Observation of Pore Growth and Self-Organization in Anodic Alumina

Mar 1, 2018 - The anodic oxidation of metals such as aluminum and titanium can lead to the development of self-ordering pores. These pores make excell...
0 downloads 0 Views 5MB Size
Article Cite This: ACS Appl. Nano Mater. 2018, 1, 1265−1271

www.acsanm.org

Observation of Pore Growth and Self-Organization in Anodic Alumina by Time-Resolved X‑ray Scattering Nikolay A. Vinogradov,†,‡ Gary S. Harlow,*,§ Francesco Carlà,† Jonas Evertsson,§ Lisa Rullik,§ Weronica Linpé,§ Roberto Felici,†,∥ and Edvin Lundgren§ †

ESRF−The European Synchrotron, 71 Avenue des Martyrs, 38000 Grenoble, France MAX IV Laboratory, SE-22594 Lund, Sweden § Division of Synchrotron Radiation Research, Lund University, SE-22100 Lund, Sweden ∥ SPIN-CNR, c/o DICII-University of Rome Tor Vergata, Via del Politecnico 1, I-00133 Roma, Italy ‡

S Supporting Information *

ABSTRACT: The anodic oxidation of metals such as aluminum and titanium can lead to the development of selfordering pores. These pores make excellent templates for a range of nanoscale objects with many applications in nanoscience. Theoretical studies on pore formation have proposed several models for the establishment, growth, and ordering of these pores; however, experimental verification has mostly been limited to ex situ measurements. Here we show that the lateral and vertical pore structure can be probed in situ with high precision, using grazing transmission X-ray scattering. By making use of the high flux available at modern synchrotrons and fitting only the difference between scattering patterns we show the nearly real-time evolution of the pore’s arrangement. We observe no dependence on the substrate crystallographic orientation for domain size or pore separation. We do however observe an anisotropy in the oxide growth rate for the different substrate surfaces. This experimental approach can be applied to the study of a large variety of electrochemically produced materials such as magnetic nanowires, novel solar cell designs, and catalysts. KEYWORDS: aluminum anodization, nano pore, self-organization, GTSAXS, small-angle scattering, porous membrane, X-ray, electrochemistry



INTRODUCTION Porous anodic alumina (PAA) is formed by the electrochemical anodization of aluminum in various electrolytes.1 Certain electrolytes and voltages have been shown to cause a selfordering of the pores into a hexagonal pattern.2,3 Pore diameters can be tuned in a range from 10 to 400 nm, whereas pore heights can be up to tens of micrometres.4 Remarkably the self-ordering is coherent over micrometer-sized domains (known as coherent domains) each with the same structural parameters (pore diameter and pore separation) but a different orientation. After long periods of anodization the orientational alignment of these coherent domains also increases.5,6 The range of possible aspect ratios of PAA has attracted much interest and has found potential applications in nanoscale fabrication,7 magnetic storage,8 catalytic membranes,9 batteries,10 and photocatalysis/solar cells11 as well as being used industrially as an adhesive base for decorative coloring processes. Despite the importance of PAA, the reason for the self-organization of the pores is not yet fully understood where the importance of both stress and dissolution for selforganization is not completely clear. © 2018 American Chemical Society

It is known that during electrochemical anodization metal ions from the oxidation of the substrate, and oxygen ions from water, migrate through the oxide layer due to the high electric field (∼1 V nm−1) created by the applied voltage.12−14 Some of this ionic current results in the field-assisted ejection of Al3+ cations into the solution, but most (60−90%) results in oxide growth.15 One theory of pore formation is that a morphological instability begins to develop in the early stages of oxide growth that leads to pore formation.15 The competition between oxide growth and dissolution processes then leads to pore formation. It has been suggested that an important driving force for pore formation and ordering is mechanical stress, which is induced by the incorporation of electrolyte species into the oxide matrix.16,17 The exact mechanism is still under debate, and these models fail to account for long-range ordering or the observation of differences in the orientational order when different aluminum single-crystal faces are anodized.18,19 Received: January 5, 2018 Accepted: March 1, 2018 Published: March 1, 2018 1265

