Paper

Integrated approach for low-temperature synthesis of high-quality silicon nitride films in PECVD using RF–UHF hybrid plasmas

, and

Published 5 January 2016 © 2016 IOP Publishing Ltd
, , Citation B B Sahu et al 2016 Plasma Sources Sci. Technol. 25 015017 DOI 10.1088/0963-0252/25/1/015017

0963-0252/25/1/015017

Abstract

This study investigates low-temperature plasma nitriding of hydrogenated silicon (SiNx:H) film in radio frequency (RF) and RF–ultra-high frequency (UHF) hybrid plasmas. To study the optimized conditions for the deposition of SiNx:H film, this work adopts a systematic plasma diagnostic approach in the nitrogen–silane and nitrogen–silane–ammonia plasmas. This work also evaluates the capability of plasma and radical formation by utilizing different plasma sources in the PECVD process. For the plasma diagnostics, we have purposefully used the combination of optical emission spectroscopy (OES), intensified CCD (ICCD) camera, vacuum ultraviolet absorption spectroscopy (VUVAS), and RF compensated Langmuir probe (LP). Data reveal that there is significant enhancement in the atomic nitrogen radicals, plasma densities, and film properties using the hybrid plasmas. Measurements show that addition of a small amount of NH3 can significantly reduce the electron temperature, plasma, and radical density. Also, optical and chemical properties of the deposited films are investigated on the basis of plasma diagnostics. Good quality SiNx:H films, with atomic nitrogen to hydrogen ratio of 4:1, are fabricated. The plasma chemistry of the hybrid plasmas is also discussed for its utility for plasma applications.

Export citation and abstract BibTeX RIS

1. Introduction

Hydrogenated silicon nitride (SiNx:H) dielectric films have shown their potential applications in the area of integrated circuits, surface passivation, layer insulation and dielectric capacitors, and wear and corrosion-resistant coatings because of their excellent physical and chemical properties such as high density and dielectric constant and good insulating properties [13]. Additionally, SiNx:H thin film has been extensively used in microelectromechanical systems due to its excellent mechanics and mechanical properties, such as friction resistance, anti-flex, and anti-fatigue [14]. Also, SiNx:H films exhibit transparency from the visible to the mid-infrared and have a reasonably high optical index (~2.0), making them useful for many photonic applications including nonlinear optics [5], photonic integrated circuits [68], on-chip biosensing [9], and high-quality optical cavities [1012]. In particular, the films prepared by the plasma enhanced chemical vapor deposition (PECVD) method have many advantages such as low deposition temperature, good uniformity, and high growth rate. The low-temperature deposition of SiNx:H films with PECVD generally uses a gas mixture of ammonia (NH3) and silane (SiH4). Numerous studies on structural, chemical, and optical properties of the PECVD grown SiNx:H films, by a different group, have recently been reported [1315].

In particular, low-temperature SiNx:H film deposition with low content of hydrogen using PECVD is of particular concern for the industry. Because of the relatively low dissociation efficiency of the gas N2, typically the deposition process uses NH3 as the nitrogen source [16, 17]. In the deposited SiNx:H film, therefore, there will be some incorporation of hydrogen from the NH3. Moreover, limited work has also been reported on the synthesis of amorphous SiNx:H (a-SiNx:H) using N2 and SiH4 as the source gases [15, 18, 19]. The foremost advantages of using N2 gas over NH3 are the rich availability at a cheaper price than NH3 and its nontoxic (harmless) nature. Interestingly, there is a report of high-rate deposition of good-quality a-SiNx:H film using very high-frequency (VHF) PECVD [19, 20]. In this context, development of new plasma processes as well as new plasma sources is necessary, for the industry to achieve high deposition rates over large areas. Thus, the growing demand for new plasma sources with high deposition rate [15, 20, 21] has received great attention.

Different deposition parameters, such as applied power, substrate temperature, gas flow ratio of the reactant gases, operating pressure, and electrode separation of the parallel plate system or capacitively coupled plasma (CCP) system, can affect the deposition process, deposition rate, and film properties. Numerous studies [15, 22], in the nitrogen-dominated PECVD plasmas, have shown the dependence of the chemical states and film properties on the PECVD process parameters. However, the relation of the radical species and the plasma parameters in the process plasma and the film properties is still not satisfactorily clear [15]. In this regard, there must be careful experimental studies of plasma parameters and their relation to the radical and plasma generation, and their effect on SiNx:H film properties in the process plasmas. This work addresses an integrated approach to systematic plasma and radical diagnostics in the PECVD synthesis of a-SiNx:H films in the N2  −  SiH4 and N2  −  SiH4  −  NH3 gas mixtures. This study also investigates different operating regimes for the low-temperature (without heating the substrate) deposition of quality SiNx:H film utilizing 13.56 MHz radio frequency (RF) and 320 MHz ultra-high frequency (UHF) CCP sources together as a hybrid source. Along with systematic plasma diagnostics, various film analysis tools are used to understand the fundamental plasma surface interactions in PECVD processes and to fabricate low-hydrogen-content SiNx:H film. The paper has been organized as follows. Section 2 illustrates the experimental conditions and the approach for this study; section 3 explains the experimental system and different diagnostics used for this work. Section 4 presents the salient results and discussion in detail. Section 5 is the conclusions.

2. Experimental situation and approach

For this work, we have used both plasma diagnostics and film analysis as the integrated approach. Note that, for the primary objective, to prepare high-density silicon nitride film one has to maximize the Si–N bond concentration and reduce the hydrogen content during the PECVD process. However, the challenge remains to maintain a near stoichiometric ratio of nitrogen and silicon, [N]/[Si]  =  1.3 [23], to avoid 'wrong bonds'. Thus, the choice of nitrogen gas as the main precursor would help to reduce the hydrogen incorporation in the film.

