Mechanical Properties, Structure, Bioadhesion and Biocompatibility of Pectin

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Mechanical properties, structure, bioadhesion and biocompatibility of pectin

hydrogels

Pavel A. Markov1, Nikita S. Krachkovsky1, Eugene A. Durnev2 , Ekaterina A. Martinson2,

Sergey G. Litvinets2 and Sergey V. Popov1

1
Institute of Physiology, Komi Science Centre, The Urals Branch of the Russian Academy of

Sciences, Syktyvkar, Russia.


2
Department of Biotechnology, Vyatka State University, Kirov, Russia

Corresponding author: Pavel A. Markov, p.a.markov@mail.ru

This article has been accepted for publication and undergone full peer review but has not been
through the copyediting, typesetting, pagination and proofreading process which may lead to
differences between this version and the Version of Record. Please cite this article as an
‘Accepted Article’, doi: 10.1002/jbm.a.36116

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 2 of 32

ABSTRACT

The surface structure, biocompatibility, textural and adhesive properties of calcium

hydrogels derived from 1, 2 and 4% solutions of apple pectin were examined in this study. An

increase in the pectin concentration in hydrogels was shown to improve their stability toward

elastic and plastic deformation. The elasticity of pectin hydrogels, measured as Young’s

modulus, ranged from 6 to 100 kPa. The mechanical properties of the pectin hydrogels were

shown to correspond to those of soft tissues. The characterization of surface roughness in terms

of the roughness profile (Ra) and the root-mean-square deviation of the roughness profile (Rq)

indicated an increased roughness profile for hydrogels depending on their pectin concentration.

The adhesion of AU2% and AU4% hydrogels to the serosa abdominal wall, liver and colon was

higher than that of the AU1% hydrogel. The adhesion of macrophages and the non-specific

adsorption of blood plasma proteins were found to increase as the pectin concentration in the

hydrogels increased. The rate of degradation of all hydrogels was higher in phosphate buffered

saline (PBS) than that in DMEM and a fibroblast cell monolayer. The pectin hydrogel was also

found to have a low cytotoxicity.

Keywords: biomaterials, pectin, hydrogels, texture, surface morphology, elemental

analysis, tissue and macrophage adhesion, protein adsorption, cytotoxicity.

INTRODUCTION

Currently, there is a need to develop biomaterials with novel properties for biomedical

applications, such as drug delivery, tissue engineering, and the creation of implantable devices.

Synthetic and natural polymers are used for the development of these materials. Many studies on

the functional properties of biomaterials were conducted using synthetic polymers to control the

polymer structure and to modify its functional properties. Furthermore, many reports have

demonstrated that the majority of synthetic polymers have a number of shortcomings, such as

high cytotoxicity and low biocompatibility.1-3 The main advantages of natural polymers are

biocompatibility, biodegradability, and low cost.

John Wiley & Sons, Inc. 2

This article is protected by copyright. All rights reserved.


Page 3 of 32 Journal of Biomedical Materials Research: Part A

Pectin is an anionic polysaccharide that constitutes the cell walls of most plants, and it

has been extensively employed in the pharmaceutical and food industries due to its gelling

properties. A wide variety of hydrocolloid pectin-based wound dressings has been patented and

is commercially available. In addition, pectin materials provide improved systems of loading and

releasing drugs (i.e., antibiotics, pain relievers and/or tissue repair factors at the site of action).4

Potential applications of pectin hydrogels have been under intense study in the field of

biomaterials for wound healing, tissue engineering, dentistry, and skin-care products. Pectin

materials were found to have great potential for bone tissue engineering applications because

they promote the nucleation of a mineral phase if they are immersed in adequate physiological

solutions, forming biomimetic constructs that better mimic the natural architecture of the bone.5,
6
Pectin hydrogels may also be used as a barrier material for the prevention of post-operative

adhesions in surgery.7

The presence of a positive or negative charge on the pectin macromolecule determines

the characteristics of the intermolecular interaction of pectin with animal tissues and cells.8-12

However, when pectin hydrogels are formed, functional groups are partially or fully engaged,

thus impeding their ability to participate in further intermolecular interactions. Biofunctional

properties of synthetic hydrogels are assumed to be determined by their surface structural 13, 14, 15

and mechanical properties.16, 17


However, little is known about the influence of the surface

structural and mechanical properties of pectin hydrogels on biocompatibility and biofunctional

properties.

The aim of this study was to evaluate the effect of the structure of pectin hydrogels on

tissues, cells, plasma protein adhesion and biocompatibility in vitro.

MATERIALS AND METHODS

Materials

Low methyl-esterified commercial apple pectin AU701 (Herbstreith & Fox, Germany;

43% degree of methylation), alpha minimum essential medium (MEM, Biolot, Russia),

John Wiley & Sons, Inc. 3

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 4 of 32

Dulbecco’s modifed eagle medium (DMEM, Biolot, Russia), HBSS (Hank’s balanced salt

solution), fetal bovine serum (HyClone), MTT (dimethylthiazol diphenyltetrazolium), 4′,6-

diamidine-2′-phenylindole dihydrochloride (DAPI, Sigma), rhodamine (Sigma), bovine serum

albumin (BSA) and sodium dodecyl sulfate (SDS) were used in the study.

Hydrogel preparation and compression measurements

A volume of 1 ml of a calcium chloride (CaCl2) solution (0.1, 0.5, 1 and 2 M) was added

to the wells of a 12-well plate and frozen at -40°C. Then, 5 ml of a 1, 2 and 4% aqueous pectin

solution (+50°C) was poured into the wells. After the plates were incubated for 24 h, the calcium

solution was removed, and the resultant hydrogel was washed three times with distilled water.

Hydrogel samples (5-mm height, 20-mm diameter) were fixed on the platform of the

texture analyzer (TA-XT plus, Stable Micro Systems, UK). The mechanical characterization of

the hydrogels was performed using a cylindrical aluminum probe P/5 (5-mm diameter). The

samples were equilibrated to room temperature prior to analysis. The determination of the

mechanical properties of the hydrogels was performed at five points for each sample and was

performed for all five samples in each group. The method settings, including the pre-test, test

and post-test speeds, were 2.0, 1.0 and 2.0 mm/s, respectively. The distance (depth of insertion)

was set as 10%, 20%, 30% and 80% of the initial height of the samples. Several mechanical

parameters were extrapolated from the force-time curves: hardness, Young’s modulus, the work

of cohesion (cohesiveness), and the compressive strength of the hydrogels. Young’s modulus

was calculated using the following equation:

E= (F/A)/(∆H/H),

where F/A is the applied force (N) per surface unit and dH/H is the uniaxial deformation.