DOI: 10.1021/acsanm.7b00303 ACS Appl. Nano Mater. 2018, 1, 1265−1271

Article

ACS Applied Nano Materials The sensitivity of in-plane orientational order to crystallographic orientation is surprising considering that the oxide layer is amorphous and the pore separation is several orders of magnitude greater than the lattice constant of the Al substrate. The average size of defect-free domains is found to follow the trend (110) < (111) < (100), after long anodization times.18 It has been suggested that the crystallographic orientation of spikes that form at the metal/oxide interface due to the hemisphere-like pore base is impacted by a difference in oxidation rates of different crystallographic planes.19 These spikes then transfer their orientational correlation to the pores as the oxide growth deepens into the metal substrate with crystallographic planes oriented closer to {111} showing the slowest rate of film growth. Experimental validation of these models of pore growth has, however, been mostly limited to ex situ measurements separated by large time intervals. Typical techniques used to examine nanoscale structures are atomic force microscopy (AFM), scanning electron microscopy (SEM), and transmission electron microscopy (TEM). These postfactum methods often require additional sample processing stages such as cleavage or metallic sputtering. Ideally measurements of pore geometry and growth should be made such that the process of interest can be observed in real time. Previous in situ measurements have used transmission small-angle X-ray scattering (SAXS);19−21 however, the need for the beam to penetrate through the substrate imposes some limitations on sample thickness. Further limitations of the transmission SAXS approach are a limited temporal resolution for changes in the oxide perpendicular to substrate, which can only be measured by tilting the sample (a rocking scan). Here we report in situ grazing-incidence transmission small-angle Xray scattering (GTSAXS) measurements during the anodization of aluminum samples with different crystallographic orientations. Despite its simplicity, the GTSAXS technique was only introduced recently,22 and our measurements demonstrate how this technique can be used for fast in situ investigations of nanoscale electrochemical systems such as PAA. The X-ray beam in the grazing transmission geometry (illustrated in Figure 1) only needs to penetrate a small amount of the substrate (compared to a standard transmission geometry); therefore, a stronger scattering signal can be detected. At 21.5 keV the attenuation length for aluminum is ∼1.4 mm, whereas our samples are 7 mm thick. Analysis of GTSAXS data is also considerably simpler than that of the reflected scattering normally measured in GISAXS, because the incident angle of the radiation is much larger than the critical for total internal reflection. Therefore, only the “Born Approximation” and not the full “Distorted Born Wave Approximation” is needed to model the data.22 In the case of GTSAXS the length of the pores is oriented nearly perpendicular to the incident beam; therefore, additional information about their out-of-plane structure (along the Qz reciprocal space direction) is available without rocking the sample. For the measurements reported here, the scattering was nearly homogeneous with rotation (a mosaic of many rotated domains is simultaneously sampled) suggesting no long-range correlational order between the different domains over the anodization time investigated. Aluminum single crystals with polished (100) and (111) faces and a polycrystalline sample have been anodized at 40 V (constant voltage) in 0.3 M oxalic acid at ∼4 °C using a twostep procedure known to give well-ordered uniform pores.23 A schematic of the electrochemical cell used is shown in Figure

Figure 1. Schematic of the experiment. (a) Illustration of the electrochemical cell, Al3+ ions migrate through the anodic alumina oxide (AAO) and are ejected into solution. O2− ions at the anode (working electrode) are incorporated into the new oxide, whereas H2 bubbles formed at the cathode (counter electrode) flow out of the cell. Polished hat-shaped aluminum working electrodes were used. (b) Illustration of the GTSAXS geometry with a vertical beam stop shown. Qz probes the out-of-plane structure, and Qy probes the in-plane structure.

