journal of membrane science -...

10
Preparation and evaluation of heparin-immobilized poly (lactic acid) (PLA) membrane for hemodialysis Ailin Gao, Fu Liu n , Lixin Xue n Ningbo Institute of Material Technology & Engineering, Chinese Academy of Sciences, 519 Zhuangshi Road, Ningbo 315201, China article info Article history: Received 30 August 2013 Received in revised form 8 October 2013 Accepted 9 October 2013 Available online 19 October 2013 Keywords: Poly (lactic acid) membrane Hemodialysis Heparinization Bio-based material abstract Bio-based poly (lactic acid) membrane with asymmetric porous structure was developed for hemodia- lysis via phase inversion for the rst time. Heparin was immobilized to PLA membrane surface through the strong adhesion ability of dopamine, as conrmed by attenuated total reectance Fourier transform infrared (ATR-FTIR) spectroscopy and X-ray photoelectron spectroscopy (XPS) respectively. The morphol- ogies variations of PLA membranes induced by dopamine coating and heparin immobilization were analyzed by scanning electron microscope (SEM) and atomic force microscopy (AFM). Hydrophilicity and permeability of PLA membranes before and after modication were characterized. Particularly, platelet adsorption, plasma recalcication time and hemolysis ratio were executed to evaluate the blood compatibility of PLA membranes decorated by heparin. The in vitro results demonstrated that surface heparinization improved the hemocompatibility of PLA membrane, suppressed the adhesion of platelet, extended plasma recalcilication time, and also decreased hemolysis ratio. The dialysis simulation experiments including urea and lysozyme clearance as well as bovine serum albumin (BSA) rejection were implemented to determine the dialysis performances. The results showed that 79% of urea and 18% of lysozyme were cleaned and over 90% of BSA was retained. This study disclosed a window of opportunity to produce novel hemodialysis membranes using bio-based materials. & 2013 Elsevier B.V. All rights reserved. 1. Introduction Biomaterials have played an enormous role in the success of biological and medical applications, such as medical devices and drug delivery systems, tissue engineering, diagnostics and array technologies etc. [1]. Hemodialysis membrane which removes blood toxins by diffusion and convection across the porous membrane and restoring uid and solute balance within the patient to some extent, also called as articial kidney, is the most important material to treat the renal failure, especially for the end-stage renal disease (ESRD). Cellulose and its derivatives are the rst generation of polymers which were used in dialyzers [2,3]; however, the weak hydraulic permeability and low molecular molecular weight cutoff near 2000 daltons limits the prospect of cellulose-based materials. Synthetic polymers such as polysul- fone, polyether sulfone, polyacrylonitrile, polyamide and poly- methylmethacrylate were widely used and researched, which can be regarded as the second generation of dialysis membrane. Compared to cellulose membrane, synthetic bers are 10 times more permeable and can be tailored to a range of molecular weight cutoffs. Nevertheless, the intrinsic hydrophobicity of synthetic polymer has adverse inuences on the biocompatibil- ity, which would directly result in the blood clotting. Therefore, surface grafting, coating or blending are usually needed to render the membrane hydrophilic and improve the biocompatibility [46]. Therefore, novel membrane materials with more comprehen- sive properties including better dialysis performances, biocom- patibility, and controlled degradation processing after use will be developed in an urgent need. Poly (lactic acid) or polylactide (PLA) is the most extensively researched and utilized biodegradable and renewable thermo- plastic bio-based polyester, with great potential to replace con- ventional petrochemical-based polymers [79]. However, few studies on PLA porous membrane preparation and application have been conducted. PLA and its copolymers, e.g. poly (lactic-co- glycolide), were fabricated into porous membranes for tissue engineering due to its eminent biocompatibility via thermally induced phase separation [1013]. Tanaka et al. attempted to prepare PLA depth lter microltration membranes serving as screen lters and bacterial retaining via a combined process including nonsolvent induced and thermally induced phase separation [14,15]. In addition, PLA and its copolymers have also been converted into hollow ber ultraltration membranes with asymmetrical structures, which exhibited high water permeability and good separation performances [16]. Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/memsci Journal of Membrane Science 0376-7388/$ - see front matter & 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.memsci.2013.10.016 n Corresponding authors. Tel.: þ86 86685831. E-mail addresses: [email protected] (F. Liu), [email protected] (L. Xue). Journal of Membrane Science 452 (2014) 390399

Upload: vuongphuc

Post on 29-Aug-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

Preparation and evaluation of heparin-immobilized poly (lactic acid)(PLA) membrane for hemodialysis

Ailin Gao, Fu Liu n, Lixin Xue n

Ningbo Institute of Material Technology & Engineering, Chinese Academy of Sciences, 519 Zhuangshi Road, Ningbo 315201, China

a r t i c l e i n f o

Article history:Received 30 August 2013Received in revised form8 October 2013Accepted 9 October 2013Available online 19 October 2013

