Next Article in Journal
Iron Sulfide Materials: Catalysts for Electrochemical Hydrogen Evolution
Previous Article in Journal
Determination of pKa Values via ab initio Molecular Dynamics and its Application to Transition Metal-Based Water Oxidation Catalysts
 
 
Correction published on 10 September 2019, see Inorganics 2019, 7(9), 112.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First Principles Study of the Vibrational and Thermal Properties of Sn-Based Type II Clathrates, CsxSn136 (0 ≤ x ≤ 24) and Rb24Ga24Sn112

Department of Physics and Astronomy, Texas Tech University, Lubbock, TX 79409-1051, USA
*
Author to whom correspondence should be addressed.
Submission received: 12 May 2019 / Revised: 6 June 2019 / Accepted: 11 June 2019 / Published: 14 June 2019

Abstract

:
After performing first-principles calculations of structural and vibrational properties of the semiconducting clathrates Rb24Ga24Sn112 along with binary CsxSn136 (0 ≤ x ≤ 24), we obtained equilibrium geometries and harmonic phonon modes. For the filled clathrate Rb24Ga24Sn112, the phonon dispersion relation predicts an upshift of the low-lying rattling modes (~25 cm−1) for the Rb (“rattler”) compared to Cs vibration in CsxSn136. It is also found that the large isotropic atomic displacement parameter (Uiso) exists when Rb occupies the “over-sized” cage (28 atom cage) rather than the 20 atom counterpart. These guest modes are expected to contribute significantly to minimizing the lattice’s thermal conductivity (κL). Our calculation of the vibrational contribution to the specific heat and our evaluation on κL are quantitatively presented and discussed. Specifically, the heat capacity diagram regarding CV/T3 vs. T exhibits the Einstein-peak-like hump that is mainly attributable to the guest oscillator in a 28 atom cage, with a characteristic temperature 36.82 K for Rb24Ga24Sn112. Our calculated rattling modes are around 25 cm−1 for the Rb trapped in a 28 atom cage, and 65.4 cm−1 for the Rb encapsulated in a 20 atom cage. These results are utilized to predict the lattice’s thermal conductivity (approximately 0.62 W/m/K) in Rb24Ga24Sn112 within the kinetic theory approximation.

1. Introduction

At present, the shortage of energy resources and air pollution have become more and more crucial around the globe. Therefore, seeking clean energy rather than old-fashioned fossil fuel has attracted attention, both from the viewpoint of theoretical investigation and experimental synthesis work. Utilizing themoelectric (TE) materials based on the clathrate phase provides an efficient way of converting waste heat to usable electric energy [1,2], according to the Seeback effect [3]. This novel technology has several advantages, such as low cost, long lifetime, as well as pollution-free properties. Motivated by these potential developments that benefit our society by exploring energy resources, growing research interest has given to clathrate, whose TE performance is evaluated by the dimensionless figure of merit (ZT) value: ZT = TS2σ/κ. Here Z denotes the figure of merit while T represents the absolute temperature. The electrical conductivity and thermal conductivity are given by σ and κ, respectively. S is the Seeback coefficient. It is noted that an efficient TE device [4,5] with prominent performance is manifested by a large value of ZT. To our knowledge, the best ZT value for TE material synthesized from a clathrate compound is around 1.0 [4].
Recently, much focus has been placed on “cage-structured” clathrate materials [5,6,7] whose empty frameworks are capable of encapsulating “rattling” guests arising from alkaline or alkaline-earth atoms. This is due to the fact that this kind of material satisfies the “phonon glass electron crystal” (PGEC) concept proposed by G. A. Slack [8,9]. In this work, we employ the term “guest” to describe the atom (e.g., Rb of Cs) that enters into the cage polyhedron made of Sn (or a mixture of Ga and Sn). Furthermore, the characteristics of the PGEC material can be connected to glass-like thermal conductivity along with the electronic performance found in the crystal. Previous calculations [10,11,12,13] have shown that the guest, acting as a “rattler”, can vibrate resonantly with the framework cage in order to effectively minimize the lattice’s thermal conductivity, thereby contributing efficiently to the enhancement of the ZT value of these TE materials.
Specifically, inorganic clathrates based on Group-IV elements, such as Si, Ge, and Sn, have so far received considerable attention [14,15,16,17], owing to their potential role of serving as thermoelectric candidate materials. Compared to silicon and germanium, the relatively large atomic weight of tin causes its cage size to stay comparably greater than its Si and Ge counterparts, leading to a reduced lattice thermal conductivity, while the “rattler” atom remains fixed in the entire compound system [18,19]. Consequently, research on Sn-based clathrate has become popular. It was reported that low thermal conductivity can be found in the Sn-based clathrate whose polyhedron building block (cage) incorporates an Sr atom as a “guest filler”, causing the material to exhibit prominent thermoelectric performance [20].
Lattice structures of the Sn-based clathrate framework are built from face-shared polyhedron cages, offering a series of ways to host encapsulated “guest” impurities. In particular, tuning the guest atom (or substitutional framework atom) type and number can produce unexpected electronic, vibrational, thermal, and transport properties for these “cage-structured” clathrate compounds. Accordingly, a rich variety of lab-synthesized clathrates, whose frameworks are made of Sn, can be normally classified as either Type-I or Type-II structures [21,22,23,24], in spite of little understanding on the Type-VIII configuration [25,26]. It is anticipated that the Type-VIII Sn-based clathrate might be one of the most efficient TE materials due to its promising transport behavior.
In this paper, we mainly explore the structural, vibrational, and thermal properties of the Type-II binary clathrate AxSn136 (0 < x ≤ 24; A = Na, K, Cs) and ternary semiconducting compound Rb24Ga24Sn112. It is necessary to mention that the framework of the Type-II clathrate comprises 20 atom (dodecahedron) cages and 28 atom (hexakeidacahedron) cages which are connected with a ratio of 4:2 in the context of a unit cell involving 136 atoms. Unlike AxSi136 (0 < x ≤ 24), AxSn136 possesses a tuned framework whose host atom varies from silicon to tin, making it impossible to have lattice contraction when x is elevated from 4 to 16 in an ascending order. Regarding the composition-dependent structural response, we always assumed that the “over-sized” cages (e.g., 28 atom tin cages or Sn28) are partially or completely filled with preferred priority. The vibrational modes with Cs vibration in the Sn28 cage, as well as Rb rattling modes in 28 atom (and 20 atom) cages, are also examined. The obtained results show that frequency of Cs vibration in the Sn28 cage of CsxSn136 (x = 12, 16, 20) is almost half as much as that of the Rb in the 28 atom cage of Rb24Ga24Sn112. Meanwhile, Rb vibration in a dodecahedron cage induces smaller atomic displacement parameters (ADPs) than that of Rb vibration in the hexakeidacahedron cage, leading to discrepant rattling frequency and different characteristic temperature of the Einstein oscillator, which pertains to specific heat investigation.

