinases
x
 
PGS2
23.72
22
56
[P07585]
Decorin
ECM component
  
DESP
2
1
24.8
[P15924]
Desmoplakinb
Cell junction
x
 
FBLN3
15.7
12
31.2
[Q12805]
EGF-containing fibulin-like extracellular matrix protein 1
Negative regulator of chondrocyte differentiation
x
x
FINC
152.34
150
58.9
[P02751]
Fibronectin
ECM component
x
x
FBLN1
2
1
11.2
[P23142]
Fibulin-1
Cell adhesion/ECM organization
x
 
FSTL1
2.1
2
16.2
[Q12841]
Follistatin-related protein 1
Cell proliferation and differentiation
  
GDN
24.73
23
54.5
[P07093]
Glia-derived nexin
Serine protease inhibitor
  
GROA
4
5
54.2
[P09341]
Growth-regulated alpha protein
Inflammatory response
  
IBP3
12.02
12
57.7
[P17936]
Insulin-like growth factor-binding protein 3
Cell proliferation and differentiation
x
 
IBP4
3.36
3
37.6
[P22692]
Insulin-like growth factor-binding protein 4
Cell proliferation and differentiation
  
IBP5
4.03
2
55.5
[P24593]
Insulin-like growth factor-binding protein 5
Cell proliferation and differentiation
  
IBP6
3.85
3
43.8
[P24592]
Insulin-like growth factor-binding protein 6
Cell proliferation and differentiation
x
 
IBP7
2.15
3
29.1
[Q16270]
Insulin-like growth factor-binding protein 7
Cell proliferation and differentiation
  
IL6
16
16
42
[P05231]
IL-6
Inflammatory response
  
IL8
4.24
6
57.6
[P10145]
IL-8
Inflammatory response
  
MMP1
35.12
38
50.1
[P03956]
Interstitial collagenase
Collagen degradation
  
K1C10
60.82
36
66.8
[P13645]
Keratin, type I cytoskeletal 10
Intermediate filament
x
x
K1C14
11.36
7
44.9
[P02533]
Keratin, type I cytoskeletal 14b
Intermediate filament
  
K1C16
15.45
8
48
[P08779]
Keratin, type I cytoskeletal 16b
Intermediate filament
  
K1C9
30.98
19
60.2
[P35527]
Keratin, type I cytoskeletal 9
Intermediate filament
x
x
K2C1
70.89
41
69.6
[P04264]
Keratin, type II cytoskeletal 1
Intermediate filament
x
x
K22E
43.26
22
69.2
[P35908]
Keratin, type II cytoskeletal 2 epidermal
Intermediate filament
x
x
K2C6B
20.77
11
53.9
[P04259]
Keratin, type II cytoskeletal 6Bb
Intermediate filament
 
x
MFGM
6.01
4
20.9
[Q08431]
Lactadherin
Angiogenesis
  
LUM
28.15
30
60.7
[P51884]
Lumican
ECM component
x
x
CSF1
4.01
2
17.3
[P09603]
Macrophage colony-stimulating factor 1b
Inflammatory response
  
TIMP1
17.8
16
57.5
[P01033]
Metalloproteinase inhibitor 1
Metalloprotease inhibitor
 
x
TIMP2
4
3
32.7
[P16035]
Metalloproteinase inhibitor 2
Metalloprotease inhibitor
  
PTX3
2.02
2
14.4
[P26022]
Pentraxin-related protein PTX3
Inflammatory response
  
PLTP
5.8
4
31
[P55058]
Phospholipid transfer protein
Lipid transport
x
x
IC1
4
2
18.8
[P05155]
Plasma protease C1 inhibitor
Immunity
x
x
POTEF
2.03
1
27.6
[A5A3E0]
POTE ankyrin domain family member Fb
Unknown
  
SAP
2
1
5
[P07602]
Proactivator polypeptideb
Lipid metabolism
  
PCOC1
5.93
4
19.8
[Q15113]
Procollagen C-endopeptidase enhancer 1
Collagen metabolism
 
x
S10A8
1.54
1
35.5
[P05109]
Protein S100-A8b
Inflammatory response
x
 
S10A9
4
2
34.2
[P06702]
Protein S100-A9b
Inflammatory response
x
 
PRG4
3.7
3
21.9
[Q92954]
Proteoglycan 4b
ECM component
x
x
TRY6
4
2
15.4
[Q8NHM4]
Putative trypsin-6b
Serine protease
  
NPT2C
2
1
10.9
[Q8N130]
Sodium-dependent phosphate transport protein 2Cb
Ion transport
  
SPRC
2
2
7.3
[P09486]
SPARC
Cell proliferation and differentiation
x
 
MMP3
48.64
47
71.9
[P08254]
Stromelysin-1
Collagen degradation
 
x
QSOX1
4.87
4
16.3
[O00391]
Sulfhydryl oxidase 1
Cell redox homeostasis
x
 
SDC4
2
1
13.6
[P31431]
Syndecan-4
ECM component
  
TARSH
8.86
9
27.4
[Q7Z7G0]
Target of Nesh-SH3
Cell proliferation and differentiation
x
 
TENA
13.36
8
22.5
[P24821]
Tenascin
Cell adhesion/ECM organization
  
TETN
4
4
34.2
[P05452]
Tetranectin
Bone mineralization
x
x
TSP1
2.36
2
18.5
[P07996]
Thrombospondin-1
Angiogenesis
x
 