Keywords:Poly (lactic acid) membraneHemodialysisHeparinizationBio-based material

a b s t r a c t

Bio-based poly (lactic acid) membrane with asymmetric porous structure was developed for hemodia-lysis via phase inversion for the first time. Heparin was immobilized to PLA membrane surface throughthe strong adhesion ability of dopamine, as confirmed by attenuated total reflectance Fourier transforminfrared (ATR-FTIR) spectroscopy and X-ray photoelectron spectroscopy (XPS) respectively. The morphol-ogies variations of PLA membranes induced by dopamine coating and heparin immobilization wereanalyzed by scanning electron microscope (SEM) and atomic force microscopy (AFM). Hydrophilicity andpermeability of PLA membranes before and after modification were characterized. Particularly, plateletadsorption, plasma recalcification time and hemolysis ratio were executed to evaluate the bloodcompatibility of PLA membranes decorated by heparin. The in vitro results demonstrated that surfaceheparinization improved the hemocompatibility of PLA membrane, suppressed the adhesion of platelet,extended plasma recalcilication time, and also decreased hemolysis ratio. The dialysis simulationexperiments including urea and lysozyme clearance as well as bovine serum albumin (BSA) rejectionwere implemented to determine the dialysis performances. The results showed that 79% of urea and 18%of lysozyme were cleaned and over 90% of BSA was retained. This study disclosed a window ofopportunity to produce novel hemodialysis membranes using bio-based materials.

& 2013 Elsevier B.V. All rights reserved.

1. Introduction

Biomaterials have played an enormous role in the success ofbiological and medical applications, such as medical devices anddrug delivery systems, tissue engineering, diagnostics and arraytechnologies etc. [1].

Hemodialysis membrane which removes blood toxins bydiffusion and convection across the porous membrane andrestoring fluid and solute balance within the patient to someextent, also called as artificial kidney, is the most importantmaterial to treat the renal failure, especially for the end-stagerenal disease (ESRD). Cellulose and its derivatives are the firstgeneration of polymers which were used in dialyzers [2,3];however, the weak hydraulic permeability and low molecularmolecular weight cutoff near 2000 daltons limits the prospect ofcellulose-based materials. Synthetic polymers such as polysul-fone, polyether sulfone, polyacrylonitrile, polyamide and poly-methylmethacrylate were widely used and researched, which canbe regarded as the second generation of dialysis membrane.Compared to cellulose membrane, synthetic fibers are 10 timesmore permeable and can be tailored to a range of molecular

weight cutoffs. Nevertheless, the intrinsic hydrophobicity ofsynthetic polymer has adverse influences on the biocompatibil-ity, which would directly result in the blood clotting. Therefore,surface grafting, coating or blending are usually needed to renderthe membrane hydrophilic and improve the biocompatibility [4–6]. Therefore, novel membrane materials with more comprehen-sive properties including better dialysis performances, biocom-patibility, and controlled degradation processing after use will bedeveloped in an urgent need.

Poly (lactic acid) or polylactide (PLA) is the most extensivelyresearched and utilized biodegradable and renewable thermo-plastic bio-based polyester, with great potential to replace con-ventional petrochemical-based polymers [7–9]. However, fewstudies on PLA porous membrane preparation and applicationhave been conducted. PLA and its copolymers, e.g. poly (lactic-co-glycolide), were fabricated into porous membranes for tissueengineering due to its eminent biocompatibility via thermallyinduced phase separation [10–13]. Tanaka et al. attempted toprepare PLA depth filter microfiltration membranes serving asscreen filters and bacterial retaining via a combined processincluding nonsolvent induced and thermally induced phaseseparation [14,15]. In addition, PLA and its copolymers have alsobeen converted into hollow fiber ultrafiltration membranes withasymmetrical structures, which exhibited high water permeabilityand good separation performances [16].

Contents lists available at ScienceDirect

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

Journal of Membrane Science

0376-7388/$ - see front matter & 2013 Elsevier B.V. All rights reserved.http://dx.doi.org/10.1016/j.memsci.2013.10.016

n Corresponding authors. Tel.: þ86 86685831.E-mail addresses: [email protected] (F. Liu), [email protected] (L. Xue).

Journal of Membrane Science 452 (2014) 390–399

The biocompatibility of bio-based PLA also stirred up consider-ably discussion for the extensive application of sutures [17],devices for bone fracture fixation [18], and meshes for tissuesengineering [19,20]. Although, previous studies have suggestedwild response between PLA device and human tissue, the out-comes may alter with the change of material feature and specialuse-pattern. Thus, to regulate response condition, a number ofmodification technologies have been presented, in which, surfaceheparinization is one of effective ways to improve anticoagulantand antithrombotic property [21–23]. Researches stated that, withthe synergistic effect of platelet factor III (AT3), heparin could notonly act on thrombin to suppress the conversion reaction fromfibrinogen to fibrin, but also prevent platelet adhesion andaggregation on the materials surface [24]. Both ionic adsorptionand covalent bonding have been explored to fix heparin. Forexample, plasma processing was adopted to make the surfaceelectropositive, followed by combine with electronegative heparin[25]. Besides, reactive groups such as amidogen or hydroxyl werealso introduced to bind heparin covalently [26–28]. To avoid thecomplex implementation process of methods summed up above,polydopamine-coating based on its self-polymerization and strongadhesion feature attracted much attention as the most practicalway to functionalize membrane surfaces [29–32].

Based on the state of art reviewed above, we developed a novelhemodialysis membrane originated from bio-based PLA for thefirst time, which can be regarded as the next generation of dialysismembrane. The PLA porous membrane was fabricated via theconventional phase inversion, and then immobilized with heparinonto the PLA membrane via polydopamine-coating. Surface chem-istry, morphology, water contact angle, and water flux of PLAmembranes were studied as a contrast after polydopamine-coating and surface heparinization were studied. The bloodcompatibility of PLA membrane was investigated in terms ofplatelet adhesion, extended plasma recalcilication time, andhemolysis ratio in details. Finally simulation experiment includingurea, lysozyme and bovine serum albumin osmosis was conductedto evaluate the dialysis performances of our PLA membranes.

