You are on page 1of 11

Colloids and Surfaces B: Biointerfaces 208 (2021) 112116

Contents lists available at ScienceDirect

Colloids and Surfaces B: Biointerfaces


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

Protocols

Nano-hybrid gradient scaffold for articular repair


Gan Xu a, b, Yao Zhao b, Yusheng Geng b, Shujun Cao a, Panpan Pan a, Jianhua Wang b,
Jingdi Chen a, *
a
Marine College, Shandong University, Weihai 264209, PR China
b
College of Biological Science and Engineering, Fuzhou University, Fuzhou 350002, PR China

A R T I C L E I N F O A B S T R A C T

Keywords: Osteoarthritis disease can easily lead to articular cartilage degeneration and subchondral bone damage, so the
Osteoarthritis disease demand for suitable articular substitutes is gradually increasing. In order to simulate the complex environment of
Continuous scaffold different layers in natural joint, we fabricate the continuous one-phase gradient scaffold. In the study, CS (chi­
Nano hydroxyapatite
tosan) was modified with SH (sodium hyaluronate) and GO (graphene oxide) to form the whole scaffold. nHAP
Articular repair
(Nano-hydroxyapatite) was in situ generated with gradient distribution in the scaffold. Continuous interface can
better imitate the combination style of cartilage and subchondral bone at joint. The diverseness of scaffold’s
different layer in water absorption/retention rate and mechanical property is similar to the difference of articular
cartilage and subchondral bone. Meanwhile, the cell experiments demonstrated that the bionic scaffold can well
promote the proliferation of bone marrow mesenchymal stem cell. Articular defect model further confirmed that
the scaffold can better induce articular regeneration. Herein, the prepared scaffold might be an excellent
candidate for endogenous articular repair.

1. Introduction bone induction ability[17–19]. It plays an important role in inducing


BMSCs to differentiate into osteoblasts, accelerating the formation of
Natural joints are constituted of cartilage and subchondral bone and subchondral bone, etc. [20–22]. GO has been shown to enhance the
connected by a dense calcified cartilage layer [1,2]. Scaffold with biomineralization ability of bone scaffolds and has a potential role in
calcified cartilage layer can play an active role in promoting the promoting the differentiation of BMSCs and bone formation [23–25].
regeneration and repair of cartilage tissue, which is related to the me­ GO modified chitosan can be used to accelerate biomineralization and
chanical properties and biological properties of different layers [3,4]. In bone formation [26,27]. For the lack of self-repair ability, the repair of
joint repair, designing a scaffold with continuous gradients in material cartilage layer relies on cells in subchondral bone migrating to the
composition, structure, mechanical properties, and biological properties cartilage layer to complete repair [28]. Cartilage materials simulating
is important [5]. the chondral environment are beneficial to the formation of cartilage
Joint injuries generally involve damage to cartilage and subchondral [29]. Hyaluronic acid is abundant in extracellular matrix (ECM) of
bone [6,7]. Cartilage is a highly specialized tissue that has no blood connective tissue[30]. For its high water retention, biocompatibility,
supply and is difficult to regenerate spontaneously once injured [8,9]. biodegradability and viscoelasticity, hyaluronic acid has good applica­
The lowest layer of cartilage is calcified cartilage and part of the signal tion prospects in cartilage repair [31,32]. Sodium hyaluronate modified
molecules from subchondral bone can pass through it [10–12]. The chitosan has large pore structure, good moisture retention and degra­
existence of calcified cartilage layer and the repair of subchondral bone dation [33–35]. In related reports, hyaluronic acid can enhance the
are conducive to the formation of cartilage [13,14]. In natural joints, the material’s adsorbability to chondrocytes and promote the proliferation
main inorganic component of subchondral bone and cartilage calcifi­ of chondrocytes [30,36]. Norimasa Iwasaki et al. [33] added hyaluronic
cation layer is nano-hydroxyapatite [15]. It presents a gradient distri­ acid to chitosan to form a fundamental material for cartilage regenera­
bution from calcified cartilage to subchondral bone and the content of tion and the cartilage analogue was get when cell was cultured in the
nHAP in subchondral bone is the highest [16]. nHAP has good scaffold.
biocompatibility, proper biodegradability, high bone conductivity and To prepare a continuous scaffold with cartilaginous layer, calcified

* Corresponding author.
E-mail address: jdchen@sdu.edu.cn (J. Chen).

https://doi.org/10.1016/j.colsurfb.2021.112116
Received 8 August 2021; Received in revised form 7 September 2021; Accepted 13 September 2021
Available online 16 September 2021
0927-7765/© 2021 Elsevier B.V. All rights reserved.
G. Xu et al. Colloids and Surfaces B: Biointerfaces 208 (2021) 112116

cartilaginous layer, and subchondral bone layer, we prepared a bone/ temperature for 4–6 h to obtain the precursor solution of subchondral
cartilage nHAP gradient scaffold using secondary freeze-drying method. bone layer.
When graphene oxide modified chitosan containing Ca2+ was added into
sodium hyaluronate modified chitosan, combining gravity and freezing 2.2.3. Preparation of integrated bone-cartilage scaffolds
technology, graphene oxide modified chitosan finally focused on the The obtained cartilage layer precursor solution was added to the
lower layer and sodium hyaluronic acid modified chitosan are mainly preformed mold. Then the solution of subchondral bone layer was
concentrated in the upper. The middle layer is the mixed solution of slowly added to the cartilage layer. The above liquid was refrigerated at
sodium hyaluronate modified chitosan and graphene oxide modified 4 ℃, − 20 ℃ and − 80 ℃. After being freeze-dried, the preliminary
chitosan. For the GO modified chitosan containing Ca2+ was mainly scaffold was soaked in the alkaline solution (2.5% NaOH), and then kept
concentrated in the lower layer and the strong adsorption of graphene in a constant temperature shaker for continuous reaction for 8 h. Later
oxide on Ca2+, when the freeze-dried scaffolds were immersed in lye, the scaffold was washed to neutral and freezed drying again. After that,
nHAP was generated in situ with gradient distribution. The porosity, an integrated scaffold with two kinds of structure was obtained. S-
water retention, mechanical properties, degradability and nHAP 0 (control), S-1, S-2, S-3 represent prepared scaffolds by adjusting SH
gradient of the prepared scaffold was tested. The experimental results content (0, 50, 100, 150 mg) in the experiment.
show that, compared with the lower layer, the upper layer has large pore
structure, high water retention, higher degradability and lower me­ 2.3. Physicochemical characterization of the scaffolds
chanical strength mechanical. SEM and EDS spectrum showed that
nHAP mainly generated in the middle and lower layer and presented The microscopic structure and elementary composition of scaffold
gradient distribution. In vitro mineralization experiments showed that were tested by scanning electron microscopy (Nova Nano SEM 230, FEI,
the lower layer of the prepared scaffold had good mineralization per­ USA) and energy-dispersive spectrometry (EDS). The crystalline phases
formance. Biocompatibility experiments showed that the scaffold can of the scaffolds were characterized by X′ Pert MPD Pro, PANalytical
promote the proliferation and differentiation of bone marrow mesen­ (Netherlands). Fourier transform infrared spectroscopy (FT-IR) (Nicolet
chymal stem cells, and the rabbit joint defect experiment further proved Nexus spectrometer) was used to study functional groups of organic
that the scaffold had a potential application prospect in promoting the components in scaffold. Renishaw invia microscope system (Invia Re­
repair of joint defects. flex) using laser source with wavelength of 532 nm detected Raman
spectrum of composite scaffold. X-ray photoelectron spectroscopy (XPS,
2. Materials and methods Manchester, England) were used to analyze binding energy state.

