213_1.pdf

Characterization of Si/SiGe Heterostructures for
Strained Si CMOS
P.M. Mooney*, S.J. Koester*, H.J. Hovel**, J.O. Chu*, K.K. Chan*,
J.L. Jordan-Sweet*, J.A. Ott*, N. Klymco** and D. M. Mocuta**
IBM Semiconductor Research and Development Center
*T.J. Watson Research Center, Yorktown Heights, NY 10598 USA
** IBM Microelectronics Division, Hopewell Junction, NY 12533 USA
Abstract. CMOS devices fabricated in a strained Si layer grown epitaxially on a strain-relaxed SiGe virtual substrate exhibit
enhanced carrier mobility compared to that of devices fabricated in bulk Si. We demonstrate that the thickness and strain state of
the Si layer, and the alloy composition and strain state of the SiGe “virtual substrate” are critical parameters for process
monitoring. The strengths and limitations of several characterization methods including x-ray diffraction, Raman spectroscopy
and spectroscopic ellipsometry for characterization of these layer structures are discussed.
buffer layer serves as a “virtual substrate” for these
devices. The electron mobility increases with
increasing Ge mole fraction in the SiGe buffer layer,
saturating at about 20% Ge [13]. The hole mobility
increases more slowly and saturates at about 35% Ge
[16]. Therefore structures consisting of strained Si on a
relaxed SiGe buffer layer containing 15-35% Ge are of
interest for CMOS applications.
Experimentally, the electron mobility of a
strained Si device on a SiGe layer having 13% Ge has
been demonstrated to be about 70% higher than the
universal MOSFET mobility and that of control Si
devices and it is 110% higher when the SiGe buffer
layer has 28% Ge [7,8]. The mobility enhancement,
which persists at high vertical field in the range
required for future CMOS generations, cannot be
explained entirely by the band structure modifications,
indicating that the tensile strain may also reduce the socalled surface roughness scattering [7,8,13]. The peak
hole mobility enhancement when the SiGe buffer layer
has 28% Ge has been shown to be 45% greater than
that of the Si control device; however, this decreases at
high effective field.
The observed mobility enhancement is very
exciting. However, there are still many materials issues
to be addressed. Key to the success of this technology
will be the ability to measure the strained Si/relaxed
SiGe heterostructures that are required for strained Si
CMOS technology. In this paper we discuss the
characterization of these structures.
INTRODUCTION
The amazing performance improvements in
CMOS circuits during the last 40 years are primarily a
result of reductions in the dimensions of the individual
transistors by orders of magnitude enabled by advances
in photolithography. However, as device dimensions
approach <100 nm, scaling becomes increasingly
difficult. Therefore, new materials that can improve
circuit performance are of interest.
Improved performance due to increased
carrier mobility has been reported for CMOS devices
fabricated on a strained Si layer [1-11]. Theoretical
calculations predict a higher mobility for both electrons
and holes in silicon under tensile strain [12-16].
Tensile strain removes the 6-fold degeneracy in the
conduction band, lowering the energy of the two valleys having electrons with a lower in-plane effective
mass with respect to the four valleys having
electrons with a higher in-plane effective mass. This
energy splitting reduces intervalley scattering and
preferential occupation of the two valleys with lower
in-plane effective mass electrons results in higher
electron mobility. Tensile strain also removes the
valence band degeneracy at the Gamma-point and
shifts the spin-orbit band [1-5], thus improving the hole
mobility.
Typically a Si layer under tensile strain is
achieved by epitaxial growth of Si on a strain-relaxed
SiGe buffer layer on a Si substrate. The relaxed SiGe
CP683, Characterization and Metrology for ULSI Technology: 2003 International Conference,
edited by D. G. Seiler, A. C. Diebold, T. J. Shaffner, R. McDonald, S. Zollner, R. P. Khosla, and E. M. Secula
© 2003 American Institute of Physics 0-7354-0152-7/03/$20.00
213
For these reasons, alternative methods to
achieve a much thinner relaxed SiGe buffer layer are
under investigation [22-27]. One approach, using
molecular beam epitaxy (MBE), employs a special low
temperature (200oC) growth step [22]. Another
approach, also by MBE, uses plasma cleaning prior to
growth [23]. Ion-implantation of hydrogen or helium
has been used to create dislocation nucleation sources
and thus enhance strain relaxation in thin SiGe/Si
structures [24-27]. In this method, a pseudomorphic
SiGe layer is first grown on Si(001) at low
temperature. The SiGe layer must be thin enough that
no strain relaxation occurs during the growth, but thick
enough that it exceeds the critical thickness for
relaxation at higher temperatures. Hydrogen or helium
is then implanted at or below the SiGe/Si interface.
After annealing, spherical defects (bubbles) were
observed at or below the SiGe/Si interface as well as a
dense network of misfit dislocations that relieve the
strain in the SiGe layer. We have investigated strain
relaxation mechanisms in He-implanted SiGe/Si
structures and found that, depending on the He dose
and implantation energy, i.e. the depth of the He atoms
below the SiGe/Si interface, three different strain
relaxation mechanisms occur [27]. For certain implant
conditions, we observed that large platelets are formed
upon annealing rather than bubbles as reported
previously. Dislocations nucleate at the platelets by a
prismatic punching mechanism and a dense regular
misfit dislocation network is formed at the SiGe/Si
interface. Threading dislocation densities in the mid
107 cm-2 range have been reported [24-26]. An
advantage of these buffer layers is the low surface
roughness, which is typically an order of magnitude
smoother than for step-graded buffer layers. And, of
course, the implanted/annealed SiGe layers are much
thinner than step-graded buffer layers.
