You are on page 1of 27

Research Signpost

37/661 (2), Fort P.O., Trivandrum-695 023, Kerala, India

Advances in Condensed Matter Physics, 2009: ISBN: 978-81-308-0336-4


Editor: Ali Hussain Reshak

8 Theoretical study of InAs, InSb


and their alloys InAsxSb1-x
K. Hachelafi1, B. Amrani2, F. El Haj Hassan3 and S. Hiadsi4
1
Département de Physique, Faculté des Sciences (U.S.T.O.), Oran 31000
Algeria; 2Département de Physique, Centre Universitaire de Mascara
Mascara 29000, Algeria; 3Laboratoire de Physique des Matériaux
Département de Physique, Faculté des Sciences, Elhadath, Beirut, Lebanon

Abstract
We have performed first-principles calculations
using full potential linearized augmented plane wave
(FP-LAPW) method within density functional theory
(DFT) to investigate the fundamental properties of
the InAs, InSb and their alloys. The ground state
properties such as lattice constant, bulk modulus,
pressure derivative of the bulk modulus and elastic
constants are in good agreement with numerous
experimental and theoretical data. Through the
quasi-harmonic Debye model, in which the phononic
effects are considered, we have obtained successfully
Correspondence/Reprint request: Dr. B. Amrani, Département de Physique, Centre Universitaire de Mascara
Mascara 29000, Algeria. E-mail: abouhalouane@yahoo.fr
2 K. Hachelafi et al.

the thermodynamic properties such as thermal expansion coefficient and


specific heats in the whole pressure range from 0 to 30 GPa and temperature
range from 0 to 1200 K.

I. Introduction
The group In-V semiconductors InAs, InSb and their ternary alloys have
been extensively used in infrared optoelectronic devices, including laser,
photodetectors and optical gas sensors [1-5]. The studies of the basic
properties are very important for the understanding of the device
characteristics and improvement of their performance. All these have over the
years helped in motivating the study of their optoelectronic properties both
experimentally [1-23] and theoretically [24 -31].
Usually, the thermodynamic properties of materials are the basis of solid-
state science and industrial applications. Furthermore, the study of the
thermodynamic properties of materials is of importance to extend our
knowledge on their specific behaviour when undergoing severe constraints of
high-pressure and high temperature environments. This is particularly true
since the coming of modern technologies (geophysics, astrophysics, particle
accelerators, fission and fusion reactors, etc.), from which we always expect
new advances and innovations in materials science to reach higher
performances. From a fundamental point of view, there is a need for
refinement of theoretical models for the computation of a more accurate
equation of state of the material [32].
It is worth noting that the theoretical calculations mentioned above only
gave material properties at zero temperature, without any thermal effects
included. Therefore, it is necessary to examine the thermal influences on the
properties of these compounds. To address this interest, in this work, we
investigate the structural and thermodynamic properties at high pressures and
temperatures, by using first-principle calculations combined with the quasi-
harmonic Debye model.
The rest of this paper is organized as follows. In section 2, we present the
theoretical background. In section 3, we present our calculated structural,
electronic and thermodynamic properties. Finally, we gave a brief conclution
in Section 4.

II. Theoretical method


The zero-temperature energy calculations are performed using the
Wien2k package [33], which is based on the density functional theory.
Exchange and correlation functional is given by the generalized-gradient
approximation of Perdew et al. (GGA96) form [34]. Atomic orbitals up to an
Short Title 3

angular momentum equal to 10 were used to expand the wave functions


inside the muffin-tin spheres.
While in the interstitial regions they are expanded in terms of PWs. The
wave functions and derivatives are made continuous at the boundary of the
spheres, and there are no shape approximations imposed on either the
crystalline charge density or potential. The PWs cut off was used with the
highly recommended condition RMTKmax =9, where RMT is the average of the
muffin-tin spheres and Kmax is the PW cut-off. The sphere radii used in the
calculations are 2.3, 2.3 and 2.2 a.u. for In, AS and Sb, respectively. A mesh of
35 special k-points for binary compounds and 27 special k-points for the alloy
were taken in the irreducible wedge of the Brillouin zone. Both the plane wave
cut-off and the number of k-points were varied to ensure total energy
convergence. To ensure proper convergence of the self consistency calculation,
the calculated total energy of the crystal converged to less than 0.5mRy.
To investigate the thermodynamic properties of InAs and InSb, we apply
the quasi-harmonic Debye model [35], in which the non-equilibrium Gibbs
function G*(V; P, T) can be written in the form of

G * (V ; P, T ) = E (V ) + PV + AVib [θ (V ); T ] (1)

where E(V) is the total energy per unit cell, PV corresponds to the constant
hydrostatic pressure condition, θ(V) is the Debye temperature, and AVib is the
vibrational term, which can be written using the Debye model of the phonon
density of states as [36, 37].

⎡ 9θ θ ⎤
Avib (θ ; T ) = nkT ⎢ + 3 ln(1 − e −θ / T ) − D ( ) ⎥ (2)
⎣ 8T T ⎦

where n is the number of atoms per formula unit, D(θ /T ) represents the
Debye integral, and For an isotropic solid, θ is expressed as [36].

