Open access peer-reviewed chapter

Characterization of Hydrogenated Amorphous Silicon Using Infrared Spectroscopy and Ellipsometry Measurements

Written By

Mounir Kassmi

Submitted: 01 May 2022 Reviewed: 12 September 2022 Published: 04 January 2023

DOI: 10.5772/intechopen.108021

From the Edited Volume

Application and Characterization of Rubber Materials

Edited by Gülşen Akın Evingür and Önder Pekcan

Chapter metrics overview

96 Chapter Downloads

View Full Metrics

Abstract

We described the primary mixed compositions of hydrogenated amorphous silicon on the surface of glass (7059) in this chapter and distinguished them optically by combining the outcomes of infrared spectroscopy and ellipsometric tests. The particular hydrogen content of the aspherical voids created determines the energy level of the optical band, which ranges from 1 eV to 4 eV depending on how passivated or unpassivated the composition is. Additionally, the dielectric response is influenced by the size and proportion of the vacuum occupation relative to the surrounding phase, and each dielectric response is based on how much the implicated components have been passivated.

Keywords

  • hydrogenated amorphous silicon
  • infraredspectroscopy
  • ellipsometry
  • band gap energy
  • void
  • passivation
  • vibrational mode

1. Introduction

Researchers and professionals have focused a lot of attention on photovoltaic power since it was first introduced in the engineering area in order to create materials with practical qualities for the conversion of solar-powered energy. For a variety of reasons, including its consistency in quantity, the ease with which it may be elaborated, and its safety, silicon emerged as the most exciting newcomer in the universe of these materials with relation to this transformation cycle [1]. This material has been researched in a number of different forms, including mono-crystalline silicon, proto-crystalline silicon, hydrogenated polymorphous silicon, and hydrogenated amorphous silicon [2]. A few endeavors used infrared spectroscopy and ellipsometry measurements to control the optical characteristics of this intriguing material [3]. In addition, it is imperative that numerous studies be conducted in order to accurately link the hydrogen concentration and the presence of microvoids within this material, using both ellipsometry measurement and other methods [3, 4, 5, 6, 7]. The latter has, however, clearly superior characteristics to the others, including a large absorption coefficient and a direct band gap energy that are easily adjustable using a variety of elaboration approaches by adjusting the temperature and hydrogen flow [7, 8, 9]. A lot of research on the a-Si: H material has been published in the literature over the past three decades, and it has produced a number of intriguing outcomes that have led to the conclusion that this material has good absorption properties. Besides, it has poor transport properties with a short carrier diffusion length of around 300 nm [10] and there is 10–30% efficiency degradation under light soaking owing to the Staebler-Wronski effect [11, 12].

The desired quality for the proper functioning of solar cells based on thin layers of hydrogenated amorphous silicon still presents a challenge for researchers despite the efforts made and the results obtained during this time because there are still a number of phenomena that are not fully understood, such as the impact of the Si▬H bond on the gap energy, microvoid density, optical index, and transport properties [5, 13]. A small-angle X-ray scattering technique (SAXS) provides details of the void covered in the array [14]. It was determined that the optoelectronic characteristics on the surface of glass are particularly influenced by the size and density of voids, and that voids reduce the mass density of such materials [15]. We are better able to comprehend the root of the enhanced transport qualities to the thorough and precise description of the film structure. Many earlier studies have established the importance of hydrogenation in the enhancement of specific properties, as hydrogen decreases defects by reducing the number of dangling bonds, which are responsible for subpar device performance [16, 17]. The physical understanding of this phenomenon is still a long way from being in a place where it can be controlled in order to optimize thin layers. The band gap may be another topic of contention in the current state of science between researchers, some of whom concur with a monotonic decreasing relationship for structural disorder [18, 19] and a monotonic increasing relationship for hydrogen concentration [20, 21]. As a result, some scientists came to the conclusion that neither the hydrogen concentration nor the structure disorder has a standalone effect on the gap.

The interdependency of the two factors is what makes these conversations challenging [22, 23]. Other earlier studies have demonstrated that as the temperature of the thin layer deposit rises, the gap and overall structural material disorder are reduced significantly [24, 25]. The hydrogen concentration and the structural disorder are both influenced by the substrate’s temperature, and both variables are required to explain an observed local minimum of gap, but the silicon monohydride (Si▬H) bond density only accounts for this dependency [26]. All of these outcomes support the assertion that even though research on the a-Si: H material has improved in terms of the latter’s ability to be doped to increase its transport capabilities [27, 28], it’s still challenging to regulate the factors that affected its optical qualities. Instead, because much less material is needed to respond and totally absorb the light, ultra-thin film optoelectronic devices, particularly those built of hydrogenated amorphous silicon a-Si: H, have the potential to be less expensive. The existence of voids and hydrogen bonding are further topics for discussion. Smets et al [4] have shown that when the amount of hydrogen connected to silicon exceeds 14%, the material may contain microvoids, while less than that, it mainly contains vacancies decorated by hydrogen.

Thus, we may determine the film microstructure and the likelihood of occurrence of such configuration, whether isolated or related to another, using the mass density of the film and the intensity of infrared absorption modes. A significant portion of earlier work relied on a trial-and-error method of hydrogenating amorphous silicon, such as using low hydrogenation gas esteems ranging from 2 to 75 sccm, which produced a variety of results. Among these are the characteristic farthest reaches of the existence or absence of vacancies, which are associated with a critical hydrogen concentration (14%), as well as the relationship between the bandgap energy and the density of Si▬H bonds [26].

In amorphous silicon, streams of hydrogenation gas that did not exceed 75 sccm were previously operated. In this chapter, a significant amount of hydrogenation gas—roughly 200 sccm—is sufficient to determine the film’s properties in terms of both its fundamental structure and its optoelectronic properties. Besides, we endeavor to follow, from one perspective, the connection between the measures of hydrogen related to the dimensions of the round voids framed in the film folds, and similarly, the relationship that influences the difference in bond density with the components of the various shapes, considering the infrared vibration frequencies of the bonds. At room temperature, ellipsometric measurements and infrared spectroscopy were used to account for the network’s hydride arrangements. For all tetrahedral configurations, experimental data were examined to clarify the spectral dependency of dielectric functions. The goal of this study is to learn more about the many ways silicon and hydrogen atoms attach using observed optical functions in a-Si: H. This research uses the well-known Bruggeman model (EMA), which has been applied to composite and heterogeneous medium. This research does suggest that optical constant measurements may be a very sensitive probe of microscopic compositions and by the very compact compound matrix within the different heterogeneous formations.

