26
Elsevier Editorial System(tm) for Acta Biomaterialia Manuscript Draft Manuscript Number: Title: Films of dextran-graft-polybutylmethacrylate to enhance endothelialization of materials Article Type: Full Length Article Keywords: Polybutylmethacrylate; Dextran; Copolymer films; Cell proliferation; Endothelial cell migration; Tissue Engineering Corresponding Author: Dr Didier Letourneur, Corresponding Author's Institution: Inserm First Author: Sidi Mohamed Derkaoui, PhD Order of Authors: Sidi Mohamed Derkaoui, PhD; thierry Avramoglou, PhD; Virginie Gueguen, PhD; Christel Barbaud, PhD; Didier Letourneur, PhD Abstract: We have synthesized new structures obtained from amphiphilic copolymers of dextran and polybutylmethacrylate with the aim of endothelialization of biomaterials. Grafting of butylmethacrylate onto dextran has been carried out using ceric ammonium nitrate as initiator. Three copolymers were obtained (11, 30 and 37 wt% dextran) and homogeneous thin films were successfully prepared. In contrast to dextran, the resulting films were stable in water, and copolymers characterized by FTIR, DSC and DMA showed evidence of hybrid properties between the parent homopolymers. Surfaces of films were smooth when analyzed by AFM (roughness 2 +/-1 nm) but greatly differed in their hydrophilicity by increasing the dextran content (water contact angle from 99° to 57°). In contrast to polybutylmethacrylate where the proliferation of vascular smooth muscle cells (VSMCs) was excellent but very low for endothelial cells, the copolymer containing 11% of dextran was excellent for endothelial cells and very limited for VSMCs. An in vitro wound assay demonstrated that copolymer with 11% dextran is even more favorable for EC migration than tissue-culture polystyrene. Increasing the dextran content in the copolymers decreased the proliferation for both vascular cell types. Altogether, these results evidenced that transparent and water-insoluble films made from copolymers of dextran and butylmethacrylate copolymers with an appropriate composition could enhance endothelial cell proliferation and migration. Therefore, a potential benefit of this approach is the availability of surfaces with tunable properties for endothelialization of materials.

Dextran

Embed Size (px)

DESCRIPTION

Dextran article

Citation preview

Page 1: Dextran

Elsevier Editorial System(tm) for Acta Biomaterialia Manuscript Draft Manuscript Number: Title: Films of dextran-graft-polybutylmethacrylate to enhance endothelialization of materials Article Type: Full Length Article Keywords: Polybutylmethacrylate; Dextran; Copolymer films; Cell proliferation; Endothelial cell migration; Tissue Engineering Corresponding Author: Dr Didier Letourneur, Corresponding Author's Institution: Inserm First Author: Sidi Mohamed Derkaoui, PhD Order of Authors: Sidi Mohamed Derkaoui, PhD; thierry Avramoglou, PhD; Virginie Gueguen, PhD; Christel Barbaud, PhD; Didier Letourneur, PhD Abstract: We have synthesized new structures obtained from amphiphilic copolymers of dextran and polybutylmethacrylate with the aim of endothelialization of biomaterials. Grafting of butylmethacrylate onto dextran has been carried out using ceric ammonium nitrate as initiator. Three copolymers were obtained (11, 30 and 37 wt% dextran) and homogeneous thin films were successfully prepared. In contrast to dextran, the resulting films were stable in water, and copolymers characterized by FTIR, DSC and DMA showed evidence of hybrid properties between the parent homopolymers. Surfaces of films were smooth when analyzed by AFM (roughness 2 +/-1 nm) but greatly differed in their hydrophilicity by increasing the dextran content (water contact angle from 99° to 57°). In contrast to polybutylmethacrylate where the proliferation of vascular smooth muscle cells (VSMCs) was excellent but very low for endothelial cells, the copolymer containing 11% of dextran was excellent for endothelial cells and very limited for VSMCs. An in vitro wound assay demonstrated that copolymer with 11% dextran is even more favorable for EC migration than tissue-culture polystyrene. Increasing the dextran content in the copolymers decreased the proliferation for both vascular cell types. Altogether, these results evidenced that transparent and water-insoluble films made from copolymers of dextran and butylmethacrylate copolymers with an appropriate composition could enhance endothelial cell proliferation and migration. Therefore, a potential benefit of this approach is the availability of surfaces with tunable properties for endothelialization of materials.

Page 2: Dextran

1

Films of dextran-graft-polybutylmethacrylate to enhance endothelialization of materials

Sidi Mohamed Derkaoui, Thierry Avramoglou, Virgine Gueguen, Christel Barbaud, and

Didier Letourneur*

Inserm, U698, Bio-ingénierie de Polymères Cardiovasculaires

Institut Galilée, Université Paris 13, 99 Av. J.B. Clément, 93430 Villetaneuse (France)

* Corresponding author :

Inserm U698, BPC; Institut Galilée, Université Paris 13, 99 Av. J.B. Clément, 93430

Villetaneuse, France

Tel/Fax : +33 1 4940 3008

Email : [email protected]

Text only

Page 3: Dextran

2

ABSTRACT

We have synthesized new structures obtained from amphiphilic copolymers of dextran and

polybutylmethacrylate with the aim of endothelialization of biomaterials. Grafting of

butylmethacrylate onto dextran has been carried out using ceric ammonium nitrate as initiator.

