Vous êtes sur la page 1sur 8

Applied Surface Science 351 (2015) 1161–1168

Contents lists available at ScienceDirect

Applied Surface Science


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

A two-step anodic method to fabricate self-organised nanopore arrays


on stainless steel
Bowei Zhang a,b , Hongwei Ni a,b,∗ , Rongsheng Chen a,b , Weiting Zhan a,b , Chao Zhang a,b ,
Rui Lei a,b , Yaxin Zha a,b
a
The State Key Laboratory of Refractories and Metallurgy, Wuhan University of Science and Technology, Wuhan 430081, China
b
Key Laboratory for Ferrous Metallurgy and Resource Utilization of Ministry of Education, Wuhan University of Science and Technology,
Wuhan 430081, China

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

Article history: We report a simple two-step anodic method to fabricate self-organized nanopore arrays on 316L stainless
Received 18 December 2014 steel surface, with the initial anodization in perchloric acid-based electrolyte followed by the second
Received in revised form 29 May 2015 anodization in sodium dihydrogen phosphate-based electrolyte. The morphology and chemical analysis
Accepted 16 June 2015
of anodic overlayers after the first and second anodizations were explored by AFM, SEM and XPS. The
Available online 22 June 2015
influence of various applied voltages, current density, various anodization times and temperatures on
nanoscaled morphology was investigated in the process of the second anodizing step. As a result, the
Keywords:
dimension of the nanopores depends linearly on the anodization time and was controlled by the applied
Two-step anodic method
Self-organization
voltage and current density in a certain range. At the premise of no appreciable expansion in the mean
Nanopore arrays pore size after the second step in contrast to the initial pores, the depth of nanopores could reach seven
Stainless steel times of the previous depth of 10 nm. The analysis of XPS spectrum performed on the nanostructure,
the energy shifts of Cr 2p3/2 , Fe 2p3/2 , O 1s, and Ni 2p3/2 levels showed that principal constituents of the
anodic overlayer are chromium and iron components. No metallic Fe, Cr, and Ni peaks were detected
on the surface after the second anodization. The depth of nanopores was roughly in agreement with the
thickness estimations obtained from depth profiling data and different chemical compositions that exist
at different depths of even a thin anodic overlayer.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction Some performances of nanostructured materials were influ-


enced strongly by structural parameters, such as pore size, depth
In recent years, tremendous scientific interest has been focused and interpore distance [11,12]. Thus, to facilitate various prac-
on fabrication of nanostructured materials due to their remarkable tical applications, fabrication of highly ordered and controllable
properties which make them very useful for many technological nanostructured materials is an essential task. Considerable sci-
applications e.g., in magnetic, electronic, photonic, sensing devices entific attention has been paid on developing new, simple and
and biomaterials [1–6]. For example, nanoporous materials with unexpensive methods of synthesis of controllable nanostructured
high surface area and pore volume have been used for implantable materials.
drug delivery systems [7–9]. Another example was shown by our Due to the excellent corrosion resistance and mechanical prop-
research, i.e., the nanopore arrays on stainless steel substrate as the erties, 316L stainless steel has been extensively employed in
template for 3D graphene platelets possesses high electrochem- cardiovascular stents, orthopedic implants, and spinal fixation
ical activity for the determination of uric acid, ascorbic acid and devices [13–15]. However, the poor adhesion of the high polish
dopamine [10]. metal surface cannot match the requirement of long-term stability
of an artificial device to modulate cell-substrate interactions. Func-
tionalization of the material surface by applying defined micro-
and nanostructures has been explored to improve osseointegra-
tion and modulation of cellular activity. For example, the nanopore
∗ Corresponding author at: The State Key Laboratory of Refractories and Metal-
surfaces fabricated on 316L stainless steel substrates are able
lurgy, Wuhan University of Science and Technology, Wuhan 430081, China.
Tel.: +86 27 68862856; fax: +86 27 68862811.
to influence the expression of integrin and direct cell behaviors
E-mail address: nihongwei@wust.edu.cn (H. Ni). such as cell migration, proliferation and focal adhesion formation

http://dx.doi.org/10.1016/j.apsusc.2015.06.083
0169-4332/© 2015 Elsevier B.V. All rights reserved.
1162 B. Zhang et al. / Applied Surface Science 351 (2015) 1161–1168

