Next Article in Journal
Integrating Solar Energy and Nature-Based Solutions for Climate-Neutral Urban Environments
Previous Article in Journal
Infiltration of CsPbI3:EuI2 Perovskites into TiO2 Spongy Layers Deposited by gig-lox Sputtering Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Laser Activation for Highly Boron-Doped Passivated Contacts

by
Saman Sharbaf Kalaghichi
1,2,*,
Jan Hoß
1,
Renate Zapf-Gottwick
2 and
Jürgen H. Werner
2
1
International Solar Energy Research Center Konstanz, Rudolf-Diesel-Straße 15, 78467 Konstanz, Germany
2
Institute for Photovoltaics, University of Stuttgart, Pfaffenwaldring 47, 70569 Stuttgart, Germany
*
Author to whom correspondence should be addressed.
Submission received: 9 June 2023 / Revised: 7 July 2023 / Accepted: 10 July 2023 / Published: 12 July 2023

Abstract

:
Passivated, selective contacts in silicon solar cells consist of a double layer of highly doped polycrystalline silicon (poly Si) and thin interfacial silicon dioxide (SiO2). This design concept allows for the highest efficiencies. Here, we report on a selective laser activation process, resulting in highly doped p++-type poly Si on top of the SiO2. In this double-layer structure, the p++-poly Si layer serves as a layer for transporting the generated holes from the bulk to a metal contact and, therefore, needs to be highly conductive for holes. High boron-doping of the poly Si layers is one approach to establish the desired high conductivity. In a laser activation step, a laser pulse melts the poly Si layer, and subsequent rapid cooling of the Si melt enables electrically active boron concentrations exceeding the solid solubility limit. In addition to the high conductivity, the high active boron concentration in the poly Si layer allows maskless patterning of p++-poly Si/SiO2 layers by providing an etch stop layer in the Si etchant solution, which results in a locally structured p++-poly Si/SiO2 after the etching process. The challenge in the laser activation technique is not to destroy the thin SiO2, which necessitates fine tuning of the laser process. In order to find the optimal processing window, we test laser pulse energy densities (Hp) in a broad range of 0.7 J/cm2Hp ≤ 5 J/cm2 on poly Si layers with two different thicknesses dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm. Finally, the processing window 2.8 J/cm2Hp ≤ 4 J/cm2 leads to the highest sheet conductance (Gsh) without destroying the SiO2 for both poly Si layer thicknesses. For both tested poly Si layers, the majority of the symmetric lifetime samples processed using these Hp achieve a good passivation quality with a high implied open circuit voltage (iVOC) and a low saturation current density (J0). The best sample achieves iVOC = 722 mV and J0 = 6.7 fA/cm2 per side. This low surface recombination current density, together with the accompanying measurements of the doping profiles, suggests that the SiO2 is not damaged during the laser process. We also observe that the passivation quality is independent of the tested poly Si layer thicknesses. The findings of this study show that laser-activated p++-poly Si/SiO2 are not only suitable for integration into advanced passivated contact solar cells, but also offer the possibility of maskless patterning of these stacks, substantially simplifying such solar cell production.

1. Introduction

Passivated emitter rear cells (PERC) are the current dominant technology in silicon-based photovoltaic (PV) production. The growing need for higher conversion efficiencies has encouraged the PV industry to seek alternatives to PERC design. Among the ideas for the mass production of highly efficient silicon solar cells, tunnel oxide passivated contact (TOPCon) solar cells are considered an upcoming replacement for PERC cells since only a few additional production steps are required to upgrade a PERC cell production line to a TOPCon production line [1]. The principle of the TOPCon design relies on separating the carrier extraction zone from the carrier generation zone by using a layer stack, which consists of a thin interfacial silicon dioxide layer and a highly doped polycrystalline silicon (SiO2/poly Si) layer [2]. The SiO2 passivates the mono-crystalline silicon (c-Si) surface and simultaneously separates the c-Si bulk carrier generation zone from the highly doped poly Si carrier transport layer [3]. Such layer stacks result in an excellent passivation quality together with low contact resistivity.
Conventional TOPCon cells employ a highly n-type-doped poly Si layer (n+-poly Si), which is usually doped in ex situ furnace diffusion [4] or deposited as an in situ doped poly Si layer followed by a furnace activation step [5,6]. Furnace-based processes allow for high throughput production of conventional TOPCon cells. However, these furnace processes have several inherent limitations:
  • Full area processing of poly Si layer: Advanced passivated contact cell concepts—such as poly Si passivated interdigitated back contact (IBC) solar cells or other cells with selective poly Si passivated contacts on the front side—rely on structuring of poly Si layers, which usually requires additional complex masking steps. Such structuring steps are expensive and time-consuming; therefore, they are not favorable industrially [7,8];
  • Doping concentrations within the poly Si layer: High doping concentrations are required within the poly Si in the poly Si/SiO2/c-Si structure to establish low resistivity of the metal/semiconductor contact [9]. However, the doping concentration in poly Si must drop sharply at the poly Si/SiO2 interface. Deep dopant diffusion, through the oxide into the underlying (single) crystalline silicon bulk substrate, would result in increased Auger recombination in the Si bulk [1]. The high required doping of the poly Si is particularly challenging, if one does not use n+-doping with phosphorus, but p+-doping with boron, because of the lower solid solubility of boron in silicon compared to phosphorus at the same temperature. The solid solubility of boron in Si is s m a x ≈ 5 × 1020 cm−3 at 1200 C, while phosphorus has a higher solubility of s m a x ≈ 1.5 × 1021 cm−3 at the same temperature [10]. Elevating the boron diffusion temperature to increase the doping concentration leads to damaged SiO2 and deep diffusion of boron into the under-laying bulk of the crystalline Si, and, consequently, to a lower passivation quality [11].
Laser irradiation—instead of furnace processing—avoids these limitations: It avoids structuring, heats the cells only locally, and, in addition, if performed properly, it also saves time. Consequently, lasers are attractive for solar cell production. For example, already, laser-doped selective emitters [12,13] and laser-doping of c-Si from precursor layers are two common production steps in the PV industry [14]. However, laser processing a poly Si/SiO2 stack is more challenging since it requires processing a thin poly Si layer without destroying the SiO2 underneath; the thickness of the tunneling oxide is only a few nanometers, while the laser-heated poly Si is more than a hundred times thicker.
This work presents a laser activation process for creating p++-poly Si/SiO2 stacks for passivated contact solar cell designs. Our laser process locally heats and melts a poly Si layer and the subsequent rapid cooling of the Si melt results in highly active boron concentrations exceeding the solid solubility limit within the poly Si layer. For our conditions, this high concentration drops sharply at the p++-poly Si/SiO2 interface, indicating that SiO2 is not damaged during the laser process. As a consequence, we observe an outstanding passivation quality with a high implied open circuit voltage (iVOC) and a low saturation current density (J0). The best sample shows iVOC = 722 mV and J0 = 6.7 fA/cm2.
The high boron doping concentration also enables the maskless patterning of laser-activated p++-poly Si/SiO2 stacks. Buchholz et al. [15] introduced a maskless process for establishing local p++-poly Si layers using the etch selectivity of a heavily boron-doped p++-poly Si in the alkaline etchant. A green laser was used in their method to create high doping concentrations locally, which resulted in structured p++-poly Si layers after etching. However, they did not show that this process could be used for highly doping the poly Si in a passivated contact solar cell design. Here, we show that this is possible. Consequently, in the next step, we will implement this approach to create local p++-poly Si/SiO2 stacks in advanced passivated contact solar cells.
Figure 1a shows the solar cell design with selective p++-poly Si/SiO2 stacks on the front side. To avoid parasitic optical losses by free-carrier absorption [16,17], the p++-poly Si/SiO2 passivation layers (indicated by a dashed circle) are only locally created underneath the front metal contacts.
Figure 1b illustrates the schematics of a poly Si passivated IBC solar cell. Such a solar cell design (with ion implantation and a masking process) achieved a remarkable conversion efficiency η = 26.1 % at laboratory scale [18]. Maskless, defect-free patterning of the p++-poly Si layers, as performed here, would substantially simplify the fabrication process and could open the door for the low-cost, large-scale, industrial production of this passivated contact solar cell design.
Figure 1. Schematic drawing of desired advanced passivated contact solar cells with local p++ poly Si/SiO2 stacks. (a) Selective p++ poly Si/SiO2 passivated front side cell. The full area n+-poly Si/SiO2 double layer passivates the rear side of the cell, while, at the front, p++-poly Si/SiO2 passivation stacks are selectively placed under the front metal contact to prevent parasitic optical losses. (b) Passivated contact, interdigitated back contact IBC solar cells. Back contacts of the cell are passivated by selectively created, laser-activated, p++-poly Si/SiO2 stack and n+-poly Si/SiO2 layers.
Figure 1. Schematic drawing of desired advanced passivated contact solar cells with local p++ poly Si/SiO2 stacks. (a) Selective p++ poly Si/SiO2 passivated front side cell. The full area n+-poly Si/SiO2 double layer passivates the rear side of the cell, while, at the front, p++-poly Si/SiO2 passivation stacks are selectively placed under the front metal contact to prevent parasitic optical losses. (b) Passivated contact, interdigitated back contact IBC solar cells. Back contacts of the cell are passivated by selectively created, laser-activated, p++-poly Si/SiO2 stack and n+-poly Si/SiO2 layers.
Solar 03 00021 g001
A laser process for dopant activation in hole-selective poly Si/SiO2 passivation stacks has also recently been introduced by Chen et al. [19]. Their main focus was the use of gallium as an alternative p-type dopant instead of boron to overcome the passivation loss arising from boron segregation at the interface between SiO2 and c-Si. The authors used an excimer laser for dopant activation on gallium-implanted poly Si layers and compared it with in situ boron-doped, plasma-enhanced chemical vapor deposition (PECVD) poly Si layers. Although they reported very high iVOC = 735 mV for gallium-doped poly Si/SiO2 layers, their boron-doped poly Si/SiO2 stacks showed only a mild passivation quality with iVOC = 700 mV.
In the present work, we show that achieving high passivation quality is possible even with boron as a dopant species for hole-selective poly Si/SiO2 layers without the need for ion implantation, which is considered a costly process step in the PV industry. We use a green laser for boron activation on low-pressure chemical vapor deposition (LPCVD) poly Si layers that results in high passivation quality with iVOC = 720 mV. Additionally, we consider here the role of interfacial SiO2 in the laser crystallization of poly Si layers.