Three copolymers were obtained (11, 30 and 37 wt% dextran) and homogeneous thin films

were successfully prepared. In contrast to dextran, the resulting films were stable in water,

and copolymers characterized by FTIR, DSC and DMA showed evidence of hybrid properties

between the parent homopolymers. Surfaces of films were smooth when analyzed by AFM

(roughness 2 +/-1 nm) but greatly differed in their hydrophilicity by increasing the dextran

content (water contact angle from 99° to 57°). In contrast to polybutylmethacrylate where the

proliferation of vascular smooth muscle cells (VSMCs) was excellent but very low for

endothelial cells, the copolymer containing 11% of dextran was excellent for endothelial cells

and very limited for VSMCs. An in vitro wound assay demonstrated that copolymer with 11%

dextran is even more favorable for EC migration than tissue-culture polystyrene. Increasing

the dextran content in the copolymers decreased the proliferation for both vascular cell types.

Altogether, these results evidenced that transparent and water-insoluble films made from

copolymers of dextran and butylmethacrylate copolymers with an appropriate composition

could enhance endothelial cell proliferation and migration. Therefore, a potential benefit of

this approach is the availability of surfaces with tunable properties for endothelialization of

materials.

KEYWORDS : Polybutylmethacrylate; Dextran; Copolymer films; Cell proliferation;

Endothelial cell migration; Tissue Engineering

Page 4: Dextran

3

1. Introduction

The development of amphiphilic copolymers based on copolymerization of hydrophilic

and hydrophobic parts is of great importance due to their versatility and potential applications

in regenerative medicine. [1-5]

For instance, methacrylic polymers have been widely studied as

a biocompatible matrix for use in tissue engineering.[6-11]

Some other synthetic polymers

showed physicochemical and mechanical properties comparable to those of biological tissues,

but were not sufficiently biocompatible. In this context, natural biological polymers such as

polysaccharides possess good biocompatibility,[12-15]

but their mechanical properties are often

inadequate. Improvements of the characteristics of synthetic polymers could be attained by

the addition of biological components in the structures. The resulting materials could combine

the appropriate mechanical properties of the synthetic polymer with the biocompatibility of

the natural polymers. The main challenge in achieving this combination relates to the

thermodynamic incompatibility of natural and synthetic polymers that is the result of enthalpy

and entropy barriers caused by the differences in hydrophilicity/hydrophobicity. This

incompatibility is a frequent phenomenon observed for many polymer mixtures and results in

phase separation.[16,17]

In order to overcome this problem, we and others have already

investigated graft copolymerization of a synthetic polymer, polymethylmethacrylate, onto the

natural dextran macromolecule using ceric ammonium nitrate as a redox initiator.[18-22]

This

strategy consisted on the creation of free-radical grafting sites on the dextran backbone by

reaction with cerium IV, the radicals so formed, initiated copolymerization with

methylmethacrylate monomers. We described here new compounds obtained from graft

polymerization of dextran with butylmethacrylate. The materials were prepared as a series of

films, differing in their wettability by varying the concentration of dextran, and we studied

their behaviors with vascular cells[23, 24]

. We found that films with a peculiar composition

Page 5: Dextran

4

were able to greatly enhance migration and proliferation of endothelial cells, but not vascular

smooth muscle cells.

2. Materials and methods

2.1 Materials

Dextran 70000 g/mol obtained from SIGMA, was dried under vacuum at 60°C for 24h.

Butylmethacrylate monomers were obtained from Acros, and were purified by washing with

NaOH 5%, NaCl 20%, followed three times by distilled water. Ce(IV) ammonium nitrate and

nitric acid were obtained from Acros.

2.2 Graft copolymerization

Three types of copolymers were synthesized by varying the amount of dextran. The

general experimental procedure was as follow: the reactions were carried out into 1L five-

neck flask equipped with stirrer and condenser then the flask was immersed into thermostated

oil bath at 40°C and purged by nitrogen gas. Three different mass ratios of dextran to BMA

(0.125/2; 0.5/2; 0.75/2 w/w) were used. Dextran was dissolved in 500mL of nitric acid 0.2M

for 10min. Five mL of solution of Ce(IV) dissolved in HNO3 0.2M (2.4 10-4

mol/L) and 2g of

monomers were simultaneously added. The system was continuously stirred for 50 min and

the solution was neutralized to pH 8 by addition of NaOH aqueous solution (10M) and

concentrated with an evaporator. Methanol was used as a non-solvent to precipitate the

polymeric material, that was washed with 50mL EDTA 10-2

M to remove cerium ions. The

three purified copolymers named DB1, DB2 and DB3 were frozen at -20°C and lyophilized

for 24h. To obtain pure graft copolymers, the resulting products were extracted with acetone

in a soxhlet for 24h. Pure copolymers were dried under vacuum and weighted.