[11]. The mouse pre-osteoblasts cell attachment, proliferation and determined by argon-ion sputtering for different time intervals.
prominent proteins can be modulated on nanometer-sized grains The reference sputtering rate was 0.78 nm/s. An inductively cou-
structure on the 316L stainless steel surface [16]. pled plasma emission spectrometer (ICP-AES, IRIS Advantage ER/S,
In the case of stainless steel, few researches have been devoted Thermo Elemental, USA) was employed to determine the compo-
to the surface of stainless steel after its anodization. Vignal et al. nents of the precipitate formed during the second anodization, the
[17] have reported that brightening of AISI 316L stainless steel precipitate was centrifuged out from the electrolyte and dried in
in perchloric acid-based electrolytes can form a quasi-hexagonal the air.
arrangement of 10 nm depth pores at the sample surface. Martin
et al. [18] have shown the chemical and periodic nanostructures of
the films growing during the electropolishing treatment of 304L 3. Results and discussion
stainless steel and duplex stainless steel (Uranus 50). Zhan and
coworkers [19] have reported that during the process of the direct 3.1. 316L stainless steel surface morphology after the first
anodization after the achievement of a microscopically smooth sur- anodizaton
face by electropolishing irregular and undesired patterns appear on
the surface of 316L stainless steel. Surface morphology after the initial anodization was character-
In spite of recent progress, some difficulties still exist in con- ized using SEM and AFM. The surface appeared smooth without
trolling important structural parameters. Fabrication of regular and defects and exhibited a quasi-periodic arrangement of pores. The
deep pores on stainless steel requires new anodic methods con- pores indeed self-organize locally in a perfect honeycomb lattice.
trast to traditional anodization. From the technical point of view, to As in images of AFM and SEM in Fig. 1, the applied voltages are
remedy these problems on stainless steel, we designed a two-step 40 V (Fig. 1a) and 50 V (Fig. 1b), respectively, which gives rise to the
anodic method: the use of perchloric acid-based electrolyte fol- nanopore’s mean dimensions of 85 nm and 120 nm, respectively.
lowed by dihydrogen phosphate-based electrolyte to form a deeper They have a depth of about 10 nm (Fig. 1c and d). The hexagonal
self-organized nanopore arrays on 316L stainless steel substrate. pattern is sensitive to the applied voltage and the hexagons size
The results of studies on the morphology and chemical analysis increases with the rise of applied voltage. Indeed, these results are
of the anodic overlayers after the first and second anodization are similar to those previously reported [17].
specified. Furthermore, the influence of several parameters such Similar to the mechanism of hexagonal pores appears on Al,
as current density, applied voltage and anodization time on the the formation of hexagonal structure on stainless steel is due to a
morphology of the surface patterns will be discussed. cracking and self-healing of the oxide layer atop preexisting ridges
on the stainless steel surface and that this forms a barrier layer of
nonuniform thickness [20].
2. Experimental methods