2. Experimental

2.1. Materials

Poly (lactic acid) (crystalline o5%, Mn¼130–140 K Dolton,Natural works, US) was used to prepare dialysis membrane. Poly(ethylene glycol) (PEG, CP, Mv20000, Sino pharm, China) wasused as pore-forming agent and flexibilizer and 1-methyl-2-pyrrolidinone (NMP, AR, Aladdin, China) as solvent. Dopaminehydrochloride, tris hydroxymethyl amino methane (tris), lysozyme,

and bovine serum albumin obtained from Aladdin with the purityof AR were all directly used. Urea with the purity of AR waspurchased from Sino pharm of China. Heparin sodium (BR) wassupplied by Solar brio, China. The anticoagulant sheep whole bloodwas purchased from Beijing Pingrui Biotechnology Company, China.

2.2. Preparation of PLA porous membrane

PLA porous membrane was prepared via phase inversion. 18 gPLA and 5 g PEG were dissolved in 77 g NMP at 80 1C. After the airbubbles removed by vacuum deaeration at 70 1C, the formedhomogenous solution was immediately casted onto a clean glassplate with a thickness of 150 μm blade, and coagulated in waterbath at room temperature for 15 min. After that, PLA porousmembranes were obtained for further modification and character-ization after phase inversion and then soaked in deionized waterto remove any residual solvent. Polysulfone (PSf) membrane as thecontrol sample was also prepared via phase inversion with thequality ratio of PSf/PEG/DMAc¼18/5/77.

2.3. Heparinization of PLA porous membranes

A certain amount of dopamine was first dissolved in tris buffersolution (10 mM, pH8.5) to obtain dopamine solution (2.0 g/L,1.0 g/L, 0.5 g/L), subsequently, PLA membranes (25 cm�15 cm)were immersed in 300 mL dopamine solution and roundly shakenat 20 1C for 8 h. Subsequently, the membranes were taken out andwashed with deionized water for 30 min. In the alkaline aqueoussolution, the pyrocatechol group of dopamine will be oxidized intobenzoquinone under the effect of dissolved oxygen. There is adisproportionation reaction between pyrocatechol and benzoqui-none. The semiquinone radical is produced consequently andfollowed by the couple reaction. The result polymerization iscalled polydopamine. Polydopamine was finally coated on PLAmembrane surface. The resultant polydopamine-coated PLA mem-branes were marked by DA2-M, DA1-M, DA05-M corresponding tothe various dopamine concentration, and used for characterizationand further surface heparinization, while the native PLA mem-brane without polydopamine coating was labeled as PLA-M.

The above polydopamine-coated PLA membranes wereimmersed into 200 mL heparin solution (2.0 g/L, PBS buffer solu-tion as solvent, pH 7.4) at 4 1C for 24 h, a long time enough for thecovalent bonding between heparin and polydopamine. Then themembranes were taken out and rinsed with deionized water for30 min to remove instable heparin. Similarly, the resultantheparin-immobilized PLA membranes were named by HEP2-M,HEP1-M, and HEP05-M. All membranes obtained were dried in theoven at 40 1C for any further characterization.

2.4. Characterization of PLA membranes

2.4.1. Surface chemistry and morphologiesAttenuated total reflectance Fourier transform infrared spectra

(ATR-FTIR, Thermo-Nicolet 6700, US) were used to detect anychange of functional groups on PLA membrane surface before andafter modification. And the chemical compositions on surface wereanalyzed by the X-ray photoelectron spectroscopy (XPS, ShimadzuAxis Utltradld, Japan), with Mg-Kα as radiation resource.

The morphologies of the cross section fractured in liquidnitrogen and top surface of the membranes were characterizedby a scanning electron microscope (SEM, Hitachi S-4800, Japan).All samples were sputtered with gold for 2 min for observation. Anatomic force microscopy (AFM, Dimension 3100V, Veeco, US) wascarried out to determine the surface roughness.Fig. 1. Schematic diagram of single membrane dialysis system.

A. Gao et al. / Journal of Membrane Science 452 (2014) 390–399 391

2.4.2. Hydrophilicity and permeability measurementContact angle change with the drop age (2 μL) of PLA mem-

branes was recorded by a water contact angle system (OCA20, Dataphysics, Germany) to measure the effects of surface modificationon the hydrophilicity of membranes.

The pure water flux was measured by a commercial filtrationapparatus (Saifei Company, China) with an effective membranearea of 24 cm2. Membranes were pre-pressured at 0.2 MPa for30 min before the flux was recorded every 5 min at 0.1 MPa until astable water flux value obtained. The average value from threetests was obtained for each sample.

2.4.3. Blood compatibility2.4.3.1. Platelet adhesion. 10ml anticoagulant sheep whole bloodwas added into centrifuge tube, followed by centrifuging at1000 rpm for 10 min. Taking out surpernatant with tubularis, theplatelet rich plasma (PRP) was obtained. All membrane samples

(1 cm�1 cm) after washed with PBS buffer solution (pH 7.4) wereadded into the 24-well plate. 100 μL PRP was dropped on each sampleusing pipette, and then maintained at 37 1C for 1 h. When incubation

Fig. 2. Schematic diagram showing dopamine-coating and heparin-grafting onto PLA porous membranes.

Fig. 3. FTIR spectra of the PLA-M, DA05-M, and HEP05-M.

Fig. 4. XPS spectra of PLA-M, DA05-M and HEP05-M.

A. Gao et al. / Journal of Membrane Science 452 (2014) 390–399392

was finished, the samples were rinsed in PBS twice to remove unstableplatelet. 2.5 wt% glutaraldehyde was added into the solution for onenight to fasten the adsorbed platelets. The samples were thendehydrated step by step with 50%, 75%, 85%, 95%, 100% (v/v)ethanol/water solution for 10 min in sequence. Platelet adhesion onPLA membrane was observed by SEM after freeze drying with liquidnitrogen.

2.4.3.2. Plasma recalcification time (PRT). 10 ml anticoagulantsheep whole blood was added into centrifuge tube and thencentrifuged at 3000 rpm for 15 min. Taking out surpernatant, theplatelet-poor plasma (PPP) was. Drop 200 μL PPP on each sample(0.2 cm�0.2 cm) in a 48-well cell culture plate and incubate theculture plate in water bath at 37 1C for 10 min. Then, 100 μL pre-heated 0.025 mol/L CaCl2 aqueous solution (37 1C, water bath) wasadded to each sample. The mixture was stirred and when anyfibrin threads formed, the time consumed was recorded as theplasma recalcification time. Experiment of each kind of samplewas carried out 3 times and the average value was figured out.

2.4.3.3. Hemolysis ratio (HR). Samples (1 cm�1 cm) were rinsedwith deionized water and 0.9 wt% NaCl aqueous solutions for10 min in sequence, and then soaked in 0.9 wt% NaCl in water bathat 37 1C for another 30 min. 200 μL sheep whole blood was addedto the NaCl solution, maintaining 37 1C for 1 h. After centrifugingat 1500 rpm for 10 min, the absorbance of top clear layer wasmeasured by Ultraviolet spectrophotometer (Lambda 950, PerkinElmer, US ) at 545 nm. Pure water was used as positive reference,while 0.9 wt% NaCl aqueous solution as negative reference. The HRwas calculated using the following equation:

HR¼ ðAS�ANÞ=ðAP�ANÞ ð1Þwhere AS is the absorption value of samples, AN is the absorptionvalue of negative reference, and AP is the absorption value ofpositive reference.

2.5. Dialysis simulation experiment

The dialysis performance of the membranes (PLA-M, HEP05-M)was measured in terms of urea and lysozyme clearance, and BSAretention. The home-made testing system was shown in Fig. 1. Thetotal effective area of every membrane is 64 cm2. 1.5 g/L ureaaqueous solution, 0.04 g/L lysozyme and 1.0 g/L BSA aqueoussolution were used as simulative blood, traversing through thedialysis mold (flow rate 200 ml/min). And 1.5 g/L dextrose solutionplayed the role of simulative dialysate, traversing through the moldfrom the opposite direction (flow rate 500 ml/min). The experimentwas continually run for 4 h and 10 mL test solution was taken outevery 1 h from the simulative blood. The concentration change ofurea solution was evaluated using TOC (Multi N/C 2100-Jena,Germany) by measuring the organic nitrogen content, and that ofBSA and lysozyme were detected with Ultraviolet spectrophot-ometer at 278 nm and 280 nm respectively. The urea clearance iscalculated using the following equation:

Urea clearance percentage¼ ðCt�C0Þ=C0� 100% ð2Þ

where C0 and Ct are the urea concentrations in the testing solutionreservoir at time t¼0 and t¼1, 2, 3, 4 h, respectively.

The lysozyme clearance is calculated using the followingequation:

Lysozyme clearance percentage¼ ðAt�A0Þ=A0� 100% ð3Þwhere A0 and At are the urea concentrations in the testingsolution reservoir at time t¼0 and t¼1, 2, 3, 4 h, respectively.

The BSA retention is calculated using the following equation:

BSA retention percentage¼ At=A0� 100% ð4Þwhere A0 and At are the BSA absorption value in the testingsolution reservoir at time t¼0 and t¼1, 2, 3, 4 h, respectively.

3. Results and discussion

3.1. Surface chemistry

PLA membrane was first coated by polydopamine due to itsstrong adhesion, and then heparin was immobilized on PLAmembrane through the covalent bonding between phenolichydroxyl group of polydopamine and amino-group on heparin.The possible reaction mechanism was depicted in Fig. 2.

The surface chemistry of PLA membrane (PLA-M), as well asdopamine coating (DA05-M) and heparinization one (HEP05-M),were characterized by ATR-FTIR. The result is shown in Fig. 3.In terms of polydopamine-coated PLA membrane, a new broadabsorbance 3550–3100 cm�1 assigned to N–H/O–H stretchingvibrations appeared compared to PLA membrane (PLA-M).The peaks appeared at 1625 and 1519 cm�1 were attributed tothe overlap of CQC resonance vibrations in the aromatic ring andthe N–H bending vibrations, respectively. The weak absorbance at806 cm�1 was caused by the C–H out-plane bending vibrations.All new appeared peaks indicated the successful coating of poly-dopamine on PLA membrane, which is in accordance with theresults reported previously [22]. As reference to heparin-immobilized PLA membrane, all peaks mentioned above appearedand just migrated slightly.