2.3 Preparation of films

Preparation of copolymers films was carried out by dissolving 6% (w/v) of copolymers in

mixture THF/H2O 90/10 (v/v). 500µL of the transparent solution obtained was molded in

Page 6: Dextran

5

Petri dishes (15mm in diameter) for 24h at room temperature in saturated atmosphere of THF

and in presence of CaCl2 to absorb water. After solvent evaporation, films were released from

the mold upon addition of distilled water, while in the case of PBMA (very fragile), the films

were soaked in distilled water for 2h and then carefully removed from the Petri dishes. The

samples were air dried for 24h to remove the solvent and obtained films are 35µm in

thickness. The films were rinsed several times with PBS. Films of copolymers were cut for

mechanical tests in the shape of 15mm x 5mm, but PBMA film was not prepared because it

was very brittle.

2.4 Infra-red spectral analysis

The infra-red spectra of copolymers films (DB1, DB2 and DB3) were obtained using a

Perkin-Elmer 1600 spectrometer. Each film was mounted on the sample holder. The

atmosphere spectrum was subtracted as background for each scan. Each spectrum was

obtained at a resolution of 1 cm-1

using 32 scans. The studied samples were not gaining

weight by moisture uptake during the acquisition, and similar data were obtained after one

hour. Omnic software was used for data acquisition and analysis. Fourier transform infra-red

(FT-IR) spectroscopy was also used to semi-quantitatively estimate the percentage of dextran

in the copolymers using physical mixtures of dextran and PBMA at different ratios. A

quantity of 15 mg of fine KBr powder was mixed with 1.5mg of physical blend of

dextran/PBMA weight ratio (5/95, 10/90, 15/85, 20/80, 25/75, 30/70, 35/65 and 40/60). We

used the integrated area of (O-H) band at 3400cm-1

with different concentrations of dextran,

PBMA and each synthesized copolymer.

2.5 Differential scanning calorimetric analysis (DSC)

Thermal behavior of homopolymers (dextran and PBMA) and all copolymers (DB1, DB2

and DB3) were performed with a Setaram DSC131 thermal analyzer, operating in

combination with the liquid nitrogen cooling accessory. A mixture of dextran + PBMA (50%,

50% w/w) was also tested. The furnaces were purged with nitrogen and samples were passed

Page 7: Dextran

6

in temperature range from -50 to +200°C at heating rate of 10°C/min. In each case, two

consecutive scans were carried out on each sample. The glass transition temperature (Tg) of

the products was taken as the midpoint in the shift of heat flow baseline.

2.6 Dynamic-mechanical analysis (DMA) on copolymer films

DMA test was used to analyze loss modulus, storage modulus and tan of all films of

copolymers (DB1, DB2 and DB3). PBMA was not tested because it was rigid and very fragile

at room temperature. Films of 15 x 5 x 0.035 mm were analyzed on dynamic-mechanical

analyzer, DMA Q800 (TA Instruments). The equipment has operated in a single cantilever

bending mode at 1Hz in frequency, amplitude 5 m and a heating rate of 3°C/min from -100

to +200°C.

2.7 Creep test

To determine the deformation of DB copolymers, a creep test at ambient temperature was

carried out by applying a constant load to DB1, DB2 and DB3 and observing the deformation

with time. Several loads (100g; 125g; 150g, 175g and 200g) were suspended to the extremity

of films with 15mm length, 5mm width and 0.035mm thickness. The expansion of films was

measured from pictures every minute until rupture. Results are obtained as creep curves

according to the following equation: “ε = a[1 − exp(−bt)] + ct”, where (ε) is the distortion and

(t) is the time. Curves were drawn for the percentage of expansion versus time, and initial

deformation velocity obtained with V0 = (d /dt) t 0 = ab + c, is then deduced for all

samples.

2.8 Contact angle measurements

The water contact angles of films were measured at room temperature using a contact

angle system (Krüss, model DSA 10- Mk2). A droplet of deionized water (2µL) was put on

the air-side surface of the films and after 10s the contact angle was measured. The process

was monitored by a video contact angle system. Ten measurements were carried out for each

Page 8: Dextran

7

simple (PBMA, DB1, DB2 and DB3; dextran gave any film) and the obtained values were

averaged.

2.9 Topography and roughness (AFM)

Atomic force microscopy (AFM) was performed to determine the surface roughness and

the uniformity of coating. All films were imaged with a Nanoscope III Digital instruments

dimension 3100 atomic force microscope (Veeco Metrology Group) and a resonance

frequency range 50-100Hz. Three areas per sample were imaged under Tapping contact mode

and analyzed with WSxM 3.0 beta 10.0 to obtain roughness (Ra) and peak heights of each

film.

