Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Materials Science & Engineering C 118 (2021) 111312

Contents lists available at ScienceDirect

Materials Science & Engineering C


journal homepage: www.elsevier.com/locate/msec

Chondroitin sulfate modified 3D porous electrospun nanofiber scaffolds T


promote cartilage regeneration
⁎ ⁎
Shuai Chena,1, Weiming Chenb,1, Yini Chenc,1, Xiumei Mod, , Cunyi Fana,
a
Department of Orthopaedics, Shanghai Jiao Tong University Affiliated Sixth People's Hospital, 200233 Shanghai, PR China
b
Shanghai Ninth People's Hospital, Shanghai Jiao Tong University School of Medicine, 200000 Shanghai, PR China
c
Department of Ultrasound in Medicine, Shanghai Jiao Tong University Affiliated Sixth People's Hospital, 200233 Shanghai, PR China
d
College of Chemistry, Chemical Engineering and Biotechnology, Donghua University, 201620 Shanghai, PR China

A R T I C LE I N FO A B S T R A C T

Keywords: 3D electrospun nanofibrous scaffolds have been developed for cartilage regeneration, however, there is no
3D scaffolds consensus on the preferable method for biocompatible scaffolds that enhance regeneration and attenuate in-
Nanofibrous scaffolds flammation. We designed a 3D porous electrospun polylactic acid (PLA) @gelatin-based scaffold by a novel
Cartilage regeneration method. Chondroitin sulfate (CS), commonly used in clinical cartilage treatment, is capable of regulating car-
Anti-inflammation
tilage formation and inhibiting inflammation. Thus we further functionalized the 3D scaffold by crosslinking of
Chondroitin sulfate
CS, assuming that CS-functionalized scaffold (CSS) would promote cartilage regeneration and modulate in-
flammation. We confirmed that CSS exhibits not only appropriate reversible compressibility and mechanical
property, but also appropriate biocompatibility, allowing cell proliferation. In vitro, the potential of CSS for
chondrogenic differentiation was improved compared to control and PLA@gelatin scaffold as chondrogenic
markers Collagen2 and Aggrecan was significantly increased. Meanwhile, significant reduction in two crucial
inflammatory factors (NO and PGE2) in CSS group demonstrated inflammation inhibition. In vivo, rabbit car-
tilage defects were created and CSS effectively promoted cartilage repair. Additionally, superior anti-in-
flammation effect of CSS was demonstrated by reduction in iNOS and PGES, enzymes producing NO and PGE2,
respectively by immunohistology. Our results indicated the preferable property of CSS for cartilage regeneration
and its potential in immunoregulation.

1. Introduction engineering by multimodal electrospinning [13] Hsu et al. [14] and


Wang et al. [15] used 3D printing technology to fabricate scaffolds for
Damaged cartilage has very limited self-repair capacity and usually cartilage regeneration. Yu et al. [16] generated 3D braid scaffolds for
progresses to destruction of the surrounding cartilage. Much effort has cartilage regeneration. Although various methods have been reported
been put into the development of a better therapeutic strategy. Various to construct 3D electrospun nanofibrous scaffolds [12,13,17–19],
electrospun nanofibers have been developed for cartilage regeneration shortcomings still exist [17]. These methods are either time-consuming
due to their outstanding performances, including high porosity and or demanding a specific device, complicated technique or strict condi-
surface-to-volume ratio [1–8]. However, as two-dimensional (2D) bio- tion. Thus, developing a suitable method for 3D nanofibrous scaffold
materials, electrospun nanofibers are inappropriate for cartilage re- fabrication, while challenging, is imperative. Recently, a novel, con-
generation because of poor mechanical property, limited thickness and venient strategy to construct 3D electrospun scaffold based on elec-
pore size. Three-dimensional (3D) electrospun nanofibers scaffolds trospinning and freeze-drying procedures was reported [20–23]. The
fabricated by a desirable method may provide a favorable micro- resultant 3D scaffold presents desirable reversible compressibility and
environment for cell adhesion and growth [9–11], and overcome the mechanics, which are important for tissue regeneration [20–23]. Al-
above obstacles. though this novel method provides superelasticity and multi-
Parker et al. constructed 3D nanofiber structures by rotary jet- functionality, scaffolds fabricated from inferior biocompatible materials
spinning [12]. Traversa et al. fabricated 3D scaffolds for soft tissue are inappropriate for tissue regeneration [21,22].


Corresponding authors.
E-mail addresses: mxmdhdx@126.com (X. Mo), cyfan@sjtu.edu.cn (C. Fan).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.msec.2020.111312
Received 6 January 2020; Received in revised form 6 July 2020; Accepted 14 July 2020
Available online 04 August 2020
0928-4931/ © 2020 Elsevier B.V. All rights reserved.
S. Chen, et al. Materials Science & Engineering C 118 (2021) 111312

