ArticlePDF Available

Abstract and Figures

This paper is motivated by the question of how residual implantation damage degrades solar cell performance. In order to avoid such degradation, annealing processes of implanted c-Si solar cells use high thermal budgets. Still, implantation-induced dislocation loops may survive these processes. We derive two models for the annealing kinetics of dislocation loops that are suitable for the study of high thermal budgets: a model that is able to describe the parallel ripening of faulted and perfect dislocation loops and a model that explicitly implements the conservative and nonconservative processes associated with Ostwald ripening. Both models lead to a better agreement with the experiment than what has been published before.
Content may be subject to copyright.
IEEE JOURNAL OF PHOTOVOLTAICS 1
Modeling the annealing of dislocation loops in
implanted c-Si solar cells
F. A. Wolf, A. Martinez-Limia, D. Stichtenoth, and P. Pichler, Senior Member, IEEE
Abstract—This paper is motivated by the question of how resid-
ual implantation damage degrades solar cell performance. In or-
der to avoid such degradation, annealing processes of implanted
c-Si solar cells use high thermal budgets. Still, implantation-
induced dislocation loops may survive these processes. We derive
two models for the annealing kinetics of dislocation loops that
are suitable for the study of high thermal budgets: A model
that is able to describe the parallel ripening of faulted and
perfect dislocation loops, and a model that explicitly implements
the conservative and non-conservative processes associated with
Ostwald ripening. Both models lead to better agreement with
experiment than what has been published before.
I. INTRODUCTION
STUDIES of implanted solar cells carried out within recent
years showed that significantly higher thermal budgets
than in microelectronics are needed to achieve satisfactory
cell performance [1, 2, 3, 4]. Typically, for such annealing
conditions (>950 C, >10 min), the primary implant damage
evolves into a defect configuration in which only dislocation
loops survive. Such loops were found to correlate with the
emitter saturation current [5, 6] although also other hypotheses
exist [1, 6]. Recently, the causation of implantation-related
performance degradation by dislocation loops, likely decorated
with metallic impurities, could be substantiated [7, 8].
The agglomeration of implantation-induced self-interstitials
(Is) via the formation of small interstitial clusters (SMICs) and
their transformation to {311}defects and even small loops has
been addressed by a variety of models, e.g. [9, 10, 11, 12, 13,
14, 15], and is well described by them. This does not hold true
for the high thermal budgets of solar cell fabrication. As we
will show later (Sec. II), even the latest of them, the model of
Zographos et al. [15] implemented in Sentaurus Process [16],
fails then to reproduce two important qualitative aspects
(i) The predicted dissolution velocity at high temperatures
is by orders of magnitude too high.
(ii) The simulated growth of the mean loop radius does not
saturate for long annealing times at high temperatures,
as would be expected from theory and experiment.
Point (i) implies that for almost all relevant annealing condi-
tions, the model of Zographos et al. [15] predicts a complete
F. A. Wolf and A. Martinez-Limia are with Corporate Research, Robert
Bosch GmbH, Robert-Bosch-Platz 1, 70839 Gerlingen-Schillerh¨
ohe, Ger-
many.
D. Stichtenoth is with Bosch Solar Energy, Arnstadt, Germany.
P. Pichler is with the Fraunhofer Institute for Integrated Systems and Device
Technology, Schottkystrasse 10, 91058 Erlangen, Germany and with the
Chair of Electron Devices, University of Erlangen-Nuremberg, Cauerstrasse
6, 91058 Erlangen, Germany.
Manuscript received April 19, 2005; revised December 27, 2012.
dissolution of loops. Thus, it cannot be used to establish cor-
relations between processing conditions and cell performance.
Point (ii) is more subtle and of lower priority: A quantitatively
wrong prediction of the mean loop radius translates directly
into a likewise wrong prediction of the proportionally related
dislocation-line density, which directly correlates to the re-
combination activity of loops [17, 18]. Two other loop models
[13, 19] implemented in Sentaurus Process [20] suffer from
even more severe problems for annealing processes with high
thermal budgets. For an in-depth discussion of the modeling of
dislocation loops the interested reader is referred to the review
article of Claverie et al. [21] and the dissertation of one of the
authors (F.A.W) [22].
In this paper, we extend the model of Zographos et al.
[15] to overcome problem (i) (Sec. III). We succeed in doing
so by accounting for the parallel ripening of faulted and
perfect dislocation loops. The new model is able to provide
meaningful information about the local densities and sizes of
insufficiently annealed dislocation loops [7, 8]. Subsequently,
we analyze the late stages of Ostwald ripening, which leads us
to sketch a model that is able to resolve problem (ii) (Sec. IV).
Finally, we conclude the paper (Sec. V). We start with a brief
discussion of the model of Zographos et al..
II. THE MODEL OF ZOGRAPHOS et al.
Zographos et al. [15] model the mean concentration of loops
DLand the concentration of self-interstitials (Is) CLcomprised
in these loops by
tCL=R311
L+RI
L,(1a)
tDL=e
R311
L+e
RI
L.(1b)
These distributions depend on time and position. The reaction
terms that describe the unfaulting reaction from {311}defects
to loops are given by
R311
L=k311
LC311C311 ,(2a)
e
R311
L=k311
Le
k311
LC311D311 ,(2b)
where D311 and C311 denote the concentration of {311}
defects and the concentration of Is bound in {311}defects,
in complete analogy to the meaning of DLand CL.k311
Lis
described by an Arrhenius law [20]. Zographos et al. [15] set
e
k311
L=1
2, which means that {311}defects with twice the
mean size unfault to loops.
The interaction of loops with free Is is described by
RI
L=kI
L2π2rLDI(CIC
I,L)DL,(3a)
e
RI
L=kI
L2π2rLDIC
I,L
DLDL
CL
=kI
L2πDI
rLnL
C
I,LDL,(3b)
IEEE JOURNAL OF PHOTOVOLTAICS 2
where CIis the concentration of free Is, kI
Lis a calibration
constant and DIthe diffusion coefficient of Is. The mean loop
radius rLand the equilibrium concentration of Is in the vicinity
of a loop C
I,L are defined by
rL=rCL
DLnLπand C
I,L =C
IeEL(rL)/kT ,(4)
where nLis the atomic areal density in the close-packed {311}
planes. For the formation energy per I in faulted dislocation
loops EL(rL), Zographos et al. [15] used the definition
EL(rL) = γ
b+Gb
4πrL(1 ν)ln(8rL/b).(5)
Therein, γis the stacking-fault energy per unit volume, the
atomic volume, bthe modulus of the Burger’s vector of the
loop, Gthe shear modulus of Si and νits Poisson ratio. With
these equations, the loop model is completely defined.
In the present paper, the focus is on the phase of the
annealing process in which the last step of the I cluster
nucleation process, the unfaulting reaction from {311}defects
to loops, is completed. At this point, almost all {311}defects
have vanished, so that R311
L'0and e
R311
L'0, and one is left
to discuss, instead of Eqs. (1), the equations tCL=RI
Land
tDL=e
RI
L.
A. Dissolution velocity of dislocation loops
The first fundamental problem that arises from the model
of Zographos et al. [15] is a qualitative overestimation of
the dissolution velocity of loops. This is illustrated in Fig. 1
for the areal density of loops NL=RtW
0dx DL(x), where
we integrated over the wafer thickness tW. We empirically
checked that this problem cannot be resolved by recalibrating
the value [16] of the sole free parameter kI
Lin Eqs. (3).
Fig. 1(a) shows that the model yields satisfying results for
temperatures up to 950 C. But at 1000 C and 1050 C,
the simulated dissolution is much too rapid. In the case of
Fig. 1(b), where many more implantation-induced Is have been
produced, the dissolution velocity is qualitatively wrong even
at 950 C.
B. Late stages of Ostwald ripening
The other fundamental problem of the model of Zographos
et al. [15] concerns the time evolution of the mean loop radius
rL. From Eq. (4), the time evolution of rLis obtained using
Eqs. (1) and Eqs. (3) as
drL
dt =kI
LπDI
nL
CI.(6)
Now consider the case CI'C
I, i.e. the time during annealing
at which the number of loops has considerably decreased. The
remaining ones are so large that they can gain relatively few
energy by exchanging Is. In this situation, Eq. (6) predicts a
linearly diverging time evolution of rLwhile, in reality, the
loop radius growth is much weaker [25, 23]. As soon as the
surface and other sinks for Is become dominant, loops finally
start shrinking again, although slowly, as Ostwald ripening
ensures that rather the number of loops decreases, than their
0 1 2 3 45678
time (min)
109
1010
1011
density of loops
N
L (cm-2 )
(a)
900 C
950 C
1000 C
1050 C
0 10 20 30 40 50 60
time (min)
109
1010
1011
density of loops
N
L (cm-2 )
(b)
950 C
Fig. 1. Time evolution of the areal loop density NLduring annealing. (a)
Rapid thermal annealing (RTA) after a Ge implant of 2·1015 cm2at 150 keV,
experimental data from [23]. (b) RTA after a B implant of 1·1015 cm2at
30 keV, experimental data from [24]. Symbols depict experimental results,
lines simulation results. Simulations were done with the model of Zographos
et al. [15] in the implementation of [16].
size. The described phenomenon is illustrated in Fig. 2. For all
the experiments shown, temperatures are sufficiently high and
annealing times sufficiently long so that Ostwald ripening is in
its final stage. In all panels of Fig. 2, the experiments shown
indicate a saturation of the loop radius with time. By contrast,
the solid lines calculated using the model of Zographos et al.