2.10 Cell adhesion and proliferation on copolymer films

Endothelial cells (human umbilical vein endothelial cells -HUVEC from ATCC) and

smooth muscle cells (aortic vascular smooth muscle cells -VSMCs, line Rb1[23, 24]

), were used

in this study. The films (15 mm in diameter) were placed on the bottom of 24-well plates, and

the two surfaces were submitted to UV at 254nm for 15min. The ECs and VSMCs cells were

seeded on each film at 5x103 cell/cm

2. Cells were cultured in Dulbecco’s Modified Eagle’s

Medium (DMEM, Gibco) with 4.5 g/L of glucose and 2% L-glutamine supplemented with

10% calf serum and 1% of penicillin and streptomycin. The cells were incubated for 5 days at

37°C in air containing 5% CO2 and in a humidified incubator. The transparency of all films

allowed seeing the cell morphology with a classical inverted phase contrast light microscope

(Zeiss, Axiovert 100). The kinetics of proliferation has been determined daily for 5 days after

starting cell culture. The cell layer adhering to the each disc was washed twice with PBS 1x to

remove remaining materials and loose cells and the cells were detached with trypsin/EDTA

0.25/0.02 (w/v), resuspended and counted with a Coulter (Coulter Counter ZM). The mean

cell number from two wells per tested material was counted. Results are averaged from two

separate experiments. Cell adhesion and proliferation was also observed by fluorescence

microscopy, with cells cultured on films of copolymers for 5 days. Cells were fixed with 10%

Page 9: Dextran

8

paraformaldehyde, permeabilized with 0.1% triton X100, and labeled with intracellular F-

actin (phalloïdin FluoProbes547) for cytoskeleton and DAPI for nucleus.

2.11 Endothelial cell migration

In vitro migration assay of ECs was performed as described.[23]

Briefly, ECs at 5x103

cell/cm2

were seeded into 24-well plates, where DB1 film has been placed on the bottom of

wells. Equivalent numbers of cells were also seeded directly into in 24-wells of tissue culture

treated polystyrene surface (TCPS) as a reference surface. After 5 days, EC monolayers were

scraped (denuded) using a 10µL plastic micropipette tip and washed by sterile phosphate

buffered saline solution (PBS, 0.01M, pH = 7.4), afterwards PBS was replaced with DMEM.

Area of migration was observed with optical microscopy and measured at 8, 24, 32, and 48

hours after taken pictures by a digital camera (Pentax istD).

3. Results and discussion

Dextran-graft-polybutylmethacrylate copolymers were synthesized in this work. A redox

initiation was performed with an acidified aqueous solution (pH= 0.32). The decrease in pH

of the reaction mixture, as well as increase of optical density (appearance of tiny white

particles shortly after the cerium and monomer addition) indicated initiation of

copolymerization (Fig. 1). In contrast, this reaction was not obtained in the absence of either

dextran or butylmethacrylate monomer, indicating that the reaction was not the formation of

homopolymers. The resulting copolymers were then purified by extensive extraction with

acetone in a soxhlet to eliminate butylmethacrylate monomer. Three copolymers (DB1, DB2,

DB3) were successfully synthesized by increasing the dextran content in the starting solution.

Interestingly, the resulting copolymers were not soluble in any pure solvents of the parent

homopolymers (water, DMSO, Dichloromethane, Acetone, Chloroform, Dioxane, THF), but

only in a mixture 90/10 of THF/water.

Page 10: Dextran

9

The absence of a common solvent between copolymers and the parent homopolymers, did

not allow NMR to be performed. However for semi-quantitative analysis, the comparison of

FT-IR spectra of PBMA and dextran showed for the three copolymers the presence of typical

absorption band of PBMA at 1730 cm−1

corresponding to C=O stretching vibrations, and

characteristic band for dextran corresponding to OH stretching absorption at 3400 cm-1

. The

intensity of O-H band increased with increasing proportion of dextran in copolymers. Semi-

quantitative analysis of the integrated area of the 3400 cm−1

peak as previously described 25

,

gave a calibration curve using physical mixtures of powders made of dextran and PBMA in

known proportions, and then the content of dextran in the copolymers : 11, 30 and 37 weight

% in dextran for DB1, DB2 and DB3, respectively.

The DSC curves of physical mixtures dextran + PBMA evidenced two separate Tg, close

to the Tg of the single homopolymers (28°C for pure PBMA, and 153°C for pure dextran).

Each DB copolymer showed a single glass transition, with an intermediate value compared to

the parent homopolymers (Fig. 2). Tg value of DB copolymers increased with increasing

dextran content (31°C, 38°C, 42°C for DB1, DB2 and DB3, respectively). These values

greatly differed from copolymers when methylmethacrylate was used instead of

butylmethacrylate.22

Thus, the existence of a single glass transition of the copolymers

confirmed the grafting of BMA onto dextran in copolymers.

Copolymers films were then obtained by dissolving copolymers in THF/H2O 90/10 (v/v),

and after controlled solvent evaporation, thin films were obtained. The films from copolymers

were homogenous, transparent and insoluble in water (in contrast to dextran). The dynamic-

mechanical analysis (DMA) gave three parameters: (a) the storage modulus (E’), (b) the loss

modulus (E”), and (c) Tan which is the ratio (E”/E’), for determining the occurrence of

molecular mobility transition. The data revealed a storage modulus E’ of 480 MPa for DB1,

680 MPa for DB2 and 1010 MPa for DB3 at 37 °C, that increased with increasing amount of

dextran. In all cases, E’ decreased with increasing temperature. The elastic modulus, i.e. the

Page 11: Dextran

10

stiffness, decreased with increasing amount of dextran in copolymers. On the other hand,

temperature dependence of the Tan showed a maximum peak for DB1 (57°C), DB2 (60°C)

and DB3 (64°C). These values were thus much lower than for pure polymethylmethacrylate

(109°C) or dextran-graft-polymethylmethacrylate (141°C).22

Similar to DSC results, each film

exhibited one glass transition relaxation confirming the graft copolymerization and the

detection of a single peak in Tan is a sufficient criterion to assume the miscibility.