The modification of synthetic scaffolds may enhance its capacity to (Fig. 1). Gelatin and PLA were all dissolved in 1,1,1,3,3,3-hexafluoro-2-
interact with cells [24]. Various factors and natural processes, such as propanol (HFIP). The concentration of gelatin was 12% (w/v), while
cell–extracellular matrix (ECM) interactions, ECM dynamics and cell–- that of PLA was 8%. The core solution (PLA) was injected at a flow rate
cell communication and positioning, and mechanical stimulation, pro- of 1 mL/h and the shell solution (gelatin) was injected at a rate of 5 mL/
vide a relatively stable microenvironment and direct cell activity [25]. h. The applied voltage was 13 kV. The obtained PLA@gelatin nanofiber
Among these, ECM not only guides the tissue development but also membranes were placed in vacuum for storage.
drives cell fate [25]. Chondroitin sulfate (CS), a member of glycosa-
minoglycan (GAG) family, is a major type of ECM primarily found in 2.3.2. Preparation of 3D scaffolds
cartilage, bone, and skin [26,27]. CS is capable of regulating cartilage First, heat-treated scaffold was prepared in the following three steps
formation [28]. Elisseeff et al. [29] confirmed the chondrogenic func- (Fig. 1): the electrospun nanofiber membranes were cut into pieces. The
tion of CS-based bioactive hydrogels. Furthermore, CS presents anti- pieces were dispersed in tert-butanol using a homogenizer. The revol-
inflammatory properties, as demonstrated by its clinical value in the ving speed was 8000–10,000 rpm for 10–30 min. Subsequently, the
treatment of cartilage degenerative disease [30,31]. Tabata et al. [32] dispersed nanofibers were poured into a cylindrical mold and frozen at
also reported the anti-inflammatory effects of CS. Therefore, CS is ef- −20 °C for 30 min. The materials were freeze-dried for 24 h, followed
fective in cartilage regeneration and inflammatory reduction [33,34]. by heating at 180 °C for 2 h in air. Different procedures were used to
In this study, we fabricated a 3D porous electrospun polylactic acid prepare different scaffolds. For SS, the heat-treated scaffold was im-
(PLA)@gelatin-based scaffold functionalized by the crosslinking of CS mersed in water for 2 h, frozen at −80 °C for 2 h, and freeze-dried for
(CS scaffold, CSS). In addition, we fabricated a sole-PLA@gelatin scaf- 48 h. For CSS, the heat-treated scaffold was crosslinked with 50 mL
fold (SS) to serve as a control. In addition to evaluating the scaffolds in solution containing 1% CS, 30 mM EDC, 8 mM NHS, and 50 mM MES
terms of reversible compressibility and mechanical properties, we as- buffer for 24 h [36], washed with distilled water to remove free CS and
sessed their potential for stimulating bone marrow stromal cell (BMSC) salt, frozen at 80 °C for 2 h, and freeze-dried for 48 h.
proliferation and chondrogenic differentiation. The anti-inflammatory
effects of scaffolds were also examined. Furthermore, we evaluated the 2.4. Characterization of scaffolds
effects of the scaffold for cartilage regeneration and immunoregulation
in a rabbit model. 2.4.1. Fourier Transform Infrared Spectroscopy (FTIR)
A Nicolet-670 FTIR spectrometer was used for attenuated total re-
2. Materials and methods flection (ATR)-FTIR of the scaffolds. All spectra were recorded over a
range from 1000 to 4000 cm−1.
2.1. Materials and reagents
2.4.2. Morphology and density of the scaffolds
CS was purchased from J&K Scientific. Poly(L-lactic acid) (PLLA)
Scanning electron microscopy (SEM; Hitachi TM-1000, Japan) was
was kindly provided by Medprin Regenerative Medical Technologies
used to observe the morphology of the scaffolds.
(Guangzhou, China). Type A gelatin was purchased from MP
The densities were calculated by using the following equation [22]:
Biomedicals. Dulbecco's Modified Eagle's Medium (DMEM), fetal bo-
vine serum (Gibco, Carlsbad, CA, USA), and trypsin (Gibco) were pur- ρ = m/v = (4 m)/(d2h) (1)
chased from Shanghai Chengpu Biotechnology N-[3-(dimethylamino)
In this equation, m, v, d, and h represent mass, volume, diameter,
propyl]-N′-ethylcarbodiimide hydrochloride (EDC) and N-hydro-
and height of the scaffold, respectively.
xysuccinimide (NHS) were purchased from J&K Scientific. The anti-
bodies used for immunohistochemistry were against the following
proteins: Collagen2 (COL2; Boster, Wuhan, China), Aggrecan (ACAN; 2.4.3. Water absorption capacity
Boster), iNOS (BIOSS, Beijing, China), and PGES (Boster). Water absorption capacity was tested as previously reported [37].
Chondrogenic medium was purchased from Cyagen Biosciences. All Dry weights (wd) were determined and recorded. Scaffolds were placed
other chemicals were of analytical grade and were used without further into distilled water and removed after 2.5, 10, 30, 60, and 120 min. The
purification. size of the specimen was 14 mm in diameter and 10 mm in height. Wet
weights (ww) were determined and recorded after the superficial water
2.2. Cell isolation and culture on the scaffolds was removed be gentle blotting with a filter paper. The
water absorption capacity (w) was calculated by using the following
Rabbit BMSCs were isolated and cultured as previously described equation:
[35]. Rabbits were provided by the Shanghai Animal Experimental w = (ww − wd)/wd × 100% (2)
Center, and all procedures were approved by the Animal Research
Committee of the Shanghai Jiao Tong University Affiliated Sixth Peo-
ple's Hospital. All experiments were conducted following the guidelines 2.4.4. Compressive mechanical property
of the Animal Research Committee of the Shanghai Jiao Tong Uni- The Instron 5969 testing system with a 200-N loaded sensor was
versity Affiliated Sixth People's Hospital. BMSCs were harvested from used for compressive mechanical property tests of SS and CSS. Five
the hind limbs of rabbits. Briefly, primary cells were cultured in DMEM samples with 13-mm diameter and 14-mm height of each group were
containing 10% fetal bovine serum and 100 U/mL penicillin. Un- used. The compression strain-stress curves of the scaffolds in wet state
attached cells were removed after a 24-h culture, and the adherent cells with compression strains (ε = 40% and ε = 60%) were measured at a
were further cultured. The medium was replaced every three days. strain rate of 10 mm/min.
When reaching 90% confluence, BMSCs were released from the culture
substratum using trypsin/EDTA (0.25% w/v trypsin, 0.02% EDTA), 2.4.5. Micro-computed tomography (CT)
reseeded into dishes, and cultured at 37 °C in a 5% CO2 atmosphere. Total porosity of the scaffolds was evaluated on the basis of 3D
micro-CT images generated with a micro-CT scanner (SkyScan 1176;
2.3. Preparation of 3D electrospun nanofiber scaffolds Bruker, Kontich, Belgium) using Dataviewer and CTAn (Bruker). 3D
reconstruction was conducted using CTvox (Bruker). Five samples of
2.3.1. Preparation of coaxial electrospun nanofibers each group were analyzed in this test. The size of the specimen was
Nanofiber membranes were fabricated by coaxial electrospinning 14 mm in diameter and 10 mm in height.

2
S. Chen, et al. Materials Science & Engineering C 118 (2021) 111312

Fig. 1. Schematic diagram of 3D electrospun scaffold fabrication.


Nanofiber membranes, fabricated by coaxial electrospinning, were cut into pieces and dispersed. The nanofiber dispersions were poured into a cylindrical mold and
frozen and, afterwards, freeze-dried, followed by heating at 180 °C for 2 h. For SS fabrication, the scaffold was subsequently disposed by water. For CSS, scaffold was
crosslinked with CS, EDC, NHS, and MES buffer.

