Next Article in Journal
Pellet Production from Pruning and Alternative Forest Biomass: A Review of the Most Recent Research Findings
Next Article in Special Issue
Investigation of the Performance of Perovskite Solar Cells with ZnO-Covered PC61BM Electron Transport Layer
Previous Article in Journal
Study on Electrochemical Performance of MnO@rGO/Carbon Fabric-Based Wearable Supercapacitors
Previous Article in Special Issue
Analyzing the Bolometric Performance of Vanadium Oxide Thin Films Modified by Carbon Nanotube Dispersions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Area-Selective Atomic Layer Deposition of ZnO on Si\SiO2 Modified with Tris(dimethylamino)methylsilane

1
Department of Chemistry and Biochemistry, Brigham Young University, Provo, UT 84602, USA
2
Department of Applied Physics, Eindhoven University of Technology, 5600 MB Eindhoven, The Netherlands
3
Institute of Physical Engineering, Brno University of Technology, Technická 2, 616 69 Brno, Czech Republic
4
CEITEC BUT, Brno University of Technology, Purkyňova 123, 612 00 Brno, Czech Republic
5
Materials Group-NA, Avery Dennison Corporation, 8080 Norton Parkway, Mentor, OH 44060, USA
6
Restek Corporation, 110 Benner Circle, Bellefonte, PA 16823, USA
7
Department of Statistics, Brigham Young University, Provo, UT 84602, USA
8
Department of Chemistry, University of Colorado, 215 UCB, Boulder, CO 80309, USA
*
Author to whom correspondence should be addressed.
Materials 2023, 16(13), 4688; https://doi.org/10.3390/ma16134688
Submission received: 25 May 2023 / Revised: 19 June 2023 / Accepted: 25 June 2023 / Published: 29 June 2023
(This article belongs to the Special Issue Feature Papers in Thin Films and Interfaces)

Abstract

:
Delayed atomic layer deposition (ALD) of ZnO, i.e., area selective (AS)-ALD, was successfully achieved on silicon wafers (Si\SiO2) terminated with tris(dimethylamino)methylsilane (TDMAMS). This resist molecule was deposited in a home-built, near-atmospheric pressure, flow-through, gas-phase reactor. TDMAMS had previously been shown to react with Si\SiO2 in a single cycle/reaction and to drastically reduce the number of silanols that remain at the surface. ZnO was deposited in a commercial ALD system using dimethylzinc (DMZ) as the zinc precursor and H2O as the coreactant. Deposition of TDMAMS was confirmed by spectroscopic ellipsometry (SE), X-ray photoelectron spectroscopy (XPS), and wetting. ALD of ZnO, including its selectivity on TDMAMS-terminated Si\SiO2 (Si\SiO2\TDMAMS), was confirmed by in situ multi-wavelength ellipsometry, ex situ SE, XPS, and/or high-sensitivity/low-energy ion scattering (HS-LEIS). The thermal stability of the TDMAMS resist layer, which is an important parameter for AS-ALD, was investigated by heating Si\SiO2\TDMAMS in air and nitrogen at 330 °C. ALD of ZnO takes place more readily on Si\SiO2\TDMAMS heated in the air than in N2, suggesting greater damage to the surface heated in the air. To better understand the in situ ALD of ZnO on Si\SiO2\TDMAMS and modified (thermally stressed) forms of it, the ellipsometry results were plotted as the normalized growth per cycle. Even one short pulse of TDMAMS effectively passivates Si\SiO2. TDMAMS can be a useful, small-molecule inhibitor of ALD of ZnO on Si\SiO2 surfaces.