Differences in values with DSC could be attributed to the presence of water tightly associated

to copolymers, which induced changes in mobility of macromolecules.

Creep test was also carried out in order to study differences between DB1, DB2 and DB3

copolymers. Elongation of copolymers under load is high (Fig. 3 left part), with values

reaching 380 %. As shown in figure 3, deformation velocity increases with strain. The

elasticity of films decreases with the increase of dextran in DB copolymer (Fig. 3 right part).

Hydrophilicity of materials was evaluated by water contact angle ( ) measurement, an

indication of surface wettability. PBMA exhibited the highest contact angle ( = 99°) and was

the most hydrophobic material. DB1 ( = 80°) and DB2 ( = 73°) film had moderate level of

wettability, whereas DB3 ( = 57°) film showed the lowest contact angle. Contact angle on

copolymer films exhibited a progressive decrease with increasing dextran content, thus

indicating increased hydrophilicity. All films were further characterized by Atomic Force

Microscope in tapping mode imaging to obtain information on the surface morphology. DB

films exhibited smooth surface, with no notable trends in topographical features upon

increasing of dextran content (Fig. 4). The values for roughness of films were similar (PBMA

Ra = 1.9 ; DB1 Ra =1.6 ; DB2 Ra = 2 ; DB3 Ra = 2.7). Roughness observed on these surfaces

is low, and suitable for comparison between samples for cell adhesion, and potential physical

immobilization of biomolecules.26

Page 12: Dextran

11

The copolymers were tested in vitro by seeding ECs and VSMCs on films. After 6 h of

culture, analysis of the VSMC morphology on PBMA film showed clear adhesion with cells

that changed from spherical to fusiform shape and spreaded out of its pseudopodia, but no

adhesion of ECs on hydrophobic PBMA. After 5 days, VSMCs attained confluence on PBMA

(never with ECs), and VSMC proliferation and morphology were comparable to the control

cultures on TCPS (Fig. 5).

VSMCs did not proliferate on DB1 (Fig. 5). Interestingly, analysis of endothelial cell

numbers indicated that on copolymer film DB1 (dextran 11%), ECs proliferated very well and

reached confluence after 5 days (Fig. 6). This marked difference for DB1 film (Fig. 7) was

somehow unexpected, but is likely to be related to a peculiar surface tension of the film and

its subsequent adhesion of serum proteins that could differentially interact with cell receptors

involved in cell adhesion. A detailed study of protein adsorption should be carried out in the

future. When the dextran content in copolymers increases to 30% or 37%, the number of

spreading cells (ECs and VSMCs) decreased considerably, and proliferation for both types of

cells was very limited (Fig. 5-7). This result indicated that EC proliferation is sensitive to the

hydrophobic/hydrophilic properties of films, the chemical components being the same

between the copolymers. In this context, many studies have proved that surface characteristics

of biomaterials such as wettability, greatly affect the cell adhesion and growth on the

materials27,28

with a preferential adsorption of some serum proteins like fibronectin and

vitronectin. In the work of Risbud et al., PMMA substrates were exposed to radio-frequency

argon and nitrogen plasmas to decrease the water contact angle.29

They found that the

untreated PMMA surface did not support endothelial growth and PMMA treated by plasma

was better but exhibited growth rates lower than the TCPS. The chemical approach reported

here by graft polymerization allowed to obtain tunable properties for excellent endothelial cell

coverage of DB1. Surface roughness is another important parameter in biomedical materials,

that may affect cell adhesion30,31

and can compromise hemocompatibility in the case of

Page 13: Dextran

12

vascular prostheses32

. According to the AFM results, the roughness was similar among the

different films. Consequently, the proportion of dextran in the copolymer that greatly affected

the hydrophilicity, seems the main parameter to obtain the favorable endothelialization of the

surface of these materials.

The migration assay, created from a wound in a confluent EC monolayer, was used to

assess the effect of DB1 on EC migration compared to TCPS (the other polymer surfaces did

not allowed a confluent EC monolayer). Cell migration was measured as the denuded area in

which ECs invading. The cells on copolymer had extensively migrated from the edge of the

scrape into the denuded space (Fig. 8 top, left panel). They reach confluence at 32h on

copolymer. In contrast, EC migration on TCPS was slower (Fig. 8, right panel), and migration

area remained non-confluent until 48h. This evidenced that surface of DB1 film was

appropriate for both EC migration and proliferation.

4. Conclusion

The aim of this study was to synthesize new copolymers having the biological properties of

polysaccharides and the mechanical properties of polyacrylate. Copolymers of dextran and

butylmethacrylate were successfully synthesized by radical polymerization and characterized.