2.1. Materials 2.4. Porosity test

Chitosan (CS) (95% deacetylated) was purchased from Shanghai Liquid displacement method was used to assess the porosity of
BioScience & Technology Co., Ltd. Graphite powder was obtained from scaffolds [39]. The dry material weighing Wd was soaked in the test tube
Sigma-Aldrich (St. Louis, MO, USA). Sodium hyaluronate (SH) was containing enough ethanol and was soaked for 5 min. The test tube was
supplied by Sigma-Aldrich (St. Louis, MO, USA). The 1-ethyl-3-(3-dime­ placed in the glass desiccants under the vacuum state to allow ethanol
thylaminopropyl) (EDC) and N-Hydroxysuccinimide (NHS) were sup­ into the scaffold space. The excess ethanol was removed from the surface
ported by GL Biochem. (Shanghai) Co., Ltd. Calcium chloride and of the scaffold and weighed (Ww). 5 parallel samples were set in each
dipotassium hydrogen phosphate trihydrate were acquired from Sigma- group. Formula (1) was used to calculate the porosity of the cartilage
Aldrich (St. Louis, MO, USA). CCK-8 and AO/EB staining reagents were layer, bony layer and overall scaffold (P).
acquired from Nanjing Keygen Biotech. Co., Ltd. Fetal bovine serum
(FBS), Dulbecco’s minimum Eagle’s medium (DMEM), penicillin/ P = (Ww− Wd)/ρ×π×h× (D⁄2)2 × 100% (1)
streptomycin, trypsin and other materials for cell experiments were
ρ represents the density of ethanol (0.789 g/cm3). π is 3.14. D is the
purchased from HyClone Co., Ltd (USA).
scaffold diameter and h is the height.
The surface density (A) is calculated by formula:
2.2. Fabrication of SH/GO/CS/nHAP scaffolds
A =Wd/π×h× (D⁄2)2× 100% (2)
2.2.1. Preparation of upper layer of cartilage
π is 3.14. D is the scaffold diameter and h was the height.
The details are as follows: SH (Sodium hyaluronate) (0, 50, 100, 150
mg) was added into 10 mL deionized water and stirred until fully dis­
solved. Chitosan solution was obtained by dissolving chitosan powder in 2.5. Swelling ratio measurement
2% acetic acid liquid. Under the condition of magnetic stirring at 37 ℃,
prepared solutions above were mixed, and the crosslinking agent EDC/ The dried scaffold (W1) was immersed in 15 mL PBS at 37 ℃ and
NHS (molar ratio 2:1) was added to form a uniform mixture. After cross- then taken out at the time points of 12, 24, 36, 48 and 60 h respectively.
linking at room temperature for 4–6 h, the precursor solution of cartilage The excess PBS liquid was removed with filter paper and the weight of
layer was obtained. the scaffold was measured (W2). Each group had 5 parallel samples. The
swelling ratio (S) was calculated by formula (3):
2.2.2. Preparation of the subchondral bone layer S = (W2− W1)/W1× 100% (3)
The GO was prepared by the improved Hummer’s method [37]. The
details are provided in the Supplementary Data. The CS/GO composite
was fabricated by published method [38]. Chitosan solution was ob­
tained by dissolving CS powder in 2% acetic acid liquid. GO was ultra­ 2.6. Water uptake and retention ratio measurement
sound in 10 mL deionized water to form a uniform GO dispersion. Under
the condition of magnetic stirring, GO was slowly added into the chi­ Water uptake and retention ratio measurement followed the process:
tosan solution and then fully stirred to form a uniform mixture. Subse­ after drying, the scaffolds weighing Wd with different components were
quently, at the ratio of 1.67, calcium and phosphorus salt solutions were immersed in distilled water for 24 h. The residual water on the surface of
added to the above mixture and stirred thoroughly to make it entirely the material was removed through dry filter paper, and then weighed
mixed. Then, cross-linking agent was added and cross-linked at room again (Ww1) to determine the water absorption rate. The soaked

2
G. Xu et al. Colloids and Surfaces B: Biointerfaces 208 (2021) 112116

Fig. 1. (A) The fabrication process of SH/GO/CS/nHAP biphysic scaffold and the chemical construction of upper section, transition interface and lower section. (B)
EDS analysis of scaffold’s element distribution.

scaffold was centrifuged at 500 rpm for 3 min, weighed (Ww2), and the and 5 parallel samples were set in each group. The degradation rate (D)
water retention rate was calculated. The water absorption rate WA and was calculated by formula (6):
water retention rate WR of the support are calculated by the formulas (4)
and (5) respectively: Weight loss ratio = (W1-W2)/W1×100% (6)

WA = (Ww1− Wd)/Wd× 100% (4)

