Sie sind auf Seite 1von 6

Solar Energy Materials and Solar Cells 189 (2019) 21–26

Contents lists available at ScienceDirect

Solar Energy Materials and Solar Cells


journal homepage: www.elsevier.com/locate/solmat

Passivation quality control in poly-Si/SiOx/c-Si passivated contact solar cells T


with 734 mV implied open circuit voltage
HyunJung Parka, Hyomin Parka, Se Jin Parka, Soohyun Baea, Hyunho Kima, Jee Woong Yanga,
Ji Yeon Hyuna, Chang Hyun Leea, Seung Hyun Shina, Yoonmook Kangb, Hae-Seok Leeb, ,

Donghwan Kima
a
Department of Material science and Engineering, Korea University, 145, Anam-ro, Seongbuk-gu, Seoul 02841, Korea
b
KU-KIST Green School Graduate School of Energy and Environment, Korea University, 145, Anam-ro, Seongbuk-gu, Seoul 02841, Korea

ARTICLE INFO ABSTRACT

Keywords: Passivation quality of poly-Si contacts with different phosphorus doping concentration were investigated in this
Polysilicon study. Intrinsic poly-Si layers were deposited by LPCVD on a tunnel oxide surface, followed by n + poly-Si
Tunnel oxide doping and hydrogenation. For lightly doped poly-Si contacts with phosphorus concentration of 2.1 × 1019 cm 3 ,
Passivated contact higher temperatures and longer times increased iVOC achieving maximum value of 734 mV, as poly-Si grain size
N-type silicon
increases from 13 nm to 40 nm. However, for heavily doped poly-Si contacts with phosphorus concentration of
High efficiency
1.1 × 1020 cm 3 , iVOC decreased from 731 mV to 696 mV as annealing time increased from 10 to 60 min because
Solar cell
Auger recombination rate increased from 9.3 fA/cm2 to 21.6 fA/cm2 as phosphorus in-diffusion occurs. The
contact resistance of poly-Si contacts was also investigated to achieve a high fill factor. Finally, a poly-Si/SiOx/c-
Si passivated contact solar cell using a poly-Si contact on the back and boron diffused emitter on the front was
fabricated. As a result, high efficiency of 21.1% solar cell was achieved with VOC of 665 mV, JSC of 40.6 mA/cm2,
and fill factor of 78.3%.

1. Introduction efficiency of 21.2% [6–10]. About iVOC values, about 730 mV was ob-
tained using crystallized a-Si passivated contact [9,11] and LPCVD
Recombination losses at the interface between the metal contact and deposited poly-Si passivated contact [12]. The highest iVOC of 749 mV
doped-semiconductor region is a major efficiency-limiting factor in was reported by ISFH [13] for the latter with phosphorus implantation.
solar cells. To reduce losses from metal-semiconductor contact re- Although many groups involved poly-Si contact, the relation between
combination, the concept of passivated contact was introduced. A the poly-Si layer and passivation is still unclear because researchers
tunnel oxide passivated contact (TOPCon) solar cell [1] is composed of focused on the tunnel oxide quality. However, characteristics of the
a n + poly-Si/SiOx/c-Si contact on the rear side and boron doped poly-Si layer also affect on passivation quality of poly-Si contacts ac-
emitter on the front side, and has a reported highest efficiency of 25.8% cording to its annealing condition and doping concentration. Thus, we
[2,3]. One major method to form poly-Si/SiOx/c-Si passivated contacts focused our study, to better understand the effect of the poly-Si layer on
is to deposit an amorphous silicon layer (a-Si) and crystallized a-Si onto passivation quality, on the grain-size growth and phosphorus in-diffu-
poly-Si with subsequent annealing. Another way is to use the low- sion. High efficiency was achieved with a tunnel-oxide-passivated
pressure chemical vapor deposition method (LPCVD) for poly-Si de- contact solar cell with a homogeneous boron emitter on the front and
position, which provides better crystallinity. Using LPCVD-deposited poly-Si contact on the rear.
poly-Si/SiOx/c-Si passivated contact (hereafter, poly-Si contact), some For the experiments, the poly-Si contacts were annealed at different
groups reported poly-Si contact solar cells using POCl3 diffusion temperatures and times, and with different phosphorus doping con-
method and the highest efficiency was 21.5% [4,5]. Also, some groups centrations, in the poly-Si layer. Passivation quality was determined
reported poly-Si contact solar cells formed by ion implantation method using iVOC and J0 measured by the quasi-steady-state photoconductance
to form n + doping on i-poly-Si layer which achieved the highest (QSSPC) method [14]. The J0 values are determined under the high


Corresponding author.
E-mail addresses: jung1029@korea.ac.kr (H. Park), solar@korea.ac.kr (H.-S. Lee).
URL: http://solar.korea.ac.kr (D. Kim).

https://doi.org/10.1016/j.solmat.2018.09.013
Received 25 October 2017; Received in revised form 25 June 2018; Accepted 10 September 2018
0927-0248/ © 2018 Elsevier B.V. All rights reserved.
H. Park et al. Solar Energy Materials and Solar Cells 189 (2019) 21–26

injection assumption. The relationship between passivation and an-


nealing temperature and time was analyzed using the grain-size growth
effect. The size of the grains of the poly-Si layer was measured using
grazing-incidence X-ray diffraction (GI-XRD) and Scherrer's equation
[15]. The impact of the doping concentration of the poly-Si layer on
phosphorus in-diffusion and contact resistance was also investigated.
Finally, a solar cell with a poly-Si contact at the rear was fabricated and
the cell efficiency was measured.

2. Experimental method

Symmetric poly-Si contact samples for lifetime measurement were


prepared on 6-in. 180 μm-thick n-type Cz wafers with resistances of
1 2 ·cm . Saw damage on the wafer surfaces were etched using KOH
solution. A thin silicon oxide layer for tunneling oxide with a thickness
of 1.2 nm was grown on the wafer surface by dipping the wafer into
H2O2 solution at 80 °C for 10 min [16]. After tunnel oxide formation, an
intrinsic poly-silicon layer with thickness of 450 nm was deposited on
both sides of the samples using LPCVD with silane precursor gas at
600 °C. To dope the intrinsic poly-Si layer, two different phosphor-si-
licate glasses (PSG) with doping concentrations of 2.8 × 10 21 cm 3 and
4.6 × 10 20 cm 3 were formed on the wafer surface. The PSG with a low
doping concentration was deposited using plasma-enhanced chemical
vapor deposition (PECVD) at 310 °C, and the PSG with a high doping Fig. 1. Experimental procedures to fabricate symmetrical tunnel oxide passi-
concentration was formed by POCl3 diffusion in a tube furnace. The vated contact samples.
subsequent annealing at 800–950 °C for 10–60 min was conducted in
the same tube furnace used for the POCl3 diffusion. Hereafter we call
the samples heavily doped (HD), lightly doped (LD), and undoped or
intrinsic (UD) poly-Si contacts, respectively. The effects of annealing on
poly-Si contacts passivation quality were investigated by measuring
iVOC , J0 , and on the lifetime using the QSSPC method after hydrogena-
tion. For the hydrogenation at the interface of n + poly-Si/SiOx/c-Si,
silicon nitride films were deposited using PECVD and subsequently
annealed in the rapid thermal process chamber (RTP) at 600 °C for
15 min. Silicon nitride layer was removed by HF dipping. We measured
the average grain size of the poly-Si layer using GI-XRD and high-re-
solution electron backscatter diffraction (HR EBSD). Doping con-
centration profiles from the surface to the crystalline silicon substrate
were measured using secondary ion mass spectroscopy (SIMS), and
contact resistance was determined using transmission line measurement
(TLM) on the lifetime measurement samples after QSSPC measurement.
Finally, a solar cell structure with a poly-Si contact on the back and B
diffused emitter with Al2O3/SiNx stack passivation layer on the front
was fabricated and the efficiency was measured Fig. 1.

3. Results and discussions

3.1. Poly-Si passivation quality as a function of annealing condition

First, the effect of annealing temperature on UD and LD poly-Si


contacts was investigated. The as-deposited (as-dep) poly-Si/SiOx/c-Si
sample before annealing showed a very low iVOC of 580 mV. As shown
in Fig. 2 (a), after undergoing the hydrogenation process without an-
nealing, the iVOC of this as-dep sample increased to 665 mV. However,
this iVOC value was still low compared with the symmetric UD sample
annealed at 875 °C without hydrogenation, which achieved an iVOC of
673 mV. This result indicates that the as-dep poly-Si contacts structure
cannot achieve interface passivation without hydrogenation or an an-
nealing process. From Fig. 2 (a), the iVOC of the UD poly-Si contacts
increased from 658 mV to 694 mV as annealing temperature increased
from 800° to 950°C. The same result was found for the LD poly-Si
contacts sample: the iVOC increased as annealing temperature increased.
Fig. 2. Implied open circuit voltage and effective lifetime of poly-Si contacts as
This two results indicate that the poly-Si contacts passivation quality is
a function of (a) annealing temperature and (b) time.

22
H. Park et al. Solar Energy Materials and Solar Cells 189 (2019) 21–26

correlated with annealing temperature, and that the high-temperature size of the poly-Si was estimated as 20 nm for the LD poly-Si contacts
annealing process is very important for achieving high-quality passi- sample annealed at 950 °C for 10 min. Also, the grain size for UD, LD
vation of the poly-Si contacts solar cell. and HD contact samples increased from 13 nm to 22 nm as annealing
Second, we changed the annealing time at a fixed temperature of temperature increased from 800° to 950°C. This result matches well
950 °C, because the highest iVOC was achieved at 950 °C, as shown in with Harbeke's result [17], which indicates that the grain size of the as-
Fig. 2 (a). As shown in Fig. 2 (b), with increase in the annealing time dep polysilicon layer deposited at 600 °C was 11 nm and it grew after
from 10 to 60 min, iVOC increased from 721 mV to 734 mV for the LD annealing at 900–1000 °C. The final grain size of polysilicon annealed
poly-Si contacts. However, the iVOC of the HD poly-Si contacts sample at 1000 °C was 18 nm from Harbeke's result [17], thus indicating a si-
decreased as the annealing time increased. This result indicates that milar grain size and growth effect in our experiments. And also, the
higher temperature and longer annealing time improves passivation result of increased grain size indicates the increased iVOC of the UD poly-
quality after low-concentration poly-Si doping. To analyze the cause of Si contact was caused by the effect of the grain size growth.
this difference in behavior, we focused on the grain size of the poly-Si Increasing the annealing time induced poly-Si grain growth. The
layer and phosphorus in-diffusion, and the effect of these factors on grain sizes of the poly-Si contacts annealed for 60 min at 950 °C were
passivation quality. twice as large as the poly-Si contacts sample annealed for 10 min
The average grain size of the poly-Si layer was measured using the (20–40 nm). Increasing the grain size of the poly-Si contacts sample
GI-XRD method and Scherrer's equation [15] resulted in an increase in iVOC of the LD poly-Si contacts. This result
indicates that the grain size of the poly-Si layer induced a high inter-
0.9
t= face-passivation quality in the poly-Si contacts between poly-Si and
Bcos (1)
SiOx while P in-diffusion decreases the iVOC of the HD poly-Si contact.
where t is the average grain size, is the wavelength of x-ray, B is full The grain size effect on passivation at the poly-Si/SiOx interface was
width half maximum and is the radians. As a result, the average grain analyzed by two physical reason. First, the energy band changed at the

Fig. 3. (a) 3D schematic of the energy band diagram for the poly-Si/SiOx/c-Si poly-Si contacts structure and 2D schematic of the energy band diagram at the interface
between poly-Si/SiOx/c-Si grains and their boundaries [18] and (b) measured J0 as a function of average grain size for LD poly-Si contacts and simulation result.

23
H. Park et al. Solar Energy Materials and Solar Cells 189 (2019) 21–26

poly-Si grain boundaries with SiOx, compared with those within the poly-Si contacts but slightly increased for LD poly-Si contacts. From the
grains, as shown in Fig. 3(a). At the grain region, the schematic energy simulation result and lifetime curve, the phosphorus in-diffusion is
band diagram is the same as the energy band of an n-type crystalline partially responsible for the Auger recombination and lowered iVOC . The
silicon/SiOx interface. However, at poly-Si grain boundary regions, the difference between the measured J0 and the calculated J0 Auger can be
energy band changes because electrons are depleted to the defect levels induced from the increase in the surface recombination caused by the
that exist at the grain boundaries. This causes the Fermi level of the
grain boundary to decrease and align with the intrinsic Fermi level.
Because of the different energy band structure, a barrier height for
electrons is formed and holes, which are minority carriers, can tunnel
through the tunnel oxide layer. Moreover, holes tunneling through
oxide move to the poly-Si grain boundaries and recombine with elec-
trons captured at the interface defect of the grain boundary. Thereby,
the poly-Si grain size affects the passivation quality at the poly-Si/SiOx
interface.
To analyze the relationship between grain size and the recombina-
tion rate of the poly-Si contacts solar cell, we made a simple assumption
that each grain is rectangular-shaped with square of length d and depth
t, resulting in a grain volume of d 2· t . From the assumption, interface
defect produced from the grain boundary is characterized by defect
concentration at the interface of SiOx/Si dangling bonds per unit length
(Ndis ) and grain size (d) and equation of interface defect becomes
Ndis
Dit =
d (2)

Because recombination current density J0 is proportional to inter-


face defect density, J0 should be proportional to 1/d if the grain
boundaries act as recombination centers according to

qni2
Jo = Seff
ND (3)

Seff = eff vth Dit (4)

From this calculation we confirmed that the average grain size of


poly-Si was inversely proportional to the recombination current den-
sity, as shown in Fig. 3 (b). This can be taken as evidence that the defect
states at the grain boundary area acted as recombination sites, and that
passivation quality can be improved by the growth in grain size. The
passivation quality of the poly-Si contacts can potentially be influenced
by other factors, including the silicon oxide interface density [19–21]
and the pin-hole density of the interface oxide layer [22]. Although we
focused on the homogeneous tunneling current that changes at the
grain, and the fact that the grain boundary and grain size affected the
passivation quality, there are other explanations that can account for
carrier transport in poly-Si passivated contact such as the localized
carrier transportation mechanism [23,24]. To understand these results,
a more detailed investigation to address the cause of the carrier
movement in poly-Si passivated contact is needed.

3.2. Poly-Si passivation quality as a function of phosphorus doping


concentration

To analyze the effect of phosphorus in-diffusion, the annealing ef-


fect of poly-Si contacts with different doping concentration was in-
vestigated. To understand the effect of phosphorus in-diffusion on iVOC ,
The relationship between phosphorus in-diffusion and recombination
rate was inferred using the PV Lighthouse EDNA 2 simulation tool [25]
to calculate J0 Auger in the doping profile of the phosphorus in-diffusion
region from the SiOx/c-Si interface to the substrate. Here, we used the
SIMS profile provided in Fig. 4 for simulation. From the simulation
result, the calculated J0 Auger for the HD poly-Si contacts increased from
9.3 to 21.6 fA/cm2 in Fig. 4 (e). This result indicates that the Auger
recombination current density can be increased because of phosphorus
in-diffusion. In addition, the lifetime graph was investigated to under- Fig. 4. (a) Phosphorus and oxygen doping concentration or intensity as a
stand the effect of phosphorus in-diffusion on Auger recombination. As function of depth measured by SIMS, and (b) its effective lifetime graph as a
shown in Fig. 4 (c), (d), the lifetimes under high injection conditions function of injection level, and (c) measured and simulated J0 values (annealing
(minority carrier density (MCD) = 2 × 1016 cm 3 ) decreased for HD temperature was fixed at 950 °C).

24
H. Park et al. Solar Energy Materials and Solar Cells 189 (2019) 21–26

poly-Si annealed under the same conditions on the back. On the front of
the solar cell, a homogeneous boron-diffused emitter was applied and
Al2O3/SiNx stack layer was deposited on the front surface as passivation
layers. For metallization, the evaporation method was used for front
and back metal contacts.
Light I-V measurements were made of the solar cell, and the results
are shown in Fig. 6. The solar cell showed 21.1% efficiency with a VOC
of 665 mV, JSC of 40.6 mA/cm2, and fill factor of 78.3% with 2 × 2 cm2
cell size. The difference between implied VOC and cell VOC is expected to
be further reduced by the optimization of the emitter, front passivation,
and metallization process. The limitation of VOC at the cell level is ex-
pected to 725 mV based of the QUOKKA simulation program [28].

4. Summary

To achieve a high-quality passivation poly-Si contact solar cell, the


annealing effects on poly-Si/SiOx/c-Si structure were investigated.
Fig. 5. Contact resistance of LD and HD poly-Si contacts structures as a function
of annealing time. Different annealing conditions, including annealing temperature, time,
and doping concentration at the poly-Si layer were applied. The iVOC

Fig. 6. (a) Schematic illustration of solar cell fabricated using HD poly-Si contact annealed at 950 °C for 30 min structure and (b) corresponding light I-V curve.

SiOx interface defect density and the oxide break-up, since the latter can results showed that the passivation quality of the poly-Si contacts im-
be enhanced when the doping concentration at poly-Si increases [26]. proved with annealing at higher temperature (950 °C) and longer time
Further, the increase in the dopant concentration in the poly-Si layer (60 min). This improvement is due to a grain growth effect on passi-
also affect the barrier height between the poly-Si/SiOx/c-Si layers, with vation quality because the grain size of poly-Si affects the recombina-
an increased barrier height from 0.217 eV to 0.259 eV by increasing the tion rate. However, high doping concentrations of poly-Si result in in-
dopant concentration from 2.1 × 1019 cm 3 (LD) to 1.1 × 1020 cm 3 (HD). creasing phosphorus in-diffusion, with increase the annealing time,
This may affect the interface passivation quality, but a deeper and more which is different from the case of low-concentration doped poly-Si
detailed study is required to better understand physical relationship contacts. This result was learned after simulation of the recombination
between the barrier height and passivation quality. current density from Auger recombination. Moreover, the lifetime
curve under the high-injection condition also decreased, meaning that
3.3. Fabrication of solar cell using poly-Si contact the Auger recombination rate increased. The conclusion is that high
temperature and long-time annealing are required to achieve large
To fabricate a poly-Si/SiOx/c-Si passivated contact solar cell using grain size and high iVOC in cases where critical degradation from
the poly-Si contacts, contact resistances were measured using Ag eva- phosphorus in-diffusion does not occur. Based on these experimental
porated TLM patterns with reactive ion etching to remove poly-Si re- results, a solar cell was fabricated, and achieved an efficiency of 21.1%.
gion without being covered by Ag. The contact resistance of poly-Si/
SiOx/c-Si was analyzed using a 2D TLM method [27].
The results indicate that the contact resistance of the LD poly-Si Acknowledgement
contacts decreased as annealing time increased to 394.27 m cm2 , as
shown in Fig. 5. Better contact characteristics were achieved with lower This work was supported by the New & Renewable Energy Core
contact resistance for HD poly-Si contacts, achieving contact resistance Technology Program of the Korea Institute of Energy Technology
below 46.31 m cm2 . The lowest contact resistance was 21.85 m cm2 , Evaluation and Planning (KETEP), which was granted financial re-
achieved for HD poly-Si contacts annealed at 950 °C for 60 min. sources from the Ministry of Trade, Industry, and Energy, Republic of
Given the optimal contact resistance and high iVOC of 731.5 mV for Korea (Nos. 20163030014020 and 20154030200760). This work was
the HD poly-Si contacts annealed at 950 °C for 30 min, we fabricated a also supported by the KU-KIST Graduate School Project, Republic of
solar cell applying the poly-Si contact structure with heavily doped Korea.

25
H. Park et al. Solar Energy Materials and Solar Cells 189 (2019) 21–26

Appendix A. Interface defect density of grain boundary

To drive the equation between the grain size of the poly-Si layer and the interface recombination rate, we made the simple assumption that each
grain is rectangular-shaped with a square of length d and depth t, resulting in a grain volume of d 2· t . The total area of the sample was set as A. The
grain boundary region at the interface is the same as the product of the grain boundary length of one grain and number of grains in the total area.
Since each grain is surrounded by four edges (length of 4d ) and the edges are shared with four adjacent grains, the grain boundary length of one grain
has a length of 2d . The number of grains are equal to the total volume divided by the grain volume, which is At / d 2t . Then, the grain boundary length
becomes A/ d and by dividing by the total area, the grain boundary length per unit area (LGB ) becomes;
At 2
LGB = 2d × ÷A=
d 2t d (A.1)
From the above equation, the grain boundary length at the interface of SiO x/ poly-Si is inversely proportional to the grain size. Since we assumed
that the grain boundary act as interface recombination centers, we set the constant NGB as the defect density per grain boundary length (NGB ). As a
result, the equation for the interface defect produced from the grain boundary (Dit ) becomes,
NGB
Dit =
d (A.2)

References [14] R.A. Sinton, A. Cuevas, Contactless determination of current-voltage characteristics


and minority-carrier lifetimes in semiconductors from quasi-steady-state photo-
conductance data, Appl. Phys. Lett. 69 (17) (1996) 2510–2512.
[1] K. Masuko, M. Shigematsu, T. Hashiguchi, D. Fujishima, M. Kai, N. Yoshimura, [15] P. Scherrer, Bestimmung der gröss kolloidteilchen mittels nachrichten von der ge-
T. Yamaguchi, Y. Ichihashi, T. Mishima, N. Matsubara, et al., Achievement of more sellschaft der wissenschaften, göttingen, Math. Phys. Kl. 2 (1918) 98–100.
than 25% conversion efficiency with crystalline silicon heterojunction solar cell, [16] H. Kim, S. Bae, K.-s. Ji, S.M. Kim, J.W. Yang, C.H. Lee, K.D. Lee, S. Kim, Y. Kang, H.-
IEEE J. Photovolt. 4 (6) (2014) 1433–1435. S. Lee, et al., Passivation properties of tunnel oxide layer in passivated contact si-
[2] A. Richter, J. Benick, F. Feldmann, A. Fell, M. Hermle, S.W. Glunz, n-type si solar licon solar cells, Appl. Surf. Sci. 409 (2017) 140–148.
cells with passivating electron contact: identifying sources for efficiency limitations [17] G. Harbeke, L. Krausbauer, E. Steigmeier, A. Widmer, H. Kappert, G. Neugebauer,
by wafer thickness and resistivity variation, Sol. Energy Mater. Sol. Cells 173 (2017) Growth and physical properties of lpcvd polycrystalline silicon films, J.
96–105. Electrochem. Soc. 131 (3) (1984) 675–682.
[3] M.A. Green, Y. Hishikawa, E.D. Dunlop, D.H. Levi, J. Hohl-Ebinger, A.W. Ho-Baillie, [18] Alan Fahrenbruch, Richard Bube, Fundamentals of Solar Cells: Photo Voltaic Solar
Solar cell efficiency tables (version 51), Progress. Photovolt.: Res. Appl. 26 (1) Energy Conversion, Academic Press, New York, USA, 1983.
(2017) 3–12. [19] K. Lancaster, S. Großer, F. Feldmann, V. Naumann, C. Hagendorf, Study of pinhole
[4] U. Römer, R. Peibst, T. Ohrdes, B. Lim, J. Krügener, E. Bugiel, T. Wietler, R. Brendel, conductivity at passivated carrier-selected contacts of silicon solar cells, Energy
Recombination behavior and contact resistance of n + and p + poly-crystalline si/ Procedia 92 (2016) 116–121.
mono-crystalline si junctions, Sol. Energy Mater. Sol. Cells 131 (2014) 85–91. [20] D. Tetzlaff, J. Krügener, Y. Larionova, S. Reiter, M. Turcu, R. Peibst, U. Höhne, J.-D.
[5] M.K. Stodolny, J. Anker, B.L. Geerligs, G.J. Janssen, B.W. van de Loo, J. Melskens, Kähler, T. Wietler, Evolution of oxide disruptions: The (w) hole story about poly-si/
R. Santbergen, O. Isabella, J. Schmitz, M. Lenes, et al., Material properties of lpcvd c-si passivating contacts, in: IEEE 44th Proceedings of the Photovoltaic Specialist
processed n-type polysilicon passivating contacts and its application in perpoly Conference (PVSC), IEEE, 2017, pp. 1–4.
industrial bifacial solar cells, Energy Procedia 124 (2017) 635–642. [21] T. Wietler, D. Tetzlaff, J. Krügener, M. Rienäcker, F. Haase, Y. Larionova,
[6] C. Reichel, F. Feldmann, R. Müller, R.C. Reedy, B.G. Lee, D.L. Young, P. Stradins, R. Brendel, R. Peibst, Pinhole density and contact resistivity of carrier selective
M. Hermle, S.W. Glunz, Tunnel oxide passivated contacts formed by ion im- junctions with polycrystalline silicon on oxide, Appl. Phys. Lett. 110 (25) (2017)
plantation for applications in silicon solar cells, J. Appl. Phys. 118 (20) (2015) 253902.
205701. [22] B.E. Deal, M. Sklar, A.S. Grove, E.H. Snow, Characteristics of the surface-state
[7] F. Feldmann, C. Reichel, R. Müller, M. Hermle, Si solar cells with top/rear poly-si charge (qss) of thermally oxidized silicon, J. Electrochem. Soc. 114 (3) (1967)
contacts, in: Proceedings of the IEEE 43rd Photovoltaic Specialists Conference 266–274.
(PVSC), IEEE, 2016, pp. 2421–2424. [23] R. Peibst, U. Römer, K.R. Hofmann, B. Lim, T.F. Wietler, J. Krügener, N.-P. Harder,
[8] G. Yang, A. Ingenito, O. Isabella, M. Zeman, Ibc c-si solar cells based on ion-im- R. Brendel, A simple model describing the symmetric i-v characteristics of p poly-
planted poly-silicon passivating contacts, Sol. Energy Mater. Sol. Cells 158 (2016) crystalline/n monocrystalline si and n polycrystalline si/p monocrystalline si
84–90. junctions, IEEE J. Photovolt. 4 (3) (2014) 841–850.
[9] G. Yang, A. Ingenito, N. van Hameren, O. Isabella, M. Zeman, Design and appli- [24] R. Peibst, U. Römer, Y. Larionova, M. Rienäcker, A. Merkle, N. Folchert, S. Reiter,
cation of ion-implanted polysi passivating contacts for interdigitated back contact c- M. Turcu, B. Min, J. Krügener, et al., Working principle of carrier selective poly-si/
si solar cells, Appl. Phys. Lett. 108 (3) (2016) 033903. c-si junctions: is tunnelling the whole story? Sol. Energy Mater. Sol. Cells 158
[10] F. Feldmann, C. Reichel, R. Müller, M. Hermle, The application of poly-si/siox (2016) 60–67.
contacts as passivated top/rear contacts in si solar cells, Sol. Energy Mater. Sol. [25] K.R. McIntosh, P.P. Altermatt, A freeware 1d emitter model for silicon solar cells, in:
Cells 159 (2017) 265–271. Proceedings of the 35th IEEE Photovoltaic Specialists Conference (PVSC), IEEE,
[11] A. Rohatgi, B. Rounsaville, Y.-W. Ok, A.M. Tam, F. Zimbardi, A.D. Upadhyaya, 2010, pp. 002188–002193.
Y. Tao, K. Madani, A. Richter, J. Benick, M. Hermle, Fabrication and modeling of [26] A. Paduthol, M.K. Juhl, Z. Hameiri, G. Nogay, P. Löper, T. Trupke, Efficient carrier
high-efficiency front junction n-type silicon solar cells with tunnel oxide passivating injection from amorphous silicon into crystalline silicon determined from photo-
back contact, IEEE J. Photovolt. 7 (5) (2017) 1236–1243. luminescence, in: Proceedings of the 33rd European Photovoltaic Solar Energy
[12] D.L. Young, W. Nemeth, V. LaSalvia, M.R. Page, S. Theingi, J. Aguiar, B.G. Lee, Conference and Exhibition, Amsterdam, Netherlands, 2017, pp. 238–241.
P. Stradins, Low-cost plasma immersion ion implantation doping for interdigitated [27] M. Rienäcker, M. Bossmeyer, A. Merkle, U. Römer, F. Haase, J. Krügener,
back passivated contact (ibpc) solar cells, Sol. Energy Mater. Sol. Cells 158 (2016) R. Brendel, R. Peibst, Junction resistivity of carrier-selective polysilicon on oxide
68–76. junctions and its impact on solar cell performance, IEEE J. Photovolt. 7 (1) (2017)
[13] R. Peibst, Y. Larionova, S. Reiter, M. Turcu, R. Brendel, D. Tetzlaff, J. Krügener, T. 11–18.
Wietler, U. Höhne, J. Kähler, et al., Implementation of n + and p + polo junctions on [28] A. Fell, K.C. Fong, K.R. McIntosh, E. Franklin, A.W. Blakers, 3-d simulation of in-
front and rear side of double-side contacted industrial silicon solar cells, in: terdigitated-back-contact silicon solar cells with quokka including perimeter losses,
Proceedings of the 32nd European Photovoltaic Solar Energy Conference and IEEE J. Photovolt. 4 (4) (2014) 1040–1045.
Exhibition, 2016, pp. 323–327.

26

Das könnte Ihnen auch gefallen