2. Results and Discussion

2.1. Structural Property of AxSn136 (0 ≤ x ≤ 24; A = Na, K, Cs) and Vibrational Property of CsxSn136 (x = 12, 16, 20)

It is necessary to point out that the binary clathrate system studied here possesses a face centered cubic (FCC) Type-II structure, for which the group notation reads No. 227, Fd3m. Accordingly, the chemical stoichiometry of the framework sites, along with the guest impurity, is manifested by the notation AxSn136. Two classes of polyhedron cages form the building blocks inside the Sn136 framework: Sn28 and Sn20 cages, whose voids are connected periodically with a 2:4 ratio. These two different sized cages that contain 28 and 20 atoms are also termed hexakaidecahedron and dodecahedron cages, respectively. For the geometry optimization process, we need to point out that the starting atomic coordinate for each atom in the primitive unit cell is a specified “guess”, indicating that it is not identical, but very close, to the perfect value, according to the definition of “Ab initio” first-principle computation. Thus, we chose the atomic coordinates that were derived from our previous research work coping with Sn136 [17]. Furthermore, on the basis of pure Sn136 framework coordinates, we make sure that the added Cs guest impurities go to the specified Wyckoff sites—for instance, 8b and 16c, in a given integer amount. In other words, it depends on how many integer numbers of Cs atoms occupy either the small cage (16c site) or the oversized cage (8b site). It seems necessary to mention that the large cages are always filled in the very high priority, meaning that the small cage begins to accommodate the Cs guest until all of the 28 atom cages are completely occupied.
Previous X-ray diffraction (XRD) work along with our VASP calculation [27] on AxSi136 reported that, as x increases, the lattice contracts for 0 < x < 8 and expands for 8 < x < 24. However, changing all silicon framework atoms into tin in AxSi136 results in no lattice contraction, according to our calculations. In other words, the lattice configuration for Sn-based clathrates experiences monotonically increasing behavior if the composition x, which spans over the range of 0 and 24, is elevated from the lower limit. The following figure illustrates our theoretical predictions on the composition-dependent lattice constant, in the context of the fact that the rattler is tuned from Na to K to Cs.
As seen from the numerical results in Figure 1, our predicted lattice constant, as a function of composition x, is 17.38 Å for Na8Sn136, 17.43 Å for K8Sn136, and 17.44 Å for Cs8Sn136. Considering the case x = 24 where all cages are completely occupied by guest impurities, the predicted lattice constants are distinct and given by 17.51 Å for Na24Sn136, 17.64 Å for K24Sn136, and 17.90 Å for Cs24Sn136. The result of the pure Sn136 gives rise to 17.34 Å for the lattice constant when x equals zero. Figure 2 shows the predicted vibrational spectrum of the CsxSn136 (x = 12, 16, 20) clathrate materials.
It is clearly seen from Figure 2 that the low-lying vibrational modes, with respect to Cs motion in Sn28 cages (written as “Cs@Sn28” in the figure), remain in the range of 11.5–12.5 cm−1 for the three clathrates. The frequency of the tops of these optic bands relating to the Cs guest seems to decrease in a slight manner, as composition x is increased from 12 to 16 to 20. Besides this, the acoustic bands are located within the range of 0–25 cm−1 while exhibiting little variation. In other words, the acoustic bandwidths in CsxSn136 (x = 12, 16, 20) are nearly unchanged, leading them to be independent of their guest concentration x.
The vibrational density of states (VDOS) obtained below (Figure 3) is shown to demonstrate the interesting features of the “rattling” modes (the top of entire optic band and the densely flat optic modes (right above the acoustic bands) in Figure 2). As seen from the low-frequency (about 5–20 cm−1) picture of VDOS, the compounds CsxSn136 are intriguing because of the existence of “rattler” vibrational modes (shown by the peaks), due to the Cs motion in the “over-sized” cage. More precisely, introducing the Cs impurities into the 28 atom framework cages can produce the “rattling” vibrational modes. These predicted VDOS curves are presented with respect to CsxSn136, for x = 12, 16, and 20. The inset figure intricately illustrates how these Cs “rattling” modes are remarkably identified by the thin peaks centered around 12 cm−1. An extremely slight decreasing behavior subjected to the Cs mode frequency (peak value) can be captured, as the number of guests who occupy Sn20 cages is elevated from 4 to 8 to 12. This result is gained on the basis of enlarged unit cell geometry (containing 136 atoms), assuming all Sn28 cages are completely occupied. Furthermore, the “red-shift” of phonon vibrational frequency is clearly illustrated when paying attention to the top of the densely flat optic band (within the range of 25 and 50 cm−1). Accordingly, there appears to be a reduction of the entire optic band’s maximum (located right below about 190 cm−1) with the elevated composition x. Specifically, the “red-shift” of the top of the entire optic band occurring at about 180 cm−1 is obtained for Cs20Sn136, in contrast to that of Cs16Sn136 (at about 183 cm−1) and Cs12Sn136 (at about 186 cm−1).