2. Sample Preparation

In this work, we design two types of test structures to study the effect of laser processing parameters on the electrical characteristics and the passivation quality of the poly Si/SiO2 double layers. To investigate the influence of the interfacial oxide and poly Si layer thickness on these test structures, we use two poly Si layer thicknesses in our samples and also prepare groups of samples without interfacial oxide.
Our process starts with commercially available as-cut n-type M2 size Czochralski (Cz) silicon substrates from LONGi with a base resistivity of 27 Ω cm. After saw-damage etching and cleaning in a mixture of sulfuric acid (H2SO4) and hydrogen peroxide (H2O2), the wafers are split into two groups. One group is thermally oxidized in a tube process to create the SiO2 with a thickness of about 2 nm, while the other group is further processed without SiO2. Accordingly, an intrinsic hydrogen-free amorphous silicon (a-Si) thin film is deposited via LPCVD. Next, the intrinsic a-Si films are lightly doped in our BBr3 diffusion furnace at diffusion temperature T diff = 820 C. The a-Si layers are partially crystallized during this high temperature step. We extract the grain size dg ≈ 6 nm after the BBr3 diffusion step from the peak analysis of X-ray diffraction (XRD) measurement. Thereafter, borosilicate glass (BSG) created during the diffusion is removed in 2% hydrofluoric acid (HF). The subsequent laser process is performed on square-shaped fields with an area A laster field = 38 × 38 mm2, using a nanosecond laser tool operating at a wavelength λ = 532 nm, with a rectangular-shaped tophat laser beam profile and a spot area A spot = 300 × 600 μ m2. The XRD measurements show the grain size growth from dg ≈ 6 nm for non-lasered layers to dg ≈ 20 nm for laser-crystallized poly Si layers (L-poly Si). Two poly Si thicknesses are examined for both groups of samples, with and without a SiO2 layer.
The thickness of the poly Si films is measured with spectral ellipsometry on dedicated test wafers. The thicknesses are determined as dpoly Si,1 = 155 nm ± 5 nm and dpoly Si,2 = 264 nm ± 9 nm.
The amount of energy carried by each laser pulse is a crucial parameter for adjusting the later doping profile and, consequently, the electrical properties of the doped layers. The area normalized energy per pulse is called the laser pulse energy density (Hp). A detailed definition of Hp is given in ref. [20]. The Hp is varied by changing the laser duty cycle at the same pulse repetition rate f = 10 kHz with a pulse duration τ p ≈ 100 ns. A laser pulse overlap of 10% is used in both the y-direction (scanning direction) and the x-direction to ensure full area coverage and short laser processing time. Two kinds of test structures are fabricated:
  • Sheet resistance Rsh samples to characterize the electrical properties of the laser-processed samples, e.g., by four-point-probe measurement of the sheet resistance and by electrochemical capacitance/voltage (ECV) measurements of the doping profile.
  • Symmetric passivation samples for assessing the passivation quality with quasi-steady-state photoconductance QSSPC measurements [21].
Figure 2a shows a schematic cross-section of a fabricated Rsh sample, while a cross-section of a passivation sample is shown in Figure 2b. Although test structures without SiO2 are also fabricated, only samples with SiO2 are shown in Figure 2a,b for the sake of simplicity.
In order to imitate a complete solar cell fabrication process with diffusion and annealing steps, an additional annealing step at temperature T anneal = 985 C for a time t anneal = 30 min at ambient nitrogen is also applied to the passivation samples. Later, using PECVD, a silicon nitride (SiNx:H) layer (which serves as a hydrogen source for passivation) is deposited on both sides of the lifetime samples. At this stage, the passivation quality of the samples is measured by QSSPC and photoluminescence (PL) measurements.

3. Laser Activation

3.1. Sheet Conductance

Figure 3a,b show the measured sheet conductances (Gsh) for different Hp c-Si/SiO2/ poly Si and c-Si/poly Si structures. The Gsh-values are obtained by converting the measured Rsh-values to
G s h = 1 / R s h .
The Gsh-values of the samples with both the tested poly Si layer thicknesses can be divided into four zones based on the different dependence on Hp:
  • Zone I, background doping: All measured Gsh values are in the range of the background doping, which is created during the BBr3 diffusion. The background doping is measured from non-lasered areas on the wafer ( Hp = 0 J/cm2). Hence, no laser activation of dopants is observed in this zone.
  • Zone II, activation: For both groups of samples, with and without interfacial oxide, Gsh starts to increase when the laser pulse begins to melt the poly Si layer at the poly Si melting threshold Hp ≈ 1 J/cm2. For Hp ≥ 1 J/cm2, Gsh linearly increases with increasing Hp, which shows an increased number of electrically active boron atoms.
  • Zone III, saturation: The Gsh-values of the samples without SiO2 start to differ from the Gsh-values of samples with SiO2. The measured difference will be discussed in more detail in Section 5. However, both groups have in common that increasing Hp further does not significantly increase the Gsh.
  • Zone IV, SiO2-destruction: A sudden increase in Gsh of samples with SiO2 is observed in this zone, rendering it identical to the value measured for the samples without SiO2.
Figure 3a compares the measured Gsh of the samples with and without SiO2 for the structures with the poly Si layer thickness dpoly Si,1 = 155 nm. In zones I and II ( Hp ≤ 2 J/cm2), the same Gsh-values are measured for both sample types with and without SiO2, indicating that the presence or absence of the SiO2 does not affect the activation process. The Gsh starts to increase linearly from the meting threshold laser pulse energy density Hp ≈ 1 J/cm2 until Hp = 2 J/cm2. The Hp at the transition from zone II to zone III is called the transition laser pulse energy density (HT) in the following. The SiO2 affects the measured Gsh in zone III, where laser pulse energy densities above the HT are used. The Gsh of samples with SiO2 saturates around the average sheet conductance Gsh-ave. = 19.6 ± 3.3 mS sq, while higher Gsh-ave. = 65.6 ± 7.3 mS sq is measured for samples without SiO2. In zone IV, the highest Hp = 5 J/cm2 results in similar Gsh-values for samples with and without SiO2.
Figure 3b compares the measured Gsh of samples with a thicker poly Si layer with the thickness dpoly Si,2 = 264 nm in the presence and absence of SiO2. Similar to the samples with a thinner poly Si layer, in zones I and II, presence of SiO2 does not significantly influence the measured Gsh-values. The activation threshold Hp remains unchanged at Hp ≈ 1 J/cm2, as was observed for samples with a poly Si layer of dpoly Si,1 = 155 nm. However, the HT from zone II to zone III is shifted from Hp = 2 J/cm2 to Hp = 2.8 J/cm2 for the samples with a thicker poly Si layer. In zone IV, the same Gsh-values for both structures are measured. Similar behavior is also observed for samples with thinner poly Si layers.
Figure 3. Sheet conductances Gsh and measured sheet resistances Rsh of samples with and without SiO2 for poly Si thicknesses dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm. (a) For both types of samples, the activation process starts at Hp ≈ 1 J/cm2, where Gsh begins to increase linearly with increasing Hp. The presence of SiO2 starts to influence the Gsh when Hp exceeds the transition energy HT = 2 J/cm2. For 2.3 J/cm2Hp ≤ 4.6 J/cm2, the Gsh of samples without SiO2 saturates at higher values compared to samples including the SiO2. The Hp = 5 J/cm2 significantly increases the Gsh of samples with SiO2, and results in the same Gsh for samples with and without SiO2. (b) For dpoly Si,2 = 264 nm samples, the activation Hp remains unchanged. However, the HT increases to Hp = 2.8 J/cm2. Similar to the samples with a thinner poly Si layer, Hp = 5 J/cm2 results in the same Gsh-values for samples with and without SiO2.
Figure 3. Sheet conductances Gsh and measured sheet resistances Rsh of samples with and without SiO2 for poly Si thicknesses dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm. (a) For both types of samples, the activation process starts at Hp ≈ 1 J/cm2, where Gsh begins to increase linearly with increasing Hp. The presence of SiO2 starts to influence the Gsh when Hp exceeds the transition energy HT = 2 J/cm2. For 2.3 J/cm2Hp ≤ 4.6 J/cm2, the Gsh of samples without SiO2 saturates at higher values compared to samples including the SiO2. The Hp = 5 J/cm2 significantly increases the Gsh of samples with SiO2, and results in the same Gsh for samples with and without SiO2. (b) For dpoly Si,2 = 264 nm samples, the activation Hp remains unchanged. However, the HT increases to Hp = 2.8 J/cm2. Similar to the samples with a thinner poly Si layer, Hp = 5 J/cm2 results in the same Gsh-values for samples with and without SiO2.
Solar 03 00021 g003