IR, DSC and DMA revealed that the graft copolymerization has been taken place with the two

different natures and non-miscible polymers. In contrast to dextran, homogenous transparent

and water-insoluble films were prepared from all DB copolymers. The mechanical properties

were enhanced as compared to our previous report with dextran-graft-

polymethylmethacrylate.22

Interestingly, DB1 was able to inhibit VSMC proliferation, and at

the same time to support endothelial cell proliferation and migration. Lastly, cell adhesion and

proliferation can be controlled by varying the content of dextran incorporated into the DB

copolymer. DB1 copolymer exhibited excellent characteristics for applications for

Page 14: Dextran

13

endothelialization of materials. Therefore, a potential benefit of this approach is the

availability of bioactive surfaces with tunable properties of interest for biomedical materials.

Acknowledgments

This work was supported by Inserm, University Paris 13, and IFR Paris Nord-Plaine de

France. The authors wish to express their gratitude to Dr. Cédric Lorthioir (CNRS Thiais,

France) for his excellent technical assistance in DMA tests.

Page 15: Dextran

14

References

[1] Rosler A, Vandermeulen GWM, Klok H-A. Advanced drug delivery devices via self-assembly

of amphiphilic block copolymers. Adv Drug Deliv Rev 2001;53:95-108.

[2] Rimmer S, German MJ, Maughan J, Sun Y, Fullwood N, Ebdon J, et al. Synthesis and

properties of amphiphilic networks 3: preparation and characterization of block conetworks of

poly(butyl methacrylate-block-(2,3 propandiol-1-methacrylate-stat-ethandiol dimethacrylate).

Biomaterials. 2005;26:2219-30.

[3] Shi R, Burt HM. Amphiphilic dextran-graft-poly([var epsilon]-caprolactone) films for the

controlled release of paclitaxel.Int J Pharm 2004;27:167-79.

[4] Wang Q, Wang L, Detamore MS, Berkland C. Biodegradable colloidal gels as moldable tissue

engineering scaffolds. Advanced Materials 2008;20:236-239, 2008

[5] Grossin L, Cortial D, Saulnier B, Felix O, Chassepot A, Decher G, Netter P, Schaaf P, Gillet P,

Mainard D, Voegel JC, Benkirane-Jessel N. Step-by-step build-up of biologically active cell-

containing stratified films aimed at tissue engineering. Advanced Materials 2009; 21: 650-655

[6] Barbani N, Bertoni F, Ciardelli G, Cristallini C, Silvestri D, Coluccio ML, et al. Bioartificial

materials based on blends of dextran and poly(vinyl alcohol-co-acrylic acid). Eur Polymer J

2005;41:3004-10.

[7] Chen YM, Shiraishi N, Satokawa H, Kakugo A, Narita T, Gong JP, et al. Cultivation of

endothelial cells on adhesive protein-free synthetic polymer gels. Biomaterials. Biomaterials

2005;26:4588-96.

[8] De Groot CJ, Van Luyn MJA, Van Dijk-Wolthuis WNE, Cadee JA, Plantinga JA, Otter WD, et

al. In vitro biocompatibility of biodegradable dextran-based hydrogels tested with human

fibroblasts. Biomaterials. 2001;22:1197-203.

[9] Kim YS, Lim JY, Donahue HJ, Lowe TL. Thermoresponsive terpolymeric films applicable for

osteoblastic cell growth and noninvasive cell sheet harvesting. Tissue Eng 2005;11:30-40.

[10] Levesque SG, Shoichet MS. Synthesis of cell-adhesive dextran hydrogels and macroporous

scaffolds. Biomaterials. 2006;27:5277-85.

[11] Moller S, Weisser J, Bischoff S, Schnabelrauch M. Dextran and hyaluronan methacrylate based

hydrogels as matrices for soft tissue reconstruction. Biomol Eng 2007;24(5):496-504.

[12] Abed A, Deval B, Assoul N, Bataille I, Portes P, Louedec L, Henin D, Letourneur D, Meddahi-

Pelle A. A biocompatible polysaccharide hydrogel-embedded polypropylene mesh for enhanced

tissue integration in rats. Tissue Eng A 2008;14:519-27.

[13] Bos GW, Hennink WE, Brouwer LA, den Otter W, Veldhuis TFJ, van Nostrum CF, et al. Tissue

reactions of in situ formed dextran hydrogels crosslinked by stereocomplex formation after

subcutaneous implantation in rats. Biomaterials. 2005;26:3901-9.

[14] Chaouat M, Le Visage C, Autissier A, Chaubet F, Letourneur D. The evaluation of a small-

diameter polysaccharide-based arterial graft in rats. Biomaterials 2006;27:5546-53

[15] Draye J-P, Delaey B, Van de Voorde A, Van Den Bulcke A, De Reu B, Schacht E. In vitro and

in vivo biocompatibility of dextran dialdehyde cross-linked gelatin hydrogel films.

Biomaterials. 1998;19:1677-87.

[16] Ashraf SM, Ahmad S, Riaz U, Alam M, Sharma HO. Miscibility behavior of blend of

polyesteramides of linseed oil and dehydrated castor oil with poly(methacrylic acid). J Appl

Polym Sci 2007;103:1367-74.

[17] Stenekes RJH, Franssen O, van Bommel EMG, Crommelin DJA, Hennink WE. The use of

aqueous PEG/dextran phase separation for the preparation of dextran microspheres. Int J Pharm

1999;183:29-32.

[18] Athawale VD, Lele V. Syntheses and characterisation of graft copolymers of maize starch and

methacrylonitrile. Carbohyd Polym. 2000;41:407-16.