One can consider the PECVD process of N2  −  SiH4 and N2  −  SiH4  −  NH3 gas mixtures, with dominant (⩾90%) contribution from N2. Also, in the process plasmas, there will be many probable reactions corresponding to electron impact dissociation, emission, and ionization [15, 2427]. Accordingly, in the SiNx:H process plasmas there can be electrons, ions ($\text{N}_{2}^{+}$ , N+, N2+), atomic nitrogen (N) radicals, excited N, and N2 neutrals. The electron impact ionization cross section data on N2 molecules [24] indicate that electrons having energy below about 18 eV, 30 eV, and 80 eV have very low ionization cross section for $\text{N}_{2}^{+}$ , N+, and N2+ ions, respectively. These data suggest that there must be a sufficient number of high-energy electrons with energy above 18 eV, preferably in the range 20–80 eV, for efficient ionization and to maintain the plasma. The power absorption in high-pressure PECVD plasmas can proceed through collisional absorption [15]. By its nature, collisional absorption can produce a thermalized population of electrons. It may be realized that such a thermal population can maintain the required ionization levels in the plasma if the electron temperature is high. Otherwise, there should be a high-energy tail of the electron energy distribution function (EEDF). Indeed, the latter case is found to be more efficient in various works [15, 28, 29]. Therefore, LP diagnostics would provide insight into the discharge in various process conditions. Additionally, electron collision with nitrogen (N2) neutral molecules can result in emitted radiation (optical emission lines) of different wavelengths [30]. The neutral nitrogen atom (N) density measurement by the vacuum ultraviolet absorption spectroscopy (VUVAS) method would be useful to investigate their role in the film fabrication. It is also necessary to monitor the excited radicals such as nitrogen and hydrogen by looking at their flow with the help of an intensified charge-coupled device (ICCD) camera during the deposition process. Moreover, adding UHF to RF excitation could contribute to a desirable chemistry for the growth of the SiNx:H film. It can later be seen that addition of the high-frequency (UHF) excitation is more efficient to produce high plasma density and radicals. Also, the plasma produced by the simultaneous operation of both RF and UHF power will be referred to as hybrid plasmas in the remainder of this article.

In the present work, SiNx:H films are fabricated utilizing RF and hybrid plasma sources. We have investigated the favorable deposition conditions using various plasma diagnostic methods. Then, we examine the film properties such as refractive index, chemical composition, and optical properties with ellipsometry, Fourier transform infrared (FTIR) spectroscopy, and UV–visible spectroscopy, respectively. The atomic concentrations of nitrogen and hydrogen in the film are evaluated from the FTIR [31, 32] data to provide qualitative information about the deposited film.

3. Experimental system and plasma diagnostics

3.1. Plasma sources

Figure 1 shows a schematic diagram of the experimental system. Figure 1(a) and figure 1(b) respectively correspond to the top and side views of the PECVD system. The process reactor is a vertical plasma chamber (figure 1(b)), made of stainless steel. It has an internal diameter of about 27.2 cm and a height of about 22.0 cm; there is a load lock chamber for introducing the sample from outside. The system is evacuated with a mechanical vacuum pump, which can provide a base pressure of less than 1  ×  10−3 Torr in the process chamber. Different gas flow rates of N2, SiH4, and NH3, regulated by mass-flow controllers (MFCs), are introduced into the reactor. A constant working pressure of about 500 mTorr is maintained in the PECVD reactor by mixing different proportions of N2, SiH4, and NH3 gases. The relative concentrations of SiH4 and NH3 are varied from 10 to 100 standard cubic centimeters per minute (sccm), with a dominant contribution from N2 (⩾900 sccm).

Figure 1. Refer to the following caption and surrounding text.

Figure 1. Schematic diagram of the PECVD system showing different components and diagnostic tools.

Standard image High-resolution image

Figure 1(a) shows that the experimental system consists of two CCP sources: the RF and the UHF source. The RF source consists of two parallel electrodes: a showerhead (as top electrode), which has a diameter of 20.0 cm and a bottom grounded electrode with a diameter of 11.5 cm that act as the substrate table. The electrode distance between the top and bottom electrode is 8 cm. Gas precursors are supplied through the showerhead (figure 1(b)). Substrates are grounded and placed on the substrate table. In the present studies, there is no additional heating by the heater for the deposition of SiNx:H film. A commercial Dressler generator (at 13.56 MHz) as the RF power source, with an L-type matching network, is used for plasma generation. The UHF (320 MHz) source is a combination of two parallel 1/4'' copper tubes. These conductors, with a length of a quarter wavelength (λ/4, where λ corresponds to the free space wavelength at the operating frequency  ≈  320 MHz), are assembled through ceramic-insulated feed-thru to insert the process chamber from side ports consisting of an ISO CF100-6'' flange. To apply the continuous-wave (CW) power we employ a low-power signal from a signal generator (hp-8656B) to a broadband high-power amplifier (BBS3C3KPQ (SKU 2037) from the Empower RF system). Both the input and output impedances of the amplifier are 50 Ω. The power amplifier has a maximum output power of 300 W. A coaxial transmission line (RG 393 50 Ω coaxial cable) of length about 23.4 cm is used from the amplifier to the antenna (copper tubes) for the power transmission. The details of the plasma sources are reported elsewhere [15, 28, 29].

3.2. Plasma diagnostics

Figure 1 also presents various systems for the plasma diagnostics. The main plasma diagnostic in the present work is using the normal and RF compensated LP, OES, VUVAS, and ICCD camera.

3.2.1. Normal and RF-compensated Langmuir probe (LP).

As explained earlier, it is important to monitor the plasma parameters to study their role in the plasma processes. It is also well known that, in an RF plasma environment, the LP IV (current–voltage) characteristic is distorted owing to oscillations of the plasma potential (Vp) at the RF frequency and its higher harmonics [33, 34]. Garscadden and Emelus [33] reported that the general feature of this distortion is to alter the LP IV characteristic towards a more negative voltage, providing incorrect readings for the LP floating potential (Vf) and Vp. For sinusoidal excursions of Vp and an LP biased to retard electrons, the electron temperature (Te) is determined accurately as long as the potential variations remain within the exponential regime of the LP characteristic (i.e. do not take the LP into the electron saturation region). Large excursions of Vp that can occur in high-density plasma can provide an overestimation of the Te and underestimation of the plasma density (n0). Numerous studies [15, 3440] with different RF compensation schemes have been proposed and reported to eliminate this distortion. For the high-frequency UHF (320 MHz) source operation, one can use an LP without RF compensation. However, in the RF frequency and the hybrid plasmas one has to use RF compensated LP. With this consideration, a new LP structure and tuning procedure, similar to that in [34], is designed, and the RF compensated LP is fabricated. The details of the RF LP and its design concept are presented elsewhere [15]. For the LP data acquisition, in the present studies, a commercial Hiden power supply system (UK) is used. The LP system uses window-based graphics ESPsoft application software, which allows acquisition, storage, and analysis of the LP IV characteristics.

LP analysis from the IV data determines various plasma parameters such as plasma density (n0), Te, Vp, electron energy distribution function (EEDF), etc. The details of LP analysis, the analysis procedure, and its validity and usage are discussed [41, 42] in great detail in the literature.

3.2.2. Optical emission spectroscopy (OES).

For the OES diagnostics (figure 1(a)), the spectral data relevant to different optical emission lines are acquired through an optical fibre (Ocean Optics) with an Acton Spectra Pro 500i spectrometer. The spectrometer has minimum resolution  ≈  0.05 nm, a 10 μm wide entrance slit, and a 1200 grooves mm−1 grating, and can be used in conjunction with a PIMAX Princeton Instruments ICCD camera connected to a PC. For the data acquisition the window based graphics software WinSpec32TM is used.