[15] display a diverging radius evolution.
III. PAR AL LE L RI PE NI NG O F FAULTED AND PERFEC T
LO OP S
The model of Zographos et al. [15] describes faulted loops
(FLs) and makes correct predictions for experiments with low
thermal budgets, while failing for high temperatures and ex-
tended annealing times. This failure can be explained with the
following experimental observation. For low thermal budgets,
mainly FLs are observed [27, 28, 29, 30]. For high thermal
budgets, during which loops undergo strong Ostwald ripening
and grow considerably, almost only perfect loops (PLs) are
observed [31, 32, 28, 33, 34, 29]. This behavior is due to
the well-known fact that for large loop sizes, perfect loops are
more stable than faulted loops. Our model extends Zographos’
model to describe not only FLs, but to also include PLs. While
modeling FLs is still necessary to correctly describe the early
stages of a high-thermal-budget process, modeling the different
reaction dynamics of PLs allows to describe the late stages.
Hence, our new model needs to physically describe both types
of loops.
IEEE JOURNAL OF PHOTOVOLTAICS 3
0 20 40 60 80 100 120
time (min)
0
10
20
30
40
50
60
70
mean loop radius r (nm)
(a) 800 C
900 C
1000 C
0 10 20 30 40 50 60
time (min)
0
10
20
30
40
50
60
70
mean loop radius r (nm)
(b)
950 C
Fig. 2. Time evolution of the mean loop radius rLduring annealing. (a) RTA
after an Si implant of 1·1015 cm2at 50 keV, experimental data from [26]. (b)
RTA after a B implant of 1·1015 cm2at 30 keV, experimental data from [24].
Symbols depict experimental results, lines simulation results. Solid lines were
calculated with the model of Zographos et al. [15] in the implementation of
[16]. Dashed lines were calculated with the same parameters using the model
presented in Sec. IV.
For intermediate thermal budgets, a transition from the state
with more faulted loops to the state with more perfect loops
takes place. This transition has traditionally been assumed
to result from an unfaulting reaction [35], meaning that a
faulted loop, in our cases always a Frank partial dislocation,
reacts with a Shockley partial dislocation to produce a perfect
dislocation loop. In contradiction to that, several authors have
observed independent time evolutions of the FL and the PL
ensembles, clearly identifying two different Ostwald ripening
mechanisms [28, 27, 29]. This observation suggests that a
second explanation for the FL-PL transistion is more likely:
Rather than by a direct unfaulting reaction, the transition
occurs as the two ensembles of different loop types exchange
Is among each other. Our model is based on this second
explanation.
A. Energy of faulted and perfect dislocation loops
The energy associated with a dislocation loop consists of the
elastic energy of the surrounding strain fields, its core energy
and a potential stacking-fault energy. The faulted loops most
frequently observed in Si are Frank partial loops, which lie
in {111}planes and have a Burger’s vector bFL =a
3[111]
with a=5.43 ˚
A denoting the size of the silicon unit cell.
Perfect loops in fcc lattices also lie in {111}planes but have
a Burger’s vector of bPL =a
2[110]. Their respective elastic
energies are [35, 27]
Eel
FL =rLGb2
FL
2(1 ν)ln(r1/r0), b2
FL =a2
3,(7a)
Eel
PL =rLGb2
PL
2(1 ν)1ν
6ln(r1/r0), b2
PL =a2
2.(7b)
Linear elasticity theory is only valid at some distance away
from the core. This distance is characterized by r0for which
we assume r0=bFL/4[27]. For r1, roughly characterizing the
extent of the lattice distortions, we assume r1= 2rL[36, 35].
The energy Ecore stored in the core can only be accessed
by atomistic calculations, albeit with a low precision. The
stacking-fault energy of faulted loops is γ'70 mJ/m2per
atomic volume. Adding these terms to Eqs. (7a) and (7b), one
obtains
EFL =γπr2
L+rLGb2
FL
2(1 ν)(ln(8rL/bFL)1 + AFL ),(8a)
EPL =rLGb2
PL
2(1 ν)1ν
6(ln(8rL/bFL)1 + APL ).(8b)
The constants AFL,PL have been introduced to represent
the core energy via Ecore
FL =AFL rLGb2
FL
2(1ν)and Ecore
PL =
APL rLGb2
PL
2(1ν)1ν
6, respectively. Their values will be discussed
in more detail below. The constant 1was introduced in
the last expressions in brackets to account for a frequent
convention in the literature, ensuring a stress-free boundary at
the upper integration limit for the elastic energy r1[37, 38].
Although the functional form of the expressions (8) is
undebated, the numerical values for the integration boundaries
r1,r0and the core energy Ecore are only roughly known [39].
If, in addition, the loop is not ciruclar but shaped like an n-
sided regular polygon, which can be a reasonable assumption
for PLs, a further constant has to be added to AFL,PL [40].
One should therefore be satisfied with specifying a numerical
range of meaningful values for these parameters. Surveying the
literature, one finds the following values (We abbreviate AFL,PL
with AL.): Some authors choose AL= 0 [15, 23, 19, 38, 41]
and others AL= 1 [27, 35] and all of these either choose
r0=b/4or r0=b. Again others employ a very different
expression AL=2ν1
4ν4[42, 13]. Having made the choice
for r0=b/4and r1= 2rLin Eqs. (8), all uncertainty is
accommodated in AL. Assuming that the true value of r0is
in the range r0[b/4,4b][43, 35, 44] and that the true
core-energy-related component of ALis in the range [0,1],
one obtains the following range of meaningful values for
AL[2.77,1].
B. Energy per self-interstitial in a dislocation loop
The energy necessary to incorporate one I in a loop is
EL=dE
dN =1
nL2πrL
dE
drL. Using Eq. (8), we obtain
EFL =γ/bFL +GbFL
4πrL(1 ν)(ln(8rL/bFL ) + AFL),(9a)
EPL =6ν
4GbFL
4πrL(1 ν)(ln(8rL/bFL ) + APL).(9b)
The result of Eqs. (9) is shown in Fig. 3(a) for AFL = 0 and
APL = 0. The stability inversion occurs for loop sizes around
IEEE JOURNAL OF PHOTOVOLTAICS 4
0 10 20 30 40 50
loop radius rL (nm)
0.0
0.1
0.2
0.3
formation energy per atom (eV)
(a)
EFL
EPL
0 10 20 30 40 50
loop radius rL (nm)
100
101
102
I supersaturation at loop C *
L,I/C *
I
(b)
C*
FL,I 850 C
C*
PL,I 850 C
C*
FL,I 1050 C
C*
PL,I 1050 C
Fig. 3. Comparison of FL and PL formation energy (a) and I supersaturation
in the vicinity of a loop (b). Parameters are given in the text.
30 nm, which is similar to the results of [27] and [29]. The pa-
rameters used for this calculation are given by γ= 70 mJ/m2,
= 2 ×1023 cm3,G= 63.28 ×105N/cm2,ν= 0.28, and
bFL = 0.3135 nm. With that, Esf =γ/bFL = 0.0279 eV and
GbFL
4π(1ν)=0.274 eV nm. The even more interesting quantity is
given by C
L,I =C
IeE(rL)/kT , see Eq. (4) and Fig. 3(b).
Here, the stability inversion of Fig. 3 is amplified to an
exponential behavior. C
L,I is responsible for the dissolution
of loops due to interactions with external sinks or the other
loop ensemble.
C. Summary of model equations
The equations for the parallel evolution of ensembles of
FLs and PLs read, in complete consistence with the model
presentation for only faulted loops in Sec. II,
tCFL =R311
FL +RI
FL,(10a)
tDFL =e
R311
FL +e
RI
FL,(10b)
tCPL =R311
PL +RI
PL,(10c)
tDPL =e
R311
PL +e
RI
PL.(10d)
where the functional forms of the terms R311
FL ,e
R311
FL ,R311
PL and
e
R311
PL have been defined in Eqs. (2) for a general loop. All of
these terms contain either the reaction rate k311
FL or k311
PL , which
determine the unfaulting rate of {311}defects to loops. In
agreement with the experimental observation that PLs are only
observed for higher temperatures, we choose k311
FL to have a
lower activation energy than k311
PL . The terms e
R311
FL and e
R311
PL
contain furthermore the factors g
k311
FL and g
k311
PL , respectively. In
agreement with the experimental observation of Stowe [29],
that perfect loops nucleate at a larger mean radius than faulted
loops, we choose g
k311
PL = 0.05 in comparison to g
k311
FL = 0.5[15]
for faulted loops. Finally, it should be kept in mind that the
expressions of the formation energies per I Eqs. (9) to be used
for C
I,L are different for faulted and perfect loops.
