You are on page 1of 10

Carbohydrate Polymers 169 (2017) 366–375

Contents lists available at ScienceDirect

Carbohydrate Polymers
journal homepage: www.elsevier.com/locate/carbpol

Biomacromolecular-based ionic-covalent hydrogels for cell


encapsulation: The atelocollagen − Oxidized polysaccharides couples
Andreea Luca a , Vasilica Maier b , Stelian S. Maier b,d,∗ , Maria Butnaru c , Maricel Danu a ,
Constanta Ibanescu a , Mariana Pinteala d , Marcel Popa a
a
Department of Natural and Synthetic Polymers, “Gheorghe Asachi” Technical University of Iasi, 700050, Romania
b
Department of Textiles and Leather Chemistry, “Gheorghe Asachi” Technical University of Iasi, 700050, Romania
c
Faculty of Medical Bioengineering, “Gr. T. Popa” University of Medicine and Pharmacy of Iasi, 700115, Romania
d
“Petru Poni” Institute of Macromolecular Chemistry, Centre of Advanced Research in Bionanoconjugates and Biopolymers, Aleea Grigore Ghica Voda 41A,
700487 Iasi, Romania

a r t i c l e i n f o a b s t r a c t

Article history: Mixed crosslinked networks of ionic-covalent entanglement type were prepared starting from ternary
Received 19 January 2017 mixtures of atelocollagen (aK; as fibrillary matrix generator), sodium hyaluronate (NaHyal; a microfib-
Received in revised form 29 March 2017 rillation assistant), and oxidized polysaccharides (OxPolys; as both cross-linkers and matrix fillers), and
Accepted 18 April 2017
were tested as hydrogels for eukaryotic cell encapsulation. Either oxidized gellan (GellOx) or pullulan
Available online 22 April 2017
(PullOx) were used. An original procedure and optimal hydrogel recipes were developed to encapsulate
fibroblasts and adipose-derived stem cells, while preserving their viability and proliferative ability dur-
Keywords:
ing ex vivo temporarily storage. Physical-chemical, rheological, and biocompatibility properties of the
Atelocollagen
Oxidized polysaccharides
prepared hydrogels were compared against the classic alginate hydrogel used for cell encapsulation. A
Hydrogels larger range of material characteristics (from lax to stiff) and better laboratory maneuverability were
Cell encapsulation demonstrated, which permit to design appropriate compositions for particular cell types. All hydrogels
Cell release undergo fast liquefaction at temperatures between 42 and 50 ◦ C, permitting the cell release after a short
Thermal liquefaction innocuous thermal shock.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction make the necessary metabolic exchanges for in vitro survival. This
is why efforts are made to produce hydrogels bearing cell instruc-
Human cells reside, evolve, and potentially move inside complex tive cues, preferably starting from biomacromolecular precursors
and dynamic morphological and biochemical environments. Except isolated from tissues, or generated by cells cultured in controlled
in the case of some epithelial tissues, non-circulating cells are environments. The envisaged applications are, among others, the
hosted (either attached or entrapped) by the extracellular matrix long term storage of cells (Chan, Zorlutuna, Jeong, Kong, & Bashir,
(ECM), which is intrinsically inhomogeneous, anisotropic and tridi- 2010), cell therapy applications (Hashemi & Kalalinia, 2015), pro-
mensionally structured. An essential role in cell accommodation ducing bioartificial organs (Iacovacci, Ricotti, Menciassi, & Dario,
inside the ECM is played by the instructive biochemical and mor- 2016), and bioprinting (Murphy & Atala, 2014). Cell encapsulation
phological cues offered by the ECM macromolecular components specifically benefits from this type of hydrogels.
(Caliari et al., 2015). In the absence of such cues, cells limit or alter One of the biomacromolecular edifices that abundantly include
their metabolism, cease their movement and/or proliferation, and cell instructive cues is represented by collagen supramolecular
may deviate from their normal functions (Janson & Putnam, 2015). aggregates, particularly of fibrillary type. Three kinds of cues
Structured hydrogels are to some extent able to mimic the nat- are common for them: (i) cell recognition domains (like the
ural environment of cells. If at least some types of cues are present GFOGER amino acid sequence (Malcor et al., 2016)), (ii) physical
in a minimal amount inside artificial or synthetic cytocompatible and mechanical peculiarities of collagen assemblies (Moeinzadeh,
hydrogels, cells “accept” to be hosted in, and remain biochemi- Shariati, & Jabbari, 2016), and (iii) microscopic-scale profile and
cally, biomechanically, and proliferative active, as long as they can orientation of fibrillary collagen entities into the tri-dimensional
substrata (Chwalek et al., 2016). An easily available form of fibril-
lar aggregates-producing collagen is the atelocollagen (aK), which,
under appropriate conditions (of purity, concentration, tempera-
∗ Corresponding author at: Department of Textiles and Leather Chemistry, “Ghe-
ture), generates physical hydrogels.
orghe Asachi” Technical University of Iasi, Romania.
E-mail address: smaier@tuiasi.ro (S.S. Maier).

http://dx.doi.org/10.1016/j.carbpol.2017.04.046
0144-8617/© 2017 Elsevier Ltd. All rights reserved.
A. Luca et al. / Carbohydrate Polymers 169 (2017) 366–375 367

To increase the versatility of cell-embedding/cell-interacting tendons, and was purified and sterilized according the protocols
substrata, aK-based mixed hydrogels are prepared. The second described in references (Maier, Maier, & Buciscanu, 2010) and (Xian,
generic component of these hydrogels is frequently a polysac- 2010), respectively. It was stored in lyophilized form, at −20 ◦ C, and
charide or a functionalized polysaccharide, whose role is: (i) to was prepared just before use, in sterile conditions, as a colloidal
modulate the rheo-mechanical characteristics of the mixed-gels, solution with a concentration of 6 mg/mL in 0.01 M HCl, which was
(ii) to diversify the morphologic entities of structured gels, (iii) to maturated for 48 h, at 4 ◦ C.
add a supplementary (bio)chemical functionality, and, (iv) in the Low acyl gellan was purchased from Kelco, and pullulan (Mw
case of crosslinking capable polysaccharides, to induce covalent of about 100 kDa) from Fluka. Prior to use, both polysaccharides
bridges between the hydrogel macromolecular entities. were purified and prepared as colloidal solutions with a concen-
The present work hypothesizes and investigates the importance of tration of 10 mg/mL, by dissolving them in double distilled water
recipe formulation on the physical-chemical and cell embedding char- (gellan at 80 ◦ C). The final viscous solutions were degassed under
acteristics of ternary hydrogels of ionic-covalent entanglement type, vacuum (100 millibar, 40 ◦ C, for 6 h), and maturated overnight, at
prepared starting from (i) aK, (ii) sodium hyaluronate (NaHyal) room temperature.
as functional polysaccharide, and (iii) oxidized polysaccharides All the general and analytic reagents were purchased from
(OxPolys) as macromolecular crosslinking agents. The rationale of Sigma. All the aqueous solutions were prepared using fresh double
working with ternary mixtures of aK − NaHyal − OxPolys derives distilled water (ddH2 O).
from the constraints imposed by ECM mimicking, when anchoring-
dependent mammalian cells must be temporarily hosted: the need 2.2. Preparation of oxidized polysaccharides in stock solid form
to generate loose but rheo-mechanically stable hydrogels, able to self-
structure in the presence of cells, and to disintegrate in cell-protective Oxidation was conducted in an opaque reaction vessel, at room
conditions. The role of aK is to produce the tri-dimensional fibrillar temperature, according to a procedure derived from that described
structure of ECM, assisted by NaHyal in small but sufficient amounts in reference (Gong et al., 2009). A solution of 0.1 M sodium meta-
to consume the co-precipitation tendency of aK against larger periodate (NaIO4 ) was added into each polysaccharides solutions,
quantities of other polysaccharides having negative zeta poten- to provide an amount of 160 mg NaIO4 for each 100 mL. After three
tials. The crosslinking active OxPolys contributes to the expansion hours, an equimolar quantity of ethylene glycol was added (relative
of the hydrogel network, by inducing both physical and cova- to the NaIO4 amount) to stop the polysaccharides oxidation, and
lent bridges between the fibrillary aK entities. As a consequence, the stirring was maintained for another half an hour. The resulted
hydrogels able to be manipulated according to cell encapsula- solutions were dialyzed against ddH2 O, for three days, at room tem-
tion/culturing/release protocols are obtained. perature, in sterile conditions, using membranes with a MWCO of
In the attempt to select mixtures with optimal cell encapsulation 12.000 Da. The oxidized polysaccharides were freeze-dried, and the
characteristics, the most important task is to perform reliable com- final products were stored under vacuum, at 4 ◦ C, until use.
parisons between hydrogels prepared according to distinct recipes.
Conferring of a “cyto-friendly” character of the embedding medium
2.3. Determination of oxidation degree of the oxidized
is of paramount importance, because “cells feel and respond to the
polysaccharides
stiffness of their substrate” (Discher, Janmey, & Wang, 2005), and
need to be assisted in their duro- and chemo-tactic migration into
A protocol derived from the one described in reference (Tang,
the ECM substitutes, especially under the urge of imminent inter-
Sun, Fan, & Zhang, 2012) was used. Aqueous colloidal solutions of
actions with other cells and with biochemical species (Herzmann,
raw and oxidized gellan and pullulan were analytically prepared
Salamon, Friedler, & Peters, 2016).
at a concentration of 10 ± 0.1 mg/mL, from the corresponding solid
If temporary cell encapsulation is required, the embedding
stock forms, and then were titrimetrically investigated.
hydrogels need to be susceptible to disaggregation under the effect
The number of moles of aldehyde groups per gram of sample
of a cell-protective and easy to apply external “aggression”. In this
was calculated considering the initial mass of polysaccharide, w
respect, we tested the ability of the investigated aK − NaHyal −
(in grams), by using the formula:
OxPolys hydrogels to be either enzymatically dissolved, or ther-
mally liquefied. The method of thermal liquefaction reported in 1 1
CCH=O = · · TNaOH · (v − v0 ) , (1)
this paper is as effective as other techniques of releasing encapsu- w ENaOH
lated cells (Mak et al., 2015; Yildirimer & Seifalian, 2014), but not as
where: ENaOH is the equivalent weight of NaOH (39.997), and
reproducible/predictable as those benefiting from the special and
TNaOH is the titre of the analytical NaOH solution used for titra-
sophisticated synthetic hydrogels (Deshayes & Kasko, 2013). It is,
tion (in g/mL), v0 and v are the volume of 0.1 M NaOH consumed
however, safe for practical encapsulation applications because cells
for the titration of raw and oxidized polysaccharide, respectively.
are able to withstand thermal shocks, commonly in a ratio larger
Theoretically, by halving the calculated value of CCH = O , the num-
than 90% after 5 min heating at about 50 ◦ C. Some of the most exi-
ber of cleaved bonds per gram of oxidized polysaccharide results.
gently conducted experiments revealed none or minimal damages
Therefore, the nondimensional theoretical value of the percentage
when cells were heated up to 48 ◦ C for 2.5 min. (Reissis, García-
degree of oxidation, Dox , is given by:
Gareta, Korda, Blunn, & Hua, 2013), up to 49 ◦ C for several minutes
in microfluidic systems (Ginet et al., 2011), and even near to the 1  
DOX = · CCH=O · Mw · 100, (2)
protein denaturation threshold (60 ◦ C, in pulses of 2 s long) (Dams, 2
Liefde-van Beest, Nuijs, Oomens, & Baaijens, 2010).
were {Mw } is the molecular mass of the repeating unit of the treated
polysaccharide (645 g/mol for gellan, and 486 g/mol in the case of
pullulan). The degree of molecule fragmentation as a consequence
2. Materials and methods
of oxidation process, FDOX , will be:
   
2.1. Materials Mw Mw
DOX 1
FDOX = = · 100 · · CCH=O = k ·
Mw,OxPolys 2 Mw,OxPolys
Type I atelocollagen was obtained according to a patented proce-
dure (Maier, Maier, Prunenanu, & Ignat, 2015), starting from bovine Mw,OxPolys · CCH=O , (3)
368 A. Luca et al. / Carbohydrate Polymers 169 (2017) 366–375