BGH3
2
1
22.6
[Q15582]
TGF-beta-induced protein ig-h3
TGF-beta signaling
x
 
TSG6
28.35
30
69.3
[P98066]
TNF-inducible gene 6 protein
Cell adhesion
  
VCAM1
2
1
12.2
[P19320]
Vascular cell adhesion protein 1
Cell adhesion
x
 
VASN
2
2
14.7
[Q6EMK4]
Vasorin
TGF-beta signaling
x
 
Proteins identified by SILAC and liquid chromatography-mass spectrometry analysis in the secretome of IL-1β-stimulated chondrocytes with or without chondroitin sulfate treatment. ECM, extracellular matrix; EGF, epidermal growth factor; TGF, transforming growth factor. aProtein accession number according to the SwissProt and TrEMBL databases. bProtein not identified in our previous analysis of chondrocyte secretome [12].

CS-mediated changes in the chondrocyte secretome

By these means we were able to relatively quantify all the identified proteins with statistical significance. To confirm our findings and exclude the possibility of any quantification differences arising from SILAC labeling [24], the whole experiment was replicated with treatment conditions crossed over (swapping the labeled state of the perturbed cells). Finally, among the identified proteins, 18 presented a significant alteration of their levels due to the pharmacological treatment (six increased and 12 decreased), which are listed in Table 2. We detected the modulation of proteins involved in several processes, such as cartilage ECM structural organization (three proteins, including noncollagenous proteins and proteoglycans), ECM remodeling (four proteins, including proteases and their inhibitors), immune response (six proteins) and angiogenesis (two proteins).
Table 2
Extracellular proteins modulated by chondroitin sulfate treatment in IL-1β-stimulated chondrocytes
Accession numbera
Name
Symbol
Ratiob
Pvalue
Error factorc
[P00751]
Complement factor B
CFAB
0.5566
0.0064
1.397
[Q08431]
Lactadherin
MFGM
0.597
0.0151
1.3855
[Q7Z7G0]
Target of Nesh-SH3
TARSH
0.7017
0.006
1.2209
[P09871]
Complement C1s subcomponent
C1S
0.7085
0.0007
1.1398
[P01024]
Complement C3
CO3
0.7241
0.0023
1.1999
[P02751]
Fibronectin
FINC
0.7321
0.0006
1.1891
[P03956]
Interstitial collagenase
MMP1
0.7113
0
1.0977
[P00736]
Complement C1r subcomponent
C1R
0.7734
0.0071
1.1755
[P08253]
72 kDa type IV collagenase
MMP2
0.7759
0.0001
1.0999
[P36222]
Chitinase-3-like protein 1
CH3L1
0.8088
0.0216
1.4096
[P10909]
Clusterin
CLUS
0.8224
0.05
1.2162
[P08254]
Stromelysin-1
MMP3
0.8256
0.0086
1.1378
[P07093]
Glia-derived nexin
GDN
1.2894
0.0031
1.1664
[O00391]
Sulfhydryl oxidase 1
QSOX1
1.3064
0.0382
1.2709
[Q92954]
Proteoglycan 4
PRG4
1.4393
0.05
1.445
[P98066]
TNF-inducible gene 6 protein
TSG6
1.7773
0.0007
1.3371
[P61769]
Beta-2-microglobulin
B2MG
2.1652
0.0125
1.4553
[P07996]
Thrombospondin-1
TSP1
89.1678
0
>2
Extracellular proteins identified by SILAC and liquid chromatography-mass spectrometry analysis as modulated by CS treatment in IL-1β-stimulated chondrocytes. aProtein accession number according to the SwissProt and TrEMBL databases. bAverage SILAC ratios (n = 3) that represent the relative protein abundance in CS-treated versus untreated cells, calculated by Protein Pilot 3.0 software (ABSciex). P ≤ 0.05 was accepted. cError factor of the quantification (measure of the error in the average quantification ratio as calculated by Protein Pilot 3.0 software).
Interestingly, we found distinctively in CS-treated cells a global decrease of immunity-related proteins, degradative enzymes (such as MMP1, MMP2, and MMP3), and some ECM structural proteins (such as fibronectin (FN1) and chitinase-3-like protein 1 (CHI3L1)). Among those proteins described in our previous work as increased by IL-1β [12], which were now decreased by CS, we found FN1 and CHI3L1, two components of normal cartilage matrix (Figure 2). Synthesis and release of both proteins and fragments is often increased in cartilage that is undergoing repair or remodeling, and they have been investigated as markers of cartilage damage in OA [2527]. Interestingly, the release of FN1 and CHI3L2 from chondrocytes was also detected in a previous proteomic analysis from our group, which aimed to evaluate the differential effect of three distinct CS molecules in chondrocytes [28]. In that work, the presence of these proteins in the chondrocyte secretomes was caused by treatment with a CS of porcine origin, which appeared to trigger catabolic effects in chondrocytes by increasing also the abundance of matrix metalloproteinases (MMP1 and MMP3). On the contrary, treatment with bovine CS (the one employed in the present study) did not have any effect on the release of these four proteins.

