You are on page 1of 17

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/332344654

First principles calculations of Elastic and Thermodynamic Properties Under


Hydrostatic Pressure of Cubic InNxP1-x Ternary Alloys

Article  in  Chinese Journal of Physics- Taipei- · April 2019


DOI: 10.1016/j.cjph.2019.04.011

CITATIONS READS

4 479

4 authors:

Hattabi Ismail Abdiche Ahmed


Université Ibn Khaldoun Tiaret Université Ibn Khaldoun Tiaret
6 PUBLICATIONS   17 CITATIONS    51 PUBLICATIONS   435 CITATIONS   

SEE PROFILE SEE PROFILE

S. H. Naqib Rabah Khenata


University of Rajshahi University Mustapha Stambouli of Mascara
266 PUBLICATIONS   2,851 CITATIONS    664 PUBLICATIONS   10,202 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

material science - experimental and theoretical investigation View project

Ternary Superconducting Compounds: Physical Properties from DFT Calculations View project

All content following this page was uploaded by Rabah Khenata on 27 August 2019.

The user has requested enhancement of the downloaded file.


Chinese Journal of Physics 59 (2019) 449–464

Contents lists available at ScienceDirect

Chinese Journal of Physics


journal homepage: www.elsevier.com/locate/cjph

First-principles calculations of elastic and thermodynamic


T
properties under hydrostatic pressure of cubic InNxP1-x ternary
alloys

I. Hattabia, , A. Abdicheb,c, S.H. Naqibd, R. Khenatab
a
Ibn Khaldoun Université de Tiaret, Post Box 78-Zaaroura, Tiaret 14000, Algeria
b
Laboratoire de Physique Quantique de la Matière et de Modélisation Mathématique (LPQ3M), Université de Mascara, Mascara 29000, Alegria
c
Electrical Engineering Department, Ibn Khaldoun University of Tiaret, Post Box 78-Zaaroura, Tiaret 14000, Algeria
d
Department of Physics, University of Rajshahi, Rajshahi 6205, Bangladesh

ABS TRA CT

We present a first-principles study of the elastic and thermodynamic properties of the zinc-blend (ZB) InNxP1-x ternary alloy in the full range of
0 ≤ x ≤ 1. Calculations were carried out using WIEN2k code based on a non-relativistic full potential linearized augmented plan wave (FP-LAPW)
method within the density functional theory (DFT). The approximation of exchange-correlation energy was performed with Generalized Gradient
Approximation (GGA). The elastic constants and related mechanical properties were investigated. We have found that the stiffness coefficients
increase with the increase of nitrogen content. The values of the bulk modulus derived from elastic stiffness coefficients are found to be very close to
the values derived directly from Murnaghan equation of state, indicating the good accuracy of the Cij coefficients. The elastic properties were also
calculated under hydrostatic pressure for (P = 0.00, 5.00, 10.0, 15.0, 20.0, 25.0 GPa). Furthermore, we have performed a quasi-harmonic Debye
model calculation using GIBBS2 package to predict the thermodynamic properties and their dependences on the temperature and pressure.
Thermodynamic parameters such as the Gibbs free energy, entropy, heat capacity and Debye temperature have been investigated. Our results show a
good agreement with available theoretical and experimental data.

1. Introduction

Semiconductor compounds are currently the focus of intense attention, particularly Group-III nitrides due to their successful
applications in different fields of electronics and optoelectronic technologies as light-emitting diodes, laser diodes, and high-electron-
mobility transistors.
In recent years, computational materials science based on first-principles calculations has emerged as an effective supplement to
experimentation, especially because it enables powerful prediction of material properties that depend only on electronic density and
atomic bonding [1]. The InN and InP compounds are interesting semiconductors materials for high-frequency electronic devices
[2–6], due to their superior speed electrons. It is of crucial importance to have an accurate knowledge of material parameters such as
structural parameters, electronic and optical properties as well as elastic coefficients, temperature and pressure dependence of
thermodynamic properties, for a reliable modelling of nitride-based devices.
Recently, we have investigated the structural, electronic and optical properties of the cubic InN and InP binary compounds
together with their related ternary InNxP1-x compounds under hydrostatic pressure [7]. To the best of our knowledge, there is neither
experimental nor theoretical data in the literature on the elastic constants and the thermodynamic properties for the InNxP1-x ternary
alloy.


Corresponding author.
E-mail address: ismail.hattabi@univ-tiaret.dz (I. Hattabi).

https://doi.org/10.1016/j.cjph.2019.04.011
Received 23 December 2018; Received in revised form 11 March 2019; Accepted 10 April 2019
Available online 25 April 2019
0577-9073/ © 2019 The Physical Society of the Republic of China (Taiwan). Published by Elsevier B.V. All rights reserved.
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Table 1
The calculated elastic constants C11, C12 and C44 for zinc blende InN1-x Px alloys for different compositions (x = 0.00, 0.25, 0.50, 0.75, 1.00).
Composition x Parameters

C11 C12 C44

InP This work 98,796 57,268 72,709


Theoretical works 101.1a 57.6a 75.3a
105.9b 56.4b 49.3b
94.87c 54.02c 54.67c
109d 55.7d 52.6d
Experimental works 101.1m 56.1m 45.6m
101n 56n 46n
102.2o 57.6o 46o
InN0.25P0.75 This work 101,949 64,605 74,030
InN0.5P0.5 This work 131,3294 69,8766 89,4421
InN0.75P0.25 This work 147,149 92,114 108,818
InN This work 174,986 130,182 139,245
Theoretical works 188.7–159.3b 102–125b 93.3b
182e 125e 177i
172f,j 124.85h 138j
183.89h 119.4l 137.6k
184i
Experimental works – – –

a
ref [31].
b
ref [32].
c
ref [33].
d
ref [34].
e
ref [35].
f
ref [36].
h
ref [37,38].
i
ref [39,40].
j
ref [39,41,42].
k
ref [43].
l
ref [44].
m
ref [12].
n
ref [61].
o
ref [62].

