scispace - formally typeset
Open AccessJournal ArticleDOI

Band structure calculations of Si Ge Sn alloys: achieving direct band gap materials

TLDR
In this article, the electronic structure of relaxed or strained Ge1−xSnx, and of strained Ge grown on relaxed Ge 1−x−ySixSny, was calculated by the self-consistent pseudo-potential plane wave method, within the mixed-atom supercell model of alloys, which was found to offer a much better accuracy than the virtual crystal approximation.
Abstract
Alloys of silicon (Si), germanium (Ge) and tin (Sn) are continuously attracting research attention as possible direct band gap semiconductors with prospective applications in optoelectronics. The direct gap property may be brought about by the alloy composition alone or combined with the influence of strain, when an alloy layer is grown on a virtual substrate of different compositions. In search for direct gap materials, the electronic structure of relaxed or strained Ge1−xSnx and Si1−xSnx alloys, and of strained Ge grown on relaxed Ge1−x−ySixSny, was calculated by the self-consistent pseudo-potential plane wave method, within the mixed-atom supercell model of alloys, which was found to offer a much better accuracy than the virtual crystal approximation. Expressions are given for the direct and indirect band gaps in relaxed Ge1−xSnx, strained Ge grown on relaxed SixGe1−x−ySny and strained Ge1−xSnx grown on a relaxed Ge1−ySny substrate, and these constitute the criteria for achieving a (finite) direct band gap semiconductor. Roughly speaking, good-size (up to ~0.5 eV) direct gap materials are achievable by subjecting Ge or Ge1−xSnx alloy layers to an intermediately large tensile strain, but not excessive because this would result in a small or zero direct gap (detailed criteria are given in the text). Unstrained Ge1−xSnx bulk becomes a direct gap material for Sn content of >17%, but offers only smaller values of the direct gap, typically ≤0.2 eV. On the other hand, relaxed SnxSi1−x alloys do not show a finite direct band gap.

read more

Content maybe subject to copyright    Report

promoting access to White Rose research papers
Universities of Leeds, Sheffield and York
http://eprints.whiterose.ac.uk/
This is an author produced version of a paper published in Semiconductor
Science and Technology.
White Rose Research Online URL for this paper:
http://eprints.whiterose.ac.uk/4976/
Published paper
Moontragoon, P., Ikonic, Z. and Harrison, P. (2007) Band structure calculations
of Si–Ge–Sn alloys: achieving direct band gap materials.
Semiconductor Science
and Technology, 22 (7). pp. 742-748.
http://dx.doi.org/10.1088/0268-1242/22/7/012
White Rose Research Online
eprints@whiterose.ac.uk

Band Structure Calculations of Si-Ge-Sn Alloys:
Achieving direct band gap materials
Pairot Moontragoon, Zoran Ikoni´c, and Paul Harrison
Institute of Microwave and Photonics, School of Electronic and Electrical
Engineering, University of Leeds, Leeds, LS2 9JT, United Kingdom
E-mail: eenpm@leeds.ac.uk
Abstract. Alloys of silicon (Si), germanium (Ge) and tin (Sn) are continuously
attracting research attention as possible direct band gap semiconductors with
prospective applications in optoelectronics. The direct gap property may be brought
about by the alloy composition alone or combined with the influence of strain, when an
alloy layer is grown on a virtual substrate of different composition. In search for direct
gap materials, the electronic structure of relaxed or strained Ge
1x
Sn
x
and Si
1x
Sn
x
alloys, and of strained Ge grown on relaxed Ge
1xy
Si
x
Sn
y
, was calculated by the
self-consistent pseudo-potential plane wave method, within the mixed-atom supercell
model of alloys, which was found to offer a much better accuracy than the virtual
crystal approximation. Expressions are given for the direct and indirect band gaps
in relaxed Ge
1x
Sn
x
, strained Ge grown on relaxed Si
x
Ge
1xy
Sn
y
, and for strained
Ge
1x
Sn
x
grown on a relaxed Ge
1y
Sn
y
substrate, and these constitute the criteria for
achieving a (finite) direct band gap semiconductor. Roughly speaking, good-size (up
to 0.5 eV) direct gap materials are achievable by subjecting Ge or Ge
1x
Sn
x
alloy
layers to an intermediately large tensile strain, but not excessive because this would
result in a small or zero direct gap (detailed criteria are given in the text). Unstrained
Ge
1x
Sn
x
bulk becomes a direct gap material for Sn content of > 17%, but offers
only smaller values of the direct gap, typically 0.2 eV. On the other hand, relaxed
Sn
x
Si
1x
alloys do not show a finite direct band gap.
PACS numbers: 71.20.Mq, 71.15.-m, 71.22.+i

Band Structure Calculations of Si-Ge-Sn Alloys: Achieving direct band gap materials 2
1. Introduction
Recent years have witnessed a widespread use of optoelectronic devices. This technology
relies on direct band gap III-V materials like GaAs, which is expensive and highly toxic.
An interesting alternative would be a direct gap alloy based on group IV materials Si,
Ge, and Sn, which are generally compatible with silicon technology, and this has been
widely investigated. However, early studies of epitaxial SiGeSn alloys have revealed
the difficulties of their growth, with the exception of Si
1x
Ge
x
binary alloys. With
large lattice mismatch between α-Sn (6.489
˚
A) and Ge (5.646
˚
A) or Si (5.431
˚
A),
approximately 15% and 17% respectively, and the instability of cubic α-Sn above 13
o
C, bulk alloys with Sn cannot be readily grown [1]. Furthermore, because of a lower
surface free energy of alpha-tin and germanium, there can be segregation on the surface
[2]. These difficulties have been overcome by low temperature molecular beam epitaxy
(MBE), which has enabled epitaxial growth of e.g. strained Ge
1x
Sn
x
superlattices [3]
and random Ge
1x
Sn
x
alloys [4] on a Ge substrate. However, these are not expected to
show an indirect-direct transition [1], because of their compressive strain: it is presently
believed that a direct gap can only app ear in tensilely strained or relaxed Ge
1x
Sn
x
,
e.g. [5]. Further advances were made by ultra-high-vacuum chemical vapor deposition
(UHV-CVD) and uniform homogeneous relaxed Ge
1x
Sn
x
alloys with x < 0.2 have
been grown on silicon[6, 7]. Experimental investigations revealed significant changes
in optical constants and redshifts in the interband transition energy as x varied [8],
indicating wide tunability of the band gap of these alloys.
The binary GeSn and ternary SiGeSn alloys are considered to b e very prosp ective
materials for infrared detectors, as pointed at by Soref and Perry [9], who used
linear interpolation scheme to calculate the electronic band structure and optical
properties of Ge
1xy
Si
x
Sn
y
alloys and concluded that these will be tunable direct
band gap semiconductors. Furthermore, both the direct and indirect band gap in
Ge decrease with tensile strain, but the former (initially 140 meV above) does so
faster, eventually delivering a direct gap material. Therefore, one can use strained
Ge, grown on ternary Ge
1xy
Si
x
Sn
y
alloys [10, 11]. There have since been a number of
theoretical investigations of the electronic structure of SiGeSn alloys and the influence
of composition fluctuations. For instance, using the tight-binding method within the
virtual crystal approximation (VCA), the bowing parameter b
GeSn
value of 0.30 eV [12]
for Ge
1x
Sn
x
was calculated, while another, pseudopotential based calculation [13] gave
the value of -0.40 eV. The latter also predicted that Ge
1x
Sn
x
alloys become direct
gap materials, with 0.55 > E
g
> 0 eV for 0.2 < x < 0.6. The results for b
GeSn
are quite remote from each other (even in sign), and both grossly deviate from the
experimental value, b
GeSn
= 2.8 eV [1, 5]. This clearly indicates that VCA cannot explain
the behaviour of disordered Ge
1x
Sn
x
alloys [14], although it is considered reasonably
accurate for Si
1x
Ge
x
. In order to take into account the alloy disorder effects, the
Coherent Potential Approximation (CPA) was employed for Si
1x
Ge
x
alloys. According
to Chibane et al. [15], who used the model developed by Zunger et al. [16], the

Band Structure Calculations of Si-Ge-Sn Alloys: Achieving direct band gap materials 3
calculated optical band gap gap bowing is in good agreement with experiment for small
Sn contents. However, so far there is no theoretical model which properly describes
the optical properties of GeSiSn alloys in a wide range of compositions. The aim of
this work is to theoretically explore various possibilities of achieving tunable direct
gap semiconductors based on group IV materials, and to investigate the composition
dependence of their electronic and optical properties.
2. Computational method
For band structure calculation of SiGeSn alloys we use the charge self-consistent
pseudopotential Xα method. It finds the self-consistent solution of the Schr¨odinger
equation, with the lattice constituents described by ionic pseudopotential formfunctions.
Compared to the first-principles density functional theory in the local density
approximation [17, 18], which perform the total energy minimization, the Xα method
is able to reproduce the electronic structure (i.e. the band gaps, or optical properties
of semiconductors) with very good accuracy, without any additional schemes like GW
approximation or “scissors correction” as are employed in total energy approaches. On
the other hand, this method would not deliver the ground state properties (e.g. the
atomic coordinates relaxation, or lattice constant bowing) very accurately, though these
can be externally supplied to the calculation. Since our interest in this work are the
electronic properties, the Xα method was adopted, and experimentally obtained lattice
parameters were used where avilable. The calculation starts with the construction of
the effective potential, including the pseudopotential, the Hartree potential and the
exchange-correlation potential
V
eff
(r) = V
ps
+ V
Hartree
(r) + V
xc
(r), (1)
where V
Hartree
(r) =
R
e
2
n(´r)
|r´r|
d´r, and n (r) is the real-space electron density. The Hartree
potential is evaluated in momentum space
V
Hartree
(q) = 4πe
2
n (q)
q
2
, (2)
where q is the wave vector and n (q) are the Fourier coefficients of charge density. This
is evaluated in two steps. First, the density was computed on a 16 × 16 × 16 grid
of the simple unit cell in real space. Second, the fast Fourier transform was used to
transform from n (r) to n (q). The exchange-correlation potential was evaluated in a
similar manner. The local exchange-correlation potential of the Slater type has been
chosen, defined as
V
xc
(q) = α
3
2
e
2
µ
3
π
1
3
[n (q)]
1
3
, (3)
where [n (q)]
1
3
are the Fourier coefficients of the cub e root of charge density. Based on
the approximation of Slater [19], α is a constant which Schl¨uter et al. [20] have set to
0.79. The calculation of this potential was similar to the calculation of n (q), except

Band Structure Calculations of Si-Ge-Sn Alloys: Achieving direct band gap materials 4
that the cube root of n (r) on the cubic grid was taken before transforming from [n (r)]
1
3
to [n (q)]
1
3
by fast Fourier transform. The V
xc
(q) was then evaluated using Eq. (3). As
for the pseudopotential, we have used the same form as Srivastava [21],
V
ps
(q) =
µ
b
1
q
2
(cos(b
2
q) + b
3
) exp(b
4
q
4
), (4)
with the parameters for Si, Ge, and Sn given in Table 1. These are slightly readjusted
values from those given in Ref. [21], as dictated by a larger energy cutoff, and a different
method of integration over the Brillouin zone, used in this work.
Parameter Si Ge Sn
b
1
(Ry) -1.213 -1.032 0.401
b
2
0.785 0.758 1.101
b
3
-0.335 -0.345 0.041
b
4
0.020 0.024 0.018
Table 1. Parameters of the pseudopotential of Si, Ge and α-Sn
The electronic structure is found by solving the Kohn-Sham equation (in atomic
units ~ = 2m
e
=
e
2
2
= 1):
£
−∇
2
+ V
eff
(r)
¤
ψ
n
(k; r) = ²
n
(k; r) ψ
n
(k; r) , (5)
which is done using the self-consistent pseudopotential plane waves method [22], with a
kinetic energy cutoff of 24 Ry, the value which gives good convergence of the calculation.
The improved linear tetrahedral method [23] was used for integration over the Brillouin
zone, with 34 k-points in the irreducible wedge. The convergence of the self-consistent
calculation was considered to be adequate when the total energy of the system was
stable to within 10
3
Ry. In these calculations we do not account for the spin-orbit
coupling, b ecause it would double the size of the problem while not being essential for
the aim of this work, which is to find whether the smallest band gap is direct or indirect
(and this is determined only by the behaviour of the conduction band). Including the
spin-orbit coupling would bring slight quantitative corrections in the calculated values
of band gaps (with the ionic pseudopotential formfunctions re-adjusted to reproduce the
known experimental values for elemental Si, Ge and Sn in this case), but this would not
affect the predicted direct-indirect crossover points. Similar conclusion on a relatively
small influence of spin-orbit coupling on the topic of interest here has been drawn in
[24], based on empirical pseudopotential calculations in the Ge-Sn alloy.
The calculated band structure of bulk Si, Ge, and α-Sn are given in Fig. 1, with the
band gap of Si (Ge) being 1.2 eV (0.71 eV), while α-Sn has zero direct band gap. The
longitudinal (m
l
) and transverse effective mass (m
t
) for the X-valley in Si are 0.87m
0
and 0.22m
0
. Similarly, the L-valley in Ge has m
l
=1.68m
0
and m
t
=0.16m
0
. All these
are in very good agreement with the published values, indicating that the parameters
can be reliably used for further calculations.

Citations
More filters
Journal ArticleDOI

Melting and homogeneity in germanium-silicon alloys and a modified micro-manufactured assembly for stable high pressure and temperature measurements

TL;DR: In this paper, a T-shaped power coupler with a ring-shaped insulator and a precision made, all-in-one single-piece crucible with mirror-image located, flat-bottomed sample and pressure marker chambers, with press-fit lids and thermocouple insert in the middle of the crucible.
Journal ArticleDOI

Evaluation of energy band offset of Si1−xSnxsemiconductors by numerical calculation using density functional theory

TL;DR: In this paper, the authors examined the numerical calculation for evaluating the energy band offset of Si1− x Sn x semiconductors and compared their calculation results with the results of previous theoretical calculation and experimental estimation.
Journal ArticleDOI

The Low Temperature Epitaxy of Strained GeSn Layers Using RTCVD System

TL;DR: In this paper, the authors investigated the low temperature (LT) growth of GeSn-GeSi structures using rapid thermal chemical vapor deposition system utilizing Ge2H6 and SnCl4 as the reactive precursors.
DissertationDOI

Modeling Key Issues in Post Silicon Semiconductors: Germanium and Gallium Nitride

Ziyang Xiao
Abstract: Title of Proposal: Modeling Key Issues in Post Silicon Semiconductors: Germanium and Gallium Nitride Ziyang Xiao, Doctor of Philosophy, 2018 Proposal directed by: Professor Neil Goldsman Department of Electrical and Computer Engineering We are rapidly approaching the end of the semiconductor roadmap with respect to silicon. To continue its growth, the semiconductor industry is therefore looking into new materials. Two primary materials that are of interest for continuing semiconductor development are germanium (Ge) and gallium nitride (GaN). Ge is of interest as a replacement for silicon, in an effort to improve electronics performances because of its high mobility and its ability to grow a native oxide. In addition, Ge is of interest because of its potential use for economical CMOS-based short wave infrared (SWIR) imaging systems. GaN is a nascent wide bandgap semiconductor and has many potential applications in high power electronics and ultraviolet imaging systems. In this thesis, the key material properties and applications of these two ”end of the roadmap” semiconductors are explored. Ge is a semiconductor material with an indirect bandgap of 0.66eV[1]. This bandgap value corresponds to a wavelength of 1.88μm, which lies in the infrared range. The Ge material itself is also compatible with the standard Si CMOS process technology. Because of these advantages, Ge is considered a candidate for the application of photo detecting in the SWIR range. Apart from the indirect bandgap of 0.66eV, Ge also has a direct bandgap of 0.8eV[1]. From early research, the relatively small offset between the indirect and direct bandgaps can be inverted either by applying strain[1, 9, 10, 11] or alloying with tin[2, 14]. GaN is a binary direct wide bandgap material with a direct bandgap of 3.4eV[26, 27]. It has a high breakdown field, and relatively high saturation velocity and carrier mobility[28]. These properties give GaN an advantage in the realm of high power application. GaN can also form a heterostructure with AlGaN, which can give rise to a 2D electron gas (2DEG) layer at the interface without intentionally doping either material. The 2DEG layer has an even higher mobility when compared to the mobility of the bulk GaN, which allows the heterostructure to be utilized for the design of high electron mobility transistors (HEMTs). The formation of the 2DEG layer also gives rise to potential well confinement at the heterostructure interface. The width of the potential well is only a few nanometers, making the interface electron gas subject to quantum confinement along the direction perpendicular to the interface. The detailed shape of the potential well is determined by the configuration of the heterostructure, as well as the applied voltage across the heterostructure. The first set of goals for this research is to investigate how the bandstructure of Ge changes: Part (1) with the applied strain, and Part (2) with alloyed tin (Sn). The empirical pseudopotential method (EPM) was utilized for the band structure calculation, together with the rules for strain translation for the investigation of Part (1). In Part (1), simulation results give the optimal orientation for different types of applied strain and also thoroughly map the influence of strain applied on any arbitrary orientations. It also reveals that for biaxial strain, there exists another orientation that is more robust against misalignment with respect to the originally desired orientation than the optimal plane, with little compromise of bandgap and slightly higher requirements for the sufficient strain. For Part (2), EPM is combined with perturbation theory for the inclusion of the influence of the Sn atoms in the Ge lattice. A new and computationally inexpensive method is developed during the research. Simulation results agree significantly when compared to reported experimental measurements, indicating the capability of the method. The second set of goals is to investigate the electron transport properties of the 2DEG layer at the interface of GaN HEMT and related power transistors. The potential well is approximated and quantified by a triangular potential well and the carrier sheet density is kept the same during the approximation. Thorough simulations are conducted by calculating the band alignment of the heterostructure with different structural configurations. A fixed correlation between the carrier sheet density and the shape of the potential well (slope of the triangular potential well and the height of the well) is revealed. This correlation is used as an input for the Monte Carlo (MC) simulation. The changes to the mobility of the electrons at the 2DEG layer with changing interface potential well shape are investigated and statistics of drift velocity, electron energy, and valley occupation are collected. Mobility information is also extracted and compares favorably with reported experimental measurements. The simulation results are used in the device simulations, which compares the performances of two GaN/AlGaN heterostructure based devices: a lateral HEMT and a current aperture vertical electron transistor (CAVET).
References
More filters
Journal ArticleDOI

Self-Consistent Equations Including Exchange and Correlation Effects

TL;DR: In this paper, the Hartree and Hartree-Fock equations are applied to a uniform electron gas, where the exchange and correlation portions of the chemical potential of the gas are used as additional effective potentials.
Journal ArticleDOI

Improved tetrahedron method for Brillouin-zone integrations

TL;DR: In this article, the tetrahedron method was used for Brillouin-zone integrations and a translational grid of k points and tetrahedral elements was proposed to obtain results for insulators identical to those obtained with special-point methods with the same number of points.
Book

Electronic Structure: Basic Theory and Practical Methods

TL;DR: In this paper, the Kohn-Sham ansatz is used to solve the problem of determining the electronic structure of atoms, and the three basic methods for determining electronic structure are presented.
Journal ArticleDOI

Band lineups and deformation potentials in the model-solid theory.

TL;DR: In this paper, a theoretical model is presented to predict the band offsets at both lattice-matched and pseudomorphic strained-layer interfaces, based on the local density functional pseudopotential formalism and the ''model solid approach'' of Van de Walle and Martin.
Related Papers (5)
Frequently Asked Questions (10)
Q1. What are the contributions in "Band structure calculations of si-ge-sn alloys: achieving direct band gap materials" ?

In this paper, the authors used the pseudo-potential plane wave ( PWP ) method to calculate the direct and indirect band gap of the Si-Ge-Snx alloys. 

The aim of this work is to theoretically explore various possibilities of achieving tunable direct gap semiconductors based on group IV materials, and to investigate the composition dependence of their electronic and optical properties. 

With large lattice mismatch between α-Sn (6.489 Å) and Ge (5.646 Å) or Si (5.431 Å), approximately 15% and 17% respectively, and the instability of cubic α-Sn above 13o C, bulk alloys with Sn cannot be readily grown [1]. 

Band structure calculations show that relaxed Ge1−xSnx alloys have an indirect-to-direct band gap crossover at a Sn content of ≈0.17, with the bowing parameter equal to 2.49 eV. 

Alloy properties can be evaluated either within the virtual crystal approximation (VCA) with identical, average-composition atoms populating the lattice sites of the minimumvolume crystalline unit cell, or by populating individual lattice sites only with pure element atoms, in proportion to the alloy composition (“mixed atom method”), in which case one has to use a supercell, with increased volume. 

In the electronic structure calculations the lattice constant of Ge1−x−ySixSny alloys was taken to depend on the Si content (x) and Ge content (y) as [35]aGeSiSn(x, y) = aGe +4SiGex + θSiGe(1− x) +4SnGey + θSnGey(1− y), (12)where 4SiGe = aSi − aGe, 4SnGe = aSn − aGe, and θSiGe = −0.026 Å. 

This system is currently believed to be of great practical interest, since it offers a direct band gap in Ge at a reasonable level of strain (>1.8%) as well as type-I heterostructure [11] (of importance for realisation of quantum well structures), together with a small thermal expansion mismatch between the two materials [6]. 

According to Chibane et al. [15], who used the model developed by Zunger et al. [16], thecalculated optical band gap gap bowing is in good agreement with experiment for small Sn contents. 

The lattice constant of an alloy can be estimated from Vegard’s law [25]a0(x) AB = (1− x)aA0 + xaB0 , (6)where aA0 and a B 0 are the lattice constants of elemental crystals of atoms A and B respectively, and a more accurate expression (with bowing) was taken where available. 

The computational method used in this work, and the conditions (strain along the [001] axis), allow meaningful extraction of:: the uniaxial deformation potential b for the valence band at Γ, the sum of valence and conduction band (at Γ) hydrostatic deformation potentials av +ac, and the uniaxial deformation potential Ξ∆u of the X (i.e. ∆) valley of the conduction band.