Lattice dynamics properties of chalcopyrite ZnSnP2: Density-functional calculations by using a linear response theory
Yu You1, †, Dong Yu-Jing2, Shen Yan-Hong1, Zhao Guo-Dong1, Zheng Xiao-Lin1, Sheng Jia-Nan1
College of Optoelectronic Technology, Chengdu University of Information Technology, Chengdu 610225, China
School of Science and Technology, Xinyang University, Xinyang 464000, China

 

† Corresponding author. E-mail: yy2012@cuit.edu.cn

Project supported by the Open Research Fund of Computational Physics Key Laboratory of Sichuan Province, Yibin University (Grant No. JSWL2014KFZ01), the Scientific Research Fund of Sichuan Provincial Education Department, China (Grant No. 16ZB0209), the Scientific Research Foundation of Chengdu University of Information Technology, China (Grant No. J201611), and the National Natural Science Foundation of China (Grant No. 11547224).

Abstract

We present a first-principles study of the structural, dielectric, and lattice dynamical properties for chalcopyrite semiconductor ZnSnP2. The structural properties are calculated using a plane-wave pseudopotential method of density-functional theory. A linear response theory is used to derive Born effective charge tensors for each atom, dielectric constants in low and high frequency limits, and phonon frequencies. We calculate all zone-center phonon modes, identify Raman and infrared active modes, and report LO–TO splitting of the infrared modes. The results show an excellent agreement with experiment and propose several predictive behaviors.

1. Introduction

Ternary ABC2 semiconducting crystals having chalcopyrite structure have been synthesized from I–III–VI2 and II–IV–V2 elements. They are derived from their binary analog zinc-blende II–VI and III–V semiconductors, respectively. ZnSnP2 is a relatively little studied ternary chalcopyrite semiconductor material of the II–IV–V2 family. Recently, ZnSnP2 has received a great deal of attention as a photovoltaic material.[1] It is a promising candidate, given that its direct band gap of 1.68 eV[2] yields a high theoretical photovoltaic conversion efficiency close to 30% in AM1.5G sunlight at the Shockley–Queisser limit.[3] Undoped ZnSnP2 typically shows p-type conductivity.[4,5]

Born effective charges, low frequency dielectric constants, and the phonon frequencies provide important information about the properties of the materials. The vibrational properties of ZnSnP2 have been widely investigated for the past 40 years by a number of experimental and theoretical methods.[610] Experimentally, the infrared (IR) reflectivity of ZnSnP2 (at 295 K) was measured by Zlatkin et al.[6] in 1969, and the corresponding optical constants and vibrational frequencies are determined. However, no Raman-active modes are given. Employing Raman (R) spectroscopy and an infrared spectrometer, Mintairov et al.[7] measured the Raman scattering and infrared reflectivity, respectively. Both Raman-active and infrared-active modes are reported in Ref. [7]. Nevertheless, some modes have not been found in previous experiments.

Several earlier calculations have been performed to study the dynamical properties of ZnSnP2 by using phenomenological models, such as rigid ion and Keating’s models. The phonon dispersion relations over the entire Brillouin zone of ZnSnP2 were calculated using the rigid ion model by Mintairov et al.[7] The zone-centered phonons of ZnSnP2 have been studied using Keating’s models by Bettini.[10] In addition, the phonon frequencies values have been calculated by transfer of the spectral pattern.[9] The theoretical research mentioned above has a limitation, that is to say, the potentials in the calculations depend on experimental value. Identification of vibration modes in experimental measurements is sometimes incomplete, partly because of the complexity of the dispersion scheme in the chalcopyrite structure and partly because of experimental errors.

