Next Article in Journal
SIRT1 Promotes Osteogenic Differentiation in Human Dental Pulp Stem Cells through Counteracting the Activation of STAT3
Previous Article in Journal
Changes of Crystal Structure and Microstructure of MoN Coatings in Accordance with Inductively Coupled Plasma Power
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigating the Nucleation of AlOx and HfOx ALD on Polyimide: Influence of Plasma Activation

by
Laura Astoreca
1,2,*,
David Schaubroeck
1,
Parinaz Saadat Esbah Tabaei
2,
Rouba Ghobeira
2,
Maaike Op de Beeck
1,
Rino Morent
2,
Herbert De Smet
1 and
Nathalie De Geyter
2
1
Centre for Microsystems Technology (CMST), imec and Ghent University, Technologiepark-Zwijnaarde 126, 9052 Ghent, Belgium
2
Research Unit Plasma Technology (RUPT), Department of Applied Physics, Faculty of Engineering and Architecture, Ghent University, Sint-Pietersnieuwstraat 41, B4, 9000 Ghent, Belgium
*
Author to whom correspondence should be addressed.
Coatings 2021, 11(11), 1352; https://doi.org/10.3390/coatings11111352
Submission received: 11 October 2021 / Revised: 25 October 2021 / Accepted: 30 October 2021 / Published: 3 November 2021
(This article belongs to the Section Thin Films)

Abstract

:
There is an increasing interest in atomic layer deposition (ALD) on polymers for the development of membranes, electronics, (3D) nanostructures and specially for the development of hermetic packaging of the new generation of flexible implantable micro-devices. This evolution demands a better understanding of the ALD nucleation process on polymers, which has not been reported in a visual way. Herein, a visual study of ALD nucleation on polymers is presented, based on the different dry etching speeds between polymers (fast) and metal oxides (slow). An etching process removes the polyimide with the nucleating ALD acting as a mask, making the nucleation features visible through secondary electron microscopy analyses. The nucleation of both Al2O3 and HfO2 on polyimide was investigated. Both materials followed an island-coalescence nucleation. First, local islands formed, progressively coalescing into filaments, which connected and formed meshes. These meshes evolved into porous layers that eventually grew to a full layer, marking the end of the nucleation. Cross-sections were analyzed, observing no sub-surface growth. This approach was used to evaluate the influence of plasma-activating polyimide on the nucleation. Plasma-induced oxygen functionalities provided additional surface reactive sites for the ALD precursors to adsorb and start the nucleation. The presented nucleation study proved to be a straightforward and simple way to evaluate ALD nucleation on polymers.

Graphical Abstract

1. Introduction

Atomic layer deposition (ALD) is a highly conformal thin film deposition technique. It consists of sequential self-limited reactions between two precursors and offers a precise control of the thickness of the layer [1,2]. A wide range of metals and metal oxides can be deposited by ALD, as compiled in the Atomic Limits database [3]. Thermal ALD processes operate at high temperatures, usually in the range of 200–400 °C [4]. Such high temperatures are not compatible with a large majority of polymers. Although there are strategies to lower the deposition temperature, such as using plasma-enhanced ALD (PEALD), herein the focus is on thermal ALD. The development of ALD is intimately related with the microelectronics industry, which in the past decades has been the main driving force of the advances in ALD technology. As such, ALD was first developed as a coating technology on solid inorganic surfaces [5,6,7], such as metals, glass or silicon wafers. At present, the reaction mechanisms during nucleation and growth of ALD materials on these surfaces are well understood. The recent evolution of the electronics and medical device industry into flexible and stretchable microelectronics boosted the interest of ALD depositions on polymeric substrates. However, at the present time, ALD nucleation mechanisms are not as well understood for depositions on polymers. ALD on polymeric substrates is widely used across many fields, such as surface modification of polymers [8,9,10], coatings [11,12,13], building 3D nanostructures (e.g., hollow fibers, hollow spheres, 3D meshes), filtration membranes [14], wearable electronics, displays, sensors, etc. [15,16,17,18,19,20,21]. However, one of the largest applications of ALD on polymers is the development of flexible glass diffusion [22,23,24,25,26,27,28], moisture [29,30,31,32,33] or copper [34,35] barriers for the microelectronics industry. Specifically, in recent years, ALD has been widely used in hermetic barriers for the packaging of implantable medical devices [36,37,38,39,40,41,42,43,44,45,46].
Compared to silicon or metal oxide surfaces, there are additional intricacies to the ALD deposition on polymers, resulting in more complex growth mechanisms during nucleation and control of the ALD process. Despite these differences, ALD processes on polymers also present a so-called nucleation stage [47]. The nucleation stage is the phase of early growth of the ALD material during which the growth per cycle (GPC) is not constant and a fully closed layer has not been formed yet. The nucleation stage finishes once a fully closed layer of the deposited material is obtained. It is affected by many variables, such as the ALD precursors, the nature of the substrate, the ALD process temperature or pressure, and exposure times to the precursors.
The initial growth and nucleation of ALD layers on polymers strongly depends on the chemistry of the polymer itself, although it is also influenced by the conditions of the ALD process and the chemistry and size of the ALD precursor molecules [48]. Chemically inert polymers do not have reactive groups at the surface to react with the ALD precursors and to initiate the ALD nucleation and growth. However, ALD precursor molecules can dissolve and/or diffuse into the polymer leading to subsurface growth of the ALD material. This type of growth, when promoted by the conditions of the deposition process and favored by the type of polymer, produces organic/inorganic hybrid subsurface layers and falls into the category of so-called sequential infiltration synthesis (SIS), sequential vapor infiltration (SVI), vapor phase infiltration (VPI), or multiple pulsed vapor phase infiltration (MPI) [49,50,51]. On the other hand, for polymers with functional groups inherently present in their chemical structure and available at the surface, the ALD nucleation and growth mostly occurs at the surface, obtaining an ALD film on top of the polymer and a distinct interface between both materials.
Functional groups can be easily introduced onto the surface of polymers by plasma activation. In the case of inert polymers, the plasma-induced functional groups can serve as reactive sites for the ALD nucleation and surface growth of a metal oxide layer, reducing or even suppressing the subsurface growth. In the case of polymers with functional groups already present on their surface, the plasma-induced functionalities can increase the number of reactive sites available at the surface for the nucleation [20,52,53,54,55,56].
The nucleation of ALD processes has been commonly studied by characterization techniques such as Fourier-transform infrared spectroscopy (FTIR), X-ray photoelectron spectroscopy (XPS), ellipsometry, quartz microbalance (QCM), both in and ex situ, but has also been investigated using other techniques such as atomic force microscopy (AFM), scanning tunneling microscopy (STM), X-ray diffraction (XRD), Rutherford backscattering spectroscopy (RBS), Auger electron spectroscopy (AES), or low energy ion scattering (LEIS) [57,58,59,60,61,62,63,64,65,66,67,68,69,70,71]. Each technique has its own advantages and disadvantages. In the case of ALD nucleation on polymers, for instance, Astaneh et al. studied the nucleation of TiO2 ALD on polydimethylsiloxane with XPS and X-ray absorption near edge structure (XANES). They also characterized TiO2 nucleation and growth via SEM/EDX, not visually, but by mapping the elemental composition of the samples [66]. On the other hand, Wilson et al. studied the nucleation of Al2O3 ALD on various polymers (polystyrene, polypropylene, poly(methyl methacrylate), polyethylene, polyvinyl chloride) via QCM [59]. All these techniques provide valuable information about changes in mass, thickness and/or optical properties, or about the concentration of species and their chemical state during the nucleation stage. Some techniques are very advanced, expensive and not easily available (LEIS), while others are limited in their depth of information, being too large for the analysis of the ultrathin ALD layer in its nucleation stage (FTIR). Others do provide precise information about the surface topography in the early nucleation stage, but they fail to detect the closure of a full layer at the end of the nucleation (AFM and STM).
In recent years, ALD has been widely used in hermetic packaging of medical devices, especially for the encapsulation of flexible devices. Traditional implantable medical devices such as pacemakers, cochlear implants or defibrillators have a rigid and rather large control unit. Nonetheless, the recent development of flexible and stretchable electronics has enabled the fabrication of miniaturized and flexible implantable medical devices [72,73,74,75,76,77]. In contrast to traditional rigid housings (sealed metal, ceramic and glass casings), the material selection for the encapsulation of flexible or soft devices is complex, being almost impossible to find all the requirements (biocompatible, biostable, hermetic, flexible) in a single material. To that end, polymers and ALD ceramic films can be combined, where the polymer provides a flexible support to the hermetic ALD films. The brittle ceramic films should be sufficiently thin (in the order of tens of nanometers) for the packaging to be flexible, which can be obtained by atomic layer deposition (ALD) [54,55,56,57,58]. This strategy has already been proven to provide long-term protection (several years) against water permeation at body temperature [78,79].
Polyimide is a high performance polymer, with excellent mechanical resistance, thermal and chemical stability, flexibility, excellent insulating properties and relatively good moisture barrier properties compared to other polymers [80,81,82,83]. This outstanding set of properties make polyimide suitable for a variety of applications ranging from electronics, medicine, membrane separation to aerospace and military industries [81,84]. Certain types of biocompatible polyimides are widely used in the packaging of implantable electronics and electrodes not only for its barrier performance [85,86,87,88,89,90,91,92,93] but also for its compatibility with microfabrication processes, being used as a structural component of devices [94,95,96,97].
Metal oxides such as Al2O3 and HfO2 are biocompatible, stable under the high temperatures of microfabrication and sterilization processes and, more importantly, excellent barriers, meeting the hermeticity requirement for encapsulating long-term implantable medical devices [98]. Polyimide and ALD deposited Al2O3/HfO2 can be combined to develop an ultra-hermetic package [78,79]. Yet, the adhesion between adjacent layers is critical for the package to retain its long-term hermeticity. If the adhesion between the layers is poor, moisture could accumulate at their interface, leading to fast lateral diffusion and an unsatisfactory performance of the barrier. When the adhesion is good, the moisture that permeates through the defects of a layer stops at the interface, unless the defects of the stacked layers are aligned. Even in such an unlikely case, since the barrier is formed by a multilayer stack, the moisture permeation would be interrupted when reaching the next interface. Thus, the nucleation and early growth of the ALD ceramic are key to the polymer/ALD interface since they determine the type of interaction between both materials. Very thin ceramic layers (5–10 nm) are deposited to obtain a hermetic encapsulation for implantable devices [29,30,31,32,33,78,79]. Although single ALD layers provide a good degree of hermeticity that could be enough for short term implantation, the combination of ALD layers provides a high degree of hermeticity as required for long-term implantable devices [78,79,91,99]. For such an application, it is crucial to ascertain the moment at which the ALD nucleation is finished, in other words, to define the minimum amount of ALD cycles necessary to obtain a fully closed and thus hermetic layer of the ceramic material.
Herein, a qualitative study of ALD nucleation on a polymeric substrate was conducted, using an innovative approach to visualize the nucleation of Al2O3 and HfO2 on a biocompatible polyimide. Selective reactive ion etching (RIE) of polyimide was applied using the nucleating ALD as a mask to reveal the features of ALD nucleation under microscopy inspection. The nucleation of Al2O3 and HfO2 was studied by field-emission-gun scanning electron microscopy (FEG–SEM) and by XPS analysis. Cross-sections of thick Al2O3 and HfO2 ALD layers were analyzed by scanning transmission electron microscopy (STEM) and energy-dispersive X-ray spectroscopy (EDS) to elucidate whether surface or subsurface nucleation and growth occurred. Lastly, the surface of polyimide was modified via non-thermal plasma activation under an air atmosphere to evaluate the influence of introducing reactive groups on the surface of polyimide on Al2O3 and HfO2 ALD nucleation.