1a. The sample and cell were tilted at glancing angles ∼5 times larger than the critical angle (αc) of aluminum in water, such that the reflected and transmitted scattering are well-separated, and the diffracted signal is at a vertical height on the detector close to the direct beam. Figure 2c shows three frames during the anodization of the Al (100) sample, where the frames are cropped to show only the right-hand side of the symmetrical pattern about the beam stop.



1ST ANODIZATION: PORE SELF-ORGANIZATION During the first anodization, for the first 60 s the scattering is confined to a horizontal streak (due to in-plane roughness) with no features along Qz. This corresponds to initial stages of pore formation (Figure 2a-I), which is characterized by a dip in the measured current (Figure 2b) due growth of the barrier layer.12,24,25 At longer times, peaks along Qz = 0 nm−1 start to appear, a signature of the quasi-hexagonal pattern of the pores. The first peak appears at Qy ≈ 0.07 nm−1 (Figure 2a-II). The entire sequence can be seen in the movie M1. The scattered intensity is a product of the average pore form factor F(Q) and a lattice structure factor S(Q). For a hexagonally packed structure S(Q) has maxima at Qhk = 4π h2 + hk + k 2 /( 3 a) where a is the unit-cell dimension (i.e., pore separation), and h and k are the Miller indices. The position of the first few peaks is indicated in Figure 2c-III (although somewhat difficult to resolve on the image by eye due to the broadening) and is in good agreement with those of a hexagonal lattice, but the 1266

DOI: 10.1021/acsanm.7b00303 ACS Appl. Nano Mater. 2018, 1, 1265−1271

Article

ACS Applied Nano Materials

Figure 2. Pore ordering during 1st anodization at 40 V. (a) Illustration of the pore alignment process, I−pore initiation, II−pore alignment, III− growth of ordered pores. (b) Current−time transient during the anodization of a polycrystalline Al sample under similar conditions. (c) GTSAXS detector images recorded after 0 s (c-I), 250 s (c-II), and 7200 s (c-III) of anodization for Al (100) (see M1); images are normalized to an ionization chamber monitor before the sample. (d) The evolution of average pore separation as a function of time obtained from fits to data along Qz = 0. (e) The domain size as a function of time for the three samples investigated; the solid line is a power-law fit to the data as described in the text.

higher-order peaks are slightly displaced, because the PAA is only quasi-hexagonally ordered (i.e., it is a paracrystal26). With sufficient time new oxide forms deeper into the metal with an increased pore order compared to that from earlier in the anodization process, with the old oxide at the oxide/electrolyte interface showing the least order (Figure 2a-III). The scattering in each GTSAXS image is the sum of the entire oxide layer that is in the path of the X-ray beam. The oxide grows deeper into the metal with time. To understand how the structure of the new oxide evolves with time it is necessary to ascertain changes in the scattered intensity. To achieve this goal, we sum the intensity of several images and then subtract the scaled intensity of the image immediately before the averaged images (see Supporting Information). This method is only applicable along the GTSAXS component parallel to the sample surface (L = 0), where the scattering sums incoherently. Here we make the assumption that as the

new oxide forms the old oxide remains unmodified; this is a reasonable approximation given how the oxide grows at low temperatures.12,13 Furthermore, it is not expected that the scattering from the old oxide will be significantly affected by any viscous flow of the oxide,27 as this mostly occurs in the new oxide close to the pore bottom being measured. In-plane structural parameters can then be determined by fitting the data along the line at Qz = 0. A model based on previous ex situ models of hexagonal nanostructures28,29 was used to fit data and obtain values for the pore separation and coherent domain length (see Supporting Information). The evolution of the pore separation and the coherent domain length for all three samples is shown in Figure 2d,e, respectively. The average pore separation at the end of the 2 h anodization was ∼100 nm, whereas the coherent domain length was ∼800 nm. These are in good agreement with our ex situ measurements (see Supporting Information). 1267

