• 검색 결과가 없습니다.

Graphene accelerates osteoblast attachment and biomineralization

N/A
N/A
Protected

Academic year: 2022

Share "Graphene accelerates osteoblast attachment and biomineralization"

Copied!
6
0
0

로드 중.... (전체 텍스트 보기)

전체 글

(1)

Graphene accelerates osteoblast attachment and biomineralization

Jia Ren, Xiaogang Zhang and Yao Chen

School of Mechanical and Electric Engineering & Collaborative Innovation Center of Suzhou Nano Science and Technology, So- ochow University, Suzhou 215123, China

Received 5 October 2016 Accepted 25 October 2016

Corresponding Author E-mail: chenyao@suda.edu.cn Tel: +86-512-6758-1872

Open Access

pISSN: 1976-4251 eISSN: 2233-4998

Article Info

Copyright © Korean Carbon Society http://carbonlett.org

Abstract

In this paper, the in vitro biocompatibility of graphene film (GF) with osteoblasts was evalu- ated through cell adhesion, viability, alkaline phosphatase activity, F-actin and vinculin ex- pressions, versus graphite paper as a reference material. The results showed that MG-63 cells exhibited stronger cell adhesion, better proliferation and viability on GF, and osteoblasts cultured on GF exhibited vinculin expression throughout the cell body. The rougher and wrinkled surface morphology, higher elastic modulus and easy out-of-plane deformation associated with GF were considered to promote cell adhesion. Also, the biomineralization of GF was assessed by soaking in simulated body fluid, and the GF exhibited enhanced miner- alization ability in terms of mineral deposition, which almost pervaded the entire GF surface.

Our results suggest that graphene promotes cell adhesion, activity and the formation of bone- like apatite. This research is expected to facilitate a better understanding of graphene-cell interactions and potential applications of graphene as a promising toughening nanofiller in bioceramics used in load-bearing implants.

Key words: graphene film, osteoblast, adhesion, viability, mineralization

1. Introduction

In recent years, carbonaceous nanomaterials including graphene and carbon nanotubes (CNTs) have attracted much attention as an important class of novel materials. They have been widely designed for structural engineering and functional device applications due to their extraordinary mechanical strength, and excellent electrical and thermal conductivities [1-4]. In particular, the choice of carbonaceous nanomaterials as toughening nanofillers in brittle ceramics seems to be an attractive option because graphene or CNT bridging around crack faces can lower the driving force for crack propagation [5-10].

Among various ceramics, the inferior fracture toughness of hydroxyapatite (HA), which otherwise has excellent bioactivity and osteointegration, limits its wide application in the load-bearing implants used in the orthopedic field. To improve their toughness, HA-based composites have been developed with some bio-inert reinforcements such as Al2O3, TiO2, and Ti. However, it is well known that an ideal reinforcement in HA should significantly improve the mechanical properties without compromising its original biocompatibility [11].

Compared with the conventional toughening agents above, it is well recognized that the main advantage of CNTs or graphene in HA-based composites is the very low filler loading required to achieve a significant toughening effect. This feature opens a promising avenue in methods for toughening HA-based composites.

Nevertheless, the biocompatibility of CNTs is still under debate due to their cytotoxic responses in organic environments, although some researchers have ascribed the cytotoxicity of CNTs to the presence of metallic catalyst particles rather than the CNT itself [11]. In this respect, advances in the synthesis of pure and large-scale graphene via chemical vapor deposition or micromechanical exfoliation of graphite are expected to fulfill the vital DOI: http://dx.doi.org/

DOI:10.5714/CL.2017.22.042 This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/

by-nc/3.0/) which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Three-dimensional porous graphene materials for environmental applications Muruganantham Rethinasabapathy, Sung-Min Kang, Sung-Chan Jang and Yun Suk Huh

KCS Korean Carbon Society carbonlett.org

pISSN: 1976-4251 eISSN: 2233-4998

REVIEWS VOL. 22 April 30 2017

(2)

Matsuoka et al. [18], in which the adhesive strength was as- sessed by the ratio of the adherent cell numbers after using 0.1% trypsin-EDTA treatment, to the untreated cell numbers.

After being cultured for 12 h, the cells were washed with phosphate buffer saline (PBS) twice and treated with 0.1%

trypsin-EDTA solution (Invitrogen) for 5 min, 10 min, 20 min, and 30 min at 37±1°C in a humidified 5% CO2 atmo- sphere. Then, each sample was washed twice with PBS to remove the detached cells. The cells adhered on the substrate were fixed with 4% paraformaldehyde and stained with rho- damine-phalloidin (Invitrogen). The cell numbers in ten pho- tographs of each substrate were counted, and measurements were performed on each sample.

The F-actin and vinculin of the adherent MG-63 cells on the GF and GP were labeled with a fluorescent dye to observe the actin cytoskeleton and focal adhesion, respectively. MG-63 cells were seeded on each sample for 12 h, and then were washed with PBS and fixed with 4% paraformaldehyde for 30 min.

Thereafter, the cells were permeabilized with 0.1% Triton X-100 for 30 min and then incubated with 2 U/mL of rhodamine-phal- loidin (Invitrogen) for 60 min to label F-actin. Subsequently, the cells were incubated with monoclonal anti-vinculin-FITC anti- body (Sigma, USA) for 12 h to label vinculin. The fluorescently stained cells were visualized using a fluorescence microscope (Olympus IX71, Japan).

Cell proliferation and viability was evaluated by WST-1 test (Cell Proliferation and Cytotoxicity Assay Kit; Beyo- time, China). After the MG-63 cells were incubated on each sample for 1 d, 3 d and 7 d, the cells were rinsed twice with PBS. Thereafter, 10 μL of WST-1 solution and 90 μL of medi- um (final dilution, 1:10) were added to each sample followed by another 2 h incubation. Supernatants were quantified using a spectrophotometer (SpectraMax Paradigm Multi-Mode De- tection Platform; Molecular Devices, USA) at 450 nm with a reference wavelength of 625 nm. Results of the WST-1 were reported as optical density.

Osteoblast differentiation was evaluated by measuring alka- line phosphatase (ALP) activity. The cells were seeded on the samples and cultured for 1 d, 3 d, and 7 d, rinsed twice with PBS, and then permeabilized with 0.1% Triton X-100 for 30 min. ALP activity was measured by using a Lab Assay ALP kit (Beyotime) according to the manufacturer’s protocol. After 4 h incubation, supernatants were quantified spectrophotometrically at 405 nm.

The osteoconductivity of the carbonaceous materials was assessed by apatite mineralization in a SBF solution, which was prepared according to the well-known Kokubo compo- sition [19]. Both samples were soaked in SBF and kept at 37±1°C in a humidified atmosphere of 5% CO2 for 1 d, 3 d, and 7 d, respectively. After the soaking period, samples were carefully washed with deionized water three times and dried at ambient temperature. Mineralization products were char- acterized by X-ray diffraction (X’Pert-Pro MPD; PANalyti- cal, the Netherlands) using Cu Kα radiation with a scanning rate of 2°/min at a range from 20° to 50°, and their morphol- ogy were examined using scanning electron microscopy (Hi- tachi S-4700, Japan).

Numerical data were analyzed using Student’s t test. Statisti- cal significance was considered to be P<0.05.

requirements for biomedical applications. In any case, graphene reinforced HA composites have been recognized as a promising material in this field [12-15].

In our previous work [12,13], we reported on the fracture toughness of graphene reinforced HA composite synthesized using spark plasma sintering, which exhibited an ~80%

improvement compared to pure HA, even when the added graphene was only 1.0% weight.

For biomaterials used in orthopedic applications, cell attachment is an initial and critical requirement, since it is essential for other cell functions, such as proliferation and mineral formation. Very recently, various investigations on the biocompatibility of graphene have demonstrated its positive effects on cell adhesion, proliferation and differentiation, while certain forms and derivatives of graphene have induced significant cell toxicity, which is considered beneficial for anticancer activity [16-19]. Kalbacova et al. [16] found human osteoblasts and mesenchymal stromal cells had improved adherence on large single layer graphene as compared with those cells on a SiO2 substrate, with the initial presence/absence of fetal bovine serum. Aryaei et al. [17] also showed that graphene is beneficial for osteoblast attachment and proliferation.

Despite the abovementioned advances in the biocompatibility of graphene, there are still concerns regarding cell functions on the surface of graphene, and especially the effect of graphene substrates on cell behavior.

Given the present scenario, the aim of this investigation was to explore the biocompatibility of a graphene film (GF) through in vitro experiments, including osteoblast adhesion, osteoblast proliferation and bone-like apatite deposited in a simulated body fluid (SBF) solution. For this study, graphite paper (GP) with the same composition as graphene was selected as the reference material.

2. Experimental

Chemical vapor deposition (CVD) grown GF (thickness, 20–

50 μm; diameter, 3 mm) from Nanjing Emperor Nano Material (China) were employed, and GP (thickness, ~1 mm; diameter, 3 mm) was selected as a reference material. Atomic force mi- croscopy (AFM; MultiView 1000, Nanonics Imaging Ltd.) was used to measure the surface morphology and roughness of the GF and GP.

Human osteoblast-like cell lines (MG-63) were collect- ed from the Orthopaedic Institute of Soochow University, China. All aqueous solutions were prepared with deionized water. MG-63 human osteoblast cells were cultured in Dul- becco’s Modified Eagle’s Medium (Thermo Scientific, USA) supplemented with 10% fetal bovine serum (Invitrogen, USA) at 37±1°C in a humidified 5% CO2 atmosphere. When the cultured cells reached confluence, they were detached us- ing 0.25% trypsin-ethylenediaminetetraacetic acid (trypsin- EDTA) solution (Invitrogen) and were used for the designed experiments described below, where the carbonaceous ma- terials were placed into a 96-well plate and seeded with a density of 2×104 cells/mL.

The adhesive strength of the MG-63 cells to a carbona- ceous material was tested based on the method proposed by

(3)

roughness and curled nanostructure of the MWCNTs, given the same chemical composition of MWCNTs and SWCNTs [18].

Cell adhesion on the surface of a solid material is mainly me- diated by the focal adhesion protein and anchor points of the cy- toskeleton. Therefore, F-actin and vinculin expressions on both carbon materials were visualized. It is clear from Fig. 3a and c that the number of osteoblasts adhered to the GF is much higher than that on the GP. Moreover, cultured osteoblasts with spindle- like morphology are spread in a flat fashion on the GF (Fig. 3b), whereas the osteoblast cultured on the surface of the GP seems to exhibit a smaller cell body and limited spreading (Fig. 3d).

This indicates that the cells attached to the graphene exhibit spreading. Furthermore, it is clear that the F-actin (orange-red) is much denser and more widely spread throughout the cell bod- ies on the GF as compared to the GP cultured osteoblasts. Most importantly, vinculin (green) is distributed throughout the cell bodies cultured on the GF, while it is only located in a limited area within a cell on the GP, as clearly depicted in Fig. 3b and d.

From the results described above, the stronger adhesion of os- teoblasts to the GF might be ascribed to the substrate structure, considering that GF and GP share the same composition. The importance of the presence of nano and submicron roughness comes from the widely accepted hypothesis that nanostructures can enhance the absorption of adhesion proteins [20], which are thought to mediate the adhesion of anchorage-dependent cells such as osteoblasts. As such, the nano and submicron textured surface of the GF offers a greater number of opportunities for these proteins to interlock with nano/submicron asperities. This is evident given the formation of cytoskeleton with F-actin, and the focal adhesion with vinculin throughout the cell body on the GF (Fig. 3).

Also, it has been demonstrated that substrate stiffness sig- nificantly affects the cell attachment and proliferation, and that stiffer substrates are more cell-friendly [21,22]. It is well known that the elastic modulus of graphene (~1 TPa) is much higher than that of graphite, and cells can sense the substrate down to several micrometers [17]. Therefore it is expected that a GF with a thickness of ~1 mm would be easily sensed by the cultured osteoblasts.

3. Results and Discussion

As shown in Fig. 1 in the AFM images of the GF and GP (capture area 50 μm×50 μm), the roughness factor of the GF was determined to be about root mean square=731±133 nm, while for the GP root mean square=336±23 nm. It is interest- ing to note that a number of wrinkles are visible on the surface of the GF (Fig. 1a), which contributed to the greater level of nano-roughness of the GF, as compared with the comparatively smooth surface of the GP (Fig. 1b).

Fig. 2 shows the ratio of adherent cells after 5–30 min treat- ment with 0.1% trypsin-EDTA. It is clear that the ratio of the adherent cells on both samples decreases with increasing treat- ment time. It should be noted that the ratio of adherent cells on the GF is higher than that on GP at each time point; in particular,

~15% of the cells cultured on the GF remain after 5 min of treat- ment. The results imply that the GF exhibits relatively stronger adhesive strength with osteoblasts.

Similarly, Matsuoka et al. [18] reported that the cell adhe- sive strength of multi-walled carbon nanotubes (MWCNTs) was higher than that of single-walled carbon nanotubes (SWCNTs), and it has been speculated that the increase is the result of the

Fig. 1. Atomic force microscopy images of as received graphene film (a) and graphite paper (b).

Fig. 2. Ratio of the adherent cell numbers after 5–30 min treatment with 0.1% trypsin-ethylenediaminetetraacetic acid treatment. Value are presented as mean±standard error (n=3). *P<0.05 between graphene film and graphite paper.

Fig. 3. Low- and high-magnification fluorescent images of F-actin (orange-red) and vinculin (green) of adherent osteoblasts on graphene film (a, b) and graphite paper (c, d).

(4)

mention that osteoblasts are anchorage-dependent cells, i.e., they will die if they do not become well-adhered to the surface of a solid material.

As for the ALP activity, it is evident from Fig. 4b that the ALP activity of the cells on the GF is relatively higher than those cultured on the GP substrate at day 3 and day 7, although it was not remarkable on day 1 of the culture. Hence, the WST and ALP results show strong evidence that the surface of the GF substrate can provide favorable conditions for osteoblast adhe- sion, growth and proliferation.

Fig. 5 shows the X-ray diffraction spectra of the two sub- strates mineralized in SBF at 37°C. It can be seen that no apatite forms on the surfaces of the GF and GP after 1 d and 3 d im- mersion (Fig. 5a and b), and the peak of HA is present when the two substrates were immersed in SBF for 7 d. Nevertheless, it should be noted that the peak intensity of HA on the surface of the GF is much higher than that on the surface of the GP (Fig.

However, it should be noted that, so far, no evidence has con- firmed that cells can sense these substrates with an elastic modu- lus higher than few hundred kilopascals, and this needs further investigation. Moreover, Aryaei et al. [17] reported that gra- phene can sustain stress and allow for easy out-of-plane defor- mation, and subsequently contribute to the formation of strong anchor points of the cytoskeleton [23]. Therefore, the enhanced cell adhesion to graphene might be ascribed to the combination of several features: the rougher and wrinkled surface morphol- ogy, higher elastic modulus and easy out-of-plane deformation associated with the GF.

The viability of osteoblasts cultured on both substrates is de- picted in Fig. 4a. Despite negligible changes in the viability of the cells cultured on both GF and GP, respectively, at day 1 and day 3, it should be pointed out that the viability of osteoblasts cultured on the GF increased by up to ~13.2% as compared with those cultured on the GP substrate at day 7. It is important to

Fig. 4. (a) Viability of MG-63 cells cultured on graphene film (GF) and graphite paper (GP) for 1 d, 3 d, and 7 d and (b) alkaline phosphatase activity of MG- 63 cells cultured on GF and GP for 1 d, 3 d, and 7 d. Value are presented as mean±standard error (n=3). *P<0.05 between GF and GP.

Fig. 5. X-ray diffraction results of (a) as-received graphene film (GF) and graphite paper (GP) samples and GF and GP soaked in simulated body fluid for (b) for 1 d and (c) for 7 d (c).

(5)

the entire surface of the graphene, imparting good osteoconduc- tivity to the graphene. These findings are expected to advance the potential application of graphene as a toughening nanofiller in bioceramics used in load-bearing implants.

Conflict of Interest

No potential conflict of interest relevant to this article was reported.

Acknowledgements

Y.C. would like to acknowledge financial supports from the National Natural Science Foundation of China (51471113, 51275326) and Science and Technology Support Program of Ji- angsu Province, China (BE2013062).

References

[1] Lee C, Wei X, Kysar JW, Hone J. Measurement of the elastic prop- erties and intrinsic strength of monolayer graphene. Science, 321, 385 (2008). https://doi.org/10.1126/science.1157996.

[2] Huang BR, Chan HW, Jou S, Chen GY , Kuo HA, Song WJ. Struc- ture and field emission of graphene layers on top of silicon nanow- ire arrays. Appl Surf Sci, 362, 250 (2016). https://doi.org/10.1016/j.

apsusc.2015.11.256.

[3] Frank IW, Tanenbaum DM, Van der Zande AM, McEuen PL. Me- chanical properties of suspended graphene sheets. J Vac Sci Tech- nol B Nanotechnol Microelectron, 25, 2558 (2007). https://doi.

org/10.1116/1.2789446.

[4] Steurer P, Wissert R, Thomann R, Mülhaupt R. Functionalized gra- phenes and thermoplastic nanocomposites based upon expanded graphite oxide. Macromol Rapid Commun, 30, 316 (2009). https://

doi.org/10.1002/marc.200800754.

[5] Zhan GD, Kuntz JD, Wan JL, Mukherjee AK. Single-wall car- bon nanotubes as attractive toughening agents in alumina-based nanocomposites. Nat Mater, 2, 38 (2003). https://doi.org/10.1038/

nmat793.

[6] Chen Y, Zhang YQ, Zhang TH, Gan CH, Zheng CY, Yu G. Carbon nanotube reinforced hydroxyapatite composite coatings produced through laser surface alloying. Carbon, 44, 37 (2006). https://doi.

org/10.1016/j.carbon.2005.07.011.

[7] Rafiee MA, Rafiee J, Wang Z, Song H, Yu ZZ, Koratkar N. En- hanced mechanical properties of nanocomposites at low graphene content. ACS Nano, 3, 3884 (2009). https://doi.org/10.1021/

nn9010472.

[8] Kun P, Tapasztó O, Wéber, F, Balázsi C. Determination of structur- al and mechanical properties of multilayer graphene added silicon nitride-based composites. Ceram Int, 38, 211 (2012). https://doi.

org/10.1016/j.ceramint.2011.06.051.

[9] Kvetková L, Duszová A, Hvizdoš P, Dusza J, Kun P, Balázsi C.

Fracture toughness and toughening mechanisms in graphene plate- let reinforced Si3N4 composites. Scr Mater, 66, 793 (2012). https://

doi.org/10.1016/j.scriptamat.2012.02.009.

[10] Walker LS, Marotto VR, Rafiee MA, Koratkar N, Corral EL.

Toughening in graphene ceramic composites. ACS Nano, 5, 3182 5c), implying that the GF seems to provide favorable conditions

for the formation of bone-like apatite.

Fig. 6 shows the representative morphologies of the bone-like apatite formed on the GF and GP at 37°C for 7 d, respectively.

It is clearly observed that the mineralization area ultimately per- vades the whole surface of the GF, while the mineral deposit is distributed in limited locations on the GP (Fig. 6a and c). As ob- served in the AFM images (Fig. 1), the higher roughness of the GF is expected to provide more HA nucleation sites, which are mainly dependent upon the roughness of a solid material surface immersed in SBF. This subsequently leads to the much greater number of HA grains on the GF than on the GP (Fig. 6a and c).

These findings indicate that graphene facilitates the formation of bone-like apatite on its surface.

Moreover, by comparison with Fig. 6b and d, the morphology of the mineralization product varies dramatically with the car- bon material substrates, i.e., hemispherical apatite consisting of nanoparticles is formed on the GF, whereas bud-shaped apatite with nanofibers forms on the graphite substrate. The difference in the morphological selection of mineralization product might be the result of precipitation behavior related to the graphene structure, which will be explored in our future investigation.

4. Conclusions

In this study, GF and GP were used as substrates to evaluate biocompatibility and biomineralization through a cell adhesion strength test, WST-1 assay, ALP activation and SBF immersion.

The results suggest that osteoblasts exhibit enhanced biocom- patibility with graphene. In particular, the cells cultured on the GF exhibited higher adhesion strength, better proliferation and differentiation owing to the improved formation of cytoskeleton with F-actin, and focal adhesion with vinculin throughout the cell body. This might be ascribed to the nano/submicron surface toughness, higher elastic modulus and easy out-of-plane defor- mation of the graphene. Moreover, SBF soaking experiments indicated that the area of apatite mineralization pervaded nearly Fig. 6. Low- and high-magnification scanning electron microscopy im- ages of apatite formation on graphene film (a, b) and graphite paper (c, d) after 7 d of soaking in simulated body fluid.

(6)

[17] Aryaei A, Jayatissa AH, Jayasuriya AC. The effect of graphene substrate on osteoblast cell adhesion and proliferation. J Biomed Mater Res Part A, 102, 3282 (2014). https://doi.org/10.1002/

jbm.a.34993.

[18] Matsuoka M, Akasaka T, Totsuka Y, Watari F. Strong adhesion of Saos-2 cells to multi-walled carbon nanotubes. Mater Sci Eng B, 173, 182 (2010). https://doi.org/10.1016/j.mseb.2009.12.044.

[19] Kokubo T, Kushitani H, Sakka S, Kitsugi T, Yamamuro T. Solu- tions able to reproduce in vivo surface-structure changes in bio- active glass-ceramic A-W3. J Biomed Mater Res Part A, 24, 721 (1990). https://doi.org/10.1002/jbm.820240607.

[20] Webster TJ, Ergun C, Doremus RH, Siegel RW, Bizios R. Spe- cific proteins mediate enhanced osteoblast adhesion on nano- phase ceramics. J Biomed Mater Res Part A, 51, 475 (2000).

https://doi.org/10.1002/1097-4636(20000905)51:3<475::AID- JBM23>3.0.CO;2-9.

[21] Discher DE, Janmey P, Wang YL. Tissue cells feel and respond to the stiffness of their substrate. Science, 310, 1139 (2005). https://

doi.org/10.1126/science.1116995.

[22] Palchesko RN, Zhang L, Sun Y, Feinberg AW. Development of Polydimethylsiloxane substrates with tunable elastic modulus to study cell mechanobiology in muscle and nerve. PLoS One, 7, e51499 (2012). https://doi.org/10.1371/journal.pone.0051499.

[23] Deepachitra R, Chamundeeswari M, Kumar BS, Krithiga G, Prabu P, Devi MP, Sastry TP. Osteo mineralization of fibrin-decorated graphene oxide. Carbon, 56, 64 (2013). https://doi.org/10.1016/j.

carbon.2012.12.070.

(2011). https://doi.org/10.1021/nn200319d.

[11] Lahiri D, Ghosh S, Agarwal A. Carbon nanotube reinforced hy- droxyapatite composite for orthopedic application: a review.

Mater Sci Eng C, 32, 1727 (2012). https://doi.org/10.1016/j.

msec.2012.05.010.

[12] Zhang L, Liu W, Yue C, Zhang T, Li P, Xing Z, Chen Y. A tough graphene nanosheet/hydroxyapatite composite with improved in vitro biocompatibility. Carbon, 61, 105 (2013). https://doi.

org/10.1016/j.carbon.2013.04.074.

[13] Zhang L, Zhang XG, Chen Y, Su JN, Liu WW, Zhang TH, Qi F, Wang YG. Interfacial stress transfer in a graphene nanosheet toughened hydroxyapatite composite. Appl Phys Lett, 105, 161908 (2014). https://doi.org/10.1063/1.4900424.

[14] Janković A, Eraković S, Mitrić M, Matić IZ, Juranić ZD, Tsui GCP, Tang CY, Mišković-Stanković V, Rhee KY, Park SJ. Bioac- tive hydroxyapatite/graphene composite coating and its corrosion stability in simulated body fluid. J Alloys Compd, 624, 148 (2015).

https://doi.org/10.1016/j.jallcom.2014.11.078.

[15] Liu Y, Dang Z, Wang Y, Huang J, Li H. Hydroxyapatite/graphene- nanosheet composite coatings deposited by vacuum cold spraying for biomedical applications: inherited nanostructures and enhanced properties. Carbon, 67, 250 (2014). https://doi.org/10.1016/j.car- bon.2013.09.088.

[16] Kalbacova M, Broz A, Kong J, Kalbac M. Graphene substrates promote adherence of human osteoblasts and mesenchymal stro- mal cells. Carbon, 48, 4323 (2010). https://doi.org/10.1016/j.car- bon.2010.07.045.

참조

관련 문서

In vitro antitumor activity accessed using 9L gliosarcoma cell line has been evaluated by assaying the viability of cells treated with BCNU released from the

Purpose: The objective of this study was to examine whether in vitro treatment with trans-10, cis-12 conjugated linoleic acid (t10c12-CLA) modulates the filamentous actin

This prevention with Aspirin was associated with a lesser decrease in colonic tissue zinc levels and maintained plasma zinc and zinc related enzymes CuZnSOD and

This paper describes in details of microalgae cell wall and How to degrade the cell wall using Ultrasound based water pretreatment methods.. Also this paper describes

12) Maestu I, Gómez-Aldaraví L, Torregrosa MD, Camps C, Llorca C, Bosch C, Gómez J, Giner V, Oltra A, Albert A. Gemcitabine and low dose carboplatin in the treatment of

Fig. Alkaline phosphatase levels. This shows that the level of alkaline phospha tase was suppressed in the presence of skin lesion and elevated after zinc supple

The proposed CAM-silk bioink was evalu- ated regarding printability, cell viability, and tissue dif- ferentiation by 3D printing a cartilage shape–mimicking scaffold for

A series of O-substituted 5-FU derivatives were synthe- sized and their cytotoxic activity was evaluated against BEL-7402, A549, HL-60 and MCF-7 cancer cell lines in