Você está na página 1de 7

JOURNAL OF APPLIED PHYSICS 110, 124513 (2011)

Comparative atom probe study of Cu(In,Ga)Se2 thin-film solar cells deposited


on soda-lime glass and mild steel substrates
Pyuck-Pa Choi,1,a) Oana Cojocaru-Mirédin,1 Roland Wuerz,2 and Dierk Raabe1
1
Max-Planck-Institut für Eisenforschung, Max-Planck-Str. 1, 40237 Düsseldorf, Germany
2
Zentrum für Sonnenenergie-und Wasserstoff-Forschung Baden-Württemberg, Industriestrasse 6,
70565 Stuttgart, Germany
(Received 26 July 2011; accepted 31 October 2011; published online 21 December 2011)
We report on a comparative study of Cu(In,Ga)Se2 solar cells deposited on soda-lime glass and
mild steel substrates, using atom probe tomography in conjunction with secondary ion mass
spectrometry, x-ray fluorescence, current density-voltage, and external quantum efficiency
measurements. Cu(In,Ga)Se2 films deposited on soda-lime glass substrates and on steel substrates
with a NaF precursor layer on top of the Mo back contact contain a significant amount of Na
impurities and yield an enhanced open circuit voltage and fill factor. Using atom probe
tomography, Na atoms are found to be segregated at grain boundaries and clustered in both bulk
and grain boundaries. The atom probe data indicate that NaCu point defects are most likely formed
at grain boundaries, reducing the number of compensating InCu point defects and thus contributing
to an enhanced cell efficiency. However, for steel substrates the positive effect of Na on the cell
performance is counterbalanced by the incorporation of Fe impurities into the Cu(In,Ga)Se2 film.
Fe atoms are homogeneously distributed inside the grains suggesting that Fe introduces point
C 2011 American Institute of Physics. [doi:10.1063/1.3665723]
defects in the bulk V

I. INTRODUCTION Na on the cell parameters are reported to be an increase in


open circuit voltage and fill factor.7,8,12–14 Furthermore, an
Cu(In,Ga)Se2 (CIGS) thin-film solar cells have reached
increase in the acceptor concentration and p-type conductiv-
the status of commercial application owing to their high
ity is commonly observed,12,14 which has been attributed to
energy conversion efficiency, long-term stable performance,
the reduction of compensating donor point defects.12,17 Na
and high radiation hardness.1–3 The absorber material CIGS
incorporation cannot only change dopant and charge carrier
is an I-III-VI2 compound semiconductor, which has a direct
concentrations but also bandgap grading10 and grain size10
bandgap and high absorption coefficients over a wide spec-
of CIGS films and thus the cell efficiency.
tral range.1,2 Thus, CIGS films of only 1 to 2 lm in thickness
Though there is no doubt about the beneficial effects of
are sufficient for efficient photovoltaic conversion, making
(moderate) Na-doping on the efficiency of CIGS solar cells
them not only advantageous in terms of material consump-
their origin is not well understood. This is mainly due to the
tion and cost reduction but also highly interesting for flexible
fact that it is extremely challenging to map the distribution
solar cell applications. The maximum energy conversion
of dopants below nanometer-scale. Na was found to be pref-
efficiency of a CIGS solar cell reported so far is 20.3%,4
erentially enriched at the grain boundaries by means of
being the highest value achieved by any thin-film photovol-
Auger electron spectroscopy15 and secondary ion mass spec-
taic technology to date. The use of steel substrates renders
trometry (SIMS),16 suggesting that the “Na effect” is associ-
CIGS solar cells light-weight and flexible while the effi-
ated with the grain boundaries rather than bulk-doping. Na
ciency can still amount up to 17.5%.5
segregation at CIGS grain boundaries could be directly pro-
Among various processing techniques, Na-doping of the
ven and quantified by means of atom probe tomography
CIGS absorber film is reported to be of uttermost importance
(APT),18–20 a high-resolution characterization technique that
for enhancing the cell efficiency.6–17 Na-doping can be
enables three-dimensional elemental mapping with sub-
inherently achieved by the usage of soda-lime glass (SLG)
nanometer resolution.21–24 APT has recently provided some
substrates,6,7,14 which contain a large amount of Na. During
unique insights into the elemental distribution in solar grade
the co-evaporation process of CIGS, Na atoms can diffuse
CIGS films (Refs. 18–20, 25) and has in particular revealed
from the SLG substrate into the absorber film, leading to
the composition at internal interfaces, but systematic APT
moderate dopant concentrations (typically about 0.1 at.%).7
studies of CIGS solar cells deposited on various types of sub-
Alternative approaches of Na-doping are the deposition of a
strates with varying amounts of Na (and other impurities
Na-containing precursor layer such as NaF prior to CIGS
such as Fe) are lacking. Furthermore, APT data have not yet
deposition,12,13 the co-evaporation of NaF during CIGS
been correlated with corresponding cell efficiencies.
deposition7,9,10 or the NaF post deposition treatment after the
In this paper, we report on a comparative study of CIGS
growth of the CIGS film.7,8 The main beneficial effects of
solar cells fabricated using the same single-stage inline pro-
cess but with different impurity contents. Na-containing
a)
Author to whom correspondence should be addressed. Electronic mail: standard cells deposited on SLG substrates are compared
choi@mpie.de. with novel types of flexible cells deposited on mild steel

0021-8979/2011/110(12)/124513/7/$30.00 110, 124513-1 C 2011 American Institute of Physics


V

Downloaded 10 Feb 2012 to 193.175.131.12. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions
124513-2 Choi et al. J. Appl. Phys. 110, 124513 (2011)

substrates. The compositions and elemental distributions To minimize beam damage, a low energy (5 keV) Ga beam
measured by APT are discussed in conjunction with the cor- was used at the final ion-milling stage. APT analyses were
responding cell efficiency data. carried out by applying laser pulses of 532 nm wavelength,
12 ps pulse length, and 0.1 nJ pulse energy. The specimen
base temperature was about 60 K. To reduce the influence of
II. EXPERIMENTAL METHODS a possible compositional inhomogeneity due to the CIGS
CIGS absorber films of 2 lm in thickness with a Ga con- deposition process, the sample pieces for XRF, SIMS, and
tent of [Ga]/([Ga]þ[In])  0.3 were grown on Mo-coated APT analyses were all taken from the same quarter of the
SLG and mild steel substrates by co-evaporation of the con- sample stripes.
stituent elements in a single-stage inline process.26 This
process was carried out at a substrate temperature of about III. RESULTS AND DISCUSSION
600  C for 30 min. Mild steel sheets having a thickness of
A. Composition of CIGS films as measured with APT,
500 lm and an average roughness of 38 nm were used as
XRF, and SIMS
substrate materials for the flexible solar cells. The CIGS
films on the steel substrates were doped with Na by evaporat- Table I lists the compositions of all three studied sam-
ing a NaF precursor layer ( 10 nm in thickness) on the Mo ples as measured with APT, XRF, and SIMS. While XRF
back contact prior to the deposition of the absorber film. was used for determining the overall concentrations of the
Complete cell structuring was done for all samples, i.e., a CIGS matrix elements, SIMS was used for determining the
CdS buffer layer was deposited by chemical bath deposition concentrations of Na and Fe impurities. APT and XRF
followed by radio-frequency sputtering of an intrinsic ZnO results are in good agreement with each other. The Se con-
layer, direct current sputtering of a ZnO:Al front contact centrations detected by APT are slightly lower than the val-
layer, and electron beam evaporation of Ni/Al contact grids. ues measured with XRF, which can be ascribed to
For convenience, the Na-containing sample on SLG and the preferential field evaporation of Se ions between the applied
Na-free and Na-containing samples on mild steel substrates laser pulses during an APT analysis. To the authors’ knowl-
are referred to as SLG, steel, and steel-NaF, respectively. edge, there is no literature value available for the evaporation
The overall compositions of the CIGS films were deter- field of Se. But due to its very low melting and boiling tem-
mined by x-ray fluorescence spectroscopy (XRF) using an perature (494 K (Ref. 28) and 963 K,29 respectively), Se is
EAGLE XXL system (0.1 mbar, Si(Li) detector, 50 kV Rh expected to field evaporate at low electric fields. The devia-
x-ray source, spot size of 1 mm in diameter). The absolute tion between APT and XRF/SIMS data for the other matrix
accuracy in determining the concentration of Cu, In, Ga, and elements as well as for Na and Fe impurities can be
Se in the CIGS film by XRF is 60.5 at.%. SIMS and second- explained by the fact that the volumes typically probed with
ary neutral mass spectrometry (SNMS) depth profiling was APT are small (50  50  200 nm3). Thus, APT is more
performed at liquid nitrogen temperature on the bare CIGS sensitive to local compositional changes than XRF and
absorber films in a LEYBOLD SSM 200 system, using 5 SIMS but is more prone to statistical errors when determin-
keV primary Arþ ions within a sputter area of 2  2 mm2. ing the overall composition. As observed in some of the APT
The measurement area after blanking was 1.2  1.2 mm2. To measurements, there are slight compositional variations
obtain quantitative data from SIMS, the Fe and Na signals between individual CIGS grains. As a result, APT analyses
were calibrated with respect to standard samples of known Fe can yield varying overall compositions, depending on the
and Na concentrations. The relative error of determining Fe local grain compositions and the number of detected grains.
and Na concentrations in CIGS by SIMS is about 25%, where Both, APT and SIMS analyses yield higher Na concen-
the detection limit is about 1 ppm. Cell efficiencies were deter- trations for the steel-NaF sample than for the SLG sample.
mined by standard current density-voltage (J-V) measurements Interestingly, the steel-NaF sample has a higher Fe content
under AM1.5 equivalent illumination. External quantum effi- than the steel sample, showing that Na impurities interact
ciency (EQE) measurements were done by measuring the with Fe impurities and promote their diffusion into the CIGS
short-circuit current with spectrally resolved monochromatic film. This may be related to the affinity between Na and Fe,
light. The APT samples were prepared using focused-ion- as reported in Ref. 30 and/or to the affinity between F and
beam milling according to the procedures described in Ref. 27. Fe.31

TABLE I. Chemical compositions of CIGS absorber films measured by APT and XRF/SIMS. While XRF was used for measuring Cu, In, Ga, and Se concen-
trations, SIMS was used for measuring Na and Fe concentrations.

Sample Cu [at.%] In [at.%] Ga [at.%] Se [at.%] Na [ppm] Fe [ppm]

SLG (Na-doped) APT 22.3 6 0.08 19.5 6 0.08 9.5 6 0.06 48.7 6 0.09 290 6 50 0
XRF/SIMS 22.9 6 0.5 17.8 6 0.5 8.2 6 0.5 51.1 6 0.5 577 6 144 3 6 0.8
Steel (non-doped) APT 22.7 6 0.05 18.7 6 0.05 10.1 6 0.04 48.5 6 0.05 0 0
XRF/SIMS 23.0 6 0.5 18.5 6 0.5 8.1 6 0.5 50.4 6 0.5 2.0 6 0.5 37 6 9
Steel-NaF (Na-doped) APT 24.7 6 0.03 21.35 6 0.03 6.5 6 0.01 47.2 6 0.03 1830 6 30 20 6 3
XRF/SIMS 22.7 6 0.5 18.8 6 0.5 8.1 6 0.5 50.4 6 0.5 2241 6 560 156 6 39

Downloaded 10 Feb 2012 to 193.175.131.12. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions
124513-3 Choi et al. J. Appl. Phys. 110, 124513 (2011)

FIG. 2. External quantum efficiency (EQE) data of the steel (dotted line),
FIG. 1. J-V curves of studied solar cells measured under illumination with
steel-NaF (dashed line), and SLG (solid line) sample.
AM 1.5. steel: deposited on mild steel without Na (dotted line), steel-NaF:
deposited on mild steel with a NaF precursor layer on the Mo back contact
(dashed line), SLG: deposited on SLG containing Na impurities (solid line).
The SNMS concentration depth profiles of the CIGS ele-
ments are shown in Fig. 3. The sum of the CuþGaþIn signal
B. Cell parameters (CIG, gray dotted line in Fig. 3) and Se signal (green dotted
line) of the steel sample decrease at nearly the same depth.
The J-V curves and corresponding cell parameters are In contrast, for the Na containing samples the Se signal is
shown in Fig. 1 and in Table II, respectively. The open cir- shifted to the right toward the Mo interface. This is a hint for
cuit voltage (Voc) and the fill factor (FF) are higher for the the formation of a MoSe2 layer at the interface between
steel-NaF compared to the steel sample (see Fig. 1 and CIGS and Mo. The Se surplus signal (red lines in Fig. 3) is
Table II), which is the well-known “Na effect”7,8,12–14. The the lowest for the Na free steel sample and enhanced for the
“Na effect” is even much more evident when comparing the Na containing substrates. Thus, the formation of MoSe2 is
SLG with the steel sample. However, the steel-NaF sample more pronounced in the presence of Na as observed by
exhibits a lower short circuit current density (Jsc) than the Caballero et al.33
steel sample, which appears to be related to its higher Fe Overall, the efficiencies of the steel and steel-NaF cells
content (see Table I). As shown by the EQE data in Fig. 2, (6.3%) are substantially lower than the efficiency of the SLG
this effect is due to reduced charge carrier collection in the cell (12.1%). This can be attributed to the diffusion of detri-
infrared region for the steel-NaF sample. Similar observa- mental Fe impurities from the substrate into the CIGS films.
tions were made by one of the authors for CIGS solar cells Voc, FF, and Jsc were reported to decrease with increasing Fe
deposited on steel substrates with varying Fe impurity con- content in CIGS, where the decrease in Jsc is more pro-
tents.32 Thus, the positive effect of Na on Voc and FF is nounced than the decrease in Voc and FF.32
counterbalanced by the negative effect of Fe on Jsc, resulting
in a similar efficiency for both the steel-NaF and steel sam-
ple. The J-V curve of the Na-free steel sample shows a
strong blocking behavior at forward bias, which is known as
the “roll-over effect.” No roll-over effect is observed for the
Na-doped cells. Such an effect was also reported by Cabal-
lero et al.33 for CIGS solar cells deposited on polyimide sub-
strates with varying NaF layer thickness on the Mo back
contact. The authors ascribed the roll-over effect to an elec-
tronic barrier at the Mo/CIGS back interface, which is
reduced in the presence of Na.33

TABLE II. Cell parameters of studied samples: total area efficiency, open-
circuit voltage Voc, fill factor FF, short-circuit current density Jsc (mean val-
ues of N cells without shunted cells), and Fe concentrations in the CIGS
films measured by SIMS.
FIG. 3. (Color) SNMS concentration profiles of Cu63 (black lines), Se80
Sample g [%] Voc [mV] FF [%] Jsc [mA cm2] c(Fe) [ppm] N (green lines), Mo98 (blue lines), and sum of Cu þ In þ Ga signal CIG
(gray lines) given in at.%. Se surplus signal Se80/CIGS (red lines, ratio of
SLG 12.1 650 72.5 25.6 3 15 the Se80 signal to the sum of all CIGS elements) given in percent for the
Steel 6.3 537 56.7 20.6 37 8 steel (dotted lines), steel-NaF (dashed lines), and SLG (solid lines) sample.
Steel-NaF 6.3 552 61.5 18.4 156 12 The thickness (x-axis) has been normalized for better comparison of the
samples.

Downloaded 10 Feb 2012 to 193.175.131.12. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions
124513-4 Choi et al. J. Appl. Phys. 110, 124513 (2011)

FIG. 4. Mass spectrum acquired from an APT analysis


of a CIGS absorber film (steel-NaF sample).

C. APT results prior to APT analysis, respectively. Only the distribution of


Cu and Na atoms is shown in the APT map for better clarity.
A typical mass spectrum acquired from an APT analysis
The TEM image reveals a grain boundary, which is enriched
of a CIGS film is shown in Fig. 4 for the steel-NaF sample.
with Na, as can be clearly seen in the APT map. Using this
The mass spectra of the other samples are very similar
correlative microscopy approach Na segregation to CIGS
except for the varying mass-to-charge peak intensities of Na
grain boundaries can be directly proven. It should be noted
and Fe. While Cu, In, and Ga ions are detected as both single
that this experiment was done on a sample with large grain
and double charged ions, Se ions are only detected in a single
sizes, processed under different conditions than the samples
charged state. Peaks belonging to complex ions such as
presented in the following.
CuSeþ, Cu2Seþ, CuSe2þ, Se2þ, and Se3þ appear in the mass
Figure 6 shows three-dimensional elemental distribution
spectrum as well. Na and Fe impurities are detected as Naþ
in the absorber film of the SLG sample as detected by APT.
and Feþ (see inset in Fig. 4). Furthermore, the mass spectrum
The front of the analyzed volume is located close to the
reveals complex ions consisting of nitrogen and hydrogen
CIGS/CdS interface (about 300 nm away). The detected Na
such as N2Hþ, (NH)2þ, (NH)3þ. These most likely originate
concentration is (290 6 50) ppm, which is significantly lower
from the residual gas molecules inside the analysis chamber,
than the overall Na concentration of the CIGS film
which are field ionized by the electric field at the APT sam-
[(577 6 144) ppm] as measured with SIMS. As mentioned,
ple. A small number of O2þ ions is detected which could
such a discrepancy can be ascribed to the small volume
stem from the residual gas inside the APT or CIGS deposi-
probed with APT (118  112  487 nm3) in comparison to
tion chamber or the SLG substrate. The peaks at 38 Da and
55 Da could not be identified but are negligibly small.
Figures 5(a) and 5(b) show an APT map and the corre-
sponding TEM image taken from the identical specimen

FIG. 6. (Color) Three-dimensional elemental map of the CIGS absorber


FIG. 5. (Color) (a) Three-dimensional elemental map showing a CIGS GB. film of the SLG sample as detected by APT. Top: all elements displayed,
Only the distribution of Cu and Na atoms is shown. The distribution of In, bottom: only Na displayed. Some of the Na clusters in CIGS grains are
Se, and Ga is omitted for better clarity. (b) Corresponding TEM image of marked with black arrows. The front of the analyzed volume is located about
the identical specimen prior to APT analysis. 300 nm away from the CIGS/CdS interface.

Downloaded 10 Feb 2012 to 193.175.131.12. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions
124513-5 Choi et al. J. Appl. Phys. 110, 124513 (2011)

SIMS (1200  1200  2 lm3), which makes the overall con-


centration measurement using APT prone to compositional
heterogeneities. Indeed, Fig. 6 shows that Na atoms are not
homogeneously distributed but are strongly segregated at
grain boundaries. Several Na-rich clusters can be resolved as
well, most of them decorating grain boundaries. The Na-
depleted zones enclosed by Na enrichments correspond to
CIGS grains. The average diameter of the grains is about
100 nm in good agreement with electron backscatter diffrac-
tion (EBSD) data (not shown). The average Na concentration
inside the grains is about 145 ppm. Kronik et al.11 explained
the limited bulk solubility of Na in CIGS by a large ionic ra-
dius mismatch between Na and Cu and the high elastic dis-
tortion energy associated with tetrahedral coordination in the
CIGS lattice. As a consequence, Na segregation at grain
boundaries, where octahedral bonds can be formed, is favor-
able.11 The APT data also reveals that some Na atoms form
clusters (marked with arrows in Fig. 6) inside the CIGS
grains due to their limited bulk solubility. The average Na
concentration in these clusters, as determined by the maxi-
mum separation method34 of solutes (defining 1.5 nm as the
maximum distance between Na solutes and 10 as the mini-
mum number of Na atoms inside a cluster), is (14.0 6 0.5)
at.%.
Figure 7(a) is an enlarged view of a part of the analyzed
volume in Fig. 6. Both, Na segregation at grain boundaries
(two of them marked as “GB1” and “GB2” and with red
arrows) and Na clusters in CIGS bulk (marked with black
arrows) can be seen. Figures 7(b) and 7(c) show composition
profiles across GB1 and GB2. Within the CIGS grains (if Na
clusters are excluded) no significant compositional fluctua-
tions of the matrix elements can be measured, which could
also be confirmed by a statistical v2-test. At the grain boun-
daries significant depletion of Cu and enrichment of Na can
be measured. In addition, there is a slight enrichment of In at
the grain boundaries as well. Hence, it is most likely that Cu
vacancies (VCu–) are occupied by Naþ ions and thus a high
concentration of NaCu point defects exist at the grain boun-
daries. The observation of Cu-depleted and slightly In-
enriched grain boundaries is in good agreement with ab initio
density functional theory (DFT) studies of point defects in
CIGS, as reported by Zunger et al.35–37 According to these
DFT calculations, neutral (2VCu– þ InCu2þ)0 defect complexes FIG. 7. (Color) (a) Enlarged view of a part of the analyzed volume in Fig.
have a very low formation enthalpy in CuInSe2. Furthermore, 6, showing Na segregation at grain boundaries (marked as “GB1” and
it was proposed that Cu vacancy rows exist at grain bounda- “GB2” with red arrows) and Na clusters in CIGS bulk (marked with black
arrows), (b) and (c) composition profiles across GB1 and GB2.
ries, resulting in an energetic barrier against holes due to
reduced repulsion between Cu d- and Se p-orbitals.36,37 Such
an effect is expected to reduce charge carrier recombination at form clusters. VCu1 clusters may be partly occupied by Naþ
grain boundaries and to be beneficial for the cell performance. ions, possibly explaining Na clustering at grain boundaries
The composition profiles in Figs. 7(b) and 7(c) also strongly and in the bulk.
suggest that Cu vacancies at the grain boundaries are partly Figures 8 and 9 show the three-dimensional elemental
occupied by Na. NaCu0 defects can inhibit the formation of distribution in the absorber films of the steel and steel-NaF
compensating InCu2þ donor levels and significantly enhance sample, respectively. The front of the analyzed volumes is
the effective p-type doping (and thus Voc) as proposed by located close to the CIGS/CdS interface (about 300 nm
Contreras et al.12 NaCu0 defects themselves are considered to away) in regions having larger grain sizes than those close to
be electrically inactive as they were reported not to introduce the Mo back contact. Both samples show a homogeneous
any gap levels.38 Rockett et al.39 proposed the existence of distribution of matrix elements. Neither Na nor Fe impurities
(VCu1 InCu2þ VCu1)0 defect complexes, which exert mutual can be detected for the steel sample. While Fe impurities
attraction because of their dipole moment and thus tend to exhibit a relatively homogeneous distribution in the steel-

Downloaded 10 Feb 2012 to 193.175.131.12. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions
124513-6 Choi et al. J. Appl. Phys. 110, 124513 (2011)

FIG. 8. (Color) Three-dimensional elemental distribution in the CIGS


absorber film of the steel sample. Neither Na nor Fe impurities are detected.
The front of the analyzed volume is located about 300 nm away from the
CIGS/CdS interface.

NaF sample, Na impurities show a strong segregation tend-


ency. Two-dimensional Na-enriched networks are clearly
visible, which can be interpreted as grain boundary segrega-
tion zones. Thus, one can conclude that irrespective of the
Na source (SLG or NaF) Na diffuses into the CIGS film dur- FIG. 10. (Color) Typical concentration profiles plotted across one of the
ing the single stage co-evaporation process and segregates at Na-enriched grain boundaries (marked with red dotted line in Fig. 9).
the grain boundaries. At some grain boundaries Na atoms are Figure 10 shows typical concentration profiles across
clustered (marked by arrows in Fig. 9). The average Na con- one of the Na-enriched grain boundaries in the steel-NaF
centration in these clusters, determined in the same way as sample (marked with a red dotted line in Fig. 9). A very sim-
for the SLG sample, is (6.0 6 0.9) at.%. The Na-depleted ilar trend as for the grain boundaries in the SLG sample is
zones (150 ppm Na) enclosed by the Na enrichments corre- observed, namely Na (2 at.%) and In enrichment (24.7
spond to CIGS grains, where the average grain size estimated at.%) and Cu depletion (17.8 at.%). While the Na concen-
from the Na map is about 50 nm in good agreement with tration inside the CIGS grains is nearly equal for both the
EBSD results (not shown). APT as well as EBSD data reveal steel-NaF and SLG sample (150 and 144 ppm, respectively),
that the steel-NaF sample has significantly smaller grains the Na concentration at the grain boundaries of the steel-NaF
than the SLG sample. This is probably related to the different sample (2 at.%) is higher compared to the SLG sample
thermal behavior (different emissivity and therefore different (0.2 at.%). Thus, it can be concluded that excess Na is
substrate temperature) of the two substrate materials during mostly accommodated at grain boundaries or is clustered in
the CIGS process, though the exact reason for this observa- the bulk. The trends shown in the composition profiles in
tion remains unclear. Figs. 7(b), 7(c), and Fig. 10 are general observations made
for all grain boundaries at which only Na but no oxygen
impurities are detected. As mentioned before, it appears that
Na-doping of the CIGS film leads to the formation of a large
number of NaCu point defects at the grain boundaries, which
may reduce the number of compensating InCu antisite
defects. The Na concentration at a CIGS grain boundary can
depend on the type of Na source (SLG or NaF) as well as the
grain boundary character and orientation. It should be men-
tioned that the measured width of the Na segregation zones
(full width at half maximum of the Na concentration peaks
in Figs. 7(b), 7(c), and (10) is about 3 nm, which is substan-
tially larger than the typical structural width of a grain
boundary (0.5 nm). This observation can be attributed to
an artifact of reconstructing the APT data. The variation of
the local tip radius results in an error in determining the lat-
eral coordinates of ions, known as the local magnification
effect.40
Na segregation at grain boundaries is considered to
enhance the cell efficiency of the steel-NaF sample in a simi-
lar way as for the SLG sample. However, the steel-NaF sam-
ple contains some Fe impurities in the CIGS film, where Fe
FIG. 9. (Color) Three-dimensional elemental distribution in the CIGS atoms are homogeneously distributed in the grains (see bot-
absorber film of the steel-NaF sample. Top: all elements, middle: Na, bot- tom map in Fig. 9). It may be speculated that Fe impurities
tom: Fe. Some of the Na clusters are marked with black arrows. The grain
boundary across which the composition profile in Fig. 10 is taken is marked
create deep defect levels in the CIGS bulk, which enhance
with the red dotted line. The front of the analyzed volume is located about charge carrier recombination and thus lower the cell effi-
300 nm away from the CIGS/CdS interface. ciency. As small amounts of Fe in CIGS are already

Downloaded 10 Feb 2012 to 193.175.131.12. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions
124513-7 Choi et al. J. Appl. Phys. 110, 124513 (2011)

7
sufficient to counterbalance the beneficial effects of Na, the D. Rudmann, “Effects of sodium on growth and properties of Cu(In,
defect levels related to Fe must be highly electrically active. Ga)Se2 thin films and solar cells”, Ph.D. thesis, ETH Zürich, 2004.
8
D. Rudmann, D. Brémaud, A. F. da Cunha, G. Bilger, A. Strohm,
To understand the exact role of Fe impurities on the cell effi- M.Kaelin, H. Zogg, and A. N. Tiwari, Thin Solid Films 480-481, 55
ciency, theoretical studies of point defects and defect levels (2005).
9
associated with Fe incorporation in the CIGS lattice are D. Rudmann, G. Bilger, M. Kaelin, F.-J. Haug, H. Zogg, and A. N. Tiwari,
Thin Solid Films 431-432, 37 (2003).
required. Finally, it should be mentioned that the “Na effect” 10
D. Güttler, A. Chirila, S. Seyrling, P. Blosch, S. Buecheler, X. Fontane, V.
is probably not only related to the Na-doping of the CIGS Izquierdo-Roca, L. Calvo-Barrio, A. Perez-Rodriguez, J. R. Morante, A.
film (change of dopant and charge carrier concentration) but Eicke, and A. N. Tiwari, Conf. Rec. 35th IEEE Photovoltaic Spec. Conf.,
also to several other phenomena. For instance, Na may seg- 3420 (2010).
11
L. Kronik, D. Cahen, and H. W. Schock, Adv. Mater. 10, 31 (1998).
regate at the CdS/CIGS41 and CIGS/Mo interface7 and sig- 12
M. A. Contreras, B. Egaas, P. Dippo, J. Webb, J. Granata, K. Ramanathan,
nificantly alter the band structure at the p/n-junction and at S. Asher, A. Swartzlander, and R. Noufi, Conf. Rec. 26th IEEE Photovol-
the back contact, respectively. To gain a full understanding taic Spec. Conf., 359 (1997).
13
of the “Na effect,” the entire solar cell needs to be analyzed K. Granath, M. Bodegard, and L. Stolt, Sol. Energy Mater. Sol. Cells 60,
279 (2000).
with respect to the electronic properties, microstructure, and 14
M. Ruckh, D. Schmid, M. Kaiser, R. Schäffler, T. Walter, and H. W.
chemical composition by applying complementary character- Schock, 24th IEEE Photovoltaic Spec. Conf., 156 (1994).
15
ization techniques. As demonstrated in this paper, APT can D. W. Niles, M. Al-Jassim, and K. Ramanathan, J. Vac. Sci. Technol. A
provide very useful information on dopant distributions and 17, 291 (1999).
16
D. Braunger, D. Hariskos, G. Bilger, U. Rau, and H. W. Schock, Thin
compositional gradients at internal interfaces. Solid Films 361-362, 161 (2000).
17
D. J. Schroeder and A. Rockett, J. Appl. Phys. 82, 4982 (1997).
18
E. Cadel, N. Barreau, J. Kessler, and P. Pareige, Acta Mater. 58, 2634
IV. CONCLUSIONS (2010).
19
R. Schlesiger, C. Oberdorfer, R. Wuerz, G. Greiwe, P. Stender, M. Artme-
CIGS solar cells fabricated with the same single-stage ier, P. Pelka, F. Spaleck, and G. Schmitz, Rev. Sci. Instr. 81, 043703
co-evaporation process but with varying dopant (Na and Fe) (2010).
20
concentrations were studied using combined APT, XRF, O. Cojocaru-Mirédin, P. Choi, R. Wuerz, and D. Raabe, Ultramicroscopy
111, 552 (2011).
SIMS, J-V, and EQE measurements. APT reveals that Na is 21
A. Cerezo, T. J. Godfrey, and G. D. W. Smith, Rev. Sci. Instrum. 59, 862
inhomogeneously distributed in Na-containing CIGS films (1988).
22
and is segregated at grain boundaries and clusters. Further- D. Blavette, A. Bostel, J. M. Sarrau, B. Deconihout, and A. Menand,
more, the APT data indicate that Na atoms at grain bounda- Nature 363, 432 (1993).
23
M. K. Miller, “Atom Probe Tomography – Analysis at the Atomic Level,”
ries partly occupy Cu vacancies and suppress the formation (Kluwer Academic/Plenum, New York, 2000).
of compensating InCu antisite defects during CIGS film 24
T. F. Kelly and M. K. Miller, Rev. Sci. Instrum. 78, 031101 (2007).
25
growth. Such a mechanism is expected to enhance the cell O. Cojocaru-Mirédin, P. Choi, R. Wuerz, and D. Raabe, Appl. Phys. Lett.
efficiency and could partly explain the “Na effect.” For the 98, 103504 (2011).
26
M. Powalla, G. Voorwinden, and B. Dimmler, Proc. 14th Europ. Photovol-
steel-NaF sample, the “Na effect” is counterbalanced by the taic Solar Energy Conf., 1270 (1997).
negative effect of Fe impurities homogeneously solved in the 27
K. Thompson, D. Lawrence, D. J. Larson, J. D. Olson, T. F. Kelly, and B.
CIGS grains. It may be speculated that Fe impurities create Gorman, Ultramicroscopy 107, 131 (2007).
28
C. Kittel, “Introduction to Solid State Physics,” 8th ed. (Wiley, New York,
deep defect levels in the bulk, which enhance charge carrier
2004).
recombination and are detrimental to the cell efficiency. 29
M. De Selincourt, Proc. Phys. Soc. 52, 348 (1940).
30
G. A. El-Shobaky, A. A. Ibrahim, and S. El-Ddfrawy, Thermochim. Acta
131, 115 (1988).
31
ACKNOWLEDGMENTS C. Schlegel, private communication (8 July 2011).
32
R. Wuerz, A. Eicke, M. Frankenfeld, F. Kessler, M. Powalla, P. Rogin,
The authors are grateful to Mr. Axel Eicke for perform- and O. Yazdani-Assl, Thin Solid Films 517, 2415 (2009).
33
ing the XRF and SIMS measurements. R. Caballero, C. A. Kaufmann, T. Eisenbarth, A. Grimm, I. Lauermann, T.
This work was funded by the German Research Founda- Unold, R. Klenk, and H. W. Schock, Appl. Phys. Lett. 96, 092104-3
(2010).
tion (DFG) (Contract No. CH 943/2-1). 34
D. Vaumousse, A. Cerezo, and P. J. Warren, Ultramicroscopy 95, 215
(2003).
1 35
B. J. Stanbery, Crit. Rev. Solid State 27, 73 (2002). S. B. Zhang, S. H. Wei, and A. Zunger, Phys. Rev. B 57, 9624 (1998).
2 36
M. Kemell, M. Ritala, and M. Leskelä, Crit. Rev. Solid State 30, 1 (2005). C. Persson and A. Zunger, Phys. Rev. Lett. 91, 266401 (2003).
3 37
L. L. Kazmerski, J. Electron Spectrosc. 150, 105 (2006). C. Persson and A. Zunger, Appl. Phys. Lett. 87, 211904 (2005).
4 38
P. Jackson, M. Powalla, E. Lotter, D. Hariskos, S. Paetel, R. Wuerz, R. S. H. Wei, S. B. Zhang, and A. Zunger, J. Appl. Phys. 85, 7214
Menner, and W. Wischmann, Prog. Photovolt: Res. Appl. 19(7), 894 (1999).
39
(2011). A. Rockett, Thin Solid Films 361-362, 330 (2000).
5 40
J. R. Tuttle, A. Szalaj, and J. Keane, Conf. Rec. 28th IEEE Photovoltaic M. K. Miller and M.G. Hetherington, Surf. Sci. 246, 442 (1991).
41
Spec. Conf., 1042 (2000). C. Heske, D. Eich, R. Fink, E. Umbach, S. Kakar, T. van Buuren, C.
6
J. Hedström, H. Ohlsen, M. Bodegård, A. Kylner, L. Stolt, D. Hariskos, M. Bostedt, L. J. Terminello, M. M. Grush, T. A. Callcott, F. J. Himpsel, D.
Ruckh, and H. W. Schock, Conf. Rec. 23rd IEEE Photovoltaic Spec. L. Ederer, R. C. C. Perera, W. Riedel, and F. Karg, Appl. Phys. Lett. 75,
Conf., 364 (1993). 2082 (1999).

Downloaded 10 Feb 2012 to 193.175.131.12. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions

Você também pode gostar