Compressive strength was calculated as follows:

Qt = F/A,

where F is the force (N) measured during compression and A is the cross-sectional area

of the probe P/5 (0.000019 м2).

John Wiley & Sons, Inc. 4

This article is protected by copyright. All rights reserved.


Page 5 of 32 Journal of Biomedical Materials Research: Part A

Surface morphology and elemental analysis

Dried gel samples were platinum coated under vacuum in an ion sputter for 30 s to render

them electrically conductive. Then, the surface morphology was observed using scanning

electron microscopy (SEM) (JEOL, JSM6510LV, USA) at 15 kV. The elemental analysis of the

cross-sections was performed using energy-dispersive X-ray spectroscopy (EDX).

Surface imaging was also performed using atomic force microscopy (AFM) (NTegra,

NT-MDT, Russia). An analysis of the surface roughness of the hydrogels was performed using

Nova PX image analysis software (Image Analysis-3.2, NT-MDT, Russia). The arithmetic mean

deviation of the roughness profile (Ra) and the root-mean-square deviation of the roughness

profile (Rq) were determined using horizontal profiles with steps of 2 µm (5 profiles for each

scan). The number of samples for each gel was 5. All of the experiments were performed in air

under ambient conditions.

Stability to degradation in vitro

Hydrogel samples (h=1 mm, d=20 mm) were placed in the wells of a 12-well plate and

incubated with 1 ml PBS or DMEM and a monolayer of fibroblast cells (NIH/3T3, concentration

8×104 - 105 cells/ml) in DMEM. The hydrogels were incubated at 37°C with 5% CO2. Hydrogels

were incubated in the indicated media for 14 days with daily replacement of the incubation

solution with fresh solution. The compressive strength of the hydrogels at 2, 4, 7 and 14 days

was evaluated using the above-described method.

Hydrogel adhesion to biological tissue

To evaluate the adhesive properties of the hydrogels, their adhesive force to the serosa of

the abdominal wall, colon and liver of white laboratory mice was measured. Adhesion strength

measurements were performed using a texture analyzer (TA-XT plus, Stable Micro Systems,

UK). A hydrogel sample measuring 5×5 mm was fixed using double-sided tape on a cylindrical

probe p/0.5R with a diameter of 12.7 mm. The adhesion was evaluated by the force of probe

John Wiley & Sons, Inc. 5

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 6 of 32

separation from the tissue after 60 s of pressing with a load of 0.01 N. The force of adhesion was

calculated using Exponent Software (Version V6.1.5.0).

Plasma protein adsorption assay

The structure of the study and the human procedures were approved by the Ethical

Committee of the Komi Science Center of the Russian Academy of Sciences. Blood samples

were obtained from healthy volunteers, and plasma was prepared by centrifuging the blood

samples at 1500 rpm for 10 min and diluting 1×10 with PBS. Diluted plasma (200 µl) was added

to the hydrogel sample (h=1 mm, d=5 mm), and the mixture was incubated for 2 h at 37°C in a

shaker incubator. Then, the hydrogel samples were removed, and the remaining protein fraction

was determined by Lowry's method using a BSA standard.18 The percentage of proteins adsorbed

onto the hydrogels was quantified with respect to the total protein content.

Adhesion of macrophages to the hydrogel surface

J-774 murine macrophages (Collection of the Institute of Cytology, Russian Academy of

Sciences, St. Petersburg, Russia) were cultured in the DMEM that was supplemented with 10%

(v/v) fetal bovine serum, 50 µg/ml gentamicin and 50 µg/ml amphotericin (Sigma-Aldrich). The

cells were incubated at 37°C with 5% CO2.

The hydrogel samples (h=1 mm, d=20 mm) sterilized with UV radiation (1 h) were

placed in a 12-well plate containing 1 ml DMEM supplemented with 10% fetal bovine serum, 50

µg/ml gentamicin and 50 µg/ml amphotericin and were incubated for 1 h at 37°C with 5% CO2.

After the incubation was complete, the medium was replaced with 1 ml of a macrophages

suspension of line J774 (1×106 cell/ml) and incubated at 37°C with 5% CO2 for 24 h. Then, non-

adherent cells were removed by washing the wells with 3 ml of PBS three times. The hydrogel

samples were removed; the cells were fixed on the hydrogels with 2.5% glutaraldehyde for

15 min, washed with PBS three times and were stained with rhodamine and DAPI. Then, the

number of cells was visually counted on each hydrogel or plastic using a fluorescent microscope

John Wiley & Sons, Inc. 6

This article is protected by copyright. All rights reserved.


Page 7 of 32 Journal of Biomedical Materials Research: Part A

(Altami, Russia) equipped with a digital camera (Canon). The number of adherent cells was

counted and was expressed as cells/100 mm2.

Cytotoxicity to fibroblasts

NIH3T3 murine fibroblasts (Collection of the Institute of Cytology, Russian Academy of

Sciences, St. Petersburg, Russia) were cultured in the MEM that was supplemented with 10%

(v/v) fetal bovine serum, 50 µg/ml gentamicin and 50 µg/ml amphotericin. Cells were plated in

96-well culture plates (100 µl, 5×104 cells/ml) and incubated for 12 h at 37°C with 5% CO2 to

allow attachment. After the incubation was complete, the medium was replaced with 100 µl of a

new medium, and the hydrogel samples (h=1 mm, d=5 mm) were added to the wells. The alpha

MEM solution (100 µl) was added to the control wells. An evaluation of the metabolic activity of

the cells was performed at 24, 48 and 96 h.

Cytotoxicity was assessed using the MTT assay; 10 µl of a solution of 5 mg/ml MTT in

PBS was added to each well and incubated at 37°C in air containing 5% CO2 for 4 h in the dark.

MTT was aspirated, 100 µl of solubilization solution (10% SDS/0.01 M HCl) was added, and the

plates were incubated overnight.19 The absorbance was read at 570 nm using a Power Wave-200

(BioTek Instruments, USA) spectrophotometer.