3.2. Progressive Melting of Poly Si

Figure 4a shows the ECV profiles of poly Si/SiO2/c-Si samples processed with low to intermediate Hp for poly Si layers with thickness dpoly Si,1 = 155 nm. The Hp = 0 J/cm2 represents the background doping taken from the samples without laser processing. The lowest used laser pulse energy density Hp = 0.7 J/cm2 does not change the background doping. Melting of the poly Si layer starts when Hp reaches the melting threshold Hp = 1 J/cm2, where the concentration at the surface is marginally increased. This is compatible with the melting threshold found in Figure 3a,b. For Hp ≥ 1 J/cm2, higher concentrations are measured with respect to the background doping, and the depth of the part of the profiles with high doping concentrations progressively increases with increasing Hp. This behavior indicates the progressive melting of the poly Si layer. After absorbing a laser pulse with sufficient energy to melt the poly-Si, the poly Si layer begins to melt from the surface, and the melt front propagates into the depth of the poly Si layer. Increasing the Hp, increases the Si melt propagation, resulting in deeper ECV profiles with high concentrations.
Figure 4b shows the doping profiles of samples including SiO2 processed with high Hp. The flat profiles with constant doping concentrations across the whole poly Si layer thickness for 2.3 J/cm2Hp ≤ 4 J/cm2 show that the laser pulse fully melts the poly Si layer. The sharp drop in the concentrations approximately at the position of the SiO2 indicates that even this very thin passivating SiO2 stops the Si melt propagation. A deeper profile inside the c-Si substrate for Hp = 4.6 J/cm2 compared to the lower laser pulse energy densities implies that the effectiveness of the SiO2 as a diffusion barrier starts to degrade at this Hp. This shows that the SiO2 is increasingly damaged for Hp > 4 J/cm2, which eventually results in destroyed SiO2 for Hp = 5 J/cm2, represented by a reduced boron concentration within the first 150 nm of the profile and the deepest profile inside the substrate. The effect of damaged and destroyed SiO2 on passivation will be shown in Section 6.
Figure 4. (a) ECV profiles of the laser-processed samples for Hp ≤ 2 J/cm2 with a poly Si layer thickness dpoly Si,1 = 155 nm. Laser activation starts at a melting threshold laser pulse energy density Hp = 1 J/cm2 where a slightly higher surface concentration than the background doping is measured. For Hp beyond the melting threshold (Hp ≥ 1.3 J/cm2), increasing Hp increases the Si melt propagation, resulting in deeper ECV profiles with high concentrations. The depth of the ECV profiles with high concentration progressively increases by increasing the Hp. (b) High laser pulse energy densities Hp ≥ 2.3 J/cm2 melt the entire dpoly Si,1 = 155 nm thick layer and create flat ECV profiles. The SiO2 prevents the Si melt transfer to the c-Si bulk up to Hp = 4 J/cm2, creating a sharp drop in high concentrations at the position of the SiO2. Higher Hp = 4.6 J/cm2 damages the SiO2 and results in a deeper doping tail compared to lower Hp. Finally, Hp = 5 J/cm2 destroys the SiO2 and pushes boron atoms into the c-Si substrate, resulting in a reduced boron concentration within the first 150 nm of the profile.
Figure 4. (a) ECV profiles of the laser-processed samples for Hp ≤ 2 J/cm2 with a poly Si layer thickness dpoly Si,1 = 155 nm. Laser activation starts at a melting threshold laser pulse energy density Hp = 1 J/cm2 where a slightly higher surface concentration than the background doping is measured. For Hp beyond the melting threshold (Hp ≥ 1.3 J/cm2), increasing Hp increases the Si melt propagation, resulting in deeper ECV profiles with high concentrations. The depth of the ECV profiles with high concentration progressively increases by increasing the Hp. (b) High laser pulse energy densities Hp ≥ 2.3 J/cm2 melt the entire dpoly Si,1 = 155 nm thick layer and create flat ECV profiles. The SiO2 prevents the Si melt transfer to the c-Si bulk up to Hp = 4 J/cm2, creating a sharp drop in high concentrations at the position of the SiO2. Higher Hp = 4.6 J/cm2 damages the SiO2 and results in a deeper doping tail compared to lower Hp. Finally, Hp = 5 J/cm2 destroys the SiO2 and pushes boron atoms into the c-Si substrate, resulting in a reduced boron concentration within the first 150 nm of the profile.
Solar 03 00021 g004

3.2.1. Supersaturation

Figure 4a also shows the peak concentrations. For Hp ≥ 1.3 J/cm2, they exceed the solid solubility limit of boron in Si, which amounts to s m a x ≈ 5 × 1020 cm−3 at 1200 C [10]. This finding indicates that the laser-activated poly Si layers are supersaturated with boron atoms. Grain boundaries in poly Si provide an easy path for boron atoms to diffuse, and, therefore, allow for high dopant diffusivity [22]. We speculate that rapid boron migration along this path strongly enhances boron diffusion during the furnace BBr3 diffusion step in our experiment. However, not all boron atoms from BBr3 diffusion are electrically active. For example, boron atoms located at grain boundaries are not electrically active and, therefore, are not detectable with ECV measurements. The background doping shown in Figure 4a represents only the electrically active boron concentration from the BBr3 diffusion step. After melting the poly Si layer with a laser pulse, boron atoms diffuse into the molten Si due to a significantly higher diffusion coefficient D of boron in molten Si (Dl ≈ 10−4 cm2/s in liquid [23]) compared to the solid Si (Ds ≈10−11 cm2/s in solid at 1150 C [24]). Directly after the laser pulse, the molten Si recrystallizes rapidly, and dopant impurities are frozen into the Si lattice where they are electrically active. This results in an increase in electrically active boron concentrations beyond the conventional solid solubility limit [25,26].
A high active boron concentration of p++-poly Si has several advantages for the processing of solar cells:
(i)
It is essential for obtaining a low contact resistance at the metal/semiconductor contact interface [9]. It was shown earlier that just the conventional maximum solid solubility of boron in silicon is a limiting factor in contacting boron-doped p+-poly Si layers, which are fabricated with a conventional furnace-based process [11].
(ii)
A high active boron concentration provides an etch stop in potassium hydroxide (KOH) solution and, therefore, enables the maskless patterning of p++ poly layers, as described in ref. [15].
Unfortunately, however, supersaturation is not a stable condition, and a high temperature post-annealing step usually results in a reduction in the electrically active dopant concentrations to an equilibrium level [27,28,29]. We also detect a reduction in the active boron concentration in our lifetime samples, which undergo an additional high temperature step. However, if the selective etching is carried out before the high temperature step, the maskless patterning of the p++-poly Si layer is still possible. At present, we are investigating this effect in a separate study, which will be reported in a future publication.

3.2.2. Melt Depth Extraction

Figure 5a shows the progressive melting of a dpoly Si,2 = 264-nm-thick layer in poly Si/SiO2/c-Si samples. Similar to the thinner poly Si layers in Figure 4a, increasing Hp results in deeper ECV profiles with high concentrations. We use this behavior to determine the approximate melt depth (dmelt) created by different laser pulse energy densities Hp. Assuming that the melting of the poly Si layer increases the doping concentration, those parts of the ECV profiles with higher concentrations than the background doping are considered the molten parts of the poly Si layer. Accordingly, the transition depth, where the measured ECV profiles drop back to the non-melted level, is extracted as the approximate dmelt. A possible diffusion of dopants within the solid part of the layer is neglected here. The determined dmelt for each Hp is indicated by a vertical dashed line in Figure 5a.
Figure 5b illustrates the extracted melt depth for poly Si layers with dpoly Si,1 = 155 nm (from Figure 4a) and dpoly Si,2 = 264 nm (from Figure 5b). For both poly Si layers, the dmelt linearly increases with increasing Hp. This behavior for poly Si agrees with the earlier observed linear dependence of dmelt on the Hp in the laser processing of conventional bulk c-Si [20,30]. However, the melting rate found in ref. [20] for c-Si is about four times higher than the melting rate found here for poly Si. This is explained by us by the different thermal conductivities of poly Si and bulk c-Si structures. The thermal conductivity kpoly Si of a boron-doped poly Si layer at T = 300 K with a grain size dg ≈ 400 nm and a doping concentration n = 1.6 × 1019 cm−3 was extracted as kpoly Si = 45.6 W/mK in [31], while the thermal conductivity of undoped c-Si at the same temperature is kc-Si = 148 W/mK [32].
In the case of our poly Si, a similar dmelt is extracted for both layers for each Hp. This behavior is expected since the melting of the poly Si layer starts from the surface. The similar extracted melt depth allows us to apply a linear fit to the average dmelt and to calculate the melting rate. The linear fit in Figure 5b shows that melting of poly Si starts at Hp = 0.7 J/cm2 and increases linearly with a rate of 122 nm/(J/cm2). This extrapolation predicts that the melt front reaches the poly Si/SiO2 interface at Hp = 1.9 J/cm2 and Hp = 2.8 J/cm2 for dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm structures, respectively. These values agree with the Gsh data shown in Figure 3a,b. The same laser pulse energy densities are found there as the HT from zone II to zone III.
Figure 6 shows the dependency of Gsh on the melt depth in zone II. The Gsh increases linearly for both poly Si layers, indicating that the number of electrically active boron atoms also increases linearly with increasing dmelt. However, the slope of the linear fit for the dpoly Si,1 = 155 nm samples is higher compared to dpoly Si,2 = 264 nm. This difference is attributed to the total number of distributed boron atoms in these poly Si layers. We assume that the total number of boron atoms stored in these two poly Si layers is equal because of the same BBr3 diffusion step. Hence, the boron atoms are distributed more densely in the dpoly Si,1 = 155 nm poly Si layer compared to the dpoly Si,2 = 264 nm poly Si layer. Thus, the same melt depth results in a higher number of activated boron atoms for the dpoly Si,1 = 155 nm poly Si layer compared to the layer with the thickness dpoly Si,2 = 264 nm. The reason for the higher Gsh of dpoly Si,2 = 264 nm samples at the beginning of the melting zone in Figure 6 is not clear and requires more investigation.