The 316L stainless steel plates (Goodfellow, 0.015 wt.% C, 3.2. Effects of two-step anodic method on the morphology of 316L
16–18 wt.% Cr, 10–14 wt.% Ni, 0.3 wt.% Mo, and Fe balance) were cut substrate
into 10 mm × 10 mm × 1 mm size. All the specimens were mechan-
ically polished with successive grades of abrasive papers (grade Evident is the result from the micrographs in Fig. 2a that the
1000, 1500 and 2000) and smoothed with diamond pastes (from3.5 direct anodization after mechanical polish results in an irregular
to 0.25 ␮m). After the above two processes, they were rinsed with pattern on the surface of 316L stainless steel. On the other hand,
ethanol and double-distilled water under ultrasonic for 10 min. The as shown in Fig. 2b, the two-step anodic method significantly
initial anodization of the mechanically polished specimens was improves the ordering degree of nanopore arrays compared with
performed in ethylene glycol (EG; 99.8% and anhydrous) solution the direct anodization. The observed ordered nanopore arrays
containing 5 vol% perchloric acid (HClO4 ; 70%) at a constant voltage could be attributed to the periodic pore nucleation sites formed by
of 40–50 V for 10 min. After the first anodization, the foil was then the first anodization in perchloric acid-based electrolyte. Thus, in
anodized in 0.3 M sodium dihydrogen phosphate solution with the nanoporous stainless steel, surface pretreatment prior to the actual
applied voltage ranged from 5 V to 45 V. After each anodization, the anodization plays a crucial role in the formation of nanopores,
samples were ultrasonically cleaned in acetone, ethanol, and dis- greatly influencing their size uniformity.
tilled water and dried in air. Anodization was conducted in a stirred As previously reported about Al and Ti, the formation of a bar-
two-electrode electrochemical cell by a direct current power sup- rier oxide layer or the prior formation of porous film by a different
ply (IT6154, ITECH, Nanjing, China). The electrolytic solution was electrolyte guarantees not only stable anodization but also homo-
stirred by a rotating magnet. Samples (the area of anodic electrode geneous growth of self-ordered oxide nanopore arrays. It has been
dipped in the electrolyte was about 1 cm2 ) were positioned verti- suggested that as anodization started, pores would start at cracks
cally in front of the counter electrode, which was a 9 cm2 graphite and imperfections in the surface, leaving an electric field concentra-
cathode, and the distance between both electrodes was kept at tion below the regions where the oxide film was thinner, because
about 6 cm. The electrolytic temperature was maintained between the thickness of the barrier oxide layer on a 316L stainless steel
0 and 10 ◦ C in an ice/water bath. surface is nonuniform, and the center of pores is the thinnest layer,
The morphology of the anodic films was characterized by field where the resistance is lowest and the electric field is highest. This
emission SEM (FE-SEM, FEI Nova 400 Nano) with an accelerat- leads to a temporarily increased dissolution rate at the bottom of
ing voltage of 20 kV. The micro images of the anodic films were pores [21,22]. As a consequence, pore nucleation for the porous
obtained in “tapping mode” by atomic force microscope (AFM) surface is easier than the flat ones.
Agilent 5500 (Agilent Technologies, Chandler, AZ) with a scan- Solution/oxide and oxide/metal interfaces exist in the pore
ning range of 5 ␮m. X-ray photoelectron spectroscopy (XPS, VG channel. Field-enhanced oxidation occurs at the metal/oxide inter-
Multilab 2000, Thermo Electron Corp., America) experiments were face near the bottom of pore when the oxygen containing ions
carried out using a VG Escalab MKII spectrometer with a non- transport from electrolyte to the oxide film, along the direction
monochromatic Al K␣ X-ray source (1486.6 eV) to investigate the of the pore growth. At the same time, metal ions migrate from
chemical composition of the anodic films on stainless steel surface, metal substrate to the solution/oxide interface and dissolve into
calibrated by C 1s at 284.5 eV. The thicknesses of the films were the electrolyte [22].
B. Zhang et al. / Applied Surface Science 351 (2015) 1161–1168 1163

Fig. 1. Morphologies of 316L stainless steel surface after the first anodization at (a) 40 V and (b) 50 V. (c) Profile along the white segment in Fig. 1a and (d) profile along the
white segment in Fig. 1b showing the relief amplitude. The measured depths both are about 10 nm, a value which may depend on the tip shape. Scale bar = 500 nm.

3.3. Morphology of 316L stainless steel surface after the second surface, but the pores are deeper. The domain boundaries are also
anodization shown in Fig. 3, as indicated by the white arrows. The domains grew
by alignments and the mergers of pores were regular at domain
Because of the nanopore diameter’s decrease toward the sub- boundaries. Maybe the bottoms of the pores could move around
strate, it is hard to define accurately the diameter and the wall [22]. As the pores moved and merged, the orientation of adjacent
thickness of the nanopores. Therefore, we will concentrate on domains could change gradually, so domains matched each other.
reporting the influence of the growth parameters on the inner In contrast to the polished 316L stainless steel, no color change
diameter measured from SEM images and the depth of pores. appears on the sample surface after anodization.
We used nanoporous surface which was generated by the ini- For a better understanding of this nanostructured surface, AFM
tial anodization to fabricate deeper self-organized nanopores in observations in air have been performed. With the AFM, the mea-
0.3 M sodium dihydrogen phosphate-based electrolyte. A perfect sure may depend on the geometry of the tip and its apex. As
self-organized nanopore arrays were obtained through the second revealed in pictures of SEM (Fig. 4a) and AFM (Fig. 4b), the voltage
anodization under suitable conditions. Similar to the first anodiza- applied at the first anodizing step was 40 V after 10 min anodiza-
tion results, the distribution of the pores was uniform on the tion in sodium dihydrogen phosphate-based electrolyte at 15 V. The