Hemolysis of red blood cells

Human blood samples from healthy donors were collected into vacutainer tubes

containing an anticoagulant. The samples were held for 2 h at room temperature, followed by

centrifugation at 350g for 15 min at 4°C. Subsequently, the red blood cells were removed,

washed 4 times in a 0.9% NaCl solution and diluted 10 fold. The hydrogels were placed in a 96-

well plate and 200 µl of red blood cells was added to each well and incubated for 1 h at 37°C

without stirring. The plate was then centrifuged (100g, 5 min), and 100 µl of the supernatant was

collected from the wells. The optical density (OD) was measured at a 540 nm wavelength using a

microplate reader (Power wave 200, BioTek instruments, USA). The positive control consisted

of blood with distilled water (20 µl), while blood with a 0.9% NaCl solution (20 µl) served as the

John Wiley & Sons, Inc. 7

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 8 of 32

negative control. The level of hemolysis was calculated in accordance with the following

equation 20:

Statistical analysis

The results were presented as the arithmetic mean ± standard deviation (SD). The

significance of the differences between mean values was evaluated using the Student's t-test. The

value of p < 0.05 was considered statistically significant.

RESULTS AND DISCUSSION

Textural properties of the pectin hydrogel

The system hardness (i.e., the maximum positive force required to attain a given

deformation) and Young’s modulus were obtained from the initial slope of the stress–strain

curve at 10% strain. The compressive strength and cohesiveness (i.e., the proportion of the

positive area under the force-time curve from zero to maximum deformation imposed by the

hydrogel) were detected at the hydrogel deformation up to 80%. All these parameters are

reported in Table 1. The hydrogel obtained from a 1% pectin solution had a soft texture and was

incompatible with the strong mechanical loads. The increase in the pectin concentration in the

hydrogel was shown to improve its stability toward elastic and plastic deformation. The elasticity

of pectin hydrogel was measured as Young’s modulus and was found to range from 6 to 100 kPa

(Table 1), which is comparable to the mechanical properties of some soft tissues.

Pectin hydrogels have been used for the reconstruction of bone tissue, as delivery systems

for osteoblasts, osteoclasts, and stem cells and for increasing the biocompatibility of wound

surfaces.21, 8, 5, 22 In this mode of application, the textural properties of pectin hydrogels should

not match the mechanical properties of the bone tissue. The mechanical properties of pectin

hydrogels are comparable to the mechanical properties of some soft tissues, which decreases the

risk of implant rejection because of the mechanical incompatibility.

John Wiley & Sons, Inc. 8

This article is protected by copyright. All rights reserved.


Page 9 of 32 Journal of Biomedical Materials Research: Part A

The elasticity of soft tissues has been recently reported to range from 0.1 to 1 kPa for

brain, 6 to 17 kPa for muscle, and to be approximately 100 kPa for cartilage and osteoids.1, 23, 24

The data obtained suggest that pectin hydrogels are appropriate for implantation into soft tissues

without mechanical incompatibility.

The maximum strength of pectin hydrogels was obtained using a 1 M CaCl2 solution. An

increase in the concentration of CaCl2 greater than 1 M failed to increase the mechanical strength

of the hydrogels (Table 1). Therefore, the pectin hydrogels that were prepared with the 1 M

CaCl2 solution were used in the remaining experiments.

[Table 1 near here]

Surface morphology and elemental analysis

The morphology of the dry gels determined by SEM at an increased magnification (x

5,000) revealed different surface appearances (Figure 1). The surface of the AU1% gel exhibited

smooth structures (Figure 1a). The gel prepared from 2% and 4% pectin solutions exhibited

fibrous and globular structures, respectively (Figure 1b, с). The surface topography of dry gels

determined by SEM was consistent with the results obtained using AFM. 3D AFM imaging of

the hydrogel surfaces revealed a smooth surface for the AU1% hydrogel (Figure 1d) and

globular domains of various sizes on the surface of AU2% and AU4% hydrogels (Figure 1e, f).

The characterization of surface roughness in terms of Ra and Rq indicated an increase in the

roughness profile for the hydrogels, depending on the pectin concentration. The surface of the

AU4% hydrogels exhibited parameters with higher values than those of hydrogels prepared from

1 and 2% pectin solutions (Figure 1d, e, f).

It was found that the surface morphology depends on the sugar composition and

macromolecular structure,25 as well as the polymer composition of the pectin gel.26 The surface

morphology of gels formed from pectin isolated from callus culture cells was dependent on the

conditions of the culture medium.27 The primary topographic elements of the surface of pectin

John Wiley & Sons, Inc. 9

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 10 of 32

gels are globular domains of various sizes. For comparison, the features observed from the gels

of synthetic polymers have a smooth surface.14, 28

The elemental analysis obtained by EDX revealed that the external layers of the dry gels

were composed of oxygen, carbon, chlorine and calcium. The calcium content was found to be

16±1, 14±2 and 13±2 wt % for gels prepared from 1, 2 and 4% pectin solutions, respectively.

Our results showed that the calcium content in the hydrogels was independent of the

pectin concentration in the solution. A previous report 25 demonstrated that the calcium content

in gels prepared from TVF (the pectin of Tanacetum vulgare L.) and CU701 (citrus pectin

CU701, Herbstreith & Fox) was 14.0±1.0 and 12±0.4 wt.%, respectively, indicating that the

calcium concentration in the gel is not dependent on the type of pectin.

[Figure 1 near here]

Hydrogel adhesion to biological tissues

The bioadhesive properties of the pectin hydrogel have not been elucidated. There are

only a few studies describing the bioadhesive quality of pectin or pectin-containing gels.12, 29

However, the design of these studies does not allow for a conclusion regarding the effect of the

structural and mechanical properties on the bioadhesive properties of pectin hydrogels.

The measurement results for hydrogel adhesion to the serosa abdominal wall, liver and

colon of white laboratory mice are shown in Figure 2. The adhesion of the AU2% and AU4%

hydrogels was higher than that of the AU1% hydrogel. The adhesion of the pectin hydrogel to

biological tissues increased as the pectin concentration in the hydrogel increased.

The five theories that are most commonly presented in conjunction with bioadhesion are

the absorption, diffusion, electronic, fracture and wetting theories.30 Data from the literature have

shown that the bioadhesive properties of pectin conform to these theories. The force and pattern

interaction between pectin and colon mucin was shown to be dependent from the amount of

COOH and NH2 groups in the galacturonan backbone.10 The mucoadhesive performance of low

methyl-esterified pectin was significantly lower than that of high methyl-esterified pectin, and

John Wiley & Sons, Inc. 10

This article is protected by copyright. All rights reserved.


