You are on page 1of 12

European Polymer Journal 119 (2019) 426–437

Contents lists available at ScienceDirect

European Polymer Journal


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

3D bio-printing of levan/polycaprolactone/gelatin blends for bone tissue T


engineering: Characterization of the cellular behavior
Busra Tugce Duymaza,b, Fatma Betul Erdilera,b, Tugba Alana,b, Mehmet Onur Aydogdua,c,
Ahmet Talat Inana,d, Nazmi Ekrena,e, Muhammet Uzunf, Yesim Muge Sahing,h, Erdi Bulush,
Faik Nüzhet Oktara,b, Sinem Selvin Selvib,i, Ebru ToksoyOnerb,i, Osman Kilica,j,

Muge Sennaroglu Bostank, Mehmet Sayip Erogluk,l, , Oguzhan Gunduza,m
a
Center for Nanotechnology & Biomaterials Application and Research at Marmara University, Goztepe Campus, 34722 Istanbul, Turkey
b
Department of Bioengineering, Faculty of Engineering, Marmara University, Goztepe Campus, 34722 Istanbul, Turkey
c
Department of Metallurgical and Materials Engineering, Master of Science, Institute of Pure and Applied Sciences, Marmara University, 34722 Istanbul, Turkey
d
Department of Mechanical Engineering, Faculty of Technology, Marmara University, 34722 Istanbul, Turkey
e
Department of Electrical and Electronics Engineering, Faculty of Technology, Marmara University, Goztepe Campus, 34722 Istanbul, Turkey
f
Department of Textile Engineering, Faculty of Technology, Marmara University, Goztepe Campus, 34722 Istanbul, Turkey
g
Biomedical Engineering Department, Faculty of Engineering & Architecture, Istanbul Arel Univ., Istanbul, Turkey
h
ArelPOTKAM (Polymer Technologies and Composite A & R Center), Istanbul Arel University, Turkey
i
IBSB, Department of Bioengineering, Faculty of Engineering, Marmara University, Goztepe Campus, 34722 Istanbul, Turkey
j
Department of Electrical and Electronics Engineering, Faculty of Engineering, Marmara University, Goztepe Campus, 34722 Istanbul, Turkey
k
Department of Chemical Engineering, Faculty of Engineering, Marmara University, Istanbul, Turkey
l
TUBITAK-UME, Chemistry Group Laboratories, PO Box 54, 41471 Gebze, Kocaeli, Turkey
m
Department of Metallurgy and Materials Engineering, Faculty of Technology, Marmara University, Goztepe Campus, 34722 Istanbul, Turkey

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

Keywords: Poly(ε-caprolactone) (PCL), gelatin (GT) and different concentrations of low molecular weight Halomonas levan
Halomonas levan (HLh) were combined and examined to develop physical networks serving as tissue scaffolds to promote cell
3D printing adhesion for biocompatibility. Three-dimensional bioprinting technique (3D bioprinting) was employed during
Tissue engineering manufacturing the test samples and their comprehensive characterization was performed to investigate the
physicochemical properties and biocompatibility. Physical properties of the printing materials such as viscosity,
surface tension, and density were measured to determine optimal parameters for 3D bioprinting. The scanning
electron microscope (SEM) was used to observe the morphological structure of scaffolds. Fourier-Transform
Infrared Spectroscopy (FT-IR) and differential scanning calorimetry (DSC) were used to identify the interactions
between the components. In-vitro cell culture assays using standard human osteoblast (Hob) cells showed in-
creased biocompatibility of the printing materials with increasing HLh content. Thus, the formulations including
the HLh are expected to be a good candidate for the production of 3D printed materials.

1. Introduction the tissue functions [3]. These fields are based on the accumulation of
body cells on porous 3D scaffold materials made of natural or bio-
Organ dysfunction is a common health problem that originated from compatible synthetic polymers, which require the contributions of
diseases, traumas, infections and aging factors in tissues or organs. biochemical and physicochemical factors enabling tissue growth
Although these problems have been decreasing by organ or tissue through cell proliferation [4]. The 3D printing technique is used in
transplantation [1], restrictions by certain limitations such as lack of various industries, and tissue engineering is focused on material com-
organ donation and tissue rejection induced by the immune system of binations to produce controlled scaffolds providing cell growth on de-
the host have been significant obstacles [2]. Regenerative medicine and fected tissues [5–7]. Cell-seeded scaffolds can act as a template for
tissue engineering are the alternative fields to biomedical sciences, tissue formation [8]. Furthermore, surface reorganization and porosity
which replaces the lost or damaged tissues by restoring or improving of the scaffolds are expected to promote cell adhesion, migration, cell


Corresponding author at: Department of Chemical Engineering, Faculty of Engineering, Marmara University, Goztepe, 34722 Istanbul, Turkey.
E-mail addresses: mehmet.eroglu@marmara.edu.tr (M.S. Eroglu), ucemogu@ucl.ac.uk (O. Gunduz).

https://doi.org/10.1016/j.eurpolymj.2019.08.015
Received 9 July 2019; Received in revised form 8 August 2019; Accepted 13 August 2019
Available online 14 August 2019
0014-3057/ © 2019 Elsevier Ltd. All rights reserved.
B.T. Duymaz, et al. European Polymer Journal 119 (2019) 426–437

