You are on page 1of 9

metals

Communication
Synthesizing Nanoporous Stainless Steel Films via Vacuum
Thermal Dealloying
Xiaotao Liu 1,2, *, Xiaomeng Zhang 2,3 , Maria Kosmidou 2 , Michael J. Detisch 2 and Thomas John Balk 2

1 State Key Laboratory of Materials Processing and Die & Mould Technology, School of Materials Science and
Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
2 Department of Chemical and Materials Engineering, University of Kentucky, Lexington, KY 40506, USA;
xzhang1222@126.com (X.Z.); maria.kosmidou86@gmail.com (M.K.); mjdeti2@uky.edu (M.J.D.);
john.balk@uky.edu (T.J.B.)
3 Gaona Aero Material Co., Ltd., Beijing 100081, China
* Correspondence: xiaotaoliu@hust.edu.cn

Abstract: Vacuum thermal dealloying is a recently developed technique and was newly introduced
to produce nanoporous metals, due to its intriguing advantages, i.e., preventing oxidation and
producing no chemical waste, etc. Here, we report on the fabrication of nanoporous stainless steel
films by vacuum thermal dealloying of sputtered stainless steel–magnesium precursor films. It was
found that crack-free nanoporous stainless steel films can be successfully attained under a broad
temperature range of 450–600 ◦ C, with a dealloying time of 0.5–2 h. The resulting structure and
ligaments were temperature- and time-dependent, and moreover, the condition of “600 ◦ C + 2 h”
generated the most homogeneous structure. Moreover, small amounts of residual Mg were found at
pore sites in the resultant structures, suggesting that the dealloying was not fully complete.

Keywords: vacuum thermal dealloying; nanoporous stainless steel; dealloying temperature; dealloy-
ing time; ligament

1. Introduction
Citation: Liu, X.; Zhang, X.;
Kosmidou, M.; Detisch, M.J.; Balk, T.J. Nanoporous materials can be applied in numerous fields, e.g., sensors [1–3], actua-
Synthesizing Nanoporous Stainless tors [4,5], catalysts [6–10], biomaterials [11–13], and energy conversion and storage [14–17],
Steel Films via Vacuum Thermal due to their unique bi-continuous open porous structure and high surface-area-to-volume
Dealloying. Metals 2023, 13, 1255. ratio. As a result, this type of material has received significant attention, leading to a
https://doi.org/10.3390/met13071255 wide variety of approaches for the fabrication of nanoporous metals, which have emerged
Academic Editors: Dooho Choi and
in the past decades. The porous nanostructure is basically obtained through dealloying,
Young-Rae Cho
which is a robust method for fabricating 3D bicontinuous porous materials. It is generally
believed that the formation of nanopores results from two kinetically competing processes
Received: 12 June 2023 at dealloying fronts, wherein a sacrificial (less stable) component is selectively removed
Revised: 4 July 2023 from the precursor alloy while leaving the remaining (more stable) component to form
Accepted: 7 July 2023 a porous microstructure [4,18,19]. In light of previous investigations, the vast majority
Published: 10 July 2023
of nanoporous metals, for instance nanoporous Au, Cu, Ni, Pd, Ti, and Si [20–23], are
synthesized through the techniques of chemical/electrochemical dealloying [4,19,24,25]
and liquid metal dealloying [10,26–29], which are also the most extensively employed
dealloying methods. However, each of them has its own limitation. More specifically,
Copyright: © 2023 by the authors.
Licensee MDPI, Basel, Switzerland.
the technique of chemical/electrochemical dealloying can be efficiently applied only for
This article is an open access article
fabrication of nanoporous noble metals, while the approach of liquid metal dealloying
distributed under the terms and inherently requires a proper liquid metal to act as the etchant, as well as high temperature
conditions of the Creative Commons experimental conditions. Furthermore, as indicated by Kosmidou et al. [30] and Zhao
Attribution (CC BY) license (https:// et al. [31], neither technique is ideal for preparing nanoporous refractory metals, as any
creativecommons.org/licenses/by/ oxidation that occurs during the dealloying process is undesired.
4.0/).

Metals 2023, 13, 1255. https://doi.org/10.3390/met13071255 https://www.mdpi.com/journal/metals