2.5. In vitro study 5 × 105BMSCs were seeded onto the scaffolds and exposed to chon-
drogenic medium (induced condition) or cultured in standard medium
2.5.1. Cell viability, morphology, and infiltration (non-induced condition) for 21 days. BMSCs grown in 6-well plates
Before cell seeding, scaffolds were sterilized by immersion in 75% with DMEM served as a control group. Total cellular RNA was isolated
ethanol for 3 h. The size of the specimen was 12 mm in diameter and using TRIzol (Invitrogen, Carlsbad, CA, USA). The mRNA was reverse-
2 mm in height. Two-hundred microliters of cell suspension containing transcribed into cDNA with the PrimeScript RT Reagent Kit (TaKaRa,
2 × 104 cells was seeded on the scaffolds. Cells viability was de- Dalian, China). The cDNA was amplified by Quantitative TaqMan® real-
termined by using a live/dead cell viability assay (KeyGEN BioTECH, time PCR. qPCR was carried out using SYBR® Premix Ex Taq™ (TaKaRa)
Nanjing, China) on days 3 and 7. Total and live cell numbers were with the following primers: COL2 A1 (forward 5′-CATCCCA CCCTCTC
counted with ImageJ (Version 1.51a; National Institutes of Health, ACAGTT-3′ and reverse 5′-GTCTCTGCCT TGACCCAAAG-3′), and ACAN
Bethesda, MD, USA). In addition, live cell percentages were calculated (forward 5′-TGGTCCACCATTCGGCATA-3′ and reverse 5′-GCCAGAAG
for comparison. Five samples of each scaffold per time point were used ATTCACCACTACCAC-3′). Data were analyzed by the 2−ΔΔCt method,
in this test. Cell morphology was observed by SEM at days 3 and 7. with normalization to the Ct of the housekeeping gene glyceraldehyde
Scaffolds were fixed with glutaraldehyde for 1 h at 4 °C, followed by 3-phosphate dehydrogenase (GAPDH). All experiments were performed
dehydration in an ethanol series. Tert-butanol was used to replace three times.
ethanol in the scaffold. Next, the scaffolds were freeze-dried in a va-
cuum at −60 °C for 24 h. Finally, the samples were gold-coated and
2.5.3. Anti-inflammatory potential of the scaffolds
characterized. For evaluation of BMSC infiltration, 2 × 104 cells were
To evaluate the anti-inflammatory potential of the scaffolds,
seeded onto scaffolds and grown for 7 days. The scaffolds were fixed in
1 × 105 BMSCs were seeded onto the scaffolds and cultured for 24 h
4% paraformaldehyde for 48 h and then embedded in paraffin. Four-
before stimulation. Human recombinant TNF-α (PeproTech) was added
micrometer-thick sections were cut and stained with hematoxylin-eosin
at a concentration of 50 ng/mL for 48 h in all three groups (control, SS,
(HE). Histological observations for infiltration were made under a light
and CSS) to induce an inflammatory micro-environment, as previously
microscope.
reported [38]. Cells cultured in normal conditions (without TNF-α sti-
mulation) were evaluated for comparison. Forty-eight hours after sti-
2.5.2. Assessment of chondrogenic induction properties of the scaffolds by mulation, supernatants were harvested, filtered, and stored at −80 °C.
quantitative (q)PCR Nitric oxide (NO) and prostaglandin E2 (PGE2) (two important in-
To study the effects of the scaffolds on chondrogenic differentiation, flammatory factors in cartilage degenerative disease pathology [28])

3
S. Chen, et al. Materials Science & Engineering C 118 (2021) 111312

levels in culture supernatants were assayed using commercial ELISA 1056 and 1127 cm−1 result from CeO and S]O stretching vibrations of
kits from MyBioSource (San Diego, CA, US) and antibodies-online hydroxyl groups on CS, respectively. The bands at 1370 and 1412 cm−1
(Aachen, Germany), respectively, following the manufacturers' in- represent C]O and CeO stretching vibrations of ionized carboxyl
structions. All experiments were performed three times. groups [40]. For scaffolds without CS (Fig. 2A, SS), a peak at
1240 cm−1, which results from CeN stretching vibration of amide III,
2.6. In vivo study was observed. However, after CS crosslinking (Fig. 2A, CSS), the peak at
1240 cm−1 was decreased due to the influence of S]O stretching vi-
2.6.1. Articular joint drilling surgery bration of CS. Because CS and gelatin showed many overlapping peaks,
The project and experiments were approved by the Animal Research IR spectra indicated limited clear differences.
Committee of the Shanghai Jiao Tong University Affiliated Sixth
People's Hospital. Animals were treated according to the standard 3.1.2. Compressive mechanical properties
guidelines of our hospital. Rabbits were divided into three treatment We observed no differences in compressive mechanical properties
groups; control, SS, and CSS (n = 6 per group). The rabbits were an- induced by CS crosslinking: as shown in Fig. 2B, SS and CSS exhibited
esthetized with pentobarbital sodium (30 mg/kg). Knee joints were similar compressive strengths, in dry state, the compressive modulus of
revealed through the lateral approach, followed by patella dislocation SS and CSS were 709 KPa and 657 KPa, respectively. Furthermore, SS
for exposure. Osteochondral drilling defects of 3-mm diameter and 4- and CSS possessed similar elastic properties in the wet state. Fig. 2C
mm depth were created in the trochlear groove of the femur. shows the compressive stress-strain curves of SS and CSS under 40%
Subsequently, two scaffolds were implanted. In the control group, the and 60% strain, respectively. Highly nonlinear and closed hysteresis
defects were untreated. Next, the joint was closed with suture. was detected. Thus, the two scaffolds exhibited the same compressive
Penicillin was given intramuscularly as a prophylactic. The rabbits were stress (ε = 40% and ε = 60%).
allowed to move freely in cages postoperatively. The sizes of the scaf-
folds used in animal study were all 3 mm in diameter and 4 mm in 3.1.3. Water absorption capacity
height. Both scaffolds presented excellent water absorption capacity
(Fig. 2D), likely owing to the hydrophilic properties of gelatin and the
2.6.2. Gross examination and histological analysis porous structure. However, CSS exhibited a water absorption behavior
Rabbits were euthanized 12 weeks postoperatively. Samples were inferior to that of SS, likely because of a decrease in the number of
observed and photographed for evaluation according to the carboxyl and amino groups after CS crosslinking with EDC/NHS.
International Cartilage Repair Society (ICRS) macroscopic assessment
scale for cartilage repair [39]. Six samples in each group were used for 3.1.4. Morphology, micro-CT evaluation, and density
evaluation. After gross examination, dissected distal femurs were fixed Photographs of the scaffolds are shown in Fig. 3A (dry state) and 3B
in 4% paraformaldehyde for 48 h, decalcified with 10% EDTA, and (wet state). Both scaffolds appeared to be porous. Micro-CT images of
paraffin-embedded. Four-micrometer-thick sections were cut and SS and CSS showed porous appearances (Fig. 3C, D). Quantification of
stained with HE or safranin O/fast green to assess GAG distribution. the total porosity (%) indicated no significant difference between SS
Histological and immunohistological observations were made under a and CSS (Fig. 3E), with porosities of 74.67% ± 8.81% and
light microscope. 66.35% ± 4.08%, respectively. The addition of CS may explain the
slightly lower porosity of CSS.
2.6.3. Immunohistological analysis SEM of SS (Fig. 3F, G) and CSS (Fig. 3H, I) demonstrated that the
To evaluate cartilage degeneration, the expression of COL2 and porous scaffolds are composed of nanofibers. All scaffolds showed an
ACAN in defects was analyzed by immunohistological staining. interconnected and cellular structure that mimicked the natural ECM.
Deparaffinized sections were hydrated in a graded series of alcohol. The As shown in the images, pores of various sizes and shapes were scat-
sections were placed in 10 mM citrate buffer solution (pH 6.0) and tered throughout the scaffolds. However, CSS shows a more compact
heated in a 95 °C bath to retrieve the antigens. Hydrogen peroxide structure than SS, which may be the result of CS crosslinking. The
(0.3%) and goat serum (1:100 dilution) were used to respectively block densities of SS and CSS were 94.5 mg/cm3 and 121.7 mg/cm3, re-
endogenous peroxidase activity and non-specific binding sites. Then, spectively. The density of CSS was higher than that of SS, most likely
the sections were incubated overnight at 4 °C with primary antibodies because of the addition of CS.
against the cartilage regeneration markers anti-COL2 and anti-ACAN,
and against iNOS and PGES, which are inducible enzymes that produce 3.2. In vitro study
NO and prostaglandins, respectively. Then, the sections were incubated
with secondary antibodies for 1 h at 37 °C and stained in DAB solution 3.2.1. Cell viability, morphology, and infiltration
(Dako, Hamburg, Germany). After BMSCs were seeded on the scaffolds for 3 or 7 days, cell vi-
abilities within SS and CSS were determined by live/dead assay, and the
2.6.4. Statistical analysis results are shown in Fig. 4A, B and C, D, respectively. On day 3, large
All quantitative data are presented as means ± standard deviations. numbers of live cells were detected on both scaffolds (Fig. 4A, C) in-
Statistical analysis was carried out using one-way analysis of variance dicating appropriate biocompatibility of SS and CSS. The total cell
(ANOVA) followed by the Tukey HSD test for post-hoc comparison number in CSS was 1.92-fold change (p < 0.01) in comparison with
(SPSS, version 22; IBM, Armonk, NY, USA). Differences were con- that in SS. The percentage of live cells was 1.54-fold change
sidered significant at p < 0.05. (p < 0.01). On day 7, the numbers of live cells were increased, and
only few dead cells were detected on both scaffolds (Fig. 4B, D). The
3. Results total cell number in CSS was 1.47-fold change (p < 0.01) than that in
SS, while the live-cell percentage was 1.32-fold change (p < 0.01).
3.1. Characterization of 3D scaffolds These results demonstrated that the scaffolds did not negatively affect
the proliferation of BMSCs. Importantly, BMSCs seeded on CSS pre-
3.1.1. ATR-FTIR sented significant higher viability than those on SS, both in total cell
We used ATR-FTIR to evaluate differences in molecular structure number and live cell percentage. BMSC adhesion and proliferation were
induced by the crosslinking of CS. In Fig. 2A (Pure CS), the character- evaluated by SEM (Fig. 4E–H). BMSCs adhered and proliferated on the
istic absorption peaks of pure CS could be observed clearly, the peaks at nanofibers and exhibited normal cell morphology. Cells nearly