θD =
=
k
[ ] 1/ 3 B
6π 2V 1/ 2 n f (σ ) s
M
(3)

where M is the molecular mass per unit cell; BS is the adiabatic bulk modulus,
which is approximated given by the static compressibility [35]

d 2 E (V )
Bs ≅ B(V ) = V (4)
dV 2
4 K. Hachelafi et al.

f (σ ) is given by Refs.[ 38, 39]; the Poisson σ is taken as 0.25 [40].


Therefore, the non-equilibrium Gibbs function G*(V; P, T) as a function of
(V; P, T) can be minimized with respect to volume V

⎡ ∂G * (V ; P, T ) ⎤ (5)
⎢ ∂V ⎥ =0
⎣ ⎦ P ,T

By solving Eq.(5), one can obtain the thermal equation-of-equation (EOS) V


(P, T). The heat capacity CV and the thermal expansion coefficient α are
given by [41].

⎡ θ 3θ / T ⎤ (6)
CV = 3nk ⎢4 D ( ) − θ / T
⎣ T e − 1⎥⎦

⎡ θ θ /T ⎤
S = nk ⎢ 4 D ( ) − 3 ln(1 − e − ) ⎥ (7)
⎣ T ⎦

γCV
α= (8)
BT V

where γ is the Grüneisen parameter, which is defined as

d ln θ (V )
γ =− (9)
d ln V

Through the quasi-harmonic Debye model, one could calculate the


thermodynamic quantities of any temperatures and pressures of InX (X=As,
Sb) from the calculated E-V data at T = 0 and P = 0.

III. Results and discussion


III. I. Binary compounds
A. Static and mechanical properties
The total energy is obtained as a function of volume and fitted to a
Murnaghan equation of state [42] to obtain the equilibrium lattice constant,
bulk modulus, and its pressure derivative. The results are summarized in
Table 1, together with some theoretical results and the available experimental
data. The calculated lattice parameters and bulk modulus are in general, in
Short Title 5

favorable agreement with the experimental and previous calculations. Our


calculated lattice parameter for InAs is closer to the FPLAPW results
reported by Rashid Ahmed et al. [43]. The small differences between the two
FPLAPW calculations are due to the different k-points sampling. For InSb,
our lattice parameter value is larger than the values obtained by Wang and Ye
[52], Wei and Zunger [51] and Kalvoda et al. [49]. The cohesive energy for
both compounds is also given in Table 1. It is clearly seen that the bulk
modulus and the absolute value of the cohesive energy decrease from InAs to
InSb, i.e. from the lower to the higher atomic number. This suggests that InSb
is more compressible than the other compound.
We focus now on the elastic properties, to study the stability of these
compounds we have calculated the elastic constants, using the method
developed by Charpin and integrated in the WIEN2K code [33]. It is well
known that a cubic crystal has only three independent elastic constants, C11,
C12 and C44. Hence, a set of three equations is needed to determine all
constants. The first equation involves calculating the elastic constants
(C11 − C12 ) which are related to the bulk modulus B

1
B = (C11 + 2C12 ) . (10)
3

The second one involves applying volume-conserving tetragonal strains

⎛ ⎞
⎜ε 0 0 ⎟
⎜ ⎟
⎜ ⎟. (11)
ε = ⎜0 ε 0 ⎟
⎜ ⎟
⎜0 1
⎜ 0 − 1⎟⎟
⎝ (1 + ε )2 ⎠

Application of this strain changes the total energy from its initial value as
follows:

E ( ε ) = (C11 − C12 )6V0ε 2 + ο (ε 3 ) , (12)

Where V0 is the volume of the unit cell.


Finally, for the last type of deformation, we used the volume-conserving
rhombohedral strain tensor given by:
6 K. Hachelafi et al.

⎛ 1 1 1⎞
ε⎜ (13)
ε = ⎜1 1 1⎟⎟
3⎜ ⎟
⎝ 1 1 1⎠

Which transform the total energy to the full elastic tensor.

V0
E (ε ) = (C11 + 2C12 + 4C44 )ε 2 + ο (ε 3 ) (14)
3

Our GGA calculated values for the elastic constants are compared with
the available results in Table 1. The traditional mechanical stability
conditions on the elastic constants in cubic crystals are known to be C11-
C12>0, C11+2C12>0, C11>0 and C44>0. Our results for elastic constants in
Table 1 obey these stability conditions.

