Next Article in Journal
Current Research Studies of Mg–Ca–Zn Biodegradable Alloys Used as Orthopedic Implants—Review
Next Article in Special Issue
Operation Mechanisms of Flexible RF Silicon Thin Film Transistor under Bending Conditions
Previous Article in Journal
Applications of Crystal Plasticity in Forming Technologies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Crystal and Electronic Structure of the New Ternary Compound Ca3InAs3

Department of Chemistry & Biochemistry, University of Delaware, Newark, DE 19716, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Current Address: Department of Chemistry, Louisiana State University, Baton Rouge, LA 70803, USA.
Crystals 2022, 12(10), 1467; https://doi.org/10.3390/cryst12101467
Submission received: 29 September 2022 / Revised: 10 October 2022 / Accepted: 12 October 2022 / Published: 17 October 2022
(This article belongs to the Special Issue Semiconductor Material Growth, Characterization, and Simulation)

Abstract

:
Crystals of a new ternary compound in the Ca-In-As family, Ca3InAs3, have been successfully synthesized via flux growth techniques. This is only the third known compound between the respective elements. As elucidated by single-crystal X-ray diffraction measurements, Ca3InAs3 crystallizes in the orthorhombic space group Pnma (No. 62, Pearson symbol oP28) with unit cell parameters a = 12.296(2) Å, b = 4.2553(7) Å, and c = 13.735(2) Å. The smallest building motifs of the structure are InAs4 tetrahedra, which are connected to one another by shared As corners, forming infinite [InAs2As2/2] chains. The latter propagate along the crystallographic b-axis. The As-In-As bond angles within the InAs4 tetrahedra deviate from the ideal 109.5° value and range from 98.12(2)° to 116.53(2)°, attesting to a small distortion from the regular tetrahedral geometry. Electronic structure calculations indicate the opening of a bandgap, consistent with the expected (Ca2+)3(In3+)(As3–)3 formula breakdown based on conventional oxidation numbers. The calculations also show that the Ca–As interactions are an intermediate between covalent and ionic, while providing evidence of strong covalent features of the In–As interactions. Weak s-p hybridization of In states was observed, supporting the experimentally found deviation of the InAs4 moiety from the ideal tetrahedral symmetry.

1. Introduction

Pnictogen-based (pnictogen, i.e., the group 15 elements P, As, Sb and Bi; abbreviated as Pn hereafter) compounds with a formula of A3MPn3 are reported to crystallize in several distinct structures, where A is an alkaline-earth or divalent rare-earth metal, M is a triel, i.e., a group 13 element [1]. The different structure types are originating from the way the MP4 tetrahedra are interconnected and arranged.
For example, Ca3AlAs3 [2], Sr3InP3 [3], Ca3InP3 [4], Ca3GaAs3 [4], Ca3AlSb3 [5], Eu3InAs3 [6], and Yb3AlSb3 [7] all crystallize in the orthorhombic Pnma space group (No. 62. Pearson symbol oP28). The MPn4 tetrahedra in the structures of the above-mentioned compounds are connected to one another with shared Pn corners, forming an infinite straight chain along the crystallographic b-axis. Surveying the published structures suggests that the In-containing A3MPn3 compounds tend to crystallize in the Pnma structure described above, although this is a requisite for the realization of infinite chains as a structural motif.
One should also note that there are A3MPn3 structures with corner-sharing tetrahedra crystallizing in different space groups. The Sr3GaSb3 compound [3] is a relevant example here since the GaSb4 tetrahedra in the structure are also connected via corner-sharing, but the resultant chains are not straight, but rather zigzagging. This difference results in an amplified structural complexity in Sr3GaSb3, evident from the increased number of repeating units per unit cell and the lowered symmetry; the latter compound crystallizes with the monoclinic space group P21/n (No. 14; Pearson symbol mP56) [3].
In addition to corner-sharing, the tetrahedra in an A3MPn3 compound can also be connected via edge-sharing. Such connections are observed in Ba3GaSb3 (space group: Pnma) [4], Ba3AlSb3 [2] (space group: Cmce), and Eu3AlAs3 (space group: P21/c) [8], in which the tetrahedra are connected via Pn–Pn edge-sharing, forming [M2Pn6]12− dimers (diborane-like units of dual tetrahedra). The different space groups of the above compounds reflect the different arrangements and orientations of the [M2Pn6]12− dimers within the unit cell.
In the present study, we synthesized a previously unknown compound in the A3MPn3 family, Ca3InAs3. We determined the crystal structure of Ca3InAs3 via single-crystal X-ray diffraction methods. Electronic structure calculations show that the bonding between Ca–As has an admixture of covalent and ionic character, while evidence of strong covalent features of the In–As interactions is observed.