As the result of a calibration based on a variety of ex-
periments, we set k311
FL = 9 ·105·e4.2eV/kT cm3/s for
faulted loops while keeping g
k311
FL ,kI
FL and AFL = 0 as in
Sentaurus Process [16]. For perfect loops, we set k311
PL = 4 ·
105·e6eV/kT cm3/s, g
k311
PL = 0.05,kI
PL = 1.5·107·e1.5eV/kT ,
and APL =2.65.
D. Comparison with the experiment
We note that for low temperatures or short annealing times,
the model extension presented gives virtually the same results
as the original model of Zographos et al. [15]. Thus, we only
discuss high temperatures and extended times. The implanta-
tion conditions for these experiments were modeled with the
native Monte Carlo implantation simulator of Sentaurus Pro-
cess, using the default parameters [20]. Sentaurus Process then
accounts for amorphization and solid-phase epitaxial regrowth
by setting the concentration of implantation-induced Is and
Vs in amorphized regions to zero. Only Is that remain behind
the amorph-crystalline boundary can therefore contribute to
the formation of dislocation loops. In Fig. 4, simulations for
1000 C are compared to experiments. At this temperature,
perfect loops become relevant for the implantation conditions
of [28] and they dominate the total density of loops for times
exceeding about 4 min. Due to the stability inversion, PLs
dissolve slowlier and have a larger mean radius than FLs.
Fig. 5 provides a further example for different implantation
conditions. Fig. 6 compares our model with experiments for
a wide range of temperatures, showing good agreement in
contrast to the results obtained with the model of Zographos
et al. shown in Fig. 1.
While the preceding experimental conditions comprised the
well-studied conditions of amorphizing Si and Ge implants,
for solar cells, we are also interested in non-amorphizing B
implants. For the example of a B implant of 1·1015 cm2at
30 keV annealed for 15 min at 900 C, Fig. 7 shows simula-
tions of the as-implanted and annealed boron concentrations.
It also compares simulated and experimentally measured val-
ues for the depth dependence of the dislocation-line density
ρL= 2πrLDL. While the experimental values correspond to a
mean value for FLs and PLs, the dislocation-line densities for
both types are resolved in the simulations.
Fig. 8(a) and (b) show the time evolution of loop densities
during RTA and furnace anneals after B implantation. Our
model reasonably compares with experiment, in contrast to
the model of Zographos et al. [15], see Fig. 1.
IV. MOD EL IN G TH E SATU RATION OF OSTWALD RIPENING
While the model of the preceding section for the first time
enables the simulation of dislocation loops in the parameter
regime of solar cell processes, this section is devoted to a more
subtle and weaker effect that arises for these processes. In
IEEE JOURNAL OF PHOTOVOLTAICS 5
0 1 2 3 45678
time (min)
109
1010
1011
1012
density of loops
N
L (cm-2 )
(a)
FL
PL
sum
0 1 2 3 45678
time (min)
0
5
10
15
20
25
30
35
40
mean loop radius r (nm)
(b)
FL
PL
mean
Fig. 4. Loop evolution during RTA annealing at 1000 C following a Si
implant of 1·1016 cm2at 50 keV. Experimental data from [28].
0510 15 20
time (min)
108
109
1010
1011
density of loops
N
L (cm-2 )
(a)
FL
PL
sum
0510 15 20
time (min)
0
10
20
30
40
50
mean loop radius r (nm)
(b)
FL
PL
mean
Fig. 5. Loop evolution during RTA annealing at 1000 C following a Ge
implant of 2·1015 cm2at 150 keV. Experimental data from [45].
Fig. 2 and the respective discussion, we showed that the model
of Zographos et al. predicts a linearly diverging loop radius
while experiments indicate a much weaker increase if not
reduction. As the mean loop radius enters the disloction line
density, which directly relates to the recombination activity
0 1 2 3 45678
time (min)
109
1010
1011
density of loops
N
L (cm-2 )
(a)
900 C
950 C
1000 C
1050 C
0 20 40 60 80 100 120
time (min)
108
109
1010
1011
1012
density of loops
N
L (cm-2 )
(b)
800 C
900 C
1000 C
Fig. 6. Time evolution of the areal loop density NL(sum of FL and PL
densities) during annealing. (a) Rapid thermal annealing (RTA) after an Ge
implant of 2·1015 cm2at 150 keV, experimental data from [23]. Compare
this to Fig. 1. (b) RTA after an Si implant of 1·1015 cm2at 50 keV,
experimental data from [26].
0 50 100 150 200 250 300 350 400
depth (nm)
1017
1018
1019
1020
1021
CB (cm-3 )
(b)
B
Bimpl
108
109
1010
1011
L (cm-2 )
FL
PL
exp
Fig. 7. The as-implanted and annealed B profile are shown on the left y-
axis. The dislocation-line densities for perfect (PL) and faulted loops (FL)
are shown on the right y-axis, together with experimental data for the mean
dislocation-line density. For this experiment a B dose of 1·1015 cm2was
implanted at 30 keV and annealed at 900 C for 15min. With this low thermal
budget, FLs can be seen to be still strongly dominant in the simulation. Only
for higher thermal budgets, as shown in Fig. 8(a), PLs start to dominate.
Experimental data from [29, p. 130].
of loops, this model artifact should reduce the quantitative
predicitive power of the model of Zographos et al.. We sketch
an idea that overcomes this artifact at the expense of an only
slightly more complicated model definition.
During annealing, Is will be exchanged among the extended
defects as well as between them and the surface or other
sources and sinks. The former processes is conservative and
leads always to an increase of the mean loop radius rL
of the ensemble. The latter process is non-conservative and
IEEE JOURNAL OF PHOTOVOLTAICS 6
0 10 20 30 40 50 60
time (min)
108
109
1010
1011
density of loops
N
L (cm-2 )
(a) FL
PL
sum
40 20 0 20 40 60
time (min)
108
109
1010
1011
1012
density of loops
N
L (cm-2 )
(b) 900 C
1000 C
Fig. 8. Time evolution of the areal loop density NLduring annealing. (a) RTA
at 950 C after a B implant of 1·1015 cm2at 30 keV, experimental data from
[24]. Compare this to Fig. 1. (b) Loop evolution during furnace annealing,
ramp-up at 10 K/min starting at 600 C. B implant of 1·1016 cm2at
30 keV. Experimental data from [29]. Negative times refer to ramp-up and
preannealing phases. Symbols depict experimental results, lines simulation
results. If not specified otherwise, data points refer to the sum of FL and PL
densities.
increases or decreases rL. Although the distinction of these two
processes has been acknowledged for a long time [25], it has
never explicitly been taken into account by process simulation
models. By explicitly accounting for these processes, we
obtain a new loop model that gives rise to different loop-loop
and loop-I reaction dynamics than the model of Zographos et
al..
A. Model equation ansatz
We make the following ansatz for the time evolution of the
mean loop radius rL
drL
dt =fnc(rL) + fc(rL),(11)
where fnc and fcare functions of rLthat correspond to
the non-conservative and conservative contributions to loop
growth. The natural choice for fnc is known to be [19]
fnc(rL) = kL
nL
(CIC
I,L(rL)),(12)
where kLis a reaction rate. In the conservative regime, when
no other comparable sinks or sources but the loop ensemble
itself are present, fnc should be zero. This holds indeed true
as then CI'C
I,L(rL).
For the conservative contribution to loop growth fc(rL), we
make the ansatz
fc(rL) = kL
nL
(C
I,L(rLσL)C
I,L(rL+σL)) (13)
0 10 20 30 40 50
mean loop radius r (nm)
10-1
100
101
102
C*
L,I/C *
I and fc/(C *
IkL/nL)
C*
FL,I 850 C
fc 850 C
C*
FL,I 1050 C
fc 1050 C
Fig. 9. Comparison of the quantities C
I,L and fc(C
I,L(rLσL)
C
I,L(rL+σL)), see Eq. (13), which govern the non-conservative and the
conservative processes during the evolution of the loop ensemble, respectively.
Here, σL
rL= 0.4.
where σLhas the meaning of the variance of the loop-size
distribution. The choice is motivated as follows. Ostwald
ripening is due to the conservative exchange of Is from small
loops to large loops. In analogy to the ansatz of Burton
and Speights [25], who described the evolution of a loop
with radius rLas a result of the interaction with the mean I
concentration C
I,L(rL), Eq. (13) describes the radius evolution
of a loop with radius rL+σLas a result of the interaction
with the emitted I concentration of a smaller loop with radius
rLσL. As the simulation cannot account for all processes
that occur between loops of the full loop-size distribution, we
assume that the exchange of Is between loops of size rLσL
and rL+σLshould be a representative approximation for these
processes. Note that also for the choice of fnc, we neglected
the details of the loop-size distribution.