Page 11 of 32 Journal of Biomedical Materials Research: Part A

the presence of amide groups in the structure enhanced the mucoadhesion of low methyl-

esterified pectin.9 The maximum adhesive force of rat and mouse peritoneal membranes

increased with increasing esterification of pectins. 11, 12

Our results showed that the values for the force of adhesion of pectin hydrogels to the

serosa abdominal wall, liver and colon are comparable. The adhesion force of hydrogels to

biological tissues increased with an increasing pectin concentration in the hydrogel. In our work,

we used hydrogels formed from low methyl-esterified pectin in which the amount of methyl-

esterified carboxylic groups was 43% of the total carboxylic groups in the galacturonan

backbone.

The Ca2+-pectin gel networks are generated via junction zones whose mechanism of

formation is mainly based on the “egg-box” model in which stretches of the non-methyl-

esterified groups of the galacturonan backbone are ionically cross-linked through Ca2+ bridges.

The methyl-esterified carboxylic groups are not involved in gel formation. We suppose that the

increased adhesion of AU2% and AU4% hydrogels to the tissue caused a change in the ratio of

methyl-esterified carboxylic groups to carboxyl groups, which reduces the negative potential of

the gel.

The low adhesion of the AU1% hydrogel may be explained by the small pectin

concentration in the hydrogel and the minor amount of methyl-esterified carboxylic groups.

Furthermore, the low adhesion of the AU1% hydrogel may be explained by a decrease in the

area of its contact with the serous membrane due to the partial destruction of the hydrogel under

compression. It is unclear why the adhesion force of the AU2% and AU4% hydrogels is

comparable.

Thus, we believe that the adhesion of hydrogels to biological tissues is due to the pectin

concentration in the hydrogels.

[Figure 2 near here]

Adhesion of macrophages to the hydrogel surface

John Wiley & Sons, Inc. 11

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 12 of 32

In the controls, macrophages efficiently colonized on the plastic surfaces (Figure 3a, b).

The number of cells that adhered to the plastic surface was 972±153 cells/100 mm2 after 24 h of

incubation. Approximately 50-60% of adherent cells had a spindle-shaped morphology; the

diameter of the cell body was 24±6 µm (n = 20). Adherent cells were not easily removed by

rinsing, and they exhibited motile phenotypes, often displaying lengthy filopodia. The length of

filopodia of these cells was 47±8 µm (n = 20). In addition, cells with filopodia lengths of 100 µm

were also found (Figure 3b). These results indicate that adherent macrophages are of a

functionally active state.

[Figure 3 near here]

The results shown in Figure 3 indicate that the surface of the pectin hydrogels fails to

support macrophage adhesion and proliferation properties. The number of adherent macrophages

was 9±2, 43±7 and 91±15 cells/100 mm2 on hydrogels obtained from 1, 2 and 4% pectin

solutions, respectively (Figure 3с, e, g). All adherent cells had spherical-shaped morphologies,

with body diameters of 20±4 µm (Figure 3d, f, h). Note that no more than 10% of all

macrophages available from the cell suspension adhered to the hydrogel surface. For

comparison, the poly(ethylene glycol) hydrogel was found to adhere to more than 300

cells/mm2.31

It was previously shown that soft gels have diffuse and dynamic adhesion to epithelial

cells and fibroblasts. In contrast, stiff gels showed cells with stable focal adhesions, which are

typical of those observed in cells attached to glass.17 The stiffness of the gel can influence
16, 32
cellular processes such as adhesion, motility and differentiation, possibly because soft gels

might be perceived as being fluid by the cells, thereby not allowing the cells to show an

anchorage-dependent response. Our results agree with the data in the literature showing that the

adhesion of macrophages to the surface of hydrogels increases as the stiffness increases.

Furthermore, it is believed that the surface roughness and topography affect cell behavior.

A very rough surface can hamper the development of focal adhesion plaques and cell

John Wiley & Sons, Inc. 12

This article is protected by copyright. All rights reserved.


Page 13 of 32 Journal of Biomedical Materials Research: Part A

spreading.13, 14 A comparison of the surface topography and the number of adhered cells showed

that increasing the magnitude of topographic elements of surface pectin hydrogels does not

prevent the attachment of macrophages.

It was previously shown that calcium ions mediate the adhesion of macrophages by

providing ligand binding lectin receptors.33, 34


In our study, hydrogels were formed

instantaneously by ionotropic gelation in which intermolecular crosslinks were formed between

divalent calcium ions and negatively charged carboxyl groups of pectin molecules. We

hypothesized that by increasing the amount of pectin in the hydrogel, the amount of calcium

would increase. However, the results of our energy-dispersive analysis found no difference in the

amount of calcium in the 1, 2 and 4% hydrogels.

The effect of pectin hydrogels on J-774 macrophages in vitro due to

monocyte/macrophage activation was studied and was assumed to represent a key stage of the

foreign body response (FBR).35 It was previously shown that after of subcutaneous implantation

of chitosan36, j-carrageenan and agarose37 hydrogels showed the presence of an inflammatory

infiltrate characterized by the presence of polymorphonuclear neutrophils (PMNs) and

macrophages. A similar reaction is caused by synthetic hydrogels. 38 We found that macrophages

had low adherence to the surface of pectin hydrogels. The adherent macrophages had a spherical

shape, which is not typical of cells in a functionally active state. Perhaps implantation of pectin

hydrogels induced minimal FBR through the inhibition of macrophage activity. The test results

may or may not be reproducible in vivo.

Plasma protein adsorption

The adsorption on the biomaterial surface of proteins, such as albumin, fibrinogen,

fibronectin, vitronectin, globulin, and others, has been known to activate host inflammatory cells,

thus inducing a subsequent FBR.35 Hydrogels obtained from the 1% pectin solution adsorbed

27±2% of blood plasma proteins. Hydrogels obtained from 2 and 4% pectin solutions were

shown to adsorb approximately 40% of blood plasma proteins (Figure 4).

John Wiley & Sons, Inc. 13

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 14 of 32

For comparison, hydrogels prepared by crosslinking sodium alginate with methylene-bis-

acrylamide have been previously reported to adsorb 27% of blood plasma proteins.18 The

molecular mechanisms of high absorption plasma proteins of pectin hydrogels remain unclear.

Note that the adhesion of pectin hydrogels to tissues and the adhesion of proteins to

hydrogels have similarities, possibly due to a reduction in the negative potential of the hydrogel,

thus increasing the adsorption of proteins such as albumin.

[Figure 4 near here]

Stability to degradation in vitro

The in vitro degradation profiles of pectin hydrogels are shown in Figure 5. The rate of

degradation of all hydrogels was higher in PBS than that in DMEM and a fibroblast monolayer.

The strength of hydrogels obtained from 1, 2 and 4% pectin solutions was reduced by 53, 30 and

20%, respectively, after a 48-h incubation in PBS (Figure 5a). The total destruction of hydrogels,

with a decrease of strength by 99%, was detected in the 1% pectin hydrogel after 4 days, whereas

2 and 4% pectin hydrogels disintegrated after 7 and 9 days of incubation in PBS, respectively.

Pectin hydrogels obtained from 1 and 2% pectin solutions were shown to disintegrate at a

very low rate in DMEM. The strength of 1 and 2% pectin hydrogels decreased by 40% after 2

and 7 days of incubation in DMEM, respectively. The mechanical properties of hydrogels

obtained from the 4% pectin solution did not change after 14 days of incubation in DMEM

(Figure 5b). The degradation behavior of pectin hydrogels in the fibroblast monolayer was

similar to that in DMEM (Figure 5c). To study hydrogel biodegradation, the hydrogels were

incubated with a monolayer of fibroblasts. In this study, we evaluated the effect of cellular

metabolites (ROS and hydrolytic enzymes) extracted by cells on the mechanical properties of the

hydrogels.

The data obtained demonstrated that hydrogel degradation was higher in saline than that

in DMEM. In PBS, the possible reason for hydrogel disintegration is due to the displacement of

calcium ions that crosslink the hydrogel matrix with sodium, potassium and phosphate ions,

John Wiley & Sons, Inc. 14

This article is protected by copyright. All rights reserved.


Page 15 of 32 Journal of Biomedical Materials Research: Part A

along with the preferential dissolution of hydrophilic polymers.39 The low rate of disintegration

of the pectin hydrogels in the DMEM and fibroblast monolayer is due to protein adhesion on the

hydrogel substrate. Both the DMEM and the fibroblast culture medium contain fetal bovine

serum. The serum proteins may adhere to the surface of the hydrogel and can impede the elution

of calcium ions.

[Figure 5 near here]

Cytotoxicity to fibroblasts

Cytotoxicity can be rated based on cell viability relative to controls, where an activity

level relative to controls of less than 30% is severe cytotoxicity, between 30 and 60% is

moderate cytotoxicity, between 60 and 90% is slight cytotoxicity, and greater than 90% is no

cytotoxicity. 40, 41

Figure 6 shows the results of fibroblast MTT assays after 24, 48 and 96 h of incubation

with pectin hydrogels. A reduction of cell viability compared to the control was detected for

hydrogels obtained from 1 and 2% pectin solutions after 48 h of incubation; these were

considered to have slight cytotoxicity. The cell viability recovered after 96 h of incubation with 1

and 2% pectin hydrogels, as shown in Figure 6. The effect of the 4% pectin hydrogel on

fibroblast viability was not time-dependent. The data obtained demonstrate that hydrogels from

pectin appear to possess low cytotoxicity, similar to hydrogels from other natural

polysaccharides such as alginate, carrageenan and chitosan.37, 42, 43


The slight cytotoxicity of

pectin hydrogels may explain, at least partially, the reduction in the functional activity of

macrophages.

[Figure 6 near here]

Hemolysis of red blood cells

During contact with red blood cells, certain materials can cause cell degradation and the

release of hemoglobin. Hemolysis levels were evaluated by measuring the change in the optical

density of the solution as a result of the release of hemoglobin from red blood cells after

John Wiley & Sons, Inc. 15

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 16 of 32

membrane destruction. The results of this test show whether a material is able to cause

hemolysis. The level of red blood cell hemolysis while interacting with pectin hydrogels is

shown in Table 2. The optical densities of the positive and negative controls were 3.8 and 0.04,

corresponding to 100 and 0% levels of hemolysis, respectively. The level of hemolysis during

the interaction with hydrogels ranged from 4 to 5.5%. These values are within the set range of

the standard ISO 10993-4: 2002.42 These results show that the developed pectin hydrogels cause

virtually no damage to the erythrocyte membrane during contact. Therefore, these hydrogels are

hemocompatible.

[Table 2 near here]

CONCLUSIONS

In our research, the biofunctional properties of pectin hydrogels with differing textural

and structural characteristics were investigated. An increase in the pectin concentration in

hydrogels was shown to enhance their cohesion, Young’s modulus, compressive strength and

stiffness. An increased concentration of pectin causes an increase in the roughness profile

surface of the pectin hydrogel.

The biofunctional properties of the pectin hydrogel, namely adhesion to serosa,

macrophage adherence, adhesion of plasma proteins and time of biodegradation, were

determined by the pectin concentration in the hydrogel. The correlation coefficient (Pearson's, r)

between the pectin concentration and these variables was calculated as 0.87-0.98. The surface of

the pectin hydrogel prevents adherence of the macrophages and inhibits their transformation to a

functionally active state. This finding indicates a potentially low immunogenicity of pectin

hydrogel-based implants. The textural properties of the pectin hydrogels were demonstrated to

correspond to those of soft tissues. The pectin hydrogel was also found to be hemocompatible in

vitro and to have a low cytotoxicity to fibroblast cells.

In biomedical applications, pectin hydrogels are primarily used as drug delivery systems,

such as those for pharmaceutical compounds 45, 46 and probiotics.47 In tissue engineering, pectin

John Wiley & Sons, Inc. 16

This article is protected by copyright. All rights reserved.


Page 17 of 32 Journal of Biomedical Materials Research: Part A

21
hydrogels have been applied for the isolation of wound surfaces , as a barrier to prevent

postoperative adhesion, 48, 7 as an extracellular matrix for reconstructing bone tissue 5, 22 and as a

coating for hard tissue implants.8 To prepare hydrogels, low methyl-esterified pectins have been

used. To modify the biofunctional properties of hydrogels, pectins with a similar composition to

natural and synthetic polymers have been used. 39, 46

Our results showed that pectin hydrogels can be used as an extracellular matrix for the

reconstruction of soft tissue. The surface of the pectin hydrogel can inhibit the adherence of

macrophages, indicating a potentially low immunogenicity for scaffolds derived from pectin

hydrogels. Furthermore, a simple method can be used to modify the mechanical properties and

biocompatibility of pectin hydrogels. All these qualities represent an advantage for using pectin

hydrogels as a scaffold for tissue engineering.