WR = (Ww2− Wd)/Wd× 100% (5) 2.9. Biomineralization study in vitro

The scaffolds were immersed in a centrifuge tube containing 15 mL


2.7. Analysis of mechanical properties SBF and incubated in oscillator at 37 ◦ C for 1 and 3 days. The biomimetic
mineralization of scaffolds was analyzed by SEM and XRD. pH value of
A universal material testing machine (INSTRON Model 1185, Shi­ SBF liquid was detected by pH meter (S20K, Mettler-Toledo,
madzu) was used to detect mechanical properties of scaffolds at upper, Switzerland) for 5 weeks. At the same time, the weight gain rate of
lower layers and the whole two-phase at a speed of 1.0 mm /min. The each sample was calculated. Each group of scaffolds were removed from
change of compressive stress was recorded, and the elastic modulus of SBF solution at 1, 3, 5, 7, 14, 21 and 28 days respectively and rinsed with
the scaffold was calculated by 30% deformation. Five parallel samples ultra-pure water for 3–5 times. Inductively coupled plasma-Atomic
were used in the measurement. Emission spectrometer, Liberty 200 (*/iCAP7400, USA), was used to
determine the content of calcium ions in immersion solution. The par­
2.8. In vitro degradation allel samples were set to 5.

The weighing W1 sample was immersed in phosphate buffer solution 2.10. Cell viability and proliferation assay
(PBS) containing lysozyme (0.5 mg/mL) to test the degradation
behavior of the composite scaffold. PBS is replaced every three days. At AO and EB reagent dye were used in Live & Dead staining. The
1, 4, 7, 14 and 28 days, the scaffolds were taken out and weighed (W2), proliferation of cell in scaffolds was investigated by Cell Counting Kit-8

3
G. Xu et al. Colloids and Surfaces B: Biointerfaces 208 (2021) 112116

Fig. 2. (A) Macro view of prepared scaffold and (B) SEM of cartilage layer (a-c) and subchondral bone layer (d-f) at magnification of 200 × , 1000 × and 10,000 × .
(C) The diagram of pore distribution. (D) Porosity analysis of different component scaffolds and (E) porosity measurement of S-2 scaffold’s different section. (F)
Apparent density of S-2 scaffold’s different section. (Statistic level of P*< 0.05, p** < 0.01 and p*** < 0.001).

(CCK-8) assay. Bone marrow mesenchyml stem cells were extracted and then slowly poured into the cartilage layer to form a one-phase
from the bone marrow of SD rat’s tibia (one month old). The detailed integration structure by gravity precipitation. After soaking in lye, it
information is included in the Supplementary Data. can be seen that nHAP was generated on freeze-dried scaffold (Fig. S2C).
The distribution of main elements of the scaffold was further studied.
2.11. Rabbit articular bone-cartilage model Through analysis of Fig. 1B, the distribution of element P and Ca has an
increasing gradient from top to bottom, while the distribution of
18 healthy New Zealand Rabbit weighing 2.1 kg were bought from element C and O is uniform. The element distribution fully confirmed
the Animal Experiment Center and were divided into 3 groups. In the that the distribution of nHAP is similar to the distribution of normal
knee patellar, defect model (Φ5mm and 4 mm high) was built. 8, 12 and human bone-cartilage inorganic components [1,40,41]. It can also be
16 weeks were as three stages to evaluate the repair degree and the observed from XPS test (Fig.S2.D).
relevant surgery procedure is described in the Fig. S3. CT was used to
evaluate bone regeneration and 3d visual modeling directly showed the 3.2. Analysis of porosity and surface density
regeneration situation. Surgical steps and histological assessments were
descripted in the Supplementary Data. All the experiments meet the From the optical image in Fig. 2A, the composite scaffold has two-
local ethical committee and laboratory animal administration rules of phase structure. Fig. 2B(a–c) is the SEM image of the cartilage layer of
China. scaffold. It can also be found that there was denser pore structure on the
subchondral bone layer (Fig. 2B (d–f)). The further enlarged image of
2.12. Statistical analysis the surface shows that the surface of the scaffold is relatively rough,
which is beneficial to cell adhesion, proliferation and differentiation in
Independent Student’s t-test was adopted to detect the difference of scaffolds [42,43]. At the same time, large number of nHAP was found on
different groups. The numerical values represent the mean ± standard the surface of the subchondral bone layer at high magnification, indi­
deviation. The different statistic level of P*< 0.05, p** < 0.01 and p*** cating that bone-like apatite had been successfully prepared by in-situ
< 0.001 was considered. Three replicates were set at least in each group. process. The distribution of nHAP in the cross-section was relatively
uniform and the amount of nHAP on lower layer was more than the
3. Result and discussion upper layer. The pore structure of natural bone-cartilage showed the
difference of large pore in cartilage layer and small pore in bone layer
3.1. Fabrication mechanism of the scaffolds [5]. After observation, it can also be found that the pore size of the upper
layer (40–200 µm) is larger than that of the lower layer (30–100 µm),
Fig. 1A shows the biomimetic construction process of scaffolds and and the pore wall of the upper layer is relatively thin. The difference in
its component of different layers. Firstly, amine groups (–NH2) of CS and pore structure has great influence on water absorption and mechanical
negative carboxyl radical (–COOH) of GO (Fig. S2A) and SH formed the properties of different layers [44]. The large and loose pore structure
amide linkage (–NHCO–) (Fig. S2) under the crosslinking of EDC and indicates that the scaffold has relatively low strength and high water
NHS to get different layers solution. Secondly, the precursor ions absorption. Besides, the pore structure of different layers determines the
including Ca2+ and PO43- were added into the subchondral bone layer heat conduction and material transfer efficiency [41]. As shown in

4
G. Xu et al. Colloids and Surfaces B: Biointerfaces 208 (2021) 112116

Fig. 3. (A) Hydrophilic diagram of scaffold’s different part in aqueous solution (B) Swelling ratio of different component scaffolds and (C) swelling ratio of S-2
scaffold’s different section. (D) Water uptake and (E) water retention ratio of scaffolds immersed in distilled water. (Statistic level of P*< 0.05, p** < 0.01 and p***
< 0.001).