DOI: 10.1021/acsanm.7b00303 ACS Appl. Nano Mater. 2018, 1, 1265−1271

Article

ACS Applied Nano Materials

with earlier studies that found similar exponents.20,31 However, both these previous reports on domain growth were made at much longer time scales, with few data points over the early stages of anodization that is investigated in this report. It is also worth mentioning the parameter measured here is an average and assumes the domains are approximately circular. This kind of power-law behavior is similar to that of grain growth in metals, albeit with a different exponent, where the driving force is a reduction in the boundary energy.31,32 It has been suggested that the growth of these domains is similar to Ostwald ripening.20 Some pores close to the domain boundaries stop growing, and the adjacent pores split into two channels at positions that merge the domains. This is of course the reason for why longer anodization times are often used, to increase the size of coherent domains; it is clear from Figure 2d and Figure 3 that the separation between the pores themselves plateaus rapidly.

The time at which the largest increase in pore separation takes place corresponds well with the rise in current after pore formation occurs (Figure 2b), suggesting considerable activity of the pores during this phase. Figure 3 shows how the pore



2ND ANODIZATION: ALIGNED PORE GROWTH

After chemical etching of the oxide layer, which was formed in the first anodization process, a second anodization process was investigated using GTSAXS. These results are summarized in Figure 4a for the Al (100) sample, where the entire time evolution can be viewed in movie M2. Because of a prepatterning of the surface in the first anodization stage, a streak of intensity along Qz at the (1 0) Bragg peak position (Qy ≈ 0.07 nm) is already visible. Since the self-ordering process has considerably slowed during the first step, we observe very little change in the parameters discussed in the previous section, over a 2 h anodization (see Supporting Information). In the early stages of oxide growth oscillations along the Qz direction are visible on the detector images (see Figure 4b). These are characteristic of the scattering from one-dimensional objects (i.e., the cylindrical pores) and are present for the second anodization, because the pores are more uniform in shape and height compared to the first anodization. The pore height (H) can be determined from the relation H = 2π/Δqz, where Δqz is the distance between adjacent oscillations. Fits of a cylindrical scattering form factor to the measured intensity along Qz are shown in Figure S2 at various time intervals (see Supporting Information). Figure 4c shows the evolution of the calculated pore heights during the early stages of the second anodization for each of the samples investigated. There are clear differences between the three aluminum faces; the fastest growth occurs with the (100) sample, which was found to grow at 1.6 nm/s (from the linear fit indicated with the red line). The slowest rate of growth in our samples was found for (111), which grew at 1.2 nm/s. The polycrystalline sample had a rate of oxide growth of 1.4 nm/s. It is already known in the very early stages of growth (where there is a large change in current) that the pore growth is not linear with time and much slower;33 therefore, the lines do not pass through the origin. Figure S7 shows a plot of the charge consumed against pore height, which does indeed pass through the origin within the error of the experiment. The relative simplicity with which growth rates, pore separation, and domain sizes can be calculated from the in situ experiments presented here shows why GTSAXS can be an important technique for understanding the growth of nanoscale materials such as PAA and their functionalization.

Figure 3. Average pore separation at different voltages. The evolution of average pore separation as a function of time obtained from fits to data along Qz = 0. The data were measured during the anodization of polycrystalline aluminum in 0.3 M oxalic acid. Solid lines show fits to the data using the model described in the text; parameters are given in Supporting Information. The increased noise in the 60 V line is an experimental artifact induced by an attenuator change and the first diffraction maxima being closer to the beam stop and is not indicative of the pore structure.