Advertisement

2. Methodologies for analysis and synthesis

Ultrathin a-Si: H films were produced using RF-PECVD. Silane (4 sccm) and hydrogen were used as gas sources for the deposition (gas mixture). A common deposition parameter uses a high radio frequency power value and a 750 mTorr starting pressure (60 W). The hydrogenation gas flow rates were 100 and 200 sccm. The substrate temperature was maintained at 300 °C for the whole 30-minute deposition period.

Using the CuKα line (λ = 1.54056A°), X-ray diffraction analysis (Bruker D8 Advance, Germany) was used to examine the amorphous structure of hydrogenated silicon. Hydrogen concentration was determined using IR absorption tests using a Perkin-Elmer FTIR in the 400–4000 cm−1. Real and imaginary components of dielectric functions were determined using ellipsometry measurements in the spectral range [1–5 eV].

In order to increase the measurement’s sensitivity, an incidence angle θ0 is chosen. However, depending on the optical constants of the samples, the incidence angle is changed. As a result, a value of 70° for θ0 may allow us to reach the maximum amount of reflected light with the sample and also produce a significant oscillation in the Cos (Δ) spectrum. Data collection and analysis were carried out utilizing WinElly II software (version 2.0.0.0), which presumes a completely flat surface.

Advertisement

3. X-ray diffraction analysis

The effects of hydrogen on the network have been extensively studied in the literature even though this topic has previously been brought up thirty years ago [29, 30]. The X-ray data revealed an amorphous structure along with a strong impact of hydrogen on the structure as seen by the numerous updated Si▬H bonds [31]. The presence of a metastable network with various optical characteristics in the film is indicated by this. Additionally, compared to tetrahedral configurations without hydrogen, those that are distinguished by their thickness are frequently much smaller in size. This is because the Si▬H bond is stronger and shorter than the Si▬Si bond.

According to the degree of hydrogenation, the spectra (Figure 1) clearly illustrate the effects of the two hydrogen esteems on the amorphous silicon lattice. Only two peaks can be found in the red spectrum, and they are 2θ1≃ 28.32°, and 2θ2 ≃ 47° in terms of diffraction angles. Further low peaks signaling the emergence of a desired trend in the materials behavior. The action of hydrogen causes an increase in crystallinity, and the size of the crystallites (≃ 71.5 pm) demonstrates that this increase is not consistent with polymorphous silicon.

Figure 1.

X-ray diffraction patterns of a-Si: H thin films.

Advertisement

4. FTIR analysis

The raw FTIR spectra were firstly corrected for incoherent and coherent reflections. In addition, the correction of the absorption of the film (a-Si: H) was obtained by subtracting the measured absorption of a bare part of the substrate. The observable peaks are due to the absorption states caused by the vibrations of various bonds and rely upon the encompassing environment which likewise doesn’t have a comparative energy level. The vast number of peaks in the frequency region between 450 cm1 and 1800 cm1 is due to the fact that its existing conformation is exceptionally wealthy in distinctively situated atomic chains and furthermore their diverse sub-joins.

It has been recognized that an increase in the Si▬H2 concentrations leads to a weakening of the photovoltaic properties which subsequently leads to poor device performance. The above physical conditions are consistent with the following specific bond units: Si▬H and Si▬H2, but we can rarely find Si▬H3 components. Infrared spectroscopy has been used to get the concentrations for the possible configurations SiSi4mHm (m=0,3) considered in the Tetrahedron model [32]. The resolve of the hydrogen content is determined through different processes like elastic recoil detection analysis (ERDA). For our situation, the hydrogen content was calculated by means of the infrared absorption range at the 640 cm1 frequency addressing the wagging mode, just as the density of the SiHx bonds and any remaining fragments asSiO2; that might be shaped inside or at the sample-substrate border. The quantitative process of the IR modes is based on the accurate measurement of the hydrogen content through the oscillator strengths. The density NSiHx (x=1,2) is proportional to the integrated absorption strength of the given mode [33]:

NSiHx=Anαeffωω=AnI,αeff=Aωd.log10eE1

The effective absorption coefficient was calculated using Eq. (1), where d is the thickness of the thin film; the integral I is the absorption strength of each absorption peak, and Aω is the absorbance as a function of the series of permissible infrared frequencies ω. The function within the integral sign (Eq. (1)) is certainly based on the infrared frequency of the vibration of a particular set. The integral traverses the whole range of frequencies proving in any case the presence of such a gathering. The understood functional difference was presented in the form of a peak which can be displayed by a Gaussian function having an area A. The completely perfect fit allows calculating the area of such function, which is the integrated absorption strength I. The proportionality constant An show a discrepancy as the inverse of the oscillator strength, and it is independent of the total hydrogen concentration, and the integral is over the absorption band of interest. The matrix constant An has the dimension of a surface concentration. Some authors have highlighted some values, which are subsequently justified passing through numerous other works. For the wagging mode (≃ 640cm1), it is2.1±0.21019cm2, and for the low stretch mode (2000cm1); it is9.0±1.01019cm2. Lastly, the matrix constant identified with the high stretch mode (≃ 2100cm1) has the worth2.2±0.21020cm2 [34]. These matrix components quantitatively characterizing the microstructure were determined by estimating the beam yield of a resonant nuclear reaction of 15N ions with hydrogen [35]. They are useful for numerically assessing the concentration of hydrogen and the densities associated to Si▬H and Si▬H2 moieties, respectively, through the integral relationship referred to above (Eq. (1)). The density of the hydride bonds (SiH andSiH2) is calculated either by analyzing the FTIR absorbance spectrum or according to the infrared absorption coefficient by applying the above mentioned method. The hydrogen concentration is obtained by dividing the concentration of groups such asSiH/SiH2,SiH2 and SiH by the atomic concentration of crystalline silicon (NcSi51022cm3):

CHat%=NH/NH+NcSiE2

Three distinctive wavenumbers ranges are recognized that compare to absorption modes of SiH bondings: wagging modes around 640 cm−1, bending modes around 860 cm−1 and stretching modes around 2000 cm−1 2100 cm−1. The wagging mode is relative to the total hydrogen concentration [36], this confirms that the components of the array of clusters as well as the distributed hydrogen are equal under this mode. We have tracked the aforementioned ways, the total hydrogen concentrations in the microstructure will inevitably increase with the level of hydrogenation gas. Accordingly, the CH values for both films are ≃ 16.18% (100 sccm) and ≃ 33.96% (200 sccm).

4.1 Bending band

The infrared spectra for the hydride setups of a-Si: H films are centered on the range of bending modes and show all sub-modes (Figures 2 and 3).

Figure 2.

Deconvolution of bending vibration modes in the range [4001000] cm−1 for two consecutive hydrogenation flow of 100 sccm (a) and 200 sccm (b).

Figure 3.

Deconvolution of vibration modes in the range [10001800] cm−1 for two consecutive hydrogenation flow of 100 sccm (a) and 200 sccm (b).