As elastic constants determine the response to an applied stress, it is very important to study the dependence of the elastic
constants and their related properties as well the physical properties dependence on strain effect. The effect of strain on electronic
properties requires knowledge of the mechanical properties, specifically the elastic constants [8].
There are a number of theoretical and experimental calculations for the elastic constants and thermal properties of binary
compounds using various methods. The elastic properties of InP were studied by K. J. Bachmann [9], Yogurtçu et al. [10], Rimai and
co-workers [11] measured experimentally the elastic constant of gallium and indium phosphide. In addition to the experimental
measurements, theoretical studies made a noteworthy development in prediction and description of the elastic properties of these
compounds, where ab initio pseudo-potential methods have been used [12–15].
The knowledge of thermodynamic properties is required for the manufacturing of advanced semiconductor materials to be used
under different environmental conditions. In fact, several significant parameters of III-nitrides, such as band-gap energy, internal
electric field and device lifetime [16–18] can be affected by thermal parameter variations. Hence, accurate knowledge of the later is
very important from both physical point of view and for device engineering.
Besides, the elastic constants, it seems to be significant to systematically study thermal stabilities, thermodynamic properties
which can provide a better basis for further theoretical and experimental investigations.
In addition, hydrostatic pressure is a thermodynamic variable notably attractive to exploit in such computational principles as it
offers the possibility of understanding the variations of solids’ properties as inter-atomic distances are changed in a systematic way
[19,20]. This provides with important complementary data to the experimental work, predicting stability of structures. Although
there exists numerous theoretical calculations on a wide variety of physical properties of III-nitrides [21–25], relatively little is known
about the pressure behaviour of these compounds.
Therefore, the aim of this paper is to give a detailed description of the behaviour of elastic and thermodynamic properties of the
ZB structure of the InNxP1−x alloys in the full range of x concentration 0 < x < 1 under hydrostatic pressure, using the FP-LAPW
method within density functional theory (DFT) as implemented in the Wien2k code.

2. Computational details

We have performed a DFT based calculation with the all-electron full-potential linearized plane wave (FP-LAPW)-based code

450
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Table 2
The calculated mechanical parameters, Bulk modulus (B), Young modulus (E), Shear modulus (G) and Passion's Ratio (ν) for zinc blende InN1-x Px
alloys for different compositions (x = 0.00, 0.25, 0.50, 0.75, 1.00).
Composition x Parameters

B E G ν B/G Hv

InP This work 71,11 109,71 44,13 0,24 1,61 6,73


Other works 71.27a 114.1a 46.1a 0.23a – 6.0–5.4g
72.9–71b 60.7e 34.5e 4.99e
67.0c
71.1d
InN0.25P0.75 This work 77,05 108,50 42,87 0,26 1,79 5,47
InN0.5P0.5 This work 90,36 142,52 57,60 0,23 1,56 8,63
InN0.75P0.25 This work 110,45 158,99 63,08 0,26 1,75 7,82
InN This work 145,11 178,26 68,81 0.29 2,10 6,23
Other works 144.53h 128.34f 62.18–65.1m 0.32f 2.54f 10g
139j 162.09m 71.4k
140j
144l

a
ref [31].
b
ref [32].
c
ref [45].
d
ref [46].
e
ref [47].
f
ref [48].
g
ref [49].
h
ref [37,38].
j
ref [39,41,42].
k
ref [43].
l
ref [44].
m
ref [50,51].

Fig. 1. Variation of the elastic constants versus composition, x, of the (ZB) InNxP1_x alloys.

Wien2K [26], where the generalized gradient approximation (WC-GGA) was considered for the exchange-correlation potential to
compute structural properties such as lattice constants and bulk modulus. Energy-volume results were fitted using the Birch-Mur-
naghan's equation of state [27].
In order to calculate elastic properties, we have used a cubic cell with eight atoms to model the InNxP1−x alloys with different
compositions x = 0.0, 0.25, 0.5, 0.75, and 1.0. Values of lattice parameters used in this work are those previously calculated in Ref.
[7] and correspond to 5.88, 5.71, 5.50, 5.26 and 4.99 (Å), respectively. The stiffness coefficients, bulk modulus, shear modulus and
young modulus were investigated for each composition under hydrostatic pressure for different values (P = 0, 5, 10, 15, 20, 25 GPa).

451
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Fig. 2. Variation of the bulk modulus versus composition, x, of the (ZB) InNxP1_x alloys.

Fig. 3. Variation of shear and Young's moduli versus composition, x, of the (ZB) InNxP1_x alloys.

The matrix size was defined with a parameter Rmt × Kmax equals to 7, where the Rmt denotes the minimum radius of the muffin-tin
sphere in unit cell, and Kmax gives the extent of the largest K vector in the plane wave expansion. The muffin-tin radii of In, N and P
are adopted to be 2.1, 1.4 and 1.7 Bohr, respectively. The integrals along the Brillouin zone are extended up to 47 k-points.
The thermal properties such as Gibbs free energy, entropy, heat capacity and Debye temperature were studied by using GIBBS2
package based on the quasi-harmonic Debye model [28].