1. Introduction

Atomic layer deposition (ALD) is a powerful, advanced, thin-film deposition technique that is based on self-terminating surface reactions [1,2]. It meets the need for atomic layer control/deposition of materials in many scientific fields, facilitating advances in fabrication processes and methods. In its most common embodiment, ALD involves two self-terminating reactions that are used in an alternating fashion to (ideally) deposit up to a monolayer of a material on a surface with each ALD cycle. Between the two ALD half reactions, unreacted reagent is pumped/purged away. When the half reactions are well chosen and properly implemented, ALD allows precise control over film thickness and uniformity. ALD is an example of bottom-up thin-film fabrication because it is literally based on the atom-by-atom growth of thin films. ALD can uniformly coat patterned/irregular surfaces that are difficult to coat by other techniques; an important advantage of ALD is the conformality of the films that are often created.
Area-selective ALD (AS-ALD) is an important focus of current research in ALD [3,4]. AS-ALD has also been referred to as substrate-dependent deposition [3]. In AS-ALD, surfaces are modified to inhibit or direct subsequent ALD growth [3,4]. Thus, surface chemistry plays a central role in AS-ALD [5,6], and differences in surface chemistry can control feature sizes. There are two main approaches to AS-ALD: activation or passivation [7]. However, while surface activation is a viable approach to AS-ALD, surface passivation is its core focus because it is cheaper and more scalable [8]. There are also other AS-ALD methods such as selective precursor and co-reactant adsorption methods [9,10]. Two general approaches for inhibiting a reaction at a surface are (i) to prevent a reactant’s diffusion to the active functional groups on a surface via self-assembled monolayers, [11] and (ii) to passivate/end-cap these groups with small inhibitor molecules [7,12]. The first approach is based on sterics and the second on chemical derivatization. Although both approaches are appropriate for AS-ALD, small inhibitor/end-capping molecules often enjoy greater success because they are often both volatile and their deposition compatible with current industrial tools and processes [8].
Recently, we demonstrated the surface passivation of the inner surfaces of silica capillary columns for gas chromatography using a small silane (tris(dimethylamino)methylsilane, TDMAMS), (see Figure 1), delivered by a near atmospheric pressure gas-phase reactor [13]. This deposition was also followed on witness silicon shards. The reaction of surface silanols with TDMAMS was confirmed by X-ray photoelectron spectroscopy (XPS), spectroscopic ellipsometry (SE), contact angle goniometry (wetting), and high-sensitivity/low-energy ion scattering (HS-LEIS). HS-LEIS showed that the reaction of Si\SiO2 with TDMAMS results in a large drop in the number of free surface silanols (from 3.66 to 0.31 silanols/nm2 on native-oxide-terminated silicon). This very large decrease in the surface silanol concentration caused us to ask whether TDMAMS-terminated silicon oxide (Si\SiO2\TDMAMS) might inhibit ALD. Accordingly, in this work, we investigate the ALD of ZnO on Si\SiO2 and Si\SiO2\TDMAMS, where TDMAMS was again deposited in the home-built, flow-through, atmospheric pressure, gas-phase reactor [14]. ZnO is an important metal oxide with applications in optoelectronics, sensing, memory devices, catalysis, transparent conducting oxides, and thin-film transistors [15,16,17,18,19]. In addition, its properties can be controlled by doping with elements such as indium (In) and gallium (Ga) [20,21]. In our work, ZnO was deposited by thermal ALD of dimethylzinc (DMZ) and H2O. For our surfaces, selectivity in AS-ALD can be attributed to blocking of active sites (silanols) and/or changing the reaction adsorption/desorption pathway. The effect of TDMAMS on the concentration of free surface silanols was further investigated using HS-LEIS. Finally, for our system, selectivity relies on the stability of adsorbed TDMAMS. To better understand this stability, we investigated the ALD of ZnO onto Si\SiO2\TDMAMS that had been deliberately damaged by heating in the air or under N2 at 330 °C.
Selectivity in AS-ALD can be represented and displayed in different ways. Selectivity is commonly represented as a selectivity window, which is the number of delayed cycles compared to growth on an uninhibited/unmodified substrate and/or a selectivity value between 0 and 1, which correspond to no selectivity and perfect selectivity, respectively [4,22]. The selectivity window can usually be determined from a plot of thicknesses versus the ALD cycle number. The selectivity value is based on a definition of chemical selectivity, which is derived from the underlying thermodynamics and first order kinetics of the desired and undesired surface reactions. The most common definition the selectivity, S , is the difference between the thicknesses of the films on the unmodified/unpassivated, θ 1 , and modified, θ 2 , substrates divided by the sum of those thicknesses (see Equation (1)) [4,22,23].
S = θ 1 θ 2 θ 1 + θ 2
where θ 1 and θ 2 are the amounts of deposited material on the growth and non-growth surfaces, respectively. Depending on the deposition method, θ can be calculated by different methods including scanning electron microscopy (SEM), transmission electron microscopy (TEM), atomic force microscopy (AFM), SE, or barriers/energy differences from computational studies [24]. Higher selectivity is always desirable. Finally, in addition to showing the selectivity as a function of the number of ALD cycles, the normalized growth per cycle was plotted. This representation of the data appears to effectively identify some of the important regimes in film growth and inhibition in AS-ALD.

2. Experimental Section

2.1. Materials

Silicon wafers, Si(100), were obtained from University Wafer Inc. (South Boston, MA, USA). Prior to functionalization, they were cleaned using a Basic Plasma Cleaner (115 V) from Harrick Plasma (Ithaca, NY, USA). Tris(dimethylamino)methylsilane, TDMAMS, 97% purity, was obtained from Gelest (Morrisville, PA, USA) and used as received. Dimethylzinc (DMZ) was obtained from Strem Chemicals (Newburyport, MA, USA). HPLC grade water was used for ALD.

2.2. TDMAMS Deposition on Silicon Wafers

TDMAMS was deposited on silicon wafers at 200 °C in a previously reported, home-built, flow-through, ambient-pressure reactor [14]. Short, 20 and 30 ms, pulses of TDMAMS and water were carried by 60 and 90 s pulses of nitrogen gas, respectively.

2.3. Atomic Layer Deposition (ALD) of ZnO on Passivated Silicon (Si\SiO2)

ALD of ZnO was performed with an ALD-150LX™ system from the Kurt J. Lesker Co. (Jefferson Hills, PA, USA) by applying 21 and 15.5 ms pulses of dimethylzinc (DMZ) and water, respectively. The purging time after each pulse was 10 s. The DMZ and water sources were held at 110 °C and 24 °C, respectively, and the substrate was at 200 °C.

2.4. Ex Situ Spectroscopic Ellipsometry (SE)

Ex situ SE was performed with an RC2 instrument from the J.A. Woollam Co. (Lincoln, NE, USA) from 200 to 1000 nm at three angles of incidence (65°, 70°, and 75°). The data from these measurements were modeled over the full wavelength range of the instrument to determine the thickness of the native oxide layer on silicon. Thick (7–8 nm) ZnO films were modeled over their (essentially) transparent region from 420–1000 nm using a Sellmeier model. All data modeling was in Woollam’s CompleteEASE software package (ver. 6.61 a). The ex situ SE modeling of thick ZnO reference sample was performed using a three-layer model that accounted for the silicon substrate; the silicon native oxide layer, which, again, was measured prior to deposition and fixed to the value obtained in this measurement; and a (Sellmeier) layer that accounted for the ZnO film.