2. Materials and Methods

2.1. Materials

Glass substrates were purchased from Präzisions Glas and Optik GmbH, (Iserlohn, Germany). Polyimides PI2611 and PI2610, both BPDA-PPD (3,3′,4′-biphenyltetracarboxylic dianhydride–p-phenylene diamine) polyimide, were purchased from HD Microsystems (Parlin, NJ, USA). Water vapor (99.999% purity) and trimethylaluminium (TMA) (99.999% purity) or tetrakis(dimethylamido)hafnium (TDMAH) (99.999% purity) from Strem Chemicals (Newburyport, MA, USA) were used as precursors for the ALD deposition, with N2 (Praxair, Belgium) as purge and carrier gas. The ALD precursors were purchased from Strem Chemicals. Dry air (Alphagaz 1) from Air Liquide (Paris, France) was used for the plasma activation.

2.2. Methods

2.2.1. Sample Preparation

Polyimide films were prepared by spin coating PI2611 and PI2610 onto 6.5 cm2 square borosilicate glass with a Polos SPIN200i (SPS Belgium, Herselt, Belgium) in two consecutive steps: 10 s at 1000 rpm and 60 s at 4000 rpm, at an acceleration rate of 500 rpm/s. The spin coating step was followed by a pre-baking step of 10 min at 200 °C on a hot plate and the samples were subsequently cured for 30 min at 200 °C and 60 min at 300 °C in an N2 atmosphere.
A selection of polyimide substrates was treated with a non-thermal plasma activation process prior to the ALD deposition to study the influence of the surface functionalization on ALD nucleation. The plasma activation was carried out in a home built dielectric barrier discharge (DBD) reactor, where the discharge occurred between two circular copper electrodes (5.5 cm diameter). The electrodes were covered with glass as dielectric material. The lower electrode was connected to ground through a 10.4 nF capacitor, while the upper electrode was connected to an AC high frequency power source (50 kHz). The substrates were placed on the lower electrode and fixed with double sided tape. The plasma activation of polyimide was carried out in a dry air atmosphere. First, the reactor was pumped down to at least 0.05 kPa. Then, air was introduced in the reactor at a flow rate of 1.0 standard litre per min (slm) and the pressure was raised to 5.0 kPa. Plasma was ignited at this moment, with a discharge power of 3.0 W. The samples were exposed to the air plasma for 1.0 s, a previously optimized parameter for BPDA-PPD polyimide [100].
All ALD depositions were carried out in a commercial reactor (Ultratech Savannah S200 G2, Cambridge NanoTech, Cambridge, MA, USA) at 150 °C and 53 kPa. The Al2O3 deposition process consisted of alternating pulses of TMA (0.03 s and 30 s N2 purge) and H2O (0.03 s and 30 s N2 purge). HfO2 was deposited by alternating pulses of TDMAH (0.4 s and 60 s N2 purge) and H2O (0.03 s and 60 s N2 purge). The GPC of the ALD processes was determined by measuring the thickness of layers deposited on witness Si wafers: 1.20 ± 0.02 Å/cycle for Al2O3 and 1.16 ± 0.01 Å/cycle for HfO2. The witness Si samples were analyzed with an FS-1EXs ellipsometer (FilmSense, Lincoln, NE, USA). The results were fitted with a Cauchy dispersion model (on Si) to obtain the thickness of the ALD films.
The reactive ion etching (RIE) process was performed in a Plasma Therm Batchtop VII (Belgium) etcher under a N2/O2 mixture atmosphere. The pressure in the chamber was kept constant at 20 Pa and the mixture flows were 5 sccm (standard cubic centimeter per minute) N2 and 20 sccm O2. The plasma was ignited at a power of 75 W and the samples were exposed to the etching process for 5 min.

2.2.2. Characterization

After the RIE step, the samples were inspected with a JEOL JSM-7000F FEG–SEM (JEOL, Tokyo, Japan) (acceleration voltage 2.7 kV) equipped with a secondary electron detector. Secondary electron images were recorded at five different spots randomly distributed in the center of the sample at 20,000× and 50,000× magnification. Prior to imaging, the samples were coated with a thin layer of gold (approximately 20 nm) using a plasma sputter coater. Two samples per condition were examined. All the images were automatically processed with Image J 1.52a software to estimate the area of polyimide covered by the nucleating ALD through particle counting. The particle analysis calculation was performed for the 20,000× magnification images, considering a minimum particle size threshold of 100 px2. The values reported per condition are the average of the images taken at five random spots, and two replicates per condition are investigated. The results of pristine and plasma-activated polyimide were analyzed with paired-samples t-test in SPSS 26.0 software, at a significance level of 0.05.
Cross-sections of polyimide/Al2O3 (40 nm)/polyimide and polyimide/HfO2 (50 nm)/polyimide stacks were analyzed with high angle annular dark field scanning transmission electron microscopy (HAADF-STEM) and EDS. Thin lamella were prepared on a Cu Omniprobe TEM grid, using a Thermo Fisher Helios focused ion beam secondary electron microscope (FIB-SEM, Thermo Fisher, Waltham, MA, USA). Prior to the FIB preparation the samples were coated with gold. The analyses were performed on a Tecnai Osiris TEM (FEI, Hillsboro, OR, USA), equipped with a super-X detector, operated at an acceleration voltage of 200 kV. One sample of Al2O3 ALD and one sample of HfO2 ALD were analyzed.
In addition to FEG–SEM inspection, the surface chemical composition was studied by XPS analysis. The measurements were performed using a PHI 5000 Versaprobe II (ULVAC-Physical Electronics, Chigasaki, Japan) spectrometer with an Al Kα X-ray source (hν = 1486.6 eV) operating at 25 W. The photoelectrons were detected with a hemispherical analyzer at 45° to the normal of the samples’ surface and the pressure in the chamber was maintained below 10−6 Pa. Survey scans were recorded with a pass energy of 187.85 eV (0.8 eV step) while high resolution spectra (C1s, O1s, Al2p, Hf4f) were recorded at 23.5 eV (0.1 eV step). The spectra were analyzed using Multipak software (v 9.6.1), applying an iterated Shirley background and correcting the final peak intensities by the relative sensitivity factors provided by the manufacturer. All the spectra were calibrated (C-C at 285.0 eV) and a Savitzky-Golay smoothing procedure was applied prior to any analysis. The instrumental error is 1%. Two samples per condition were analyzed at four randomly located points in the center of the sample. The values reported are the average and standard deviation of these 8 measurements.

3. Results and Discussion

3.1. Introduction

Instead of an analytical, quantitative approach, herein we propose a visual approach to study the ALD nucleation on polymers. Polymers are organic materials with much faster dry etching (also known as RIE) speeds compared to metal oxides. The approach in this work is based on this difference in etching speed between polyimide and the ALD metal oxide. A closed metal oxide layer has yet to be formed for the nucleation phase of the ALD deposition on the polyimide, so the local metal oxide sites act as a hard mask for the underlying polymer when subjected to RIE. This means that the polyimide, which is not covered by the nucleating metal oxide, is etched away to a certain degree. The RIE step is key for visualizing the features of the ALD nucleation under SEM. The topographical differences between the polymeric substrate and the ultrathin nucleating ALD material are very small, making it impossible to distinguish the ALD material from the polyimide substrate. The RIE process, etching the polyimide not covered by the metal oxide, increases the topographical difference between the polymer and the metal oxide, enabling the analysis of the ALD nucleation behavior under an SEM. The process is schematically represented in Figure 1. This approach allows to visualize and observe the progression of the nucleation on polymers; however, it is not possible to discern whether surface or subsurface growth occurs, as only a top-view image is obtained with SEM inspection. Subsurface growth and nucleation could be inspected by analyzing a cross-section of the polymer/ALD layers under STEM combined with EDS analysis.