3.2.3. Vacuum ultraviolet absorption spectroscopy (VUVAS).

The VUVAS system (figure 1) is essentially a combination of a VUV monochromator (ARC VM-502) and a high-pressure micro-discharge hollow cathode lamp (MHCL), symmetrically positioned with respect to the chamber. Figure 1(a) also shows that a magnesium fluoride (MgF2) lens (plano-convex) is placed on each side through the chamber atmosphere. To generate VUV light from the MHCL an on–off modulated dc power of about 10 Hz is supplied. VUV light from the MHCL is focused parallel to the MgF2 lens and is made to pass through the monochromator across the plasma chamber. The emitted radiation relevant to the resonant emission of the N atoms (in the MHCL) is absorbed by the N atoms (ground state) in the plasma. The light transmitted through the plasma is then focused on the monochromator by another MgF2 lens. This UV light is converted to visible radiation by the sodium salicylate scintillator of the monochromator. This light is then detected by a photomultiplier tube, and the amplitude of this is recorded as a voltage signal in the digital oscilloscope.

For the VUVAS measurements, it is necessary to select and detect the relevant, resonant emission of the N atoms of the light source, which is absorbed by the ground state N in the process plasmas. The emission profile of the light source also has to be determined to estimate the absorption profile of N and the N atom density (nA). Additionally, one has to consider the effective absorption path length accurately. In the process plasmas, the guess is, however, not trivial due to the nature of the spatial distribution of N atom density. Also, one needs the information of the gas temperature in the plasma and MHCL light source. Therefore, these aspects should be considered for the development of the VUVAS diagnostic system to make an accurate measurement of nA.

According to Mitchell and Zemansky [43], the density of the absorbers (or the radical density, nA) can be connected to the parameter α called the background absorption coefficient in the plasma, by integrating α over the absorption line profile as

Equation (1)

where λ0, A21, g1, and g2 are respectively the line center wavelength, the transition probability from an upper to a lower level, and the statistical weights of the lower and upper states of the transition. Note that the coefficient α is dependent on the plasma conditions. Additionally, depending upon the particular radicals or absorbing species and their interaction with other species in the background plasma, the parameter α can vary. In particular, at a given pressure, plasma density, and gas temperature, it can be accurately predicted according to the theory of line broadening in plasmas. In the present study, one can calculate the electron–neutral collision frequency at pressure  =  500 mTorr. Considering a typical plasma process with electron temperature, Te  =  4 eV (velocity Ve  =  1.2  ×  106 m s−1), neutral density (at 500 mTorr, gas temperature  =  300 K) ~ 1.6  ×  1022 m−3), and collision cross section [24], σ ~ 13  ×  10−20 m2, the collision frequency ~σVe (neutral density) can acquire a value of 1.5  ×  109 Hz (a few GHz). In this sense, and from the spectroscopy standpoint, the present operating conditions can be treated as high-pressure conditions. In the high-pressure environment, the line absorption has to be described by a Voigt profile (see [24] for more details), which is a combination of Gaussian and Lorentzian profiles. Further, the absorption coefficient A of the absorbed light in the plasma is a function of various parameters such as absorption path length, gas temperature in the plasma, gas temperature in the MHCL light source, and line absorption profiles [44]. Note that the details of the VUVAS theoretical formulation, the measurement, and the analysis procedure are reported elsewhere [45]. In the present study, the reference line used for the VUVAS measurement is 120.7 nm.

3.2.4. ICCD camera measurement.

During the deposition process, images of the flow of radical species in plasma are simultaneously recorded (figure 1(a)) using an ICCD camera (PI-MAX, Princeton Instruments) with the objective UV-NIKKOR 105 mm. We use two bandpass filters with FWHM (full width at half maximum) of 7.5 nm. The center wavelengths of the filters are 656.2 nm (F10-656.2, Korea Electro-Optics) and 354.7 nm (F10-354.7, Korea Electro-Optics). They are respectively used for measuring the two-dimensional (2D) gaseous flows of hydrogen and nitrogen with the ICCD camera. A cage filter wheel is used to install these filters so that they can be changed quickly and easily without disturbing the setup. The ICCD camera used has a resolution of 512 pixels  ×  512 pixels. It is focused on a typical field of view of 25 mm  ×  25 mm, which gives a magnification of about 20.5 pixels mm−1. The delay time between the plasma generation and the actual measurements is an essential parameter. For all the measurements, the time delays are individually adjusted with respect to the laser Q-switch delays having an accuracy of 2 ns. The ICCD camera is operated in shutter mode to allow very small (up to 2 ns) and precise integration times.

4. Results and discussion

4.1. Plasma diagnostics

4.1.1. Process optimization using OES.

For the reference, we consider the optimized power conditions of the RF (13.56 MHz) and UHF (320 MHz) sources investigated earlier [28, 29] for deposition of nanocrystalline silicon films using PECVD processes. For this purpose, a gas mixture of SiH4 (10 sccm) and H2 (2000 sccm) is utilized. Then both the plasma sources are operated independently by increasing applied power, and the emission intensities from the various excited states of atomic hydrogen (Hα (656.2 nm), Hβ (486.1 nm)) and molecular hydrogen along with the Si (288 nm) and SiH (410–425 nm) emission lines are monitored using the OES diagnostic. Measurements [28, 29] reveal (not shown here) that line intensities produced by the RF source are much lower than those of the UHF source. It is seen that the emission intensities increased with increasing RF and UHF power up to 180 W, and then saturated beyond 180 W. This experiment clearly suggests that, at high power (beyond 180 W), the individual operation of these (RF and UHF) sources would not be effective. However, one can combine RF and UHF sources simultaneously in dual-frequency (hybrid) operation for the PECVD process. Note that, in the hybrid operation, the total power is about twice that of the individual (RF/UHF) source power. However, for efficient radical and plasma production, the hybrid operation would be quite useful for the plasma process. Additionally, the emission intensities measured (not shown [28, 29]) in the hybrid plasmas are significantly (about two to four times) higher than that of individual source operated at high powers. Therefore, in the present experiment hybrid plasma refers to the simultaneous operation of both RF and UHF sources at about 180 W power.

