Vijay Kumar Gilla,
Sucheta Juneja*b,
Shiv Kumar Dixit*a,
Shruti Vashista and
Sushil Kumarb
aDepartment of Electronics and Communication Engineering, Manav Rachna University, Aravalli Hills, Faridabad, Haryana – 121004, India. E-mail: shivkumardixit.7@gmail.com
bCSIR Network of Institutes for Solar Energy, CSIR – National Physical Laboratory, Dr K. S. Krishnan Marg, New Delhi 110012, India. E-mail: suchetajuneja@gmail.com
First published on 30th July 2024
Initially hydrogenated silicon (Si:H) thin films have been deposited using a plasma-enhanced chemical vapor deposition technique (PECVD) using silane (SiH4) as a precursor gas diluted in an inert gas argon (Ar) environment. Subsequently phosphine gas (PH3) was used as the n-type dopant and the deposition was carried out at a fixed substrate temperature of 200 °C. The PH3 flow rate was varied in the range of 0–1 sccm. The effect of PH3 flow rates on optical, electrical, and structural properties of hydrogenated amorphous and micro/nanocrystalline silicon films has been investigated and detailed analysis is presented. These films may find application in heterojunction solar cells as an emitter layer. Further, a crystalline silicon (c-Si) based simple p–n junction solar cell is simulated using an SCAP-1D tool to observe the effect of layer thickness and doping density on solar cell parameters.
The essential requirement is to develop high quality n type nanocrystalline silicon layer (nc-Si:H) on p type wafer via efficient doping. The amount of doping in thin films controls the performance of the devices. The higher conductivity in n type silicon films can be obtained under controlled PH3 doping. But heavy doping also reduces crystalline nature in silicon network.11 Thus, at low temperature, high hydrogen dilution in the plasma is mostly used to achieve crystallinity. A combination of both hydrogen and noble gas can also be used to obtain crystallinity.
Scientists have used different techniques to obtain good quality n type nc-Si:H films under controlled doping such as plasma enhanced chemical vapor deposition (PECVD), microwave PECVD,12 electron beam evaporation,13 electron cyclotron resonance PECVD,14 Si ion implantation,15 hot wire CVD,16 cathodic vacuum arc,17 RF magnetron sputtering18 etc. Furthermore, in case of selection of wafer, n type possess lower defect density, low recombination rate of charge carriers at the interface and even high resistance against degradation10 than p type wafer. But the major concern is high production cost of n type ingots and hence less availability in the market. Thus, it is better to explore heterojunction solar cells with p type wafers.
The n-type hydrogenated micro/nanocrystalline silicon films can be achieved by precisely controlling the process parameters during deposition by PECVD. However, surplus doping atoms generates defects that hinder crystallization in the network and results in lower doping efficiency. Accordingly, optimization of parameters and integration of optimum dopants remains primary task to maintain crystallinity. There are only few reports available giving comprehensive investigation and influence of various deposition parameters on electrical, optical and structural properties with the correlation among other properties of PH3 doped silicon films developed by PECVD.19
In the present work PH3 doped n type nc-Si:H has been developed at low substrate temperature of 200 °C. The deposition of n-type nc-Si:H films with high conductivity of 0.18 S cm−1 has been carried out at very high frequency (VHF) PECVD process with appreciably high deposition rate. The PECVD process supplies low defect density and high doping efficiency in large area thin films at low temperature.20 We systematically investigate the influence of PH3 doping on the microstructural and optoelectronic properties of n-type nc-Si:H thin films grown by PECVD at low substrate temperature of 200 °C. The electrical as well as structural properties were found to correlate very well. The use of nc-Si:H films with high conductivity would replace a-Si:H films to overcome the Staebler–Wronski effect, improving diffusion length, more carrier life time and charge carrier mobility for high PCE solar cells.21,22 Structural morphology at different doping of PH3 has been studied to obtain suitable optoelectronic properties of nanocrystalline silicon network for solar cell.
Further numerical modelling of crystalline silicon (c-Si) based simple p–n junction solar cell is carried out using solar cell capacitance simulator (SCAP-1D) tool to observe the effect of layer thickness and doping density on solar cell parameters.