where Mw,OxPolys represents the molecular mass of the oxidized 2.6.3. Direct encapsulation procedure
polysaccharide. The value of FDOX (in moles/g) also offers a mea- If the hydrogels must be prepared in the presence of the cells
sure of the amount of molecules of functionalized polysaccharide to be embedded in, culture medium (Dulbecco’s Modified Eagle’s
bearing aldehyde groups. medium, DMEM D6046, Sigma) is used for aK dilution. Cells are
added at a seeding density of 4·105 cells/mL relative to the final
hydrogel volume just before the solution of oxidized polysaccha-
2.4. Chromatographic characterization of oxidized ride. The compositions are incubated at 37 ◦ C, in an atmosphere
polysaccharides with 5% CO2 and 95% relative humidity.
Table SD-2 (in Supplementary Data) resumes the recipes of the
The average molecular mass and the molar mass dispersity of prepared hydrogels variants.
both raw and oxidized gellan and pullulan were determined by size
exclusion chromatography (SEC), on a PL-GPC 120 Varian system
equipped with refractive index detector and three PL-aquagel-OH 2.7. Evaluation of crosslinking between aK and the oxidized
packed columns (8 ␮m particle size, and 20, 40 and 60 Å pore aver- polysaccharides
age diameters), connected in series. Polysaccharide stock solutions
were diluted with a saline neutral buffer (0.2 M NaNO3 , 0.01 M To put in evidence the covalent crosslinking mediated by the
NaH2 PO4 , pH 7) to a concentration of 0.1 mg/mL. The operating condensation reaction between the carbonyl groups of oxidized
parameters were as follows: the eluent was the same as the dilution polysaccharides and the primary amino groups of lysine in ate-
buffer, flow rate 1 mL/min, injected sample volume 100 ␮L, work- locollagen, the amount of free ␧- and ␣-amino functions was
ing temperature 25 ◦ C. Columns calibration was performed using a quantitatively determined, after gelation completion. The photo-
pullulan-containing standard (P-82, Shodex Denko K.K., Japan, lot colorimetric method based on the reaction of the remnant primary
01101) which includes the following standard molecular weights: amino groups with 2,4,6-trinitrobenzen sulfonic acid (TNBS), was
6.1, 9.6, 21.1, 47.1, 107, 194, 337 and 642 kDa. A five order polyno- used (Fields, 1972; Hermanson, 2008). The final solutions absorban-
mial equation best fitted the calibration curve. cies were measured at 346 nm using a Anthos Zenyth 200rt
microplate reader (Biochrom Ltd., Cambridge, UK). Glycine was
used for method calibration and the crosslinking degree was
2.5. NMR characterization of oxidized gellan estimated as the percentage of free amino groups content of
crosslinked hydrogels (recipe variants R2–R7) against the content
Solid samples of raw and oxidized gellan were dried in a vac- in the hydrogel prepared according to recipe R1.
uum oven (100 mbar, 45 ◦ C, overnight) and dissolved in deuterated
water (D2 O). The solutions were degassed for 24 h, under vacuum,
and were analyzed using a Bruker Avance DRX400NMR spectrom- 2.8. Rheological characterization of prepared hydrogels
eter. 1 H NMR spectra were recorded at a working frequency of
400.13 MHz. Amplitude, frequency and temperature sweep, and also creep-
recovery tests were conducted using a Physica MCR501 rheometer
(Anton Paar) fitted with PP25 parallel plate system. Tests were per-
2.6. Preparation of atelocollagen − oxidized polysaccharides formed for the seven prepared hydrogels, and for a 0.5% alginate
hydrogels hydrogel, considered as reference for the ability to be manipu-
lated in cell encapsulation protocols. The rheometric parameters
2.6.1. Recipes formulation were as follows: (i) shear strain between 0 and 1% in the ampli-
All the recipe variants consider the same final weight tude sweep tests, (ii) angular velocity between 500 and 0.05 rad/s,
of biomacromolecular components into the hydrogel sample under a shear strain of 0.3%, in frequency sweep tests, (iii) tempera-
(1.92 mg), with a constant amount of NaHyal (0.06 mg) as the ture variation between 5 and 65 ◦ C, and 10 Hz oscillation frequency
unit part in the biomacromolecules mixtures. Table 1 presents the in temperature sweep tests, and (iv) 5 Pa shear stress, 750 s in
tested recipe variants. Scheme 1 graphically describes the hydro- loaded followed by 600 s in unloaded state, in creep-recovery tests.
gels preparation protocol. Excepting the temperature sweep analyses, investigations were
conducted at 37 ◦ C.
Two distinct domains were considered for processing creep-
2.6.2. Preparation procedure
recovery data: the long-time creep and the early creep evolution. In
Hydrogel samples were prepared by mixing the macromolecular
accordance, two types of parametric rheological models were fit-
components solutions and some reagents solutions, in a pre-
ted: the classic Burgers model, and the Jeffreys inertial model (both
cise order, in the central well of sterile organ culture dishes (BD
described in Supplementary Data file, Figure SD-1, and equations
FalconTM 353037, 60 × 15 mm). Based on preliminary tests, the pre-
Eq. SD-1 and Eq. SD-2).
scribed volume of aK solution is diluted with a precise volume of
ddH2 O (for the mixed volumes, see Table SD-1, in Supplementary
Data file) (Lefter, Maier, Maier, Popa, & Desbrieres, 2013). After pH 2.9. Proteolytic degradation of the prepared hydrogels
adjustment of the aK solution at 7.0, using 1 M NaOH, this was sta-
bilized with 40 ␮L of 10 x PBS (pH 7.0). Then a constant amount The ability of the prepared hydrogels to be degraded and
(20 ␮L) of NaHyal is added. In a separate sterile glass vessel, the destructured by proteolytic enzymes was tested using bacterial col-
oxidized polysaccharide solution is adjusted to pH 8.0, and then lagenase. A ninhydrin-based assay to quantify the liberated amino
is pregelified with 0.1% v/v of a 1 mg/mL sterile solution of CaCl2 . acids, adapted form reference (Quek et al., 2004), was applied
Immediately after the start of the agglutination process, the pre- to the hydrogels prepared according to recipe variants R1 (not
scribed volume is added to the mixture of aK and NaHyal, under crosslinked) and R4 (with best behavior in the preliminary tests
efficient stirring, then the prepared compositions are incubated at of cell encapsulation).
37 ◦ C, for 30 min. During incubation, the mixture gelifies and turns A collagenase solution (Gibco, Life Technologies, U.S.A., 220
into a perfect translucent matrix, able to be manipulated according units/mg) with a final proteolytic activity of 5 units/mL, and a 4%
to cell culturing protocols. ninhydrin solution in 2-methoxyethanol were used. The final solu-
A. Luca et al. / Carbohydrate Polymers 169 (2017) 366–375 369

Table 1
The amounts of biomacromolecular components in the recipe variants for hydrogels preparation.

Recipe variant Components amounts (mg) Total amount (mg) Parts ratio aK:NaHyal:OxPolys (w/w/w) Total parts in the mixture

aK NaHyal OxPolys

R1 1.86 0.06 0.00 1.92 31:1:0 32


R2 1.77 0.06 0.09 29.5:1:1.5
R3 1.56 0.06 0.30 26:1:5
R4 1.38 0.06 0.48 23:1:8
R5 1.14 0.06 0.72 19:1:12
R6 0.93 0.06 0.93 15.5:1:15.5
R7 0.66 0.06 1.20 11:1:20

Scheme 1. The blueprint of the hydrogels preparing protocol.