2.5. In Situ Multi-Wavelength Ellipsometry

In situ SE measurements during ALD of ZnO were performed with a four-wavelength FS-1 ellipsometer from Film Sense (Lincoln, NE, USA), and the data were modeled using Film Sense’s software (Ver 2.66W) [8,25,26]. Two approaches were used to model the data. In the first, the optical constants for ZnO were obtained from an ex situ SE analysis of the ZnO film, which yielded values for n at 464, 523, 597, and 638 nm (the wavelengths of the Film Sense ellipsometer) of 2.01, 2.01, 1.99, and 1.99, respectively [27], and k was taken to be zero at these wavelengths. In the second, the ZnO film was modeled with a Sellmeier function. Both approaches produced similar thicknesses. Only the results from the first approach are shown in this work. In the modeling of the in situ data, the optical constants for heated silicon (at 200 °C) were used.

2.6. X-ray Photoelectron Spectroscopy (XPS)

XPS spectra were acquired with a Thermo Scientific Model ESCALAB 250Xi XPS instrument equipped with a monochromatic Al K-alpha X-ray source. All XPS spectra were acquired using a focused X-ray spot size of 650 μ m at the sample. XPS survey spectra (0–1350 eV) were acquired using an analyzer pass energy of 150 eV with a 1.0 eV step size and a 10 ms dwell time. High-resolution XPS spectra were acquired with an analyzer pass energy of 20 eV, a 0.1 eV step size, and a 50 ms dwell time. Charge neutralization of the samples during analysis was accomplished using a low-energy electron source that is in-lens in the analyzer column plus an auxiliary combination low-energy electron/low-energy argon ion source in the analysis chamber. The base pressure in the analysis chamber was typically 2.0 × 10−8 mbar or better prior to XPS data acquisition. During the XPS data acquisition, the pressure in the analysis chamber was typically 2.0 × 10−7 mbar because of the low-energy argon ions leaked into the auxiliary change neutralization source.

2.7. High-Sensitivity/Low-Energy Ion Scattering (HS-LEIS)

Ion scattering experiments were performed with a Qtac100 (IonTof GmbH, Münster, Germany) instrument. Small (ca. 100 μ m) diameter beams of He+ (3.0 keV) were scanned over selected 2 × 2 mm2 areas. Surface charging of the samples was compensated with an electron flood gun. During the transport of the samples to the LEIS, they will pick up some (organic) contamination. Therefore, before the LEIS analyses, they were cleaned with atomic oxygen (50 min). Since this may form some ZnO2, Zn(IV), this peroxide is converted to the regular oxide (ZnO, Zn(II)) by annealing at 270 °C. All results reported herein are based on three measurements at two different spots.