The electrical conductivity of samples measured at room temperature varies in the order of 10−5 S cm−1 (minimum) to 0.18 S cm−1 (maximum) for the n-doped films. Meanwhile, the activation energy decreased to 0.18 eV from maximum value of 0.7 eV for the PH3 doped film and hence higher conductivity can be achieved in Si films with doping. These n doped films may increase charge generation in solar cells and also PCE in amorphous and micro/nano crystalline silicon solar cells. These deposited n layer film can be useful as an emitter layer in crystalline silicon bulk heterojunction solar cells to achieve high PCE. For simulation purpose SCAP-1D simulation tool has been used under AM1.5 G (100 mW cm−2) solar spectrum with 300 Kelvin temperature. This SCAP software which is one dimensional can be used to solve basic semiconductor equations, boundary conditions, Poisson's equation etc., to study and simulate properties of photovoltaic devices.23,24
Yue Kuo25 demonstrated the effect of SiH4 (1% PH3) flow rates on phosphorus doped n+ silicon films and observed that the volume fraction of microcrystalline phase in the film decreases with the increase in flow rates along with shifting of peak in Raman spectrum. Debajyoti Das et al.10 reported the overall crystallinity of the n type nc-Si network reduced from 78.5% to 51% at higher PH3 flow rates.
These films prepared at PH3 flow rates of 0.5 and 0.75 sccm have a sharp peak at 520 cm−1, corresponding to the transverse optic (TO) phonon vibration, which is due to the presence of nano-crystalline phase.26 With an increase in the PH3 flow rate beyond 0.75, the intensity of TO peak is decreased and broader. This is due to reduction in the crystalline volume fraction as the PH3 flow rate increased. From these results, it is observed that the addition of PH3 beyond limit in nc-Si:H network causes nano-crystalline to amorphous transition in the films. At zero doping level (no doping), the film is mostly amorphous while with subsequent inclusion of PH3 in the deposition process enhances the overall crystallinity. There is a reduction in crystalline volume fraction beyond 0.75 sccm and this is due to lattice distortion and more disordered atomic arrangement. The higher PH3 doping also does a change in structural properties of the material. Therefore, optimized PH3 concentration is essential for making thin films electronic devices.
The variation in structural morphology with respect to doping of n type nc-Si:H films is examined by X-ray diffraction pattern, revealed in Fig. 2(a). The low peak intensity was observed at zero doping in the film which reveals that the deposited film is dominantly amorphous in nature. However, with doping at 0.5 sccm transition in film structure observed in Raman results are also correlated with the emergence of sharp crystalline peak at 28° along 〈1 1 1〉 diffraction plane in XRD. The crystalline nature in doped samples are observed by the appearance of characteristic peaks corresponding to different diffraction planes 〈1 1 1〉 and 〈2 2 0〉 at 2θ ∼28° and 47°respectively. These results signify the formation of nanocrystalline silicon structure embedded in the silicon matrix. However, it has been observed that the intensity of the 〈1 1 1〉 plane peak at 28° is much higher than the intensity of the 〈2 2 0〉 peak, which indicates the preferential crystalline growth along the 〈1 1 1〉 crystallographic orientation. Though, with further increase in doping beyond 0.75 sccm less intensity peak shows reduction in crystalline nature and shifting towards amorphous. The average crystallite size (D) along with the 〈1 1 1〉 orientation in the nc-Si:H film has been calculated from full width at half maximum (FWHM) of the XRD peak using Debye Scherrer's formula27 where λ is the wavelength of X-ray radiation, β is full width at half maximum (FWHM) of XRD peak at diffraction angle θ.
Fig. 2 (a) XRD spectra of PH3 doped n type nc-Si:H films deposited at different PH3 flow rates (b) variation in crystallite size with PH3 flow rates. |
The variation in average crystallite size (D) with respect to PH3 doping is shown in Fig. 2(b). The 〈1 1 1〉 oriented nanocrystallite size ‘D’ reduced from ∼25 nm to ∼15 nm on increasing PH3 flow rates from 0.75 to 1 sccm. However from 0.5 to 0.75 sccm no change in size was observed. But initially the size increases from 20 nm to 25 nm at 0.0 to 0.5 sccm respectively due to introduction of dopant atoms. The reduction in nanocrystallite size with increase in doping might be induced amorphization in the crystalline structure. An elevated doping leads to the formation of P–Si–H clusters and disordered amorphous region generating voids between the crystalline matrix.28
The plane view of field emission scanning electron microscope (FESEM) of n type nc-Si:H films is shown in Fig. 3(a–d). At elevated doping the distribution of crystalline grains and surrounding grain boundaries can be seen in Fig. 3(b and c). The smooth surface morphology of undoped and doped at 1 sccm in Fig. 3(a and d) respectively confirms the formation of reduced nature of crystallinity and noticeable reduction in grain size forming amorphous matrix. The small nano-sized crystallites are embedded in silicon matrix at doping of 0.5 and 0.75 sccm. The amorphous to crystalline transformation takes place at 0.5 sccm that is also confirmed with the help of Raman results. However, in Fig. 3(c) it is seen that grains are more or less of same size and distributed uniformly in the matrix.