The assessment of the infrared tops at the level of the absorption most extreme and at the level of the half-width shows critical contrasts coming chiefly from the hydrogenation. The subsequent peaks arrangement is decreased in size and on number as the hydrogenation stream parts (100 sccm → 200 sccm). The peak showed up at the frequency 450 cm1 (Figure 2a) moves towards 500 cm1 (Figure 2b) with decrease of the most extreme absorption and of the half-width and shows a stretch mode relative to the vibrating SiSi bond.

It also shows a critical decrease of the deformities, as the dangling bonds and in contrast the prompt expansion in the number of SiH dipoles. The volume fraction involved by SiH dipole is lower than that of a homopolar SiSi dipole seen that the SiH bond is shorter than that of SiSi, which implies that the number of SiSi bonds establishes absorption field which widens its range by means of their void/aSi arrangements. The chains SiH2n have a lot of similar bonds, and the vibrations can be conjugated, which leads to infrared absorptions at characteristic frequencies linked to the vibrations of chemical clusters. The doublet appeared between 800 cm1 and 890 cm1 indicating the presence of vibrating SiH2n chains in scissors mode. Moreover, the doublet showed up between 950 cm1 and 1000 cm1 (Figure 2a) indicating a degenerate mode connected to the presence of isolated SiH2 hydrides in transverse vibration (Scissoring and Rocking). Consequently, it was reconstructed to make a single peak for a similar part of hydride in scissors mode vibration (Figure 2b). This decrease in the instability of the blended compositions is due to the buildup of the substance by hydrogen. In the frequency interval [600cm1, 700 cm1] (Figure 2), two convoluted peaks appear with the exception of the size and the absorption extremum which undergoes a slight variation due to the concentration of hydrogen bound. It appears that a degenerate out-of-plane longitudinal vibration of the wagging/twist type is established and persists for the SiH2 fragments [13]. We conclude that bound hydrogen reduces the degree of degeneration of vibrational dynamic states of hydrides. Conversely, in the frequency margin [900–1000cm1] (Figure 2a) two convoluted peaks reflecting the breaking of the chains SiH2n for turn out to be isolated (Figure 2b). These two degenerate modes combine to be in a single high extremum absorption mode reflecting the decrease in vibrational instability from the fragments of the polymeric hydrides. Therefore, increasing the hydrogenated configurations then reorganizes the microstructure by reducing the modes of vibrational degeneration due to the SiH2 units. In contrast, a structure with low hydrogen content removes the vibrational degeneration of the hydrogenated fragments, reflecting fractional levels of apparent mixed-phase energies that may act as an induced dipole.

Likewise, for a flow of 200 sccm, Figure 3b show an extreme reduction of IR peaks resulting from the formation of SiH dipoles (Figure 3a). Here again, the role of hydrogen seems to passivate the unordered aSi/Void arrangements by in turn increasing the characteristic gap (Eg ≃ 4 eV), which has already been demonstrated by simulating the data via the Tauc- Lorentz model (Figure 4).

Figure 4.

Tauc-Lorentz parameters as a function of tetrahedra concentration: A, C, E0 and Eg respectively.

The peaks (Figure 3a) appeared near 1208cm1, 1400 cm1 and 1800 cm1 showing the formation of SiO bonds by developing clusters SiOx in which the non-passivated part adheres to the vitreous substrate. These boundary configurations were further marked by the XRD spectrum resulting from oxidation at the aSi:H/glass interface following locally required areas. Hydrogen reduces absorption and widens the characteristic range, showing only a peak near 1250 cm1 refers to a rolling absorption center (Figure 3b). The peak centered in the vicinity of 1600 cm1 relates to a low hydrogenation Void/SiO2 composition, but when more bound hydrogen is added, such a composition has become bilateral and is less energetic. Hence, the identical fragments (OSiO,orOSiH2O) are in fact crossed and have not yet had a dynamic conservative mode [37]. Infrared absorption spectra relating to bending mode show the reduction of the microstructure of hydrogenated amorphous silicon via the increase in the hydrogen flow which pushes the multiplication of SiH andSiH2 bonds. This multiplication rearranges the tetrahedral conformations by replacing a silicon atom with a hydrogen atom, and thus an extreme reduction in the density of the dangling bonds. On the other hand, the absorption rate is more increasing when the structure is purely amorphous formed by silicon atoms. The gain of one hydrogen atom in the new microstructure lowers the absorption power. Figure 3b clearly shows the minimization of the number of absorption peaks when switching to a more enhanced hydrogenation flow. Previous work has shown that hydrogen can push the atomic grid back to a more ordered state through enhanced minimization of the built-in void and thus through the change in electrical properties caused by the entrainment of heteropolar bonds [38].

4.2 Stretching band

While the aggregate sum of hydrogen is regularly assessed from the integrated intensity of the wagging modes, the stretching modes contain more data about the microstructure and the voids fraction in a-Si: H. There is general agreement that there are at any rate two stretching modes. To resolve the densities of the hydrogenated SiH-type configurations and that of SiH2 -type in the bulk environment, the stretch band deconvolution measure is performed by two Gaussians focused separately at 2000cm1 and 2100cm1. Given the 7059 glass substrate used to deposit the film, it is possible to have connections between the mass of the film and the boundary particles of the glass substrate. Hence, to find out the intensity of these atomic configurations responsible for this interconnection that locally brings a new aspect to the film such as a nanocrystal or a microcrystal, it is necessary to decompose the spectral portion by Gaussians centered on all the Eigen frequencies, and thus calculate the required intensities. Accordingly, at the substrate interface, the limit bunches act like a little part of the film perceiving the monocrystalline or microcrystalline profile. There, this perspective can react with absorption modes reaching out around 1890cm1–1975 cm1 (ELSM). In the mid-frequency range located at 2030 cm12040 cm1 (MSM), the limit bunches basically causes absorption groups via platelet surfaces (chain shut by the succession of patterns:SiSiHn and most likely reflect restricting fragments in the substrate-film interface. What’s more, it ought to be noticed that there are different groups beginning in the frequency positions 2083 cm1 and 2137 cm1, and which exhibit the presence of the SiHx configurations (x = 1, 2 and 3). These groups so-called NHSM were set apart on little glasslike regions situated at the bulk material grains of the film [37, 38]. We consider that these groups are because of hydrogenated crystal surfaces, which are basically present in the less thick μc-Si: H with high crystallinity (see Figure 5). Besides, NHSMs modes show the presence of smooth surfaces containing accretions of mono-, di, and tri-hydride [39], which thus accentuate the edges of voids drawn with valuable patterns in the folds of the thin film (Table 1) [38]. This outcome is in concurrence with the investigation of the stretch modes in the frequency range 1900–2250 cm1 (Figure 6). It relates to HSM mode a hydrogen bonds to the inner surface of the microcavities [4].