differentiation, and proliferation. Additionally, materials used for scaffolds are very promising for future development. The scaffolds were
scaffold construction should possess the suitable level of biodegradation characterized by their surface tension, viscosity, and density properties.
to prevent losing the scaffold form with extremely rapid biodegradation Moreover, their morphological, thermal and chemical properties along
and should provide enough time for cells to produce their extracellular with in-vitro cell culture, using HOb cells for the modeling of bone
matrix (ECM) while their presence on the scaffolds [3,9]. tissue, were studied.
Bone tissue is known for self-repairing ability. However, large bone
defects cannot be entirely treated without external operations such as 2. Materials and methods
autografts or allografts. Today, bone tissue engineering is focused on
techniques to synthesize bone-like scaffolds to preserve or enhance its 2.1. Materials
functions [10]. Fibrous structure, porosity, pore size and modifications
of selected polymer scaffolds are important to become a good candidate Dimethyl sulfoxide (DMSO), Dichloromethane (DCM) and poly(ε-
for bone tissue engineering [11]. Pore size higher than 300 µm is re- polycaprolactone) (PCL, Molecular weight 80,000 g/mol), and gelatin
ported to provide bone formation and enhance vascularisation. (GT, Type A, Gel strength ~ 300 g Bloom) were obtained from
Poly(ε-caprolactone) (PCL) is a biodegradable and biocompatible Sigma–Aldrich (St. Louis, MO, USA). Halomonas levan (HL) was mi-
semi-crystalline synthetic polyester, which is mostly used to be 3D bone crobially produced by Halomonas smyrnensis bioreactor cultures and
scaffold material due to its high mechanical strength and relatively low purified as described [22]. Sodium nitrite (NaNO2) and sodium nitrate
melting temperature. The high pore volume of PCL scaffolds provides (NaNO3) were supplied from JT Baker. Acetic acid was purchased from
enhanced bone regeneration and increased strength and biocompat- Sigma-Aldrich. Sodium carbonate (Na2CO3) was supplied from Riedel-
ibility on HOb cells. On the other hand, due to its slow degradation deHaen. For cell culture experiments; WST-1(4-[3-(4-iodophenyl)-2-(4-
capability, PCL appears more promising over the other biodegradable nitrophenyl)-2H-5-tetrazolio]-1,3-benzene disulfonate) was obtained
equivalents [12,13]. from Roche Applied Science(Germany). DMEM(Dulbecco’s Modified
Gelatin (GT) is a naturally originated material consisting of small Eagle Medium), Trypsin, Fetal Bovine Serum (FBS) and Penicillin-
peptides formed by hydrolysis of collagen fibers. Gelatin can promote Streptomycin were purchased from PAN Biotech (UK Ltd). HOb cell line
wound repair and has remarkable biocompatibility, which overcomes was kindly provided by Gamze Torun Kose (Yeditepe University,
some limitations of using synthetic polymeric materials in bioprinting Turkey). DAPI (4′,6-diamidino-2-phenylindole) was purchased from
[14]. AppliChem(Germany), Trypan Blue, Phosphate Buffered Saline(PBS)
Halomonas levan (HL) is a β (2 → 6) - linked fructan type poly- tablets were obtained from MP Biomedicals.
saccharide, produced by various microorganisms and plants. It is a
water-soluble, non-toxic, biocompatible and strongly adhesive poly- 2.2. Instrumentation
saccharide. Halophilic Halomonas smyrnensis AAD6T is classified as the
first levan producing extremophile, which provides unsterile and low- Scanning electron microscope (SEM) images of the printed samples
cost production. Antioxidant, anticoagulant and heparin mimetic ac- at different magnifications were obtained using a Zeiss EVO MA10
tivity together with peptide and protein-based drug delivery properties (Germany) SEM instrument which was operated at 20 kV acceleration
of HL were reported [15]. In our previous study, improved bio- voltage. Samples were sputter-coated with a thin layer of gold. Melting
compatibility and cell proliferation at the HL/N-isopropyl acrylamide behavior and temperature of pure PCL and the blend samples were
(PNIPAm) hydrogel surface was observed with increasing HL portion, comparatively determined by differential scanning calorimetry (DSC)
which was attributed to the enhancement of HL on the hydrogel surface using Perkin-Elmer Thermal Analyzer System equipped with Jude DSC
at close to body temperature [8]. In our separate study, we observed system and Pyris software. Approximately 5–7 mg of each sample was
that HL induced mouse fibroblast L929 cell viability and proliferation tested with two successive heating and cooling cycle under an argon
when included in poly(ethylene oxide)/chitosan (PEO/Ch) blend films atmosphere (200 ml.min−1 flow rate). Heating and cooling rates were
[16]. Additionally, HL suppressed the PEO crystallinity via hydrogen 10 °C/min and 5 °C/min, respectively. 2D 13C–1H NMR spectra were
bonding formed between hydroxyl groups of HL and etheric C-O-C recorded using a Varian 600 MHz NMR instrument at room temperature
groups of PEO, which resulted in more amorphous, and thus, more in D2O. Molecular weight determination was performed using gel per-
homogeneous and flexible blend films [16]. HL is a high molecular meation chromatography-light scattering system (GPC-LS) equipped
weight (4,200 kD) fructan [8], which could result in viscosity problems with Perkin Elmer 200 high-pressure pump, serial connected column
during the 3D bioprinting process. Therefore, low molecular weight system (Ultrahydrogel 120 + Ultrahydrogel 250 + Ultrahydrogel
Halomonas levan (HLh) was prepared by nitrous acid hydrolysis of HL. 500 + Ultrahydrogel 100), Wyatt Optilab differential refractometer
It is noteworthy that, the presence of HLh in the scaffold compositions (654 nm) and Dawn Heleos multi-angle light-scattering detector.
results in a more homogeneous and flexible structure with induced Hob Mobile phase was 0.1 M NaNO3 solution in 2% acetic acid/water (v/v)
cell proliferation. Considering the hydrogen bonding formed between mixture with a flow rate of 1.0 ml/min. A JASCO FT-IR-4700
PCL, GT, and HLh, the scaffold matrix is expected to keep their uniform Spectrometer (USA) was used to determine intermolecular interactions
3D shape for an extended period of time with more bio-availability as between the components. FT-IR analyses were performed between 400
reported in the previous studies explaining the characteristics of Poly and 4000 cm−1. Mechanical testing of the scaffolds was conducted
(ethylene oxide) blends with sodium alginate [17,18]. using a universal testing machine (Devotrans, Turkey) with a cross-
The present study aims to develop highly biocompatible 3D scaf- head speed of 1.0 mm/min. The surface area of the samples was mea-
folds with improved biocompatibility and physical properties. Different sured using a digital clipper (500, Mitutoyo, USA).
combinations of PCL, GT, and HLh were used for this purpose. Although
many articles concerning PCL/GT based formulations for 3D printing 2.3. Preparation of the hydrolyzed Halomonas levan (HLh)
are available in the literature [19,20], to the best of our knowledge, this
is the first study reporting 3D printing results of HLh containing PCL Halomonas levan (HL) was obtained from halophilic Halomonas
and GT scaffold formulations. Including the high bioadhesive HLh into smyrnensis AAD6 T cultures according to the method described in our
the 3D printing compositions is expected to provide increasing bioa- previous study [8]. Low molecular weight levan (HLh) was obtained by
vailability of the scaffolds. Additionally, the hydroxymethyl groups of the acid hydrolysis of high molecular weight HL (4,200 kDa) under mild
HLh can interact with carbonyl groups of PCL and primary amine condition. HL (3.0 g) was dissolved in 150 ml of 2% acetic acid/water
groups of GT, forming more stable physical networks in 3D bioprinting (v/v) solution. To this solution, NaNO2 (0.5 g dissolved in 5 ml water)
[21]. The initial results of the mechanical properties of the developed was added dropwise under vigorous stirring, which was continued for

427
B.T. Duymaz, et al. European Polymer Journal 119 (2019) 426–437

Table 1
Compositions, 3D printing conditions and mechanical properties of the samples.
Compositions (wt%) 3D printing conditions Mechanical properties

Sample no PCL GT HLh Print head travel speed (mm/s) Build plate temperature (°C) Nozzle diameter (mm) Compressive strength (MPa)

S1 20 0.5 0 100 45 0.4 35.20 ± 0.2


S2 20 0.5 0.2 150 55 0.4 33.12 ± 0.3
S3 20 0.5 1.2 150 47 0.4 32.93 ± 0.4
S4 20 0.5 1.6 130 55 0.4 30.54 ± 1.2

one day at 60 °C and then additional two days at 25 °C. The solution was 2.7. Biocompatibility of the scaffolds
neutralized with 1.0 M of Na2CO3 and dialyzed against ultrapure water
using SpectraPor®Dialysis Membrane cut of 2 kDa. The product was The WST-1 viability test was performed using HOb cell line for 24,
freeze-dried (at −60 °C and 0.001 atm), which yielded low molecular 48 and 72 h. Before seeding, the cells were cultured in DMEM complete
weight levan (HLh) as a white solid (2.1 g, yield 70%). The molecular medium (10% FBS, 1% penicillin-streptomycin) at 37 °C in 5%CO2(6–8
weight of HL and HLh was determined using GPC-LS and NMR spec- passages). Sterilization was provided under UV light for each side,
troscopy was used to determine if any structural change after hydrolysis followed by incubation in 70% ethanol for 30 min and 2% penicillin-
of HL [8]. streptomycin/PBS for 2 h. The scaffolds were placed into 24 well plates
and pre-saturated with DMEM complete medium for 2 h. After the
Trypsinization process, cells were dyed by trypan blue and counted by
2.4. Preparation of solutions and production of bone scaffolds
creating a colorimetric difference since the dye used is meant to be
absorbed only by the cells. HOb cells were seeded on the scaffolds at the
Different compositions were prepared to determine the optimum
density of 1.0 × 104 and incubated for 24, 48, and 72 h at 5%CO2.
concentration for bone tissue printing. Four different scaffold compo-
WST-1 test reagent was added onto wells and incubated at a dark place
sitions were prepared from stock solutions of PCL/GT (20:0.5 wt%) in a
for 37 °C and 5% CO2 for 2 h. The absorbance was measured at 405 nm.
mixture of DCM/DMSO (60:40 wt%). PCL/GT and HLh were mixed at
Adhesion of HOb cells on scaffolds was determined with fluorescence
desired proportions given in Table 1 and then stirred at room tem-
microscopy. Scaffolds with cells and control wells were incubated for
perature for 24 h. The compositions were printed by using a modified
24, 48, and 72 h, and fixation was performed with 4% PFA/PBS solu-
3D printer (Ultimaker® 2+, Netherlands), which utilized a fused de-
tion and then washed with PBS for 2 times. Cells on the scaffold samples
position modeling (FDM) system. It included a computer-aided design
were observed under a Leica DMLB2, (Leica Microsystems, Wetzlar,
(CAD) technology using a heatable build-plat. A digital syringe pump
Germany) fluorescent microscopy.
was used to feed the solutions. Nozzle diameter was 0.4 mm, the build-
plate temperature was varied between 45 and 55 °C, and print head
2.8. Statistical analysis of the biocompatibility tests
travel speed was kept between at 100–150 mm.s−1. The 3D bioprinting
compositions and their printing parameters are given in Table 1.
The statistical analysis was determined by Graph Pad V 5.0 prims
program. The statistically significant mean differences between scaf-
2.5. Solution characterization folds and controls were estimated by One-Way ANOVA with Tukey’s
method. The confidence interval (CI) for all data was assumed to be
The density of the PCL/GT and the PCL/GT/HLh solutions at dif- 95% and p-values below 0.05 were regarded as significant.
ferent concentrations were measured using a standard 10 ml pycn-
ometer. The surface tension of the solutions was determined using a 3. Results and discussion
Sigma 700, DYNE, tensiometer (UK). The solution viscosity of the
samples was determined using a Brookfield type viscometer 3.1. Characterization of hydrolysed Halomonas levan (HLh)
(Brookfield, DV-E, Massachusetts, USA). Measurements were performed
at 25 °C. HLh was prepared by acid hydrolysis of native Halomonas levan
(HL). After hydrolysis, the number average molecular weight (Mn) re-
markably reduced from 4.247 × 106 g/mol to 1.051 × 105 g/mol with
2.6. Biodegradation studies
better water solubility. The specific refractive index (dn/dc) of HLh in
0.1 M NaNO3 in 2% (v/v) acetic acid/water solution at 25 °C was de-
10 mg of each dry scaffold sample was placed in a glass test tube
termined to be 0.1370 ± 0.0028 ml/g. GPC-LS chromatograms (Fig.
containing 10 ml of PBS solution (pH 7.0). The test tubes were in-
S1) and molecular weight values (Table S1) of HL and HLh are given in
cubated in a thermostatic oven at 37 °C. After 1, 3, 5 and 7 days in-
the supporting file. To determine if any structural change that occurred
cubation, samples were removed from the test tubes, quickly blotted
in hydrolysis, NMR characterization of HL and HLh was performed.
with a filter paper and weighted to calculate water absorption capa-
Fig. 1A and B show 1H NMR spectra of HL and HLh. Fructose rings of
cities using Eq. (1). Subsequently, each swollen sample was dried and
HLh showed similar 1H NMR signals at the same shift values observed
weighed again to calculate the biodegradability using Eq. (2).
for HL at (δ, ppm): H1(3.54, 3.62), H3(4.04), H4 (3.96), H5(3.81),
Ww − Wr H6(3.42, 3.75). For further characterization, the 2D 13C–1H NMR cor-
Waterabsorption (%) = × 100
Wr (1) relation spectrum of HLh was recorded (Fig. 1C). 13C NMR signals of
HLh (δ, ppm); C1(59.8), C2(1 0 4), C3(76.2), C4(75.0), C5(80.1)
W0 − Wr C6(63.3) and corresponding 1H NMR signals (δ, ppm); H1(3.54, 3.62),
Weightloss (%) = × 100 H3(4.04), H4(3.96), H5(3.81), H6(3.42, 3.75) are observed. Since C2
W0 (2)
carbon is quaternary, the corresponding 1H NMR signal was not ob-
where Ww is the weight of the swollen sample, Wr is the weight of the served in the spectrum. NMR characterization showed that the nitrous
dry sample after degradation and W0 is the initial weight of the dry acid hydrolysis of HL did not cause any decomposition in fructose rings
sample. of HL, while reducing the molecular weight.

428
B.T. Duymaz, et al. European Polymer Journal 119 (2019) 426–437

Fig. 1. 1H NMR spectrum of native Halomonas levan (A); 1H NMR spectrum of hydrolysed Halomonas levan (B); 1H–13C correlated NMR spectrum of hydrolysed
Halomonas levan (C).

3.2. SEM-XSEM interconnectivity of the fabricated 3D scaffolds were displayed in this


figure. The HLh portion of the scaffolds increased by 0.2, 1.20, and
Fig. 2 shows the production setup, SEM and XSEM images of the 3D 1.60 wt% to optimize the printing parameters to obtain smoother bio-
printed scaffolds. The structural integrity, porosity, and printing process and better scaffold features. The viscosity of printing

429
B.T. Duymaz, et al. European Polymer Journal 119 (2019) 426–437

Fig. 1. (continued)

materials is a particularly important parameter. Lower viscosity results PCL with a compressive strength value of 21.5 MPa. However, me-
in deformation and collapse in printed scaffolds. On the other hand, chanical properties can be improved by using higher molecular weight
higher viscosity generally results in an undesirable defect in scaffold polymer [24]. The compressive strength of sample-S1 (Table 1), con-
and nozzle jam during printing, which leads to morphological differ- sisting of PCL and GT, was measured to be 35.20 MPa. It was reported
ences between the scaffolds. In our formulations, while the percentages that the GT incorporation into the PCL increased the flexibility of PCL/
of PCL and GT were kept constant, the viscosity of the printing material GT blend [26]. However, the compressive strength gradually decreased
increased with increasing HLh portion. As shown in Fig. 2, the scaffold as the HLh was included in the formulation. It was measured to be
formulation containing 1.20 wt% of HLh concentration showed uniform 30.54 MPa for the formulation containing 1.6 wt% of HLh (sample-S4).
shape with dense appearance and pores compared to the others. This could be due to the more brittle nature of HLh than PCL and GT
Moreover, the addition of HLh into the PCL/GT solution exhibited finer [27]. Additionally, the average compressive strength value of the
and ordered scaffold structures as well as an increment in the number of human femur cortical bone was reported to be between 96.35 and
pores. As reported in a previous study [23], a ternary composite na- 143.35 MPa for female and 122.86 to 161.44 MPa for male [28].
nofibrous scaffold from PCL/GT/chitosan was fabricated for tissue en-
gineering applications. SEM images showed that the fiber morphology 3.4. FT-IR
of the scaffolds was found to be influenced by the concentrations of
PCL/GT/chitosan in the polymer solution. Increasing the PCL and de- Fig. 3 shows the FT-IR spectra of native components and the scaffold
creasing the GT concentrations were suggested to obtain better mor- sample-S3. In the FT-IR spectrum of pure PCL, the characteristic ab-
phology with uniform structure [23]. In the present study, similar re- sorption bands at 1749 cm−1 and 1185 cm−1 are due to the carbonyl
sults were obtained consistent with the outcome of the previous study stretching and axial symmetric stretching of CeOeC groups, respec-
[23]. The higher concentration of PCL compared to GT in the PCL/GT tively. The peaks in the range of 2850–3000 cm−1 were assigned to
scaffold provided better results such as morphological integrity and different CeH vibrational stretching absorption of PCL [29]. In the FT-
improved solidification while bioprinting as well as the smooth bio- IR spectrum of pure GT, the bands at 1636 cm−1 and 1524 cm−1 are
printing phase. due to the characteristic peptide amide-1 and amide-2 absorptions,
respectively [26]. In the FT-IR spectrum of pure HLh, a broad absorp-
3.3. Mechanical properties tion peak observed between 3200 cm−1–3500 cm−1 is due to the OeH
stretching of OH and CH2eOH groups of fructose rings. The strong
The ultimate compressive strength of a sample means the load ap- absorption peak at 1120 cm−1 was attributed to the CeOeC symmetric
plied per unit area at which the sample will fail in compression. The bending vibration of the fructose ring of HLh [8,30]. FT-IR spectra of
compressive strength value of a scaffold should be close to the value of the scaffold formulations exhibited the characteristic absorption peaks
the native bone to maintain a certain level of mechanical compatibility. of the pure components. As an example, in the FT-IR spectrum of the
The mechanical properties of pure PCL scaffolds produced by different sample-S3, which contains 1.2% of HLh, the peaks of PCL at 1749 cm−1
methods were reported in various studies. The compressive modulus, and 1185 cm−1 shifted to 1720 cm−1 and 1165 cm−1, respectively.
resulting in a wide range of compressive strength (14.9 MPa − Moreover, the CeOeC band of HLh and the amide I band of GT shifted
215 MPa), is believed to be directly related to the production method to 1106 cm−1 and 1626 cm−1, respectively. These remarkable shifts of
and the morphology of the scaffolds [24]. It has also been reported that characteristic absorption peaks of the pure components could be con-
the compressive strength of the pure and bulk PCL was 38.7 MPa [25]. sidered as evidence of their energetic interactions in scaffold formula-
Tissue scaffolds were also manufactured using low molecular weight tions. This confirmed the hydrogen bond formation between scaffold

430
B.T. Duymaz, et al. European Polymer Journal 119 (2019) 426–437

Fig. 2. (a) Schematic representation of scaffolds and 3d printing system with surface and cross-sectional SEM microscope of sample-S1 (b, b1); sample-S2 (c, c1);
sample-S3 (d, d1); sample-S4 (e, e1) with high and low magnifications for each sample respectively, showing the morphological structure of the scaffolds and the
porous structure of it (f) with the mechanical properties of the scaffolds.

components, resulting in physical crosslinking leading a more di- melting behavior could be due to the hindrance of the segmental
mensionally stable 3-D printing scaffold. movement of PCL chains in sample-S3 to some extent, which was
probably due to the hydrogen bonding formed between the components
in the presence of GT and HLh [17]. This was another confirmation of
3.5. DSC the intermolecular interactions of the components and, thus, the for-
mation of physical networks between the components, which provided
DSC curves of pure PCL and sample-S3 were recorded to determine
more dimensionally stable 3D printing scaffolds with an effective
the interaction of the blend components (Fig. 4). Successive heating and
bioavailability.
cooling cycle was used for the removal of possible impurities and the
elimination of possible error arising from PCL crystallization condition.
Although sample-S3 and pure PCL displayed the same melting tem- 3.6. Solution characterization
perature (approx. 54 °C), they exhibited different melting behavior.
While sample-S3 had a narrow and sharp endotherm, pure PCL ex- Surface tension, viscosity, and density of the printing solutions are
hibited relatively wide melting endotherm. HLh and GT do not have a displayed in Fig. 5. The solution viscosity of the samples-S2-S4 was
melting temperature at around 54 °C. By considering the different found to be higher than that of the sample-S1. Surface tension had a

431
B.T. Duymaz, et al. European Polymer Journal 119 (2019) 426–437

Fig. 3. FT-IR spectra of native components and sample-S3.

high value for sample-S1 and sample-S2 solutions, which was in the increasing HLh portion. This was probably due to the effective hy-
range of 47.56–84.50 mN.m−1 and decreased with the increasing of drogen bonding capability of HLh with PCL and GT. As a result of ob-
HLh content (for sample-S3 and sample-S4). The density was measured servations, 1.1 ± 0.05 mm (ranging from 1.00 to 1.30 mm) diameters
between the range of 1.08–1.16 g.ml−1 and increased with HLh con- for sample-S3 have a more uniform pore size distribution. In a previous
tent. It can be concluded that as the HLh concentration increases the study, PCL/GT nanofiber membranes were prepared at different PCL/
viscosity of the sample solutions increases to some extent. Sample-S4, GT ratios, and the average pore sizes of the membranes were found to
which has a higher amount of HLh presented the highest viscous be remarkably less than L929 fibroblast cells (in the range of 1–3 µm)
structure, which is not considered to be ideal for the printing process. [32]. Much smaller membrane pore sizes than L929 fibroblast cells
Because of the evaporation of the solvent during the printing process, it might seem favorable, which allows the cell to proliferate on the
was very difficult to use highly viscous printing solutions. The density membrane surface rather than inside of the membrane. However, this
of the solutions also increased with the increase of HLh. Pore size is an can create limitations for the area whereas cells can find a place to
important parameter since the pores in the scaffold provide cell mi- attach. Our pore sizes are found to be bigger compared to the previous
gration and efficient exchange of nutrients or wastes between the studies [32] due to the nature of the selected production method and
scaffold and its environment [31]. According to the measurements in yet they can provide a suitable environment for proper proliferation.
Fig. 6a1-a4, the macrostructure of scaffolds becomes arranged with an Since the scaffold is biodegradable, either guiding cells to inside of the

Fig. 4. DSC thermograms of (a) sample-S3 and (b) pure PCL.

432
B.T. Duymaz, et al. European Polymer Journal 119 (2019) 426–437

entanglements and cyclization. Therefore, some differences in both


rates of swelling and equilibrium swelling values of the networks can be
observed [33,34]. Despite the decreasing water absorption of these
scaffolds, the sample-S1 scaffold started to increase its water absorption
rate after 3 days of the incubation period. As a result of the biode-
gradation tests, similar to the previous studies, it has been observed that
the degradation of samples decreased the water absorption behavior of
the samples [35]. Fig. 6c shows the weight loss of the scaffold samples
at different time intervals (1, 3, 5, and 7 days) in PBS solution
(pH = 7.4) at 37 °C. On 1, 3, 5, and 7 days, the samples were in-
vestigated for any apparent weight loss of scaffolds in the PBS solution.
During the first day of incubation, all of the scaffolds had a negligible
amount of weight loss. After 3 days of incubation, the samples-S1 and
S2 scaffolds showed slightly fast degradation compared to others. After
5 days of the incubation period, the sample-S3 scaffold started to de-
grade drastically and kept losing weight after day 7. Other scaffold
samples were degraded very slowly and exhibited a slight linear slope
throughout the seven days of study. The weight losses were found to be
10.18%, 7.18%, 46.54%, and 3.45% according to the Eq. (2) for the
samples-S1, S2, S3, and S4, respectively after 7 days of incubation.
In a previous study, PCL/GT biocomposite scaffolds were fabricated
by varying the ratios of PCL/GT concentrations for nerve tissue en-
gineering. PCL/GT nanofibers with a ratio of 50:50 was displayed the
fastest degradation rate in PBS over 2 weeks, and biodegradability of
PCL/ GT nanofibrous scaffolds increased by increasing the GT content
[36].

3.8. Cell proliferation and viability

As can be seen in Fig. 7, the cell proliferation and viability of HOb


cells on 3D printed scaffolds were determined for 24, 48, and 72 h.
Viability results of samples-S1, S2, S3 and S4 were 86.18, 89.76, 88.54,
and 91.11% at 24 h, 82.77, 80.63, 81.93, and 81.29% at 48 h whereas,
77.76, 84.79, 84.62, and 90.19% at 72 h, respectively (Fig. 7a1-a3).
Among the synthetic biocompatible polymers, PCL has been widely
used as a porous scaffold material for tissue engineering particularly
bone and cartilage tissue repairing [19,37,38]. On the second day, all
viability results decreased and at 72 h, increased aside from sample-S1.
Viability results of the scaffold samples consisting of HLh indicate that
after 48 h and viability increased. Additionally, chemicals could be
released from scaffolds that can affect cell viabilities during the first
Fig. 5. Solution properties of blends; (a) surface tension, (b) viscosity and (c) 48 h and cells could proliferate after adjusting the new environment. On
density. the other hand, viability proceeds decreasing through 24–72 h on the
sample-S1 scaffold. Notably, the results show that HLh enhances the
porous membrane or having them on the surface of the membrane does viability of more than PCL and GT. Additionally, DMSO/DCM solution,
not change anything and this versatility increases the value of the which used as a solvent for the scaffold samples, might have toxic ef-
scaffolds. fects on cells. DMSO is generally used as a freezing agent in cell culture
and it has a toxic impact on concentrations above 10% at room tem-
3.7. Water absorption and biodegradability perature [39]. Cells on the scaffolds tolerate these cytotoxic effects with
its high biocompatible and proliferation properties. In previous studies,
The water absorption percentage of the scaffolds was calculated microspheres composed of GT and hydroxyapatite were produced with
using Eq. (1), and a graph was drawn as a function of degradation time uniform morphology by a wet-chemical method. Cell viability tests of
(Fig. 6b). The absorption rate varied depending on the content of HLh microspheres were performed by MTT test with HOb cell line G-292.
during the incubation period. The maximum water absorption rate was The materials were diluted into 1:4 and 1:1 with culture medium and
observed after the first day of incubation for each scaffold samples and, only 1% and 2% cell damage were observed, respectively. This result
among them, the sample-S4 scaffold had the highest absorption rate. illustrates that the gelatine-hydroxyapatite blend was non-toxic for
After one day, water absorption started to decrease. After three days, bone tissue engineering [40]. In another study, 5 and 10% (w/v) PCL
sample-S2 and sample-S4 scaffolds showed nearly a linear slope until coated 3D-printed scaffolds were prepared to enhance the osteogenic
the end of 7 days of incubation, which indicated that these scaffolds potential and mechanical strength. In vitro tests of 3D scaffolds, con-
reached to a water absorption equilibrium. However, the water ab- sisting of calcium sulfate-based hydroxyapatite were performed by MTS
sorption of the sample-S3 scaffold dropped severely at the end of 5 days assay with the MG-63 cell line. On the first day, there were no differ-
and increased slightly at the end of 7 days. Physically formed hydrogels ences in cell proliferation results between 5% PCL coated and uncoated
of biomaterials are spontaneously formed networks via intermolecular scaffolds. On the 10% PCL coated scaffolds, proliferation was higher
interaction of the components. In the course of network formation, than other scaffolds and cell growth was increased rapidly until day 5.
while most of the polymer chains join the network in gel form, others On day 15, 5% of PCL coated scaffolds were the highest for cell growth.
remain in sol form as a result of chain irregularities such as physical These results indicated that PCL coating did not cause toxicity [41]. In

433
B.T. Duymaz, et al. European Polymer Journal 119 (2019) 426–437

Fig. 6. Pore size distribution of (a1) sample-S1, (a2) Sample-S2, (a3) Sample-S3, (a4) Sample-S4 scaffolds. Weight loss (b) and water absorption (c) graphs of scaffolds.

our previous study, we synthesized HLh based PNIPAM hydrogels and with the viability results. Adherent cells increased after 48 h in HLh
performed their viability tests with mouse fibroblast L929 for 24, 48, scaffolds and maximum adherent cells were observed at the sample-S4
and 72 h. Viability increased with increasing HLh concentration, where scaffold, which was also consistent with viability result at 72 h. Cells
the result is consistent with previous studies [8]. Morphology of ad- were highly aggregate on some specific areas, which were considered as
herent cells on controls and scaffolds were also determined by fluor- high HLh concentration spots on scaffolds. Scaffolds were floating over
escent microscopy, and cell nuclei were observed after DAPI staining on the culture medium due to their lightweight and it was the main pro-
24, 48, and 72 h of culture time (Fig. 7b1-b3). The scaffold samples and blem on cell culture studies. The cells were not able to contact with the
control cells were seeded on the glass in well plates, where the cell scaffolds completely, which was observed on fluorescent microscopy
population was observed to increase with increasing culture time. After images (Fig. 7b1-b3), but the intolerable amount of toxic chemicals
24 h adherence of cells on sample-S1 and sample-S4 were found to be were not released from scaffolds.
significantly higher than the other scaffolds but less than the control Adhesion of HOb cells on the scaffolds was observed at 24, 48, and
group. At 48 h and 72 h, cell adherence on the sample-S1 scaffold de- 72 h (Fig. 7c1i-c4ii) of culture by using SEM to examine in depth the
creased, which was consistent with the viability results. Cell adherence biocompatibility of these scaffolds. The SEM images indicated that HOb
increased with increasing concentration of HLh and at 48 h, the number cells adhered tightly to the scaffold surface and grew actively by
of adherent cells decreased on all scaffolds, which was also consistent showing typical morphological features of osteoblast. Additionally, the

434
B.T. Duymaz, et al. European Polymer Journal 119 (2019) 426–437

Fig. 7. HOb cell viability results after (a1) 24, (a2) 48 and (a3) 72 h of cultivation on scaffolds with fluorescent microscopy images of HOb cells after (b1) 24, (b2) 48,
and (b3) 72 h of cultivation on control, along with bioprinted scaffolds with SEM images of (c1, c1i, c1ii) sample-S1, (c2, c2i, c2ii) sample-S2, (c3, c3i, c3ii) sample-S3,
and (c4, c4i, c4ii) sample-S4 after cell culture test.

435
B.T. Duymaz, et al. European Polymer Journal 119 (2019) 426–437

cells were proven to be worthy of vitality and fast proliferation at the regeneration in 3D printing bioactive ceramic scaffolds with improved tissue/ma-
end of 72 h, which indicated the biocompatibility of these scaffolds. The terial interface pore architecture in thin-wall bone defect, Biofab 9 (2017) 025003.
[8] A. Osman, E.T. Oner, M.S. Eroglu, Novel levan and pNIPA temperature sensitive
distribution of pores and toughness of the scaffold affected the mor- hydrogels for 5-ASA controlled release, Carbohydr. Polym. 165 (2017) 61–70.
phology of the cells on the scaffold. The SEM images showed that the [9] Cao, Z. Dou, C.S. Dong, Scaffolding biomaterials for cartilage regeneration, J.
cells were randomly distributed on the surface and inside of the scaf- Nanomater. 8 (2014) 489128.
[10] S. Bose, S. Vahabzadeh, A. Bandyopadhyay, Bone tissue engineering using 3D
folds. The cell microfilaments, which are the thinnest filaments of the printing, Mater. Today 16 (2013) 496–504.
cell skeleton, were shown as linkages to the scaffold material. When [11] S. Hosseinpour, M.G. Ahsaie, M.R. Rad, M. Taghi-Baghani, S.R. Motamedian,
comparing the scaffolds, the ones with HLh had more uniform pore size A. Khojasteh, Application of selected scaffolds for bone tissue engineering: a sys-
tematic review, OMFS 21 (2017) 109–129.
distribution and shape. The addition of HLh to the PCL/GT solution [12] A.V. Do, B. Khorsand, S.M. Geary, A.K. Salem, 3D printing of scaffolds for tissue
increased its compatibility and had a significant effect on cell growth of regeneration applications, Adv. Healthc. Mat. 4 (2015) 1742–1762.
the scaffold. It is notable that the increasing HLh portion of the com- [13] L. Dong, S.J. Wang, X.R. Zhao, Y.F. Zhu, J-K., Yu3D-, printed poly(ε-caprolactone)
Scaffold integrated with cell-laden chitosan hydrogels for bone tissue engineering,
positions increased the number of cells on the scaffold and showed
Sci. Rep. 7 (2017) 13412.
better proliferation rates. In our previous study, we showed that HL [14] J.F. O’Brien, Biomaterials and scaffolds for tissue engineering, Mater. Today 14
enhanced the biocompatible, cell attachment and proliferation prop- (2011) 88–95.
erties of polyethylene oxide/chitosan/HL polymer blend films, and HL [15] E.T. Oner, L. Hernandez, J. Combie, Review of levan polysaccharide: from a century
of past experiences to future prospects, Biotechnol. Adv. 34 (2016) 827–844.
increased the viability [16]. These results indicated that the scaffolds [16] M.S. Bostan, E.C. Mutlu, H. Kazak, S. Sinan Keskin, E.T. Oner, M.S. Eroglu,
have excellent biocompatibility. Moreover, adequate growing of cells Comprehensive characterization of chitosan/PEO/levan ternary blendfilms,
on the scaffold reveals that there is no toxic substance release to the Carbohydr. Polym. 102 (2014) 993–1000.
[17] T. Caykara, S. Demirci, M.S. Eroglu, O. Guven, Poly(ethylene oxide) and its blends
medium. with sodium alginate, Polymers 46 (2005) 10750–10757.
[18] N.J. Chang, Y.R. Jhung, C.K. Yao, M.L. Yeh, Hydrophilic gelatin and hyaluronic
4. Conclusion acid-treated PLGA scaffolds for cartilage tissue engineering, JABFM 11 (2013)
45–52.
[19] R. Zheng, H. Duan, J. Xue, Y. Liu, B. Feng, S. Zhao, Y. Zhu, Y. Liu, A. He, W. Zhang,
The 3D scaffolds of the PCL/GT and PCL/GT/HLh blends were W. Liu, Y. Cao, G. Zhou, The influence of Gelatin/PCL ratio and 3-D construct shape
successfully fabricated using novel 3D bioprinting formulations. The of electrospun membranes on cartilage regeneration, Biomaterials 35 (2014)
152–164.
effect of HLh content on the properties of the polymer solution and 3D [20] Y. Zhang, H. Ouyang, C.T. Lim, S. Ramakrishna, Z.-M. Huang, Electrospinning of
scaffolds was comprehensively tested and analyzed. The increasing HLh gelatin fibers and gelatin/PCL composite fibrous scaffolds, J. Biomed. Mater. Res. B.
ratio in PCL/GT composite formulation led to an increase in density and Appl. Biomater. 72 (2005) 156–165.
[21] A. Saarai, T. Sedlacek, V. Kasparkova, T. Kitano, P. Saha, On the characterization of
viscosity. The surface tension of the solution decreased as well as the
sodium alginate/gelatine-based hydrogels for wound dressing, J. Appl. Polm. Sci.
compressive strength of the scaffolds. The obtained scaffolds had uni- (2012) 79–88.
form pore size with homogeneous distribution and shape with dense [22] H.K. Sarilmiser, O. Ates, G. Ozdemir, K.Y. Arga, E.T. Oner, Effective stimulating
appearance. The hydrogen bonding between the scaffold components factors for microbial levan production by Halomonas smyrnensis AAD6 T, J. Biosci.
Bioeng. 119 (2015) 455–463.
was observed, which resulted in more dimensional stability of scaffolds. [23] S. Gautam, C.F. Chou, A.K. Dinda, P.D. Potdar, N.C. Mishra, Fabrication and
PCL/GT/HLh scaffold formulations provided better proliferation to characterization of PCL/gelatin/chitosan ternary nanofibrous composite scaffold
cells than PCL/GT scaffold. These results indicated that the scaffold has for tissue engineering applications, J. Mater. Sci. 49 (2014) 1076–1089.
[24] A.D. Olubamiji, Z. Izadifar, J.L. Si, D.M.L. Cooper, B.F. Eames, D.X. Chen,
good biocompatibility and potential for the preparation of effective Hob Modulating mechanical behaviour of 3D-printed cartilage-mimetic PCL scaffolds:
cell scaffolds. influence of molecular weight and pore geometry, Biofabrication 8 (2016) 025020.
[25] S. Eshraghi, S. Das, Mechanical and microstructural properties of polycaprolactone
scaffolds with 1-D, 2-D, and 3-D orthogon, ally oriented porous architectures pro-
Acknowledgment duced by selective laser sintering, Acta. Biomat. 6 (2010) 2467–2476.
[26] J.W. Jung, H. Lee, J.M. Hong, J.H. Park, J.H. Shim, T.H. Choi, D.W. Cho, A new
This study has been funded by BAPKO, Marmara University, Turkey; method of fabricating a blend scaffold using an indirect three dimensional printing
technique, Biofabrication 7 (2015) 045003.
grant no: SAG-B-090217-0036 and Fen-C-YLP-120917-0549; grand no:
[27] J.R. Barone, M. Medynets, Thermally processed levan polymers, Carbohydr. Polym.
FEN-E-130515-0175. 69 (2007) 554–561.
[28] R. Havaldar, S. Pilli, B. Putti, Insights into the effects of tensile and compressive
loadings on human femur bone, Adv. Biomed. Res. 3 (2014) 101.
Appendix A. Supplementary material
[29] T. Elzein, M. Nasser-Eddine, C. Delaite, S. Bistac, P. Dumas, FTIR study of poly-
caprolactone chain organization at interfaces, J. Colloid. Interface. Sci. 273 (2004)
Supplementary data to this article can be found online at https:// 381–387.
doi.org/10.1016/j.eurpolymj.2019.08.015. [30] F. Kucukkasik, H. Kazak, D. Guney, I. Fionre, A. Poli, O. Yenigun, B. Nicolaus,
E.T. Oner, Molasses as fermentation substrate for levan production by Halomonas
sp, Appl. Microbiol. Biotechnol. 89 (2010) 1729–1740.
References [31] E. Díaz, I. Sandonis, M.B. Valle, In vitro degradation of poly(caprolactone)/nHA
composites, J. Nanomater. 8 (2014) 802435.
[32] J. Xue, M. He, H. Liu, Y. Niu, A. Crawford, P.D. Coates, D. Chen, R. Shi, L. Zhang,
[1] Y. Bozkurt, A. Sahin, A. Sunulu, M.O. Aydogdu, E. Altun, F.N. Oktar, N. Ekren, Drug loaded homogeneous electrospun PCL/gelatin hybrid nanofiber structures for
O. Gündüz, Electrospun nanocomposite materials, a novel synergy of polyurethane anti-infective tissue regeneration membranes, Biomaterials 35 (2014) 9395–9405.
and bovine derived hydroxyapatite, J. Phys. Conf. Ser. 829 (2016) 012015. [33] M.S. Eroglu, Characterization of the network structure of hydroxyl terminated poly
[2] M.-Y. Shie, W.-C. Chang, L.-J. Wei, Y.-S. Huang, C.-H. Chen, C.-T. Shih, Y.-W. Chen, (butadiene) elastomers prepared by different reactive systems, J. Appl. Polym. Sci.
Y.-F. Shen, 3D printing of cytocompatible water-based light-cured polyurethane 70 (1998) 1129–1135.
with hyaluronic acid for cartilage tissue engineering applications, Materials 10 [34] Y. Moukbil, F.N. Oktar, B. Ozbek, D. Ficai, A. Ficai, E. Andronescu, M.S. Eroglu,
(2017) 136. O. Gunduz, Biohydrogels for medical applications: A short review, Org. Commun.
[3] N.S. Remya, P.D. Nahir, Modulation of chondrocyte phenotype by bioreactor as- 11 (3) (2018) 123–141.
sisted static compression in a 3D polymeric scaffold with potential implications to [35] N. Sultana, M.R.A. Kadir, Study of in vitro degradation of biodegradable polymer
functional cartilage tissue engineering, tissue science and engineering, J. Tissue Sci. based thin films and tissue engineering scaffolds, Afr. J. Biotechnol. 10 (2011)
Eng. 6 (2015) 148. 18709–18715.
[4] J. Kundu, J.H. Shim, J. Shang, S.W. Kim, D.W. Cho, An additive manufacturing- [36] L.G. Mobarakeh, M.P. Prabhakaran, M. Morshed, M.H.N. Esfahani, S. Ramakrishna,
based PCL–alginate– chondrocyte bioprinted scaffold for cartilage tissue en- Electrospun poly(3-caprolactone)/gelatin nanofibrous scaffolds for nerve tissue
gineering, J. Tissue Eng. 9 (2013) 1286–1297. engineering, Biomaterials 29 (2008) 4532–4539.
[5] F. Asghari, M. Samiei, K. Adibkia, A. Akbarzadeh, S. Davaran, Biodegradable and [37] A.M. Haaparanta, E. Jarvinen, I.F. Cengiz, V. Ella, H.T. Kokkonen, I. Kiviranta,
biocompatible polymers for tissue engineering application: a review, Artif. Cells M. Kellomaki, Preparation and characterization of collagen/pla, chitosan/pla, and
Blood Substit. Biotechnol. 45 (2017) 185–192. collagen/chitosan/pla hybrid scaffolds for cartilage tissue engineering, J. Mater.
[6] V. Karageorgiou, D. Kaplan, Porosity of 3D biomaterial scaffolds and osteogenesis, Sci. Mater. Med. 25 (2014) 1129–1136.
Biomaterials 26 (2005) 5474–5491. [38] X. Li, R. Cui, L. Sun, K.E. Aifantis, Y. Fan, Q. Feng, F. Watari, 3D-printed biopoly-
[7] H. Shao, X. Ke, A. Liu, M. Sun, Y. He, X. Yang, J. Fu, Y. Liu, L. Zhang, G. Yang, Bone mers for tissue engineering application, Int. J. Polym. Sci. 13 (2014) 829145.

436
B.T. Duymaz, et al. European Polymer Journal 119 (2019) 426–437

[39] J. Galvao, B. Davis, M. Tilley, E. Normamdo, M.R. Duchen, M.F. Cordeiro, Eng. C 57 (2015) 113–122.
Unexpected low-dose toxicity of the universal solvent DMSO, FASEB J. 28 (2014) [41] B.S. Kim, S.S. Yang, H. Park, S.H. Lee, Y.S. Cho, J. Lee, Improvement of mechanical
1317–1330. strength and osteogenic potential of calcium sulfate-based hydroxyapatite 3-di-
[40] S.C. Chao, M.J. Wang, N.S. Pai, S.K. Yen, Preparation and characterization of ge- mensional printed scaffolds by ε-polycarbonate coating, J. Biomater. Sci. Polym. Ed.
latin–hydroxyapatite composite microspheres for hard tissue repair, Mater. Sci. 28 (2017) 1256–1270.

437

You might also like