Declaration of conflicting interests

The author(s) declared no potential conflicts of interest with respect to the research,

authorship, and/or publication of this article.

Funding

The study was supported the Program of UD RAS, project 15-3-4-38.

John Wiley & Sons, Inc. 17

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 18 of 32

REFERENCES

1. Santos E, Pedraz JL, Hernandez RM, Orive G. Therapeutic cell encapsulation: Ten steps

towards clinical translation. J Control Release 2013; 170: 1–14.

2. Chen Q, Lianga S, Thouas A. Elastomeric biomaterials for tissue engineering. Prog

Polym Sci 2013; 38: 584– 671.

3. de Vos P, Lazarjani H, Poncelet D, Faas M. Polymers in cell encapsulation from an

enveloped cell perspectives. Adv Drug Deliver Rev 2014; 67–68: 15–34.

4. Munarin F, Tanzi MC, Petrini P. Advances in biomedical applications of pectin gels. Int

J Biol Macromol 2012; 51: 681–689.

5. Jahromi SH, Grover LM, Paxton JZ, et al. Degradation of polysaccharide hydrogels

seeded with bone marrow stromal cells. J Mech Behave Biomed 2011; 1157-1166.

6. Amirian J, Linh N, Min YK., LeeB-T. Bone formation of a porous gelatin-pectin-

biphasic calcium phosphate composite in presence of BMP-2 and VEGF. Int J Biol

Macromol 2015; 76: 10–24.

7. Popov SV, Popova GYu, Nikitina IR, Markov PA, Latkin DS, Golovchenko VV, Patova

OA, Krachkovsky N, Smirnov VV, Istomina EA, Shumikhin KV, Burkov AA, Martinson

EA, Litvinets SG. Injectable hydrogel from plum pectin as a barrier for prevention of

postoperative adhesion. J Bioact Compat Polym 2016; 1: 1–17.

8. Kokkonen H, Niiranen H, Schols HA, Morra M, Stenback F Tuukkanen J. Pectin-coated

titanium implants are well-tolerated in vivo. J Biomed Mater Res 2010; 93A: 1404–1409.

9. Thirawong N, Nunthanid J, Puttipipatkhachorn S, Sriamornsak P. Mucoadhesive

properties of various pectins on gastrointestinal mucosa: An in vitro evaluation using

texture analyzer. Eur J Pharm Biopharm 2007; 67: 132–140.

10. Liu LS, Fishman ML, Hicks KB, Kende M. Interaction of various pectin formulations

with porcine colonic tissues. Biomat 2005; 26: 5907–5916.

John Wiley & Sons, Inc. 18

This article is protected by copyright. All rights reserved.


Page 19 of 32 Journal of Biomedical Materials Research: Part A

11. Miyazaki S, Kawasaki N, Nakamura T, Iwatsu M, Hayashi T, Hou W-M., Attwood D.

Oral mucosal bioadhesive tablets of pectin and HPMC: in vitro and in vivo evaluation Int

J Pharm 2000; 204: 127–132.

12. Takeda C, Takahashi Y, Seto I, Kawano G, Takayama K, Onishi H, Machida Y.

Influence of pectins on preparation characteristics of lactoferrin bioadhesive tablets.

Chem Pharm Bull 2007; 55: 1164—1168.

13. Stevens MM and George JH. Exploring and engineering the cell surface interface.

Science 2005; 310: 18-25.

14. Orchel A, Jelonek K, Kasperczyk J, Dobrzynski P, Marcinkowski A, Pamula E, Orchel J,

Bielecki I, Kulczycka A. The influence of chain microstructure of biodegradable

copolyesters obtained with low-toxic zirconium initiator to in vitro biocompatibility.

Biomed Res Int 2013; 1-12.

15. Thapa A, Webster TJ, Haberstroh KM. Polymers with nano-dimensional surface features

enhance bladder smooth muscle cell adhesion. J Biomed Mater Res 2003; 67A: 1374 –

1383.

16. Blakney AK, Swartzlander MD, Bryant SJ. The effects of substrate stiffness on the in

vitro activation of macrophages and in vivo host response to poly(ethylene glycol)-based

hydrogels. J Biomed Mater Res A 2012; 100: 1375–1386.

17. Discher DE, Janmey P, Wang Yu. Tissue cells feel and respond to the stiffness of their

substrate. Science 2005; 310: 1139-43.

18. Thankam F, Muthu J. Influence of physical and mechanical properties of amphiphilic bio

synthetic hydrogels on long-term cell viability. J Mech Behave Biomed 2014; 35: 111–

122.

19. Yang C, Xu L, Zhou Y, Zhang X, Huang X, Wang M, Han Y, Zhai M, Wei S, Li J. A

green fabrication approach of gelatin/CM-chitosan hybrid hydrogel for wound healing.

Carbohydr Polym 2010; 82: 1297–1305.

John Wiley & Sons, Inc. 19

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 20 of 32

20. Seo KH, You SJ, Chun HJ, Who CK, Lee K, Lim YM. In Vitro and In Vivo

Biocompatibility of γ -ray Crosslinked Gelatin-Poly-(vinyl Alcohol) Hydrogels. Tissue

Eng Regen Med 2009; 6: 414–8.

21. Boateng JS, Matthews KH, Stevens H, Eccleston GM. Wound healing dressings and drug

delivery systems: a review. J Pharm Sci 2008; 97: 2892–2923.

22. Munarin F, Guerreiro SG, Grellier MA, Tanzi MC, Barbosa MA, Petrini P and Granja

PL. Pectin-based injectable biomaterials for bone tissue engineering. Biomacromolecules

2011; 12: 568–577.

23. Engler A, Sen S, Sweeney H, Discher DE. Matrix elasticity directs stem cell lineage

specification. Cell 2006; 126: 677–689.

24. Discher D, Mooney D, Zandstra P. Growth factors, matrices, and forces combine and

control stem cells. Science 2009; 324: 1673–1681.