Table 1. Cohesive energy Ecoh, Lattice constant a, bulk modulus B, pressure


derivative of bulk modulus B’ and elastic constants parameters of InAs and InSb at
zero pressure and zero temperature, compared with the experimental data and other
theoretical works.

a b c d
Ref. [46] Ref. [47] Ref. [48] Ref. [44] e Ref. [49] f Ref. [50] g
Ref. [51]
h i j k
Ref. [52] Ref. [43] Ref. [53] Ref. [54] l Ref. [45]
Short Title 7

B. Temperature and pressure effects


The thermal properties are determined in the temperature range from
0 to 1200 K, where the quasi-harmonic model remains fully valid. The
pressure effect is studied in the 0–30 GPa range. The temperature effects
on the lattice parameters are shown in Fig. 1. The lattice parameter
increases with increasing temperature but the rate of increase is very
moderate. On the other hand, it is noted from Fig. 2 that the relationships
between bulk modulus and pressure are all nearly linear at various
temperatures of 300, 600, 900, and 1200 K, respectively. The bulk
modulus increases with pressure at a given temperature and decreases
with temperature at given pressure. These results are due to the fact that
the effect of increasing pressure on the material is the same as decreasing
temperature of the material.
The investigation on the heat capacity of crystals is an old topic of
condensed matter physics with which illustrious names are associated [55,
56, 57]. Knowledge of the heat capacity of a substance not only provides
essential insight into its vibrational properties but is also mandatory for many
applications. Two famous limiting cases are correctly predicted by the
standard elastic continuum theory [57]. At high temperatures, the constant
volume heat capacity Cv tends to the Petit and Dulong limit [58]. At
sufficiently low temperatures, Cv is proportional to T3 [57]. At intermediate
temperatures, however, the temperature dependence of Cv is governed by the
details of vibrations of the atoms and for a long time could only be
determined from experiments. Fig. 3 and Fig. 4 represent the variation of the
volume expansion coefficient, α(T), and the heat capacity, Cv (T) as function
of the temperature, respectively. These two quantities indicate a sharp
increase up to ~ 400 K which is due to the anharmonic approximation of the
Debye model used here. However, at higher temperature, the anharmonic
effect on Cv is suppressed, and Cv is very close to the Dulong–Pettit limit
(Cv (T) ~ 3R for mono-atomic solids), which is common to all solids at high
temperatures.
It is obvious from Fig. 4, that, for a given pressure, α increases with
temperature at low temperatures especially at zero pressure and
gradually tends to a linear increase at higher temperatures. As pressure
increases, the increase of α with temperature becomes smaller. While
for a given temperature, α decreases strongly with increasing pressure,
and it is very small at higher temperatures and higher pressures. Finally,
in Fig. 5, we have plotted our results for the entropy S. It is shown that
the high temperature dependence of entropy S is nearly insensitive to
pressure.
8 K. Hachelafi et al.

6,85 0 Kbar
10 Kbar InSb
6,80 20 Kbar
30 Kbar
Lattice Parameter (Å )
3

6,75

6,70

6,65

6,60

6,55

6,50

0 200 400 600 800 1000 1200

Temperature (K)

6,45 InAs 0 Kbar


6,40
10 Kbar
20 Kbar
6,35
30 Kbar
Lattice Parameter (Å )
3

6,30

6,25

6,20

6,15

6,10

6,05
0 200 400 600 800 1000 1200

Temperature (K)

Figure 1. The relationships between lattice parameters and temperature at pressures of


0, 10, 20 and 30 Kbar, respectively.
Short Title 9

80
75 InAs
70
65
60
Bulk modulus (GPa)

55
50
45
40
35 0K
300 K
30
600 K
25 900 K
20 1200 K
15

0 1 2 3 4 5

Pressure (Kbar)

60
InSb
55

50
Bulk modulus (GPa)

45

40

35 0K
300 K
30 600 K
900 K
25 1200 K

0 1 2 3 4 5

Pressure (Kbar)

Figure 2. The relationships between bulk modulus and pressure at temperatures of


300, 600, 900, and 1200 K, respectively.
10 K. Hachelafi et al.

Dulong-petit limit
50
Heat capacity (J mol K )

40
-1

0 Kbar
10 Kbar
-1

20 Kbar
30 30 Kbar
40 Kbar
50 Kbar
20

10

InAs
0
0 200 400 600 800 1000 1200

Temperature (K)

60
Dulong-petit limit

50
Heat capacity (J mol K )
-1

40
0 Kbar
-1

10 Kbar
20 Kbar
30
30 Kbar
40 Kbar
20 50 Kbar

10

InSb
0
0 200 400 600 800 1000 1200

Temperature (K)

Figure 3. The heat capacity versus temperature at pressures of 0, 10, 20, and 30 GPa,
respectively.
Short Title 11

0 Kbar
10 Kbar
10 20 Kbar
30 Kbar
40 Kbar
Thermal expansion (10 K )
-1

8 50 Kbar
-5

InAs
0
0 200 400 600 800 1000 1200

Temperature (K)

10

InSb
8
Thermal expansion (10 K )
-1
-5

4
0 Kbar
10 Kbar
20 Kbar
2 30 Kbar
40 Kbar
50 Kbar
0
0 200 400 600 800 1000 1200

Temperature (K)

Figure 4. The thermal expansion versus temperature at pressures of 0, 10, 20, and 30
GPa, respectively.
12 K. Hachelafi et al.