Fig. 2D, the porosity of S-2 scaffold was higher than others, reaching bone layer (74%), and the overall porosity is between the upper carti­
83%, while the control group approaches 60%. Therefore, comparing lage layer and subchondral bone layer. In addition, the surface density
with other groups, S-2 scaffold can provide a good physiological space and porosity were opposite (Fig. 2F). As a whole, the prepared scaffolds
environment for bone cell infiltration, which is conducive to the adhe­ have different porosity and pore size in different layers, which is similar
sion and proliferation of cells. However, because of the independence to the physiological structure of natural joint [45].
and specialty of two layer structure, the experiment further investigated
the optimal S-2 scaffold. It can be seen from Fig. 2E that the porosity of
upper cartilage layer (96%) was much higher than that of subchondral

Fig. 4. (A) Load-distance curves, (B) compressive strength and elastic modulus of different component scaffolds. (C) Compressive strength and elastic modulus of S-2
scaffold’s different section. (D) Degradation in vitro of different scaffolds immersed in PBS containing lysozyme and (E) degradability of S-2 scaffold’s different
section. (Statistic level of P*< 0.05, p** < 0.01 and p*** < 0.001).

5
G. Xu et al. Colloids and Surfaces B: Biointerfaces 208 (2021) 112116

Fig. 5. (A) SEM images of control (a, c) and S-2 (b, d) scaffold after mineralization for 1 day and 3 days and element distribution analysis in S-2 scaffold with 3 days
mineralization. (B) XRD analysis of control and S-2 scaffolds immersed in SBF for 1 & 3 days. (C) Mass change of S-2 and control groups during biomineralization
process of 4 weeks. (D) pH value changes of SBF in S-2 and control groups during 5 weeks. (E) Analysis of compressive strength and elastic modulus of S-2 and control
group after mineralizing for 1 & 3 days. (F) The Ca2+ content of SBF in S-2 and control group for 1 & 3 days analyzed with ICP-OES. (Statistic level of P*< 0.05, p** <
0.01 and p*** < 0.001).

3.3. Swelling ratio, water uptake and water retention measurement 3.4. Mechanical property and degradation analysis

Low swelling rate is considered to be one of the factors of a good As a tissue replacement material, the scaffold should have enough
scaffold. Fig. 3A displays the hydrophilia of different layers. As shown in mechanical properties to support the defect and its surroundings orga­
Fig. 3B, the variation of swelling rate of S-2 scaffold is relatively small nization. The good mechanical properties of the subchondral bone layer
compared with other scaffolds, and the inflation equilibrium is relatively can support the defect and the high water content and low strength of
long. Mainly due to the appropriate addition of SH, the effective cross­ the cartilage layer can reduce the damage caused by wear [47]. Fig. 4A
linking points of SH/GO/CS/nHAP scaffold increased. In swelling pro­ shows the load-distance curve of the composite scaffold. It can be seen
cess, complete network structure is not easy to produce great changes that the elastic deformation of the scaffold has a certain range, mainly
and the absorbable amount of liquid is limited. In view of the relatively occurring in the range of 0–20%. Since the scaffold still has certain
stable swelling rate of S-2 scaffold, each part was studied independently bearing capacity at 30% deformation, the compressive strength of the
in the experiment. As shown in Fig. 3C, the upper layer of the scaffold scaffold at 30% deformation was experimentally studied (Fig. 4B). It can
showed a large swelling rate, while the lower layer changed small. The be found that the S-2 scaffold has the highest compressive strength
difference of swelling rate indicates that the scaffold can better fit (0.28 ± 0.024mpa) and elastic modulus (0.272 ± 0.213mpa). When
osteochondral environment. Good water retention rate can not only appropriate SH was added, effective crosslink point was formed between
effectively avoid the accumulation of secretions, but also prevent tissue SH and CS, which enhanced the mechanical properties of the substrate
wound dehydration [46]. The cartilage layer constructed in the study framework. Meanwhile, due to the effective doping of nHAP, the me­
was not significantly different from commonly used cartilage materials. chanical properties of the scaffold were greatly improved. According to
But compared with the commonly used bone repair materials, the water the analysis in Fig. 4C, the mechanical properties of the lower layer are
retention of the cartilage layer was significantly enhanced. When two superior to those of the whole and the upper part, with the highest
layers were constructed through one-step process, there was no obvious compressive strength (0.38 ± 0.021Mpa) and elastic modulus
difference in water retention with common bone and cartilage material, (0.369 ± 0.030Mpa), which is similar to the mechanical difference of
indicating that the scaffolds had good water retention. Fig. 3D shows bone /cartilage. Meanwhile, implanted tissue replacement material
that, with the increase of SH content, the water absorption rate of the should be biodegradable and satisfy the needs of different tissue. The
scaffold also increases relatively. In particular, after being modified by degradation curve of the scaffold in PBS solution containing lysozyme
SH, the scaffold’s water absorption rate was higher than that of the was shown in Fig. 4D. It can be seen that S-2 scaffolds have a relatively
control group. At the same time, through further analysis of the water fast degradation rate of 11.5% in about 14 days. However, the degra­
retention rate, as shown in Fig. 3E, it was found that the water retention dation rates of the control group and S-3 changed significantly within 7
rate of the scaffold modified with SH was also higher than that of the days, with the degradation rates of 12.1% and 15.1%, respectively. The
scaffold without SH modification. In particular, the water retention rate degradation rates of the S-1 group changed significantly in different
of S-2 group was the highest, reaching about 15%. The results show that periods. Therefore, compared with other scaffolds, the degradation rate
SH is beneficial to improve the absorbency and water retention of the of S-2 scaffolds is relatively low and stable in the same time. To analyze
scaffold materials. The reason may be due to the high hydrophilicity of each part and the overall degradation rate, as shown in Fig. 4E, it was
SH molecule. To sum up, the S-2 scaffold has a higher water retention found that, at the time of day 1, the degradation rate of the upper part is
performance and is more suitable for in vivo application. larger and the overall and lower layer at the time of 14 days to appear
relatively large degradation rate, which illustrated that the degradation

