Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Influence of temperature and plasma parameters on the properties of PEALD HfO2

Open Access Open Access

Abstract

HfO2 has promising applications in semiconductors and optics due to its high dielectric constant and high refractive index. In this work, HfO2 thin films were deposited by plasma enhanced atomic layer deposition (PEALD) using tetrakis-dimethylamino hafnium (TDMAH) and oxygen plasma. The process optimization to obtain high quality HfO2 thin films with excellent uniformity over a 200 mm diameter is thoroughly discussed. The effects of deposition temperature and plasma parameters on the structural, mechanical, and optical properties, and the chemical composition of the films were investigated. Optimized process parameters yielding a high refractive index, high density, low impurities, low OH incorporation, low absorption in the UV spectral range, and high laser-induced damage threshold (LIDT) were selected for antireflection coatings. The HfO2 thin films were incorporated into antireflection coatings designed for the fundamental wavelength at 1064 nm and its higher harmonics up to the 4th order.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Downscaling of semiconductor devices necessitates a thin dielectric gate oxide layer (typically < 2 nm) [1,2]. Traditionally used SiO2-based gate oxides have reached their operational limit due to a high gate leakage current of ultrathin layers [3]. HfO2 is a promising alternative to SiO2 in semiconductor technology, e.g., in metal-oxide-semiconductor field-effect transistors (MOSFET) due to its high dielectric constant and high density [46]. Due to the shrinking size of memory devices, HfO2-based resistive random access memory (RRAM) devices are promising candidates for the new generation of non-volatile memory units because of complementary metal-oxide-semiconductor (CMOS) compatibility, high switching speed, and increased endurance [7]. HfO2 is also widely used in optical coatings, e.g., in antireflection coatings, mirrors, and optical filters due to its high refractive index in the ultraviolet, visible, and infrared spectral range [811]. HfO2 has a high melting point, as well as high thermal and chemical stability; hence it is widely studied experimentally and theoretically [1214].

A low deposition temperature is often desired for optical applications to obtain relatively thick, amorphous, and smooth layers. High-power laser applications mandate thin films with low impurities to guarantee low optical losses and a high laser-induced damage threshold (LIDT) [15]. The current development of quantum technologies relies on optical components that put even more stringent requirements on thin film properties and optical performance. Traditionally, HfO2 thin films are deposited by physical vapor deposition [1620] or chemical vapor deposition techniques [17,18]. However, for optimal performance of nanostructured components or strongly curved substrates, conformal and uniform coatings are demanded. Since atomic layer deposition (ALD) is based on self-limiting surface reactions and cyclic layer-by-layer growth with alternately pulsed precursors into the reaction chamber, it provides precise thickness control, a uniform coating on large-area substrates [2125] and conformal growth on 3D and high aspect-ratio surfaces [22,2629].

In conventional thermal ALD processes, the surface reactions are activated by heating the substrate [2931]. Various metal-organic precursors with functional groups like halides [3236], alkyls [3638], alkoxides [36,3941], amides [3638,4245], β-diketonates [3638], and cyclopentadienyls [3638,4648] have been used to develop thermal ALD HfO2 thin films. High deposition temperatures are required to break the relatively strong metal-halide and metal-oxide bond in HfO2 precursors containing halide and alkoxide functional groups [38]. Additionally, the by-product resulting from the metal halide reaction with surface functional groups are highly corrosive and might damage the substrate or contaminate the film [38]. The precursors containing β-diketonates and cyclopentadienyl functional groups are relatively unstable at low temperatures. Moreover, they are bulky and cause steric hindrance [38]. In contrast, amide precursors are volatile and relatively reactive at low deposition temperatures [38]. Hence, they have found broad interest in the growth of HfO2 thin films. Thermal ALD processes have been reported for various alkyl amide precursors and oxidizing agents such as H2O [4245,4954] and D2O [55,56], H2O2 [57,58], molecular O2 [59,60], or ozone [6164].

High carbon and nitrogen impurities have been observed in HfO2 thin films deposited by thermal ALD at high deposition temperatures (>300°C) due to decomposition when using amide precursors [43,45,65,66] and at low deposition temperatures (<200°C) due to insufficient activation energy of precursor reactions [60,64,67]. At high temperatures, the chemical bonds break as the energy of surface species exceeds the activation energy. On the other hand, a high hydrogen content has been observed in the films deposited using an alkyl amide precursor and water at low deposition temperatures due to incomplete surface reactions [45,66,68,69]. The hydrogen impurities can be reduced by introducing longer purging durations after the alkyl amide precursor pulse or/and by post-deposition annealing [49,68,70]. In this context, plasma enhanced atomic layer deposition (PEALD) can provide dielectric films with high purity, high density, and high refractive index at lower deposition temperature compared to thermal ALD [30,69,71,72]. In PEALD, highly energetic plasma reactants are used as an oxidizing agent, enabling a low-temperature deposition with good thin film quality. Plasma enhanced ALD processes of HfO2 using metal-amides, metal-cyclopentadienyls, metal-alkoxides, and metal β-diketonates have previously been reported [9,13,24,26,28,36,50,64,7379].

In this study, we present a thorough approach in the development of HfO2 thin films grown by PEALD using tetrakis-dimethylamino-hafnium (TDMAH) and oxygen plasma. The influence of the deposition temperature and plasma parameters were studied. So far, structural, optical, and mechanical properties of HfO2 layers deposited by PEALD have not been systematically investigated in the literature. Most ALD process development articles report on optimizing the pulse/purge durations of the precursors towards a constant growth rate per cycle (GPC) [64,74,76,78]. Detailed optimization procedures to achieve an optimum thickness distribution, i.e., minimum non-uniformity (NU), are rarely considered. Due to the strong influence of the reactor geometry on the NU, specific gas flow rates and pulse durations are required for each reactor design [80]. Thin film properties strongly depend on deposition parameters like temperature, plasma conditions, reactants, gas flow rates, and pulse and purge durations. To meet the optical coatings requirements with respect to optical constants, surface roughness, impurities, and optical losses, a reactor-specific process development and an in-depth investigation of the thin film properties are required. Furthermore, we discuss the impact of different plasma configurations on these properties to prepare an optimal basis for the incorporation of hafnium oxide thin films in optical interference coatings.

High power laser systems warrant high-resistance coatings on optical elements. Conventionally these coatings are prepared using e-beam evaporation [8187], ion assisted deposition [82,84,87], ion beam sputtering [88,89], magnetron sputtering [84,88,90], RF sputtering [91], and sol-gel methods [84,92,93]. In a thin film laser damage competition held at the Boulder Damage Symposium 2010, antireflection coatings manufactured with six different techniques were submitted and tested at 355 nm laser wavelength with a 7.5 nm pulse length [84]. It was shown that the coatings exhibited a high laser induced damage threshold (LIDT), but these values strongly depend on the selected materials and the deposition method.

Recently, ALD has raised interest for high power laser applications due to its ability to coat 3D substrates with precise thickness control, good uniformity and conformality [88,9496]. Although the LIDT of ns-pulsed lasers is mainly defect driven (particles, defects, etc.), the laser beam parameters like wavelength, beam diameter, or the absorption in the film also play a significant role. The absorption in the film can cause thermal effects where heat diffusion resulting from the high energy of lasers increases laser damage probability and lowers LIDT. We demonstrate highly efficient and low absorption antireflection coatings consisting of hafnium oxide, silicon oxide, and aluminum oxide prepared by PEALD. The antireflection coatings were designed for the fundamental wavelength of a Nd:YAG laser at 1064 nm and the higher harmonics at 532 nm, 355 nm, and 266 nm.

2. Experimental techniques

Plasma enhanced ALD was employed to deposit hafnium oxide thin films. Tetrakis-dimethylamino hafnium (TDMAH) [Strem Chemicals Inc., Bischheim, France] was used as a metal-organic precursor and oxygen as a plasma gas using a Silayo ICP330 equipment [Sentech Instruments GmbH, Berlin, Germany]. A detailed description of the deposition tool is given elsewhere [97,98]. Various substrates were used to satisfy the requirements for the experimental investigations. For the initial optimization of the non-uniformity and control of the GPC, 9 single-side polished, p-doped and (100) oriented Si substrates were arranged in a concentric circular pattern on the heating plate in the reactor: 1 in the center, 4 at 100 mm diameter, and the remaining 4 at 200 mm diameter around the center of the heating plate. The UV/VIS spectroscopy, LIDT, and the absorption loss measurements were carried out on HfO2 thin films deposited on vertically (double-side coating) and horizontally (single-side coating) placed fused silica substrates. For the atomic force microscopy (AFM), X-ray diffraction (XRD), X-ray reflectometry (XRR), Fourier transform infrared (FTIR) spectroscopy, spectroscopic ellipsometry (SE), and Auger electron spectroscopy (AES) analysis, silicon substrates with a native SiO2 layer were used. Furthermore, 75 mm Si (100) double-side polished wafers were employed for determining the mechanical stress.

The process optimization was carried out applying 200 ALD cycles. Further, the effect of temperature and plasma parameters on the properties of HfO2 were investigated on thicker films deposited with 1482 ALD cycles, whereas the density of the films was investigated from thin HfO2 films grown with 184 ALD cycles (∼20 nm thickness). The film thicknesses and optical constants were determined by spectroscopic ellipsometry using the 850 spectroscopic ellipsometer and the SpectraRay 3 software [Sentech Instruments GmbH, Berlin, Germany]. The measurements were performed in the spectral range from 200 to 980 nm wavelength. The Kramers-Kronig consistent Tauc-Lorentz dispersion model using three oscillators was used to fit the refractive index and the extinction coefficient. For film thicknesses over ≈100 nm, an additional effective medium approximation (EMA) layer using the Bruggemann model was added to improve the fit by taking into account the surface roughness into the model. The fraction of void inclusion was assumed to be 0.5. During the process optimization, the thickness non-uniformity (NU) across 200 mm diameter area was estimated by (dmax-dmin)/2daverage, whereby dmax, dmin and daverage are the maximum, minimum and average thickness of HfO2 films, respectively. Additionally, optical constants of HfO2 films deposited on fused silica substrates have been estimated via spectrophotometry by modeling the transmittance and reflectance spectra using the LCalc software with a Lorentzian multi-oscillator dispersion model [99]. The model consists of fused silica substrate, HfO2 thin film, and a roughness layer with 50% void inclusion. The hafnia dispersion curves were modelled by using 7–9 Lorentzian oscillators.

The reflectance (R) and transmittance (T) spectra were recorded with a Lambda 950 spectrophotometer [PerkinElmer Inc., Waltham, MA, USA] in a wavelength range from 200 to 1200 nm and at an angle of incidence at 6°. The optical losses (OL), the sum of the scattering and the absorption, were calculated from the R/T data using OL = 100 - R - T. For validation of the film thickness, x-ray reflectometry (XRR) was performed using a Bruker D8 Discover system [Bruker AXS, Karlsruhe, Germany]. The mechanical stress was determined by measuring the change in curvature of a double-side polished Si (100) wafer (400 µm thickness, 75 mm diameter) before and after deposition [Model FLX-2320, KLA-Tencor, San Jose, CA, USA]. The stress was calculated using the Stoney equation [100]:

$$\sigma = \; \frac{1}{6}\frac{{{E_s}}}{{({1 - {\vartheta_s}} )}}\left( {\frac{1}{{{R_f}}} - \; \frac{1}{{{R_s}}}} \right)\frac{{t_s^2}}{{{t_f}}}$$
where Es is the Young's modulus, ϑS is the Poisson’s ratio of the substrate, Rs and Rf are the radii of curvature of the substrate before and after coating, respectively, ts is the substrate thickness, and tf is the film thickness. Positive values of stress represent tensile stress and negative values correspond to compressive stress. The surface roughness and topography of HfO2 thin films were investigated by atomic force microscopy (AFM). The measurements were performed by a Veeco Dimension 3100 [Veeco Instruments, Santa Barbara, CA, USA]. The crystallinity of the films was investigated by x-ray diffraction (XRD) using a Bruker D8 Discover system [Bruker AXS, Karlsruhe, Germany] in a Bragg-Brentano geometry using Cu-Kα radiation (λ = 0.154 nm). The 2θ measurement range was from 10° to 65°. The chemical structure, particularly the amount of –OH bonds, was investigated by Fourier transform infrared spectroscopy (FTIR) [Varian Inc., Palo Alto, CA, USA] in a measurement range from 400 to 4000 cm−1. To investigate the stoichiometry, carbon and nitrogen impurities, the depth profiles were recorded by Auger electron spectroscopy (AES) using an Auger cylindrical mirror spectrometer [Varian Inc., Palo Alto, CA, USA]. Thin films were sputtered by Kr ions (2 keV energy, 10 µA current), and a focused electron beam with 5 keV energy was directed on the samples at an incidence angle of 30°.