160 InAs
140

120
Entropy (J/mol,K)

100

0 Kbar
80
10 Kbar
20 Kbar
60
30 Kbar
40 Kbar
40
50 Kbar

20

0
0 200 400 600 800 1000 1200

Temperature (K)

160 InSb
140

120
Entropy (J/mol,K)

100

0 Kbar
80
10 Kbar
20 Kbar
60
30 Kbar
40 Kbar
40
50 Kbar

20

0
0 200 400 600 800 1000 1200

Temperature (K)

Figure 5. The entropy versus temperature at pressures of 0, 10, 20, and 30 GPa,
respectively.
Short Title 13

C. Electronic properties
Electronic band structures of InAs and InSb based on self consistent
scalar relativistic FP-LAPW calculations in which the exchange and
correlation were treated in the GGA scheme along the various symmetry lines
of the face centred cubic Brillouin zone are plotted in Fig. 6. In both cases the
overall band profiles are found to be in fairly good agreement with previous
theoretical results. Our calculations show that the valence band maximum
and conduction band minimum are both at the Γ point of the Brillouin zone
for both compounds. As such, they possess direct bandgaps. The calculated
energy gap of InAs and InSb is about 0.00 eV lower than the experimental
value. Underestimation of band gaps is the normal result of first-principles
GGA calculations of the band structures of semiconductors and insulators.
This effect is well known and understood as arising from the fact that
eigenvalues of the Kohan-Sham equations are not excitation energies of the
system. Accurate band gaps can be obtained from solutions of the
quasiparticle equations in which the exchange and correlation effects are
described by the self energy, which is a nonlocal, energy-dependent effective
potential.
In Tables 2 we summarize the most important electronic structure
parameters of InAs and InSb, respectively. From Table 2, we note that our
energy values are in good agreement with the corresponding values by PL
spectra [59].
Turning to the effective masses, we are interested to study the effective
masses of electrons and holes, which are important for the excitonic
properties. We have computed the effective masses of electrons and holes
both at the conduction-band minima and valence-band maxima. The different
effective masses are obtained from the expression

−1
2 ⎛ d Ek ⎞
2
m = ± = ⎜⎜
*
2
⎟⎟ (15)
⎝ dk ⎠

Table 3 lists our theoretical results along with the available experimental
values, for convenience some theoretical results of other methods are also
listed. It is clear that our calculated values are somewhat higher than the other
reported values.
Turning now to the bonding properties, to visualise the nature of the
bond character and to explain the charge transfer and bonding properties of
InAs and InSb, we calculated the total valence charge density. We show in
Figs. 7-8 the total valence charge densities in the (1 1 1) plane for each
14 K. Hachelafi et al.

material. The case of ionicity is interesting since it can be related to


properties of the charge density on the whole and features in the band
structures.

10

5
Energy (eV)

0 EF

-5

InAs
-10

-15
W L Λ Γ ∆ X W K

10

5
Energy (eV)

0 EF

-5

InSb
-10

-15
W L Λ Γ ∆ X W K

Figure 6. Band structure of InAs and InSb. The energy zero is taken at Ef.
Short Title 15

Table 2. A summary of the important features, energy gaps and valence bandwidths
of InAs and InSb, compared with the experimental data and other theoretical works.
All energies are in eV.
16 K. Hachelafi et al.

The ionicity which is directly associated to the character of the chemical


bond provides us a means for explaining and classifying the properties of
these III-V compounds. It is well known that the ionicity character is highly
dependent on the total valence charge densities by calculating the charge
distribution. To obtain an estimated value of the ionicty factor for InAs and
InSb compounds, we have used an empirical formula [70]. In this approach,
the total area occupied by the valence charge density is divided into parts
with respect to the centre. SC and SA are the areas of the cation and anion
sides, respectively, and the ionicity factor is defined as

λ
⎛ SA ⎞
f i = ⎜⎜ ⎟⎟ (16)
⎝ S A + λS C ⎠

Where λ is a parameter separating the highly ionic elements from the


weakly ionic ones. λ=-1 for elemental and III-V semiconductors; λ=1 for II-
VI and I-VII semiconductors. Our results indicate that these compounds
exhibit small ionicity character of 0.372, and 0.316 for InAs and InSb,
respectively.
Our results are consistent with the fact that InAs is a more ionic crystal
than InSb [71].

Table 3. Conduction and valence band-edge electron and heavy-hole effective masses
(in units of free electron mass) at the Γ point of the Brillouin zone in InAs and InSb.
Short Title 17

Table 4. Calculated lattice parameter a and bulk modulus B compared to experiment


and other theoretical works of InAsxSb1-x.

a b c d e f g
Ref. [46] Ref. [47] Ref. [53] Ref. [54] Ref. [49] Ref. [50] Ref. [51]
h i
Ref. [52] Ref. [43]

1,1
As
1,5
0,45
0,10
0,13
0,45
0,33
In0,33
0,13
0,45

1,5

As 0,45
1,1
0,33
0,45
0,13 0,33
0,10
0,45 In
0,13

0,45
1,5
1,1 As

Figure 7. Total valence charge densities of InAs in the (1 1 1) plane.