6
G. Xu et al. Colloids and Surfaces B: Biointerfaces 208 (2021) 112116

Fig. 6. (A) Live & Dead staining (a-j) for the proliferation of rBMSCs cells cultured with extract liquid of scaffolds for 3 and 5 days. Live cells are stained in green and
dead cells in orange with AO/EB staining. (B) The quantitative analysis by CCK-8 assay for the proliferation of rBMSCs cells cultured with extract liquid of scaffolds
for different times. (Statistic level of P*< 0.05, p** < 0.01 and p*** < 0.001). (For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

of the scaffold in different parts exist certain differences. The relatively found that the peak intensity of XRD increased with culture time
slow degradation process in lower layer is advantageous to the growing increasing. However, at 3 days, 222 and 322 crystal plane peaks
and maturing of bone [48]. appeared in the S-2 group, but not in the control group, indicating that
SH had some influence on the growth direction of nHAP, making it grow
3.5. Biomineralization in vitro mainly along the orientation of 222 and 322 crystal planes. In addition,
due to the formation of nHAP, the scaffold quality also changed
As shown in Fig. 5A, the lower layer of S-2 scaffold was immersed in accordingly. As shown in Fig. 5C, comparing with control group, the
SBF liquid at 37 ℃ for 1 and 3 days to study the mineralization of nHAP growth of nHAP in S-2 group has a continuous and slow process, which is
on the surface of scaffolds. According to the SEM figure after minerali­ consistent with the Ca content test results (Fig. 5F). In addition, the mass
zation, the content of nHAP on the surface of the scaffold increased with increase rate of group S-2 reached 36% at 21 days, indicating that there
the extension of incubation time. After 3 days of incubation, the distri­ was large amount of nHAP formation on its surface. As shown in Fig. 5D,
bution of nHAP in the two groups increased significantly (Fig. 5A (c, d)). pH values of each group showed an upward trend in the first 7 days,
It is worth noting that, compared with control group, the nHAP particle mainly caused by residual lye on the synthetic scaffold. Alkaline envi­
size on the surface of S-2 scaffold was smaller during the same miner­ ronment is the main condition for in-situ synthesis of nHAP [50].
alization time and a complete and uniform apatite layer appeared. EDS Therefore, the increase of pH in mineralization solution is conducive to
energy spectrum was used to analyze the surface elements of S-2 promoting the regrowth of bone-like apatite, which is consistent with
mineralization scaffold, and found that the main elements were Ca and P the results observed in SEM images after mineralization. After soaking
(Fig. 5A). Meanwhile, the coating structure can induce the formation of for 7 days, the pH value of the control group gradually decreased and
a large number of nHAP in physiological environment and improve the reached acidity at 35 days, which indicated that the scaffolds gradually
biological activity of the material. XRD was used to analyze nHAP for­ degraded. However, the pH value of the S-2 group reached the highest at
mation on the surface of scaffolds in control and S-2 groups after 18 days, indicating that the environment of the S-2 group was conducive
mineralization. As shown in Fig. 5B, by comparing the standard cards, it to the massive generation of nHAP on the scaffold surface. Therefore, in
was found that 2θ values respectively corresponded to (002), (102), general, the environment made by S-2 scaffold has a relatively high pH
(211), (310), (222) and (322) crystal planes of nHAP [49]. It was also value at the same time point and can continuously promote the synthesis

7
G. Xu et al. Colloids and Surfaces B: Biointerfaces 208 (2021) 112116

Fig. 7. (A) CT analysis of articular osteochondral repairs after 8, 12 and 16 weeks postimplantation. (B) Defect repairing images of articular osteochondral after 8, 12
and 16 weeks postimplantation.

of nHAP in simulated body fluids. At the same time, the mechanical basically covered the whole interface. Although there existed orange
properties of the two scaffolds were evaluated after soaking in SBF liquid spots, but the number was very small, indicating that the scaffold had
for 1 and 3 days. As shown in Fig. 5E, the mechanical properties of the good cell compatibility and non-toxicity. At the same time, the quanti­
scaffolds were time-dependent after soaking in the mineralized solution. tative proliferation of rBMSCs cell co-cultured with bone-cartilage
The compressive strength and elastic modulus of each scaffold increased scaffolds was also studied. As shown in Fig. 6B, it is worth noting that
with the increase of time. Especially for the S-2 group, the compressive OD values of all groups were significantly higher than the minimum
strength (4.5mpa) and elastic modulus (3.9mpa) were about 3 times and standard of non-toxic. With time prolongation, the corresponding OD
2 times higher than those of the control group at 3 days. At the same value also increased. Comparing with control group and blank group,
time, due to the production of nHAP, the content of Ca2+ in the the proliferation of cells in the experimental group was not obvious on
mineralization liquid changed accordingly (Fig. 5F). The content of Ca2+ the first day. At the 7th day, the OD values of each group increased
in the solution showed a decreasing trend with the extension of time. At significantly and the OD values of the S-1, S-2 and S-3 groups were
1 day, the content of Ca2+ in the SBF solution in S-2 group was significantly higher than those of other groups, indicating that the
88.6 mg/kg which was higher than that in the control group addition of SH improved the cell proliferation ability. Therefore, CCK-8
(44.4 mg/kg), indicating that the amount of nHAP on the surface of the experimental results showed that all two-phase integrated scaffolds had
control group was relatively high. At 3 days after mineralization, the good cell compatibility. The survival rate of S-2 component scaffold cells
content of calcium ions in SBF solution in the S-2 group was 63.6 mg/kg, was the highest.
while the control group was still 42.2 mg/kg. The stable and slow for­
mation of nHAP is beneficial to its growth. The size of nHAP was rela­ 3.7. In vivo bone repair of calvarial defects
tively small, which could be clearly identified by SEM images. In vivo
environment, because of the slow growth of nHAP, it is not easy to form As can be seen from Fig. 7B, compared with initial defect area, the
local aggregation. When the scaffold is implanted in vivo, the physio­ tissues of the three groups were all repaired and the surface was filled
logical microenvironment will change dynamically due to the with soft tissue. The depression degree of the control group in the defect
biomineralization. area slowed down and the sag of the S-2 group was further reduced, but
the defect area was not completely filled. At 12 weeks, the experimental
3.6. . In vitro cellular assays group was basically completely repaired. The surface was relatively
rough and only the new tissue and boundary circle existed, indicating
In order to study the growth conditions of bone marrow mesen­ that the fusion of the new tissue and the interface still needed to further
chymal stem cells in the contact with bone-cartilage scaffold. The rBMSC matured. As can be seen from the 16-week repair diagram, the defect
cell was cultured in the bone-cartilage scaffold leaching solution for 3 area of the control and experimental group has been completely covered
and 5 days to study the cell growth state through dyeing. As shown in by the new tissue. Compared with the 12-week repair effect, the surface
figure Fig. 6A, AO/EB staining after 3 days in inverted fluorescence of the experimental group has a good fusion between the new tissue and
microscope showed that the cells in each scaffold grow with good con­ the host boundary. The contour line has basically disappeared and the
dition and visible spindle shape can easily be found. At the same time, surface is relatively smooth and flat. At the same time, the repair process
the number of cells increased with the extension of incubation time in of articular bone-cartilage defects in rabbits was tested by CT. The tissue
the scaffold extracting solution. After 5 days of culture, the cells repair was analyzed by microscopic reconstruction at 8, 12 and 16