The laser-induced damage threshold (LIDT) was measured at Layertec GmbH under normal incidence using a Nd:YAG, frequency-tripled, Q-switched LITRON NanoTRL-650-10 laser [Litron Lasers, Rugby, UK] at 355 nm with 7 ns pulse duration (FWHM), 10 Hz repetition rate and approximately 240 µm beam diameter (1/e2). Because of the limited test area available, the R-on-1 test method was used. This testing method is useful for production monitoring as well as comparison of manufacturing processes for vendors. In this investigation, each test site was irradiated with 10 pulses of 5 J/cm2 initial fluence. The pulse burst was reapplied with fluence increments of 5 J/cm2 until damage occurred. Damage onset was detected in-situ using the scattering method. For each test sample, roughly 40 test sites were irradiated to ensure statistical reliability. From these results, damage probability over fluence was calculated. Each pulse was interpretated as an individual test. This way, the number of available datapoints, i.e., pulse fluence and the information whether or not damage occurred, greatly increases and the statistical error is further reduced. The damage probability P for a given fluence F is then determined using equation: P(F) = n1(F)/(n1(F) + n0(F)), where n1(F) is the number of damaged test sites and n0(F) is the number of undamaged ones at that fluence. In accordance with ISO 21254, a linear regression of P(F) in the transition region from 0 to 100% damage probability was performed and LIDT was determined by the intersection of the fit function and the fluence axis. However, due to aging and conditioning effects, the R-on-1 regime results may differ from those in the S-on-1 regime.

Absorption losses were measured at 355, 532, and 1030 nm wavelengths via photothermal common path interferometry (PCI), using a custom-built set-up of Layertec GmbH. In PCI, optical layers are irradiated by a pump beam, and part of the incident radiation is absorbed by the layers. Due to thermal conduction, the absorbed energy propagates from the layers into the underlying substrate, creating a thermal lens. The phase-front deviations of a probe beam resulting from the thermal lens were interferometrically measured. By employing a modulated pump beam and a lock-in amplifier, resolution in the sub-ppm range can be achieved.

3. Results and discussion

3.1 Process development

The initial development of the hafnium oxide processes aimed at finding adequate deposition conditions for the precursor pulse and purge durations, whereby a low thickness non-uniformity (NU) and a constant growth per cycle (GPC) were desired. Therefore, a series of processes varying different parameters like TDMAH pulse duration, carrier gas (Ar) flow rate, purge duration after TDMAH pulse, and O2 plasma pulse duration were carried out at 100°C deposition temperature and 200 ALD cycles. Additionally, the focus of the work was to keep the total cycle time as short as possible without compromising the film quality. The saturated growth rate and thickness uniformity of PEALD films depends on several factors such as reactor type, precursor exposure time, reaction mechanisms, and sticking probabilities [22]. The GPC and NU values obtained in the initial development processes are summarized in Figs. 1(a)–1(d).

 figure: Fig. 1.

Fig. 1. The effect of process parameters on the non-uniformity (NU) HfO2 thin films on a 200 mm diameter surface and growth per cycle (GPC) deposited varying a) the precursor (TDMAH) pulse with a carrier gas (Ar) flow rate of 80 sccm b) the carrier gas (Ar) flow rate with TDMAH pulse of 3.12 s c) the precursor (TDMAH) pulse with a carrier gas (Ar) flow rate of 160 sccm, and d) the precursor and plasma purge with precursor (TDMAH) pulse of 3.12 s and Ar flow rate of 160 sccm. All thin films were deposited at 100°C with 200 ALD cycles. The sequence of the PEALD cycle is given where x denotes the varied parameters.

Download Full Size | PDF

The process optimization started with a constant argon gas flow rate of 80 sccm while the precursor pulse duration was varied from 0.62 to 4.12 s (Fig. 1(a)). The durations of the plasma pulse, precursor purge, and plasma purge were kept constant at 5 s. At a TDMAH pulse of 0.62 s, a relatively high NU (18.4%) and a GPC of 1.88 Å/cycle were observed. This indicates the absence of the self-limiting surface reaction regime and poor gas distribution in the reactor. The lowest NU of 3.4% was observed at 3.12 s TDMAH pulse. Additionally, at this TDMAH pulse duration, the GPC reached the highest value of 2.55 Å/cycle. Although the NU decreased significantly at 3.12 s, the high GPC may be due to a lower film density and the formation of voids resulting in higher incorporation of impurities. The gas flow rate of 80 sccm (TDMAH pulse < 3 s) resulted in a thickness gradient across the surface area (see Supplement 1 Fig. S1). A higher thickness was observed close to the precursor inlet and lower thickness at the maximum distance from the inlet. It is important to note that the same gas flow rate is also used for purging reactant by-products after the TDMAH pulse. Hence, a carrier (and purging) gas flow rate of 80 sccm and 3.12 s of TDMAH pulse may be sufficient to cover the surface but inadequate to efficiently remove reactant by-products for this reactor design [45].

In the second optimization phase (step 2), the TDMAH carrier gas flow was varied (Fig. 1(b)) while the durations of the precursor pulse, the plasma pulse, the precursor purge, and the plasma purge were 3.12 s, 5 s, 5 s and 5 s, respectively. Due to the optimal distribution of the precursor and effective flow rate during purging, the lowest NU of 0.5% was achieved for 160 sccm argon gas flow and 3.12 s precursor pulse duration. The average GPC decreased from 2.58 Å/cycle to 1.92 Å/cycle. While a relatively constant GPC was obtained in the first optimization step with the precursor pulse variation, a significant influence of the gas flow rate on the GPC and NU is observed in the second optimization phase. These results are critical towards ALD process optimization since gas flow optimization is rarely considered in the literature. Further details on the film properties were not discussed here, since the NU is already a main process selection criterion in the development of optical and semiconductor devices.

In step 3, considering the optimized GPC and NU obtained from step 1 and step 2, the precursor pulse duration was varied again from 0.62 s to 4.12 s, using the higher Ar flow rate of 160 sccm (Fig. 1(c)). The durations of the plasma pulse, precursor purge, and plasma purge were 5 s. The NU decreased significantly with increasing TDMAH pulse duration and attained the lowest value of 0.5% at 3.12 s of TDMAH pulse. This indicates a uniform distribution of the precursor molecules in the chamber and hence uniform surface coverage of the substrate. The GPC increased significantly from 1.51 Å/cycle to 1.88 Å/cycle on increasing the duration of TDMAH pulse from 0.62 s to 4.12 s. The low GPC at lower precursor pulse duration may be due to the unsaturated surface as the result of an insufficient TDMAH dose. The surface saturation occurs when the duration of the TDMAH pulse increased beyond 1.12 s, and consequently, a higher GPC of 1.88 Å/cycle is observed. A similar trend was reported by Kim et al. for HfO2 films deposited at 80°C [64]. The GPC of 1.88 Å/cycle at 3.12 s TDMAH pulse is lower compared to the GPC obtained using 80 sccm Ar gas flow rate. This may be the result of a more efficient removal of residual impurities after the TDMAH pulse or loosely bound precursor residuals.

Finally, the durations of the precursor and plasma purge times were decreased from 5 s to 2 s to investigate the shortest purge duration required for the complete removal of reaction by-products (Fig. 1(d)). Shorter purge durations are desirable in ALD processes for the sake of cost and time efficiency. It is important to note that, in this investigation, the duration of the plasma pulse was reduced to 3 s as compared to 5 s used in the optimization steps 1–3, as it has a negligible impact on the NU and GPC. However, the detailed investigation of the effect of plasma pulse duration on material properties of HfO2 will be discussed in a later section.

Initially, the plasma purge duration was kept at 5 s, and the precursor purge duration was varied. With decreasing precursor purge duration from 5 s to 4 s, the GPC remained reasonably constant at 1.89 Å/cycle, whereas the NU increased slightly from 0.5% to 0.7%. For a precursor purge duration below 4 s, another minor increase of the NU to 0.8% was observed, whereas the GPC increased significantly, indicating an increase of impurities and inefficient removal of precursor residuals. Hence, the precursor purge of 5 s was considered as an optimum parameter, and the effect of the plasma purge duration was investigated.

For the plasma purge duration of 5 s and 4 s, the GPC and NU remained reasonably constant at 1.90 Å/cycle and 0.5%, respectively. On further decreasing plasma purge duration below 4 s, there was a slight increase in GPC, whereas a significant increase in NU was observed. However, NU remained below 1%. The higher GPC below 4 s of precursor purge and plasma purge indicate incomplete removal of residual precursor molecules and reaction by-products.

Consequently, for the optimized process for PEALD hafnia using TDMAH and oxygen plasma in the Silayo Sentech ICPEALD tool, the following process parameters were considered: 3.12 s precursor pulse, 160 sccm Ar carrier gas flow, 5 s precursor purge, and 5 s plasma purge. Using the above-mentioned optimized parameters and 5 s plasma pulse, the GPC of HfO2 thin films was found to be 1.88 Å/cycle, and thickness NU was 0.5% over a 200 mm diameter surface.

The HfO2 thin films deposited using various precursors, oxygen sources, and deposition parameters have been summarized by Blaschke et al. and Sharma et al. [28,45]. The GPC value observed in our study was significantly higher than 1.50 Å/cycle, as reported by Kim et al. using the same precursor and oxygen plasma at 100°C [71]. This may be due to the higher reactivity of plasma species generated by the uniquely designed planar triple spiral antenna (PTSA) source [97,98,101]. Additionally, there was a constant supply of Ar and oxygen during the entire duration of deposition. The generated plasma consists of O and Ar radicals and ions. Increased GPC of HfO2 films in the presence of Ar during the O2 plasma pulse was reported earlier [76]. It was shown that HfO2 thin films deposited using tris(dimethylamido) cyclopentadienylhafnium (TDMACpHf) and Ar/O2 plasma had a 1.6 times higher GPC compared to pure oxygen plasma. A fraction of argon in the mixed plasma is excited to metastable Ar* or ionized Ar+ by electron collisions. The non-radiative momentum transfer after collisions of Ar* atoms with electrons enhances the kinetic energy of low energy electrons of the O2 plasma. Hence, the addition of Ar increases the amount of high-energy electrons in the plasma, which further increases the amount of oxygen radicals and thereby the GPC [76,102]. It is important to emphasize that the GPC and NU depend on the reactor configurations, even when the same chemicals and similar deposition temperatures are employed. The pressure conditions, which are also influenced by the purge gas flow rates, pumping configurations, etc., will affect the surface functional groups and consequently the chemical reactions.

3.2 Influence of the deposition temperature

It is widely known that thin films of several high refractive index oxides may undergo significant morphological changes with increasing film thickness. The deposition temperature also critically influences the film morphology. Hence, the properties of a thin (≈30 nm/184 ALD cycles) and a thick (≈250 nm/1482 ALD cycles) HfO2 film were investigated. HfO2 samples were deposited at a temperature ranging from 100 to 250°C using 160 sccm carrier (Ar) gas flow. Figure 2 shows the variation of the NU and GPC of the HfO2 films deposited using 1482 cycles and an ALD cycle sequence of 3.12s/5s/3s/5s as function of the deposition temperature. The HfO2 film deposited at 100°C shows a GPC of 1.80 Å/cycle and a NU of 1.3%. Noteworthy, the NU deteriorated slightly for the thick film compared to the thin film. There is also a slight variation of the GPC, although all process conditions are kept constant. This could be related to the influence of the substrate and reactor conditioning, or to the precursor conditioning. The GPC significantly decreases to 1.71 Å/cycle, and the NU slightly increases to 1.6% at a deposition temperature of 150°C. On further increasing deposition temperature to 200°C, the GPC increases slightly to 1.78 Å/cycle, whereas the NU increases to 1.8%. At the highest deposition temperature of 250°C, the GPC increases further and reaches a maximum value of 2.05 Å/cycle, whereas the NU decreases slightly to 1.7%. The NU for thick films is in the range of 1.3% to 1.8% and has not been further optimized since for antireflection coatings rather thin films (d<50 nm) are required (see section 3.4).

 figure: Fig. 2.