18 K. Hachelafi et al.

Sb
0,92
0,70
0,10
0,100,23
0,10
0,30 In
0,10

0,70

0,23
0,92 Sb

0,30
0,30
0,10 0,100,30
0,10
In
0,10
0,23
0,23

0,70
Sb

Figure 8. Total valence charge densities of InSb in the (1 1 1) plane.

III. II. Alloys


A. Structural parameters of the alloys
The alloys were modeled at some selected compositions with ordered
structures described in terms of periodically repeated supercells. This can be
done with little effort (i.e. with eight-atoms per unit cell) for the composition
x = 0.25, 0.5 and 0.75. For the composition x = 0.25 and 0.75 the simplest
structure is an eight-atom simple cubic lattice (luzonite): the cations with the
lower concentration form a regular simple cubic lattice. For x = 0.5, the
smallest ordered structure is a four-atom tetragonal cell, corresponding to the
(001) superlattice. We have also checked the chalcopyrite structure, which
has a 16-atom tetragonal cell for x=0.5, and the results were found to be
similar to those of the (001) superlatice. For the considered structures, we
perform the structural optimization by minimizing the total energy with
respect to the cell parameters. The total energies calculated as a function of
unit cell volume were fitted to the Murnaghan’s equation of state [42]. Our
Short Title 19

results for the materials of interest are compared with the available
experimental and theoretical predictions in table 4. It is clear that our GGA
results are in reasonable agreement with experimental values for the end
alloys.
In addition to theoretical studies based on first principle, some
phenomenological models also explain various aspects of physical properties
of alloys. In a phenomenological approach [72] physical properties of an
alloy AxB1-xC can be obtained, in good approximation, by a linear relation
between the physical property F(AxB1-xC) and the composition x, known as
Vegard’s Law, by taking the weighted average of the corresponding physical
properties FAC and FBC of the end-point compounds:

F ( x ) = xFAC + (1 − x) FBC . (17)

However, many of the alloys deviate from linear behavior. A general


form of their behaviour is expressed by a semi empirical quadratic
relationship [73]

F ( x, k ) = F ( x) + bx(1 − x), (18)

where b is the bowing parameter which accounts for the deviation from a
linear behavior. Experiments show that the lattice parameter a (x) follow
relation (17) implying that the bowing parameter b is zero. For the lattice
constants, we can then write equation (17) as

a ( x) = xa AB + (1 − x)a AC . (19)

The results obtained for the composition dependence of the calculated


equilibrium lattice parameter for InAsxSb1-x alloy are shown in Fig. 9. It is
found that the deviation of the optimized lattice constants from linearity is
relatively small. Therefore, Vegard’s linear rule can be a good approximation
to obtain the lattice constants of InAsxSb1−x alloys. The small bowing
parameter due to the small mismatching between the lattice constants of InSb
(6.635 Å) and InAs (6.192 Å) compounds is calculated to be -0.073 Å, which
is very close to the theoretical value of -0.063 Å [30].
The overall behaviors of the variation of bulk modulus as a function of the
composition x is presented in Fig. 10. A significant deviation from the linear
concentration dependence with upward bowing equal to -3.72 GPa. In fact, we
find that the bulk modulus of InAs is 24 % higher than of InSb. This result
suggests the existence of a significant upward bowing of the bulk modulus.
20 K. Hachelafi et al.

6,7

InAsxSb1-x
6,6
Lattice parameter a(Å)

6,5

6,4

6,3

6,2

0,00 0,25 0,50 0,75 1,00

Composition x

Figure 9. Composition dependence of the calculated lattice constants (solid squares)


of InAsxSb1-x alloy compared with Vegard’s prediction (dashed line). A bowing
parameter equal to -0.073 Å is obtained by fitting the calculated values with a
polynomial function (solid curve).
50
InAs xSb 1-x
48

46
Bulk modulus (GPa)

44

42

40

38

36

0,00 0,25 0,50 0,75 1,00

Composition x

Figure 10. Composition dependence of the calculated bulk modulus (solid squares) of
InAsxSb1-x alloy compared with LCD prediction (dashed line). The bowing parameter
equal to -3.72 GPa was obtained by polynomial fitting (solid curve).
Short Title 21

B. Thermodynamic properties
In this section we present a rigorous theoretical study of the
thermodynamic properties of InAsxSb1−x, we calculated the phase diagram
based on the regular-solution model [74-76]. The Gibbs free energy of
mixing, ∆Gm, for alloys is expressed as

∆Gm = ∆H m − T∆S m , (20)

where

∆H m = Ωx(1 − x ), (21)

∆S m = − R[x ln x + (1 − x) ln(1 − x)]. (22)

∆Hm and ∆Sm are the enthalpy and entropy of mixing, respectively; Ω is the
interaction parameter and depends on material; R is the gas constant; and T is
the absolute temperature. Only the interaction parameter Ω depends on the
material.
The mixing enthalpy of alloys can be obtained from the calculated total
energies as, ∆H m = E InAsx Sb1− x − xE InAs − (1 − x) E InSb , were E InAsx Sb1− x , E InAs and
E InSb are the respective energies of InAsxSb1-x alloy and the binary
compounds InAs and InSb. We then calculated ∆Hm to obtain Ω as a function
of concentration. Fig. 11 shows Ω versus x in this manner for InAsxSb1-x
alloys. Ω increases almost linearly with increasing x. From a linear fit we
obtained

Ω( kcal / mol ) = 37.737 x − 8.096 (23)

The average values of the x-dependent Ω in the range 0 ≤ x ≤ 1 obtained


from this equation is 10.773 Kcal/mol.
We can now calculate accurately the free energy of mixing ∆Gm using
Eqs. (20)–(22). Fig. 12 shows the free energy of mixing for the InAsxSb1-x
alloy versus composition x at several selected temperatures. Then we use
∆Gm at different concentrations to calculate the T–x phase diagram which
shows the stable, metastable and unstable mixing regions of the alloy. At a
temperature lower than the critical temperature Tc, the two binodal points are
determined as those points at which the common tangent line touches the
∆Gm curves. The two spinodal points are determined as those points at which
the second derivative of ∆Gm is zero; ∂2(∆Gm)/∂x2 = 0.
22 K. Hachelafi et al.

InAsxSb 1-x
20
Interaction Parameter Ω (kcal/mole)

15

10

0
0,25 0,50 0,75

Composition x

Figure 11. Interaction parameter, x, as a function of the composition x calculated for


InAsxSb1-x. The dashed line is the linear fit to the x values.

0,08
0,07
0,06 T=1000 K
Mixing free energy∆Gm(eV/pair)

0,05
0,04
0,03 T=1500 K
0,02
0,01
0,00 T=2000 K
-0,01
-0,02
-0,03 T=2500 K
-0,04
-0,05
-0,06 T=3000 K
-0,07
0,00 0,25 0,50 0,75 1,00

Composition x

Figure 12. Mixing free energy ∆Gm (eV/pair) for InAsxSb1-x alloy versus composition x.
Short Title 23

Fig. 13 shows the calculated phase diagrams including the spinodal and
binodal curves of InAsxSb1-x, alloy. We observed a critical temperature Tc of
2708 K. The spinodal curve in the phase diagram marks the equilibrium
solubility limit, i.e., the miscibility gap. For temperatures and compositions
above this curve a homogeneous alloy is predicted. The wide range between
spinodal and binodal curves indicates that the alloy may exist as metastable
phase. Hence, our results indicate that the InAsxSb1-x alloy is stable at high
temperature.

3000
Tc =2708 K

2500
Temperature (K)

2000

1500

1000

500

0
0.00 0.25 0.50 0.75 1.00

Composition x

Figure 13. T-x phase diagram of InAsxSb1-x alloy. Solid line: spinodal curve; dashed
line: binodal curve.

V. Conclusions
In summary, using the FP-LAPW, we have presented an analysis of the
structural, elastic, electronic and thermodynamic properties of zinc blende
InAs, InSb and their alloys. Our main results are as follows:

(i) The calculated structural parameters are found to be in reasonable


agreement with available theoretical and experimental work. The
calculated cohesive energy and bulk modulus suggest that InSb is
more compressible than InAs.
(ii) A numerical first-principles method was used to calculate the elastic
constants C11, C12 and C44.
24 K. Hachelafi et al.

(iii) The quasi-harmonic Debye model which combines with the ab


initio calculation of the static total energy was adopted to
calculate the temperature and pressure dependence of the thermal
properties. The thermal expansion and heat capacity are shown to
converge to a nearly constant value at high pressures and
temperatures.
(iv) Our calculated band gap is smaller than that of experimental result
due to the limitation of GGA.
(v) The effective masses of the electrons and the holes at valence and
conduction bands were calculated and compared with previous
investigation.
(vi) The bonding character has been discussed in terms of the charge
density and shows a localization of charge around the anion. Our
GGA calculations predicted the ionicity factor value of 0.372, and
0.316 for InAs and InSb, respectively.
(vii) For InAsxSb1-x alloy the lattice constant closely follows Vegard’s
law, while a significant deviation of the bulk modulus from the
linear concentration dependence. This deviation is mainly due to the
mismatch of the bulk modulus of binary compounds.
(viii)The calculated phase diagram indicates that the InAsxSb1-x alloy is
stable at high temperature.

References
1. Vurgftman, I., Meyer, J.R., Ram-Mohan, L.R. 2001, J.Appl. Phys., 89, 5815.
2. Casey, H.C., Panish Jr, M.B. 1978, Heterostructure Lasers, Part A: Fundamental
Principles, Academic, New York.
3. Luo, L.F., Beresford, R., Wang, W.I. 1998, Appl. Phys. Lett., 53, 2320.
4. Bolognesi, C.R., Dvorak, M.W., Chow, D.H. 1998, J. Vac. Sci. Technol., A 16,
843.
5. Volin, C.E., Garcia, J.P., Dereniak, E.L., Descour, M.K., Hamilton, T.,
McMillan, R. 2001, Appl. Opt., 40, 4501.
6. Kirby, B.J., and Hanson, R.K. 2002, Appl. Opt., 41, 1190.
7. Osiński, M. 2003, Opto-electron. Rev., 11, 321.
8. Aschley, T., Burhe, T.M., Pryce, G.J., Adams, A.R., Andreev, A., Murdin, B.N.,
O’Reilly, E.P., Pidgeon, C.R. 2003, Solid State Electron., 47, 387.
9. Yen, M.Y., Levine, B.F., Bethea, C.G., Choi, K.K., Cho, A.Y. 1987, Appl. Phys.
Lett., 50, 927.
10. Yen, M.Y., People, R., Wecht, K.W., Cho, A.Y. 1988, Appl. Phys. Lett., 52, 489.
11. Fang, Z.M., Ma, K.Y., Jaw, D.H., Cohen, R.M., and Stringfellow, G.B. 1990, J.
Appl. Phys., 67, 7034.
12. Elies, S., Krier, A., Cleverley, I.R., and Singer, K. 1993, J. Phys. D., 26, 159.
13. Huang, K.T., Chiu, C.T., Cohen, R.M., and Stringfellow, G.B. 1994 J. Appl.
Phys., 75, 2857.
Short Title 25

14. Dixit, V.K., Bansal, B., Venkataraman, V., and Bhat, H. L. 2002, Appl. Phys.
Lett., 81, 1630.
15. Gozu, S.-I., Akahane, K., Yamamoto, N., Ueta, A., and Ohtani, N. 2006, Phys. E
32, 230.
16. Neogi, A., Yoshida, H., Mozume, T., Georgiev, N., Akiyama, T., and Wada, O.
2000, Phys. E., 7, 183.
17. Hall, E., Kroemer, H., and Coldren, L.A. 1999, J. Cryst. Growth., 203, 447.
18. Wang, J.-B., Johnson, S.R., Chaparro, S.A., Ding, D., Cao, Y., Sadofyev, Y.G.,
Zhang, Y.-H., Gupta, J.A., and Guo, C.Z. 2004, Phys. Rev. B., 70, 195339.
19. Batyrev, I.G., Norman, A.G., Zhang, S.B., and Wei, S.-H. 2003, Phys. Rev. Lett.
90, 026102.
20. Shahid, M.A., Mahajan, S., Laughlin, D.E., and Cox, H.M. 1987, Phys. Rev.
Lett., 58, 2567.
21. Jiang, W.Y., Liu, J.Q., Zhang, X., Thewalt, M.L.W., Kavanagh, K.L. and
Watkins, S.P. 2006, J. Cryst. Growth., 287, 541.
22. Dorin, C., Mirecki Millunchick, J., Pearson, C., and Wauchope, C. 2005, J. Cryst.
Growth., 283, 8.
23. Sigmund, J., Hartnagel, H.L. 2005, J. Cryst. Growth., 278, 209.
24. Gilsson, T.H., Hauser, J.R., Littlejohn, M.A., and Williams, C.K. 1978, J.
Electron. Mater., 7, 1.
25. Wei, S.-H., Ferreira, L.G., Bernard, J.E., and Zunger, A. 1990, Phys. Rev. B., 42,
9622.
26. Van Vechten, J.A., Bergstresser, T.K. 1970, Phys. Rev. B 1, 3351.
27. Ferhat, M. 2004, Phys. Status Solidi (b) 241, 38.
28. Vinter, B. 2002, Phys. Rev. B., 66, 045324
29. Bouarissa, N., Aourag, H. 1999, Infrared Phys. Technol., 40, 343; Bouarissa, N.,
Amrane, N., Aourag, H. 1995, Infrared Phys. Technol., 36, 755.
30. Rezek mohamed, şenay Katırcıoğlu, Musa El-Hasan 2008, J Mater Sci., 43,
2935; Rezek mohamed, şenay Katırcıoğlu 2008, J. Alloys Compd.,
doi:10.1016/j.jallcom.2008.02.016
31. Belabbes, A., Zaoui, A., Ferhat, M. 2007, Mater. Sci. Eng. B., 137, 210.
32. Yoo, C.S., Akella, J., Campbell, A.J., Mao, H.K., Hemley, R.J. 1995, Science
270, 1473.
33. Blaha, P., Schwarz, K., Madsen, G.K.H., Kvasnicka, D., Luitz, J. 2001, WIEN2k,
An Augmented Plane Wave Plus Local Orbitals Program for Calculating Crystal
Properties, Vienna University of technology, Vienna, Austria; Schwarz, K.,
Blaha, P. 2002, Quantum Mechanical Computations at the Atomic Scale for
Material Sciences, WCCM V, Vienna, Austria; Schwarz, K., and Blaha, P. 2003,
Comput. Mat. Sci., 28, 259; http://www.wien2k.at.
34. Perdew, J. P., Burke, S., Ernzerhof, M. 1996, Phys. Rev. Lett., 77, 3865.
35. Blanco, M. A., Francisco, E., and Luaña, V. 2004, Comput. Phys. Commun., 158, 57.
36. Blanco, M. A., Martín Pendás, A., Francisco, E., Recio, J. M., and Franco, R.
1996, J. Molec. Struct. Theochem., 368, 245.
37. Flórez, M., Recio, J. M., Francisco, E., Blanco, M. A., and Martín Pendás, A.
2002, Phys. Rev. B., 66, 144112.
26 K. Hachelafi et al.

38. Francisco, E., Recio, J. M., Blanco, M. A., Martín Pendás, A. 1998, J. Phys.
Chem., 102, 1595.
39. Francisco, E., Blanco, M.A. and Sanjurjo, G. 2001, Phys. Rev. B., 63, 094107.
40. Poirier, J. P. 2000, Introduction to the Physics of the Earth's Interior, Oxford:
Cambridge University Press, 39.
41. Hill, R. 1952, Proc. Phys. Soc. Lond. A., 65, 349.
42. Murnaghan, F.D. 1944, Proc. Natl. Acad. Sci. USA 30, 5390.
43. Ahmed, R., Javad Hashemifar, S., Akbarzadeh, H., Ahmed, M., Fazal-e-Aleem
2007, Comput. Mater. Sci., 39, 580.
44. Adachi, S. 2005, Properties of Group IV, III-V and II-VI Semiconductors, Wiley,
Cherchester.
45. Boucenna, M., Bouarissa, N. 2007, Mater. Sci. Eng. B., 138, 228.
46. Hellwege, K. H., and Madelung, O. 1982, Landolt–Börnstein, New Series Group
III, vol 17, pt. A, Berlin: Springer.
47. Wyckoff, R.W.G. 1986 Crystal Structures, 2nd Edition Krieger, Malabar.
48. Vubcevich, M. 1972, Phys. Status Solidi b 54, 219.
49. Kalvoda, S., Paulus, B., Fulde, P., and Stoll, H. 1997, Phys. Rev. B., 55, 4027.
50. Mujica, A., Needs, R.J. 1997, Phys. Rev. B., 55, 9659.
51. Wie, S.H., and Zunger, A., 1999, Phys. Rev. B., 60, 5404.
52. Wang, S. Q., Ye, H.Q. 2002, J. Phys.; Cond. Mat., 14, 9579.
53. Paulus, B., Fulde, P., and Stoll, H. 1996, Phys. Rev. B., 54, 2556.
54. Causà, M., Dovesi, R., and Roetti, C. 1991, Phys. Rev. B., 43, 11937.
55. Einstein, A. 1907, Ann. Phys., 22, 180.
56. Nernst, W., Lindemann, A. F., and Elektrochem, Z. 1977, Angew. Phys. Chem.,
17, 817.
57. Debye, P. 1912, Ann. Phys., 39, 789.
58. Petit, A. T., Dulong, P. L. 1819, Ann. Chim. Phys., 10, 395.
59. Ley, L., Pollak, R.A., McFeely, F.R., Kowalczyk, S.P., Shirley, D.A. 1974, Phys.
Rev. B., 9, 600.
60. Chelikowsky, J.R., Cohen, M.L. 1976, Phys. Rev. B., 14, 556.
61. Huang, M., Ching, W.Y. 1985, J Phys Chem Solids., 46, 977.
62. Talwar, D.N., Ting, C.S. 1982, Phys. Rev. B., 25, 2660.
63. Kittel, C. 1976, Introduction to Solid State Physics, 5th edn., Wiley, New York.
64. Harrison, W.A. 1980, Electronic Structure and the Properties of Solids, Freeman,
San Francisco.
65. Boguslawski, P., Baldereshi, A. 1989, Solid State Commun., 70, 1085.
66. Ukhanov, Y.I., Mal’tsev, Y.V. 1963, Sov. Phys. Solid State., 5, 1124.
67. Kazakova, L.A., Kostsova, V.V., Karymshakov, R.K., Ukhanov, Y.I., Yagup’ev,
V.P. 1972, Sov. Phys. Semicond., 5, 1495.
68. Chasmar, R.P., Stratton, R. 1956, Phys. Rev., 102, 1686.
69. Nakwaski, W. 1995, Physica B., 210, 1, and references therein.
70. Zaoui, A., Ferhat, M., Khelifa, B., Dufour, J.P., Aourag, H. 1994, Phys. Status
Solidi b., 185, 163.
71. Phillips, J.C. 1970, Rev. Mod. Phys., 42, 317.
72. Vegard, L. 1921, Z. Phys., 5, 17.
Short Title 27

73. El Haj Hassan, F., Javad Hashemifar, S., and Akbarzadeh, H. 2006, Phys. Rev.
B., 73, 195202; El Haj Hassan, F., and Akbarzadeh, H. 2006, Comput. Mater.
Sci., 35, 423.
74. Swalin, R. A. 1961, Thermodynamics of Solids,Wiley, New York.
75. Ferreira, L.G., Huai Wie, S., Zunger, A. 1989, Phys. Rev. B., 40, 3197.
76. Teles, L.K., Furthmüller, J., Scolfaro, L.M.R., Leite, J.R., Bechstedt, F. 2000,
Phys. Rev. B., 62, 2475.

You might also like