25. Popov SV, Markov PA, Patova OA, Vityazev FV, Bakutova LA, Borisenkov MF,

Martinson EA, Ananchenko BA, Durnev EA, Burkov AA, Litvinets SG. In vitro

gastrointestinal-resistant pectin hydrogel particles for β-glucuronidase adsorption. J

Biomater Sci Polym Edit 2017; 28 (3): 293-311.

26. Borisenkov MF, Karmanov AP, Kocheva LS, Markov PA, Istomina EI, Bakutova LA,

Litvinets SG, Martinson EA, Durnev EA, Vityazev FV & Popov SV. Adsorption of β-

glucuronidase and estrogens on pectin/lignin hydrogel particles. Int J Polym Mater 2016;

65: 433-441.

27. Gunter EA, Popeyko OV, Markov PA, Martinson EA, Litvinets SG, Durnev EA, Popov

SV, Ovodov YS. Swelling and morphology of calcium pectinate gel beads obtainedfrom

Silene vulgaris callus modified pectins. Carb Polym 2014; 103: 550– 557.

John Wiley & Sons, Inc. 20

This article is protected by copyright. All rights reserved.


Page 21 of 32 Journal of Biomedical Materials Research: Part A

28. Subramani K, Birch MA. Fabrication of poly(ethylene glycol) hydrogel micropatterns

with osteoinductive growth factors and evaluation of the effects on osteoblast activity and

function. Biomed. Mater 2006; 1: 144–154.

29. Konovalova MV, Markov PA, Durnev EA, Kurek DV, Popov SV, Varlamov VP

Preparation and biocompatibility evaluation of pectin and chitosan cryogels for

biomedical application. J Biomed Mat Res A 2017; 105(2): 547-556.

30. Edsman K, Hagerstrom H. Pharmaceutical applications of mucoadhesion for the non-oral

routes. JPP 2005; 57: 3–22.

31. Waldeck H, Wang X, Joyce E, Kao WJ. Active leukocyte detachment and

apoptosis/necrosis on PEG hydrogels and the implication in the host inflammatory

response. Biomat 2012; 33: 29-37.

32. Fereol S, Fodil R, Labat B, Galiacy S, Laurent,VM, Louis B, Isabey D, Planus E.

Sensitivity of alveolar macrophages to substrate mechanical and adhesive properties. Cell

Motil Cytoskel 2006; 10: 1–20.

33. Lis H. and Sharon N. Lectins: carbohydrate-specific proteins that mediate cellular

recognition. Chem Rev 1998; 98: 637-674.

34. Cummings RD, McEver RP. C-type Lectins. In: Varki A, Cummings RD, Esko JD, et al.,

editors. Essentials of Glycobiology. 2nd edition. Cold Spring Harbor (NY): Cold Spring

Harbor Laboratory Press; 2009. Chapter 31.

35. Anderson JM, Rodriguez A, Chang DT. Foreign body reaction to biomaterials. Semin

Immunol 2008; 20: 86–100.

36. Azab AK, Doviner V, Orkin B, Kleinstern J, Srebnik M, Nissan A, Rubinstein A.

Biocompatibility evaluation of crosslinked chitosan hydrogels after subcutaneous and

intraperitoneal implantation in the rat. J Biomed Mater Res 2007; 83A: 414–422.

John Wiley & Sons, Inc. 21

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 22 of 32

37. Popa E, Carvalho P, Dias A, Santos TC, Santo VE, Marques AP, Viegas CA, Dias IR,

Gomes ME, Reis RL. Evaluation of the in vitro and in vivo biocompatibility of

carrageenan-based hydrogels. J Biomed Mater Res Part A 2014; 102: 4087–4097.

38. Park JH and Bae YH. Hydrogels based on poly(ethylene oxide) and poly(tetramethylene

oxide) or poly(dimethyl siloxane). In vivo biocompatibility and biostability. J Biomed

Mater Res 2003; 64A: 309–319.

39. Ninan N, Muthiah M, Park IK, Elain A, Thomas S, Grohens Y. Pectin/carboxymethyl

cellulose/microfibrillated cellulose composite scaffolds for tissue engineering. Carbohydr

Polym 2013; 98: 877–885.

40. Lönnroth E. Toxicity of medical glove materials. Int J Occup Saf Ergon 2005; 11: 131-

139

41. Barrioni B, Maria S, Orefice R, Oliveira AAR, Pereira MM. Synthesis and

characterization of biodegradable polyurethane films based on HDI with hydrolyzable

cross linked bonds and a homogeneous structure for biomedical applications. Mater Sci

Eng C 2015; 52: 22–30.

42. Morais DS, Rodrigues MA, Lopes MA, Coelho MY, Maurıcio AC, Gomes R, Amorim I,

Ferraz MP, Santos JD, Botelho CM. Biological evaluation of alginate-based hydrogels,

with antimicrobial features by Ce(III) incorporation, as vehicles for a bone substitute. J

Mater Sci: Mater Med 2013; 24: 2145–2155.

43. Kumar PTS, Ramya C, Jayakumar R, Nair SV, Lakshmanan VK. Drug delivery and

tissue engineering applications of biocompatible pectin–chitin/nano CaCO3 composite

scaffolds. Colloids Surf B Biointerfaces 2013; 106: 109–116.

44. ISO 10993-4:2002. Biological evaluation of medical devices—Part 4: Selection of tests

for interactions with blood.

45. Liu LS, Fishman ML, Kost J, Hicks KB. Pectin-based systems for colon-specific

drugdelivery via oral route. Biomaterials 2003; 24: 3333–3343.

John Wiley & Sons, Inc. 22

This article is protected by copyright. All rights reserved.


Page 23 of 32 Journal of Biomedical Materials Research: Part A

46. Noreen A, Nazlic ZH, Akram J, Rasul I, Mansha A, Yaqoo N, Iqbal R, Tabasum S,

Zuber M, Zi KM. Pectins functionalized biomaterials; a new viable approach for

biomedical applications: A review. Int J Biol Macromol 2017; 101: 254–272

47. Dafe A, Etemadi H, Dilmaghani A, Mahdavinia GR. Investigation of pectin/starch

hydrogel as a carrier for oral delivery of probiotic bacteria Int J Biol Macromol 2017;

97: 536–543.

48. Giusto G, Vercelli C, Iussich S, Audisio A, Morello E, Odore R, Gandini M. A pectin-

honey hydrogel prevents postoperative intraperitoneal adhesions in a rat model. BMC

Vet Res 2017; 13:55.

John Wiley & Sons, Inc. 23

This article is protected by copyright. All rights reserved.