3.2. ALD Nucleation Study of Al2O3 and HfO2 on Polyimide

Using an optimized RIE process of 5 min, the evolution of Al2O3 and HfO2 nucleation on polyimide was studied by depositing an increasing amount of ALD cycles on a set of substrates followed by FEG-SEM and XPS analyses. Prior to studying the ALD nucleation on polyimide, the surface of pristine polyimide, polyimide after the RIE process and polyimide with a fully closed (c.a. 100 nm) Al2O3 layer on top after RIE were inspected (see Figure 2). As expected, the surface of pristine spin coated polyimide was smooth. After RIE, the surface topography and morphology of polyimide clearly changed, a roughness commonly referred to as grass-like roughness was created as a consequence of the etching process [101]. The surface topography of polyimide with a closed ALD layer on top (Al2O3 100 nm) was the same as that of pristine polyimide, due to the high conformality of ALD depositions combined with the high resistance of the thick AlOx layer to the RIE process.
Figure 3 shows the progress of Al2O3 and HfO2 nucleation on polyimide. The features of the grass-like roughness of polyimide and of the nucleating ALD were also visibly different, especially at higher magnifications, making it possible to differentiate the underlying polyimide (dark grey) from the nucleating ALD (light grey). As the nucleation progressed, the grass-like roughness of polyimide was no longer visible. The features of the nucleating ALD differed considerably from one ALD material to another, to be expected since the adsorption and reaction of the precursors is different for each ALD process.
At five cycles, islands were observed for both Al2O3 and HfO2 homogeneously distributed over the sample; however, the islands were considerably smaller for HfO2. In the case of Al2O3, the islands were not only larger, but also presented branches. At seven cycles, Al2O3 continued to nucleate on local sites over the sample, forming more islands. Some of them started to connect one to another, forming some sort of filaments. HfO2 nucleation progressed in the same way, but the features were more easily distinguishable as they did not present as many small branches as for Al2O3. HfO2 nucleated on more sites on polyimide, and neighboring islands merged into very well defined filaments. At 10 cycles, the difference between Al2O3 and HfO2 were even more obvious. For Al2O3, barely any change compared to seven cycles could be appreciated, but the filaments observed at seven cycles in HfO2 samples had connected with each other, forming meshes. At 13 cycles, the nucleating Al2O3 had formed an incipient mesh structure, while for HfO2 a porous layer was already formed. From 13 to 15 cycles, Al2O3 nucleation progressed substantially and grew into a porous layer. The HfO2 film presented visibly smaller pores at 15 cycles. For Al2O3, the pores were round and homogeneous in shape and size. On the other hand, the pores of HfO2 had various sizes and shapes, more elongated rather than round. Eventually, at 20 cycles, a fully closed layer was obtained for both ALD processes, marking the end of the nucleation stage.
Despite the differences, the same phases could be identified in the nucleation of both materials. Initially, Al2O3 and HfO2 ALD processes nucleated on specific sites homogeneously distributed over the polyimide substrate, forming isolated islands. As more cycles were being deposited, the islands grew into filaments. There are two possible routes of formation of these filaments: (i) Precursor molecules adsorbed on reactive sites that were still available forming islands next to the already existing ones until these coalesced transforming into the filamentous shapes; (ii) Lateral growth of the islands occurred through random growth of Al2O3 or HfO2 on the already grown islands rather than through the adsorption of the precursors on polyimide. Via visual inspection, it is difficult to determine which of the two routes occurred, but most likely both mechanisms contributed to the early stage nucleation. These filaments grew with each ALD cycle and coalesced forming mesh structures, which progressively gained density leading to the formation of porous films. Eventually, the pores closed and a continuous, pin-hole free layer was formed at 20 cycles, both for Al2O3 and HfO2. At the last stages of the nucleation, the growth mechanisms were dominated by the second mechanism of lateral growth on the already deposited Al2O3 and HfO2, as the likelihood of reactive sites on the polyimide to be vacant or even physically available was rather low. These observations clearly pointed to a nucleating mechanism based on the adsorption of precursors at reactive sites of the polyimide and further lateral growth, commonly referred to as island-coalescence nucleation or growth [34,65,70,71,102,103]. Island growth of Al2O3 on H-terminated Si was already observed by selective etching of SiOx through the “defects” of the nucleating Al2O3 layer [104].
The recorded FEG–SEM images were processed with Image J software to estimate the area of polyimide covered by the nucleating ALD, as a way to evaluate the evolution of the nucleation in a qualitatively but numerical manner. Figure 4 shows how the ALD coverage increased as the nucleation progressed, that is, as more cycles were being deposited. Although the features of the nucleating ALD were different for Al2O3 and HfO2, the progress of ALD coverage of polyimide was similar, as depicted in Figure 4. At five cycles, the coverage was estimated to be 24.3 ± 0.4% for Al2O3 and 22.0 ± 2.4% for HfO2 during the islands stage. At seven cycles, the coverage increased to 34.4 ± 0.6% (Al2O3) and 32.3 ± 2.5% (HfO2). At 10 and 13 cycles, the coverage of the nucleating ALD was higher for HfO2 (42.9 ± 2.3% and 55.7 ± 2.2%, respectively) than for Al2O3 (34.5 ± 1.0% and 48.9 ± 2.2%, respectively), but both presented a similar value after 15 cycles (74.7 ± 6.2% for Al2O3, 74.1 ± 2.9% for HfO2).
An approximate range of the area of polyimide covered by the nucleating ALD could be derived for each stage of the nucleation based on these results. During the islands phase, the coverage was within 0–30%, increasing to 31–40% when filaments were formed. The coalescence of the filaments into meshes corresponded to a coverage of 41–50%, while the coverage of the porous layers was within a broader range, namely 51–99%.
In addition to the visual study of the nucleation, XPS analyses were also performed. The surface concentrations of Al and Hf (Figure 5a) were evaluated over an increasing amount of Al2O3 and HfO2 ALD cycles, respectively. The concentration of both elements progressively increased as more cycles were being deposited, following the same trend as the ALD coverage values. For Al2O3, at the very early nucleation (5–10 cycles) the Al concentration barely changed, and remained within 6–8 at.%, corresponding to the island/filament stages, as observed in Figure 3. Conversely, the concentration of Hf increased after each amount of cycles: 3.0 ± 0.1 at.% at five cycles (island stage), 4.6 ± 0.1 at.% at seven cycles (filament stage) and 6.9 ± 0.1 at.% at 10 cycles (mesh stage). The Al content went up to 10.8 ± 0.3 at.% as the Al2O3 ALD reached the mesh stage of the nucleation, and to 11.1 ± 0.2 at.% at 15 cycles (porous film stage). The Hf content also increased as the nucleation continued into the formation of porous HfO2 layers (10.2 ± 0.3 at.% at 13 cycles, 11.3 ± 0.3 at.% at 15 cycles). The Al and Hf concentrations increased even after the Al2O3 and HfO2 nucleation finished (after 20 cycles): 13.8 ± 0.1 at.% Al, 14.4 ± 0.3 at.% Hf at 20 cycles, 19.2 ± 0.1 at.% Al, 15.9 ± 0.4 at.% Hf at 30 cycles and 27.0 ± 0.5 at.% Al, 17.0 ± 0.5 at.% Hf at 40 cycles. The Al and Hf concentrations would be expected to become constant once the nucleation finished; however, the thickness of the deposited ALD layers after such few cycles (below 40 cycles) is in the order of 1–2 nm at the GPC of the ALD processes (1.20 Å/cycle for Al2O3 and 1.16 Å/cycle for HfO2). The depth of information of the XPS technique at the conditions that the measurements were performed is approximately 5–10 nm for polymer surfaces. Therefore, if a thin ALD layer of 1–2 nm deposited on polyimide is being measured, the signal of the underlying polyimide will still be present and considerably affect the elemental composition that is measured. Once the nucleation is finished and more cycles are deposited, the thicker the Al2O3 or HfO2 layer becomes, and the lower the signal of the underlying polyimide is, until only the signal of the ALD layers is being detected and a constant Al or Hf concentration is observed (not shown).
Figure 5b–e shows how the high resolution O1s, Al2p and Hf4f spectra changed as Al2O3 or HfO2 cycles, respectively, were deposited on polyimide. The peak fitting of the O1s, Al2p and Hf4f high resolution spectra after 5 and 40 cycles are shown in Figures S1 and S2 of the Supplementary Materials. The intensity of the O1s peaks progressively increased with increasing cycles of Al2O3 and HfO2 deposition. The oxygen ratio in polyimide (c.a. 10%) is much lower than in Al2O3 (c.a. 60%) or HfO2 (c.a. 67%), thus, the O1s signal became more intense as the ALD layers were nucleating and growing. From five up to 40 ALD cycles, the O1s peak shifted approximately 1.0 eV towards lower binding energies (from 532.2 eV to 531.4 eV for Al2O3 and from 532.4 eV to 530.7 eV for HfO2), as a consequence of the growing O–Al or O–Hf component of Al2O3 and HfO2, respectively. The width of the O1s peak also progressively increased due to the growing contribution of the O–Al/O–Hf component. This was particularly evident in the case of HfO2, for which the polyimide and metal oxide components were more clearly distinguished. The intensity of the Al2p and Hf4f peaks also increased as more cycles were being deposited. The Hf4f peak presented the characteristic spin orbits of HfO2 at 17.8 eV and 19.5 eV. While no shift was observed on the Hf4f peak, the Al2p peak presented a small shift towards lower binding energies of approximately 0.7 eV from five cycles to 40 cycles, that could be ascribed to the change from CO–Al at the polyimide/Al2O3 interface to the bulk O–Al of Al2O3.
In conclusion, it can be stated that the XPS and the FEG–SEM imaging information presented parallel trends and matching results. This agreement confirmed that the RIE process did not significantly damage the deposited ALD, as otherwise there would have been a considerable disagreement between the observations from both techniques. This proved that the visualization approach was a reliable way of studying the nucleation of ALD layers on polymeric substrates. However, as already mentioned in the introduction section, the approach is limited to a top-view imaging of the nucleating process, not being possible to identify whether surface or sub-surface ALD growth is occurring.
For that reason, cross-sections of polyimide/Al2O3 (40 nm)/polyimide and polyimide/HfO2 (50 nm)/polyimide stacks were analyzed with HAADF-STEM and EDS. Figure 6 shows the analysis of Al2O3 and HfO2 cross-sections. The HAADF-STEM images show the Al2O3 or HfO2 ALD layer sandwiched between the thicker polyimide layers. Polyimide is an oxygen-containing polymer; therefore, the oxygen signal (in blue) is present through the whole area of analysis in the EDS maps, but it is much more intense on the area corresponding to the Al2O3 and HfO2 layers, where the respective Al or Hf are also present. There were no signs of sub-surface growth of the ALD, as a sharp interface between polyimide and the ALD layers was observed. These observations thus imply that the nucleation and growth of both Al2O3 and HfO2 ALD processes occurred at the surface of polyimide only. The polyimide used in this work (BPDA-PPD polyimide) inherently contains reactive groups (C=O, C–O, N–C=O), which are consequently available at the surface as reactive sites for the nucleation during the ALD process. It seems that the nucleation started by the adsorption of precursor molecules at those reactive sites.

3.3. Influence of Plasma Activation on the ALD Nucleation

Previous research has shown that surface functionalization of the substrate prior to ALD deposition provides (additional) reactive sites during ALD nucleation [44,45,46,47,48,49]. The HAADF-STEM cross-section analysis revealed that no sub-surface growth occurred as pristine BPDA-PPD polyimide has reactive groups available for the nucleation of Al2O3 and HfO2. Here, the influence of plasma-activating polyimide prior to the ALD deposition on the nucleation behavior was explored. The additional reactive sites introduced by the plasma activation can increase the formation of covalent bonds between polyimide and Al2O3 or HfO2 during the nucleation. A higher chemical interaction between both layers would increase their adhesion in a multilayer barrier.
Polyimide surface was activated under air plasma and an increasing amount of Al2O3 or HfO2 ALD cycles was deposited on it. Figure 7 and Figure 8 show the evolution of Al2O3 and HfO2 nucleation, respectively, on both plasma-activated and pristine polyimide. The surface characterization of the activated polyimide with the plasma process used in this work has been described in detail elsewhere [64]. The air plasma activation introduced oxygen-containing groups such as C–O, C=O and O–C=O on the surface of polyimide, increasing the surface oxygen content by approximately 10 at.%. These oxygen functional groups were found to act as nucleating sites for both Al2O3 and HfO2 ALD processes, as evidenced by the images shown in Figure 7 and Figure 8.
The features of the nucleating ALD on plasma-activated polyimide were visible on a more advanced nucleation stage compared with pristine polyimide. In Figure 4 it can also be observed that Al2O3 and HfO2 ALD coverage was higher on plasma-activated polyimide, except for 15 cycles of Al2O3. In the case of Al2O3, the influence of the plasma activation was not so evident at lower amounts of cycles (see Figure 7). At five and seven cycles, it was difficult to distinguish differences in the FEG–SEM images of plasma-activated and pristine polyimide. The islands were slightly larger and filaments were starting to form on the plasma-activated polyimide. The coverage of Al2O3 ALD was almost the same on plasma-activated polyimide as on pristine polyimide after seven cycles, 34.2 ± 1.2% and 34.5 ± 1.0%, respectively. On the other hand, the coverage after five cycles was significantly higher on plasma-activated polyimide (25.5 ± 1.0% versus 24.3 ± 0.4%). As more cycles were deposited, differences between Al2O3 nucleation on plasma-activated and pristine polyimide became more patent. After 10 and 13 cycles, Al2O3 had already formed a mesh on plasma-activated polyimide while it was only beginning to do so after 13 cycles on pristine polyimide. Likewise, the coverage of 10 (from 34.5 ± 1.0% to 44.5 ± 0.8%) and 13 cycles (from 48.9 ± 2.2% to 56.9 ± 1.3%) also significantly increased on plasma-activated polyimide (see Figure 4). After 15 cycles, a porous Al2O3 layer had been formed on both plasma-activated and pristine polyimide. The pores on plasma-activated polyimide were smaller, and of different sizes, as they were being progressively closed. The coverage after 15 cycles was very similar, 70.0 ± 4.5% on plasma-activated polyimide and 74.7 ± 6.2% on pristine polyimide. A fully closed, pin-hole free layer was formed at 20 cycles on either plasma-activated or pristine polyimide.
The differences of the nucleation on plasma-activated and pristine polyimide were more obvious in the case of HfO2 ALD. While only HfO2 islands were formed after five cycles on pristine polyimide, filaments were already visible on plasma-activated polyimide (see Figure 8), reaching a significantly higher coverage (27.0 ± 1.1% compared with 22.0 ± 2.4%). After a few more cycles, at seven cycles, HfO2 had grown into a mesh on plasma-activated polyimide, while only filaments were observed on pristine polyimide. As more HfO2 cycles were deposited, the nucleation on plasma-activated polyimide remained in a more advanced stage, showing a denser mesh at 10 cycles, and less porous layers at 13 and 15 cycles. HfO2 coverage (see Figure 4) was also considerably and significantly higher on plasma-activated polyimide: 45.6 ± 0.9% versus 42.9 ± 2.3% (10 cycles), 68.6 ± 2.8% versus 55.7 ± 2.2% (13 cycles) and 86.6 ± 4.5% versus 74.1 ± 2.9% (15 cycles). After 20 cycles a fully closed HfO2 layer had been formed on both plasma-activated and pristine polyimide.
The marked improvement of HfO2 nucleation on plasma-activated polyimide, in comparison to Al2O3 nucleation on plasma-activated polyimide, denotes a high influence of the surface composition of the substrate on the HfO2 ALD process. TDMAH nucleation mechanisms and early growth seem more susceptible to the surface composition of the substrate, that is, to the availability of reactive sites. This could be related to the nucleation and early growth mechanisms of both materials. Both Al2O3 and HfO2 initially nucleate on the reactive sites at the surface of polyimide, but, if after the very first cycles, TMA molecules preferentially react with the already grown clusters of Al2O3, further growth of Al2O3 would not be considerably influenced by the availability of surface reactive sites. Whereas, if HfO2 nucleation is mostly driven by the chemisorption of TDMAH molecules on surface reactive sites of polyimide, it would be largely influenced by the availability of these reactive sites.
The coverage of the nucleating ALD was higher on plasma-activated polyimide, and the FEG–SEM imaging revealed a more advanced stage of the nucleation on plasma-activated polyimide than on pristine polyimide. However, despite these observations, Al2O3 and HfO2 nucleations were finished at 20 cycles on both plasma-activated and pristine polyimide. This does not necessarily mean that the number of cycles needed to complete the nucleation stage is not affected, but rather that 15–20 cycles is a wide range and the exact number of cycles to obtain a pin-hole free Al2O3 or HfO2 layer on plasma-activated polyimide was not determined. Based on the observed results, the availability of functional groups at the surface of polyimide influences the nucleation, not in terms of speed, but in terms of the chemical interaction between the ALD layers and the polymer. A higher concentration of functional groups at the surface of the polymer provides more reactive sites for the nucleation, increasing the amount of covalent bonds formed between the polymer and ALD layers. The oxygen groups introduced by the plasma activation acted as reactive sites for the nucleation, resulting in an increased chemical linkage between the successive ALD layer and the polymeric substrate. This chemical interaction can result in a better adhesion between the multiple layers of hermetic packaging based on polymers and ALD metal oxides [105].

4. Conclusions

In this research, an approach to visualize the nucleation of ALD processes on polymeric substrates was introduced. It is based on the dry etch selectivity between the deposited ALD metal oxide and the polymer substrate under an RIE process.
The approach was successfully applied to study the nucleation of Al2O3 and HfO2 ALD on polyimide, in the framework of developing flexible and hermetic barriers for the packaging of implantable devices. Al2O3 and HfO2 ALD processes produced a full, pin-hole free layer on polyimide after 20 cycles. Both processes presented an island-coalescence nucleation model, that is, the nucleation started at reactive sites forming islands and further lateral growth led to coalescence of the islands through the progressive formation of filaments-meshes-porous layers, and eventually pin-hole free layers. This evolution could also be observed in the surface Al and Hf content, and in the O1s, Al2p and Hf4f high resolution XPS spectra.
Cross-sections of thick Al2O3 (40 nm) and HfO2 (50 nm) layers embedded in polyimide were analyzed by HAADF-STEM and EDS, revealing no sign of sub-surface growth. This confirmed the initial surface nucleation of the ALD processes at the reactive groups inherently present on polyimide.
For the envisioned application of the barriers, the adhesion at the polymer/ALD interface is key. For that reason, the beneficial influence of plasma activating the surface of polyimide prior to the ALD depositions was explored. The oxygen groups introduced by the plasma activation (C–O, C=O, O–C=O) acted as additional reactive sites for the adsorption of the precursor molecules, as evidenced by the FEG–SEM images that revealed a more advanced stage of the nucleation on plasma-activated polyimide than on pristine polyimide, for the same number of cycles. This ensured the formation of additional covalent bonds between the polymer and the successive ALD layers.
The presented results can be considered as one step forward in order to obtain a better insight and understanding of the nucleation behavior of ALD metal oxides on a biocompatible polyimide. However, the exact influence on the improvement of the barrier properties of metal oxide-polymer stacks still needs to be determined.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/coatings11111352/s1. Figure S1: Peak fitting of the O1s and Al2p high resolution spectra after 5 and 40 ALD cycles of Al2O3. Figure S2: Peak fitting of the O1s and Hf4f high resolution spectra after 5 and 40 ALD cycles of HfO2.