Nowadays the first-principles calculation has proved to be a powerful tool for the study of lattice dynamical properties. First-principles calculations of dynamical properties of ZnSnP2 have been studied by Lazewski et al.[8] The calculations were done by the direct method[11,12] from HF forces arising when a single atom is displaced from its equilibrium position. Generally, the first-principles investigation of phonon frequencies is mainly divided into two categories: the direct method and the linear-response method.[13,14] The longitudinal–optical/transverse–optical (LO–TO) splitting of the infrared active modes based on effective charges and dielectric tensors are not directly given for the direct method. However, the two quantities can be calculated directly by a linear response method based on density functional perturbation theory (DFPT).[15,16] Although there are some studies of lattice dynamical properties for ZnSnP2, they are still an important part of a full understanding of the phonon in the material.

In this paper we present first-principles linear response theory to investigate the lattice dynamics properties, such as Born effective charge tensors for each atom, dielectric constants in low and high frequency limits, and phonon frequencies. This paper is organized as follows. In Section 2, we briefly review the underlying theoretical methods. In the following sections, results are presented and discussed for geometry parameters (Subsection 3.1), Born effective charge tensor (Subsection 3.2), phonon frequencies (Subsection 3.3), and dielectric tensor (Subsection 3.4), respectively. Finally, conclusions are given in Section 4.

2. Computational method

Our calculations were performed with the ABINIT package,[17] which is based on density-functional theory (DFT) using the pseudopotential method and a plane-wave expansion of the wave functions. The interactions between the ions and valence electrons were described using norm-conserving local density approximation (LDA) pseudopotentials which are generated in the scheme of Troullier–Martins (TM).[1821] For the exchange–correlation potential we have used the local density approximation data of Ceperly–Alder, as parametrized by Perdew and Wang.[22] The Zn (3d10, 4s2), Sn (5s2, 5p2), and P (3s2, 3p3) orbitals are treated as valence states. The pseudopotentials used in the present calculations are soft potentials of the TM type, available on the ABINIT website. The density-functional perturbation theory is the basis of the linear-respond approach. Phonon frequencies and atoms displacements are obtained using the linear-respond method, which avoids the use of supercells and allows the calculation of dynamical matrix at arbitrary vectors. Technical details on the computation of responses to atoms displacements can be found in Refs. [23] and [24], while reference [14] presents the subsequent computation of phonon frequencies, Born effective charge tensors for each atom, and dielectric constants.

The calculations were carried out with a 36-hartree plane-wave energy cutoff, and the tetragonal Brillouin zone (BZ) was sampled with a regular and shifted 4× 4× 4 k-point mesh. Convergence tests show that the BZ sampling and the kinetic energy cutoff are sufficient to insure an excellent convergence within 1 cm−1 for the calculated phonon frequencies.

3. Results and discussion
3.1. Structural optimization

The body-centered tetragonal chalcopyrite structure of ZnSnP2 is shown in Fig. 1. Each unit cell contains four formula units and the point group is . The chalcopyrite structure is deduced from that of the zinc-blende by the replacement of the cationic sublattice by two different atomic species. In general, II–V and IV–V bond lengths, denoted by dII–V and dIV–V, respectively, are not equal, which is mentioned in the substitution results in two different structural deformations: the first one is the relocation of anions in the xy plane which is characterized by parameter . Here, a is the lattice constant in the x or y direction. The second consequence of differing anion-cation bond lengths is a deformation of the unit cell along the z direction to a length c, which is generally different from 2a. This tetragonal distortion is characterized by the quantity c/a. For real compounds of pnictide II–IV–V2 groups in most cases c/a = 1.769–2.016 and u = 0.214–0.304.[25] The unit vectors of the primitive cell are (a, 0, 0), (0, a, 0), (a/2, a/2, c/2). The cations are located at 4a and 4b while anions are located at 8d Wyckoff positions. Ion positions can be generated using the following minimum set of (x, y, z) coordinates expressed in units of the a and c constants.

Group II: (0, 0, 0), (0, 1/2, 1/4);

Group IV: (0, 0, 1/2), (0, 1/2, 3/4);

Group V: (u, 1/4, 1/8), (−u, 3/4, 1/8), (3/4, u, 7/8), (1/4,−u, 7/8).

Fig. 1. (color online) Crystal structure of ZnSnP2 (chalcopyrite-type; body-centered tetragonal- .