4
S. Chen, et al. Materials Science & Engineering C 118 (2021) 111312

Fig. 2. Mechanical properties of the scaffolds.


(A) ATR-FTIR spectra of SS, CSS, and pure CS. (B) Compressive stress−strain curves of SS and CSS. (C) Compressive stress−strain curves of the SS and CSS under
compressing and releasing cycles (ε = 40% or 60%). (D) Water absorption rate of the SS and CSS.

contacted each other and formed a cell sheet. These findings demon- infiltration was examined by HE, which demonstrated better pro-
strated an appropriate biocompatibility of both scaffolds. In addition, in liferation on CSS as indicated by the large number of cells (violet blue)
comparison with those on SS (Fig. 4E, F), cells seeded on CSS (Fig. 4G, observed throughout the scaffold (Fig. 4I–L).
H) presented a better morphology, exhibiting a spindle shape. Cell

Fig. 3. Appearance, micro-CT, total porosity, and SEM of 3D scaffolds.


(A, B) The appearances of scaffolds in dry (A) and wet (B) state. (C, D) Micro-CT images of SS (C) and CSS (D). (E) Total porosity of the scaffolds. Results are
mean ± S.D. n = 5 per group. NS, no significant difference. (F–I) SEM images of SS (F, G) and CSS (H, I). Magnification: 200 × (F, H), 1000 × (G, I).

5
S. Chen, et al. Materials Science & Engineering C 118 (2021) 111312

Fig. 4. CSS exhibits favorable cell viability, morphology, and infiltration in comparison with SS.
(A–D) Live/dead assay of BMSCs seeded on SS (A, B) and CSS (C, D). Magnification, 100×. (E–H) SEM images of BMSCs seeded on SS (E, F) and CSS (G, H).
Magnification, 500×. (I–L) The infiltration of BMSCs cultured on SS (I, J) and CSS (K, L) by HE staining. BMSCs in scaffolds were stained with violet blue.
Magnification: 40 × (I, K), 100 × (J, L). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

3.2.2. Chondrogenic differentiation (Fig. 5C). These findings indicated that CSS may have preferable im-
mRNA expression of two chondrogenic genes, COL2 and ACAN, was munosuppressive potential than SS under TNF-α stimulation, which
assessed by qPCR to evaluate the chondrogenic inductive potential of imitated the inflammatory environment in cartilage damage.
the scaffolds after 21 days (Fig. 5). BMSCs cultured in plates served as a Similarly, under normal condition, no significant difference in PGE2
control group. In standard medium (non-induced condition), the ex- level was found among the three groups, while after stimulation, PGE2
pression of COL2 in the CSS group was higher than that in the control production in all groups increased (Fig. 5D). PGE2 release after TNF-α
(8.00-fold, p < 0.05) and SS (8.30-fold, p < 0.05) groups. Further- treatment was the highest in the SS group. The level of PGE2 in the CSS
more, in chondrogenic medium (induced condition), COL2 expression group was significantly lower than that in the control (0.87-fold,
was significantly higher in the CSS than in the control (4.56-fold, p < 0.05) and SS (0.83-fold, p < 0.01) groups (Fig. 5D). PGE2 pro-
p < 0.01) and SS (2.29-fold, p < 0.05) groups. In addition, COL2 duction was 1.04-fold higher in the SS than in the control group, but the
expression in the SS group was significantly higher than that in the difference was not significant. These results demonstrated that CSS may
control group (1.99-fold, p < 0.05) (Fig. 5A). Similarly, ACAN mRNA have superior immunosuppressive effect in comparison with SS. Taken
expression was significantly higher in the CSS group than in the control together, although the anti-inflammatory effect of SS was uncertain, all
(12.97-fold, p < 0.05) and SS (13.65-fold, p < 0.05) groups in findings demonstrated the potential immunosuppressive property of
standard medium. In chondrogenic medium, ACAN expression was CSS, which indicated that crosslinking of CS may be effective for anti-
significantly higher in the CSS than in the control (3.02-fold, p < 0.01) inflammation.
and SS (1.60-fold, p < 0.05) groups. In addition, ACAN expression was
significantly higher in the SS than in the control group (1.89-fold, 3.3. In vivo study using a rabbit model
p < 0.05) (Fig. 5B). These findings suggested that CSS is effective in
inducing chondrogenic differentiation of BMSCs in both standard and 3.3.1. Gross appearance and ICRS score
chondrogenic medium. Twelve weeks after transplantation, an obvious defect and little
white neo-tissue was observed in the boundary area in the control
3.2.3. Immunosuppressive potential of scaffolds in vitro group (Fig. 6A). In the SS and CSS groups, neo-tissue appeared to be
To test the immunosuppressive potential of the scaffolds in vitro, we integrated with the surrounding tissue (Fig. 6B, C). In the SS group, the
examined the expression of NO and PGE2 in BMSCs seeded onto the defect was filled with white tissue with small fissures (Fig. 6B). The
different scaffolds, with and without TNF-α stimulation (Fig. 5C, D). defect in the CSS group was filled with cartilage-like tissue that was
Under normal condition (without stimulation), no significant difference similar to the surrounding cartilage. The defect was well integrated
was found among three groups with regard to NO production (Fig. 5C). with the surrounding tissue, with a smooth and flat surface (Fig. 6C).
After stimulation, the NO level increased in all groups. CSS produced ICRS scores obtained from macroscopic evaluation were 1.7 ± 0.6,
the lowest level of NO, with significant differences as compared to the 7.3 ± 2.5, and 9.7 ± 2.5 in the control, SS, and CSS groups, re-
control (0.82-fold, p < 0.05) and SS (0.74-fold, p < 0.01) groups spectively. The ICRS scores showed substantial improvement in the CSS

6
S. Chen, et al. Materials Science & Engineering C 118 (2021) 111312

Fig. 5. CSS exhibits superior chondrogenic inductive and immunosuppressive potential in vitro.
(A, B) mRNA expression levels of COL2 (A) and ACAN (B). Gene expression levels were expressed as fold change compared to GAPDH expression. n = 5; *p < 0.05,
**p < 0.01. (C, D) NO (C) and PGE2 (D) concentrations in the supernatants of BMSCs cultured on scaffolds in normal and TNF-α stimulation condition. n = 5;
*p < 0.05, **p < 0.01.

group as compared to the non-treated group at 12 weeks (p < 0.01). indicated by the intense red staining on the defect surfaces. The neo-
The ICRS score for the SS group was significantly higher that for the tissues were integrated with the native tissue and showed a similar
non-treated group (p < 0.05). In addition, the CSS group presented a color and organized architecture, demonstrating favorable cartilage
higher score than the SS group (Fig. 6D). These findings demonstrated regeneration (Fig. 7I, L).
that CSS presents superior outcomes for cartilage regeneration in rab-
bits.
3.3.3. Immunohistochemistry
The expression levels of COL2 and ACAN in vivo were examined by
immunohistochemistry (Fig. 8A). In the non-treated group, limited
3.3.2. HE and safranin O/fast green staining
COL2 and ACAN expression was observed, and only at the bottom of the
Six weeks after implantation, the defects still existed in the non-
defects. Compared with the non-treated group, the SS group presented
treated group, with thin fibrous-like tissue appearing on the surface as
an increased expression level of both markers. However, they were even
shown by HE staining (Fig. 6E, H). However, in the SS group, loose neo-
more strongly expressed in the CSS group, indicating superior cartilage
tissues were found in the defect (Fig. 6F, I). In the CSS group, dense
regeneration. Immunohistochemistry largely matched the findings of
fibrous-like tissue was found in defects infiltrated by large numbers of
HE and safranin O/fast green staining, and corroborated that CSS ex-
cells (Fig. 6G, J). Moreover, safranin O/fast green showed few light-red-
hibited the best properties allowing cartilage regeneration.
stained neo-tissues, indicating that almost no GAG was deposited in the
To evaluate the anti-inflammatory effect of the scaffolds, we ex-
non-treated group (Fig. 6K, N). In contrast, GAG was deposited in the
amined the expression of iNOS and PGES in vivo (Fig. 8B). Compared to
defects of SS and CSS groups, with the CSS group showing the highest
the non-treated group, the expression levels of iNOS and PGES were
deposition (Fig. 6L, M, O, P).
decreased in the SS group as shown by immunohistochemistry. In the
At 12 weeks after implantation, the defects in the non-treated group
CSS group, both markers were significantly decreased as compared to
were not healed, and we observed little thin connective tissue on the
the non-treated and SS group. These results indicated that CSS possesses
surface (Fig. 7A, D). In the SS group, HE staining showed that the de-
anti-inflammatory properties in vivo.
fects were filled up with dense neo-tissues, which was likely connective
tissue with a disordered structure. The cells were mostly fibroblast-like
cells (Fig. 7B, E). In contrast, the defects in the CSS group were almost 4. Discussion
entirely filled up with thick, more cartilage-like tissue. The cells were
mostly chondrocyte-like cells (Fig. 7C, F). In addition, safranin O/fast In this study, we successfully fabricated 3D electrospun scaffolds
green staining demonstrated poor GAG deposition in the non-treated possessing distinctive mechanical properties and evaluated their po-
group (Fig. 7G, J). Limited red staining was detected on the defect tential for cartilage regeneration and immunoregulation. The results
surfaces, indicating inferior cartilage regeneration (Fig. 7H, K). How- demonstrated that BMSCs seeded on CSS showed better chondrogenic
ever, in the CSS group, GAG formation was clearly detectable, as differentiation than cells seeded on SS and the control scaffold. In

7
S. Chen, et al. Materials Science & Engineering C 118 (2021) 111312

(caption on next page)

8
S. Chen, et al. Materials Science & Engineering C 118 (2021) 111312

Fig. 6. Gross observation and ICRS score at 12 weeks after implantation, and preferable cartilage repair in CSS as compared SS and non-treated scaffold by
histological evaluation at 6 weeks.
(A–C) Representative gross observations of cartilage defects. The black dotted lines show the defects. (D) ICRS evaluation of cartilage repair. n = 6; *p < 0.05,
**p < 0.01. (E–J) HE of cartilage regeneration in defects. (K–P) Safranin O/fast green staining of cartilage regeneration. The edges of defects are indicated by black
arrows. Blue dotted lines indicate the implantation areas. (For interpretation of the references to color in this figure legend, the reader is referred to the web version
of this article.)

addition, CSS showed better anti-inflammatory properties than control studies reported cartilage regeneration by 3D printed or braid scaffolds
and SS in vitro. Furthermore, in a rabbit model, CSS enhanced cartilage that were fabricated with custom-made machinery [14–16]. The
regeneration as compared to control and SS. Finally, the anti-in- printed scaffolds were efficient in chondrogenic regeneration [14,15],
flammatory effect of CSS is also superior. and the braid scaffold displayed mechanical properties of cartilage re-
Scaffolds play a vital role in controlling cell growth and differ- pair in vitro [16]. These findings indicated the potential of 3D scaffolds
entiation for tissue regeneration [41–43]. Karbasi et al. [4–8] used in cartilage repair. The complicated production process and relatively
multi-walled carbon nanotubes to fabricate electrospun scaffolds which low porosity of materials remain challenge for scaffolds fabrication.
showed appropriate potential for cartilage regeneration. However, However, our 3D scaffolds were fabricated by a more convenient and
these materials are relatively two dimensional which may limit its economic method. Additionally, our previous study, in which the same
porosity to some degree in comparison with 3D structure. Previous method was applied, showed the feasibility of 3D scaffolds in cartilage

Fig. 7. The favorable cartilage regeneration in CSS by histological evaluation at 12 weeks after implantation.
(A–F) HE of cartilage regeneration in defects. (G–L) Safranin O/fast green staining of cartilage regeneration. The edges of defects are indicated by black arrows. (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

