Home Equation of State, Nonlinear Elastic Response, and Anharmonic Properties of Diamond-Cubic Silicon and Germanium: First-Principles Investigation
Article Publicly Available

Equation of State, Nonlinear Elastic Response, and Anharmonic Properties of Diamond-Cubic Silicon and Germanium: First-Principles Investigation

  • Chenju Wang , Jianbing Gu , Xiaoyu Kuang EMAIL logo and Shikai Xiang
Published/Copyright: May 5, 2015

Abstract

Nonlinear elastic properties of diamond-cubic silicon and germanium have not been investigated sufficiently to date. Knowledge of these properties not only can help us to understand nonlinear mechanical effects but also can assist us to have an insight into the related anharmonic properties, so we investigate the nonlinear elastic behaviour of single silicon and germanium by calculating their second- and third-order elastic constants. All the results of the elastic constants show good agreement with the available experimental data and other theoretical calculations. Such a phenomenon indicates that the present values of the elastic constants are accurate and can be used to further study the related anharmonic properties. Subsequently, the anharmonic properties such as the pressure derivatives of the second-order elastic constants, Grüneisen constants of long-wavelength acoustic modes, and ultrasonic nonlinear parameters are explored. All the anharmonic properties of silicon calculated in the present work also show good agreement with the existing experimental results; this consistency not only reveals that the calculation method of the anharmonic properties is feasible but also illuminates that the anharmonic properties obtained in the present work are reliable. For the anharmonic properties of germanium, since there are no experimental result and other theoretical data till now, we hope that the anharmonic properties of germanium first offered in this work would serve as a reference for future studies.

1 Introduction

As most important technological materials in fundamental science and practical applications, diamond-cubic silicon (Si) and germanium (Ge) have attracted growing theoretical and experimental attention. However, most of the studies focus on their phase transition at high pressure; only a few papers explore their nonlinear elastic response which is a natural reflection when a material undergoes a finite deformation. Moreover, knowledge of the nonlinear elastic response not only can help us to understand the nonlinear mechanical effects when large stress and strain are involved, but also can assist us to have an insight into the related anharmonic properties such as the Grüneisen parameter, ultrasonic nonlinear parameter, thermal expansion, and phonon–phonon interaction [1]. Therefore, we investigate the nonlinear elastic response of Si and Ge and try to shed some light on their related anharmonic properties.

To date, a few theoretical works [2–4] have been performed on the nonlinear elastic response and anharmonic properties of Si and Ge. In the year 1966, Keating [2] first investigated the nonlinear elastic properties of Si and Ge by calculating their second- and third-order elastic constants which were defined as the second and third derivatives of the total energy with respect to the Lagrangian strain, whereas he did not refer the related anharmonic properties. Similarly, Zhao et al. [3] also merely studied the nonlinear elastic properties of Si by the method of homogeneous deformation based on the first-principles total-energy calculations [5, 6]. Employing the method provided in [3], Lopuszynski et al. [4] calculated the third-order elastic constants of Si and obtained its related anharmonic properties such as the pressure derivatives of the second-order elastic constants and Gruneisen parameters. However, they did not mention the ultrasonic nonlinear parameters [7] which are indispensable quantities in the second-harmonic generation experiments. Moreover, no work on the nonlinear elastic response and anharmonic properties of Ge has been carried out by using the homogeneous deformation method which seems to be the most promising approach to handle such tasks at present. In view of the facts mentioned above, we investigate the nonlinear elastic response and related anharmonic properties of Si and Ge by using the method of the homogeneous deformation in the present work.

The rest of the paper is organized as follows. In Section 2, the nonlinear elastic theory and computational details are described briefly. Subsequently, the calculated results and discussion are given in Section 3. In Section 4, we present the applications of the third-order elastic constants. Finally, the conclusions are drawn in Section 5.

2 Nonlinear Elasticity Theory and Computational Details

2.1 Nonlinear Elasticity Theory

Nonlinear elasticity theory is the theory of elasticity under finite deformation. Suppose the underformed state X moves to the deformed state x when a solid body undergoes a finite deformation. We denote the undeformed coordinates of a point in configuration X by (X1, X2, X3) and represent that in configuration x by (x1, x2, x3). Then, the deformation gradient matrix can be defined as

(1)αij=xi/Xj, (1)

where the subscripts i and j vary from 1 to 3. The Lagrangian strain ηij can be expressed as [8]

(2)ηij=12(kαkiαkjδij), (2)

where δij is the Kronecker delta (unity when i=j, zero otherwise). And the free energy F and internal energy U per unit mass can be expanded respectively as Taylor series in terms of the Lagrangian strain [9–11]

(3a)F(X,ηij,T)=F(X,  0, T)+V(X)ijTijT(X)ηij+12!V(X)ijklCijklT(X)ηijηkl+13!V(X)ijklmnCijklmnT(X)ηijηklηmn+ (3a)
(3b)U(X,ηij,S)=U(X, 0, S)+V(X)ijTijS(X)ηij+12!V(X)ijklCijklS(X)ηijηkl+13!V(X)ijklmnCijklmnS(X)ηijηklηmn+. (3b)