Journal of Biomedical Materials Research: Part A Page 24 of 32

Captions

FIGURE 1. Scanning electron micrographs (a, b, c) and 3D height AFM images (d, e, f) of the

surface AU1% (a, d), AU2% (b,e) and AU4% (c,f) pectic gels. Ra and Rq are represented as the

arithmetic mean ± SD (n = 5).

FIGURE 2. The adhesive properties of hydrogels. Data are presented as the arithmetic mean ±

SD (n = 6). * – p < 0.05 compared with AU1% hydrogel.

FIGURE 3. Microscopic images of J-774 macrophage adherent on the plastic (a, b) and pectin

hydrogel (c-h) surface after 24 h of incubation; a, c, e, g: phase contrast photomicrographs. Cell

number adhered on the plastic was equal to 972±153 cell/100 mm2 (×200); b, d, f, h:

corresponding fluorescent micrographs (×400), arrows indicate cell filopodia. C, d - AU1%; e, f

- AU2% and g, h - AU4%.

FIGURE 4. The adsorption of blood plasma proteins by hydrogels. Data shown are the

arithmetic mean ± SD; n=9. * – p < 0.05 compared with AU1% hydrogel.

FIGURE 5. In vitro degradation profiles of pectins hydrogels during incubation in PBS (a),

DMEM (b) and cell monolayer (c). Data shown are the arithmetic mean ± SD, n=9. * - p < 0.05

compared to zero point.

FIGURE 6. Viability of the NIH3T3 fibroblast cells in the presence of the pectin hydrogels.

Data shown are the arithmetic mean ± SD; n=9. a - p < 0.05 compared to 24 h, b - p < 0.05

compared to 48 h.

John Wiley & Sons, Inc. 24

This article is protected by copyright. All rights reserved.


Page 25 of 32 Journal of Biomedical Materials Research: Part A

TABLE 1. Mechanical properties of pectin hydrogels

Young Compressive
Samples Cohesion, mJ Hardness, mN
modulus, kPa strength, kPa
AU 1%
0.1 М 0.26±0.0 6±1 15±3 34±2
0.5 М 1 8±1 33±4 45±3
1М 0.81±0.1 10±1 32±4 58±5
2М 8 9±1 34±4 50±7
0.79±0.0
6
0.82±0.1
3
AU 2%
0.1 М 1.32±0.1 13±2 39±3 75±10
0.5 М 9 28±2 96±3 165±12
1М 3.35±0.1 23±2 104±6 127±17
2М 9 30±3 106±7 171±16
3.64±0.2
3
3.56±0.3
3
AU 4%
0.1 М 1.69±0.4 20±3 51±8 118±23
0.5 М 1 70±15 204±20 424±70
1М 6.78±0.7 101±1 219±15 598±66
2М 7 1 236±10 570±67
8.34±0.6 96±11
2
9.31±0.6
9

Data shown are the arithmetic mean ± SD, n = 5.

John Wiley & Sons, Inc.


This article is protected by copyright. All rights reserved.
Journal of Biomedical Materials Research: Part A Page 26 of 32

FIGURE 1. Scanning electron micrographs (a, b, c) and 3D height AFM images (d, e, f) of the surface AU1%
(a, d), AU2% (b,e) and AU4% (c,f) pectic gels. Ra and Rq are represented as the arithmetic mean ± SD (n
= 5).

119x117mm (300 x 300 DPI)

John Wiley & Sons, Inc.


This article is protected by copyright. All rights reserved.
Page 27 of 32 Journal of Biomedical Materials Research: Part A

The adhesive properties of hydrogels. Data are presented as the arithmetic mean ± SD (n = 6). * – p <
0.05 compared with AU1% hydrogel.

68x43mm (600 x 600 DPI)

John Wiley & Sons, Inc.


This article is protected by copyright. All rights reserved.
Journal of Biomedical Materials Research: Part A Page 28 of 32

FIGURE 3. Microscopic images of J-774 macrophage adherent on the plastic (a, b) and pectin hydrogel (c-h)
surface after 24 h of incubation; a, c, e, g: phase contrast photomicrographs. Cell number adhered on the
plastic was equal to 972±153 cell/100 mm2 (×200); b, d, f, h: corresponding fluorescent micrographs
(×400), arrows indicate cell filopodia. C, d - AU1%; e, f - AU2% and g, h - AU4%.

119x180mm (300 x 300 DPI)

John Wiley & Sons, Inc.


This article is protected by copyright. All rights reserved.
Page 29 of 32 Journal of Biomedical Materials Research: Part A

FIGURE 4. The adsorption of blood plasma proteins by hydrogels. Data shown are the arithmetic mean ±
SD; n=9. * – p < 0.05 compared with AU1% hydrogel.

67x46mm (600 x 600 DPI)

John Wiley & Sons, Inc.


This article is protected by copyright. All rights reserved.
Journal of Biomedical Materials Research: Part A Page 30 of 32

FIGURE 5. In vitro degradation profiles of pectins hydrogels during incubation in PBS (a), DMEM (b) and cell
monolayer (c). Data shown are the arithmetic mean ± SD, n=9. * - p < 0.05 compared to zero point.

130x249mm (600 x 600 DPI)

John Wiley & Sons, Inc.


This article is protected by copyright. All rights reserved.
Page 31 of 32 Journal of Biomedical Materials Research: Part A

FIGURE 6. Viability of the NIH3T3 fibroblast cells in the presence of the pectin hydrogels. Data shown are
the arithmetic mean ± SD; n=9. a - p < 0.05 compared to 24 h, b - p < 0.05 compared to 48 h.

78x56mm (600 x 600 DPI)

John Wiley & Sons, Inc.


This article is protected by copyright. All rights reserved.
Journal of Biomedical Materials Research: Part A Page 32 of 32

TABLE 2. Hemolysis of red blood cells

Samples Optical densities, (545 nm) Hemolysis , %

DW (Positive) 3.7±0.85 100±22

0.9 % NaCl 0.04±0.01 1±0.18

(Negative)

AU1% 0.15±0.024 4.0±0.5

AU2% 0.19±0.031* 4.8±0.7*

AU4% 0.21±0.012* 5.5±0.3*

Data shown are the arithmetic mean ± SD, n=9. * - p < 0.05 compared to AU1%. DW –

distilled water.

John Wiley & Sons, Inc.


This article is protected by copyright. All rights reserved.

You might also like