Metals 2023, 13, 1255 2 of 9

In contrast, the technique of vacuum thermal dealloying adeptly overcomes these


obstacles. As reported in literature [30], Kosmidou et al. succeeded in synthesizing
unoxidized nanoporous refractory metals by employing this method. Unlike traditional
dealloying approaches, vacuum thermal dealloying takes advantage of the vapor pressure
difference between the two components in the precursor materials to selectively remove
one and leave the other to form a nanoporous structure. This technique was initially
developed over 10 years ago and has been recently introduced to materials synthesis,
especially for fabricating nanoporous metals [30]. As demonstrated by Han et al. [22], the
mechanism of vacuum thermal dealloying is a surface diffusion-governed process. This
differs from the techniques of chemical/electrochemical dealloying, which are interface
diffusion-controlled processes. Additionally, the pores that form in vacuum thermal-
dealloyed structures are internal closed pores, which can be attributed to the Kirkendall
effect. The pores generated in chemical/electrochemical dealloyed structures are, however,
open pores. To date, vacuum thermal dealloying has not yet been widely applied in the
fabrication of nanoporous materials, though this technique possesses very conspicuous
virtues and advantages, compared to the conventional dealloying approaches.
Nanoporous stainless steel has been put into use in the field of energy harvesting,
an example being the anode in microbial fuel cells, due to its enhanced bacterial catalytic
activity and stability [15]. In recent years, this material has also started to be applied
in the biomedical field, such as being used as body implants, because cell selectivity
can be achieved by its specific surface nanofeatures [32–34]. According to a literature
study regarding the fabrication of nanoporous stainless steel materials, Mokhtari et al. [11]
succeeded in preparing a fully porous ferritic stainless steel via the approach of liquid
metal dealloying while utilizing the Incoloy 800 precursor alloy. Abbas et al. [15] also
fabricated the nanoporous stainless steel 316L. Nevertheless, the preparation process was
rather complicated. The sample needed to be ultrasonically cleaned, anodized, then
rinsed thoroughly with DI water and dried in air, followed by annealing in a tube furnace.
Moreover, the anodization condition was required to be finely optimized for the generation
of a nanoporous structure. It should be noted that both forms of nanoporous stainless steels
were bulk materials rather than thin films. So far, the fabrication of a nanoporous stainless
steel film has not been reported.
In this paper, we, for the first time, prepared nanoporous stainless steel films via
the vacuum thermal dealloying approach. More specifically, stainless steel–magnesium
precursor films were obtained by magnetron sputtering, where highly purified stainless
steel and magnesium target materials were utilized. After that, nanoporous stainless steel
thin films were generated by dealloying the precursor films at elevated temperatures and
under vacuum conditions. In addition, microstructural characteristics of the resulting
films and the chemical evolution taking place during the dealloying process were investi-
gated and discussed. This work not only provides a novel way to synthesize nanoporous
stainless steel films, it also offers an in-depth understanding of the relation between the
microstructure of resultant films and the dealloying parameters.

2. Materials and Methods


The fabrication of nanoporous stainless steel involves two procedures, namely prepar-
ing the precursor stainless steel–Mg films through the technique of magnetron sputtering
and then vacuum thermal dealloying the obtained precursor films. The precursor films
were deposited using the facility of the magnetron sputtering system (ORION sputter-
ing system, AJA International Inc., Scituate, MA, USA). The high-purity magnesium and
316 stainless steel (99.95%) target materials were purchased from the Kurt J. Lesker com-
pany. The nominal compositions of the 316 stainless steel target materials are listed in
Table 1. Before deposition, the main chamber of the sputtering system was evacuated
to ~10−7 torr and filled with Ar gas until it reached a pressure of 2.5 mtorr. During the
preparation of precursor stainless steel–Mg films, the substrate was rotated to obtain a
uniform single-composition film. As measured by EDS, the composition of the prepared
Metals 2023, 13, 1255 3 of 9

precursor film was Fe 37.3%, Cr 9.5%, Ni 6.7%, Mo 1.4%, and Mn 1.1% (with the balance of
Mg), expressed in wt.%, which was roughly equal to (stainless steel)25 Mg75 , apart from the
trace elements of Si, P, C, and S, whose concentrations were not able to be quantified by
EDS. It is noted that the substrate utilized was a commercially purchased (100) Si wafer,
and a thin interlayer of Ta (~10 nm) was deposited to enhance adhesion between the film
and the substrate.

Table 1. Nominal compositions of 316 stainless steel target materials (wt. %).

Ni Cr Mo Si Mn C P S Fe
12 17 2.5 1 2 0.08 0.045 0.03 bal.

The vacuum thermal dealloying process was performed in the main chamber of the
sputtering system under high vacuum (less than 10−6 torr), and the heater inside the
chamber was a quartz lamp (AJA model SHQ-X). A total of 9 sets of precursor films were
dealloyed under the conditions of 450 ◦ C, 525 ◦ C, and 600 ◦ C for 0.5 h, 1 h, and 2 h,
respectively. During the vacuum thermal dealloying, the chamber was heated to a set
temperature, with a constant ramp rate of 10 ◦ C/min, and then held for about 20 min to
stabilize the temperature. As explained in a separate study by the authors [30], the sample
temperature was about 100 ◦ C lower than the set temperature of the quartz heater. Upon
the completion of dealloying, samples were cooled to room temperature inside the chamber.
Focused ion beam-scanning electron microscopy (FIB-SEM, Nanolab 660 and Helios
G4) was applied to study the morphology of the sample before and after dealloying. Chem-
ical composition was determined through the energy dispersive spectroscopy technique
(EDS, X-MaxN 80 mm2 detector). The TEM lamella was prepared using the FIB (Helios
G4) and was then investigated by transmission electron microscopy (FEI TITAN X). High-
resolution EDS mapping was acquired using the Bruker EDS detector attached to the TEM
system (FEI TITAN X).