[19] Bertholon I, Labarre D, Besnard M, Vauthier C. Characterization of dextran-

poly(isobutylcyanoacrylate) copolymers obtained by redox radical and anionic emulsion

polymerization. Macromolecules. 2006;39:3559-67.

[20] Tauer K, Müller H, Schellenberg C, Rosengarten L. Evaluation of heterophase polymerizations

by means of reaction calorimetry Colloids Surf A Physicochem Eng Asp 1999;153:143-151.

[21] Yagci C, Yildiz U. Redox polymerization of methyl methacrylate with allyl alcohol 1,2-

butoxylate-block-ethoxylate initiated by Ce(IV)/HNO3 redox system. Eur Polymer J

Page 16: Dextran

15

2005 ;41(1):177-84.

[22] Derkaoui SM, Avramoglou T, Barbaud B, Letourneur D. Synthesis and characterization of a

new polysaccharide-graft-polymethacrylate copolymer for three-dimensional hybrid hydrogels.

Biomacromolecules 2008;9:3033-38.

[23] Hlawaty H, San Juan A, Jacob MP, Vranckx R, Letourneur D, Feldman LJ. Inhibition of MMP-

2 gene expression with small interfering RNA in rabbit vascular smooth muscle cells. Am J

Physiol Heart Circ Physiol 2007;293:3593-3601

[24] San Juan A, Ducrocq G, Hlawaty H, Bataille I, Guenin E, Letourneur D, Feldman LJ. Tubular

cationized pullulan hydrogels as local reservoirs for plasmid DNA. J Biomed Mater Res A

2007;83:819-27

[25] McCool B, Murphy L, Tripp CP. A simple FTIR technique for estimating the surface area of

silica powders and films. J Colloid Interface Sci 2006;295:294-298

[26] Simon J, Carl G., Eidelman N, Kennedy SB, Sehgal A, Khatri CA, Washburn NR.

Combinatorial screening of cell proliferation on poly(l-lactic acid)/poly(d,l-lactic acid) blends.

Biomaterials. 2005;26:6906-15.

[27] Knetsch MLW, Aldenhoff YBJ, Koole LH. The effect of high-density-lipoprotein on thrombus

formation on and endothelial cell attachement to biomaterial surfaces. Biomaterials.

2006;27:2813-9.

[28] Tsuda Y, Kikuchi A, Yamato M, Sakurai Y, Umezu M, Okano T. Control of cell adhesion and

detachment using temperature and thermoresponsive copolymer grafted culture surfaces. J

Biomed Mater Res 2004;69:70-8.

[29] Risbud MV, Dabhade R, Gangal S, Bhonde RR. Radio-frequency plasma treatment improves

the growth and attachment of endothelial cells on poly(methyl methacrylate) substrates:

implications in tissue engineering . J Biomater Sci Polym Ed 2002;13:1067-80.

[30] Chung T-W, Liu D-Z, Wang S-Y, Wang S-S. Enhancement of the growth of human endothelial

cells by surface roughness at nanometer scale. Biomaterials. 2003;24:4655-61.

[31] Faucheux N, Schweiss R, Lutzow K, Werner C, Groth T. Self-assembled monolayers with

different terminating groups as model substrates for cell adhesion studies. Biomaterials.

2004;25:2721-30.

[32] Zheng CL, Cui FZ, Meng B, Ge J, Liu DP, Lee I-S. Hemocompatibility of C-N films fabricated

by ion beam assisted deposition. Surface Coatings Technol 2005;193:361-5.

Page 17: Dextran

16

Legends for Figures

Figure 1. Variation of pH (▲) and optical density (■) during copolymerization. Dextran-

graft-polybutylmethacrylate copolymers were synthesized via redox initiation. The decrease

in pH (left), as well as increase of optical density at 650 nm (right) was plotted as a function

of time. No change was observed in the absence of dextran or butylmethacrylate monomer.

Figure 2. DSC curves showing the glass transition temperature (Tg). Curves for (a) dextran,

(b) PBMA, (c) mixture of dextran + PBMA (50% + 50%), (d) DB1, (e) DB2 and (f) DB3 are

reported in temperature range from -50 to +200°C at heating rate of 10°C/min. In each case,

two consecutive scans were carried out on each sample. The Tg was taken as the midpoint in

the shift of heat flow baseline.

Figure 3. Left) Representative example of elongation of copolymer DB1; Right) Comparison

of initial deformation velocity of DB1 (♦), DB2 (■) and DB3 (▲). A constant load (100g;

125g; 150g, 175g and 200g) was applied to DB1, DB2 and DB3 films (15mm length, 5mm

width and 0.035mm thickness). Elongation of films was measured from pictures every minute

until rupture. Curves were obtained for the percentage of expansion versus the time, and

initial deformation velocity is then deduced for all copolymers. PBMA was too fragile to be

characterized, and dextran gave any film.

Figure 4. Atomic Force Microscopy of PBMA (top left), copolymers DB1 (top right), DB2

(bottom left) and DB3 (bottom right). Scale X, Y = 2 µm/div. All films were imaged under

tapping contact mode with a resonance frequency range of 50-100Hz. Three areas per sample

were imaged and analyzed to obtain roughness (Ra) of each film.

Figure 5. Effect of film copolymers on endothelial cell proliferation. ECs were cultured on

films of PBMA or copolymers DB1, DB2, DB3. Top panel : Pictures were taken after 5 days

in culture (Original magnification 12.5 x). Bottom panel : Endothelial cell growth curve for

PBMA and copolymers DB1, DB2, DB3. Values are means of duplicate wells in two

experiments.

Figure 6. Effect of film copolymers on vascular smooth muscle cell proliferation in vitro.

VSMCs were cultured on films of PBMA or copolymers DB1, DB2, DB3. Top panel :

Page 18: Dextran

17

Pictures were taken after 5 days in culture (Original magnification 12.5 x). Bottom panel :

Cell growth curve for PBMA and copolymers DB1, DB2, DB3. Values are means of duplicate

wells in two experiments.

Figure 7. Effect of DB1 film on endothelial cell (left), or vascular smooth muscle cell

proliferation in vitro (right). Cells were cultured on films of copolymer DB1 for 5 days. Cells

were fixed, permeabilized, and labeled with intracellular F-actin (phalloïdin FluoProbes547 in

red) for cytoskeleton and DAPI (in blue) for nucleus. Cells were observed under a fluorescent

microscope. Pictures are representative of three experiments.

Figure 8. EC migration of a denuded EC monolayer on DB1 film (top, left panel) or on TCPS

(top, right panel). Bar is 100µm for each picture. Bottom panel : Quantitative analysis of EC

migration versus time. Area of migration was observed with optical microscopy and measured

at 8, 24, 32, and 48 hours. Values are representative data obtained in two independent

experiments.

Page 19: Dextran

Figure 1. Variation of pH (▲) and optical density (■) during copolymerization. Dextran-

graft-polybutylmethacrylate copolymers were synthesized via redox initiation. The decrease

in pH (left), as well as increase of optical density at 650 nm (right) was plotted as a function

of time. No change was observed in the absence of dextran or butylmethacrylate monomer.

Figure(s)

Page 20: Dextran

Figure 2. DSC curves showing the glass transition temperature (Tg). Curves for (a) dextran,

(b) PBMA, (c) mixture of dextran + PBMA (50% + 50%), (d) DB1, (e) DB2 and (f) DB3 are

reported in temperature range from -50 to +200°C at heating rate of 10°C/min. In each case,

two consecutive scans were carried out on each sample. The Tg was taken as the midpoint in

the shift of heat flow baseline.

Page 21: Dextran

Figure 3. Left) Representative example of elongation of copolymer DB1; Right) Comparison

of initial deformation velocity of DB1 (♦), DB2 (■) and DB3 (▲). A constant load (100g;

125g; 150g, 175g and 200g) was applied to DB1, DB2 and DB3 films (15mm length, 5mm

width and 0.035mm thickness). Elongation of films was measured from pictures every minute

until rupture. Curves were obtained for the percentage of expansion versus the time, and

initial deformation velocity is then deduced for all copolymers. PBMA was too fragile to be

characterized, and dextran gave any film.

Page 22: Dextran

Figure 4. Atomic Force Microscopy of PBMA (top left), copolymers DB1 (top right), DB2

(bottom left) and DB3 (bottom right). Scale X, Y = 2 µm/div. All films were imaged under

tapping contact mode with a resonance frequency range of 50-100Hz. Three areas per sample

were imaged and analyzed to obtain roughness (Ra) of each film.

Page 23: Dextran

Figure 5. Effect of film copolymers on endothelial cell proliferation. ECs were cultured on

films of PBMA or copolymers DB1, DB2, DB3. Top panel : Pictures were taken after 5 days

in culture (Original magnification 12.5 x). Bottom panel : Endothelial cell growth curve for

PBMA and copolymers DB1, DB2, DB3. Values are means of duplicate wells in two

experiments.

Page 24: Dextran

Figure 6. Effect of film copolymers on vascular smooth muscle cell proliferation in vitro.

VSMCs were cultured on films of PBMA or copolymers DB1, DB2, DB3. Top panel :

Pictures were taken after 5 days in culture (Original magnification 12.5 x). Bottom panel :

Cell growth curve for PBMA and copolymers DB1, DB2, DB3. Values are means of duplicate

wells in two experiments.

Page 25: Dextran

Figure 7. Effect of DB1 film on endothelial cell (left), or vascular smooth muscle cell

proliferation in vitro (right). Cells were cultured on films of copolymer DB1 for 5 days. Cells

were fixed, permeabilized, and labeled with intracellular F-actin (phalloïdin FluoProbes547 in

red) for cytoskeleton and DAPI (in blue) for nucleus. Cells were observed under a fluorescent

microscope. Pictures are representative of three experiments.

Page 26: Dextran

Figure 8. EC migration of a denuded EC monolayer on DB1 film (top, left panel) or on TCPS

(top, right panel). Bar is 100µm for each picture. Bottom panel : Quantitative analysis of EC

migration versus time. Area of migration was observed with optical microscopy and measured

at 8, 24, 32, and 48 hours. Values are representative data obtained in two independent

experiments.