You are on page 1of 16

DOI: 10.1002/celc.

201901634 Articles

1
2
3 A Thermodynamic Approach for Selection of Anodizing
4
5
Electrolytes in Aluminium-Holmium System
6
7
Khurram Shahzad,[a] Cezarina Cela Mardare,[a, b] Andrei Ionut Mardare,*[a] and
8 Achim Walter Hassel[a, b]
9
10
11 In this paper, an approach is described for the electrolyte For the application of this approach, the most widely inves-
12 selection to grow optimally anodic films on thermally co- tigated valve metal, aluminium, and a non-valve metal rare
13 evaporated aluminium and holmium combinatorial alloys. This earth, holmium, have been chosen as a model case. Based on
14 method based on thermodynamic and electrochemical princi- self-constructed Eh-pH diagrams and solubility data, suitable
15 ples proves that electrolyte selection for anodizing of alloys electrolytes and pH range leading to the formation of barrier-
16 having different microstructures and electrochemical behaviour type anodic films on aluminium and holmium as well as on
17 of individual constituents can be decided by superimposing their alloys are proposed.
18 potential-pH diagrams of individual constituents of the alloy.
19
20
1. Introduction splitting, energy storage devices, biomedical, architecture,
21
nanotechnology, and as templates for the fabrication of nano-
22
One of the most challenging tasks in electrochemistry is the materials in the form of nanowires, nanopores, and
23
fabrication of new materials with a desired set of properties. In nanotubes.[3,9] In addition to the fabrication of new materials
24
particular, the surface modification for providing various using novel electrochemical approaches, the development of
25
functionalities to underlying metals and alloys via electro- new engineering alloys with peculiar chemistries (which can
26
chemical anodizing would require a sufficient understanding of serve as raw materials for electrochemical fabrication) is a time
27
electrolyte composition, pH range, and material stability under consuming and iterative process. One approach to accelerate
28
acidic and/or alkaline environments. It is now well known that the investigation of many new alloys chemistry at the same
29
surface layers in the form of oxides, hydroxides, and basic salts time is to use specimens containing a gradient in composition
30
formed by the interaction of various metals with water i. e. the combinatorial and computational approaches.[10–12] Addi-
31
represent the most important source of passivation in aqueous tionally, the final properties of the product, for engineering
32
electrolytes.[1] Electrochemical anodizing has many advantages alloys not only depend on the parent alloy chemistry but also
33
such as providing a cost-effective synthesis of oxide film, ease on the synthesis methods used to develop a unique set of
34
of fabrication and effective usage of underlying metal as an properties and microstructures.[13]
35
anode material when compared with expensive synthesis It is now a well-known fact that the stability of metals and
36
methods of amorphous oxide such as chemical vapour alloys against dissolution in aqueous electrolytes depends
37
deposition (CVD), molecular beam epitaxy (MBE), and atomic primarily on the pH value which often plays a critical role in the
38
layer deposition (ALD) which demand high fabrication resistance of metals against corrosion processes. In a certain
39
temperatures.[2] Moreover, the discovery of self-organized nano- pH-range, either the metal dissolves or protective films may be
40
porous oxide structure by H. Masuda et al extended the formed. In other words, some metals dissolve easily under
41
applications of anodizing process in aerospace, military, water alkaline conditions, while some metals are susceptible to acidic
42
dissolution[14] or even doing both, thus being amphoteric. The
43
[a] Dr. K. Shahzad, Assist. Prof. Dr. C. C. Mardare, Assoc. Prof. Dr. A. I. Mardare, situation becomes more complicated during the anodizing of
44 Prof. Dr. A. W. Hassel alloys having constituents with different electrochemical sus-
45 Institute for Chemical Technology of Inorganic Materials,
Johannes Kepler University Linz,
ceptibility to dissolve in a particular electrolyte.[15] It is, therefore,
46
Altenberger Str. 69, 4040 Linz, Austria crucial to know the dissolution susceptibility of alloys as well as
47 Fax.: + 43 732 2468 8702 their individual constituents in a certain pH range. Considering
48 E-mail: andrei.mardare@jku.at
[b] Assist. Prof. Dr. C. C. Mardare, Prof. Dr. A. W. Hassel
the widely varying surface morphology, structure, and composi-
49
Christian Doppler Laboratory for Combinatorial Oxide Chemistry (COM- tion of individual elements in an alloy, it is an extremely tedious
50 BOX), and time-consuming task to find a common pH range and
51 Institute for Chemical Technology of Inorganic Materials,
Johannes Kepler University Linz,
electrolyte where individual constituents of the alloy are
52
Altenberger Str. 69, 4040 Linz, Austria thermodynamically stable.[15] This occurs particularly, when an
53
© 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA. This amorphous structure is induced into the parent alloy owing to
54 is an open access article under the terms of the Creative Commons the lattice mismatch as a result of the large difference in the
55 Attribution Non-Commercial NoDerivs License, which permits use and
distribution in any medium, provided the original work is properly cited, the atomic radii of individual constituents. Unfortunately, the Eh-pH
56
use is non-commercial and no modifications or adaptations are made. or Pourbaix diagrams (where Eh describes the potential against
57

ChemElectroChem 2020, 7, 1342 – 1357 1342 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles
SHE) provides thermodynamic information only for pure metal- Experimental Section
1
H2O systems but not for alloys, although superimposing Eh-pH
2 Metallic thin films were deposited by thermal evaporation on
diagrams approach has been successfully used in directionally borosilicate float glass substrates (VWR International GmbH,
3
solidified alloys for a selective dissolution of phases.[16] As an Germany) with dimensions of 2.6 · 7.6 cm2. A substrate cleaning
4
example, the rare-earth element holmium (Ho) reveals the procedure based on sequential ultrasonication in acetone, isopro-
5 panol and water was applied before each deposition. The
highest stability between pH 8.0–12.0, while the stability
6 deposition system with a base pressure of 1 · 10 5 Pa was designed
window for aluminium (Al) oxidation terminates at around
7 and self-developed for thin film combinatorial libraries. For this
pH 9.0. As a result, identification of a compromise pH range and purpose, two independent W thermal boats containing either pure
8
hence the electrolyte selection, which allows anodic polar- Al (99.95 %, Goodfellow) or Ho, (99.99 % Smart Elements) were
9
ization of alloys having elements of different electrochemical placed eccentrically in relation to the centre axis normal to the
10
behaviour is still very challenging. sample. This allowed the fabrication of an Al Ho compositional
11 gradient along the substrate. The compositional gradient control is
Therefore, the aim of the present work is to introduce an
12 obtained by modifying the individual deposition rates which are
approach that will allow identifying the pH range and confer
13 monitored in-situ by quartz crystal microbalances (Inficon). Deposi-
the choice of electrolyte for alloys anodizing rather than tion rates of 0.80 and 0.18 nm s 1 were used for Al and Ho,
14
performing hundreds of tedious attempts. For this purpose, we respectively. A final film thickness of 500 nm was obtained for the
15
superimpose the Eh-pH diagrams of individual elements in Al Ho library. Additionally, pure Al and Ho films were deposited in
16 identical conditions. Pure holmium is generally considered to be of
order to find the common pH range of oxidation for alloys. For
17 low toxicity and does not react at normal temperatures and
the applicability of this approach, two metals of different
18 depositing very minute amount of Ho under high vacuum and at
chemical properties were chosen namely the widely explored
19 ambient temperature would further reduce any health hazardous
valve metal, Al and rarely explored non-valve metal, Ho due to effect. The ultra-high vacuum would also prevent formation of
20
its numerous promising applications.[17–23] Here, it is important oxides of Al or Ho on depositing film surface, although oxides of
21
to differentiate the terms “valve and non-valve metals”. both Al and Ho are non-toxic in nature.