Figure 5.

Growth of Si-Hx bond fragments until the formation of fully passivated spherical microcavities: Linear chain of low density in SiH and SiH2 (a), Atomic surface with SiH2 bonds at the outer limits (b) And a microvoid with internal walls of SiH2 bonds (c).

ELSM
∼1890–1970 cm−1
SiHn located near the interface substrate-film with mutual hydride dipole-dipole interactions: bond centered H SiSiHSiSi
LSM
∼2000 cm−1
MHs (SiH) restricted in the internal tissue of a-Si: H and form divacancies.
MSM
∼2030–2040 cm−1
Platelet surfaces: A chain shut by the succession of patterns
SiSiHn, located in the bulk mass.
HSM
∼2099±2 cm−1
DHs (SiH2) limited in the bulk mass of a-Si: H, unscreened configurations grafted into the internal walls of nanovoids.
NHSM
∼2083–2103–2137 cm−1
Micro-surfaces of SiHxx=123 with less dense hydrogenated configurations at high level of crystallinity, recognized in crystal grain boundaries surrounded by mass.

Table 1.

Infrared absorption modes reflecting all potential arrangements in hydrogenated amorphous silicon.

Figure 6.

Frequency deconvolution of the stretching mode for two hydrogenation flow, 100 sccm and 200 sccm, respectively.

To envisage the change of the density of the hydrogenated setups in the range of the stretching mode frequencies and in the case of higher hydrogen stream, it is certain that they change as per them dimensions and the light absorption rate (Figure 7).

Figure 7.

Evolution of the density of monohydride and dihydride configuration as a function of all Eigen frequency.

Nstrp=03pωpp+3E3

p is a dimensional measurement of the gathering vibrating at the frequencyωp. Eq. (2) was expected through the experimental curve (Figure 7) to attempt to clarify how the density of SiHx (x = 1, 2, 3) formations increments. We assessed that these advance because of the kind of coupling made in the microstructure. The components of the molded chains were entered into the equation, and this is referred to as the power p, whether it is open or shut.

In the case that they are open, the proven density is minor, and when these congregations are shut, the density is more contrasted with the situation where the chains are open. With respect to the related frequencyωp, it considers by its worth the sort of arrangements, i.e., SiH,SiH2 orSiH2n. As well, the dimension p addresses the normal length of SiH2 and SiH groups in the amorphous silicon lattice. For p = 2 (separately p = 3), p mirrors the typical region and the normal size involved by the restructured round Nano-Configurations [40].

This semi noticed depiction consolidates dimensional boundaries such as the changes in the vacancy size and the diameter of nanovoids; that advance the development of hydrogenated microstructures [41]. Nevertheless, absorption sub-modes mirroring the presence of hydrogenated bunches close to the limit substrate-film or in bulk mass; concede different frequencies not the same as those referred to for the stretching mode as a whole, chiefly LSM and HSM. The frequency shift ωSM restricted by the edge of a characteristic frequency ωp and a particularly unknown frequency ωp (ωp±ωp) related with such bunches yet to be noticed gives, among others, the bulk positions of monohydrides SiH.

It is likewise connected to the nanoenvironnement of SiH configurations in stretch mode. This frequency shift between such two stretching sub-modes is given by the following relation [38]:

ωp10424π2c2mpω0,pε0Npqp,e2E4

The frequency ω0,p assigns the p-eigen frequency; it is taken as the first limit of the frequency shift, mp is the mass of a SiH dipole in the hydrogenated packs dwelling in a micro-cavity of volume V, Np is the density of the dipoles (SiH or SiSi) in the screened phase at the p-mode, and qp,e is the effective charge of p-mode. In this way, for any shift that happens, it is in every case a lot of lower than the frequency ωp (i.e.ωpωp1). In this manner, we can develop Eq. (2) in arranged power series:

Nstr=p=03pωp+ωpp+3p=03pωpp+31+ωpωpp+3p=03pωpp+31+p+3ωpωpP=03pωpp+3+ppωpp+2ωp+3ωppωpp+2E5

If p traverses these values granted, then, the density of the hydrogenated phase (SiHx) in the stretching mode is given by the sum above showing a set of eleven terms, each of which contributes to a certain density of sequences of dipoles located at well-suited sites. The only difference between all stretch sub-modes is the area of ELSM (Extreme Low Stretching) modes around the following position 1943 cm−1, 1950 cm−1, 1975 cm−1 (Figure 6). Even so, it can be seen that the relationship between the two batches of ELSMs may result from a change in a single material composition. Hence, this is due to a very low fraction of hydrogen bonded. Additionally, these absorption modes show a net frequency shift which appeared also with respect to the modes (LSM and HSM) reflecting the presence of configurations of the SiH, SiH2 or in like manner tri-hydride (SiH3) type on crystalline surfaces credited to the limits of the vitreous grains in the mass. Such atomic configurations with hydrogen are relied upon to connect to extremely high hydride densities commonly responding through dipole-dipole interactions. Figure 6 shows also that a more noteworthy extent of structures having the SiH bond shows up with the LSM absorption mode. While the density way of the hydrogenated combinations goes through an upward sunken change focused on edges of 2100 cm−1, this exhibits that the SiH2 proportion is low contrasted with SiH. This outcome was seen through the upgraded combinations affirmed by the reproduction of the SE results (Figure 8).

Figure 8.

Hydrogen content versus the Eigen frequencies of the stretching mode.

The variation of hydrogen content as a function of the Eigen frequencies of the stretching mode follows a polynomic function. The error which follows the end of the adjustment is minimal, and the data regression ends with a good degree of fit (R2 = 0.90).

CH%imaiωjiE6

According to the order of each of the multiplying coefficients for the j-frequency, we note that they sometimes have an approximate weighting of the lengths of the dissimilar bonds, and at other times the extent of the spherical microcavities of surface S which assumes the size of the voids recreated at a concentration more prominent than 14%. We suppose the factor a1 as the approximate length of the chainsSiHn. Then, a2 is a surface and can also be a closed chain by weak extended SiHx fragments. After that, a3 is a volume, so we can envisage that there is a microvoid configured by adjusting a small offset of the surface of the chains SiHxn below the ω3-frequency. We will set up the reliance of the multiplier coefficients on the cross part of the microcavities and along these lines with its mean degree.

This outcome is in exceptionally legitimized concurrence with the reasonable mainstays of the ECMR model proposed by Drabold et al. [42]. Although the relation that considers the change in hydrogen concentration as part of the Eigen frequency of stretching mode, it is composed roughly as a component of the leveled surfaces with a given cross over degree:

CH%n=1Nsn2ωnE7

This empirical equation is proportional to the comparing frequency of the infrared vibration of SiHx-type composition (x = 1, 2). Then, an expansion in hydrogen concentration permits the distinctive complex gatherings to bend or shift in the permitted directions, trailed by a brief change in the angles between the diverse substituents of the first framed straight series. As per a few works did for over 40 years, this additionally implies that the overall concentration is controlled by the statistics of the infrared vibration frequencies of the trademark clusters of n-request surfaces that are framed by the hydrogenated groups that are presently arrangement. Subsequently, the level of hydrogenation is the operating factor liable for the arrangement of spherical nanovoids in the overall matrix. Beyond 14% in hydrogen content which does not surpass a restriction of 23%, the multiplication coefficient a3 is equivalent tos3/2. This limit is consistent when we have it in the state where the absolute bound hydrogen concentration accurately surpasses 14%. It identifies with the measure of the volume of void shaped by the hydrogen holding chains, their region, and the level of hydrogen in the HSM mode as follows [4]:

VvoidSvoid32CHSM32E8

The all setups considered show the process of advancement of the hydrogenated arrangements from a state where the density of SiH bonds is very low (Figure 5a), and until partial saturation by the formation of two-dimensional micro-chains mainly made up of slight zones (Figure 5b). The increment in hydrogen combined with silicon atoms enhances the creation of SiH2n chains which thusly start to squirm under the additional forces of electrodynamic states and eventually form vacuoles with embedded internal walls of SiH2 fragments (Figure 5c).

Advertisement

5. Ellipsometry measurements

An optical model is addressed as a function of the refractive index and the thickness of the thin film (N, d). Knowing the refractive index of the substrate and its thickness, those of the thin film can be determined by treatment the ellipsometric data (ψ,) by means of suitable adjustment model. Optical angles are interrelated by Snell’s laws as a light shaft passes starting with one layer then onto the next. A very important fact, in the event that the absorption’s film is low, this compares to a low refraction angle, which causes optical interference by a few reflections of light from the film’s surface. In particular, there is a superposition of the interferometric fringes in a destructive manner. In this case, the total intensity of the reflected light becomes smaller and smaller towards high photon energies because of the surface properties including roughness. This explains the fact of the wide oscillations in the spectrum of the optical phase shift Cos between the parts in the (p, s) plane of the light reflected (rp,s) from the surface. To maximize the amount of reflected light, the incidence angle θ0 should surpass the limited angle of refraction.

However, part of the incident light undergoes a surface scattering phenomenon depending on the roughness of the latter. Along these lines, the use of an incidence angle around 70°∼80° makes it possible to recover the minimum fractions of the reflected light absorbed by the surface and to avoid the light scattering by the effect of roughness [43, 44, 45, 46, 47]. The BEMA model shows more exact and nitty gritty outcomes than those of Maxwell Garnet (MG), Lorentz-Lorenz (LL), and provides the best fit in the analysis of surface roughness layers [48].

The countless materials compositions at last do not acknowledge the smallest change in their arrangement with one another (type or amount). The SE data never shows an arrangement such asSi3H/Si2H2, and the reproduction sequence never approved such configuration during the relapse of the conceivable proposed compositions, which shows that the film does not contain any setups framed by monohydrides and dihydrides chains stable.

These outcomes have been proven after 6,000 emphases. Film thicknesses were determined by SE measurement. Appropriately, for a hydrogenation gas stream FH equal to 100 sccm, the current film has a thickness around 213 nm; the second has 183 nm for FH equal to 200 sccm.

Brugemman’s approximation show nine phase materials, each with its own dielectric response and that make up the whole sample (see Figure 9). The dielectric response for every mixed phase was acquired through setback of SE measurements. So as to get and investigate its optical properties, we present our optical examinations dependent on the SE, just as the determination of dielectric function. In fact, hydrogen has numerous impacts in silicon matrix; such as, the remaking of frail SiSi bonds, the nearby densification of the amorphous network because of shorter length of the SiH bond in contrast with the SiSi bond, the bringing down of network cross-linking, since of an improvement of void growth, and an expansion of the density of states in the conduction band because of the antibonding states of the SiH bond. The void used as an explanatory factor of the optical properties of hydrogenated amorphous silicon has not been threw in the folds of the atomic network with proportions known by a particular procedure, but is rather present in its impressions at each corner of the film, it is a very important characteristic of the hydrogenated amorphous silicon, and we have shown the extent of its effect on the optoelectronics properties.

Figure 9.

Diagram ofϵ1, ϵ2 for all mixed phase in hydrogenated amorphous silicon: (a) and (b) relate respectively to the real and imaginary parts for a 100 sccm hydrogenation flow, (c) and (d) for the two dielectric parts respectively for a flow of 200 sccm.

5.1 Dielectric function

The dielectric spectra were defined by the system of Eq. (9) which is the theoretical foundation of the Bruggeman model [2]:

iviϵiϵϵi+2ϵ=0avecivi=1E9

For the various mixed compositions of each thin film, the diagrams of the real part ϵ1 and imaginary part ϵ2 of the dielectric function were calculated as functions of photon energy (Figure 9). The parameters ϵi and vi separately address the dielectric functions; the volume part of the nth component, and ϵ is the measured effective dielectric function of the film. Therefore, in the case of a random distribution of hydrogen bonds, each mixed composition has a volume fraction fm given by the following relative expressions:

fm=PmxVmmPmxVmϱ=mϱmvmE10

The mass density of the whole film ϱ is expressed as a function of the elementary densities ϱm of each tetrahedral composition. In essence, the volume that each material occupies relies on the number of bonds formed with the tetrahedral silicon, with the material compositions inside each having a sufficient opportunity to form based on the atomic percentage of hydrogen (SiH orSiSi). Every composition, thus, has a distinct density. The flow of hydrogen affects how many bonds (SiHx) there are. As a result, the optical index will change as a result of the qualities it has developed into in order to detect the impact caused by the disorder in the material and the different composition (V/Si4mHm).

The dominant proportion of hydrogenated clusters can be calibrated according to the ratio p1/2 = (NSiH/NSiH2). For the situation that the hydrogen concentration is less than 14%, implying that p1/2 is more than 16, the proportion of monohydride configurations dominates on the silicon network. Interestingly, the dihydrides dominate the folds of the microstructure, which thus create a low rate of stable divacancy as expected by numerous scientific works; this is the thing that we got through the regression of ellipsometric data.

The SE measurements provided very precise values that could be linked to some facts.

The variational shapes depend on the distribution of the nanovoids and on the extent of the change in its size as a function of the hydrogen content. Liu et al [49] show that the highest refractive index value indicates a very dense microstructure, which is the case of the following hydride composition: aSi/SiSi2H2 or aSi/SiSi3H. In the Near infrared, the refractive index diminishes with expanding dose. This is a contention and proof of hydrogen passivation of the sample. If the index increments in UV and diminishes in infrared shafts, around then, this mirrors that the substance is passivated by hydrogen [13]. The density of the different bonds and the change in the voids size are the first makes driving unraveling the optical spectra. Just as, the dielectric functions are acquired for various individual coordinated amorphous silicon, which ought to be available in the movies as SiSi4mHm joined by some fraction of voids. In the event that a layer is completely made out of hydrogenated amorphous silicon whatever his state of hybridization; at that point the level of control of the void is zero (the case of SiSi2H2 or SiSi3H). The curves identifying with the structure Void/a-Si have the equivalent nearly variety profile along the energy extend. However, the distinction at a noteworthy quality (most extreme and least) is clearly because of the extents of the arrangementsSiSi4. The maxima of variation as indicated by the energy of these configurations diminished in sufficiency when the level of the void increments. It has the most reduced amplitude (≃ 7.48 SI), and which relates to the most elevated level of void (≃ 52%). Oppositely, the least worth (≃ 7%) is related with a similar mixture having amplitude of the request for ≃ 22.104 SI.

The dielectric response of V/Si4mHm (m [0, 3[ ) compositions showed up at exceptionally high energy, unequivocally past 4.35 eV, they are gotten by supplanting a SiSi bond by SiH in the tetrahedral configuration SiSi4. By disregarding the compositions of aSi, it is seen that the profiles are requested then again regarding the presence of the setup having SiH2 and that conveying the SiH. This is additionally because of the strength of SiH bonds, which need more energy (4.55eV) compared with SiSi bonds (Figure 9b and d). The substitution of more grounded (SiSi) bonds diminished the maximum real and imaginary part of dielectric function. This could likewise be proof that the dispersion of hydrogenated clusters in the matrix of a-Si: H is exceptionally slight and far separated, isolated by silicon gatherings and various extents of the void [50]. The configuration SiSi2H2 deprived of void has the most elevated estimation of dielectric amplitude (23.26 SI 4.82 eV, Figure 9d) trailed by SiSi3H (18.95 SI 4.84 eV, Figure 9d). As a result, the higher part of void is related to lower hydrogen content.

5.2 Mass density

Film density was obtained using the classical Clausius-Mossotti relation. It describes the dealing between the refraction index and film density. It is given by the following formula when it is 100% amorphous [51]:

ϱ=3mSin21n2+14π2αSiSi+CH1CHαSiH12αSiSiE11

n is the refractive index for the state where the wavelength is exceptionally high, and CH is the total hydrogen concentration. Furthermore, similarly as the mass of the film should include the polarizability of each sort of bond present, here we have two kinds of bond: SiH and SiSi. The polarizability is equal to 1.36 10−24 cm3 (respectively 1.96 10−24 cm3).

In such manner, every material composition was viewed as a whole film stacked by bonds of the SiHx (x = 1, 2) type, containing a specific concentration of hydrogen [52]. The previously mentioned equation was reapplied to compute the mass density ρm (m = 1… 9) of every one of them along with of percent occupancy versus void scattered in the film microstructure. We likewise thought to be in this setting that when we have a tetrahedral composition which doesn’t contain SiH bond, we can’t hold the worth of the polarizability against zero. Instead, the equation must be taken into account counts as being equal to zero, as is the case with bonds of the SiSi type. From Figure 10, it can be seen that there is a slight decrement in certain values of mass density for some mixed phases compared to roughly in the film (100 sccm), then, these values are the ones that contain the highest percentage of hydrogen bonded to silicon, or rather, a huge number of SiH bonds. This fact shows that the completely passivated composition occupies the smallest parts of the waist in the folds of the film. Again, the film with the highest hydrogenation level was set up to be the most flexible and the most elastic (less stressful microstructure). Consequently, the inclusion of hydrogen in the silicon matrix causes a relative change in the a-Si cross-section.

Figure 10.

Mass densities ρm for each tetrahedral configuration as a function of the volume fraction fm.

It has as well been shown that it is potential to ascertain the mass density of each blended configuration using the Clausius-Mossotti equation, and since we are attempting in this work to conclusively decide the degree of changes that can occur in the atomic structure, according to its developmental real properties, it is expected that all mixed phase have around a similar mass, and as we demonstrate that the two film contain nine blended configurations (Figure 9), the absolute mass of the film will be the amount of every single sub-phase (see section 5.1 above). This critical prototypical assumes that in each case there is complete consistency of structural and discretionary properties.

5.3 Tauc-Lorentz Parameters

The counts were performed using Tauc-Lorentz model indicating the impact of void and hydrogen content on the bandgap of every hydride configurations. The gap of a-Si: H is changed by the disorder and with the hydrogen content, this implies that it is pretended by the all previously factors. According to Jellison et al. [53], the imaginary part ϵ2 of the dielectric function can be written as:

ϵ2ETL=1EAE0CEEg2E2E022+C2E2ԊEEgE12

TL model contain four constants are treated as fitting parameters, Ԋ is the Heaviside distribution, E0 is the energy of peak transition, C is the broadening term which is related to the disorder of the film and represents also a full width half maximum (FWHM) value of the normalized broadening Lorentz function, A is a factor related to the film density and proportional to the height ofϵ2.

All data supporting the relative changes of the band gap energy were computed (Figure 4). Apparently the increment in the progress of hydrogen gas has encircled the band gap to the expansion. In fact, the hydrogen clearly influences the band gap only in the case while the a-Si: H is unsaturated in hydrogen (the case of void/a-Si) [54]. Instead, the gap retains a constant mean value independent of the concentration, assuming that it is fully saturated (SiSi2H2). We once noticed that the gap increases when we have a completely hydrogen-free setup, again, if we had the opposite case.

The original reason why the gap differs in a monotonous manner could be the percentage of the void, that is, its size which affects the expansion of the internal range. As we have mentioned, an increase in the optical gap can result not only from an increase in the hydrogen concentration, but, also from the decrease of void ratio for each configuration [40].

Figure 4 shows this fact that the band gap reaches 4 eV when the tetrahedral configuration is fully passivated. Since the distortion of Si▬Si bonds in hydrogenated amorphous silicon is random, it results in the matrix of a-Si: H several possible values of the gap. Instead, a passivated bond removes a valence S-band in the material leaving valences formed primarily by deeper P-bands. This impact causes a critical broadening of the gap which can arrive at values more noteworthy than 3 eV. Moreover, it has been indisputably affirmed that the gap expansions with respect to the bound hydrogen content, that is, in entirely passivated tetrahedral configurations, as well, to the chance of bond among silicon and oxygen (=Si=O) invading from the substrate areas which improves this fact [55].

It can be concluded that the relevant hydrogen concentration has been changed, and this also brings a change in the thickness of the intended composition. Smets et al. [40] have shown that voids induced anisotropic volumetric compressive stress in the a-Si network leading to higher values of band gap energy. The disorder parameter (C) was not affected much by the rate of hydrogen more than the percentage of void in each composition.

All optoelectronic parameters evolve in a manner that varies according to the factor CtCt02 who Ct is the concentration of a specific tetrahedra composition and Ct0 is a threshold concentration. This fact has already been acknowledged by Likhachev et al. [56], and a similar variation was also noted for the refractive index. The differences seen regarding the Tauc-Lorentz parameters may have a more rationale accommodating the output of the three parameters that generate the overall variation ofϵ2.

Advertisement

6. Conclusion

As a function of the hydrogen and void concentration in each material composition, this chapter focuses on the dielectric and vibrational aspect of hydrogenated amorphous silicon. For each of the passivated configurations, a further relationship was established between the hydrogen content and the void proportion in order to explain how each configuration’s characteristics relate to vibrational frequencies. Additionally, in connection to the eigenfrequencies of the stretching mode, the relationship between the densities of the SiHxx=12 bonds and their spatial dimensions was taken into consideration and characterized. It has been found that, at the necessary synthesis circumstances, the completely passivated structures change precisely in terms of the degree of void, indicating a high gap that can reach multiple eV.

References

  1. 1. Konagai MJJ. Present status and future prospects of silicon thin-film solar cells. Japanese Journal of Applied Physics. 2011;50(3R):030001
  2. 2. Morral AF, Cabarrocas PR, Clerc CJP. Structure and hydrogen content of polymorphous silicon thin films studied by spectroscopic ellipsometry and nuclear measurements. Physical Review B. 2004;69(12):125307
  3. 3. Saha JK, Bahardoust B, Leong K, Gougam AB, Kherani NP, Zukotynski SJTSF. Spectroscopic ellipsometry studies on hydrogenated amorphous silicon thin films deposited using DC saddle field plasma enhanced chemical vapor deposition system. Thin Solid Films. 2011;519(9):2863-2866
  4. 4. Smets A, Kessels W, Van de Sanden MJA. Vacancies and voids in hydrogenated amorphous silicon. Applied physics letters. 2003;82(10):1547-1549
  5. 5. Stuckelberger M, Biron R, Wyrsch N, Haug F-J, Ballif CJR. Progress in solar cells from hydrogenated amorphous silicon. Renewable and Sustainable Energy Reviews. 2017;76:1497-1523
  6. 6. Joannopoulos JD, Lucovsky G. The physics of Hydrogenated Amorphous silicon II: Electronic and Vibrational properties. 2008
  7. 7. Robertson JJJ. Deposition mechanism of hydrogenated amorphous silicon. Journal of Applied Physics. 2000;87(5):2608-2617
  8. 8. Carlson DE, Wronski CRJAPL. Amorphous silicon solar cell. Applied Physics Letters. 1976;28(11):671-673
  9. 9. Yang J, Banerjee A, Guha SJAPL. Triple-junction amorphous silicon alloy solar cell with 14.6% initial and 13.0% stable conversion efficiencies. Applied Physics Letters. 1997;70(22):2975-2977
  10. 10. Shah A et al. Thin-film silicon solar cell technology. Progress in photovoltaics: Research and applications. 2004;12(2–3):113-142
  11. 11. Kołodziej AJO. Staebler-Wronski effect in amorphous silicon and its alloys. Opto-electronics review. 2004;12(1):21-32
  12. 12. Shimizu TJJ. Staebler-Wronski effect in hydrogenated amorphous silicon and related alloy films. Japanese journal of applied physics. 2004;43(6R):3257
  13. 13. Kassmi M, Samti R, Dimassi W, Amlouk MJJS. Optical properties of material compositions in hydrogenated amorphous silicon through the degree of passivation. Journal of Non-Crystalline Solids. 2021;562:120771
  14. 14. Paudel D, Atta-Fynn R, Drabold DA, Elliott SR, Biswas PJPRB. Small-angle X-ray scattering in amorphous silicon: A computational study. Physical Review B. 2018;97(18):184202
  15. 15. Haage T, Schmidt U, Fath H, Hess P, Schröder B, Oechsner HJJ. Density of glow discharge amorphous silicon films determined by spectroscopic ellipsometry. Journal of applied physics. 1994;76(8):4894-4896
  16. 16. Carlson DEJ. Amorphous silicon solar cells. Applied Physics Letters. 1977;24(4):449-453
  17. 17. Takeda T, Sano SJM. Amorphous silicon position sensor for telephone terminal. MRS Online Proceedings Library (OPL). 1988;118
  18. 18. Cody G, Tiedje T, Abeles B, Brooks B, Goldstein YJPRL. Disorder and the optical-absorption edge of hydrogenated amorphous silicon. Physical Review Letters. 1981;47(20):1480
  19. 19. Sokolov A, Shebanin A, Golikova O, Mezdrogina MJJ. Structural disorder and optical gap fluctuations in amorphous silicon. Journal of Physics: Condensed Matter. 1991;3(49):9887
  20. 20. Yamasaki SJPMB. Optical absorption edge of hydrogenated amorphous silicon studied by photoacoustic spectroscopy. Philosophical Magazine B. 1987;56(1):79-97
  21. 21. Daouahi M, Othmane AB, Zellama K, Zeinert A, Essamet M, Bouchriha HJS. Effect of the hydrogen bonding and content on the opto-electronic properties of radiofrequency magnetron sputtered hydrogenated amorphous silicon films. Solid state communications. 2001;120(5-6):243-248
  22. 22. Maley N, Lannin JJPRB. Influence of hydrogen on vibrational and optical properties of a-Si 1− x H x alloys. Physical Review B. 1987;36(2):1146
  23. 23. Gupta S, Katiyar R, Morell G, Weisz S, Balberg IJA. The effect of hydrogen on the network disorder in hydrogenated amorphous silicon. Applied physics letters. 1999;75(18):2803-2805
  24. 24. Bertran E, Andujar J, Canillas A, Roch C, Serra J, Sardin GJT. Effects of deposition temperature on properties of rf glow discharge amorphous silicon thin films. Thin solid films. 1991;205(2):140-145
  25. 25. Müllerová J, Fischer M, Netrvalová M, Zeman M, Šutta PJOP. Influence of deposition temperature on amorphous structure of PECVD deposited a-Si: H thin films. Central European Journal of Physics. 2011;9(5):1301-1308
  26. 26. Steffens J, Rinder J, Hahn G, Terheiden BJJ. Correlation between the optical bandgap and the monohydride bond density of hydrogenated amorphous silicon. Journal of Non-Crystalline Solids: X. 2020;5:100044
  27. 27. Sánchez P et al. Characterization of doped amorphous silicon thin films through the investigation of dopant elements by glow discharge spectrometry: A correlation of conductivity and bandgap energy measurements. International Journal of Molecular Sciences. 2011;12(4):2200-2215
  28. 28. de Lima Jr M, Freire F Jr, Marques FJB. Boron doping of hydrogenated amorphous silicon prepared by rf-co-sputtering. Brazilian Journal of Physics. 2002;32:379-382
  29. 29. Biswas P, Atta-Fynn R, Elliott SRJPRB. Metadynamical approach to the generation of amorphous structures: The case of a-Si: H. Physical Review B. 2016;93(18):184202
  30. 30. Johanson RE, Guenes M, Kasap SOJIPC. Devices, and Systems, 1/f noise in hydrogenated amorphous silicon–germanium alloys. Devices, and Systems, 1/f noise in hydrogenated amorphous silicon–germanium all. IEE Proceedings-Circuits, Devices and Systems. 2002;149(1):68-74
  31. 31. Mahan A, Williamson D, Nelson B, Crandall RJSC. Small-angle X-ray scattering studies of microvoids in a-SiC: H and a-Si: H. Solar Cells. 1989;27(1-4):465-476
  32. 32. Mui K, Smith FJPRB. Optical dielectric function of hydrogenated amorphous silicon: Tetrahedron model and experimental results. Physical Review B. 1988;38(15):10623
  33. 33. Mei J, Chen H, Shen W, Dekkers HJJ. Optical properties and local bonding configurations of hydrogenated amorphous silicon nitride thin films. Journal of applied physics. 2006;100(7):073516
  34. 34. Alvarez HDS, Silva AR, Cioldin FH, Espindola LC, Diniz JAJ. Hydrogenated amorphous silicon films deposited by electron cyclotron resonance chemical vapor deposition at room temperature with different radio frequency chuck powers. Thin Solid Films,. 2019;690:137534
  35. 35. Langford A, Fleet M, Nelson B, Lanford W, Maley NJPRB. Infrared absorption strength and hydrogen content of hydrogenated amorphous silicon. Physical Review B. 1992;45(23):13367
  36. 36. Ouwens JD, Schropp RJPRB. Hydrogen microstructure in hydrogenated amorphous silicon. Physical Review B. 1996;54(24):17759
  37. 37. Smets A, Matsui T, Kondo MJAPL. Infrared analysis of the bulk silicon-hydrogen bonds as an optimization tool for high-rate deposition of microcrystalline silicon solar cells. Applied Physics Letters. 2008;92(3):033506
  38. 38. Smets A, Van De Sanden MJP. Relation of the Si H stretching frequency to the nanostructural Si H bulk environment. Physical Review B. 2007;76(7):073202
  39. 39. Burrows VJAPL, Chabal YJ , Higashi GS, Raghavachari K, Christman SB. Applied Physics Letters. 1988;53:998
  40. 40. Smets AH et al. The relation between the bandgap and the anisotropic nature of hydrogenated amorphous silicon. IEEE Journal of Photovoltaics. 2012;2(2):94-98
  41. 41. Sekimoto T, Matsumoto M, Sagara A, Hishida M, Terakawa AJJ. Changes in the vacancy size distribution induced by non-bonded hydrogens in hydrogenated amorphous silicon. Journal of Non-Crystalline Solids. 2016;447:207-211
  42. 42. Biswas P, Atta-Fynn R, Drabold DAJPRB. Experimentally constrained molecular relaxation: The case of hydrogenated amorphous silicon. Physical Review B. 2007;76(12):125210
  43. 43. Archer R, Gobeli GJJ, Solids CO. Measurement of oxygen adsorption on silicon by ellipsometry. Journal of Physics and Chemistry of Solids.1965;26(2):343-351
  44. 44. Herzinger C, Johs B, McGahan W, Woollam JA, Paulson WJJ. Ellipsometric determination of optical constants for silicon and thermally grown silicon dioxide via a multi-sample, multi-wavelength, multi-angle investigation. Journal of Applied Physics. 1998;83(6):3323-3336
  45. 45. Wakagi M, Fujiwara H, Collins RJTSF. Real time spectroscopic ellipsometry for characterization of the crystallization of amorphous silicon by thermal annealing. Thin Solid Films. 1998;313:464-468
  46. 46. Fujiwara H, Koh J, Lee Y, Wronski C, Collins RJM. Real time spectroscopic ellipsometry studies of the solid phase crystallization of amorphous silicon. MRS Online Proceedings Library (OPL). 1998;507
  47. 47. Koh J, Fujiwara H, Lu Y, Wronski C, Collins RJT. Real time spectroscopic ellipsometry for characterization and optimization of amorphous silicon-based solar cell structures. Thin solid films. 1998;313:469-473
  48. 48. Fujiwara H, Koh J, Rovira P, Collins RJPRB. Assessment of effective-medium theories in the analysis of nucleation and microscopic surface roughness evolution for semiconductor thin films. Physical Review B. 2000;61(16):10832
  49. 49. Liu W et al. Characterization of microvoids in thin hydrogenated amorphous silicon layers by spectroscopic ellipsometry and Fourier transform infrared spectroscopy. Scripta Materialia. 2015;107:50-53
  50. 50. Timilsina R. Microstructure, vacancies and voids in hydrogenated amorphous silicon. Mississippi: The University of Southern; 2012
  51. 51. Remeš Z, Vaněček M, Torres P, Kroll U, Mahan A, Crandall RJJ. Optical determination of the mass density of amorphous and microcrystalline silicon layers with different hydrogen contents. Journal of non-crystalline solids. 1998;227:876-879
  52. 52. Shaik H, Sheik AS, Rachith S, Rao GMJ. Microstructure dependent opto-electronic properties of amorphous hydrogenated silicon thin films. Materials Today: Proceedings. 2018;5(1):2527-2533
  53. 53. Jellison G Jr, Modine FJAPL. Parameterization of the optical functions of amorphous materials in the interband region. Applied Physics Letters. 1996;69(3):371-373
  54. 54. Legesse M, Nolan M, Fagas GJT. Revisiting the dependence of the optical and mobility gaps of hydrogenated amorphous silicon on hydrogen concentration. The Journal of Physical Chemistry C. 2013;117(45):23956-23963
  55. 55. Hanyecz I, Budai J, Szilágyi E, Tóth ZJTSF. Characterization of pulsed laser deposited hydrogenated amorphous silicon films by spectroscopic ellipsometry. Thin Solid Films. 2011;519(9):2855-2858
  56. 56. Likhachev DV, Malkova N, Poslavsky LJTSF. Modified Tauc–Lorentz dispersion model leading to a more accurate representation of absorption features below the bandgap. Thin Solid Films. 2015;589:844-851

Written By

Mounir Kassmi

Submitted: 01 May 2022 Reviewed: 12 September 2022 Published: 04 January 2023