Then the isothermal and adiabatic elastic constants at state X can be defined as [12]

(4a)CijklT(X)=1V(X)2Fηijηkl|XηmnT,CijklmnT(X)=1V(X)2Fηijηklηmn|XηpqT, (4a)

and

(4b)CijklS(X)=1V(X)2Uηijηkl|XηmnS,CijklmnS(X)=1V(X)2Uηijηklηmn|XηpqS, (4b)

where V(X) is the volume of the system at state X, and the subscripts ηmn denotes that the elements of the strain tensor other than those involved in the partial derivative are held constant. T and S represent the isothermal and adiabatic processes, respectively. Tij(X) appeared in formulas (3a) and (3b) is the component of the stress tensor applied on the state X, which can be defined as

(5)TijT(X)=1V(X)Fηij|XηmnT,TijS(X)=1V(X)Uηij|XηmnS. (5)

Specifically, they can be expressed as

(6a)TijS=TijT=Pδijhydrostaticpressure, (6a)
(6b)TijS=TijT=Ptitjuniaxialstressalongt, (6b)

where t is a unit vector. Here, we do not distinguish the two types of elastic constants, since the first-principles calculations are performed at 0 K, i.e., F=UTS=U and CS=CT. So is Tij(X).

If we introduce the density ρ0 of the unstrained crystal and apply the Voigt convention (η11η1, η22η2, η33η3, η23η4, η13η5, and η12η6), (3a) can be simplified as

(7)ρ0F(X,ηij)=ρ0F(X,0)+iTi(X)ηi+12!ijCij(X)ηiηj+13!ijkCijk(X)ηiηjηk+. (7)

For cubic crystals, there are three independent second-order elastic constants C11, C12, and C44 and six independent third-order elastic constants C111, C112, C123, C144, C166, and C456. Thus, (7) can be further written as

(8)ρ0[F(X,ηij)ρ0F(X,0)]=iTi(X)ηi+φ2+φ3, (8)

where

(9)φ2=12C11(η12+η22+η32)+C12(η1η2+η2η3+η1η3)+12C44(η42+η52+η62) (9)

and

(10)φ3=16C111(η13+η23+η33)+12C112(η12η2+η12η3+η22η1+η22η3+η32η1+η32η2)+C123η1η2η3+12C144(η42η1+η52η2+η62η3)+12C166(η42η2+η42η3+η52η1+η52η3+η62η1+η62η2)+C456η4η5η6. (10)

Using (8)–(10), we can obtain the elastic constants by applying a series of specific homogeneous deformations on a crystal and calculating the total energies of the undeformed and deformed states, respectively. In the following, the special deformation types and calculation details of the total energies are presented.

2.2 Deformation Types

Since the single Si and Ge at 0 K and 0 GPa belong to the cubic structure, and there are six independent third-order elastic constants, six special deformation types at the least are needed to calculate their complete set of the third-order elastic constants. Moreover, the six deformation types should be as simple as possible to reduce the number of the strain components that will appear in (9) and (10). Following this principle, we select the six deformation types [3] ηA, ηB, ηC, ηD, ηE, and ηF shown in Table 1. Then, we obtain the corresponding energy expressions fα by inserting the six different deformation strains into (8)–(10),

Table 1

Strain types used for calculating the second- and third-order elastic constants; their corresponding energy expressions are fα(fα=fA, fB, fC, fD, fE, and fF) as shown in (11).

Strain typesP2P3
ηA=(η, 0, 0, 0, 0, 0)12C1116C111
ηB=(η, η, 0, 0, 0, 0)C11 + C1213C111+C112
ηC=(η, η, η, 0, 0, 0)32C11+3C1212C111+3C112+C123
ηD=(η, 0, 0, 2η, 0, 0)12C11+2C4416C111+2C144
ηE=(η, 0, 0, 0, 2η, 0)12C11+2C4416C111+2C166
ηF=(0, 0, 0, 2η, 2η, 2η)6C448C456

The coefficients P2 and P3 are the linear combinations of the second- and third-order elastic constants.

(11)fα=ρ0[F(X,η)ρ0F(X,0)]=P2η2+P3η3+o(η4), (11)

where fα=fA, fB, fC, fD, fE, and fF; the coefficients P2 and P3 are related to the second- and third-order elastic constants; in practice, the strain magnitude η is set from −|ηmax| to +|ηmax| in steps of 0.0005. Here, it should be pointed out that the term ijTij(X)ηij does not appear in (11), for Tij(X) representing that the stress applied on the reference state is zero at 0 GPa.

For each deformation type listed in Table 1, the deformed state can be obtained by using the equation of ri=αijrj, where ri and rj represent respectively the deformed and undeformed lattice vectors; αij can be obtained by the following formula:

(12)αij=δij+ηij12kηkiηkj+12rkηrkηriηkj58kmnηkjηmkηmnηni+. (12)

In general, for a given η, αij is not unique, but this is not a problem since the Lagrange strain brings the rotational invariance of total energy [9]. And for a system without rotation, (12) will provide a unique relation.

2.3 Computational Details

2.3.1 Density Functional Theory Calculations

Within the framework of the density functional theory (DFT), the calculations of the total energies of Si and Ge are performed using the Vienna ab initio simulation package (VASP) developed by Hafner Research Group at the University of Vienna [13–15]. The projector augmented wave method [16] is employed in its variant available in the VASP [17]. To choose the appropriate exchange-correlation functional, we calculate the lattice parameters, equilibrium volume, bulk modulus B0, and its pressure derivation B0 using the local-density approximation (LDA) and generalised gradient approximation (GGA), respectively, and compare them with other theoretical results [18–25] and available experimental data [26–32] shown in Table 2. From Table 2, we can clearly see that results calculated by LDA show better agreement with the experimental data [26–32] on the whole than those obtained through the GGA. Hence, the LDA is used in the calculations of the present work. In addition, careful convergence tests (tolerance for the total-energy difference in the self-consistent-field calculation is 1.0×10−9 eV and tolerance for the maximum force in the geometry optimization is 1.0×10−5 eV/Å) are performed to determine the cutoff energy and the number of k-points. Finally, the reasonable cutoff energy (650 eV for Si and 600 eV for Ge) and rational k-points (15×15×15 for Si and 17×17×17 for Ge) are ascertained.

Table 2

Lattice parameters, unit-cell volume, bulk modulus B0, and its pressure derivation B0 of the diamond-cubic Si and Ge, together with the experimental results and other theoretical data.

a (Å)V03/atom)B0 (GPa)B0
Si – diamond-cubic phase
 LDA (present work)5.4119.7596.344.083
 GGA (present work)5.4720.5187.294.06
 Theoretical references5.45a, 5.40b19.57c, 19.69d98.00a, 93.00b, 97.00c, 95.25d5.11d
 Experimental references5.43i20.01j99.20j3.80–4.24k
Ge – diamond-cubic phase
 LDA (present work)5.6322.2674.884.82
 GGA (present work)5.7623.9158.734.88
 Theoretical references5.59b, 5.76e21.84b, 21.70f, 23.88g72.00b, 76.00e, 80.09f, 58.40g, 71.00–78.00h5.19f, 4.75g, 3.70–3.90h
 Experimental references5.66e, 5.65i22.55i77.00e, 76.80j, 74.90l, 67.40m, 75.00n, 77 ± 4.00o4.70l, 5.00m, 4.55n, 4.30 ± 1.00o

aReference [18], bReference [19], cReference [20], dReference [21], eReference [22], fReference [23], gReference [24], hReference [25], iReference [26], jReference [27], kReference [28], lReference [29], mReference [30], nReference [31], oReference [32].

2.3.2 Determination of the Third-Order Elastic Constants

Since the third-order elastic constants are very sensitive to the k-points, the maximum strain magnitude, and the order of the polynomial used to fit (11), we take Si as an example to test the dependence of the third-order elastic constants on the three parameters. First of all, we investigate the dependence of the third-order elastic constants on the k-points and display it in Figure 1. As shown in Figure 1, all the third-order elastic constants of Si converge after the k-point mesh size reaches 15×15×15. And the relative difference between two successive values of an examined constant is <1.0 %, which is exactitude enough for the calculations of the third-order elastic constants. Such a test is also performed for Ge, and results show that EcutoffGe=650 eV and 17×17×17 are sufficient to calculate the third-order elastic constants. Subsequently, the dependence of the third-order elastic constants on the maximum strain magnitude |ηmax| is tested and illustrated in Figure 2. From Figure 2, it can be clearly found that the third-order elastic constants converge well after the maximum strain parameters |ηmax| up to 0.009. Therefore, |ηmax|=0.009 is employed to calculate the third-order elastic constants of Si. Similarly, |ηmax|=0.015 is determined for Ge. Based on the parameters determined above, we obtain the relations of η and fα(η) (α=A, B, C, D, E, and F). Then the remaining step is to fit these relationships using proper polynomials. For ascertaining the suitable order of the polynomials, the residuals obtained respectively from the third- and fourth-order polynomial fits are compared in Figure 3a–f. Figure 3a–f show that the residuals obtained from the fourth-order polynomial fit are clearly smaller than those obtained from the third-order polynomial fit. Thus, the fourth-order polynomial is employed to fit the relations of fα(η) and η. Then, we present the fα(η) ∼ η data of Si in Figure 4 and fit them using the fourth-order polynomial. From Figure 4, the second- and third-order elastic constants of Si are obtained. In the same way, the elastic constants of the Ge are obtained.

Figure 1: Dependence of the third-order elastic constants C111 and C112 (a), C123 and C144 (b), and C166 and C456 (c) on the k-points (energy cutoff of 650 eV is applied for all points).
Figure 1:

Dependence of the third-order elastic constants C111 and C112 (a), C123 and C144 (b), and C166 and C456 (c) on the k-points (energy cutoff of 650 eV is applied for all points).

Figure 2: Convergence tests of the third-order elastic constants of Si as a function of the maximum strain |ηmax|.
Figure 2:

Convergence tests of the third-order elastic constants of Si as a function of the maximum strain |ηmax|.

Figure 3: Comparisons of the residuals obtained respectively from the third- and fourth-order polynomial fit. (a), (b), (c), (d), (e), and (f) represent respectively the relations between the residual of fA, fB, fC, fD, fE, fF and η.
Figure 3:

Comparisons of the residuals obtained respectively from the third- and fourth-order polynomial fit. (a), (b), (c), (d), (e), and (f) represent respectively the relations between the residual of fA, fB, fC, fD, fE, fF and η.

Figure 4: The strain energy density as a function of the Lagrangian strain of Si, where the discrete points and the solid lines represent respectively the results obtained from the first-principles calculations and the nonlinear theory; fA, fB, fC, fD, fE, and fF denote different energy expressions corresponding to various deformation modes.
Figure 4:

The strain energy density as a function of the Lagrangian strain of Si, where the discrete points and the solid lines represent respectively the results obtained from the first-principles calculations and the nonlinear theory; fA, fB, fC, fD, fE, and fF denote different energy expressions corresponding to various deformation modes.

3 Results and Discussion

3.1 Equation of State

Within the framework of the LDA, we calculate a series of total energies of Si and Ge in both diamond-cubic and β-tin structures and fit them using the fourth-order Brich–Murnaghan equation [33]

(13)E(V)=n=14anV2n/3. (13)

Then, the equation of states of Si and Ge are obtained and displayed in Figure 5. It is known that the diamond-cubic structure will coexist with the β-tin phase at the pressure where the Gibbs free energies, G=E + PVTS, of the two structures are equal. At zero temperature, the Gibbs free energy is equal to the enthalpy H=E + PV, where the pressure is given by the identity P=−∂E/∂V. That is, the phase transition pressures from the diamond-cubic structure to the β-tin phase can be determined by using the enthalpy method. Therefore, we display their enthalpy difference versus pressure in Figure 5.

Figure 5: Energy versus primitive volume for Si and Ge in the diamond-cubic and β-tin structures. Inset: enthalpy differences ΔH of the β-tin structures relative to diamond phases as a function of the pressure.
Figure 5:

Energy versus primitive volume for Si and Ge in the diamond-cubic and β-tin structures. Inset: enthalpy differences ΔH of the β-tin structures relative to diamond phases as a function of the pressure.

Figure 5 shows that the phase transition (diamond cubic → β-tin) of Si occurs at about 7.2 GPa, which is exactly between the previous theoretical data 7.0 [34] and 8.4 GPa [20, 35, 36], but less than the available experimental values 11.7 GPa [37] and 10.0–12.5 GPa [38, 39]. For Ge, the corresponding phase transformation is near 9.7 GPa, which is in general agreement with the previous theoretical results, 9 GPa [40], and also lower than the existing experimental results, 10.5 ± 0.2 [41] and 8.1 ± 0.3–10.6 ± 0.5 GPa [29]. Here, it is worth pointing out that such underestimation of the phase transition pressure is typical in the LDA calculations.

3.2 Results of the Third-Order Elastic Constants

The second- and third-order elastic constants of the diamond-cubic Si and Ge are listed in Table 3 and compared with other theoretical data [3, 4, 19, 42] and the available experimental results [27, 43–48]. For the second-order elastic constants, it is obvious that the present data for both Si and Ge are well consistent with the existing experimental results [27, 43, 45, 46] and foregoing theoretical values [3, 4, 19]. The differences do not exceed the typical error (10 %), which is considered the typical error in elasticity calculations based on the DFT. Such agreement suggests that the elasticity calculations are performed at high accuracy and the elastic constants obtained in the present work are reliable. Here, it should be pointed out that some elastic constants can be determined by several fits, while the obtained values are slightly different (e.g., for Si, C44=77.06, 77.33, and 78.20 from fD, fE, and fF, respectively). In such cases, the average of these data is given in Table 3. In addition, Table 3 clearly shows that the second-order elastic constants are in better agreement with the available experimental results [27, 43, 45, 46] than those of other theoretical values [3, 4, 19].

Table 3

The second- and third-order elastic constants (GPa) of the diamond-cubic Si and Ge, in comparison with the experimental and previous calculations.

Present workTheoretical referencesExperimental references
Si – diamond cubic phase
C11161.86153a, 162.07b, 159c165.64e, 165.77f, 167g
C1263.5865a, 63.51b, 61c63.94e, 63.92f, 65g
C4477.5373a, 77.26b, 85c79.51e, 79.62f, 80g
C111−841−698a, −810b, −750c, 816d,−795 ± 10e, −825 ± 10f, −880h, −834i
C112−489−451a, −422b, −480c, −446d−445 ± 10e, −451 ± 5f, −515h, −531i
C12330−112a, −61b, 0c, −79d−75 ± 5e, −64 ± 10f, 27h, −2i
C14435−74a, 31b, −14d15 ± 5e, 12 ± 25f, 74h, −95i
C166−287−253a, −293b, −344d−310 ± 5e, −310 ± 10f, −385h, −296i
C456−58−57a, −61b, −80c, −76d−86 ± 5e, −64 ± 20f, −40h, ± 7i
Ge – diamond cubic phase
C11129.86130c128.35j, 128.53k
C1247.3945c48.23j, 48.26k
C4465.7363c66.66j, 66.80k
C111−708−716 ± 20j, −710 ± 6k, −696 ± 108l, −732 ± 50m
C112−346−403 ± 10j, −389 ± 3k, −340 ± 62l, −290 ± 30l
C123−26−18 ± 30j, −18 ± 6k, 25 ± 43l, −22 ± 20m
C144−10−53 ± 5j, −23 ± 16k, 18−21l, −8.3 ± 9m
C166−279−315 ± 5j, −292 ± 8k, −296 ± 22l, −303 ± 9m
C456−40−47 ± 10j, −53 ± 7k, −42 ± 6l, −41 ± 5m

aReference [3], bReference [4], cReference [19], dReference [42], eReference [43], fReference [27] (T=300 K), gReference [27] (T=73 K), hReference [44] (T=4 K), iReference [44] (T=298 K), jReference [45], kReference [46], lReference [47], mReference [48].

As regards the third-order elastic constants of Si and Ge, overall agreement of our calculated data and available experimental [27, 43–48] and other theoretical results [3, 4, 19, 42] can be seen from Table 3. The discrepancies between our calculated values and experimental results may be derived from the following two aspects. First, the experimental measurements of the third-order elastic constants are still difficult to date; i.e., the reported experimental results [27, 43–48] are determined with considerable uncertainties and even exhibit a difference between different groups (e.g., [43] and [44] shown in Table 3). Second, theoretical results of the third-order elastic constants are obtained at 0 K, while the experimental values are often determined in conditions which are far from the ideal case of 0 K. Particularly, for the cases of Si and Ge, the temperature effects on the third-order elastic constants can be identified and should not be ignored (see [31, 44, 49]). Moreover, the distinctions between the values obtained in the present work and other theoretical data, in our opinion, are understandable, since the theoretical results obtained at different times by different numerical procedures often have different and unknown accuracies.

4 Applications

4.1 Ascertainment of the Pressure Derivatives of the Second-Order Elastic Constants

For describing the nonlinear elastic properties of a material under large hydrostatic pressure, we need to introduce the effective elastic constants Cij(P), which can be obtained by the following formula [10]:

(14)Cij(P)Cij+CijP+, (14)

where Cij denotes the pressure derivative of Cij. If Cij is known, the effective elastic constants Cij(P) of a material under high pressure can be calculated readily through (14). For cubic crystals, the pressure derivatives Cij can be expressed as [10]

(15)C11=2C112+C111+2C12+2C112C12+C11C12=C123+2C112C12C112C12+C11C44=2C166+C144+C44+2C12+C112C12+C11. (15)

Using (15) and the results of the second- and third-order elastic constants, we calculate the pressure derivatives Cij of Si and Ge. To compare them with other theoretical data [3, 50] and experimental results [49, 51], we list them in Table 4.

Table 4

Pressure derivatives of the second-order elastic constants, together with the experimental values.

Present resultsTheoretical valuesExperimental data
Si – diamond cubic phase
C114.734.09a, 4.0b4.19c, 4.33d, 4.29e
C124.064.34a, 3.0b4.02c, 4.19d, 4.20e
C440.600.27a, 1.1b0.80c, 0.8d, 0.75e
Ge – diamond cubic phase
C114.65
C123.99
C441.24

aReference [3], bReference [50], cReference [51], dReference [49] (T=300 K), eReference [49] (T=79K).

With regard to the diamond-cubic Si, Table 4 shows that our calculated pressure derivatives are in reasonable agreement with the existing experimental results [49, 51]. The discrepancies between our data and experimental values respectively are 9.2 %–12.9 % for C11, 1.0 %–3.3 % for C12, and 20 %–25 % for C44. Though the difference between the present result of C44 and the experimental data is large, it is much smaller than the discrepancies (64 %–66 % for [3] and 37.5 %–46.7 % for [50]) between the experimental data and the values offered in [3, 50]. We guess that the difference between theoretical data and experimental values may be originated from the temperature effects which have a large influence on the elastic constants of Si. For the pressure derivatives of the diamond-cubic Ge, however, there are no experimental and other theoretical values to date unfortunately. So it is hoped that the results obtained in this work would serve as a reference for future studies.

4.2 Calculations of the Grüneisen Constants of Long-Wavelength Acoustic Modes

The Grüneisen constant is a most important parameter for describing the anharmonic behaviour of solids such as thermal expansivity, shock deformation, and ultrasonic attenuation. In the quasiharmonic approximation, the generalized Grüneisen constants can be defined as [52]

(16)γ(q,j)=lnω(q,j)lnV=Vω(q,j)ω(q,j)V, (16)

where V is the crystal volume, q denotes the propagation direction, j represents the polarization vector, and ω stands for the frequency of the ith mode. In addition, Brugger defined the strain Grüneisen tensor as [53]

(17)γαβ(q,j)=1ω(q,j)ω(q,j)ηαβ(α,β=x,y,z), (17)

where ηαβ is the Lagrangian strain. For a cubic crystal, the generalized Grüneisen constant is equal to one third of the trace of the strain Grüneisen tensor. Namely,

(18)γ(q,j)=13[γxx(q,j)+γyy(q,j)+γzz(q,j)]. (18)

According to the motion equation [9] of a long-wavelength elastic wave, the element of the strain Grüneisen tensor for a given q can be written as

(19)γαβ(q,jj)=14ρ0vqjvqjμνςξ(Sμν|ςξ|αβ+Sμν|ςξ|βα)×qνqξeμ(qj)eς(qj), (19)

where ρ0 is the density of the undeformed crystal, v represents the sound velocity, e(q, j) denotes the polarization vector of branch j and propagation direction q, and S is the combination of the second- and third-order elastic constants

(20)Sαμ|βν|ςξ=Cαμ|βν|ςξ+δαβCμν|ςξ+δαςCβν|μξ+δβςCαμ|νξ. (20)

For a specific q and j, γ(q, j) can be expressed in terms of the second- and third-order elastic constants (see the appendix of [52]) by combining (18)–(20) and a series of equations listed in Tables 13 of [8]. Using these equations displayed in the appendix of [52] and our results of the elastic constants, we compute the thermodynamic Grüneisen constants of Si and Ge and list them in Table 5, together with other theoretical data [3, 52] and available experimental results [44, 46]. With regard to Si, the comparisons show that our calculated results are in better agreement with the experimental values [44, 46] than those of other theoretical data [3, 52]. This indicates that our results of the elastic constants are reliable and can be used to further calculate the ultrasonic nonlinear parameters. In addition, this phenomenon also illuminates that the Grüneisen constants are very sensitive to the values of the elastic constants, since our elastic constants are also in better agreement with the experimental data than that of the other theoretical calculations. To the best of our knowledge, there are no experimental and theoretical data on the Grüneisen constants of Ge till now. Hence, the Grüneisen constants of Ge calculated in this work may be severed as a reference for future studies.

Table 5

Grüneisen constants of the long-wavelength acoustic modes and the nonlinear parameters of Si and Ge, together with the experimental results for Si.

SiGe
This workTheoretical referencesExperimental referencesThis work
q(ε, 0, 0)
γ (LA)1.2421.098a, 0.983b1.108c1.175
γ (TA)0.2040.006a, 0.456b0.324c0.537
β12.1962.034d2.452
q(ε, ε, 0)
γ (LA)1.0980.999a, 1.063b1.109c1.181
γ(TAxy̅)0.163−0.301a, −0.153b−0.049c0.137
γ(TAz)0.2040.006a, 0.456b0.324c0.537
β24.5584.689d5.250
q(ε, ε, ε)
γ (LA)1.0590.973a, 1.084b1.081c1.1824
β33.1823.750d4.036

aReference [3], bReference [52], cReference [46], dReference [44] (300 K).

4.3 Determination of the Ultrasonic Nonlinear Parameters

The ultrasonic nonlinear parameters, obtained by the formula [9]

(21)β(q,jj)=1ρvqj2αβμνςξSαμ|βν|ςξqμqνqξ×eα(q,j)eβ(q,j)eς(q,j), (21)

are the indispensable quantities in the second-harmonic generation experiments. Combining with (19), we obtain the general relation between the ultrasonic nonlinear parameters and the Grüneisen constants, which is written as

(22)β(q,jj)=2αβγαβ(q,jj)qαeβ(qj). (22)

For the longitudinal (L) modes along the main-symmetry direction of cubic crystals, (22) can be reduced as the following three formulas:

(23)β([1,0,0],LL)=2γxx([ε,0,0],L),β(12[1,1,0],LL)=2{γxx([ε,ε,0],L)+γxy([ε,ε,0],L)}andβ(13[1,1,1],LL)=2{γxx([ε,ε,ε],L)+2γxy([ε,ε,ε],L)}. (23)

Thus, the three ultrasonic nonlinear parameters can be explicitly expressed in terms of the second- and third-order elastic constants by using the equations given in the appendix of [52]. Then we calculate the three ultrasonic nonlinear parameters of Si and Ge by using the results of the second- and third-order elastic constants and list them in Table 5. As displayed in Table 5, three ultrasonic nonlinear parameters of Si are in reasonable agreement with the existing experimental results [44] measured at 300 K, and the discrepancies for β1, β2, and β3, respectively, are 7.9 %, 2.8 %, and 15.1 %. It is a pity that there are no experimental and theoretical results on the ultrasonic nonlinear parameters of the Ge up to now. Actually, the comparison between our results and experimental data as well as other theoretical values is not our aim. Instead, our aim is to illuminate some applications of the third-order elastic constants and introduce the way of calculating the ultrasonic nonlinear parameters on the basis of the elastic constants.

5 Conclusions

In the present work, the nonlinear elastic response of the diamond-cubic Si and Ge is investigated by calculating their third-order elastic constants. Our theoretical results of the elastic constants for both Si and Ge are in excellent agreement with previous theoretical results and available experimental data. Based on the results of the elastic constants, we compute the pressure derivatives of the second-order elastic constants, the mode Grüneisen constants of long-wavelength acoustic modes, and the ultrasonic nonlinear parameters generated in the second-harmonic generation experiments. All the results of the diamond-cubic Si show good agreement with the existing experimental values and other theoretical calculations. Unfortunately, there are no related experimental and theoretical data on the anharmonic properties of Ge at present. Hence, our calculated values may serve as a reference for future related studies.


Corresponding author: Xiaoyu Kuang, Institute of Atomic and Molecular Physics, Sichuan University, Chengdu 610065, China, E-mail:

Acknowledgments

This work was supported by the Industrial Technology Development Program (grant no B1520110001), the National Natural Science Foundation of China [grant nos 10904133, 11304294, and U1230201 (NSAF)], and the Development Foundation of CAEP (grant no 2013B0401062).

References

[1] Y. Hiki, Annu. Rev. Mater. Sci. 11, 51 (1981).10.1146/annurev.ms.11.080181.000411Search in Google Scholar

[2] P. N. Keating, Phys. Rev. 149, 674 (1966).10.1103/PhysRev.149.674Search in Google Scholar

[3] J. J. Zhao, J. Winey, and Y. Gupta, Phys. Rev. B 75, 094105 (2007).10.1103/PhysRevB.75.094105Search in Google Scholar

[4] M. Lopuszynski and J. Majewski, Phys. Rev. B 76, 045202 (2007).10.1103/PhysRevB.76.045202Search in Google Scholar

[5] M. Born and K. Huang, Dynamical Theory of Crystal Lattices, Oxford University Press, London 1956.Search in Google Scholar

[6] R. Srinivasan, Phys. Rev. 144, 620 (1966).10.1103/PhysRev.144.620Search in Google Scholar

[7] J. Cantrell, Phys. Rev. B 21, 4191 (1980).10.1103/PhysRevB.21.4191Search in Google Scholar

[8] R. N. Thurston and K. Brugger, Phys. Rev. 133, A1604 (1964).10.1103/PhysRev.133.A1604Search in Google Scholar

[9] D. C. Wallace, Thermodynamics of Crystals, Wiley, New York 1972.10.1119/1.1987046Search in Google Scholar

[10] F. Birch, Phys. Rev. 71, 809 (1947).10.1103/PhysRev.71.809Search in Google Scholar

[11] F. Murnaghan, Finite Deformation of an Elastic Solid, Wiley, New York 1951.Search in Google Scholar

[12] K. Brugger, J. Appl. Phys. 36, 768 (1965).10.1063/1.1714215Search in Google Scholar

[13] G. Kresse and J. Hafner, Phys. Rev. B 48, 3115 (1993).10.1103/PhysRevB.48.13115Search in Google Scholar

[14] G. Kresse and J. Furthmuller, Comput. Mater. Sci. 6, 15 (1996).10.1016/0927-0256(96)00008-0Search in Google Scholar

[15] G. Kresse and J. Furthmuller, Phys. Rev. B 54, 11169 (1996).10.1103/PhysRevB.54.11169Search in Google Scholar PubMed

[16] P. E. Blochl, Phys. Rev. B 50, 17953 (1994).10.1103/PhysRevB.50.17953Search in Google Scholar

[17] G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).10.1103/PhysRevB.59.1758Search in Google Scholar

[18] M. T. Yin and M. L. Cohen, Phys. Rev. Lett. 45, 1004 (1980).10.1103/PhysRevLett.45.1004Search in Google Scholar

[19] O. H. Nielsen and R. M. Martin, Phys. Rev. B 32, 3792 (1985).10.1103/PhysRevB.32.3792Search in Google Scholar

[20] N. Moll, M. Bockstedte, M. Fuchs, E. Pehlke, and M. Scheffler, Phys. Rev. B 52, 2550 (1995).10.1103/PhysRevB.52.2550Search in Google Scholar

[21] K. Gaal-Nagy, A. Bauer, P. Pavone, and D. Strauch, Comput. Mater. Sci. 30, 1 (2004).10.1016/j.commatsci.2004.01.002Search in Google Scholar

[22] V. V. Struzhkin, H. K. Mao, J. F. Lin, R. J. Hemley, J. S. Tse, et al. Phys. Rev. Lett. 96, 137402 (2006).10.1103/PhysRevLett.96.137402Search in Google Scholar

[23] M. Durandurdu, Phys. Rev. B 71, 054112 (2005).10.1103/PhysRevB.71.054112Search in Google Scholar

[24] K. Gaal-Nagy, P. Pavone, and D. Strauch, Phys. Rev. B 69, 134112 (2004).10.1103/PhysRevB.69.134112Search in Google Scholar

[25] K. J. Chang and M. L. Cohen, Phys. Rev. B 34, 8581 (1986).10.1103/PhysRevB.34.8581Search in Google Scholar

[26] J. Donohue, The Structure of the Elements, Wiley, New York 1974.Search in Google Scholar

[27] H. J. McSkimin, J. Appl. Phys. 24, 988 (1953).10.1063/1.1721449Search in Google Scholar

[28] M. Senoo, H. Mii, I. Fujishiro, and T. Fujikawa, J. Appl. Phys. 15, 871 (1976).10.1143/JJAP.15.871Search in Google Scholar

[29] C. S. Menoni, J. Z. Hu, and I. L. Spain, Phys. Rev. B 34, 362 (1986).10.1103/PhysRevB.34.362Search in Google Scholar

[30] G. Queisser and W. B. Holzapfel, Appl. Phys. A 53, 114 (1991).10.1007/BF00323869Search in Google Scholar

[31] H. J. McSkimin and P. Andreatch, J. Appl. Phys. 34, 651 (1963).10.1063/1.1729323Search in Google Scholar

[32] A. Di Cicco, A. C. Frasini, M. Minicucci, E. Principi, J. P. Itie, et al. Phys. Status Solidi B 240, 19 (2003).10.1002/pssb.200301847Search in Google Scholar

[33] E. Birch, J. Geophys. Res. 83, 1257 (1978).10.1029/JB083iB03p01257Search in Google Scholar

[34] R. J. Needs and R. M. Martin, Phys. Rev. B 30, 5390 (1984).10.1103/PhysRevB.30.5390Search in Google Scholar

[35] L. L. Boyer, E. Kaxiras, J. L. Feldman, J. Q. Broughton, and M. J. Mehl, Phys. Rev. Lett. 67, 715 (1991).10.1103/PhysRevLett.67.715Search in Google Scholar

[36] R. J. Needs and A. Mujica, Phys. Rev. B 51, 9652 (1995).10.1103/PhysRevB.51.9652Search in Google Scholar PubMed

[37] J. C. Jamieson, Science 139, 762 (1963).10.1126/science.139.3556.762Search in Google Scholar PubMed

[38] H. Olijnyk, S. K. Sikka, and W. B. Holzapfel, Phys. Lett. A 103, 137 (1984).10.1016/0375-9601(84)90219-6Search in Google Scholar

[39] J. Z. Hu and I. L. Spain, Solid State Commun. 51, 263 (1984).10.1016/0038-1098(84)90683-5Search in Google Scholar

[40] P. Modak, A. Svane, N. E. Christensen, T. Kotani, and M. van Schilfgaarde, Phys. Rev. B 79, 153203 (2009).10.1103/PhysRevB.79.153203Search in Google Scholar

[41] S. B. Qadri, E. F. Skelton, and A. W. Webb, J. Appl. Phys. 54, 3609 (1983).10.1063/1.332434Search in Google Scholar

[42] E. Anastassakis, A. Cantarero, and M. Cardona, Phys. Rev. B 41, 7529 (1990).10.1103/PhysRevB.41.7529Search in Google Scholar PubMed

[43] J. J. Hall, Phys. Rev. 161, 756 (1967).10.1103/PhysRev.161.756Search in Google Scholar

[44] J. Philip and M. Breazeale, J. Appl. Phys. 52, 3383 (1981).10.1063/1.329162Search in Google Scholar

[45] E. H. Bogardus, J. Appl. Phys. 36, 2504 (1965).10.1063/1.1714520Search in Google Scholar

[46] H. J. McSkimin and P. Andreatch, J. Appl. Phys. 35, 3312 (1964).Search in Google Scholar

[47] W. P. Mason and T. B. Bateman, J. Acoust. Soc. Am. 36, 644 (1964).10.1121/1.1919031Search in Google Scholar

[48] T. Bateman, W. P. Mason, and H. J. Mcskimin, J. Appl. Phys. 32, 928 (1961).10.1063/1.1736135Search in Google Scholar

[49] H. J. McSkimin and P. Andreatch, J. Appl. Phys. 35, 2161 (1964).10.1063/1.1713214Search in Google Scholar

[50] L. Pizzagalli, J. L. Demenet, and J. Rabier, Phys. Rev. B 79, 045203 (2009).10.1103/PhysRevB.79.045203Search in Google Scholar

[51] A. G. Beattie and J. Schirber, Phys. Rev. B 1, 1548 (1970).10.1103/PhysRevB.1.1548Search in Google Scholar

[52] A. Mayer and R. Wehner, Phys. Status Solidi B 126, 91 (1984).10.1002/pssb.2221260112Search in Google Scholar

[53] K. Brugger, Phys. Rev. 137, A1826 (1964).10.1103/PhysRev.137.A1826Search in Google Scholar

Received: 2015-1-23
Accepted: 2015-3-31
Published Online: 2015-5-5
Published in Print: 2015-6-1

©2015 by De Gruyter

Downloaded on 22.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/zna-2015-0027/html
Scroll to top button