22
Historically, the term “valve metal” was applied to those metals Scanning energy-dispersive X-ray (SEDX) spectroscopy was used for
23
which were used for making vacuum valves in early days of compositionally mapping the Al Ho thin film library. The quantita-
24
electron tube industry and were commonly known as elec- tive analysis of different alloys located at discreet positions along
25 the sample was performed using IDFix software (remX GmbH).
tronics or Vacuum tubes.[24] To make an effective cathode in
26 Characteristic X-rays resulting from surface irradiation with 20 keV
these tubes, a combination of conducting metal and a metal
27 electrons in a spot size of 500 μm were analyzed by a Si drift
oxide substance good for electron emitter (thermionic emis-
28 detector (SDD, remX GmbH). As a result, the SEDX mapping
sion) was needed. In other words, certain metals were found to precisely provides the location of individual Al Ho alloys along the
29
have a special property of forming a thin film of their oxide to entire library. A total compositional spread of 10 at.% was obtained.
30
significantly increase the rate of electron emission. However, The Ho amount varied between 2 and 12 at.% along the library.
31 Complementary, the Al concentration varied between 98 and
specifically for electrochemical anodizing, Güntherschulze and
32 88 at.%. This describes a compositional resolution of
Betz[25] categorized many metals into complete valve effect (Al,
33 0.13 at.% mm 1. If (typical) 1 at.% precision is desired for Al Ho
Ta, Bi, and Sb) and those with incomplete valve effect (Si, Ag, alloy identification along the library, then a virtual stripe more than
34
Mg, Cd, Fe, Sn, Zr, W, and Zn). L. young[26] further classified the 7 mm long may be defined as hosting a single alloy. This rather
35
various metals into the “most typical members” such as Bi, Al, large surface is quite convenient for various follow-up measure-
36
Hf, Nb, W, Zr, Sb, and Ta and “other elements” including Ti, Sn, ments in order to characterize compositionally induced changes in
37 Al Ho alloys. The phases in the evaporated films were identified
V, Ge, Be, Mg, Si. Moreover, depending on the electrolyte used,
38 using X-ray diffraction (XRD) in θ-2θ (Bragg-Brentano) and ω-2θ
a certain metal may exhibit complete or incomplete valve
39 with ω = 3° (grazing incidence) geometries. CuKα radiation was
effect. As the list of valve metals is uncertain, therefore, in the employed for these measurements. Additionally, surface morpholo-
40
present manuscript, the term “valve metals” is reserved for gies of Al Ho alloy thin films before and after anodizing were
41
“most typical member” i. e. Al as reported by L. young[26] and examined using scanning electron microscopy (SEM - FEI Philips
42 XL30 ESEM FEG).
“non-valve metal” is used for metals excluding both “most
43
typical member” and “other elements” such as Ho in the The electrochemical measurements on evaporated films were
44
present case. carried out using a three-electrode scanning droplet cell micro-
45
In addition to fundamental aspects, oxides of rare-earths in scope (SDCM) in contact mode.[31] The SDCM contains gold as
46 counter and Ag/AgCl/3 M KCL as μ-reference electrodes, respec-
conjunction with Al are also considered as promising candi-
47 tively while the glass substrate coated with the evaporated film
dates in electronic industry owing to their good thermal and
48 served as working electrode. The contact area of SDCM was
chemical stability, high permittivity, and excellent insulating carefully examined under scanning electron microscope and a
49
and dielectric properties such as low dielectric losses, relatively value of 0.0346 cm2 was measured. Potentiodynamic anodizing was
50
high dielectric breakdown field strength as well as high performed to 10 V at a sweep rate of 100 mV s 1 in a 0.2 M
51
dielectric constant.[27–30] Therefore, such approach may also phosphate buffer (pH 8.3). The buffer solution was obtained by
52 mixing anhydrous potassium phosphate monobasic and sodium
provide the suitable electrolyte selection for the formation of
53 phosphate dibasic heptahydrate. The electrochemical impedance
anodic films on other alloy systems and broaden the existing
54 spectroscopy performed using a Compact Stat electrochemical
applications domain of rare-earths oxides as well. interface system (IVIUM Technology) with an AC perturbation of
55
50 mV in a frequency range of 105 to 10 1 Hz. The thermodynamic
56
data and electrochemical reactions used in the present study were
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1343 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles

1 Table 1. Literature survey of chemical and electrochemical reactions and standard Gibb’s free energy of formation for Al and Ho systems.[32–35]
2 Chemical and electrochemical reactions ΔG (kJ mol 1) Species o
ΔG (kJ mol 1)
3 f
4 H2O
5 2H2O(l)!4H + + 4e + O2 (g) 474.27 H2O 237.37
6
Ho H2O system
7 Ho3 + + 3e !Ho 674.2 Ho3 + 674.2
8 Ho3 + + 3H2O!3H + + Ho(OH)3 (s) 87.87 Ho(OH)3 (s) 1298.44
9 Ho3 + + 4H2O!4H + + Ho(OH)4 191.3 Ho(OH)4 1432.36
2Ho3 + + 3H2O!Ho2O3 + 6H + 268.91 Ho2O3 1791.6
10 2Ho3 + + 3H2O!Ho2O3 + 6H + + 6e 1091.49
11 Ho H2O PO43 system
12 Ho3 + + 2PO43 !Ho(PO4)23 121.53 Ho(PO4)23 2845.2
2H + + Ho3 + + PO43 !HoH2PO42 + 127.53 HoH2PO42 + 1826.5
13 Ho3 + + PO43 !HoPO4(s) 143.04 HoPO4 (s) 1835.04
14 Al H2O system
15 3e + Al3 + !Al(s) 481.0 Al3 + 481.0
Al3 + + 3H2O!3H + + Al(OH)3(s) 59.23 Al(OH)3(s) 1137.6
16 Al3 + + 4H2O!4H + + Al(OH)4 129.52 Al(OH)4 310.92
17 2Al3 + + 3H2O!Al2O3 + 6H + 91.81 Al2O3 1582.3
18 Al H2O cit3 PO43 system
AlHPO4 + !Al3 + + H + + PO43 49.576 AlHPO4 + 1555.34
19 Al3 + + cit3 !Al(cit) 57.401 Al(cit) 1700.20
20 Al3 + + cit3 + H + !Al(Hcit) + 73.61 Al(Hcit) + 1716.41
21 Al3 + + cit3 !H + + Al(cit)OH 37.24 Al(cit)OH 1680.04
2Al3 + + 2H2O + 2cit3 !2H + + Al2(cit)2(OH)22 94.50 Al2(cit)2(OH)22 3854.84
22 Ho H2O 0.1 M cit3 0.2 M PO43 system
23 Cit3 + Ho3 + !Ho(cit) 53.54 Ho(cit) 1889.58
24 2H + + Ho3 + + PO43 !HoH2PO42 + 127.55 HoH2PO42 + 1826.50
2PO32 + Ho3 + !Ho(PO4)23 121.53 Ho(PO4)23 2845.20
25 cit3 -H2O system
26 cit3 (aq) 1161.80
3 2
27
+
H + cit !H(cit) 36.50 H(cit)2 1198.30
2H + + cit3 !H2(cit) 63.66 H2(cit) 1225.46
28 PO43 H2O system
29 P (cr) + 4H2O!8H + + 5e + PO43 75.26 PO43 1024.74
30 H + + PO43 !HPO42 70.44 HPO42 509.020
2H + + PO43 !H2PO42 111.57 H2PO42 550.180
31 B(OH)3 H2O system
32 2B(OH)3!B2O3 (cr) + 3H2O 32.78 B2O3 (cr) 1194.79
33 H3BO3 (s)!B(OH)3 0.40 H3BO3 (s) 969.440
B (cr) + 3H2O!3H + + 3e + B(OH)3 257.73 B(OH)3 969.840
34 3B(OH)3 ! H + + B3O3(OH)4 40.112 B3O3(OH)4 2869.40
35
36
37 taken from the Geochemist’s workbench software[32–33] and Medusa Table 2. Summary of different electrolytes used for Al-3.0 at. % Ho
38 software[34–35] database and Eh-pH diagrams were calculated at corrosion studies.
39 298 K and atmospheric pressure for dissolved Al and Ho concen- Electrolytes pH
40 trations per litre of water. These details are summarized in Table 1.
The electrochemical protocol followed for corrosion experiments is Borate (BA NaOH) buffer 9.0
41 (0.1 M boric acid + 0.03 sodium hydroxide)
as follows: (i) open circuit potential (OCP) measurements for 60 s,
42 0.2 M Na2HPO4 (non-buffered) 8.8
(ii) potentiodynamic anodizing up to 10 V with a sweep rate of 0.2 M Phosphate buffer 8.3
43 100 mV s 1, (iii) measurements of polarization resistance (Rp) and (Potassium phosphate + disodium Phosphate)
44 corrosion rate (Crate) of the resultant anodic films in various Citrate-Phosphate buffer 7.6
45 electrolytes (indicated in Table 2), and (iv) repetition of steps (i-iii) (0.1 M citric acid monohydrate +
46 after each 150 s for a maximum period of 3600 s. 0.2 M disodium phosphate)
0.2 M Phosphate buffer 7.3
47 (Potassium phosphate + disodium Phosphate)
48 Citrate buffer (0.1 M citric acid monohydrate + 6.0
0.1 M trisodium citrate dihydrate)
49
2. Results and discussion
50
51
2.1 Eh-pH and Solubility Diagrams for Al and Ho
52
experimentally determined Gibbs free energy of formation for
53
The formation of species as a function of pH and potential is various species, therefore, literature data shows several different
54
well documented in literature.[1] Partial substitution of hydroxyl Pourbaix diagrams for metal oxide and hydroxide due to
55
groups for waters of hydration results in modification of the different free energies of formation.
56
solution species. As these diagrams are constructed from
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1344 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles
In order to find a common pH range for alloy oxidation, the ( 2.32 V vs SHE) compared with Al ( 1.66 V vs SHE) and unlike
1
Eh-pH diagrams of Al and Ho were simulated at 4 different aluminium,[1] holmium reveals different domains of passivation
2
concentrations of dissolved Al and Ho at 298 K and 1013 hPa. and corrosion. Looking into the Eh-pH diagram in Figure 1a, it is
3
All used chemical and electrochemical reactions together with clear that both Al and Ho at 1 mol L 1 concentration present
4
standard Gibbs free energy of formation for Al and Ho are large regions of oxide/hydroxide formation, where anodic
5
summarized in Table 1. As a direct result of the simulations, oxidation of both Al, Ho, and their alloys can theoretically be
6
Figure 1 shows the superimposed Eh-pH diagrams at 100, 10 3, feasible. Their common domain of passivation gradually shrinks
7
10 6, and 10 9 mol L 1 H2O concentrations. Using the formulae with the reduction in concentration of dissolved Al and Ho.
8
given in Table 1, the Eh-pH diagrams of Al and Ho are Starting from a pH range of 5.2 to 14 for 1 mol L 1, the common
9
superimposed on each other considering the Al(OH)3 and Ho passivation domain shrinks to pH 6.1 11.6 for 10 3 mol L 1
10
(OH)3 as stable forms. These equilibrium diagrams are only valid (Figure 1b) reaching a narrow pH range of 7.2–8.6 (Figure 1c)
11
in the absence of substances with which Al and Ho can form before ultimately vanishing for 10 9 mol L 1 (Figure 1d). For
12
complexes or salts. Ho has a large driving force for oxidation obvious reasons, the optimum choice of electrolyte selection
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55 Figure 1. Superimposed Eh-pH diagrams at different concentrations of dissolved Ho and Al in water at 298 K. The shaded yellow region indicates stability
domain where oxidation of both Ho and Al is thermodynamically possible. These diagrams are constructed for molar concentrations of dissolved Ho and Al
56 with exponents of 0, 3, 6 and 9.
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1345 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles
should be within overlapping pH passivation ranges of both Al
1
and Ho. It is also worth mentioning here that the amount of
2
dissolved Al and Ho is very critical for oxide formation. Too high
3
concentration of dissolved species may reduce the efficiency of
4
resulting film owing to the consumption of a large amount of
5
Al and Ho, since growth efficiency is directly associated with
6
the number of cations within the film (Figure 1a). At the same
7
time, a too small concentration of dissolved species may
8
enlarge the corrosion region without formation of oxide, as can
9
be seen in Figure 1d. Therefore, in the present study a relatively
10
low concentration of 10 6 mol L 1 was considered for electrolyte
11
selection in order to grow the anodic film at high efficiency.
12
In order to get further insight of oxidation stability regions,
13
solubility diagrams of Al and Ho are also simulated in aqueous
14
electrolytes and these results are summarized in Figure 2. In
15
agreement with the Eh-pH diagrams from Figure 1, it is clear
16
that the solubility of both Al and Ho is highly dependent on pH.
17
Moreover, both elements share different pH ranges depending
18
on the concentration of dissolved Al and Ho, where the
19
solubility diagrams display an optimum (minimum) overlap, as
20
indicated by the greyed zones in Figure 2. The identified pH
21
region coincides with the common passivation domains
22
previously described during analysis of Eh-pH diagrams from
23
Figure 1. Due to the common nature of both calculations, the
24
same narrowing of the optimum solubility overlap region is
25
observable also here with decreasing concentration of both
26
species. Consequently, based on Eh-pH and solubility diagrams,
27
an aqueous solution with a pH value around 8.0 was suggested
28
(see Figure 2c). Therefore, 0.2 M phosphate buffer rendering a
29
pH 8.3 solution was selected for testing the validity of our
30
present approach. Additionally, the choice of phosphate-
31
containing electrolytes is reasonable considering the previous
32
reports which show that rare earths chlorides, perchlorates,
33
carbonates, and nitrates generally exhibit high solubility, where-
34
as rare-earth containing phosphates are highly insoluble in
35
aqueous electrolytes.[36–37] Corrosion studies in the final part of
36
present study further elucidate the choice of electrolyte and pH
37
range based on our present approach.
38
After deciding upon the pH range and electrolyte selection,
39
it was also necessary to further investigate the electrochemical
40
response of both Al and Ho with and without the presence of
41
selected phosphate electrolytes. For this purpose, Pourbaix
42
diagrams of Al and Ho are also simulated in various oxide-
43
forming substances such as citrate, borate, or citrate-phosphate
44
mixture at different concentrations as can be seen in Figures 3
45
and 4. From a thermodynamic point of view, the presence of
46
phosphate and borate do not affect the Pourbaix diagram of Al.