Author Contributions

Conceptualization, L.A. and D.S.; Formal analysis, L.A.; Funding acquisition, H.D.S. and N.D.G.; Investigation, L.A., P.S.E.T. and R.G.; Methodology, L.A.; Resources, M.O.d.B., R.M., H.D.S. and N.D.G.; Supervision, D.S., M.O.d.B., H.D.S. and N.D.G.; Validation, L.A.; Visualization, L.A.; Writing–original draft, L.A.; Writing–review & editing, D.S., P.S.E.T., R.G., M.O.d.B., R.M., H.D.S. and N.D.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research study was supported by a research grant (G086717N) from the Research Foundation Flanders (FWO).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Leskelä, M.; Niinistö, J.; Ritala, M. 4.05—Atomic Layer Deposition. In Comprehensive Materials Processing; Hashmi, S., Batalha, G.F., van Tyne, C.J., Yilbas, B., Eds.; Elsevier: Oxford, UK, 2014; pp. 101–123. [Google Scholar]
  2. Knoops, H.C.M.; Potts, S.E.; Bol, A.A.; Kessels, W.M.M. 27—Atomic Layer Deposition. In Handbook of Crystal Growth, 2nd ed.; Kuech, T.F., Ed.; North-Holland: Boston, MA, USA, 2015; pp. 1101–1134. [Google Scholar]
  3. Atomic Limits. Available online: https://www.atomiclimits.com/alddatabase/ (accessed on 1 October 2021).
  4. Pakkala, A.; Putkonen, M. Chapter 8—Atomic Layer Deposition. In Handbook of Deposition Technologies for Films and Coatings, 3rd ed.; Martin, P.M., Ed.; William Andrew Publishing: Boston, MA, USA, 2010; pp. 364–391. [Google Scholar]
  5. Pessa, M.; Mäkelä, R.; Suntola, T. Characterization of surface exchange reactions used to grow compound films. Appl. Phys. Lett. 1981, 38, 131–132. [Google Scholar] [CrossRef]
  6. Tanninen, V.P.; Oikkonen, M.; Tuomi, T.O. X-ray diffraction study of thin electroluminescent ZnS films grown by atomic layer epitaxy. Phys. Status Solidi (A) 1981, 67, 573–583. [Google Scholar] [CrossRef]
  7. Busse, W.; Gumlich, H.E.; Törnqvist, R.O.; Tanninen, V.P. Zero-phonon lines in electroluminescence and photoluminescence of ZnS:Mn thin films grown by atomic layer epitaxy. Phys. Status Solidi (A) 1983, 76, 553–558. [Google Scholar] [CrossRef]
  8. Napari, M.; Malm, J.; Lehto, R.; Julin, J.; Arstila, K.; Sajavaara, T.; Lahtinen, M. Nucleation and growth of ZnO on PMMA by low-temperature atomic layer deposition. J. Vac. Sci. Technol. A 2014, 33, 01A128. [Google Scholar] [CrossRef] [Green Version]
  9. Kemell, M.; Färm, E.; Ritala, M.; Leskelä, M. Surface modification of thermoplastics by atomic layer deposition of Al2O3 and TiO2 thin films. Eur. Polym. J. 2008, 44, 3564–3570. [Google Scholar] [CrossRef]
  10. Hyde, G.K.; Scarel, G.; Spagnola, J.C.; Peng, Q.; Lee, K.; Gong, B.; Roberts, K.G.; Roth, K.M.; Hanson, C.A.; Devine, C.K.; et al. Atomic Layer Deposition and Abrupt Wetting Transitions on Nonwoven Polypropylene and Woven Cotton Fabrics. Langmuir 2010, 26, 2550–2558. [Google Scholar] [CrossRef]
  11. Pessoa, R.S.; Santos, V.P.d.; Cardoso, S.B.; Doria, A.C.O.C.; Figueira, F.R.; Rodrigues, B.V.M.; Testoni, G.E.; Fraga, M.A.; Marciano, F.R.; Lobo, A.O.; et al. TiO2 coatings via atomic layer deposition on polyurethane and polydimethylsiloxane substrates: Properties and effects on C. albicans growth and inactivation process. Appl. Surf. Sci. 2017, 422, 73–84. [Google Scholar] [CrossRef]
  12. Lindholm, N.F.; Zhang, J.; Minton, T.K.; O’Patchen, J.; George, S.M.; Groner, M.D. Protection of polymers from the space environment by atomic layer deposition. In AIP Conference Proceedings; American Institute of Physics: College Park, MD, USA, 2009; pp. 407–418. [Google Scholar]
  13. Cooper, R.; Upadhyaya, H.P.; Minton, T.K.; Berman, M.R.; Du, X.; George, S.M. Protection of polymer from atomic-oxygen erosion using Al2O3 atomic layer deposition coatings. Thin Solid Film. 2008, 516, 4036–4039. [Google Scholar] [CrossRef]
  14. Weber, M.; Julbe, A.; Kim, S.S.; Bechelany, M. Atomic layer deposition (ALD) on inorganic or polymeric membranes. J. Appl. Phys. 2019, 126, 041101. [Google Scholar] [CrossRef]
  15. Sweet, W.J.; I, I.I.; Oldham, C.J.; Parsons, G.N. Conductivity and touch-sensor application for atomic layer deposition ZnO and Al:ZnO on nylon nonwoven fiber mats. J. Vac. Sci. Technol. A Vac. Surf. Film. 2015, 33, 01A117. [Google Scholar] [CrossRef]
  16. Sun, C.; Zhang, J.; Zhang, Y.; Zhao, F.; Xie, J.; Liu, Z.; Zhuang, J.; Zhang, N.; Ren, W.; Ye, Z.-G. Design and fabrication of flexible strain sensor based on ZnO-decorated PVDF via atomic layer deposition. Appl. Surf. Sci. 2021, 562, 150126. [Google Scholar] [CrossRef]
  17. Ras, R.H.A.; Kemell, M.; de Wit, J.; Ritala, M.; ten Brinke, G.; Leskelä, M.; Ikkala, O. Hollow Inorganic Nanospheres and Nanotubes with Tunable Wall Thicknesses by Atomic Layer Deposition on Self-Assembled Polymeric Templates. Adv. Mater. 2007, 19, 102–106. [Google Scholar] [CrossRef] [Green Version]
  18. Oldham, C.J.; Gong, B.; Spagnola, J.C.; Jur, J.S.; Senecal, K.J.; Godfrey, T.A.; Parsons, G.N. Atomic layer deposition on polymers: Applications to physical encapsulation of electrospun nylon nanofibers. ECS Trans. 2010, 33, 279–290. [Google Scholar] [CrossRef]
  19. Lee, J.Y.; Shin, C.M.; Heo, J.H.; Kim, C.R.; Park, J.H.; Lee, T.M.; Ryu, H.; Son, C.S.; Shin, B.C.; Lee, W.J. Effects of O2 plasma pre-treatment on ZnO thin films grown on polyethersulfone substrates at various deposition temperatures by atomic layer deposition. Curr. Appl. Phys. 2010, 10, S290–S293. [Google Scholar] [CrossRef]
  20. Heo, J.H.; Ryu, H.; Lee, W.J. Effect of O2 plasma pretreatment on structural and optical properties of ZnO films on PES substrate by atomic layer deposition. J. Ind. Eng. Chem. 2013, 19, 1638–1641. [Google Scholar] [CrossRef]
  21. Choi, S.W.; Park, J.Y.; Lee, C.; Lee, J.G.; Kim, S.S. Synthesis of highly crystalline hollow TiO2 fibers using atomic layer deposition on polymer templates. J. Am. Ceram. Soc. 2011, 94, 1974–1977. [Google Scholar] [CrossRef]
  22. Vähä-Nissi, M.; Pitkänen, M.; Salo, E.; Kenttä, E.; Tanskanen, A.; Sajavaara, T.; Putkonen, M.; Sievänen, J.; Sneck, A.; Rättö, M.; et al. Antibacterial and barrier properties of oriented polymer films with ZnO thin films applied with atomic layer deposition at low temperatures. Thin Solid Film. 2014, 562, 331–337. [Google Scholar] [CrossRef]
  23. Matsumae, T.; Dushatinski, T.; Abdel-Fattah, T.M.; Suga, T.; Zhang, K.; Chen, X.; Baumgart, H. Room Temperate Bonding of Al2O3 Layers by Atomic Layer Deposition on Polyimide Substrates. ECS Trans. 2015, 69, 99–105. [Google Scholar] [CrossRef]
  24. Kääriäinen, T.O.; Maydannik, P.; Cameron, D.C.; Lahtinen, K.; Johansson, P.; Kuusipalo, J. Atomic layer deposition on polymer based flexible packaging materials: Growth characteristics and diffusion barrier properties. Thin Solid Film. 2011, 519, 3146–3154. [Google Scholar] [CrossRef]
  25. Groner, M.D.; George, S.M.; McLean, R.S.; Carcia, P.F. Gas diffusion barriers on polymers using Al2O3 atomic layer deposition. Appl. Phys. Lett. 2006, 88, 1–3. [Google Scholar] [CrossRef]
  26. Ferrari, S.; Perissinotti, F.; Peron, E.; Fumagalli, L.; Natali, D.; Sampietro, M. Atomic layer deposited Al2O3 as a capping layer for polymer based transistors. Org. Electron. 2007, 8, 407–414. [Google Scholar] [CrossRef]
  27. Ferguson, J.D.; Weimer, A.W.; George, S.M. Atomic Layer Deposition of Al2O3 Films on Polyethylene Particles. Chem. Mater. 2004, 16, 5602–5609. [Google Scholar] [CrossRef]
  28. Chawla, V.; Ruoho, M.; Weber, M.; Chaaya, A.A.; Taylor, A.A.; Charmette, C.; Miele, P.; Bechelany, M.; Michler, J.; Utke, I. Fracture mechanics and oxygen gas barrier properties of Al2O3/ZnO nanolaminates on PET deposited by atomic layer deposition. Nanomaterials 2019, 9, 88. [Google Scholar] [CrossRef] [Green Version]
  29. Langereis, E.; Creatore, M.; Heil, S.B.S.; van de Sanden, M.C.M.; Kessels, W.M.M. Plasma-assisted atomic layer deposition of Al2O3 moisture permeation barriers on polymers. Appl. Phys. Lett. 2006, 89, 081915. [Google Scholar] [CrossRef]
  30. Carcia, P.F.; McLean, R.S.; Walls, D.J.; Reilly, M.H.; Wyre, J.P. Effect of early stage growth on moisture permeation of thin-film Al2O3 grown by atomic layer deposition on polymers. J. Vac. Sci. Technol. A Vac. Surf. Film. 2013, 31, 061507. [Google Scholar] [CrossRef]
  31. Carcia, P.F.; McLean, R.S.; Reilly, M.H.; Groner, M.D.; George, S.M. Ca test of Al2O3 gas diffusion barriers grown by atomic layer deposition on polymers. Appl. Phys. Lett. 2006, 89, 031915. [Google Scholar] [CrossRef] [Green Version]