8
G. Xu et al. Colloids and Surfaces B: Biointerfaces 208 (2021) 112116

Fig. 8. Histological analysis of defect area after 12 weeks (A) and 16 weeks (B) implantation. Blank defect group (a, d, g, j); CS/nHAP group (b, e, h, k); S-2 SH/GO/
CS/nHAP group(c, f, i, l). The yellow arrows are osteocytes and the red arrows are new blood vessels. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)

weeks, respectively. As shown in Fig. 7A, through three-dimensional physiological requirements of articular cartilage and subchondral bone,
reconstruction and comprehensive sagittal analysis, it was found that and effectively promote the growth and maturation of new tissue.
the repair process of each group was enhanced with the passage of time,
and the repair effect of the S-2 group was better than that of the control 3.8. HE staining analysis
group. According to the 3D reconstruction, compared with other groups,
the articular cartilage defects in the 16-week S-2 repair group had been H&E staining was used to analyze the formation process of cartilage/
completely repaired. The contour lines were basically disappeared and bone. The regenerated tissue could be identified by the morphology in
the surface was smooth and flat, indicating the rapid growth of new defect area. The blank group still had loose structure after 12 weeks
tissue. According to the 2d-x sectional analysis, it was found that the implantation (Fig. 8A (a, d)). Only a few osteocytes grow and bone
central defect had been filled with new tissue. The background color and matrix formation was not obvious. Additionally, in the control group,
bone density were consistent with the surrounding bone. The 2d-y sec­ osteocyte increased significantly and calcium accumulation appeared. In
tion also showed that the upper part of the defect was completely S-2 group (Fig. 8A (c, f)), calcium accumulated significantly and blood
overgrown with newborn tissue, which was close to the height of the vessels began to form. After 16 weeks implantation, mature collagen
host on both sides. Further 2d-z examination revealed that the border could be observed in S-2 group and much more blood vessel formed in
cortical bone was also largely closed and intact, with no fractures or regenerated bony layer (Fig. 8B (i, l)). The collagen fiber density and
fractures. The results of CT analysis showed that the prepared integrated calcium accumulation in blank (Fig. 8B (g, j)) and control (Fig. 8B (h, k))
scaffold could effectively repair the tissue defect after implantation in group was significantly lower than the S-2 group. The results demon­
the defect area for a certain period of time, realize the bidirectional strated that S-2 scaffold can better promote the regeneration of joint.