3. Results and discussion

3.1. Mechanical properties

Knowledge of the basic mechanical properties of intermetallic compound allows to estimate and evaluate its mechanical stability
and to clearly find out its purpose.
Elastic properties supply a link between the mechanical and dynamical behaviours of the crystals and give important information
which correlates with many physical properties such as the Debye temperature, specific heat capacity, melting point and thermal

452
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Fig. 4. Variation of B/G ratio and Poisson's ratio ν, versus the composition x of the (ZB) InNxP1_x alloy.

expansion coefficient. These parameters are vital for understanding of the thermo-mechanical behaviour. Moreover, studying the
aforementioned parameters at the high pressure also provides information about the material response, mechanical stability, and
strength under compression.
For a cubic system, there are three independent elastic constants denoted by C11, C12 and C44 and can be determined by the
analysis of changes in calculated stress values resulting from changes in the strain [29,30]. In order to determine them, the cubic unit
cell is deformed using an appropriate strain tensor to yield an energy-strain relation. In this paper, we have used the FP-LAPW
method implemented in the Wien2k package [26].
The calculated elastic constants C11, C12 and C44 of cubic InNxP1-x alloys using the WC-GGA approximation, for different com-
positions (x = 0.00, 0.25, 0.50, 0.75, 1.00), are computed and summarized in Table 1, together with the available experimental
values and other theoretical data for comparison. The elastic behaviour of a lattice is described by its matrix of second-order elastic
constants given by:

1 ⎛ ∂ 2E ⎞
Cij = ⎜ ⎟
V 0 ⎝ ∂εi ∂εj ⎠ (1)

where E denotes the energy of the crystal, V0 is the equilibrium volume and ε represents the strain component. This elastic matrix has
size 6 × 6 and is symmetric. For arbitrary homogeneous deformation by an infinitesimal strain, the energy of the crystal is expressed
as:
6
1
E = E0 + V0 ∑ Cij εi εj + O (ε 3)
2 i, j = 1 (2)

the elastic energy, given by the quadratic form of Eq. (2), is always positive
(E > 0, ∀ ε ≠ 0). Cii > 0, ∀ i (3)

(Cij)2 < Cii Cjj, ∀ i, j (4)

Eqs. (3) and (4) represent the generalized criteria of mechanical stability for a given crystal.
For a cubic crystal, the case of this study, the mechanical stability criteria are given by:
C11 – C12 > 0, C11> 0, C44> 0, C11> |C12|, (C11 + 2C12) > 0. According to the above criteria, elastic constants of InNxP1-x clearly
satisfy all of the mechanical stability criteria, indicating their structural stability [52,53].
From the calculated elastic constants Cij, and in order to get a deeper insight, one can determine mechanical properties that
include bulk modulus (B), shear modulus (G), Young modulus (Y) and Poisson's ratio (n). These quantities are calculated using the
VRH method [52], where the arithmetic average of Voigt and Reuss's bounds is written as:
GV + GR
G=
2 (5)

where GV = (C11 − C12 + 3C44)/5 is Voigt's shear modulus corresponding to the upper bound of G value and GR = 5(C11 − C12) C44/
[4C44 + 3(C11 − C12)] is Reuss's shear modulus conforming to the lower bound of G value. Besides, we have-

453
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Fig. 5. The dependence of the elastic constants on pressure for InN and InP binary compounds.

C11 + 2C12
B=
3 (6)

9BG
E=
3B + G (7)

3B − E
ν=
6B (8)

Vicker's Hardness, HV, is related to the elastic properties of crystalline materials, and can be predicted by a semi-empirical
equation given as [52]:
Hv = 2(K2G)0.5853 (9)

where, K = G/B.
The obtained mechanical parameters of InNxP1-x compounds are listed in Table 2. As mentioned before, there exists no report of
the elastic properties of InNxP1-x ternary compounds, which does not allow for making comparisons. However, it can be seen that the
calculated bulk modulus from the elastic constants is in excellent concordance with that obtained from the total energy minimization

454
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Fig. 6. The dependence of the elastic constants on pressure for InN0.5P0.5 ternary alloy.

Table 3
The Elastic constants Cij of the InN, InP and InN0.50P0.50 compounds at different pressure values.
Hydrostatic pressure P (GPa)

Composition x Elastic constants Cij 0 5 10 15 20 25

InP C11 98.79 120.36 140.58 161.68 174.84 189.29


C12 57.26 78.26 96.56 112.31 130.11 146.29
C44 72.7 93.88 113.58 135.85 150.43 168.99
InN0.25P0.75 C11 101.94 121.98 150.36 168.33 178.33 193.99
C12 64.6 86.63 101.32 118.68 137.95 153.9
C44 74.03 95.3 122.11 141.22 151.79 170.07
InN0.5P0.5 C11 165.33 147.66 170.72 179.51 200.61 210.59
C12 114.77 92.45 110 132.88 148.25 168.4
C44 133.82 113.82 134.56 149.77 166.66 178.85
InN0.75P0.25 C11 147.14 167.2 183.32 197.69 214.36 237.15
C12 92.11 113.28 134.03 154.62 171.98 186.14
C44 108.81 128.31 148.67 165.11 181.93 196
InN C11 174.98 204.23 216.7 250.06 274.59 262.14
C12 130.18 146.8 170.18 183.39 200.23 233.87
C44 139.24 149.7 162.84 193.27 215.62 226.33

and this might be a consideration for the reliability and accuracy of our predicted elastic constants for the InNxP1-x alloy [7].
We show in Fig. 1 the variation of the elastic constants versus composition, x, of the InNxP1-x alloy. From this figure it is seen that
the elastic constants increase in all directions with increasing x. This is due to the decrease of the anionic radius and the ionicity
(related to the Bader charge) decrease as the anion radius decreases. Therefore, it implies that the elastic constant increases when the
lattice parameter of the crystal structure decreases [32]. As it is the case of InNxP1-x alloy, in fact, the gradual substitution of
phosphorus with nitrogen implies a decrease of lattice parameter with an increase of elastic coefficients.
The calculated bulk modulus of InNxP1-x alloys are plotted as a function of the composition x in Fig. 2. Moreover, we have plotted
the variation of the shear and Young's moduli as function of composition, x, of the (ZB) InNxP1-x alloys in Fig. 3. One can observe the
increase of the shear and young moduli in InNxP1-x ternary compounds with increasing x composition and this can be explained by the
effect of nitrogen on bond strength [54]. This allows us to note that the incorporated nitrogen forms stronger bonds than that formed
by the substituted phosphorus
Pugh's criterion [55] is used to classify the material's nature as ductile or brittle, where the B/G ratio was calculated and plotted as
function of composition x, for the InNxP1-x ternary alloys in Fig. 4. If B/G > 1.75, the material will be ductile, otherwise it will be of
the brittle in nature. It is found that InN is the more ductile material (B/G = 2.1), while the most fragile and the hardest compound is
the InN0.5P0.5 ternary alloy, corresponding to B/G = 1.57 and Hv = 8.64 GPa (Table 2). The results indicate rather impressive al-
loying effects for 50% nitrogen doping with a fragile and ductile mixed kind of behaviour. Moreover, as it is shown in Fig. 4 and as
well B/G ratio variations, Poisson's ratio reproduce the same trend versus composition x, which confirm the ductile/brittle behaviour
of the InNxP1-x alloys.

455
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Fig. 7. The dependence of the elastic moduli on pressure for InN and InP binary compounds.

Furthermore, we have investigated the mechanical behavior of the latter compounds under the effect of hydrostatic pressure using
the FP-LAPW method, firstly, the lattice constant has been calculated as function of pressure at P equal to 0, 5, 10, 15, 20 and 25 GPa,
using the following formula [56]:

2
( 3−B1′ )
B′
a (P ) = a0 ⎡1 + P ⎤
⎣ B⎦ (10)

where B is the bulk modulus, B' is the pressure derivative of the bulk modulus, and a(P) is the lattice parameter at the pressure P.
The pressure dependent elastic constants of the InNxP1-x compounds were computed and plotted in Figs. 5 and 6 for InN, InP
binary and InN0.5P0.5 ternary alloys.
In the considered range, the hydrostatic pressure dependence of stiffness coefficients displays a linear dependence, when the
applied pressure is increased, an increase of elastic constants is observed for InNxP1-x alloys in the full range of composition x, and
that is due to the decrease of the lattice constant. The values are summarized in Table 3. Besides, we have investigated the variations
of shear, Young and bulk moduli as function of pressure, where the variations curves are depicted in Figs. 7 and 8 respectively, for
InN, InP binary and InN0.5P0.5 ternary alloys. Results obtained here are compared to the available data where a good agreement is
observed for binary constituents; contrariwise, we have to note that we have not found any report in literature on these parameters

456
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Fig. 8. The dependence of the elastic moduli on pressure for InN0.5P0.5 ternary alloy.

Fig. 9. Gibbs free energy versus temperature for InP binary compound at different pressures.

for ternary compounds.

3.2. Thermal properties

In this section, we investigate the thermodynamic properties of InNxP1-x ternary alloys using the GIBBS2 program [57,58] within
the framework of the quasi-harmonic approximation. This model requires only the input of the static E(V) fitted Data from birch-
Murnaghan equation, to compute different thermal properties such as variations of Gibbs free energy, heat capacity, entropy and
Debye temperature, as function of temperature and pressure respectively, in the range of 0 to 1200 K and 0 to 25 GPa.
The non-equilibrium Gibbs function G*(V, P, T) is expressed as [57]:
G (V , T , P ) = E (V ) + PV + Avib [θD (V ), T ] (11)

Where E(V) is the total energy per unit cell, PV is the constant hydrostatic pressure condition, θD(V) is the Debye temperature and Avib
is the vibrational Helmhotz free energy, which can be written using the Debye model of the phonon density of states as:

9θ θD θ
Avib (θD , T ) = nKB T ⎡ D + 3ln ⎛1 − e− T ⎞ − D ⎛ D ⎞ ⎤

⎣ 8T ⎝ ⎠ ⎝ T ⎠⎥⎦ (12)

457
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Fig. 10. Gibbs free energy versus temperature for InN binary compound at different pressures.

Fig. 11. Gibbs free energy versus temperature for InN0.50P0.50 ternary alloy at different pressures.

Where D(θD/T) represents the Debye integral, n is the number of atoms per formula unit,
For an isotropic solid, θD is expressed as [57]:
1
ℏ 1 BS
θD = ⎡6π 2V 2 n ⎤3 f (σ )
k⎣ ⎦ M (13)
M being the molecular mass per unit cell and BS the adiabatic bulk modulus, approximated by the static compressibility [59]

d 2E (V ) ⎞
BS ≅ B (V ) = V ⎛ ⎜
2

⎝ dV ⎠ (14)
The thermodynamic quantities such as heat capacity at constant volume (CV) and entropy S were calculated using the following
relationships [57]:

θ 3θ / T ⎤
CV = 3nk ⎡4D ⎛ ⎞ − θ / T

⎣ ⎝T ⎠ e − 1⎥
⎦ (15)

θ
S = nk ⎡4D ⎛ ⎞ − 3ln (1 − e−θ / T ) ⎤

⎣ ⎝ T ⎠ ⎥
⎦ (16)
Firstly, we have studied the temperature and pressure dependence of the Gibbs free energy, G. Results are shown for both InP, InN

458
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Fig. 12. Debye temperature θD as function of temperature for InP binary compound at different pressures.

Fig. 13. Debye temperature θD as function of temperature for InN binary compound at different pressures.

Fig. 14. Debye temperature θD as function of temperature for InN0.50P0.50 ternary alloy at different pressures.

459
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Fig. 15. Variation of the heat capacities Cv versus temperature for InP binary compound at different pressures.

Fig. 16. Variation of the heat capacities Cv versus temperature for InN binary compound at different pressures.

Fig. 17. Variation of the heat capacities Cv versus temperature for InN0.50P0.50 ternary alloy at different pressures.

460
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Fig. 18. Variation of the entropy S versus temperature for InP binary compound at different pressures.

Fig. 19. Variation of the entropy S versus temperature for InN binary compound at different pressures.

binary and InN0.5P0.5ternary alloys in Figs. 9–11 respectively. One can see that the Gibbs free energy, increases when pressure
increases at fixed temperature, per contra it decreases slightly with increasing temperature at a given pressure.
The Debye temperature is an important characteristic temperature of solids. When the temperature rises above absolute zero, the
atoms of the solid gradually vibrate to Debye's temperature. This represents the temperature at which the vibrations reach their
maximum of possible modes. It is a good approximation of the hardness of solids. Figs. 12–14 show the Debye temperature θD as
function of temperature, respectively for InP, InN binary and InN0.5P0.5 ternary alloys at different pressures. At a given pressure,
Debye temperature θD is found to be nearly constant from 0 K to 200 K. In this temperature range the anharmonicity is low and the
crystal expands very little with increasing temperature. This results in almost a constant Debye temperature. At low temperatures one
expects the high-frequency modes to be frozen, i.e., vibrational excitations arise solely from acoustic vibrations, it is proportional to
the sound velocity and directly related to the elastic constants, θD decreases gradually and almost linearly with increasing tem-
perature, indicating the change in the vibrational spectrum of the atoms with the temperature. θD is associated with the maximum
crystal vibration frequency for a given structure and symmetry. At room temperature and zero pressure the values of Debye tem-
perature are 317.38, 457.58 and 350.59 K for InP, InN and InN0.5P0.5, respectively.
The variations of heat capacity Cv versus temperature at different pressures of InP, and InN binaries and InN0.5P0.5 ternary alloys
are presented in Figs. 15–17, respectively. From these figures, for low temperature, the proportionality of heat capacity to T3 is clear
as given in Debye model. Cv increases rapidly with temperature due to the enhanced anharmonic effect at elevated temperatures,
while for higher temperatures it tends to the classical Dulong-Petit limit [60], approximately to 49.7, 49.6 and 186,5586 (J mol−1
K−1) for InP, InN and InN0.5P0.5, respectively. In addition, Cv increases with the temperature rise at a given pressure and slightly
decreases with the pressure rise at a given temperature. The effects of temperature on the heat capacity at low temperatures (below

461
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

Fig. 20. Variation of the entropy S versus temperature for InN0.50P0.50ternary alloy at different pressures.

Table 4
Thermodynamic properties, Gibbs free energy (G), Debye temperature θD, isochoric heat capacity Cv and entropy S, of the InN, InP and InN0.50P0.50
compounds at room temperature and different pressure values.
Hydrostatic pressure P (GPa)

Composition x Thermodynamic parameters 0 5 10 15 20 25

InP G (KJ/mol) −1,63423E7 −1,63422E7 −1,6342E7 −1,63419E7 −1,63418E7 −1,63417E7


θD (K) 317,62 362,49 406,95 435,15 454,59 486,71
Cv (J/mol.K) 47,19907 46,42685 45,58239 45,00905 44,59796 43,89219
S (J/mol.K) 65,03933 58,85164 53,52724 50,49193 48,53381 45,51292
InN0.25P0.75 G (KJ/mol) −6,46151E7 −6,46145E7 −6,4614E7 −6,46135E7 −6,4613E7 −6,46126E7
θD (K) 328,14 377,33 416,35 448,9 477,96 504,53
Cv (J/mol.K) 188,1024 184,6138 181,5778 178,878 176,3501 173,9485
S (J/mol.K) 254,0172 227,9777 209,9548 196,3837 185,2421 175,7664
InN0.5P0.5 G (KJ/mol) −6,38609E7 −6,38604E7 −6,38599E7 −6,38595E7 −6,3859E7 −6,38586E7
θD (K) 350,59 397,26 435,47 464,93 493,22 524,58
Cv (J/mol.K) 186,5586 183,0914 180,0094 177,4973 174,9808 172,0827
S (J/mol.K) 241,6165 218,5135 201,835 190,1342 179,7213 169,0223
InN0.75P0.25 G (KJ/mol) −6,31068E7 −6,31064E7 −6,3106E7 −6,31056E7 −6,31052E7 −6,31048E7
θD (K) 387,36 428,91 462,5 493,06 521,75 548,13
Cv (J/mol.K) 183,855 180,5535 177,7085 174,9951 172,349 169,8373
S (J/mol.K) 223,1425 204,573 191,0637 179,7773 169,9549 161,5134
InN G (KJ/mol) −1,55759E7 −1,55758E7 −1,55757E7 −1,55756E7 −1,55755E7 −1,55754E7
θD (K) 457,58 492,65 523,64 551,48 576,96 600,8
Cv (J/mol.K) 44,53351 43,75806 43,04283 42,37849 41,75385 41,15686
S (J/mol.K) 48,24115 44,98069 42,33287 40,12029 38,21966 36,54163

θD) are very significant. The entropy is a measure of the dispersion of matter and energy which can be described as the number of
arrangements of particles, bosonic quasiparticles or disorder of a system [61].
Based on the quasi-harmonic Debye model, another important thermodynamic parameter, entropy S, is also computed at different
pressures in the range from 0.0 to 30 GPa with the step of 5.0 GPa for InP, InN and InN0.5P0.5 in the temperature range of 0–1200 K, as
shown in Figs. 18–20, respectively. At normal condition, the obtained values of entropy are 65.039, 48.24 and 241.61 J mol−1 K−1
for InP, InN and InN0.5P0.5, respectively. It is established that the entropy S increases in a non-linear way with the increase of the
temperature and decreases with the increase of pressure. We can also observe that the decrease of the entropy S with pressure is
slower and slower when the temperature increases. One can also remark that the entropy S of the ternary alloy obtained by mixing
the binary InP and InN is significantly higher than those obtained separately for the binary compounds. This is expected, since mixing
invariably adds to the disorder and increases the number of possible microstates which leads to higher entropy. It is also interesting to
note that as the pressure is progressively increased, its effect on entropy at a given temperature decreases monotonically. The value of
S also decreases with pressure. These can be explained qualitatively by considering the fact that at high applied pressures the crystal
structure becomes stiffer and at any given temperature the number of phonon modes is decreased as the amplitude of the atomic
vibrations are reduced. The calculated values at room temperature of the thermodynamic properties such as Gibbs free energy (G),
Debye temperature θD, isochoric heat capacity Cv and entropy S, of the InN, InP and InN0.50P0.50 compounds at different pressure

462
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

values are listed in Table 4.

4. Conclusion

First principles calculations based on the DFT theory have been employed, using the FP-LAPW method, to investigate the elastic
and thermodynamic properties of zinc-blende InN, InP and InNxP1-x ternary alloys. The hydrostatic pressure dependence on stiffness
coefficients as well different elastic moduli and ratios have also been considered.
It is found that elastic constants of InNxP1-x clearly satisfy all of the mechanical stability criteria, indicating their structural
stability. In addition, the incorporation of nitrogen increases the stiffness coefficients, therefore, the increase of shear and young
moduli, given that a considerable alloying effects was found with brittle and ductile mixed behaviour for 50% nitrogen substitution.
Further, the hydrostatic pressure effects on stiffness coefficients display an increasing linear dependence. Finally, temperature and
pressure dependent thermal properties have been investigated using the quasi-harmonic approximation, providing a good description
of free Gibbs energy G, Debye temperature θD, entropy S, and heat capacity Cv in the pressure range from 0 to 25 GPa and tem-
perature range from 0 to 1200 K.

References

[1] G. Chris, V. Walle, J. Neugebauer, J. Appl. Phys. 95 (2004) 3851 https://doi.org/10.1063/1.1682673.


[2] T. Matsuoka, Superlatt. Microstruc. 37 (1) (2005) 19–32 https://doi.org/10.1016/j.spmi.2004.06.003.
[3] W.J. Schaff, H. Lu, J. Hwang, H. Wu, Proceedings of the 7th IEEE/ Cornell Conf. “Advanced Concepts in High Performance Devices,”, 2000, p. 225 7–9 Aug.
[4] Y.C. Kong, Y.D. Zheng, C.H. Zhou, Y.Z. Deng, B. Shen, S.L. Gu, R. Zhang, P. Han, R.L. Jiang, Y. Shi, Solid-State Electron 49 (2) (2005) 199–203 https://doi.org/
10.1016/j.sse.2004.08.016.
[5] W. Deal, X.B. Mei, Kevin M.K.H. Leong, V. Radisic, S. Sarkozy, IEEE Trans. Terahertz Sci. Technol. 1 (1) (2011) 25–32 https://doi.org/10.1109/TTHZ.2011.
2159539.
[6] I. Moerman, P.P. Van Daele, P.M. Demeester, IEEE J. Select. Topics Quant. Electron. 3 (6) (1997) 1308–1320 https://doi.org/10.1109/2944.658785.
[7] I. Hattabi, A. Abdiche, R. Moussa, R. Riane, K. Hadji, F. Soyalp, Dinesh Varshney, S.V. Syrotyuk, R. Khenata, Zeitschrift für Naturforschung A 71 (9) (2016)
783–796 https://doi.org/10.1515/zna-2016-0100.
[8] R.W. Keyes, J. Appl. Phys. 33 (1962) 3371 https://doi.org/10.1063/1.1931171.
[9] K.J. Bachmann, Rev. Mater. Sci. 11 (1) (1981) 441–484 https://doi.org/10.1146/annurev.ms.11.080181.002301.
[10] Y.K. Yogurtçu, A.J. Miller, G.A. Saunders, J. Phys. Chem. Solids 42 (1) (1981) 49–56 https://doi.org/10.1016/0022-3697(81)90010-X.
[11] D.S. Rimai, R.J. Sladek, Solid State Commun. 30 (10) (1979) 591–594 https://doi.org/10.1016/0038-1098(79)90102-9.
[12] D.N. Nichols, D.S. Rimai, R.J. Sladek, Solid State Commun. 36 (8) (1980) 667–669 https://doi.org/10.1016/0038-1098(80)90205-7.
[13] P. Rodríguez-Hernández, A. Munoz, Semicond. Sci. Technol. 7 (1992) 1437 http://iopscience.iop.org/article/10.1088/0268-1242/7/12/002/pdf.
[14] P. Rodríguez-Hernández, M. Gonzalez-Diaz, A. Munoz, Phys. Rev. B 51 (1995) 14705 https://doi.org/10.1103/PhysRevB.51.14705.
[15] K. Kabita, J. Maibam, B.I. Sharma, R.K. Thapa, R.K. Brojen Singh, Ind. J. Phys. 89 (2015) 1265 https://doi.org/10.1007/s12648-015-0701-0.
[16] L. Dong, S.K. Yadav, R. Ramprasad, S.P. Alpay, Appl. Phys. Lett. 96 (2010) 202106https://doi.org/10.1063/1.3431290.
[17] L. Zhang, K. Cheng, S. Degroote, M. Leys, M. Germain, G. Borghs, J. Appl. Phys. 108 (2010) 073522https://doi.org/10.1063/1.3493115.
[18] L.C. Xu, R.Z. Wang, X. Yang, H. Yan, J. Appl. Phys. 110 (2011) 043528https://doi.org/10.1063/1.3627237.
[19] A. Mujica, A. Rubio, A. Muñoz, R.J. Needs, Rev. Mod. Phys. 75 (2003) 863 https://doi.org/10.1103/RevModPhys.75.863.
[20] G.J. Ackland, Rep. Prog. Phys. 64 (2001) 483 http://iopscience.iop.org/article/10.1088/0034-4885/64/4/202/meta.
[21] J.W. Orton, C.T. Foxon, Rep. Prog. Phys. 61 (1998) 1 http://iopscience.iop.org/article/10.1088/0034-4885/61/1/001/meta.
[22] S.C. Jain, M. Willander, J. Narayan, R. van Overstraeten, J. Appl. Phys. 87 (2000) 965 https://doi.org/10.1063/1.371971.
[23] I. Vurgaftman, J.R. Meyer, J. Appl. Phys. 94 (2003) 3675 https://doi.org/10.1063/1.1600519.
[24] N. Bouarissa, Eur. Phys. J. B 26 (2002) 153 https://doi.org/10.1140/epjb/e20020076.
[25] N. Bouarissa, Philos. Mag. B 80 (2000) 1743 https://doi.org/10.1080/13642810008216503.
[26] K. Schwarz, P. Blaha, Comput. Mater. Sci. 28 (2003) 266, https://doi.org/10.1016/S0927-0256(03)00112-5.
[27] F.D. Murnaghan, Proc. Natl. Acad. Sci. U.S.A. 30 (1944) 244.
[28] A. Otero-de-la-Roza, V. Luaña, “Gibbs2: a new version of the quasi-harmonic model. I. Robust treatment of the static data”, Comput. Phys. Commun. 182 (2011)
1708–1720 https://doi.org/10.1016/j.cpc.2011.04.016.
[29] B.B. Karki, L. Stixrude, S.J. Clark, M.C. Warren, G.J. Ackland, J. Crain, Am. Mineral. 82 (1997) 51 https://doi.org/10.2138/am-1997-1-207.
[30] R.M. Wentzcovitch, N.L. Ross, G.D. Price, Phys. Earth Planet. Inter. 90 (1995) 101 https://doi.org/10.1016/0031-9201(94)03001-Y.
[31] A. Bouhemadou, R. Khenata, M. Kharoubi, T. Seddik, Ali H.Reshak, Y. Al-Douri, Comput. Mater. Sci. 45 (2009) 474–479 https://doi.org/10.1016/j.commatsci.
2008.11.013.
[32] C.M.O. Bastos, F.P. Sabino, G.M. Sipahi, J.L.F. Da Silva, J. Appl. Phys. 123 (2018) 065702https://doi.org/10.1063/1.5018325.
[33] M.J. Herrera-Cabrera, P. Rodríguez-Hernández, A. Munoz, Phys. Stat. Sol. (b) 223 (2001) 411 https://doi.org/10.1002/1521-3951(200101)223:2%3C411::AID-
PSSB411%3E3.0.CO;2-H.
[34] S.Q. Wang, H.Q. Ye, Phys. Stat. Sol. (b) 240 (1) (2003) 45–54.
[35] M. van Schilfgaarde, A. Sher, A.B. Chen, J. Cryst. Growth 178 (1997) 8 https://doi.org/10.1016/S0022-0248(97)00073-0.
[36] M.E. Sherwin, T.J. Drummond, J. Appl. Phys. 69 (1991) 8423 https://doi.org/10.1063/1.347412.
[37] H.J. Hou, S.F. Zhu, B.J. Zhao, Y. Yu, S. Ru Zhang, L.H. Xie, Phys. B. Condensed Matter 407 (3) (2012) 408–411 https://doi.org/10.1016/j.physb.2011.11.007.
[38] M. FlÓrez, J.M. Recio, E. Francisco, M.A. Blanco, A.M. Pendás, Phys. Rev. B 66 (2002) 144112https://doi.org/10.1103/PhysRevB.66.144112.
[39] K. Kim, W.R.L. Lambrecht, B. Segall, Phys. Rev. B 53 (1996) 16310 https://journals.aps.org/prb/abstract/10.1103/Phys.Rev.B.56.7018.2.
[40] V. Yu. Davydov, A.A. Klochikhin, V.V. Emtsev, D.A. Kurdyukov, S.V. Ivanov, V.A. Vekshin, F. Bechstedt, J. Furthmüller, J. Aderhold, J. Graul, A.V. Mudryi,
H. Harima, A. Hashimoto, A. Yamamoto, E.E. Haller, Phys. Stat. Sol. b 234 (2002) 787 https://doi.org/10.1002/1521-3951(200212)234:3%3C787::AID-
PSSB787%3E3.0.CO;2-H.
[41] A.F. Wright, J. Appl. Phys. 82 (1997) 2833 https://doi.org/10.1063/1.366114.
[42] K. Kim, W.R.L. Lambrecht, B. Segall, Phys. Rev. B 50 (1994) 1502 https://doi.org/10.1103/PhysRevB.50.1502.
[43] D.N. Talwar, D. Sofranko, C. Mooney, S. Tallo, Mater. Sci. Eng. B 90 (2002) 269–277 https://doi.org/10.1016/S0921-5107(01)00939-4.
[44] M.B. Kanoun, A.E. Merad, G. Merad, J. Cibert, H. Aourag, Solid State Electron. 48 (9) (2004) 1603 https://doi.org/10.1016/j.sse.2004.03.007.
[45] M.L. Cohen, Phys. Rev. B 32 (1985) 7988 https://doi.org/10.1103/Phys.Rev.B.32.7988.
[46] Landolt-Börstein, Semiconductors: Physics of Group IV Elements and III–V Compounds, III/17a, Springer-Verlag, Berlin, 1992.
[47] P. Pluengphon, T. Bovornratanaraks, U. Pinsook, J. Alloys Compd. 700 (2017) 98–105 https://doi.org/10.1016/j.jallcom.2017.01.063.
[48] R. Moussa, A. Abdiche, R. Khenata, X.T. Wang, D. Varshney, X.W. Sun, S.B. Omran, A. Bouhemadou, D.P. Rai, J. Phys. Chem. Solid (2018) 36–49 https://doi.org/
10.1016/j.jpcs.2018.03.035.

463
I. Hattabi, et al. Chinese Journal of Physics 59 (2019) 449–464

[49] F. Gao, J. He, E. Wu, S. Liu, D. Yu, D. Li, S. Zhang and Y. Tian. Phys. Rev. Lett.91, 015502. https://doi.org/10.1103/PhysRevLett.91.015502.
[50] A.R. Degheidy, E.B. Elkenany, Chin. Phys. B 26 (8) (2017) 086103http://iopscience.iop.org/article/10.1088/1674-1056/26/8/086103/meta.
[51] M. Sarwan, S. Singh, J. Alloys Compd. 550 (2013) 150 https://doi.org/10.1016/j.jallcom.2012.09.097.
[52] M. Zhang, M. Wang, T. Cui, Y. Ma, Y. Niu, G. Zou, J. Phys. Chem. Solid 69 (2008) 2096–2102 https://doi.org/10.1016/j.jpcs.2008.03.008.
[53] F. Peng, D. Chen, H. Fu, X. Cheng, Phys. Status Solid B 246 (2009) 71 https://doi.org/10.1002/pssb.200844332.
[54] edited by, S. Nakamura, S.F. Chichibu (Eds.), Introduction to Nitride Semiconductor Blue Lasers and Light Emitting Diodes, CRC Press, 2000edited by.
[55] S.F. Pugh, Philos. Mag. 45 (1954) 823 https://doi.org/10.1080/14786440808520496.
[56] S. Adachi, Properties of Group III -V and II-VI Semiconductors, Wiley, New York, 2005 (Chapter 2).
[57] M.A. Blanco, E. Francisco, V. Luaña, “GIBBS: isothermal-isobaric thermodynamics of solids from energy curves, Comp. Phys. Commun. 158 (1) (2004) 57–72
https://doi.org/10.1016/j.comphy.2003.12.001.
[58] A. Otero-de-la-Roza, V. Luaña, Comput. Phys. Commun. 182 (2011) 2232–2248 https://doi.org/10.1016/j.cpc.2011.04.016.
[59] A.T. Petit, P.L. Dulong, Ann. Chim. Phys. 10 (1819) 395.
[60] O. Nemiri, S. Ghemid, Z. Chouahda, H. Meradji, F. El Haj Hassan, Int. J. Mod. Phys. B 27 (2013) 12 https://doi.org/10.1142/S021797921350166X.
[61] edited by, W. Martienssen, H. Warlimont (Eds.), Springer Handbook of Condensed Matter and Materials Data, Springer Nature, 2005edited by.
[62] R.M. Martin, Phys. Rev. B 1 (1970) 4005 https://doi.org/10.1103/PhysRevB.1.4005.

464

View publication stats

You might also like