The characteristic peaks of SO3� ascribed to heparin could notbe identified from FTIR spectra of HEP2-M in Fig. 3, due to theoverlap with PLA membrane at 1205, 1158 and 1037 cm�1. There-fore, XPS was further employed to characterize the heparincomponent on PLA membrane surface. XPS spectra of PLA-M,DA05-M and HEP05-M are shown in Fig. 4, and their correspond-ing elemental compositions are shown in Table 1. PLA membranehad no nitrogen and sulfur elements just as shown in Fig. 4,whereas, the nitrogen element (mass ratio 2.51%) was observed forpolydopamine-coated membrane DA-05, and both nitrogen (massratio 3.22%) and sulfur elements (mass ratio 0.54%) were observedfor heparin-immobilized membrane HEP05-M. All results con-firmed the successful covalent immobilization of heparin on PLAmembrane through polydopamine anchor effect.

3.2. Surface morphology of membranes

SEM images were used to observe the surface morphology asshown in Fig. 5. It can be seen that the native PLA membrane((A) PLA-M) possesses smooth and compact surfaces and finger-like cross sections, which is a typical asymmetric structureinduced by an delayed liquid–liquid demixing mechanism. Incomparison, all polydopamine-coated membranes ((C) DA2-M;(E) DA1-M; (G) DA05-M) exhibited porous surfaces despite thedopamine concentration. Furthermore, all heparin-immobilizedPLA membranes ((D) HEP2-M; (F) HEP1-M; (H) HEP05-M) showedmuch larger pores on the surface after dipping in 2.0 g/L heparin

Table 1Element mass concentration determined by XPS.

Samples C (%) O (%) N (%) S (%)

PLA-M 70.54 29.46 – –

DA05-M 60.78 36.71 2.51 –

HEP05-M 59.63 36.61 3.22 0.54

A. Gao et al. / Journal of Membrane Science 452 (2014) 390–399 393

Fig. 5. SEM images of the surface morphologies and cross-section views for different membranes. (A) PLA-M; (B) cross-section views of PLA-M; (C) DA2-M; (D) HEP2-M;(E) DA1-M; (F) HEP1-M; (G) DA05-M; (H) HEP05-M.

A. Gao et al. / Journal of Membrane Science 452 (2014) 390–399394

buffer solution for 24 h. Chong et al. have reported similarphenomenon that polyethersulfone (PES) ultra-filtration mem-brane modified by polydopamine displayed more porous surface[33], in that article the reason was deemed to the deposition ofpolydopamine spherical nanoparticles. However, different possiblereasons were first unveiled for the appearances of more poroussurfaces. In our experiments, different pH value phosphate buffersolutions (0.2 M, pH 5.5, 7.0, 8.5) without addition of dopamine

were prepared. Subsequently, PLA membranes were put into thebuffer solution, maintaining at the same time and temperature aspolydopamine-coating process. The membranes were dried andobserved by SEM. As shown in Fig. 6, significant pores appeared onthe membrane surfaces despite the pH value of buffer solution.It can be inferred that the surface pores were formed due to the

pH8.5pH7.0pH5.5

Fig. 6. SEM images of surface morphology on PLA membranes treated by different pH value phosphate buffer solution.

PLA-M: Rq8.71nm HEP05-M: Rq9.26nm

HEP2-M:Rq11.7nmHEP1-M:Rq14.2nm

Fig. 7. Surface AFM images of different PLA membranes.

Fig. 8. Water contact angle of different membranes versus drop age.

Fig. 9. Pure water flux of different membranes (0.1 MPa, 25 1C).

A. Gao et al. / Journal of Membrane Science 452 (2014) 390–399 395

buffer solution. The possible mechanism was that the salt ingre-dients in buffer solution caused polymer degradation and corro-sion for both PES and PLA membranes. Some oligomer ormolecules were evolved to induce the formation of pores due tothe mild degradation of PLA substrate. More importantly, themembrane performance was maintained and due to the briefaction time by buffer solution. It is worth noting that acidic buffersolution seems to create more porous surfaces, indicating thevarious structure regulation way and possible controlled mem-brane degradation post-treatment.

The surface roughness (Rq) of native PLA membrane and heparin-immobilized membranes are shown in Fig. 7. The native PLA mem-brane surface is mainly flat with a roughness Rq of 8.71 nm. However,the roughness increased slightly after polydopamine coating andsubsequent heparinization. Large bulges emerged as 1 g/L dopaminesolution deposited to form plentiful polydopamine aggregation. How-ever, there is no absolutely positive correlation between the roughnessRq and the dopamine solution. The roughness of HEP2-M decreased to11.7 nm compared to HEP1-M (14.2 nm), implying that excess poly-dopamine may smooth the membrane surface to some extent.

5μm

PLA-M PSf-M

HEP2-MDA2-M

DA1-M HEP1-M

HEP05-MDA05-M

Fig. 10. SEM images of platelets adhesion to all the membrane samples.

A. Gao et al. / Journal of Membrane Science 452 (2014) 390–399396

3.3. Hydrophobicity and water flux

For materials contacting human blood, the balance betweenhydrophilic and hydrophobic was important. The dynamic watercontact angle variation of PLA membranes before and aftermodification was measured and shown in Fig. 8. The contact angleof all membranes decreased slightly with drop age, despite theaddition of hydrophilic polymer PEG. The native PLA presents aninitial contact angle 841, and that of polydopamine-coated PLAmembranes decreased to DA05-M (731), DA1-M (751), and DA2-M(741) with the increase of dopamine concentration due to thehydrophilic group benzene hydroxyl on polydopamine. The con-tact angle of heparin-immobilized PLA membranes decreasedsignificantly to 411, 431and 511 for HEP2-M, HEP1-M, andHEP05-M respectively due to the introduction of more hydrophilicgroups, such as hydroxyl, carboxyl, and sulfuric acid groups. Thehydrophilic surface of PLA membrane by heparinization was infavor of the improvement of blood compatibility.