9
S. Chen, et al.

10
Fig. 8. The potential immunosuppressive property of CSS at 12 weeks after implantation.
Immunohistochemistry images of cartilage regeneration (A) and anti-inflammation (B). The edges of the defect are indicated by black arrows.
Materials Science & Engineering C 118 (2021) 111312
S. Chen, et al. Materials Science & Engineering C 118 (2021) 111312

repair [44]. In this study, CSS showed appropriate compressive me- confirmed that CSS exhibits not only appropriate mechanical property,
chanical properties and water absorption capacity. In addition, micro- but also appropriate biocompatibility. In vitro, CSS exhibited superior
CT and SEM images indicated a favorable porous structure. Further, SS chondrogenic differentiation potential and better anti-inflammation
exhibited appropriate properties for cell proliferation and infiltration in effect than the control and SS. In vivo, CSS effectively promoted carti-
vitro. lage repair in rabbit cartilage defects. Additionally, superior anti-in-
Native ECM is an attractive choice for scaffold modification owing flammation effect of CSS was found in animal study. Our results in-
to the inherent cues that regulate specific cell proliferation and differ- dicated the preferable property of CSS for cartilage regeneration and its
entiation [45–47]. Among native ECMs, CS plays an important role in potential in immunoregulation.
cell communication, which regulates crucial physiological processes
such as induction of differentiation [33,48]. Our findings showed that Author statement
after the crosslinking of CS, COL2 and ACAN mRNA expression in
BMSCs seeded in CSS was significantly higher than that in the SS and Shuai Chen: Conceptualization, Methodology, Investigation,
control groups in vitro, which suggested a better chondrogenic effect of Software, Data curation, Writing-Original draft preparation. Weiming
CSS. Further, CSS was superior to SS in terms of cartilage regeneration Chen: Methodology, Visualization, Investigation, Validation. Yini Chen:
with higher COL2 and ACAN expression in our rabbit model, as shown Software, Validation. Xiumei Mo: Supervision. Cunyi Fan: Supervision,
by immunohistology. These findings are in accordance with previous Writing- Reviewing and Editing,
studies showing that CS regulates cartilage regeneration by promoting
collagen and proteoglycan synthesis [27,49]. Moreover, the functio- Declaration of competing interest
nalization of CS did not significantly influence the physical properties
of the scaffold, as shown by the compressive mechanical properties and The authors declare that they have no known competing financial
water absorption capacity, as well as micro-CT image and SEM ana- interests or personal relationships that could have appeared to influ-
lyses. ence the work reported in this paper.
CS not only is a major component of the ECM, it also plays a more
important role in inflammation and immunity than other components, Acknowledgements
owing to its intrinsic stability and low immunogenicity [50,51]. Our
results showed that CSS reduce the generation of pro-inflammatory This work was supported by the Youth Program of National Natural
cytokines, PGE2 and NO, in vitro. NO is an important factor in the pa- Science Foundation of China (81802157), National Natural Science
thology of cartilage damage because it can induce inflammatory com- Foundation of China (81830076, 81672146), and clinical project of
ponents and may lead to tissue injury [28]. Meanwhile, the crucial role Shanghai Jiao Tong University Affiliated Sixth People's Hospital
played by PGE2 in cartilage damage has been demonstrated in animal (ynlc201821).
models [52–54]. PGE2 increases the generation of catabolic factors,
which in turn induce further deterioration of the cartilage [28]. Our Appendix A. Supplementary data
results are consistent with those of previous studies showing that CS
efficiently suppresses the production of pro-inflammatory cytokines Supplementary data to this article can be found online at https://
[55,56]. We further examined the immunomodulatory properties of doi.org/10.1016/j.msec.2020.111312.
CSS in a rabbit model, which demonstrated reduced expression of PGES
and iNOS, inducible enzymes producing PGE2 and NO, respectively References
[57], in the CSS group. Taken together, these findings indicate that
functionalization by CS potentially improve the immunomodulatory [1] K. Shimomura, A.C. Bean, H. Lin, N. Nakamura, R.S. Tuan, In vitro repair of me-
properties of the scaffold. As chronic inflammation in cartilage de- niscal radial tear using aligned electrospun nanofibrous scaffold, Tissue Eng. A 21
struction may provide an inferior environment for regeneration [58], (13–14) (2015) 2066–2075.
[2] S. Khorshidi, A. Solouk, H. Mirzadeh, S. Mazinani, J.M. Lagaron, S. Sharifi,
the regulation of inflammation could therefore create favorable con- S. Ramakrishna, A review of key challenges of electrospun scaffolds for tissue-en-
ditions for cartilage repair [59]. Thus, the immunomodulatory effect of gineering applications, J. Tissue Eng. Regen. Med. 10 (9) (2016) 715–738.
CSS may have positive impacts on cartilage regeneration. For cartilage [3] T.J. Sill, H.A. von Recum, Electrospinning: applications in drug delivery and tissue
engineering, Biomaterials 29 (13) (2008) 1989–2006.
degenerative diseases that are accompanied by persistent and intense [4] P. Zadehnajar, B. Akbari, S. Karbasi, M.H. Mirmusavi, Preparation and character-
inflammation, such as osteoarthritis, CSS holds promise for cartilage ization of poly ε-caprolactone-gelatin/multi-walled carbon nanotubes electrospun
repair. scaffolds for cartilage tissue engineering applications, Int. J. Polym. Mater. Polym.
Biomater. 69 (5) (2020) 326–337.
In this study, we fabricate 3D porous electrospun nanofiber scaffolds [5] M. Nikbakht, S. Karbasi, S.M. Rezayat, Biological evaluation of the effects of hya-
by a convenient method, and functionalize it by CS to promote cartilage luronic acid on poly (3-hydroxybutyrate) based electrospun nanocomposite scaf-
regeneration. However, there are limitations to this study. Further folds for cartilage tissue engineering application, Mater. Technol. 35 (3) (2020)
141–151.
studies are required to compare our method with conventional elec-
[6] Z. Mohammadalizadeh, S. Karbasi, S. Arasteh, Physical, mechanical and biological
trospinning. In addition, the feasibility of CSS for cartilage tissue re- evaluation of poly (3-hydroxybutyrate)-chitosan/MWNTs as a novel electrospun
generation needs to be tested. The mechanisms underlying the chon- scaffold for cartilage tissue engineering applications, Polym.-Plast. Technol. Mater.
drogenic and anti-inflammatory effects of CSS were not studied, and 59 (4) (2020) 417–429.
[7] M.H. Mirmusavi, P. Zadehnajar, D. Semnani, S. Karbasi, F. Fekrat, F. Heidari,
signaling pathways related to the pharmacological functions of CS were Evaluation of physical, mechanical and biological properties of poly 3-hydro-
not evaluated. Future studies will be conducted focusing on the ma- xybutyrate-chitosan-multiwalled carbon nanotube/silk nano-micro composite
nipulation of CS dosage to achieve optimal cartilage repair. In addition, scaffold for cartilage tissue engineering applications, Int. J. Biol. Macromol. 132
(2019) 822–835.
evaluation of CSS in big animal models, such as dogs, is needed because [8] Z. Shahali, S. Karbasi, M.R. Avadi, D. Semnani, E. Naghash Zargar, B. HashemiBeni,
these are more suitable for the evaluation of cartilage repair than Evaluation of structural, mechanical, and cellular behavior of electrospun poly-3-
rabbits. hydroxybutyrate scaffolds loaded with glucosamine sulfate to develop cartilage
tissue engineering, Int. J. Polym. Mater. Polym. Biomater. 66 (12) (2017) 589–602.
[9] W. Cui, Y. Zhou, J. Chang, Electrospun nanofibrous materials for tissue engineering
5. Conclusion and drug delivery, Sci. Technol. Adv. Mater. 11 (1) (2010) 014108.
[10] T. Jiang, E.J. Carbone, K.W.-H. Lo, C.T. Laurencin, Electrospinning of polymer
nanofibers for tissue regeneration, Prog. Polym. Sci. 46 (2015) 1–24.
In this study, we designed a 3D porous electrospun scaffold by a [11] X. Wang, B. Ding, B. Li, Biomimetic electrospun nanofibrous structures for tissue
novel method and further functionalized it by the crosslink of CS to engineering, Mater. Today 16 (6) (2013) 229–241.
promote cartilage regeneration and modulate inflammation. We [12] M.R. Badrossamay, H.A. McIlwee, J.A. Goss, K.K. Parker, Nanofiber assembly by