3. Results and Discussion


For tuning the dealloying parameters, nine sets of precursor films were dealloyed at
temperatures of 450 ◦ C, 525 ◦ C, and 600 ◦ C for 0.5 h, 1 h, and 2 h, individually. Note that
the highest selected dealloying temperature was 600 ◦ C because at this temperature, the
corresponding vapor pressure of Mg was on the order of 1 torr, which was significantly
higher than the chamber pressure (less than 10−6 torr), to ensure that dealloying would
occur. Plan view micrographs of the resulting structures after vacuum thermal dealloying
are presented in Figure 1. For the sake of comparison, all of the SEM micrographs were
acquired at the same magnification. It appeared that the nanoporous structures exhibited
isolated islands and pores between ligaments under all of the given experimental conditions.
However, the resulting structures were clearly different in terms of the sizes of the pores
and islands, indicating that the dealloying process was temperature- and time-dependent.
At a temperature of 450 ◦ C, pores in the resultant structure after dealloying for 0.5 h
(Figure 1a) could barely be distinguished, implying a very low level of dealloying. Under
the same temperature, the nanoscale pores became much more distinct as the dealloying
time increased to 1 h and 2 h (Figure 1b,c). As the temperature rose to 525 ◦ C, the sizes of the
pores and islands became larger, as displayed by Figure 1d–f. One interesting phenomenon
was that partial pores and island surfaces were covered by a darker material exhibiting
contrast with the ligaments, especially for the sample that was dealloyed for 1h (Figure 1e),
which was probably the MgO. When the dealloying temperature increased to 600 ◦ C, the
surface of the dealloyed films became cleaner, and moreover, the pores and islands tended
to be more homogeneous, as exhibited in Figure 1g–i. Aside from the sample corresponding
to the dealloying time of 0.5 h, on whose surface a certain quantity of residual Mg was
found, the other two films, which were dealloyed for 1 h and 2 h individually, presented
debris-free surfaces, indicating that the dealloying process was nearly complete. It is also
worth noting that no cracks were observed on the planar surfaces of any of the dealloyed
Metals 2023, 13, x FOR PEER REVIEW 4 of 9

Metals 2023, 13, 1255 4 of 9

individually, presented debris-free surfaces, indicating that the dealloying process was
nearly complete. It is also worth noting that no cracks were observed on the planar sur-
films,of
faces suggesting
any of thethat the surface
dealloyed films,tensile stressthat
suggesting during the dealloying
the surface process
tensile stress was not
during the
excessive. Otherwise,
dealloying process was micro-cracks would
not excessive. likely appear.
Otherwise, micro-cracks would likely appear.

Figure 1. Plan view SEM micrographs of nanoporous stainless steel films after thermal dealloying
Figure 1. Plan view SEM micrographs of nanoporous stainless steel films after thermal dealloying
under various conditions. The conditions were (a) 450 °C, 0.5 h; (b) 450 °C, 1 h; (c) 450 °C, 2 h; (d)
◦ C, 0.5 h; (b) 450 ◦ C, 1 h; (c) 450 ◦ C, 2 h;
under
525 °C,various
0.5 h; (e)conditions.
525 °C, 1 h;The conditions
(f) 525 were
°C, 2 h; (g) 600(a)
°C,450
0.5 h; (h) 600 °C, 1 h; and (i) 600 °C, 2 h.
(d) 525 ◦ C, 0.5 h; (e) 525 ◦ C, 1 h; (f) 525 ◦ C, 2 h; (g) 600 ◦ C, 0.5 h; (h) 600 ◦ C, 1 h; and (i) 600 ◦ C, 2 h.
Cross-sections
Cross-sections of of the
the obtained
obtained structures
structures were were investigated
investigated using
using FIB-SEM,
FIB-SEM, as as shown
shown
in the micrographs displayed in Figure 2. The morphologies
in the micrographs displayed in Figure 2. The morphologies and sizes of the ligaments and sizes of the ligaments in
the resulting structures corresponding to various dealloying conditions
in the resulting structures corresponding to various dealloying conditions were different, were different,
implying
implying thatthat dealloying
dealloyingtemperature
temperatureand andtimetime together
together governed
governed thethe final
final microstruc-
microstructure.
ture. More specifically, for a dealloying time of 0.5 h ◦at 450 °C,

More specifically, for a dealloying time of 0.5 h at 450 C, 525 C, and 600 C, ligaments 525 °C, and
◦ 600 °C, liga-
of
ments
dealloyedof dealloyed
films were films wereand
smaller smaller
tightlyand tightly arranged,
arranged, and moreover, and moreover, partial liga-
partial ligaments had
ments
not yethadgrownnot across
yet grown across the
the sample sampleas
thickness, thickness,
shown inasFigure
shown2a,d,g.
in FigureAs the2a,d,g. As the
dealloying
dealloying time increased to 1 h, dealloyed samples exhibited
time increased to 1 h, dealloyed samples exhibited more sparsely distributed ligaments more sparsely distributed
ligaments
(Figure (Figure
2b,e,h), 2b,e,h), suggesting
suggesting that morethat Mgmorewas Mg was removed.
removed. It is worth It is worth
notingnoting that
that some
some ligaments in these three fabricated films were “abnormally”
ligaments in these three fabricated films were “abnormally” larger in size, which can larger in size, which
can probably
probably be attributed
be attributed to inhomogeneous
to inhomogeneous dealloying.
dealloying. WhenWhen the dealloying
the dealloying time in-
time increased
creased
to 2 h, thetoresulting
2 h, the specimens
resulting specimens
displayed moredisplayed more homogeneously
homogeneously nanoporousnanoporous
microstructures,mi-
crostructures,
especially for especially
the samplefor the sample
dealloyed 600 ◦ C (Figure
at dealloyed at 6002i).
°C (Figure 2i). The microstructure
The microstructure presented
presented
in Figure 2c in (450
Figure ◦ C,2c2 (450 °C, 2 h) exhibited
h) exhibited a more compacta more compact
morphology, morphology,
implyingimplying
that less that
Mg
less Mg was evaporated than the microstructures corresponding
was evaporated than the microstructures corresponding to dealloying temperatures to dealloying tempera-of
525 ◦ C
tures of and
525 °C600 and◦ C.
600 °C. Since
Since the films
the films were were constrained
constrained by by
thetheSiSisubstrate,
substrate,their
their lateral
dimensions
dimensions were werefixed,
fixed,andandtheythey could
could shrink
shrink onlyonly in thickness
in thickness (without
(without cracking).
cracking). As
As such,
such, it provided
it provided a good a good
way way to evaluate
to evaluate the extent
the extent of dealloying
of dealloying by measuring
by measuring thethe change
change in
in thickness
thickness before
before versus
versus after
after vacuum
vacuum thermal
thermal dealloying.
dealloying. WeWe firstfirst measured
measured the the thick-
thickness
ness
of theofprecursor
the precursor film film (Figure
(Figure 2j), which
2j), which waswas ± 2 nm,
173173±2 nm,as asindicated
indicatedby by the
the red
red line in
Figure 2k. The thicknesses of dealloyed films were determined using Image J, and are
shown in the histogram in Figure 2k. Each column represents the average thickness based
Metals 2023, 13, x FOR PEER REVIEW 5 of 9