47
This can be seen by viewing the corresponding Eh-pH diagram
48
(Figure 3a) and electrochemical reactions in the Al PO43 H2O,
49 Figure 2. Solubility diagrams for Ho and Al at three different concentrations
PO43 H2O, and B(OH)3 H2O systems (Table 1) neglecting all
50 of dissolved Ho and Al per litre of water at 298 K and at 1013 hPa pressure.
kinetic aspects. Therefore, it is likely that the Pourbaix diagram
51
of Al PO43 H2O or Al B(OH)3 H2O systems will resemble in
52
principle the Al H2O system. The obvious reason for this is the
53
formation of soluble complexes AlHPO4 +, Al2PO43 +, B3O3(OH)4 phosphate addition not only expands the passivation region by
54
with a large free energy change of 1555.3, 2003.8, and the formation HoPO4 but also inhibits the migration of Ho3 +
55
2869.40 kJ mol 1respectively. On the other hand, the Eh pH ions in the low pH range. At the same time, a reduced
56
diagram of Ho PO43 H2O system (Figure 3b) shows that concentration of dissolved Ho shrinks the stability field of Ho
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1346 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
Figure 3. Eh-pH diagrams for (a) Al in Al-H2O-0.2 M PO43 and Al-H2O-0.1 M B(OH)3, (b) Ho in Ho-H2O-0.2 M PO43 . These diagrams are constructed at molar
40 concentrations of dissolved Ho and Al with exponents of 0, 3, 6 as indicated by red, green, and blue lines respectively. For simplification, the species at
41 1 M concentrations are highlighted as red bold. (c, d) superimposed Eh-pH diagrams for Al and PO43 at molar concentration per litre of water with exponent
42 6.
43
44
45
(OH)3 due to the formation of more stable HoPO4 (for HoPO4, As mentioned above, presence of phosphate species does
46
o not affect the Eh-pH diagrams of Al, therefore, the addition of
47 ΔG = 1835.04 kJ mol 1) compared with Ho(OH)3 (for Ho
f citrate either alone or in combination with phosphate induces
48
o similar effect for Al and Ho as in the presence of citrate free
49 (OH)3, ΔG = 1298.44 kJ mol 1). Thus, it is expected that
f phosphate electrolyte except the gradual formation of Al(cit)
50
anodizing both Al and Ho to be favourable in an aqueous and Ho(cit) and narrowing of Al(OH)3 and Ho(OH)3 stability
51
electrolyte containing phosphate buffer having a pH of 8.3. The domains (see Figures 3a, 3 b, 4 a, 4 b, 4 d). This occurs due to the
52
influence of phosphate on Eh-pH diagrams of Al was also more negative free energy of Al(cit) and Ho(cit) compared with
53
confirmed by superimposing individual Pourbaix diagrams of their oxides/hydroxides counterparts (Table 1). However, the
54
both Al and P (Figure 3c) and from recently introduced first formation of Ho(cit) and Al(cit) occurs either below water
55
principle DFT method (Figure 3d).[38] stability window or anodizing voltage (< 0 V for Ho) or below
56
pH 4 (for Al) and may not hinder the growth of anodic films
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1347 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40 Figure 4. Eh-pH diagrams for (a) Al in Al-H2O-0.2 M PO43 0.1 M cit3 , (b) Ho in Ho-H2O-0.2 M PO43 0.1 M cit3 , (c) Ho in Ho-H2O-0.1 M B(OH)3 buffer, (d) Ho
in Ho-H2O-0.1 M cit3 .. These diagrams are constructed at molar concentrations of dissolved Ho and Al with exponents of 0, 3, 6 as indicated by red, green,
41 and blue lines respectively. For simplification, the species at 1 M concentrations are highlighted as red bold.
42
43
44
between pH 7.0–9.0 at 0–10 V. Formation of strong soluble of Ho(OH)3 as can be seen from the Ho-H2O system by the
45
complexes like Al(Hcit) + and Al(cit)OH may affect the Eh-pH addition of B(OH)3 in Figure 4c. Therefore, thermodynamically,
46
diagram of Al (Figure 4a), however, when superimposing with electrolytes containing either borate buffer (pH 9.0) or
47
holmium in phosphate-citrate containing electrolyte, this effect ammonium pentaborate tetrahydrate (pH 7.6) can be treated in
48
might be wiped out owing to formation of more stable Ho(cit) a similar manner for predicting the oxidation stability of Ho.
49
and HoPO4 (Table 1). Lastly, the stability of H3BO3 and B(OH)3 in Though, we must keep in mind that, the formation of soluble
50
aqueous electrolyte is largely dependent on pH of the electro- complexes of various nature in the presence of orthoboric acid
51
lyte and the solubility of H3BO3 and B(OH)3 increases with pH and ammonium pentaborate may significantly influence the
52
increase, resulting in the formation of soluble complexes such electrolyte pH as well as the kinetics of the electrochemical
53
as B3O3(OH)4 in the Ho(OH)3 stability field. Due to relatively system. Following above discussion, superimposing Al and Ho
54
stable rare-earth oxides/hydroxides compared with B(OH)3/B2O3 Eh-pH diagrams (Figures 4a and 4 d) in the presence of citrate
55
(Table 1), it is likely that electrolytes containing either borate or only will promote soluble citrate complexes containing both Al
56
ammonium pentaborate do not influence the stability domain and Ho and would also result in significant reduction of growth
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1348 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles
efficiency due to large consumption of cations in the form of Figure 5a. For investigations of crystal phases along the Al Ho
1
soluble complexes, while the combining Eh-pH diagrams for Al compositional spread, both θ-2θ and ω-2θ modes were
2
and Ho in the presence B(OH)3 (Figures 3a and 4 c) would not employed. As can be seen in Figure 5b, typical Al fcc phase is
3
grow film of high efficiency owing to similar thermodynamic evidenced at high Al content, e. g. 3 at.% of Ho. Increasing the
4
behavior as in simple water. On the other hand, mixing Eh-pH Ho amount in the library, the Al-7 at.% Ho alloy, in the middle
5
diagrams of Al and Ho in the presence of phosphate alone of the compositional range, displays an amorphous or mixed
6
(Figures 3a and 3 b) or with the combination of phosphate and structure of Al and intermetallic HoAl3. Finally, at high Ho
7
citrate (Figures 4a and 4 b) would produce anodic films with concentrations the Al-11 at.% Ho alloy shows the appearance of
8
maximum efficiency. Phosphate and citrate are also considered a HoAl3-rich phase. The crystal structure of this phase is
9
among many electrolytes due to their large anion sizes which presented in Figure 6 where the atomic environment of Al and
10
may also help in growing relatively anion free oxide films. Ho species can be observed. Both Al and Ho atoms are indexed
11
with numerals from 1 to 3 for easier identification, each having
12
a coordination number of 12. The existence of Al-rich, HoAl3-
13
2.2 Crystallography and Microstructure of Al, Ho and their rich, mixed structures, and evolution of phases together with
14
Alloys the change in Ho concentration was reported in a separate
15
study.[39] The presence of mixed structures was previously
16
Prior to electrochemical studies, crystallographic diffraction reported and is fully consistent with the available literature
17
patterns and the microstructure of Al, Ho, and selected Al Ho data.[40–43]
18
alloys were examined by means of XRD and SEM. The obtained Surface features of Al, Ho and selected Al Ho alloys before
19
crystallographic data are summarized in Figure 5. Figure 5a and after anodizing were examined by means of scanning
20
represents the diffractograms corresponding to as-evaporated electron microscopy. A display of selected surfaces is demon-
21
Al and Ho thin films (~ 500 nm thick). A typical fcc structure, strated in Figure 7. The surface features analysis shows the
22
with a strong main (111) peak was identified for Al. Even significantly different morphologies as the composition changes
23
though the presence of (200) and (220) peaks can still be from pure Al, along the Al Ho library to pure Ho. The Al surface
24
observable, their intensity is almost negligible as compared to demonstrates a typical morphology for evaporated films. Larger
25
the main (111) peak. This indicates a high-quality film and such grains (~ 150 nm) are dispersed on top of an underlying
26
texturing is a direct consequence of the evaporation conditions, compact and smooth film with ~ 70 nm grains. The Ho surface
27
common for Al condensation on glass. Pure Ho film was appears as a uniform fine-grained morphology with an average
28
identified as hexagonal α-Ho with an hcp structure. A grain size of ~ 34 nm. The further microstructural examination
29
polycrystalline film is observable with the peak (011) being the of Al Ho alloys disclosed a remarkable change in morphology
30
most intense one. Several other peaks are also identifiable in with the increase in Ho concentration. At 3 at.% Ho, the
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55 Figure 5. (a) XRD patterns of thermally evaporated Ho and Al thin films indicating hcp and fcc phases respectively (b) XRD scans obtained at three selected
positions along the Al Ho alloys library. The appearance of different patterns with compositions corresponds well with SEM examination of the same
56 compositions.
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1349 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 Figure 6. Cell projection of crystal structure of HoAl3 and atomic environ-
22 ment of Al and Ho indicating the coordination number of 12 for Al1, Al2,
23 Al3, Ho1, Ho2, and Ho3.
24
25
26
morphology resembles that of Al, with grains that start to
27
elongate with the addition of Ho. Distinct elongated and much
28
smaller grains can be visible at 7 and 11 at.% Ho when the
29
entire surface morphology transforms into a new surface. The
30
gradual increase of the HoAl3 amount may be responsible for
31
the significant change in surface features with Ho addition. The
32
morphological change agrees well with structural analysis
33
indicating a crystal structure distortion induced by the differ-
34
ences between the cubic Al and hexagonal HoAl3. A smoothen-
35 Figure 7. Surface morphologies of Al, Ho and Al Ho alloys (a) before and (b)
ing of the surface cannot be readily observable after anodizing
36 after anodizing at selected atomic concentrations of Ho. The Ho concen-
due to the limited thickness of the anodic films (~ 10–15 nm) tration in the alloy is given at the upper left corner of each figure.
37
indicating that surface morphology is controlled by the texture
38
of the original parent surface.[44,45] Nevertheless, low magnifica-
39
tion SEM images (not shown here) clearly reveal the evidence of
40
anodic film formation irrespective of the Al Ho alloy composi- response is readily observable on Al and Al-3.0 at.% Ho, where
41
tion and anodizing metals, which supports the idea of electro- the film structure and surface features have a resemblance to
42
lyte selection based on thermodynamic considerations. Al. However, a non-steady state current behaviour can be seen
43
in pure Ho, Al-7 at. % Ho, and Al-11 at.% Ho. In the case of pure
44
Al and Al-3 at. % Ho alloy, an initial increase in current is
45
2.3 Potentiodynamic Anodizing observed, which eventually falls to a constant current plateau.
46
This is immediately observable in Figure 8a for pure Al, when a
47
After electrolyte selection based on thermodynamic consider- single cyclic voltammogram was recorded up to 10 V. The
48
ations, electrochemical studies were performed on Al, Ho, and plateau is also identified in the series of voltammograms from
49
three selected Al Ho alloys as shown in Figure 8a and Fig- Figure 8b corresponding to low Ho concentrations (3 at.%). The
50
ure 8b. For the complete picture of electrochemical response sudden rise in current density at low voltage is the result of the
51
from the specimens of different structure, composition, and kinetic hindrance of oxide formation as explained by M. M.
52
morphology, EIS investigations were simultaneously done (as Lorengel in the extended high field model.[46] The full over-
53
will be briefly discussed later in Figure 9). In all cases, typical lapping of positive and negative space charge layers within the
54
valve metal behaviour was observed as evidenced by the oxide (under the influence of applied electric field) marks the
55
rectification of current upon reversal of potential scan due to maximum of the current density peak/overshoot. After that, the
56
the weakening of the electric field. The steady state current growth of new anodic oxide is directly described by the
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1350 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
1
17 Figure 8. (a) Cyclic voltammograms of thermally evaporated Ho and Al in 0.2 M phosphate buffer (pH 8.3) at a sweep rate of 100 mV s (b) Potentiodynamic
anodizing of selected Al Ho alloys at a sweep rate of 100 mV s 1 in 0.2 M phosphate buffer (pH 8.3).
18
19
20
21
mentioned current density plateau, when the rate of potential grain density and surface area. This might be responsible for
22
increase is kept constant. the sudden current rise at the onset of anodizing. Additionally,
23
In the case of Ho anodizing, up to ~ 3 V, the nearly constant very high current density is needed to cover the entire surface
24
current plateau typical of steady state condition was examined. of parent alloy even at 1 V. Nonetheless, it is very interesting
25
However, the steady state region rapidly diminishes above ~ 3 V that even at these non-steady state current plateaus, Ho and
26
and the current density rises continuously without the appear- Al Ho alloys still follow the high field conduction model as can
27
ance of any additional current plateau up to 10 V. Such be seen on the reversal of potential scan after each 1 V in
28
behaviour has been recently characterized by J. P. Kollender Figure 8b. The valve metal behaviour is observable for all
29
et al. during the dissolution of titanium in sulphuric acid under investigated samples containing Al.
30
similar anodizing conditions. The current density evolution was
31
explained in terms of ionic current through the oxide (steady
32
state) and electronic current (non-steady state) originating from 2.4. Electrochemical Impedance Spectroscopy
33
side reactions at oxide/electrolyte interface (i. e. oxygen
34
evolution).[47] At the same time, the previously described After each 1 V step-wise growth of anodic oxide, EIS measure-
35
extremely fined-grained Ho surface (see Figure 7) may provide a ments were performed to analyze the dielectric properties of
36
relatively large grain boundary area and inherent defective anodic films. EIS is often used to investigate a broad range of
37
structure to induce such anomalous current response. The experimental systems with very different electrochemical prop-
38
sudden rise of current density upon anodizing is frequently erties as well as for the advancement of many areas of science
39
observed when the formation of nanoporous/ nanotubes and engineering including new product development, materials
40
occurs.[48,49] However, the absence of any porosity and lack of analysis, and mechanistic studies. It allows for the separation of
41
surface features in SEM imaging annuls the supposition of system components such as electrolyte resistance, system
42
current rise due to the formation of porous structures (Fig- capacitance, and film resistance, which otherwise cannot be
43
ure 7). achieved through steady-state measurements. The determina-
44
At higher concentration of Ho in the Al Ho library, cyclic tion of system capacitance is crucial as it provides information
45
voltammograms display an initial increase in current density regarding layer thickness, active surface area, and material
46
(Figure 8b). The current rises to its maximum value at the end permittivity. The dielectric properties estimated from analysing
47
of the first anodizing step (1 V) and gradually decreases during EIS data of oxide films can be used in the design and fabrication
48
next anodizing steps to a minimum value at 10 V. Furthermore, of semiconductors and integrated circuits.
49
the initial current rise increases with the increase of the Ho Figures 9a–e show the series of Bode plots measured for
50
concentration in the Al matrix. This suggests a strong influence pure Al, pure Ho, and selected Al Ho alloys. The direction of
51
of Ho in the electrochemical response of Al Ho alloys. XRD and arrows shows the increase in anodizing potential (in 1 V steps
52
SEM observations support well this idea by emphasizing the during CV measurements). The Bode plots indicate a typical
53
influence of Ho addition to the structure and film morphology, capacitive response with a negative slope. In all cases, the slope
54
as previously discussed. The elongated and smaller grain gradually reaches a value close to 1 as it can be seen in
55
morphology at 7 at.% Ho, which transforms into the thick Figure 10 for Al, Ho, and Al-11 at. % Ho after anodization up to
56
needle-like structure at 11 at.% Ho, possibly provides a high 10 V. It is obvious that Al, Ho, and Al-3 at.% Ho alloy reveal
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1351 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 Figure 9. (a, e) Bode plots during simultaneous CV and EIS measurements along the thermally evaporated Al Ho alloys as well as on pure Al and Ho in 0.2 M
phosphate buffer (pH 8.4). The EIS spectra were recorded after each 1 V increase in CV up to a maximum of 10 V, (f) comparison of reciprocal capacitance
42 obtained as a function of formation voltage at selected compositions of Al Ho alloys. The inverse capacitance for Ho and Al is also added as a reference.
43
44
45
similar impedance spectra, even though both Al and Ho possess electrolyte or Ohmic resistance Re and a parallel combination of
46
different morphology, structure, and electrochemical behaviour. film resistance Rf and CPE parameters α and Q as:
47
In a narrow frequency range, the phase angle shift shows a
48
non-ideal capacitive response as can be seen from the Rf
49 Z ¼ Re þ (1)
magnitude of the phase shift. In such situations, the constant 1 þ QðjwÞa Rf
50
phase element (CPE) is commonly used to obtain a suitable
51
fitting of impedance spectra. However, the phase angle shifts where ω is the angular frequency in units of Hz, whereas the
52
away from the capacitive region to the positive direction in the impedance associated with CPE in Equation (1) can be ex-
53
frequency range of 101 to 100 Hz indicating that the constant pressed as [Eq. (2)]:
54
phase element exists only in a narrow frequency domain.
55
The impedance spectra of anodic film covered electrode
56
exhibiting CPE behaviour is generally expressed in terms of
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1352 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Figure 10. (a–c) Slope obtained at 1 V and 10 V for Al, Al-11 at. % Ho alloy, and Ho in log f vs log Z plot.
20
21
22 ð2pf Þ a �ap� ð2pf Þ a
ap � � fZmax
Z CPE ¼ Q 1 ðjwÞ a
¼ cos þj sinð Þ (2) 9:2
23 Q 2 Q 2 Rf ¼ Z img dlog10 f (7)
p
24 1

25
Here Q has a unit of S cm 2 sα. When α = 1, Q has the unit of
26
capacitance (F cm 2). Therefore, CPE with parameters Q and α After a careful examination of the frequency dependence of
27
cannot represent the unit of true capacitance for α < 1. The the imaginary part of the impedance (not shown here) it was
28
origin of CPE behaviour is generally attributed to porosity, concluded that CPE exists in a wider frequency domain for the
29
surface roughness and heterogeneity, fractal geometry, non- case of Al anodizing and shrinks to a narrow range at the
30
uniform current distributions, and the presence of grain highest concentration of Ho in the alloy. Furthermore, the
31
boundaries. Independent of the cause of CPE, the phase angle magnitude of phase shift remains constant for Al anodizing,
32
associated with a CPE is independent of frequency, while the whereas for Ho and Al Ho alloys this phase shift depends on
33
parameter Q depends on frequency as shown in the following the anodizing potential and film resistance. The exact value of
34
equations [Eqs. (3)–(4)] phase angle can be estimated from the following equation
35
[Eq. (8)]
36 � �
Z img ap
37 �CPE ¼ arctan ¼ (3) � � ��
Z re 2 �d log�Zimg ��
38 a ¼ �� � (8)
d log f �
39 �ap� 1
40 Q ¼ sin (4)
2 Z img ð2pf Þa Finally, the EIS analysis is summarized in the form of inverse
41
capacitance as a function of formation potential in Figure 9f.
42
The non-ideal behaviour of the anodic film as indicated by The dielectric permittivity from inverse capacitance can be
43
phase shift is fitted well with CPE and the true capacitance, C is estimated by considering the following relation [Eq. (9)];
44
obtained from logf-Zimg plot using the approach described by
45
the Hsu and Mansfeld[50] and B. Hirschorn et al.[51] The Hsu- 1 1 k d0
46 C ¼ ðd þ d0 Þ ¼ Vþ (9)
Mansfeld and B. Hirschorn equations used to convert Q into Aer e0 Aer e0 Aer e0
47
true capacitance C can be expressed as follows [Eqs. (5)-(6)]:
48
in which do is the thickness of natural oxide prior to anodizing,
49 1
C ¼ Qðwmax Þa (5) V is the formation potential, k is the oxide formation factor, and
50
ɛo is the permittivity of free space and ɛr is the relative
51 1 a
C ¼ Qa ðRf Þð a Þ
1
(6) permittivity number to describe the film material, respectively.
52
The k-factor can be estimated either from Faraday’s equation or
53
where ωmax is the frequency at which Zimg has the maximum directly from film thickness measurement in EIS analysis. A
54
value and Rf is film resistance which is also verified using the coulometric analysis or determination of current plateau in
55
following Mansfeld and Kending based integration method[50,52] potentiodynamic sweeps allows the calculation of k-factor by
56
and shows close approximation with EIS fitting data [Eq. (7)]. considering the following equations [Eqs. (10)–(11)]
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1353 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles

q ¼ Kox1 kE (10) 2.5. Corrosion Studies


1
2
In order to further validate our approach for the selection of
3
Where Kox is a material constant defined as: suitable electrolytes and pH range used for the formation of
4
anodic films, corrosion studies were also performed on an
5
M exemplary Al-3.0 at.% Ho alloy with identical alloy preparation
6 K ox ¼ (11)
z1d F history and storage time. The corrosion studies of anodic films
7
were performed in different electrolytes (see Table 2). Potentio-
8
Here, M is the molar mass, F is Faraday’s constant and 1d is dynamic anodizing was repeatedly performed after each
9
the density of oxide. M and 1d of composite anodic films corrosion and Rp measurements and all important parameters,
10
containing oxides of Al and Ho are calculated from mixed namely OCP, potentiodynamic anodizing behaviour, polariza-
11
matter theory assuming a similar migration rate of Al and Ho tion resistance Rp and corrosion rate Crate were continuously
12
within the anodic film. It is important to point out that the k- monitored for 3600 s. The OCP values which give the
13
factor determined from Faraday’s law requires the anodizing information on the ongoing processes are plotted in Figure 11.
14
charge used for oxide formation, while the EIS measurements The OCP values recorded before anodizing do not provide any
15
involve the thickness determination directly resulting from the relationship with the pH of the electrolyte used. In the electro-
16
anodic film. The k-factor obtained from Faradays’s equation for lyte containing phosphates, such as pH 7.3, pH 7.6, and pH 8.3,
17
the anodic film involving non-steady state current response OCP values shift towards 0 V after repeatedly anodizing to 10 V.
18
may not provide a true value. Therefore, the k-factor for the This indicates a higher stability of the films against aqueous
19
anodic film having a linear response of inverse capacitance and corrosion (ESHE = 0 V), while the electrolytes without phosphate
20
formation voltage in Figure 9f is obtained from the slope of film reveal relatively negative OCP values. The change of OCP over a
21
thickness d (estimated from the plot of EIS analysis) as a certain time period during repeated passivation provides
22
function of anodizing potential, E. The analysis provides the k- valuable information on the stability of the resulting passive
23
values of 1.19, 1.22, and 1.0 nm V 1 for pure Al, Al-3 at.% Ho film. The nearly stable and higher OCP values for the pH values
24
alloy and pure Ho, respectively, while the do values vary from 7.6 and 8.3 show that anodic films might be more stable in the
25
1.0 nm to 1.36 nm in above three cases. Dielectric permittivity, electrolytes containing phosphate and in the pH range around
26
which is the ratio between the amounts of electrical energy 8. Figure 12 shows the current-voltage response for various
27
stored in a material to that of vacuum is directly deduced from time intervals in four selected electrolytes exhibiting different
28
the slope of inverse capacitance vs potential plot. In the present pH values. In agreement with the thermodynamic assumptions
29
case, values of 10.1 and 16.0 are calculated for Al and Ho anodic of pH range (7.2 to 8.6) for Al Ho alloys, the electrolytes falling
30
films, respectively, in 0.2 M phosphate buffer. The non-linear outside this pH range are either unsuitable to form anodic films
31
dependency of reciprocal capacitance on voltage makes it on Al Ho alloys or the resulting anodic film might not be
32
difficult to compute the dielectric constant of Al Ho alloys at protective as it is directly observable in Figure 12a and Fig-
33
higher Ho concentrations. In other words, the presence of HoAl3 ure 12d. As it can be seen, the citrate buffer having a pH 6.0
34
rich phase not only affects the structure and morphology of the produces an anodic current density with high magnitude,
35
Al Ho alloys but also influences the resulting properties of the indicating a large consumption of metal cations. On the other
36
anodic films, as dielectric strength largely depends on the hand, although boric acid-NaOH buffer (pH 9.0) displays a 10
37
composition of the anodic film. times lower peak anodic current compared with citrate buffer,
38
One additional aspect may be discussed observing Figure 9. the anodic film does not show typical passive film behaviour as
39
The values of the components found in the impedance spectra seen after repeated anodizing in Figure 12 d. The physical
40
are different in nature. While the film resistance (Rf) scales examination of the specimens (optical images not shown here)
41
inversely and film capacity (C) scales directly with the addressed
42
area, the condition is different for the so-called uncompensated
43
resistance (RΩ) resulting from the use of a 3 electrode arrange-
44
ment. This does not scale with area in a straightforward manner
45
but is rather a result of the solution resistance, the distance
46
between reference electrode and working electrode as well as
47
the area of the counter electrode, among other factors. This is
48
accentuated in the present study by the use of an SDCM with
49
complex 3D geometry. An important observation in all EIS
50
spectra is the ratio of the impedance at very low and very high
51
frequencies. While the low impedance at high frequencies is a
52
direct measure of RΩ the high impedance at low frequencies is
53
the numerical sum of Rf and RΩ. In the present case RΩ ! Rf by
54 Figure 11. Open circuit potential (OCP) during simultaneous potentiodynam-
2.5-3.5 orders of magnitude and therefore the RΩ is not a ic anodizing and corrosion studies in different electrolytes. The OCP is
55
limiting factor in the anodization. examined continuously after each polarization resistance and corrosion rate
56 measurement for a period of 3600s.
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1354 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
Figure 12. Repeated potentiodynamic scans to 10 V at 100 mV s 1 in (a) citrate buffer (pH 6.0), (b) phosphate-citrate buffer (pH 7.6), (c) phosphate buffer
31 (pH 8.4), and (d) boric acid-NaOH buffer (pH 9.0. The anodizing is repeatedly performed after each polarization and corrosion measurement.
32
33
34
was consistent with the magnitude of anodic current density. Figure 12c with diminishing leakage current magnitudes. In
35
The anodic films were completely dissolved in citrate buffer other words, repeated film re-passivation can be used to further
36
(pH 6.0) after 3600 s, while boric acid-NaOH buffer (pH 9.0) improve the passivity of the anodic film by completely
37
partially dissolves the film. Owing to this reason, the polar- removing defects present during a first anodizing cycle. Physical
38
ization curve in Figure 12a display a negligible current density, examination after repeated anodizing and corrosion measure-
39
while the simultaneous film dissolution during anodizing ments with naked eye also revealed that anodic films were
40
induces anodic polarization response without complete passiva- present on the substrate without any film dissolution. Besides
41
tion in Figure 12d. As mentioned earlier in the theoretical pH range, the role of phosphate in a certain pH window is also
42
explanation of selecting Eh-pH diagrams, a relatively high important. Literature data show that phosphate containing
43
concentration (10° mol L 1) of dissolved Al and Ho provides the electrolytes provide advantages over other acidic or alkaline
44
wider oxidation window than lower concentrations of dissolved electrolytes. They are relatively less corrosive compared with
45
Al and Ho (Figure 1a). At the same time, such large oxidation other aqueous electrolytes owing to the strong absorption of
46
window may enhance the dissolution rate by ejecting more phosphate anions on the metal oxide surface and display strong
47
cations into the electrolytes without appreciable film formation. complex forming affinity towards most metal cations (Table 1).
48
Furthermore, it is a well-known fact that the growth of anodic Likewise, F. H. Firsching et al. investigated that solubility of
49
films only proceeds at high Faradaic efficiency when the rate of phosphates containing rare-earth elements are extremely low in
50
oxide formation overrules the rate of film dissolution. Figur- aqueous electrolytes.[37] It is therefore likely that phosphate
51
es 12b and 12 c shows that phosphate containing electrolytes containing electrolytes in a certain pH range are suitable to
52
having pH values 7.6 and 8.3 produce films with appropriate form passive anodic films on Al, Ho, and their alloys. Finally, the
53
passive behaviour even after repeated anodizing for 3600s. The obtained polarization resistance, Rp and corrosion rate, Crate as
54
anodic films remain passive without dissolution and even show well as their time-dependent changes are shown in Figure 13.
55
improved passivation as it can be seen after comparing Phosphate-containing electrolytes having a pH 7.6 and 8.3 show
56
polarization curves for 324 and 3600 s in Figure 12b and a nearly linear rise of Rp with time (Figure 13a). The increase in
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1355 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 Figure 13. (a) Polarization resistance and (b) corrosion rate examined after repeated potentiodynamic anodizing to 10 V in different electrolytes having wide
range of pH.
17
18
19
20
Rp is associated with the improvement in passive behaviour and 3. Conclusions
21
film stability as explained earlier in Figures 12b and 12 c. It is
22
obvious in Figure 13a that only phosphate-containing electro- In this study, a systematic investigation of electrolyte selection
23
lytes having pH values in the range of around 8 provides for alloy anodizing has been carried out on an exemplary valve
24
highest Rp while all other electrolytes similar to Figures 12a and metal, Al and a non-valve metal, Ho and their alloys obtained in
25
12 d give very low Rp values due to either partial or complete identical conditions using a combinatorial approach. Based on
26
dissolution of resulting anodic films during repeated anodizing the results obtained, the following conclusions are drawn:
27
cycles. The obvious reason is the instability of either Al or Ho in 1. The pH range and electrolyte selection prior to anodizing
28
the electrolyte if we move away from the Al or Ho stability can be decided by considering the Gibbs free energy and
29
domains. Apparently, the corrosion rate, Crate in Figure 13b solubility data.
30
appears to be similar for citrate buffer (pH 6.0), boric acid-NaOH 2. The combined Eh-pH diagram of Ho and Al in aqueous
31
buffer (pH 9.0), citrate-Phosphate buffer (pH 7.6), and phosphate electrolyte discloses a pH range where anodizing of Ho and
32
buffer (pH 8.3). However, careful examination (inset of Fig- Al as well as their alloys can be thermodynamically feasible.
33
ure 13b) discloses that only phosphate-containing electrolytes 3. Cyclic voltammograms show that the anodizing behaviour
34
having pH values 7.6 and 8.3 show the negligible Crate with time, largely depends on the inherent composition and structure
35
whereas the Crate of citrate and boric acid-NaOH buffers of the parent alloys. We found out that our present
36
continuously fluctuate during re-passivation cycles due to the approach is suitable even on films with different structures
37
continuous dissolution of the anodic film from the alloy surface. and morphologies and we can anodize alloys having
38
For the determination of Crate from corrosion current Icorr, gradient in composition using a single electrolyte based on
39
equivalent weight, density and number of electrons are also our thermodynamic explanations.
40
defined assuming mixed matter theory. For instance, a value of
41
18.3 g eq 1 and a density of 3.26 g cm 3 was assigned for Al-3 at.
42
% Ho alloy, whereas 1 for amorphous Al2O3 = 3.1 g cm 3, 1 for Acknowledgments
43
Ho2O3 = 8.41 g cm 3 were selected. In summary, the present
44
approach of superimposing Eh-pH diagrams of individual The financial support by the Austrian Federal Ministry of Science,
45
constituents of Al Ho alloy allows us to rapidly identify the pH Research and Economy and the National Foundation for Research,
46
range for anodic oxidation. This, in turn, helps to select a Technology, and Development through funding of the Christian
47
suitable electrolyte, rather than performing several experiments Doppler Laboratory for Combinatorial Oxide Chemistry (COMBOX)
48
on a trial and error basis. This simple approach based on is gratefully acknowledged.
49
theoretical consideration of thermodynamic data (Gibbs free
50
energy) and electrochemical principles enables successful anod-
51
izing not only of Al Ho alloys but also of individual constitu- Conflict of Interest
52
ents, having widely varying structure and surface morphology.
53
Rp and Crate measurements are fully consistent with the results The authors declare no conflict of interest.
54
and further support our theoretical considerations of the
55
electrochemical behaviour of Al and Ho prior to anodizing.
56
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1356 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA
Articles
Keywords: anodic film · aluminium-holmium alloys · scanning [25] A. Güntherschulze, H. Betz, Elektrolytkondensatoren, M. Krayn Verlag,
1 Berlin, 1937.
droplet cell microscopy · potential-pH diagrams
2 [26] L. Young, Anodic Oxide Films, Academic Press, London, 1961.
[27] T. Wiktorczyk, K. Nitsch, Z. Bober, IEEE 1992 87–91.
3
[28] T. Wiktorczyk, IEEE 1992, 27, 807–812.
4 [1] M. Pourbaix, Atlas of electrochemical equilibria in aqueous solutions, [29] T. Wiktorczyk, J. Electrost. 2001, 52, 131–136.
5 Pergamon Press Ltd., Houston, Texas 1974. [30] M. El-Hagary, M. Emam-Ismail, S. H. Mohamed, A. S. Hamid, S. Althoyaib,
[2] P. Katiyar, C. Jin, R. J. Narayan, Acta Mater. 2005, 53, 2617–2622. Thin Solid Films 2010, 518, 4058–4065.
6
[3] H. Masuda, K. Fukuda, Science 1995, 268, 1466–1468. [31] A. W. Hassel, M. M. Lohrengel, Electrochim. Acta. 1997, 42, 3327–3333.
7 [4] Y. Konno, A. A. Farag, E. Tsuji, Y. Aoki, H. Habazaki, J. Electrochem. Soc. [32] https://www.gwb.com/.
8 2016, 163, C386-C393. [33] https://www.gwb.com/software_overview.php.
[5] H. Habazaki, S. Koyama, Y. Aoki, N. Sakaguchi, S. Nagata, ACS Appl. [34] http://www.kemi.kth.se/medusa.
9
Mater. Interfaces 2011, 3, 2665–2670. [35] https://sites.google.com/site/chemdiagr/.
10 [6] M. Forsyth, B. Hinter, Rare Earth based Corrosion Inhibitors, Elsevier, [36] D. R. Lide, CRC Handbook of Chemistry and Physics, 71st edition, CRC
11 Cambridge, Uk 2014, 143–162. Press, Boston 1990.
[7] I. Paramasivam, H. Jha, N. Liu, P. Schmuki, Small 2012, 8, 3073–3103. [37] F. H. Firsching, J. Chem. Eng. Data. 1992, 37, 497–499.
12
[8] N. Baram, D. Starosvetsky, J. Starosvetsky, M. Epshtein, R. Armon, Y. Ein- [38] K. A. Persson, B. Waldwick, P. Lazic, G. Ceder, Phys. Rev. B - Condens.
13 Eli, Electrochim. Acta 2009, 54, 3381–3886. Matter Mater. Phys. 2012, 85, 1–12.
14 [9] J. Ferre-Borrull, J. Pallares, G. Macias, L. F. Marsal, Materials 2014, 7, [39] K. Shahzad, C. C. Mardare, D. Recktenwald, A. I. Mardare, A. W. Hassel,
5225–5253. Electrochim. Acta 2019, 297, 888–904.
15
[10] F. De Geuser, M. J. Styles, C. R. Hutchinson, A. Deschamps, Acta Mater. [40] T. B. Massalski, Binary Alloys Phase Diagrams, ASM, Ohio 1986, 122.
16 2015, 101, 1–9. [41] K. A. Gschneidner, F. W. Calderwood, Bull. Alloy Phase Diagrams 1988, 9,
17 [11] A. I. Mardare, A. Savan, A. Ludwig, A. D. Wieck, A. W. Hassel, Electrochim. 684–685.
Acta 2009, 54, 5973–5980. [42] A. Meyer, J. Less-Common Met. 1965, 10, 121–129.
18
[12] C. Nyshadham, C. Oses, J. E. Hansen, I. Takeuchi, S. Curtarolo, G. L. W. [43] K. A. Gschneidner, F. W. Calderwood, Bull. Alloy Phase Diagrams 1988, 9,
19 Hart, Acta Mater. 2017, 122, 438–447. 658–668.
20 [13] S. Kirklin, J. E. Saal, V. I. Hegde, C. Wolverton, Acta Mater. 2016, 102, [44] H. Masuda, K. Fukuda, Science 1995, 268, 1466–1468.
125–135. [45] Y. Konno, A. A. Farag, E. Tsuji, Y. Aoki, H. Habazaki, J. Electrochem. Soc.
21
[14] C. Fenster, M. Rohwerder, A. W. Hassel, Mater. Corros. 2009, 60, 855–858. 2016, 163, C386–C393.
22 [15] A. I. Mardare, C. D. Grill, I. Pötzelberger, T. Etzelstorfer, J. Stangl, A. W. [46] M. M. Lohrengel, Mater. Sci. Eng. R 1993, 11, 243–294.
23 Hassel, J. Solid State Electrochem. 2016, 20, 1673–1681. [47] J. P. Kollender, A. W. Hassel, ChemElectroChem. 2017, 4, 1–4.
[16] S. Milenkovic, V. Dalbert, R. Marinkovic, A. W. Hassel, Corros. Sci. 2009, [48] C.-Y. Lee, L. Wang, Y. Kado, M. S. Killian, P. Schmuki, ChemSusChem.
24
51, 1490–1495. 2014, 7, 934–940.
25 [17] P. J. Gilling, C. B. Cass, A. R. Malcolm, M. R. Fraundorfer, J. Endourol. [49] H. Habazaki, K. Shahzad, T. Hiraga, E. Tsuji, Y. Aoki, ECS Trans. 2015, 69,
26 1995, 9, 151–153. 211–223.
[18] S. Han, Phys. B 1990, 164, 17–20. [50] C. H. Hsu, F. Mansfeld, Corros. Sci. Sect. 2001, 57, 747–748.
27
[19] L. B. Lerner, M. D. Tyson, Urol. Clin. 2009, 36, 485–495. [51] B. Hirschorn, M. E. Orazem, B. Tribollet, V. Vivier, I. Frateur, M. Musiani,
28 [20] G. S. Young, R. Ball, Urol. Nurs. 1999, 19, 254–257. Electrochim. Acta 2010, 55, 6218–6227.
29 [21] M. J. Erhard, D. H. Bagley, J. Endourol. 1995, 9 383–386. [52] M. Kendig, F. Mansfeld, Corrosion 1983, 39, 466–467.
30 [22] F. D. Natterer, K. Yang, W. Paul, P. Willke, T. Choi, T. Greber, A. J.
Heinrich, C. P. Lutz, Nature 2017, 543, 226–228.
31 [23] A. A. Khajetoorians, A. J. Heinrich, Toward single- atom memory, Science
32 2016, 352, 296–297. Manuscript received: October 1, 2019
33 [24] J. F. Rider, Inside the Vacuum Tube, John F. Rider Publisher, INC, New Revised manuscript received: November 19, 2019
York 1945. Accepted manuscript online: December 11, 2019
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57

ChemElectroChem 2020, 7, 1342 – 1357 www.chemelectrochem.org 1357 © 2019 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA

You might also like