Fig. 3 Top view FESEM images of PH3 doped n type nc-Si:H films grown at different PH3 flow rates (a) 0.0 (b) 0.5 (c) 0.75 and (d) 1 sccm. |
From the SEM analysis it is observed that grain size initially increases from 20 nm (at 0 sccm) to 35 nm (at 0.5 sccm) and then starts decreasing. Similar observations can be seen from the XRD results also. The variation in grain size (±5 nm) with respect to PH3 doping is shown in Fig. 4. The reduction in grain size with increase in doping from 0.75 to 1 sccm may be due to induced amorphization in the crystalline structure and formation of P–Si–H clusters and disordered amorphous region.28
The change in optical bandgap (Eg) for undoped (0 sccm) and doped at 0.5, 0.75 and 1 sccm of n type nc-Si:H films was determined using the well-known Tauc's relation (αhν)1/2 = B(hν − Eg), where B stands for Tauc's constant and α is absorption coefficient as a function of photon energy hν, see Fig. 5(a) and (b).29,30 The absorption coefficient α was calculated from UV-visible absorbance data (see Fig. S2†).
From these results it has been found that the Eg decreases from 1.9 eV to 1.71 eV, on increasing PH3 flow rates from 0 to 1 sccm. The change in energy gap may be due to localized defects and accumulation of donor impurity atoms below conduction band leading to narrowing energy band gap.10,31 The reduction in crystalline volume fraction due to rise in PH3 flow rates may be attributed to decrease in optical band gap.32
The change in deposition rate as a function of PH3 flow rates for n type nc-Si:H films is shown in Fig. 6. The deposition rate was calculated from the ratio of film thickness and deposition time. The film thickness was measured using a talystep surface profilometer. The deposition rate was ∼8.5 Å s−1 at ∼0.1 sccm PH3 flow rate. As the PH3 flow rate was further increased, the deposition rate decreased to minimum ∼2.2 Å s−1 for PH3 flow rate of 0.29 sccm. The increment in density of film not forming radicals and more hydrogen content may be the reason for initial decrease in deposition rate.
Fig. 6 Variation of deposition rate as a function of PH3 flow rates for n type nc-Si:H films deposited by PECVD technique. |
The maximum deposition rate ∼10 Å s−1 was achieved for films deposited at 0.5 sccm and beyond 0.5 sccm, the deposition rate decreased. The films were deposited using very high frequency (VHF > 50 MHz) at 20 W and hence the use of VHF results in increased density of precursors and high deposition rate. However beyond certain limit at high frequency the ion bombardment growing surface increases which results in reduction of crystalline growth. It is observed that deposition rate first decreases and then increases again with PH3 flow because with increase in pressure, the density of film forming precursor increases thus reducing mean free path for precursors in the plasma. For the application purpose, higher deposition rate is required with reduced processing time and cost.
The temperature dependent dark conductivity of doped and undoped film was measured from 300 to 477 Kelvin to elucidate electrical behavior. The conductivity of undoped film (at 0 sccm) increases on increasing temperature due to generation of more intrinsic charge carrier generation and hence leading to more intrinsic conductivity as shown in Fig. 7. But more variation in conductivity due to temperature of doped films is not observed from 300 to 477 Kelvin being temperature independent majority carriers. The dark conductivities are calculated from the relation
(1) |
Fig. 8 Variation in dark conductivity (σD) and charge carrier activation energy (Ea) as a function of the PH3 flow rates. |
As can be seen from Fig. 8, significant changes were observed in the dark conductivity (σD) and activation energy of films with different PH3 concentration. The mean value of σD and Ea with error bar was obtained from 5 different sample measurements. It has been observed that as the PH3 gas flow rate increases, the conductivity initially starts increasing and reaches a maximum value of 0.18 S cm−1 and then decreases to minimum 0.07 S cm−1. The initial increase in conductivity of nanocrystalline films is due to presence of large number of defects and impurities such as oxygen and nitrogen that are compensated by PH3 doping.33 This confirms that the PH3 doping compensates the defects and impurities grown in the material and therefore the film becomes n-type with improved conductivity. The room temperature σD was measured in the coplanar geometry by evaporating the Al contacts on the films using vacuum thermal evaporation. As shown in Fig. 8, for pure film, the dark conductivity and activation energy (Ea) is ∼10−5 S cm−1 and 700 meV respectively. When the PH3 flow rate is maintained at 0.2 sccm, the dark conductivity is increased to 0.072 S cm−1. With a further increase in the PH3 flow rates, the dark conductivity reaches a maximum value of 0.18 S cm−1 along with lowest Ea of 190 meV at 0.5 sccm. This may be due to shifting of Fermi level more close to conduction band edge resulting high conductivity and lowering activation energy.