3. Results and Discussion

Passivation of native-oxide-terminated silicon (Si\SiO2) with TDMAMS (Si\SiO2\TDMAMS) was confirmed by spectroscopic ellipsometry (SE), X-ray photoelectron spectroscopy (XPS), and contact angle goniometry. For the SE, a simple, two-layer model accounted for the silicon substrate and the native oxide layer and was used to obtain the thickness of the native oxide layer on a plasma-cleaned silicon wafer. This same model was then used to find the combined thickness of the native oxide and TDMAMS layers. This approximation should be very good because these layers are very thin and chemically similar [28]. The thickness of the SiO2 layer from the unpassivated surface was subtracted from the combined thickness of the SiO2 and TDMAMS layers on the passivated surface to yield the thickness of the TDMAMS film, which was 0.54 ± 0.01 nm. The XPS Si 2p narrow scans of clean, bare Si\SiO2 before and after TDMAMS passivation are shown in Figure 2a. These narrow scans show the expected bulk Si and oxide Si 2p peaks, where evidence of the underlying spin-orbit splitting is present in the bulk Si 2p peaks [29]. Consistent with the deposition of a thin film of TDMAMS on the native oxide, the ratio of the oxide Si 2p peak to the bulk Si 2p peak increases after TDMAMS deposition. Contact angle goniometry of bare Si\SiO2 and Si\SiO2\TDMAMS showed that, as expected, the clean Si\SiO2 surface was wet with water, e.g., a typical water contact angle was 8°, while, after reaction of Si\SiO2 with TDMAMS, the advanced water contact angle (average of three different surfaces) increased dramatically to 81.3 ± 1.5°. It is remarkable that such a thin film/layer of hydrocarbon can have such a profound influence on the wetting of a surface.
To test their ability to resist ALD, clean, bare Si\SiO2 and Si\SiO2\TDMAMS underwent 50 ALD cycles of dimethylzinc (DMZ) and water under conditions that are expected to grow ZnO. XPS shows good evidence for ZnO growth on the unpassivated surface and less growth on the passivated one. For example, after 50 ALD cycles to deposit ZnO, the Zn 2p signals from bare Si\SiO2 are noticeably higher than those from Si\SiO2\TDMAMS (see Figure 2b). In addition, the Si 2p peaks from the Si\SiO2 surface are essentially gone, which is consistent with it being covered with a moderately thick film of ZnO, while the Si 2p peaks from Si\SiO2\TDMAMS are still pronounced, suggesting inhibition of ZnO growth (see Figure 2c). The thickness of the ZnO film on uninhibited Si\SiO2 was measured by SEM (see Figure 3) and ex situ SE (see Experimental Section) to be 8.0 nm (average of two measurements) and 8.39 nm, respectively. These values are in reasonable agreement.
While it is often possible to determine the chemical states of elements by XPS, the Zn(0) and Zn(II) 2p3/2 peaks at 1021.4 and 1021.7 eV, respectively, are not widely spaced in energy, which makes chemical state speciation of Zn(0) or Zn(II) from this peak challenging [30]. Accordingly, to confirm the chemical state of zinc in this film, we calculated its modified Auger parameter. The Auger parameter was first introduced by Wagner as the kinetic energy (KE) of a specified photoemission peak subtracted from the KE of a specified Auger peak [31,32]. In many cases, it is both more sensitive to the oxidation states of elements and occurs over a wider range of energies than the positions and differences of corresponding photoemission signals. Auger peaks are often more sensitive to sample chemistry than photoemission peaks because they involve higher energy electrons, i.e., electrons closer in energy to those used for chemical bonding. Gaarenstroom and Winogard added the photon energy to the Auger parameter, which made it independent of the X-ray energy and removed the possibility of it returning negative values [33]. Their form of the Auger parameter is referred to as the modified Auger parameter. Because the modified Auger parameter is the difference between two signals, it is not generally affected by sample charging. To calculate the modified Auger parameter for Zn in our film, we found the positions of the Zn 2p3/2 and Zn LMM signals by fitting them in CasaXPS [34,35] with asymmetric Lorentzian (LA) functions (see Figure 4) [35,36]. The modified Auger parameter was then obtained from Equation (2):
Modified Auger parameter (eV) = BE of Zn 2p3/2 + KE of Zn LMM
where ‘BE’ stands for ‘binding energy’. The modified Auger parameter we obtained for the zinc in our film is 2010.2 eV, which is the same as that reported elsewhere for ZnO, and very different from that of Zn metal/Zn(0) (2013.8 eV) [37].
Figure 3. SEM image of a layer of ZnO deposited on a bare Si\SiO2 substrate via 50 ALD cycles of DMZ and water. The bright region on top of the image is a protective, conductive Au-Pd\Pt coating. The contrast of this image was enhanced using the ImageJ software package (Java 1.8.0_172) [38].
Figure 3. SEM image of a layer of ZnO deposited on a bare Si\SiO2 substrate via 50 ALD cycles of DMZ and water. The bright region on top of the image is a protective, conductive Au-Pd\Pt coating. The contrast of this image was enhanced using the ImageJ software package (Java 1.8.0_172) [38].
Materials 16 04688 g003