Fig. 2. The effect of deposition temperature on NU and GPC. The thin films were deposited using 1482 ALD cycles, 160 sccm Ar flow and an ALD sequence of 3.12s/5s/3s/5s.

Download Full Size | PDF

A similar result of decreasing GPC with increasing deposition temperature from 230°C to 350°C was reported by Heil et al. for PEALD HfO2 films deposited using tetrakis(ethylmethylamido)hafnium (TEMAH) precursor and O2 plasma [78]. The slightly higher GPC at low deposition temperature was attributed to a low film density and a high H content. Hence, the films deposited at 100°C in this study may also possess higher hydrogen content originating from the incorporation of residual OH groups. Due to a higher film density, the H content could have decreased when the temperature was increased to 150°C (Table 1). A similar parabolic dependence of the GPC on the deposition temperature of HfO2 thin films deposited by thermal ALD (TALD) using TDMAH and H2O has been observed [49]. The minimum GPC of 1.0 Å/cycle was found at 250°C for HfO2 deposited by thermal ALD for TDMAH + H2O and a GPC of 1.2 Å/cycle using TEMAH + H2O at 275°C. Hence, it is important to mention that the ALD window for TALD HfO2 films using TDMAH precursor is at lower temperatures compared to TEMAH precursor, which is critical for precursor selection for low-temperature deposition. Additionally, it is well known that the temperature window of thermal ALD can be shifted to lower deposition temperatures by using PEALD [71]. The high GPC of HfO2 at the deposition temperature of 250°C may occur either due to an increased surface reactivity and/or precursor decomposition [46]. This increase in the surface reaction rate may also result in an altered surface morphology as observed in AFM and SEM images (Figs. 3(a)–3(d)).

 figure: Fig. 3.

Fig. 3. AFM and SEM images of the film deposited at 100°C and 250°C with 1482 ALD cycles; a) AFM image of a hafnium oxide thin film, d ≈ 273 nm, deposited at 100°C with a roughness RMS = 0.5 nm; b) AFM image of hafnium oxide deposited at 250°C with d = 309 nm and RMS = 7.3 nm; SEM images of hafnium oxide at c) 100°C and d) 250°C, respectively; crystalline grains are observed at 250°C while the films grow rather smooth at 100°C. The thin films were deposited using an ALD sequence of 3.12s/5s/3s/5s.

Download Full Size | PDF

Tables Icon

Table 1. Thin film properties of the HfO2 deposited at 100°C - 250°C.a

The HfO2 film deposited at 100°C (Figs. 3(a) and 3(c)) and 150°C shows a smooth surface with nearly constant root-mean-square (RMS) surface roughness (Table 1). It increases linearly from 0.5 to 10.1 nm on increasing deposition temperature from 100°C to 200°C (Table 1) due to the onset of crystallization, as evident from the XRD pattern shown in Fig. 4. However, at the deposition temperature of 250°C, the roughness decreases to 7.25 nm due to the coalescence of islands [45] (Fig. 3(d)) and formation of laterally oriented grains of different orientations [28,33,53] confirmed in the XRD pattern shown in Fig. 4. HfO2 is well known to be prone to crystallization when deposited at elevated temperatures [28,45,49,53,61,64,68,74,96]. The SEM image in Fig. 3(d) shows polycrystalline HfO2 with the grain size of 22 to 25 nm (calculated using the Scherrer equation from XRD data).

 figure: Fig. 4.

Fig. 4. XRD diffractogram of HfO2 films (1482 ALD cycles) deposited at deposition temperature from 100°C to 250°C. The peak positions corresponding to tetragonal (ICDD: PDF card No. 08–0342), orthorhombic (ICDD: PDF card No. 21-0904) and monoclinic (ICDD: PDF card No. 43-1017) phase are given as reference. The thin films were deposited using an ALD sequence of 3.12s/5s/3s/5s.

Download Full Size | PDF

The effect of deposition temperature on the surface roughness due to crystallization is further supported by the XRD measurements shown in Fig. 4. The HfO2 films deposited at 100°C and 150°C show a broad feature around 2θ = 32°, indicating amorphous films [28,45,64]. At a deposition temperature of 200°C, the onset of crystallization occurs. At the deposition temperature of 250°C, intense peaks with different phase orientations are visible, indicating an increase in the degree of crystallization. Contrary to previous studies, the microstructure transition of HfO2 thin films from amorphous to polycrystalline at deposition temperatures below 250°C in this study may be due to increased surface energy via momentum transfer of high energetic plasma species in PEALD [27,103]. Hence, a balance between the thermal energy and the energy of the plasma species must be found to grow smooth but pure and dense thin films. The increase in crystallization with an increase in deposition temperature also influences the density of HfO2 films.

The density of HfO2 films deposited at 100°C is 7.7 g/cm3 as determined by XRR. A similar value was reported by Provine et al. for HfO2 films deposited at 200°C [100] by TDMAH precursor and O2 plasma [59]. The density obtained here at 100°C is slightly higher than the reported value of 7.3 g/cm3 for HfO2 film deposited at 230°C using TEMAH and O2 plasma [78]. Kim et al. have reported a density of 8.1 g/cm3 for HfO2 film deposited at 80°C using TEMAH and oxygen plasma. On increasing deposition temperature from 100 to 250°C, the density of the HfO2 films increased and approached the bulk value of 9.68 g/cm3, which is comparable to the densities observed in HfO2 films deposited by IBS, IAD, PIAD with xenon as working gas and ion plating [104,105]. A similar effect of an increase in density with increasing deposition temperature has been reported earlier [28,74,78]. The increase in density has been attributed to decreased C, N, and H impurities in the film. The hydrogen impurities in the films are mainly present in the form of hydroxyl groups. Therefore, on increasing substrate temperature, the peaks corresponding to isolated OH groups decreased significantly, as shown in the FTIR spectra (Fig. 5). Additionally, the impurities, such as carbon and nitrogen in the HfO2 films, act as defects and result in deteriorated film properties [64]. The decrease in the impurity content with increasing deposition temperature (Table 1) is confirmed by AES measurements. The carbon content in the films decreased from 12.1% to 5.2%, and the nitrogen content decreased from 3.6% to 1.2% on increasing deposition temperature from 100°C to 250°C. Although the impurities content decreases in HfO2 films deposited at a higher temperature, the absorption (Table 2) increases significantly, possibly due to defects in the film in the form of oxygen vacancies [94]. Noteworthy, the density could be increased at 100°C by applying a longer plasma pulse.

 figure: Fig. 5.

Fig. 5. The region of FTIR spectra corresponding to -OH stretching for the HfO2 films (1482 ALD cycles) deposited at a temperature from 100°C to 250°C.

Download Full Size | PDF

Tables Icon

Table 2. Thin film properties of HfO2 films (1482 cycles) deposited at 100°C to 250°C.

The residual tensile stress of HfO2 films increases from 588 MPa to 916 MPa with increasing deposition temperature from 100°C to 250°C. The residual intrinsic stress in thin films can be influenced by several factors like thickness, density, grain growth, grain boundary generation, surface roughness, voids, and impurities [81,106,107]. As the deposition temperature increases, crystallization occurs, as confirmed by larger grains observed in Fig. 3(d). The change in the structure and morphology of ALD HfO2 with the increase in deposition temperature is similar to the effect of annealing temperature on HfO2 films reported by Shen et al. [108]. As the temperature increases, the crystallite grain size increases, and the grain boundary area decreases. Since the density of the film at the grain boundaries is low, the reduction in grain boundaries results in the lateral film contraction, which leads to an increase in tensile stress [109,110]. This can be further explained in terms of the total energy change of the film. The two main contributions leading to the change in the total energy of the film are the free surface energy of the crystallites and grain boundary free energy. As the grain growth in the amorphous matrix proceeds, crystallites snap together to form grain boundaries; therefore, the energy gained by grain boundary formation between the crystallites replaces their free surface energy; thereby leading to high tensile stress [111,112].

Additionally, a decrease in micro defects and pores containing OH (loosely bound hydroxyl groups or H2O, which are not entirely removed at low temperature) can lead to an increase in tensile stress. This is confirmed by the reduction in OH content with an increase in deposition temperature, as shown in Fig. 5. A similar observation of decreasing OH content with increasing deposition temperature for PEALD HfO2 was previously reported [9]. Thin films with incorporated water mainly possess compressive stress due to the repulsion of the permanent dipole moments of the adsorbed water molecules on the walls of the pores [113115]. On increasing the deposition temperature or a post-deposition annealing, the films densify, and porosity decreases. As a result, the adsorbed water and OH amount decreases with increasing temperature to 250°C, and consequently, tensile stress increases from 588 MPa to 916 MPa [9,49,116]. The change in density on varying deposition temperature is also reflected in the refractive index of HfO2 films (see Table 1).

A relatively low refractive index of 1.92 (at 633 nm wavelength) was observed for the HfO2 films deposited at 100°C (Table 1 and Fig. 6(a)). At a deposition temperature of 250°C, the refractive index reaches the maximum value of 2.07 as the consequence of increased density in accordance with previous studies [68,96,117]. The absorption edge of HfO2 thin films deposited at higher temperatures shifts towards lower wavelengths as seen by the shift in the extinction coefficient k towards lower values (Fig. 6(b)). It indicates an increase in optical band gap with increase in deposition temperature. The Tauc-Lorentz analysis of the ellipsometry data in contrast, indicates increasing extinction coefficient with increasing temperature. It is important to note that the increase in extinction coefficient already starts below 400 nm wavelength for HfO2 films deposited at 200 and 250 °C which can result in high absorption and high scattering losses (data not shown). Defects like oxygen vacancies in HfO2 thin films can increase absorption losses, which significantly affects the laser-induced damage threshold (LIDT) [96,118]. Absorption losses determined with high accuracy at selected wavelengths are summarized in Table 2. The absorption for all HfO2 thin films at 1030 nm before annealing was 2 ppm, which is comparable to 3 ppm reported by Liu et al. at 1064 nm for the HfO2 films deposited using TEMAH and H2O at 200°C [95]. The actual absorption was reported to be between 1 and 2 ppm for the thermal process as the slightly higher absorption of 3 ppm was contributed by the fused silica substrate. The lowest absorption is observed for PEALD HfO2 films deposited at 150°C at all three wavelengths (Table 2). The absorption in the HfO2 films increases with increasing deposition temperature and annealing, which may be attributed to the increase in crystallization. As the crystallization increases, the absorbing defects at grain boundaries and the substrate interface also increase, thereby increasing the absorption [81].

 figure: Fig. 6.

Fig. 6. Dispersion of HfO2 thin films (1482 cycles) deposited at different deposition temperatures determined from a) ellipsometry and b) transmission-reflection measurement using UV-VIS spectroscopy. The thin films were deposited using an ALD sequence of 3.12s/5s/3s/5s. Details of the analysis are provided in Supplement 1.

Download Full Size | PDF

The LIDT measured for HfO2 deposited at 100°C is not shown because the film was entirely removed at minimal fluence values. For HfO2 films deposited at 150°C and 200°C with a post process annealing for 4 h at 400–450°C, an LIDT value of 8 J/cm2 was measured. Jena et al. showed that the LIDT at 532 nm and 1064 nm laser wavelength of annealed HfO2 thin films increased with increasing density [81]. It was shown that increased density leads to low thermal barriers and superior thermal conductivity and, thereby, improvement of LIDT. However, LIDT decreases due to an increase in absorbing defect concentration with increase in deposition temperature. In this study, the LIDT decreases to 2 J/cm2 for HfO2 deposited at 250°C and annealed for 12 h at 400°C. This may be due to an increase in crystallinity at elevated temperatures and will be investigated in detail in the future. Wei et al. have shown that the reflection and refraction of the incident laser can occur at the grain boundaries where a higher concentration of defects is present, which increases the absorption due to the increased optical path [119].