Lucovsky et al.34 proposed a model to explain origin of reverse-Meyer–Neldel rule (MNR) phenomena which supports that the Fermi energy level moves in n type semiconductor above the bottom of conduction band in the crystalline phase and simultaneously it may shift more into the band tail states of the surrounding disordered amorphous regions. K. P. Chil et al.35 reported the highly degenerate microcrystalline phosphorus-doped silicon films grown by thermal LPCVD method and found the position of Fermi level inside the conduction band.
However, the conductivity is decreased further with increase in the PH3 flow rate beyond 0.75 sccm. The decrease in dark conductivity for higher PH3 concentration at 1 sccm may be due to reduction in crystallinity and more defect generation. Low angle XRD and Raman analysis also confirms the deterioration of crystallinity at 1 sccm as compared to 0.5 sccm. The deposition parameters used to develop the films are given in Table 1.
Process parameters | Value |
---|---|
Base pressure | 3.8 × 10−6–4.1 × 10−6 torr |
Deposition pressure | 0.25 torr |
Gas flow rate | |
SiH4 | 4 sccm |
PH3 | 0.0–1.0 sccm |
Very high frequency (VHF) power | 20 Watt |
Deposition time | 10 min |
Substrate temperature | 200 °C |
However for fabrication of solar cell, optimization for parameters is important. Crystalline silicon (c-Si) dominates the wafer-based solar cells while amorphous silicon (a-Si) plays a vital role in thin-film solar cells. In this theoretical work crystalline silicon p–n junction solar cell has been simulated using SCAP-1D simulation tool. The schematic of simulated structure of the solar cells is shown in Fig. 9.
The parameters required in the simulation such as initial thickness of n and p layer, charge carrier mobility, band gap, density of states etc. are taken from different published sources and input file given by SCAP software, as shown in Table 2.
Parameters | p layer | n layer |
---|---|---|
Initial thickness (nm) | 300000 | 100 |
Bandgap (eV) | 1.12 | 1.12 |
Electron affinity (eV) | 4.050 | 4.050 |
Dielectric permittivity (relative) | 11.90 | 11.90 |
CB effective DOS (cm−3) | 2.8 × 1019 | 2.8 × 1019 |
VB effective DOS (cm−3) | 1.04 × 1019 | 1.04 × 1019 |
Electron thermal velocity (cm s−1) | 2.3 × 107 | 2.3 × 107 |
Hole thermal velocity (cm s−1) | 1.65 × 107 | 1.65 × 107 |
Electron mobility (cm2 V−1 s−1) | 1.5 × 103 | 1.5 × 103 |
Hole mobility (cm2 V−1 s−1) | 4.5 × 102 | 4.5 × 102 |
ND (donor density) (cm−3) | 1010 | 1019 |
NA (acceptor density) (cm−3) | 1015 | 1010 |
Density of defects | 1013 | 1013 |
These results are also supported by external quantum efficiency (EQE) curves. For n type layer, at lower wavelength the % of EQE is better at 50 nm while it decreases on increasing thickness. At lower wavelength due to front surface recombination and thermalization losses the EQE decreases. On the other hand, for p type layer there is no effect of thickness on EQE and JV. The EQE results are shown in Fig. 11(a) and (b) for n type and p type respectively.
As the n type emitter layer thickness increases, the PCE decreases due to reduction in current density and dip in EQE. However, there is no major variation is observed in VOC but slight change may be due to front surface recombination. The effect of thickness on output parameters in n type layer are shown in Fig. 12(a–d). The effect of thickness variation in n type and p type layer on solar cell parameters is given in Tables 3 and 4 respectively.
Fig. 12 Impact of thickness variation in n-type layer on output parameters (a) PCE (b) VOC (c) current density (d) FF. |
Thickness (nm) | JSC (mA cm−2) | VOC (V) | FF% | PCE% |
---|---|---|---|---|
50 | 35.31481 | 0.54321 | 79.63348 | 15.27648 |
138 | 34.04482 | 0.54324 | 79.69935 | 14.73996 |
225 | 33.10142 | 0.54282 | 79.72825 | 14.32553 |
313 | 32.29507 | 0.54234 | 79.74795 | 13.96787 |
400 | 31.60949 | 0.5419 | 79.76215 | 13.66258 |
Thickness (μm) | JSC (mA cm−2) | VOC (V) | FF% | PCE% |
---|---|---|---|---|
250 | 34.31629 | 0.54266 | 80.04386 | 14.90581 |
313 | 34.5614 | 0.54346 | 79.5832 | 14.94792 |
375 | 34.67956 | 0.54379 | 79.06316 | 14.91019 |
438 | 34.74799 | 0.54394 | 78.48913 | 14.83517 |
500 | 34.78447 | 0.54402 | 77.86074 | 14.73382 |
The thickness of p type layer is varied from 250 to 500 μm but no major change in output parameters has been observed in this range. The results are shown in Fig. 13(a–d).
Fig. 13 Impact of thickness variation in p type layer on output parameters (a) PCE (b) VOC (c) current density (d) FF. |
Similarly, the effect of defect density variation in p type layer is shown in Fig. 15 in which all the output parameters decreases on increasing defect density. The defect density is varied from 1013 to 1016 cm−3 in p type layer. The effect of defect density variation in n type and p type layer on output parameters is given in Tables 5 and 6 respectively.
Defect density (cm−3) | JSC (mA cm−2) | VOC (V) | FF% | PCE% |
---|---|---|---|---|
1.00 × 1013 | 34.5275 | 0.54335 | 79.68022 | 14.94853 |
2.58 × 1013 | 32.5275 | 0.52335 | 71.68022 | 13.94853 |
5.05 × 1013 | 30.5275 | 0.51335 | 60.68022 | 12.94853 |
7.53 × 1013 | 28.5275 | 0.44335 | 58.68022 | 11.94853 |
1.00 × 1014 | 27.5275 | 0.39335 | 55.68022 | 10.94853 |
Defect density (cm−3) | JSC (mA cm−2) | VOC (V) | FF% | PCE% |
---|---|---|---|---|
1.00 × 1013 | 35.94573 | 0.55954 | 80.97943 | 16.28738 |
2.50 × 1015 | 23.34534 | 0.45893 | 73.54753 | 7.87988 |
5.00 × 1015 | 21.68631 | 0.44431 | 71.79455 | 6.9177 |
7.50 × 1015 | 20.73422 | 0.4347 | 70.26206 | 6.33288 |
1.00 × 1016 | 20.07179 | 0.42737 | 69.46169 | 5.95844 |
Doping (cm−3) | JSC (mA cm−2) | VOC (V) | FF% | PCE% |
---|---|---|---|---|
1.00 × 1016 | 35.20517 | 0.44602 | 74.44422 | 11.68944 |
2.50 × 1019 | 34.52734 | 0.54389 | 79.7035 | 14.9676 |
5.00 × 1019 | 34.52651 | 0.54407 | 79.71079 | 14.97359 |
7.50 × 1019 | 34.52497 | 0.54413 | 79.71319 | 14.97502 |
1.00 × 1020 | 34.52281 | 0.54416 | 79.7144 | 14.97511 |
Doping (cm−3) | JSC (mA cm−2) | VOC (V) | FF% | PCE% |
---|---|---|---|---|
1.00 × 1014 | 36.34676 | 0.47377 | 62.12877 | 10.69867 |
2.50 × 1019 | 18.29967 | 0.61468 | 82.78289 | 9.31176 |
5.00 × 1019 | 14.2573 | 0.60793 | 82.42996 | 7.14454 |
7.50 × 1019 | 12.0726 | 0.60373 | 82.4101 | 6.00649 |
1.00 × 1020 | 10.64165 | 0.60065 | 82.30128 | 5.26061 |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra02429j |
This journal is © The Royal Society of Chemistry 2024 |