To assess the inhibitory effects of TDMAMS in ALD, ALD of ZnO was followed by in situ ellipsometry on Si\SiO2\TDMAMS and bare Si\SiO2. On the unmodified surface (see Figure 5a), there is an induction period of slower growth during the first 5–10 cycles followed by steady, consistent thin-film growth. In contrast, on Si\SiO2\TDMAMS, only minimal growth of ZnO takes place during the first 30–35 ALD cycles. Not even 0.1 nm of ZnO has been deposited on Si\SiO2\TDMAMS after 30 DMZ-water cycles. These results suggest that TDMAMS interferes with and inhibits ALD of ZnO from DMZ and water. This inhibition is probably a result of the low silanol concentration on the Si\SiO2\TDMAMS surface [13], i.e., there are few reactive groups/sites where reaction with DMZ can occur. Furthermore, the relatively few silanols on the Si\SiO2\TDMAMS surface may be sterically hindered/inaccessible. Nevertheless, an enlarged view of the in situ ellipsometry of the ALD of ZnO on Si\SiO2\TDMAMS suggests very slow, linear growth of ZnO for the first ca. 25 ALD cycles (Figure 5b). In other words, ZnO is being slowly deposited on Si\SiO2\TDMAMS. After ca. 30 ALD cycles, ZnO deposition takes place in a more noticeable way. The results in Figure 5 showcase the value of in situ ellipsometry in ALD.
Because many ALD processes take place at higher temperatures (100–400 °C), it is important that inhibited surfaces show good thermal stability. Accordingly, Si\SiO2\TDMAMS was deliberately stressed/damaged by heating to 330 °C for 20 min in an inert atmosphere (N2) or in the air prior to ALD of ZnO. Of course, exposure to the air under these conditions is expected to be more damaging. The results of ZnO growth on these surfaces are also shown in Figure 5a so that they can be directly compared to those from bare Si\SiO2 and undamaged Si\SiO2\TDMAMS. Figure 5a shows that the TDMAMS-terminated surface that was heated in nitrogen still has significant inhibitory power, although there is some suggestion that the surface has been damaged. ALD occurs somewhat earlier on this surface than on undamaged Si\SiO2\TDMAMS. In contrast, the Si\SiO2\TDMAMS surface heated in the air appears to have been substantially damaged. While ALD on this surface is initially slower than on bare Si\SiO2, it then takes place quite readily. That is, this ALD shows (i) an induction period with a low but increasing growth rate for ZnO; (ii) a region of rapid/accelerated film growth, which suggests the presence of islands on this surface; and finally, (iii) the same steady-state growth that was observed on the unsilanized surface [22].
The growth per cycle (GPC) is an important figure of merit for ALD. Figure 6a shows a representative example of how we calculated the GPC from our in situ ellipsometry results, and Figure 6b shows the GPCs for all the cycles in the deposition of DMZ and H2O on bare Si\SiO2 in Figure 5a. As shown in Figure 6b, the GPC for the ALD of ZnO on bare Si\SiO2 plateaus after 20 cycles (and maybe after 15). We define the parameter ' G P C ' for our deposition as the average GPC at and beyond 20 cycles in Figure 6b. All the results in Figure 5a were then normalized to G P C to yield the normalized GPCs shown in Figure 7a. This representation of the information appears to facilitate its interpretation. First, since the normalized GPCs for bare Si\SiO2 in Figure 7a were obtained from a parameter derived from the same set of data, they do not fundamentally change, except in scale. The normalized GPCs for Si\SiO2\TDMAMS that was stressed/damaged by heating to 330 °C for 20 min in the air are more interesting. They show (i) an induction period for the first ca. 18 cycles, (ii) a region of accelerated growth up to ca. cycle 32 that is indicative of island growth where the substrate surface area is higher than after island coalescence (this type of growth is commonly observed in substrate-inhibited ALD [22]), and (iii) a normalized GPC value for the deposition approximately equal to G P C thereafter, which suggests that island coalescence has taken place at this point. Finally, the GPC of the Si\SiO2\TDMAMS surface that was stressed/damaged by heating to 330 °C in N2 reaches G P C after ca. 47 cycles, while the unstressed Si\SiO2\TDMAMS surface has still not reached G P C at 50 ALD cycles.
The traditional selectivities [3] for the depositions shown in Figure 5a are presented in Figure 7b, where the threshold for selectivity (0.9) that is commonly used for comparison purposes is indicated [22]. As expected, the selectivity is highest (32 cycles) for the intact TDMAMS surface. It is moderately diminished for the surface heated to 330 °C in N2 (25 cycles), and strongly diminished for the surface heated to 330 °C in the air (2 cycles). Both intact Si\SiO2\TDMAMS and Si\SiO2\TDMAMS stressed under N2 show good selectivity.
Our previous work had shown that ALD between TDMAMS and water is not possible [13]. That is, previously, multiple cycles of TDMAMS and H2O did not increase the film thickness beyond what was obtained in the first deposition of TDMAMS. We follow up on these results by testing whether there is a difference in the number of silanols remaining on Si\SiO2 after exposing it to one or ten pulses of TDMAMS, where no water was applied until after the end of these pulses. After tagging with Zn, HS-LEIS shows little difference between the number of silanols, i.e., Zn atoms (see Figure 8), on these surfaces. These results (ca. 0.20 silanols/nm2) are in reasonable agreement with our previous LEIS results (0.31 silanols/nm2). Thus, one short (30 ms) pulse of TDMAMS, or even multiple pulses of this reagent, passivate Si\SiO2 to a significant extent.