The ground state structural properties are obtained by minimization of the total energy with respect to the unit cell volume. The structural parameters are obtained by optimizing lattice constants and atomic coordinates until all force components are below 5× 10−5 hartree/Bohr (1 hartree=27.2114 eV). Table 1 summarizes our results obtained after relaxation of the lattice constants, as well as the available calculated[8,26,27] and experimental values.[5,28] For ZnSnP2, optimization of unit cell geometry within LDA leads to a = 5.622 Å, c = 11.233 Å. The calculated lattice constants are underestimated with the maximal error of 0.7%, which is typically the expected precision for the LDA. Even comparison with experimental values at the lowest temperatures available is, strictly speaking, not correct, since it neglects the anharmonic effect of the zero-point vibrations. Even so, the calculated lattice parameters are found to be in good agreement with the experimental ones. The optimized results show our calculation method is feasible and we will adopt the optimized structure parameters to calculate other properties.

Table 1.

The equilibrium lattice constant a (c) (in unit Å), axial ratio c/a and internal parameter for ZnSnP2.

.
3.2. Born effective charge tensors

The Born effective charge tensor quantifies the macroscopic electric response of a crystal to the internal displacements of its atoms. They are important quantities in obtaining the LO–TO phonon splitting, also they provide some information about the ionicity of the material. For atom κ, the Born effective charge tensor is defined as the proportionality coefficient relating, at linear order, the polarization per unit cell, created along the direction β, and the displacement along the direction α of the atoms belonging to the sublattice κ, under the condition of a zero electric field.[14] In this work, the Born effective charge tensors for each atom of ZnSnP2 are calculated by DFPT.

Values of the calculated Born effective charge of ZnSnP2 are presented in Table 2. To the best of our knowledge, no other first-principles calculations of the Born effective charge for ZnSnP2 exist. The form of for atom κ depends on the site symmetry of the ions. The point symmetries at the site of Zn ion and Sn ion are S4 (4a sites) and S4 (4b sites), respectively. For P ion, the point symmetry at the site is C2 (8d sites). The effective charge tensors for cations are diagonal and obey , as required by symmetry for a tetragonal crystal. According to the calculation results, cations have diagonal Born effective charges with , , and , . Z* of cations are almost diagonal with an anisotropy of 2% for Zn and 5% for Sn, respectively. The effective charge tensors have nearly spherical symmetry, with small tetragonal distortion caused by . P ions, located at lower symmetry sites, have nondiagonal and anisotropic Z*. For all anions, while and take the value −1.99 or −2.27 depending on the distortion of u. Also, depending on the u distortion being along the x or y direction, the nondiagonal components , or , are different than zero. We can see that the shape of Z* for P ions is far from a sphere. This behavior has also been observed in the case of other chalcopyrite semiconductors.[2932]

Table 2.

Calculated Born effective charge tensors, Z*, eigenvalues of the symmetric part of , average of eigenvalues , and force-field model effective charges of Ref. [10].

.

The average eigenvalues for cations and anions are comparable to the values in Ref. [10]. The calculated average eigenvalues ( in our work for Zn, Sn, and P are 1.75, 2.55, and −2.15, respectively. Obviously, the three atoms have smaller effective charges than their formal charges, +2, +4, and −3 for Zn, Sn, and P, respectively. Such difference can be derived from the strong covalent characteristic for bonds. To the best of our knowledge, no other first-principle calculations and experimental data of Z* for ZnSnP2 exist. We have compared our results with other literature data calculated with a different force-field model.[10] This discrepancy is pronounced for the effective charges in Ref. [10] obtained by a least square fit of the phonon energies and Coulomb splittings, probably because of the use of approximate eigenvectors and assumptions about the dynamical charges. Earlier results by Akdogan et al.[31] gave Born effective charges of CuGaS2 with the same linear response as ours, and average of eigenvalues displayed a good agreement with experimental data.

3.3. Phonons

The crystal symmetry of ZnSnP2 is body-centered tetragonal with a centro-symmetric space group and a corresponding point group. Positions of nonequivalent atoms are indicated in Fig. 2. There are four distinct P, two Zn, and two Si atoms. Transformation properties of these atoms under symmetry operations of the space group are noted in Table 3. As a tetragonal crystal, ZnSnP2 has phonon dispersion relations consisting of 24 branches whose group theoretical analysis at the BZ center ( yields a decomposition into

where E is a double degenerate mode. The acoustic modes correspond to one B2 and one E mode. For optical modes, the irreducible representation is

Fig. 2. (color online) (001) projection of unit cell showing positions of four nonequivalent P atoms, the two nonequivalent Zn atoms and the two nonequivalent Sn atoms. Symmetry coordinates for A modes ( ; displacements in the plane are represented by arrows and displacements in the z direction by open and closed circles.
Table 3.

Character table for the D2d point group (tetragonal).

.

The A1 and B1 modes are Raman-active, B2 and E modes are both Raman- and IR-active, and the A2 mode is silent. These IR modes are polar modes and subject to an LO–TO splitting. Symmetry coordinates for the onefold A and B representations are given in Figs. 2 and 3, respectively, and for the twofold E representations in Fig. 4. A detailed discussion can be found in Ref. [33].

Fig. 3. (color online) Symmetry coordinates for B modes ( ; displacements in the plane are represented by arrows and displacements in the z direction by open and closed circles.
Fig. 4. (color online) Symmetry coordinates for E modes (7 ; displacements in the plane are represented by arrows and displacements in the z direction by open and closed circles.

Complete phonon branches of ZnSnP2 are plotted for the high-symmetry lines in the BZ together with the corresponding phonon density of states (DOS) in Fig. 5. An interesting feature of acoustic branches around the point is observed, especially along X and Z directions. In order to obtain the phonon dispersion curves throughout the BZ, the dynamical matrices are obtained 4× 4× 4 grid of q points, and real space force constants are then found by Fourier transform of the dynamical matrices. The acoustic sum rule is applied to force the three acoustic phonon frequencies at the point equal to zero strictly as being implied by translation symmetry.

Fig. 5. Calculated phonon dispersion curves along symmetry lines in the Brillouin zone and the corresponding phonon density of states (DOS) for ZnSnP2.

In Table 4, we compare calculated phonon frequencies of ZnSnP2 at the point with the measured Raman[7] and infrared[6,7] values. For comparison, we cite also the phonon frequencies from theoretical calculations through direct methods[8] and a classical model.[7,9,10] As shown in Fig. 5 and Table 4, three groups of modes can be identified. A clear gap exists between the twelve upper energy branches and the remaining twelve branches. The low frequency group of modes has a range of values from 82 cm−1 to 115 cm−1 and the medium frequency group of modes has a range of values from 180 cm−1 to 207 cm−1. The highest frequency group of modes can be up to 370 cm−1. Unfortunately, some vibration modes have not been observed in previous experiments, including infrared and Raman measurements. To get more accurate measurements, inelastic neutron scattering experiments should be adopted, but it is not available in previous works. Otherwise, for some of the modes there are large differences among the reported theoretical frequencies, for example the highest frequency B2 mode ranges from 350 cm−1 to 375 cm−1. Our calculation shows that the mode is at 365 cm−1, which is in very good agreement with experiments.[6,7] It makes us very confident in the prediction of the frequencies for the others.

Table 4.

A comparison of calculated phonon frequencies (in unit cm−1) at the point with Raman and infrared data as well as with other theory values. Two numbers in a row correspond to LO/TO frequencies.

.

It should be noted that the direct method was performed to calculate the phonon frequencies of ZnSnP2 by Lazewski et al.[8] Although the calculated frequencies show a better agreement with ours, the LO–TO splittings are not discussed. This is because the splitting depends on the effective charges and dielectric tensors of the system, which are not directly accessible to the direct method. The IR modes group into modes with displacements either in the x, y plane or along the z direction. The E mode has a displacement pattern in the x, y plane, and the B2 mode has displacements along the z direction. The LO–TO splittings for B2 and E modes are presented in Table 4 and the values of LO–TO splitting agree well the experimental results.[6,7] The missing modes in the experiment are predicted successfully. For ZnSnP2, the LO–TO splitting is small, ranging from 0 cm−1 to 10 cm−1. The splitting is even zero in the low frequency region, because the low frequency modes are essentially the folded acoustic modes, they correspond to whole molecular units moving relative to each other. The largest difference appears on the E mode with the highest frequency. The large LO–TO splittings of these modes suggest they involve large effective charges and make large contributions to the static dielectric tensor of ZnSnP2. A similar splitting case occurs for CuGaSe2 and the results can be found in Ref. [29].

3.4. Lattice dielectric properties

The dielectric tensor for a tetragonal system such as this will be diagonal with two distinct elements. For ZnSnP2, the calculated electronic ( and static ( dielectric tensors are diagonal and have two independent components and perpendicular to and along the c axis, respectively. In Table 5 we compare the values calculated for the dielectric tensor of ZnSnP2 with theoretical and experimental data available. The electronic dielectric constants are calculated to be and . The static dielectric tensor can be decomposed into contributions of different modes and calculated by the generalized Lyddane–Sachs–Teller (LST) relation:[34]

where the frequencies and are the long wavelength limits for longitudinal and transverse vibrations, respectively. According to the LST relation, the static dielectric constants of ε0 both directions are and , respectively, showing a nearly isotropic character. The averages of and ε0, obtained from the expression (or = are also shown in Table 5. The components of ε0 are much larger than those of as a consequence of the significant contribution to the low-frequency dielectric permittivity tensor due to ionic displacements.

Table 5.

Static and high frequency dielectric tensor components of ZnSnP2.

.

The dielectric properties of ZnSnP2 have not been studied very much experimentally. We are only aware of the measurement of the dielectric constant of ZnSnP2 in which a value of (or has been reported in Ref. [6]. On the theoretical side, the dielectric constant has been calculated by Verma et al.[34] using a plasma oscillations theory. ε0 from Ref. [35] is 9.90 and seems to be more consistent with the experimental value 10.0. Our calculated values and previous theoretical values[27,36] are calculated by using the DFT and larger than the experimental results. The overestimation of the calculation is a well-known fact of the DFT due to the underestimation of the band gap.[34,37,38] We note that a smaller energy gap yields a larger ε0 value. This can be explained on the basis of the Penn model.[39] The Penn model is based on the expression . It is clear that ε0 is inversely proportional to . The dielectric constants calculated in our work deviate from other theoretical data[27,35] because a small band gap was adopted in the calculation.

4. Conclusions

To summarize, we have presented first-principles calculations of the structural, dielectric, and, in particular, of the lattice dynamical properties of the chalcopyrite semiconductor ZnSnP2within the density-functional theory and pseudopotential methods. The optimal ground state structure is studied within the LDA approaches. The relaxed lattice constants are found to be in good agreement with the experimental ones with the maximal error of 0.7%. Born effective charge and dielectric constants are calculated within the linear response theory. The calculated average eigenvalues for Zn, Sn, and P are 1.75, 2.55, and , respectively. They are smaller than their formal charges because of the strong covalent characteristic for bonds. Our theoretical values of dielectric constants tend to be overestimated in LDA calculations due to the underestimation of the band gap. The phonon frequencies at the point of the BZ are calculated and their assignments are given. The LO–TO splittings for B2 and E modes are presented and the values of LO–TO splitting agree well the infrared and Raman measurements. We believe that our theoretical predictions should be highly valuable for the experimental community in the framework of the characterization of ZnSnP2.

Reference
[1] Folmer J C W Tuttle J R Tu J A Parkinson B A 1985 J. Electrochem. Soc. 132 1608
[2] St-Jean P Seryogin G A Francoeur S 2010 Appl. Phys. Lett. 96 231913
[3] Shockley W Queisser H J 1961 J. Appl. Phys. 32 510
[4] Rubenstein M Ure R W 1968 J. Phys. Chem. Solids 29 551
[5] Vaipolin A A Goryunova N A Kleshchinskii L I Loshakova G V Osmanov E O 1968 Phys. Status Solidi 29 435
[6] Zlatkin L B Markov Yu F Stekhanov A I Shur M S 1969 Phys. Stat. Solidi 32 473
[7] Mintairov A M Sadchikov N A Sauncy T Holtz M 1999 Phys. Rev. 59 15197
[8] Lazewski J Parlinski K 2001 J. Alloys Compd. 328 162
[9] Ohrendorf F W Haeuseler H 1999 Cryst. Res. Technol. 34 339
[10] Bettini M 1975 Phys. Stat. Solidi 69 201
[11] Parlinski K Li Z Q Kawazoe Y 1997 Phys. Rev. Lett. 78 4063
[12] Kunc K Martin R M 1982 Phys. Rev. Lett. 48 406
[13] Giannozzi P Gironcoli S de Pavone P Baroni S 1991 Phys. Rev. 43 7231
[14] Gonze X Lee C 1997 Phys. Rev. 55 10355
[15] Baroni S Giannozzi P Testa A 1987 Phys. Rev. Lett. 58 1861
[16] Baroni S Gironcoli S de Dal Corso A Giannozzi P 2001 Rev. Mod. Phys. 73 515
[17] Gonze X Beuken J M Caracas R Detraux F Fuchs M Rignanese G-M Sindic L Verstraete M Zerah G Jollet F Torrent M Roy A Mikami M Ghosez Ph Raty J Y Allan D C 2002 Comput. Mater. Sci. 25 478
[18] Goedecker S 1997 SIAM J. Sci. Comput. 18 1605
[19] Payne M C Teter M P Allan D C Arias T A Joannopoulos J D 1992 Rev. Mod. Phys. 64 1045
[20] Gonze X 1996 Phys. Rev. 54 4383
[21] Fuchs M Scheffler M 1999 Comput. Phys. Commun. 119 67
[22] Perdew J P Wang Y 1992 Phys. Rev. 45 13244
[23] Gonze X 1997 Phys. Rev. 55 10337
[24] Hamann D R Wu X Rabe K M Vanderbilt D 2005 Phys. Rev. 71 035117
[25] Jaffe J E Zunger A 1984 Phys. Rev. 29 1882
[26] Hinuma Y Oba F Nose Y Tanaka I 2013 J. Appl. Phys. 114 043718
[27] Sahin S Ciftci Y O Colakoglu K Korozlu N 2012 J. Alloys Compd. 529 1
[28] Shaposhnikov V L Krivosheeva A V Borisenko V E 2012 Phys. Rev. 85 205201
[29] Parlak C Eryigit R 2006 Phys. Rev. 73 245217
[30] Lazewski J Jochym P T Parlinski K 2002 J. Chem. Phys. 117 2726
[31] Akdogan M Eryigit R 2002 J. Phys.: Condens. Matter 14 7493
[32] Yu Y Zhao B J Zhu S F Gao T Hou H J 2011 Solid State Sci. 13 422
[33] Kaminow I P Buehler E Wernick J H 1970 Phys. Rev. 2 960
[34] Martin R M Ortiz G 1997 Phys. Rev. 56 1124
[35] Verma A S Bhardwaj S R 2006 Phys. Stat. Solidi 243 2858
[36] Chiker F Abbar B Bresson S Khelifa B Mathieu C Tadjer A 2004 J. Solid State Chem. 177 3859
[37] Gonze X Ghosez Ph Godby R W 1995 Phys. Rev. Lett. 74 4035
[38] Ghosez Ph Michenaud J P Gonze X 1998 Phys. Rev. 58 6224
[39] Penn D R 1962 Phys. Rev. 128 2093