9
G. Xu et al. Colloids and Surfaces B: Biointerfaces 208 (2021) 112116

4. Conclusion [9] M. Liu, X. Zeng, C. Ma, H. Yi, Z. Ali, X. Mou, S. Li, Y. Deng, N. He, Injectable
hydrogels for cartilage and bone tissue engineering, Bone Res. 5 (2017)
17014–17033.
For the special physiological structure of joint, the experiment con­ [10] K.P. Arkill, C.P. Winlove, Solute transport in the deep and calcified zones of
structed a continuous gradient scaffold. The scaffold had different pore articular cartilage, Osteoarthr. Cartil. 16 (2008) 708–714.
size and porosity on different layers. And EDS spectrum showed that [11] H.X. Cai, P.L. Wang, Y. Xu, Y. Yao, J. Liu, T. Li, Y. Sun, J. Liang, Y.J. Fan, X.
D. Zhang, BMSCs-assisted injectable Col I hydrogel-regenerated cartilage defect by
nHAP generated with gradient distribution. Other relevant tests also reconstructing superficial and calcified cartilage, Regen. Biomater. 7 (2020) 35–46.
showed that the scaffold largely imitated the environment of different [12] W.D. Lee, M.B. Hurtig, R.M. Pilliar, W.L. Stanford, R.A. Kandel, Engineering of
layers in joint. In cell co-culture system, the bone marrow mesenchymal hyaline cartilage with a calcified zone using bone marrow stromal cells,
Osteoarthr. Cartil. 23 (2015) 1307–1315.
stem cells showed high proliferation vigor. In rabbit defect experiment, [13] R. Longley, A.M. Ferreira, Recent approaches to the manufacturing of biomimetic
CT three-dimensional reconstruction and comprehensive sagittal anal­ multi-phasic scaffolds for osteochondral regeneration, Int. J. Mol. Sci. 19 (2018)
ysis of each group showed that the tissue regeneration increased over 1755–1771.
[14] C.H. Lee, J.L. Cook, A. Mendelson, E.K. Moioli, H. Yao, J.J. Mao, Regeneration of
time and S-2 group was the best. HE staining further confirmed that S-2 the articular surface of the rabbit synovial joint by cell homing: a proof of concept
scaffold can better promote joint repair. To sum up, the continuous study, Lancet 376 (2010) 440–448.
scaffold can better imitate the functions of different layers comparing [15] Y.Z. Zhang, J.R. Venugopal, A. El-Turki, S. Ramakrishna, B. Su, C.T. Lim,
Electrospun biomimetic nanocomposite nanofibers of hydroxyapatite/chitosan for
with traditional scaffold and is expected to be a proper candidate for bone tissue engineering, Biomaterials 29 (2008) 4314–4322.
joint. [16] A. Di Luca, C. Van Blitterswijk, L. Moroni, The Osteochondral Interface as a
Gradient Tissue: From Development to the Fabrication of Gradient Scaffolds for,
Regener. Med. Birth Defects Res. Part C-Embryo Today-Rev. 105 (2015) 34–52.
CRediT authorship contribution statement [17] J. Chen, P. Pan, Y. Zhang, S. Zhong, Q. Zhang, Preparation of chitosan/nano
hydroxyapatite organic-inorganic hybrid microspheres for bone repair, Colloids
Jingdi Chen conceived the project, and Gan Xu developed the Surf. B Biointerfaces 134 (2015) 401–407.
[18] Y.L. Li, C.S. Liu, Nanomaterial-based bone regeneration, Nanoscale 9 (2017)
experimental plan, performed and evaluated the in vitro experiments. 4862–4874.
Jingdi Chen gave the guidance on animal surgeries in vivo. Gan Xu and [19] G. Campi, F. Cristofaro, G. Pani, M. Fratini, B. Pascucci, P.A. Corsetto,
Yao Zhao planned and established animals’ protocols. Yusheng Geng, B. Weinhausen, A. Cedola, A.M. Rizzo, L. Visai, G. Rea, Heterogeneous and self-
organizing mineralization of bone matrix promoted by hydroxyapatite
Shujun Cao and Jianhua Wang helped with the operation of animal
nanoparticles, Nanoscale 9 (2017) 17274–17283.
surgeries. Gan Xu recorded and analyzed all data of the in vitro and vivo [20] Z. Li, W. Mi, H. Wang, Y. Su, C. He, Nano-hydroxyapatite/polyacrylamide
experiments. In addition, Gan Xu drafted the manuscript, Yao Zhao and composite hydrogels with high mechanical strengths and cell adhesion properties,
Colloids Surf. B Biointerfaces 123 (2014) 959–964.
Panpan Pan helped to revise the manuscript, and all authors contributed
[21] R. Ying, H. Wang, R. Sun, K. Chen, Preparation and properties of a highly dispersed
in discussion of the results. nano-hydroxyapatite colloid used as a reinforcing filler for chitosan, Mater. Sci.
Eng. C Mater. Biol. Appl. 110 (2020) 110689–110696.
[22] C. Koski, A.A. Vu, S. Bose, Effects of chitosan-loaded hydroxyapatite on osteoblasts
Declaration of Competing Interest and osteosarcoma for chemopreventative applications, Mater. Sci. Eng. C Mater.
Biol. Appl. 115 (2020) 111041–111049.
[23] Y.W. Zhu, S. Murali, W.W. Cai, X.S. Li, J.W. Suk, J.R. Potts, R.S. Ruoff, Graphene
The authors declare that they have no known competing financial and graphene oxide: synthesis, properties, and applications, Adv. Mater. 22 (2010)
interests or personal relationships that could have appeared to influence 3906–3924.
the work reported in this paper. [24] P. Yu, R.Y. Bao, X.J. Shi, W. Yang, M.B. Yang, Self-assembled high-strength
hydroxyapatite/graphene oxide/chitosan composite hydrogel for bone tissue
engineering, Carbohydr. Polym. 155 (2017) 507–515.
Acknowledgement [25] G. Wei, C. Gong, K. Hu, Y. Wang, Y. Zhang, Biomimetic hydroxyapatite on
graphene supports for biomedical applications: a review, Nanomaterials 9 (2019)
1435–1454.
The research was supported by National Key Research and Devel­ [26] Y. Zhao, J. Chen, L. Zou, G. Xu, Y. Geng, Facile one-step bioinspired mineralization
opment Program of China (2017YFC1104102), National Natural Science by chitosan functionalized with graphene oxide to activate bone endogenous
Foundation of China (31370958) and Key Program of Natural Science regeneration, Chem. Eng. J. 378 (2019) 122174–122187.
[27] M. Li, Y.B. Wang, Q. Liu, Q.H. Li, Y. Cheng, Y.F. Zheng, T.F. Xi, S.C. Wei, In situ
Foundation of Fujian Province (2018Y0056). synthesis and biocompatibility of nano hydroxyapatite on pristine and chitosan
functionalized graphene oxide, J. Mater. Chem. B 1 (2013) 475–484.
[28] A.R. Poole, T. Kojima, T. Yasuda, F. Mwale, M. Kobayashi, S. Laverty, Composition
Appendix A. Supporting information
and structure of articular cartilage - a template for tissue repair, Clin. Orthop.
Relat. Res. (2001) S26–S33.
Supplementary data associated with this article can be found in the [29] Y. Huang, D. Seitz, F. Konig, P.E. Muller, V. Jansson, R.M. Klar, Induction of
articular chondrogenesis by chitosan/hyaluronic-acid-based biomimetic matrices
online version at doi:10.1016/j.colsurfb.2021.112116.
using human adipose-derived stem cells, Int. J. Mol. Sci. 20 (2019) 4487–4512.
[30] H. Tan, C.R. Chu, K.A. Payne, K.G. Marra, Injectable in situ forming biodegradable
References chitosan-hyaluronic acid based hydrogels for cartilage tissue engineering,
Biomaterials 30 (2009) 2499–2506.
[31] S. Yamane, N. Iwasaki, T. Majima, T. Funakoshi, T. Masuko, K. Harada, A. Minami,
[1] H. Zhang, H. Huang, G. Hao, Y. Zhang, H. Ding, Z. Fan, L. Sun, 3D printing
K. Monde, S. Nishimura, Feasibility of chitosan-based hyaluronic acid hybrid
hydrogel scaffolds with nanohydroxyapatite gradient to effectively repair
biomaterial for a novel scaffold in cartilage tissue engineering, Biomaterials 26
osteochondral defects in rats, Adv. Funct. Mater. 31 (2020) 2006697–2006709.
(2005) 611–619.
[2] C.D. Hoemann, C.H. Lafantaisie-Favreau, V. Lascau-Coman, G.P. Chen, J. Guzman-
[32] H. Park, K.Y. Lee, Cartilage regeneration using biodegradable oxidized alginate/
Morales, The cartilage-bone interface, J. Knee Surg. 25 (2012) 85–97.
hyaluronate hydrogels, J. Biomed. Mater. Res. Part A 102 (2014) 4519–4525.
[3] Z. Liu, M.A. Meyers, Z. Zhang, R.O. Ritchie, Functional gradients and
[33] N. Iwasaki, Y. Kasahara, S. Yamane, T. Igarashi, A. Minami, Chitosan-based
heterogeneities in biological materials: design principles, functions, and
hyaluronic acid hybrid polymer fibers as a scaffold biomaterial for cartilage tissue
bioinspired applications, Prog. Mater. Sci. 88 (2017) 467–498.
engineering, Polymers 3 (2010) 100–113.
[4] B.A. Harley, A.K. Lynn, Z. Wissner-Gross, W. Bonfield, I.V. Yannas, L.J. Gibson,
[34] T. Kutlusoy, B. Oktay, N.K. Apohan, M. Süleymanoğlu, S.E. Kuruca, Chitosan-co-
Design of a multiphase osteochondral scaffold III: fabrication of layered scaffolds
Hyaluronic acid porous cryogels and their application in tissue engineering, Int. J.
with continuous interfaces, J. Biomed. Mater. Res. A 92 (2010) 1078–1093.
Biol. Macromol. 103 (2017) 366–378.
[5] J.M. Oliveira, V.P. Ribeiro, R.L. Reis, Advances on gradient scaffolds for
[35] N. Mohan, P.V. Mohanan, A. Sabareeswaran, P. Nair, Chitosan-hyaluronic acid
osteochondral tissue engineering, Prog. Biomed. Eng. 3 (2021) 033001–033022.
hydrogel for cartilage repair, Int. J. Biol. Macromol. 104 (2017) 1936–1945.
[6] C. Deng, J. Chang, C. Wu, Bioactive scaffolds for osteochondral regeneration,
[36] A. Di Martino, M. Sittinger, M.V. Risbud, Chitosan: a versatile biopolymer for
J. Orthop. Translat. 17 (2019) 15–25.
orthopaedic tissue-engineering, Biomaterials 26 (2005) 5983–5990.
[7] D. Yang, J. Xiao, B. Wang, L. Li, X. Kong, J. Liao, The immune reaction and
[37] D.C. Marcano, D.V. Kosynkin, J.M. Berlin, A. Sinitskii, Z.Z. Sun, A. Slesarev, L.
degradation fate of scaffold in cartilage/bone tissue engineering, Mater. Sci. Eng. C
B. Alemany, W. Lu, J.M. Tour, Improved synthesis of graphene oxide, ACS Nano 4
Mater. Biol. Appl. 104 (2019) 109927–109943.
(2010) 4806–4814.
[8] R.A. Muzzarelli, F. Greco, A. Busilacchi, V. Sollazzo, A. Gigante, Chitosan,
[38] D. Depan, B. Girase, J.S. Shah, R.D. Misra, Structure-process-property relationship
hyaluronan and chondroitin sulfate in tissue engineering for cartilage regeneration:
of the polar graphene oxide-mediated cellular response and stimulated growth of
a review, Carbohydr. Polym. 89 (2012) 723–739.