3.3. Role of Interfacial Oxide in Electrical Properties

Figure 7 shows the doping profiles for the four different laser activation zones for samples with and without SiO2. In zone I (Hp = 0 J/cm2), no laser processing is performed, and, hence, the active boron concentration after the initial (furnace) boron diffusion is shown. The measured value agrees with the solubility limit of boron in Si at the BBr3 diffusion temperature T diff = 820 C [24]. The presence of the SiO2 creates only a minor difference between the ECV profiles of samples with and without SiO2 in this zone. Diffusion of boron into c-Si is more pronounced for samples without SiO2. In zone II, where the used Hp exceeds the threshold Hp of poly Si melting, the active boron concentrations above the background doping are measured at the surface of poly Si. However, the concentration decreases from the surface to the depth of the poly Si layer, implying a partial melting of the poly Si layers. The poly Si layer is only partially melted by Hp = 1.3 J/cm2, and the melt front is stopped before reaching the poly Si/SiO2 or the poly Si/c-Si interface. Thus, the presence or absence of SiO2 does not influence the doping of the poly Si layers in this zone. In zone III, for both sample types with and without SiO2, Hp = 3 J/cm2 results in a consistent and high doping concentration in the depth of the entire poly Si layer, which indicates melting of the entire poly Si layer. For samples with SiO2, the boron concentration is strongly reduced at the approximate position of the SiO2, showing that SiO2 blocks the Si melt front from propagating into the bulk c-Si wafer. A deep profile in the c-Si substrate is measured for samples without SiO2, which leads to higher measured Gsh, as shown in Figure 3a,b. In zone IV, similar ECV profiles for samples with and without SiO2 indicate that the Si melt front created by Hp = 5 J/cm2 destroys the SiO2 and propagates into the c-Si bulk wafer.

4. Charge Carrier Mobility in Laser-Activated Poly Si

Due to the presence of grain boundaries, the charge carrier mobility in poly Si is expected to be lower than that for monocrystalline silicon for the same doping concentrations [33]. The average mobility μ in our doped semiconductor layers with a diffusion profile is calculated by us from
μ = G s h q 0 x n ( x ) d x ,
where Gsh is the sheet conductance, q is the elementary charge, and n ( x ) is the depth-depending doping concentration. With the depth integrated average doping dose or sheet concentration N ¯ sh = 0 x n ( x ) d x , Equation (2a) becomes
μ = G s h q N ¯ s h .
We calculate the hole mobility in laser-activated poly Si layers with Equation (2b), using the Gsh-values from four-point-probe measurements and by integrating the ECV profiles over the depth, which provides N ¯ sh. Due to the different doping types of the substrate and laser-processed poly Si layers, the substrate does not contribute to the Rsh measurements for the sample structures used here (see Figure 2a).
Figure 8 shows the N ¯ sh versus the Hp for structures with SiO2. For both poly Si thicknesses, N ¯ sh increases for Hp ≥ 1 J/cm2 and saturates in zone III, which is similar to the behavior of Gsh in Figure 3a,b. Based on the constant values of N ¯ sh in the regime 2.3 J/cm2Hp ≤ 4.6 J/cm2, we use the average N ¯ sh value for the hole mobility calculation. The mentioned Hp range also results in a similar Gsh-values, as is shown in Figure 3a,b, which makes it possible to use the average Gsh-values for hole mobility calculation.
Table 1 shows the utilized data for the hole mobility calculation in poly Si and compares the calculated hole mobility in poly Si with the hole mobility in c-Si for a similar average boron concentration around n ¯ = 4 × 1020 cm−3. The average boron concentration n ¯ in poly Si layers is estimated by dividing N ¯ sh by the respective thickness of the poly Si layers through
n ¯ = N ¯ s h d poly Si .
The calculated hole mobility values are comparable for both poly Si layer thicknesses. Hence, we consider the average value of μ poly Si ≈ 15 cm2V−1s−1 as the hole mobility in this work, which compares well to that calculated in ref. [34] as μ ≈ 14 cm2V−1s−1 for boron-doped poly Si layers.
Figure 8. Extracted N ¯ sh for samples with SiO2 and poly Si layer thicknesses of dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm. For both poly Si layers, N ¯ sh saturates in zone III. The average N ¯ sh values in this zone are used for the hole mobility derivation.
Figure 8. Extracted N ¯ sh for samples with SiO2 and poly Si layer thicknesses of dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm. For both poly Si layers, N ¯ sh saturates in zone III. The average N ¯ sh values in this zone are used for the hole mobility derivation.
Solar 03 00021 g008
Table 1. Input data for hole mobility μ calculation and calculated hole mobility in poly Si for two different thicknesses.
Table 1. Input data for hole mobility μ calculation and calculated hole mobility in poly Si for two different thicknesses.
Structure n ¯ [cm−3] N ¯ sh  [cm−2]Gsh [mS sq] μ [cm2V−1s−1]
155 nm poly Si(4.8 ± 0.9) × 1020(7.6 ± 1) × 101519.6 ± 3.316.0± 2.6
264 nm poly Si(2.4 ± 0.4) × 1020(6.6 ± 1) × 1015 14.1 ± 1.313.1± 1.2
c-Si4 × 1020--40 [35]

5. Role of SiO2 in Poly Si Crystallization in the Saturation Zone

Figure 3a,b, in zone III, show significantly higher measured Gsh-values for samples without SiO2 compared to Gsh-values of samples with SiO2. In this section, we formulate two hypotheses to explain the measured difference in the Gsh-values of samples with and without SiO2. Accordingly, we calculate the expected Gsh from ECV profiles for each hypothesis and compare them with the measured Gsh-values. We perform the analysis for samples with a poly Si layer thickness dpoly Si,1 = 155 nm.

5.1. Hypothesis A

The higher Gsh-value in zone III for samples without SiO2 (see Hp = 3 J/cm2 in Figure 7a) stems from the extent of the doping tail into the bulk material. Therefore, the doping tail in the bulk leads to higher measured Gsh. For the test of this hypothesis A, we assume that the ECV profile is composed of two parts, (i) a highly doped laser-crystallized poly Si L-poly Si with a thickness dpoly Si,1 = 155 nm, and (ii) a sheet of highly doped mono-crystalline Si. Because of a different charge carrier mobility in L-poly Si and c-Si, the contribution of each structure in the total Gsh is calculated individually.

5.2. Hypothesis B

We ascribe the higher sheet conductance for samples without oxide to a seeded crystallization of the poly Si to become single crystalline Si with a higher mobility: In the absence of SiO2 for Hp > HT in zone III, the Si melt reaches the poly Si/c-Si interface and melts a portion of the c-Si substrate. In the crystallization process, the single-crystalline bulk silicon acts as a seed layer. Then, via liquid phase epitaxy, the molten poly Si regrows as a single crystalline layer. Due to the lack of grain boundaries in the solidified single crystal Si, the charge carrier mobility of the solidified Si layer approaches the carrier mobility of single crystalline Si and results in high Gsh-values.

5.2.1. Gsh-Data from Hypothesis A

Figure 9a visualizes the Gsh-calculation from the ECV profiles using hypothesis A. For estimating the contribution of the highly doped L-poly Si layer to the total Gsh, we use the derived hole mobility value in Section 4 ( μ poly Si = 15 cm2V−1s−1) within the thickness dpoly Si,1 = 155 nm of the poly Si layer. For calculating the share of a highly doped mono-crystalline Si sheet from the part of the ECV profiles beyond 155 nm, we use the PV lighthouse Rsh calculator tool, which calculates the corresponding Rsh of an ECV profile for a c-Si structure [36].

5.2.2. Gsh-Data from Hypothesis B

Figure 9b depicts the Gsh calculation utilizing hypothesis B. In hypothesis B, we assume the poly Si/c-Si structure recrystallizes to a c-Si/c-Si structure when the Hp is higher than the HT. Hence, to calculate the Gsh-values using hypothesis B, we consider the same hole mobility of c-Si for a recrystallized c-Si/c-Si structure, which enables us to also use the Rsh calculator tool of PV lighthouse.
Figure 9. Employed hypotheses in the Gsh calculation from the ECV profiles for samples without SiO2. (a) Hypothesis A considers the recrystallized structures as the combination of L-poly Si and c-Si. The Gsh in the L-poly Si part, indicated in cyan, is calculated for the first 155 nm of the doping profile from the surface using the determined hole mobility of μ poly Si ≈ 15 cm2V−1s−1. The contribution of the doped c-Si part, highlighted in magenta, is determined with the PV lighthouse Rsh calculator tool. (b) Hypothesis B assumes that, for Hp ≥ 2.3 J/cm2, the Si melt extends to the c-Si substrate seed layer and recrystallizes to a mono-crystal Si layer via liquid phase epitaxy. This results in a c-Si/c-Si structure with the same hole mobility as c-Si. The final Gsh(c-Si) is also determined via the PV lighthouse Rsh calculator tool.
Figure 9. Employed hypotheses in the Gsh calculation from the ECV profiles for samples without SiO2. (a) Hypothesis A considers the recrystallized structures as the combination of L-poly Si and c-Si. The Gsh in the L-poly Si part, indicated in cyan, is calculated for the first 155 nm of the doping profile from the surface using the determined hole mobility of μ poly Si ≈ 15 cm2V−1s−1. The contribution of the doped c-Si part, highlighted in magenta, is determined with the PV lighthouse Rsh calculator tool. (b) Hypothesis B assumes that, for Hp ≥ 2.3 J/cm2, the Si melt extends to the c-Si substrate seed layer and recrystallizes to a mono-crystal Si layer via liquid phase epitaxy. This results in a c-Si/c-Si structure with the same hole mobility as c-Si. The final Gsh(c-Si) is also determined via the PV lighthouse Rsh calculator tool.
Solar 03 00021 g009

5.2.3. Comparison of Hypotheses for Samples without SiO2

The comparison in Figure 10 shows that hypothesis B explains the measured data better. The figure compares the expected Gsh values considering hypothesis A with the measured Gsh data shown in Figure 3a. For Hp less than the HT (Hp < 2.3 J/cm2), where the Si melt does not propagate into the c-Si substrate, hypothesis A is in close agreement with the measured Gsh data. However, for Hp ≥ 2.3 J/cm2, where the Si melt extends to the underlying c-Si substrate, hypothesis A predicts lower Gsh-values than the measured data, indicating that the high Gsh-values in zone III for samples without SiO2 cannot solely be explained by the deeper doping profile. For Hp ≥ 2.3 J/cm2, the prediction by hypothesis B is in better agreement with the measured data than the prediction by hypothesis A. From this comparison, we conclude that the recrystallized structure is a c-Si/c-Si structure, as assumed by hypothesis B. Similar re-crystallization of the pre-amorphized poly Si layers to a single-crystal structure via liquid phase epitaxy during the laser annealing is also described elsewhere [28,37].

5.2.4. Comparison of Hypotheses for Samples with SiO2

To examine the influence of SiO2 in the crystallization process, we apply the same methodology as described above for samples with SiO2. Figure 11 shows a comparison of the measured Gsh-values and the values predicted by hypotheses A and B. In zones I, and II, estimated values by hypothesis A show better agreement with the measured values, as also found above. However, here, hypothesis A also fits better in zone III, in contrast to the result for structures without SiO2. This difference shows that the SiO2 prevents the Si melt propagation to the c-Si substrate. Hence, no lattice information from the substrate is available, and the Si melt recrystallizes to an L-poly Si structure.
In zone IV, for the highest Hp = 5 J/cm2, the suggested values from hypothesis B agree with the measured Gsh-values, indicating that the Si melt destroys the SiO2 and extends to the c-Si seed layer. Therefore, the Si melt recrystallizes into a single-crystal layer.
Figure 11. Comparison of the measured Gsh from layers including SiO2 with the expected Gsh from hypothesis A and hypothesis B. For Hp ≤ 4.6 J/cm2, where SiO2 acts as a barrier against the Si melt, the measured Gsh-values agree well with hypothesis A, suggesting that the Si melt is recrystallized to an L-poly Si structure. However, for Hp = 5 J/cm2, the measured value agrees with hypothesis B, which implies that the Si melt destroys the SiO2 and melts a part of a c-Si substrate. As a result, a mono-crystalline Si layer grows by liquid phase epitaxy.
Figure 11. Comparison of the measured Gsh from layers including SiO2 with the expected Gsh from hypothesis A and hypothesis B. For Hp ≤ 4.6 J/cm2, where SiO2 acts as a barrier against the Si melt, the measured Gsh-values agree well with hypothesis A, suggesting that the Si melt is recrystallized to an L-poly Si structure. However, for Hp = 5 J/cm2, the measured value agrees with hypothesis B, which implies that the Si melt destroys the SiO2 and melts a part of a c-Si substrate. As a result, a mono-crystalline Si layer grows by liquid phase epitaxy.
Solar 03 00021 g011
Figure 12 summarizes the findings and illustrates the role of SiO2 in a crystallization process. In the poly Si/SiO2/c-Si structures, the SiO2 separates a poly Si layer from a c-Si substrate and prevents the Si melt propagation to the c-Si bulk which acts as a seed layer. This results in a seedless crystallization of Si melt into an L-poly Si structure. However, for structures without SiO2 and for structures where the SiO2 is destroyed by excessive Hp, the Si melt reaches the c-Si substrate, and a mono-crystalline Si regrows by liquid phase epitaxy from the bottom c-Si substrate. The recrystallized single crystal structure has increased mobility and, hence, higher Gsh-values compared to structures where the Si melt recrystallizes to an L-poly Si structure.

6. Passivation

We quantify the passivation quality of the fabricated lifetime samples (see Figure 2b) in terms of the implied open circuit voltage (iVOC) and the saturation current density (J0), which are derived from the QSSPC measurements using a Sinton WCT-120 tool.
Figure 13a compares the measured iVOC of the samples with and without SiO2 for a poly Si layer with dpoly Si,1 = 155 nm. For Hp < 4.6 J/cm2, the majority of samples with SiO2 show a good passivation quality with the highest iVOC = 721 mV, which is comparable with the highest reported values for p+-poly Si/SiO2 passivation stacks [38,39]. The reason for the slight reduction in iVOC for 2 J/cm2 < Hp < 2.6 J/cm2 is not clear and requires further scrutiny. Nearly 100 mV higher iVOC is measured for most samples with SiO2 compared to those without the SiO2. Such a high difference in a measured iVOC is attributed to the double role of SiO2 in the passivation quality. The SiO2 in a poly Si/SiO2/c-Si structure chemically passivates the dangling bonds on the c-Si surface. It simultaneously limits the dopant diffusion from the poly Si layer to the c-Si bulk and creates a large doping gradient at the interface, which leads to increased field-effect passivation and reduced Auger recombination. For Hp = 4.6 J/cm2, the measured iVOC of the samples with SiO2 is reduced from iVOC =720 mV to iVOC = 661 mV, which hints at damaged SiO2. Damage of the SiO2 is also suggested by the increased depth of the doping profile for Hp = 4.6 J/cm2 with respect to the lower laser pulse energy densities in Figure 4b. Increasing the used Hp to Hp = 5 J/cm2 leads to further degradation of the passivation, resulting in a low iVOC =624 mV. This value matches the iVOC values measured for samples without SiO2, which indicates a destroyed SiO2. This finding is also supported by the ECV profile in Figure 4b. The ECV profile of Hp = 5 J/cm2 implies that the Si melt destroyed the SiO2 and extended to the c-Si bulk.
Figure 13b shows the measured iVOC of samples with and without SiO2 with a poly Si layer thickness dpoly Si,2 = 264 nm. The highest measured iVOC = 722 mV is comparable with the maximum measured iVOC of samples with thinner poly Si layers, implying that the passivation quality does not significantly depend on the tested poly Si layer thickness. This behavior is expected and also supported by other studies of poly Si/SiO2 passivation layers [34,40,41]. Similar to the samples with a thinner poly Si layer, in samples with a poly Si layer of dpoly Si,2 = 264 nm thickness, the presence of the SiO2 leads to a higher iVOC compared to samples without SiO2 for Hp < 4.6 J/cm2. However, the iVOC loss for Hp = 4.6 J/cm2 is less pronounced in samples with increased poly Si layer thickness compared to the samples with thinner poly Si layers. This is plausible since the increased poly Si layer reduces the damage from the Si melt to the SiO2. Nevertheless, the same measured iVOC of samples with and without SiO2 shows that increasing the poly Si layer thickness from dpoly Si,1 = 155 nm to dpoly Si,2 = 264 nm does not prevent the SiO2 destruction and the passivation loss for Hp = 5 J/cm2.
Figure 14a,b shows the extracted saturation current densities J0 of symmetrical passivation samples with SiO2 from the effective lifetime ( τ eff ) measurement. Due to the symmetry of the lifetime samples, we show J0/2, representing a saturation current density per side. The J0 values follow the trend of iVOC shown in Figure 13a,b. For the Hp range, that does not damage the interfacial oxide (Hp < 4.6 J/cm2) samples with both poly Si layers dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm providing low J0, with the best sample achieving J0 = 6.7 fA/cm2. The Hp = 4.6 J/cm2 damages the SiO2 in samples with a thinner poly Si layer and sharply increases the saturation current density to J0 = 112 fA/cm2. In contrast, the increased poly Si thickness in the dpoly Si,2 = 264 nm structure reduces the damage caused by Hp = 4.6 J/cm2 and provides only a moderately degraded passivation quality with J0 = 17 fA/cm2. The highest used Hp = 5 J/cm2 destroys the SiO2 for both poly Si thicknesses and results in J0 = 501 fA/cm2 and J0 = 482 fA/cm2 for dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm, respectively.

7. Summary

This work has shown the feasibility of a laser activation process for p++-poly Si/SiO2 passivation stacks. We tested various Hp in a wide range of 0.7 J/cm2Hp ≤ 5 J/cm2 on samples with two poly Si layer thicknesses of dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm. The poly Si layers already contain boron due to previous boron furnace diffusion. For both poly Si layers, the activation process begins when the laser pulse starts to melt the poly Si layer at the melting threshold laser pulse energy density Hp ≈ 1 J/cm2. For Hp ≥ 1 J/cm2, the melt depth of Si and, consequently, the dopant activation, linearly increases with increasing Hp. The activation process is completed when the Hp is high enough to melt the entire poly Si layer. Laser pulse energy densities Hp = 2 J/cm2 and Hp = 2.9 J/cm2 fully melt the dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm layers, respectively. The activation process becomes saturated for higher Hp up to Hp = 4 J/cm2. Further increase in the Hp results in damaged or destroyed SiO2 by the Si melt. The amount of damage induced by the high Hp depends on the poly Si layer thicknesses. The Hp = 4.6 J/cm2 severely damages the dpoly Si,1 = 155 nm poly Si layer, while the damage is less pronounced for the samples with thicker poly Si layers. However, for both poly Si layers, the highest used Hp = 5 J/cm2 destroys the SiO2.
Further, we have investigated the recrystallization process after melting the poly Si layers with the laser process. The SiO2 inhibits the transfer of the Si melt to the c-Si bulk up to Hp = 4.6 J/cm2, resulting in seedless crystallization of the Si melt into a laser-crystallized L-poly Si structure. In the absence of SiO2, and for structures where excessive Hp destroys the SiO2, mono-crystalline Si grows by liquid phase epitaxy from the bottom c-Si bulk, which acts as a seed layer.
Additionally, the passivation study on the symmetrical lifetime samples shows that the passivation quality is independent of the thickness of the tested poly Si layers. The p++-poly Si/SiO2 structures with both poly Si layer thicknesses show a good passivation quality, with the best sample achieving iVOC = 722 mV and J0 = 6.7 fA/cm2. Only samples processed with high Hp = 4.6 J/cm2 show a degraded passivation quality due to the damaged SiO2, especially for the samples with thinner poly Si layers where the damage to SiO2 is more significant.

8. Conclusions

By choosing the appropriate laser pulse energy density, a green laser can be used to locally melt a poly Si layer in a poly Si/SiO2 passivation stack, resulting in a high active boron concentration around n ≈ 5 × 1020 cm−3 within the poly Si layers. After a high-temperature annealing step and capping with SiNx:H, these laser-activated stacks achieve very high passivation quality with iVOC = 722 mV and J0 = 6.7 fA/cm2. The achieved passivation quality shows that the interfacial SiO2 layer is not destroyed during the laser process.
This work shows that high active boron concentrations is achieved in the laser-activated parts of poly Si layers. As described in ref. [15], high active boron concentrations in poly Si create the etch stop layer and result in the local p++-poly Si/SiO2 stacks after the etching process in an alkaline solution. In the future, we will use this maskless patterning method to create local p++-poly Si/SiO2 layers for the advanced passivated contact solar cells with p++-poly Si/SiO2 passivation stacks on the front side. In this cell structure, the p++-poly Si/SiO2 passivation stacks should be locally created under the front metal contacts to avoid parasitic absorption from the heavily doped poly Si layer.
One potential challenge is the observed reduction in high active boron concentration upon the high-temperature post-annealing step. Since solar cell production includes several high-temperature steps, like thermal diffusion, subsequent research in this regard is needed to maintain the high active boron concentration throughout the whole processing sequence.

Author Contributions

Conceptualization and experiment design, S.S.K., J.H. and J.H.W.; sample preparation, S.S.K.; measurements S.S.K.; modeling, S.S.K.; writing, S.S.K.; review and editing, J.H., R.Z.-G. and J.H.W.; supervision, J.H. and J.H.W.; data analysis and visualization, S.S.K.; project administration, J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Bundesministerium für Wirtschaft und Klimaschutz (BMWK), funding reference: 03EE1138A.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not available.

Acknowledgments

This work was supported by the German Federal Ministry for Economic Affairs and Climate Action under contract number 03EE1138A (SelFi). The corresponding author also thanks the DAAD organization for the Ph.D. scholarship. A special thanks to Radovan Kopecek, managing director of ISC Konstanz, for organizing the corresponding author’s transfer to ISC Konstanz. We would like to thank all colleagues at ISC Konstanz who helped with the sample preparation and processing, Mahdi Malekshahi from ipv Stuttgart for the XRD measurements, and acknowledge the support of Michael Saliba the director of ipv Stuttgart. We also thank Jan Lossen from ISC Konstanz for the valuable insights.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kafle, B.; Goraya, B.S.; Mack, S.; Feldmann, F.; Nold, S.; Rentsch, J. TOPCon—Technology options for cost efficient industrial manufacturing. Sol. Energy Mater. Sol. Cells 2021, 227, 111100. [Google Scholar] [CrossRef]
  2. Glunz, S.; Steinhauser, B.; Polzin, J.I.; Luderer, C.; Grübel, B.; Niewelt, T.; Okasha, A.; Bories, M.; Nagel, H.; Krieg, K.; et al. Silicon-based passivating contacts: The TOPCon route. Prog. Photovoltaics Res. Appl. 2021, 31, 341–359. [Google Scholar] [CrossRef]
  3. Glunz, S.W.; Feldmann, F. SiO2 surface passivation layers—A key technology for silicon solar cells. Sol. Energy Mater. Sol. Cells 2018, 185, 260–269. [Google Scholar] [CrossRef]
  4. Chen, Y.; Chen, D.; Liu, C.; Wang, Z.; Zou, Y.; He, Y.; Wang, Y.; Yuan, L.; Gong, J.; Lin, W.; et al. Mass production of industrial tunnel oxide passivated contacts (i-TOPCon) silicon solar cells with average efficiency over 23% and modules over 345 W. Prog. Photovoltaics Res. Appl. 2019, 27, 827–834. [Google Scholar] [CrossRef]
  5. Nandakumar, N.; Rodriguez, J.; Kluge, T.; Groβe, T.; Fondop, L.; Padhamnath, P.; Balaji, N.; König, M.; Duttagupta, S. Approaching 23% with large-area monoPoly cells using screen-printed and fired rear passivating contacts fabricated by inline PECVD. Prog. Photovoltaics Res. Appl. 2019, 27, 107–112. [Google Scholar] [CrossRef]
  6. Chaudhary, A.; Hoß, J.; Lossen, J.; Huster, F.; Kopecek, R.; van Swaaij, R.; Zeman, M. Screen Printed Fire-Through Contact Formation for Polysilicon-Passivated Contacts and Phosphorus-Diffused Contacts. IEEE J. Photovoltaics 2022, 12, 462–468. [Google Scholar] [CrossRef]
  7. Yang, G.; Guo, P.; Procel, P.; Limodio, G.; Weeber, A.; Isabella, O.; Zeman, M. High-efficiency black IBC c-Si solar cells with poly-Si as carrier-selective passivating contacts. Sol. Energy Mater. Sol. Cells 2018, 186. [Google Scholar] [CrossRef]
  8. Yuan, S.; Cui, Y.; Zhuang, Y.; Chen, P.; Hu, Y.; Yang, B.; Yu, Y.; Ren, Y.; Wang, W.; Chen, W.; et al. PERC Silicon Solar Cells with Front Polysilicon Passivating Contacted Selective Emitter. Phys. Status Solidi (RRL) Rapid Res. Lett. 2021, 15, 2100057. [Google Scholar] [CrossRef]
  9. Sze, S.; Lee, M. Semiconductor Devices: Physics and Technology. In Semiconductor Devices, Physics and Technology; Wiley: Hoboken, NJ, USA, 2012; p. 239. [Google Scholar]
  10. Trumbore, F.A. Solid solubilities of impurity elements in germanium and silicon. Bell Syst. Tech. J. 1960, 39, 205–233. [Google Scholar] [CrossRef]
  11. Padhamnath, P.; Nampalli, N.; Khanna, A.; Nagarajan, B.; Aberle, A.G.; Duttagupta, S. Progress with passivation and screen-printed metallization of Boron-doped monoPoly™ layers. Solar Energy 2022, 231, 8–26. [Google Scholar] [CrossRef]
  12. Carlsson, C.; Esturo-Bretón, A.; Ametowobla, M.; Köhler, J.; Werner, J. Laser doping for selective silicon solar cell emitter. In Proceedings of the 21st European Photovoltaic Solar Energy Conference, Dresden, Germany, 4–8 September 2006; pp. 938–940. [Google Scholar]
  13. Zhang, C.; Shen, H.; Sun, L.; Yang, J.; Wu, S.; Lu, Z. Bifacial p-Type PERC Solar Cell with Efficiency over 22% Using Laser Doped Selective Emitter. Energies 2020, 13, 1388. [Google Scholar] [CrossRef]
  14. Zapf-Gottwick, R.; Seren, S.; Fernandez-Robledo, S.; Wete, E.P.; Schiliro, M.; Hassan, M.; Mihailetchi, V.; Buck, T.; Kopecek, R.; Köhler, J.; et al. Solar Cells with Laser Doped Boron Layers from Atmospheric Pressure Chemical Vapor Deposition. Solar 2022, 2, 274–292. [Google Scholar] [CrossRef]
  15. Buchholz, F.; Linke, J.; Hoß, J.; Chu, H.; Mihailetchi, V.; Chaudhary, A.; Arumughan, J.; Lossen, J.; Kopecek, R.; Wefringhaus, E. Local Passivating Contacts from Laser Doped P+ Polysilicon. In Proceedings of the 38th European Photovoltaic Solar Energy Conference and Exhibition, Online, 6–10 September 2021; pp. 140–143. [Google Scholar] [CrossRef]
  16. Feldmann, F.; Nicolai, M.; Müller, R.; Reichel, C.; Hermle, M. Optical and electrical characterization of poly-Si/SiOx contacts and their implications on solar cell design. Energy Procedia 2017, 124, 31–37. [Google Scholar] [CrossRef]
  17. Feldmann, F.; Reichel, C.; Müller, R.; Hermle, M. The application of poly-Si/SiOx contacts as passivated top/rear contacts in Si solar cells. Sol. Energy Mater. Sol. Cells 2017, 159, 265–271. [Google Scholar] [CrossRef]
  18. Haase, F.; Hollemann, C.; Schäfer, S.; Merkle, A.; Rienäcker, M.; Krügener, J.; Brendel, R.; Peibst, R. Laser contact openings for local poly-Si-metal contacts enabling 26. 1Sol. Energy Mater. Sol. Cells 2018, 186, 184–193. [Google Scholar] [CrossRef]
  19. Chen, K.; Napolitani, E.; De Tullio, M.; Jiang, C.S.; Guthrey, H.; Sgarbossa, F.; Theingi, S.; Nemeth, W.; Page, M.; Stradins, P.; et al. Pulsed Laser Annealed Ga Hyperdoped Poly-Si/SiOx Passivating Contacts for High-Efficiency Monocrystalline Si Solar Cells. Energy Environ. Mater. 2023, 6, e12542. [Google Scholar] [CrossRef]
  20. Hassan, M.; Dahlinger, M.; Köhler, J.R.; Zapf-Gottwick, R.; Werner, J.H. Unified Model for Laser Doping of Silicon from Precursors. Materials 2021, 14, 2322. [Google Scholar] [CrossRef]
  21. Sinton, R.A.; Cuevas, A.; Stuckings, M. Quasi-steady-state photoconductance, a new method for solar cell material and device characterization. In Proceedings of the Conference Record of the Twenty Fifth IEEE Photovoltaic Specialists Conference—1996, Washington, DC, USA, 13–17 May 1996; pp. 457–460. [Google Scholar]
  22. Kamins, T. Dopant Diffusion and Segregation. In Polycrystalline Silicon for Integrated Circuit Applications; Springer US: Boston, MA, USA, 1988; pp. 91–123. [Google Scholar] [CrossRef]
  23. Kodera, H. Diffusion Coefficients of Impurities in Silicon Melt. Jpn. J. Appl. Phys. 1963, 2, 212. [Google Scholar] [CrossRef]
  24. Vick, G.L.; Whittle, K.M. Solid Solubility and Diffusion Coefficients of Boron in Silicon. J. Electrochem. Soc. 1969, 116, 1142. [Google Scholar] [CrossRef]
  25. Poate, J.M.; Brown, W.L. Laser annealing of silicon. Phys. Today 1982, 35, 24–30. [Google Scholar] [CrossRef]
  26. White, C.W.; Wilson, S.R.; Appleton, B.R.; Young, F.W. Supersaturated substitutional alloys formed by ion implantation and pulsed laser annealing of group-III and group-V dopants in silicon. J. Appl. Phys. 1980, 51, 738–749. [Google Scholar] [CrossRef]
  27. Fiory, A.; Bourdelle, K.K. A comparative study of laser and thermal annealing of boron-implanted silicon. J. Appl. Phys. 1978, 49, 3912–3917. [Google Scholar] [CrossRef]
  28. Lim, S.; Williams, J. Electrical and Optical Doping of Silicon by Pulsed-Laser Melting. Micro 2021, 2, 1–22. [Google Scholar] [CrossRef]
  29. Goetzlich, J.; Tsien, P.; Ryssel, H. Relaxation Behavior of Metastable As and P Concentrations in Si After Pulsed and CW Laser Annealing. MRS Online Proc. Libr. (OPL) 1983, 23, 235. [Google Scholar] [CrossRef]
  30. Esturo-Breton, A.; Ametowobla, M.; Köhler, J.; Werner, J. Laser Doping for Crystalline Silicon Solar Cell Emitters; European Photovoltaic Solar Energy Conference, 20: München, Germany, 2005; pp. 851–854. [Google Scholar]
  31. McConnell, A.; Uma, S.; Goodson, K. Thermal Conductivity of Doped Polysilicon Layers. Microelectromechanical Syst. 2001, 10, 360–369. [Google Scholar] [CrossRef]
  32. Haynes, W.M. Thermal Conductivity of Metals and Semiconductors as a Function of Temperature. In CRC Handbook of Chemistry and Physics; CRC: Boca Raton, FL, USA, 2014; pp. 212–221. [Google Scholar]
  33. Seto, J.Y.W. The electrical properties of polycrystalline silicon films. J. Appl. Phys. 1975, 46, 5247–5254. [Google Scholar] [CrossRef]
  34. Çiftpınar, H.E.; Stodolny, M.K.; Wu, Y.; Janssen, G.J.; Löffler, J.; Schmitz, J.; Lenes, M.; Luchies, J.M.; Geerligs, L. Study of screen printed metallization for polysilicon based passivating contacts. Energy Procedia 2017, 124, 851–861. [Google Scholar] [CrossRef]
  35. Masetti, G.; Severi, M.; Solmi, S. Modeling of carrier mobility against carrier concentration in arsenic-, phosphorus-, and boron-doped silicon. IEEE Trans. Electron Devices 1983, 30, 764–769. [Google Scholar] [CrossRef]
  36. Sheet Resistance Calculator. Available online: https://www2.pvlighthouse.com.au/calculators/sheet (accessed on 20 December 2022).
  37. Chong, Y.F.; Pey, K.L.; Lu, Y.F.; Wee, A.T.S.; Osipowicz, T.; Seng, H.L.; See, A.; Dai, J.Y. Liquid-phase epitaxial growth of amorphous silicon during laser annealing of ultrashallow p+/n junctions. Appl. Phys. Lett. 2000, 77, 2994–2996. [Google Scholar] [CrossRef]
  38. Ma, D.; Liu, W.; Xiao, M.; Yang, Z.; Liu, Z.; Liao, M.; Han, Q.; Cheng, H.; Xing, H.; Ding, Z.; et al. Highly improved passivation of PECVD p-type TOPCon by suppressing plasma-oxidation ion-bombardment-induced damages. Solar Energy 2022, 242, 1–9. [Google Scholar] [CrossRef]
  39. Stuckelberger, J.; Yan, D.; Phang, S.P.; Samundsett, C.; Wang, J.; Antognini, L.; Haug, F.J.; Wang, Z.; Yang, J.; Zheng, P.; et al. Pre-annealing for improved LPCVD deposited boron-doped poly-Si hole-selective contacts. Solar Energy Mater. Sol. Cells 2023, 251, 112123. [Google Scholar] [CrossRef]
  40. Chaudhary, A.; Hoß, J.; Lossen, J.; Huster, F.; Kopecek, R.; van Swaaij, R.; Zeman, M. Influence of Polysilicon Thickness on Properties of Screen-Printed Silver Paste Metallized Silicon Oxide/Polysilicon Passivated Contacts. Phys. Status Solidi (a) 2021, 218, 2100243. [Google Scholar] [CrossRef]
  41. Padhamnath, P.; Buatis, J.K.; Khanna, A.; Nampalli, N.; Nandakumar, N.; Shanmugam, V.; Aberle, A.G.; Duttagupta, S. Characterization of screen printed and fire-through contacts on LPCVD based passivating contacts in monoPoly™ solar cells. Solar Energy 2020, 202, 73–79. [Google Scholar] [CrossRef]
Figure 2. (a) Schematic cross-section of the Rsh samples. Laser irradiation increases the doping concentration by activating boron atoms in the p+ poly Si layer. The electrical properties of the irradiated regions are characterized by Rsh and ECV measurements. (b) Symmetric passivation samples for QSSPC measurements. The front and rear sides of the samples are irradiated symmetrically, which is followed by an extra high temperature annealing step and SiNx:H deposition on both the front and rear sides. Note that test structures without SiO2 are also fabricated. However, for the sake of simplicity, only samples containing SiO2 are shown.
Figure 2. (a) Schematic cross-section of the Rsh samples. Laser irradiation increases the doping concentration by activating boron atoms in the p+ poly Si layer. The electrical properties of the irradiated regions are characterized by Rsh and ECV measurements. (b) Symmetric passivation samples for QSSPC measurements. The front and rear sides of the samples are irradiated symmetrically, which is followed by an extra high temperature annealing step and SiNx:H deposition on both the front and rear sides. Note that test structures without SiO2 are also fabricated. However, for the sake of simplicity, only samples containing SiO2 are shown.
Solar 03 00021 g002
Figure 5. (a) Doping profiles used for the dmelt determination of samples with a poly Si layer thickness of dpoly Si,2 = 264 nm. The melt depth (indicated by dashed lines) of each Hp is selected as a depth where the high concentration at the surface drops back to the non-melted background doping level. (b) The extracted and average dmelt for dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm poly Si layers in zone II (1 J/cm2Hp ≤ 2 J/cm2). For both poly Si layers, the dmelt increases linearly with increasing Hp.
Figure 5. (a) Doping profiles used for the dmelt determination of samples with a poly Si layer thickness of dpoly Si,2 = 264 nm. The melt depth (indicated by dashed lines) of each Hp is selected as a depth where the high concentration at the surface drops back to the non-melted background doping level. (b) The extracted and average dmelt for dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm poly Si layers in zone II (1 J/cm2Hp ≤ 2 J/cm2). For both poly Si layers, the dmelt increases linearly with increasing Hp.
Solar 03 00021 g005
Figure 6. The measured Gsh of the c-Si/SiO2/p+-poly Si samples in zone II. The Gsh of both poly Si layers linearly increases with increase in the dmelt, suggesting that the quantity of incorporated boron atoms in conduction also increases by increasing dmelt. Due to the equal total number of boron atoms from the same BBr3 diffusion step, boron atoms are distributed more densely in the dpoly Si,1 = 155 nm samples, resulting in a higher number of active boron atoms in the dpoly Si,1 = 155 nm samples compared to dpoly Si,2 = 264 nm for the same melt depth.
Figure 6. The measured Gsh of the c-Si/SiO2/p+-poly Si samples in zone II. The Gsh of both poly Si layers linearly increases with increase in the dmelt, suggesting that the quantity of incorporated boron atoms in conduction also increases by increasing dmelt. Due to the equal total number of boron atoms from the same BBr3 diffusion step, boron atoms are distributed more densely in the dpoly Si,1 = 155 nm samples, resulting in a higher number of active boron atoms in the dpoly Si,1 = 155 nm samples compared to dpoly Si,2 = 264 nm for the same melt depth.
Solar 03 00021 g006
Figure 7. ECV profiles of samples with and without SiO2 (position of SiO2 is marked) with the poly Si layer thicknesses dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm. In zone I (Hp = 0 J/cm2), for both poly Si layers in (a,b), similar doping profiles are measured for structures with and without SiO2. These profiles represent the background doping after the BBr3 furnace diffusion. The presence of SiO2 plays a minor role in the background doping by limiting the boron diffusion in c-Si for samples with SiO2 in this zone. In zone II, Hp = 1.3 J/cm2 results in ECV profiles with high active boron concentrations at the surface of the poly Si layers, which decrease towards the depth and reach the background doping level. Thus, the SiO2 does not significantly influence the electrical properties of dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm samples. In Zone III, for samples with both poly Si layers, samples with SiO2 show a significant reduction in boron concentration at the position of SiO2, while for samples without SiO2, a deep profile inside c-Si is measured. In zone IV, Hp = 5 J/cm2 creates similar ECV profiles for samples with and without SiO2 for both poly Si layers, which suggests the destruction of SiO2.
Figure 7. ECV profiles of samples with and without SiO2 (position of SiO2 is marked) with the poly Si layer thicknesses dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm. In zone I (Hp = 0 J/cm2), for both poly Si layers in (a,b), similar doping profiles are measured for structures with and without SiO2. These profiles represent the background doping after the BBr3 furnace diffusion. The presence of SiO2 plays a minor role in the background doping by limiting the boron diffusion in c-Si for samples with SiO2 in this zone. In zone II, Hp = 1.3 J/cm2 results in ECV profiles with high active boron concentrations at the surface of the poly Si layers, which decrease towards the depth and reach the background doping level. Thus, the SiO2 does not significantly influence the electrical properties of dpoly Si,1 = 155 nm and dpoly Si,2 = 264 nm samples. In Zone III, for samples with both poly Si layers, samples with SiO2 show a significant reduction in boron concentration at the position of SiO2, while for samples without SiO2, a deep profile inside c-Si is measured. In zone IV, Hp = 5 J/cm2 creates similar ECV profiles for samples with and without SiO2 for both poly Si layers, which suggests the destruction of SiO2.
Solar 03 00021 g007
Figure 10. Comparison of measured Gsh from layers without SiO2 with expected Gsh when considering the mobility of L-poly Si within the first 155 nm of the sample (hypothesis A) and when considering the mobility of c-Si for the entire samples structure (hypothesis B). For Hp less than the HT, in which the Si melt is not extended to the c-Si substrate, the measured Gsh-values agree with the predicted Gsh-values of hypothesis A, suggesting that the Si melt recrystallizes to an L-poly Si structure. However, when Hp exceeds the HT, the predicted Gsh-values by hypothesis B agree with the measured Gsh-values indicating that the Si melt recrystallizes to a mono-crystal layer with comparable carrier mobility to a single-crystal structure.
Figure 10. Comparison of measured Gsh from layers without SiO2 with expected Gsh when considering the mobility of L-poly Si within the first 155 nm of the sample (hypothesis A) and when considering the mobility of c-Si for the entire samples structure (hypothesis B). For Hp less than the HT, in which the Si melt is not extended to the c-Si substrate, the measured Gsh-values agree with the predicted Gsh-values of hypothesis A, suggesting that the Si melt recrystallizes to an L-poly Si structure. However, when Hp exceeds the HT, the predicted Gsh-values by hypothesis B agree with the measured Gsh-values indicating that the Si melt recrystallizes to a mono-crystal layer with comparable carrier mobility to a single-crystal structure.
Solar 03 00021 g010
Figure 12. Role of SiO2 in the crystallization mechanism of laser-activated samples. (a) A laser pulse with sufficient energy melts the poly Si layer, but the SiO2 blocks the Si melt transfer to the c-Si bulk, resulting in the seedless crystallization of the Si melt into a laser-crystallized L-poly Si structure. (b) In the absence of an interfacial oxide, the Si melt propagates to the substrate and melts a portion of the c-Si bulk. After stopping the laser pulse, the molten Si recrystallizes by liquid phase epitaxy from the underlying c-Si substrate with the same lattice structure.
Figure 12. Role of SiO2 in the crystallization mechanism of laser-activated samples. (a) A laser pulse with sufficient energy melts the poly Si layer, but the SiO2 blocks the Si melt transfer to the c-Si bulk, resulting in the seedless crystallization of the Si melt into a laser-crystallized L-poly Si structure. (b) In the absence of an interfacial oxide, the Si melt propagates to the substrate and melts a portion of the c-Si bulk. After stopping the laser pulse, the molten Si recrystallizes by liquid phase epitaxy from the underlying c-Si substrate with the same lattice structure.
Solar 03 00021 g012
Figure 13. (a) Implied open circuit voltage iVOC of samples with and without SiO2 with a poly Si layer thickness dpoly Si,1 = 155 nm. For Hp < 4.6 J/cm2 samples with SiO2 mostly show high iVOC = 721 mV. But, Hp = 4.6 J/cm2 reduces the iVOC by nearly 60 mV by damaging the SiO2. The Hp = 5 J/cm2 destroys the SiO2 and lowers the iVOC to values comparable with samples without SiO2. (b) Increasing the tested poly Si thickness to dpoly Si,2 = 264 nm does not affect the passivation quality. For Hp < 4.6 J/cm2, the majority of the samples show a good passivation quality with the highest iVOC = 722 mV. However, a thicker poly Si layer reduces the damage caused by Hp = 4.6 J/cm2 to the SiO2 and results in higher iVOC values compared to samples with a thinner poly Si layer. Still, Hp = 5 J/cm2 destroys the SiO2 and strongly reduces the iVOC.
Figure 13. (a) Implied open circuit voltage iVOC of samples with and without SiO2 with a poly Si layer thickness dpoly Si,1 = 155 nm. For Hp < 4.6 J/cm2 samples with SiO2 mostly show high iVOC = 721 mV. But, Hp = 4.6 J/cm2 reduces the iVOC by nearly 60 mV by damaging the SiO2. The Hp = 5 J/cm2 destroys the SiO2 and lowers the iVOC to values comparable with samples without SiO2. (b) Increasing the tested poly Si thickness to dpoly Si,2 = 264 nm does not affect the passivation quality. For Hp < 4.6 J/cm2, the majority of the samples show a good passivation quality with the highest iVOC = 722 mV. However, a thicker poly Si layer reduces the damage caused by Hp = 4.6 J/cm2 to the SiO2 and results in higher iVOC values compared to samples with a thinner poly Si layer. Still, Hp = 5 J/cm2 destroys the SiO2 and strongly reduces the iVOC.
Solar 03 00021 g013
Figure 14. Saturation current densities J0 per side of symmetrical lifetime samples with SiO2. (a) A poly Si layer with dpoly Si,1 = 155 nm: For Hp < 4.6 J/cm2 where the SiO2 is not damaged, the majority of samples show good surface passivation with J0 < 10 fA/cm2. The Hp = 4.6 J/cm2 leads to damaged SiO2 and, consequently, a higher J0 = 112 fA/cm2. The Hp = 5 J/cm2 destroys the SiO2 and results in J0 = 482 fA/cm2. (b) A poly Si layer with dpoly Si,2 = 264 nm: Similar to the samples with a thinner poly Si layer, Hp < 4.6 J/cm2 results in low J0, with the best sample achieving J0 = 6.7 fA/cm2. The increased thickness of the poly Si layer reduces the damage of the Si melt to the SiO2 for Hp = 4.6 J/cm2, which results in a lower J0 = 17 fA/cm2 compared to the samples with a thinner poly Si layer. The high measured J0 for Hp = 5 J/cm2 shows that increasing the thickness of the poly Si layer does not prevent the SiO2 destruction with the highest Hp.
Figure 14. Saturation current densities J0 per side of symmetrical lifetime samples with SiO2. (a) A poly Si layer with dpoly Si,1 = 155 nm: For Hp < 4.6 J/cm2 where the SiO2 is not damaged, the majority of samples show good surface passivation with J0 < 10 fA/cm2. The Hp = 4.6 J/cm2 leads to damaged SiO2 and, consequently, a higher J0 = 112 fA/cm2. The Hp = 5 J/cm2 destroys the SiO2 and results in J0 = 482 fA/cm2. (b) A poly Si layer with dpoly Si,2 = 264 nm: Similar to the samples with a thinner poly Si layer, Hp < 4.6 J/cm2 results in low J0, with the best sample achieving J0 = 6.7 fA/cm2. The increased thickness of the poly Si layer reduces the damage of the Si melt to the SiO2 for Hp = 4.6 J/cm2, which results in a lower J0 = 17 fA/cm2 compared to the samples with a thinner poly Si layer. The high measured J0 for Hp = 5 J/cm2 shows that increasing the thickness of the poly Si layer does not prevent the SiO2 destruction with the highest Hp.
Solar 03 00021 g014
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sharbaf Kalaghichi, S.; Hoß, J.; Zapf-Gottwick, R.; Werner, J.H. Laser Activation for Highly Boron-Doped Passivated Contacts. Solar 2023, 3, 362-381. https://0-doi-org.brum.beds.ac.uk/10.3390/solar3030021

AMA Style

Sharbaf Kalaghichi S, Hoß J, Zapf-Gottwick R, Werner JH. Laser Activation for Highly Boron-Doped Passivated Contacts. Solar. 2023; 3(3):362-381. https://0-doi-org.brum.beds.ac.uk/10.3390/solar3030021

Chicago/Turabian Style

Sharbaf Kalaghichi, Saman, Jan Hoß, Renate Zapf-Gottwick, and Jürgen H. Werner. 2023. "Laser Activation for Highly Boron-Doped Passivated Contacts" Solar 3, no. 3: 362-381. https://0-doi-org.brum.beds.ac.uk/10.3390/solar3030021

Article Metrics

Back to TopTop