2.2. Vibrtional and Thermal Properties of Rb24Ga24Sn112 and Cs24Ga24Sn112

The crystal structure for the framework of the clathrate-II system is explained in the way that the [512614] and [512] polyhedron cages alternate with a 1:2 ratio, generally. For our studied material, Rb24Ga24Sn112, its stoichiometry can be associated with Type-II semiconducting clathrate due to the “Zintl-phase” concept, which was named after Edward Zintl who pioneered their investigation on the classical NaTl material [28]. It is theoretically noted that each constituent of this ternary compound attains a closed s3p1 valence shell by establishing a required charge transfer with covalent bonds. The charge compensation is achieved between the substitutional cage atoms (Ga) and guest impurities (Rb). In other words, the Rb atoms donate twenty-four electrons to the host so that the octet rule for the twenty-four Ga atoms is satisfied in terms of Rb24Ga24Sn112. All framework bonds are, therefore, saturated. Additionally, the covalent configuration demonstrating sp3 coordination makes it possible to form zeolite-like expanded volume phases in an open framework given by Ga24Sn112. This four-fold-coordinated system possesses a local bonding picture, which remains similar to that of the diamond phase structure (e.g., α-Sn).
In order to explore the effect of the encapsulated Rb rattler on the cages of the open framework Ga24Sn112 in our study, we show the vibrational dispersion curves for Rb24Ga24Sn112 and Cs24Ga24Sn112 in Figure 4 and Figure 5. Our eigenvector analysis shows that the acoustic phonon region (below about 45 cm−1) contains flat Rb-associated phonon bands lying around 25 cm−1. Accordingly, the optic mode region ranges from about 45 cm−1 to about 220 cm−1. The average value of the “localized”, flat, vibrational modes of Rb is used to characterize the guest, which is loosely confined in a 28 atom cage (“over-sized” cage), leading to a comparably small effective force constant (ϕ). Here, the vibrational frequency ω is given by ω = (ϕ/m)1/2, where m can be numerically replaced by the atomic mass of Rb (85.47 a.u). Contrary to this, the Rb guest, which is strongly bound to the small 20 atom cage, results in a vibrational frequency that is labeled by the red arrow (showing about 65.4 cm−1 in Figure 4). The discussed modes of Rb in the 28 and 20 atom cages are indicated by the notation “Rb@28” along with “Rb@20”. The downshift of the labeled frequency is expected to arise from the comparably large cage size of the 28 atom cage, in contrast to that of the 20 atom cage.
In contrast to vibrational property found in Figure 4, we illustrate the dispersion relation of Cs24Ga24Sn112 in the following figure after conducting the DFT calculation. As seen from Figure 5, the Cs-associated phonon bands lying around 22 cm−1 denote the localized, flat, optic phonons. This means that the Cs guest vibration restricted within the 28 atom cage produces a slightly smaller “rattling” frequency than that of Rb guest mode in Rb24Ga24Sn112. Similar to the result in Figure 4, the Cs guest, which is strongly bound in the 20 atom cage, relates to the vibrational mode, leading to a value of about 55 cm−1, labeled by the red arrow. Considering the relation ω = (ϕ/m)1/2 while taking the Rb mass (85.47 a.u) and Cs mass (132.91 a.u.) into account, the calculated force constant ratio of Rb to Cs yields (85.472 × 25)/(132.912 × 22) ≈ 0.47. Therefore, the Rb seems to be loosely bound with respect to the 28 atom cage when compared to the Cs resonant vibration with the 28 atom cage in Cs24Ga24Sn112.
In Figure 6, we present our calculated isotropic atomic displacement parameters (ADPs) for the Rb “rattler”, which is localized to vibrate in the 20 and 28 atom cages. This work is done after we have theoretically determined the vibration mode of Rb from lattice dynamics computation, using the LDA scheme. Our work was originally motivated by the temperature-dependent isotropic mean-square displacement (Uiso), defined as Uiso = (h/(8π2mv))coth(hv/(2kBT)). This defined parameter evaluates displaced distance for an atom that deviates from its equilibrium position. From the numerical definition, v denotes the vibrational frequency, m represents the reduced mass of the oscillator, h is Plank’s constant, and kB is Boltzmann’s constant. Combined with the high-temperature approximation (/2π << kBT; ω ≡ 2πv), we use kBT/ϕ to estimate Uiso according to the relation ϕ2. In Figure 6, the temperature variable is restricted to the range of 0 and 400 K. Meanwhile, the notations Rb(1) and Rb(2) represent the guest rattlers confined in the 28 atom and 20 atom cages, respectively. The results show that Uiso parameters for the Rb atoms in the 28 atom cages are larger than those for the same “rattlers” confined within the Sn20 polyhedron cages. These calculations can be compared, to some extent, to the old investigations on Uiso of Rb in Na16Rb8Ge136 [29], as well as Uiso of Rb in Na16Rb8Si136 [30].
In addition, the slope of each curve of ADPs vs. T predicts a different magnitude of the “rattler”’s oscillator strength. Weak bonding gives rise to a small ϕ, as found in the inverse of the slope of curve (with notation “Rb(1)”) in Figure 6. The Rb impurity vibrating inside the 28 atom cage induces the smallest ϕ compared to the same “rattler”’s motion in the 20 atom cage. Similar results can be linked to the powder XRD and single-crystal XRD data regarding the calculated ϕ at room temperature (RT) in the case of Na24Si136 [31]. Other similar results reported elsewhere relate to the Type-I clathrates [32,33,34] and skutterudites [35,36].
Before determining the heat capacity within the framework of harmonic approximation (HA), the temperature-dependent Helmholtz free energy F is identified as F(T) = kBTqj In {2sinh(j(q)/4πkBT)}, where kB is Boltzmann’s constant and h is Planck’s constant. ωj(q) remains to be the jth-phonon mode frequency. It was realized that we should ignore the volume-dependent static contribution to the internal energy acquired by LDA because we solely care about the temperature derivative with respect to free energy. Consequently, using the same HA scheme, we theoretically obtain the specific heat at a constant volume of Cv = −T(ə2F(T)/ə2T)V, in the context of the first-principles determination on the eigenenergy of each phonon mode. According to this introduced method to compute the specific heat, the following Figure 7 gives rise to our predicted Cv(T) for Rb24Ga24Sn112 and Cs24Ga24Sn112. Values for the heat capacity of Rb24Ga24Sn112 are always smaller than those of Cs24Ga24Sn112, when the temperature roughly satisfies the condition 100 K < T ≤ 1000 K.