11
S. Chen, et al. Materials Science & Engineering C 118 (2021) 111312

rotary jet-spinning, Nano Lett. 10 (6) (2010) 2257–2261. [36] J.S. Mao, Y.J. Yin, K. De Yao, The properties of chitosan–gelatin membranes and
[13] S. Soliman, S. Pagliari, A. Rinaldi, G. Forte, R. Fiaccavento, F. Pagliari, O. Franzese, scaffolds modified with hyaluronic acid by different methods, Biomaterials 24 (9)
M. Minieri, P. Di Nardo, S. Licoccia, Multiscale three-dimensional scaffolds for soft (2003) 1621–1629.
tissue engineering via multimodal electrospinning, Acta Biomater. 6 (4) (2010) [37] P. Jithendra, A.M. Rajam, T. Kalaivani, A.B. Mandal, C. Rose, Preparation and
1227–1237. characterization of aloe vera blended collagen-chitosan composite scaffold for
[14] K.C. Hung, C.S. Tseng, L.G. Dai, S.H. Hsu, Water-based polyurethane 3D printed tissue engineering applications, ACS Appl. Mater. Interfaces 5 (15) (2013)
scaffolds with controlled release function for customized cartilage tissue en- 7291–7298.
gineering, Biomaterials 83 (2016) 156. [38] B. Hegyi, G. Kudlik, É. Monostori, F. Uher, Activated T-cells and pro-inflammatory
[15] Q. Yao, B. Wei, N. Liu, C. Li, Y. Guo, A.N. Shamie, J. Chen, T. Cheng, C. Jin, Y. Xu, cytokines differentially regulate prostaglandin E2 secretion by mesenchymal stem
Chondrogenic regeneration using bone marrow clots and a porous poly- cells, Biochem. Biophys. Res. Commun. 419 (2) (2012) 215–220.
caprolactone-hydroxyapatite scaffold by 3D printing, Tissue Eng. A 21 (7–8) [39] M. Van Den Borne, N. Raijmakers, J. Vanlauwe, J. Victor, S. de Jong, J. Bellemans,
(2014). D. Saris, International Cartilage Repair Society (ICRS) and Oswestry macroscopic
[16] H. Ahn, K.J. Kim, S.Y. Park, J.E. Huh, H.J. Kim, W.R. Yu, 3D braid scaffolds for cartilage evaluation scores validated for use in Autologous Chondrocyte
regeneration of articular cartilage, J. Mech. Behav. Biomed. Mater. 34 (6) (2014) Implantation (ACI) and microfracture, Osteoarthr. Cartil. 15 (12) (2007)
37–46. 1397–1402.
[17] B. Sun, Y.Z. Long, H.D. Zhang, M.M. Li, J.L. Duvail, X.Y. Jiang, H.L. Yin, Advances [40] H. Tian, Y. Chen, C. Ding, G. Li, Interaction study in homogeneous collagen/
in three-dimensional nanofibrous macrostructures via electrospinning, Prog. Polym. chondroitin sulfate blends by two-dimensional infrared spectroscopy, Carbohydr.
Sci. 39 (5) (2014) 862–890. Polym. 89 (2) (2012) 542–550.
[18] Q.P. Pham, U. Sharma, A.G. Mikos, Electrospun poly (ε-caprolactone) microfiber [41] D.L. Butler, S.A. Goldstein, F. Guilak, Functional tissue engineering: the role of
and multilayer nanofiber/microfiber scaffolds: characterization of scaffolds and biomechanics, J. Biomech. Eng. 122 (6) (2000) 570–575.
measurement of cellular infiltration, Biomacromolecules 7 (10) (2006) 2796–2805. [42] B.O. Diekman, F. Guilak, Stem cell-based therapies for osteoarthritis: challenges and
[19] S.H. Park, T.G. Kim, H.C. Kim, D.-Y. Yang, T.G. Park, Development of dual scale opportunities, Curr. Opin. Rheumatol. 25 (1) (2013) 119.
scaffolds via direct polymer melt deposition and electrospinning for applications in [43] S. Panseri, A. Russo, C. Cunha, A. Bondi, A. Di Martino, S. Patella, E. Kon,
tissue regeneration, Acta Biomater. 4 (5) (2008) 1198–1207. Osteochondral tissue engineering approaches for articular cartilage and sub-
[20] Y. Si, Q. Fu, X. Wang, J. Zhu, J. Yu, G. Sun, B. Ding, Superelastic and super- chondral bone regeneration, Knee Surg. Sports Traumatol. Arthrosc. 20 (6) (2012)
hydrophobic nanofiber-assembled cellular aerogels for effective separation of oil/ 1182–1191.
water emulsions, ACS Nano 9 (4) (2015) 3791–3799. [44] W. Chen, J. Ma, L. Zhu, Y. Morsi, E.-H. Hany, S.S. Al-Deyab, X. Mo, Superelastic,
[21] Y. Si, J. Yu, X. Tang, J. Ge, B. Ding, Ultralight nanofibre-assembled cellular aerogels superabsorbent and 3D nanofiber-assembled scaffold for tissue engineering,
with superelasticity and multifunctionality, Nat. Commun. 5 (2014). Colloids Surf. B: Biointerfaces 142 (2016) 165–172.
[22] G. Duan, S. Jiang, V. Jérôme, J.H. Wendorff, A. Fathi, J. Uhm, S. Agarwal, [45] S.F. Badylak, The extracellular matrix as a biologic scaffold material, Biomaterials
Ultralight, soft polymer sponges by self-assembly of short electrospun fibers in 28 (25) (2007) 3587–3593.
colloidal dispersions, Adv. Funct. Mater. 25 (19) (2015) 2850–2856. [46] G.C. Reilly, A.J. Engler, Intrinsic extracellular matrix properties regulate stem cell
[23] T. Xu, J.M. Miszuk, Y. Zhao, H. Sun, H. Fong, Bone tissue engineering: electrospun differentiation, J. Biomech. 43 (1) (2010) 55–62.
polycaprolactone 3D nanofibrous scaffold with interconnected and hierarchically [47] K.G. Cornwell, A. Landsman, K.S. James, Extracellular matrix biomaterials for soft
structured pores for bone tissue engineering, Adv. Healthc. Mater. 4 (15) (2015) tissue repair, Clin. Podiatr. Med. Surg. 26 (4) (2009) 507–523.