separation evolves in time for different anodizing voltages. For higher voltages (e.g., 50 and 60 V) the etching conditions at the pore bottoms are strongly time-dependent, and this leads a decline in measured current over time, accompanied by worse ordering of the pores.30 Since we only anodize for a relatively short time period this does not significantly affect our measurements. The solid lines in Figure 3 are a fit to the data using a sigmoidal function (see Supporting Information). This is the first time we are aware of that the early dynamics of pore separation have been measured in situ or investigated at different voltages. These time-resolved data on pore ordering should be useful in testing the various theories that currently exist to explain the self-ordering phenomena. Interestingly we found no significant difference between the different crystal faces (Figure 2e) or voltages (Figure S1) with respect to the average domain size. In contrast, reported ex situ measurements of anodization over much longer time scales show a difference in the long-range ordering and density of defects depending on crystallographic orientation.6,18 However, the order parameter we describe here is not equivalent, and such changes only become apparent after much longer anodization times, where a longer-range order has been reported.19 The coherent domain length (Figure 2e) is found to follow a power law (L = Btn, where B is a function of temperature); for our measurements, a value of n = 0.25 ± 0.04 best fit the complete data, as indicated by the black line. This implies that the domain area grows with a value n = 0.5 and is in agreement 1268

DOI: 10.1021/acsanm.7b00303 ACS Appl. Nano Mater. 2018, 1, 1265−1271

Article

ACS Applied Nano Materials

Figure 4. Oxide growth during 2nd anodization stage. (a) GTSAXS images recorded after I−0 s, II−250 s, and III−7200 s of the 2nd anodization for Al (100) at 40 V in 0.3 M oxalic acid; images are normalized to an ionization chamber monitor before the sample. (b) Oscillations in the intensity along Qz at the first Bragg peak (Qy ≈ 0.07 nm) for the Al (100) sample at various times. (c) The calculated pore length from the period of the oscillations in Qz for the three samples. The lines show the linear fits used to determine growth rates.





CONCLUSIONS

EXPERIMENTAL PROCEDURES

The aluminum samples were Al(100) and Al(111) single crystals and polycrystalline Al (from Goodfellow, 99.995% purity) extruded from an Al rod, machined in a hat shape with the diameter of 7 mm on the polished side. The samples were mechanically polished by Surface Preparation Laboratory, Netherlands (SPL) with final polishing step less than or equal to 0.03 μm. The alignment of the crystallographic faces was to ∼0.1°. These were anodized in a home-built polyether ether ketone (PEEK) electrochemical cell under potentiostatic conditions, at 3−5 °C and constant electrolyte flow; the bath temperature was −7 °C, but the ambient temperature increased the temperature of the electrolyte before it reached the cell. A 0.3 M oxalic acid (H2C2O4) + 5 wt % ethanol (C2H5OH) solution in ultrapure water was used as an electrolyte. The presence of ethanol was used to prevent the electrolyte freezing in the bath. For each sample, we have employed the well-known two-step anodization process,2 where after an initial anodization the oxide layer is stripped in a mixture of chromic and phosphoric acida subsequent anodization then produces well-ordered and aligned pores. Constant pumping of the electrolyte through the cell also allows for controlling the temperature of the electrolyte in the electrochemical cell and for removing the reaction heat from the small volume of the cell. For each sample the first anodization step lasted ∼3 h, the second anodization step was done for a shorter time, ∼2 h, due to fewer changes in the scattering pattern at higher anodization times. An energy of 21.5 keV photons (λ = 0.057 77 nm) at the ID03 (ESRF) beamline34 was used for the experiment. A Maxipix-X5 detector was placed 2.2 m away from the sample, in the path of the direct X-ray beam, and used for capturing the scattering pattern. The detector dimensions are 256(v) × 1296(h) pixels, where each pixel has a square shape and dimensions of 55 × 55 μm. Approximately 3 mm in front of the detector a 15(v) × 0.5(h) mm tantalum beam stop was mounted to prevent the detector oversaturation by the direct and reflected beams. A framerate of up to 1 Hz was achievable but reduced to 0.1 Hz to