10
G. Xu et al. Colloids and Surfaces B: Biointerfaces 208 (2021) 112116

osteoblasts on hybrid chitosan network structure nanocomposite scaffolds, Acta [45] W.M. Groen, P. Diloksumpan, P.R. van Weeren, R. Levato, J. Malda, From intricate
Biomater. 7 (2011) 3432–3445. to integrated: biofabrication of articulating joints, J. Orthop. Res. 35 (2017)
[39] S. Liu, C. Zhou, S. Mou, J. Li, M. Zhou, Y. Zeng, C. Luo, J. Sun, Z. Wang, W. Xu, 2089–2097.
Biocompatible graphene oxide-collagen composite aerogel for enhanced stiffness [46] J. Ruan, X.S. Wang, Z. Yu, Z. Wang, Q. Xie, D.D. Zhang, Y.Z. Huang, H.F. Zhou, X.
and in situ bone regeneration, Mater. Sci. Eng. C Mater. Biol. Appl. 105 (2019) P. Bi, C.W. Xiao, P. Gu, X.Q. Fan, Enhanced physiochemical and mechanical
110137–110147. performance of chitosan-grafted graphene oxide for superior osteoinductivity, Adv.
[40] P.A. Mouthuy, H. Ye, J. Triffitt, G. Oommen, Z. Cui, Physico-chemical Funct. Mater. 26 (2016) 1085–1097.
characterization of functional electrospun scaffolds for bone and cartilage tissue [47] S. Nair, N.S. Remya, S. Remya, P.D. Nair, A biodegradable in situ injectable
engineering, Proc. Inst. Mech. Eng. Part H 224 (2010) 1401–1414. hydrogel based on chitosan and oxidized hyaluronic acid for tissue engineering
[41] X. Miao, D. Sun, Graded/Gradient porous biomaterials, Materials 3 (2009) 26–47. applications, Carbohydr. Polym. 85 (2011) 838–844.
[42] C. Deng, R. Lin, M. Zhang, C. Qin, Q. Yao, L. Wang, J. Chang, C. Wu, Micro/ [48] B.Q. Zhang, H. Sun, L.N. Wu, L. Ma, F. Xing, Q.Q. Kong, Y.J. Fan, C.C. Zhou, X.
Nanometer-structured scaffolds for regeneration of both cartilage and subchondral D. Zhang, 3D printing of calcium phosphate bioceramic with tailored
bone, Adv. Funct. Mater. 29 (2019) 1806068–1806082. biodegradation rate for skull bone tissue reconstruction, Bio-Des. Manuf. 2 (2019)
[43] D.D. Deligianni, N.D. Katsala, P.G. Koutsoukos, Y.F. Missirlis, Effect of surface 161–171.
roughness of hydroxyapatite on human bone marrow cell adhesion, proliferation, [49] T. Kokubo, H.M. Kim, M. Kawashita, Novel bioactive materials with different
differentiation and detachment strength, Biomaterials 22 (2001) 87–96. mechanical properties, Biomaterials 24 (2003) 2161–2175.
[44] B. Kreklau, M. Sittinger, M.B. Mensing, C. Voigt, G. Berger, G.R. Burmester, [50] V.M. Rusu, C.H. Ng, M. Wilke, B. Tiersch, P. Fratzl, M.G. Peter, Size-controlled
R. Rahmanzadeh, U. Gross, Tissue engineering of biphasic joint cartilage hydroxyapatite nanoparticles as self-organized organic-in organic composite
transplants, Biomaterials 20 (1999) 1743–1749. materials, Biomaterials 26 (2005) 5414–5426.

11

You might also like