You are on page 1of 16

ActaActa

Biomaterialia 99 (2019)
Biomaterialia 479–494
xxx (xxxx) xxx

Contents lists available at ScienceDirect

Acta Biomaterialia
journal homepage: www.elsevier.com/locate/actabiomat

Full length article

A manufacturing and annealing protocol to develop a cold-sprayed


Fe-316L stainless steel biodegradable stenting material
Jennifer Frattolin a, Ranjan Roy b, Sriraman Rajagopalan c, Michael Walsh d, Stephen Yue c,
Olivier F. Bertrand a,e, Rosaire Mongrain a,f,⇑
a
Department of Mechanical Engineering, McGill University, Macdonald Engineering Building, Montreal, Quebec H3A 0C3, Canada
b
Department of Chemical Engineering, McGill University, M.H. Wong Building, Montreal, Quebec H3A 0C5, Canada
c
Department of Mining and Materials Engineering, McGill University, M.H. Wong Building, Montreal, Quebec H3A 0C5, Canada
d
Biomaterials Cluster, Bernal Institute, School of Engineering, Health Research Institute, University of Limerick, Ireland
e
Interventional Cardiology Laboratories, Quebec Heart and Lung Institute, Laval University, Quebec City, Quebec G1V 4G5, Canada
f
Montreal Heart Institute, Montreal, Quebec H1T 1C8, Canada

a r t i c l e i n f o a b s t r a c t

Article history: Biodegradable stents show promise to revolutionize coronary artery disease treatment. Its successful
Received 23 March 2019 implementation in the global market remains limited due to the constraints of current generation
Received in revised form 20 August 2019 biodegradable materials. Cold gas dynamic spraying (CGDS) has been proposed as a manufacturing
Accepted 21 August 2019
approach to fabricate a metallic biodegradable amalgamate for stent application. Iron and 316L stainless
Available online 23 August 2019
xxxx
steel powders are combined in a 4:1 ratio to create a novel biomaterial through cold spray. Cold spray
processing however, produces a coating in a work hardened state, with limited ductility, which is a crit-
Keywords:
ical mechanical property in stent design. To this end, the influence of annealing temperature on the
Biodegradable stent
Cold spray
mechanical and corrosion performances of the proposed Fe-316L amalgamate is investigated. It was
Bioresorbable scaffold found that annealing at 1300 C yielded a complex material microstructure, with an ultimate tensile
Biomaterials strength of approximately 280 MPa and ductility of 23%. The static corrosion rate determined at this
Coronary artery disease annealing temperature was equal to 0.22 mg cm2 day1, with multiple corrosion species identified
within the degradation layers. Precipitates were observed throughout the microstructure, which
appeared to accelerate the overall corrosion behaviour. It was shown that cold-sprayed Fe-316L has sig-
nificant potential to be implemented in a clinical setting.

Statement of Significance

Biodegradable stents have potential to significantly improve treatment of coronary artery disease by
decreasing or potentially eliminating late-term complications, including stent fracture and in-stent
restenosis. Current generation polymer biodegradable stents have led to poorer patient outcomes in com-
parison to drug-eluting stents, however, and it is evident that metallic biomaterials are required, which
have increased strength. To this end, a novel iron and stainless steel 316L biomaterial is proposed, fabri-
cated through cold-gas dynamic spraying. This study analyses the effect of annealing on the Fe-316L bio-
material through corrosion, mechanical, and microstructural investigations. The quantitative data
presented in this work suggests that Fe-316L, in its annealed condition, has the mechanical and corrosion
properties necessary for biodegradable stent application.
 2019 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

1. Introduction conventional thermal powder deposition methods. Cold spray


allows for the deposition of pure metals, metallic alloys, and com-
Cold-gas dynamic spraying (CGDS) has emerged as an posite powders, creating a wide range of possible materials with
alternative powder processing technique to other more varied properties. It is a high-pressure powder processing tech-
nique, utilizing compressed gas (nitrogen or helium) at a pressure
⇑ Corresponding author at: Macdonald Engineering Building, 817 Sherbrooke St. of 1–5 MPa. The high-pressure gas is heated to 100–800 C and
West, Room 270, Montreal, QC H3A 0C3, Canada. travels to the gun nozzle, where it is mixed with the powder
E-mail address: rosaire.mongrain@mcgill.ca (R. Mongrain). carried by the gas jet stream from the powder feeder [1]. A

https://doi.org/10.1016/j.actbio.2019.08.034
1742-7061/ 2019 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
480
2 J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx

converging/diverging nozzle is implemented to generate gas flow maris stent, though it has been recently associated with rapid
with high velocities, propelling the powder particles from 300 to degradation, resulting in poor stent scaffolding [15]. The use of
1200 m s1. Since the powder impacts the substrate at a high cold spray presents an alternative means to fabricate metallic
velocity, the particles deform, adhere to the substrate and form a biodegradable stents. Cold spray allows for the deposition of reac-
coating [2]. The temperature of the powder remains low (below tive powders without oxidation, resulting in a biodegradable coat-
typical solid-state powder sintering temperatures), and effects nor- ing. Multiple metallic powders can be sprayed to create an
mally associated with high temperature powder processing tech- intermixed metal–metal composite with a controlled corrosion
niques, such as casting or sintering, including oxidation, thermal rate, by modifying the relative weight percentages of the powders.
stresses, and grain growth, are avoided. This is advantageous for In addition, by utilizing dissimilar metals with a large electropo-
biomaterial applications as oxygen-sensitive powders can be tential gap, galvanic corrosion can be used to further accelerate
sprayed. Furthermore, no melting is observed, and the particles the corrosion behaviour of the material. One of the remaining chal-
are able to yield and mix under very high strain rates, producing lenges associated with CGDS is the work-hardened coatings. Since
a complex microstructure [1]. The resulting microstructure has cold spray is a deformation-based technique, the available defor-
grain sizes in the micron to submicron range, resulting in high mation existing within the metallic particles is used during depo-
strength, as well as improved fatigue resistance and wear proper- sition. Furthermore, the particle–particle interface is weak, where
ties [3]. particles are largely only mechanically interlocked. This results in
The application of cold spray to the development of biomateri- overall poor ductility, with brittle fracture commonly occurring
als is an interesting alternative for the fabrication of coronary [16]. Ductility is a critical property in stent application, in order
stents, a scaffolding device used to treat coronary artery disease. to facilitate stent expansion and resist deformation under heart
Traditional manufacturing techniques, such as casting or sintering, pulsation.
utilize elevated temperatures to form the materials, operating Frattolin et al. [17] previously studied the effect of different
above the transformation temperature of the material. This is powder mixtures of iron and 316L stainless steel on the corrosion
avoided in cold spray due to its low operating temperature. In and mechanical properties, including 20% Fe and 80% 316L, 50% Fe
addition, the refined microstructure obtained through cold spray and 50% 316L, and 80% Fe and 20% 316L. It was evident from this
would allow for a thinner stent strut profile, without sacrificing study that the combination of 80% Fe and 20% 316L had the most
the scaffolding performance of the stent. This is of considerable suitable degradation properties of the different mixtures. However,
importance since strut thickness is an independent predictor of the particle interfaces were weakly bonded, and the overall mate-
restenosis risk [4]. Cold spray has already been demonstrated as rial ductility was insufficient for stent application. Therefore, the
a potential fabrication technique for permanent stents through objective of this work was to develop a manufacturing and anneal-
the work of Al-Mangour et al. [5]. ing protocol that would improve the ductility of the cold-sprayed
The application of CGDS to biodegradable stent fabrication may Fe-316L material, validating its potential for biodegradable stent
help address some of the deficiencies of current generation application.
biodegradable stents. Biodegradable stents provide temporary
scaffolding support to prevent acute recoil post-implantation but 2. Materials and methods
degrade fully to avoid late-term complications like late stent
thrombosis and stent fracture. The potential advantages of 2.1. Cold-gas dynamic spraying and sample fabrication
biodegradable stents over traditional drug-eluting stents include
decreased incidence or potential elimination of in-stent restenosis Commercially available iron and 316L stainless steel powders
and stent fracture, restoration of more native tissue behaviour, were utilized for cold spray. Pure iron powder (ATOMET 195SP
including cyclic strain and vessel pulsatility, as well as physiolog- classification) was obtained from Quebec Metal Powders (Sorel,
ical shear stress, among other benefits [6–9]. Polymers were the Canada) and 316L stainless steel powder was purchased from
material of choice for biodegradable stent application, due to their SandVik Osprey (Neath, United Kingdom). The iron and 316L stain-
known bioresorbable properties in clinical applications, such as less steel powders were characterized using multiple techniques.
biodegradable sutures. However, polymer stents pose a significant First, X-ray diffraction (XRD), coupled with quantitative phase
number of challenges. The comparatively poor mechanical proper- analysis, was used to confirm the composition of the powders. Sec-
ties of polymer stents have required thicker and wider struts than ond, inductively coupled plasma optical emission spectrometry
traditional DES to increase scaffolding strength and stiffness. This (ICP-OES) was utilized to determine the element composition of
leads to reduced deliverability due to a larger catheter profile, in the powders. Third, to examine the morphology of the powder, a
addition to established precision constraints due to the poor radio- scanning electron microscope (SEM) was used to image the pow-
graphic visualization of polymers [10,11]. The reduced stiffness of ders at 2000 magnification. Lastly, the particle size and distribu-
polymer stents also results in greater scaffold recoil in complex tion were measured with a LA-920 laser diffraction analyzer
lesions. From a clinical perspective, there are several factors (HORIBA Scientific, France).
revealed by the recent clinical trials of polymer stents that are con- The powders were prepared in a 4:1 ratio, by weight, of iron to
cerning, including the relatively high target lesion failure, the pres- 316L stainless steel. To ensure a homogeneous particle distribu-
ence of late-stent thrombosis, and poor outcomes in smaller calibre tion, the powders were then mixed utilizing a rolling mixer, with
and tortuous vessels [12,13]. Furthermore, due to the poor 2 mill balls, for 1 h. The Plasma Giken PCS-800 cold spray system
mechanical properties of polymer stents, attaining firm strut appo- (Plasma Giken Co., Ltd., Japan) was used to fabricate flat samples of
sition against the vessel wall is problematic. This has resulted in a Fe-316L. Prepared powder was preheated in a vacuum furnace as
phenomenon referred to as intraluminal scaffold dismantling, humidity may affect coating adhesion to the substrate when spray-
where stent struts protrude into the arterial lumen as they ing [18]. Two 1018 carbon steel sheets, 15 cm in length and width
degrade. This is of significant clinical concern as the protruding and approximately 0.6 cm thick, were utilized as substrates. Nitro-
struts are highly thrombogenic, increasing the risk of late scaffold gen (N2) was used as the propellant gas at a pressure of 4.0 MPa
thrombosis [14]. and a temperature of 700 C. The substrates were grit-blasted prior
These limitations may be resolved with the implementation of to spraying, as well as heated with 2 passes of nitrogen gas, to
biodegradable metallic materials. Currently, the only metallic improve powder deposition efficiency. The deposition efficiency,
biodegradable stent approved for clinical use is BioTronik’s Mag- DE, of the coating was determined utilizing the following equation:

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx 481
3

DE ¼ Dms =mp;total ð1Þ ing. Engineering ultimate tensile strength (UTS), yield stress, and
elongation at fracture were determined. Yield stress was calculated
where Dms, in grams, is the difference between the initial and final
using the 0.2% offset method. After testing, the morphology of the
mass of the substrate and mp,total, in grams, is the total mass of the
fractured surface was characterized using SEM.
sprayed particles, which is obtained directly from the CGDS system.
After cold spraying, the samples were cut into the dog-bone
tensile specimens, according to the dimensions outlined of the sub- 2.4. Corrosion testing
size sample in ASTM E8/E8M-16a [19], with a gauge length of
25 mm and a width of 6 mm. Electrical discharge machining Static and galvanic corrosion tests were implemented to quan-
(EDM) was used to remove the steel substrate from the Fe-316L tify the impact of annealing on the corrosion behaviour of Fe-316L.
coating. Samples were 10 mm � 10 mm, with an approximate thickness of
The prepared samples were then vacuum annealed at 1100 C, 1 mm. Prior to corrosion testing, samples were progressively
1200 C, or 1300 C. The vacuum annealing setup employed a Type ground with 400, 600, 800, and 1200 grit SiC paper. First, baseline
46100 high temperature furnace (Thermolyne Corporation, USA), static immersion tests were conducted to calculate the corrosion
with the samples contained in an alumina tube insert. The furnace rate of each sample, followed by galvanic corrosion tests to quan-
temperature was increased at a rate of 250 C h-1 until the required tify the impact of heat treatment on the galvanic corrosion reaction
temperature was reached, after which it was held at temperature between iron and 316L stainless steel. Both tests were completed
for 2 h. Samples were then allowed to cool at a controlled rate to in triplicate.
room temperature. After annealing, all samples were subsequently
ground and polished to improve surface finish. 2.4.1. Static corrosion testing
Static corrosion tests were conducted of each sample following
2.2. Coating characterization ASTM standard G31-12a [20]. Physiological test conditions were
implemented, with modified HyCloneTM Hanks’ Balanced Salt Solu-
The chemical composition and microstructure of the Fe-316L tion (HBSS) (Fisher Scientific, Canada) used as the corrosive media,
amalgamates exposed to different annealing conditions were which was changed every 24 h. Modified HBSS is used as its ionic
assessed and subsequently compared to as-sprayed Fe-316L. Sam- composition is similar to that of blood plasma [21]. The solution
ples (10 mm � 10 mm) were ground with up to 1200 silicon car- was buffered with HEPES to maintain a physiological pH of 7.4.
bide paper and subsequently polished with 3 mm and 1 mm The solution was held at physiological temperature of 37 C, by
diamond paste, according to standard metallographic preparation means of an ISOTEMP 202 hot water bath (Fisher Scientific,
procedures. Prior to microscopic observation, samples were Canada). As outlined by ASTM G31-12a, the solution volume to
cleaned in an ultrasonic bath with isopropanol for 10 min. sample surface area ratio (V/S) was maintained between 20 and
40 mL cm�2. The test duration was 7 days, and the samples were
2.2.1. X-ray diffraction removed from the test bench every 24 h, cleaned and weighed.
Phase analysis was performed by XRD on a D8 DISCOVER X-ray The corrosion rate of each sample, CRstatic, in mg cm�2 day�1, was
Diffractometer (Bruker Corporation, Madison, WI, USA), with a Co then calculated according to the following equation [20]:
Ka source. Scans were conducted at a voltage and current of
35 kV and 45 mA, respectively, with a 0.02 s�1 scan rate. The back- CRstatic ¼ W=At ð2Þ
ground noise was subtracted from the original scan directly within
the DIFFRAC.SUITE EVA software (Bruker Corporation, Billerica, where W is the total mass loss in mg, A is the surface area in cm2,
MA, USA). The quantitative composition of the detected austenite and t is the total exposure time in days.
and ferrite phases of each annealing condition was determined
by Rietveld refinement. In addition, the full width at half maximum 2.4.2. Galvanic corrosion testing
(FWHM) of the (1 1 0) peak was determined to assess peak broad- Galvanic corrosion tests were conducted in compliance with
ening, which was computed within the DIFFRAC.SUITE EVA ASTM standard G71-81 (2014) [22]. To induce a galvanic couple,
software. a VersaStat3 potentiostat (Princeton Applied Research, TN, USA)
was utilized. The annealed and as-sprayed samples were used as
2.2.2. Optical and scanning electron microscopy the working electrode, with the other electrode defined as pure
Specimens were characterized by optical microscopy with a 316L stainless steel, with a 1:1 surface area ratio implemented. A
Nikon Epiphot 200 microscope (Nikon Instruments Inc., Melvile, saturated calomel electrode was used as the reference electrode.
NY, USA), with Clemex Vision software (Clemex Technologies Samples were immersed for 6 h at 37 C within a jacketed reaction
Inc., Longueuil, QC, Canada) at 200� magnification. The polished beaker with 700 mL of buffered modified HBSS. Samples were sus-
samples were also observed utilizing a Hitachi SU3500 SEM (Hita- pended on wires and were exposed to agitation through magnetic
chi, Ltd., Tokyo, Japan) at an accelerating voltage of 15 kV. The ele- mixing. Galvanic current and potential were measured with
ment distribution of the samples was assessed by energy- respect to time.
dispersive X-ray spectroscopy (EDS) utilizing AZtec software The galvanic corrosion rate (CRgalv), in mg cm�2 day�1, was cal-
(Oxford Instruments Ltd., Abingdon, United Kingdom) at 500� culated by the following equation:
magnification.
CRgalv ¼ W=At ð3Þ
2.3. Mechanical testing
where A is the surface area in cm2, t is the elapsed time in days, and
W is the mass loss, in mg, equal to:
Uniaxial tensile tests were conducted on an MTS hydraulic test-
ing machine with an extensometer. Tensile tests were performed W ¼ MIt=zF ð4Þ
with the prepared dog-bone specimens at room temperature
according to ASTM E8/E8M-16a, at a strain rate of 0.001 s�1 [19]. where M is the molar mass of iron in mg mol�1, I is the stabilized
For each annealing condition, three specimens were tested. Sample galvanic current in Ampere, t is the total time in seconds, z is ion
width and thickness were re-measured after testing to assess neck- valence (z = 2), and F is Faraday’s constant, 96 485.34 C mol�1.

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
482
4 J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx

2.5. Corrosion product characterization 2.6. Statistical analysis

The corrosion products were analyzed utilizing five different Mean values were determined for relevant data sets. Statistical
techniques, four of which specifically analyzed the corrosion pro- analysis was completed in GraphPad Prism 7.04 (GraphPad Soft-
duct layer on the sample, including XRD, X-ray photoelectron spec- ware Inc., La Jolla, CA, USA). A one-way ANOVA was used to com-
troscopy (XPS), SEM, and EDS. The final characterization technique, pare data sets. Tukey’s multiple comparisons test was selected as
ICP-OES, analyzed the composition of the loosed corrosion prod- a post-hoc test, with all possible pairwise combinations tested. A
ucts in the testing solution. p-value of < 0.05 was considered significant for both tests.

3. Results
2.5.1. X-ray diffraction and scanning electron microscopy
X-ray diffraction was used to assess the composition of the cor- 3.1. Cold-gas dynamic spraying and sample fabrication
rosion products, following a similar procedure outlined in Sec-
tion 2.2.1. SEM was used to assess the morphology of the The iron and 316L stainless steel powders were characterized
degradation product layer, and the element distribution of the cor-
using SEM, laser diffraction particle size analysis, XRD, and ICP-
rosion products on the specimen surface was evaluated through OES. The SEM micrographs obtained at 2000 magnification are
EDS. Analysis was conducted at 100 magnification and at an
presented in Fig. 1a. The morphology of the iron powder was more
accelerating voltage of 20 kV. irregular in comparison to the 316L stainless steel powders spher-
ical morphology, which could influence cold sprayability. The par-
ticle size analysis indicated a dimension of 33 ± 23 mm for the iron
2.5.2. X-ray photoelectron spectroscopy powder, with a dimension of 21 ± 14 mm for the 316L stainless steel
X-ray photoelectron spectroscopy of the accumulated corrosion powder. The corresponding frequency percentage with respect to
products was performed using a ThermoScientific K-Alpha XPS. X- diameter is presented in Fig. 1b for both powders. X-ray diffraction
ray photoelectron spectroscopy is a surface characterization tech- confirmed the chemical composition and crystalline structure of
nique, which allows for depth profiling through etching to under- the powders (Fig. 1c), with a-Fe identified for the pure iron pow-
stand the degree of oxidation or change in structures, which is der, and a-Fe and c-Fe for the 316L stainless steel. Quantitative
not possible with XRD. First, survey spectra were recorded from 0 phase analysis of the 316L powder indicated a relative phase per-
to 1400 eV, implementing an aluminum K-alpha source with a centage of 19.1% and 80.9% of ferrite and austenite, respectively.
beam width of 400 mm. High-resolution scans of O1s, Fe2p, C1s, It is presumed that the presence of ferrite in the 316L powder is
Cl2p, P2p, and Ca2p, were then performed. For all samples, the sur- a result of strain-induced transformation during powder process-
face was etched for 225 s to analyze the corrosion products in ing. The element concentrations determined through ICP-OES are
depth with an argon ion gun at an emission angle of 30. The ref- presented in Tables 2 and 3 for iron and 316L stainless steel pow-
erence binding energies used for identification are presented in ders, respectively.
Table 1. The deposition efficiency of the two cold-sprayed plates was
equal to 28.9 ± 0.2%. The reduced deposition efficiency was associ-
ated with particle bounce back, due to insufficient deformation. It
2.5.3. Inductively coupled plasma optical emission spectrometry was hypothesized that this was a result of the crystalline structure
The testing solution from the static corrosion tests was col- of pure iron, body-centred cubic (BCC), which has a lower particle
lected every 24 h for the duration of the 7-day test. To prepare packing density than FCC and HCP materials (0.68 vs 0.74). In addi-
the solution for testing, each 100-mL sample was split into two tion, the irregular morphology of the powder particles could
50-mL tubes and weighed. Each sample was centrifuged for one impact deformation kinetics as irregular powders have been
hour at 1000 RPM and 25 C. The supernatant was removed, and reported to yield less dense coatings with greater porosity [30].
any remaining solution was evaporated at 65 C prior to re- The resulting cold-sprayed coatings were 2.5 mm thick. The dog-
weighing. Nitric acid (5 mL) was added to each sample, followed bone subsize specimens were then machine cut and ground, with
by a digestion period of 4 h at 95 C. After digestion, all samples an average sample thickness of 1.3 ± 0.2 mm.
were completed to volume with deionized water and re-weighed.
A total of 14 samples were tested, two samples per day, for each
3.2. Coating characterization
annealing condition. Testing standards for transitional metals (Fe,
Ni, and Cr) were made at concentrations of 100, 10, and 1 mg L-1.
3.2.1. X-ray diffraction
The average element percentage, per day, was determined (n = 2).
X-ray diffraction was used to identify the annealed and as-
sprayed samples. As-sprayed Fe-316L was a mixture of a-Fe and
Table 1 c-Fe. After annealing, the peaks of c-Fe were significantly reduced
Reference binding energies in eV for Fe2p, O1s, C1s, P2p, Cl2p, and Ca2p. in comparison to the as-sprayed Fe-316L and could not be confi-
dently differentiated from the noise of the signal (Fig. 2a). Quanti-
Element Compound Binding Energy Reference
tative phase analysis indicated that the amount of austenite had
Fe2p Fe3O4 710.6 [23]
decreased significantly to trace amounts in comparison to the as-
724
O1s Fe-O 530.2 [24] sprayed sample (Table 4). The decrease in austenite is a result of
Fe-OH/Fe-OOH/Fe-PO4 531.5–531.6 [24] a transformation of austenite to ferrite, presumably due to nickel
C1s Hydrocarbons 284.6/284.8 [24] diffusion from 316L stainless steel to iron during annealing. Iron
C-O bonds 286.6 [25] transforms to austenite above 900 C, which may have facilitated
288.3 [25]
P2p Ca3(PO4)2 133.6 [26,27]
nickel diffusion. Any austenite detected could be a result of incom-
Cl2p FeCl2 201.0 [28] plete nickel diffusion (i.e. not at equilibrium), where local concen-
199.3 [28] trations give small amounts of austenite.
Ca2p Ca3(PO4)2 347.8–347.9 [26,27,29] The FWHM was determined for the annealed Fe-316L samples
350.6
and compared to the as-sprayed material, as shown in Fig. 2b.

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx 483
5

Fig. 1. Characterization of iron (top) and 316L stainless steel (bottom) powders with (a) SEM; (b) laser diffraction particle size analysis; and (c) XRD.

Table 2
Element concentrations of the iron powder determined through ICP-OES. Elements denoted with an (*) indicate that a concentrated solution was used for ICP testing (dilute
solutions were used for all non-marked elements).

Element Fe P* Ti* S* Pb* Total


Weight Percentage (%) 99.3 0.00966 0.693 0.00670 0.00391 100

Table 3
Element concentrations of the 316L powder determined through ICP-OES. Elements denoted with an (*) indicate that a concentrated solution was used for ICP testing (dilute
solutions were used for all non-marked elements).

Element Fe Cr Ni Mo Mn Si* P* S* Total


Weight Percentage (%) 67.4 16.7 12.5 2.44 0.550 0.370 0.0277 0.00641 100

The FWHM is typically sensitive to variation in microstructure, 3.2.2. Optical and scanning electron microscopy
such as grain refinement, as well as residual stress and plastic The optical micrographs of annealed and as-sprayed Fe-316L
deformation within the material [31,32]. Peak broadening, and are presented in Fig. 3a. In the optical image of the as-sprayed
therefore an increase in FWHM, can be attributed to an accumula- specimen, the darker matrix is composed of iron and the lighter
tion of stacking faults and structural disorders, as well as increased dispersed particles are 316L stainless steel. The most significant
stress. Overall, there was a decrease in the FWHM of the (1 1 0) observation of the annealed samples, in comparison to the as-
peak between the annealed and as-sprayed samples. This reduc- sprayed, is the absence of the lighter phase, i.e. the stainless-
tion is attributed to the release of internal stress within the steel particles. This confirms the XRD analysis that indicated a
microstructure and recrystallization during annealing. An increase transformation of austenitic stainless steel to ferritic low-alloyed
in annealing temperature resulted in a decrease in FWHM, though steel. In addition, a decrease in porosity was observed between
the FWHM did not change significantly between 1200 C and all annealed and as-sprayed samples. A scanning electron micro-
1300 C. Moravej et al. [33] observed a similar relationship scope was utilized to further assess the transformation of 316L
between annealing temperature and FWHM for electroformed iron, stainless steel by EDS (Fig. 3b). Distinct particles of 316L stainless
where an increase in the former resulted in a decrease of the latter. steel can be observed within the iron matrix of the as-sprayed

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
484
6 J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx

sample, indicated by the bright areas of nickel and chromium. In


contrast, atomic diffusion is evident in all annealed samples and
confirms the transformation of the stainless steel observed in the
optical micrographs. The degree of diffusion across particle bound-
aries was not uniform for all annealed samples, however, and clus-
ters of chromium and nickel could still be observed when
annealing at 1100 C.
Precipitates were observed in the optical micrographs of the
annealed specimens. Formation of precipitates and intermetallic
phases are a result of sensitization, defined as long term exposure
at high temperatures; for 316L stainless steel, the sensitization
temperature range is between 600 and 950 C [34]. In 316L stain-
less steel, an intermetallic phase precipitates at grain boundaries,
where a chromium rich region develops, resulting in chromium
depletion surrounding the precipitate [35]. EDS images confirmed
the presence of a chromium rich precipitate within the annealed
samples, particularly for samples annealed at 1300 C.

3.3. Mechanical testing

Uniaxial tensile tests were conducted to assess the impact of


annealing temperature on the mechanical behaviour of cold-
sprayed Fe-316L. Engineering stress–strain curves were generated
for all annealed Fe-316L samples. A representative stress–strain
curve of a sample from each annealing condition is presented in
Fig. 5. As the annealing temperature increased, the allowable
deformation prior to fracture increased, and visible necking was
detected for samples annealed at higher temperatures, particularly
at 1300 C. The mechanical behaviour of the as-sprayed coupons
was brittle, with limited deformation occurring prior to failure.
One of the three as-sprayed samples tested failed during pre-load
and no data was recorded for this sample.
The fracture surfaces of the differently annealed samples were
analyzed under SEM at 2000 magnifications (Fig. 4b–e). Both
ductile and brittle fracture (cleavage) were observed on the surface
of the heat-treated samples. Substantial microvoid coalescence
(MVC) were observed for all annealed samples, evident by the net-
work of spherical ‘‘dimples” in the microstructure. These micro-
voids are the mechanism of ductile fracture, which initiates crack
formation under loading. In contrast, the surface of the as-
sprayed sample indicated a complete brittle failure mode, with
no MVC observed.
Engineering UTS, elongation, and yield stress were determined
for each annealing condition; the mean values are presented in
Fig. 5 with standard deviation and relevant p-values. The engineer-
ing UTS was highest for the samples annealed at 1300 C (Fig. 5a).
Multiple comparisons testing showed that the tensile strengths of
the differently annealed specimens were all significantly different
from the as-sprayed condition, p = 0.0001 for samples annealed
at 1100 C and p < 0.0001 for samples annealed at 1200 C and
1300 C. Specimens annealed at 1300 C had the greatest elonga-
Fig. 2. (a) XRD patterns of the annealed and as-sprayed samples; (b) FWHM of the tion at fracture, with significant differences found between the dif-
(1 1 0) peak of the as-sprayed and annealed samples. ferently annealed specimens, as shown in Fig. 5b. In addition, the
elongation fractures determined for all annealed samples were sig-
nificantly higher than that of the as-sprayed, with p = 0.0011 for
samples annealed at 1100 C and p < 0.0001 for samples annealed
Table 4
Quantitative phase composition of ferrite and austenite in the annealed and as-
at 1200 C and 1300 C. One-way ANOVA tests found no significant
sprayed samples. difference between the yield strengths of the annealed samples
with p = 0.1314.
Sample Ferrite (%) Austenite (%)
Due to the deposition process of cold-spray, which yields work-
1100 C 99.8 0.2 hardened coatings and poor particle–particle bonds, limited ductil-
1200 C 99.9 0.1
ity was determined for the as-sprayed condition (0.047%). Low
1300 C 99.8 0.2
As-Sprayed 91.1 8.9 strength was observed for the as-sprayed tensile samples, due to
limited ductility (i.e. premature fracture), with a UTS of 84 MPa.

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx 485
7

Fig. 3. (a) Optical micrographs of representative polished annealed and as-sprayed specimens at 200 magnification; (b) Concentration of select elements by EDS of as-
sprayed and annealed samples at 1100 C, 1200 C, 1300 C.

The combination of low strength and low ductility was associated despite increased material softening as ductility (i.e. fracture resis-
with incomplete bonding at particle–particle interfaces, which was tance) is increased.
confirmed by the as-sprayed optical and SEM images. Incomplete
particle bonding results in fracture at low forces, with limited 3.4. Corrosion testing
deformation. During the annealing process, recrystallization
occurs, which removes dislocations existing within the crystalline 3.4.1. Static corrosion testing
structure, softening the material and improving ductility. More The average static corrosion rate, CRaverage, was calculated in mg
importantly for a cold-sprayed material, annealing also improves cm2 day1. The relationship between corrosion rate and time is
interparticle bonding through ‘sintering’, which is accompanied presented in Fig. 6a with standard deviation. Elevated degradation
by atom diffusion across particle boundaries (Fig. 3) [5,36]. rates were observed during the first 24–48 h of immersion, due to
Improved interparticle bonding results in greater tensile strength the exposure of the fresh sample surface to the testing solution. A

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
486
8 J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx

Fig. 4. (a) Engineering stress–strain curves of representative annealed and as-sprayed specimens. The fracture surface at 2000 magnification using SEM for representative
samples annealed at (b) 1100 C; (c) 1200 C; (d) 1300 C; as well as (e) as-sprayed condition.

Fig. 5. The engineering UTS, elongation at fracture, and yield stress, with standard deviation, of the differently annealed samples (n = 3), as compared to the as-sprayed
condition (n = 2). A statistically significant increase in both UTS and ductility was observed after heat treatment at all temperatures. NS signifies a non-significant difference.

one-way ANOVA test found that the average corrosion rates of all cold-sprayed ferrous alloy coatings degraded in 3.5% NaCl solution,
samples, after 7 days of immersion time, were significantly differ- where corrosion products were first observed around interparticle
ent, p < 0.0001. Tukey’s multiple comparisons test indicated that boundaries and micro-pores. In contrast, the sintering that
the corrosion rates of the annealed samples were all significantly occurred at the particles boundaries of the annealed samples
less from that of the as-sprayed sample. Similarly, the corrosion would slow the rate of the corrosion reaction, due to decreased
rate of samples annealed at 1200 C and 1300 C were found to corrosion initiation sites. This difference between the annealed
be significantly different from that of specimens heat treated at and as-sprayed samples was clearly observed when analyzing the
1100 C (p = 0.0393 and p = 0.0399, respectively). There was no sig- corroded surface under SEM (Fig. 6b-e). From a biodegradable stent
nificant difference between samples annealed at 1200 C and application perspective, the severe corrosion at the boundaries of
1300 C, p > 0.9999. the as-sprayed specimen is problematic as it could lead to a severe
The highest corrosion rate observed was 0.22 mg cm2 day1 reduction of the radial strength and stiffness of the stent. This may
(0.10 mm year1), determined for samples annealed at 1200 C result in premature failure of the device before complete vessel
and 1300 C. This corrosion rate is approximately 43% less than remodelling can occur [38]. Decreased corrosion around particle
that of the as-sprayed sample. This decrease in corrosion rate is boundaries of the annealed samples is advantageous in this regard.
attributed to several factors. Increased porosity and poor interpar-
ticle bonding of the as-sprayed samples allowed for preferential 3.4.2. Galvanic corrosion testing
corrosion to occur at the particle boundaries and surface pores. The galvanic potential of the as-sprayed sample was measured
Choi et al. [37] observed similar behaviour when analysing at approximately 700 mV, consistent to that of pure iron. In con-

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx 487
9

Fig. 6. (a) Average static corrosion rate versus time for the annealed and as-sprayed samples (n = 3); SEM images at 500 magnification of the corroded sample surfaces after
cleaning annealed at (b) 1100 C; (c) 1200 C; (d) 1300 C; and (e) as-sprayed specimens.

trast, the potential range of the annealed samples was between calcium, and sulfur all found in varying concentrations (refer to
550 and 600 mV, which is similar to that of mild steel on the Supplementary Figs. 1–4). Decreased iron density was observed
galvanic potential series chart [39]. This discrepancy between the in highly concentrated areas of chlorine, indicative of an active cor-
potentials of the heat treated and as-sprayed samples confirms a rosion reaction occurring on the sample surface (pitting corrosion).
material transformation due to Ni and Cr diffusion during anneal- Overlapping areas of concentrated phosphorous and calcium were
ing. Despite this, there was no discernable influence of heat treat- observed in the EDS maps, which coincided with the agglomerate
ment on the determined galvanic corrosion rate (Table 5). structure on the corrosion sample surface. The concentration of
Statistical significance testing indicated that there was no signifi- this layer was higher for the annealed samples than the as-
cant difference between the galvanic corrosion rates of the tested sprayed sample but was not homogeneous across the entire sur-
samples (One-way ANOVA, p = 0.9856). face. This could be attributed to the propensity of the as-sprayed
sample to actively corrode in the presence of chlorine, due to its
poor interparticle bonding and high porosity, which act as initia-
3.5. Corrosion product characterization tion sites for chlorine attack.
The corrosion products on the sample surface were analyzed
3.5.1. Scanning electron microscopy and X-ray diffraction using XRD. The resulting XRD patterns were plotted for represen-
The corrosion products were analyzed with multiple character- tative samples (Fig. 8). The XRD patterns had low intensity counts,
ization techniques, including XRD, SEM, EDS, XPS, and ICP-OES. The which was attributed to the decreased thickness of the corrosion
accumulated corrosion products on the sample surface were pre- product layer that was prone to flaking and poorly adherent to
served and dried. All samples were covered in an orange-brown the specimen surface. The predominant compound identified was
layer, which was prone to flaking and had poor surface adherence. iron phosphate, which was observed for the heat treated and as-
This is typically associated with the formation of ferrous sprayed specimens. In addition, iron oxide (Fe2O3) was detected
hydroxides/oxide-hydroxides [24]. The morphology of the layer in samples annealed at 1200 C and 1300 C. The presence of
was analyzed with SEM at 3000 magnification (Fig. 7). The Fe2O3 in the 1300 C specimen was identified after depositing some
mineral-like structure observed is similar to that observed by Her- of the corrosion products onto carbon tape and reanalysing with
mawan et al. [40], when conducting corrosion tests of Fe-35Mn XRD. The identification of iron phosphates and iron oxides are con-
specimens in modified HBSS. These agglomerates were found to sistent to that reported by Tolouei et al. [24], which investigated
contain higher concentrations of calcium and phosphorus, which corrosion products of pure iron in different pseudo-physiological
was similarly observed when analysing the Fe-316L samples by solutions.
EDS. The EDS images illustrated the heterogeneous phenomena
occurring on the sample surface during corrosion. All samples were
densely covered in oxygen, with chlorine, sodium, phosphorus, 3.5.2. X-ray photoelectron spectroscopy
A survey spectral scan was completed for each annealing condi-
tion, which positively identified elements Fe, O, C, Cr, Na, Ca, Cl, P,
and K. As expected, many of these elements are associated with the
Table 5 components of the testing solution, modified HBSS. The identifica-
Galvanic corrosion rates, in mg cm2 day1, determined for the annealed and as- tion of these elements was consistent with EDS mapping analysis
sprayed specimens.
of the corrosion product layer. High-resolution scans of Fe2p,
Sample Galvanic Corrosion Rate (mg cm2 day1) O1s, C1s, Cl2p, Ca2p, and P2p were conducted of the surface and
1100 C 2.14 ± 0.22 after 225 s etching (Supplementary Figs. 5–8). Some minor shifting
1200 C 2.17 ± 0.31 was observed in the peak binding energies after etching, which
1300 C 2.16 ± 0.36 was anticipated as oxygen disappears in depth profile and the sur-
As-Sprayed 2.09 ± 0.16
rounding atoms differ. Overall, good agreement was found

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
488
10 J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx

Fig. 7. SEM micrograph at 3000 of the corrosion products of samples annealed at (a) 1100 C; (b) 1200 C; (c) 1300 C; as well as the (d) as-sprayed sample.

Fig. 8. X-ray diffraction pattern of corroded samples annealed at (a) 1100 C; (b) 1200 C; (c) 1300 C; as well as the (d) as-sprayed condition. The primary corrosion product
identified was iron phosphate, with iron oxides detected in two of the four samples analysed (specimens annealed at 1200 C and 1300 C).

between the selected reference binding energies and peak binding trast, the intensity of all other analyzed elements increased after
energies determined for the samples, with some limited variation etching, emphasizing the importance of depth analysis when iden-
occurring between different specimens. tifying corrosion products. Analysis of the Fe2p spectra identified
Since the corrosion product layer was dried in air, surface the presence of Fe3O4 at the Fe2p3/2 and Fe2p1/2 peaks, though
hydrocarbons and C-O compounds (carbon dioxide and glucose) the Fe2p3/2 peak is also associated with FeCl2 at a binding energy
were observed for all samples on the surface, which were signifi- of 710.6 eV according to Grosvenor et al. [28]. The O1s spectra anal-
cantly reduced after etching for 225 s, as shown in Table 6. In con- ysis indicated the presence of both Fe-O and Fe-OH/Fe-OOH on the

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx 489
11

Table 6
The percent change in counts s1 on the sample surface and after etching for 225 s, as determined by high-resolution XPS, for the annealed and as-sprayed specimens. The raw
intensity data, in counts s1, is provided in parentheses, where the counts on the surface is listed first, followed by the counts after etching.

Element Reference Binding Energy (eV) 1100 C 1200 C 1300 C As-Sprayed


Fe2p 710.6 37.2% (17890/24550) 86.4% (20123/37512) 119.8% (16086/35362) 92.3% (7733/14871)
724 27.3% (15823/20135) 36.2% (29658/21769) 90.0% (15634/29701) 152.7% (3225/8149)
O1s 530.2 9.7% (21216/23278) 6.6% (47920/44759) 56.5% (22759/35620) 13.8% (24631/28019)
531.6 14.2% (23605/26957) 28.0% (50158/64209) 5.9% (38096/35841) 14.0% (26183/29841)
C1s 284.4/284.6 40.5% (5797/3447) 60.0% (11060/4428) 60.6% (13091/5160) 54.8% (11229/5074)
286.6 26.1% (4495/3320) 46.30% (7688/4129) 36.6% (9799/6217) 37.0% (7677/4834)
288.3 15.4% (3273/2770) – 19.8% (5865/4706) 15.7% (4413/3718)
Cl2p 199.3 22.2% (4922/6017) 5.27% (3927/4134) 61.8% (5627/9105) 42.6% (5480/7816)
201 20.2% (3712/4463) – 71.5% (4031/6912) 35.1% (3818/5159)
P2p 133.6–133.6 18.0% (3502/4133) 25.0% (7075/8843) 60.6% (2968/4766) 65.8% (2199/3646)
Ca2p 347.8 – 63.4% (8009/13086) 17.9% (5029/5930) 31.9% (3619/4774)
350.6 – 40.1% (6322/8857) 15.0% (4776/5494) 37.4% (4322/5937)

surface for all samples. The peak intensity was higher for the Fe- tenfold increase in concentration in comparison to the as-sprayed
OH/Fe-OOH species than Fe-O, where Fe-OH and Fe-OOH are inter- data was found. This is a further indication of the transformation of
mediate compounds formed in the corrosion reaction pathway, the Fe-316L amalgamate after annealing, due to atomic diffusion.
which lead to the formation of Fe3O4. High resolution scans of More specifically, it suggests that both the iron and nickel are
Cl2p revealed the presence of FeCl2 (rokuhnite), which is a precur- degrading concurrently in the annealed specimens, whereas the
sor compound to c-FeOOH and Fe3O4. This species forms when iron matrix is only degrading in the as-sprayed sample. However,
conducting corrosion tests in HBSS, which has a high concentration the change in chromium concentration with time did not follow
of Cl- ions [41]. the same trend as nickel and iron for the annealed samples, and
Analysis of the Ca2p spectra identified the composition of the the reported concentration was often lower than the value deter-
agglomerates observed under SEM and EDS as Ca3(PO4)2. The pres- mined in HBSS. While both nickel and chromium are substitutional
ence of Ca3(PO4)2 is a result of the reactivity of the material when atoms during atomic diffusion, the atomic radius of chromium is
exposed to simulated body fluid, such as HBSS, and not necessarily larger than that of nickel (166 vs. 149 pm), which could slow the
due to the corrosion of Fe-316L. Other studies conducting immer- diffusion rate of chromium across particle boundaries during
sion tests with Fe-based materials, including Fe-35Mn and 316L annealing [46]. Furthermore, EDS mapping of the annealed sam-
stainless steel, have identified calcium phosphates on the surface ples revealed chromium-rich precipitates throughout the matrix,
after testing in HBSS [40,42]. Variation in binding energies for Ca3(- rather than a uniform distribution like that observed with nickel.
PO4)2 was found in the literature, particularly when identifying This may explain why chromium was not observed in the corrosion
Ca2p3/2 peak [26,29]. The peak binding energies identified in the products within the testing solution.
tested samples matched well with studies conducted by França
et al. [26], with a binding energy of 347.8 eV for the Ca2p3/2 peak.
However, the binding energy the Ca2p1/2 peak was not reported. 4. Discussion
An investigation by Chusuei et al. [29] reported binding energies
for both Ca2p3/2 and Ca2p1/2, but shifts in the peak binding ener- Microstructural, mechanical, and corrosion investigations were
gies of Ca2p3/2 and Ca2p1/2 were observed of the tested samples conducted to assess the impact of annealing on the overall material
in comparison. It is possible that the Ca2p1/2 peak may be associ- behaviour of Fe-316L, with heat treatment at 1100 C, 1200 C, and
ated with CaCO3, which has a more similar binding energy 1300 C. Annealing resulted in substantial atomic diffusion
(351.5 eV) to that observed in the samples tested [43]; CaCO3 can between the Fe and 316L particles, and ultimately the transforma-
precipitate due to the bicarbonate ions present in modified HBSS. tion of austenitic stainless steel to low-alloyed ferrite, which was
However, it has been shown by Kirkland et al. [44] that the use confirmed by EDS and XRD. Precipitates with a high element con-
of HEPES buffer in HBSS may decrease or inhibit the formation of centration of chromium were observed with increasing concentra-
CaCO3 in vitro. Due to the low intensity of Ca2p spectra, it was tion as the annealing temperature increased. Interestingly, the
not possible to conclude whether both Ca3(PO4)2 and CaCO3 were presence of precipitates within the microstructure, which tradi-
present. Analysis of the P2p spectra further suggested the presence tionally have embrittling characteristics, did not have an impact
of Ca3(PO4)2, with tested samples matching well to reference bind- on the mechanical behaviour of annealed Fe-316L [47,48]. In com-
ing energies of calcium phosphate in literature, with minor varia- parison to the as-sprayed condition, the mechanical properties of
tion [26,27]. This P2p peak may also be associated with iron Fe-316L significantly improved after heat treatment. This was
phosphates, which were identified in the XRD analysis [27,45]. attributed to two primary factors: an increase in particle–particle
bonding due to sintering at the boundaries and recrystallization,
which removed residual stress, and work hardening induced by
3.5.3. Inductively coupled plasma optical emission spectrometry powder deposition. A statistically significant increase was
The change in concentration of Fe, Ni, and Cr in the corrosion observed in the ductility between the annealed samples. It is inter-
solution as measured by ICP-OES are presented in Fig. 9. Overall, esting to note, however, that the austenite levels were very similar
the iron concentration decreased with time for all samples, which between the annealed specimens and significantly lower than the
reflects a similar trend when plotting the corrosion rate with time. as-sprayed sample. Since austenite is stabilized by Ni, then exten-
This was attributed to the formation of a stable degradation layer sive Ni diffusion had occurred at all annealing conditions, which
on the sample surface, which would have slowed ion diffusion at should lead to strong interparticle bonds [49]. The ductility is
the surface, as well as to the development of the calcium- low when annealed at 1100 C, however, and high at 1300 C. This
phosphate passive film. Interestingly, the concentration of Ni for suggests that large scale diffusion and particle sintering are not the
all annealed samples followed the same trend as that of iron; such only influence on elongation and perhaps atomic diffusion is not
a pattern was not observed for the as-sprayed data. Furthermore, a the sole indicator of effective sintering for cold-sprayed coatings.

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
490
12 J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx

Fig. 9. Change in percent of iron, nickel, and chromium within the testing solution for the static corrosion tests for samples annealed at (a) 1100 C; (b) 1200 C; (c) 1300 C;
and (d) as-sprayed samples. The dotted line represents the concentration of iron, nickel, and chromium determined within the Hanks’ Balanced Salt Solution and HEPES buffer
only, which was equal to 0.0042%, 0.00002%, and 0.0006%, respectively.

This suggests that annealing of cold sprayed mixed metal powders to the corrosion rate of the as-sprayed material, equal to
could provide an interesting rapid prototyping technique for bio- 0.39 mg cm2 day1 (0.18 mm year-1). Interestingly, the corrosion
material fabrication. This could be used to generate a new series rate increased with higher annealing temperatures, where the
of low alloy steels with incremental changes of Ni and Cr to control opposite would necessarily be expected due to improved particle
material properties. From the perspective of manufacturing bonding [50]. The increase in corrosion rate observed between
biodegradable stents, this would allow for the modification and samples annealed at 1100 C and 1200/1300 C may be attributed
control of the corrosion rate of the material, potentially creating to microgalvanic corrosion, due to increased precipitate concentra-
a more patient-specific treatment approach to reflect disease tion in samples annealed at 1200/1300 C. The intermetallic phase
severity. observed in the microstructure can act as a cathode in an anodic
A peak corrosion rate of 0.22 mg cm2 day1 (0.10 mm year-1) Fe-based matrix. It has been documented within literature that
was determined for the samples annealed at 1300 C, in contrast intermetallic particles, such as those observed in this work, as well

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx 491
13

as inclusions and secondary phases, act as micro-cathodes to initi- developing biodegradable stents (Table 7). It has good tensile
ate galvanic corrosion in an anodic matrix [48,51–53]. The imple- strength for biodegradable stent application, and the UTS deter-
mentation of intermetallic phases has been used to accelerate the mined is higher than magnesium-based biodegradable materials.
behaviour of iron-based biomaterials [51,52]. Schinhammer et al. While the ductility is less than that of pure iron or 316L stainless
[51] introduced manganese and palladium to an iron matrix, which steel, the elongation at fracture determined for Fe-316L is suffi-
produced Pd-based intermetallic phases that acted as cathodic cient for coronary stents and is considerably improved from its
sites resulting in microgalvanic corrosion, increasing the overall as-sprayed state. Interestingly, the yield stress of Fe-316L was
corrosion rate of the Fe-Mn-Pd alloy. In low-alloy steels, MnS inclu- the lowest of all compared materials, which is an important prop-
sions have been shown to act as cathodes to preferentially degrade erty in stent design as decreased yield stress will lower the balloon
the nearby steel matrix [54–56]. In addition, microgalvanic corro- pressure required for stent expansion [59].
sion has been observed between Cr-rich secondary phases and a The corrosion rate of Fe-316L annealed at 1300 C is less than
steel matrix, which promotes the Fe degradation within the steel other developing biodegradable stents [62]. The reduced corrosion
[57]. M secondary phase has also been shown to act as a cathode rate of annealed Fe-316L can be viewed as a compromise to
in low carbon bainitic steel, resulting in accelerated corrosion in achieve the mechanical properties required for biodegradable stent
NaCl solution [58]. Similar behaviour has been observed with application. Recent clinical studies suggest that a reduced corro-
sigma phase in stainless steels, where preferential corrosion occurs sion rate may be advantageous, however, to maintain device integ-
of c-austenite [48]. Such secondary phases are typically a result of rity and scaffolding support in vivo over a longer duration of time
sensitization due to heat treatment. Therefore, the precipitates [15]. In particular, a recent case report indicated that the rapid
formed during annealing may have acted as cathodic sites to accel- degradation of the BioTronik’s Magmaris stent after only
erate the overall corrosion behaviour of the annealed Fe-316L sam- 4 months led to inadequate stent scaffolding, resulting in severe
ples. This may explain why the static corrosion rate was highest for restenosis [15].
samples annealed at 1300 C as microstructural observation indi- The identification of the corrosion products in this work allows
cated that it had the highest concentration of such precipitates. for an important understanding of the chemical reactions occur-
EDS mapping of a sample annealed at 1300 C revealed an ring on the sample surface during degradation. By understanding
increased oxygen concentration around the precipitates (Fig. 10), how and why the corrosion products form, the change in corrosion
suggestive of a corrosion reaction. Point identification by EDS behaviour over time can be explained. Furthermore, the accumula-
revealed that the precipitates had a higher concentration of Cr than tion of corrosion products can influence the local pH environment
the surrounding matrix, the majority of which had Cr content of of the surrounding tissues in vivo [64]. To identify the corrosion
27–32%. Microvoids were also identified within the Fe-316L products formed on the sample surface, various characterization
matrix. techniques were used, including SEM, EDS, XRD, XPS, and ICP-
From a more global perspective, the material properties of Fe- OES. A mixture of iron oxides/oxyhydroxides and phosphates were
316L annealed at 1300 C are consistent with or better than other found, as well as FeCl2, which form a multi-layer structure, typi-
cally composed of three layers [65,66]:

(1) Iron (II) oxide (FeOnH2O) or iron hydroxide (Fe(OH)2)


(2) Magnetite (Fe3O4)
(3) Iron (III) oxide (Fe2O3nH2O)
This is consistent with the corrosion products of other iron-
based biomaterials reported in literature [24,33,40]. It has been
documented within literature that the accumulation of these cor-
rosion products in vivo has not led to local or systemic toxicity
[67–70].
While an in-depth approach was used to analyze the corrosion
products of Fe-316L, identification of the species was challenging,
which was associated with the decreased thickness of the corro-
sion product layer. Furthermore, despite the increased resolution
of XPS in comparison to XRD, it was not possible to confidently
identify multiple species within the Fe2p high resolution scan,
which was attributed to the broad spectra and relatively undefined
satellite structures. Furthermore, the overlapping binding energies
of iron species, as well as substantial variation in binding energies
reported within literature, make interpretation difficult [71]. For
Fig. 10. Element concentration by EDS mapping of a sample annealed at 1300 C at
instance, the Fe2p3/2 peak, identified as Fe3O4, may also be associ-
3000 magnification. Chromium-rich precipitates, as well as microvoids, were
observed within the surrounding Fe-316L matrix. Increased concentration of ated with FeCl2 at a similar binding energy, which was identified
oxygen was found around the precipitates. within the chlorine spectra. There was an indication of multiple

Table 7
Comparison of the mechanical properties of Fe-316L, annealed at 1300 C for two hours, with other biodegradable stent material properties reported within literature.

Material UTSeng (MPa) Yield Stress (MPa) Elongation (%) Corrosion Rate (mm year1) Reference
Fe-316L (Annealed: 1300 C, 2 h) 280 140 23 0.10 (0.22 mg cm2 day1) –
316L Stainless Steel 490 190 40 – [60]
Pure Iron 210 150 40 0.2 [61]
Mg-Zn 280 250 22 0.92 [62]
WE43 magnesium alloy 250 150 4 1.35 [63]

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
492
14 J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx

species within the Fe2p spectra, which was established by the O1s It is critical that newly developed biomaterials intended for
high resolution analysis, but the reduced intensity of the Fe2p biodegradable stent application address the limitations of polymer
spectra made this interpretation less reliable. The limitations of stents, which have largely underperformed in a clinical setting. It
the analysis outlined here could be addressed through additional was demonstrated in this work that cold-gas dynamic spraying is
analysis of a thicker corrosion product layer, assessing multiple a viable alternative for biomaterial fabrication and that a cold-
locations, for a more precise identification. sprayed iron and 316L stainless steel material has potential to be
Due to the bioactivity of the sample material, the development implemented in a clinical setting. It was found that annealing at
and stabilization of a passive Ca3(PO4)2 film was facilitated when 1300 C for 2 h yielded the mechanical and corrosion properties
exposed to modified HBSS. The passive layer may be disrupted, necessary for biodegradable stent application, with an ultimate
however, by Cl adsorption. This is accomplished by decreasing tensile strength of 280 MPa, and a ductility of 23%. The corrosion
its resistance to dissolution, by preventing the amalgamation and rate was equal to 0.22 mg cm2 day1, and a multi-layer structure
crystallization of the film [24]. Furthermore, the use of HEPES buf- of ferrous oxides, hydroxides and oxide hydroxides was observed
fer is known to weaken the passive film by influencing the nucle- after degradation. Ongoing investigations include the effect of
ation processes of Ca3(PO4)2, ultimately producing a less dense degradation on the mechanical properties of Fe-316L, as well as
corrosion product layer [44]. This may explain why the protective in vivo studies to assess device biocompatibility and
layer observed in this work was not homogeneous over the entire biodegradation.
surface, ultimately permitting the diffusion of Cl ions.
The results of this investigation are presented in the context of Acknowledgements
their limitations. General limitations of the work include the local
nature of microscopy images presented, as well as potential varia- The authors wish to acknowledge funding sources from the Natural
tion when measuring sample dimensions. More specifically, the Sciences and Engineering Research Council of Canada (Discovery
identification of the precipitates observed within the annealed Grant NSERC RGPIN 217183-13), as well as the McGill University
specimens was challenging, and the techniques presented in this Engineering Doctoral Award.
work were unable to confidently identify its composition. Future
work involves additional microstructural investigations to confi-
Declaration of Competing Interest
dently identify the intermetallic phase observed. Next, the tech-
nique utilized to assess the galvanic corrosion reaction was
The authors declare that they have no known competing finan-
limited to a macro-characterization. It cannot directly quantify
cial interests or personal relationships that could have appeared to
the microgalvanic reaction within the Fe-316L coating, only that
influence the work reported in this paper.
of the annealed samples coupled with pure 316L. To quantify
microgalvanic corrosion, future work will develop a protocol to
implement scanning Kelvin probe force microscopy (SKPFM) to Appendix A. Supplementary data
measure potential differences across particle boundaries. This will
confirm the presence of micro-galvanic couples, as well as any Supplementary data to this article can be found online at
electropotential variations within the structure of annealed Fe- https://doi.org/10.1016/j.actbio.2019.08.034.
316L material.
In this work, static corrosion tests were utilized as it is a conve- References
nient method to assess the global corrosion behaviour of
[1] B. Al-Mangour, The Use of Cold Sprayed Alloys for Metallic Stents (Master’s
biodegradable materials. However, this corrosion environment Thesis), Department of Mechanical Engineering, McGill University, Montreal,
has its limitations. Typically, it underestimates the corrosion Canada, 2012.
occurring on the surface as the convective contribution due to fluid [2] A. Papyrin, V. Kosarev, S. Klinkov, A. Alkimov, V. Fomin, Chapter 2 - High-
velocity interaction of particles with the substrate. Experiment and modeling,
flow is not considered. Conversely, proteins were not included Cold Spray Technology, Elsevier, Oxford, 2007, pp. 33–118.
within the testing solution, nor was the presence of cells on the [3] B.P. Murphy, P. Savage, P.E. McHugh, D.F. Quinn, The stress-strain behavior of
material surface considered, both of which will affect material coronary stent struts is size dependent, Ann. Biomed. Eng. 31 (6) (2003) 686–
691.
degradation [21]. In addition, the atmospheric conditions were
[4] A. Kastrati, J. Mehilli, J. Dirschinger, F. Dotzer, H. Schühlen, F.J. Neumann, M.
not controlled, i.e. implementing a testing environment with 5% Fleckenstein, C. Pfafferott, M. Seyfarth, A. Schömig, Intracoronary stenting and
CO2. Future investigations should incorporate these factors to pro- angiographic results: strut thickness effect on restenosis outcome (ISAR-
vide an improved prediction of the in vivo corrosion environment. STEREO) trial, Circulation 103 (23) (2001) 2816–2821.
[5] B. AL-Mangour, P. Vo, R. Mongrain, E. Irissou, S. Yue, Effect of heat treatment on
In addition, the use of polarization corrosion tests would comple- the microstructure and mechanical properties of stainless steel 316L coatings
ment the corrosion investigations completed in this work. produced by cold spray for biomedical applications. J. Thermal Spray Technol.
The formation of Ca3(PO4)2 on the sample surface due to the 23(4) (2014) 641–652
[6] P.W. Serruys, H.M. Garcia-Garcia, Y. Onuma, From metallic cages to transient
reactivity of Fe-316L in HBSS slowed the overall corrosion beha- bioresorbable scaffolds: change in paradigm of coronary revascularization in
viour. To truly quantify its impact on Fe-316L degradation, as well the upcoming decade?, Eur Heart J. 33 (1) (2012) 16–25b.
as the other environmental conditions mentioned previously, an [7] N. Resnick, H. Yahav, A. Shay-Salit, M. Shushy, S. Schubert, L.C. Zilberman, E.
Wofovitz, Fluid shear stress and the vascular endothelium: for better and for
in vivo study should be conducted, which would also allow for an worse, Prog. Biophys. Mol. Biol. 81 (3) (2003) 177–199.
assessment of Fe-316L biocompatibility. The use of cellular tests [8] S. Brugaletta, B.D. Gogas, H.M. Garcia-Garcia, V. Farooq, C. Girasis, J.H. Heo, R.J.
would also be beneficial to evaluate the biocompatibility of Fe- van Geuns, B. de Bruyne, D. Dudek, J. Koolen, P. Smits, S. Veldhof, R. Rapoza, Y.
Onuma, J. Ormiston, P.W. Serruys, Vascular compliance changes of the
316L, which should be conducted in future investigations.
coronary vessel wall after bioresorbable vascular scaffold implantation in the
treated and adjacent segments, Circ. J. 76 (7) (2012) 1616–1623.
[9] Y. Onuma, D. Dudek, L. Thuesen, M. Webster, K. Nieman, H.M. Garcia-Garcia, J.
A. Ormiston, P.W. Serruys, Five-year clinical and functional multislice
5. Conclusions
computed tomography angiographic results after coronary implantation of
the fully resorbable polymeric everolimus-eluting scaffold in patients with de
There are numerous challenges associated with the develop- novo coronary artery disease: the ABSORB cohort A trial, J. Am. Coll. Cardiol.
ment and implementation of biodegradable stents in a clinical set- Cardiovasc. Interventions 6 (10) (2013) 999–1009.
[10] J. Foerst, M. Vorpahl, M. Engelhardt, T. Koehler, K. Tiroch, R. Wessely, Evolution
ting, particularly when comparing current biodegradable stent of coronary stents: From bare-metal stents to fully biodegradable, drug-
performance to the gold standard treatment of drug-eluting stents. eluting stents, Combination Prod. Ther. 3 (1) (2013) 9–24.

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx 493
15

[11] H.Y. Ang, Y.Y. Huang, S.T. Lim, P. Wong, M. Joner, N. Foin, Mechanical behavior [38] H.Y. Ang, J. Ng, H. Bulluck, P. Wong, S. Venkatraman, Y. Huang, N. Foin, 5 -
of polymer-based vs. metallic-based bioresorbable stents, J. Thoracic Dis. 9 Fundamentals of bioresorbable stents, in: H. Podbielska, M. Wawrzyńska
(Suppl 9) (2017) S923–S934. (Eds.), Functionalised Cardiovascular Stents, Woodhead Publishing,
[12] T. Toyota, T. Morimoto, H. Shiomi, Y. Yoshikawa, H. Yaku, Y. Yamashita, T. Cambridge, UK, 2018, pp. 75–97.
Kimura, Very late scaffold thrombosis of bioresorbable vascular scaffold: [39] H.P. Hack, Galvanic Corrosion, Reference Module in Materials Science and
systematic review and a meta-analysis, JACC Cardiovasc. Interv. 10 (1) (2017) Materials, Engineering 2 (2010) 828–856.
27–37. [40] H. Hermawan, A. Purnama, D. Dube, J. Couet, D. Mantovani, Fe–Mn alloys for
[13] Y. Onuma, Y. Sotomi, H. Shiomi, Y. Ozaki, A. Namiki, S. Yasuda, T. Ueno, K. metallic biodegradable stents: degradation and cell viability studies, Acta
Ando, J. Furuya, K. Igarashi, K. Kozuma, K. Tanabe, H. Kusano, R. Rapoza, J.J. Biomater. 6 (5) (2010) 1852–1860.
Popma, G.W. Stone, C. Simonton, P.W. Serruys, T. Kimura, Two-year clinical, [41] C.S. Obayi, R. Tolouei, A. Mostavan, C. Paternoster, S. Turgeon, B.A. Okorie, D.O.
angiographic, and serial optical coherence tomographic follow-up after Obikwelu, D. Mantovani, Effect of grain sizes on mechanical properties and
implantation of an everolimus-eluting bioresorbable scaffold and an biodegradation behavior of pure iron for cardiovascular stent application,
everolimus-eluting metallic stent: insights from the randomised ABSORB Biomatter 6 (1) (2016) 1–9.
Japan trial, EuroIntervention 12 (9) (2016) 1090–1101. [42] T. Hanawa, S. Hiromoto, A. Yamamoto, D. Kuroda, K. Asami, XPS
[14] Y. Sotomi, P. Suwannasom, P.W. Serruys, Y. Onuma, Possible mechanical characterization of the surface oxide film of 316L stainless steel samples
causes of scaffold thrombosis: insights from case reports with intracoronary that were located in quasi-biological environments, Mater. Trans. 43 (12)
imaging, EuroIntervention 12 (14) (2017) 1747–1756. (2002) 3088–3092.
[15] H. Yang, F. Zhang, J. Qian, J. Chen, J. Ge, Restenosis in Magmaris stents due to [43] M. Ni, B.D. Ratner, Differentiation of calcium carbonate polymorphs by surface
significant collapse, J. Am. Coll. Cardiol. Cardiovas. Interventions 11 (10) analysis techniques – an XPS and TOF-SIMS study, Surf. Interface Anal. 40 (10)
(2018) e77–e78. (2008) 1356–1361.
[16] A. Papyrin, V. Kosarev, S. Klinkov, A. Alkimov, V. Fomin, Chapter 1 - Discovery [44] N.T. Kirkland, J. Waterman, N. Birbilis, G. Dias, T.B.F. Woodfield, R.M.
of the cold spray phenomenon and its basic features, Cold Spray Technology, Hartshorn, M.P. Staiger, Buffer-regulated biocorrosion of pure magnesium, J.
Elsevier, Oxford, 2007, pp. 1–32. Mater. Sci. - Mater. Med. 23 (2) (2012) 283–291.
[17] J. Frattolin, R. Barua, H. Aydin, S. Rajagopalan, L. Gottellini, R. Leask, S. Yue, D. [45] M. Eglin, A. Rossi, N.D. Spencer, X-ray Photoelectron Spectroscopy Analysis of
Frost, O.F. Bertrand, R. Mongrain, Development of a novel biodegradable Tribostressed Samples in the Presence of ZnDTP: A Combinatorial Approach,
metallic stent based on microgalvanic effect, Ann. Biomed. Eng. 44 (2) (2016) Tribol. Lett. 15 (3) (2003) 199–209.
404–418. [46] E. Clementi, D.L. Raimondi, W.P. Reinhardt, Atomic screening constants from
[18] V.K. Champagne, The Cold Spray Materials Deposition Process: Fundamentals SCF functions: atoms with 37 to 86 electrons, J. Chem. Phys. 47 (4) (1967)
and Applications, Woodhead Publishing, Cambridge, UK, 2007. 1300–1307.
[19] Standard, test methods for tension testing of metallic materials, ASTM [47] F. Qin, Y. Li, W. He, X. Zhao, H. Chen, Aging precipitation behavior and its
International, West Conshohocken, PA, 2016. influence on mechanical properties of Mn18Cr18N austenitic stainless steel,
[20] Standard guide for laboratory immersion corrosion testing of metals, ASTM Met. Mater. Int. 23 (6) (2017) 1087–1096.
International, West Conshohocken, PA, 2012. [48] C.-C. Hsieh, W. Wu, Overview of intermetallic sigma phase precipitation in
[21] Z. Zhen, T.-F. Xi, Y.-F. Zheng, A review on in vitro corrosion performance test of stainless steels, ISRN Metallurgy 2012 (2012) 1–16.
biodegradable metallic materials, Trans. Nonferr. Met. Soc. China. 23 (8) [49] T. Michler, Austenitic Stainless Steels, Reference Module in Materials Science
(2013) 2283–2293. and Materials Engineering, Elsevier, 2016.
[22] Standard guide for conducting and evaluating galvanic corrosion tests in [50] S.M. Hassani-Gangaraj, A. Moridi, M. Guagliano, Critical review of corrosion
electrolytes, ASTM International, West Conshohocken, PA, 2014. protection by cold spray coatings, Surf. Eng. 31 (11) (2015) 803–815.
[23] T. Yamashita, P. Hayes, Analysis of XPS spectra of Fe2+ and Fe3+ ions in oxide [51] M. Schinhammer, A.C. Hänzi, J.F. Löffler, P.J. Uggowitzer, Design strategy for
materials, Appl. Surf. Sci. 254 (8) (2008) 2441–2449. biodegradable Fe-based alloys for medical applications, Acta Biomater. 6 (5)
[24] R. Tolouei, J. Harrison, C. Paternoster, S. Turgeon, P. Chevallier, D. Mantovani, (2010) 1705–1713.
The use of multiple pseudo-physiological solutions to simulate the [52] Y.P. Feng, A. Blanquer, J. Fornell, H. Zhang, P. Solsona, M.D. Baró, S. Suriñach, E.
degradation behavior of pure iron as a metallic resorbable implant: a Ibáñez, E. García-Lecina, X. Wei, R. Li, L. Barrios, E. Pellicer, C. Nogués, J. Sort,
surface-characterization study, Phys. Chem. Chem. Phys. 18 (29) (2016) Novel Fe–Mn–Si–Pd alloys: insights into mechanical, magnetic, corrosion
19637–19646. resistance and biocompatibility performances, J. Mater. Chem. B 4 (39) (2016)
[25] K. Idczak, R. Idczak, R. Konieczny, An investigation of the corrosion of 6402–6412.
polycrystalline iron by XPS, TMS and CEMS, Physica B: Condensed Matter 491 [53] T. Huang, Y. Zheng, Uniform and accelerated degradation of pure iron
(2016) 37–45. patterned by Pt disc arrays, Sci. Rep. 6 (2016) 23627.
[26] R. França, T.D. Samani, G. Bayade, L.H. Yahia, E. Sacher, Nanoscale surface [54] G. Wranglen, Pitting and sulphide inclusions in steel, Corros. Sci. 14 (5) (1974)
characterization of biphasic calcium phosphate, with comparisons to calcium 331–349.
hydroxyapatite and b-tricalcium phosphate bioceramics, J. Colloid Interface [55] R. Avci, B.H. Davis, M.L. Wolfenden, I.B. Beech, K. Lucas, D. Paul, Mechanism of
Sci. 420 (2014) 182–188. MnS-mediated pit initiation and propagation in carbon steel in an anaerobic
[27] Y. Barbaux, M. Dekiouk, D. Le Maguer, L. Gengembre, D. Huchette, J. Grimblot, sulfidogenic media, Corros. Sci. 76 (2013) 267–274.
Bulk and surface analysis of a Fe-P-O oxydehydrogenation catalyst, Appl. Catal. [56] Y. Wang, G. Cheng, Y. Li, Observation of the pitting corrosion and uniform
A 90 (1) (1992) 51–60. corrosion for X80 steel in 3.5wt.% NaCl solutions using in-situ and 3-D
[28] A.P. Grosvenor, B.A. Kobe, M.C. Biesinger, N.S. McIntyre, Investigation of measuring microscope, Corros. Sci. 111 (2016) 508–517.
multiplet splitting of Fe 2p XPS spectra and bonding in iron compounds, Surf. [57] J. Zhang, Z.L. Wang, Z.M. Wang, X. Han, Chemical analysis of the initial
Interface Anal. 36 (12) (2004) 1564–1574. corrosion layer on pipeline steels in simulated CO2-enhanced oil recovery
[29] C.C. Chusuei, D.W. Goodman, M.J. Van Stipdonk, D.R. Justes, E.A. Schweikert, brines, Corros. Sci. 65 (2012) 397–404.
Calcium Phosphate Phase Identification Using XPS and Time-of-Flight Cluster [58] J. Wei, J. Dong, Y. Zhou, X. He, C. Wang, W. Ke, Influence of the secondary phase
SIMS, Anal. Chem. 71 (1) (1999) 149–153. on micro galvanic corrosion of low carbon bainitic steel in NaCl solution,
[30] B. Jodoin, L. Ajdelsztajn, E. Sansoucy, A. Zúñiga, P. Richer, E.J. Lavernia, Effect of Mater. Charact. 139 (2018) 401–410.
particle size, morphology, and hardness on cold gas dynamic sprayed [59] T.V. How, R.K. Fisher, S.R. Vallabhaneni, C.K. Chong, 8 - Vascular implants for
aluminum alloy coatings, Surf. Coat. Technol. 201 (6) (2006) 3422–3429. peripheral arterial bypass and aortic aneurysm repair, in: T. Gourlay, R.A.
[31] V. Roy, B. Nahak, M. Zaheer, M. Vashista, Assessment of plastic deformation Black (Eds.), Biomaterials and Devices for the Circulatory System, Woodhead
upon grinding using x-ray diffraction profiles, Int. J. Eng. Res. 2 (2) (2013) 119– Publishing, 2010, pp. 217–248.
124. [60] Standard specification for wrought 18chromium-14nickel-2.5molybdenum
[32] Ö. Cem, L.E. Dirk, An experimental investigation into strain and stress stainless steel bar and wire for surgical implants (UNS S31673), ASTM
partitioning of duplex stainless steel using digital image correlation, X-ray International, West Conshohocken, PA, USA, 2013.
diffraction and scanning Kelvin probe force microscopy, J. Strain Anal. Eng. [61] H. Hermawan, Biodegradable Metals: From Concept to Applications, Springer,
Des. 51 (3) (2016) 207–219. Berlin; New York, 2012.
[33] M. Moravej, A. Purnama, M. Fiset, J. Couet, D. Mantovani, Electroformed pure [62] P.K. Bowen, E.R. Shearier, S. Zhao, R.J. Guillory, F. Zhao, J. Goldman, J.W. Drelich,
iron as a new biomaterial for degradable stents: in vitro degradation and Biodegradable metals for cardiovascular stents: from clinical concerns to
preliminary cell viability studies, Acta Biomater. 6 (5) (2010) 1843–1851. recent Zn-alloys, Adv. Healthcare Mater. 5 (10) (2016) 1121–1140.
[34] J. Ding, E.-H. Han, Z. Zhang, S. Wang, J. Wang, Influence of sigma phase on [63] M. Moravej, D. Mantovani, Biodegradable metals for cardiovascular stent
corrosion behavior of 316L stainless steel in high temperature and high application: interests and new opportunities, Int. J. Mol. Sci. 12 (7) (2011)
pressure water, Mater. High Temp. 34 (1) (2017) 78–86. 4250–4270.
[35] Z. StonawskÁ, M. Svoboda, M. SozaŃSka, M. KŘÍStkovÁ, J. Sojka, C. Dagbert, L. [64] N. Eliaz, Corrosion of Metallic Biomaterials: A Review, Materials (Basel,
HyspeckÁ, Structural analysis and intergranular corrosion tests of AISI 316L Switzerland) 12(3) (2019) 407.
steel, J. Microsc. 224(1) (2006) 62-64 [65] P.R. Roberge, Handbook of corrosion engineering, McGraw-Hill, Toronto,
[36] M.R. Rokni, C.A. Widener, V.K. Champagne, G.A. Crawford, S.R. Nutt, The effects Canada, 2000.
of heat treatment on 7075 Al cold spray deposits, Surf. Coat. Technol. 310 [66] H. Hermawan, D. Dube, D. Mantovani, Degradable metallic biomaterials:
(2017) 278–285. Design and development of Fe-Mn alloys for stents, J. Biomed. Mater. Res. Part
[37] S.J. Choi, H.S. Lee, J.W. Jang, S. Yi, Corrosion behavior in a 3.5 wt% NaCl solution A 93 (1) (2010) 1–11.
of amorphous coatings prepared through plasma-spray and cold-spray coating
processes, Met. Mater. Int. 20 (6) (2014) 1053–1057.

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034
494
16 J. J.Frattolin
Frattolinetet
al.al.
/ Acta
/ ActaBiomaterialia 99xxx
Biomaterialia (2019) 479–494
(xxxx) xxx

[67] M. Peuster, C. Hesse, T. Schloo, C. Fink, P. Beerbaum, C. von Schnakenburg, [70] M. Peuster, P. Wohlsein, M. Brugmann, M. Ehlerding, K. Seidler, C. Fink, H.
Long-term biocompatibility of a corrodible peripheral iron stent in the porcine Brauer, A. Fischer, G. Hausdorf, A novel approach to temporary stenting:
descending aorta, Biomaterials 27 (28) (2006) 4955–4962. degradable cardiovascular stents produced from corrodible metal—results 6–
[68] W. Lin, L. Qin, H. Qi, D. Zhang, G. Zhang, R. Gao, H. Qiu, Y. Xia, P. Cao, X. Wang, 18 months after implantation into New Zealand white rabbits, Heart 86 (5)
W. Zheng, Long-term in vivo corrosion behavior, biocompatibility and (2001) 563–569.
bioresorption mechanism of a bioresorbable nitrided iron scaffold, Acta [71] M.C. Biesinger, B.P. Payne, A.P. Grosvenor, L.W.M. Lau, A.R. Gerson, R.S.C.
Biomater. 54 (2017) 454–468. Smart, Resolving surface chemical states in XPS analysis of first row transition
[69] M.P. P., A. Sylvia, B. Muhammad, B. Dirk, B. Friedrich-Wilhelm, D. Andreas, M.- metals, oxides and hydroxides: Cr, Mn, Fe, Co and Ni, Appl. Surface Sci. 257 (7)
L. Andrea, H. Hansjörg, P. Matthias, Histological and molecular evaluation of (2011) 2717–2730.
iron as degradable medical implant material in a murine animal model, J.
Biomed. Mater. Res. Part A 100A(11) (2012) 2881-2889

Please cite this article as: J. Frattolin, R. Roy, S. Rajagopalan et al., A manufacturing and annealing protocol to develop a cold-sprayed Fe-316L stainless steel
biodegradable stenting material, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2019.08.034

You might also like