Nagendra Singh Chauhana and
Takao Mori*ab
aInternational Center for Materials Nanoarchitectonics (WPI-MANA), National Institute for Materials Science (NIMS), Namiki 1-1, Tsukuba, 305-0047, Japan. E-mail: mori.takao@nims.go.jp; Fax: +81-29-851-6280; Tel: +81-29-860-4323
bGraduate School of Pure and Applied Sciences, University of Tsukuba, 1-1-1 Tennodai, Tsukuba, Ibaraki 305-8573, Japan
First published on 18th November 2024
The off-centering phenomenon manifests as locally distorted configurations with broken symmetry in a crystal structure due to the displacement of constituent atoms from their ideal coordination centers within the lattice. In contrast to the anticipated formation of anionic solid solutions of Mg3(Sb1−x−yBixGey)2, herein we report β-Mg3(Sb, Bi)2 based superionic phase formation (space group – Ia, 206) with off-centering of the dominant trigonal α-Mg3(Sb, Bi)2 phase and segregation of nanophase Mg3Ge upon equiatomic (Bi, Ge) alloying. The discordant nature of Ge is unveiled within the layered α-Mg3(Sb, Bi)2 structure and is assessed employing (3 + 1) dimensional superspace to reveal an off-centering (dz) along the z direction for the constituent atoms in the range of ±0–0.02 Å. The (Bi, Ge) alloying results in favourable tuning of the desired p-type conduction for attaining higher power factors by band engineering and synergistic reduction of lattice thermal conductivity. The stable superionic polymorph co-existing in an anionic solid solution of Mg3(Sb, Bi)2 provides a renewed basis for understanding the crystal structure and its transformation in CaAl2Si2-type Zintl compounds.
The evaluated Mg–Sb phase diagram (Fig. 1) based on thermal analysis suggests the occurrence of a phase transformation from α-Mg3Sb2 to β-Mg3Sb2 (Ia, 206) at high temperature (∼1200 K), prior to melting congruently at ∼1500 K.16–18 However, studies undertaken on the Mg–Sb binary system largely show the realisation of only α-Mg3Sb2 experimentally, thus suggesting its retention even at high temperature.14 Presumably the high-temperature β-phase exists only above ∼1200 K for Mg3Sb2, which poses a synthesis challenge due to the high reactivity and volatility of Mg atoms at elevated temperatures. Owing to the inherent difficulty in synthesizing high-quality β-phase Mg3Sb2 crystals, their structural description has thus remained unclear so far. Nonetheless, the high-temperature β-phase for the structural analogue Mg3Bi2 as evaluated by Barnes et al.19 is a disordered cubic structure (Imm, 229), which demonstrates superionic properties.
Fig. 1 The Mg–Sb phase diagram near Mg3Sb2 indicating the phase transformation of α-Mg3Sb2 (Pm1, 164) ↔ β-Mg3Sb2 (Ia, 206).16–18 |
The structural flexibility and rich chemistry enabled by the Zintl–Klemm concept have allowed tuning of the coordination environment of Mg and Sb atoms (or their substitutes) by incorporating different dopants at the anionic and cationic sites.14,20 In this work, we report the implication of co-substituting equiatomic (Bi, Ge) in Mg3(Sb1−x−yBixGey)2 based anionic solid solutions, to achieve higher power factors by band engineering for p-type conduction and synergistic reduction of lattice thermal conductivity. As a discordant dopant, Ge atoms induce local distortions and broken symmetry within the lattice, effectively enhancing phonon scattering. This mechanism significantly reduces lattice thermal conductivity, contributing to an improved thermoelectric figure of merit (zT) of 0.3 (±0.1) at 673 K. Polycrystalline Mg3(Sb1−x−yBixGey)2 materials with (0 < x < 0.1) were prepared by a high-energy milling process, followed by spark milling sintering. Upon equiatomic (Bi, Ge) alloying in the anionic framework [Mg2Sb2]2− and high temperature (∼1023 K) sintering, the formation of a high-temperature β-phase of Mg3(Sb, Bi)2 was observed, with off-centering of constituent atoms within the dominant and co-existing α-Mg3Sb2 phase.
Composition | Space group | Lattice parameters | Volume | R-factorsa | ||
---|---|---|---|---|---|---|
GOF | Rp | Rwp | ||||
a | ||||||
x = 0; y = 0.1 | Pm1 | a = b = 4.5797(3), c = 7.2758(6), α = β = 90°, γ = 120° | 132.16(1) | 2.95 | 6.06 | 8.55 |
Ia | a = b = c = 9.2634(7), α = β = γ = 90° | 794.90(1) | ||||
P | a = b = 12.557(1), c = 4.252(1), α = β = 90°, γ = 120° | 580.76(8) | ||||
x = y = 0.025 | Pm1 | a = b = 4.5840(2), c = 7.2531(4), α = β = 90°, γ = 120° | 131.99(1) | 1.84 | 4.13 | 5.72 |
Ia | a = b = c = 9.3502(4), α = β = γ = 90° | 817.45(5) | ||||
P | a = b = 12.629(2), c = 4.166(1), α = β = 90°, γ = 120° | 575.60(6) | ||||
x = y = 0.05 | Pm1 | a = b = 4.5785(2), c = 7.2494(4), α = β = 90°, γ = 120° | 131.60(1) | 2.38 | 5.06 | 7.25 |
Ia | a = b = c = 9.3504(4), α = β = γ = 90° | 817.52(4) | ||||
P | a = b = 12.627(2), c = 4.162(1), α = β = 90°, γ = 120° | 574.68(5) | ||||
x = y = 0.1 | Pm1 | a = b = 4.5798(2), c = 7.2542(4), α = β = 90°, γ = 120° | 131.77(1) | 2.20 | 4.65 | 6.45 |
Ia | a = b = c = 9.3317(3), α = β = γ = 90° | 812.63(1) | ||||
P | a = b = 12.613(2), c = 4.268(1), α = β = 90°, γ = 120° | 588.09(4) |
Fig. 2(d) shows the observed, calculated, and difference profiles of the powder diffraction profiles of the representative synthesized Mg3(Sb0.95Bi0.025Ge0.025)2 samples measured at 295 K. The short vertical lines below the patterns indicate the peak positions of possible Bragg reflections for α and β-phases of Mg3(Sb, Bi)2, implying their coexistence due to the locally distorted configurations with broken symmetry. The Mg–Sb phase diagram suggests that the β-phase melts above ∼1500 K and has a cubic structure, whose details beside the space group (i.e. Ia, 206) are unknown.16–18 The rapid transition to the superionic state, often occurring at elevated temperatures, presents challenges in determining intermediate states or precise structural details during the transition. This complexity is further aggravated by dynamic disorder, thermal expansion, and increased atomic vibrations, which obscure the characterization of these phases. In the superionic phase, a subset of ions, typically cations, becomes highly mobile, creating a “liquid-like” molten sublattice within the solid framework. Such ion redistribution can dramatically alter the scattering intensity of certain reflections. Moreover, the superionic state is typically characterized by the strongly anharmonic vibrations of the mobile ions, leading to unusual changes in peak intensities and positions. Such changes are often difficult to interpret using conventional crystallographic models, complicating the structural analysis. Thus, superspace formalism is well-suited for describing the complex structures in the synthesized Mg3(Sb1−x−yBixGey)2 samples using the superspace group (Pm1(00g)0s0), which effectively restores the translational symmetry that is lost in the superionic transition due to its incommensurate nature.22–24
The superionic transitioning of a lower-symmetry (Pm1, 164) phase to a higher-symmetry (Ia, 206) phase represents a notable increase in crystal symmetry. This phase transition is consistent with the Mg–Sb phase diagram reported in previous studies, further supporting the observed structural evolution.16–18 Refinements of the powder X-ray diffraction evaluate the cubic β-Mg3Sb2 unit cell (Formula unit = 8) with a lattice parameter i.e. a = b = c ≈ 9.350 Å, having a volume per formula unit of ≈817.5 Å3 as also presented in Table 1. A more isotropic arrangement of atoms in the superionic β-Mg3Sb2 phase likely facilitates easier ion movement in multiple directions, compared to the trigonal structure, which has distinct atomic positions and potentially layered or anisotropic features. It is also noteworthy that the synthesized polycrystals exhibit an irreversible and co-existing β-phase, in contrast to the pressure-induced displacive and reversible (trigonal ↔ monoclinic) phase transition at above 7.8 and 4.0 GPa, respectively, for Mg3Sb2 and Mg3Bi2, as reported previously.25 The constituting superionic phases exhibit a bigger, and highly disordered superionic cell, indicating their complex anionic structural motifs, while demonstrating the structural transition prevalent at elevated temperatures in similar Zintl compounds.
The SE-SEM micrograph shown in Fig. 4(a) reveals a typical morphology of the fractured surface indicating a transgranular mode of fracture (i.e. crack propagation through the grains). The cleavage facets and aligned lamellae patterns were observed largely on the fracture surfaces, which are often associated with plastic deformation within the grains. The average grain size is evaluated to be larger than ∼1 μm, suggesting their contribution towards lowering of electrical resistivity due to reduced grain boundary scattering.28 At higher magnification, the Mg3Ge nanoinclusion segregation within the microstructure becomes distinctly visible. This can be attributed to local compositional fluctuations induced by discordant Ge atoms. The dark-appearing nanoprecipitates, observed embedded within the grains, further indicate the role of Ge in promoting phase separation at the nanoscale. We anticipate Mg3Ge nucleation and segregation from the super-saturated Mg3(Sb, Bi)2 solid solution to be an outcome of limited solid solubility of Ge. The EDX line scan (Fig. 4(c)) for the zoomed-in region confirms the presence of Mg and Ge excess in the same regions corresponding to dark appearing nanoprecipitations embedded in the grains. Similar to the polished surface of the bulk samples, the elemental mapping (Fig. 4(d)) of the fractured surface also confirms the presence of Mg and Ge excess in dark appearing embedded nanoprecipitates. The EDS mapping revealing spatial distribution indicates that all the constituting elements, i.e. Mg, Sb, Bi and Ge, are well distributed uniformly throughout the sample.
Interestingly, the incorporation of Ge in n-type Mg3.2Sb1.49−2xBi0.5Te0.01+xGex has led to the formation of Bi/Ge-rich Janus nanoprecipitates,29 which may possess similarity to the observed secondary phase formation. As understood previously,29 despite negligible mutual solid-state solubilities of Ge and Bi, during sintering (at a sintering temperature of ≈923 K) they become fully miscible through a co-melting process. During crystallization, the α-Mg3(Sb, Bi)2 and Bi–Ge liquid phases compete for formation, which we anticipate has induced Mg3Ge phase formation and a partial displacive structural transition from the trigonal (Pm1) phase to the cubic (Ia) phase during sintering. As the nominal composition of the synthesized Mg3(Sb1−x−yBixGey)2 nanocomposites lies well within the solid solubility range (i.e. x ∼ 0.4) of Bi in Mg3Sb2−xBix compounds,30 the superionic phase formation and observation of Mg3Ge are indicative of the discordant nature of Ge atoms, which exhibit limited solubility with Bi and undergo local co-melting of Bi and Ge during sintering.29 These findings suggest that the synthesized nanocomposites possess a complex microstructure that necessitates high-resolution techniques to fully characterize and understand the origins of these compositional variations. The coexistence of these phases has a significant impact on the microstructure, deformation behaviour, and transport properties of the sintered material, as discussed in the subsequent sections.
As conventional 3-dimensional structural models overlook local and random distortions within the crystal structure, we employ a (3 + 1) dimensional superspace group (Pm1(00γ)0s0) to characterize inherent modulation in the synthesized samples.22–24 Modulations can arise from displacive modulations (atomic positions modulated) or occupational modulations (site occupancies modulated). The superspace approach can model both types by introducing modulation functions for atomic coordinates and occupancies to capture the complex structural variations. Fig. 5(b) reveals the evaluated off-centering in angstroms of Mg(I), Mg(II) and Sb atoms along the z axis with varying alloying content, evaluated upon Rietveld refinement of the dominant α-Mg3Sb2 phase from their ideal lattice positions, i.e., 1a (0, 0, 0), 2d (1/3, 2/3, 0.664), and 2d (1/3, 2/3, 0.225), respectively, against the fourth superspace coordinate t. All the displacements (in angstroms) are periodic in the interval 0 ≤ t ≤ 1.0, considering up to the fourth order of cosine and sine components of the Fourier terms, with an isotropic displacement parameters of each atoms. The off-centering phenomenon in a crystal structure involves the displacement of atoms from their ideal coordination centers, leading to locally distorted configurations and broken symmetry, while maintaining the overall crystallographic symmetry of the material. Interestingly, the displacement occurs only along the z direction, which is likely triggered by the presence of Ge atoms in layered anionic solid solutions of Mg3(Sb1−x−yBixGey)2. This counterintuitive site preference of Ge atoms arises from factors such as ionicity, covalency, and the flexibility of the host structure to accommodate the dopant atoms in energetically favourable sites. The off-centering of the constituting atoms is only allowed along the z direction in all the alloyed compositions, indicative of a low ideal shear strength of ∼1.95 GPa found for Mg3Sb2.32
The overlapping periodicity of displacive modulations in the alloyed compositions, which corresponds to the atomic coordinates in the α-Mg3(Sb, Bi)2 phase, suggests off-centering as an inherent characteristic of the synthesized anionic solid solutions, which are independent of alloyed content as revealed in Fig. 5(b). The evaluated positional modulations of the atomic coordinates from their ideal, symmetric lattice position are shown in Fig. 5(c–e). Substituting Sb with Ge in Mg3(Sb1−x−yBixGey)2 will lead to an uncommon and destabilizing coordination environment for Sb. Intuitively, Ge might tend to shift its lattice position from the anionic site to lower its energy, thereby disrupting local symmetry and enabling orbital hybridization that would otherwise be forbidden. Thus, even for the isostructural compounds, the local coordination and bond distances may vary, and detailed structural analysis is therefore required to draw any conclusions. The high symmetry superionic phase of Mg3(Sb, Bi)2 and Mg3Ge rich nanoprecipitates accommodates alterations in the anionic structural motifs within their large unit cells, wherein the valence electron count per anionic atom determines the type of motif formed. Through electron transfer and the resulting electronic configurations, the trigonal α-Mg3(Sb, Bi)2 phase facilitates Bi-dopant solubility, as indicated by its relatively expanded unit cell. This highlights the versatility of Zintl phases in coexisting with diverse anionic substructures that can be explained well through the Zintl–Klemm concept.
High temperature and pressure sintering leads to superionic phase formation, off-centering of constituting atoms in the trigonal α-Mg3Sb2, and Mg3Ge phase nanoprecipitation. The lattice thermal conductivity (κL) being intimately related to the microstructure is significantly lowered (∼30%) in comparison to the Ge doped Mg3Sb2 as shown in Fig. 6(a). Nanoprecipitates and the superionic phase owing to their large and complex unit cell are anticipated to effectively scatter the phonons, contributing to κL reduction. The off-centering feature of the trigonal α-Mg3Sb2, wherein all the constituting atoms undergo displacive modulation, can be considered as a primary cause for κL reduction as it disrupts the overall periodic arrangement of atoms. The off-centering ranges from ∼0 to 0.02 Å at 300 K and occurs periodically in the structure, wherein locally off-centered atoms are expected to hinder the smooth propagation of phonons.
Pure and intrinsic p-type Mg3Sb2 exhibits poor electrical performance quantified as the power factor (S2σ), primarily due to its high electrical resistivity. However, the synthesized Mg3(Sb1−x−yBixGey)2 (0 < x,y < 0.1) nanocomposites display a significant improvement in temperature dependent electrical conductivity σ(T), as shown in Fig. 6(b). For comparison, the TE properties of the synthesized Ge doped Mg3Sb2 are also shown alongside, wherein at 300 K the σ(T) of Mg3Sb1.8Ge0.2 is ∼1.74 × 103 S m−1, and it improved significantly to ∼1.45 × 104 S m−1. All samples indicate degenerate semiconducting behaviour, wherein σ decreases with increasing temperature. The temperature-dependent S(T) shown in Fig. 6(c) displays p-type conduction, which gradually increases with rising temperature and decreases with increasing alloying content showcasing an inverse correlation with σ(T). A higher power factor as displayed in Fig. 6(d) was attained for all the alloyed compositions with the maximum value approaching 3 × 10−4 W m−1 K−1 for a higher alloy content (x > 0.05). The enhancement can be understood as an outcome of improved weighted mobility (μw)37 derived S(T) and σ(T) measurements, as shown in Fig. 6(e). It provides a weighted average contribution of all energy levels to the overall mobility of the carrier in the crystal and provides good results at room temperature and above, and for mobilities as low as 10−3 cm2 V−1 s−1. All the alloyed compositions exhibit higher μw and correspond well with the evaluated power factor enhancement. The temperature-dependent zT presented in Fig. 6(f) indicates comparatively higher values for alloyed off-centered compositions with a maximum zT ∼ 0.3(±0.05) at 673 K. Optimizing the Fermi level by cation site doping or band bending is anticipated to further enhance the zT of the synthesized alloyed compositions, which exhibits an inherently lowered κL. Thus, the high-symmetry superionic phase in Mg3(Sb, Bi)2-based compositions enhances electrical transport properties, providing a renewed basis for exploring structure–property relationships. This insight paves the way for designing Mg3(Sb, Bi)2 based Zintl phases as advanced functional materials with improved performance.
Footnote |
† Electronic supplementary information (ESI) available: Rietveld refined XRD patterns of Mg3(Sb1−x−yBixGey)2; temperature dependent thermal properties; SEM micrographs; off-centering along x, y, and z directions. See DOI: https://doi.org/10.1039/d4ta06173j |
This journal is © The Royal Society of Chemistry 2024 |