2237. [48] K. Sugahara, T. Mikami, T. Uyama, S. Mizuguchi, K. Nomura, H. Kitagawa, Recent
[24] J.A. Rowley, G. Madlambayan, D.J. Mooney, Alginate hydrogels as synthetic ex- advances in the structural biology of chondroitin sulfate and dermatan sulfate, Curr.
tracellular matrix materials, Biomaterials 20 (1) (1999) 45–53. Opin. Struct. Biol. 13 (5) (2003) 612–620.
[25] A.W. Lund, B. Yener, J.P. Stegemann, G.E. Plopper, The natural and engineered 3D [49] K. Ishizeki, Y. Hiraki, M. Kubo, T. Nawa, Sequential synthesis of cartilage and bone
microenvironment as a regulatory cue during stem cell fate determination, Tissue marker proteins during transdifferentiation of mouse Meckel’s cartilage chon-
Eng. B Rev. 15 (3) (2009) 371. drocytes in vitro, Int. J. Dev. Biol. 41 (1) (1997) 83–89.
[26] J.D. Esko, K. Kimata, U. Lindahl, Proteoglycans and Sulfated Glycosaminoglycans, [50] N. Volpi, Anti-inflammatory activity of chondroitin sulphate: new functions from an
(2009). old natural macromolecule, Inflammopharmacology 19 (6) (2011) 299–306.
[27] M. McCarty, A. Russell, M. Seed, Sulfated glycosaminoglycans and glucosamine [51] M. Vallieres, P. Du Souich, Modulation of inflammation by chondroitin sulfate,
may synergize in promoting synovial hyaluronic acid synthesis, Med. Hypotheses Osteoarthr. Cartil. 18 (2010) S1–S6.
54 (5) (2000) 798–802. [52] L.K. Myers, A.H. Kang, A.E. Postlethwaite, E.F. Rosloniec, S.G. Morham,
[28] J. Monfort, J.-P. Pelletier, N. Garcia-Giralt, J. Martel-Pelletier, Biochemical basis of B.V. Shlopov, S. Goorha, L.R. Ballou, The genetic ablation of cyclooxygenase 2
the effect of chondroitin sulphate on osteoarthritis articular tissues, Ann. Rheum. prevents the development of autoimmune arthritis, Arthritis Rheum. 43 (12) (2000)
Dis. 67 (6) (2008) 735–740. 2687–2693.
[29] S. Varghese, N.S. Hwang, A.C. Canver, P. Theprungsirikul, D.W. Lin, J. Elisseeff, [53] J.M. McCoy, J.R. Wicks, L.P. Audoly, The role of prostaglandin E2 receptors in the
Chondroitin sulfate based niches for chondrogenic differentiation of mesenchymal pathogenesis of rheumatoid arthritis, J. Clin. Invest. 110 (5) (2002) 651–658.
stem cells, Matrix Biol. 27 (1) (2008) 12–21. [54] B. Averbeck, K. Rudolphi, M. Michaelis, Osteoarthritic mice exhibit enhanced
[30] A. Kahan, D. Uebelhart, F. De Vathaire, P.D. Delmas, J.Y. Reginster, Long-term prostaglandin E2 and unchanged calcitonin gene-related peptide release in a novel
effects of chondroitins 4 and 6 sulfate on knee osteoarthritis: the study on os- isolated knee joint model, J. Rheumatol. 31 (10) (2004) 2013–2020.
teoarthritis progression prevention, a two-year, randomized, double-blind, placebo- [55] C. Jomphe, M. Gabriac, T.M. Hale, L. Héroux, L.É. Trudeau, D. Deblois, E. Montell,
controlled trial, Arthritis Rheum. 60 (2) (2009) 524–533. J. Vergés, P. Du Souich, Chondroitin sulfate inhibits the nuclear translocation of
[31] B.A. Michel, G. Stucki, D. Frey, F. De Vathaire, E. Vignon, P. Bruehlmann, nuclear factor-κB in interleukin-1β-stimulated chondrocytes, Basic Clin. Pharmacol.
D. Uebelhart, Chondroitins 4 and 6 sulfate in osteoarthritis of the knee: a rando- Toxicol. 102 (1) (2008) 59–65.
mized, controlled trial, Arthritis Rheum. 52 (3) (2005) 779–786. [56] M.A.-S.R. Largo, E. Calvo, J. Egido, G. Herrer-Beaumont, Differential Anticatabolic
[32] G.-K. Tan, Y. Tabata, Chondroitin-6-sulfate attenuates inflammatory responses in Profile of Glucosamine Sulfate Versus Other Anti-osteoarthritic Drugs on Human
murine macrophages via suppression of NF-κB nuclear translocation, Acta Osteoarthritic Chondrocytes and Synovial Fibroblast in Culture, (2005).
Biomater. 10 (6) (2014) 2684–2692. [57] B. Corradetti, F. Taraballi, S. Minardi, J. Van Eps, F. Cabrera, L.W. Francis,
[33] W.H. Liang, B.L. Kienitz, K.J. Penick, J.F. Welter, T.A. Zawodzinski, H. Baskaran, S.A. Gazze, M. Ferrari, B.K. Weiner, E. Tasciotti, Chondroitin sulfate immobilized on
Concentrated collagen-chondroitin sulfate scaffolds for tissue engineering applica- a biomimetic scaffold modulates inflammation while driving chondrogenesis, Stem
tions, J. Biomed. Mater. Res. A 94 (4) (2010) 1050–1060. Cells Transl. Med. 5 (5) (2016) 670–682.
[34] I. Yannas, J. Burke, P. Gordon, C. Huang, R. Rubenstein, Design of an artificial skin. [58] S.C. Mastbergen, D.B. Saris, F.P. Lafeber, Functional articular cartilage repair: here,
II. Control of chemical composition, J. Biomed. Mater. Res. 14 (2) (1980) 107–132. near, or is the best approach not yet clear? Nat. Rev. Rheumatol. 9 (5) (2013)
[35] Z. Shao, X. Zhang, Y. Pi, X. Wang, Z. Jia, J. Zhu, L. Dai, W. Chen, L. Yin, H. Chen, 277–290.
Polycaprolactone electrospun mesh conjugated with an MSC affinity peptide for [59] A. Mobasheri, The future of osteoarthritis therapeutics: emerging biological
MSC homing in vivo, Biomaterials 33 (12) (2012) 3375–3387. therapy, Curr. Rheumatol. Rep. 15 (12) (2013) 1–9.

12

You might also like