4. Conclusions

TDMAMS deposition on Si\SiO2 and AS-ALD of ZnO on this surface were confirmed by various surface analytical techniques, including XPS, in situ and ex situ ellipsometry, wetting, and HS-LEIS. This characterization showed that TDMAMS deposits onto silicon surfaces in a self-limiting way and that it dramatically reduces the number of free silanols on the surface. TDMAMS-passivated Si\SiO2 delayed ALD of ZnO for 32 cycles at a selectivity value of 0.9. To test its stability, Si\SiO2\TDMAMS was stressed/heated to 330 °C in the air and in nitrogen. The surface was not considerably damaged by heating in N2 but was seriously damaged in the air. The surface damaged in the air showed island growth of ZnO at early times. Plots of the normalized GPC facilitated the interpretation of the ALD of ZnO on TDMAMS-terminated surfaces. Overall, TDMAMS is reasonably effective at preventing ALD of ZnO. It is, potentially, an industrial-friendly, small, surface-reactive molecule that may be useful for AS-ALD.

Author Contributions

Conceptualization, B.M. and M.R.L.; Methodology, B.M. and M.R.L.; Formal analysis, B.M.; Investigation, B.M., T.G.A., H.H.B., S.P., P.B., E.V., B.R.S., D.S.B., D.E. and S.M.G.; Writing—original draft, B.M. and M.R.L.; Writing—review & editing, B.M. and M.R.L.; Project administration, M.R.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Restek Corporation, USA. CzechNanoLab project LM2023051 funded by MEYS CR is gratefully acknowledged for the financial support of the measurements/sample fabrication at CEITEC Nano Research Infrastructure.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data used in this study will be made available upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Puurunen, R.L. A Short History of Atomic Layer Deposition: Tuomo Suntola’s Atomic Layer Epitaxy. Chem. Vap. Depos. 2014, 20, 332–344. [Google Scholar] [CrossRef] [Green Version]
  2. Kääriäinen, T.; Cameron, D.; Kääriäinen, M.-L.; Sherman, A. Atomic Layer Deposition: Principles, Characteristics, and Nanotechnology Applications; John Wiley & Sons: Hoboken, NJ, USA, 2013. [Google Scholar]
  3. Mackus, A.J.M.; Merkx, M.J.M.; Kessels, W.M.M. From the Bottom-Up: Toward Area-Selective Atomic Layer Deposition with High Selectivity. Chem. Mater. 2019, 31, 2–12. [Google Scholar] [CrossRef] [Green Version]
  4. Parsons, G.N.; Clark, R.D. Area-Selective Deposition: Fundamentals, Applications, and Future Outlook. Chem. Mater. 2020, 32, 4920–4953. [Google Scholar] [CrossRef]
  5. Mameli, A.; Karasulu, B.; Verheijen, M.A.; Mackus, A.J.M.; Kessels, W.M.M.; Roozeboom, F. (Invited) Area-Selective Atomic Layer Deposition: Role of Surface Chemistry. ECS Trans. 2017, 80, 39. [Google Scholar] [CrossRef]
  6. Chen, R.; Bent, S.F. Chemistry for Positive Pattern Transfer Using Area-Selective Atomic Layer Deposition. Adv. Mater. 2006, 18, 1086–1090. [Google Scholar] [CrossRef]
  7. Mackus, A.J.M. Approaches and Opportunities for Area-Selective Atomic Layer Deposition. In 2018 International Symposium on VLSI Technology, Systems and Application (VLSI-TSA); IEEE: New York, NY, USA, 2018; pp. 1–2. [Google Scholar] [CrossRef]
  8. Mameli, A.; Karasulu, B.; Verheijen, M.A.; Barcones, B.; Macco, B.; Mackus, A.J.M.; Kessels, W.M.M.E.; Roozeboom, F. Area-Selective Atomic Layer Deposition of ZnO by Area Activation Using Electron Beam-Induced Deposition. Chem. Mater. 2019, 31, 1250–1257. [Google Scholar] [CrossRef] [Green Version]
  9. Weber, M.J.; Mackus, A.J.M.; Verheijen, M.A.; van der Marel, C.; Kessels, W.M.M. Supported Core/Shell Bimetallic Nanoparticles Synthesis by Atomic Layer Deposition. Chem. Mater. 2012, 24, 2973–2977. [Google Scholar] [CrossRef]
  10. Mackus, A.J.M.; Mulders, J.J.L.; Van De Sanden, M.C.M.; Kessels, W.M.M. Local Deposition of High-Purity Pt Nanostructures by Combining Electron Beam Induced Deposition and Atomic Layer Deposition; American Institute of Physics: College Park, MD, USA, 2010. [Google Scholar]
  11. Chang, C.-W.; Hsu, H.-H.; Hsu, C.-S.; Chen, J.-T. Achieving Area-Selective Atomic Layer Deposition with Fluorinated Self-Assembled Monolayers. J. Mater. Chem. C 2021, 9, 14589–14595. [Google Scholar] [CrossRef]
  12. Xu, W.; Haeve, M.G.N.; Lemaire, P.C.; Sharma, K.; Hausmann, D.M.; Agarwal, S. Functionalization of the SiO2 Surface with Aminosilanes to Enable Area-Selective Atomic Layer Deposition of Al2O3. Langmuir 2022, 38, 652–660. [Google Scholar] [CrossRef]
  13. Moeini, B.; Joshua, W.P.; Avval, T.G.; Jacobsen, C.; Brongersma, H.H.; Prusa, S.; Bábík, P.; Vaníčková, E.; Argyle, M.D.; Strohmeier, B.R.; et al. Controlling the Surface Silanol Density in Capillary Columns and Planar Silicon via the Self-Limiting Gas-Phase Deposition of Tris(dimethylamino)Methylsilane, and Quantification of Surface Silanols after Silanization by Low Energy Ion Scattering. 2023. Submitted. [Google Scholar]
  14. Patel, D.I.; Major, G.H.; Jacobsen, C.; Shah, D.; Strohmeier, B.R.; Shollenberger, D.; Bell, D.S.; Argyle, M.D.; Linford, M.R. Flow-Through Atmospheric Pressure-Atomic Layer Deposition Reactor for Thin-Film Deposition in Capillary Columns. Anal. Chem. 2022, 94, 7483–7491. [Google Scholar] [CrossRef]
  15. Tonezzer, M.; Dang, T.T.L.; Bazzanella, N.; Nguyen, V.H.; Iannotta, S. Comparative Gas-Sensing Performance of 1D and 2D ZnO Nanostructures. Sens. Actuators B Chem. 2015, 220, 1152–1160. [Google Scholar] [CrossRef]
  16. Ding, M.; Guo, Z.; Zhou, L.; Fang, X.; Zhang, L.; Zeng, L.; Xie, L.; Zhao, H. One-Dimensional Zinc Oxide Nanomaterials for Application in High-Performance Advanced Optoelectronic Devices. Crystals 2018, 8, 223. [Google Scholar] [CrossRef] [Green Version]
  17. Sushma, C.; Girish Kumar, S. Advancements in the Zinc Oxide Nanomaterials for Efficient Photocatalysis. Chem. Pap. 2017, 71, 2023–2042. [Google Scholar] [CrossRef]
  18. Ellmer, K.; Klein, A.; Rech, B. Transparent Conductive Zinc Oxide: Basics and Applications in Thin Film Solar Cells; Springer: Berlin/Heidelberg, Germany, 2007. [Google Scholar]
  19. Hirao, T.; Furuta, M.; Furuta, H.; Matsuda, T.; Hiramatsu, T.; Hokari, H.; Yoshida, M.; Ishii, H.; Kakegawa, M. Novel Top-gate Zinc Oxide Thin-film Transistors (ZnO TFTs) for AMLCDs. J. Soc. Inf. Disp. 2007, 15, 17–22. [Google Scholar] [CrossRef]
  20. Guziewicz, E.; Krajewski, T.A.; Przezdziecka, E.; Korona, K.P.; Czechowski, N.; Klopotowski, L.; Terziyska, P. Zinc Oxide Grown by Atomic Layer Deposition: From Heavily n-Type to p-Type Material. Phys. Status Solidi 2020, 257, 1900472. [Google Scholar] [CrossRef]
  21. Chen, M.-C.; Chang, T.-C.; Tsai, C.-T.; Huang, S.-Y.; Chen, S.-C.; Hu, C.-W.; Sze, S.M.; Tsai, M.-J. Influence of Electrode Material on the Resistive Memory Switching Property of Indium Gallium Zinc Oxide Thin Films. Appl. Phys. Lett. 2010, 96, 262110. [Google Scholar] [CrossRef] [Green Version]
  22. Parsons, G.N. Functional Model for Analysis of ALD Nucleation and Quantification of Area-Selective Deposition. J. Vac. Sci. Technol. A 2019, 37, 20911. [Google Scholar] [CrossRef]
  23. Gladfelter, W.L. Selective Metalization by Chemical Vapor Deposition. Chem. Mater. 1993, 5, 1372–1388. [Google Scholar] [CrossRef]
  24. Longo, R.C.; McDonnell, S.; Dick, D.; Wallace, R.M.; Chabal, Y.J.; Owen, J.H.G.; Ballard, J.B.; Randall, J.N.; Cho, K. Selectivity of Metal Oxide Atomic Layer Deposition on Hydrogen Terminated and Oxidized Si(001)-(2 × 1) Surface. J. Vac. Sci. Technol. B 2014, 32, 03D112. [Google Scholar] [CrossRef] [Green Version]
  25. Hilfiker, J.N. 5-In Situ Spectroscopic Ellipsometry (SE) for Characterization of Thin Film Growth. In Woodhead Publishing Series in Electronic and Optical Materials; Woodhead Publishing: Cambridge, UK, 2011; pp. 99–151. [Google Scholar] [CrossRef]
  26. Maynard, H.L.; Layadi, N.; Lee, J.T.C. Plasma Etching of Submicron Devices: In Situ Monitoring and Control by Multi-Wavelength Ellipsometry. Thin Solid Film. 1998, 313–314, 398–405. [Google Scholar] [CrossRef]
  27. Fang, L.; Li, H.; Ma, X.; Song, Q.; Chen, R. Optical Properties of Ultrathin ZnO Films Fabricated by Atomic Layer Deposition. Appl. Surf. Sci. 2020, 527, 146818. [Google Scholar] [CrossRef]
  28. Tompkins, H.G. A User’s Guide to Ellipsometry; Courier Corporation: North Chelmsford, MA, USA, 2006. [Google Scholar]
  29. Jensen, D.S.; Kanyal, S.S.; Madaan, N.; Vail, M.A.; Dadson, A.E.; Engelhard, M.H.; Linford, M.R. Silicon (100)/SiO2 by XPS. Surf. Sci. Spectra 2013, 20, 36–42. [Google Scholar] [CrossRef]
  30. Biesinger, M.C.; Lau, L.W.M.; Gerson, A.R.; Smart, R.S.C. Resolving Surface Chemical States in XPS Analysis of First Row Transition Metals, Oxides and Hydroxides: Sc, Ti, V, Cu and Zn. Appl. Surf. Sci. 2010, 257, 887–898. [Google Scholar] [CrossRef]
  31. Wagner, C.D. Chemical Shifts of Auger Lines, and the Auger Parameter. Faraday Discuss. Chem. Soc. 1975, 60, 291–300. [Google Scholar] [CrossRef]
  32. Wagner, C.D.; Joshi, A. The Auger Parameter, Its Utility and Advantages: A Review. J. Electron Spectros. Relat. Phenom. 1988, 47, 283–313. [Google Scholar] [CrossRef]
  33. Gaarenstroom, S.W.; Winograd, N. Initial and Final State Effects in the ESCA Spectra of Cadmium and Silver Oxides. J. Chem. Phys. 2008, 67, 3500–3506. [Google Scholar] [CrossRef]
  34. Ortega, S.; Halicek, M.; Fabelo, H.; Callico, G.M.; Fei, B. Hyperspectral and Multispectral Imaging in Digital and Computational Pathology: A Systematic Review [Invited]. Biomed. Opt. Express 2020, 11, 3195–3233. [Google Scholar] [CrossRef] [PubMed]
  35. Fairley, N.; Fernandez, V.; Richard-Plouet, M.; Guillot-Deudon, C.; Walton, J.; Smith, E.; Flahaut, D.; Greiner, M.; Biesinger, M.; Tougaard, S.; et al. Systematic and Collaborative Approach to Problem Solving Using X-Ray Photoelectron Spectroscopy. Appl. Surf. Sci. Adv. 2021, 5, 100112. [Google Scholar] [CrossRef]
  36. Hesse, R.; Streubel, P.; Szargan, R. Product or Sum: Comparative Tests of Voigt, and Product or Sum of Gaussian and Lorentzian Functions in the Fitting of Synthetic Voigt-Based X-Ray Photoelectron Spectra. Surf. Interface Anal. 2007, 39, 381–391. [Google Scholar] [CrossRef]
  37. Dake, L.S.; Baer, D.R.; Zachara, J.M. Auger Parameter Measurements of Zinc Compounds Relevant to Zinc Transport in the Environment. Surf. Interface Anal. 1989, 14, 71–75. [Google Scholar] [CrossRef]
  38. Collins, T.J. ImageJ for Microscopy. Biotechniques 2007, 43, S25–S30. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structure of tris(dimethylamino)methylsilane (TDMAMS).
Figure 1. Structure of tris(dimethylamino)methylsilane (TDMAMS).
Materials 16 04688 g001
Figure 2. XPS Si 2p narrow scans of (a) bare Si\SiO2 and Si\SiO2\TDMAMS (inhibited Si\SiO2), (b) Zn 2p peaks of Si\SiO2 and Si\SiO2\TDMAMS after 50 ALD cycles of DMZ and water, and (c) Si 2p signals from the samples in (b). In these panels, the solid orange lines correspond to bare Si\SiO2, and the dotted blue lines correspond to Si\SiO2\TDMAMS.