With green color are marked the optional steps performed only when cells are embedded into the hydrogel during preparation.

tions absorbencies were determined at 570 nm, using a plate reader 2.11. Scanning electron microscopy
(Tecan, Sunrise). Isoleucine was used for method calibration.
The microscopic morphology of the biomacromolecular matrix
and the accommodation of adipose stem cells to the hydrogel
2.10. Cytotoxicity tests embedding environment were studied by the SEM technique. The
samples of uncultured and cell embedding hydrogels were dried
The cytocompatibility of aK: OxPolys hydrogels prepared at the critical point of carbon dioxide. Briefly, samples were first
according to R4 recipes was proved by MTT cell viability test, fixed (2% glutaric dialdehyde, 12 h), abundantly washed with sterile
performed in triplicate on adipose-derived stem cells, and on pri- ddH2 O, dehydrated through a series of graded ethanol baths (70%,
mary culture fibroblasts. Cells were seeded at a density of 5000 80%, 90% and 100%) and finally dried using an EMS 850 Critical Point
cells/well, in 24 well plates, in DMEM supplemented with 10% BFS Dryer. The dried samples were sputter-coated with gold, prior to
and 1% antibiotic. After 24 h of incubation (37 ◦ C, 5% CO2 ), cells be visualized using a Tescan Vega II SBH electron microscope.
were directly exposed to 50 ␮L of hydrogel composition. As con-
trol, unexposed cells were further cultured. Tests were performed
after 24, 48 and 72 h of culture, in triplicate, using a 5% MTT solution.
370 A. Luca et al. / Carbohydrate Polymers 169 (2017) 366–375

2.12. Fluorescence microscopy

The viability and the morphology of the rabbit adipose-derived


stem cells cultured both in the presence of, and encapsulated in,
the R4 hydrogel were put in evidence by fluorescence microscopy.
After 7 days of culture, the medium was replaced with a solution of
2 ␮M Calcein AM in Hank’s Balanced Salt Solution (HBSS), and left
in contact for 30 min, at 37 ◦ C.
To put in evidence the cell morphology cells were stained
with rhodamine-phalloidin (Molecular Probes, Life Technologies,
U.S.A.), an indicator for actin filaments, and with 4 ,6-diamidino-2-
phenylindole dihydrochloride (DAPI), for nuclei observation. After
solution removal, the cells were visualized with a Leica DMI3000
microscope, using specific filters.

2.13. Release of cells from hydrogels by a short thermal shock

Fibroblasts of primary culture and adipose-derived stem cells


were separately encapsulated in hydrogels prepared according to
R4 recipe, at a seeding density of 1·105 cells/mL. After 72 h of
Fig. 1. The time evolution of the oxidation degree of gellan and pullulan. The isolated
incubation, embedding hydrogels were transferred, in sterile con-
points marked with unfilled symbols represent the values of aldehyde content after
ditions, in a glass Petri dish (60 mm in diameter) preheated at 48 ◦ C two months of storage of the oxidized products.
on a thermoblock (HLC Ditabis), and immediately covered with a
glass plate (20 × 20 × 3 mm) preheated at the same temperature.
After 4 min, the cell suspension was collected, re-suspended in cul- (by using a limited amount of the acidic polysaccharide) and proce-
ture media, and further cultured. After cell adherence, the hydrogel dure (by a distinctive stage of NaHyal admixing), in order to obtain
debris were washed by changing the culture medium. After 48 h macroscopically homogenous hydrogels.
of culture, the cells viability, proliferation, and morphology were
investigated using calcein AM, DAPI and rhodamine-phalloidin, 3.2. Physical-chemical characteristics of oxidized gellan and
respectively. pullulan

2.14. Statistical analysis The oxidation process of polysaccharide with metaperiodate


concluded with a noticeable decrease of the viscosity of gellan solu-
Experimental data (three to five parallel measured values) were tion (due to the decrease of average molar mass from 1486.3 kDa
statistically processed as follows: (i) possible outliers were iden- and 1.79 mass dispersity, Di , to 848.1 kDa and 1.64 Di ), but with
tified and removed using Grubbs test; the removed outliers were only a slight decrease in the case of pullulan (associated with a
replaced by performing supplemental determinations; (ii) differ- molar mass decrease from 85.9 kDa and 1.76 Di , to 50.8 kDa and
ences between the means of experimental data sets were evaluated 2.07 Di ). Also, the intrinsic gelation ability of GellOx significantly
using one-way ANOVA, not assuming equal variances; (iii) where decreased.
applicable, post-hoc test was used to identify the specific data sets The proof of the presence of carbonyl groups on the chain of
which significantly differ; Games-Howell procedure for sets with GellOx was done by analyzing its 1 H NMR spectrum (Figure SD-3,
unequal variances and unequal sample sizes was considered. In in Supplementary Data file) in comparison with the spectrum of
all cases, the evidence against the null hypothesis was declared native low acyl gellan (Figure SD-2). The peak at about 9.5 ppm is
at p < 0.05. characteristic for aldehyde proton. The oxidation process resulted
in some degradation of the polysaccharide chain, which is empha-
3. Results and discussion sized by the abundance of the peaks in the range 3.0–0.5 ppm.
In order to estimate the crosslinking ability of the oxidized
3.1. General remarks polysaccharides, the content of sterically available and reactive
aldehyde groups of the samples was analytically determined
When atelocollagen is mixed in appropriate conditions with through a titrimetric technique (Zhao & Heindel, 1991). Based on
raw or functionalized polysaccharides, two processes evolve con- the evolutions in Fig. 1, the optimal duration of oxidation is con-
currently: coacervation and gelation. The first process generates sidered to be of 45 min for gellan functionalization, and of 75 min
tight associations between collagen and polysaccharides, which for pullulan processing, when the relative yield of the oxidation
may result in co-precipitation. The second leads to extended processes (reported as percentage degree of oxidation) is of about
supramolecular edifices, and concludes with a morphologically 48.96% for GellOx, and of 13.52% for PullOx. The percentage degree
defined network of local aggregates. In addition to coacervation and of oxidation determined after two months of storage (hermeti-
gelation, the collagen microfibrillation continues to take place as a cally sealed, at 4 ◦ C) was of 45.97± 0.11%, in the case of GellOx,
parallel process, supplementary affecting the results of physical- and of 11.81± 0.26% in the case of PullOx. Accordingly, when used
chemical interactions. as crosslinker, the prepared GellOx contains three times as much
The inevitable interferences between the microfibrillation, free aldehyde groups (1.42 ± 0.025 mM/g), as compared with Pul-
coacervation, and gelation, were considered when the recipes lOx (0.47 ± 0.0124 mM/g).
and procedures of preparing aK-NaHyal-OxPolys hydrogels were
designed. While aK microfibrillation must be favored through 3.3. Quantitation of crosslinking degree of the prepared hydrogels
the procedure by slow temperature increase, the associative co-
precipitation of aK and the admixed polysaccharides (NaHyal and The characteristics of the prepared hydrogels are directly influ-
OxPolys) must be prevented and/or controlled through both recipe enced by the degree of crosslinking between the atelocollagen and
A. Luca et al. / Carbohydrate Polymers 169 (2017) 366–375 371

eter a was calculated as being 19.55 for GellOx, and 42.19 for
PullOx. Accordingly, the difference is of 22.64% in the benefit of
PullOx, which means that, in practical terms, the same degree of
crosslinking can be obtained using about 20% less PullOx, than Gel-
lOx, provided the two polysaccharides were functionalized using
the same recipe, and were further post-processed identically.
In correlation with the final remark of Section 3.2, even if, in
abstracto and per se, PullOx should be about four times more active
as crosslinker when compared with GellOx, in the real conditions of
hydrogels preparing it is only 1.2 fold more efficient. Paradoxically,
the measured amounts of aldehyde groups are in opposite rela-
tion (1.42 ± 0.025 mM/g for GellOx, and only 0.47 ± 0.0124 mM/g
in the case of PullOx). The reason of the discordance between
the abundance of reactive groups and the efficacy as crosslinkers
resides in the difference of fragmentation degrees of the function-
alized polysaccharides. The smaller PullOx molecules are prone
to self-disentanglement, and to better entangle with the fibrillary
structured aK, thus maximizing the reciprocal interaction.

Fig. 2. The evolution with temperature of the internal organization of investigated 3.4. Rheological behavior of the prepared hydrogels
hydrogels, as reflected by their complex viscosity. Hydrogels prepared according to
R1 and R4 recipes are compared.
Organoleptic observation made during the maneuverability
tests (like those presented in Figure SD-19) revealed that the R4
the functionalized polysaccharides. The number of covalent bridges hydrogels from the GellOx-crosslinked series, and the R3 hydro-
developed during the preparation is pointed out by the consump- gels from the PullOx-crosslinked series, were the most versatile
tion of the free amino groups of protein partner, and is correlated in responding to various physical-mechanical stresses. In order to
with the mass ratio of oxidized polysaccharide in the mixture recipe substantiate this finding, the variation of gels strength was first
(see Figure SD-4). Figure SD-5 depicts the values of the relative determined, by comparing the ratio between G’ and G” shear mod-
covalent crosslinking degree determined for the hydrogels pre- uli with the value accepted as threshold between the weak and
pared according to the recipe variants presented in Table SD-1, the strong covalently crosslinked gels (which is G’/G” = 10 (Clark
considering the physical crosslinked hydrogel (which only contains & Ross-Murphy, 1987), or, equivalently, G”/G’ = 0.1 (Borzacchiello
aK and NaHyal) as reference for the absence of covalent bridges. & Ambrosio, 2009). Figure SD-6 depicts the linear variation of the
Because of the low volume of PullOx molecules, the mixing with aK conventional strength of the prepared hydrogels in relation with
molecules is more intimate, and crosslinking evolve more rapidly, the recipes details. With a G’/G” ratio of 7.1998, the purely phys-
at smaller amount in recipes. To facilitate the comparison between ically crosslinked aK − NaHyal hydrogel represents the reference
the crosslinking efficacy of GellOx and PullOx, the dependence of in evaluating the increase of hydrogels strength. The increase of
crosslinking degrees on the amount of functionalized polysaccha- the strength of PullOx-based hydrogels is about three times faster
ride was fitted by using a power growth law (y = a·xp ; see Figure in comparison with those of the GellOx-based ones. GellOx- and
SD-5). Then, for the amount related with the recipe of the most PullOx-crosslinked hydrogels from the R4 series produce gels with
versatile hydrogels (which is 1.039 mg/mL OxPolys), the param- relatively similar rheological behavior. At PullOx amounts higher

Fig. 3. Dependence of the values of fitted Burgers models, according the preparing recipes of the investigated hydrogels. Lateral isolated segments correspond to the values
calculated for 5 mg/mL alginate hydrogel.
372 A. Luca et al. / Carbohydrate Polymers 169 (2017) 366–375

than 1 mg/mL, resulting hydrogels become increasingly stronger At 37 ◦ C, the complex viscosity of aK and of 3: 1 aK: Gel-
due to the increasing number of covalent linkages. Compara- lOx hydrogels are quasi-equal, suggesting that the regularity of
tively, all the GellOx-based hydrogels seem to be predominantly their internal organization is the same, being dominated by the
crosslinked by means of physical forces, to the detriment of cova- resulted aK microfibrils. Advantageously, up to about 45 ◦ C, the 3:1
lent bridging (even if chemical crosslinking is present, as confirmed aK:GellOx hydrogel maintains its ordered organization in the par-
by the relative crosslinking degree; see Figure SD-5). For compari- ticular state which was achieved at physiologic temperature. Such
son purpose, the strength of an alginate hydrogel of physical type, an effect could be exploited in cell encapsulation techniques, when
produced in the presence of Ca2+ ions, is also represented in Figure cell release is to be done by a short thermal shock (as it will be
SD-6. described in a further section).
Classical frequency sweep tests were performed to put in evidence Creep-recovery tests were performed to confirm the covalent
the ranges of oscillatory stresses on which elastic responses still crosslinking between the aK and OxPolys, and to evaluate the
occur under low shear strain (see the amplitude sweep data, Fig- parameters describing the viscoelastic behavior of the investigated
ure SD-7). The first column of Figures SD-8 and SD-9 includes the hydrogels. Because all samples exhibited specific “tremor” immedi-
results. All the investigated samples reveal specific behaviors of ately after the constant shear stress was applied, the creep-ringing
stable gels, dominantly elastic (G’ > G” up to 80–100 rad/sec, with formalism was also considered (Ewoldt & McKinley, 2007). Two
a poor dependence on the frequency), and mostly crosslinked by classes of idealized rheomechanical models were fitted: Burgers
physical interactions. At higher oscillation frequency, the trends of model for long term creep, and Jeffreys model for the early time
the two moduli reverse, indicating that the macromolecular matrix period after shear stress application. Figure SD-12 comparatively
of the hydrogels cease to respond elastically (as a solid), but dump presents the creep-recovery curves determined for the hydrogels
the propagation of the shears train, acting as highly viscous fluids. prepared according to the seven recipes. In Figure SD-13, the early
The rheological characteristics of the covalent crosslinked hydro- stage of creep is zoomed, to put in evidence the creep-ringing.
gels with best maneuverability, the alginate-based hydrogel mostly Tables SD-3 to SD-6 systematize the values of the fitted models
used for mammalian cell encapsulation, and the purely physically coefficients, together with the time and rheological constants cal-
crosslinked aK hydrogel are comparatively presented in Figure SD- culated based on these coefficients.
10. Fig. 3 depicts the dependence of Burgers coefficients values on
Temperature sweep tests, included in the second columns of the amount of OxPolys crosslinker used to prepare hydrogels. For
Figures SD-8 to SD-10, allow the evaluation of the variation of comparison, the corresponding values fitted for the alginate hydro-
hydrogels complex viscosity in the range of temperature in which gel (5 mg/mL) are also represented.
cell encapsulation is usually performed (30–40 ◦ C). Complex viscos- Based on the evolution of Burgers parameters with the aK:
ity is proportionally related with the degree of regular organization OxPolys ratio (Fig. 3), optimal recipes can be formulated according
of the macromolecular networks, especially of those physically to different types of constraints imposed to the prepared hydro-
crosslinked (Soares et al., 2014). gels. Ratio values of 3: 1, in the case of using GellOx as crosslinker,
Fig. 2 graphically resumes the results of temperature sweep and of 1.6: 1 in the case of working with PullOx, seem to repre-
tests performed on the most representative hydrogels discussed sent particular compositions of aK − NaHyal − OxPolys hydrogels.
in the present paper. Two peculiarities are evident, both due to the Even if covalent crosslinking bridges were induced, these hydro-
presence of atelocollagen in the gelling mixtures: (i) an increase gels tend to resemble with the pure physical aK hydrogel, from
of structural order in the temperature range of collagen fibrillation the point of view of their rheological behavior. For both lower and
(30–42 ◦ C), and (ii) a liquefaction threshold at about 50 ◦ C. larger ratios, hydrogels are stiffer (larger G1 values), tend to pro-
The supramolecular association of triple-helical aK macro- long the duration of reversible flow in parallel with a slow evolution
molecules is promoted by the slow heating, with a slope of of deformation during this period (larger ␩2 values), also tend to
0.5 ◦ C/min. The presence of GellOx and PullOx has negligible influ- slowly reach the irreversibly flow domain (larger G2 ), and slowly
ence on the process as far as the physical crosslinking between deform irreversibly afterwards (larger ␩1 values). The retardation
polysaccharides and protein remain predominant (up to an aK: time (␭ret ) is the shortest in the case of the above mentioned ratios
OxPolys ratio of about 3:1), but increasingly interfere when cova- (see Table SD-3), but still more than three times longer as compared
lent crosslinking becomes significant (see the series of SD-8 to with the same parameter measured for alginate hydrogel (see Table
SD-10 Figures). In the latter case, a noticeable difference between SD-4).
GellOx and PullOx crosslinking action is revealed. Even if both tend Fitting of Jeffreys mathematical model to the early-creep data
to diminish the supramolecular aggregation of aK, the smaller and allows a finer discrimination between the real rheological char-
more active PullOx reagent delay the aK microfibrillation and the acteristics of hydrogels. Therefore, because of the complicated
protein network generation, mainly due to earlier and more intense dependency between the recipe formulations and the results of
crosslinking. Larger GellOx macromolecules exhibit a lower chem- classical creep measurements, we decided to check whether the
ical reactivity, but entangle in a more efficient manner with aK, general allure and the peculiarities of the curves in Fig. 3 are also
maximizing the physical crosslinking. found for early-creep stages. Of special interest was the signifi-
The hydrogels liquefaction is induced by the thermal denatu- cant decline of elastic moduli for R4/R5 recipes (aK:OxPolys ratio of
ration of atelocollagen, concluded with the complete unfolding of about 3:1). The results are resumed in Figures SD-14 and SD-15 and
helical domains, by helix-coil transition. The liquefaction starting Tables SD-5 and SD-6, confirming the general evolution of Burgers
temperature is practically the same for both pure aK and mixed parameters over the recipes series.
hydrogels, suggesting that, after gelation, OxPolys are unable to A graphical comparison between the creep-recovery curves of
protect the supramolecular aggregation of protein. The slope of liq- the most representative hydrogels is presented in Figure SD-16. The
uefaction process seems to be related with the ratio between the investigated covalent aK − NaHyal − OxPolys hydrogels have inter-
physical and of covalent crosslinkages. For comparison purposes, mediate behavior between two extrema of purely physical ones: (i)
Figure SD-11 also includes the temperature sweep behavior of the alginate-based, which are stiffer, and flow in limited but essentially
alginate hydrogel. In this case, thermal liquefaction does not occur, irreversible amounts, and (ii) protein-based, which are approxi-
but contrary, the complex viscosity increases at temperature val- mately six fold more elastic, and significantly recover more than
ues larger than 32 ◦ C, which supplementary constrict the embedded half of the induced deformation, after stress unloading.
cells.
A. Luca et al. / Carbohydrate Polymers 169 (2017) 366–375 373

imental one, of shorter duration, by applying a thermal shock on


the cell populated macromolecular matrices.

3.7.1. Degradation of hydrogels under collagenolytic attack


The action of collagenase was investigated for three types of
non-populated hydrogels: one prepared according R1 recipe, and
two according R4 recipe, using, respectively, GellOx and PullOx as
crosslinker agents. The amount of solubilized polypeptides of colla-
gen origin during the proteolytic treatment is represented in Figure
SD-17. All three hydrogels are completely destructured after 24 h
of enzyme attack, but during the first eight hours, their behavior
is quite different. The investigated hydrogels partially lose their
encapsulating action after about three hours, and are completely
destructured after more than five hours. Such long durations could
be detrimental for the encapsulated cells, this is why we decided
to capitalize the liquefaction ability of the aK − NaHyal − OxPolys
hydrogels, under short thermal shocks.
Fig. 4. Results of cytotoxicity evaluation, performed on the investigated hydrogels
by using the MTT assay. 3.7.2. Cell release by the liquefaction of aK − OxPolys hydrogels
under thermal shock
Temperature sweep rheological tests revealed that a complete
3.5. Cytotoxicity assays loss of physical integrity of aK − NaHyal − OxPolys hydrogels occur
in a relatively narrow thermal range (between 44 and 50 ◦ C), and
The in vitro cytotoxicity tests were performed according to the in short intervals of time (2–8 min.). The irreversible liquefaction
ISO 10993-5:2009 standard, with the results depicted in Fig. 4. ability was exploited to rapidly release the encapsulated cells, and
Because the viability of adipose stem cells cultured in the presence the effect of the necessary thermal shock on the viability of released
of aK − GellOx hydrogel slightly decreased after 72 h of exposure cells was assessed.
(from 98.75% to 95.72% viability compared to control), the same The viability of released cells was assessed by means of their
hydrogel was further evaluated against fibroblasts of primary cul- biochemical activity and morphologic and ultrastructural charac-
ture. All investigated hydrogels proved to be non-cytotoxic. As a teristics, in comparison with those not thermally exposed. The
general remark, it seems that, in time, hydrogels crosslinked with tested fibroblasts and adipose-derived stem cells preserved their
GellOx tend to limit the cells metabolic activity, in comparison with viability, and kept their proliferative ability unaltered.
those containing PullOx. This effect could be also induced by the Figure SD-18 summarizes the morphological and structural
increased ability of functionalized gellan to retain cations, both by characteristics of heat exposed adipose-derived stem cells, after
complexation or as counter-ions. re-culturing. The microscopic aspect of cells does not differ from
To confirm that the hydrogels do not release cytotoxic com- that of not-exposed ones (see Figure SD-18.a; hematoxylin − eosin
pounds in the culture medium, cells were incubated for seven more staining). The cellular metabolic activity and the integrity of cells
days in the presence of the samples. The incubated cells preserved membrane were put in evidence by calcein staining of cytoplasm
their normal morphology and viability, and proliferated to the cul- (Figure SD-18.b). DAPI (4 ,6-diamidino-2-phenylindole) staining
ture plate surface. indicates that cells nuclei are intact (see Figure SD-18.c), without
nuclear blebbing that usually suggests the start of apoptotic pro-
cesses. Rhodamine −phalloidin staining (Figure SD-18.c) revealed
3.6. Microscopic investigation
that the heat shock did not affected cells integrity, fact confirmed
by the presence and abundance of the intracellular actin filaments.
The morphological details of the hydrogels fibrillary matrix are
Figure SD-20 graphically summarizes the liquefaction tests.
revealed by the SEM images in Figs. 5 a and 6 a. Fibers with mean
diameters of 200 nm compose a dense but little entangled network.
4. Conclusions
The apparent density of the microscopically observed fibrillary net-
work usually increases during the drying protocol we used (at
Atelocollagen − oxidized polysaccharides hydrogels are of ionic-
critical point of CO2 ) (Ulrich, Jain, tanner, MacKay, & Kumar, 2010),
covalent entanglement type. Their biomacromolecular network is
but the gross morphology is preserved (Branco da Cunha et al.,
predominantly based on physical interactions, but covalent bridges
2014). Salts crystals are abundantly present on the samples pre-
consolidate the tri-dimensional edifice, and provide the rheo-
pared using cell culture media. Adipose stem cells in active state
mechanical properties imposed by cell encapsulation protocols.
were visualized at the surface of hydrogels samples and in rupture
The present work investigated the dependence of the char-
areas, as can be observed in Figs. 5b and 6b.
acteristics of aK − NaHyal − OxPolys hydrogels on the recipe
formulation, in order to identify the optimal substrata for cell
3.7. Tests on cells release from the embedding hydrogels embedding. Considering a constant amount of NaHyal in the recipes
series, the hydrogels prepared at an aK: OxPolys ratio of 3: 1
Generally, cell encapsulation aims to physically isolate cells for offered the best correlations between the physical-chemical and
a certain period of time, without altering their viability and spe- rheo-mechanical characteristics.
cific functionality, and frequently, in conditions that they can be In the case of optimal recipes, oxidized gellan with relatively
“quantitatively” released in “active” state, by simple techniques. low amount of free aldehyde groups (1.42 ± 0.025 mM/g) is able to
Therefore, some applications require that cells encapsulation be consolidate the atelocollagen-based hydrogels by ensuring a cova-
reversible. Two feasible methods to extract cells from their host- lent crosslinking yield of about 20%, and the rheological behavior
ing matrices were tested: (i) a conventional one, by slow enzyme of a weak hydrogel (expressed by the G’/G” ratio of about 8). When
erosion/hydrolysis of the embedding hydrogels, and (ii) an exper- compared with the hydrogels crosslinked with oxidized pullulan,
374 A. Luca et al. / Carbohydrate Polymers 169 (2017) 366–375

Fig. 5. GellOx crosslinked hydrogel (Recipe 4). The fibrillary structure and the accommodation of the encapsulated cells.

Fig. 6. PullOx crosslinked hydrogel (Recipe 5). The fibrillary structure and the accommodation of the encapsulated cells.

or with the 5 mg/mL alginate hydrogel, the GellOx crosslinked ones The cells encapsulated in aK − NaHyal − OxPolys hydrogels can
have the most similar rheological characteristics to those prepared be released by applying a short thermal shock (48 ◦ C, 4 min), with-
with aK only, but an increased maneuverability during cell encap- out compromising their viability and their proliferation ability. This
sulation and culturing. effect is based on the liquefaction capacity of the investigated gels,
induced by the irreversible denaturation of aK component.
A. Luca et al. / Carbohydrate Polymers 169 (2017) 366–375 375

Acknowledgements Hashemi, M., & Kalalinia, F. (2015). Application of encapsulation technology in


stem cell therapy. Life Sciences, 143, 139–146.
Hermanson, G. T. (2008). Bioconjugate techniques (2nd ed., pp. 127–128). New
The research works have been supported by a grant of the Roma- York: Academic Press.
nian National Authority for Scientific Research, CNCS-UEFISCDI, Herzmann, N., Salamon, A., Fiedler, T., & Peters, K. (2016). Analysis of migration
project number PNII-ID-PCCE-2011-2-0028. rate and chemotaxis of human adipose-derived mesenchymal stem cells in
response to LPS and LTA in vitro. Experimental Cell Research, 342, 95–103.
Iacovacci, V., Ricotti, L., Menciassi, A., & Dario, P. (2016). The bioartificial pancreas
Appendix A. Supplementary data (BAP): Biological, chemical and engineering challenges. Biochemical
Pharmacology, 100, 12–27.
Janson, I. A., & Putnam, A. J. (2015). Extracellular matrix elasticity and topography:
Supplementary data associated with this article can be found, in Material-based cues that affect cell function via conserved mechanisms.
the online version, at http://dx.doi.org/10.1016/j.carbpol.2017.04. Journal of Biomedical Materials Research, A, 103, 1246–1258.
046. Lefter, C. M., Maier, S. S., Maier, V., Popa, M., & Desbrieres, J. (2013). Engineering
preliminaries to obtain reproducible mixtures of atelocollagen and
polysaccharides. Materials Science and Engineering C, 33, 2323–2331.
References Maier, S. S., Maier, V., & Buciscanu, I. (2010). Novel procedure for large-scale
purification of atelocollagen by selective precipitation. Journal of American
Borzacchiello, A., & Ambrosio, L. (2009). Structure-property relationships in Leather Chemists’ Association, 105, 1–8.
hydrogels. In R. Barbucchi (Ed.), Hydrogels: Biological properties and applications Maier, S.S., Maier, V., Pruneanu, M., & Ignat, C.M. (2015). Procedure for the
(pp. 9–20). Milan: Springer-Verlag. extraction and purification of biologically active atelocollagen, RO Patent
Branco da Cunha, C., Klumpers, D. D., Li, W. A., Koshy, S. T., Weaver, J. C., Chaudhuri, 126403 B1/2015.
O., et al. (2014). Influence of the stiffness of three-dimensional Mak, W. C., Olesen, K., Sivlér, P., Lee, C. J., Moreno-Jimenez, I., Edin, J., et al. (2015).
alginate/collagen-I interpenetrating networks on fibroblast biology. Controlled delivery of human cells by temperature responsive microcapsules.
Biomaterials, 35, 8927–8936. Journal of Functional Biomaterials, 6, 439–453.
Caliari, S. R., Weisgerber, D. W., Grier, W. K., Mahmassani, Z., Boppart, M. D., & Malcor, J. D., Bax, D., Hamaia, S. W., Davidenko, N., Best, S. M., Cameron, R. E., et al.
Harley, B. A. C. (2015). Collagen scaffolds incorporating coincident gradations (2016). The synthesis and coupling of photoreactive collagen-based peptides
of instructive structural and biochemical cues for osteotendinous junction to restore integrin reactivity to an inert substrate, chemically crosslinked
engineering. Advanced Healthcare Materials, 4, 831–837. collagen. Biomaterials, 85, 65–77.
Chan, V., Zorlutuna, P., Jeong, J. H., Kong, H., & Bashir, R. (2010). Three-dimensional Moeinzadeh, S., Shariati, S. R. P., & Jabbari, E. (2016). Comparative effect of
photopatterning of hydrogels using stereolithography for long-term cell physicomechanical and biomolecular cues on zone-specific chondrogenic
encapsulation. Lab on a Chip, 10, 2062–2070. differentiation of mesenchymal stem cells. Biomaterials, 92, 57–70.
Chwalek, K., Dening, Y., Hinüber, C., Brünig, H., Nitschke, M., & Werner, C. (2016). Murphy, S. V., & Atala, A. (2014). 3D bioprinting of tissues and organs. Nature
Providing the right cues in nerve guidance conduits: Biofunctionalization Biotechnology, 32, 773–785.
versus fiber profile to facilitate oriented neuronal outgrowth. Materials Science Quek, C. H., Li, J., Sun, T., Chan, M. L. H., Mao, H. Q., Gan, L. M., et al. (2004).
and Engineering, C, 61, 466–472. Photo-crosslinkable microcapsules formed by polyelectrolyte copolymer and
Clark, A. H., & Ross-Murphy, S. B. (1987). Structural and mechanical properties of modified collagen for rat hepatocyte encapsulation. Biomaterials, 25,
biopolymer gels. In K. Kamide, M. Saito, A. H. Clark, & S. B. Ross-Murphy (Eds.), 3531–3540.
Advances in polymer science (pp. 57–192). Heidelberg: Springer-Verlag. Reissis, Y., García-Gareta, E., Korda, M., Blunn, G. W., & Hua, J. (2013). The effect of
Dams, S. D., de Liefde-van Beest, M., Nuijs, A. M., Oomens, C. W. J., & Baaijens, F. P. T. temperature on the viability of human mesenchymal stem cells. Stem Cell
(2010). Pulsed heat shocks enhance procollagen type I and procollagen type III Research Therapy, 4, 139.
expression in human dermal fibroblasts. Skin Research Technology, 16, 354–364. Soares, P. A. G., Bourbon, A. I., Vicente, A. A., Andrade, C. A. S., Barros, W., Jr., Correia,
Deshayes, S., & Kasko, A. M. (2013). Polymeric biomaterials with engineered M. T. S., et al. (2014). Development and characterization of hydrogels based on
degradation. Journal of Polymer Science, Part A: Polymers Chemistry, 51, natural polysaccharides: Policaju and chitosan. Materials Science and
3531–3566. Engineering C, 42, 219–226.
Discher, D. E., Janmey, P., & Wang, Y. (2005). Tissue cells feel and respond to the Tang, Y., Sun, J., Fan, H., & Zhang, X. (2012). An improved complex gel of modified
stiffness of their substrate. Science, 310, 1139–1143. gellan gum and carboxymethyl chitosan for chondrocytes encapsulation.
Ewoldt, R. H., & McKinley, G. H. (2007). Creep ringing in rheometry or How to deal Carbohydrate Polymers, 88, 46–53.
with oft-discarded data in step stress tests. Rheology Bulletin, 76, 4–9. Ulrich, T. A., Jain, A., Tanner, K., MacKay, J. L., & Kumar, S. (2010). Probing cellular
Fields, R. (1972). The rapid determination of amino groups with TNBS. In C. H. W. mechanobiology in three-dimensional culture with collagen–agarose matrices.
Hirs, & S. N. Timasheff (Eds.), Methods in enxymology, part B −Enzymes structure Biomaterials, 31, 1875–1884.
(pp. 464–468). New York: Academic Press. Xian, W. (2010). A laboratory course in biomaterials. Boca Raton FL: CRC Press
Ginet, P., Montagne, K., Akiyama, S., Rajabpour, A., Taniguchi, A., Fujii, T., et al. [Chapter six, Apendix A, 157–159]
(2011). Towards single cell heat shock response by accurate control on thermal Yildirimer, L., & Seifalian, A. M. (2014). Three-dimensional biomaterial degradation
confinement with an on-chip microwire electrode. Lab on a Chip, 11, − Material choice: Design and extrinsic factor considerations. Biotechnological
1513–1520. Advances, 32, 984–999.
Gong, Y., Wang, C., Lai, R. C., Su, K., Zhang, F., & Wang, D. (2009). An improved Zhao, H., & Heindel, D. N. (1991). Determination ofdegree of substitution of formyl
injectable polysaccharide hydrogel: Modified gellan gum for long-term groups in polyaldehydedextran by the hydroxylamine hydrochloride method.
cartilage regeneration in vitro. Journal of Materials Chemistry, 19, 1968–1977. Pharmaceutical Research, 8, 400–402.

You might also like