Few articles report on the LIDT values of HfO2 at 355 nm wavelength. For HfO2 films deposited using e-beam evaporation, Zhang et al. have shown that nanometer-scale absorber defects are a crucial factor for the LIDT at 355 nm [83]. Also oxygen deficiency creates sub bandgap states which increases coupling of laser energy and thereby reduces LIDT [118,120]. Gallais et al. have shown a comparison of the LIDT at 355 nm for HfO2 films deposited using different deposition techniques. The reported LIDT values at 355 nm were 2.1 J/cm2 (electron beam deposition using HfO2 source - EBD-HfO2), 2.8 J/cm2 (electron beam deposition using Hf source - EBD-Hf), 2.3 J/cm2 (reactive low voltage ion plating - RLVIP), and 0.22 J/cm2 (dual ion beam sputtering - DIBS) [82]. The low LIDT for DIBS films was attributed to a high extinction coefficient and damage morphology due to the high intrinsic absorption level as a result of high oxygen deficiency. This is in agreement with the low LIDT and higher absorption in HfO2 deposited at 100°C in this study. Additionally, the absorption of HfO2 deposited at 100°C using a higher plasma pulse duration of 7 s shows a significant decrease compared to HfO2 deposited at the same temperature using 3 s plasma pulse (Table 2), and consequently an improved LIDT.

However, these LIDT values cannot be properly compared with our PEALD films because of different test regimes (1-on-1 vs. R-on-1) and different laser parameters (beam diameter, pulse length and repetition rate). Therefore, Layertec GmbH has produced single layer HfO2 layers using magnetron sputtering (MS) and ion assisted deposition (IAD). These single layers have LIDT values of 20 J/cm2 and 11 J/cm2, respectively, as measured with the same system as the ALD coatings. The corresponding reference value for the uncoated substrate is 85 J/cm2. The LIDT values of the PEALD films depend on the deposition conditions, with the highest value of 16 J/cm2, when a longer plasma pulse was applied. Thus, the effect of plasma parameters was investigated and is discussed in detail in the next section.

3.3 Influence of the plasma parameters

In the influence of temperature section 3.2, the duration of the TDMAH pulse was 3.12 s, the precursor and plasma purge were 5 s, the carrier (Ar) gas flow rate was 160 sccm, the ICP power was 100 W, and the oxygen gas flow was 200 sccm. The effect of plasma parameters (ICP power, oxygen flow rate and plasma pulse) on HfO2 properties were investigated on films deposited at 100°C using 1482 ALD cycles, which leads to films with a thickness above 250 nm. The variation of parameters is marked in Table 3. The thickness non-uniformity (NU) over 200 mm diameter area surface was 1.3% except the HfO2 films deposited using 20 sccm oxygen flow. The NU of HfO2 films deposited using 20 sccm oxygen flow was 2.4%, probably due to insufficient oxidation of surface functional groups. Various plasma parameters were varied here, but the study is punctual in order to limit the number of samples and experiments. Tables 3 and 4 summarize the film properties discussed in this section as function of these parameters.

Tables Icon

Table 3. Thin film properties of the HfO2 deposited at different plasma conditions.a

Tables Icon

Table 4. The effect of plasma parameters on refractive index and absorption in HfO2 thin films (1482 ALD cycles).

The plasma pulse duration was varied from 3 s to 7 s. While the duration of the precursor and the purge pulses as discussed in the section 3.1 had a significant influence on the NU, no considerable effects of the variation of plasma pulse durations were observed. The NU of HfO2 remained under 1% on varying plasma pulse duration from 3 s to 7 s due to the uniform O2 gas distribution in the reactor. However, the thin film refractive index increased from 1.90 to 1.96 (Fig. 7) with the increase in plasma pulse duration from 3 s to 7 s. The prolonged exposure of plasma enables the energetic reaction of plasma species with intermediate surface functional groups leading to a densification of the thin films. The increase in density and the refractive index can be correlated to a decrease in GPC from 1.90 to 1.80 Å/cycle with increasing plasma pulse duration. A similar effect of decreasing GPC with increasing density was observed with increasing deposition temperature, as mentioned in the section 3.2. In contrast, an increase in GPC with an increase in plasma exposure time was reported earlier [28,64]. The low GPC at low plasma pulse durations in those studies was associated with unsaturated growth. Hence, the effects of variation of plasma pulse duration, O2 gas flow, and ICP power on thin film growth, mechanical properties, optical properties, and chemical composition of HfO2 films were investigated thoroughly in this study.

 figure: Fig. 7.

Fig. 7. Variation of the refractive index at 632.8 nm and growth per cycle (GPC) with plasma pulse time. The ALD cycles of 200 were used in these depositions.

Download Full Size | PDF

At the initial plasma conditions (3 s plasma pulse), HfO2 films show a smooth surface (See Supplement 1 Fig. S4a) with a roughness of ≈2 nm estimated from EMA using spectroscopic ellipsometry fits and an RMS roughness of 0.51 nm obtained from AFM measurements (See Supplement 1 Fig. S5a). The RMS roughness increases slightly from 0.51 to 0.68 nm on increasing plasma pulse duration from 3 s to 7 s (Table 3), which is visible as the slightly rougher texture of the film deposited at 7 s plasma (See Supplement 1 Fig. S4). A significant decrease in the GPC from 1.80 to 1.70 Å/cycle is observed. This may be due to the densification of the film, as confirmed by XRR measurements (Table 3). The density of the HfO2 films increases from 7.7 to 8.2 g/cm3 on increasing plasma pulse duration. The higher plasma pulse duration provides sufficient time for the oxygen to react with intermediate functional groups, thereby removing impurities associated with carbon and nitrogen. The effect of increasing density is reflected in an increase in the refractive index. For thicker HfO2 films (> 260 nm), the refractive index at 632.8 nm increases from 1.92 to 1.96 on increasing plasma pulse duration from 3 s to 7 s. The longer plasma pulse duration of 7 s (Table 4) leads to higher quality films considering the film composition and the optical losses. The tensile stress increases from 588 MPa to 653 MPa. This can be explained by the onset of crystallization (See Supplement 1 Fig. S6), which should be carefully monitored in a process development.

The impurities in the form of N decrease from 3.6% to 2.0% and C from 12.1% to 8.4% on increasing plasma pulse duration to 7 s (Figs. 8(a), 8(d)). Additionally, the O:Hf ratio remains relatively constant at around 2.1. The high impurity content in the HfO2 films deposited at 3 s compared to 7 s plasma exposure correlates very well with the higher absorption losses at 355 nm wavelength, as shown in Table 4. Moreover, the optical losses in the UV range decrease (Fig. 9) due to a reduction of the impurity concentration. The FTIR measurement confirms the decrease in residual OH groups with increasing plasma exposure to 7 s (See Supplement 1 Fig. S7). As explained in the previous section, an increase in crystallinity results in a decrease in grain boundaries. Hence, the tensile stress increases with an increase in crystallization (See Supplement 1 for Figs. S4 and S6).

 figure: Fig. 8.

Fig. 8. AES depth profiles of HfO2 thin films (1482 ALD cycles) deposited a) using an initial condition with 3 s plasma pulse, 200 sccm O2 gas flow, 100 W ICP power b) 3 s plasma pulse, 20 sccm O2 gas flow, 100 W ICP power c) 3 s plasma pulse, 200 sccm O2 gas flow, 500 W ICP power d) 7 s plasma pulse, 200 sccm O2 gas flow, 100 W ICP power.

Download Full Size | PDF

 figure: Fig. 9.

Fig. 9. Comparison of optical losses (OL = 100% - R – T) of hafnium oxide (thickness 260 - 275 nm/1482 ALD cycles) thin films deposited at various plasma conditions. The varied parameters are marked in bold.

Download Full Size | PDF

We also investigated the effect of the oxygen flow rate on HfO2 properties. The HfO2 films deposited using 20 sccm of oxygen flow show smoother surface with RMS roughness of 0.38 nm compared to 0.51 nm (See Supplement 1 for Fig. S5a,b) for the films deposited using higher oxygen flow (200 sccm). For HfO2 deposited using 3 s plasma pulse, the GPC decreases from 1.80 to 1.72 Å/cycle on decreasing O2 gas flow from 200 sccm to 20 sccm. The lower GPC can be attributed to the low oxidation of the surface ligands. Hence, less dense films with a low refractive index contain a higher amount of carbon and nitrogen impurities, as summarized in Table 3. As a result of the high impurity level, the absorption at 355 nm wavelength increases drastically from 463 to 3920 ppm (Table 4). Additionally, under stoichiometric films may also be responsible for the higher absorption. Correspondingly, high optical losses in the UV range are observed in Fig. 9. The mechanical stress in the film is lower compared to HfO2 deposited using 200 sccm O2 gas flow due to the low density and a porous microstructure. The incomplete oxidation of surface ligands due to insufficient oxygen flow (20 sccm) leads to higher residual OH groups (See Supplement 1 Fig. S7). Hence, a higher O2 flow (200 sccm) is necessary for the plasma species to sufficiently react with the surface ligands, thereby achieving lower impurity content and absorption losses.

The impact of the ICP power on HfO2 deposited at 100°C using a 3 s plasma pulse and 200 sccm O2 gas flow were also investigated in this study. On increasing ICP power from 100 W to 500 W, the GPC increases slightly from 1.80 to 1.84 Å/cycle, and the density increases from 7.7 g/cm3 to 8.9 g/cm3. The increase in density correlates with an increase in the refractive index of HfO2 deposited at 500 W ICP power. For example, the refractive index at 632.8 nm increases from 1.92 to 2.02 on increasing ICP power. Also, the surface roughness estimated from the EMA layer increases from 2 to 14 nm. Similarly, the RMS surface roughness measured using AFM also shows an increase from 0.51 to 9.55 nm (See Supplement 1 for Figs. S5a,d). These effects of increasing roughness, density, and refractive index may be attributed to an increase in crystallization, as evident in the surface morphology (See Supplement 1 for Figs. S4a,c) and XRD pattern (See Supplement 1 Fig. S6). The increase in crystallization can be due to higher ad-atom mobility facilitated by increased ion energy with increasing ICP power. The increased surface energy through highly energetic oxygen ions reacting at the surface assists the removal of C% and N% impurities, as shown in Table 3. The HfO2 deposited using 500 W ICP power also showed lowest incorporation of residual -OH groups (See Supplement 1 Fig. S7). The lower amount of OH groups may be attributed to higher density of the HfO2 films. The absorption at 355 nm increases significantly from 463 ppm to 1570 ppm and at 532 nm from 12 to 110 ppm with increasing ICP power from 100 W to 500 W (Table 4). The impurities in the thin films, as well as defects at grain boundaries, can be the source of absorption [121,122]. Sub-bandgap absorption was observed in polycrystalline films [122]. These defects can exist in the form of oxygen vacancies or interstitials. However, AES measurements show a nearly constant O:Hf ratio of 2.2 for HfO2 deposited at 500 W and 100 W ICP power. Theoretical calculation by Lyons et al. has shown that oxygen interstitials are more stable in oxygen-rich conditions [123]. The tensile stress in the film deposited at 500 W ICP power is low (460 MPa) compared to the film deposited at 100 W power (588 MPa), possibly due to stress-release because of crack formation.

It is important to note that the standard deviations of the N and C impurities in the HfO2 films deposited by 500 W plasma power (3 s plasma pulse, 200 sccm O2) and 7 s plasma pulse (100W ICP power, 200 sccm O2) are lower compared to the other two plasma conditions (Table 3). In-depth insight into this phenomenon is given by the AES depth profiles shown in Fig. 8. At the initial conditions of 3 s plasma pulse duration, a significant increase of the carbon content and a slight increase in the nitrogen content towards the substrate were observed (Fig. 8(a)). This effect is reduced when the plasma pulse duration was increased to 7 s. However, a minor increase in the carbon content towards the substrate is seen in Fig. 8(d). In contrast, the atomic impurity content in the depth profile of the HfO2 deposited with 20 sccm oxygen flow was relatively uneven due to the poor gas flow while showing the same increasing tendency of the C and N content towards the substrate. As a result, the standard deviation is high. This observation shows the influence of the substrate and the available functional surface groups, as explained below.

For the HfO2 deposited at a low oxygen flow rate (20 sccm) or short plasma pulse duration (3 s), more TDMAH ligands, i.e., more nitrogen and carbon, remain on the surface during the plasma step and do not participate in chemical reactions during the first cycles until the film reaches a critical thickness. At this point, the influence of the substrate is no longer prominent, and the film solely grows from the new hydroxyl groups formed on the growing film. The chemical environment and the availability of functional groups are different in the vicinity of the substrate, and the self-limiting reaction regime during the plasma step changes with thickness during the thin film growth. Consequently, an increase of the plasma power (to 500 W) or the plasma pulse duration (7 s) leads to more ligand oxidation during the first few hundreds of PEALD cycles and a more uniform atomic nitrogen and carbon content throughout the film. However, despite the broad variation of the plasma parameters (20 sccm O2 flow vs. 200 sccm, 100 W vs. 500 W), the Hf:O ratio remained relatively constant.

The HfO2 film using 7 s plasma pulse duration, 100 W ICP power, and 200 sccm O2 gas flow showed low impurity and low absorption losses. Hence, these conditions were considered further for antireflection coatings, as discussed in the following section.

3.4 Antireflection coatings

Antireflection (AR) coatings were realized for the wavelengths of a Nd:YAG laser and its higher harmonics (up to the 4th order) consisting of hafnium oxide in combination with silicon oxide and aluminum oxide. The individual layers were deposited using plasma enhanced ALD (PEALD) at 100°C temperature. The Al2O3 thin films were deposited using trimethyl aluminum (TMA) precursor and oxygen plasma with an ALD sequence of 0.08s/2s/3s/2s, 80 sccm carrier gas (Ar) flow, 100 sccm oxygen flow and 100 W ICP power. The SiO2 thin films were deposited using bis(diethylamino)silane (BDEAS) precursor and oxygen plasma with an ALD sequence of 0.32s/5s/3s/2s, 30 sccm carrier gas (Ar) flow, 200 sccm oxygen flow and 100 W ICP power. The ALD cycle for the HfO2 thin films consists of 3.12 s of precursor pulse, 5 s of precursor purge, 7 s of plasma pulse, and 5 s of plasma purge. The O2 and Ar gas flow rates of 200 sccm and 160 sccm, respectively, were maintained for the entire HfO2 deposition. The depositions were carried out on fused silica (FS) substrates placed in different orientations. Most FS substrates were clipped in a vertical orientation on the substrate holder for a double-side coating, while one FS substrate was placed directly on a 200 mm Si wafer (to prevent backside deposition) for single-side deposition. The AR designs and the total physical thicknesses are summarized in Table 5. Figure 10(a) shows the reflection spectra of the AR1 sample with an antireflection (AR) coating for 1ω and 2ω of the Nd:YAG laser. The measured reflectance spectrum matches very well with the design (data shown for single side coating, without backside reflection). The thickness gradient of the vertically aligned sample is small and the AR is successfully achieved for single and double-side deposition. All designs have been well achieved, and the total reflectance of the double side coatings are below 0.7% at all target wavelengths; however, the absorbance at 266 nm is larger than at 355 nm wavelength. The minima of the AR2 and AR3 samples is slightly shifted towards lower wavelengths than the target spectral positions. The ARs have been not further optimized. The mechanical stress of AR1 was 282 MPa, which was lower than 326 MPa and 380 MPa for AR2 and AR3 because of fewer HfO2 layers in the AR stack.

 figure: Fig. 10.

Fig. 10. Reflectance spectra of an antireflection coating consisting of HfO2, SiO2, and Al2O3 on single and double-side coated fused silica substrates at a) 1064 nm and 532 nm b) 1064 nm, 532 nm, and 355 nm c) 1064 nm, 532 nm, 355 nm, and 266 nm. A comparison of the designed and deposited thin-film system is shown. The backside reflection of the single-side coated sample is subtracted.

Download Full Size | PDF

Tables Icon

Table 5. The composition of Al2O3[HfO2/Al2O3]xSiO2 thin films and design used in antireflection coatings.a

The absorption in all AR coatings (single and double sides) at 1ω is 5 ± 2 ppm and at 2ω is 10 ± 4 ppm. Similar absorptions of 1 ppm (1ω) and 9.3 ppm (2ω) in AR coatings deposited on the single-side fused silica was reported by Liu et al. [95]. The authors also reported an absorption of 597.1 ppm at 3ω [95], which is significantly higher than the values obtained here 188 ± 14 ppm and 170 ± 14 ppm for single-side AR2 and AR3. Hence, a PEALD process might be more appropriate in terms of absorption losses than a thermal process.

The LIDT at 355 nm wavelength for all the three AR coatings is 27–28 J/cm2 (see Table 5). Liu et al. reported LIDT of 4.9 J/cm2 at 3ω wavelength (10 pulses, S-on-1, 10 ns, 100 Hz, beam focus Ø 300 µm) [95]. In a similar study, Yin et al. have shown that LIDT (1-on-1, 7.8 ns, beam focus 0.28 mm2) of 24.4 J/cm2 at 3ω for AR coatings deposited by PEALD HfO2/SiO2 at 150°C was higher than AR coatings (20.6 J/cm2) deposited by electron beam deposition [94], which was attributed to the lower absorption of PEALD AR coating. Comparable AR systems consisting of HfO2/SiO2 multilayers have been deposited at Layertec GmbH by magnetron sputtering and ion assisted deposition on similar substrates. The LIDT values of these coatings reached values of 31 J/cm2 and 28 J/cm2, respectively. This indicates that the main limiting factor on the LIDT performance are the intrinsic properties of the oxides, specifically of the high refractive index component. Additionally, macroscopic defects such as nodules or particles are also low within the PEALD processes. The influence of different materials and designs for AR coatings at 355 nm made by PEALD is currently under investigation and will be presented in a sub-sequent article.

In this study, high-quality AR coatings with low absorption and high LIDT deposited using an optimized PEALD process at 100°C demonstrate a promising application in high power laser systems. The deposition processes must be carefully optimized to achieve the desired film properties to meet the performance of AR coatings.

4. Conclusion

The requirements of thin film coatings are a high density with pinhole-free morphology, amorphous materials with low defect density, low impurity and OH incorporation and a very high uniformity. This study provides a thorough approach towards the process development of HfO2 coatings by plasma enhanced atomic layer deposition (PEALD). The process parameters, mainly gas flow rates, have been demonstrated to have a major influence on the non-uniformity of the coating. The deposition temperature has a high influence on the film density and refractive index, OH incorporation, and C and N impurities; the effect of crystallization was prominent by high absorption in the UV range, high surface roughness, and high tensile stress. Correspondingly, the films deposited above the crystallization threshold showed lower LIDT. The deposition process was further optimized by varying plasma conditions and keeping the deposition temperature at 100°C. The HfO2 films deposited using 7 s of plasma exposure time, 100 W ICP power, and 200 sccm oxygen flow rate showed a high refractive index in the UV range, high density, low C and N impurities, and low absorption. The effects of crystallization on deteriorating thin film properties were also visible by depositing HfO2 films at 500 W ICP power. Eventually, the shortcomings of HfO2 films deposited using 3 s plasma exposure time were rectified using 7 s of plasma exposure. This also improved the LIDT value. Thus, oxygen plasma species with sufficient energy were successful in improving material properties. Subsequently, HfO2 thin films deposited under optimized conditions were incorporated into antireflection coatings designed for 1ω, 2ω, 3ω, and 4ω wavelength of the Nd:YAG laser. The deposited films were in excellent agreement with the design, and a reflectance below 0.4% was observed for all the coatings. Besides, very good LIDT values were realized for these coatings. Consequently, it was shown that high-quality films at a deposition temperature of 100°C are achievable by using PEALD.

Funding

Bundesministerium für Wirtschaft und Energie (ZF4309604SY8, ZF4596101SY8); Deutsche Forschungsgemeinschaft (Priority Programme Fields Matter PP1959 SZ253/2-1, and SFB NOA 1375 Project-ID 398816777); Fraunhofer-Gesellschaft (Attract 066-601020) and by the German Research Foundation and the Open Access Publication Fund of the Thueringer Universitaetsund Landesbibliothek Jena Projekt (Nr. 433052568).

Disclosures

The authors declare no conflicts of interest.

Supplemental document

See Supplement 1 for supporting content.

References

1. G. D. Wilk, R. M. Wallace, and J. M. Anthony, “High-κ gate dielectrics: current status and materials properties considerations,” J. Appl. Phys. 89(10), 5243–5275 (2001). [CrossRef]  

2. S. Siddiqui, T. Ando, R. K. Pandey, and D. Schepis, “Limits of gate dielectrics scaling,” in Handbook of thin film deposition, K. Seshan and D. Schepis, eds. (Elsevier, William Andrew Applied Science Publishers, 2018), pp. 107–145.

3. A. I. Kingon, J.-P. Maria, and S. K. Streiffer, “Alternative dielectrics to silicon dioxide for memory and logic devices,” Nature 406(6799), 1032–1038 (2000). [CrossRef]  

4. J. Sheng, J.-H. Lee, W.-H. Choi, T. Hong, M. Kim, and J.-S. Park, “Review Article: Atomic layer deposition for oxide semiconductor thin film transistors: Advances in research and development,” J. Vac. Sci. Technol. A 36(4), 060801 (2018). [CrossRef]  

5. J. Robertson and R. M. Wallace, “High-K materials and metal gates for CMOS applications,” Mater. Sci. Eng., R 88, 1–41 (2015). [CrossRef]  

6. A. Nishiyama, “Hafnium-based gate dielectric materials,” in High Permittivity Gate Dielectric Materials, Vol. 43, S. Kar, ed. (Springer, 2013), pp. 153–181.

7. L.-G. Wang, X. Qian, Y.-Q. Cao, Z.-Y. Cao, G.-Y. Fang, A.-D. Li, and Di Wu, “Excellent resistive switching properties of atomic layer-deposited Al2O3/HfO2/Al2O3 trilayer structures for non-volatile memory applications,” Nanoscale Res. Lett. 10(1), 135 (2015). [CrossRef]  

8. P. Torchio, A. Gatto, M. Alvisi, G. Albrand, N. Kaiser, and C. Amra, “High-reflectivity HfO2/SiO2 ultraviolet mirrors,” Appl. Opt. 41(16), 3256–3261 (2002). [CrossRef]  

9. S. Shestaeva, A. Bingel, P. Munzert, L. Ghazaryan, C. Patzig, A. Tünnermann, and A. Szeghalmi, “Mechanical, structural, and optical properties of PEALD metallic oxides for optical applications,” Appl. Opt. 56(4), C47–C59 (2017). [CrossRef]  

10. T. Begou, F. Lemarchand, F. Lemarquis, A. Moreau, and J. Lumeau, “High-performance thin-film optical filters with stress compensation,” J. Opt. Soc. Am. A 36(11), C113–C121 (2019). [CrossRef]  

11. K. Pfeiffer, S. Shestaeva, A. Bingel, P. Munzert, L. Ghazaryan, C. van Helvoirt, W. M. M. Kessels, U. T. Sanli, C. Grévent, G. Schütz, M. Putkonen, I. Buchanan, L. Jensen, D. Ristau, A. Tünnermann, and A. Szeghalmi, “Comparative study of ALD SiO2 thin films for optical applications,” Opt. Mater. Express 6(2), 660 (2016). [CrossRef]  

12. J. T. Gaskins, P. E. Hopkins, D. R. Merrill, S. R. Bauers, E. Hadland, D. C. Johnson, D. Koh, J. H. Yum, S. Banerjee, B. J. Nordell, M. M. Paquette, A. N. Caruso, W. A. Lanford, P. Henry, L. Ross, H. Li, L. Li, M. French, A. B. Mei, and S. W. King, “Review—Investigation and review of the thermal, mechanical, electrical, optical, and structural properties of atomic layer deposited high- k dielectrics: Beryllium oxide, aluminum oxide, hafnium oxide, and aluminum nitride,” ECS J. Solid State Sci. Technol. 7(5), X3 (2018). [CrossRef]  

13. H. Jeon and Y. Won, “The reaction pathways of the oxygen plasma pulse in the hafnium oxide atomic layer deposition process,” Appl. Phys. Lett. 93(12), 124104 (2008). [CrossRef]  

14. M. Cortez-Valadez, C. Fierro, J. R. Farias-Mancilla, A. Vargas-Ortiz, M. Flores-Acosta, R. Ramírez-Bon, J. L. Enriquez-Carrejo, C. Soubervielle-Montalvo, and P. G. Mani-Gonzalez, “Comparison of HfCl4, HfI4, TEMA-Hf, and TDMA-Hf as precursors in early growing stages of HfO2 films deposited by ALD: A DFT study,” Chem. Phys. 472, 81–88 (2016). [CrossRef]  

15. Vitaly Gruzdev, “Self-focusing and nonlinear effects,” in Laser-Induced Damage in Optical Materials, D. Ristau, ed. (CRC Press, 2014), pp. 90–141.

16. M. F. Al-Kuhaili, S. M. A. Durrani, and E. E. Khawaja, “Characterization of hafnium oxide thin films prepared by electron beam evaporation,” J. Phys. D: Appl. Phys. 37(8), 1254–1261 (2004). [CrossRef]  

17. M. Niwa, “Hf-based high-k gate dielectric processing,” in High Permittivity Gate Dielectric Materials, S. Kar, ed. (Springer, 2013), pp. 183–234.

18. R. Thielsch, “Optical coatings for the DUV/VUV,” in Optical Interference Coatings, N. Kaiser and H. K. Pulker, eds. (Springer International Publishing, 20), pp. 257–279.

19. R. Thielsch, A. Gatto, J. Heber, and N. Kaiser, “A comparative study of the UV optical and structural properties of SiO2, Al2O3, and HfO2 single layers deposited by reactive evaporation, ion-assisted deposition and plasma ion-assisted deposition,” Thin Solid Films 410(1-2), 86–93 (2002). [CrossRef]  

20. O. Stenzel, M. Schürmann, S. Wilbrandt, N. Kaiser, A. Tünnermann, M. Mende, H. Ehlers, D. Ristau, S. Bruns, M. Vergöhl, W. Riggers, M. Bischoff, and M. Held, “Optical and mechanical properties of oxide UV coatings, prepared by PVD techniques,” Proc. SPIE8168 (2011).

21. W. He, “ALD: Atomic layer deposition – Precise and conformal coating for better performance,” in Handbook of Manufacturing Engineering and Technology, A. Y. C. Nee, ed. (Springer Reference, 2015), pp. 2959–2996.

22. V. Cremers, R. L. Puurunen, and J. Dendooven, “Conformality in atomic layer deposition: Current status overview of analysis and modelling,” Appl. Phys. Rev. 6(2), 021302 (2019). [CrossRef]  

23. K.-I. Liu, C.-C. Kei, M. Mishra, P.-H. Chen, W.-S. Liu, and T.-P. Perng, “Uniform coating of TiO2 on high aspect ratio substrates with complex morphology by vertical forced-flow atomic layer deposition,” RSC Adv. 7(55), 34730–34735 (2017). [CrossRef]  

24. P. Schindler, M. Logar, J. Provine, and F. B. Prinz, “Enhanced step coverage of TiO2 deposited on high aspect ratio surfaces by plasma-enhanced atomic layer deposition,” Langmuir 31(18), 5057–5062 (2015). [CrossRef]  

25. P. O. Oviroh, R. Akbarzadeh, D. Pan, R. A. M. Coetzee, and T.-C. Jen, “New development of atomic layer deposition: processes, methods and applications,” Sci. Technol. Adv. Mater. 20(1), 465–496 (2019). [CrossRef]  

26. K. Arts, M. Utriainen, R. L. Puurunen, W. M. M. Kessels, and H. C. M. Knoops, “Film conformality and extracted recombination probabilities of O atoms during plasma-assisted atomic layer deposition of SiO2, TiO2, Al2O3, and HfO2,” J. Phys. Chem. C 123(44), 27030–27035 (2019). [CrossRef]  

27. T. Faraz, H. C. M. Knoops, M. A. Verheijen, Cristian A. A. van Helvoirt, S. Karwal, A. Sharma, V. Beladiya, A. Szeghalmi, D. M. Hausmann, J. Henri, M. Creatore, and W. M. M. Kessels, “Tuning material properties of oxides and nitrides by substrate biasing during plasma-enhanced atomic layer deposition on planar and 3D substrate topographies,” ACS Appl. Mater. Interfaces 10(15), 13158–13180 (2018). [CrossRef]  

28. A. Sharma, V. Longo, M. A. Verheijen, A. A. Bol, and W. M. M. Kessels, “Atomic layer deposition of HfO2 using HfCp(NMe2)3 and O2 plasma,” J. Vac. Sci. Technol., A 35(1), 01B130 (2017). [CrossRef]  

29. K. Pfeiffer, U. Schulz, A. Tünnermann, and A. Szeghalmi, “Antireflection coatings for strongly curved glass lenses by atomic layer deposition,” Coatings 7(8), 118 (2017). [CrossRef]  

30. B. C. Mallick, C.-T. Hsieh, K.-M. Yin, Y. A. Gandomi, and K.-T. Huang, “Review - On atomic layer deposition: Current progress and future challenges,” ECS J. Solid State Sci. Technol. 8(4), N55–N78 (2019). [CrossRef]  

31. R. L. Puurunen, “Surface chemistry of atomic layer deposition: A case study for the trimethylaluminum/water process,” J. Appl. Phys. 97(12), 121301 (2005). [CrossRef]  

32. R. L. Puurunen, “Analysis of hydroxyl group controlled atomic layer deposition of hafnium dioxide from hafnium tetrachloride and water,” J. Appl. Phys. 95(9), 4777–4786 (2004). [CrossRef]  

33. R. Rammula, J. Aarik, H. Mändar, P. Ritslaid, and V. Sammelselg, “Atomic layer deposition of HfO2: Effect of structure development on growth rate, morphology and optical properties of thin films,” Appl. Surf. Sci. 257(3), 1043–1052 (2010). [CrossRef]  

34. I.-K. Oh, B.-E. Park, S. Seo, B. C. Yeo, J. Tanskanen, H.-B.-R. Lee, W.-H. Kim, and H. Kim, “Comparative study of the growth characteristics and electrical properties of atomic-layer-deposited HfO2 films obtained from metal halide and amide precursors,” J. Mater. Chem. C 6(27), 7367–7376 (2018). [CrossRef]  

35. J. Aarik, A. Aidla, A.-A. Kiisler, T. Uustare, and V. Sammelselg, “Influence of substrate temperature on atomic layer growth and properties of HfO2 thin films,” Thin Solid Films 340(1-2), 110–116 (1999). [CrossRef]  

36. J. S. Ponraj, G. Attolini, and M. Bosi, “Review on atomic layer deposition and applications of oxide thin films,” Crit. Rev. Solid State Mater. Sci. 38(3), 203–233 (2013). [CrossRef]  

37. R. G. Gordon, “ALD precursors and reaction mechanisms,” in Atomic Layer Deposition for emiconductors, C. S. Hwang, ed. (Springer, 2014), pp. 15–46.

38. C. B. Musgrave and R. G. Gordon, “Precursors for atomic layer deposition of high-k dielectrics,” Future Fab. Int. 18, 126 (2005).

39. W. Cho, K.-S. An, T.-M. Chung, C. G. Kim, B.-S. So, Y.-H. You, J.-H. Hwang, D. Jung, and Y. Kim, “ALD of hafnium dioxide thin films using the new alkoxide precursor hafnium 3-methyl-3-pentoxide, Hf(mp)4,” Chem. Vap. Dep. 12(11), 665–669 (2006). [CrossRef]  

40. K. Kukli, M. Ritala, M. Leskelä, T. Sajavaara, J. Keinonen, A. C. Jones, and J. L. Roberts, “Atomic layer deposition of hafnium dioxide films from 1-methoxy-2-methyl-2-propanolate complex of hafnium,” Chem. Mater. 15(8), 1722–1727 (2003). [CrossRef]  

41. R. J. Potter, P. R. Chalker, T. D. Manning, H. C. Aspinall, Y. F. Loo, A. C. Jones, L. M. Smith, G. W. Critchlow, and M. Schumacher, “Deposition of HfO2, Gd2O3 and PrOx by liquid injection ALD techniques,” Chem. Vap. Dep. 11(3), 159–169 (2005). [CrossRef]  

42. K. Kukli, M. Ritala, J. Lu, A. Hårsta, and M. Leskelä, “Properties of HfO2 thin films grown by ALD from Hafnium tetrakis(ethylmethylamide) and Water,” J. Electrochem. Soc. 151(8), F189 (2004). [CrossRef]  

43. D. M. Hausmann, E. Kim, J. Becker, and R. G. Gordon, “Atomic layer deposition of hafnium and zirconium oxides using metal amide precursors,” Chem. Mater. 14(10), 4350–4358 (2002). [CrossRef]  

44. A. Deshpande, R. Inman, G. Jursich, and C. Takoudis, “Atomic layer deposition and characterization of hafnium oxide grown on silicon from tetrakis(diethylamino)hafnium and water vapor,” J. Vac. Sci. Technol. A 22(5), 2035–2040 (2004). [CrossRef]  

45. D. Blaschke, F. Munnik, J. Grenzer, L. Rebohle, H. Schmidt, P. Zahn, and S. Gemming, “A correlation study of layer growth rate, thickness uniformity, stoichiometry, and hydrogen impurity level in HfO2 thin films grown by ALD between 100°C and 350°C,” Appl. Surf. Sci. 506, 144188 (2020). [CrossRef]  

46. S. Consiglio, R. D. Clark, G. Nakamura, C. S. Wajda, and G. J. Leusink, “Evaluation of high thermal stability cyclopentadienyl Hf precursors with H2O as a co-reactant for advanced gate logic applications,” J. Vac. Sci. Technol. A 30(1), 01A119 (2012). [CrossRef]  

47. C. L. Dezelah, J. Niinistö, K. Kukli, F. Munnik, J. Lu, M. Ritala, M. Leskelä, and L. Niinistö, “The atomic layer deposition of HfO2 and ZrO2 using Advanced Metallocene Precursors and H2O as the oxygen source,” Chem. Vap. Dep. 14(11-12), 358–365 (2008). [CrossRef]  

48. R. O’Kane, J. Gaskell, A. C. Jones, P. R. Chalker, K. Black, M. Werner, P. Taechakumput, S. Taylor, P. N. Heys, and R. Odedra, “Growth of HfO2 by liquid injection MOCVD and ALD using new hafnium-cyclopentadienyl precursors,” Chem. Vap. Dep. 13(11), 609–617 (2007). [CrossRef]  

49. J. C. Hackley and T. Gougousi, “Properties of atomic layer deposited HfO2 thin films,” Thin Solid Films 517(24), 6576–6583 (2009). [CrossRef]  

50. R. Lo Nigro, E. Schilirò, C. Tudisco, G. G. Condorelli, P. Fiorenza, H. Gargouri, and F. Roccaforte, “Thermal and plasma-enhanced atomic layer deposition of hafnium oxide on semiconductor substrates,” IEEE 9th Nanotechnology Materials and Devices Conference (NMDC), 112–115 (2014).

51. W. Chen, Q.-Q. Sun, M. Xu, S.-J. Ding, D. W. Zhang, and L.-K. Wang, “Atomic layer deposition of hafnium oxide from tetrakis(ethylmethylamino)hafnium and water precursors,” J. Phys. Chem. C 111(17), 6495–6499 (2007). [CrossRef]  

52. F. Hirose, Y. Kinoshita, K. Kanomata, K. Momiyama, S. Kubota, K. Hirahara, Y. Kimura, and M. Niwano, “IR study of fundamental chemical reactions in atomic layer deposition of HfO2 with tetrakis(ethylmethylamino)hafnium (TEMAH), ozone, and water vapor,” Appl. Surf. Sci. 258(19), 7726–7731 (2012). [CrossRef]  

53. X. Nie, F. Ma, D. Ma, and K. Xu, “Growth mode evolution of hafnium oxide by atomic layer deposition,” J. Vac. Sci. Technol. A 32(1), 01A109 (2014). [CrossRef]  

54. B. A. Sperling, W. A. Kimes, and J. E. Maslar, “Reflection absorption infrared spectroscopy during atomic layer deposition of HfO2 films from tetrakis(ethylmethylamido)hafnium and water,” Appl. Surf. Sci. 256(16), 5035–5041 (2010). [CrossRef]  

55. M.-T. Ho, Y. Wang, R. T. Brewer, L. S. Wielunski, Y. J. Chabal, N. Moumen, and M. Boleslawski, “In situ infrared spectroscopy of hafnium oxide growth on hydrogen-terminated silicon surfaces by atomic layer deposition,” Appl. Phys. Lett. 87(13), 133103 (2005). [CrossRef]  

56. T. Lee, H.-K. Ko, J. Ahn, I.-S. Park, H. Sim, H. Park, and H. Hwang, “Electrical properties of atomic layer deposited HfO2 gate dielectric film using D2O as oxidant for improved reliability,” Jpn. J. Appl. Phys. 45(9A), 6993–6995 (2006). [CrossRef]  

57. M.-H. Tseng, H.-H. Yu, K.-Y. Chou, J.-H. Jou, K.-L. Lin, C.-C. Wang, and F.-Y. Tsai, “Low-temperature gas-barrier films by atomic layer deposition for encapsulating organic light-emitting diodes,” Nanotechnology 27(29), 295706 (2016). [CrossRef]  

58. M.-J. Choi, H.-H. Park, D. S. Jeong, J. H. Kim, J.-S. Kim, and S. K. Kim, “Atomic layer deposition of HfO2 thin films using H2O2 as oxidant,” Appl. Surf. Sci. 301, 451–455 (2014). [CrossRef]  

59. J. Provine, P. Schindler, J. Torgersen, H. J. Kim, H.-P. Karnthaler, and F. B. Prinz, “Atomic layer deposition by reaction of molecular oxygen with tetrakisdimethylamido-metal precursors,” J. Vac. Sci. Technol. A 34(1), 01A138 (2016). [CrossRef]  

60. J. C. Kim, Y. S. Cho, and S. H. Moon, “Atomic layer deposition of HfO2 onto Si using Hf(NMe2)4,” Jpn. J. Appl. Phys. 48(6), 066515 (2009). [CrossRef]  

61. M. Seo, Y.-S. Min, S. K. Kim, T. J. Park, J. H. Kim, K. D. Na, and C. S. Hwang, “Atomic layer deposition of hafnium oxide from tert-butoxytris(ethylmethylamido)hafnium and ozone: rapid growth, high density and thermal stability,” J. Mater. Chem. 18(36), 4324 (2008). [CrossRef]  

62. J. Price, P. S. Lysaght, S. C. Song, H.-J. Li, and A. C. Diebold, “Identification of sub-band-gap absorption features at the HfO2∕Si(100) interface via spectroscopic ellipsometry,” Appl. Phys. Lett. 91(6), 061925 (2007). [CrossRef]  

63. R. D. Clark, C. S. Wajda, G. J. Leusink, L. F. Edge, J. Faltermeier, P. Jamison, B. P. Linder, M. Copel, V. Narayanan, M. A. Gribelyuk, R. Loesing, and R. Murphy, “Process and electrical characteristics of MO-ALD HfO2 films for high-k gate applications grown in a production worthy 300 mm deposition system,” ECS Trans. 11(3), 55–69 (2019). [CrossRef]  

64. K.-M. Kim, J. S. Jang, S.-G. Yoon, J.-Y. Yun, and N.-K. Chung, “Structural, optical and electrical properties of HfO2 thin films deposited at low-temperature using plasma-enhanced atomic layer deposition,” Materials 13(9), 2008 (2020). [CrossRef]  

65. N. K. Oh, J.-T. Kim, J.-K. Ahn, G. Kang, S. Y. Kim, and J.-Y. Yun, “The effects of thermal decomposition of tetrakis-ethylmethylaminohafnium (TEMAHf) precursors on HfO2 film growth using atomic layer deposition,” Appl. Sci. Converg. Technol. 25(3), 56–60 (2016). [CrossRef]  

66. K. Kukli, M. Ritala, T. Sajavaara, J. Keinonen, and M. Leskelä, “Atomic layer deposition of hafnium dioxide films from hafnium tetrakis(ethylmethylamide) and Water,” Chem. Vap. Dep. 8(5), 199–204 (2002). [CrossRef]  

67. J. H. Kim, T. J. Park, S. K. Kim, D.-Y. Cho, H.-S. Jung, S. Y. Lee, and C. S. Hwang, “Chemical structures and electrical properties of atomic layer deposited HfO2 thin films grown at an extremely low temperature (≤100°C) using O3 as an oxygen source,” Appl. Surf. Sci. 292, 852–856 (2014). [CrossRef]  

68. K. Kukli, T. Pilvi, M. Ritala, T. Sajavaara, J. Lu, and M. Leskelä, “Atomic layer deposition of hafnium dioxide thin films from hafnium tetrakis(dimethylamide) and water,” Thin Solid Films 491(1-2), 328–338 (2005). [CrossRef]  

69. S. Chu, “ALD Machines,” in Atomic Layer Deposition for Semiconductors, C. S. Hwang, ed. (Springer, 2014), pp. 241–255.

70. K. Kukli, S. Dueñas, H. Castán, H. García, J. Barbolla, J. Aarik, A. Aidla, M. Ritala, and M. Leskelä, “Disordered structure and density of gap states in high-permittivity thin solid films,” in Defects in High-κ Gate Dielectric Stacks, Nano-electronic Semiconductor Devices, edited by Evgeni Gusev, vol. 220 of NATO science series, Series II, Mathematics, Physics, and Chemistry (Springer, 2006), pp. 123–134.

71. H. Kim, “Characteristics and applications of plasma enhanced-atomic layer deposition,” Thin Solid Films 519(20), 6639–6644 (2011). [CrossRef]  

72. H. C. M. Knoops, T. Faraz, K. Arts, and W. M. M. Kessels, “Status and prospects of plasma-assisted atomic layer deposition,” J. Vac. Sci. Technol. A 37(3), 030902 (2019). [CrossRef]  

73. K. M. Price, S. Najmaei, C. E. Ekuma, R. A. Burke, M. Dubey, and A. D. Franklin, “Plasma-enhanced atomic layer deposition of HfO2 on monolayer, bilayer, and trilayer MoS2 for the integration of high-κ dielectrics in two-dimensional devices,” ACS Appl. Nano Mater. 2(7), 4085–4094 (2019). [CrossRef]  

74. S. Park, B.-E. Park, H. Yoon, S. Lee, T. Nam, T. Cheon, S.-H. Kim, H. Cheon, S. Im, T. Seong, and H. Kim, “Comparative study on atomic layer deposition of HfO2 via substitution of ligand structure with cyclopentadiene,” J. Mater. Chem. C 8(4), 1344–1352 (2020). [CrossRef]  

75. I. Kim, S. Kuk, S. Kim, J. Kim, H. Jeon, M.-H. Cho, and K.-B. Chung, “Thermodynamic properties and interfacial layer characteristics of HfO2 thin films deposited by plasma-enhanced atomic layer deposition,” Appl. Phys. Lett. 90(22), 222101 (2007). [CrossRef]  

76. H. Jung, I.-K. Oh, C. M. Yoon, B.-E. Park, S. Lee, O. Kwon, W. J. Lee, S.-H. Kwon, W.-H. Kim, and H. Kim, “Effects of Ar addition to O2 plasma on plasma-enhanced atomic layer deposition of oxide thin films,” ACS Appl. Mater. Interfaces 10(46), 40286–40293 (2018). [CrossRef]  

77. J. Joo and S. M. Rossnagel, “Plasma modeling of a PEALD system for the deposition of TiO2 and HfO2,” J. Korean Phy. Soc. 54(3), 1048–1053 (2009). [CrossRef]  

78. S. B. S. Heil, J. L. van Hemmen, C. J. Hodson, N. Singh, J. H. Klootwijk, F. Roozeboom, M. C. M. van de Sanden, and W. M. M. Kessels, “Deposition of TiN and HfO2 in a commercial 200 mm remote plasma atomic layer deposition reactor,” J. Vac. Sci. Technol. A 25(5), 1357 (2007). [CrossRef]  

79. J. Kim, S. Kim, H. Kang, J. Choi, H. Jeon, M. Cho, K. Chung, S. Back, K. Yoo, and C. Bae, “Composition, structure, and electrical characteristics of HfO2 gate dielectrics grown using the remote- and direct-plasma atomic layer deposition methods,” J. Appl. Phys. 98(9), 094504 (2005). [CrossRef]  

80. K. -E. Elers, T. Blomberg, M. Peussa, B. Aitchison, S. Haukka, and S. Marcus, “Film uniformity in atomic layer deposition,” Chem. Vap. Dep. 12(1), 13–24 (2006). [CrossRef]  

81. S. Jena, R. B. Tokas, S. Tripathi, K. D. Rao, D. V. Udupa, S. Thakur, and N. K. Sahoo, “Influence of oxygen partial pressure on microstructure, optical properties, residual stress and laser induced damage threshold of amorphous HfO2 thin films,” J. Alloys Compd. 771, 373–381 (2019). [CrossRef]  

82. L. Gallais, J. Capoulade, J.-Y. Natoli, M. Commandré, M. Cathelinaud, C. Koc, and M. Lequime, “Laser damage of silica and hafnia thin films made with different deposition technologies,” Pros. SPIE6720 (2007).

83. D. Zhang, C. Wang, P. Fan, X. Cai, Z. Zheng, J. Shao, and Z. Fan, “Influence of plasma treatment on laser-induced damage characters of HfO2 thin films at 355 nm,” Opt. Express 17(10), 8246–8252 (2009). [CrossRef]  

84. C. J. Stolz, M. D. Thomas, and A. J. Griffin, “BDS thin film damage competition,” Proc. SPIE 7842 (2010).

85. Meiping Zhu, Nuo Xu, Behshad Roshanzadeh, S. T. P. Boyd, Wolfgang Rudolph, Yingjie Chai, and Jianda Shao, “Nanolaminate-based design for UV laser mirror coatings,” Light: Sci. Appl. 9(1), 1–6 (2020). [CrossRef]  

86. C. Li, Y. A. Zhao, Y. Cui, Y. Wang, X. Peng, C. Shan, M. Zhu, J. Wang, and J. Shao, “Investigation on picosecond laser-induced damage in HfO2/SiO2 high-reflective coatings,” Opt. Laser Technol. 106, 372–377 (2018). [CrossRef]  

87. A. Melninkaitis, M. Maciulevicius, T. Rakickas, D. Miksys, R. Grigonis, V. Sirutkaitis, A. Skrebutenas, R Buzelis, R. Drazdys, and G. Abromavicius, “Comparison of optical resistance of ion assisted deposition and standard electron beam deposition methods for high reflectance dielectric coatings,” Proc. SPIE5963 (2005).

88. H. Liu, L. Jensen, H. Mädebach, P. Ma, and D. Ristau, “Comparison of ALD and IBS HfO2/Al2O3 anti-reflection coatings for the harmonics of the Nd: YAG laser,” Proc. SPIE10691 (2018).

89. G. Abromavičius, S. Kičas, and R. Buzelis, “High temperature annealing effects on spectral, microstructural and laser damage resistance properties of sputtered HfO2 and HfO2-SiO2 mixture-based UV mirrors,” Opt. Mater. 95, 109245 (2019). [CrossRef]  

90. J. M. Khoshman, A. Khan, and M. E. Kordesch, “Amorphous hafnium oxide thin films for antireflection optical coatings,” Surf. Coat. Technol. 202(11), 2500–2502 (2008). [CrossRef]  

91. M. L. Grilli, F. Menchini, A. Piegari, D. Alderighi, G. Toci, and M. Vannini, “Al2O3/SiO2 and HfO2/SiO2 dichroic mirrors for UV solid-state lasers,” Thin Solid Films 517(5), 1731–1735 (2009). [CrossRef]  

92. Y. J. Guo, X. T. Zu, B. Y. Wang, X. D. Jiang, X. D. Yuan, and H. B. Lv, “Preparation of sol–gel ZrO2–SiO2 highly reflective multilayer films and laser-induced damage threshold characteristic,” Optik 122(13), 1140–1142 (2011). [CrossRef]  

93. X. Wang and J. Shen, “A review of contamination-resistant antireflective sol–gel coatings,” J. Sol-Gel Sci. Technol. 61(1), 206–212 (2012). [CrossRef]  

94. C. Yin, M. Zhu, T. Zeng, C. Song, Y. Chai, Y. Shao, R. Zhang, J. Zhao, D. Li, and J. Shao, “HfO2/SiO2 anti-reflection films for UV lasers via plasma-enhanced atomic layer deposition,” J. Alloys Compd. 859, 157875 (2021). [CrossRef]  

95. H. Liu, L. Jensen, P. Ma, and D. Ristau, “ALD anti-reflection coatings at 1ω, 2ω, 3ω, and 4ω for high-power ns-laser application,” Adv. Opt. Technol. 98, 23–32 (2018). [CrossRef]  

96. Q. Zhang, F. Pan, J. Luo, Q. Wu, Z. Wang, and Y. Wei, “Optical and laser damage properties of HfO2/Al2O3 thin films deposited by atomic layer deposition,” J. Alloys Compd. 659, 288–294 (2016). [CrossRef]  

97. V. Beladiya, M. Becker, T. Faraz, W. M. M. Kessels, P. Schenk, F. Otto, T. Fritz, M. Gruenewald, C. Helbing, K. D. Jandt, A. Tünnermann, M. Sierka, and A. Szeghalmi, “Effect of an electric field during the deposition of silicon dioxide thin films by plasma enhanced atomic layer deposition: an experimental and computational study,” Nanoscale 12(3), 2089–2102 (2020). [CrossRef]  

98. F. Naumann, J. Reck, H. Gargouri, B. Gruska, A. Blümich, A. Mahmoodinezhad, C. Janowitz, K. Henkel, and J. I. Flege, “In situ real-time and ex situ spectroscopic analysis of Al2O3 films prepared by plasma enhanced atomic layer deposition,” J. Vac. Sci. Technol. B 38(1), 014014 (2020). [CrossRef]  

99. O. Stenzel, S. Wilbrandt, K. Friedrich, and Kaiser Norbert, “Realistische Modellierung der NIR/VIS/UV-optischen Konstanten dünner optischer Schichten im Rahmen des Oszillatormodells,” Vak. Forsch. Prax. 21(5), 15–23 (2009). [CrossRef]  

100. G. G. Stoney, “The tension of metallic films deposited by electrolysis,” Proc. R. Soc. Lond. A 82(553), 172–175 (1909). [CrossRef]  

101. M. Kot, K. Henkel, F. Naumann, H. Gargouri, L. Lupina, V. Wilker, P. Kus, E. Poźarowska, S. Garain, Z. Rouissi, and D. Schmeißer, “Comparison of plasma-enhanced atomic layer deposition AlN films prepared with different plasma sources,” J. Vac. Sci. Technol. A 37(2), 020913 (2019). [CrossRef]  

102. J. Kim, S. Kim, H. Jeon, M.-H. Cho, K.-B. Chung, and C. Bae, “Characteristics of HfO2 thin films grown by plasma atomic layer deposition,” Appl. Phys. Lett. 87(5), 053108 (2005). [CrossRef]  

103. C. Vallée, M. Bonvalot, S. Belahcen, T. Yeghoyan, M. Jaffal, R. Vallat, A. Chaker, G. Lefèvre, S. David, A. Bsiesy, N. Possémé, R. Gassilloud, and A. Granier, “Plasma deposition - Impact of ions in plasma enhanced chemical vapor deposition, plasma enhanced atomic layer deposition, and applications to area selective deposition,” J. Vac. Sci. Technol. A 38(3), 033007 (2020). [CrossRef]  

104. O. Stenzel, S. Wilbrandt, S. Yulin, N. Kaiser, M. Held, A. Tünnermann, J. Biskupek, and U. Kaiser, “Plasma ion assisted deposition of hafnium dioxide using argon and xenon as process gases,” Opt. Mater. Express 1(2), 278–292 (2011). [CrossRef]  

105. R. Thielsch, T. Feigl, N. Kaiser, S. Martin, S. Scaglione, F. Sarto, M. Alvisi, and A. Rizzo, “Comparison of the optical properties and UV radiation resistance of HfO2 single layers deposited by reactive evaporation, IAD, and PIAD,” Proc. SPIE3902 (2000).

106. G. Abadias, E. Chason, J. Keckes, M. Sebastiani, G. B. Thompson, E. Barthel, G. L. Doll, C. E. Murray, C. H. Stoessel, and L. Martinu, “Review article: Stress in thin films and coatings: Current status, challenges, and prospects,” J. Vac. Sci. Technol. A 36(2), 020801 (2018). [CrossRef]  

107. G. N. Strauss, “Mechanical stress in optical coatings,” in Optical Interference Coatings, N. Kaiser and H. K. Pulker, eds. (Springer International Publishing, 20), pp. 207–229.

108. S. Y. Ming, S. S. Ying, D. Z. Xia, H. H. Bo, S. J. Da, and F. Z. Xiu, “Influences of annealing on residual stress and structure of HfO2 films,” Chin. Phys. Lett. 24(10), 2963–2966 (2007). [CrossRef]  

109. E. Chason, “A kinetic analysis of residual stress evolution in polycrystalline thin films,” Thin Solid Films 526, 1–14 (2012). [CrossRef]  

110. M. F. Doerner and W. D. Nix, “Stresses and deformation processes in thin films on substrates,” Crit. Rev. Solid State Mater. Sci. 14(3), 225–268 (1988). [CrossRef]  

111. P. Chaudhari, “Grain Growth and Stress Relief in Thin Films,” J. Vac. Sci. Tech. 9(1), 520–522 (1972). [CrossRef]  

112. W. D. Nix and B. M. Clemens, “Crystallite coalescence: A mechanism for intrinsic tensile stresses in thin films,” J. Mater. Res. 14(8), 3467–3473 (1999). [CrossRef]  

113. E. H. Hirsch, “Stress in porous thin films through absorption of polar molecules (and relevance to optical coatings),” J. Phys. D: Appl. Phys. 13(11), 2081–2094 (1980). [CrossRef]  

114. S. Jena, R. B. Tokas, S. Thakur, and D. V. Udupa, “Study of aging effects on optical properties and residual stress of HfO2 thin film,” Optik 185, 71–81 (2019). [CrossRef]  

115. T. Zeng, M. Zhu, Y. Chai, C. Yin, N. Xu, K. Yi, Y. Wang, Y. Zhao, G. Hu, and J. Shao, “Effects of water adsorption on properties of electron-beam HfO2/SiO2 high-reflection coatings,” Thin Solid Films 697, 137826 (2020). [CrossRef]  

116. H. Leplan, B. Geenen, J. Y. Robic, and Y. Pauleau, “Residual stresses in evaporated silicon dioxide thin films: Correlation with deposition parameters and aging behavior,” J. Appl. Phys. 78(2), 962–968 (1995). [CrossRef]  

117. R. L. Nigro, E. Schilirò, G. Mannino, S. Di Franco, and F. Roccaforte, “Comparison between thermal and plasma enhanced atomic layer deposition processes for the growth of HfO2 dielectric layers,” J. Cryst. Growth 539, 125624 (2020). [CrossRef]  

118. S. Papernov, M. D. Brunsman, J. B. Oliver, B. N. Hoffman, A. A. Kozlov, S. G. Demos, A. Shvydky, F. H. M. Cavalcante, L. Yang, C. S. Menoni, B. Roshanzadeh, S. T. P. Boyd, L. A. Emmert, and W. Rudolph, “Optical properties of oxygen vacancies in HfO2 thin films studied by absorption and luminescence spectroscopy,” Opt. Express 26(13), 17608–17623 (2018). [CrossRef]  

119. Y. Wei, Q. Xu, Z. Wang, Z. Liu, F. Pan, Q. Zhang, and J. Wang, “Growth properties and optical properties for HfO2 thin films deposited by atomic layer deposition,” J. Alloys Compd. 735, 1422–1426 (2018). [CrossRef]  

120. C. Harthcock, S. R. Qiu, P. B. Mirkarimi, R. A. Negres, G. Guss, M. G. Menor, G. Bhowmik, and M. Huang, “Origin and effect of film sub-stoichiometry on ultraviolet, ns-laser damage resistance of hafnia single layers,” Opt. Mater. Express 10(4), 937–951 (2020). [CrossRef]  

121. K. Kukli, M. Ritala, J. Sundqvist, J. Aarik, J. Lu, T. Sajavaara, M. Leskelä, and A. Hårsta, “Properties of hafnium oxide films grown by atomic layer deposition from hafnium tetraiodide and oxygen,” J. Appl. Phys. 92(10), 5698–5703 (2002). [CrossRef]  

122. N. V. Nguyen, A. V. Davydov, D. Chandler-Horowitz, and M. M. Frank, “Sub-bandgap defect states in polycrystalline hafnium oxide and their suppression by admixture of silicon,” Appl. Phys. Lett. 87(19), 192903 (2005). [CrossRef]  

123. J. L. Lyons, A. Janotti, and C. G. van de Walle, “The role of oxygen-related defects and hydrogen impurities in HfO2 and ZrO2,” Microelectron. Eng. 88(7), 1452–1456 (2011). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplementary Material

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1.
Fig. 1. The effect of process parameters on the non-uniformity (NU) HfO2 thin films on a 200 mm diameter surface and growth per cycle (GPC) deposited varying a) the precursor (TDMAH) pulse with a carrier gas (Ar) flow rate of 80 sccm b) the carrier gas (Ar) flow rate with TDMAH pulse of 3.12 s c) the precursor (TDMAH) pulse with a carrier gas (Ar) flow rate of 160 sccm, and d) the precursor and plasma purge with precursor (TDMAH) pulse of 3.12 s and Ar flow rate of 160 sccm. All thin films were deposited at 100°C with 200 ALD cycles. The sequence of the PEALD cycle is given where x denotes the varied parameters.
Fig. 2.
Fig. 2. The effect of deposition temperature on NU and GPC. The thin films were deposited using 1482 ALD cycles, 160 sccm Ar flow and an ALD sequence of 3.12s/5s/3s/5s.
Fig. 3.
Fig. 3. AFM and SEM images of the film deposited at 100°C and 250°C with 1482 ALD cycles; a) AFM image of a hafnium oxide thin film, d ≈ 273 nm, deposited at 100°C with a roughness RMS = 0.5 nm; b) AFM image of hafnium oxide deposited at 250°C with d = 309 nm and RMS = 7.3 nm; SEM images of hafnium oxide at c) 100°C and d) 250°C, respectively; crystalline grains are observed at 250°C while the films grow rather smooth at 100°C. The thin films were deposited using an ALD sequence of 3.12s/5s/3s/5s.
Fig. 4.
Fig. 4. XRD diffractogram of HfO2 films (1482 ALD cycles) deposited at deposition temperature from 100°C to 250°C. The peak positions corresponding to tetragonal (ICDD: PDF card No. 08–0342), orthorhombic (ICDD: PDF card No. 21-0904) and monoclinic (ICDD: PDF card No. 43-1017) phase are given as reference. The thin films were deposited using an ALD sequence of 3.12s/5s/3s/5s.
Fig. 5.
Fig. 5. The region of FTIR spectra corresponding to -OH stretching for the HfO2 films (1482 ALD cycles) deposited at a temperature from 100°C to 250°C.
Fig. 6.
Fig. 6. Dispersion of HfO2 thin films (1482 cycles) deposited at different deposition temperatures determined from a) ellipsometry and b) transmission-reflection measurement using UV-VIS spectroscopy. The thin films were deposited using an ALD sequence of 3.12s/5s/3s/5s. Details of the analysis are provided in Supplement 1.
Fig. 7.
Fig. 7. Variation of the refractive index at 632.8 nm and growth per cycle (GPC) with plasma pulse time. The ALD cycles of 200 were used in these depositions.
Fig. 8.
Fig. 8. AES depth profiles of HfO2 thin films (1482 ALD cycles) deposited a) using an initial condition with 3 s plasma pulse, 200 sccm O2 gas flow, 100 W ICP power b) 3 s plasma pulse, 20 sccm O2 gas flow, 100 W ICP power c) 3 s plasma pulse, 200 sccm O2 gas flow, 500 W ICP power d) 7 s plasma pulse, 200 sccm O2 gas flow, 100 W ICP power.
Fig. 9.
Fig. 9. Comparison of optical losses (OL = 100% - R – T) of hafnium oxide (thickness 260 - 275 nm/1482 ALD cycles) thin films deposited at various plasma conditions. The varied parameters are marked in bold.
Fig. 10.
Fig. 10. Reflectance spectra of an antireflection coating consisting of HfO2, SiO2, and Al2O3 on single and double-side coated fused silica substrates at a) 1064 nm and 532 nm b) 1064 nm, 532 nm, and 355 nm c) 1064 nm, 532 nm, 355 nm, and 266 nm. A comparison of the designed and deposited thin-film system is shown. The backside reflection of the single-side coated sample is subtracted.

Tables (5)

Tables Icon

Table 1. Thin film properties of the HfO2 deposited at 100°C - 250°C.a

Tables Icon

Table 2. Thin film properties of HfO2 films (1482 cycles) deposited at 100°C to 250°C.

Tables Icon

Table 3. Thin film properties of the HfO2 deposited at different plasma conditions.a

Tables Icon

Table 4. The effect of plasma parameters on refractive index and absorption in HfO2 thin films (1482 ALD cycles).

Tables Icon

Table 5. The composition of Al2O3[HfO2/Al2O3]xSiO2 thin films and design used in antireflection coatings.a

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

σ = 1 6 E s ( 1 ϑ s ) ( 1 R f 1 R s ) t s 2 t f
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.