2. Materials and Methods

2.1. Synthesis

Single crystals of Ca3InAs3 were synthesized by using indium as molten flux. Ca, In, and As were purchased from Alfa Aesar with stated purity of 99.9 wt % or higher, and were used as received. Manipulations were carried out in an argon-filled glovebox. A starting composition of Ca:In:As of 1:30:1 was used. Excessive In was used as the medium for the crystal growth. The corresponding amounts of elements were loaded into 2 mL alumina crucibles, which were then put inside fused silica ampoules. A piece of quartz wool was placed on top of the crucible without touching the elements inside. The ampoule was sealed under a vacuum level of ca. 30 millitorrs.
All samples in this study were heated in standard muffle furnaces with a temperature profile as follows: 100 °C → 200 °C/h to 1000 °C → hold at 1000 °C for 15.5 h → 5 °C/h to 650 °C. At this point of the crystal growth, the sealed tube was taken out from the furnace, flipped, and the molten metallic flux was separated from the grown crystals by centrifugation. The sealed ampoule was break-opened in the glovebox afterwards. Inspection of the specimen under an optical microscope revealed the presence of many small crystallites, the majority of which were CaIn2As2 [9] with only a minor fraction of crystals of Ca3InAs3.
The crystals were brittle and had a metallic luster; they were found to degrade in ambient air within about ten minutes. During the brief exposure, the crystals’ surface becomes tarnished; after several hours in air the material becomes amorphous.

2.2. Single-Crystal X-ray Diffraction (SC-XRD)

Single-crystal X-ray diffraction (SC-XRD) was performed on single crystals using a Bruker APEX II CCD diffractometer, equipped with Mo Kα radiation. Single-crystal morphologies for Ca3InAs3 were typically thin rod-like, but were cut (under a microscope in dry Paratone-N oil) to block shape, with each side smaller than 100 µm in length. The measurements were conducted at a temperature of 200 K under a N2 atmosphere to achieve optimal data quality and avoid decomposition of the sample during the measurement. Data integration and semiempirical absorption correction were performed with the Bruker-supplied software [10]. The crystal structure was solved with ShelXT [11] using the intrinsic phasing method. Structure refinements were carried out using the ShelXL software, which employs full-matrix least-squares minimization on F2 [12]. Olex2 software was used as a graphical interface [13]. Atomic coordinates of all compounds reported in this paper were standardized with the Structure Tidy program [14]. All sites were refined with anisotropic displacement parameters. The site occupation factors (SOF) were checked by freeing an individual SOF, while other variables were kept fixed. No statistically significant deviations were observed for the SOF on any of the atomic sites. Final difference Fourier map was flat and featureless. Selected crystallographic data are summarized in Table 1.

2.3. Powder X-ray Diffraction

X-ray powder diffraction patterns were collected at room temperature on a Rigaku MiniFlex powder diffractometer, using filtered Cu Kα radiation. Because Ca3InAs3 was the minor phase, and because the experiment was carried out at ambient conditions, there was a very high background. The just a few experimentally observed peaks matched mostly with the calculated positions for CaIn2As2 [9] (major phase) and elemental In (left over flux).

2.4. Electronic Structure Calculations

To investigate the chemical bonding of all compositions, electronic structure calculations were performed within the local density approximation (density functional theory) using the TB-LMTO-ASA program [15]. Experimental unit cell parameters and atomic coordinates from Table 1 and Table 2 were used as input parameters in the calculations. In order to satisfy the atomic sphere approximation (ASA), we employed the von-Barth-Hedin functional [16] and introduced empty spheres. The Brillouin zone was sampled by a 1000 k-point grid. Electronic density of states (DOS), atom-projected electronic density of states (PDOS), and crystal orbital Hamilton population (COHP) were calculated with modules in the LMTO program [17].