SiGe “VIRTUAL SUBSTRATES” FOR
STRAINED Si DEVICES
An essential feature of strained Si CMOS
device structures is the strain-relaxed SiGe buffer layer
or virtual substrate. The alloy composition and degree
of strain relaxation of the SiGe buffer layer determines
the magnitude of the strain in the pseudomorphic Si
cap layer. Moreover, the defect density in the SiGe
buffer layer must be sufficiently low to achieve the
necessary yield of devices with good characteristics.
Strain-Relaxed SiGe Buffer Layers
When a strained Si1-xGex layer having x<0.5
grown epitaxially on Si(001) exceeds a critical
thickness, strain relaxation occurs by the introduction
of 60o misfit dislocations near the SiGe/Si interface.
Since Si(001) substrates are essentially dislocation
free, the critical thickness in this case is that needed for
dislocation nucleation; the critical thickness decreases
as the Ge mole fraction of the SiGe layer is increased
and also depends on the growth temperature [18-20].
A relatively thick SiGe layer may remain fully strained
(pseudomorphic) when grown at low temperature;
however, strain relaxation occurs when a metastable
structure is subsequently heated, e.g. during device
fabrication.
Since dislocations must terminate at a free
surface, each 60o misfit dislocation nucleated at the
initial growth interface or in the epitaxial layer ends
with two threading arms that go through the SiGe layer
to the wafer surface. Relatively low threading
dislocation densities have been achieved by means of
compositional grading. The graded buffer typically
consists of a compositionally graded SiGe layer
followed by a uniform composition SiGe layer. The
alloy composition may be linearly graded or step-wise
graded. Threading dislocation densities in the range
105-108 cm–2 have been reported, with the lowest
densities occurring for grading rates as slow as 10% Ge
per micron of layer thickness [21]; thus graded Si1-xGex
buffer layers for strained Si CMOS applications may
be as thick as 2-4 m. Because SiGe has a lower
thermal conductivity than Si, self-heating effects are
observed in strained Si CMOS devices fabricated on
step-graded buffer layers [7,8]. Graded buffer layers
also have relatively rough surfaces; a cross-hatch
surface pattern is created by the surface steps
associated with the 60o misfit dislocations. The surface
of the relaxed SiGe buffer layer can be planarized prior
to the growth of the active device layers using
chemical-mechanical polishing methods. However, the
polishing step adds to the cost of fabrication, as does
the growth of such a thick SiGe buffer layer.
SiGe-on-Insulator Substrates
The advantages of strained Si can be
combined with the advantages of SOI substrates by
fabricating SiGe-on-insulator (SGOI) substrates.
Several different methods to fabricate SGOI using a
relaxed SiGe buffer layer on Si(001) have been
implemented. One group fabricated SGOI substrates
by means of a SIMOX-like process, i.e. implantation of
oxygen below the surface of the SiGe buffer layer and
then annealing at high temperature (>1300oC) to form a
layer of SiO2 [28,29]. Apparently this method works
only for SiGe buffer layers having low (~10%) Ge.
Another approach uses dry oxidation at high
temperature after deposition of a pseudomorphic SiGe
layer with low Ge content [30]. During oxidation, Ge
atoms are rejected from the oxide and condensed in the
214
remaining SiGe layer, which is partially relaxed
without introducing a significant number of
dislocations. Raman spectroscopy shows that a strained
Si cap layer grown on such a structure has about 1%
tensile strain.
Wafer bonding and layer transfer is a very
promising method to fabricate SGOI substrates with
higher Ge content [9,31-33]. The fabrication steps
involve implantation of hydrogen into the relaxed SiGe
buffer layer, then bonding the implanted wafer to a Si
wafer with an SiO2 layer on top, then annealing to split
the SiGe layer and strengthen the bonded interface, and
finally epitaxial growth of the strained Si device layers
on the bonded SGOI substrate. Key process steps also
needed are chemical-mechanical polishing to planarize
the surface of the relaxed SiGe buffer layer and
subsequent cleaning of the polished surface prior to
bonding and also after layer splitting before growing
the device layers.
Temperature
dependent
Hall
effect
measurements performed on modulation-doped
structures grown simultaneously on a polished SiGe
buffer layer and on a polished SGOI substrate having
the same SiGe alloy composition show the presence of
a two dimensional electron gas and identical electron
mobility, ~2000 cm2/Vs at 300 K [31]. This indicates
that the quality of the SiGe layer on the bonded SGOI
wafer is the same as that of the relaxed SiGe buffer
layer used to fabricate the SGOI wafer. An
enhancement of about 50% in the effective electron
mobility was observed in strained Si MOSFETS
fabricated on a bonded SGOI substrate having 15% Ge
[9,32], comparable to the results for strained Si
MOSFETs on thick relaxed SiGe buffer layers on Si
substrates. Hole mobility enhancements of 15-20%
were found at low vertical field [9,32]; at high vertical
field, however, there is negligible mobility
enhancement, again comparable to results for “bulk”
strained Si pFETs.
We note that as device dimensions are
reduced it will be interesting to transfer only a thin
strained Si layer to form strained Si-on-insulator
(SSOI) wafers. Initial experiments show that this can
be done [11].
addition to the thickness and alloy composition of the
SiGe buffer layer and the thickness of the strained Si
layer, the strain state of these layers must also be
evaluated.
High-resolution x-ray diffraction is the usual
method to measure both the alloy composition and
strain state of the relaxed SiGe buffer layer. The lattice
parameter of bulk (cubic) SiGe increases almost
linearly as the Ge fraction of the alloy increases. The
strain in the epitaxial film can be determined by
measuring the difference between the in-plane and outof plane lattice parameters. When a SiGe layer is
pseudomorphic, i.e. when the defect density is
negligible and thus the in-plane lattice parameter of the
SiGe layer is equal to that of the Si substrate, the Ge
fraction of the alloy is determined from the
measurement of the out-of-plane lattice parameter of
SiGe layer, i.e. from the angular separation between
the SiGe layer peak and that of the Si substrate on the
symmetric 004 x-ray rocking curve. The strong
thickness fringes that are present on the x-ray rocking
curve in this case allow the layer thickness to be
determined as well.
When the SiGe layer relaxes, the misfit
dislocations that are present result in mosaic
broadening of the x-ray diffraction peak and the
thickness fringes are washed out. In this case both the
in-plane and out-of-plane lattice parameters must be
measured and the alloy composition and the strain are
determined using both the symmetric 004 rocking
curve and an asymmetric rocking curve, typically the
grazing exit 224 [34]. The 004 reflection measures the
crystal lattice spacing in the direction of the surface
normal, i.e. the out-of-plane lattice spacing. The 224
reflection measures the spacing of lattice planes at an
angle to the surface normal, i.e. a set of planes whose
lattice spacing has both an out-of-plane and an in-plane
component. The composition and the strain are
calculated from the angular separation between the
SiGe layer and the Si substrate peaks in the two
measurements [34]. In the case of a graded SiGe buffer
layer, it is necessary to use triple-axis x-ray diffraction,
i.e. to insert an analyzer crystal in front of the detector
to reduce the angular aperture. This suppresses the
mosaic broadening of the diffraction peak thus
allowing the various different lattice parameters
corresponding to the different SiGe layers to be
detected. Here again both the symmetric and
asymmetric reflections must be measured to determine
the alloy composition and the strain of the various
SiGe layers.
Measurement of a thin strained Si layer grown
on a graded SiGe buffer layer is much more difficult.
Even though the strained Si layer is pseudomorphic to
the underlying SiGe buffer layer, the mosaic structure
of the SiGe layer is replicated in the strained Si layer,
CHARACTERIZATION OF STRAINED
Si DEVICE STRUCTURES BY HIGHRESOLUTION X-RAY DIFFRACTION
As emphasized in the previous sections, the
electrical properties of strained Si MOSFETs depend
strongly on the degree of strain in the strained Si layer.
This is usually determined by the composition and
strain state the underlying SiGe layer. Therefore in
215
which is typically <30 nm thick. Triple-axis x-ray
diffraction measurements were performed at beamline
X20 at the National Synchrotron Light Source at
Brookhaven
National
Laboratory [35].
The
composition and strain of the uniform Si1-xGex layer
were determined as described above. Fig. 1(a) shows
the 004 data for a 21 nm-thick Si cap layer on a
Si0.72Ge0.28 buffer layer taken at two different regions,
one with and one without the strained Si cap layer.
Comparing the two scans, the features to the right of
the Si substrate peak are from the strained Si cap layer.
Although the Si layer thickness can be determined
from the solid curve in Fig. 1(a), comparison with the
simulation is more precise when the two scans are
subtracted. The difference data obtained by subtracting
the two scans in Fig. 1(a) is shown in Fig. 1(b), along
with a simulated diffraction scan from RADS [36]. The
thickness of the Si cap layer determines the spacing of
the oscillations and the strain determines their position
with respect to the Si substrate peak. The shift of the
experimental data from the Si cap layer with respect to
the Si substrate peak is clearly visible. With this
method of analysis the relaxation in the Si cap layer
with respect to the underlying SiGe layer can be
determined to within 1 or 2%, depending on the
sample.
Intenity (arb. units)
3
004
Si
substrate
5
10
Si1-xGex
layer
10
Si cap layer
1
10
10
-1
10
-3
Si cap layer removed
34.0
34.5
35.0
35.5
36.0
θB (deg.)
Fig. 1(a). 004 x-ray scans taken at regions with and without
the 21nm-thick strained Si cap layer on Si0.72Ge0.28. Reprinted
with permission from [35].
4
10
Intenity (arb. units)
Si
substrate
Thermal stability of strained Si structures
Strained Si CMOS devices, implemented by
the epitaxial growth of a 10-30 nm-thick strained Si
layer on a relaxed Si1-xGex buffer layer on Si(001), are
typically grown by a variety of epitaxial methods at
relatively low temperatures. However, device
fabrication includes standard processes, e.g. the
activation of ion-implanted dopant atoms and gate
oxidation, requiring temperatures as high as 1000oC.
When threading dislocations are present in the
substrate, an epitaxial strained layer that exceeds the
critical thickness for dislocation glide will relax by the
formation of misfit dislocations at the Si/SiGe interface
[37]. The rate of misfit dislocation formation increases
exponentially with temperature. Interdiffusion at the
Si/SiGe interface also occurs at this temperature,
resulting in a graded interface and effectively reducing
the thickness of the strained Si layer. We have
investigated these effects by annealing Si/Si1-xGex
structures, consisting of a strained Si cap layer of
varying thickness on a step-graded Si1-xGex buffer layer
of varying Si1-xGex alloy concentration, at 1000oC for
5, 30 and 300 seconds, times which far exceed those
normally used for Si CMOS fabrication [35]. The goal
of this work was to measure the magnitude of the strain
relaxation of the Si layer and the interdiffusion that
occurs at the Si/SiGe interface during annealing and
to be sure that the measurement methods used could
simulation
difference
data
2
10
Si cap layer
0
10
-2
10
peak
shift
004
-4
10
-0.28 0.00
fringes
0.28
0.56
0.83
1.11
1.39
θB (rel. deg.)
Fig. 1(b). The difference of the two x-ray scans shown in Fig.
1(a) with the simulated x-ray data. Reprinted with permission
from [35].
detect changes in these structures.
The x-ray data of Fig. 1(b) are compared with
those from annealed pieces of the same wafer in Fig. 2,
where the scans are offset along the vertical axis for
clarity. Here we see essentially no change after
annealing for 5 seconds, but a small shift of the main
diffraction peak toward the Si substrate peak is seen
after 30 sec and a larger shift is seen after 300 sec. The
thickness fringes are more spread out after 30 sec and
even more so after 300 sec. These results clearly show
that the Si layer becomes thinner with longer annealing
times, and also that the strain in the layer decreases as
the annealing time is increased.
Similar changes were observed in other
samples. The thickness change for several different
216
samples is summarized in Fig. 3. The change in the Si
cap layer thickness showed no dependence on the
initial thickness of the Si cap layer or the alloy
composition of the SiGe layer over the ranges
investigated [35]. Since these samples were annealed in
an inert atmosphere, the most likely explanation is that
interdiffusion occurs at the Si/Si1-xGex interface,
effectively reducing the thickness of the pure Si layer.
These changes in thickness are consistent with
experimental values for the diffusion coefficients
reported for Ge in Si [38-39]. Interdiffusion results in
a graded interface rather than the abrupt one achieved
during epitaxial growth at low temperature. This is an
advantage for the pMOSFETs, since a graded
interface reduces the probability of parallel hole
conduction in the top of the relaxed Si1-xGex buffer
layer [3].
Fig. 4 shows the time dependence of the strain
relaxation for two samples. The Si cap layer is 0%
relaxed when the in-plane lattice parameter matches
that of the underlying Si1-xGex layer and it is fully
relaxed when the in-plane lattice parameter has the
value of bulk Si. Even in the worst case, very little
strain relaxation occurs (~10%). The trends observed
are exactly as expected; greater relaxation occurs for
higher Ge mole fraction and for thicker Si layers. The
x-ray results agree well with the strain relaxation
calculated from the misfit dislocation density
determined from planar-view transmission electron
microscopy (TEM) images, such as the one shown in
Fig. 5 [35]. Other TEM images (not shown) clearly
show that the misfit segments terminate at threading
dislocations in the upper part of the Si1-xGex buffer
layer, indicating that strain relaxation occurs by the
glide of threading dislocations in the relaxed Si1-xGex
layer.
21 nm strained Si/relaxed Si0.72Ge0.28
1
10
o
1000 C anneal
0
10
-1
10
-2
10
-3
10
-4
10
-5
10
-6
as-grown
5 sec
30 sec
300 sec
% Strain Relaxation in Si Layer
Intensity (cts/sec)
10
bulk
Si
0
1000
2000
3000
4000
5000
Bragg Angle(deg.)
Fig. 2. Difference scans from a sample taken after
annealing for various times. The curves are off-set in the
vertical direction for clarity.
12
10
8
21 nm cap
14 nm cap
TEM data
6
4
2
0
Si0.72 Ge 0.28 buffer layer
-2
10
100
o
Change in Layer Thickness (nm)
Time at 1000 C (sec)
2
Fig. 4. Percent strain relaxation as a function of
annealing time. Strain relaxation calculated from the
misfit dislocation density from planar view TEM images
is also plotted. Reprinted with permission from [35].
0
-2
-4
0.28/14nm
0.28/21nm
0.27/18nm
0.19/23nm
0.30/18nm
0.23/23nm
-6
-8
-10
1
10
To avoid the formation of misfit dislocations
at the strained Si/SiGe interface during device
fabrication, the strained Si/SiGe heterostructures must
be thermodynamically stable. The SiGe graded buffer
layers are essentially fully relaxed and no change in
their lattice parameter was detected after annealing.
The strained Si layer is stable provided the layer
thickness is less than the critical thickness for
dislocation glide. If metastable heterostructures are
used, device fabrication process conditions must be
restricted to temperatures and times for which
dislocation motion is negligible. Although the
100
o
Time at 1000 C (sec)
Fig. 3. Change in strained Si layer thickness with annealing
time in several samples having initial Si layer thickness and
Si1-xGex alloy composition indicated. The uncertainty is
+1.5 nm. Reprinted with permission from [35].
217
Depending upon the wavelength of the incident
radiation, surface sensitivities in the range of 10 nm to
1000 nm, can also be achieved.
We have previously described Raman analysis
using λ = 488 nm on strained Si on relaxed SiGe
heterostructures [40]. Those results showed that no
significant strain relaxation occurred in Si/SiGe layers
at annealing temperatures up to 1000 oC for 300 sec.
We also showed that Ge diffusion into the Si cap could
be detected in the Raman spectra by a drop in the
intensity of the Si-Si phonon mode arising from the
strained Si layer. More recently we showed that the
change in the Si thickness extracted from Raman
spectra agreed quantitatively with results from highresolution x-ray diffraction (HRXRD) measurements
[35].
Here we describe new measurements using
Raman spectroscopy at λ = 442 nm excitation and
compare with our previous results at λ = 488 nm as
well as with SIMS and HRXRD measurements. We
show that the shorter wavelength improves the surface
sensitivity by a factor of about 2, and therefore allows
measurement of thinner Si capping layers. We also
discuss the limitations of Raman measurements for
determining the changes in the Si cap thickness that
result from interdiffusion at the Si/SiGe interface.
Finally, we describe the limitations in the accuracy of
Raman measurements for determining the strain
relaxation in the Si cap layer.
Fig. 6 shows a comparison of Raman spectra
on several Si/Si1-xGex samples at excitation
wavelengths of 488 nm and 442 nm. The thickness, d,
of the Si surface layer is 12, 21 or 31 nm and the Ge
mole fraction, x, of the underlying SiGe layer is 23% as
determined by HRXRD. In each case, the plots show
the Si-Si phonon modes arising from the SiGe layer
and the strained Si layer. The plot taken at 442 nm
shows a stronger signal arising from the strained Si
layer relative to the SiGe layer. The greater surface
sensitivity at shorter wavelengths allows samples with
thinner caps to be analyzed. At λ = 488 nm, the
strained Si layer is barely visible at d = 20 nm, and at
d = 12 nm, cannot be resolved, even with curve fitting.
On the other hand, at λ = 442 nm, all thickness of the
Si cap can be detected.
Our previous analysis showed that the Si cap
thickness can be determined using Raman
spectroscopy, given a suitable calibration point, and
changes in the Si cap thickness (e.g., as a function of
annealing time or temperature), can also be
determined. We have subsequently performed highresolution SIMS characterization of these structures
and compared the results with Raman spectra taken at
442 nm. We have utilized the following methodology
for determining the cap thickness. After subtracting
the baseline intensity, the Raman data are fit using a
Fig. 5. Planar view TEM image of a 21 nm-thick strained
Si layer on a relaxed Si0.72Ge0.28 buffer layer annealed for
30 sec at 1000 oC. The scale bar is 0.5 m. The misfit
density is 2.5 m-1. Reprinted with permission from [35].
absolute strain in the Si layer is important for carrier
mobility, it is the change in mismatch strain with
respect to the underlying SiGe buffer layer that
determined the misfit dislocation density at that
interface. Therefore it is important to monitor both the
SiGe buffer layer and the strained Si cap layer.
OPTICAL CHARACTERIZATION
METHODS
While laboratory x-ray methods are useful for
routine characterization of strained Si CMOS device
wafers, x-ray measurements that require a synchrotron
source are not suitable for timely measurements of a
large number of wafers. Thus x-ray diffraction can be
used to monitor relaxed SiGe buffer layers, but not for
monitoring the thin strained Si cap layer. Although the
thickness of the strained Si layer can be measured by
cross-sectional TEM, the lengthy sample preparation
required for TEM measurements makes this method
unsuitable for routine characterization. We have
therefore investigated optical methods to determine
their potential for metrology of these layer structures.
Raman Spectroscopy
Micro-Raman spectroscopy is a useful
technique for the analysis of strained Si on relaxed
SiGe heterostructures because it is non-destructive,
allows high spatial resolution and has high throughput.
218
double Lorentzian line shape in order to the determine
the intensity, Raman shift and full-width halfmaximum (FWHM) of the peaks arising from the Si-Si
vibrational modes in the SiGe buffer layer and the
strained Si cap layer. After curve fitting, the ratio of
the strained Si peak area, ASS, to the total fit area, ASS +
ASiGe, is used to extract the thickness change of the Si
cap, ∆d, as a function of annealing condition, y,
according to the following formulas:
∆d(y) = [1 - r(y)/r(0)] d(0),
(1)
r(y) = {ASS(y) / [ASiGe(y) + ASS(y)]}.
(2)
for various times. The initial layer thickness was
determined using HRXRD as described in previous
sections. The thickness change from SIMS was
determined as the point at which the Ge concentration
dropped by half. The results are shown in Fig. 7,
where the change in the Si cap thickness is plotted vs.
annealing time in (a) and temperature in (b). The data
show good agreement between the SIMS and Raman
data, though the accuracy in the cap thickness
measurement determined by Raman is still limited to
about + 2 nm.
The analysis described above to determine the
Si cap thickness by Raman assumes a simple two-layer
system, when it is known that a significant amount of
interdiffusion occurs after annealing at high
temperature. To test the validity of this assumption, a
three-layer system was modeled and simulated spectra
were generated. The three-layer system consists of a Si
cap layer, with thickness dSi, a Si1-xGex substrate and an
Two sets of samples and annealing conditions were
utilized. Sample A has x = 23%, d = 31 nm, and was
annealed at 1000oC for varying times, and also
annealed for 30 sec at various temperatures. Sample B
has x = 30%, d = 18 nm, and was annealed at 1000 oC
10
Si thickness Change (nm)
Intensity (arb. units)
6000
tSi = 31 nm
tSi = 21 nm
tSi = 12 nm
3000
0
6
4
2
o
T = 1000 C
0
500
510
520
SIMS, sample A
Raman, sample A
SIMS, sample B
Raman, sample B
8
530
1
10
100
Anneal Time (sec)
-1
Wavenumber (cm )
Fig. 6(a). Raman spectra with λ = 488 nm excitation of
Si/SiGe samples with different thickness of the strained Si
layer.
Fig. 7(a). Comparison of thickness changes extracted from
Raman data and SIMS data as a function of annealing time
for two different sample sets.
20
Si thickness Change (nm)
Intensity (arb. units)
6000
tSi = 31 nm
tSi = 21 nm
tSi = 12 nm
3000
15
10
5
t = 30 sec
0
0
500
510
520
SIMS, sample A
Raman, sample A
1000
1100
o
Anneal Temperature ( C)
530
-1
Wavenumber (cm )
Fig. 7(b). Comparison of thickness changes extracted from
Raman data and SIMS as a function of annealing
temperature.
Fig. 6(b). Raman spectra of same samples as in (a) using
442 nm excitation.
219
25
25
20
20
15
15
10
10
dislocations, it is important for a characterization
method to be extremely sensitive to small variations in
the in-plane lattice constants of the two layers.
However, due to several uncertainties in the system
parameters this is difficult task using Raman
spectroscopy. First of all, the strained Si peaks must be
compared to a Si control. Secondly, the strained Si
peak is sometimes small and can be difficult to fit
accurately. As a function of thermal annealing, the
strained Si peak position can be shifted due to
interdiffusion at the Si/SiGe interface, as shown in Fig.
8. Finally, the SiGe in-plane lattice constant cannot be
uniquely determined by solely examining the Si-Si
peak arising from the SiGe layer. This is because the
peak position is determined not only by the Ge
concentration of the SiGe layer, but also by the residual
strain in the layer. If the SiGe layer is not fully
relaxed, relaxation can occur after annealing, causing
greater strain in the Si layer, and counteracting any
apparent loss of strain caused by misfit dislocations.
Our analysis on Si/Ge layers using Raman
spectroscopy has shown that the net relaxation of the Si
cap layer, relative to the underlying layer, can only be
determined with an accuracy of +2-4%, and is
therefore much less accurate than other methods such
as HRXRD and XTEM. Therefore, for routine analysis
of non-patterned Si/SiGe layers, Raman spectroscopy
is of limited value for determination of the misfit
formation and plastic relaxation due to annealing.
However, Raman measurements could be useful for
analyzing strain relaxation in patterned structures,
where much greater relaxation is expected to occur and
where other analysis methods do not have sufficient
spatial resolution.
Extracted peak shift (cm )
Extracted dSi (nm)
intermediate Si1-yGey region with Ge content linearly
graded between 0 and x, and thickness dint. The
simulated Raman data from this system, was
nevertheless fit using a double Lorentzian line shape to
determine at what thickness of the interdiffused region
the double Lorentzian fit becomes invalid. Since we
assume that the Si cap thickness is defined at the point
where the Ge concentration decreases by 50%, the Si
cap thickness should not change as long as the
interdiffusion is symmetric. Also, since no strain
relaxation is assumed, the position of the Si-Si peak
arising from the Si cap should also remain constant.
The results are shown in Fig. 8, which shows the
extracted Si cap thickness and strained Si peak position
as a function of dint/dSi. The plot shows that when the
thickness of the interdiffused region is over roughly
50% of the original cap thickness, significant error in
the peak area and position is seen to occur, setting an
upper limit on the validity of the two-layer model. A
more sophisticated three-layer model can be used to
account for the interdiffusion, but the number of
adjustable parameters makes accurate curve fitting
difficult. For the results shown in Fig. 7, the
interdiffused region is much thinner than the limit
described above, indicating that the two-layer fit is a
valid assumption for these layers.
Spectroscopic Ellipsometry
-1
0.0
0.2
0.4
0.6
0.8
Film
thickness
measurements
by
spectroscopic ellipsometry have become a valuable
part of semiconductor metrology. This method relies
on analyzing the reflection of polarized light for
wavelengths from the UV to near-IR in terms of the
Fresnel coefficients of each material in a film stack.
For Si1-xGex and Si layers, the optical properties of the
two materials are sufficiently different, even at
relatively low values of the Ge mole fraction x, that
layers of the two materials are easily distinguished.
Graded buffer layers, whether smoothly graded or stepwise graded, are modeled by a series of layers each of
fixed composition and thickness, such that the total
thickness equals that of the buffer layer. Intermixing at
boundaries resulting from heat treatments can be
handled the same way. Roughness at the upper surface
is handled by modeling as a thin layer comprised of a
mixture of native oxide and “void.”
1.0
dint / dSi
Fig. 8. Plot showing the extracted Si thickness and peak
position as a function of the relative thickness of the
interdiffused region.
Raman spectroscopy is also useful in
determining the amount of strain in strained Si on
relaxed SiGe layers, and the strain relative to bulk Si
can be determined with good accuracy. However, what
is perhaps most important for CMOS applications is to
determine whether or not the strained Si layer is
relaxing via the formation of misfit dislocations at the
Si/SiGe interface. These misfit dislocations can
possibly lead to yield and reliability problems for
devices. Since only a small degree of strain relaxation
is needed to cause a significant degree of misfit
220
The ellipsometer measures the reflectances rp
and rs of light polarized parallel and perpendicular to
the plane of incidence, respectively. The data are
usually displayed as tan Ψ and cosine ∆, where
rp / rs = tan Ψ exp( i ∆ ).
values measured by other methods. For the annealed
samples discussed above, ellipsometry measurements
show comparable thickness changes for the strained Si
layer, but the measured thickness values are typically
somewhat lower that those determined from x-ray
diffraction and the composition of the SiGe layer is a
lot lower than that determined from x-ray
measurements. This difference may arise because the
optical constants for SiGe layers taken from the
literature are not accurate for the dislocated SiGe layers
in the graded buffer layer structures.
(3)
As an example, tan Ψ and cosine ∆ for a strained
Si/SiGe stack consisting of a native oxide, a 20 nm
strained Si layer, and several layers of SiGe with
different thickness and composition are shown in Fig.
9. The layer thickness and alloy composition, but not
the strain, are readily obtained from such data,
provided the correct optical properties are used in the
model. Grading in either of the SiGe layers requires a
more complicated model. If the SiGe buffer layers are
made thicker, more peaks appear in the spectra and the
analysis is facilitated by extension of measurement
wavelengths farther into the infrared.
CONCLUSIONS
Strained Si CMOS devices fabricated on 200
mm wafers using standard CMOS fabrication processes
exhibit increased carrier mobility primarily due to the
modified band structure in Si under tensile strain.
These devices are fabricated in a strained Si layer
grown epitaxially on a strain-relaxed SiGe buffer layer
that acts as a “virtual substrate”. The relaxed SiGe
layer may also be the thin SiGe layer of an SGOI
substrate. High-resolution x-ray diffraction is very
useful for evaluating the alloy composition and strain
state of the relaxed SiGe buffer layer, but cannot
evaluate the strained Si cap layer on these samples
unless a synchrotron x-ray source is used. Raman
spectroscopy and spectroscopic ellipsometry are
interesting for metrology applications. The absolute
strain in the Si cap layer can be measured using Raman
spectroscopy and the layer thickness can be determined
when a suitable calibration is performed. The thickness
of the strained Si layer is easily measured by
spectroscopic ellipsometry, as is the composition and
thickness of the SiGe buffer layer in some types of
structures. We note that the measurement error inherent
in some of these methods is as large as 3-4 nm, a
significant fraction of the thickness of the strained Si
cap layer. We also note that x-ray reflectivity is another
possible method for measuring the strained Si layer
thickness; however, this method can only be used for
structures having very flat surfaces and interfaces.
TAN PSI
0.8
0.6
0.4
0.2
0
0.25
0.35
0.45
0.55
0.65
0.75
WAVELENGTH, m icron
COSINE DEL
1
0.5
0
-0.5
-1
0.25
0.35
0.45
0.55
0.65
0.75
WAVELENGTH, micron
Fig. 9. Ellipsometric simulation for a structure (model)
consisting of 1 nm SiO2 / 20 nm Si / 30 nm Si0.85Ge0.15 /
500 nm Si0.65Ge0.35 on Si.
ACKNOWLEDGEMENTS
The work at NSLS was partially supported by
the US Department of Energy, Materials Science and
Chemical Science Divisions.
In practice, we have found that this method
gives good results for some types of Si/SiGe layer
structures, including strained Si on SGOI and strained
Si on implanted/annealed buffer layers. For example,
ellipsometry measured a strained Si layer thickness of
6.5 nm compared to 7.9 nm measured by HRXRD and
7.2 nm by cross-sectional TEM. The difference is
within the error of the x-ray measurement. However,
ellipsometry results obtained for strained Si on stepgraded SiGe buffer layers do not agree well with the
REFERENCES
1.
2.
221
Welser, J., Hoyt, J.L., Takagi, S., and Gibbons,
J.F., IEDM Tech Dig. 373-6 (1994).
Nayak, D.K., Goto, K., Yutani, A., Murota, J., and
Shiraki, Y., IEEE Trans. Elect. Dev. 43 1709-16
(1996).
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
Rim, K., Welser, J., Hoyt, J.L., and Gibbons, J.F.,
IEDM Tech Dig. 517-20 (1995).
Rim, K., Hoyt, J.L., and Gibbons, J.F., IEDM
Tech. Dig. 707-10 (1998).
Mizuno, T., Takagi, S., Sugiyama, N., Koga, J.,
Tezuka, T., Usuda, K., Hatakeyama, T., Kurobe,
A., and Toriumi, A., IEDM Tech. Dig. 934-6
(1999).
Mizuno, T., Sugiyama, N., Satake, H., Takagi, S.,
Symposium on VLSI Technology Dig. 210-11
(2000).
Rim, K., Koester, S., Hargrove, M., Chu, J.,
Mooney, P.M., Ott, J., Kanarsky, T., Ronsheim, P.,
Ieong, M., Grill, A., Wong, H.-S.P., Symposium on
VLSI Technology Dig. 59-60 (2001).
Rim, K., Chu, J., Chen, H., Jenkins, K.A.,
Kanarsky, T., Lee, K., Mocuta, A., Zhu, H., Roy,
R., Newbury, J., Ott, J., Petrarca, K., Mooney, P.,
Lacey, D., Koester, S., Chan, K., Boyd, D., Ieong,
M., Wong, H.-S., Symposium on VLSI Technology
Dig. 98-9, (2002).
Huang, L.J., Chu, J.O., Goma, S.A., D’Emic, C.P.,
Koester, S.J., Canaperi, D.F., Mooney, P.M.,
Cordes, S.A., Speidell, J.L., Anderson, R.M.,
Wong, H.-S.P., Symposium on VLSI Technology
Dig. 57-58 (2001).
Sugii, N., Hisamoto, D., Washio, K., Yokoyama,
N. and Kimura, S., IEDM Tech. Dig. 737-40
(2001).
Rim, K., Shi, L., Chan, K., Boyd, D., Jenkins, K.,
Ott, J., Lacey, D., Mooney, P., Cobb, M., Klymco,
N., Jamin, F., Koester, S. and Kararsky, T., paper
presented at the 1st International SiGe Technology
and Device Meeting, Nagoya Japan, January 15-17,
2003.
Fischetti, M.V., and Laux, S.E., J. Appl. Phys. 80,
2234-52 (1996).
Takagi, S., Hoyt, J.L., Welser, J.J., and Gibbons,
J.F., J. Appl. Phys. 80, 1567-77 (1996).
Roldán, J.B., Gámiz, F., López-Villanueva, J.A.,
and Carceller, J.E., J. Appl. Phys. 80, 5121-28
(1996).
Formicone, G., Vasileska, D., and Ferry, D.K.,
Solid State Elec. 41, 879-85 (1997).
Oberhuber, R., Zandler, G., Vogl, P., Phys. Rev. B
58, 9941-8 (1998).
Nayak, D.K., Woo, J.C.S., Park, J.S., Wang, K.L.,
and MacWilliams, K.P., Appl. Phys. Lett. 62,
2853-5 (1993).
Matthews, J.W., Mader, S., and Light, T.B., J.
Appl. Phys. 70, 3800-04 (1970); Matthews, J.W.,
Blakeslee, A.E., and Mader, S., Thin Solid Films
33, 253-66 (1976).
Bean, J.C., Feldman, L.C., Fiory, A.T., Nakahara,
S., and Robinson, I.K., J. Vac. Sci. Technol. A 2
(1984) 436-440; Hull, R., in Silicon Germanium
and SiGe:Carbon, EMIS Datareviews Series No.
24, INSPEC, London (2002) pp.21-41.
Stiffler, S.R., Comfort, J.H., Stanis, C.L., Harame,
D.L., and de Fresart, E., J. Appl. Phys. 70, 1416-20
(1991).
21. Mooney, P.M., Materials Science and Engineering
Reports, R17, 105-146 (1996), and references
therein.
22. Bauer, M., Lytovich, K., Oehme, M., Kasper, E.,
Herzog, H.J., Ernst, F., Thin Solid Films 369, 152156 (2000).
23. Kuchenbecker, J., Kibbel, H., Muthsam, P., Konig,
U., Thin Solid Films 389, 146-152 (2001).
24. Follstaedt, D.M., Meyers, S.M., Lee, S.R., Appl.
Phys. Lett. 69, 2059-61 (1996).
25. Mantl, S., Hollaender, B., Leidtke, R., Mesters, S.,
Herzog, H.J., Kibble, H., Hackbarth, T., Nucl.
Instr. and Meth. Phys. Res. B 147, 29-34 (1999).
26. Luysberg, M., Kirch, D., Trinkhaus, H.,
Hollaender, B., Lenk, S., Mantl, S., Herzog, H.J.,
Hackbarth, T., Fichtner, P.F., J. Appl. Phys. 92,
4290-5 (2002).
27. Christiansen, S.H., Mooney, P.M., Chu, J.O., and
Grill, A., in Materials Issues in Novel Si-Based
Technology, Mat. Res. Soc. Symp. Proc. 686, 2732 (2002).
28. Sugiyama, N., Mizuno, T., Takagi, S., Koike, M.,
Kurobe, A., Thin Solid Films 369, 199-202 (2002).
29. Mizuno, T., Takagi, S., Sugiyama, N., Satake, H.,
Kurobe, A., Toriumi, A., IEEE Elect. Device Lett.
21, 230-2 (2000).
30. Tezuka, T., Sugiyama, N., and Takagi, S., Appl.
Phys. Lett. 79, 1798-1800 (2001).
31. Huang, L.J., Chu, J.O., Canaperi, D.F, D’Emic,
C.P., Anderson, R.M., Koester, S.J., and Wong, H.S.P., Appl. Phys. Lett. 78, 1267-9 (2001).
32. Huang, L., Chu, J.O., Goma, A., D’Emic, C.P.,
Koester, S.J., Canaperi, D., Mooney, P.M., Cordes,
S.A., Speidell, J.L., Anderson, R.M., and Wong,
H.-S.P., IEEE Trans. Electron Devices 49, 156671 (2002).
33. Cheng, Z.-Y., Currie, M.T., Leitz, C.W., Taraschi,
G., Fitzgerald, E.A., Hoyt, J.L., and Antoniadis,
D.A., IEEE Electron. Device Lett. 22, 321-3
(2001).
34. Mooney, P.M., Jordan-Sweet, J.-L., Stephenson,
G.B., LeGoues, F.K., and Chu, J.O., Advances in
X-Ray Analysis 38, 181-93 (1995).
35. Mooney, P.M., Koester, S.J., Ott, J.A., JordanSweet, J.L., Chu, J.O., and Chan, K.K., in
Materials Issues in Novel Si-Based Technology,
Mat. Res. Soc. Symp. Proc. 686, 3-8 (2002).
36. Rocking Curve Analysis by Dynamical Simulation,
Bede Scientific, Inc.
37. Matthews, J.W., J. Vac. Sci. Technol. 12, 126-33
(1975).
38. Baucaud, P., Wu, L., Guedji, C., Julien, H.F.,
Sahnes, I., Campidelli, Y., and Garchery, L., J.
Appl. Phys. 80, 1414-21 (1996).
39. Griglione, M., Anderson, T.J., Haddara, Y.M.,
Law, M.E., Jones, K.S., and van den Bogaard, A.
J., Appl. Phys. 88, 1366-72 (2000).
40. Koester, S.J., Rim, K, Chu, J.O., Mooney, P.M.,
and Ott, J.A., Appl. Phys. Lett. 79, 2148-50 (2001).
222