Fig. 2. SEM images of 316L surface after two different treatments: (a) the direct anodization after mechanical polish and (b) the nanopore arrays that formed through
two-step anodic method. The inset in (a) is the image of mechanically polished surface. Samples were anodized at 25 V.
1164 B. Zhang et al. / Applied Surface Science 351 (2015) 1161–1168

of anodization at 40 V and the second step at 20 V for 15 min. The


profiles present the relative concentrations of Cr, Fe, O, and Ni. The
precise stoichiometry of the oxide cannot be determined due to the
ion-sputtering-induced effects, such as the reduction of the oxi-
dation state and the influence of nanostructured morphology of
the surface [18,23]. Oxygen is enhanced toward the outer regions,
where the signals from Fe species are increased. As the depth ranges
from the surface to 15 s (∼10 nm) and 80 s (∼60 nm), the XPS quan-
titative analysis shows that the content of O element is nearly zero
compared with about 52 at% and 78 at% of the surface. Obviously,
the O fraction in the second anodized sample was higher than the
initial surface. In the first anodized sample, the Cr fraction was
slightly decreased. In contrast, the Cr fraction was slightly increased
in the second anodized sample.
Fig. S1 showed the overview XPS spectra of 316L stainless steel
surfaces after mechanical polishing, the first anodization, and the
second anodization, respectively. The peaks of Fe 2p, Cr 2p, and Ni
2p are obvious on the bare 316L stainless steel surface. In contrast
Fig. 3. Domain boundaries and pore merging along the domain boundaries. The with bare 316L flat, the intensity of O 1s signal increases, Cr 2p peak
white arrows show the merging of pores.
is almost consistent, and Fe, Ni peaks decreases on the first anodized
sample surface. After the second anodization, O 1s signal increases
measured mean dimension of pores was about 90 nm and the depth greatly, while the Cr, Ni peaks are almost completely vanished from
was of 30–40 nm (Fig. 4c), which is three to four times that of initial the sample surface.
pores. Self-organized pores were shown in Fig. 4d and e that was The high-resolution spectra of Cr 2p3/2 , Fe 2p3/2 and O 1s at dif-
obtained at 20 V potential for 15 min after the initial achievement ferent depths from the surface were all analyzed in Fig. 5. As shown
of a 120 nm nanoporous surface at 50 V in perchloric acid-based in Fig. 5c, e and g, the spectra of Cr 2p3/2 , Fe 2p3/2 and O 1s for the first
electrolyte. The mean size of nanopores is about 130 nm and they anodized film reveals the presence of metallic Cr (573.6 eV), Cr2 O3
are about 70–80 nm deep (Fig. 4f), which is about seven to eight (576.1 eV), Fe (706.3 eV), FeO (708.4 eV) and Fe2 O3 (710.9 eV) on the
times that of the initial depth of the pores. Obviously, there is no surface [24–26]. The spectrum shows visible changes in chemical
appreciable difference between the pore dimensions after the sec- composition of Cr and Fe after 10 s or 20 s etching [27].
ond anodization and the initial one. The measured thickness of the The Cr 2p3/2 , Fe 2p3/2 and O 1s spectrum obtained from the sec-
walls is 20–50 nm without correction due to the tip shape [18]. As ond anodized overlayer was revealed in Fig. 5d, f and h. For Cr,
shown in SEM (Fig. 4a and d) and AFM (Fig. 4b and e) images, all the Fe, no metallic Fe and Cr peaks were detected at the surface while
white spots can be associated with defects in the nanopore arrays. another peak of FeOOH (711.8 eV) emerges. The peaks of metallic
Fe (706.4 eV), FeO (708.4 eV), Fe2 O3 (710.9 eV) were detected after
3.4. Chemical analysis of the nanostructure and the mechanisms 45 s etching [27–29]. The Cr 2p3/2 spectra show two types of bonds,
of anodization such as metallic Cr and Cr2 O3 . After 90 s etching, the main peak in Cr
2p3/2 can be separated into two possible peaks. As shown in Fig. 5f,
Fig. 5a and b revealed some qualitative information on the com- similar case is also found in the spectra of Fe, which are related to
position depth profiles for 316L stainless steel after the first step metallic Fe and FeO and two kinds of Fe oxide components appear

Fig. 4. (a) SEM and (b) AFM images of stainless steel anodized at 15 V. (c) Profile along the white segment in Fig. 4b showing the relief amplitude, the measured depth is
about 30–40 nm. (d) SEM and (e) AFM images show the surface morphology anodized at 20 V for 15 min after the first anodization at 50 V. (f) Profile along the white segment
in Fig. 4e showing the relief amplitude, the measured depth is about 70–80 nm, a value which may depend on the tip shape. Scale bar = 500 nm.
B. Zhang et al. / Applied Surface Science 351 (2015) 1161–1168 1165

Fig. 5. (a) The XPS depth profiles and high resolution XPS spectra of (c) Cr 2p3/2 , (e) Fe 2p3/2 , and (g) O 1s peaks collected from the anodic films that formed after the first
anodization at 50 V for 10 min. (b) The XPS depth profiles and detailed XPS spectra of (d) Cr 2p3/2 , (f) Fe 2p3/2 , and (h) O 1s peaks collected from the anodic films that formed
after the second anodization at 20 V for 15 min, the first anodization was performed at 50 V for 10 min.

in Fe 2p3/2 peaks after 45 s etching. These two components corre- composition exist at different depths of even a thin anodic overlayer
spond to Fe oxide stoichiometry changing with depth from FeOOH [31].
to FeO and Fe2 O3 . As showed in Fig. S2, nickel existing in different depth of the
Fig. 5g and h displays the detailed XPS spectra of O 1s generated anodic films that formed during the process of the first or the second
from anodic films that formed during the first and second anodizing anodization in the form of metallic nickel and NiO. No metallic Ni
steps. The O 1s spectra show two types of bonds in Fig. 5g, such as 2p peak was detected on the surface after the second anodization
O2− in oxide species and OH− in hydroxide species. Fig. 5b shows [30]. During the first step of anodization, the formation of oxides is
the spectra of O 1s fitted by OH− in hydroxide species on the surface. described as the following reactions [32]:
After 45 s or 90 s etching, two peaks of O 1s correspond to O2− in
oxide species at binding energies of 530.2 eV and OH− in hydroxide 2H2 O → O2 + 4H+ + 4e−
species at binding energies of 531.5 eV, respectively [30]. It appears
that the nanostructured surface was consistent with a mixture of y
Cr and Fe components. This may suggest that different chemical xM(Fe, Cr, Ni) + O2 → Mx Oy (FeO, Fe2 O3 , Cr2 O3 , NiO)
2
1166 B. Zhang et al. / Applied Surface Science 351 (2015) 1161–1168

Fig. 6. (a) Examples of j–t curves characteristic of nanopores that formed and morphology evolution of the anodized 316L stainless steel surface upon the experimental
conditions. (b) Relationship between the applied voltages and pore dimensions during the second anodizing step after the first anodization at 40 V and 50 V, respectively.
(c) Inner dimension-time anodization curve during the second anodizing step. The dimension increases linearly with bias voltage in a certain range. Scale bar = 500 nm. The
error bars represent the standard deviations from the mean of each measured quantity.

The anodization process was further investigated by determina-


tion of the metal ions in the electrolyte after the first anodization, The dissolved metallic cations (Cr3+ , Ni2+ ) can react with hydrogen
as shown in Fig. S3. The color of the electrolyte turned to brown phosphate, and form phosphates [36,37]:
yellow after the first step of anodization, indicating the existence Cr3+ + 2HPO2− −
4 → [Cr(HPO4 )2 ]
of Fe3+ in the electrolyte. The results suggest that the dissolution of
these metal oxides including Fe2 O3 , NiO, and Cr2 O3 , appears during Ni2+ + 2HPO2−
4 → [Ni(HPO4 )2 ]
2−

the first step of anodization [18,33–35]:


The precipitate should be attributed to the hydrolysis of Fe3+ in
Fe2 O3 + 6H+ → 2Fe3+ + 3H2 O the solution of sodium dihydrogen phosphate [36]:

Cr2 O3 + 6H+ → 2Cr3+ + 3H2 O Fe3+ + 2H2 O → FeOOH + 3H+

NiO + 2H+ → Ni2+ + H2 O


3.5. Nanostructure growth parameters
The enrichment of Cr atoms in the anodic layer is attributed to
the less proportion of Cr atoms in the solution than that in the bulk The morphology of the nanostructured surface is influenced by
material of 316L stainless steel. the physico-chemical conditions of the second anodizing process
The formation of the nanostructures during the second anodiza- [38]. Besides the applied voltage, the current density and temper-
tion is described as the following reactions [32]: ature have been investigated. By changing the temperature of the
electrolyte from 0 ◦ C up to 10 ◦ C, no change in the morphology of
2H2 O → O2 + 4H+ + 4e− the nanostructured surface was noticed. Further experiments have
been performed to investigate the influence of anodizing voltage
and the current density changes on the formation and morphology
y of nanostructure. The applied voltage was regulated from 5 V up to
xM(Fe, Cr, Ni) + O2 → Mx Oy (FeO, Fe2 O3 , Cr2 O3 , NiO) 45 V. Average values of the current density recorded during steady-
2
state growth of porous anodic 316L stainless steel in 0.3 M sodium
According to the chemical composition of precipitate after the dihydrogen phosphate solution show a dependency on the anodiz-
second anodization step (Fig. S4), the precipitate was mainly com- ing potential independently of the electrolyte temperature. The
posed of iron compound. The trace amount of Cr and Ni in the reaction was strongly affected by the current density. Fig. 6a depicts
precipitate should be ascribed to the absorption on the precipitate. the influence of applied voltages on the current density and the
B. Zhang et al. / Applied Surface Science 351 (2015) 1161–1168 1167

morphology of nanostructures in the second anodization process. [2] D. Ho, D. Garcia, C.M. Ho, Nanomanufacturing and characterization modalities
The current density increased with the voltage. At 5 V the current for bio-nano-informatics systems, J. Nanosci. Nanotechnol. 6 (2006) 875–891.
[3] D. Grieshaber, R. MacKenzie, J. Voros, E. Reimhult, Electrochemical biosensors
was very steady; the morphology has almost no obvious changes – sensor principles and architectures, Sensors 8 (2008) 1400–1458.
compared with the first anodization. At 45 V the current density [4] X. Li, Strain induced semiconductor nanotubes: from formation process to
showed large fluctuations and the interpore walls are etched during device applications, J. Phys. D: Appl. Phys. 41 (2008) 193001–193013.
[5] J.B. Dai, J.K. Tang, S.T. Hsu, W. Pan, Magnetic nanostructures and materi-
the anodization process. Following the applied voltages increase in als in magnetic random access memory, J. Nanosci. Nanotechnol. 2 (2002)
range of 5–45 V, the chemical dissolution rate at the layer/solution 281–291.
interface increases. Local temperature may rise and rates of heat [6] S. Oh, K.S. Brammer, Y.S. Julie Li, D. Teng, A.J. Engler, S. Chien, S. Jin, Stem cell
fate dictated solely by altered nanotube dimension, Proc. Natl. Acad. Sci. U. S.
dissipation may become nonuniform, thus causing fluctuations in
A. 106 (2009) 2130–2135.
the anodization current density [22]. As shown in Fig. 6b, it was [7] D. Losic, S. Simovic, Self-ordered nanopore and nanotube platforms for drug
found that the applied voltage has a strong influence on the mor- delivery applications, Exp. Opin. Drug Deliv. 6 (2009) 1363–1381.
[8] E.A. Jackson, M.A. Hillmyer, Nanoporous membranes derived from block
phology of the surface, and that the pore dimension increases with
copolymers: from drug delivery to water filtration, ACS Nano 4 (2010)
the potential in a certain range. 3548–3553.
Surprisingly, as we can see from Fig. 6c, the dimension of pore [9] E. Gultepe, D. Nagesha, S. Sridhar, M. Amiji, Nanoporous inorganic membranes
increases linearly with the second anodization time in the process or coatings for sustained drug delivery in implantable devices, Adv. Drug Deliv.
Rev. 62 (2010) 305–315.
of the second step at 20 V, the first step was performed at 50 V [10] Z. Wang, H. Ni, R. Chen, W. Zhan, C. Zhang, R. Lei, B. Zhang, Enhanced perfor-
in perchloric acid-based electrolyte. In the range of 5–15 min, the mance of multilayer graphene platelet film via three dimensional configuration
dimension of pores increases from about 90 nm to 130 nm. How- with efficient exposure of graphitic edge planes, Electrochem. Commun. 47
(2014) 75–79.
ever, the walls were etching when the anodization time amounts [11] H.-A. Pan, J.-Y. Liang, Y.-C. Hung, C.-H. Lee, J.-C. Chiou, G.S. Huang, The spa-
to 20 min. This fact is enough to demonstrate that these parame- tial and temporal control of cell migration by nanoporous surfaces through
ters have significant effects on the achieved nanostructure on 316L the regulation of ERK and integrins in fibroblasts, Biomaterials 34 (2013)
841–853.
stainless steel surface. [12] S.H. Chung, S.J. Son, J. Min, The nanostructure effect on the adhesion and
growth rates of epithelial cells with well-defined nanoporous alumina sub-
strates, Nanotechnology 21 (2010) 125104–125111.
4. Conclusions [13] L. Cardenas, J. MacLeod, J. Lipton-Duffin, D.G. Seifu, F. Popescu, M. Siaj, D. Man-
tovani, F. Rosei, Reduced graphene oxide growth on 316L stainless steel for
It was found that a porous anodic film that formed on initial medical applications, Nanoscale 6 (2014) 8664–8670.
[14] P. Elter, F. Sickel, A. Ewald, Nanoscaled periodic surface structures of medi-
anodized 316L stainless steel before the second anodization pro- cal stainless steel and their effect on osteoblast cells, Acta Biomater. 5 (2009)
cess can effectively suppress undesired surface events, and that the 1468–1473.
initial pores left in the first anodized surface actually guided pore [15] A. Latifi, M. Imani, M.T. Khorasani, M.D. Joupari, Electrochemical and chem-
ical methods for improving surface characteristics of 316L stainless steel for
nucleation during the second step. The two-step anodic method
biomedical applications, Surf. Coat. Technol. 221 (2013) 1–12.
significantly improves the ordering degree of nanopore arrays com- [16] R.D.K. Misra, C. Nune, T.C. Pesacreta, M.C. Somani, L.P. Karjalainen, Understand-
pared to the direct anodization after mechanically polished surface. ing the impact of grain structure in austenitic stainless steel from a nanograined
In the second anodization, our results indicated that a linearly rela- regime to a coarse-grained regime on osteoblast functions using a novel metal
deformation–annealing sequence, Acta Biomater. 9 (2013) 6245–6258.
tionship exists between the mean inner dimension and anodization [17] V. Vignal, J.C. Roux, S. Flandrois, A. Fevrier, Nanoscopic studies of stainless steel
time, and that the surface morphology is also influenced by the electropolishing, Corros. Sci. 42 (2000) 1041–1053.
applied voltage and current density in a certain range. At the [18] F. Martin, D. Del Frari, J. Coustya, C. Bataillon, Self-organisation of nanoscaled
pores in anodic oxide overlayer on stainless steels, Electrochim. Acta 54 (2009)
premise of no significant diameter expansion in contrast to that 3086–3091.
of the initial pores, we found that the measured mean depth after [19] W.T. Zhan, H.W. Ni, R.S. Chen, Z.Y. Wang, Y.W. Li, J.H. Li, One-step hydrothermal
the second anodization is about three to seven times of the initial preparation of TiO2 /WO3 nanocomposite films on anodized stainless steel for
photocatalytic degradation of organic pollutants, Thin Solid Films 548 (2013)
pores; this is also a clear improvement over previous syntheses. 299–305.
The analysis of shifts of Cr 2p 3/2 , Fe 2p 3/2 , O 1s, and Ni 2p 3/2 [20] K. Shimizu, K. Kobayashi, G.E. Thompson, Development of porous anodic films
levels in XPS spectra suggested that the depth of nanopores were on aluminium, Philos. Mag. A 66 (1992) 643–652.
[21] W. Lee, K. Nielsch, U. Gösele, Self-ordering behavior of nanoporous anodic alu-
roughly consistent with thickness estimations obtained from depth minum oxide (AAO) in malonic acid anodization, Nanotechnology 18 (2007)
profiling data, and that different chemical composition of Cr and Fe 475713–475721.
exist at different depths of even a thin oxide layer. At the same [22] F. Li, L. Zhang, R.M. Metzger, On the growth of highly ordered pores in anodized
aluminum oxide, Chem. Mater. 10 (1998) 2470–2480.
time, the possible mechanisms of the anodization were proposed
[23] W.H. Hocking, F.W. Stanchell, E. McAlpine, D.H. Lister, Mechanisms of corro-
by us. We believe that the present results may facilitate the further sion of stellite-6 in lithiated high temperature water, Corros. Sci. 25 (1985)
investigations of other porous oxides. 531–557.
[24] A. Kocijan, Č. Donik, M. Jenko, Electrochemical and XPS studies of the passive
film formed on stainless steels in borate buffer and chloride solutions, Corros.
Acknowledgement Sci. 49 (2007) 2083–2098.
[25] S.J. Yuan, S.O. Pehkonen, Microbiologically influenced corrosion of 304 stain-
less steel by aerobic Pseudomonas NCIMB 2021 bacteria: AFM and XPS study,
This work was supported by the National Natural Science Foun- Colloids Surf. B: Biointerfaces 59 (2007) 87–99.
dation of China (No. 51471122, No. 51171133) and the Program for [26] J. Xu, X. Wu, E.-H. Han, The evolution of electrochemical behaviour and oxide
New Century Excellent Talents in University (No. NCET-07-0650). film properties of 304 stainless steel in high temperature aqueous environment,
Electrochim. Acta 71 (2012) 219–226.
[27] A.L. Neal, K. Lowe, T.L. Daulton, J. Jones-Meehan, B.J. Little, Oxidation state of
Appendix A. Supplementary data chromium associated with cell surfaces of Shewanella oneidensis during chro-
mate reduction, Appl. Surf. Sci. 202 (2002) 150–159.
[28] C.Q. Cheng, J. Zhao, T.S. Cao, Q.Q. Fu, M.K. Lei, D.W. Deng, Facile chromaticity
Supplementary data associated with this article can be found, in approach for the inspection of passive films on austenitic stainless steel, Corros.
the online version, at http://dx.doi.org/10.1016/j.apsusc.2015.06. Sci. 70 (2013) 235–242.
[29] W. Fredriksson, K. Edström, XPS study of duplex stainless steel as a possible
083 current collector in a Li-ion battery, Electrochim. Acta 79 (2012) 82–94.
[30] X. Cheng, Z. Feng, C. Li, C. Dong, X. Li, Investigation of oxide film formation on
316L stainless steel in high-temperature aqueous environments, Electrochim.
References Acta 56 (2011) 5860–5865.
[31] C. Donik, D. Mandrino, M. Jenko, Depth profiling and angular dependent XPS
[1] V. Balzani, Nanoscience and nanotechnology: the bottom-up construction of analysis of ultra thin oxide film on duplex stainless steel, Vacuum 84 (2010)
molecular devices and machines, Pure Appl. Chem. 80 (2008) 1631–1650. 1266–1269.
1168 B. Zhang et al. / Applied Surface Science 351 (2015) 1161–1168

[32] M. Paulose, K. Shankar, S. Yoriya, H.E. Prakasam, O.K. Varghese, G.K. Mor, T.A. [36] L. Fu, H. Yu, Y. Li, C. Zhang, X. Wang, Z. Shao, B. Yia, Ethylene glycol adjusted
Latempa, A. Fitzgerald, C.A. Grimes, Anodic growth of highly ordered TiO2 nano- nanorod hematite film for active photoelectrochemical water splitting, Phys.
tube arrays to 134 ␮m in length, J. Phys. Chem. B 110 (2006) 16179–16184. Chem. Chem. Phys. 16 (2014) 4284–4290.
[33] A.V. Kuzin, I.G. Gorichev, Y.A. Lainer, Stimulating effect of phosphate ions on the [37] B.F. Alfonso, C. Piqué, C. Trobajo, J.R. García, E. Kampert, U. Zeitler, J. Rodríguez
dissolution kinetics of iron oxides in an acidic medium, Russ. Metall. (Metally) Fernández, M.T. Fernández-Díaz, J.A. Blanco, Double magnetic phase transition
9 (2013) 652–657. in ND4 Fe(DPO4 )2 and NH4 Fe(HPO4 )2 , Phys. Rev. B 82 (2010) 1–10.
[34] T. Grygar, P. Bezdička, Electrochemical dissolution of CrIII and CrIV oxides, J. [38] S. Yoriya, W. Kittimeteeworakul, N. Punprasert, Effect of anodization parame-
Solid State Electrochem. 3 (1998) 31–38. ters on morphologies of TiO2 nanotube arrays and their surface properties, J.
[35] E.J. Underwood, J.F. Filmer, Enzootic Marasmus: the determination of the bio- Chem. Chem. Eng. 6 (2012) 686–691.
logically potent element (cobalt) in limonite, Aust. Vet. J. 11 (1935) 84–92.

Vous aimerez peut-être aussi