The pure water flux of different dry membranes was measured.The flux of native PLA membrane is around 120 L/h m2, anddecreased gradually with increasing dopamine content in buffersolution (DA05-M 98 L/h m2, DA1-M 90 L/h m2, DA2-M 82 L/h m2)after polydopamine coating. HEP05-M presents better hydrody-namic permeability performance compared to the rest. Effects ofpolydopamine nano-layer to the pure water flux of polymericmembranes have been studied before; however, the results were

inconsistent. It might decrease sharply only after coating forseveral hours [22], or also decrease mildly [34]. For PLA dialysismembrane, 70% flux is kept after coating for 8 h, which is due tothe slightly improved hydrophilic and newly formed nano-scalepores. As shown in Fig. 9, the surface immobilized heparinizationis helpful to improve water flux, which is mainly benefited fromthe improvement of hydrophilicity.

3.4. Blood compatibility

PLA has been used as tissue engineering materials for a longtime. So its histocompatibility has been researched in depth,including cell adhesion, anti-inflammatory, cytotoxicity etc. Thoseresults proved that PLA shows outstanding biocompatibility prop-erty. However, the biocompatibility research on PLA dialysismembrane was not reported before. As relating to hemodialysis,platelet thrombus forming caused by platelet adhesion and activa-tion was the most crucial interaction between material surface andblood ingredients. In this study, platelet adhesion and plasmarecalcification time were employed to evaluate the anticoagulationproperties of native PLA membrane and the heparin-immobilizedPLA membranes. PSf ultrafiltration membrane was prepared andadopted as the control sample due to its extensive application inhemodialysis.

Platelet adhesion results of all samples are revealed in Fig. 10. PSfmembrane adsorbed more platelets in the same time quantum andmore serious aggregation occurred compared to native PLA mem-brane, whereas, the shape of platelets adsorbed on PLA membraneseemed more irregular. Notably, the amount of platelets did notdecrease for DA2-M, DA1-M and DA05-M, which is quite differentfrom the previous reports. They assumed that the membranes withpolydopamine layer will adsorb fewer platelets [22], but strongerplatelet adhesion was occurred to all polydopamine-coated PLAmembrane within our study. In consideration of the strong adhesion

Fig. 11. Plasma recalcification time of different membranes samples.

PLA: Rq8.71nmPSF: Rq4.90nm

Fig. 12. The surface roughness of PSF and PLA native membranes.

Table 2Hemolysis ratio of different membrane samples.

Sample Absorbance (545 nm) Hemolytic rate (%)

Positive reference 0.98670.012Negative reference 0.03270.0020PLA-M 0.06370.0015 3.24PSF-M 0.06170.0020 3.04DA2-M 0.05470.0025 2.30DA1-M 0.05770.0015 2.62DA05-M 0.05270.0005 2.10HEP2-M 0.04670.0010 1.46HEP1-M 0.04870.0015 1.68HEP05-M 0.04570.0010 1.36

A. Gao et al. / Journal of Membrane Science 452 (2014) 390–399 397

and combination property of polydopamine, it is also reasonable thatpolydopamine nano-layer can promote the platelet adhesionalthough the improvement of the hydrophilicity for a certain extent.As expected, there were only very few platelets adsorbed on HEP2-M,HEP1-M and HEP05-M as shown in Fig. 10. The surface heparinizationsignificantly suppressed the aggregation of platelets and improvedthe biocompatibility of PLA membranes due to the good anti-coagulation ability of heparin and better hydrophilicity [35].

Plasma recalcification time (PRT), also named partial thrombo-plastic time (PTT), was used to monitor the time taken for clottingof blood or to determine the deficiency of factor responsible forclotting. If coagulation factor VII is activated, the thrombin will beproduced from thrombinogen via the stepped active process. Andthen the thrombin will spur the transform from fibrinogen in theplasma into fibrous protein, which is the foundation of thrombusformation. With the assistance of Ca2� and activated factor VIII,fibrous protein will be crosslinking, to be steady and insoluble.From Fig. 11, we can see that native PLA membrane exhibitedshorter plasma recalcification time (223 s) compared to PSf mem-brane (272 s), possibly attributed to the higher roughness asshown in Fig. 12 and release of bits of acid substances producedby the degradation of native PLA. However, heparin-immobilizedPLA membranes exhibited excellent prolongation of plasma recal-cification time (270–280 s) compared to native PLA membranes,PSf membranes and polydopamine-coated PLA membranes. Thesulfonic groups of heparin have been thought to play an importantrole to restrict the transform of fibrinogen [23].

Hemolysis ratio (HR) is another aspect of blood compatibility.Generally, it is used to detect the erythrocyte damage caused bymaterials. Table 2 shows the HR value of different membranesamples. We can find that PSf and PLA native membrane haveconsiderable HR, 3.24% and 3.04% respectively, less than 5%, withinsafety levels for biomaterials. The HR of polydopamine-coated PLAmembranes decreases slightly, and that of heparin-immobilized

PLA membranes decreases to 1.46%, 1.68%, and 1.36% for HEP2-M,HEP1-M and HEP05-M respectively. All results indicate thatheparin can also weaken the damage to erythrocyte in additionto preventing platelet adhesion and blood clotting.