  32. Carcia, P.F.; McLean, R.S.; Reilly, M.H. Permeation measurements and modeling of highly defective Al2O3 thin films grown by atomic layer deposition on polymers. Appl. Phys. Lett. 2010, 97, 221901. [Google Scholar] [CrossRef]
  33. Carcia, P.F.; McLean, R.S.; Li, Z.G.; Reilly, M.H.; Marshall, W.J. Permeability and corrosion in ZrO2/Al2O3 nanolaminate and Al2O3 thin films grown by atomic layer deposition on polymers. J. Vac. Sci. Technol. A Vac. Surf. Film. 2012, 30, 041515. [Google Scholar] [CrossRef]
  34. Hoyas, A.M.; Schuhmacher, J.; Shamiryan, D.; Waeterloos, J.; Besling, W.; Celis, J.P.; Maex, K. Growth and characterization of atomic layer deposited WC0.7N0.3 on polymer films. J. Appl. Phys. 2004, 95, 381–388. [Google Scholar] [CrossRef]
  35. Elam, J.W.; Wilson, C.A.; Schuisky, M.; Sechrist, Z.A.; George, S.M. Improved nucleation of TiN atomic layer deposition films on SiLK low-k polymer dielectric using an Al2O3 atomic layer deposition adhesion layer. J. Vac. Sci. Technol. B Microelectron. Nanometer Struct. 2003, 21, 1099–1107. [Google Scholar] [CrossRef]
  36. Song, E.; Li, R.; Jin, X.; Du, H.; Huang, Y.; Zhang, J.; Xia, Y.; Fang, H.; Lee, Y.K.; Yu, K.J.; et al. Ultrathin Trilayer Assemblies as Long-Lived Barriers against Water and Ion Penetration in Flexible Bioelectronic Systems. ACS Nano 2018, 12, 10317–10326. [Google Scholar] [CrossRef]
  37. Jeong, J.; Laiwalla, F.; Lee, J.; Ritasalo, R.; Pudas, M.; Larson, L.; Leung, V.; Nurmikko, A. Conformal Hermetic Sealing of Wireless Microelectronic Implantable Chiplets by Multilayered Atomic Layer Deposition (ALD). Adv. Funct. Mater. 2019, 29, 1806440. [Google Scholar] [CrossRef]
  38. Peron, M.; Cogo, S.; Bjelland, M.; Afif, A.B.; Dadlani, A.; Greggio, E.; Berto, F.; Torgersen, J. On the evaluation of ALD TiO2, ZrO2 and HfO2 coatings on corrosion and cytotoxicity performances. J. Magnes. Alloy. 2021, 9, 1806–1819. [Google Scholar] [CrossRef]
  39. Xie, X.; Rieth, L.; Merugu, S.; Tathireddy, P.; Solzbacher, F. Plasma-assisted atomic layer deposition of Al2O3 and parylene C bi-layer encapsulation for chronic implantable electronics. Appl. Phys. Lett. 2012, 101, 093702. [Google Scholar] [CrossRef] [Green Version]
  40. Xie, X.; Rieth, L.; Caldwell, R.; Diwekar, M.; Tathireddy, P.; Sharma, R.; Solzbacher, F. Long-Term Bilayer Encapsulation Performance of Atomic Layer Deposited Al2O3 and Parylene C for Biomedical Implantable Devices. IEEE Trans. Biomed. Eng. 2013, 60, 2943–2951. [Google Scholar] [PubMed]
  41. Minnikanti, S.; Diao, G.; Pancrazio, J.J.; Xie, X.; Rieth, L.; Solzbacher, F.; Peixoto, N. Lifetime assessment of atomic-layer-deposited Al2O3–Parylene C bilayer coating for neural interfaces using accelerated age testing and electrochemical characterization. Acta Biomater. 2014, 10, 960–967. [Google Scholar] [CrossRef]
  42. Xie, X.; Rieth, L.; Williams, L.; Negi, S.; Bhandari, R.; Caldwell, R.; Sharma, R.; Tathireddy, P.; Solzbacher, F. Long-term reliability of Al2O3 and Parylene C bilayer encapsulated Utah electrode array based neural interfaces for chronic implantation. J. Neural Eng. 2014, 11, 026016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Xie, X.; Rieth, L.W.; Sharma, R.; Negi, S.; Bhandari, R.; Caldwell, R.; Tathireddy, P.; Solzbacher, F. Atomic Layer Deposited Al2O3 and Parylene C Bi-layer Encapsulation for Utah Electrode Array Based Neural Interfaces. MRS Online Proc. Libr. 2014, 1621, 259–265. [Google Scholar] [CrossRef]
  44. Caldwell, R.; Rieth, L.; Xie, X.; Sharma, R.; Solzbacher, F.; Tathireddy, P. Failure mode analysis of Al2O2-parylene c bilayer encapsulation for implantable devices and application to penetrating neural arrays. In Proceedings of the 2015 Transducers—2015 18th International Conference on Solid-State Sensors, Actuators and Microsystems (TRANSDUCERS), Anchorage, AK, USA, 21–25 June 2015; pp. 1747–1750. [Google Scholar]
  45. Xie, X.; Rieth, L.; Caldwell, R.; Negi, S.; Bhandari, R.; Sharma, R.; Tathireddy, P.; Solzbacher, F. Effect of bias voltage and temperature on lifetime of wireless neural interfaces with Al2O3 and parylene bilayer encapsulation. Biomed. Microdevices 2015, 17, 1. [Google Scholar] [CrossRef]
  46. Caldwell, R.; Mandal, H.; Sharma, R.; Solzbacher, F.; Tathireddy, P.; Rieth, L. Analysis of Al2O3—parylene C bilayer coatings and impact of microelectrode topography on long term stability of implantable neural arrays. J. Neural Eng. 2017, 14, 046011. [Google Scholar] [CrossRef]
  47. Guo, H.C.; Ye, E.; Li, Z.; Han, M.-Y.; Loh, X.J. Recent progress of atomic layer deposition on polymeric materials. Mater. Sci. Eng. C 2017, 70, 1182–1191. [Google Scholar] [CrossRef]
  48. Parsons, G.N.; Atanasov, S.E.; Dandley, E.C.; Devine, C.K.; Gong, B.; Jur, J.S.; Lee, K.; Oldham, C.J.; Peng, Q.; Spagnola, J.C.; et al. Mechanisms and reactions during atomic layer deposition on polymers. Coord. Chem. Rev. 2013, 257, 3323–3331. [Google Scholar] [CrossRef]
  49. Losego, M.D.; Peng, Q. Atomic layer deposition and vapor phase infiltration. In Surface Modification of Polymers: Methods and Applications; Wiley: Hoboken, NJ, USA, 2019; pp. 135–159. [Google Scholar]
  50. Waldman, R.Z.; Mandia, D.J.; Yanguas-Gil, A.; Martinson, A.B.F.; Elam, J.W.; Darling, S.B. The chemical physics of sequential infiltration synthesis—A thermodynamic and kinetic perspective. J. Chem. Phys. 2019, 151, 190901. [Google Scholar] [CrossRef]
  51. Ashurbekova, K.; Ashurbekova, K.; Botta, G.; Yurkevich, O.; Knez, M. Vapor phase processing: A novel approach for fabricating functional hybrid materials. Nanotechnology 2020, 31, 342001. [Google Scholar] [CrossRef]
  52. Brandt, E.S.; Grace, J.M. Initiation of atomic layer deposition of metal oxides on polymer substrates by water plasma pretreatment. J. Vac. Sci. Technol. A 2011, 30, 01A137. [Google Scholar] [CrossRef]
  53. Lee, G.B.; Son, K.S.; Park, S.W.; Shim, J.H.; Choi, B.-H. Low-temperature atomic layer deposition of Al2O3 on blown polyethylene films with plasma-treated surfaces. J. Vac. Sci. Technol. A 2012, 31, 01A129. [Google Scholar]
  54. Park, S.W.; Bae, K.; Kim, J.W.; Lee, G.B.; Choi, B.-H.; Lee, M.H.; Shim, J.H. Chemical Protection of Polycarbonate Surfaces by Atomic Layer Deposition of Alumina with Oxygen Plasma Pretreatment. Adv. Mater. Interfaces 2016, 3, 1600340. [Google Scholar] [CrossRef]
  55. Song, S.H.; Lee, M.Y.; Lee, G.B.; Choi, B.-H. Characterization of Al2O3 and ZnO multilayer thin films deposited by low temperature thermal atomic layer deposition on transparent polyimide. J. Vac. Sci. Technol. A 2016, 35, 01B110. [Google Scholar] [CrossRef]
  56. Krumpolec, R.; Cameron, D.C.; Homola, T.; Černák, M. Surface chemistry and initial growth of Al2O3 on plasma modified PTFE studied by ALD. Surf. Interfaces 2017, 6, 223–228. [Google Scholar] [CrossRef]
  57. Frank, M.M.; Chabal, Y.J.; Wilk, G.D. Nucleation and interface formation mechanisms in atomic layer deposition of gate oxides. Appl. Phys. Lett. 2003, 82, 4758–4760. [Google Scholar] [CrossRef]
  58. Wilson, C.A.; Grubbs, R.K.; George, S.M. Nucleation and Growth during Al2O3 Atomic Layer Deposition on Polymers. Chem. Mater. 2005, 17, 5625–5634. [Google Scholar] [CrossRef]
  59. Kirsch, P.D.; Quevedo-Lopez, M.A.; Li, H.J.; Senzaki, Y.; Peterson, J.J.; Song, S.C.; Krishnan, S.A.; Moumen, N.; Barnett, J.; Bersuker, G.; et al. Nucleation and growth study of atomic layer deposited HfO2 gate dielectrics resulting in improved scaling and electron mobility. J. Appl. Phys. 2006, 99, 023508. [Google Scholar] [CrossRef]
  60. Elam, J.W.; Zinovev, A.V.; Pellin, M.J.; Comstock, D.J.; Hersam, M.C. Nucleation and growth of noble metals on oxide surfaces using atomic layer deposition. ECS Trans. 2007, 3, 271–278. [Google Scholar] [CrossRef]
  61. Lee, J.S.; Kaufman-Osborn, T.; Melitz, W.; Lee, S.; Delabie, A.; Sioncke, S.; Caymax, M.; Pourtois, G.; Kummel, A.C. Atomic imaging of nucleation of trimethylaluminum on clean and H2O functionalized Ge(100) surfaces. J. Chem. Phys. 2011, 135, 054705. [Google Scholar] [CrossRef]
  62. Lei, Y.; Lu, J.; Zhao, H.; Liu, B.; Low, K.-B.; Wu, T.; Libera, J.A.; Greeley, J.P.; Chupas, P.J.; Miller, J.T.; et al. Resolving Precursor Deligation, Surface Species Evolution, and Nanoparticle Nucleation during Palladium Atomic Layer Deposition. J. Phys. Chem. C 2013, 117, 11141–11148. [Google Scholar] [CrossRef]
  63. Alivio, T.E.G.; de Jesus, L.R.; Dennis, R.V.; Jia, Y.; Jaye, C.; Fischer, D.A.; Singisetti, U.; Banerjee, S. Atomic Layer Deposition of Hafnium(IV) Oxide on Graphene Oxide: Probing Interfacial Chemistry and Nucleation by using X-ray Absorption and Photoelectron Spectroscopies. ChemPhysChem 2015, 16, 2842–2848. [Google Scholar] [CrossRef]
  64. Walter, T.N.; Lee, S.; Zhang, X.; Chubarov, M.; Redwing, J.M.; Jackson, T.N.; Mohney, S.E. Atomic layer deposition of ZnO on MoS2 and WSe2. Appl. Surf. Sci. 2019, 480, 43–51. [Google Scholar] [CrossRef]
  65. Gakis, G.P.; Vahlas, C.; Vergnes, H.; Dourdain, S.; Tison, Y.; Martinez, H.; Bour, J.; Ruch, D.; Boudouvis, A.G.; Caussat, B.; et al. Investigation of the initial deposition steps and the interfacial layer of Atomic Layer Deposited (ALD) Al2O3 on Si. Appl. Surf. Sci. 2019, 492, 245–254. [Google Scholar] [CrossRef]
  66. Astaneh, S.H.; Jursich, G.; Sukotjo, C.; Takoudis, C.G. Surface and subsurface film growth of titanium dioxide on polydimethylsiloxane by atomic layer deposition. Appl. Surf. Sci. 2019, 493, 779–786. [Google Scholar] [CrossRef]
  67. Su, D.-Y.; Kuo, Y.-H.; Tseng, M.-H.; Tsai, F.-Y. Effects of surface pretreatment and deposition conditions on the gas permeation properties and flexibility of Al2O3 films on polymer substrates by atomic layer deposition. J. Coat. Technol. Res. 2019, 16, 1751–1756. [Google Scholar] [CrossRef]
  68. Reif, J.; Knaut, M.; Killge, S.; Albert, M.; Bartha, J.W. In vacuo investigations on the nucleation of TaCN by plasma enhanced atomic layer deposition. Microelectron. Eng. 2019, 211, 13–17. [Google Scholar] [CrossRef]
  69. Zhu, H.; Addou, R.; Wang, Q.; Nie, Y.; Cho, K.; Kim, M.J.; Wallace, R.M. Surface and interfacial study of atomic layer deposited Al2O3 on MoTe2 and WTe2. Nanotechnology 2020, 31, 055704. [Google Scholar] [CrossRef] [PubMed]
  70. Puurunen, R.L.; Vandervorst, W.; Besling, W.F.A.; Richard, O.; Bender, H.; Conard, T.; Zhao, C.; Delabie, A.; Caymax, M.; de Gendt, S.; et al. Island growth in the atomic layer deposition of zirconium oxide and aluminum oxide on hydrogen-terminated silicon: Growth mode modeling and transmission electron microscopy. J. Appl. Phys. 2004, 96, 4878–4889. [Google Scholar] [CrossRef]
  71. Thian, D.; Yemane, Y.T.; Xu, S.; Prinz, F.B. Methodology for Studying Surface Chemistry and Evolution during the Nucleation Phase of Atomic Layer Deposition Using Scanning Tunneling Microscopy. J. Phys. Chem. C 2017, 121, 27379–27388. [Google Scholar] [CrossRef]
  72. Kim, D.-H.; Ghaffari, R.; Lu, N.; Rogers, J.A. Flexible and Stretchable Electronics for Biointegrated Devices. Annu. Rev. Biomed. Eng. 2012, 14, 113–128. [Google Scholar] [CrossRef] [Green Version]
  73. Poppendieck, W.; Sossalla, A.; Krob, M.-O.; Welsch, C.; Nguyen, T.A.K.; Gong, W.; DiGiovanna, J.; Micera, S.; Merfeld, D.M.; Hoffmann, K.-P. Development, manufacturing and application of double-sided flexible implantable microelectrodes. Biomed. Microdevices 2014, 16, 837–850. [Google Scholar] [CrossRef]
  74. Lee, H.C.; Ejserholm, F.; Gaire, J.; Currlin, S.; Schouenborg, J.; Wallman, L.; Bengtsson, M.; Park, K.; Otto, K.J. Histological evaluation of flexible neural implants; flexibility limit for reducing the tissue response? J. Neural Eng. 2017, 14, 036026. [Google Scholar] [CrossRef] [Green Version]
  75. Kassanos, P.; Anastasova, S.; Chen, C.M.; Yang, G.-Z. Sensor Embodiment and Flexible Electronics. In Implantable Sensors and Systems: From Theory to Practice; Yang, G.-Z., Ed.; Springer International Publishing: Cham, Switzerland, 2018; pp. 197–279. [Google Scholar]
  76. Verplancke, R.; Cauwe, M.; Schaubroeck, D.; Cuypers, D.; Vandecasteele, B.; Mader, L.; Vanhaverbeke, C.; Ballini, M.; O’Callaghan, J.; Goikoetxea, E.; et al. Development of an active high-density transverse intrafascicular micro-electrode probe. J. Micromech. Microeng. 2019, 30, 015010. [Google Scholar] [CrossRef]
  77. Vandekerckhove, B.; Missinne, J.; Vonck, K.; Bauwens, P.; Verplancke, R.; Boon, P.; Raedt, R.; Vanfleteren, J. Technological Challenges in the Development of Optogenetic Closed-Loop Therapy Approaches in Epilepsy and Related Network Disorders of the Brain. Micromachines 2021, 12, 38. [Google Scholar] [CrossRef]
  78. Li, C.; Cauwe, M.; Mader, L.; Schaubroeck, D.; de Beeck, M.O. Accelerated Hermeticity Testing of Biocompatible Moisture Barriers Used for the Encapsulation of Implantable Medical Devices. Coatings 2020, 10, 19. [Google Scholar] [CrossRef] [Green Version]
  79. Li, C.; Cauwe, M.; Yang, Y.; Schaubroeck, D.; Mader, L.; de Beeck, M.O. Ultra-Long-Term Reliable Encapsulation Using an Atomic Layer Deposited HfO2/Al2O3/HfO2 Triple-Interlayer for Biomedical Implants. Coatings 2019, 9, 579. [Google Scholar] [CrossRef] [Green Version]
  80. Fahim, M.; Bijwe, J.; Nalwa, H.S. Chapter 8—Polyimides for Microelectronics and Tribology Applications. In Supramolecular Photosensitive and Electroactive Materials; Nalwa, H.S., Ed.; Academic Press: San Diego, CA, USA, 2001; pp. 643–726. [Google Scholar]
  81. Liaw, D.J.; Wang, K.L.; Huang, Y.C.; Lee, K.R.; Lai, J.Y.; Ha, C.S. Advanced polyimide materials: Syntheses, physical properties and applications. Prog. Polym. Sci. 2012, 37, 907–974. [Google Scholar] [CrossRef]
  82. Qin, Y.; Howlader, M.M.R.; Deen, M.J.; Haddara, Y.M.; Selvaganapathy, P.R. Polymer integration for packaging of implantable sensors. Sens. Actuators B Chem. 2014, 202, 758–778. [Google Scholar] [CrossRef]
  83. Ghosh, M.K.; Mittal, K.L. Polyimides: Fundamentals and Applications; CRC Press: Boca Raton, FL, USA, 2018. [Google Scholar]
  84. Bryant, R.G. Polyimides. In Ullmann’s Encyclopedia of Industrial Chemistry. 2014. Available online: https://en.wikipedia.org/wiki/Ullmann’s_Encyclopedia_of_Industrial_Chemistry (accessed on 1 October 2021).
  85. Patrick, E.; Sankar, V.; Rowe, W.; Sheng-Feng, Y.; Sanchez, J.C.; Toshikazu, N. Flexible polymer substrate and tungsten microelectrode array for an implantable neural recording system. In Proceedings of the 2008 30th Annual International Conference of the IEEE Engineering in Medicine and Biology Society, Vancouver, BC, Canada, 21–24 August 2008; pp. 3158–3161. [Google Scholar]
  86. Winkin, N.; Mokwa, W. Flexible multi-electrode array with integrated bendable CMOS-chip for implantable systems. In Proceedings of the 2012 Annual International Conference of the IEEE Engineering in Medicine and Biology Society, San Diego, CA, USA, 28 August–1 September 2012; pp. 3882–3885. [Google Scholar]
  87. Kirsten, S.; Schubert, M.; Braunschweig, M.; Woldt, G.; Voitsekhivska, T.; Wolter, K.J. Biocompatible packaging for implantable miniaturized pressure sensor device used for stent grafts: Concept and choice of materials. In Proceedings of the Proceedings of the 16th Electronics Packaging Technology Conference, EPTC 2014, Singapore, 3–5 December 2014; pp. 719–724. [Google Scholar]
  88. Schubert, M.; Kirsten, S.; Voitsekhivska, T.; Bock, K. Characterization of polymeric encapsulation for implantable microsystems applying dynamic fluidic and electrical load. In Proceedings of the Proceedings of the International Spring Seminar on Electronics Technology, Eger, Hungary, 6–10 May 2015; pp. 129–133. [Google Scholar]
  89. Tolstosheeva, E.; Biefeld, V.; Lang, W. Accelerated soak performance of BPDA-PPD polyimide for implantable MEAs. Procedia Eng. 2015, 120, 36–40. [Google Scholar] [CrossRef] [Green Version]
  90. Lee, J.H.; Kim, H.; Kim, J.H.; Lee, S.-H. Soft implantable microelectrodes for future medicine: Prosthetics, neural signal recording and neuromodulation. Lab Chip 2016, 16, 959–976. [Google Scholar] [CrossRef]
  91. Beeck, M.O.d.; Verplancke, R.; Schaubroeck, D.; Cuypers, D.; Cauwe, M.; Vandecasteele, B.; Callaghan, J.O.; Braeken, D.; Andrei, A.; Firrincieli, A.; et al. Ultra-thin biocompatible implantable chip for bidirectional communication with peripheral nerves. In Proceedings of the 2017 IEEE Biomedical Circuits and Systems Conference (BioCAS), Torino, Italy, 19–21 October 2017; pp. 1–4. [Google Scholar]
  92. Bleck, L.; Steins, H.; von Metzen, R. Interface Adhesion in Implantable Chip-in-Foil Systems. In Proceedings of the Proceedings of the Annual International Conference of the IEEE Engineering in Medicine and Biology Society, EMBS, Dubai, United Arab Emirates, 12–15 November 2018; pp. 2981–2984. [Google Scholar]
  93. Li, C. Long-Term Miniaturized Flexible Hermetic Encapsulation for Implantable Medical Devices Using Polymers and Atomic Layer Deposited Ceramics; Department of Electronics and Information Systems, Ghent University: Ghent, Belgium, 2020. [Google Scholar]
  94. Steckl, A.J.; Balakrishnan, S.; Jin, H.S.; Corelli, J.C. Micromachining of polyimide films with focused ion beams. Microelectron. Eng. 1986, 5, 461–462. [Google Scholar] [CrossRef]
  95. al Hashimi, H.; Chaalal, O. Flexible temperature sensor fabrication using photolithography technique. Therm. Sci. Eng. Prog. 2021, 22, 100857. [Google Scholar] [CrossRef]
  96. Rajawat, S.; Hübner, M.; Kempen, L.; Lang, W. Flexible passive LC resonator for wireless measurement during curing of thermosets. J. Phys. Conf. Ser. 2021, 1837, 012001. [Google Scholar]
  97. Kuliasha, C.A.; Judy, J.W. The Materials Science Foundation Supporting the Microfabrication of Reliable Polyimide–Metal Neuroelectronic Interfaces. Adv. Mater. Technol. 2021, 6, 2100149. [Google Scholar] [CrossRef]
  98. Ahn, S.-H.; Jeong, J.; Kim, S.J. Emerging Encapsulation Technologies for Long-Term Reliability of Microfabricated Implantable Devices. Micromachines 2019, 10, 508. [Google Scholar] [CrossRef] [Green Version]
  99. Forssell, M.; Ong, X.C.; Fedder, G.K. Multilayer ALD ceramic films for enhancement of parylene barrier properties in compliant neural probes with bonded chips. In Proceedings of the 2018 Solid-State Sensors, Actuators and Microsystems Workshop, Hilton Head Island, SC, USA, 3–7 June 2018; pp. 272–275. [Google Scholar]
  100. Astoreca, L.; Cools, P.; Schaubroeck, D.; Asadian, M.; Aliakbarshirazi, S.; Declercq, H.; de Beeck, M.O.; Morent, R.; de Smet, H.; de Geyter, N. Non-thermal plasma activation of BPDA-PPD polyimide for improved cell-material interaction. Polymer 2020, 205, 122831. [Google Scholar] [CrossRef]
  101. Buder, U.; von Klitzing, J.P.; Obermeier, E. Reactive ion etching for bulk structuring of polyimide. Sens. Actuators A Phys. 2006, 132, 393–399. [Google Scholar] [CrossRef]
  102. Zhang, Y.; Guerra-Nuñez, C.; Utke, I.; Michler, J.; Rossell, M.D.; Erni, R. Understanding and Controlling Nucleation and Growth of TiO2 Deposited on Multiwalled Carbon Nanotubes by Atomic Layer Deposition. J. Phys. Chem. C 2015, 119, 3379–3387. [Google Scholar] [CrossRef]
  103. Schilirò, E.; Nigro, R.L.; Panasci, S.E.; Gelardi, F.M.; Agnello, S.; Yakimova, R.; Roccaforte, F.; Giannazzo, F. Aluminum oxide nucleation in the early stages of atomic layer deposition on epitaxial graphene. Carbon 2020, 169, 172–181. [Google Scholar] [CrossRef]
  104. Grigoras, K.; Sainiemi, L.; Tiilikainen, J.; Säynätjoki, A.; Airaksinen, V.M.; Franssila, S. Application of ultra-thin aluminum oxide etch mask made by atomic layer deposition technique. J. Phys. Conf. Ser. 2007, 61, 369–373. [Google Scholar] [CrossRef] [Green Version]
  105. Kim, H.G.; Lee, J.G.; Kim, S.S. Surface modification of polymeric substrates to enhance the barrier properties of an Al2O3 layer formed by PEALD process. Org. Electron. 2017, 50, 239–246. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the approach to visualize the ALD nucleation on polymers.
Figure 1. Schematic representation of the approach to visualize the ALD nucleation on polymers.
Coatings 11 01352 g001
Figure 2. FEG–SEM images (magnification ×20,000 (top) and ×50,000 (bottom)) of pristine spin coated polyimide (PI), polyimide after the RIE process and polyimide with a closed ALD layer on top after the RIE process.
Figure 2. FEG–SEM images (magnification ×20,000 (top) and ×50,000 (bottom)) of pristine spin coated polyimide (PI), polyimide after the RIE process and polyimide with a closed ALD layer on top after the RIE process.
Coatings 11 01352 g002
Figure 3. FEG–SEM images showing the evolution of Al2O3 and HfO2 ALD nucleation on polyimide after 5 min of RIE. The ALD deposition processes had a GPC of 1.20 ± 0.02 Å/cycle for Al2O3 and 1.16 ± 0.01 Å/cycle for HfO2 on Si.
Figure 3. FEG–SEM images showing the evolution of Al2O3 and HfO2 ALD nucleation on polyimide after 5 min of RIE. The ALD deposition processes had a GPC of 1.20 ± 0.02 Å/cycle for Al2O3 and 1.16 ± 0.01 Å/cycle for HfO2 on Si.
Coatings 11 01352 g003
Figure 4. Estimated coverage of polyimide as the nucleation of (a) Al2O3 ALD and (b) HfO2 ALD evolved. The coverage was calculated on both pristine and plasma-activated polyimide. * Denotes a statistically significant difference between the coverage values of identical number of ALD cycles on pristine and plasma-activated polyimide at a significance level of 0.05.
Figure 4. Estimated coverage of polyimide as the nucleation of (a) Al2O3 ALD and (b) HfO2 ALD evolved. The coverage was calculated on both pristine and plasma-activated polyimide. * Denotes a statistically significant difference between the coverage values of identical number of ALD cycles on pristine and plasma-activated polyimide at a significance level of 0.05.
Coatings 11 01352 g004
Figure 5. (a) Surface concentration of Al and Hf detected by XPS after the deposition of increasing Al2O3 and HfO2 ALD cycles, respectively, on polyimide. Evolution of (b) O1s and (c) Al2p high resolution peaks as successive Al2O3 cycles were deposited on polyimide, from 5 to 40 cycles. Evolution of (d) O1s and (e) Hf4f high resolution peaks as successive HfO2 cycles were deposited on polyimide, from 5 to 40 cycles. The arrows in (be) show the direction of increment of the deposited ALD cycles.
Figure 5. (a) Surface concentration of Al and Hf detected by XPS after the deposition of increasing Al2O3 and HfO2 ALD cycles, respectively, on polyimide. Evolution of (b) O1s and (c) Al2p high resolution peaks as successive Al2O3 cycles were deposited on polyimide, from 5 to 40 cycles. Evolution of (d) O1s and (e) Hf4f high resolution peaks as successive HfO2 cycles were deposited on polyimide, from 5 to 40 cycles. The arrows in (be) show the direction of increment of the deposited ALD cycles.
Coatings 11 01352 g005
Figure 6. HAADF-STEM and EDS maps of PI/Al2O3/PI and PI/HfO2/PI cross-sections. The scale bar on the Al2O3 images corresponds to 50 nm and the scale bar on the HfO2 images to 30 nm. Blue represents oxygen. Red represents Al in the Al2O3 EDS maps and Hf in the HfO2 EDS maps.
Figure 6. HAADF-STEM and EDS maps of PI/Al2O3/PI and PI/HfO2/PI cross-sections. The scale bar on the Al2O3 images corresponds to 50 nm and the scale bar on the HfO2 images to 30 nm. Blue represents oxygen. Red represents Al in the Al2O3 EDS maps and Hf in the HfO2 EDS maps.
Coatings 11 01352 g006
Figure 7. FEG–SEM images showing the evolution of Al2O3 ALD nucleation on plasma-activated polyimide and pristine polyimide.
Figure 7. FEG–SEM images showing the evolution of Al2O3 ALD nucleation on plasma-activated polyimide and pristine polyimide.
Coatings 11 01352 g007
Figure 8. FEG–SEM images showing the evolution of HfO2 ALD nucleation on plasma-activated polyimide and pristine polyimide.
Figure 8. FEG–SEM images showing the evolution of HfO2 ALD nucleation on plasma-activated polyimide and pristine polyimide.
Coatings 11 01352 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Astoreca, L.; Schaubroeck, D.; Esbah Tabaei, P.S.; Ghobeira, R.; Op de Beeck, M.; Morent, R.; De Smet, H.; De Geyter, N. Investigating the Nucleation of AlOx and HfOx ALD on Polyimide: Influence of Plasma Activation. Coatings 2021, 11, 1352. https://doi.org/10.3390/coatings11111352

AMA Style

Astoreca L, Schaubroeck D, Esbah Tabaei PS, Ghobeira R, Op de Beeck M, Morent R, De Smet H, De Geyter N. Investigating the Nucleation of AlOx and HfOx ALD on Polyimide: Influence of Plasma Activation. Coatings. 2021; 11(11):1352. https://doi.org/10.3390/coatings11111352

Chicago/Turabian Style

Astoreca, Laura, David Schaubroeck, Parinaz Saadat Esbah Tabaei, Rouba Ghobeira, Maaike Op de Beeck, Rino Morent, Herbert De Smet, and Nathalie De Geyter. 2021. "Investigating the Nucleation of AlOx and HfOx ALD on Polyimide: Influence of Plasma Activation" Coatings 11, no. 11: 1352. https://doi.org/10.3390/coatings11111352

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop