You are on page 1of 4

Applied Surface Science 224 (2004) 193–196

Local atomic structure in Czochralski-grown


Ge1xSix bulk alloys
I. Yonenaga*, M. Sakurai, M.H.F. Sluiter, Y. Kawazoe
Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan

Abstract

We investigated the local atomic structure in GeSi alloy experimentally and theoretically. Extended X-ray absorption fine
structure measurements in Czochralski-grown Ge1xSix bulk alloys showed that the Ge–Ge and Ge–Si bond lengths maintain
distinctly different lengths and vary linearly with alloy composition across the whole composition range 0 < x < 1. This is in
good agreement with the expectation derived from ab initio electronic structure calculations. The result indicates that the bond
lengths and bond angles are distorted with alloy composition in GeSi.
# 2003 Elsevier B.V. All rights reserved.

PACS: 61.10.Ht; 61.43.Dq; 61.66.Dk

Keywords: Germanium–silicon; Silicon–germanium; Bond length; Bond angle; Extended X-ray absorption fine structure; Bulk alloy

1. Introduction spectroscopy has been performed to determine bond


lengths for different types of neighbor atoms and the
Ge1xSix or SixGe1x (where x indicates the mole corresponding fractional occupancy of each type of
fraction of silicon) is a complete solid solution system, neighbor [1–6]. However, the local structure of GeSi
having the diamond-cubic structure. The 4.2% differ- crystal is still in question. Previously, the GeSi alloys
ence in the lattice constants of the constituent Ge and used for these studies were grown as thin films on Si
Si atoms leads to various unique alloying effects on the substrates by various epitaxial techniques. In such
electronic, optical, thermodynamic, and mechanical hetero-structures, the large inherent biaxial stress,
properties. GeSi alloys, thus, are important for micro- even when relaxed through the induction of disloca-
electronics and opto-electronics in view of the possi- tions, hinders the quantitative study of the native
bilities of band gap engineering. Knowing accurately, material properties. From such a viewpoint, we started
the local atomic structure in the context of the local XAFS investigations and reported preliminary results
strain relaxation is essential in order to clarify the on the bond length of the nearest-neighbor Ge–Ge and
origins of such properties and to utilize the device Ge–Si atom pair in bulk GeSi crystals grown by the
potential of GeSi. Up to now, much work on the Czochralski method across the whole composition
extended X-ray absorption fine structure (XAFS) range [7].
This paper reports the results, including those on
*
Corresponding author. Tel.: þ81-22-215-2042;
Si–Si atom pairs recently obtained, in comparison
fax: þ81-22-215-2041. with those derived using ab initio electronic structure
E-mail address: yonenaga@imr.edu (I. Yonenaga). calculations [8].

0169-4332/$ – see front matter # 2003 Elsevier B.V. All rights reserved.
doi:10.1016/j.apsusc.2003.08.035
194 I. Yonenaga et al. / Applied Surface Science 224 (2004) 193–196

2. Experimental 5

Coordination number
High quality bulk crystals of Ge1xSix alloys across 4 N-Ge(Ge)
the whole composition range 0 < x < 1 were grown
3
by the Czochralski technique at very low pulling rates N-Si(Ge)
ranging from 1 to 8 mm/h in a flowing Ar gas atmo- 2
sphere [9–13]. The composition of the alloy was
determined by means of energy dispersive X-ray 1
(EDX) analysis.
X-ray absorption measurements were conducted by 0
0 0.5 1
utilizing an XAFS station BL10B in the photon fac- Si content
tory at the High Energy Acceleration Research Orga-
nization (KEK, Tsukuba) with a double Si(1 1 1) Fig. 2. Ge and Si coordination numbers around Ge atoms derived
monochromator at 20 K. Ge K-edge XAFS of the from XAFS data as a function of Si content. The symbol (^) is for
pure Si crystal. The dashed lines show the coordination numbers
GeSi alloys were measured in transmission mode at predicted from a random mixture of Si and Ge atoms.
20 K together with a comparison of that for pure Ge
single crystal. The details are described elsewhere [7].
shortened. The first-shell Ge and Si coordination
numbers around Ge atoms derived from the XAFS
3. Results and discussion data are shown as a function of alloy composition in
Fig. 2. The dashed lines show the coordination num-
Fig. 1 shows the Fourier transform of the Ge K-edge bers predicted for a random mixture of Si and Ge
XAFS data, k3w(k), at 20 K as a function of Si content atoms. The XAFS results are in good agreement with
x for the Ge1xSix alloys. The intensity of the main the random-site occupancy of Si and Ge atoms across
peak around r ¼ 2 Å, which is due to nearest-neighbor the whole composition range. Diffraction studies of
atoms around Ge, decreases and becomes asymmetric GeSi thin films grown by various epitaxial methods
with decreasing Ge content. The peak position is also indicate the existence of ordering [14]. Aubry et al. [5]
pointed out the tendency of clustering of Ge–Ge
dimers within Si-rich GeSi thin films. Conversely,
there is no report detecting an ordered structure
within bulk GeSi crystals [15]. In the present study,
the ordering parameter ð0:25NGe  ð1  xÞÞ=x (Ref.
[16]) estimated from the Ge coordination number NGe
around the Ge atoms in bulk Ge1xSix crystals
(0:44 < x < 0:82), is around 0.06–0.11, but not close
to 1. This means that no preferential ordering of the
Ge–Ge dimer seems to occur in bulk GeSi crystals
across the whole composition range. These XAFS
results are in good agreement with the prediction by
ab initio calculation [8].
The Ge–Ge bond length and the Ge–Si bond length
in GeSi alloys are plotted against the alloy composi-
tion in Fig. 3. Although the uncertainty on the derived
bond length is less than 0.015 Å, it seems that the
Ge–Ge and Ge–Si bond lengths in the GeSi alloys are
distinctly different and vary in a linear fashion as a
Fig. 1. Fourier transform |F(r)| for the Ge K-edge XAFS data, function of Si content. At the composition x ¼ 0:5 the
k3w(k), of GeSi for various Si content at 20 K. Ge–Si bond length is 2.40–2.41 Å, corresponding to
I. Yonenaga et al. / Applied Surface Science 224 (2004) 193–196 195
Distance of 1st nearest atom (nm)

0.25 bonded to four neighbor atoms pushes the nearest-


neighbor atoms away. A roughly estimated topological
Ge => Ge
rigidity parameter a according to the model by Cai
and Thorpe [22] is 0.60 in agreement with the values
of 0.63–0.70 reported experimentally [3,5] and found
0.24
theoretically [23,24] for the Ge–Ge bond in GeSi
Ge => Si
alloys. Since the parameter a varies from 0 in the
Vegard limit to 1 in the Pauling limit, our results mean
Si => Si
that both the bond lengths and bond angles are dis-
0.23 torted by changes in the composition of GeSi alloys,
0 0.5 1 similar to the reported results for most ternary alloy
Si content semiconductors. Typically for GaInAs a  0:78 is
Fig. 3. Ge–Ge bond length (Ge ! Ge), Ge–Si bond length (Ge ! estimated from the change of the InAs and GaAs bond
Si), and Si–Si bond length (Si ! Si) in GeSi alloys as a function of lengths across the alloy composition range [25]. Thus,
Si content. The dashed lines are the predicted compositional the strain in GeSi alloys is accommodated mostly by
dependence of the bond lengths from ab initio electronic structure changes of both the bond angle and bond length.
calculation.

the sum of the Si and Ge covalent radii. The Si–Si 4. Conclusion


bond length increases with decreasing Si content,
though the TEM–electron energy loss spectroscopy We have investigated the local atomic structure of
data is limited to Si-rich GeSi [17]. Possibly the Si–Si GeSi alloys by XAFS measurements in comparison
bonds have the same dependence on composition as with ab initio electronic structure calculation. From Ge
the Ge–Ge and Ge–Si bonds. K-edge spectra measured at 20 K Ge–Ge and Ge–Si
The observed feature is similar to other semicon- bond lengths of Ge1xSix grown by the Czochralski
ductor alloys as GaInAs [18], ZnCdTe [19], etc. Our method in the whole composition range 0 < x < 1 have
results confirm that the bonding in crystalline GeSi been determined. The bond lengths are found to main-
alloys grown by the Czochralski method is close to, tain distinctly different lengths and vary linearly with
but not completely corresponding to, the so-called alloy composition and are close to, but not completely
Pauling limit [20] rather than the Vegard limit [21]. at, the Pauling limit in good accordance with the
The dependence of the bond length on the bonding theoretical expectation. The estimated topological
types indicates that bond bending occurs. These rigidity parameter 0.6 suggests that both the bond
results agree rather well with those theoretically esti- lengths and bond angles distort with alloy composition.
mated from ab initio electronic structure calculations.
The dotted lines in Fig. 3 are the slope of the calcu-
lated results of [8]. Acknowledgements
However, the bonds of Ge–Ge, Ge–Si, and possibly
Si–Si show a dependence of the bonding length on the This work was partially supported by JSPS Grants-
alloy composition, i.e. they do not follow a strict in Aid (no. 13450001).
Pauling model, which is consistent with other XAFS
studies on GeSi alloys [4,5]. The fact that their
dependence has the same slope can be explained with References
regard to the topological rigidity model, similar to the
results by Woicik et al. [4] but contrary to those by [1] M. Matsuura, J.M. Tonnerre, G.S. Cargill III, Phys. Rev. B 44
(1991) 3842.
Aubry et al. [5].
[2] H. Kajiyama, S. Muramatsu, T. Shimada, Y. Nishino, Phys.
From the slope of the bond length dependences on Rev. B 45 (1992) 14005.
the composition shown in Fig. 3, we can estimate the [3] D.B. Aldrich, R.J. Nemanich, D.E. Sayers, Phys. Rev. B 50
topological rigidity parameter a how a central atom (1994) 15026.
196 I. Yonenaga et al. / Applied Surface Science 224 (2004) 193–196

[4] J.C. Woicik, K.E. Miyano, C.A. King, R.W. Johnson, J.G. [15] D. Stekamp, W. Jäger, Philos. Mag. A 65 (1992) 1369.
Pellegrino, T.-L. Lee, Z.H. Lu, Phys. Rev. B 57 (1998) 14592. [16] G.S. Cargill, F. Spaepen, J. Non-Cryst. Solids 43 (1981) 91.
[5] J.C. Aubry, T. Tyliszczak, A.P. Hitchcock, J.-M. Baribeau, [17] S. Muto, I. Yonenaga, unpublished work.
T.E. Jackman, Phys. Rev. B 59 (1999) 12872. [18] J.C. Mikkelsen Jr., J.B. Boyce, Phys. Rev. Lett. 49 (1982)
[6] S. Wei, H. Oyanagi, K. Sakamoto, Y. Takeda, T.P. Pearsall, 1412.
Phys. Rev. B 62 (2000) 1883. [19] A. Balzarotti, in: S.K. Deb, A. Zunger (Eds.), Ternary and
[7] I. Yonenaga, M. Sakurai, Phys. Rev. B 64 (2001) 113206. Multinary Compounds, Materials Research Society, Pittsburg,
[8] M.H.F. Sluiter, Y. Kawazoe, Mater. Trans. 42 (2001) 2201. 1987, p. 333.
[9] I. Yonenaga, A. Matsui, S. Tozawa, K. Sumino, T. Fukuda, J. [20] L. Pauling, The Nature of the Chemical Bond, Cornell
Cryst. Growth 154 (1995) 275. University Press, Ithaca, New York, 1967.
[10] I. Yonenaga, M. Nonaka, J. Cryst. Growth 191 (1998) 393. [21] L. Vegard, Z. Phys. 5 (1921) 17.
[11] I. Yonenaga, J. Cryst. Growth 198/199 (1999) 404. [22] Y. Cai, M.F. Thorpe, Phys. Rev. B 46 (1992) 15872.
[12] I. Yonenaga, J. Mater. Sci.: Mater. Electr. 10 (1999) 329. [23] N. Mousseau, M.F. Thorpe, Phys. Rev. B 48 (1993) 5172.
[13] I. Yonenaga, J. Cryst. Growth 226 (2001) 47. [24] P. Venezuela, G.M. Dalpian, A.J.R. da Silva, A. Fazzio, Phys.
[14] J.Z. Tischler, J.D. Budai, D.E. Jesson, G. Eres, P. Zschack, J.M. Rev. B 64 (2001) 193202.
Baribeau, D.C. Houghton, Phys. Rev. B 68 (1995) 10947. [25] J.L. Martins, A. Zunger, Phys. Rev. B 30 (1984) 6217.

You might also like