Sunteți pe pagina 1din 9

Materials Science and Engineering C 69 (2016) 631639

Contents lists available at ScienceDirect

Materials Science and Engineering C


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

Highly porous, low elastic modulus 316L stainless steel scaffold prepared
by selective laser melting
Jaroslav apek a,b,, Markta Machov a, Michaela Fousov a, Ji Kubsek a, Dalibor Vojtch a, Jaroslav Fojt a,
Eva Jablonsk c, Jan Lipov c, Tom Ruml c
a
b
c

Department of Metals and Corrosion Engineering, University of Chemistry and Technology Prague, Technick 5, 166 28 Prague 6, Czech Republic
Institute of Physics, Academy of Sciences of the Czech Republic (AS CR), Na Slovance 1999/2, 182 21 Prague 8, Czech Republic
Department of Biochemistry and Microbiology, University of Chemistry and Technology Prague, Technick 5, 166 28 Prague 6, Czech Republic

a r t i c l e

i n f o

Article history:
Received 26 May 2016
Received in revised form 8 July 2016
Accepted 12 July 2016
Available online 14 July 2016
Keywords:
Selective laser melting
316L stainless steel
Porous implants
Scaffolds

a b s t r a c t
Recently, porous metallic materials have been extensively studied as candidates for use in the fabrication of scaffolds and augmentations to repair trabecular bone defects, e.g. in surroundings of joint replacements. Fabricating
these complex structures by using common approaches (e.g., casting and machining) is very challenging. Therefore, rapid prototyping techniques, such as selective laser melting (SLM), have been investigated for these applications. In this study, we characterized a highly porous (87 vol.%) 316L stainless steel scaffold prepared by SLM.
316L steel was chosen because it presents a biomaterial still widely used for fabrication of joint replacements and,
from the practical point of view, use of the same material for fabrication of an augmentation and a joint replacement is benecial for corrosion prevention. The results are compared to the reported properties of two representative nonporous 316L stainless steels prepared either by SLM or casting and subsequent hot forging. The
microstructural and mechanical properties and the surface chemical composition and interaction with the cells
were investigated. The studied material exhibited mechanical properties that were similar to those of trabecular
bone (compressive modulus of elasticity ~0.15 GPa, compressive yield strength ~3 MPa) and cytocompatibility
after one day that was similar to that of wrought 316L stainless steel, which is a commonly used biomaterial.
Based on the obtained results, SLM is a suitable method for the fabrication of porous 316L stainless steel scaffolds
with highly porous structures.
2016 Elsevier B.V. All rights reserved.

1. Introduction
The increasing number of various orthopedic surgeries each year has
resulted in an increased demand for new implants with suitable properties for specic application. Different types of bone scaffolds and augmentations are examples of relatively recently designed orthopedic
implants [13]. These implants should mimic the microstructural and
mechanical properties of natural bone to signicantly enhance the
healing process [1,2,46].
In general, human bones can be divided into two types (i.e., cortical
and trabecular (or cancellous) bone). Important characteristics of both
types of bone are listed in Table 1. Cortical bone forms the outer part
of bones and is relatively dense. Its density is only reduced by the presence of longitudinally oriented Haversian canals. Due to the relatively
high density and uniaxially oriented Haversian canals, cortical bone
possesses relatively high mechanical property values with anisotropic
Corresponding author at: Department of Metals and Corrosion Engineering,
University of Chemistry and Technology Prague, Technick 5, 166 28 Prague 6, Czech
Republic.
E-mail address: capekj@fzu.cz (J. apek).

http://dx.doi.org/10.1016/j.msec.2016.07.027
0928-4931/ 2016 Elsevier B.V. All rights reserved.

behavior [1,7]. However, trabecular bone, which is located in the core


of bones, contains a large amount of interconnected and approximately
equiaxed pores with a diameter of hundreds of micrometers. The high
porosity of trabecular bone leads to signicantly lower mechanical
property values compared to those of cortical bone. As shown in Table 1,
the compressive modulus of elasticity of trabecular bone is as low as a
few tenths of GPa. Therefore, a low modulus scaffold material that does
not cause the stress shielding effect is needed [1,810].
The implants that serve as reinforcement or replacement of trabecular bone should possess similar structural and mechanical properties,
especially the Young's modulus. The similarity in the Young moduli of
the bone and implant is important for reduction of the stress shielding
effect, which often leads to degradation and loosening of the tissue in
the vicinity of the implant. Moreover, the interconnected porous structure of the implant allows ingrowth of new tissue into the implant, providing direct biological xation of the implant [1,2,8,1114].
Although the microstructural and mechanical characteristics of trabecular bone have been extensively studied, preparing materials that
mimic its properties remains a challenge. Metallic materials, such as
stainless steels, Co-Cr alloys or titanium alloys, have been used to fabricate different types of orthopedic implants for many years because

J. apek et al. / Materials Science and Engineering C 69 (2016) 631639

632
Table 1
Selected properties of human bone [1,8,9].
Cortical bone
Density/gcm3
Porosity/%

Compressive mechanical
properties
Tensile mechanical
properties
Trabecular bone
Density/gcm3
Porosity/%

Compressive mechanical
properties
Tensile mechanical
properties

~1.52
~312
Loading
direction
Longitudinal
Cross
Longitudinal
Cross

Ultimate
strength/MPa
~190245
~30170
~130190
~4060

Modulus of
elasticity/GPa
~1428
~725

0.20.6
~3095
Ultimate
strength/MPa
~112

Modulus of elasticity/GPa

~2

~614

~0.10.4

Next, a precise laser beam scans the powder bed according to the slice
data of a CAD model. The laser beam completely melts the powder particles and the melt quickly solidies to form a solid layer. Then, the piston moves down a dened distance, and the next powder layer is
deposited. This process is repeated layer by layer until the product is
complete. Finally, the non-sintered powder is removed, and the product
is complete or can be treated (e.g., annealing and polishing). The SLM
factors that affect the quality of the nal products include the powder
particle size and distribution, thickness of the deposited powder layer,
beam power and thickness, scanning rate and distance between scanning points [3238].
The main purpose of this study was to characterize a model 316L
stainless steel scaffold with a high interconnected porosity (i.e., nearly
90%) and mechanical properties that are similar to those of trabecular
bone. In addition, the results were compared to nearly pore-free material that was prepared by SLM [39] and hot forged 316L stainless steel
[39], which is not typically used for implant fabrication.
2. Experimental setup

these materials possess sufcient strength and ductility [15]. Unfortunately, the fabrication of these materials with designed porosity is
very difcult due to their high melting points that prevent the use of
casting methods, such as foaming of a melt or casting into a removable
mold [16].
Rapid prototyping (RP) techniques aim to be the most suitable
method for fabricating materials with structures as sophisticated as trabecular bone [17,18]. Selective laser melting (SLM), which is one of
these techniques, allows for fabrication of materials with a very-well dened external shape and negligible porosity or materials containing a
nearly nonporous matrix and pores with a designed shape, size, volume
fraction and interconnectivity. Therefore, SLM has been extensively investigated as a candidate for fabrication of different types of orthopedic
scaffolds or augmentations [17,1922]. A lot of scientic papers have
been dealing with using of SLM for fabrication of different metallic materials both bulk and porous [2325]. In the case of biomaterials, most of
those papers have been focused on Ti-based alloys or Co-based alloys
because of their enhanced biocompatibility and high strength, respectively [23,2631]. Although 316L stainless steel is still widely used biomaterial for fabrication of joint replacements and other implants,
there is a lack of publications concerning 316L stainless steel porous
scaffolds and augmentations prepared by SLM as a biomaterial. Therefore, fabrication and characterization of highly porous 316L steel structures are important because, from the corrosion point of view, it is
necessary for augmentations to be made of the same material as
implants.
Fig. 1 schematically shows the mechanism of SLM. First, a powder
layer with a dened thickness is deposited on a movable base plate.

In this study, stainless steel AISI 316L (1.4404) with a composition


consisting of 18 wt.% Cr, 12 wt.% Ni, 2.5 wt.% Mo and less than
0.03 wt.% C was used as the studied material. A gas-atomized powder
with a spherical particle shape and diameter of 1550 m was used as
the initial material. The studied material was prepared using an M2
CUSING SLM instrument made by ConceptLaser. The SLM equipment
was equipped with an Yb: YAG laser with a maximum power of
200 W and a spot diameter of 200 m. The preparation conditions
were as follows: laser power of 200 W, continual mode, layer thickness
30 m, scanning rate of 200 mm/s and hatching space 1 mm. As a result,
a porous specimen with a size of 52 52 47 mm3 was prepared. A
macrograph of this specimen is shown in Fig. 2. The porosity of the studied material was calculated using Eq. (1) [9],

P


m
 100%
abc

where P is the porosity, m a weight, a, b and c external dimensions of the


sample and is a density of a bulk 316L stainless steel (7.99 g/cm3). The
porosity of the material was 87 vol.%, and the square-shaped pores had a
side length of approximately 750 m (Fig. 2).
The material structure and the surface of the material struts were
observed using a TESCAN VEGA-3 LMU scanning electron microscope
equipped with an Oxford instruments INCA 350 EDX analyzer (SEMEDX). Then, the as-prepared material was cut using a metallographic diamond blade into samples that were used for tensile, compressive and
three-point bending tests. For each test, three rectangular samples
were used. The tests were performed at room temperature using a
LabTest 5.250SP1-VM universal loading machine. The tensile tests

Fig. 1. Schematic representation of the SLM process [39].

J. apek et al. / Materials Science and Engineering C 69 (2016) 631639

Fig. 2. Macrograph of the as-prepared porous material.

were performed according to the CSN EN ISO 6892-1 standard and the
exural tests according to the CSN EN ISO 7438 standard. For compressive tests, there are no CSN EN ISO standard for ductile materials; therefore, we used loading conditions usually used in literature (i.e. the initial
strain rate was 103 s1) [4,40,41]. From the measured data, average
values and standard deviations were calculated.
A metallographic cross-section was prepared using the standard
method. The sample was mounted into an epoxide resin, ground using
SiC papers (P220 - P4000) and polished using a diamond paste with a
particle size of 0.7 m. The nal polishing was performed using a colloidal alumina. Then, the sample was electrochemically etched in 10 wt.%
oxalic acid at a potential of 6 V for 80 s and observed using an Olympus
PME3 light metallographic microscope (LM) and SEM-EDX.
The surface composition of the samples was studied using an X-ray
photoelectron spectrometer (XPS) ESCAprobe P (Omicron Nanotechnology Ltd.) equipped with an Al K ( = 1486.7 eV) X-ray source.
The spectra were measured with an energy step of 0.05 eV and normalized to the binding energy of the C1s peak (285.0 eV). The data for the
chemical state evaluation were obtained from the NIST X-ray

633

Photoelectron Spectroscopy Database [42]. The surfaces of the samples


were prepared in the same fashion as the samples for biological tests
(i.e., grinding, polishing, cleaning and sterilization).
The SLM scaffold was cut into samples with a square base with dimensions of approximately 7 7 mm2 and 3 mm in height. The as-prepared surface was preserved to determine the inuence of the SLM
procedure on cell attachment, and the rest of the dimensions were
ground on SiC papers (up to P4000). For use as a control, samples
with the same dimensions were prepared from corresponding but
wrought and compact material. Next, all of the samples were thoroughly cleaned and degreased by sonication in hexane and acetone, respectively. Dry heat sterilization was carried out for 2 h at 180 C. Then,
the tested and control samples were placed in 24-well plates. Human
osteosarcoma cells U-2 OS (ATCC HTB-96) were seeded directly
onto the samples and placed in a 24-well plate using DMEM supplemented with 10% FBS. The seeding density was 24,000 cells/cm2. U-2
OS cells were used for cytocompatibility testing due to their resemblance to osteoblasts. After 24 h of incubation, the medium was removed, and the samples were rinsed with PBS. The cells growing on
the samples were examined using uorescence (FM) and electron microscopy (SEM). The rst set of samples was xed with 4% formaldehyde for 20 min at RT. The cell membranes were permeabilized using
Tween 20 (0.1%, 15 min), the nuclei were stained using DAPI
(0.5 gml 1, 5 min), and F-actin was visualized with phalloidinTRITC (5 gml1, 15 min). These samples were observed using an
Olympus IX81 inverted uorescence microscope (FM) equipped with
a disk scanning unit. The z-stacks were processed using the maximal
intensity projection method (several photos were taken in different
confocal planes and then assembled into the resulting picture). The
second set of samples was xed with Karnovsky's xative (2% formaldehyde, 2.5% glutaraldehyde and 2.5% sucrose in 0.2 M cacodylate
buffer) for 1.5 h for electron microscopy observation. Then, the samples were rinsed with 0.1 M cacodylate buffer and sequentially
dehydrated in ethanol (50%, 70%, 80% and 100%, 15 min). Finally,
these samples were covered with hexamethyldisilazane (100%,
2 10 min) and dried at 45 C for 4 h. Prior to the SEM observation,
the samples were sputter coated with a 10-nm-thick layer of gold.
Due to the porous nature of the 3D printed samples, quantication
of live cells and subsequent comparison to wrought material were
disabled. Therefore, only the cell morphology was evaluated by
both FM and SEM.

Fig. 3. SEM macrographs of the porous 316L stainless steel prepared by SLM taken from different angels: a) perpendicular and b) skew to the surface. White arrows show the balls on the
strut surface.

634

J. apek et al. / Materials Science and Engineering C 69 (2016) 631639

3.2. Microstructure
Fig. 4 shows micrographs of metallographic cross-section of a strut
obtained by LM. The material possessed a microstructure that is typical
of laser-treated stainless steels [35,44,45]. The cross-sections of the
melted scan tracks are visible as arcs and marked by black arrows in
Fig. 4. The distance between these arcs ranges between approximately
80 and 150 m, which suggests that the laser beam melted the material
to a depth of more than one layer of added powder (30 m). The material consisted of elongated austenitic grains that were oriented in the solidication direction. This orientation is due to heat transport during
solidication. The grains were tens of micrometers wide and several
hundred micrometers long. The grains grew throughout several melted
scan tracks in the building direction, which suggests that the grains
grew epitaxially. Fig. 5 shows SEM micrographs at higher magnication.
In this image, the grains possessed a very ne cellular-dendritic substructure. This substructure is formed due to rapid cooling of the melt
and is typical of laser-treated austenitic steels [35,36]. Neither contamination nor segregation were observed on the grain or cell boundaries
using EDS analysis.
3.3. Mechanical behavior

Fig. 4. Microstructure of the porous 316L stainless steel (LM). The black arrows show the
scan tracks.

3. Results and discussions


3.1. Macrostructure
The macrostructure of the porous 316L stainless steel acquired from
different angles is shown in Fig. 3. A shown in the image, the squareshaped pores had a side length of approximately 750 m. The struts
were cylindrically shaped with a diameter of approximately 250 m
and formed an angle of 45 with the building direction. The struts
were rough with respect to the large amount of balls observed on the
strut surface. Balling is considered as a negative phenomenon in structural materials because it increases the roughness, which may decrease
the mechanical properties and fatigue life [33,35,37,38,43]. Therefore,
other experiments should be done in order to optimize SLM parameters
and prevent the balling effect.

The tensile, compressive and exural curves are shown in Fig. 6,


where the results indicate that the material reached high levels of plasticity under all types of loading. For compression, plastic deformation
between approximately 2 and 40% was connected to an almost no increase in stress. This part of the stress-strain curve, which is typically
called the plateau, is often observed in compressive stress-strain curves
of highly porous metallic materials [14,46] and caused by the collapse of
the porous structure and subsequent compaction of the material. Plastic
deformation was also observed visually, and this result is shown in the
macrographs of the deformed samples in Fig. 7. The fracture surfaces obtained from the tensile test are represented in Fig. 8. Necking of the
struts (Fig. 8a) and a dimple-like surface morphology (Fig. 8b) were observed. These observations indicate that a plastic and transcrystalline
fracture occurred due to the deformation.
The mechanical characteristics obtained from the curves shown in
Fig. 6 are listed in Table 2. In addition, Table 2 also lists the mechanical
properties of a nearly nonporous 316L stainless steel prepared by SLM
under the same conditions as the studied material [39] and a 316L hot
forged stainless steel [39,47]. Although Fousov et al. did not evaluate
the moduli of elasticity in their study [39], the linear portion of the
stress-strain curves of the SLM and hot forged 316L stainless steels

Fig. 5. Cellular microstructure of the grains (SEM).

J. apek et al. / Materials Science and Engineering C 69 (2016) 631639

635

Fig. 6. Stress strain curves of the porous 316L stainless steel: a) tensile, b) compressive and c) exural.

Fig. 7. Macrographs of the samples deformed in the a) tensile, b) compressive and c) exural tests.

exhibited nearly the same slope. Therefore, the moduli of elasticity can
be assumed to be the same (i.e., approximately 200 GPa) [47].
The compressive behavior of the porous 316L stainless steel was
comparable to that of trabecular bone (Table 1) but the tensile modulus
of elasticity was lower. However, the UTS and even TYS of the porous
316L stainless steel were higher than the UTS of trabecular bone.
Based on the obtained data, the prepared material exhibited mechanical
properties that were suitable for fabrication of trabecular bone scaffolds
and augmentations.

The mechanism of structure collapse during compressive deformation can be estimated from the value of the density factor (n) in the Gibson-Ashby model for compressive yield strength (CYS), which is
expressed by Eq. (2) [6,14,41,48],
X rel C  rel n

where C is a constant inuenced by the pore shape and size, Xrel is a relative value for the mechanical property, rel is the relative density and n

Fig. 8. Fracture surface obtained from the tensile test: a) strut necking and b) dimple-like morphology of the fracture surface.

J. apek et al. / Materials Science and Engineering C 69 (2016) 631639

636

predominant deformation mechanism during compressive deformation


[48].

Table 2
Mechanical properties of the studied and reference materials [39,47].

Tensile modulus of elasticity


[GPa]
Tensile yield strength [MPa]
Ultimate tensile strength [MPa]
Compressive modulus of
elasticity [GPa]
Compressive yield strength
[MPa]
Flexural modulus of elasticity
[GPa]
Flexural yield strength [MPa]
References

Porous SLM
316L

Nonporous SLM
316L

Hot forged
316L

0.12 0.02

193 3

198 4

3.46 0.25
14.55 1.56
0.15 0.03

567 15
635 23
187 5

622 7
717 13
195 2

3.01 0.13

497 24

505 11

0.20 0.05

3.82 0.42
This study

[29]

[29,35]

is the density factor. The Xrel and rel values can be calculated as a fraction of X or of the porous and compact material [14,41,48]. Three
mechanisms of strut deformation are considered based on the
value of n, and these deformation mechanisms are schematically
shown in Fig. 9. The deformation mechanism is inuenced not only
by the orientation of the pores and cell struts but also by the mechanical properties of the strut material [48]. A combination of deformation mechanisms is observed frequently because some struts can be
reinforced by neighboring struts while others are not. Therefore,
some struts can deform by yielding, and other struts deform by
bending or buckling [48].
When n is equal to 1, yielding of the struts occurs. At n = 1.5, strut
bending is the main deformation mechanism, and at n = 2, buckling
of the struts predominates [48]. The obtained constants (i.e., C and n)
can also be used to estimate the CYS or compressive modulus of elasticity (Ec) of materials with an unknown porosity [14,48,49]. Although the
Gibson-Ashby model is typically used to describe compressive behavior
in the literature, in our study [6], this mode provided a good approximation of exural behavior. Similarly, we assumed that the modied Gibson-Ashby model could also be used to estimate the estimate tensile
properties of materials with different porosities.
To calculate the constants in the Gibson-Ashby model, the data listed
in Table 2 was used. The compressive and tensile yield stresses (CYS and
TYS) as well as the ultimate tensile stress (UTS) of the nonporous SLM
316L stainless steel were used as the values of the compact material because the microstructure of this material was more similar to the strut
microstructure of the porous material. Therefore, these data were assumed to be more appropriate for the approximation than those of
the hot forged 316L stainless steel. The C and n constants were very similar for all of the analyzed properties and are listed in Table 3. The n
value for CYS (1.14) suggests that yielding of the pore struts was the

3.4. Surface composition


The surface composition of the porous 316L stainless steel that was
prepared by SLM was compared to that of the hot forged one, which
possesses good biocompatibility. This analysis was performed because
the surface composition strongly inuences the biological compatibility
of the material with the living system. The surface composition determined by XPS is summarized in Table 4. In addition to the matrix elements, contamination carbon and oxidic oxygen were detected on the
surface of both materials. Nickel and molybdenum were only detected
on the surface of the hot forged 316L stainless steel with binding energies corresponding to NiO (855.6 eV) and MoO3 (232.6 eV), respectively
[50]. Based on comparison of only the concentrations of iron, chromium
and other alloying elements (Table 4), the surface of the SLM sample
was strongly enriched by chromium. This surface enrichment by chromium occurred due to the high afnity of chromium for oxygen,
whose residues were present in the protective argon atmosphere, at
high temperatures during the SLM process. The hot forged 316L stainless steel was also subjected to high temperatures but the surface
layer was removed by machining, and the oxidic surface layer was
formed during sterilization at 180 C. This low temperature is not sufcient to form an oxidic layer with a chromium content as high as that in
the SLM 316L stainless steel.
The oxidation states of iron and chromium were the same for both
materials (Figs. 10 and 11). The Fe 2p 3/2 spectrum was tted to three
peaks with binding energies corresponding to a mixture of Fe2O3,
Fe3O4 and chromite (FeCr2O4) [50,51]. The binding energy of Cr 2p indicates the presence of Cr2O3 or chromite (FeCr2O4) [50,51]. Based on a
comparison of the Fe and Cr spectra, the signal from the surface of the
SLM material was weaker than that from the hot forged one. The high
content of Cr in the surface layer of the SLM 316L stainless steel should
not have a negative effect on the cytocompatibility because it was
bound as Cr2O3 or chromite, which plays a main role in the corrosion resistivity of stainless steels and is barely soluble in the body. Therefore, a
minimum amount of metallic ions would be released into the environment and be harmful for the cells. This assumption was conrmed by
the in vitro cytocompatibility tests, and these results are discussed in
the next section.
3.5. Contact in vitro cytocompatibility tests
Fig. 12a-c show the cells observed by uorescence microscopy. The
nuclei that were stained by DAPI are shown in blue, and F-actin visualized by TRITC-phalloidin is shown in red. Fig. 12c and C represent cells
seeded on control slides, which serve as a second control and an

Fig. 9. Schematic representation of the mechanisms of strut deformation: a) yielding, b) bending and c) buckling [48].

J. apek et al. / Materials Science and Engineering C 69 (2016) 631639

637

Table 3
Gibson-Ashby constants (i.e., C and n) (Eq. (2)) for different mechanical properties.

Tensile modulus of elasticity [GPa]


Tensile yield strength [MPa]
Ultimate tensile strength [MPa]
Compressive modulus of elasticity [GPa]
Compressive yield strength [MPa]

0.71
0.72
0.75
0.71
0.75

1.15
1.14
1.12
1.15
1.14

Table 4
Surface composition determined by XPS (only metallic elements included; weight %).

Hot forged AISI 316L


SLM AISI 316L

Fe

Cr

Ni

Mo

74.4
58.2

15.0
41.8

2.8

7.8

example of a healthy morphology. Based on these images, the cells that


attached to the SLM porous material look healthy, and their morphology
after one day is comparable to those on the control material as well as
on the control slide. The cells were well spread with developed actin
bers.
The lower number of cells was caused by the porous nature of the
SLM scaffold. The cells were concentrated, especially in the joints of
struts that form shallow depressions. The rest of the cell suspension,
which was not retained in these locations during seeding, passed
through the porous structure to the bottom of the testing vessel. Therefore, few cells were observed on the inclined struts. Moreover, the struts
were very rough due to the presence of balls on their surface, and therefore, the cells could not be easily distinguished using SEM.

Fig. 11. Detailed Cr 2p spectrum: a) Hot forged and b) SLM porous 316L stainless steel.

Figs. 12A-C represent the cells observed by SEM. The cells were
widespread, and lopodia attaching the cells to the material surface
were observed.
Our experiment indicated that the SLM process does not affect the
cytocompatibility of 316L stainless steel after one day of incubation.
Therefore, porous scaffolds, implants and other facilities used in medicine prepared by this additive technique as well as by other common
techniques can be further tested in vitro.

4. Conclusion
In this study, the properties of a highly porous 316L stainless steel
scaffold prepared by SLM were comprehensively investigated and its
microstructure, mechanical properties, surface chemistry and
cytocompatibility were compared to those of a nonporous 316L stainless steel prepared by SLM or casting and subsequent hot forging. The
scaffold possessed an open-cellular porous structure with squareshaped pores with a side length of approximately 750 m. The individual struts were approximately 250 m in diameter. The struts consisted
of austenitic grains elongated in the building direction. Due to rapid
cooling, the austenitic grains possessed a very ne cellular-dendritic
substructure. The porous scaffold exhibited mechanical properties
(Ec = 0.15 GPa and CYS = 3 MPa) that were relatively close to those
of trabecular bone. Although the surface of the SLM 316L stainless
steel was enriched with chromium compared to that of the hot forged
one, the cell morphology after one day of incubation on both types of
materials was comparable. Based on the obtained results, SLM is a suitable technique for fabricating highly porous implants from 316L stainless steel.

Acknowledgments

Fig. 10. Detailed Fe 2p3/2 spectrum: a) Hot forged and b) SLM porous 316L stainless steel.

The authors would like to thank the Czech Science Foundation (project no. P108/12/G043) for supporting this research.

J. apek et al. / Materials Science and Engineering C 69 (2016) 631639

638

Fig. 12. Cell morphologies on (a, A) 3D printed sample, (b, B) wrought sample and (c, C) control slide observed by uorescence microscopy and SEM.

References
[1]
[2]
[3]
[4]
[5]

S. Wu, X. Liu, K.W.K. Yeung, et al., Mater. Sci. Eng. R 80 (2014) 136.
K.H. Frosch, K.M. Strmer, Eur. J. Trauma 32 (2006) 149159.
S.J. Hollister, Nat. Mater. 4 (2005) 518524.
J. Capek, D. Vojtech, Mater. Sci. Eng. C 35 (2014) 2128.
J. apek, D. Vojtch, Mater. Sci. Eng. C 43 (2014) 494501.

[6]
[7]
[8]
[9]
[10]
[11]
[12]

J. apek, D. Vojtch, A. Oborn, Mater. Des. 83 (2015) 468482.


J. Jowsey, J. Anat. 100 (1966) 857864.
K. Alvarez, H. Nakajima, Materials 2 (2009) 790832.
V. Karageorgiou, D. Kaplan, Biomaterials 26 (2005) 54745491.
J. apek, D. Vojtch, Manuf. Technol. 15 (2015) 964969.
X. Zhang, X.W. Li, J.G. Li, et al., Mater. Sci. Eng. C 42 (2014) 362367.
B. Arifvianto, J. Zhou, Materials 7 (2014) 35883622.

J. apek et al. / Materials Science and Engineering C 69 (2016) 631639


[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]

S. Bose, S. Vahabzadeh, A. Bandyopadhyay, Mater. Today 16 (2013) 496504.


X. Wang, Y. Li, J. Xiong, et al., Acta Biomater. 5 (2009) 36163624.
Q. Chen, G.A. Thouas, Mater. Sci. Eng. R 87 (2015) 157.
T. Murakami, T. Akagi, E. Kasai, Proc. Mater. Sci. 4 (2014) 3035.
B. Vandenbroucke, J.-P. Kruth, Rapid Prototyp. J. 13 (2007) 196203.
G.E. Ryan, A.S. Pandit, D.P. Apatsidis, Biomaterials 29 (2008) 36253635.
C. Yan, L. Hao, A. Hussein, et al., J. Mech. Behav. Biomed. Mater. 51 (2015) 6173.
I. Tolosa, F. Garciandia, F. Zubiri, et al., Int. J. Adv. Manuf. Technol. 51 (2010)
639647.
A. Fukuda, M. Takemoto, T. Saito, et al., Acta Biomater. 7 (2011) 23272336.
D.K. Pattanayak, A. Fukuda, T. Matsushita, et al., Acta Biomater. 7 (2011) 13981406.
E. Liverani, A. Fortunato, A. Leardini, et al., Mat. Des. 106 (2016) 6068.
Z. Sun, X. Tan, S.B. Tor, et al., Mat. Des. 104 (2016) 197204.
A.M. Khorasani, I. Gibson, M. Goldberg, et al., Mat. Des. 103 (2016) 348355.
S. Amin Yavari, R. Wauthle, A.J. Bttger, et al., Appl. Surf. Sci. 290 (2014) 287294.
J. Vaithilingam, E. Prina, R.D. Goodridge, et al., Mater. Sci. Eng. C 67 (2016) 294303.
J.-y. Xu, X.-s. Chen, C.-y. Zhang, et al., Mater. Sci. Eng. C 68 (2016) 229240.
M. Fischer, D. Joguet, G. Robin, et al., Mater. Sci. Eng. C 62 (2016) 852859.
S. Amin Yavari, R. Wauthle, J. van der Stok, et al., Mater. Sci. Eng. C 33 (2013)
48494858.
L. Ren, K. Memarzadeh, S. Zhang, et al., Mater. Sci. Eng. C 67 (2016) 461467.
S. Bremen, W. Meiners, A. Diatlov, Laser. Technik. J. 9 (2012) 3338.
J.P. Kruth, L. Froyen, J. Van Vaerenbergh, et al., J. Mater. Process. Technol. 149 (2004)
616622.
A. Laohaprapanon, P. Jeamwatthanachai, M. Wongcumchang, et al., Mater. Manuf.
Technol. II Pts 1 and 2 (2012) 816820.

639

[35] J.A. Cherry, H.M. Davies, S. Mehmood, et al., Int. J. Adv. Manuf. Technol. 76 (2014)
869879.
[36] I. Yadroitsev, P. Krakhmalev, I. Yadroitsava, et al., J. Mater. Process. Technol. 213
(2013) 606613.
[37] R. Li, Y. Shi, Z. Wang, et al., Appl. Surf. Sci. 256 (2010) 43504356.
[38] R. Li, J. Liu, Y. Shi, et al., Int. J. Adv. Manuf. Technol. 59 (2012) 10251035.
[39] M. Fousova, D. Vojtch, J. Kubsek, et al., Manuf. Technol. 15 (2015) 809814.
[40] H. Zhuang, Y. Han, A. Feng, Mater. Sci. Eng. C 28 (2008) 14621466.
[41] X.-h. Wang, J.-s. Li, R. Hu, et al., Trans. Nonferrous Metals Soc. China 23 (2013)
23172322.
[42] NIST X-ray Photoelectron Spectroscopy Database, Version 4.0, http://srdata.nist.gov/
xps2National Institute of Standards and Technology, Gaithersburg, 2008.
[43] N.W. Hrabe, P. Heinl, B. Flinn, et al., J. Biomed. Mater. Res. Part B-Appl. Biomater. 99B
(2011) 313320.
[44] E. Yasa, J.P. Kruth, Process. Eng. 19 (2011) 389395.
[45] R. Li, J. Liu, Y. Shi, et al., J. Mater. Eng. Perform. 19 (2010) 666671.
[46] W. Niu, C. Bai, G. Qiu, et al., Mater. Sci. Eng. A 506 (2009) 148151.
[47] M.B. Nasab, M.R. Hassan, B.B. Sahari, Trends Biomater. Artif. Organs 24 (2010)
6982.
[48] Y. Yamada, C. Wen, K. Shimojima, et al., Mater. Trans. 43 (2002) 12981305.
[49] L. Gibson, Annu. Rev. Mater. Sci. 30 (2000) 191227.
[50] M.C. Biesinger, B.P. Payne, A.P. Grosvenor, et al., Appl. Surf. Sci. 257 (2011)
27172730.
[51] T. Yamashita, P. Hayes, Appl. Surf. Sci. 254 (2008) 24412449.

S-ar putea să vă placă și