Putative mediators of CS anti-inflammatory and anti-catabolic effects

We also performed a database search, using STRING software, to visualize protein interactions on the set of CS-modulated proteins and further elucidate its effect on chondrocytes (Figure 3). The role of CS in counteracting the IL-1β-mediated increase of some proteins was also detected for three degradative enzymes and three members of the complement pathway (Figure 2 and Table 2). Recently, a central role for the inflammatory complement system in the pathogenesis of OA has been identified [29]. Expression and activation of complement is abnormally high in human osteoarthritic joints. We show in this study how CS could reduce inflammation directly by decreasing the presence of several complement components (CFAB, CLUS, CO3, C1S and C1R), and also indirectly by increasing proteins such as TSG6. This protein plays a crucial role in ECM formation, inflammatory cell migration and cell proliferation. TSG6 is also a key component of a negative feedback loop operating through the protease network that reduces matrix degradation during the OA process [30]. The mechanism driven by TSG6 leads to a decrease in pro-matrix metalloproteinase activation, which might protect cartilage from extensive degradation even in the presence of acute inflammation (represented in our case by a high level of IL-1β). Western blot analyses were performed to confirm the detected increase of TSG6 caused by CS treatment. As shown in Figure 4, CS increased the amount of TSG6 secreted by chondrocytes, and this increase correlates with a decline in MMP1 and MMP3 levels. These results point to the increase of TSG6 as a putative mediator of the reduction in pro-matrix metalloproteinase activation, suggesting an important role of this mechanism for the anti-catabolic effect of CS.

Modulation of thrombospondin-1 by CS

A remarkable increase of TSP1, an angiogenesis inhibitor, was detected as a consequence of the CS treatment and counteracting the effect of IL-1β. This result is consistent with our previously observed increase of TSP1 protein driven by CS in the absence of IL-1β stimulation, although in osteoarthritic chondrocytes [28]. TSP1 overexpression reduces inflammation and neovascularization in the OA joint [31]. In our previous study on IL-1β-stimulated chondrocytes, TSP1 presented a ratio of zero [12], indicating a cytokine-dependent dramatic decrease of its release from these cells. IL-1β is a well-recognized angiogenic factor, so the possibility that an increased concentration of IL-1β in OA synovial fluid may reduce the TSP1 expression in severe stages of OA cannot be excluded. The selective inhibition of angiogenesis - also confirmed by the decrease of lactadherin, a protein that promotes vascular endothelial growth factor-dependent neovascularization [32] - demonstrates a novel mechanism of action of CS according to recent results obtained in synoviocytes [33].
The data obtained in the SILAC analysis need to be validated for differences in protein expression profiles before the biological roles of the modulated proteins are extensively studied. We therefore performed additional studies in order to verify the altered abundance of TSP1 in CS-treated chondrocytes. Interestingly, TSP1 is a multifunctional adhesive glycoprotein present in articular cartilage and synthesized by articular chondrocytes [34], whose gene transfer suppresses the disease progression of experimental OA [31]. The inhibitory effect of TSP-1 on angiogenesis has been largely described [35]. Owing to the pivotal role of angiogenesis in OA physiopathology [36], we decided to verify TSP1 gene expression level in CS-treated chondrocytes stimulated with IL-1β by real time-PCR analysis, and also in cells without cytokine stimulation. As shown in Figure 5A, CS upregulates TSP1 already in the absence of IL-1β. When the cytokine is present, CS is capable of counteracting its suppressive effect on TSP1 in chondrocytes. Furthermore, TSP1 protein levels were also evaluated in chondrocyte conditioned media (secretome) and cellular extracts (proteome) by western blot analyses and in cartilage explant culture by immunohistochemistry (Figure 5). The increase of TSP1 protein observed both in cell and tissue cultures following CS treatment suggests the possible mechanism through which this drug could exert an anti-angiogenic action.

Conclusion

Our work provides a comprehensive quantitative analysis of the effects of CS in IL-1β-stimulated chondrocyte secretome, as well as novel molecular evidence for its anti-angiogenic, anti-inflammatory, and anti-catabolic properties. Proteins modulated by this drug are potential new targets for OA treatment (for example, TSP1). These findings might provide a rationale for targeting angiogenesis as a disease-modifying therapy for OA.

Acknowledgements

The authors express appreciation to the Pathology Service from the Orthopaedics Department of CHU A Coruña for providing cartilage samples, and to Purificación Filgueira and Noa Goyanes for their help in the histochemistry assays. This study was supported by grants from Fondo Investigación Sanitaria - Spain (CIBER-CB06/01/0040, PI08/2028, PI11/02397), Ministerio de Ciencia e Innovacion PLE2009-0144 and Secretaría I+D+I Xunta de Galicia (10CSA916058PR). BR (FI10/866), JM (CA11/00050) and PF-P (CA09/00458) are supported by Fondo Investigación Sanitaria - Spain. CR-R is supported by the Miguel Servet program from Fondo Investigación Sanitaria-Spain (CP09/00114).

Competing interests

The authors declare that they have no competing interests.

Authors' contributions

VC carried out the experimental work, analyzed the data and drafted the manuscript. LL and BR helped to collect and process protein samples, participated in western blot experiments and helped with statistical data analysis. PF-P and JM carried out the MS analysis and database search. EM and JV provided CS and helped with the study design. CR-R participated in the study design, interpretation of the data and manuscript preparation. FJB conceived and coordinated the project and revised the manuscript. All authors read and approved the final manuscript.
Literatur
1.
Zurück zum Zitat Martel-Pelletier J, Lajeunesse D, Pelletier J: Etiopathogenesis of osteoarthritis. Arthritis & Allied Conditions. A Textbook of Rheumatology. Edited by: Koopman WJ, Moreland LW. 2005, Baltimore, MD: Lippincott, Williams & Wilkins, 2199-2226. 15 Martel-Pelletier J, Lajeunesse D, Pelletier J: Etiopathogenesis of osteoarthritis. Arthritis & Allied Conditions. A Textbook of Rheumatology. Edited by: Koopman WJ, Moreland LW. 2005, Baltimore, MD: Lippincott, Williams & Wilkins, 2199-2226. 15
2.
Zurück zum Zitat Alcaraz MJ, Megías J, García-Arnandis I, Clérigues V, Guillén MI: New molecular targets for the treatment of osteoarthritis. Biochem Pharmacol. 2010, 80: 13-21. 10.1016/j.bcp.2010.02.017.CrossRefPubMed Alcaraz MJ, Megías J, García-Arnandis I, Clérigues V, Guillén MI: New molecular targets for the treatment of osteoarthritis. Biochem Pharmacol. 2010, 80: 13-21. 10.1016/j.bcp.2010.02.017.CrossRefPubMed
3.
Zurück zum Zitat Imada K, Oka H, Kawasaki D, Miura N, Sato T, Ito A: Anti-arthritic action mechanisms of natural chondroitin sulfate in human articular chondrocytes and synovial fibroblasts. Biol Pharm Bull. 2010, 33: 410-414. 10.1248/bpb.33.410.CrossRefPubMed Imada K, Oka H, Kawasaki D, Miura N, Sato T, Ito A: Anti-arthritic action mechanisms of natural chondroitin sulfate in human articular chondrocytes and synovial fibroblasts. Biol Pharm Bull. 2010, 33: 410-414. 10.1248/bpb.33.410.CrossRefPubMed
4.
Zurück zum Zitat Monfort J, Pelletier JP, Garcia-Giralt N, Martel-Pelletier J: Biochemical basis of the effect of chondroitin sulphate on osteoarthritis articular tissues. Ann Rheum Dis. 2008, 67: 735-740. 10.1136/ard.2006.068882.CrossRefPubMed Monfort J, Pelletier JP, Garcia-Giralt N, Martel-Pelletier J: Biochemical basis of the effect of chondroitin sulphate on osteoarthritis articular tissues. Ann Rheum Dis. 2008, 67: 735-740. 10.1136/ard.2006.068882.CrossRefPubMed
5.
Zurück zum Zitat Kahan A, Uebelhart D, De Vathaire F, Delmas PD, Reginster JY: Long-term effects of chondroitins 4 and 6 sulfate on knee osteoarthritis: the study on osteoarthritis progression prevention, a two-year, randomized, double-blind, placebo-controlled trial. Arthritis Rheum. 2009, 60: 524-533. 10.1002/art.24255.CrossRefPubMed Kahan A, Uebelhart D, De Vathaire F, Delmas PD, Reginster JY: Long-term effects of chondroitins 4 and 6 sulfate on knee osteoarthritis: the study on osteoarthritis progression prevention, a two-year, randomized, double-blind, placebo-controlled trial. Arthritis Rheum. 2009, 60: 524-533. 10.1002/art.24255.CrossRefPubMed
6.
Zurück zum Zitat Sawitzke AD, Shi H, Finco MF, Dunlop DD, Harris CL, Singer NG, Bradley JD, Silver D, Jackson CG, Lane NE, Oddis CV, Wolfe F, Lisse J, Furst DE, Bingham CO, Reda DJ, Moskowitz RW, Williams HJ, Clegg DO: Clinical efficacy and safety of glucosamine, chondroitin sulphate, their combination, celecoxib or placebo taken to treat osteoarthritis of the knee: 2-year results from GAIT. Ann Rheum Dis. 2010, 69: 1459-1464. 10.1136/ard.2009.120469.PubMedCentralCrossRefPubMed Sawitzke AD, Shi H, Finco MF, Dunlop DD, Harris CL, Singer NG, Bradley JD, Silver D, Jackson CG, Lane NE, Oddis CV, Wolfe F, Lisse J, Furst DE, Bingham CO, Reda DJ, Moskowitz RW, Williams HJ, Clegg DO: Clinical efficacy and safety of glucosamine, chondroitin sulphate, their combination, celecoxib or placebo taken to treat osteoarthritis of the knee: 2-year results from GAIT. Ann Rheum Dis. 2010, 69: 1459-1464. 10.1136/ard.2009.120469.PubMedCentralCrossRefPubMed
7.
Zurück zum Zitat McAlindon TE, LaValley MP, Gulin JP, Felson DT: Glucosamine and chondroitin for treatment of osteoarthritis: a systematic quality assessment and meta-analysis. JAMA. 2000, 283: 1469-1475. 10.1001/jama.283.11.1469.CrossRefPubMed McAlindon TE, LaValley MP, Gulin JP, Felson DT: Glucosamine and chondroitin for treatment of osteoarthritis: a systematic quality assessment and meta-analysis. JAMA. 2000, 283: 1469-1475. 10.1001/jama.283.11.1469.CrossRefPubMed
8.
Zurück zum Zitat Clegg DO, Reda DJ, Harris CL, Klein MA, O'Dell JR, Hooper MM, Bradley JD, Bingham CO, Weisman MH, Jackson CG, Lane NE, Cush JJ, Moreland LW, Schumacher HR, Oddis CV, Wolfe F, Molitor JA, Yocum DE, Schnitzer TJ, Furst DE, Sawitzke AD, Shi H, Brandt KD, Moskowitz RW, Williams HJ: Glucosamine, chondroitin sulfate, and the two in combination for painful knee osteoarthritis. N Engl J Med. 2006, 354: 795-808. 10.1056/NEJMoa052771.CrossRefPubMed Clegg DO, Reda DJ, Harris CL, Klein MA, O'Dell JR, Hooper MM, Bradley JD, Bingham CO, Weisman MH, Jackson CG, Lane NE, Cush JJ, Moreland LW, Schumacher HR, Oddis CV, Wolfe F, Molitor JA, Yocum DE, Schnitzer TJ, Furst DE, Sawitzke AD, Shi H, Brandt KD, Moskowitz RW, Williams HJ: Glucosamine, chondroitin sulfate, and the two in combination for painful knee osteoarthritis. N Engl J Med. 2006, 354: 795-808. 10.1056/NEJMoa052771.CrossRefPubMed
9.
Zurück zum Zitat Calamia V, Ruiz-Romero C, Rocha B, Fernandez-Puente P, Mateos J, Montell E, Verges J, Blanco FJ: Pharmacoproteomic study of the effects of chondroitin and glucosamine sulfate on human articular chondrocytes. Arthritis Res Ther. 2010, 12: R138-10.1186/ar3077.PubMedCentralCrossRefPubMed Calamia V, Ruiz-Romero C, Rocha B, Fernandez-Puente P, Mateos J, Montell E, Verges J, Blanco FJ: Pharmacoproteomic study of the effects of chondroitin and glucosamine sulfate on human articular chondrocytes. Arthritis Res Ther. 2010, 12: R138-10.1186/ar3077.PubMedCentralCrossRefPubMed
10.
Zurück zum Zitat Fernandes JC, Martel-Pelletier J, Pelletier JP: The role of cytokines in osteoarthritis pathophysiology. Biorheology. 2002, 39: 237-246.PubMed Fernandes JC, Martel-Pelletier J, Pelletier JP: The role of cytokines in osteoarthritis pathophysiology. Biorheology. 2002, 39: 237-246.PubMed
11.
Zurück zum Zitat Piersma SR, Fiedler U, Span S, Lingnau A, Pham TV, Hoffmann S, Kubbutat MH, Jimenez CR: Workflow comparison for label-free, quantitative secretome proteomics for cancer biomarker discovery: method evaluation, differential analysis, and verification in serum. J Proteome Res. 2010, 9: 1913-1922. 10.1021/pr901072h.CrossRefPubMed Piersma SR, Fiedler U, Span S, Lingnau A, Pham TV, Hoffmann S, Kubbutat MH, Jimenez CR: Workflow comparison for label-free, quantitative secretome proteomics for cancer biomarker discovery: method evaluation, differential analysis, and verification in serum. J Proteome Res. 2010, 9: 1913-1922. 10.1021/pr901072h.CrossRefPubMed
12.
Zurück zum Zitat Calamia V, Rocha B, Mateos J, Fernández-Puente P, Ruiz-Romero C, Blanco FJ: Metabolic labeling of chondrocytes for the quantitative analysis of the interleukin-1-beta-mediated modulation of their intracellular and extracellular proteomes. J Proteome Res. 2011, 10: 3701-3711. 10.1021/pr200331k.CrossRefPubMed Calamia V, Rocha B, Mateos J, Fernández-Puente P, Ruiz-Romero C, Blanco FJ: Metabolic labeling of chondrocytes for the quantitative analysis of the interleukin-1-beta-mediated modulation of their intracellular and extracellular proteomes. J Proteome Res. 2011, 10: 3701-3711. 10.1021/pr200331k.CrossRefPubMed
13.
Zurück zum Zitat Ruiz-Romero C, López-Armada MJ, Blanco FJ: Proteomic characterization of human normal articular chondrocytes: a novel tool for the study of osteoarthritis and other rheumatic diseases. Proteomics. 2005, 5: 3048-3059. 10.1002/pmic.200402106.CrossRefPubMed Ruiz-Romero C, López-Armada MJ, Blanco FJ: Proteomic characterization of human normal articular chondrocytes: a novel tool for the study of osteoarthritis and other rheumatic diseases. Proteomics. 2005, 5: 3048-3059. 10.1002/pmic.200402106.CrossRefPubMed
14.
Zurück zum Zitat Tat SK, Pelletier JP, Mineau F, Duval N, Martel-Pelletier J: Variable effects of 3 different chondroitin sulfate compounds on human osteoarthritic cartilage/chondrocytes: relevance of purity and production process. J Rheumatol. 2010, 37: 656-664. 10.3899/jrheum.090696.CrossRefPubMed Tat SK, Pelletier JP, Mineau F, Duval N, Martel-Pelletier J: Variable effects of 3 different chondroitin sulfate compounds on human osteoarthritic cartilage/chondrocytes: relevance of purity and production process. J Rheumatol. 2010, 37: 656-664. 10.3899/jrheum.090696.CrossRefPubMed
16.
Zurück zum Zitat Livak KJ, Schmittgen TD: Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) method. Methods. 2001, 25: 402-408. 10.1006/meth.2001.1262.CrossRefPubMed Livak KJ, Schmittgen TD: Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) method. Methods. 2001, 25: 402-408. 10.1006/meth.2001.1262.CrossRefPubMed
17.
Zurück zum Zitat Dudhia J: Aggrecan, aging and assembly in articular cartilage. Cell Mol Life Sci. 2005, 62: 2241-2256. 10.1007/s00018-005-5217-x.CrossRefPubMed Dudhia J: Aggrecan, aging and assembly in articular cartilage. Cell Mol Life Sci. 2005, 62: 2241-2256. 10.1007/s00018-005-5217-x.CrossRefPubMed
18.
Zurück zum Zitat Martel-Pelletier J, Kwan Tat S, Pelletier JP: Effects of chondroitin sulfate in the pathophysiology of the osteoarthritic joint: a narrative review. Osteoarthritis Cartilage. 2010, 18: S7-S11.CrossRefPubMed Martel-Pelletier J, Kwan Tat S, Pelletier JP: Effects of chondroitin sulfate in the pathophysiology of the osteoarthritic joint: a narrative review. Osteoarthritis Cartilage. 2010, 18: S7-S11.CrossRefPubMed
19.
Zurück zum Zitat Polacek M, Bruun JA, Johansen O, Martinez I: Differences in the secretome of cartilage explants and cultured chondrocytes unveiled by SILAC technology. J Orthop Res. 2010, 28: 1040-1049.PubMed Polacek M, Bruun JA, Johansen O, Martinez I: Differences in the secretome of cartilage explants and cultured chondrocytes unveiled by SILAC technology. J Orthop Res. 2010, 28: 1040-1049.PubMed
20.
Zurück zum Zitat Clutterbuck AL, Smith JR, Allaway D, Harris P, Liddell S, Mobasheri A: High throughput proteomic analysis of the secretome in an explant model of articular cartilage inflammation. J Proteomics. 2011, 74: 704-715. 10.1016/j.jprot.2011.02.017.PubMedCentralCrossRefPubMed Clutterbuck AL, Smith JR, Allaway D, Harris P, Liddell S, Mobasheri A: High throughput proteomic analysis of the secretome in an explant model of articular cartilage inflammation. J Proteomics. 2011, 74: 704-715. 10.1016/j.jprot.2011.02.017.PubMedCentralCrossRefPubMed
21.
Zurück zum Zitat Kapoor M, Martel-Pelletier J, Lajeunesse D, Pelletier JP, Fahmi H: Role of proinflammatory cytokines in the pathophysiology of osteoarthritis. Nat Rev Rheumatol. 2011, 7: 33-42. 10.1038/nrrheum.2010.196.CrossRefPubMed Kapoor M, Martel-Pelletier J, Lajeunesse D, Pelletier JP, Fahmi H: Role of proinflammatory cytokines in the pathophysiology of osteoarthritis. Nat Rev Rheumatol. 2011, 7: 33-42. 10.1038/nrrheum.2010.196.CrossRefPubMed
22.
Zurück zum Zitat Fernández-Puente P, Mateos J, Fernández-Costa C, Oreiro N, Fernández-López C, Ruiz-Romero C, Blanco FJ: Identification of a panel of novel serum osteoarthritis biomarkers. J Proteome Res. 2011, 10: 5095-5101. 10.1021/pr200695p.CrossRefPubMed Fernández-Puente P, Mateos J, Fernández-Costa C, Oreiro N, Fernández-López C, Ruiz-Romero C, Blanco FJ: Identification of a panel of novel serum osteoarthritis biomarkers. J Proteome Res. 2011, 10: 5095-5101. 10.1021/pr200695p.CrossRefPubMed
23.
Zurück zum Zitat Mateos J, Lourido L, Fernández-Puente P, Calamia V, Fernández-López C, Oreiro N, Ruiz-Romero C, Blanco FJ: Differential protein profiling of synovial fluid from rheumatoid arthritis and osteoarthritis patients using LC-MALDI TOF/TOF. J Proteomics. 2012, 75: 2869-2878. 10.1016/j.jprot.2011.12.042.CrossRefPubMed Mateos J, Lourido L, Fernández-Puente P, Calamia V, Fernández-López C, Oreiro N, Ruiz-Romero C, Blanco FJ: Differential protein profiling of synovial fluid from rheumatoid arthritis and osteoarthritis patients using LC-MALDI TOF/TOF. J Proteomics. 2012, 75: 2869-2878. 10.1016/j.jprot.2011.12.042.CrossRefPubMed
24.
Zurück zum Zitat Ong SE, Mann M: A practical recipe for stable isotope labeling by amino acids in cell culture (SILAC). Nat Protoc. 2006, 1: 2650-2660.CrossRefPubMed Ong SE, Mann M: A practical recipe for stable isotope labeling by amino acids in cell culture (SILAC). Nat Protoc. 2006, 1: 2650-2660.CrossRefPubMed
25.
Zurück zum Zitat Zivanović S, Rackov LP, Vojvodić D, Vucetić D: Human cartilage glycoprotein 39 - biomarker of joint damage in knee osteoarthritis. Int Orthop. 2009, 33: 1165-1170. 10.1007/s00264-009-0747-8.PubMedCentralCrossRefPubMed Zivanović S, Rackov LP, Vojvodić D, Vucetić D: Human cartilage glycoprotein 39 - biomarker of joint damage in knee osteoarthritis. Int Orthop. 2009, 33: 1165-1170. 10.1007/s00264-009-0747-8.PubMedCentralCrossRefPubMed
26.
Zurück zum Zitat Homandberg GA, Wen C, Hui F: Cartilage damaging activities of fibronectin fragments derived from cartilage and synovial fluid. Osteoarthritis Cartilage. 1998, 6: 231-244. 10.1053/joca.1998.0116.CrossRefPubMed Homandberg GA, Wen C, Hui F: Cartilage damaging activities of fibronectin fragments derived from cartilage and synovial fluid. Osteoarthritis Cartilage. 1998, 6: 231-244. 10.1053/joca.1998.0116.CrossRefPubMed
27.
Zurück zum Zitat Zack MD, Arner EC, Anglin CP, Alston JT, Malfait AM, Tortorella MD: Identification of fibronectin neoepitopes present in human osteoarthritic cartilage. Arthritis Rheum. 2006, 54: 2912-2922. 10.1002/art.22045.CrossRefPubMed Zack MD, Arner EC, Anglin CP, Alston JT, Malfait AM, Tortorella MD: Identification of fibronectin neoepitopes present in human osteoarthritic cartilage. Arthritis Rheum. 2006, 54: 2912-2922. 10.1002/art.22045.CrossRefPubMed
28.
Zurück zum Zitat Calamia V, Fernandez-Puente P, Mateos J, Lourido L, Rocha B, Montell E, Verges J, Ruiz-Romero C, Blanco FJ: Pharmacoproteomic study of three different chondroitin sulfate compounds on intracellular and extracellular human chondrocyte proteomes. Mol Cell Proteomics. 2012, 11: M111.013417-10.1074/mcp.M111.013417.PubMedCentralCrossRefPubMed Calamia V, Fernandez-Puente P, Mateos J, Lourido L, Rocha B, Montell E, Verges J, Ruiz-Romero C, Blanco FJ: Pharmacoproteomic study of three different chondroitin sulfate compounds on intracellular and extracellular human chondrocyte proteomes. Mol Cell Proteomics. 2012, 11: M111.013417-10.1074/mcp.M111.013417.PubMedCentralCrossRefPubMed
29.
Zurück zum Zitat Wang Q, Rozelle AL, Lepus CM, Scanzello CR, Song JJ, Larsen DM, Crish JF, Bebek G, Ritter SY, Lindstrom TM, Hwang I, Wong HH, Punzi L, Encarnacion A, Shamloo M, Goodman SB, Wyss-Coray T, Goldring SR, Banda NK, Thurman JM, Gobezie R, Crow MK, Holers VM, Lee DM, Robinson WH: Identification of a central role for complement in osteoarthritis. Nat Med. 2011, 17: 1674-1679. 10.1038/nm.2543.PubMedCentralCrossRefPubMed Wang Q, Rozelle AL, Lepus CM, Scanzello CR, Song JJ, Larsen DM, Crish JF, Bebek G, Ritter SY, Lindstrom TM, Hwang I, Wong HH, Punzi L, Encarnacion A, Shamloo M, Goodman SB, Wyss-Coray T, Goldring SR, Banda NK, Thurman JM, Gobezie R, Crow MK, Holers VM, Lee DM, Robinson WH: Identification of a central role for complement in osteoarthritis. Nat Med. 2011, 17: 1674-1679. 10.1038/nm.2543.PubMedCentralCrossRefPubMed
30.
Zurück zum Zitat Wisniewski HG, Hua JC, Poppers DM, Naime D, Vilcek J, Cronstein BN: TNF/IL-1-inducible protein TSG-6 potentiates plasmin inhibition by inter-alpha-inhibitor and exerts a strong anti-inflammatory effect in vivo. J Immunol. 1996, 156: 1609-1615.PubMed Wisniewski HG, Hua JC, Poppers DM, Naime D, Vilcek J, Cronstein BN: TNF/IL-1-inducible protein TSG-6 potentiates plasmin inhibition by inter-alpha-inhibitor and exerts a strong anti-inflammatory effect in vivo. J Immunol. 1996, 156: 1609-1615.PubMed
31.
Zurück zum Zitat Hsieh JL, Shen PC, Shiau AL, Jou IM, Lee CH, Wang CR, Teo ML, Wu CL: Intraarticular gene transfer of thrombospondin-1 suppresses the disease progression of experimental osteoarthritis. J Orthop Res. 2010, 28: 1300-1306. 10.1002/jor.21134.CrossRefPubMed Hsieh JL, Shen PC, Shiau AL, Jou IM, Lee CH, Wang CR, Teo ML, Wu CL: Intraarticular gene transfer of thrombospondin-1 suppresses the disease progression of experimental osteoarthritis. J Orthop Res. 2010, 28: 1300-1306. 10.1002/jor.21134.CrossRefPubMed
32.
Zurück zum Zitat Silvestre JS, Théry C, Hamard G, Boddaert J, Aguilar B, Delcayre A, Houbron C, Tamarat R, Blanc-Brude O, Heeneman S, Clergue M, Duriez M, Merval R, Lévy B, Tedgui A, Amigorena S, Mallat Z: Lactadherin promotes VEGF-dependent neovascularization. Nat Med. 2005, 11: 499-506. 10.1038/nm1233.CrossRefPubMed Silvestre JS, Théry C, Hamard G, Boddaert J, Aguilar B, Delcayre A, Houbron C, Tamarat R, Blanc-Brude O, Heeneman S, Clergue M, Duriez M, Merval R, Lévy B, Tedgui A, Amigorena S, Mallat Z: Lactadherin promotes VEGF-dependent neovascularization. Nat Med. 2005, 11: 499-506. 10.1038/nm1233.CrossRefPubMed
33.
Zurück zum Zitat Lambert C, Mathy-Hartert M, Dubuc JE, Montell E, Vergés J, Munaut C, Noël A, Henrotin Y: Characterization of synovial angiogenesis in osteoarthritis patients and its modulation by chondroitin sulfate. Arthritis Res Ther. 2012, 14: R58-10.1186/ar3771.PubMedCentralCrossRefPubMed Lambert C, Mathy-Hartert M, Dubuc JE, Montell E, Vergés J, Munaut C, Noël A, Henrotin Y: Characterization of synovial angiogenesis in osteoarthritis patients and its modulation by chondroitin sulfate. Arthritis Res Ther. 2012, 14: R58-10.1186/ar3771.PubMedCentralCrossRefPubMed
34.
Zurück zum Zitat Miller RR, McDevitt CA: Thrombospondin is present in articular cartilage and is synthesized by articular chondrocytes. Biochem Biophys Res Commun. 1988, 153: 708-714. 10.1016/S0006-291X(88)81152-5.CrossRefPubMed Miller RR, McDevitt CA: Thrombospondin is present in articular cartilage and is synthesized by articular chondrocytes. Biochem Biophys Res Commun. 1988, 153: 708-714. 10.1016/S0006-291X(88)81152-5.CrossRefPubMed
35.
36.
Zurück zum Zitat Haywood L, McWilliams DF, Pearson CI, Gill SE, Ganesan A, Wilson D, Walsh DA: Inflammation and angiogenesis in osteoarthritis. Arthritis Rheum. 2003, 48: 2173-2177. 10.1002/art.11094.CrossRefPubMed Haywood L, McWilliams DF, Pearson CI, Gill SE, Ganesan A, Wilson D, Walsh DA: Inflammation and angiogenesis in osteoarthritis. Arthritis Rheum. 2003, 48: 2173-2177. 10.1002/art.11094.CrossRefPubMed
Metadaten
Titel
Secretome analysis of chondroitin sulfate-treated chondrocytes reveals anti-angiogenic, anti-inflammatory and anti-catabolic properties
verfasst von
Valentina Calamia
Lucía Lourido
Patricia Fernández-Puente
Jesús Mateos
Beatriz Rocha
Eulalia Montell
Josep Vergés
Cristina Ruiz-Romero
Francisco J Blanco
Publikationsdatum
01.10.2012
Verlag
BioMed Central
Erschienen in
Arthritis Research & Therapy / Ausgabe 5/2012
Elektronische ISSN: 1478-6362
DOI
https://doi.org/10.1186/ar4040

Weitere Artikel der Ausgabe 5/2012

Arthritis Research & Therapy 5/2012 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Triglyzeridsenker schützt nicht nur Hochrisikopatienten

10.05.2024 Hypercholesterinämie Nachrichten

Patienten mit Arteriosklerose-bedingten kardiovaskulären Erkrankungen, die trotz Statineinnahme zu hohe Triglyzeridspiegel haben, profitieren von einer Behandlung mit Icosapent-Ethyl, und zwar unabhängig vom individuellen Risikoprofil.

Gibt es eine Wende bei den bioresorbierbaren Gefäßstützen?

In den USA ist erstmals eine bioresorbierbare Gefäßstütze – auch Scaffold genannt – zur Rekanalisation infrapoplitealer Arterien bei schwerer PAVK zugelassen worden. Das markiert einen Wendepunkt in der Geschichte dieser speziellen Gefäßstützen.

Vorsicht, erhöhte Blutungsgefahr nach PCI!

10.05.2024 Koronare Herzerkrankung Nachrichten

Nach PCI besteht ein erhöhtes Blutungsrisiko, wenn die Behandelten eine verminderte linksventrikuläre Ejektionsfraktion aufweisen. Das Risiko ist umso höher, je stärker die Pumpfunktion eingeschränkt ist.

Wie managen Sie die schmerzhafte diabetische Polyneuropathie?

10.05.2024 DDG-Jahrestagung 2024 Kongressbericht

Mit Capsaicin-Pflastern steht eine neue innovative Therapie bei schmerzhafter diabetischer Polyneuropathie zur Verfügung. Bei therapierefraktären Schmerzen stellt die Hochfrequenz-Rückenmarkstimulation eine adäquate Option dar.

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.