The technical feasibility of in situ GTSAXS during the electrochemical growth of porous oxides, such as PAA, is demonstrated. We have introduced a difference GTSAXS technique that allows changes in the pore separation and average coherent domain size to be followed throughout the anodization process and compared for different crystal faces as these parameters evolve. Previous SAXS measurements have only considered how the overall average throughout the oxide layer evolves in time. We have shown how pore separation and average domain lengths for PAA evolve in time during the early stages of anodization for three different crystal faces and as well as with different voltages. Our measurements gave similar results for all the crystallographic orientations investigated, despite a difference in the oxide growth rates. This suggests that the positional correlation of the pores is independent of the height of the oxide layer above. We find, in agreement with previous studies,31 that domain growth follows a power law like that of grain growth in metals, albeit with a different exponent, where there is a tendency to minimize the boundary energy. It has also been shown how the growth rates of the aluminum oxide (i.e., pore heights) for the various crystallographic orientations can be measured in situ with GTSAXS. The growth rate was found to be greatest for the (100) face. Furthermore, our approach demonstrates how GTSAXS can be used for real-time in situ measurements of this important class of materials, opening the door to further in situ studies in harsh conditions during the electrochemical production of nanoscale solar cells, catalysts, and membranes. 1269

DOI: 10.1021/acsanm.7b00303 ACS Appl. Nano Mater. 2018, 1, 1265−1271

Article

ACS Applied Nano Materials minimize the effect of beam damage. The setup with such parameters allows the detection of the momentum transfer from Qmin = 0.0136 ± 0.0027 nm−1 to Qmax = 1.765 ± 0.0027 nm−1 with the resolution in the reciprocal space of 2.72 × 10−3 nm−1. This allows for observing the structural features from the objects with typical sizes in the range from 462 to 3.55 nm, respectively. The diffraction pattern was corrected for refraction (which was in any case small due to the high incidence angle) by assigning the center pixel Qy,z = 0 from the symmetry of the diffraction pattern after anodization.



(5) Napolskii, K. S.; Roslyakov, I. V.; Eliseev, A. A.; Petukhov, A. V.; Byelov, D. V.; Grigoryeva, N. A.; Bouwman, W. G.; Lukashin, A. V.; Kvashnina, K. O.; Chumakov, A. P.; et al. Long-Range Ordering in Anodic Alumina Films: A Microradian X-Ray Diffraction Study. J. Appl. Crystallogr. 2010, 43, 531−538. (6) Roslyakov, I. V.; Koshkodaev, D. S.; Eliseev, A. A.; HermidaMerino, D.; Ivanov, V. K.; Petukhov, A. V.; Napolskii, K. S. Growth of Porous Anodic Alumina on Low-Index Surfaces of Al Single Crystals. J. Phys. Chem. C 2017, 121, 27511−27520. (7) Chun, H.; Hahm, M. G.; Homma, Y.; Meritz, R.; Kuramochi, K.; Menon, L.; Ci, L.; Ajayan, P. M.; Jung, Y. J. Engineering Low-Aspect Ratio Carbon Nanostructures: Nanocups, nanorings, and Nanocontainers. ACS Nano 2009, 3, 1274−1278. (8) Berganza, E.; Bran, C.; Jaafar, M.; Vázquez, M.; Asenjo, A. Domain Wall Pinning in FeCoCu Bamboo-like Nanowires. Sci. Rep. 2016, 6, 29702. (9) Pellin, M. J.; Stair, P. C.; Xiong, G.; Elam, J. W.; Birrell, J.; Curtiss, L.; George, S. M.; Han, C. Y.; Iton, L.; Kung, H.; et al. Mesoporous Catalytic Membranes: Synthetic Control of Pore Size and Wall Composition. Catal. Lett. 2005, 102, 127−130. (10) Gowda, S. R.; Leela Mohana Reddy, A.; Zhan, X.; Ajayan, P. M. Building Energy Storage Device on a Single Nanowire. Nano Lett. 2011, 11, 3329−3333. (11) Roy, P.; Berger, S.; Schmuki, P. TiO2 Nanotubes: Synthesis and Applications. Angew. Chem., Int. Ed. 2011, 50, 2904−2939. (12) Diggle, J. W.; Downie, T. C.; Goulding, C. W. Anodic Oxide Films on Aluminum. Chem. Rev. 1969, 69, 365−405. (13) O’Sullivan, J. P.; Wood, G. C. The Morphology and Mechanism of Formation of Porous Anodic Films on Aluminium. Proc. R. Soc. London, Ser. A 1970, 317, 511−543. (14) Singh, G. K.; Golovin, A. A.; Aranson, I. S.; Vinokur, V. M. Formation of Nanoscale Pore Arrays during Anodization of Aluminum. Europhys. Lett. 2005, 70, 836−842. (15) Hebert, K. R.; Albu, S. P.; Paramasivam, I.; Schmuki, P. Morphological Instability Leading to Formation of Porous Anodic Oxide Films. Nat. Mater. 2011, 11, 162−166. (16) Ç apraz, Ö . Ö .; Shrotriya, P.; Skeldon, P.; Thompson, G. E.; Hebert, K. R. Role of Oxide Stress in the Initial Growth of SelfOrganized Porous Aluminum Oxide. Electrochim. Acta 2015, 167, 404−411. (17) Ç apraz, Ö . Ö .; Shrotriya, P.; Skeldon, P.; Thompson, G. E.; Hebert, K. R. Factors Controlling Stress Generation during the Initial Growth of Porous Anodic Aluminum Oxide. Electrochim. Acta 2015, 159, 16−22. (18) Beck, G.; Bretzler, R. Regularity of Nanopores in Anodic Alumina Formed on Orientated Aluminium Single-Crystals. Mater. Chem. Phys. 2011, 128, 383−387. (19) Napolskii, K. S.; Roslyakov, I. V.; Romanchuk, A. Y.; Kapitanova, O. O.; Mankevich, A. S.; Lebedev, V. A.; Eliseev, A. A. Origin of Long-Range Orientational Pore Ordering in Anodic Films on Aluminium. J. Mater. Chem. 2012, 22, 11922−11926. (20) Napolskii, K. S.; Roslyakov, I. V.; Eliseev, A. A.; Byelov, D. V.; Petukhov, A. V.; Grigoryeva, N. A.; Bouwman, W. G.; Lukashin, A. V.; Chumakov, A. P.; Grigoriev, S. V. The Kinetics and Mechanism of Long-Range Pore Ordering in Anodic Films on Aluminum. J. Phys. Chem. C 2011, 115, 23726−23731. (21) Roslyakov, I. V.; Koshkodaev, D. S.; Eliseev, A. A.; HermidaMerino, D.; Petukhov, A. V.; Napolskii, K. S. Crystallography-Induced Correlations in Pore Ordering of Anodic Alumina Films. J. Phys. Chem. C 2016, 120, 19698−19704. (22) Lu, X.; Yager, K. G.; Johnston, D.; Black, C. T.; Ocko, B. M. Grazing-Incidence Transmission X-Ray Scattering: Surface Scattering in the Born Approximation. J. Appl. Crystallogr. 2013, 46, 165−172. (23) Masuda, H.; Satoh, M. Fabrication of Gold Nanodot Array Using Anodic Porous Alumina as an Evaporation Mask. Jpn. J. Appl. Phys. 1996, 35, L126. (24) Thamida, S. K.; Chang, H.-C. Nanoscale Pore Formation Dynamics during Aluminum Anodization. Chaos 2002, 12, 240−251.

ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsanm.7b00303. First anodization process (MPG) Second anodization process (MPG) Domain growth as a function of applied potential, fitting of oscillations along Qz, subtraction procedure, fitting of experimental data, fits to evolution of pore separation, current−time transients, consumed charge, and pore growth, pore height as a function of consumed charge, ex situ investigations, evolution of average parameters for 2nd anodization step (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Gary S. Harlow: 0000-0001-6644-0648 Roberto Felici: 0000-0001-9897-5866 Edvin Lundgren: 0000-0002-3692-6142 Author Contributions

N.A.V., F.C., and R.F. conceived and designed the experiments; G.S.H. and N.A.V. analyzed the data and wrote the manuscript. All authors preformed the X-ray experiments and discussed the results. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The experiments were performed on the ID03 beamline at the European Synchrotron Radiation Facility (ESRF). This work was financially supported by the Foundation for Strategic Research (SSF) Project No. RMA11-0090, and the Swedish Research Council by the Röntgen-Ångströom Cluster “In-situ High Energy X-ray Diffraction from Electrochemical Interfaces (HEXCHEM)” (Project No. 2015-06092). N.V. gratefully acknowledges the financial support from Knut and Alice Wallenberg Foundation.



REFERENCES

(1) Lee, W.; Park, S.-J. Porous Anodic Aluminum Oxide: Anodization and Templated Synthesis of Functional Nanostructures. Chem. Rev. 2014, 114, 7487−7556. (2) Masuda, H.; Fukuda, K. Ordered Metal Nanohole Arrays Made by a Two-Step Replication of Honeycomb Structures of Anodic Alumina. Science 1995, 268, 1466−1468. (3) Nielsch, K.; Choi, J.; Schwirn, K.; Wehrspohn, R. B.; Gösele, U. Self-Ordering Regimes of Porous Alumina: The 10 Porosity Rule. Nano Lett. 2002, 2, 677−680. (4) Ono, S.; Masuko, N. Evaluation of Pore Diameter of Anodic Porous Films Formed on Aluminum. Surf. Coat. Technol. 2003, 169− 170, 139−142. 1270

DOI: 10.1021/acsanm.7b00303 ACS Appl. Nano Mater. 2018, 1, 1265−1271

Article

ACS Applied Nano Materials (25) Parkhutik, V. P.; Shershulsky, V. I. Theoretical Modelling of Porous Oxide Growth on Aluminium. J. Phys. D: Appl. Phys. 1992, 25, 1258. (26) Cheng, W.-D.; et al. GISAXS and SAXS Studies on the Spatial Structures of Co Nanowire Arrays. Chin. Phys. C 2011, 35, 875. (27) Houser, J. E.; Hebert, K. R. The Role of Viscous Flow of Oxide in the Growth of Self-Ordered Porous Anodic Alumina Films. Nat. Mater. 2009, 8, 415. (28) Förster, S.; Timmann, A.; Konrad, M.; Schellbach, C.; Meyer, A.; Funari, S. S.; Mulvaney, P.; Knott, R. Scattering Curves of Ordered Mesoscopic Materials. J. Phys. Chem. B 2005, 109, 1347−1360. (29) Engel, M.; Stühn, B.; Schneider, J. J.; Cornelius, T.; Naumann, M. Small-Angle X-Ray Scattering (SAXS) off Parallel, Cylindrical, Well-Defined Nanopores: From Random Pore Distribution to Highly Ordered Samples. Appl. Phys. A: Mater. Sci. Process. 2009, 97, 99−108. (30) Jessensky, O.; Müller, F.; Gösele, U. Self-Organized Formation of Hexagonal Pore Structures in Anodic Alumina. J. Electrochem. Soc. 1998, 145, 3735−3740. (31) Li, F.; Zhang, L.; Metzger, R. M. On the Growth of Highly Ordered Pores in Anodized Aluminum Oxide. Chem. Mater. 1998, 10, 2470−2480. (32) Hillert, M. On the Theory of Normal and Abnormal Grain Growth. Acta Metall. 1965, 13, 227−238. (33) Kim, M.; Kim, H.; Bae, C.; Lee, J.; Yoo, H.; Moreno, J. M. M.; Shin, H. Initial Self-Ordering of Porous Anodic Alumina: Transition from Polydispersity to Monodispersity. J. Phys. Chem. C 2014, 118, 26789−26795. (34) Ferrer, S.; Comin, F. Surface Diffraction Beamline at ESRF. Rev. Sci. Instrum. 1995, 66, 1674−1676.

1271

DOI: 10.1021/acsanm.7b00303 ACS Appl. Nano Mater. 2018, 1, 1265−1271