Figure 2(a) presents various optical emission intensities, over a wide wavelength range of 300  −  700 nm, relevant to different excited species in the RF and hybrid plasma processes with the gas mixture of N2 (1000 sccm) and SiH4 (10 sccm). Spectra acquisitions are made from a viewport (figure 1(a)) with an exposure time of 0.5 s to diagnose the plasma emission using the spectrometer. It can be seen that the line intensities of the emission lines (figure 2(a)) produced in the hybrid plasma are much higher than those of the RF plasma. The OES diagnostic is carried out further (figure 2(b)) with the sequential addition of SiH4 gas from 10 sccm to 100 sccm in the N2 (1000–930 sccm) plasmas to investigate the optimize process conditions in the SiH4–NH3 processes. For clarity, only N2 (337.1 nm) and N2 (357.7 nm) line intensities are monitored. It can be seen from figure 2(b) that emission intensity rapidly increases up to the SiH4 flow rate of 20 sccm. Further, the intensity does not change much, and it tends to saturate, at the higher flow rates. Accordingly, the gas flow rates of N2 at 1000 sccm and SiH4 at 20 sccm are considered as the optimized/favorable conditions and marked as a dashed line in figure 2(b) for the N2–SiH4–NH3 deposition process. Then, NH3 gas is added successively with different flow rates to the reactive mixture of N2 (1000 sccm) and SiH4 (20 sccm). Figure 2(c) presents the overall variation of intensities (of N2 (337.1 nm) and N2 (357.7 nm) emission lines) as a function of NH3 flow rates. It is evident from figure 2(c) that even a small addition of NH3, about 1% (10 sccm), can reduce the N2 intensity significantly. Thus, the OES diagnostic suggests that the mixture of N2 (1000 sccm)  +  SiH4 (20 sccm)  +  NH3 (10 or 20 sccm), would be the favorable conditions for the SiNx:H process. Note also that the higher emission intensities, in figure 2, of the hybrid plasmas than of RF plasmas, correspond to more excitation and dissociation of radicals and molecules in the process plasmas. This nature of variation in the N2–SiH4–NH3 plasmas is further investigated using LP and VUVAS diagnostics, which will be discussed later.

Figure 2. Refer to the following caption and surrounding text.

Figure 2. (a) Typical OES spectra obtained in the RF and hybrid plasmas indicating the effectiveness of the hybrid plasma; (b), (c) the emission intensity variations of N2 (337.1 nm) and N2 (357.7 nm) lines as a function of SiH4 and NH3 flow rate.

Standard image High-resolution image

4.1.2. Diagnostics in N2–SiH4 and N2–SiH4–NH3 processes.

Additionally, during the deposition process by addition of N2, SiH4, and NH3, there will also be formation of hydrogen (H2 gas and H radicals) and nitrogen species in the plasmas. Thus, to examine the gaseous flows or distribution of various species such as N2 and H, we carry out measurements using the ICCD camera. As explained above, image intensities are measured using the camera corresponding to the favorable process conditions for the mixture of N2 (1000 sccm) and SiH4 (20 sccm). Then the effect of NH3 addition to the N2 (1000 sccm) and SiH4 (20 sccm) mixture is also monitored. Figures 3(a) and (b) show the image intensities in two dimensions using filters of nitrogen and hydrogen, respectively. In the figures, the upper colored scale bar represents the magnitude of image intensities of the gaseous flows. All figures are shown with the same scale in the range from 1000 to 90 000 (in arbitrary units) to make a visual comparison between the RF and hybrid plasmas. The figure also depicts the relevant positions A, B, and C for the showerhead (of the RF source), UHF electrodes, and substrate, respectively. One can locate the relative intensities from the background color corresponding to flows of N2 and H, represented by the black dotted circles, at the top of each scale bar.

Figure 3. Refer to the following caption and surrounding text.

Figure 3. Image intensities using the ICCD camera with the (a) nitrogen filter and (b) hydrogen filter, to visualize the gaseous flows by utilizing RF and hybrid plasmas at different flow rates of NH3. The details are given in the text.

Standard image High-resolution image

Figure 3(a) clearly shows the effectiveness of using hybrid plasmas. The two panels on the left-hand side present the image intensities in the absence of NH3 (0 sccm flow rate). The panel at the bottom shows distinct regions of yellow and red colors that represent high image intensities of nitrogen flows generated by the hybrid plasmas. The corresponding intensities in RF plasmas, represented by the green and light-blue colors, are lower than those of hybrid plasmas. However, at the substrate location (point C) the figure shows moderate intensity (with light-green background) in the hybrid plasmas. With the addition of NH3 with flow rate 10 sccm, there is a major change in the intensity. The yellow and red colors are replaced by a green background above the UHF electrodes. Further, the green background under the UHF electrodes is replaced by a semi-blue background (with the dotted circle in the substrate location), which corresponds to the reduction in the image intensity. With further increase of flow rate to 20 sccm, the N2 species are confined in between the showerhead and the UHF electrodes. There is a noteworthy reduction in the intensity (represented by the dark-blue color) at the higher flow rate of 30 sccm. Note that these 2D plots represent images of the flow of radical species in plasmas during the deposition process. Further, the appreciable change in the intensity indicates that the plasma is only confined to nearby regions of the RF and UHF electrodes with higher NH3 flow rates. Note also that the reduction in the image intensities is consistent with the OES data, as explained in figure 2(c) above. This modification in the image intensities indicates a considerable change in the plasma parameters that is to be explained further by LP data. In contrast to the intensities relevant to the N2 flows, the corresponding intensities using the hydrogen filter are much lower, as shown in figure 3(b). In the absence of NH3, there is a small trace of hydrogen intensities (represented by the green and yellow colors), which is confined to the UHF electrode only. Figure 3(b) shows that most parts of the figures are filled with the blue background color, which corresponds to a very low value of intensity in the scale bar.

For a better understanding of the plasma and radical formation during the PECVD process, figures 4(a)(e) present the LP and VUVAS data taken in N2–SiH4 plasmas by using RF and hybrid power. The process condition is changed by varying the SiH4 flow rate at a fixed operating pressure of 500 mTorr. It is obvious from figure 4(a) that Te increases linearly from 3.0 to 9.0 eV, and 2.0 to 7.5 eV, in RF and hybrid plasmas, respectively. Figure 4(a) also shows that the value of Te in RF plasmas is higher than that in hybrid plasmas, and the overall behavior of Te variation is quite similar in the two cases. Note that the measured Te is a thermalized population in the thermal equilibrium, which means that electrons with energy higher than Te are present in the process plasmas. Additionally, to maintain the ionization and to sustain the discharge one requires the minimum energy to be approximately the ionization potential of nitrogen (~18 eV) [24]. However, the measured Te values are much less than this ionization energy. In this sense, one can expect here the role played by the high-energy electrons to maintain the required ionization and dissociation processes. Note that, in the previous work [15], it has been shown that there is a significant high-energy electron tail in the EEDF using the hybrid plasmas (by adding the UHF source) that enables dissociation of N2 and ionization of N atoms much better in the hybrid plasma than in the RF plasma. It can be seen that nA (figure 4(b)) has a lower value in the RF plasmas than in hybrid plasmas. As expected, the absorption coefficient A (figure 4(c)) also acquires a lower value in the RF plasmas than in the hybrid case. Also, the profiles of nA follow the corresponding profiles of A. Figure 4(d) shows that the n0 value is much higher in the hybrid plasmas. Thus, LP and VUVAS measurements clearly show that both the ionization and dissociation in the hybrid plasmas are quite effective as compared to the RF plasmas. There is also similar variation of Vp (figure 4(e)) from about 108 to 125 V and 105 to 115 V in the RF and hybrid plasmas, respectively. One can translate this Vp into an equivalent energy (=  eVp; e is the electron charge) for ions in plasmas. Now, consider the typical case of 20 sccm SiH4 flow rate of figure 4. From figure 4(e), one can see that Vp  ≈  115 and 111 V in RF and hybrid plasmas, respectively. The corresponding ion energies (=  eVp) are about 115 and 111 V in RF and hybrid plasmas, respectively. Note that the cross sections for charge exchange collisions of N ions (N+) with N2 can be of three types: slow ions, $\text{N}_{2}^{+}$ , and fast N (see figure 1 and table 1 of [46]). It can be seen (figure 1 of [46]) that the charge exchange cross section (σi) of slow ions is significantly higher (about one order of magnitude) than that of $\text{N}_{2}^{+}$ . Also, below ion energy ~600 eV, σi is negligible [46]. Therefore, in the present study the ion charge exchange collision will mostly generate slow N+ ions. Further, the value of σi relevant to ion energy ~111 eV and 115 eV is about 5  ×  10−19 m2, and the value of neutral nitrogen (Nneutral) at an operating pressure of 500 mTorr is about 1.6  ×  1022 m−3. Using these values of σi and Nneutral, one can calculate the ion charge exchange path length for nitrogen [46] as Lch,ex  =  1/(σiNneutral)  ≈  0.1 mm. Corresponding to the plasma parameters, the values of Debye length (λd  =  √ (ε0Te/(n0e2) and ε0 is the permittivity of free space) are about 0.07 mm and 0.04 mm in the RF and hybrid plasmas, respectively. Considering a sheath thickness (ws)  ≈  2–3 λd, one can see that Lch,ex ~ ws. As mentioned earlier, the substrate is grounded in the present experiments, and there is no externally applied bias. Thus, the accelerated ions (by the Vp) crossing the sheath will undergo charge-exchange collisions in the sheath and do not affect the substrate much by losing their charge effectively.

Figure 4. Refer to the following caption and surrounding text.

Figure 4. (a)  −  (f) Detailed information on plasma parameters, radical density, and power coupling in RF and hybrid plasmas in the N2–SiH4 PECVD process. All figures clearly dictate the effectiveness of the hybrid plasmas for the radical and plasma generation.

Standard image High-resolution image

Note that, at 100 sccm SiH4 flow rate, there is a significant improvement in nA (figure 4(b)) due to the rapid enhancement of the absorption (high value of A, figure 4(c)) in the plasma. The high value of nA can be attributed to high values of Te. However, in figure 4(d), n0 values of both RF and hybrid plasmas are close to each other at a very high flow rate (100 sccm) of SiH4. To obtain a better understanding of the PECVD processes, one can examine the plasma process by the power balance estimation. It is obvious that, during the electron impact ionization and dissociation processes, power in the plasma is coupled through collisions. We recall that, in the process plasmas, the main precursor is the N2 gas. Consider that an electron undergoes a collision with a neutral nitrogen gas molecule. After the collision event, its oscillatory motion will be affected, and momentum will also be randomized. In the steady state condition, the average energy (electron temperature)  ≈  Te (in the unit of eV) is such that the energy imparted to the neutral gas in a collision is equal to the energy gained by electrons in the time between two successive collisions. Accordingly, the former will acquire a value of about Te(2me/Mn) (here Mn and me are respectively, the mass of an electron and N2). The latter can be written as ~Pa/νc (where Pa and νc are the power absorbed from the applied ac fields and the collision frequency respectively). Equating these terms, one will obtain Pa  =  Te(2me/Mn)νc. Then, the power loss in the plasma per unit volume (Ploss) through collision can be expressed as

To determine the power loss or power coupling in the plasma, consider the case of electron–neutral collision at a pressure of 500 mTorr (relevant to the N2 flow rate variation from 1000 to 930 sccm for a SiH4 variation of 0 to 100 sccm). Using the collision cross section data [24] for nitrogen and plasma parameters (figures 4(a) and (d)), one can determine the value of Ploss. For a simple calculation, one can assume the plasmas to be uniform and homogeneous. Relevant to the chamber height of about 22 cm and diameter about 27.2 cm, the volume of the plasma is V  =  0.013 m3. The product of Ploss and V can give the value of power loss Pcol (in the unit of W) due to collision in plasmas. We recall here that the applied RF and hybrid power are 180 W and 360 W respectively. Figure 4(f) presents the estimated power loss in the plasma due to electron–N2 collisions. For simplicity, the power loss in the coaxial transmission line (coaxial cable) and heating of chamber wall are ignored in the power estimation. Figure 4(f) shows that power loss in the plasma increases with increase in SiH4 flow rate up to 20 sccm, and then it tends to saturate at a higher flow rate of SiH4. It indicates that, with an increase in flow rate, most of the power will be spent in the dissociation of SiH4 gas. This may be the reason why the n0 values (figure 4(d)) of both RF and hybrid plasmas are somewhat closer at 100 sccm of SiH4 due to closer values of Pcol (figure 4(f)). From figure 4(f), it is seen that the significant (>60% of applied power) power loss takes place at a SiH4 flow rate of 20 sccm. Accordingly, NH3 is sequentially mixed with a fixed background of 20 sccm of SiH4 for the N2–SiH4–NH3 process, as shown in figure 5.

Figure 5. Refer to the following caption and surrounding text.

Figure 5. Plots similar to figure 4 showing the plasma and radical formation information in the N2–SiH4–NH3 PECVD process.

Standard image High-resolution image

Additionally, note that the information on plasma parameters is determined from IV curves below the electron saturation region using the LP analysis procedure reported earlier [41]. Further, with an increase in the gas flow rate of SiH4 there is a decrease in n0 values and increase in Te values (figure 4(a)). Moreover, the value of Pcol (figure 4(f)) also decreases beyond the flow rate of 20 sccm. This variation indicates that, apart from the electron impact process with N2, substantial power is also spent for the dissociation of SiH4. Note also that sequential addition of SiH4 would increase the electronegativity and produce dust particles. Thus, the negatively charged dust would enhance the bulk electric field and cause the electron energy (Te) to increase [47, 48]. However, there is no trace of any oscillatory nature in the probe current of the IV characteristic (not shown here) observed near the electron saturation region (which can occur due to negatively charged dust). Thus, the concern of dust particles in the LP data is believed to be not serious.

Note that, for the LP measurements, the LP is inserted into the discharge chamber from the side port (figure 1(a)). Note also that the LP is symmetrically placed with respect to the UHF and RF electrodes at the mid-plane of the chamber. Specifically, the LP is positioned in between the UHF electrodes (at a distance of 2.0 cm from each electrode of the UHF source and 4 cm from the shower head), which is well outside the sheath. In this sense, in the hybrid discharge, the LP is located near the UHF power source. One may consider here the possibility of the UHF source position compared to the LP position. However, our previous plasma diagnostic study [29], in the plasma processes (for the synthesis of crystalline), shows that the n0 and Te values measured independently in the RF and UHF plasmas are quite comparable. In particular, in the hybrid discharges, n0 values are significantly higher, and Te acquires lower values than that of RF and UHF discharges. Of course, the total applied power in the hybrid plasma is about twice that for individual operation of RF and UHF plasmas.

The plasma chemistry in the hybrid discharges can be visualized further by considering the concept of the plasma sheath in CCPs. A CCP can be modeled (not shown) as a series LCR resonance circuit [48, 49]. In this sense, one can consider the sheaths as capacitors in series with a resistive load (not shown). The LCR circuit consists of an inductor (L) representing electron inertia and sheaths as capacitors (C) in series with a resistive load (R) resembling power dissipation (ohmic heating) via elastic collisions. The characteristic resonance frequency (ωres) of the circuit is given by [50]

Equation (3)

where the terms δ, Leff, and ωpe are respectively the mean sheath thickness, the effective bulk length between the electrodes, and the electron plasma frequency. Considering the situation of plasma excitation directly at resonance frequency ωres, where the plasma can be efficiently heated, for typical n0 ~ 1011 cm−3 (in the present experiment), δ ~ 11 cm (=  c/ωpe, c speed of light in vacuum), and discharge geometries (with Leff ~ 9 cm), ωres ~ 314 MHz. In this case, ωres is much higher than the RF exciting frequency at 13.56 MHz. Therefore, electron resonance heating by low-frequency RF (13.56 MHz) can be avoided in the present scenario. This value (314 MHz) is closer to the UHF excitation frequency (320 MHz) used for the present case. Thus, there can be resonant heating by the UHF source when operated in the hybrid plasmas. Indeed, the EEDF in the case of hybrid plasmas has a significant high-energy tail, as reported earlier [15]. Also, efficient heating of the plasmas at ωres depends on the nonlinearity (plasma sheath) of the system, which can allow the generation of higher-order harmonics of the RF (at fundamental) driving frequency in the plasma currents [51]. The detailed investigation of the electron heating in CCPs is beyond the scope of this paper.

In contrast to N2–SiH4 process plasmas (figure 4), Te values (figure 5(a)) in N2–SiH4–NH3 processes rapidly decrease on adding NH3 even for a small amount of 10 sccm. At higher NH3 flow rate (>10 sccm), Te tends to saturate. Accordingly, nA and A values in figures 5(b) and (c) are also much lower. Moreover, n0 values, in both RF and hybrid plasmas (figure 5(d)), monotonically decrease with increasing NH3 flow rate. Similar variations of Vp are also observed in figure 5(e). The power expended or power lost (Pcol) by the electron with N2, in figure 5(f), has a similar profile to Te (figure 5(a)). The saturation of Pcol at higher NH3 flow rates (figure 5(f)) indicates that most of the power will be spent in the dissociation of NH3 due to its lower dissociation energy. Further, with the increase in the NH3 flow rate, Te becomes small. Probably for this reason the n0 value decreases significantly with the enhancement of NH3 flow rate. Note that the high-energy tail of the EEDF is drastically suppressed with the addition of NH3 [15]. This result suggests that films deposited with higher NH3 flow rate would contain a higher fraction of –NH or –NH2 groups. Note moreover that LP and VUVAS measurements are also consistent with those measured by the ICCD camera (figure 3), where the image intensities of the gas flows are very small even at low flow rate of NH3. Therefore, the LP and VUVAS result indicates that low content of NH3 would be a favorable condition for the synthesis SiNx:H film.

4.2. Film analysis

On the basis of systematic plasma diagnostics, SiNx:H films are deposited in the N2–SiH4 and N2–SiH4–NH3 plasmas. Accordingly, SiNx:H films are deposited on high-quality Corning EAGLE XG AMLCD glass substrates. We change the operating conditions by changing flow rates of SiH4 and NH3 in N2–SiH4 and N2–SiH4–NH3 plasmas, respectively. For the present study, the film thickness is about 200 nm. Note also that high temperature causes desorption of hydrogen [5254] from PECVD SiNx:H. Also, at high temperature N–H bonds are especially reduced rather than the Si–H bonds. However, we employ the low-temperature deposition process for this work. Initially, films without adding NH3 are fabricated.

The optical properties of the deposited a-SiNx:H films are investigated using UV–vis spectroscopy (spectra not shown) in the 200–1000 nm wavelength range. The estimation of the band gap (Eg) of the deposited film is done using the Tauc equation [55, 56]

Equation (4)

where α, h, ν, and B are, respectively, the absorption coefficient, Planck's constant, the frequency of the incident photon, and the optical density of states. Note that α is directly determined from the optical transmission data on the basis of the Beer–Lambert law,

Equation (5)

where I0, IT, and t are, respectively, the intensity of the incident light, the intensity of the transmitted light, and the thickness of the a-SiNx:H thin film. The thickness of the films is measured with a KLA-Tencor Alpha-Step IQ profiler. Figure 6(a) presents the information of Eg (eV) of the deposited film in the SiH4–N2 plasmas, as a function of SiH4 flow rates. The figure shows that Eg gradually reduces with increasing SiH4 flow rates. For all the films in RF plasmas, Eg decreases linearly from about 3.3 to 2.2 eV. This variation is according to the variation of transmittance from about 81 to 56%, for the SiH4 flow rate ranging from 10 to 50 sccm. In figure 6(a), there are small dotted circles, which help in comparing the film properties between RF and hybrid plasmas. Note that, in the hybrid plasmas, the Eg have relatively higher values than in RF plasmas. The films deposited using the hybrid plasmas shows a marginal variation in the transmittance (~89 to 80%) and band gap (Eg ~ 4.7 eV to 3.2 eV). One can correlate these data with the nA profile of figure 4(b). Figure 4(b) shows that nA does not change much, even with higher SiH4 flow rates up to 40 sccm. However, one can also expect here the effect of hydrogen content in the film along with the nitrogen content. It can be noted that, at lower SiH4 flow rate, the variation of the band gap is due to the formation of more Si–N network rather than the presence of SiH or NH groups. Figure 6(a) clearly shows that the films with higher value of the band gap can be produced in the hybrid plasmas due to effective dissociation of N2 and formation of higher content of atomic N than that of RF plasmas. Further, to investigate the optical features, spectroscopic ellipsometry is used with the variable angle spectroscopic ellipsometer (VASE) instrument (Woollam) that is attached to an HS-190 monochromator and a VB-400 control module. Measurements are made using spectroscopic scans (with wavelengths ranging from 300 to 1000 nm) at an incident angle of 70°. The results (not shown) are fitted to a model consisting of an approximately 300 μm thick silicon substrate provided with a transparent Cauchy layer on top that represent the a-SiNx:H film. Further, the optical property of matter is described by the complex refractive index, n  =  n  +  iκ. Here n and κ are, respectively the real part of the refractive index and the extinction coefficient. The parameter κ is a measure of the absorption (in the material) and is neglected in the Cauchy model. Then, by fitting the measured value of ellipsometry data to the model, the real part of n as a function of wavelength is extracted. In the present study, the value at 633 nm is used as the representative value of n for simplicity. Figure 6(b) presents the value of n, which show its dependence on SiH4 dilution in the hybrid plasmas. Corresponding to the lower value of optical transmittance at a SiH4 flow rate of 10 sccm, figure 6(b) shows that the values of n are higher in the RF plasma than the hybrid plasma. Also, the figure shows that the n value increases with increasing SiH4 flow rate.

Figure 6. Refer to the following caption and surrounding text.

Figure 6. (a) Optical bandgap and transmittance of the deposited films in N2–SiH4 plasmas, using the RF source, with changing SiH4 flow rate. Both values decrease with increasing SiH4 flow rates. However, there is enhancement and control of their values even at high flow rates of NH3, in the hybrid plasmas. (b) Effect of SiH4 flow on the refractive index using the RF and hybrid plasmas.

Standard image High-resolution image

To study the chemical bonding states of the SiNx:H films Fourier transform infrared (FTIR) absorption spectra are measured with a Bruker IFS 66v/S spectrometer in the mid-infrared range from 600 to 4000 cm−1. FTIR spectra (in figure 7), taken at a different flow rate of the SiH4, show various peaks representing different bonding states. The figure shows the absorption coefficient (αabs) calculated by dividing the intensity of the infrared absorbance (A) by the film thickness (t) [57].

Equation (6)

where (αabs) represents the Napierian absorption coefficient and the coefficient 2.303 arises from the conversion factor ln(10). The dominant peaks in the spectra are attributed to the Si–N stretching mode, Si–N rocking mode, Si–H stretching mode, N–H stretching mode, and O–H stretching mode, located at about 880–900 cm−1, about 1190 cm−1, about 2100 cm−1, about 3350 cm−1, and 3550 cm−1, respectively [58]. The figure also presents NSi3 asymmetric stretching around 970 cm−1, which indicates the stoichiometric configuration. There is also a region representing N–H2 scissors, and noise (especially in higher SiH4 flow rate) may be attributed to a defect in the substrate or CO2 noise [59]. It is also noticeable that the peaks corresponding to OH and NH stretching vibrations have very low intensities. Comparison between figures 6 and 7 shows that the higher value of the band gap at low SiH4 dilution  ⩽20 sccm may be due to the formation of silicon nitride. This may be due to the replacement of Si–Si bonds with strong Si–N or Si–H bonds along with the higher fraction of NHn content. Indeed, the peak position of the Si–N stretching mode upshifts from 840 cm−1 (at SiH4 flow rate  =  10, 30, 50 sccm) to 878 cm−1 (at SiH4 flow rate  =  20 sccm). This observation corresponds to the induction effect due to the higher electronegativity of N atoms than Si atoms [58]. Note that all the prepared films contained very few OH groups, which are assumed to be due to the moisture during the process and after air exposure.

Figure 7. Refer to the following caption and surrounding text.

Figure 7. FTIR spectra of the silicon nitride film deposited with different SiH4 flow rates in hybrid plasmas with N2–SiH4 mixtures.

Standard image High-resolution image

Further, it may be noted that, using different operation parameters, one can develop different stoichiometries (e.g. Si–H and N–H ratios in the film can change significantly) of the SiNx:H film, creating an impact on the performance of microelectronic devices [5254]. FTIR data in figure 7, corresponding to Si–H bonding, show a distinct nature (showing that the peak position of the absorption coefficient is different at a different SiH4 flow rate). Note that the deposition rate (to be discussed later) at a SiH4 flow rate of 10 sccm is much smaller than that at 20 sccm. As predicted by the plasma diagnostic result, the SiH4 flow rate of 20 sccm can be a favorable condition for silicon nitride film preparation in the N2–SiH4–NH3 plasmas. Accordingly, film properties in the hybrid plasmas are studied by changing the NH3 flow rate, in the fixed background of N2 (1000 sccm) and SiH4 (20 sccm) gases for the N2–SiH4–NH3 plasmas.

As compared to NH3-free (without adding NH3) deposition (in figure 7), figure 8 clearly shows that NH3 addition enables two major changes: enhancement of the peak corresponding to the Si–N rocking mode and flattening of the peak corresponding to the Si–H stretching mode, with increasing NH3 flow rate. This variation indicates that there is a significant reduction in the Si–H bonding fraction due to the addition of NH3 (in figure 7). It may be noted that the electron temperature (Te) is also very low (figure 5(a)), in the range of 2–3 eV, and the collisional power loss also low (figure 5(f)), with increasing NH3 addition. This plasma characteristic suggests that appreciable power in the plasma is expended on the dissociation of NH3 due to its lower dissociation energy. The defect in the substrate or CO2 noise is expected due to heating because of the higher reflected power of the UHF source. The figure also shows that there is a negligible trace of OH bonding fraction. Moreover, to investigate the effect of NH3 incorporation in the film using hybrid plasmas, the bond concentrations (N) are evaluated from the FTIR data using the following equation, where the proportionality factors KX-Y are reported by different groups [32, 60] and are presented in table 1:

Equation (7)

where α(Λ) and Λ are respectively the absorption coefficient and the angular frequency.

Table 1. Bond and mode configuration, proportionality factors KX-Y, and maximum of absorption peaks for various bonds.

Bonding (X-Y) Mode Λmax (cm−1) KX-Y  ×  (1019 cm−2) Configuration
Si–N Asymmetric stretching 875 2.07
N–H Stretching 3350 12
Si–H Stretching 2140 11 H–Si–SiN2
2175 40 H–Si–HN2
2220 20 H–Si–N3
Figure 8. Refer to the following caption and surrounding text.

Figure 8. FTIR spectra of the deposited film with different NH3 flow rates, in hybrid plasmas, in N2–SiH4–NH3 processes.

Standard image High-resolution image

Figure 9(a) illustrates various bond concentrations of Si–N, Si–H and N–H of SiNx:H films, deposited at different NH3 gas flow rates, calculated from the FTIR spectra (figure 8) using equation (4) and table 1. The baseline is modeled by inspection in Origin (software) and subtracted, and respective bond concentrations are calculated from the peak magnitudes. The figure clearly shows that N corresponding to Si–N bonding is at least one order of magnitude higher than that of other two. The decrease in the N values with NH3 incorporation is consistent with the earlier diagnostic result in figures 3 and 5. Figure 9(a) also shows that the curves corresponding to Si–N and Si–H bonds are similar. The total concentrations of atomic N and H, calculated using the work of Yin and Smith [32] with the relation

Equation (8)

where N(N) and N(H) correspond to the atomic concentrations of N and H in the film, respectively. Figure 9(b) shows that, for all the films, the ratio of atomic N to atomic H is about 4:1. Note that N(N) values, corresponding to atomic nitrogen, do not change much (they remain in between 1.7  ×  1022 and 2.0  ×  1022 cm−3) when the NH3 flow rates change from 10 sccm to 100 sccm.

Figure 9. Refer to the following caption and surrounding text.

Figure 9. (a) Different bond concentrations and (b) atomic nitrogen (N) and hydrogen (H) concentrations in the deposited film of figure 8, with changing NH3 flow rates.

Standard image High-resolution image

Moreover, the optical transmittance measurement of the deposited films, in the N2–SiH4–NH3 processes, also does not change much (figure 10). Figure 10 shows a very high transmittance in the range of about 85–90% on addition of NH3. Moreover, the optical bandgap of the films deposited with NH3 addition does not change much (remains in between 4.3 and 4.7 eV). Comparison with the film data without the addition of NH3, in figure 6 (except at NH4 flow rate of 20 sccm), shows that there is a significant improvement in the transmittance of the deposited films by the addition of NH3 even at high flow rates. The probable reason for high transparency may be correlated with the FTIR data (figures 8 and 9). Even though the formation of N–H bonding with the increase of NH3 appears to be increased, this bond concentration does not vary much (figure 9). Also, both Si–H and N–H bonding fractions are very low (an order of magnitude lower than that of Si–N).

Figure 10. Refer to the following caption and surrounding text.

Figure 10. Optical transmittance and bandgap of the deposited films in N2–SiH4–NH3 plasma processes, using hybrid plasmas.

Standard image High-resolution image

Additionally, the presented result using RF–UHF hybrid plasmas can be compared with the other work present in the literature. For ease of understanding, table 2 presents the comparison of film properties deposited by different plasma-based processes. Table 2 shows that most of the studies [13, 14, 16, 17, 58, 61] have utilized substrate heating or high-temperature post-annealing. It is evident that the present study based on a low-temperature deposition process using hybrid discharges can be useful for the deposition of a-SiNx:H films for industry. The optical transmittance and the optical band gap of the deposited films are also in the same range as the literature [13, 16, 17, 58, 63]. Further, due to application of UHF power in the hybrid plasmas, there is higher dissociation of NHn into nitrogen species, which maintains the concentration of Si–N bonding in the range around 5–6  ×  1022 cm−3 (figure 9(a)) throughout the NH3 variation from 0 to 100 sccm. The VUVAS measurements (figures 4 and 5) show that there is a significant enhancement of the N atom radicals in the hybrid plasmas. Thus, impurities such as carbon and hydrogen in the film could be minimized by the higher content of N2, which, in turn, causes the higher optical transparency of the film. Also, the presented data show that the hybrid source would be beneficial for a better film property. In the hybrid plasmas, the low frequency (RF 13.56 MHz) component controls the ion energy while the high frequency (UHF 320 MHz) component controls the electron energy and plasma density [28, 29].

Table 2. Comparison between plasma based processes used for SiNx:H film deposition process.

Deposition method/frequency: PECVD Magnetron sputtering
Low/mid-frequency RF VHF/microwave RF–UHF (present work)
Source and excitation parameters 10–50 kHz [62] atmospheric pressure plasma 13.56 MHz [13, 17] 50 MHz [20] CCP RF power ~ 4–10 W pressure ~ 1 Torr 2.45 GHz [16, 63] power ~ 500 W pressure ~ 0.5–1.5 Torr ne ~ 1010–1011 cm−3 [63] Te ~ 2–6 eV [63] 13.56 MHz/320 MHz CCP RF–UHF power ~ 180 W pressure ~ 500 mTorr 13.56 MHz [14]
RF power ~ 400–450 W [14], 100 W [63]
400–500 kHz [58] high ICP and CCP power ~ 2 kW pressure ~ 10 mTorr CCP RF power ~ 20–100 W pressure ~ 650 mTorr [13] pressure ~ 1 Torr [17] pressure ~ 2–3 mTorr [14] ~ 3–33 mTorr [63]
ne ~ 1010–1011 cm−3 Te ~ 2–10 eV
Deposition rate (nm s−1) 0.05–0.1 [62] 0.1–0.5 [13] 1.0–2.0 [15, 16] 0.5–1.6 0.3–0.8 [14]
0.1–1.1 [58] 2.0–2.6 [17] 0.5–3.5 [63]   0.5–1.3 [63]
Deposition/annealing temperature (°C) 400–800 [62] >300–800 [13] 350 [20] <80 >200 [14]
200 [58] 350 [17] 300 [16] <60 [63]
600 [63]
Refractive index 1.85–2.2 [13] 1.9–2.6 [20] 1.7–1.9 2.0–2.5 [14]
1.9–2.3 [17] 1.6–2.2 [14]
Optical transmittance (%) 85–90
80–90 [58] 65–80 [13, 17] 80–85 [16] 80–85 [63]
Optical bandgap (eV) 2.0–4.0 [13] 2.2–4.7
2.0–5.1 [58] 1.5–3.3 [16] 4.14.9 [63]
Si:N ratio 1:3.5 [58] 1:3 [13] 1:1.2 [16]
N:H ratio 4:1

5. Conclusion

This work investigates effects of the RF and hybrid plasmas for the radical and plasma formation in the SiNx:H PECVD processes. With careful analysis, systematic diagnostic tools are utilized to study the deposition process, and the favorable deposition conditions are investigated for quality SiNx:H film fabrication. The presented result shows that the addition of UHF power efficiently enhances the plasma and atomic nitrogen radical density. Also, we have studied the correlation between the various PECVD process parameters of N2–SiH4 and N2–SiH4–NH3 plasmas on the compositional and optical properties of silicon nitride films. The plasma chemistry using hybrid plasmas is also studied. The deposited films, using the hybrid plasmas, show a very low (~25%) content of atomic H as compared to atomic N (~75%). The suitability of the dual frequency hybrid plasma for plasma application is investigated using various plasma diagnostic and film analysis tools.

Acknowledgments

This work was supported by the Leading Foreign Research Institute Recruitment Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT and Future Planning (MSIP) (2011-0031643). The authors sincerely thank both reviewers for their valuable comments, which helped us to significantly improve the overall outlook of the paper.

Please wait… references are loading.
10.1088/0963-0252/25/1/015017