Metals 2023, 13, 1255 5 of 9

Figure 2k. The thicknesses of dealloyed films were determined using Image J, and are
shown in the histogram in Figure 2k. Each column represents the average thickness based
on 10
on 10 measurements,
measurements, and and the
the Y
Y error
error bar
bar denotes
denotes standard
standard deviation.
deviation. This
This showed
showed that
that
the thickness change in the films dealloyed at a temperature of 450 ◦ C or 525 ◦ C was
the thickness change in the films dealloyed at a temperature of 450 °C or 525 °C was in-
insensitive
sensitive toto thedealloying
the dealloyingtime,
time,suggesting
suggestingthat
thatthese
thesetwo
twotemperatures
temperatures were,
were, perhaps,
perhaps,
not the optimum temperatures
not the optimum temperatures for driving the thermal dealloying process. However,
for driving the thermal dealloying process. However, at at aa
temperature of 600 ◦ C, the thickness variation in the resulting films was clearly a function
temperature of 600 °C, the thickness variation in the resulting films was clearly a function
of the
of the dealloying
dealloying time. With aa longer
time. With longer dealloying
dealloying time,
time, the
the resultant
resultant film
film became
became much
much
thinner, indicating
thinner, indicating aa higher
higher level
level of
of completion
completion for
for the
the dealloying
dealloying process.
process.

Figure 2. Cross-sectional micrographs of the precursor films after vacuum thermal dealloying under
Figure 2. Cross-sectional micrographs of the precursor films after vacuum thermal dealloying under
various conditions. The conditions were (a) 450 °C, 0.5 h; (b) 450 °C, 1 h; (c) 450 °C, 2 h; (d) 525 °C,
various conditions. The conditions were (a) 450 ◦ C, 0.5 h; (b) 450 ◦ C, 1 h; (c) 450 ◦ C, 2 h; (d) 525 ◦ C,
0.5 h; (e) 525 °C, 1 h; (f) 525 °C, 2 h; (g) 600 °C, 0.5 h; (h) 600 °C, 1 h; and (i) 600 °C, 2 h. (j) Cross-
section 525 ◦ C,
0.5 h; (e)image of1the
h; (f) 525 ◦ C, 2film
precursor h; (g)
and
◦ C, 0.5 h; (h) 600 ◦ C, 1 h; and (i) 600 ◦ C, 2 h. (j) Cross-section
600(k) the thickness change in films due to thermal dealloying.
image of the precursor film and (k) the thickness change in films due to thermal dealloying.
For nanoporous materials, their unique bi-continuous microstructure makes them
For nanoporous materials, their unique bi-continuous microstructure makes them
suitable for a wide range of applications. Normally, the ligament size is the most relevant
suitable for a wide range of applications. Normally, the ligament size is the most relevant
parameter, since it directly governs materials’ mechanical properties or other functionali-
parameter, since it directly governs materials’ mechanical properties or other functionalities.
ties. As discussed above, the degree of completion of the vacuum thermal dealloying pro-
As discussed above, the degree of completion of the vacuum thermal dealloying process is
cess is dependent upon both the dealloying temperature and time. This is, however, rea-
dependent upon both the dealloying temperature and time. This is, however, reasonable
sonable because the thermal dealloying process fundamentally relies on the surface diffu-
because the thermal dealloying process fundamentally relies on the surface diffusion of the
sion of the metal adatoms. Variation in the ligament size of the developed, nanoporous
metal adatoms. Variation in the ligament size of the developed, nanoporous microstructure
microstructure was determined quantitatively. For each dealloyed structure, no less than
was determined quantitatively. For each dealloyed structure, no less than 200 ligaments
200 ligaments were measured, and then the average value coupled with standard devia-
were measured, and then the average value coupled with standard deviation were calcu-
tion
lated,were calculated,inas
as presented presented
Figure 3. Asincan
Figure 3. Asthe
be seen, can be seen,
films the films
dealloyed dealloyed
at 450 at 450
◦ C possessed
°C possessed smaller ligament sizes than those in films dealloyed
◦ at 525 °C.
smaller ligament sizes than those in films dealloyed at 525 C. This was possibly due This was pos-
to
sibly due to their more tightly arranged structures (see Figure 2a,c), despite
their more tightly arranged structures (see Figure 2a,c), despite roughly equivalent degreesroughly
equivalent
of dealloyingdegrees of dealloying
completion at these completion at theseas
two temperatures, two temperatures,
evidenced as evidenced
by Figure by
2j. Note that
the case of “450 ◦ C, 1 h” was an exception, which could be attributed to the abnormally
13, x FOR PEER REVIEW 6 of 9
Metals 2023, 13, 1255 6 of 9

Figure 2j. Note that the case of “450 °C, 1 h” was an exception, which could be attributed
coarse ligaments
ligaments (Figure
(Figure 2b).
2b). In
In addition,
addition, the ◦
to the abnormally coarse the ligament
ligament size
size of
ofthe
thefilms
filmsdealloyed at 600 C

dealloyed at 600 °Cwas
wassmaller
smaller than
than that of the
that of the films
filmsdealloyed
dealloyedatat525525°C.C. This
This waswas probably because the
prob-
structure resulting from dealloying at 600 ◦ C was more uniform and without the abnormal
ably because the structure resulting from dealloying at 600 °C was more uniform and
coarse
without the abnormal ligaments.
coarse Furthermore,
ligaments. Furthermore, for for
each of the
each dealloying
of the dealloyingtemperatures,
tempera- the ligament size
became larger as the dealloying time increased.
tures, the ligament size became larger as the dealloying time increased.

Figure
Figure3.3.Variation
Variationin
inligament
ligamentsize
sizeof
ofthe
thedealloyed
dealloyedfilms.
films.

Conventionally, theConventionally,
precursor materials utilizedmaterials
the precursor for fabricating
utilizednanoporous metals
for fabricating nanoporous metals are
are binary solid solution
binaryalloys,
solid e.g., AuxAg
solution l-x. The
alloys, precursor
e.g., Aux Aglfilm in the
−x . The current study
precursor film inwas,the current study was,
however, sputteredhowever,
stainless sputtered
steel–magnesium,
stainless where the former component
steel–magnesium, contained
where the former component contained
multiple elements, multiple
as listed in Table 1. as
elements, It was,
listedtherefore,
in Table worth studying
1. It was, the element
therefore, worth mi- studying the element
migration/removal
gration/removal before and after vacuum before
thermal anddealloying,
after vacuum which thermal dealloying,
is crucial for under- which is crucial for
understanding
standing the mechanism the mechanism
of this dealloying technique. of this dealloying
As discussed technique.
above, the filmAsdeal-discussed above, the
film dealloyed at a2temperature of 600 ◦ C for 2 h exhibited the most ideal version of the
loyed at a temperature of 600 °C for h exhibited the most ideal version of the desired
desired
nanoporous structure. A TEM nanoporous
lamella was structure.
extracted A TEM
from lamella
this sample,was extracted
followed from this sample, followed by
by milling
millingthickness
to an electron-transparent to an electron-transparent
using the FIB-SEM. thickness
Afterwards using the FIB-SEM.
(S)TEM Afterwards (S)TEM and
and high-res-
olution EDS mapping high-resolution
were appliedEDS mapping
to study were applied
the morphology of to
thestudy the morphology
nanoporous structure of the nanoporous
structurecaused
and the chemical evolution and the bychemical
the dealloyingevolution caused
process. Theby the dealloying
results are summarizedprocess. The results are
and presented in Figure 4, in which the first two high-angle annular dark-field (HAADF) annular dark-field
summarized and presented in Figure 4, in which the first two high-angle
STEM micrographs(HAADF)
(Figure 4a,b)STEM micrographs
revealed that most (Figure
of the4a,b)
poresrevealed
extended thatacross
most ofthethe
sam-pores extended across
the sample thickness, while some isolated voids
ple thickness, while some isolated voids were also visible in the inner part of the sample.were also visible in the inner part of the
In addition, the nanostructure of the obtained nanoporous stainless steel film was rather stainless steel film
sample. In addition, the nanostructure of the obtained nanoporous
was ratherby
homogeneous, as evidenced homogeneous,
open porosityasand evidenced
ligaments by with
opencomparable
porosity andsizesligaments
[30]. with comparable
As suggested by Han et al. [21,22], the appearance of these internal isolated pores canthese
sizes [30]. As suggested by Han et al. [21,22], the appearance of be internal isolated
ascribed to the Kirkendall effect, which originates from differing diffusion rates between differing diffusion
pores can be ascribed to the Kirkendall effect, which originates from
rates between participating elements.
participating elements.
In addition
In addition to morphology to morphology
analysis, analysis,
high-resolution EDS high-resolution
mapping wasEDS mapping was performed
performed
with the aim of probing the chemical variation after dealloying, as exhibited in Figureexhibited
with the aim of probing the chemical variation after dealloying, as 4c– in Figure 4c–j.
Note that the elements C, P, and S were not mapped,
j. Note that the elements C, P, and S were not mapped, since the initial concentrations of since the initial concentrations of
these trace elements were less than 0.1 wt%. As shown by the EDS mapping results in
these trace elements were less than 0.1 wt%. As shown by the EDS mapping results in
Figure 4c–j, it appeared that the ligaments primarily consisted of Fe, Cr, Ni, Mn, and Mo,
Figure 4c–j, it appeared that the ligaments primarily consisted of Fe, Cr, Ni, Mn, and Mo,
which came from the stainless steel. Moreover, some residual Mg was also found in the
which came from the stainless steel. Moreover, some residual Mg was also found in the
pores (Figure 4d), indicating that the dealloying process, even for the case of 600 ◦ C + 2 h,
pores (Figure 4d), indicating that the dealloying process, even for the case of 600 °C + 2 h,
was not fully complete. By comparing the elemental maps of Fe, Mg, and O (Figure 4c–e),
was not fully complete. By comparing the elemental maps of Fe, Mg, and O (Figure 4c–e),
it was observed that oxygen coincided with Mg, suggesting that the chemical state of the
it was observed that oxygen coincided with Mg, suggesting that the chemical state of the
residual Mg was oxide (MgO). This can be attributed to the higher chemical affinity of
residual Mg was oxide (MgO). This can be attributed to the higher chemical affinity of
oxygen to magnesium versus iron. As a result, oxygen was preferentially alloyed with Mg
oxygen to magnesium versus iron. As a result, oxygen was preferentially alloyed with Mg
rather than Fe (or the other metallic elements in the stainless steel). Furthermore, Figure 4j
Metals 2023, 13, x FOR PEER REVIEW 7 of 9
Metals 2023, 13, 1255 7 of 9

rather than Fe (or the other metallic elements in the stainless steel). Furthermore, Figure
4j suggests
suggests thatthat
thethe Si the
Si in in the dealloyed
dealloyed film
film hadhad possibly
possibly disappeared,
disappeared, although
although this
this maymaybe
be artifact.
an an artifact.
OneOne possible
possible explanationfor
explanation forthis
thisobservation
observationisisthat
that the
the Si signal generated
generated
from the
from the substrate
substrate was much stronger than that emanating
emanating from
from the
the nanoporous
nanoporous stainless
stainless
steel
steel film,
film, resulting
resulting in
in almost
almost zero
zero Si
Si contrast
contrast in the film area.

Figure 4. STEM micrographs and high-resolution EDS mapping results of nanoporous films deal-
Figure 4. STEM micrographs and high-resolution EDS mapping results of nanoporous films dealloyed
loyed ◦at 600 °C for 2 h. (a) Low-magnification HAADF micrograph; (b) zoomed image correspond-
at 600 C for 2 h. (a) Low-magnification HAADF micrograph; (b) zoomed image corresponding to
ing to the red box in (a) and indicating the area for EDS mapping; (c–j) elemental maps of Fe, Mg,
the redNi,
O, Cr, boxMo,
in (a)
Mn,and indicating
and the area for EDS mapping; (c–j) elemental maps of Fe, Mg, O, Cr, Ni,
Si, respectively.
Mo, Mn, and Si, respectively.
4. Conclusions
4. Conclusions
Nanoporous stainless steel films were successfully fabricated by vacuum thermal
Nanoporous stainless steel films were successfully fabricated by vacuum thermal
dealloying of magnetron-sputtered stainless steel–magnesium precursor films. Various
dealloying of magnetron-sputtered stainless steel–magnesium precursor films. Various
structures, in terms of ligament size and morphology, were achieved under different deal-
structures, in terms of ligament size and morphology, were achieved under different
loying temperatures and times, which also indicates that the nanostructure of the deal-
dealloying temperatures and times, which also indicates that the nanostructure of the
loyed nanoporous stainless steel films can be tailored. Based on the analysis and discus-
dealloyed nanoporous stainless steel films can be tailored. Based on the analysis and
sion above,above,
discussion severalseveral
conclusions can becan
conclusions drawn.
be drawn.
Metals 2023, 13, 1255 8 of 9

(1) Crack-free nanoporous stainless steel films were successfully prepared, with deal-
loying temperatures ranging from 450 ◦ C to 600 ◦ C and dealloying times of 0.5 h to
2 h.
(2) The resulting structures were influenced both by the dealloying temperature and the
dealloying time. The ligament size increased with dealloying temperature for a given
dealloying time. Higher dealloying temperature generally resulted in the coarsest
ligaments, and deviations from this trend may reflect the existence of abnormally
thick ligaments within the overall distribution of ligament sizes for a given film.
(3) Magnesium was not completely removed, even in the dealloyed film corresponding
to “600 ◦ C + 2 h”, i.e., the film that experienced the highest degree of dealloying.
Additionally, the elemental maps of residual Mg and oxygen matched very well,
implying that residual magnesium exists as an oxide.

Author Contributions: X.L.: data curation, formal analysis, funding acquisition, investigation,
methodology, project administration, resources, supervision, validation, writing—original draft, and
writing—review and editing; X.Z.: formal analysis, investigation, and methodology; M.K.: formal
analysis and investigation; M.J.D.: formal analysis, investigation, and writing—review and editing;
T.J.B.: project administration, resources, supervision, and writing—review and editing. All authors
have read and agreed to the published version of the manuscript.
Funding: This research was funded by the Young Elite Scientist Sponsorship Program of the China As-
sociation for Science and Technology (No. YESS20210054) and the Independent Innovation Research
Fund of Huazhong University of Science and Technology (2172021XXJS010).
Data Availability Statement: The data presented in this study are available on request from the
corresponding author. The data are not publicly available due to privacy.
Acknowledgments: The authors acknowledge the support from the Electron Microscopy Center at
the University of Kentucky.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Huang, J.; Fang, X.; Liu, X.; Lu, S.; Li, S.; Yang, Z.; Feng, X. High-Linearity Hydrogen Peroxide Sensor Based on Nanoporous Gold
Electrode. J. Electrochem. Soc. 2019, 166, B814–B820. [CrossRef]
2. Wittstock, A.; Biener, J.; Bäumer, M. Nanoporous Gold: A New Material for Catalytic and Sensor Applications. Phys. Chem. Chem.
Phys. 2010, 12, 12919–12930. [CrossRef]
3. Wang, J.; Gao, H.; Sun, F.; Xu, C. Nanoporous PtAu Alloy as an Electrochemical Sensor for Glucose and Hydrogen Peroxide. Sens.
Actuators B Chem. 2014, 191, 612–618. [CrossRef]
4. Sun, Y.; Kucera, K.P.; Burger, S.A.; John Balk, T. Microstructure, Stability and Thermomechanical Behavior of Crack-Free Thin
Films of Nanoporous Gold. Scr. Mater. 2008, 58, 1018–1021. [CrossRef]
5. Jin, H.J.; Wang, X.L.; Parida, S.; Wang, K.; Seo, M.; Weissmüller, J. Nanoporous Au-Pt Alloys as Large Strain Electrochemical
Actuators. Nano Lett. 2010, 10, 187–194. [CrossRef]
6. Zielasek, V.; Jürgens, B.; Schulz, C.; Biener, J.; Biener, M.M.; Hamza, A.V.; Bäumer, M. Gold Catalysts: Nanoporous Gold Foams.
Angew. Chem.—Int. Ed. 2006, 45, 8241–8244. [CrossRef] [PubMed]
7. McCurry, D.A.; Kamundi, M.; Fayette, M.; Wafula, F.; Dimitrov, N. All Electrochemical Fabrication of a Platinized Nanoporous
Au Thin-Film Catalyst. ACS Appl. Mater. Interfaces 2011, 3, 4459–4468. [CrossRef]
8. Detisch, M.J.; Balk, T.J.; Bhattacharyya, D. Synthesis of Catalytic Nanoporous Metallic Thin Films on Polymer Membranes. Ind.
Eng. Chem. Res. 2018, 57, 4420–4429. [CrossRef]
9. Wan, Q.; Li, J.; Liu, Z.; Han, L.; Huang, S.; Wang, Z. Fabrication of Nanoporous Gold via an Improved Solid-Phase Method for
Non-Enzymatic Detection of Aniline. Metals 2023, 13, 754. [CrossRef]
10. Song, R.; Han, J.; Okugawa, M.; Belosludov, R.; Wada, T.; Jiang, J.; Wei, D.; Kudo, A.; Tian, Y.; Chen, M.; et al. Ultrafine
Nanoporous Intermetallic Catalysts by High-Temperature Liquid Metal Dealloying for Electrochemical Hydrogen Production.
Nat. Commun. 2022, 13, 1–12. [CrossRef]
11. Mokhtari, M.; Wada, T.; Le Bourlot, C.; Duchet-Rumeau, J.; Kato, H.; Maire, E.; Mary, N. Corrosion Resistance of Porous Ferritic
Stainless Steel Produced by Liquid Metal Dealloying of Incoloy 800. Corros. Sci. 2020, 166, 108468. [CrossRef]
12. Okulov, A.V.; Volegov, A.S.; Weissmüller, J.; Markmann, J.; Okulov, I.V. Dealloying-Based Metal-Polymer Composites for
Biomedical Applications. Scr. Mater. 2018, 146, 290–294. [CrossRef]
Metals 2023, 13, 1255 9 of 9

13. Joo, S.H.; Okulov, I.V.; Kato, H. Unusual Two-Step Dealloying Mechanism of Nanoporous TiVNbMoTa High-Entropy Alloy
during Liquid Metal Dealloying. J. Mater. Res. Technol. 2021, 14, 2945–2953. [CrossRef]
14. Eessaa, A.K.; El-shamy, A.M. Microelectronic Engineering Review on Fabrication, Characterization, and Applications of Porous
Anodic Aluminum Oxide Films with Tunable Pore Sizes for Emerging Technologies. Microelectron. Eng. 2023, 279, 112061.
[CrossRef]
15. Abbas, A.A.; Farrag, H.H.; El-Sawy, E.; Allam, N.K. Microbial Fuel Cells with Enhanced Bacterial Catalytic Activity and Stability
Using 3D Nanoporous Stainless Steel Anode. J. Clean. Prod. 2021, 285, 124816. [CrossRef]
16. Ramesh, R.; Niauzorau, S.; Sampath, V.K.; Wang, L.; Azeredo, B.P. In Situ Temperature Dependent Optical Characterization and
Modeling of Dealloyed Thin-Film Nanoporous Gold Absorbers. Adv. Opt. Mater. 2022, 10, 1–9. [CrossRef]
17. Sang, Q.; Hao, S.; Han, J.; Ding, Y. Dealloyed Nanoporous Materials for Electrochemical Energy Conversion and Storage. Energy
Chem. 2022, 4, 100069. [CrossRef]
18. Xia, Y.; Lu, Z.; Han, J.; Zhang, F.; Wei, D.; Watanabe, K.; Chen, M. Bulk Diffusion Regulated Nanopore Formation during Vapor
Phase Dealloying of a Zn-Cu Alloy. Acta Mater. 2022, 238, 118210. [CrossRef]
19. Sun, Y.; Burger, S.A.; Balk, T.J. Controlled Ligament Coarsening in Nanoporous Gold by Annealing in Vacuum versus Nitrogen.
Philos. Mag. 2014, 94, 1001–1011. [CrossRef]
20. Maxwell, T.L.; Balk, T.J. The Fabrication and Characterization of Bimodal Nanoporous Si with Retained Mg through Dealloying.
Adv. Eng. Mater. 2018, 20, 1–9. [CrossRef]
21. Lu, Z.; Li, C.; Han, J.; Zhang, F.; Liu, P.; Wang, H.; Wang, Z.; Cheng, C.; Chen, L.; Hirata, A.; et al. Three-Dimensional Bicontinuous
Nanoporous Materials by Vapor Phase Dealloying. Nat. Commun. 2018, 9, 1–7. [CrossRef]
22. Han, J.; Li, C.; Lu, Z.; Wang, H.; Wang, Z.; Watanabe, K.; Chen, M. Vapor Phase Dealloying: A Versatile Approach for Fabricating
3D Porous Materials. Acta Mater. 2019, 163, 161–172. [CrossRef]
23. Hsieh, S.R.; Lu, N.H.; Chen, C.H.; Lee, Y.L.; Cheng, I.C. Morphology, Ligament Strength, and Energy Absorption of Nanoporous
Copper via Vapor Phase Dealloying. Mater. Sci. Eng. A 2022, 857, 144131. [CrossRef]
24. Briot, N.J.; Balk, T.J. Developing Scaling Relations for the Yield Strength of Nanoporous Gold. Philos. Mag. 2015, 95, 2955–2973.
[CrossRef]
25. Wang, L.; Briot, N.J.; Swartzentruber, P.D.; Balk, T.J. Magnesium Alloy Precursor Thin Films for Efficient, Practical Fabrication of
Nanoporous Metals. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 2014, 45, 1–5. [CrossRef]
26. Zhao, C.; Wada, T.; De Andrade, V.; Williams, G.J.; Gelb, J.; Li, L.; Thieme, J.; Kato, H.; Chen-Wiegart, Y.C.K. Three-Dimensional
Morphological and Chemical Evolution of Nanoporous Stainless Steel by Liquid Metal Dealloying. ACS Appl. Mater. Interfaces
2017, 9, 34172–34184. [CrossRef]
27. Wada, T.; Geslin, P.A.; Kato, H. Preparation of Hierarchical Porous Metals by Two-Step Liquid Metal Dealloying. Scr. Mater. 2018,
142, 101–105. [CrossRef]
28. McCue, I.; Ryan, S.; Hemker, K.; Xu, X.; Li, N.; Chen, M.; Erlebacher, J. Size Effects in the Mechanical Properties of Bulk
Bicontinuous Ta/Cu Nanocomposites Made by Liquid Metal Dealloying. Adv. Eng. Mater. 2016, 18, 46–50. [CrossRef]
29. Guo, X.; Zhang, C.; Tian, Q.; Yu, D. Liquid Metals Dealloying as a General Approach for the Selective Extraction of Metals and
the Fabrication of Nanoporous Metals: A Review. Mater. Today Commun. 2021, 26, 102007. [CrossRef]
30. Kosmidou, M.; Detisch, M.J.; Maxwell, T.L.; Balk, T.J. Vacuum Thermal Dealloying of Magnesium-Based Alloys for Fabrication of
Nanoporous Refractory Metals. MRS Commun. 2019, 9, 144–149. [CrossRef]
31. Zhao, C.; Yu, L.C.; Kisslinger, K.; Clark, C.; Chung, C.C.; Li, R.; Fukuto, M.; Lu, M.; Bai, J.; Liu, X.; et al. Kinetics and Evolution of
Solid-State Metal Dealloying in Thin Films with Multimodal Analysis. Acta Mater. 2023, 242, 118433. [CrossRef]
32. Benčina, M.; Junkar, I.; Vesel, A.; Mozetič, M.; Iglič, A. Nanoporous Stainless Steel Materials for Body Implants—Review of
Synthesizing Procedures. Nanomaterials 2022, 12, 2924. [CrossRef] [PubMed]
33. Hassan, A.; Ali, G.; Park, Y.J.; Hussain, A.; Cho, S.O. Formation of a Self-Organized Nanoporous Structure with Open-Top
Morphology on 304L Austenitic Stainless Steel. Nanotechnology 2020, 31, 315603. [CrossRef] [PubMed]
34. Wang, Y.; Guo, R.; Zhou, X.; Hu, G. Experimental Investigation on Optimal Anodising Parameters of Nanopore Preparation
Process on the Stainless Steel Surface. Corros. Eng. Sci. Technol. 2020, 55, 513–519. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like