Conference PaperPDF Available

Boron-doped PECVD silicon oxides as diffusion sources for simplified high-efficiency solar cell fabrication

Authors:
  • WAVELABS Solar Metrology Systems GmbH

Figures

Content may be subject to copyright.
BORON-DOPED PECVD SILICON OXIDES AS DIFFUSION SOURCES FOR SIMPLIFIED
HIGH-EFFICIENCY SOLAR CELL FABRICATION
Nadine Wehmeier1, Gunnar Schraps1, Hannes Wagner3, Bianca Lim1,
Nils-Peter Harder2*, and Pietro P. Altermatt3
1 Institute of Solar Energy Research Hamelin (ISFH), Am Ohrberg 1, 31860 Emmerthal, Germany
2 Institute of Electronic Materials and Devices, Univ. Hannover, Schneiderberg 32, 30167 Hannover, Germany
*Now with Total New Energies, USA; seconded to SunPower Corp., 77 Rio Robles, San Jose, California 95134, USA
3 Dep. Solar Energy, Leibniz Universität Hannover, Appelstr. 2, 30167 Hannover, Germany
ABSTRACT: We develop plasma-deposited boron-doped silicon oxide layers with multifunctional capabilities: Firstly,
they are used as local diffusion sources for attaining low saturation current densities of J0 = 84 fA/cm2 at sheet resistances
as low as 67 / and surface boron concentrations in the range of 4–71019 cm3, suitable for efficient metallization.
Secondly, they act as diffusion barriers against other dopants. Therefore a co-diffusion process consisting of a boron
drive-in and a phosphorus diffusion in a POCl3 furnace in only one high temperature step potentially allows to simplify
the fabrication of high-efficiency n-type PERT cells. Process and device simulations indicate that a co-diffusion
temperature near 900C and a co-diffusion time of about 20 minutes are optimum for reaching cell efficiencies near 21%.
Keywords: PECVD, oxide, boron, diffusion, simulation
1 INTRODUCTION
During conventional fabrication of high-efficiency
solar cells, a high number of process steps is usually
necessary for creating structured or single-side emitter
and back surface field (BSF) regions. The process steps
include at least two high-temperature furnace diffusions,
each accompanied with depositing and removing capping
layers.
Resource, cost and time efficient production processes, as
shown in figure 1, may be achieved by using oxide layers
deposited by means of plasma-enhanced chemical vapour
deposition (PECVD): the doped oxide layers are
deposited one-sided, and in only one co-diffusion
process, the dopant atoms from the oxide are driven in
while simultaneously a second doping type is diffused
from the gas phase as already demonstrated in [1].
Alternatively, PECVD oxide layers with different doping
types can be deposited on both sample sides and act as
diffusion sources in a high-temperature step, so that the
different doping profiles can be controlled and optimised
independently as shown in [2]. Multifunctional PECVD
layer stacks can also be used for surface passivation and
as anti-reflection coating as demonstrated in [3].
In this paper, we investigate how the boron and
phosphorus doping profiles are influenced and can be
controlled independently during the co-diffusion process.
This is done by applying a nitride capping layer on top of
the boron-doped PECVD oxide and by varying the
nitrogen gas-flow and thus modifying the atmosphere in
the furnace. In addition, experimental and simulated
doping profiles resulting from different diffusion times
and temperatures are investigated both by experiments
and simulations. This allows us to achieve a deeper
understanding of the co-diffusion optimisation with
regard to high efficiencies of solar cells fabricated with
doped oxides.
2 EXPERIMENTAL PRCEUDRES
Doped oxide layers, between 40 nm and 100 nm thick,
are grown on monocrystalline silicon substrates by means
of PECVD within about four minutes, using an
inductively coupled plasma (ICP) source. As precursor
gases, silane (SiH4) and nitrous oxide (N2O) are used. To
produce doped oxides, the precursor gas diborane (B2H6)
or phosphine (PH3) is added. In contrast, depositing
PECVD borosilicate glass (BSG) layers from
trimethylborate (TMB) has been investigated e.g. in [1].
In this paper, we first focus on growing boron-doped
silicon oxides on planar p-type monocrystalline silicon
substrates. Then we present boron-doped oxides on n-
type alkaline textured silicon surfaces, relevant for
fabricating n-type PERT cells.
Figure 1: Simplified fabrication of n-type passivated
emitter and rear totally diffused (PERT) solar cells. Using
boron-doped PECVD oxides allows us to form the boron
emitter and the phosphorus BSF simultaneously in a co-
diffusion process. Only one high temperature step and
one protection layer are needed.
To achieve a diffusion barrier for other dopants, some
samples are subsequently deposited with a SiNx capping
layer, having a refractive index of n = 1.9 and a thickness
of about 100 nm.
In a high-temperature step, the dopant atoms diffuse from
the PECVD oxide into the silicon substrate. This boron
drive-in step is combined with a POCl3 furnace diffusion
in a co-diffusion process, in order to form a n-type
emitter and a p-type BSF on a p-type substrate
28th European Photovoltaic Solar Energy Conference and Exhibition
1980
simultaneously. This process enables us to skip a separate
BBr3 diffusion within a quartz furnace.
Note that phosphorus atoms diffuse effectively at
temperatures of about 800-900°C, while the boron
diffusion is only effective above about 900°C. As a
consequence, the co-diffusion step within the furnace is
splitted into two segments: first, the drive-in of B at
1000°C for 30-60 min in a nitrogen atmosphere (with N2
gas flows between 3 l/min and 12 l/min) and, second, the
diffusion of P after cooling down to 852°C with POCl3
(400 ml/min), O2 (2 l/min) and N2 (9 l/min) gases for 20
min.
The resulting doped emitter and BSF regions are
characterised by their sheet resistivities and saturation
current densities that are determined by means of four
point probe (FPP) and inductively-coupled
photoconductance measurements. In addition
electrochemical capacitance voltage (ECV)
measurements are performed on both sample sides to
determine the B and P doping profiles, especially the
surface concentrations and the profile depths.
First PERT cells with a boron-doped PECVD oxide as
diffusion source are fabricated with the simplified
process-flow shown in figure 1. An n-type silicon
substrate with a specific resistance of 5.5 cm and a
thickness of about 160 µm after alkaline texture is used.
After PECVD-BSG and SiNx capping layer deposition
the co-diffusion as described above is applied. Resulting
BSG and PSG layers are removed in hydrofluoric acid
and the front and rear surfaces are passivated by a 100
nm thick AlOx/SiNx stack and a SiNx layer, respectively.
For metallisation AgAl and Ag screen-printing are used
and co-fired. For the characterisation of the cells, current-
voltage (IV) measurements are performed.
3 EXPERIMENTAL RESULTS
3.1 Doped oxides on planar p-type Si wafers
The diffusion profiles are known to be strongly
influenced by the temperature and time during the drive-
in step. We observe that also the applied nitrogen flow
influences the resulting sheet resistivities: an increasing
N2 flow leads to decreasing sheet resistivities of the P
emitters as exemplarily shown in figure 2.
Figure 2: Doping profiles measured by ECV on the
sample side that was uncovered and therefore was P-
diffused during the co-diffusion process with N2 gas-
flows of either 8 l/min or 12 l/min during the 1000°C
step. The ECV measurements indicate that n-type doped
emitters with sheet resistivities of about 13-125 / are
reached, depending on the N2 flow and the sample
position within the furnace.
Figure 3 demonstrates that by the co-diffusion
process, BSF regions with sheet resistances of about 67-
97 / are successfully produced. Note the strong
influence of the SiNx capping layer: only oxide/nitride
stacks result in p-type doped regions. If the doped oxide
layers are exposed to the co-diffusion uncapped, a n-type
doped surface region is observed, indicating
overcompensation by P.
Figure 3: Doping profiles measured by ECV on the
sample sides that were covered with boron-doped
PECVD oxides, (a) with and (b) without SiNx capping
layer. A co-diffusion process with N2 gas
flows of either 8 l/min or 12 l/min during the 1000°C step
was performed. (a) P-type doped BSF regions with sheet
resistivities of about 67-97 /□, measured by ECV and
FPP are reached, nearly unaffected by the N2 flow and
the sample position within the furnace. (b) In the case of
the uncapped oxide layers, n-type doping is measured at
the sample surface indicating that phosphorus was
diffused through the boron-doped oxide. At a gas
flow of 8 l/min, the measured doping type changes to p-
type due to boron diffusion, while at 12 l/min of N2 no p-
type doping is observed at all.
3.2 Doped oxides on textured n-type Si wafers
The doped oxides deposited on an alkaline textured
surface are investigated by means of scanning electron
microscopy (SEM) directly after PECVD deposition and
after co-diffusion and BSG removal in order to
investigate the homogeneity of the doped oxide and the
resulting boron emitter.
The SEM images presented in figure 4 show that the
PECVD of a boron-doped oxide with a deposition
thickness of 80 nm on a planar silicon surface leads to a
closed oxide layer with a thickness of about 40-65 nm.
This oxide thickness is to be expected due to the about
1.7 times higher surface of textured silicon. After the co-
diffusion a continuous p-type boron emitter is formed on
(a)
(b)
28th European Photovoltaic Solar Energy Conference and Exhibition
1981
the n-type silicon, that is clearly visible in the SEM
image due to the doping contrast.
Figure 4: SEM images of boron doped PECVD oxides
on alkaline textured n-type silicon. (a) A closed oxide
layer is formed, its thickness varies between about 65 nm
at the top and 40 nm at the bottom of a pyramid. (b) A
continuous p-type emitter with a depth of 500-900 nm is
formed during the boron drive-in at 1000°C (apparent as
bright area).
The boron concentration in the doped oxide layers is
increased by changing the gas flows of the precursor
gases. The resulting doping profiles measured by ECV
shown in figure 5 reveal high surface concentrations of
about Nsurf = 3.5-141019 cm-3 resulting in sheet
resistivities of Rsheet = 61-73 / after 30 min of boron
drive-in. Increasing the drive-in time to 60 min results in
deeper doping profiles and thus lower sheet resistivities
of 21-53 /. Corresponding emitter saturation current
densities of J0e = 125-145 fA/cm² have been measured on
planar p-type 200 cm silicon substrates, using the
intrinsic density ni = 8.6109 cm-3.
If the boron concentration in the doped oxide is high
enough to exceed the solubility limit of boron in silicon
of about 1.01020 cm-3 at the drive-in temperature of
1000°C [4], a boron rich layer (BRL) formation at the
silicon/oxide interface occurs during the high temperature
step. The presence of a BRL is indicated by the observed
hydrophilic behaviour of the sample surface after HF
etch. It makes a low temperature oxidation (LTO)
necessary to remove the BRL in order to prevent bulk
carrier lifetime degradation as observed in [5].
Figure 5: Boron doping profiles measured by ECV.
Higher surface concentrations and diffusion depths and
lower sheet resistivities are reached by increasing the
B2H6/SiH4 gas flow ratio and the boron drive-in time.
3.3 Cell results
From the IV measurements on the first fabricated n-
type PERT cells with an area of 4 cm² the cell parameters
shown in table I are extracted.
Table I: Results of the first n-type PERT cells we
fabricated in a simplified manner with boron-doped
PECVD oxide as diffusion source.
[%]
FF
[%]
pFF
[%]
Jsc
[mA/cm²]
Voc
[mV]
A
[cm²]
17.5 72.9 78.2 37.9 633 4.0
4 DISCUSSION
We attribute the influence of different N2 flows on the
sheet resistivities to the following effect: Nitrogen
displaces oxygen that is inserted in the furnace while
loading the quartz boat containing the samples. For low
N2 flows during the 1000°C step, a high concentration of
unwanted oxygen reacts with silicon on the uncoated
substrate surface leading to a thick thermal oxide layer.
For the rear position, the oxygen concentration and
therefore the oxide thickness is highest leading to sheet
resistivities of about 125 /□. But for high N2 flows
and thin thermal oxide layers, the amount of P2O5 from
the furnace walls dissolved in the oxide is sufficient to
decrease the melting temperature of the SiO2- P2O5-phase
such that this mixed phase becomes liquid at diffusion
temperatures 850°C [6]. Thus P-diffused regions with
very low emitter sheet resistivities of about 13 /are
formed during the 1000°C step and the subsequent POCl3
diffusion.
Boron-doped oxide layers grown by PECVD do not
act as diffusion barriers against phosphorus diffusion at
temperatures as high as 1000°C. The measured ECV
doping profiles, which reveal a change of doping type
shown in figure 3 (b), may be explained as follows:
during the 1000°C step, oxygen diffuses through the
PECVD oxide and forms a thermal oxide layer at the
silicon interface. In the case of a high nitrogen flow and a
thin oxide layer, P2O5 can react at the silicon surface so P
diffuses into the substrate. Simultaneously, boron is
diffused from the doped oxide, but due to the higher P
concentration and solubility in silicon compared to B, a
n-type doped region is formed. In the case of a lower N2
gas flow of 8 l/min, a thicker thermal oxide is grown that
retards the P diffusion, while boron is already driven in
so only the surface region becomes P doped. See figure 3
(b), co-diffusion 1.
5 PROCESS AND DEVICE SIMULATION
We use the software SENTAURUS PROCESS [7] to
reproduce the experimental boron profile shown in
figure 3 (a) (co-diffusion number 3, middle position). We
use a simulation approach comparable to [8] and adjust
the boron concentration in the 80 nm thick oxide layer,
from where the boron diffuses into silicon. We use the
same temperature and time sequence as in the experiment
and obtain the result shown in figure 6 as lines. The
experiment (symbols) is reproduced very well without the
need to adjust any other model parameters. For
comparison, we also simulate a standard boron silicate
glass (BSG) layer, shown in figure 6 as dashed line. A
(a)
(b)
28th European Photovoltaic Solar Energy Conference and Exhibition
1982
standard BSG has a higher boron concentration than our
boron-doped oxide, resulting in a higher boron peak
concentration in silicon. To achieve the same profile
depths, we need to reduce the diffusion time by about
50%. The reason is that the diffusivity of boron in Si
depends on the boron concentration.
Figure 6: Boron dopant profiles, fabricated with a boron-
doped oxide (symbols), as simulated (solid line), and a
simulation of a profile fabricated with a standard boron
silicate glass (BSG) layer (dashed line).
We use the software SENTAURUS DEVICE (former
DESSIS) with the models described in [9] to compute the
saturation current density J0 of these two profiles with the
method of Kane and Swanson [10]. Because J0 depends
sensitively on the surface recombination velocity S, we
update the data on Al2O3-passivated, boron-diffused
surfaces of Hoex et al. [11] by including the recent data
from Richter at al. [12], as shown in figure 7. The line is
a least-square fit to the data (symbols), expressed as:
(1)
similarly to [13]. Using this updated parameterisation for
S, we obtain J0 = 84 fA/cm2 for the doped-oxide profile,
compared to J0 = 183 fA/cm2 for the BSG profile (at
25C, using 8.312109 cm-3 as intrinsic density ni, given
in equation 3 of [14]). The reason for this difference in J0
is the reduced amount of Auger recombination in the
doped-oxide profile. By doping the oxide weaker, J0 may
be reduced further without reaching unfeasibly high sheet
resistivities.
Figure 7: Symbols: surface recombination velocity on
boron-diffused Al2O3-passivated surfaces, extracted by
device simulations from published dopant profiles and J0
values [8,9]. Line: the parameterisation given in equation
(1).
Finally, we model the performance of PERT cells in
two dimensions, again by using SENTAURUS DEVICE. As
p+-emitters, we use the simulated boron dopant profiles
shown in figure 8. The boron density in the oxide is
assumed to be sufficiently high, so the solid-solubility of
boron in Si is just reached at the surface. This is done
because, with higher boron concentrations, lower sheet
resistivities are attained within shorter diffusion times
and at lower temperatures, which is industrially relevant.
However, care must be taken that the solubility limit is
not exceeded, otherwise a boron rich layer is formed,
which would need to be oxidised for removal. At the rear
side of the cell, the phosphorus profiles measured and
modeled in [15] are used as BSFs, shown in figure 9. The
simulation domain is two-dimensional, as described in
[7] in detail. We add a lumped external series resistance
of 0.5 cm2 due to the metallisation, and an effective
shading loss of 3% [16] in the module, which
corresponds to a metallisation fraction of the front
surface in the range of 5-6%. The SRH lifetime in the 1
cm base is assumed to be 1 ms. The resulting efficiency
for the three different diffusion temperatures and
diffusion times is shown in figure 10. It is apparent that
the cells, having a front finger distance of 2 mm, perform
considerably worse than if the finger distance is lowered
to 1 mm with accordingly finer metal fingers. The reason
is that boron has a small diffusivity and solid-solubility
compared to phosphorus, leading to a high sheet
resistivity of the emitter. To keep the manufacturing costs
down, it is important to keep both the co-diffusion
temperature and time as low as possible. The simulations
indicate (with the assumptions outlined above) that a co-
diffusion temperature near 900C and a co-diffusion time
near 20 minutes are optimum. In practice, often higher
temperatures and longer diffusion times need to be
chosen to achieve an optimum. This discrepancy between
modeling and experiment may point to relevant non-
idealities in fabrication, such as: (i) the oxide near the Si
surface may start to act as diffusion barrier during the
diffusion; (ii) inhomogeneity is a notorious problem in
boron diffusions, especially at lower temperatures; (iii)
the boron diffusion is defect-mediated, so the defects
introduced by oxidisation or texturing may play an
important role.
Regarding the strategy for PERT cell development,
the rather high sheet resistivity of boron diffusions
implies that, in the first place, special attention should be
given to fine front finger metallisation to keep the front
finger spacing small. With improved screen printing
pastes, it is likely that a fine front finger metallisation is
cheaper than long diffusion times at high diffusion
temperatures. Secondly, care must be taken that the
recombination losses in the phosphorus-diffused BSF do
not dominate the total recombination losses. This easily
occurs because phosphorus has a higher solubility and
diffusivity in Si than boron, and phosphorus leads to
precipitation. The losses in the BSF are most likely to
dominate at co-diffusion temperatures above 900°C.
Therefore, low POCl3-flows should be used when
designing PERT process sequences. Thirdly, our
simulations also indicate that a p-type wafer resistivity
near 1 cm is optimum (not shown here). Wafer
resistivities near 5 cm, for example, lower the
efficiency by 0.3% absolute, assuming that the SRH
lifetime is independent of wafer resistivity.
28th European Photovoltaic Solar Energy Conference and Exhibition
1983
Figure 8: Boron dopant-profiles, simulated with
SENTAURUS PROCESS for the indicated diffusion
temperatures and diffusion times, using highly-doped
oxides.
Figure 9: Phosphorus dopant-profiles, modeled in [15]
based on experiments, for the indicated diffusion
temperatures and diffusion times.
Figure 10: Simulated PERT cell efficiencies in
dependence of the co-diffusion time, for the various co-
diffusion temperatures as indicated. The solid lines are
with a front metal finger distance of 1 mm, the dashed
lines with 2 mm.
6 CONCLUSIONS
Doped oxide layers have a high potential to simplify
the fabrication of high-efficiency Si cells like PERT and
interdigitated back contact (IBC) cells. High temperature
steps and protection layers and thus process time and
costs are likely to be reduced, which is necessary for an
efficient industrial production.
First n-type PERT cells with boron-doped PECVD
oxides have been fabricated and reached efficiencies of
= 17.5%. Simulations of dopant-profiles and PERT cell
efficiencies show that by optimising the co-diffusion
process concerning diffusion time and temperature
efficiencies in the range of 20-21% can be achieved.
There are some requirements to the emitter and BSF
of a high-efficiency PERT cell with reduced process
steps that we fulfil by optimising the deposition of doped
PECVD oxides and the co-diffusion process: the oxides
have to be deposited homogeneously to form a
continuous boron emitter on alkaline textured n-type
silicon. The emitter needs to have a sufficient thickness
and surface concentration to allow efficient metallisation.
But the boron concentration in the doped oxide layer
must be kept low enough to prevent BRL formation so
that no additional LTO and HF etch are needed. Also the
J0 value of a weaker boron doped oxide is further reduced
allowing for improved cell performance. We control the
doping profile of the phosphorus BSF efficiently by the
gas flows during the POCl3 co-diffusion while the boron
profile is unaffected by the atmosphere in the furnace and
is adjusted by temperature and time during the drive-in.
[1] B. Bazer-Bachi, C. Oliver, B. Semmache, Y.
Pellegrin, M. Gauthier, N. Le Quang, M. Lemiti, 26th
EU PVSEC Proceedings (2011), p. 1155.
[2] P. Rothardt, R. Keding, A. Wolf, D. Biro, Phys.
Status Solidi RPL (2013) 1-4.
[3] J. Seiffe, F. Pillath, D. Trogus, A. Brand , C. Savio,
M. Hofmann, J. Rentsch, R. Preu, IEEE Journal of
Photovoltaics, vol.3, no.1 (2013) p.224-229.
[4] R. Hull, Properties of cristalline silicon, 10.4
Solubility of B, Al, Ga, In, Tl, P, As and Sb in c-Si
(1999).
[5] M. Kessler, T. Ohrdes, B. Wolpensinger, N.-P.
Harder, Semicond. Sci. Technol. 25 (2010) 055001.
[6] P. Balk, J. M. Elderidge, Phosphosilicate Glass
Stabilization of FET Devices, Proceedings of the
IEEE, vol.57, no.9 (1969) p. 1558- 1563.
[5] Sentaurus Userguide, Synopsys, Mountain View, CA,
2013.
[8] H. Wagner et al., Proc. 37th IEEE PV Specialists
Conf. (2011).
[9] P. P. Altermatt, J. Comput. Electron. 10 (2011) 314.
[10] D.E. Kane and R.M. Swanson, 18th IEEE PV
Specialists Conf. (1985), 578.
[11] B. Hoex, et al., Appl. Phys. Lett. 91 (2007) 112107.
[12] A. Richter et al., IEEE J. PV3 (2013) 236.
[13] P.P. Altermatt et al., J. Appl. Phys. 3187 (2002) 92.
[14] K. Misiakos and D. Tsamakis, J. Appl. Phys. 74
(1993) 3293.
[15] A. Bentzen et al., J. Appl. Phys. 99 (2006) 064502.
[16] R. Woehl, et al., 23rd European PV Solar Energy
Conf. (2008) 1377.
28th European Photovoltaic Solar Energy Conference and Exhibition
1984
... Cabal et al. [35] adapted boron-doped silicon oxide (SiOx:B) layers, also denoted as BSG, by means of plasma-enhanced CVD (PECVD) to the fabrication of Si solar cells in 2009. Since then, considerable research has been carried out on this topic by several research groups [53][54][55][56][57][58][59][60][61][62][63][64][65][66]. It was shown that SiOx:B layers, manufactured by PECVD, can form an appropriate boron diffusion source, resulting in boron emitters with high electrical quality in terms of a low j0e [62][63][64]. ...
... The highly-doped region can be preserved by selective masking with an etch-resistant mask, whereas the dead layer is removed in the non-masked regions. Thus, the electrical losses on the front surface can be minimized [57], screen-printing of etch-resistant pastes or inkjetting of etch-resistant inks [102]. ...
... Recent research has focused on the formation process of PECVD silicate glass emitters [7][8][9], monitoring the quality of these PECVD emitters [10], co-diffusion processes with PECVD emitters [11,12], contacting of these emitters [13,14], passivation through the same BSG layer [15], and the fabrication of solar cells using PECVD silicate glass [7][8][9]12,14,[16][17][18][19]. Solar cell efficiencies of above 20% have been achieved with PECVD silicate glass processes for n-type small area back junction IBC cell concepts [16], large area back junction [18,19], and front junction bifacial solar cell concepts [14]. ...
... Solar cell efficiencies of above 20% have been achieved with PECVD silicate glass processes for n-type small area back junction IBC cell concepts [16], large area back junction [18,19], and front junction bifacial solar cell concepts [14]. Evidence for interacting diffusion phenomena have also been reported using a BSG layer and phosphorus oxychloride (POCl 3 ) diffusion [15,17]. ...
Article
Boron and phosphorus doping of crystalline silicon using a borosilicate glass (BSG) layer from plasma‐enhanced chemical vapor deposition (PECVD) and phosphorus oxychloride diffusion, respectively, is investigated. More specifically, the simultaneous and interacting diffusion of both elements through the BSG layer into the silicon substrate is characterized in depth. We show that an overlying BSG layer does not prevent the formation of a phosphorus emitter in silicon substrates during phosphorus diffusion. In fact, a BSG layer can even enhance the uptake of phosphorus into a silicon substrate compared with a bare substrate. From the understanding of the joint diffusion of boron and phosphorus through a BSG layer into a silicon substrate, a model is developed to illustrate the correlation of the concentration‐dependent diffusivities and the emerging diffusion profiles of boron and phosphorus. Here, the in‐diffusion of the dopants during diverse doping processes is reproduced by the use of known concentration dependences of the diffusivities in an integrated model. The simulated processes include a BSG drive‐in step in an inert and in a phosphorus‐containing atmosphere. Based on these findings, a PECVD BSG/capping layer structure is developed, which forms three different n ⁺⁺ −, n ⁺ − and p ⁺ −doped regions during one single high temperature process. Such engineered structure can be used to produce back contact solar cells. Copyright © 2016 John Wiley & Sons, Ltd.
... Another option is the use of one or more pre-deposited dopant sources with or without an additional gaseous precursor in the process atmosphere of a high temperature step. As all dopants enter the crystal in the same process, it has been termed "co-diffusion" [9][10][11]. The feasibility of this approach has been shown for doped silicate layers from both Atmospheric Pressure and Plasma Enhanced Chemical Vapor Deposition ("APCVD", "PECVD") [12][13][14][15], as well as Spin-On Glasses ("SOG") [16][17][18]. ...
Conference Paper
We report recent progress in the adoption of an optimized screen-printable boron dopant material, which enables the cost-competitive fabrication of novel n-type silicon solar cells such as “nPERT” and “IBC” cells. We manufactured first “IBC” devices with an early version of our dopant material, a co-diffusion approach and evaporated electrodes, achieving an efficiency of 20.9 % in 2016. Since then we optimized the dopant material and co-diffusion processes, as reported in this work. These optimizations now enable a long-term printing of the dopant material, suited for mass production, while maintaining crucial performance parameters. Moreover, we demonstrate a co-diffusion setup with POCl3, which does not require any additional dopant sources and can instead be adjusted with a wet chemical etch back. Combined with new commercially available electrode pastes, which have been evaluated with regard to simultaneous contacting of n- and p-type dopings, all screen printed n-type solar cells become a mass market possibility in the near future.
... Different approaches are being pursued to improve the emitter in terms of recombination and with it, the performance of solar cells, by reducing the emitter's surface doping concentration N surf [3][4][5][6]. One widespread option is to use a diluted source, meaning a reduction of the phosphorus concentration in the PSG/SiO 2 stack layer [3,4,7]. This is realized by, e.g., a reduced N 2 -POCl 3 to oxygen (N 2 -POCl 3 :O 2 ) ratio during the growth of the PSG/SiO 2 stack layer. ...
Article
Full-text available
The phosphosilicate glass (PSG) layer system grown on the silicon surface during diffusion processes with phosphorus oxychloride (POCl3) is a two-layer stack system consisting of a PSG and a silicon dioxide (SiO2) layer. Understanding the stack layers’ structure and composition is essential for further optimizing POCl3 diffusion processes. For diffusion processes with in-situ oxidation (i.e. a high oxygen gas flow during the drive-in step), we find that the intermediate SiO2 layer thickness increases significantly from 8 nm (no in-situ oxidation) to 43 nm, while the PSG layer thickness remains constant at about 8 nm. This thick SiO2 layer seems to hinder the diffusion of phosphorus atoms from the PSG through the SiO2 layer into the silicon. Implementation of a second deposition step with active N2-POCl3 flow at the end of another diffusion process type increases the thickness of the PSG layer from 14 nm to 25 nm, while at the same time the intermediate SiO2 layer thickness decreases by about 40% to 5 nm. The total phosphorus dose within the PSG/SiO2 stack layer thereby increases by a factor of three and is determined to be about 1.7·10¹⁵ cm⁻². These findings make the approach of a second deposition step very interesting for, e.g., laser doping applications to form selective emitters. For diffusion processes with different N2-POCl3 to oxygen (N2-POCl3:O2) ratios during deposition, we find only small influences on the PSG/SiO2 stack layers present after drive-in. However, the resulting emitter sheet resistances are strongly impacted by the N2-POCl3:O2 ratio; the measured sheet resistances continuously decrease from 343 Ω/sq for a N2-POCl3:O2 ratio of 0.8 to 72 Ω/sq for a ratio of 4.0.
... The wafers obtain an alkaline texture on both sides and co-diffused B emitters from doped plasma-enhanced chemical vapor deposition (PECVD) layers. The B emitter and the P front surface field are formed in one high temperature diffusion step (co-diffusion) in a POCl 3 tube furnace using a two-step process similar to the one described in [7][8][9]. Four different co-diffusions for emitter formation are conducted to investigate the impact of the Ag/Al spikes on the solar cell characteristics for different emitter profiles. Sheet resistances R SH measured by a four point probe setup are in the range of 40-70 /sq. ...
Article
Full-text available
In this study we investigate metal spike formation of screen-printed Ag/Al pastes during contact firing in an infrared belt furnace and its influence on the characteristics of n-type bi-facial silicon solar cells. The boron emitters are formed in a co-diffusion step using boron doped PECVD layers. It is demonstrated that the formation of Ag/Al spikes results in strong FF and VOC losses limiting the solar cell efficiency. This can mainly be attributed to an increased saturation current density of the second diode which is strongly increasing with increasing set peak firing temperature. A detailed scanning electron microscopy analysis reveals that this j02 increase can be attributed to an increasing area density and depth of the Ag/Al spikes for increasing peak firing temperatures.
... Wehmeier et al. simulated a diffusion profile generated from a BSG source deposited by ICP-PECVD using SiH 4 , B 2 H 6 and N 2 O as precursor gases. They detected a small difference in diffusion profile comparing a BBr 3based diffusion with the one from a CVD BSG [12,13]. Investigation of the B concentration profile within the CVD BSG showed for optimum PECVD parameters with regard to solar cell efficiency an increased concentration at the interface to Si. Averaged B concentrations of ICP-PECVD with CO 2 as oxygen source are slightly higher than those from N 2 O as precursor gas. ...
Article
Full-text available
This review presents the current state of the art and interesting questions with regard to CVD BSG layers. The advantages of CVD doping sources over the conventional POCl3 and BBr3 or BCl3 gaseous sources are the simple way to deposit a diffusion source on only a single side of the wafer and structuring the diffusion source to achieve dopant concentration profiles next to each other on the same side of the wafer. In addition, these CVD glasses are multifunctional. The same CVD BSG can serve e. g. as doping source, passivation layer, antireflective coating and as electrical insulator.
... During the codiffusion, the capping layer functions as a diffusion barrier against phosphorus. When an uncapped PECVD BSG layer is exposed to co-diffusion, a P in-diffusion disturbs the boron emitter formation as reported in [15]. ...
Article
The manufacturing process of Passivated Emitter and Rear Totally diffused (PERT) solar cells on n-type crystalline silicon is significantly simplified by applying multifunctional layer stacks acting as diffusion source, etching and diffusion barrier. We apply boron silicate glasses (BSG) capped with silicon nitride (SiNz) layers that are deposited by means of plasma enhanced chemical vapor deposition (PECVD). Optimum PECVD deposition parameters for the BSG layer such as the gas flow ratio of the precursor gases silane and diborane SiH4/B2H6=8% and the layer thickness of 40 nm result in a boron diffusion with saturation current density J0,B below 10 fA/cm2 applying an AlOx/SiNy passivation and firing. The PECVD BSG diffusion source is integrated into the n-type PERT back junction (BJ) solar cell process with screen-printed front and rear contacts. The only high temperature step is a POCl3 co-diffusion for the formation of the boron emitter from the PECVD BSG layer and for the formation of the phosphorus-doped front surface field (FSF). An independently confirmed energy conversion efficiency of 21.0% is achieved for a 156×156 mm2 large n-PERT BJ cell with this simplified process flow. This is the highest efficiency reported for a large-area co-diffused n-type PERT BJ solar cell using a PECVD BSG as diffusion source. For comparison, reference n-type PERT BJ cells with separate POCl3 and BBr3 diffusions reach an efficiency of 21.2% in our lab. A synergistic efficiency gain analysis (SEGA) for the co-diffused n-PERT BJ cell shows that the main possible efficiency gain of 1.1%abs. originates from recombination in the phosphorus-diffused front surface field while the PECVD BSG boron-doped emitter accounts for only 0.1%abs. efficiency gain. We evaluate the use of the PECVD BSG/SiNz stack as a rear side passivation as a replacement of the AlOx/SiNy stack in order to further simplify the process flow. We obtain J0,B values of 40 fA/cm2, an implied open-circuit voltage of 682 mV and a simulated n-PERT BJ cell efficiency of 21.1%.
Article
In this paper a simplified approach for the generation of laterally p‐ and n‐doped structures applicable for cost‐effective production of interdigitated back contact (IBC) solar cells is presented. We use a stack of doping glasses deposited by atmospheric pressure chemical vapor deposition (APCVD), consisting of borosilicate glass (BSG) and phosphosilicate glass (PSG) on Czochralski‐grown (Cz) silicon substrates. A laser process creates the p‐doped regions by local liquid phase diffusion of boron from the BSG layer into the underlying molten Cz‐Si substrate. Simultaneously, the BSG‐PSG stack is removed by laser ablation. In a subsequent high‐temperature step, phosphorus diffuses from the remaining PSG‐BSG layer into the crystalline silicon substrate under inert gas atmosphere, creating complementary to laser doped areas n+‐doped regions. By the use of APCVD, phosphorus and boron contents of the doping glasses can be adjusted freely to vary the resulting p‐ and n‐doped profiles. A higher boron content in the BSG layer enhances the diffusion of phosphorus through the BSG, especially at lower diffusion temperatures. The resulting doping profiles are characterized using electrochemical capacitance‐voltage measurements and the resulting sheet resistances using the four‐point probe method. The amount of minority dopant contamination in n‐ and p‐doped regions is investigated by secondary ion mass spectrometry. Furthermore, transfer length method (TLM)‐measurements indicate contactability of the generated doped regions. In this work, a novel and simplified approach for the generation of laterally p‐ and n‐doped structures applicable for cost‐effective production of interdigitated back contact (IBC) solar cells is presented. A chemical vapour deposition (CVD) doping glass stack consisting of BSG and PSG is used to create the p‐doped regions by laser doping and simultaneous laser ablation. In a subsequent high temperature step, phosphorus diffuses from the remaining PSG‐BSG stack and forms the n‐doped regions complementary to the p‐doped regions.
Article
The diffusion profiles of the front floating emitter (FFE) and front surface field (FSF) in a bifacial interdigitated back contact solar cell are optimized. The optimization results revealed that the FFE and FSF schemes are beneficial for enhancing the cell performance at the front and rear sides, respectively. Lighter doping is particularly better for the FSF scheme, and the FFE scheme requires a large diffusion depth for improving the performance. Increasing the area of the rear emitter boosts the performance of the cell, and an FFE scheme with 90% rear emitter area is found to be the best design. Quantum efficiency mapping demonstrated that the FFE scheme suppresses the loss at the back surface field region, thereby enhancing the performance of the total cell. The FSF scheme improves the quantum efficiency for the entire region by enhancing the carrier transport in the vertical direction. Furthermore, loss analysis revealed that the FFE scheme suppresses the recombination loss at the maximum power point, which is an important advantage over the FSF scheme.
Article
Full-text available
We investigate boron silicate glasses (BSG) deposited by plasma-enhanced chemical vapor deposition (PECVD) as a boron diffusion source on n-type wafers for the simplified fabrication of crystalline Si solar cells by the codiffusion processes. By varying the SiH 4/B 2H 6 gas flow ratio and the layer thickness of the PECVD BSG layers, we obtain sheet resistivities in a wide range from 30 to 500 Ω/□ after thermal B drive-in. Emitter saturation current densities as low as J0e = 4 fA/cm2 (for Rsheet = 236 Ω/□) are demonstrated using PECVD BSG layers as diffusion sources. A boron concentration in the PECVD BSG of up to 6.4 × 1021 cm−3 is measured by plasma profiling time-of-flight mass spectrometry (PP-TOFMS). A process simulation model of the B diffusion from the PECVD BSG into the Si substrate reproduces the experimental B concentration profile in the Si, measured both by PP-TOFMS and electrochemical capacitance–voltage (ECV) measurements. We fabricate industrial-type passivated emitter and rear totally diffused back-junction (PERT BJ) solar cells, as well as back-junction back-contact cells on n-type wafers. Applying codiffusion from PECVD BSG layers and thus a lean process flow including only one high-temperature step, we demonstrate n-PERT BJ cells with conversion efficiencies of up to 19.85%.
Article
Full-text available
Current issues of numerical modeling of crystalline silicon solar cells are reviewed. Numerical modeling has been applied to Si solar cells since the early days of computer modeling and has recently become widely used in the photovoltaics (PV) industry. Simulations are used to analyze fabricated cells and to predict effects due to device changes. Hence, they may accelerate cell optimization and provide quantitative data e.g. of potentially possible improvements, which may form a base for the decision making on development strategies. However, to achieve sufficiently high prediction capabilities, several models had to be refined specifically to PV demands, such as the intrinsic carrier density, minority carrier mobility, recombination at passivated surfaces, and optical models. Currently, the most unresolved issue is the modeling of the emitter layer on textured surfaces, the hole minority carrier mobility, Auger recombination at low dopant densities and intermediate injection levels, and fine-tuned band parameters as a function of temperature. Also, it is recommended that the widely used software in the PV community, PC1D should be extended to Fermi-Dirac statistics. KeywordsSilicon solar cells–Numerical modeling–Device simulation–Raytracing
Article
The intrinsic carrier density in silicon has been measured by a novel technique based on low‐frequency capacitance measurements of a p<sup>+</sup>‐i‐n<sup>+</sup> diode biased in high injection. The major advantage of the method is its insensitivity to uncertainties regarding the exact values of the carrier mobilities, the recombination parameters, and the doping density. The intrinsic carrier density was measured in the temperature range from 78 to 340 K. At 300 K the value of n i was found to be (9.7±0.1)×10<sup>9</sup> cm<sup>-3</sup>.
Article
From lifetime measurements, including a direct experimental comparison with thermal Si O <sub>2</sub> , a- Si : H , and as-deposited a- Si N <sub>x</sub>: H , it is demonstrated that Al <sub>2</sub> O <sub>3</sub> provides an excellent level of surface passivation on highly B-doped c- Si with doping concentrations around 10<sup>19</sup> cm <sup>-3</sup> . The Al <sub>2</sub> O <sub>3</sub> films, synthesized by plasma-assisted atomic layer deposition and with a high fixed negative charge density, limit the emitter saturation current density of B-diffused p<sup>+</sup> -emitters to ∼10 and ∼30 fA / cm <sup>2</sup> on ≫100 and 54 Ω/ sq sheet resistance p<sup>+</sup> -emitters, respectively. These results demonstrate that highly doped p -type Si surfaces can be passivated as effectively as highly doped n -type surfaces.
Article
The threshold voltage of MOSFET devices can be effectively stabilized from changes due to field-assisted motion of Na<sup>+</sup>in the gate oxide by the addition of a phosphosilicate glass (PSG) layer. The effectiveness of the glass for this purpose is markedly enhanced by increasing the P 2 O 5 concentration of the PSG. However, polarization of the PSG layer can, in turn, cause an appreciable instability of the threshold voltage. It is shown that detailed knowledge of the behavior of PSG layers permits prediction of the threshold stability of P 2 O 5 -treated FET devices. Thus, threshold stability can be maintained to within 0.1 V/1000 Å under device operating conditions by making a proper compromise on PSG thickness and P 2 O 5 concentration. Such stabilizing films offer satisfactory protection against realistic Na<sup>+</sup>contamination levels. Quantitative data on these phenomena are presented, and a simple structural model is given to account for the polarization and the Na<sup>+</sup>trapping behavior of the films. The formation of PSG films by doping of SiO 2 with P 2 O 5 at elevated temperatures is discussed.
  • P Rothardt
  • R Keding
  • A Wolf
  • D Biro
P. Rothardt, R. Keding, A. Wolf, D. Biro, Phys. Status Solidi RPL (2013) 1-4.
  • P P Altermatt
P.P. Altermatt et al., J. Appl. Phys. 3187 (2002) 92.
  • A Bentzen
A. Bentzen et al., J. Appl. Phys. 99 (2006) 064502.
  • M Kessler
  • T Ohrdes
  • B Wolpensinger
  • N.-P Harder
M. Kessler, T. Ohrdes, B. Wolpensinger, N.-P. Harder, Semicond. Sci. Technol. 25 (2010) 055001.
  • A Richter
A. Richter et al., IEEE J. PV3 (2013) 236.