Figure 2. XPS Si 2p narrow scans of (a) bare Si\SiO2 and Si\SiO2\TDMAMS (inhibited Si\SiO2), (b) Zn 2p peaks of Si\SiO2 and Si\SiO2\TDMAMS after 50 ALD cycles of DMZ and water, and (c) Si 2p signals from the samples in (b). In these panels, the solid orange lines correspond to bare Si\SiO2, and the dotted blue lines correspond to Si\SiO2\TDMAMS.
Materials 16 04688 g002
Figure 4. XPS of ZnO deposited on a bare Si\SiO2 substrate via 50 ALD cycles of DMZ and water. (a) Narrow scan of the Zn 2p region and (b) the Zn LMM Auger signal. These signals were fit with LA functions in CasaXPS. These peak fits were only performed to find the positions of the maxima of the Zn 2p3/2 and Zn LMM signals from which we obtained the modified Auger parameter. We do not claim that these fits have any chemical significance.
Figure 4. XPS of ZnO deposited on a bare Si\SiO2 substrate via 50 ALD cycles of DMZ and water. (a) Narrow scan of the Zn 2p region and (b) the Zn LMM Auger signal. These signals were fit with LA functions in CasaXPS. These peak fits were only performed to find the positions of the maxima of the Zn 2p3/2 and Zn LMM signals from which we obtained the modified Auger parameter. We do not claim that these fits have any chemical significance.
Materials 16 04688 g004
Figure 5. In situ ellipsometry of ZnO grown by thermal ALD of DMZ and H2O on (a) bare Si\SiO2, Si\SiO2\TDMAMS, Si\SiO2\TDMAMS damaged by heating to 330 °C in N2, and Si\SiO2\TDMAMS damaged by heating to 330 °C in the air. (b) An enlarged view of the results in (a) of the in situ ellipsometry of ZnO grown by ALD on Si\SiO2\TDMAMS. The inset of this panel shows the ZnO growth for the first 25 cycles of the deposition and a linear fit to these thicknesses.
Figure 5. In situ ellipsometry of ZnO grown by thermal ALD of DMZ and H2O on (a) bare Si\SiO2, Si\SiO2\TDMAMS, Si\SiO2\TDMAMS damaged by heating to 330 °C in N2, and Si\SiO2\TDMAMS damaged by heating to 330 °C in the air. (b) An enlarged view of the results in (a) of the in situ ellipsometry of ZnO grown by ALD on Si\SiO2\TDMAMS. The inset of this panel shows the ZnO growth for the first 25 cycles of the deposition and a linear fit to these thicknesses.
Materials 16 04688 g005
Figure 6. (a) An enlarged view of a portion of the in situ ellipsometry results in Figure 5a for the ALD of DMZ and H2O on bare Si\SiO2 showing an example of the determination/calculation of the growth per cycle (GPC). (b) The GPC values for the ALD of DMZ and H2O on bare Si\SiO2. The red line through the data is a guide for the eye. The parameter ' G P C ' is defined as the average GPC at and beyond 20 cycles where growth is nearly constant.
Figure 6. (a) An enlarged view of a portion of the in situ ellipsometry results in Figure 5a for the ALD of DMZ and H2O on bare Si\SiO2 showing an example of the determination/calculation of the growth per cycle (GPC). (b) The GPC values for the ALD of DMZ and H2O on bare Si\SiO2. The red line through the data is a guide for the eye. The parameter ' G P C ' is defined as the average GPC at and beyond 20 cycles where growth is nearly constant.
Materials 16 04688 g006aMaterials 16 04688 g006b
Figure 7. (a) Normalized growth per cycle and (b) selectivities as a function of the number of ALD cycles for the ZnO ALD growth on bare Si\SiO2, Si\SiO2\TDMAMS, and Si\SiO2\TDMAMS heated to 330 °C in N2 or the air in Figure 5a.
Figure 7. (a) Normalized growth per cycle and (b) selectivities as a function of the number of ALD cycles for the ZnO ALD growth on bare Si\SiO2, Si\SiO2\TDMAMS, and Si\SiO2\TDMAMS heated to 330 °C in N2 or the air in Figure 5a.
Materials 16 04688 g007
Figure 8. LEIS spectra of Si\SiO2 passivated with one pulse of TDMAMS and then reacted with DMZ (Si\SiO2\TDMAMS(1)\Zn(1)) and Si\SiO2 passivated with ten pulses of TDMAMS and then reacted with DMZ (Si\SiO2\TDMAMS(10)\Zn(1)). The inset of this figure is an enlargement of the Zn region of the spectra shown in the main panel.
Figure 8. LEIS spectra of Si\SiO2 passivated with one pulse of TDMAMS and then reacted with DMZ (Si\SiO2\TDMAMS(1)\Zn(1)) and Si\SiO2 passivated with ten pulses of TDMAMS and then reacted with DMZ (Si\SiO2\TDMAMS(10)\Zn(1)). The inset of this figure is an enlargement of the Zn region of the spectra shown in the main panel.
Materials 16 04688 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Moeini, B.; Avval, T.G.; Brongersma, H.H.; Průša, S.; Bábík, P.; Vaníčková, E.; Strohmeier, B.R.; Bell, D.S.; Eggett, D.; George, S.M.; et al. Area-Selective Atomic Layer Deposition of ZnO on Si\SiO2 Modified with Tris(dimethylamino)methylsilane. Materials 2023, 16, 4688. https://doi.org/10.3390/ma16134688

AMA Style

Moeini B, Avval TG, Brongersma HH, Průša S, Bábík P, Vaníčková E, Strohmeier BR, Bell DS, Eggett D, George SM, et al. Area-Selective Atomic Layer Deposition of ZnO on Si\SiO2 Modified with Tris(dimethylamino)methylsilane. Materials. 2023; 16(13):4688. https://doi.org/10.3390/ma16134688

Chicago/Turabian Style

Moeini, Behnam, Tahereh G. Avval, Hidde H. Brongersma, Stanislav Průša, Pavel Bábík, Elena Vaníčková, Brian R. Strohmeier, David S. Bell, Dennis Eggett, Steven M. George, and et al. 2023. "Area-Selective Atomic Layer Deposition of ZnO on Si\SiO2 Modified with Tris(dimethylamino)methylsilane" Materials 16, no. 13: 4688. https://doi.org/10.3390/ma16134688

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