3.5. Dialysis performance

The urea and lysozyme clearance, BSA retention of HEP05-Mand PLA-M were determined to evaluate the dialysis performancesof PLA membranes. From Fig. 13, we can figure out that 79% of ureahas been cleaned for HEP05-M after simulating dialysis for 4 h,which is slightly higher than 74.6% of PLA-M. With respect tomiddle molecule clearance, 13.7% of lysozyme was cleaned out forPLA-M after simulating dialysis for 3 h, and 18.5% for HEP05-M.After running for 4 h, lysozyme concentration was raised to someextents for both membrane samples as shown in Fig. 13(b),indicating a saturation clearance for lysozyme. Whereas, the BSArejection value of PLA-M was slightly higher than that of HEP05-M(93.7% and 90.8%) possibly due to its denser membrane surface.

4. Conclusion

In this work, PLA porous membrane aimed to hemodialysis wasprepared firstly via the phase inversion process. The surfaceheparinization was realized through the self-polymerization andstrong adhesion of dopamine. It can be concluded from the resultsthat: (1) the surface heparinization chemistry was confirmed byboth FTIR and XPS, all PLA membranes exhibited asymmetric finger-like pore structure; (2) the heparin-immobilized PLA membranesshowed smooth surface, better hydrophilicity and permeability;(3) the native PLA membrane exhibits better platelet adhesion,shorter plasma recalcilication time and considerable hemolysis ratiocompared to PSf membrane; (4) the surface heparinization on PLA

Fig. 13. Solute transmission efficiency of PLA-M and HEP05-M: (a) urea clearance percentage; (b) lysozyme clearance percentage; (c) BSA retention percentage.

A. Gao et al. / Journal of Membrane Science 452 (2014) 390–399398

membrane improved the platelet adhesion, plasma recalcilicationtime and decreased hemolysis ratio significantly; (5) the dialysissimulation test revealed that heparin-immobilized PLA membranecleaned almost 80% urea, 18% lysozyme and retained over 90% BSA.All results demonstrate there is a great potential of heparin-immobilized PLA membranes in the application of hemodialysis,and more work e.g. membrane micro-structure regulation, steriliza-tion, controlled degradation post-treatment etc. need to be done toachieve the clinical application of PLA dialysis membranes in nearfuture.

Acknowledgment

The authors would like to thank National Natural ScienceFoundation of China (51273211), National 863 Foundation of China(2012AA03A605), the International Cooperation Project from Min-istry of Science and Technology of China (2012DFR50470).

References

[1] L. Robert, David A. Tirrell, Designing materials for biology and medicine,Nature 428 (2004) 487–492.

[2] C.M. Kee, A. Idris, Permeability performance of different molecular weightcellulose acetate hemodialysis membrane, Sep. Purif. Technol. 75 (2010)102–113.

[3] K. Ishihara, R. Takayama, N. Nakabayashi, K. Fukumoto, J. Aoki, Improvement ofblood compatibility on cellulose dialysis membrane: 2. Blood compatibility ofphospholipid polymer grafted cellulose membrane, Biomaterials 13 (1992)235–239.

[4] L.F. Hancock, S.M. Fagan, M.S. Ziolo, Hydrophilic, semipermeable membranesfabricated with poly(ethylene oxide)–polysulfone block copolymer, Biomater-ials 21 (2000) 725–733.

[5] J.Y. Park, M.H. Acar, A. Akthakul, W. Kuhlman, A.M. Mayes, Polysulfone-graft-poly(ethylene glycol) graft copolymers for surface modification of polysulfonemembranes, Biomaterials 27 (2006) 856–865.

[6] T. Hasegawa, Y. Iwasaki, K. Ishihara, Preparation and performance of protein-adsorption-resistant asymmetric porous membrane composed of polysulfone/phospholipid polymer blend, Biomaterials 22 (2001) 243–251.

[7] R.M. Rasal, A.V. Janorkar, D.E. Hirt, Poly(lactic acid) modifications, Prog. Polym.Sci. 35 (2010) 338–346.

[8] B. Gupta, N. Revagade, J. Hilborn, Poly(lactic acid) fiber: an overview, Prog.Polym. Sci. 32 (2007) 455–462.

[9] S. Saeidlou, M.A. Huneault, H. Li, C.B. Park, Poly(lactic acid) crystallization,Prog. Polym. Sci. 37 (2012) 1657–1677.

[10] Z. Ma, C. Gao, Y. Gong, J. Shen, Cartilage tissue engineering PLLA scaffold withsurface immobilized collagen and basic fibroblast growth factor, Biomaterials26 (2005) 1253–1259.

[11] S. Shi, X.H. Wang, G. Gao, M. Fan, Preparation and characterization ofmicroporous poly(D,L-lactic acid) film for tissue engineering scaffold, Int. J.Nanomed. 5 (2010) 1049–1055.

[12] N.M.S. Bettahalli, H. Steg, M. Wessling, D. Stamatialis, Development of poly(l-lactic acid) hollow fiber membranes for artificial vasculature in tissueengineering scaffolds, J. Membr. Sci. 371 (2011) 117–126.

[13] M.J. Ellis, J.B. Chaudhuri, Poly(lactic-co-glycolic acid) hollow fibre membranesfor use as a tissue engineering scaffold, Biotechnol. Bioeng. 96 (2007) 177–187.

[14] T. Tanaka, M. Ueno, Y. Watanabe, T. Kouya, M. Taniguchi, D.R. LLoyd, Poly(L-lactic acid) microfiltration membrane formation via thermally inducedphase separation with drying, J. Chem. Eng. Jpn. 44 (2011) 467–475.

[15] T. Tanaka, T. Nishimoto, K. Tsukamoto, M. Yoshida, T. Kouya, M. Taniguchi,et al., Formation of depth filter microfiltration membranes of poly(l-lacticacid) via phase separation, J. Membr. Sci. 396 (2012) 101–109.

[16] A. Moriya, T. Maruyama, Y. Ohmukai, T. Sotani, H. Matsuyama, Preparation ofpoly(lactic acid) hollow fiber membranes via phase separation methods,J. Membr. Sci. 342 (2009) 307–312.

[17] W. Hu, Z.M. Huang, Biocompatibility of braided poly(L-lactic acid) nanofiberwires applied as tissue sutures, Polym. Int. 59 (2010) 92–99.

[18] J.E. Bergsma, W.C.D. Bruijn, F.R. Rozema, R.R.M. Bos, G. Boering, Late degrada-tion tissue response to poly(l-lactide) bone plates and screws, Biomaterials 16(1995) 25–31.

[19] H.C. Liu, I. Lee, J.H. Wang, S.H. Yang, T.H. Young, Preparation of PLLAmembranes with different morphologies for culture of MG-63 Cells, Bioma-terials 25 (2004) 4047–4056.

[20] J.P. Chen, C.H. Su, Surface modification of electrospun PLLA nanofibers byplasma treatment and cationized gelatin immobilization for cartilage tissueengineering, Acta Biomater. 7 (2011) 234–243.

[21] R.A. Hoshi, R. Van Lith, M.C. Jen, J.B. Allen, K.A. Lapidos, G. Ameer, The bloodand vascular cell compatibility of heparin-modified ePTFE vascular grafts,Biomaterials 34 (2013) 30–31.

[22] J.H. Jiang, L.P. Zhu, X.L. Li, Y.Y. Xu, B.K. Zhu, Surface modification of PE porousmembranes based on the strong adhesion of polydopamine and covalentimmobilization of heparin, J. Membr. Sci. 364 (2010) 194–202.

[23] F.Y. Tong, X.Q. Chen, L.B. Chen, P.Y. Zhu, J.F. Luan, C. Mao, et al., Preparation,blood compatibility and anticoagulant effect of heparin-loaded polyurethanemicrospheres, J. Mater. Chem.: B 1 (2013) 447–453.

[24] I. Bjork, U. Lindahl, Mechanism of the anticoagulant action of heparin, Mol.Cell. Biochem. 48 (1982) 161–162.

[25] Y.P. Jiao, F.Z. Cui, Surface modification of polyester biomaterials for tissueengineering, Biomed. Mater. 2 (2007) 24–27.

[26] T. Sharkawi, V. Darcos, M. Vert, Poly(DL-lactic acid) film surface modificationwith heparin for improving hemocompatibility of blood-contacting bioresorb-able devices, J. Biomed. Mater. Res.: A 98 (2011) 80–87.

[27] G. Rohman, S.C. Baker, J. Southgate, N.R. Cameron, Heparin functionalisation ofporous PLGA scaffolds for controlled, biologically relevant delivery of growthfactors for soft tissue engineering, J. Mater. Chem. 19 (2009) 9265–9273.

[28] D.J. Lin, D.T. Lin, T.H. Young, F.M. Huang, C.C. Chen, L.P. Cheng, Immobilizationof heparin on PVDF membranes with microporous structures, J. Membr. Sci.245 (2004) 137–146.

[29] J. Jiang, L. Zhu, Zhu Le, Y. Xu, Surface characteristics of a self-polymerizeddopamine coating deposited on hydrophobic polymer films, Langmuir 27(2011) 14180–14187.

[30] B.D. McCloskey, H.B. Park, H. Ju, B.W. Rowe, D.J. Miller, B.D. Freeman,A bioinspired fouling-resistant surface modification for water purificationmembranes, J. Membr. Sci. 413 (2012) 82–90.

[31] R. Li, H. Wang, W. Wang, Y. Ye, Immobilization of heparin on the surface ofpolypropylene non-woven fabric for improvement of the hydrophilicity andblood compatibility, J. Biomater. Sci. – Polym. E 24 (2013) 15–20.

[32] L.P. Zhu, J.H. Jiang, B.K. Zhu, Y.Y. Xu, Immobilization of bovine serum albuminonto porous polyethylene membranes using strongly attached polydopamineas a spacer, Colloids Surf. B Biointerfaces 86 (2011) 111–118.

[33] C. Cheng, S. Li, W. Zhao, Q. Wei, S. Nie, S. Sun, et al., The hydrodynamicpermeability and surface property of polyethersulfone ultrafiltration mem-branes with mussel-inspired polydopamine coatings, J. Membr. Sci. 417–418(2012) 228–236.

[34] Q. Wei, F.L. Zhang, J. Li, B.J. LiC.S. Zhao, Oxidant-induced dopamine polymer-ization for multifunctional coatings, Polym. Chem. 1 (2010) 1430–1433.

[35] F. Ran, S. Nie, J. Li, B.H. Su, S.D. Sun, C.S. Zhao, Heparin-like macromolecules forthe modification of anticoagulant biomaterials, Macromol. Biosci. 12 (2012)116–125.

A. Gao et al. / Journal of Membrane Science 452 (2014) 390–399 399