In the early stages of the annealing process, when many
small loops are present, the conservative exchange of Is (13)
will govern the loop evolution, while, at a stage at which loops
have become large, the non-conservative contributions (12)
may become more important. Our choice for fcreflects this
transition: fctakes high values for low values of rLand tends
to zero for high values of rL, i.e. in the late-stage regime.
This behavior is obvious in Fig. 9. Furthermore, fcreflects
the fact that a loop-size distribution with a large size variance
σLdisplays faster Ostwald ripening than a distribution with a
small variance.
We note that we always choose σLrL. As the size
distribution function of loops, normalized to the mean loop
radius rL, is time-independent for conservative Ostwald ripen-
ing [46, 25] as well as in the early stage when the conservative
exchange of Is dominates, the normalized variance σL/rLof
the distribution must be time-independent, too. In the late-
stage phase, fc'0holds and, by that, the value of σL
becomes unimportant. An additional equation for the time
evolution of σL, which corresponds to the third moment of the
loop-size distribution, is therefore unnecessary. Consequently,
the preceding definition of a two-moment model automatically
corresponds to a three-moment model.
IEEE JOURNAL OF PHOTOVOLTAICS 7
B. Summary of model equations and comparison to experi-
ment
Before, we obtained the time evolution of rLin Eq. (6)
from Eqs. (1) with RI
Land e
RI
Ldefined in Eqs. (3). Now we
use the expression for rLin Eq. (11) and the expression of
RI
Lof Eqs. (3) to obtain an expression for e
RI
L. Adopting the
notation of [15], we set the constant kL=πkI
LDIand obtain
RI
L=kI
L2π2rLDI(CIC
I,L(rL))DL,(14a)
e
RI
L=kI
L2πDI
rLnL
(C
I,L(rLσL)C
I,L(rL+σL))DL.(14b)
This is to be compared with the model equations of [15] given
in Eqs. (3). The difference between the model we propose and
the one of [15] is that instead of the term (C
I,L(rLσL)
C
I,L(rL+σL)) in the equation for e
RI
L, [15] employ C
I,L(rL).
We have already compared these terms in Fig. 9.
We compare the results of our model (14) with the results
of Zographos’ model in Fig. 2. Our model clearly allows to
identify the early and late stages of Ostwald ripening via the
observation of two time-scales in the reaction dynamics: In the
early stage, when the conservative exchange of Is dominates,
the loop radius increases strongly t[25], while for
long times where the non-conservative exchange dominates,
it increases only weakly or even saturates. Zographos’ model,
by contrast, leads to a strong loop-radius growth also for long
times when considering high temperatures.
For the calculation of Fig. 2 we used σFL/rFL = 0.4with
all other parameters as in the preceding section. Only for k311
FL
the original value of [16] was used.
V. CONCLUSION
We presented a dislocation-loop model for the concurrent
growth of faulted and perfect dislocation loops and sketched
a model that reproduces the saturation of Ostwald ripening.
In comparison to established work, both approaches improve
the agreement with experiments particularly for high thermal
budgets. The model has already proven its viability for solar-
cell processing conditions [7, 8].
REFERENCES
[1] B. J. Pawlak, T. Janssens, S. Singh, I. Kuzma-Filipek,
J. Robbelein, N. E. Posthuma, J. Poortmans, F. Cristiano,
and E. M. Bazizi, “Studies of implanted boron emitters
for solar cell applications,” Progress in Photovoltaics:
Research and Applications, vol. 20, no. 1, pp. 106–110,
2012. URL: http://dx.doi.org/10.1002/pip.1106
[2] M. G. Kang, J.-H. Lee, H. Boo, S. J. Tark,
H. C. Hwang, W. J. Hwang, H. O. Kang, and
D. Kim, “Effects of annealing on ion-implanted Si for
interdigitated back contact solar cell,” Current Applied
Physics, vol. 12, no. 6, pp. 1615–1618, 2012. URL:
http://dx.doi.org/10.1016/j.cap.2012.05.035
[3] J. Benick, R. M¨
uller, N. Bateman, M. Hermle, and
S. Glunz, “Fully implanted n-type PERT solar cells, EU
PVSEC Proceedings, vol. 27, pp. 676–679, 2012. URL:
http://dx.doi.org/10.4229/27thEUPVSEC2012-2BO.7.5
[4] W. Ho, Y.-H. Huang, W.-W. Hsu, Y.-Y. Chen, and
C. Liu, “Ion implanted boron emitter n-silicon solar
cells with wet oxide passivation, in Photovoltaic
Specialists Conference (PVSC), 2011 37th IEEE, june
2011, pp. 001 058 –001 060. URL: http://dx.doi.org/10.
1109/PVSC.2011.6186134
[5] A. Florakis, T. Janssens, E. Rosseel, B. Douhard,
J. Delmotte, E. Cornagliotti, J. Poortmans, and
W. Vandervorst, “Simulation of the anneal of
ion implanted boron emitters and the impact on
the saturation current density, Energy Procedia,
vol. 27, no. 0, pp. 240–246, 2012. URL:
http://dx.doi.org/10.1016/j.egypro.2012.07.058
[6] K. R. C. Mok, R. C. G. Naber, and L. K. Nanver,
“Insights to emitter saturation current densities of boron
implanted samples based on defects simulations,” AIP
Conference Proceedings, vol. 1496, no. 1, pp. 245–248,
2012. URL: http://dx.doi.org/10.1063/1.4766534
[7] J. Kr¨
ugener, F. A. Wolf, R. Peibst, F. Kiefer,
C. Sch¨
ollhorn, A. Grohe, R. Brendel, and H. J. Os-
ten, “Correlation of dislocation line densities and emit-
ter saturation current densities of ion implanted boron
emitters,” in 23rd Photovoltaic Science and Engineering
Conference (PVSEC), 2013, pp. 1–0–31.
[8] J. Kr¨
ugener, R. Peibst, F. A. Wolf, E. Bugiel, T. Ohrdes,
F. Kiefer, C. Sch¨
ollhorn, A. Grohe, R. Brendel, and
H. J. Osten, “Electrical and structural analysis of crystal
defects after high-temperature rapid thermal annealing of
highly boron ion-implanted emitters,” to be submitted to
IEEE Journal of Photovoltaics, 2014.
[9] A. H. Gencer and S. T. Dunham, A predictive model for
transient enhanced diffusion based on evolution of {311}
defects,” Journal of Applied Physics, vol. 81, no. 2, pp.
631–636, 1997. URL: http://dx.doi.org/10.1063/1.364204
[10] N. E. B. Cowern, G. Mannino, P. A. Stolk,
F. Roozeboom, H. G. A. Huizing, J. G. M. van
Berkum, F. Cristiano, A. Claverie, and M. Jara´
ız,
“Energetics of self-interstitial clusters in Si, Phys.
Rev. Lett., vol. 82, pp. 4460–4463, May 1999. URL:
http://dx.doi.org/10.1103/PhysRevLett.82.4460
[11] M. Law and K. Jones, A new model for {311}defects
based on in situ measurements,” in Electron Devices
Meeting (IEDM), Technical Digest. International, 2000,
pp. 511–514. URL: http://dx.doi.org/10.1109/IEDM.
2000.904367
[12] D. Stiebel, P. Pichler, and N. E. B. Cowern, “A reduced
approach for modeling the influence of nanoclusters and
{113}defects on transient enhanced diffusion, Applied
Physics Letters, vol. 79, no. 16, pp. 2654–2656, 2001.
URL: http://dx.doi.org/10.1063/1.1406147
[13] I. Avci, M. E. Law, E. Kuryliw, A. F. Saavedra, and
K. S. Jones, “Modeling extended defect ({311}and
dislocation) nucleation and evolution in silicon, Journal
of Applied Physics, vol. 95, no. 5, pp. 2452–2460, 2004.
URL: http://dx.doi.org/10.1063/1.1645644
[14] C. Zechner, N. Zographos, D. Matveev, and A. Erlebach,
Accurate and efficient TCAD model for the formation
and dissolution of small interstitial clusters and {311}
IEEE JOURNAL OF PHOTOVOLTAICS 8
defects in silicon,” Materials Science and Engineering:
B, vol. 124-125, no. 0, pp. 401–403, 2005. URL:
http://dx.doi.org/10.1016/j.mseb.2005.08.010
[15] N. Zographos, C. Zechner, and I. Avci, “Efficient
TCAD model for the evolution of interstitial clusters,
{311}defects, and dislocation loops in silicon,” MRS
Proceedings, vol. 994, p. 0994 F10 01, April 2007.
URL: http://dx.doi.org/10.1557/PROC-0994-F10-01
[16] SPAC, Advanced Calibration for Sentaurus Process User
Guide, Version G-2012.06, Synopsys, Inc., Mountain
View, CA, 2012.
[17] C. Donolato, “Modeling the effect of dislocations on the
minority carrier diffusion length of a semiconductor,
Journal of Applied Physics, vol. 84, no. 5, pp. 2656–
2664, 1998. URL: http://dx.doi.org/10.1063/1.368378
[18] V. Kveder, M. Kittler, and W. Schr¨
oter, “Recombination
activity of contaminated dislocations in silicon: A
model describing electron-beam-induced current contrast
behavior, Phys. Rev. B, vol. 63, p. 115208, Mar 2001.
URL: http://dx.doi.org/10.1103/PhysRevB.63.115208
[19] R. Y. S. Huang and R. W. Dutton, “Experimental
investigation and modeling of the role of extended
defects during thermal oxidation,” Journal of Applied
Physics, vol. 74, no. 9, pp. 5821–5827, 1993. URL:
http://dx.doi.org/10.1063/1.355306
[20] SP, Sentaurus Process User Guide, Version G-2012.06,
Synopsys, Inc., Mountain View, CA, 2012.
[21] A. Claverie, B. Colombeau, B. de Mauduit, C. Bonafos,
X. Hebras, G. Ben Assayag, and F. Cristiano, “Extended
defects in shallow implants, Applied Physics A, vol. 76,
pp. 1025–1033, 2003. URL: http://dx.doi.org/10.1007/
s00339-002-1944-0
[22] F. A. Wolf, “Modeling of annealing processes for ion-
implanted single-crystalline silicon solar cells,” Ph.D.
dissertation, University of Erlangen-Nuremberg, 2014.
[23] C. Bonafos, D. Mathiot, and A. Claverie, “Ostwald
ripening of end-of-range defects in silicon,” Journal of
Applied Physics, vol. 83, no. 6, pp. 3008–3017, 1998.
URL: http://dx.doi.org/10.1063/1.367056
[24] M. Milosavljevi´
c, M. Lourenco, G. Shao, R. Gwilliam,
and K. Homewood, “Optimising dislocation-engineered
silicon light-emitting diodes,” Applied Physics B,
vol. 83, pp. 289–294, 2006. URL: http://dx.doi.org/10.
1007/s00340-006-2149-6
[25] B. Burton and M. V. Speight, “The coarsening and
annihilation kinetics of dislocation loop,” Philosophical
Magazine A, vol. 53, no. 3, pp. 385–402, 1986. URL:
http://dx.doi.org/10.1080/01418618608242839
[26] J. Liu, M. Law, and K. Jones, “Evolution of dislocation
loops in silicon in an inert ambient I,” Solid-State
Electronics, vol. 38, no. 7, pp. 1305–1312, 1995. URL:
http://dx.doi.org/10.1016/0038-1101(94)00257-G
[27] F. Cristiano, J. Grisolia, B. Colombeau, M. Omri,
B. de Mauduit, A. Claverie, L. F. Giles, and N. E. B.
Cowern, “Formation energies and relative stability of
perfect and faulted dislocation loops in silicon, Journal
of Applied Physics, vol. 87, no. 12, pp. 8420–8428,
2000. URL: http://dx.doi.org/10.1063/1.373557
[28] G. Z. Pan, K. N. Tu, and A. Prussin, “Size-distribution
and annealing behavior of end-of-range dislocation
loops in silicon-implanted silicon,” Journal of Applied
Physics, vol. 81, no. 1, pp. 78–84, 1997. URL:
http://dx.doi.org/10.1063/1.364099
[29] D. J. Stowe, An investigation of efficient room
temperature luminescence from silicon which contains
dislocations,” Ph.D. dissertation, University of Oxford,
2006. URL: http://ora.ox.ac.uk/resolve/info:fedora/uuid:
9ee073b7-9e3c-4637-9ce1-62e9e4ade69d/THESIS01
[30] M. Sztucki, T. H. Metzger, I. Kegel, A. Tilke,
J. L. Rouvi`
ere, D. L¨
ubbert, J. Arthur, and J. R.
Patel, “X-ray analysis of temperature induced defect
structures in boron implanted silicon,” Journal of
Applied Physics, vol. 92, no. 7, pp. 3694–3703, 2002.
URL: http://dx.doi.org/10.1063/1.1505982
[31] K. S. Jones, S. Prussin, and E. R. Weber, A
systematic analysis of defects in ion-implanted silicon,”
Applied Physics A: Materials Science And Processing,
vol. 45, pp. 1–34, 1988. URL: http://dx.doi.org/10.1007/
BF00618760
[32] J. J. Comer, “Electron microscope study of stackingfault
formation in boron implanted silicon,” Radiation
Effects, vol. 36, no. 1-2, pp. 57–61, 1978. URL:
http://dx.doi.org/10.1080/00337577808233171
[33] Y. L. Huang, M. Seibt, and B. Plikat, “Nonconservative
Ostwald ripening of dislocation loops in silicon, Applied
Physics Letters, vol. 73, no. 20, pp. 2956–2958, 1998.
URL: http://dx.doi.org/10.1063/1.122642
[34] M. Milosavljevi´
c, G. Shao, M. A. Lourenco, R. M.
Gwilliam, and K. P. Homewood, “Engineering of
boron-induced dislocation loops for efficient room-
temperature silicon light-emitting diodes,” Journal of
Applied Physics, vol. 97, no. 7, p. 073512, 2005. URL:
http://dx.doi.org/10.1063/1.1866492
[35] D. Hull and D. J. Bacon, Introduction to Dislocations,
5th ed. Amsterdam: Butterworth-Heinemann, 2011.
[36] G. Schoeck and W. A. Tiller, “On dislocation formation
by vacancy condensation, Philosophical Magazine,
vol. 5, no. 49, pp. 43–63, 1960. URL: http://dx.doi.org/
10.1080/14786436008241199
[37] N. Martsinovich, “Theory of defects arising from
hydrogen in silicon and diamond,” Ph.D. dissertation,
University of Sussex, 2004. URL: http://www.lifesci.
sussex.ac.uk/research/tc/thesis
[38] F. Kroupa, “Circular edge dislocation loop,”
Cechoslovackij fiziceskij zurnal B, vol. 10, pp. 284–293,
1960. URL: http://dx.doi.org/10.1007/BF02033533
[39] R. LeSar, Ambiguities in the calculation of dislocation
self energies, physica status solidi (b), vol. 241, no. 13,
pp. 2875–2880, 2004. URL: http://dx.doi.org/10.1002/
pssb.200302054
[40] D. J. Bacon and A. G. Crocker, “The elastic
energies of symmetrical dislocation loops, Philosophical
Magazine, vol. 12, no. 115, pp. 195–198, 1965. URL:
http://dx.doi.org/10.1080/14786436508224960
[41] W. Cai and C. R. Weinberger, “Energy of a prismatic
dislocation loop in an elastic cylinder, Mathematics and
IEEE JOURNAL OF PHOTOVOLTAICS 9
Mechanics of Solids, vol. 14, no. 1-2, pp. 192–206, 2009.
URL: http://dx.doi.org/10.1177/1081286508092611
[42] C. D. Meekison, A model of diffusion-controlled
annealing of dislocation loops in pre-amorphized
silicon,” Philosophical Magazine A, vol. 69, no. 2,
pp. 379–396, 1994. URL: http://dx.doi.org/10.1080/
01418619408244350
[43] H. F¨
oll, Defects in Crystals. Germany: Lecture Notes,
University of Kiel, 2012. URL: http://www.tf.uni- kiel.
de/matwis/amat/def en/
[44] D. B. Holt and B. G. Yacobi, Extended Defects in Semi-
conductors Electronic Properties, Device Effects and
Structures, 1st ed. Cambridge: Cambridge University
Press, 2007.
[45] C. Bonafos, A. Martinez, M. Faye, C. Bergaud,
D. Mathiot, and A. Claverie, “Transient enhanced
diffusion of dopant in preamorphised Si: The role of EOR
defects,” Nuclear Instruments and Methods in Physics
Research Section B, vol. 106, no. 1-4, pp. 222–226, 1995.
URL: http://dx.doi.org/10.1016/0168-583X(95)00707-5
[46] I. Lifshitz and V. Slyozov, “The kinetics of precipitation
from supersaturated solid solutions,” Journal of Physics
and Chemistry of Solids, vol. 19, no. 1-2, pp. 35–
50, 1961. URL: http://dx.doi.org/10.1016/0022-3697(61)
90054-3
F. Alexander Wolf obtained a M. Sc. in Physics in 2011 from the University
of Augsburg. At the time of this work, he was a PhD student at Bosch
Corporate Research, Gerlingen, Germany. He now is with the Ludwigs-
Maximilians-Universit¨
at (LMU), Munich.
Alberto Martinez-Limia obtained his diploma in Nuclear Physics in 1994
from the Institute of Nuclear Sciences and Technology of Havana and his
PhD in 2002 from the Institute of Theoretical Chemistry of the University
of Erlangen-Nuremberg. Later he worked in several academic and scientific
organizations: the University of South Carolina, the University of Chemnitz
and the Fraunhofer Institute IISB in Erlangen. He investigated topics related
to material modeling and process simulation. In October 2008 he joined the
Robert Bosch GmbH as a Research and Development scientist.
Daniel Stichtenoth studied physics at the University of G¨
ottingen, Germany
and Uppsala, Sweden. He received his diploma degree in physics in 2005 and
his Dr. rer nat in 2008 from the University of G¨
ottingen, Germany. Since 2009
he is with BOSCH Solar Energy AG, Arnstadt, Germany and works in the
field of silicon solar cell development. His main focus is on the interaction
of silicon with high-efficiency silicon solar cell processes.
Peter Pichler obtained the Dip.-Ing. degree in Electrical Engineering in
1982 and the Dr. techn. degree 1985 both from the Technical University
of Vienna. Since 1986 he has been Group Manager at Fraunhofer IISB,
responsible now for the doping and device simulation activities. In 2004
he obtained the venia legendis from the University of Erlangen-Nuremberg.
Dr. Pichler contributed to various European projects. coordinated the EC
projects RAPID, FRENDTECH, and ATOMICS on diffusion and activation
phenomena in silicon and silicon-based materials, and coordinates the ICT
Project ATEMOX on the modeling of leakage currents and technologies for
low-leakage ultrashallow junctions. He is the author or coauthor of some
120 publications in international journals and conference proceedings, and
author of the book “Intrinsic Point Defects, Impurities, and Their Diffusion
in Silicon” published by Springer Wien-New York.
... It was extended by Zographos et al. [3] by including a moment-based method for dislocation loops (without distinguishing between FDLs and PDLs). Wolf et al. [4] found that the model by Zographos et al. [3] resulted in too high dissolution velocities of loops at high temperatures and a diverging mean loop radius for long annealing times at high temperatures, contradicting experimental findings [5]. To account for that, they extended the model by Zographos et al. [3] to include PDLs and a different model approach for the evolution of the mean loop radii, implementing the saturation of Ostwald ripening [5]. ...
... Wolf et al. [4] found that the model by Zographos et al. [3] resulted in too high dissolution velocities of loops at high temperatures and a diverging mean loop radius for long annealing times at high temperatures, contradicting experimental findings [5]. To account for that, they extended the model by Zographos et al. [3] to include PDLs and a different model approach for the evolution of the mean loop radii, implementing the saturation of Ostwald ripening [5]. For its calibration, presented in Wolf's thesis [4], the focus was predominantly on conditions relevant for solar cell processing (long high-temperature anneals), where they mainly considered the sum of dislocation loops rather than the distinction between the two types. ...
... In this work, the focus was to recalibrate the model by Wolf et al. [5] to better describe measured data where FDLs and PDLs coexist after post-implantation annealing. The better the individual defect types can be described, the clearer the overall picture will be, contributing to better predictions of enhanced dopant diffusion and performance deterioration due to leakage currents in final devices. ...
Article
Full-text available
Predictive models of the damage evolution during post-implantation annealing are important for predicting the performance of final devices. An existing model of the thermal evolution of dislocation loops in silicon during post-implantation annealing has been recalibrated to better capture the distinction between faulted and perfect dislocation loops. The calibration was based on experimental findings from the literature on the simultaneous presence of both faulted and perfect dislocation loops after post-implantation annealing. Graphical abstract
... To simulate both the residual bulk iron point defect concentration as well as the iron precipitate size distribution in the implanted layer, we use as inputs the implantation parameters, the time-temperature profile of the gettering anneal, and the iron contamination level. Our model contains three parts: 1) An implantation damage model, which describes the formation of silicon interstitials during ion-implantation and their evolution into dislocation loops during subsequent thermal processing [52]; 2) A heterogeneous iron precipitation model, which describes the nucleation and growth of iron precipitates at the implantation damage [53]; 3) A segregation model, which describes iron pairing with boron, which increases the solubility of iron in the heavily boron-doped layer, reducing bulk iron point defect concentration [28]. The uncertainty of the simulated values is estimated by changing the iron solid solubility to lower or higher by a factor of 2, which corresponds to the scatter in the literature data of iron solubility [54,55]. ...
... The implantation damage model, described thoroughly in a previous report [52], calculates the average areal dislocation loop density NL (in cm -2 ) and the mean loop radius rL, using the implantation parameters and the time-temperature profile as inputs. For the implantation parameters used in this study, the simulated loop density was 2.0×10 7 cm -2 and the average loop radius was 305.6 nm. ...
... where n is the number of Fe atoms per precipitate. The precise depth distribution of the implantation damage is unknown, but it is generally much closer to the surface than the junction depth [52]. As long as the areal density of dislocations is held constant, the dislocations can be distributed arbitrarily along the first 10 -1000 nm of the wafer without changing the simulation results presented here. ...
Article
Full-text available
To facilitate cost-effective manufacturing of boron-implanted silicon solar cells as an alternative to BBr 3_{3} diffusion, we performed a quantitative test of the gettering induced by solar-typical boron-implants with the potential for low saturation current density emitters (<50 fA/cm2). We show that depending on the contamination level and the gettering anneal chosen, such boron-implanted emitters can induce more than a 99.9% reduction in bulk iron point defect concentration. The iron point defect results as well as synchrotron-based nano-X-ray-fluorescence investigations of iron precipitates formed in the implanted layer imply that, with the chosen experimental parameters, iron precipitation is the dominant gettering mechanism, with segregation-based gettering playing a smaller role. We reproduce the measured iron point defect and precipitate distributions via kinetics modeling. First, we simulate the structural defect distribution created by the implantation process, and then we model these structural defects as heterogeneous precipitation sites for iron. Unlike previous theoretical work on gettering via boron- or phosphorus-implantation, our model is free of adjustable simulation parameters. The close agreement between the model and experimental results indicates that the model successfully captures the necessary physics to describe the iron gettering mechanisms operating in boron-implanted silicon. This modeling capability allows high-performance, cost-effective implanted silicon solar cells to be designed.
... 35 The resultant crystal defect was caused by boron implantation, which necessitated high-temperature annealing to rebuild the lattice and activate the dopant (>950°C). 36 During annealing, the interactions of the dopant atom (B) with the host atom (Si) and the defect formation kinetics (from small interstitial clusters to dislocation loops) are accounted for using the models listed in Table II. Along with defect formation, some unwanted phenomena, such as transient enhanced diffusion (TED) and oxidation enhanced diffusion (OED), during annealing lead to enhanced dopant diffusion and dopant segregation (reducing dopant activation), which are also accounted for using respective models. ...
... The small interstitial clusters will develop further into dislocation loops (DLs) through the Ostwald ripening process. 20,36,37,46 These defects are stable even after consecutive high-temperature annealing at 1050°C in an inert atmosphere, as evident in Fig. 3(a), and they are replicated in components of J 0 in Fig. 4(a). ...
Article
The effect of process-induced defects on the photo-generated charge-carrier lifetime and solar cell performance is critical, which will help optimize the process recipe. In this work, we attempt to quantify the effects of process-induced defects during boron implantation on the n-type silicon wafer in different annealing ambiences. We have evaluated the role of defects that can be formed during oxygen and inert ambience annealing on n-type bifacial passivated emitter rear totally diffused solar cells using a recombination current prefactor (J 0 ). The numerically calculated J 0 is calibrated with the reported experimental J 0 values using two different methods: (i) Shockley–Read–Hall lifetime and (ii) effective trap-density method. In the latter method, we used the simulated defect density profiles. Both methods capture the process-induced degradation. We observed that the process-induced defects could deteriorate by almost 1% absolute efficiency for the considered annealing conditions. We found that dislocation loops alone cause an ignorable effect on terminal characteristics, but other process-induced mechanisms could dominantly degrade the cell's performance. To further support, we show that independent defects (apart from coupled defects) other than dislocation loops could explain the experimentally reported boron-implanted diodes’ J–V curves under reverse bias conditions.
... The process simulations, which calculated the evolution of extended defects, including 311 linear defects and dislocation loops, during thermal annealing, have been carried out using the Sentaurus Process software of Synopsys with Advanced Calibration [5]. The model for the evolution of dislocation loops was extended according to the work of Wolf et al. [6]. It was implemented and verified at IISB using the Alagator language of Sentaurus Process version S-2021.06. ...
... Even a longer ramp-down time cannot reduce the dislocation loop density in furnace annealing to the low level of the of RTP annealing as can be seen in Fig. 4b. The reduction of the concentration of dislocation loops in case of RTP annealing compared to furnace annealing is due to a faster transition to temperatures over 950 °C at which the annealing process of the dislocation loops begins, as shown in Fig. 4b [6]. The mean radius for dislocations reaches a 35 nm for 20-min RTP, while it settles at 42 nm for furnace annealing. ...
Article
Full-text available
Low leakage diodes are necessary in order to manufacture high-quality variable capacitance diodes (varicaps), which are used in voltage-controlled oscillators. Junction leakage current affects the single sideband noise of the oscillator by up-conversion of 1/ f and shot noise (Chan et al. in IEEE Trans Electron Devices 54(9):2570–2573, 2007, https://doi.org/10.1109/TED.2007.903201 ). Several sources show higher leakage current for RTP compared to furnace anneal (Lunnon et al. in J Electrochem Soc 132(10):2473, 1985, https://doi.org/10.1149/1.2113602 , Gramenova et al. in J Electrochem Soc 146(1):359, 1999, https://doi.org/10.1149/1.1391613 , Mikoshiba et al. in Jpn J Appl Phys, 1986, https://doi.org/10.1143/jjap.25.l631 ). In our experiments, we found lower leakage currents for RTP compared to furnace annealing. We present results from annealing experiments where we compare three annealing conditions with and without oxidizing annealing conditions. Graphical abstract
... 27 However, recent work on the impact of amorphizing implantation on the performance of solar cells has suggested that extended thermal budgets, used both in solar fabrication and in this work, are not sufficient to remove clustered defects even for annealing temperatures >950 C and duration of >10 min. 28 An alternative (and perhaps more convincing) explanation for our observation of enhanced diffusion is the phenomenon of concentration dependent diffusion. For example, Ref. 29 describes an experimental study of Ge diffusion in Si 1-x Ge x with x ranging from 0 to 0.5. ...
Article
Full-text available
Germanium condensation is demonstrated using a two-step wet oxidation of germanium implanted Silicon-On-Insulator (SOI). Samples of 220 nm thick SOI are implanted with a nominal fluence of 5 × 10¹⁶ cm⁻² Ge⁺ at an energy of 33 keV. Primary post-implantation wet oxidation is performed initially at 870 °C for 70 min, with the aim of capping the sample without causing significant dose loss via Ge evaporation through the sample surface. This is followed by a secondary higher temperature wet oxidation at either 900 °C, 1000 °C, or 1080 °C. The germanium retained dose and concentration profile, and the oxide thickness is examined after primary oxidation, and various secondary oxidation times, using Rutherford backscattering analysis. A mixed SiGe oxide is observed to form during the primary oxidation followed by a pure silicon oxide after higher temperature secondary oxidation. The peak germanium concentration, which varies with secondary oxidation condition, is found to range from 43 at. % to 95 at. %, while the FWHM of the Ge profile varies from 13 to 5 nm, respectively. It is also observed that both the diffusion of germanium and the rate of oxidation are enhanced at 870 and 900 °C compared to equilibrium expectations. Transmission electron microscopy of a representative sample with secondary oxidation at 1080 °C for 20 min shows that the SiGe layer is crystalline in nature and seeded from the underlying silicon. Raman spectroscopy is used to determine residual strain in the SiGe region following secondary oxidation. The strain is compressive in nature and increases with Ge concentration to a maximum of approximately 1% in the samples probed. In order to elucidate the physical mechanisms, which govern the implantation-condensation process, we fit the experimental profiles of the samples with a model that uses a modified segregation boundary condition; a modified linear rate constant for the oxidation; and an enhanced diffusion coefficient of germanium where the enhancement is inversely proportional to the temperature and decays with increasing time. Comparison of the modeled and experimental results shows reasonable agreement and allows conclusions to be made regarding the dominant physical mechanisms, despite the semi-empirical nature of the model used.
... In order to treat this question from a simulation perspective, a predictive model for the annealing kinetics of DLs will be derived. Parts of this chapter have been published in Wolf et al. (2014b). ...
Thesis
Full-text available
Ion implantation technology has become economically competitive for solar cell doping in about 2012. The decisive reduction in production costs was achieved by using plasma implantation tools, which allow a throughput of more than 2000 wafers per hour. Implantation technology now opens the door for sophisticated cell concepts with high conversion efficiencies. The present thesis clarifies key questions in the simulation of the associated annealing processes. This is achieved by developing models for the reaction and diffusion kinetics of implanted phosphorus, implanted boron and implantation-induced dislocation loops. These models are based on conceptual ideas that originate from the field of microelectronics but have hitherto not been applied to the considerably different fabrication conditions of solar cells. Compared to standard solar cell doping using POCl3 diffusion, ion implantation allows a much better control of the doping process. Besides its immediate advantages though, it comes with the drawback of inevitable implantation damage. This work investigates the question of how implantation damage, in particular dislocation loops, has to be annealed to obtain high-performing solar cells. To address this question, a model for the reaction kinetics of dislocation loops has been developed and shown to be viable for characteristic situations of solar-cell processing. This model is the first one to capture the late stages of Ostwald ripening and the transition from faulted to perfect dislocation loops. The two primary dopant species of interest for solar-cell doping by implantation are phosphorus and boron. Due to the complexity of the diffusion and segregation behavior of phosphorus, its annealing behavior is still under debate today and no generally predictive models exist. In this thesis, a predictive model for phosphorus annealing suitable for the particular requirements of solar cell processing has been developed, calibrated and applied. As part of the model development, a general strategy for an effcient calibration of models describing impurity diffusion via two species has been found. The annealing behavior of boron, on the other hand, is well understood and highly predictive models exist. The plasma implantation tools of the solar cell industry though usually implant BF3 instead of atomic boron. The diffusion behavior of boron in presence of uorine differs fundamentally from that without uorine and no generally valid model for this situation has been available previously. In the present thesis, a comprehensive model for this problem has been implemented and successfully tested. Finally, the models for the annealing behavior of implantation damage, phosphorus and boron were studied with respect to their relevance for the electronic properties of solar cells. In particular, simulated predictions of the dislocation-line density and its associated recombination activity have been compared with recombination-current measurements from a large-scale experimental study performed at the University of Hannover and the ISFH Hameln. The quantitative agreement obtained provides, for the first time, strong evidence that dislocation loops are responsible for implantation damage-related degradation of solar cells. Combining futhermore process simulation with device simulation, main factors that in uence solar-cell performance could be investigated. This allowed to verify assertions from the literature as well as suggesting processing options.
Chapter
In this chapter an outline of the very diverse topic of process simulation is given. This includes the process steps of ion implantation and thermal annealing, which introduce, activate, and modify dopant distributions, and the process steps lithography, deposition, and etching, which are used to structure the semiconductor wafer. The section on oxidation deals with the simulation of the oxide growth, whereas the aspect of dopant segregation is included in the section on diffusion. Finally, the impact of process variations is outlined. Overall, the process simulation chapter primarily deals with the physics and the related models for the various process steps, whereas the discussion of generic algorithms, e.g., for the solution of partial differential equations, is left for another dedicated chapter of this book. However, some algorithms which are specific for process simulation are also briefly described in this chapter. Due to the diversity of the area of process simulation, this chapter could not strive for completeness in the presentation of the physical models. We largely refer to silicon technology, whereas most models can also be applied for or adapted to other top-down semiconductor technologies where, in contrast to bottom-up technologies based on self-assembling, patterning steps, ion implantation and high-temperature process steps are used to generate three-dimensional geometries and dopant distributions.
Article
Boron junction and its passivation is an active topic in photovoltaic research due to its importance to passivated emitter and rear totally‐diffused (PERT) bifacial Si solar cell. In this paper, a systematic study on boron‐implanted junction, its passivation and ohmic contact formation, as well as application to p‐PERT bifacial cells has been presented. More specifically, the impact of junction profile and surface passivation on boron junction quality, which can be influenced by implantation and in situ oxidation anneal parameters is studied. Good‐quality boron emitter and metal/p⁺‐Si ohmic contact are achieved, as demonstrated by emitter saturation current of 5–30 fA cm⁻² and specific contact resistance of 3–6 mΩ cm⁻¹. A roadmap for a p‐PERT bifacial cell using fully ion implanted (boron and phosphorus) technology, based on which p‐PERT bifacial cells demonstrats front side efficiency of 20.6% (open circuit voltage of 658 mV), rear side efficiency of ≈17% and bifaciality factor of 0.82–0.87 is also presented. Effective cell efficiency of ≈22% is achieved with a Halm IV‐tester with white transporting belts that enhance reflection about 10% from the cell rear.
Conference Paper
Ion implantation of boron is a promising technique for the preparation of p-type emitters in n-type silicon solar cells. Here, experimental emitter saturation current densities are compared with simulated defect densities, namely boron interstitial clusters and dislocation loops. We report on experimental conditions, which allow separating the specific impact of both defect types on the resulting electrical properties after annealing at 1050°C and surface passivation. In that way, dislocation loops are identified to be the dominating defect species under the used implant and annealing conditions.
Thesis
Full-text available
Ion implantation technology has become economically competitive for solar cell doping in about 2012. The decisive reduction in production costs was achieved by using plasma implantation tools, which allow a throughput of more than 2000 wafers per hour. Implantation technology now opens the door for sophisticated cell concepts with high conversion efficiencies. The present thesis clarifies key questions in the simulation of the associated annealing processes. This is achieved by developing models for the reaction and diffusion kinetics of implanted phosphorus, implanted boron and implantation-induced dislocation loops. These models are based on conceptual ideas that originate from the field of microelectronics but have hitherto not been applied to the considerably different fabrication conditions of solar cells. Compared to standard solar cell doping using POCl3 diffusion, ion implantation allows a much better control of the doping process. Besides its immediate advantages though, it comes with the drawback of inevitable implantation damage. This work investigates the question of how implantation damage, in particular dislocation loops, has to be annealed to obtain high-performing solar cells. To address this question, a model for the reaction kinetics of dislocation loops has been developed and shown to be viable for characteristic situations of solar-cell processing. This model is the first one to capture the late stages of Ostwald ripening and the transition from faulted to perfect dislocation loops. The two primary dopant species of interest for solar-cell doping by implantation are phosphorus and boron. Due to the complexity of the diffusion and segregation behavior of phosphorus, its annealing behavior is still under debate today and no generally predictive models exist. In this thesis, a predictive model for phosphorus annealing suitable for the particular requirements of solar cell processing has been developed, calibrated and applied. As part of the model development, a general strategy for an effcient calibration of models describing impurity diffusion via two species has been found. The annealing behavior of boron, on the other hand, is well understood and highly predictive models exist. The plasma implantation tools of the solar cell industry though usually implant BF3 instead of atomic boron. The diffusion behavior of boron in presence of uorine differs fundamentally from that without uorine and no generally valid model for this situation has been available previously. In the present thesis, a comprehensive model for this problem has been implemented and successfully tested. Finally, the models for the annealing behavior of implantation damage, phosphorus and boron were studied with respect to their relevance for the electronic properties of solar cells. In particular, simulated predictions of the dislocation-line density and its associated recombination activity have been compared with recombination-current measurements from a large-scale experimental study performed at the University of Hannover and the ISFH Hameln. The quantitative agreement obtained provides, for the first time, strong evidence that dislocation loops are responsible for implantation damage-related degradation of solar cells. Combining futhermore process simulation with device simulation, main factors that in uence solar-cell performance could be investigated. This allowed to verify assertions from the literature as well as suggesting processing options.
Article
Full-text available
Ion implantation of boron is a promising technique for the preparation of p-type emitters in n-type cells. We use rapid thermal annealing with temperatures up to 1250 °C and annealing durations between 6 s and 20 min to anneal the implant-induced crystal defects. Experimental J 0 e is compared with simulated and measured defect densities. Perfect dislocation loops are identified to be the dominating defect species after rapid thermal annealing (RTA) above 1000 °C. Even for emitters with J 0 e values around 40 fA/cm 2 , defects are present within the valleys of the textured sur-faces after annealing. On textured Al 2 O 3 -passivated boron emit-ters, we measure J 0 e of 38 fA/cm 2 for a sheet resistance around 80 Ω/ after very short annealing processes (1 min at 1200 °C). Index Terms—Boron, crystal defects, ion implantation, photo-voltaic, rapid thermal annealing (RTA), silicon.
Conference Paper
Full-text available
We present detailed investigations of the influence of high temperature (T) annealing on the resulting emitter saturation current densities (J0e) of ion implanted boron emitters. We use rapid thermal annealing (RTA) with T up to 1250 °C and annealing durations between 6 s and 20 min. The experimental J0e values are compared with dislocation line densities (ρL) obtained from SENTAURUS process [1] simulations using an extended model of dislocation loop formation for T up to 1250 °C. We find clear correlations between simulated ρL and measured J0e. For ρL below 2∙108 cm-2, J0e is no longer dominated by ρL. On textured surfaces, we obtain J0e values of 38 fA/cm² on 80 Ω/sq Al2O3 passivated boron emitters after very short RTA processes (1 min at 1200 °C).
Article
Full-text available
Recently, the interest for the implementation of local ion implantation for emitter formation has been renewed, as it results in high cell efficiencies [1], while the amount of process steps is reduced. After implantation, a thermal diffusion (annealing) is necessary for the removal of the ion-induced damage, the activation of dopants and the formation of the desired profile shape. The goal is to change the implanted retrograde near surface (10-100 nm) inactive profile, into a damage free, box-like shaped, and fully active emitter (or BSF) with a low saturation current density (J0e).Here, we study different post-implantation annealing schemes, to achieve high performance boron emitters for n-type IBC solar cells. We show experimentally, that the use of oxygen gas during the boron anneal has an impact on the electrical performance (J0e) of the emitter. On the other hand, the obtained J0e values are higher than those obtained by the standard diffusion approach. Hence, in order to understand these results and optimize the implant/anneal process, the modeling of the profile shape and the damage evolution would be extremely valuable. In this work, we report on our initial simulation results, and we demonstrate the impact of oxygen on the evolution of the profile shapes of the implanted B-emitters after anneal, taking into account boron diffusion kinetics mechanisms such as Transient Enhanced Diffusion (TED) and Oxidation Enhanced Diffusion (OED). Moreover, we monitor the residual damage and we make a first effort to correlate the increase on the J0e values with the residual damage levels. Throughout this work, we adopt the process simulator tool Sentaurus Process.
Article
Ion implantation is the technique of choice for introducing dopant species into semiconductors in CMOS devices. In photovoltaics ion implantation could be an interesting alternative to tube furnace diffusion processes. In this work we thus show that the damage introduced during implantation can be completely removed during annealing and that a co-annealing process can be applied for the annealing of boron and phosphorus profiles. Boron emitter implanted PassDop as well as fully implanted PERT solar cells have been fabricated. High conversion efficiencies could be achieved for both, the PassDop (22.2%, 694 mV) as well as for the fully implanted PERT (22.3%, 684 mV) solar cells, proving the high quality of the applied boron and phosphorus implantation.
Article
Dislocations are defects in the crystal structure of important engineering materials, such as metals, ceramics and semi-conductors. They strongly influence the properties of these materials. The first edition of this classic academic reference was published in the 1960s when research on dislocations was intense and in the years since the book has carved its niche as a respected, reliable and understandable introduction to the subject area for students, graduates and researchers. No other resource is as credible, accessible and up to date as this title. Introduction to Dislocations is widely recognized as providing the essential knowledge base required for further study, research and technological application. This update brings the book's coverage, references and context in-line with the latest developments to ensure it remains a core reference for anyone studying and working in the field. Classic, long-established academic reference by an expert author team, highly regarded for their contributions to the field. Presents the theory, related problems and applications in a highly detailed yet easy-to-read text, making this an ideal, understandable introduction to a complex topic. Uses minimal mathematics to make the subject more accessible and includes all-important, up-to-date references to essential further reading and research sources. © 2011 D. Hull and D. J. Bacon Published by Elsevier Ltd All rights reserved.
Article
1. Semiconducting materials; 2. An introduction to extended defects; 3. Characterization of extended defects in semiconductors; 4. Core structures and mechanical effects of extended defects specific to semiconductors; 5. The electrical, optical and device effects of dislocations and grain boundaries; 6. Point defect materials problems.
Article
Emitter saturation current densities, Joe have been investigated with different boron implantation dose and annealing conditions. The higher thermal budgets used here are shown experimentally to improve Joe, implying more complete defect dissolution. Simulations show that significant degradation in Joe can be attributed to the presence of dislocation loops. In addition, in cases where dislocation loops have been annealed, high dose boron implantation still results in stable boron interstitial clusters, which contributes to Joe degradation.
Article
Effects of annealing on the properties of P- and B-implanted Si for interdigitated back contact (IBC) solar cells were investigated with annealing temperature of from 950 to 1050 °C. P-implanted samples annealed at 950 °C were enough to activate dopants and recover the damage by implantation. As the annealing temperature was increased, the diode properties of P-implanted samples were degraded, while that of B-implanted samples were improved. However, in order to activate an implanted B ion, B-implanted samples needed an annealing of above 1000 °C. The implied Voc of lifetime samples by quasi-steady-state photoconductance decay followed the trend of diode properties on annealing temperature. Finally, IBC cell was fabricated with a two-step annealing at 1050 °C for B of the emitter and 950 °C for P of the front and back surface fields. The IBC cell had Voc of 618 mV, Jsc of 35.1 mA/cm2, FF of 78.8%, and the efficiency of 17.1% without surface texturing.
Article
An ion implanted boron emitter silicon solar cell is demonstrated. By using appropriate annealing condition, the implanted dopants and damage introduced by the implantation can be activated and repaired, respectively. Both the rapid thermal annealing (RTA) and furnace annealing were investigated within this work. For the surface passivation and the antireflection coating, the wet oxide (SiO2) was grown during the furnace annealing step. Since the wet oxidation process has about one order of magnitude faster than the dry oxidation process, it could be a suitable technology for industrial silicon solar cell processing. The rapid thermal annealing with various temperatures and annealing time were also investigated. The open-circuit voltage increased with increasing the temperature and time of the RTA process. The planar p+nn+ solar cell with wet oxide passivaiton achieves the Voc of 0.631 V and the efficiency of 15.43%.