3. Results and Discussion

3.1. Crystal Structure

To date, only two compounds have been structurally characterized for the Ca-In-As ternary phase diagram, namely CaIn2As2 [9] and Ca3In2As4 [18]. The Ca3InAs3 reported in the present study is a previously unknown phase and is described in the following paragraphs for the first time.
Ca3InAs3 crystallizes in the orthorhombic Pnma space group (No. 62. Pearson symbol oP28) with the Ca3InP3 structure type (also referred to as Ca3AlAs3 structure type) [19]. The unit cell parameters are a = 12.296(2) Å, b = 4.2553(7) Å and c = 13.735(2) Å. Other relevant crystallographic parameters are tabulated in Table 1. The factional coordinates of the seven crystallographically unique atomic positions are tabulated in Table 2. All positions are of the same symmetry, m, with the same Wyckoff position 4c.
A view of the crystal structure of Ca3InAs3 is shown in Figure 1. As shown there, we prefer to visualize the structure as Ca2+ cations and In–As polyanionic sub-lattice. The InAs4 tetrahedra are connected to one another with shared As corners, forming an infinite straight chain along the crystallographic b-axis.
Selected interatomic distances obtained from the single-crystal refinements can be found listed in Table 3. The In–As distances range from 2.642 to 2.685 Å. These values match very well the metrics reported for another ternary arsenides such as Eu3InAs3 [6], Ba3InAs3 [20], CaIn2As2 [9], Ca3In2As4 [18] and Sr3In2As4 [21], among others. It should also be noted the experimentally observed bond length values are comparable to the sum of the respective single-bonded covalent radii (rIn and rAs are 1.42 Å and 1.21 Å, respectively, yielding a summation value of 2.63 Å [22], which is an indication of the strong covalent bonding nature of the In–As interactions. One may also notice that there are some Ca–As distances that are shorter than 3.0 Å; this also points at substantial covalency of these interactions since the single-bonded covalent radius of Ca is 1.74 Å [22]. Further evidence of this supposition can be found in the DOS and COHP discussions in Section 3.2.
In addition, it is observed that the As-In-As bond angles within the InAs4 tetrahedra deviate from the 109.5° value expected for a regular tetrahedron. Similar tetrahedral distortions are also reported for other “3-1-3” compounds [6,9,20,21].
Lastly, considering the lack of any homoatomic bonding, we can rationalize the structure and the bonding in this new ternary arsenide as a typical valence-precise compound, i.e., a Zintl phase. There are two complimentary ways to understand the electronic charge balance of Ca3InAs3. The first way is to simply exaggerate the ionicity of the bonding and view the formula as [Ca2+]3[In3+][As3−]3. Considering the close electronegativities of In (χP 1.7) and As (χP 2.0) [22], assigning an oxidation number of +3 for In and –3 for As may be hard to justify. Therefore, the other way may be preferred. From the point of view of the Zintl formalism [23], the formal charge of an In atom, covalently bonded to four As atoms will be 1–. In the [InAs2As2/2] tetrahedral chain, formally [InAs3]6−, two of the As atoms at the shared-corner positions have two bonds (2b) and their formal charge is also 1–. The other two As atoms (terminal ones) only have one covalent bond (1b), and need two more valence electrons to satisfy their octets; the formal charge of 1b-As is thus 2–. Therefore, the formula Ca3InAs3 could be expressed as (Ca2+)3[(In1−)(1b-As2−)2(2b-As1−)].
Either the fully ionic or the bonding picture within the Zintl formalism capture the essence of the valance electron count in Ca3InAs3, which is expected to be an intrinsic semiconductor. The notion of semiconducting behavior is confirmed by the electronic structure calculations, presented next.

3.2. Electronic Structure

The bonding characteristics of Ca3InAs3 were investigated via electronic structure calculations using the LMTO program [15]. We evaluated the electronic properties of the new material with calculations employing the experimental data listed in Table 1 and Table 2.
The stacked atom-projected electronic density of states (DOS) curves for Ca3InAs3 are shown in Figure 2a, in which a calculated bandgap of approx. 0.46 eV can be observed. This value is reconcilable with those of other ternaty indium arsenides, such as Eu3InAs3 and CaIn2As2 which have calculated bandgaps on the order of 0.2 and 0.5 eV [6,9]. The existence of a bandgap is in agreement with the formal electron count as discussed in the previous section. Knowing that DFT methods typically underestimate the bandgap, the actual bandgap of Ca3InAs3 is speculated to be larger.
The majority of the states in the conduction band are dominated by Ca d orbitals, with minor As and In contributions. The states close to the valence band maximum are contributed mainly by the As p orbitals with minor Sr and In contributions. Specifically, in the region of −1 < EEF < 0 eV, the contribution of the As p orbitals is most considerable. The minor overlap of Ca and As states suggest that there are some covalent features between Ca and As. Therefore, due to the incomplete electron transfer of the cation, strictly speaking, Ca3InAs3 does not exhibit the textbook characteristics of a Zintl phase.
A few insights can be observed for the bonding of the InAs4 tetrahedra. In-p and As-p states showed extensive overlapping in the range of between −4.1 < EEF < −2 eV (Figure 2c,d). This means the existence of a strong covalent feature of the In–As interactions. In addition, the In atoms show a clear separation between the s and p orbitals, indicating a weak s-p hybridization. This insight agrees well with the InAs4 tetrahedra distortion given by the crystal refinements.
Figure 3 shows the electronic band structure of Ca3InAs3 in momentum space. As evident from the plot, there are no bands crossing from the valence band to the conduction band. The lowest energy separation between the top of the valence band and the bottom of the conduction band occurs at the Γ-point. This observation is suggestive of a direct bandgap semiconductor.
The crystal orbital Hamilton population curves (COHP) for the averaged selected interactions of Ca3InAs3 are plotted in Figure 4. Both the In–As and Ca–As interactions are optimized at the Fermi level. For the Ca–As interactions, some bonding features can still be observed above the Fermi level when 0.5 < EEF < 2 eV. This indicates the possibility of electron doping on the Ca site. In contrast, In–As interactions only show antibonding features above the Fermi level.

4. Conclusions

With this study, we reported on the discovery of the new compound, Ca3InAs3, extending the knowledge in the Ca-In-As compositional space, and expanding the variety of compounds that crystallize with this quasi 1D-structure. The bonding characteristics of Ca3InAs3 were studied both experimentally and computationally. We found that Ca3InAs3 is an intrinsic semiconductor, where the InAs4 tetrahedral building units are lightly distorted. A possibility of doping on the Ca site is also suggested.
This materials system can providee promising thermoelectric material candidates and/or topological insulators, but for quality property measurements, the synthetic challenges making a phase-pure bulk sample must be resolved.

Author Contributions

W.P.: Investigation, Methodology, Formal analysis, Visualization, Writing—original draft; S.B. (Sviatoslav Baranets): Conceptualization, Investigation, Formal analysis, Writing—review and editing; S.B. (Svilen Bobev): Conceptualization, Supervision, Project administration, Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge financial support from the US Department of Energy through a grant DE-SC0008885.

Informed Consent Statement

Not applicable.

Data Availability Statement

The corresponding crystallographic information file (CIF) for Ca3InAs3 has been deposited with CSD, and the data for this paper can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html (accessed on 28 September 2022) (or from the CCDC, 12 Union Road, Cambridge CB2 1 EZ, UK; Fax: +44-1223-336033; E-mail: deposit@ccdc.cam.ac.uk). Depository number is 2210325.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stoyko, S.; Voss, L.; He, H.; Bobev, S. Synthesis, crystal and electronic structures of the pnictides AE3TrPn3 (AE = Sr, Ba; Tr = Al, Ga; Pn = P, As). Crystals 2015, 5, 433–446. [Google Scholar] [CrossRef] [Green Version]
  2. Cordier, G.; Savelsberg, G.; Schäfer, H. Zintl phasen mit komplexen anionen: Zur kenntnis von Ca3AlAs3 und Ba3AlSb3. Z. Naturforsch. B 1982, 37, 975–980. [Google Scholar] [CrossRef]
  3. Cordier, G.; Schäfer, H.; Stelter, M. Sr3GaSb3 und Sr3InP3, zwei neue Zintl phasen mit komplexen anionen. Z. Naturforsch. B 1982, 42, 1268–1272. [Google Scholar] [CrossRef]
  4. Cordier, G.; Schäfer, H.; Steher, M. Neue Zintl phasen: Ba3GaSb3, Ca3GaAs3 und Ca3InP3. Z. Naturforsch. B 1985, 40, 1100–1104. [Google Scholar] [CrossRef] [Green Version]
  5. Cordier, G.; Schäfer, H.; Stelter, M. Ca3AlSb3 und Ca5Al2Bi6, zwei neue Zintl phasen mit kettenförmigen anionen. Z. Naturforsch. B 1984, 39, 727–732. [Google Scholar] [CrossRef]
  6. Rajput, K.; Baranest, S.; Bobev, S. Observation of an unexpected n-type semiconducting behavior in the new ternary Zintl phase Eu3InAs3. Chem. Mater. 2020, 32, 9444. [Google Scholar] [CrossRef]
  7. Shang, R.; Nguyen, A.T.; He, A.; Kauzlarich, S.M. Crystal structure characterization and electronic structure of a rare-earth-containing Zintl phase in the Yb–Al–Sb family: Yb3AlSb3. Acta Crystallogr. C 2021, 77, 281–285. [Google Scholar] [CrossRef] [PubMed]
  8. Radzieowski, M.; Stegemann, F.; Klenner, S.; Zhang, Y.; Fokwa, B.P.; Janka, O. On the divalent character of the Eu atoms in the ternary Zintl phases Eu5In2Pn6 and Eu3MAs3 (Pn = As–Bi; M = Al, Ga). Mater. Chem. Front. 2020, 4, 1231–1248. [Google Scholar] [CrossRef]
  9. Ogunbunmi, M.O.; Baranets, S.; Childs, A.B.; Bobev, S. The Zintl phases AIn2As2 (A = Ca, Sr, Ba): New topological insulators and thermoelectric material candidates. Dalton Trans. 2021, 50, 9173–9184. [Google Scholar] [CrossRef] [PubMed]
  10. Bruker Analytical X-ray Systems; Inc. SMART & SAINT: Madison, WI, USA, 2003.
  11. Sheldrick, G.M. SHELXT–Integrated space group and crystal structure determination. Acta Crystallogr. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  12. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  14. Gelato, L.; Parthé, E. STRUCTURE TIDY—A computer program to standardize crystal structure data. J. Appl. Crystallogr. 1987, 20, 139–143. [Google Scholar] [CrossRef]
  15. Tank, R.; Jepsen, O.; Burkhardt, A.; Andersen, O. TB-LMTO-ASA Program; Max-Planck-Institut für Festkörperforschung: Stuttgart, Germany, 1994. [Google Scholar]
  16. von Barth, U.; Hedin, L. A local exchange-correlation potential for the spin polarized case. J. Phys. C Solid State Phys. 1972, 5, 1629. [Google Scholar] [CrossRef]
  17. Steinberg, S.; Dronskowski, R. The crystal orbital Hamilton population (COHP) method as a tool to visualize and analyze chemical bonding in intermetallic compounds. Crystals 2018, 8, 225. [Google Scholar] [CrossRef] [Green Version]
  18. Cordier, G.; Schäfer, H.; Stelter, M. Sr3In2P4 and Ca3In2As4 Zintl phases with strings of InP4 and InAs4 tetrahedra, resp., sharing edges and corners. Z. Naturforsch. B 1986, 41, 1416–1419. [Google Scholar] [CrossRef]
  19. Villars, P.; Calvert, L.D. (Eds.) Pearson’s Handbook of Crystallographic Data for Intermetallic Compounds, 2nd ed.; American Society for Metals: Materials Park, OH, USA, 1991. [Google Scholar]
  20. Somer, M.; Carrillo-Cabrera, W.; Peters, K.; von Schnering, H.-G.; Cordier, G. Crystal structure of tribarium triarsenidoindate, Ba3InAs3. Z. Kristallogr. New Cryst. Struct. 1996, 211, 632. [Google Scholar] [CrossRef]
  21. Childs, A.; Baranets, S.; Bobev, S. Five new ternary indium–arsenides discovered. Synthesis and structural characterization of the Zintl phases Sr3In2As4, Ba3In2As4, Eu3In2As4, Sr5In2As6 and Eu5In2As6. J. Solid State Chem. 2019, 278, 120889. [Google Scholar] [CrossRef]
  22. Pauling, L. The Nature of the Chemical Bond, 3rd ed.; Cornell University Press: Ithaca, NY, USA, 1960. [Google Scholar]
  23. Schäfer, H.; Eisenmann, B.; Müller, W. Zintl Phases: Transitions between metallic and ionic bonding. Angew. Chem. Int. Ed. Engl. 1973, 12, 694–712. [Google Scholar] [CrossRef]
Figure 1. (a) A schematic polyhedral representation of the orthorhombic crystal structure of Ca3InAs3. (b) Distinct coordination environments for Ca1, Ca2, Ca3, As1, As2, and As3. (c) The InAs4 tetrahedra and the way they are linked (corner-sharing of the As1 atoms). The As-In-As angles deviate from the ideal 109.5° of a regular tetrahedron. As1-In-As2, As1-In-As3, and As3-In-As2 angles measure 108.81(2)°, 116.53(2)°, and 98.12(2)°, respectively. Interatomic distances can be found in Table 3.
Figure 1. (a) A schematic polyhedral representation of the orthorhombic crystal structure of Ca3InAs3. (b) Distinct coordination environments for Ca1, Ca2, Ca3, As1, As2, and As3. (c) The InAs4 tetrahedra and the way they are linked (corner-sharing of the As1 atoms). The As-In-As angles deviate from the ideal 109.5° of a regular tetrahedron. As1-In-As2, As1-In-As3, and As3-In-As2 angles measure 108.81(2)°, 116.53(2)°, and 98.12(2)°, respectively. Interatomic distances can be found in Table 3.
Crystals 12 01467 g001
Figure 2. (a) The stacked atom-projected electronic density of states (DOS) for Ca3InAs3. The Fermi level is taken as a reference at 0 eV and represented with a dashed line. A bandgap of 0.46 eV can be observed. (b) Calculated partial DOS of different orbitals of for Ca. (c) Calculated partial DOS of different orbitals of for In. (d) Calculated partial DOS of different orbitals of for As.
Figure 2. (a) The stacked atom-projected electronic density of states (DOS) for Ca3InAs3. The Fermi level is taken as a reference at 0 eV and represented with a dashed line. A bandgap of 0.46 eV can be observed. (b) Calculated partial DOS of different orbitals of for Ca. (c) Calculated partial DOS of different orbitals of for In. (d) Calculated partial DOS of different orbitals of for As.
Crystals 12 01467 g002
Figure 3. The electronic band structure of Ca3InAs3. The Fermi level is taken as a reference at 0 eV and represented with a dashed line. The opening of a bandgap of ca. 0.46 eV at the Γ point is suggestive of an intrinsic, direct gap semiconductor.
Figure 3. The electronic band structure of Ca3InAs3. The Fermi level is taken as a reference at 0 eV and represented with a dashed line. The opening of a bandgap of ca. 0.46 eV at the Γ point is suggestive of an intrinsic, direct gap semiconductor.
Crystals 12 01467 g003
Figure 4. The crystal orbital Hamilton population curves (COHP) for two averaged interactions in Ca3InAs3. The In–As interactions are optimized at the Fermi level. Ca–As interactions are almost fully optimized at the Fermi level.
Figure 4. The crystal orbital Hamilton population curves (COHP) for two averaged interactions in Ca3InAs3. The In–As interactions are optimized at the Fermi level. Ca–As interactions are almost fully optimized at the Fermi level.
Crystals 12 01467 g004
Table 1. Selected crystal data and structure refinement parameters for Ca3InAs3.
Table 1. Selected crystal data and structure refinement parameters for Ca3InAs3.
FormulaCa3InAs3
Formula weight (g·mol−1)459.82
Radiation, λMo Kα, 0.71073 Å
Temperature (°C)–70(2)
Crystal systemorthorhombic
Space groupPnma (No. 62)
Z4
a (Å)12.296(2)
b (Å)4.2553(7)
c (Å)13.735(2)
V3)718.7(2)
ρcalc (g·cm−3)4.25
µMoKα (cm−1)190.1
Reflections: parameters1151: 44
R1 [I > 2σ(I)] a0.0222
R1 (all data) a0.0330
wR2 [I > 2σ(I)] a0.0463
wR2 (all data) a0.0496
Largest peak; deepest hole (e/Å−3)0.83; –1.05
[a] R1 = ∑||Fo| – |Fc||/∑|Fo|; wR2 = [∑[w(Fo2 – Fc2)2]/∑[w(Fo2)2]]1/2, where w = 1/[σ2Fo2 + (0.0213·P)2 + (0.301·P)], and P = (Fo2 + 2Fc2)/3.
Table 2. Atomic coordinates and equivalent isotropic displacement parameters (Ueq) for Ca3InAs3.
Table 2. Atomic coordinates and equivalent isotropic displacement parameters (Ueq) for Ca3InAs3.
AtomWyckoff SymbolSite SymmetryxyzUeq2) a
As14c.m0.04891(4)1/40.36528(4)0.0103(1)
As24c.m0.25779(4)1/40.62613(4)0.0086(1)
As34c.m0.60837(4)1/40.61098(4)0.0085(1)
In4c.m0.05554(3)1/40.69988(3)0.0103(2)
Ca14c.m0.06164(7)1/40.10532(8)0.0104(2)
Ca24c.m0.26873(7)1/40.28697(8)0.0092(2)
Ca34c.m0.34436(8)1/40.00296(8)0.0104(2)
[a] Ueq is defined as one third of the trace of the orthogonalized Uij tensor.
Table 3. Selected interatomic distances (Å) in Ca3InAs3. Contacts longer than 3.9 Å are not shown.
Table 3. Selected interatomic distances (Å) in Ca3InAs3. Contacts longer than 3.9 Å are not shown.
AtomsDistance (Å)AtomsDistance (Å)
In–As1 (×2)2.6415(5)As2–In2.6853(7)
In–As22.6853(7)As2–Ca1 (×2)3.0885(8)
In–As32.6779(7)As2–Ca2 (×2)3.0845(9)
In–Ca1 (×2)3.710(1)As2–Ca3 (×2)2.9945(8)
In–Ca2 (×2)3.2600(8)As3–In2.6779(7)
In–Ca3 (×2)3.655(1)As3–Ca1 (×2)2.984(8)
In–Ca33.809(1)As3–Ca13.026(1)
As1–In (×2)2.6415(5)As3–Ca2 (×2)2.9623(8)
As1–Ca13.574(1)As3–Ca33.297(1)
As1–Ca22.9091(1)Ca1–Ca23.565(1)
As1–Ca33.099(1)Ca1–Ca33.750(1)
As1–Ca3 (×2)3.1345(9)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Peng, W.; Baranets, S.; Bobev, S. Synthesis, Crystal and Electronic Structure of the New Ternary Compound Ca3InAs3. Crystals 2022, 12, 1467. https://doi.org/10.3390/cryst12101467

AMA Style

Peng W, Baranets S, Bobev S. Synthesis, Crystal and Electronic Structure of the New Ternary Compound Ca3InAs3. Crystals. 2022; 12(10):1467. https://doi.org/10.3390/cryst12101467

Chicago/Turabian Style

Peng, Wanyue, Sviatoslav Baranets, and Svilen Bobev. 2022. "Synthesis, Crystal and Electronic Structure of the New Ternary Compound Ca3InAs3" Crystals 12, no. 10: 1467. https://doi.org/10.3390/cryst12101467

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