The atomic vibrations of guest atoms produce glass-like low-lying modes, accounting for the “hump” found in the temperature-dependent specific heat profile. Typically, such excess specific heats due to guest “rattling” can be associated with the peak occurring in a temperature range whose center is normally 10–100 times smaller than the Debye temperature [37]. Previous reports identified the characteristic humps that were determined around T ~ 5 K, in the case of Type-I β-Rb8Ga16Sn30 (β-RGS) and Type-I Sr8Ga16Ge30 [38,39,40,41,42,43] compounds. Our computational work presents this form of hump in the Type-II Rb24Ga24Sn112 clathrate, owing primarily to two sorts of isotropic Einstein oscillators that originate from the 28 atom and 20 atom cages, respectively. Making use of the definition of specific heat regarding the Einstein solid, we realize that the characteristic temperature of the Rb oscillator encapsulated in the 28 atom cage is calculated to be θE1 = 36.82 K, while the characteristic temperature of the Rb oscillator trapped in the 20 atom cage is θE2 = 84.00 K.
Besides the isotropic ADP study in Figure 6, we have found that excess specific heat results can be attributed to the introduced isotropic Einstein oscillators. Consequently, three characteristic temperature regions in Figure 8 might be of great interest. Initially, the low-T (<6 K) behavior of CV(T) is numerically governed by the relation [44,45,46]: CV(T) ≈ αT1+δ + βT3. The usage of the three coefficients, α, β, δ, will be discussed later. Second, excess specific heat at about T ~ 10 K is thought to correlate with the “rattling” modes in the presence of guest vibration in Sn20 and Sn28 cages. This excess hump in the CV/T3 curve is characterized to remain nearly identical to the Einstein peaks. The third temperature-regime is the one for which CV(T) approaches a vanishingly small magnitude above about 70 K. Additional intriguing thermal properties involving inspection of the lattice dynamics quantity CP(T) − CV(T) = KTVα2 agree with the previous work [47,48], while CP(T) pertains to the heat capacity at a constant pressure. Here, K is the bulk modulus and α denotes the volume coefficient of thermal expansion.
It is necessary to point out that the thermal influence of CV(T) arising from conduction electrons is neglected for Rb24Ga24Sn112 in our computation. In Figure 8, DFT data (black dotted line) plotted as CV(T)/T3 vs. T show a remarkable hump appearing around T ~ 10 K. This observed unusual peak was simulated by displaying the CV(T) calculated from encapsulated “rattlers”, according to the theory of the isotropic Einstein oscillator. Therefore, our work, shown by the red dashed and blue dashed curves, was deduced via the fitting parameters given by the characteristic temperature of the Einstein oscillator (θE1, θE2) and the number of oscillators (NE1, NE2)). Here, the obtained parameters were NE1 = 2 guest oscillators in 28 atom cages and NE2 = 4 rattler oscillators inside 20 atom cages of primitive unit cells (consisting of 34 atoms). Apparently, such excess specific heat is shown to be more heavily dependent on Rb vibration in the hexakaidecahedron, rather than dodecahedron, cage.
Figure 9 shows the low-T (<6 K) specific heat (green triangle from DFT calculation) fitted by the theoretical relation: CV(T) ≈ αT1+δ + βT3. This scaled form of CV(T) was initially proposed by Zeller and Pohl [44], who focused on silica glass. Here, we note that the deviation from a linear dependence on temperature of the first term (αT1+δ) is given by a power exponent increment, δ. However, αT1+δ has no connection with the specific heat contributed by the conduction electron, γT. Therefore, our first-principle generated CV(T) was fitted by taking the pristine silica glass model (SGM) into account. Then, the fitting process lead to α ≈ 7.075 mJ·mol−1·K2, β ≈ 15.886 mJ·mol−1·K−4 along with the small exponential factor determined as δ ≈ 0.21. As a whole, our predicted values (α, β parameters) were relatively small when compared to the measurement of Type-I Ba6Ga16Sn30 (β-BGS) showing α ≈ 30 mJ·mol−1·K2, β ≈ 50 mJ·mol−1·K−4 and δ ≈ 0.20 [41,42].
Figure 10 displays the evaluated lattice’s thermal conductivity κL for Rb24Ga24Sn112 in terms of the guest–host distance, which is approximated as the phonon’s mean free path. Using the average value for a range of mean free paths, our computation was guided by the simple kinetic model κL = (1/3)CVvsλ, which yields 0.25 W/m/K ≤ κL ≤ 1.08 W/m/K in the vicinity of room temperature when the mean free path spans over the range of 2 and 8 Å. It is mentioned that λ is the average mean free path while vs. manifests the sound speed of acoustic phonon bands. On the other hand, previous work predicted a κL of the Type-I tin-containing compound K8Al8Sn38, which turned out to be 1.2 W/m/K [49]. In our work, λ has been approximated as a covalent-like Rb–Sn bond whose length ranges from 4.56 Å to 4.76 Å in terms of the DFT computation. The small box (formed by blue solid lines) in the shaded area of Figure 10 represents the coverage of the possible bond length range (4.56 Å < λ < 4.76 Å). Therefore, the correct range of lattice thermal conductivity should be κL(λ = 4.56 Å) ≈ 0.60 W/m/K < κL < κL(λ = 4.76 Å) ≈ 0.64 W/m/K. It is also expected that the formation of those covalent-like Rb–Sn bonds links to the coupling of the Rb “rattling” modes with host framework vibrations. On the whole, our predicted κL at RT seems to be around 0.62 W/m/K for Rb24Ga24Sn112. Previous work reports the figure of merit of Type-I β-Ba8Ga16Sn30 (abbreviated as β-BGS) to be smaller than 0.6 [41]. Meanwhile, the lattice thermal conductivity of this β-BGS remains between 0.6 and 0.7 W/K/m at T = 150 K, approximately. According to our DFT-calculated heat capacity in Figure 7, we found that the CV at T = 150 K for Rb24Ga24Sn112 is about half as much as the value at T = 300 K. Assuming the existence of a temperature-independent phonon mean free path and the sound velocity of acoustic phonons, we concluded that the lattice thermal conductivity at T = 150 K is nearly half as much as the value at T = 300 K, in the context of Rb24Ga24Sn112. Therefore, our predicted κL at T = 150 K is about equal to 0.31 W/K·m. Considering that the figure of merit is approximately proportional to the inverse of the lattice’s thermal conductivity while taking the negligible impact from the Seeback coefficient, electrical conductivity, and electronic thermal conductivity into account, the lower lattice thermal conductivity always indicates a larger ZT value. Consequently, Rb24Ga24Sn112 shows better thermoelectric performance when compared to the β-Ba8Ga16Sn30 and several other Sn-based clathrate compounds, including Type-VIII Ba8Ga16Sn30, or Type-I n-type (or p-type) Ba8Ga16Sn30 [40,41,50].

3. Computational Approach

Utilizing first-principles code named the Vienna ab initio simulation package (VASP) [51], we performed our research work to study the structural and vibrational properties of Sn-based Type-II clathrate compounds. Normally, the entire computation process makes use of ultra-soft pesudopotential (US-PP) [52,53] or generalized gradient approximation (GGA) PW91 to describe the correlation found between electron and ion cores, as well as the functional exchange-correlation, with respect to electrons. In contrast to these formalisms, our study is based on an ab initio density functional planewave method, which relies on local density approximation (LDA) and density functional theory (DFT). Additionally, the Celperly-Alder functional [54] is used to explain electron correlation effects and the electron exchange-correlation energy. This technique has been tested a few times on various clathrate systems [30,55,56,57], in order to ensure the accuracy of computational results while achieving a reduced time load for calculation. The cutoff-energy of the plane-wave basis is selected to be 300 eV, when the 4 × 4 × 4 Monkhorst-pack k-point grid [58] confined within the first Brillouin zone (BZ) is used to establish the cutting grid for acquiring the structural features. The detailed pseudopotential contains information on the valence electrons of the tin framework atom (denoted by 5s25p2), while the rest of the electrons in the Sn atom are treated as core electrons. Similarly, the valence electron configurations of Cs and Rb, as well as the Ga atom, read 5p66s1, 4p65s1, and 4s24p1.
To determine the optimized configuration, a fixed unit cell volume is selected at the beginning. Next, optimizing the ionic positions is done by minimizing the total binding energy with a conjugate gradient (CG) algorithm that uses atomic forces. This energy minimization process is repeated for a range of unit cell volumes, from which an equation of state (EOS) [59] and an equilibrium state configuration possessing the least energy can be established. Then, the internal coordinates are optimized at the minimum energy configuration. Once the equilibrium geometry has been fully gained, knowledge of the structural parameters corresponding to the minimum energy state can guide us to continuously explore the volume-dependent binding energy, the lattice constant, the dispersion relationships, and the heat capacity trends in a quantitative manner. For binary CsxSn136 (0 ≤ x ≤ 24) and ternary Rb24Ga24Sn112, the LDA-determined energy vs. volume curve, which arises from the above procedure, has been fit with the Birch-Murnaghan EOS. This fitting process determines the minimum binding energy E0, the volume V0 at optimized geometry, the equilibrium bulk modulus K0, and the pressure (P) derivative of K0 written as dK0/dP.
It is noticed that three inequilibrium sites exist for framework atoms that seek stable locations, namely, 8a, 32e and 96g. It is also known that selecting a different number of Group-III atom (e.g., Ga) that occupy some places in these Wyckoff positions may have an impact on tuning the physical properties of those Sn-based ternary clathrates. Generally, the balance between the number of excess electrons resulting from alkaline atoms and the number of compensation donors arising from Ga atoms needs to be carefully considered, when taking the “semiconducting compound” idea into account.
According to harmonic approximation, the diagonalization process applied to the wavevector (q)-dependent dynamic matrix D(q) produces eigenvalues for vibrational modes, including acoustic phonons and optic ones. Determination of the dynamic matrix needs a second derivative of the LDA-binding energy with respect to atomic displacement. Normally, we utilize the finite atomic displacement parameter U0 to compute these forces, which linearly scale with the displacements. Both +U0 and −U0 variables are used to calculate D(q). Two wavevector-dependent phonon modes are examined before gaining a complete D(q) in the vicinity of the Γ point confined within the q-space. Firstly, D(q) is constructed at the gamma (Γ) point [q = (0,0,0)]. On the other hand, D(q) can still be calculated for nonzero wavevectors if we assumed that the matrix elements of D(q) vanish for atoms separated by a distance greater than the third nearest neighbor distance [60,61]. Additionally, group theory [17] has been used to reduce the computational effort when determining the energies and forces during structural optimization. It also reduces the effort of examining vibrational dispersion relationships. Considering the symmetries found with the 102 (3 × 34) Γ-point modes that exist in the GaxSn136−x framework (x = 0, 24), 42 distinct frequencies that possess symmetries among them are of crucial importance.

4. Conclusions

Previously, theoretical investigations of the properties of the Type-II clathrate CsxSi136 (0 < x ≤ 24) have found that, as the composition x increases, the Si136 lattice exhibits framework contraction upon filling (0 < x ≤ 8), followed by an expansion of the unit cell volume (9 < x ≤ 24). However, tuning all the host atoms from Si to Sn cause no lattice contraction to occur according to our determination on composition-dependent lattice parameters in CsxSn136 (0 < x ≤ 24). In addition to these geometry characteristics, the calculated behavior of the guest’s “rattling” phonon modes, along with the vibrational density of states in these materials, are also presented. Specifically, the determined frequency describing “localized” flat optic bands emerging from Cs guest seems to decrease in a slight manner as composition x is elevated from 12 to 16 to 20 from the viewpoint of dispersion relations. Meanwhile, an extremely slight decreasing behavior relating to the VDOS peak of Cs rattling modes is also captured as the number of guests who occupy Sn20 cages is elevated from 4 to 8 to 12 in the presence of enlarged unit cell (136 atoms) geometry, assuming all Sn28 cages are completely filled. In contrast to the predicted low-lying modes of guests appearing in a 28 atom cage of the CsxSn136 (x = 12, 16, 20) system, an upshift of the low-lying optic modes (centered around 25 cm−1) is found for vibration of the light guest (“rattler”) Rb in a 28 atom cage of Rb24Ga24Sn112. It is also expected that the guest vibration in the “over-sized” cage is mainly responsible for effectively lowering the lattice’s thermal conductivity. This is due to the fact that the Einstein-peak-like hump arising from the CV/T3 vs. T relation is shown to originate mainly from the contribution of the guest oscillator in a 28 atom cage rather than that in the 20 atom cage of Rb24Ga24Sn112. Within the kinetic theory approximation, our lattice’s predicted thermal conductivity in Rb24Ga24Sn112 is approximately equal to 0.62 W/m/K at room temperature.

Author Contributions

Conceptualization, D.X.; methodology, D.X.; software, C.W.M.; validation, D.X., C.W.M.; formal analysis, D.X.; investigation, D.X.; resources, C.W.M.; data curation, D.X.; writing—original draft preparation, D.X.; writing—review and editing, D.X.; visualization, D.X.; supervision, C.W.M.; project administration, C.W.M.

Funding

This research received no external funding.

Acknowledgments

We would like to acknowledge grateful discussions with Mahdi Sanati (Texas Tech University) about the use of VASP. We appreciate many hours of computing time at the High-Performance Computing Center of Texas Tech University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Prokofiev, A.; Sidorenko, A.; Hradil, K.; Ikeda, M.; Svagera, R.; Winkler, H.; Neumaier, K.; Paschen, S. Thermopower enhancement by encapsulating cerium in clathrate cages. Nat. Mater. 2013, 12, 1096–1101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Kleinke, H. New bulk Materials for Thermoelectric Power Generation: Clathrates and Complex Antimonides. Chem. Mater. 2010, 22, 604–611. [Google Scholar] [CrossRef]
  3. Seeback. Magnetische Polarisation der Metalle und Erze durch Temperatur-Differenz; Abhandlungen der Königlichen Akademie der Wissenschaften zu Berlin: Berlin, Germany, 1822; pp. 265–373. [Google Scholar]
  4. Saiga, Y.; Suekuni, K.; Deng, S.K.; Yamamoto, T.; Kono, Y.; Ohya, N.; Takabatake, T. Optimization of Thermoelectric Properties of Type-VIII Clathrate Ba8Ga16Sn30 by Carrier Tuning. J. Alloys Compd. 2010, 507, 1–5. [Google Scholar] [CrossRef]
  5. Christensen, M.; Johnsen, S.; Juranyi, F.; Iversen, B.B. Clathrate guest atoms under pressure. J. Appl. Phys. 2009, 105, 073508. [Google Scholar] [CrossRef] [Green Version]
  6. Yanagisawa, T.; Tateiwa, N.; Mayama, T.; Saito, H.; Hidaka, H.; Amitsuka, H.; Haga, Y.; Nemoto, Y.; Goto, T. Ultrasonic Measurements on the Cage-Structured Clathrate Compound U3Pd20Si6. J. Phys. Soc. Jpn. 2011, 80 (Suppl. A), SA105. [Google Scholar] [CrossRef]
  7. Khabibullin, A.R.; Huan, T.D.; Nolas, G.; Woods, L.M. Cage disorder and gas encapsulation as routes to tailor properties of inorganic clathrates. Acta Mater. 2017, 131, 475–481. [Google Scholar] [CrossRef] [Green Version]
  8. Slack, G.A. CRC Handbook of Thermoelectrics; Rowe, D.M., Ed.; CRC: Boca Raton, FL, USA, 1995; pp. 407–440. [Google Scholar]
  9. Slack, G.A. Proceedings of Thermoelectric Materials: New Directions and Approaches; MRS Symposia Proceedings No. 478; Materials Research Society: Pittsburgh, PA, USA, 1997; p. 47. [Google Scholar]
  10. Suekuni, K.; Takasu, T.; Hasegawa, T.; Ogita, N.; Udagawa, M.; Avila, M.A.; Takabatake, T. Off-center rattling modes and glasslike thermal conductivity in the type-I clathrate Ba8Ga16Sn30. Phys. Rev. B 2010, 81, 205207. [Google Scholar] [CrossRef]
  11. Tadano, T.; Gohda, Y.; Tsuneyuki, S. Impact of Rattlers on Thermal Conductivity of a Thermoelectric Clathrate: A First-Principles Study. Phys. Rev. Lett. 2015, 114, 095501. [Google Scholar] [CrossRef]
  12. Nolas, G.S.; Cohn, J.L.; Slack, G.A.; Schujman, S.B. Semiconducting Ge clathrates: Promising candidates for thermoelectric applications. Appl. Phys. Lett. 1998, 73, 178. [Google Scholar] [CrossRef]
  13. Dong, J.; Sankey, O.F.; Myles, C.W. Theoretical Study of the Lattice Thermal Conductivity in Ge Framework Semiconductors. Phys. Rev. Lett. 2001, 86, 2361. [Google Scholar] [CrossRef]
  14. Beekman, M.; Nolas, G.S. Inorganic clathrate-II materials of group 14: Synthetic routes and physical properties. J. Mater. Chem. 2008, 18, 842–851. [Google Scholar] [CrossRef]
  15. Guloy, A.M.; Ramlau, R.; Tang, Z.; Schnelle, W.; Baitinger, M.; Grin, Y. A guest-free germanium clathrate. Nature 2006, 43, 320–323. [Google Scholar] [CrossRef] [PubMed]
  16. Biswas, K.; Myles, C.W.; Sanati, M.; Nolas, G.S. Thermal Properties of Guest-Free Si136 and Ge136 Clathrates. J. Appl. Phys. 2008, 104, 033535. [Google Scholar] [CrossRef]
  17. Myles, C.W.; Dong, J.; Sankey, O.F.; Nolas, G.S. Vibrational Properties of Tin Clathrate Materials. Phys. Rev. B 2002, 65, 235208. [Google Scholar] [CrossRef]
  18. Nolas, G.S.; Slack, G.A.; Schujman, S.B. Semiconductor clathrates: A phonon glass electron crystal material with potential for thermoelectric applications. Semicond. Semimet. 2000, 69, 255–300. [Google Scholar]
  19. Nolas, G.S.; Chakoumakos, B.C.; Mahieu, B.; Long, G.J.; Weakley, T.J.R. Structural Characterization and Thermal Conductivity of Type-I Tin Clathrates. Chem. Mater. 2000, 12, 1947–1953. [Google Scholar] [CrossRef]
  20. Qiu, L.; Swainson, I.P.; Nolas, G.S.; White, M.A. Structure, thermal, and transport properties of the clathrates Sr8Zn8Ge38, Sr8Ga16Ge30, and Ba8Ga16Si30. Phys. Rev. B 2004, 70, 035208. [Google Scholar] [CrossRef]
  21. Wilkinson, A.P.; Lind, C.; Young, R.A.; Shastri, S.D.; Lee, P.L.; Nolas, G.S. Preparation, Transport Properties, and Structure Analysis by Resonant X-ray Scattering of the Type I Clathrate Cs8Cd4Sn42. Chem. Mater. 2002, 14, 1300–1305. [Google Scholar] [CrossRef]
  22. Dubois, F.; Fässler, T.F. Ordering of Vacancies in Type-I Tin Clathrate: Superstructure of Rb8Sn442. J. Am. Chem. Soc. 2005, 127, 3264–3265. [Google Scholar] [CrossRef]
  23. Schäfer, M.C.; Bobev, S. Tin Clathrates with the Type II Structure. J. Am. Chem. Soc. 2013, 135, 1696–1699. [Google Scholar] [CrossRef]
  24. Wei, K.; Zeng, X.; Tritt, T.M.; Khabibullin, A.R.; Woods, L.M.; Nolas, N.S. Structure and Transport Properties of Dense Polycrystalline Clathrate-II (K,Ba)16(Ga,Sn)136 Synthesized by a New Approach Employing SPS. Materials 2016, 9, 732. [Google Scholar] [CrossRef] [PubMed]
  25. Huo, D.; Sakata, T.; Sasakawa, T.; Avila, M.A.; Tsubota, M.; Iga, F.; Fukuoka, H.; Yamanaka, S.; Aoyagi, S.; Takabatake, T. Structural, transport, and thermal properties of the single-crystalline type-VIII clathrate Ba8Ga16Sn30. Phys. Rev. B 2005, 71, 075113. [Google Scholar] [CrossRef]
  26. Kono, Y.; Ohya, N.; Taguchi, T.; Suekuni, K.; Takabatake, T.; Yamamoto, S.; Akai, K. First-principles study of type-I and type-VIII Ba8Ga16Sn30 clathrates. J. Appl. Phys. 2010, 107, 123720. [Google Scholar] [CrossRef]
  27. Xue, D.; Myles, C.W.; Higgins, C. Effect of Guest Atom Composition on the Structural and Vibrational Properties of the Type II Clthrate-Based Materials AxSi136, AxGe136 and AxSn136 (A = Na, K, Rb, Cs; 0 ≤ x ≤ 24). Materials 2016, 9, 691. [Google Scholar] [CrossRef] [PubMed]
  28. Zintl, E.; Brauer, G. Über die Valenzelektronenregel und die Atomradien unedler Metalle in Legierungen. Z. Für Phys. Chem. 1933, 20, 245–271. [Google Scholar] [CrossRef]
  29. Biswas, K.; Myles, C.W. Density-functional investigation of Na16A8Ge136 (A = Rb, Cs) clathrates. J. Phys. Condens. Matter 2007, 19, 466206. [Google Scholar] [CrossRef]
  30. Biswas, K.; Myles, C.W. Electronic and Vibrational Properties of Framework-Substituted Type II Si Clathrates. Phys. Rev. B 2007, 75, 245205. [Google Scholar] [CrossRef]
  31. Beekman, M.; Hermann, R.P.; Möchel, A.; Juranyi, F.; Nolas, G.S. A study of low-energy guest phonon modes in clathrate-II NaxSi136 (x = 3, 23, and 24). J. Phys. Condens. Matter 2010, 22, 355401. [Google Scholar] [CrossRef]
  32. Dong, J.; Sankey, O.F.; Ramachandran, G.K.; McMillan, P. Chemical trends of the rattling phonon modes in alloyed germanium clathrates. J. Appl. Phys. 2000, 87, 7726. [Google Scholar] [CrossRef]
  33. Machon, D.; Toulemonde, P.; McMillan, P.F.; Amboage, M.; Munoz, A.; Rodriguez-Hernandez, P.; Miguel, A.S. High-pressure phase transformations, pressure-induced amorphization, and polyamorphic transition of the clathrate Rb6.15Si46. Phys. Rev. B 2009, 79, 184101. [Google Scholar] [CrossRef]
  34. Imai, Y.; Imai, M. Chemical trends of the band gaps of idealized crystal of semiconducting silicon clathrates, M8Si38A8 (M = Na, K, Cs, Rb; A = Ga, Al, In), predicted by first-principle pseudopotential calculations. J. Alloys Compd. 2011, 509, 3924–3930. [Google Scholar] [CrossRef]
  35. Mutka, H.; Koza, M.M.; Johnson, M.R.; Hiroi, Z.; Yamaura, J.; Nagao, Y. Generalized density-of-states ad anharmonicity of the low-energy phonon bands from coherent inelastic neutron scattering response in the pyrochlore osmates AOs2O6 (A = K, Rb, Cs). Phys. Rev. B 2008, 78, 104307. [Google Scholar] [CrossRef]
  36. Patschke, R.; Zhang, X.; Singh, D.; Schindler, J.; Kannewurf, C.R.; Lowhorn, N.; Tritt, T.; Nolas, G.S.; Kanatzidis, M.G. Thermoelectric Properties and Electronic Structure of the Cage Compounds A2BaCu8Te10 (A = K, Rb, Cs): System with Low Thermal Conductivity. Chem. Mater. 2001, 13, 612–621. [Google Scholar] [CrossRef]
  37. Takabatake, T.; Suekuni, K.; Nakayama, T.; Kaneshita, E. Phonon-glass electron-crystal thermoelectric clathrates: Experiments and theory. Rev. Mod. Phys. 2014, 86, 669. [Google Scholar] [CrossRef]
  38. Umeo, K.; Avila, M.A.; Sakata, T.; Suekuni, K.; Takabatake, T. Probing Glasslike Excitations in Single-Crystalline Sr8Ga16Ge30 by Specific Heat and Thermal Conductivity. J. Phys. Soc. Jpn. 2005, 74, 2145–2148. [Google Scholar] [CrossRef]
  39. Suekuni, K.; Avila, M.A.; Umeo, K.; Takabatake, T. Cage-size control of guest vibration and thermal conductivity in Sr8Ga16Si30−xGex. Phys. Rev. B 2007, 75, 195210. [Google Scholar] [CrossRef]
  40. Avila, M.A.; Suekuni, K.; Umeo, K.; Fukuoka, H.; Yamanaka, S.; Takabatake, T. Ba8Ga16Sn30 with type-I clathrate structure: Drastic suppression of heat conduction. Appl. Phys. 2008, 92, 041901. [Google Scholar] [CrossRef]
  41. Suekuni, K.; Avila, M.A.; Umeo, K.; Fukuoka, H.; Yamanaka, S.; Nakagawa, T.; Takabatake, T. Simulations structure and carrier tuning of dimorphic clathrate Ba8Ga16Sn30. Phys. Rev. B 2008, 77, 235119. [Google Scholar] [CrossRef]
  42. Suekuni, K.; Yamamoto, S.; Avila, M.A.; Takabatake, T. Universal Relation between Guest Free Space and Lattice Thermal Conductivity Reduction by Anharmonic Rattling in Type-I Clathrates. J. Phys. Soc. Jpn. 2008, 77, 61–66. [Google Scholar] [CrossRef] [Green Version]
  43. Xu, J.; Tang, J.; Sato, K.; Tanabe, Y.; Heguri, S.; Miyasaka, H.; Yamashita, M.; Tanigaki, K. Heat Capacity Study on Anharmonicity in Ae8Ga16Ge30 (Ae = Sr and Ba). J. Electron. Mater. 2011, 40, 879. [Google Scholar] [CrossRef]
  44. Zeller, R.C.; Pohl, R.O. Thermal Conductivity and Specific Heat of Noncrystalline Solids. Phys. Rev. B 1971, 4, 2029. [Google Scholar] [CrossRef]
  45. Stephens, R.B. Low-Temperature Specific Heat and Thermal Conductivity of Noncrystalline Dielectric Solids. Phys. Rev. B 1973, 8, 2896. [Google Scholar] [CrossRef]
  46. Freeman, J.J.; Anderson, A.C. Thermal conductivity of amorphous solids. Phys. Rev. B 1986, 34, 5684. [Google Scholar] [CrossRef]
  47. Qiu, L.; White, M.A.; Li, Z.; Tse, J.S.; Ratcliffe, C.I.; Tulk, C.A.; Dong, J.; Sankey, O.F. Thermal and lattice dynamical properties of Na8Si46 clathrate. Phys. Rev. B 2001, 64, 024303. [Google Scholar] [CrossRef]
  48. Nenghabi, E.N.; Myles, C.W. First Principles Calculations of the Vibrational and Thermal Properties of the Type-I Clathrates Ba8Ga16SixGe30−x and Sr8Ga16SixGe30−x. Phys. Rev. B 2008, 78, 195202. [Google Scholar] [CrossRef]
  49. Hayashi, M.; Kishimoto, K.; Akai, K.; Asada, H.; Kishio, K.; Koyanagi, T. Preparation and thermoelectric properties of sintered n-type K8M8Sn38 (M = Al, Ga and In) with the type-I clathrate structure. J. Phys. D Appl. Phys. 2012, 45, 45. [Google Scholar] [CrossRef]
  50. Avilla, M.A.; Suekuni, K.; Umeo, K.; Fukuoka, H.; Yamanaka, S.; Takabatake, T. Glasslike versus crystalline thermal conductivity in carrier-tuned Ba8Ga16X30 clathrates (X = Ge, Sn). Phys. Rev. B 2006, 74, 125109. [Google Scholar] [CrossRef]
  51. Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  52. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892. [Google Scholar] [CrossRef]
  53. Laasonen, K.; Car, R.; Lee, C.; Vanderbilt, D. Implementation of ultrasoft pseudopotentials in ab initio molecular dynamics. Phys. Rev. B 1991, 43, 6796. [Google Scholar] [CrossRef]
  54. Ceperley, D.M.; Alder, B.J. Ground State of the Electron Gas by a Stochastic Method. Phys. Rev. Lett. 1980, 45, 566. [Google Scholar] [CrossRef]
  55. Biswas, K.; Myles, C.W. Electronic Structure of the Na16Rb8Si136 and K16Rb8Si136 Clathrates. Phys. Rev. B 2006, 74, 115113. [Google Scholar] [CrossRef]
  56. Myles, C.W.; Biaswas, K.; Nenghabi, E.N. Rattling “Guest” Impurities o in Si and Ge Clathrate Semiconductors. Phys. B 2007, 401, 695. [Google Scholar] [CrossRef]
  57. Myles, C.W.; Dong, J.; Sankey, O.F. Structural and Electronic Properties of Tin Clathrate Materials. Phys. Rev. B 2001, 64, 165202. [Google Scholar] [CrossRef]
  58. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  59. Birch, F. Elasticity and constitution of the Earth’s interior. J. Geophys. Res. 1952, 57, 227. [Google Scholar] [CrossRef]
  60. Dong, J.; Sankey, O.F.; Kern, G. Theoretical study of the vibrational modes and their pressure dependence in the pure clthrate-II silicon framework. Phys. Rev. B 1999, 60, 950. [Google Scholar] [CrossRef]
  61. Dong, J.; Sankey, O.F. Theoretical study of two expanded phases of crystalline germanium: Clathrate-I and clathrate-II. J. Phys. Condens. Matter 1999, 11, 6129. [Google Scholar] [CrossRef]
Figure 1. Predicted x dependence of the lattice constant for the Type-II clathrate-based compounds: AxSn136 (A = Na, K, Cs; 0 ≤ x ≤ 24).
Figure 1. Predicted x dependence of the lattice constant for the Type-II clathrate-based compounds: AxSn136 (A = Na, K, Cs; 0 ≤ x ≤ 24).
Inorganics 07 00074 g001
Figure 2. Predicted dispersion relations of (a) Cs12Sn136, (b) Cs16Sn136 and (c) Cs20Sn136 with the usage of the Vienna ab initio simulation package (VASP).
Figure 2. Predicted dispersion relations of (a) Cs12Sn136, (b) Cs16Sn136 and (c) Cs20Sn136 with the usage of the Vienna ab initio simulation package (VASP).
Inorganics 07 00074 g002
Figure 3. Vibrational density of states (VDOS) given in an arbitrary unit as a function of frequency for Cs12Sn136 (black dashed line), Cs16Sn136 (red dashed line), and Cs20Sn136 (green dashed line). Illustration of a detailed VDOS in the frequency range of 0 and 25 cm−1 (inset figure).
Figure 3. Vibrational density of states (VDOS) given in an arbitrary unit as a function of frequency for Cs12Sn136 (black dashed line), Cs16Sn136 (red dashed line), and Cs20Sn136 (green dashed line). Illustration of a detailed VDOS in the frequency range of 0 and 25 cm−1 (inset figure).
Inorganics 07 00074 g003
Figure 4. Dispersion relation of Rb24Ga24Sn112 that was obtained via VASP.
Figure 4. Dispersion relation of Rb24Ga24Sn112 that was obtained via VASP.
Inorganics 07 00074 g004
Figure 5. Dispersion relation of Cs24Ga24Sn112 that was obtained via VASP.
Figure 5. Dispersion relation of Cs24Ga24Sn112 that was obtained via VASP.
Inorganics 07 00074 g005
Figure 6. Predicted theoretical values of the isotropic atomic displacement parameters (Uiso) for the Rb vibration in a 28 atom cage (Rb(1)) and the Rb vibration in a 20 atom cage (Rb(2)).
Figure 6. Predicted theoretical values of the isotropic atomic displacement parameters (Uiso) for the Rb vibration in a 28 atom cage (Rb(1)) and the Rb vibration in a 20 atom cage (Rb(2)).
Inorganics 07 00074 g006
Figure 7. Predicted specific heat CV(T) as a function of temperature for Rb24Ga24Sn112 (black square) and Cs24Ga24Sn112 (red ball).
Figure 7. Predicted specific heat CV(T) as a function of temperature for Rb24Ga24Sn112 (black square) and Cs24Ga24Sn112 (red ball).
Inorganics 07 00074 g007
Figure 8. The specific heat data plotted as a CV(T)/T3 vs. T diagram (black dotted line) using density functional theory (DFT) code. The red dashed and blue dashed lines are obtained via the Einstein model’s contribution to the Rb vibration in the 28 atom and 20 atom cages. Temperature is given in a logarithmic scale.
Figure 8. The specific heat data plotted as a CV(T)/T3 vs. T diagram (black dotted line) using density functional theory (DFT) code. The red dashed and blue dashed lines are obtained via the Einstein model’s contribution to the Rb vibration in the 28 atom and 20 atom cages. Temperature is given in a logarithmic scale.
Inorganics 07 00074 g008
Figure 9. The silica glass model (SGM) fit of the DFT-determined CV/T curve of the Rb24Ga24Sn112 in the low-temperature range (0 < T < 6 K).
Figure 9. The silica glass model (SGM) fit of the DFT-determined CV/T curve of the Rb24Ga24Sn112 in the low-temperature range (0 < T < 6 K).
Inorganics 07 00074 g009
Figure 10. Lattice thermal conductivity based on Rb24Ga24Sn112 in terms of the average mean free path λ. The shaded region includes the possible range of Rb–Sn distance, described graphically in the structure (inset figure). The blue box denotes the possible coverage of the mean free path (4.56 Å < λ < 4.76 Å; 1 Å = 10−10 m).
Figure 10. Lattice thermal conductivity based on Rb24Ga24Sn112 in terms of the average mean free path λ. The shaded region includes the possible range of Rb–Sn distance, described graphically in the structure (inset figure). The blue box denotes the possible coverage of the mean free path (4.56 Å < λ < 4.76 Å; 1 Å = 10−10 m).
Inorganics 07 00074 g010

Share and Cite

MDPI and ACS Style

Zahid, H.; Xue, D.; Myles, C.W. First Principles Study of the Vibrational and Thermal Properties of Sn-Based Type II Clathrates, CsxSn136 (0 ≤ x ≤ 24) and Rb24Ga24Sn112. Inorganics 2019, 7, 74. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7060074

AMA Style

Zahid H, Xue D, Myles CW. First Principles Study of the Vibrational and Thermal Properties of Sn-Based Type II Clathrates, CsxSn136 (0 ≤ x ≤ 24) and Rb24Ga24Sn112. Inorganics. 2019; 7(6):74. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7060074

Chicago/Turabian Style

Zahid, Hadeel, Dong Xue, and Charles W. Myles. 2019. "First Principles Study of the Vibrational and Thermal Properties of Sn-Based Type II Clathrates, CsxSn136 (0 ≤ x ≤ 24) and Rb24Ga24Sn112" Inorganics 7, no. 6: 74. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7060074

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop