Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX
pubs.acs.org/IC
Synthesis and Characterization of the Tin Iodide Borate Sn3[B3O7]I Sandra Schönegger,† Florian Pielnhofer,‡ Andreas Saxer,§ Klaus Wurst,† and Hubert Huppertz*,† †
Institut für Allgemeine, Anorganische und Theoretische Chemie, Universität Innsbruck, Innrain 80-82, 6020 Innsbruck, Austria Institut für Anorganische Chemie, Universität Regensburg, Universitätsstraße 31, D-93053 Regensburg, Germany § Institut für Konstruktion und Materialwissenschaften, Universität Innsbruck, Technikerstraße 13, A-6020 Innsbruck, Austria ‡
Inorg. Chem. Downloaded from pubs.acs.org by UNIV OF LOUISIANA AT LAFAYETTE on 04/09/19. For personal use only.
S Supporting Information *
ABSTRACT: The tin iodide borate Sn3[B3O7]I was synthesized via a hydrothermal synthesis and crystallizes in the centrosymmetric space group Pbca (no. 61) possessing lattice parameters of a = 1071.8(3), b = 852.3(2), and c = 2016.8(5) pm and Z = 8. Characteristic for the structure are infinite chains along the b axis, built up of three membered B3O8 rings consisting of one BO3 unit and two corner-sharing BO4 tetrahedra. The three tin cations are oriented differently: one cation is located layer-like between the infinite chains, and the other two cations show an orientation in a row with the infinite chain. In this structure, only one of the three tin cations exhibits a coordination to the halogen anion. The new centrosymmetric tin iodide borate Sn3[B3O7]I was investigated by single-crystal diffraction, vibrational spectroscopy, powder X-ray diffraction data, thermogravimetry, differential scanning calorimetry, and DFT calculations.
1. INTRODUCTION In 2010, Zhang et al. synthesized the first iodide borate Pb2B5O9I via a solid-state reaction.1 Eight years later, a second compound in the class of halogen borates was obtained through a hydrothermal synthesis leading to Pb2BO3I possessing a large second-harmonic generation response.2 Halogen borates received more and more attention and became excellent materials for nonlinear optical (NLO) materials. Hitherto, mainly lead fluoride, chloride, and bromide borates were synthesized and investigated. For example, Pb3B6O11F2 was the first lead borate fluoride (noncentrosymmetric) with a large second harmonic generation response that is 4× that of KDP.3 In the field of chlorides and bromides, the compounds M2B5O9X (M = Ca, Sr, Ba, Pb, Eu; X = Cl, Br) were reported in the literature.4−6 Interestingly, their SHG signal increases significantly along the isostructural series M = Ca < Sr < Ba < Pb and X = Cl < Br.3 The synthesis of new iodide borates is a challenge and simple synthetic access to such compounds became successful only recently with the compound Pb2BO3I.2 In 2018, the first lead iodide borate Pb2BO3I was synthesized via a hydrothermal synthesis at 220 °C with a reaction time of 48 h,2 and in this work we want to present the first tin iodide borate Sn3[B3O7]I obtained through a hydrothermal synthesis. A view on the excellent SHG intensities exhibits similar values for both iodide borates, Pb2BO3I and Pb2B5O9I. Among borates, Pb2B5O9I shows the largest SHG response (measured on a powder) with about 13.5× that of potassiumdihydrogen phosphate and a transparency from the near-UV to the middle-IR area.1 Pb2BO3I exhibits a structural similarity to the well-known compound KBe2BO3F2 (KBBF),2 showing at 1064 nm a SHG signal 10× that of KDP. The combination of the Pb2+ cations, © XXXX American Chemical Society
the iodide anions, and the special crystal structure is responsible for the high SHG responses of both iodide borates. Generally, borates are very well suited for NLO materials. In borates, the boron atom shows the ability to coordinate four or three oxygen atoms and form [BO3]3− groups (trigonal planar) and [BO4]5− units (tetrahedral) promoting SHG properties and birefringence. Several borates with SHG properties are known in the literature, e.g., CsB3O5 (CBO),7 β-BaB2O4 (BBO),8 CsLiB6O10 (CLBO),9,10 and LiB3O5 (LBO).11 A few fluoride borates have been presented in the literature so far showing an SHG response, including Sn2B7O12F,12 Pb3B6O11F2,3 Cd5(BO3)3F,13 and Ba3B6O11F2.14 The fluorooxoborates NaB4O6F,15 BiB2O4F,16 Sn[B2O3F2],17 CsB4O6F,18 and SrB5O7F319 are also literature-known compounds. We already were able to synthesize a new tin fluoride borate Sn3[B3O7]F,20 which exhibits an SHG signal intensity 0.4× that of KDP. Aiming on the synthesis of the first tin iodide borate using hydrothermal conditions, we successfully obtained the novel compound Sn3[B3O7]I. However, the structure of this new iodide borate Sn3[B3O7]I differs significantly from the previous reported fluoride borate Sn3[B3O7]F.20
2. EXPERIMENTAL SECTION 2.1. Synthesis. Sn3[B3O7]I was synthesized hydrothermally using the starting materials SnO [174 mg, 1.3 mmol, Strem Chemicals, Inc. (≥99.8%, Newburyport, USA)], SnI2 [55 mg, 0.2 mmol], and H3BO3 [79 mg, 1.3 mmol, Roth GmbH + Co. KG (≥99.8%, Karlsruhe, Germany)]. The necessary SnI2 was prepared by a precipitation Received: February 8, 2019
A
DOI: 10.1021/acs.inorgchem.9b00377 Inorg. Chem. XXXX, XXX, XXX−XXX
Article
Inorganic Chemistry reaction of an aqueous SnCl2 solution with a KI solution. The mixture was placed in a small alumina inlet and the ceramic inlet in an 8 mL Teflon inlet. The ceramic inlet was backfilled with 1.5 mL of deionized water. Finally, the Teflon inlet was positioned in a stainlesssteel autoclave and heated up to a temperature of 220 °C and kept there for 2 days. Afterward, the reaction mixture was cooled down at a rate of 3° per hour down to 63 °C and finally quenched to room temperature. A yellow green powder was obtained and analyzed by Xray powder diffraction. Furthermore, on a single-crystal X-ray diffractometer, isolated platelets of Sn3[B3O7]I were measured. 2.2. X-ray Structure Determination. Powder X-ray Diffraction. The product Sn3[B3O7]I was identified by X-ray powder diffraction data collected with a Stoe Stadi P powder diffractometer in transmission geometry and Mo Kα1 (λ = 70.93 pm) radiation applying a focusing Ge(111) primary beam monochromator and a Mythen2 1K microstrip detector (1280 strips, Dectris, Baden, Switzerland). Figure 1 shows the experimental powder pattern of
Table 1. Crystal Data and Structure Refinement of Sn[B3O7]I empirical formula −1
molar mass, g mol crystal system space group single-crystal diffractometer radiation; wavelength, pm a, pm b, pm c, pm V, nm3 formula units per cell, Z calculated density, g cm−3 crystal size, mm3 temperature, K absorption coefficient, mm−1 F(000), e 2θ range, deg range in hkl total no. of reflections independent reflections/Rint data/refined parameters absorption correction goodness-of-fit on Fi2 final R1/wR2 [I > 2σ(I)] R1/wR2 (all data) largest diff. peak/hole, e Å −3
Sn3[B3O7]I 627.40 orthorhombic Pbca (no. 61) Bruker D8 QUEST PHOTON 100 Mo Kα; λ = 71.07 1071.8(3) 852.3(2) 2016.8(5) 1842.4(8) 8 4.52 0.110 × 0.050 × 0.015 299.0 11.42 2192 5.5−72.8 −17 ≤ h ≤ 17, −12 ≤ k ≤ 13, −32 ≤ l ≤ 32 34 141 4053/0.0391 4053/127 multiscan (Bruker SADABS 2014/5) 1.073 0.0381/0.0979 0.0450/0.1011 2.87/−5.86
Reflection) was used. The spectrum was performed in the spectral range of 400−4000 cm−1 using a Bruker ALPHA Platinum-ATR spectrometer (Bruker, Billerica, USA). The spectrometer is equipped with a 2 × 2 mm diamond ATR-crystal and a DTGS detector. A total of 320 scans of the powder sample were acquired, and the Opus 7.2 software24 was used for the correction of atmospheric influences. 2.4. Thermoanalytical Investigation. The DTA-TG curve of Sn3[B3O7]I was obtained on a simultaneous thermal analysis (STA) 449 F5 Jupiter from Netzsch. STA enables differential thermal analysis (DTA) to be performed simultaneously with a thermogravimetric analysis (TGA). Sample weight, the rate of weight increase/ decrease, and the heat flow were continuously measured as a function of temperature. A difference in heat flow (DTA) between sample and reference corresponds to exothermic or endothermic reactions inside of the sample. A total of 20.3 mg of the sample was heated from room temperature up to 800 °C under a N2 atmosphere with a heating rate of 10 °C/min. 2.5. High Temperature XRD Powder Pattern. On an Empyrean powder diffractometer (Panalytical), the high temperature X-ray diffraction patterns were obtained. The diffractometer is equipped with a Pixcel 1D detector and Cu tube with 40 kV tube tension and 40 mA tube current. A variable divergence slit was used with an irradiated length of 6 mm, and the sample was applied to the platinum strip (thickness of the platinum strip: 1 mm). The sample was measured under air and in a temperature range between 25 and 700 °C in an Anton Paar HTK 16N high temperature strip-heater chamber. 2.6. DFT Calculations. With the program CRYSTAL17,25,26 quantum chemical calculations were performed in the framework of density functional theory (DFT). Furthermore, with the GGA (PBE)27 functional, the total energy calculations including full structural optimizations were performed and the hybrid functional was range separated with the HSE06.28 The convergence criterion considering the energy was set to 1 × 10−8 a.u. with a k-mesh sampling of 6 × 6 × 6. All-electron basis sets for Sn, B, and O were taken from refs 29−31, and an effective core potential basis set was applied for I.32 On the basis of the relaxed structures, the electronic
Figure 1. Experimental powder pattern (Mo Kα radiation, λ = 71.073 pm; top), compared with the theoretical powder pattern from singlecrystal data of Sn3[B3O7]I (bottom). The reflections marked with a red asterisk derive from an unknown impurity. Sn3[B3O7]I at the top compared to the theoretical pattern derived from the single-crystal data; the reflections marked with a red asterisk stem from an unknown impurity. Single-Crystal X-ray Diffraction. A platelet of Sn3[B3O7]I was isolated through polarization contrast microscopy and analyzed via single-crystal X-ray diffraction. The intensity data were collected at 180 K with a Bruker D8 Quest diffractometer (Photon 100) equipped with an Incoatec Microfocus source generator (multilayered optics monochromatized Mo Kα radiation, λ = 71.073 pm). Multiscan absorption corrections were applied with the program SADABS2014/5.21 During structure solution and parameter refinement with anisotropic displacement parameters for all atoms using the SHELXS/ L-13 software suite,22,23 the space group Pbca was found to be correct. The measured crystal was a nonmerohedral twin with two domains in a ratio of around 2:1 and few overlays. Therefore, the crystal was approximately refined as a normal single crystal with the reflections of the main component. All relevant details of the data collection are presented in Table 1. The positional parameters, the anisotropic displacement parameters, interatomic distances, and bond angles are given in the Supporting Information in Tables S1−S4. CCDC 1893973 (Sn3[B3O7]I) contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac. uk/data_request/cif. 2.3. Vibrational Spectroscopy. For the characterization of the Sn3[B3O7]I powder sample, an FTIR-ATR (Attenuated Total B
DOI: 10.1021/acs.inorgchem.9b00377 Inorg. Chem. XXXX, XXX, XXX−XXX
Article
Inorganic Chemistry
Figure 2. (a) The fundamental building block formed by the three membered ring ⟨Δ2□⟩ (encircled in red) and its linkage to endless chains along the b axis. O2−, small blue spheres; B3+, small red spheres; BO4 groups, blue polyhedra; BO3 groups, yellow triangles. (b) The black rectangle represents the unit cell with the infinite chains along the b axis. Sn2+ cations: Sn2, turquoise; Sn1, gray; and Sn3, purple; I−, small orange spheres. band structures were performed. By calculating the electron localization function (ELF),33,34 a direct space analysis of the charge density was carried out, and with TOPOND35 interfaced to CRYSTAL17, 3D plots were visualized with XCrysDen.36
3. RESULTS AND DISCUSSION 3.1. Crystal Structure. Sn3[B3O7]I crystallizes in the centrosymmetric space group Pbca (no. 61) with a = 1071.8(3), b = 852.3(2), and c = 2016.8(5) pm (see Table 1). The fundamental building block (see Figure 2a, red circle) in the structure of Sn3[B3O7]I is built up of B3O8 groups, which form three membered rings. This FBB is formed by one BO3 group and two corner sharing BO4 tetrahedra described as ⟨Δ2□⟩37 with the square and triangle assigned to the tetrahedral [BO4]5− and the trigonal-planar [BO3]3− groups, respectively. Furthermore, this FBB is connected to chains running along the b axis. The oxygen atom O5 (encircled in red in Figure 4c) shows the connection to further FBBs. Interestingly, the tips of the BO3 groups of alternating infinite chains are always oriented in different directions, once to the right and once to the left as presented in Figure 2b. The Sn22+ are located layer-like between the infinite chains along the b axis. The Sn12+ and Sn32+ are oriented in a row with the infinite chains (see Figure 2b). There exist three unique boron atoms in the structure of Sn3[B3O7]I, and the B2 atom shows a 3-fold coordination to oxygen atoms forming BO3 groups. The O−B−O bond angles are found with values of 118.5(4) to 122.8(4)° and an average value of 120.0°. Furthermore, the B−O distances are in the range from 136.5(6) to 137.1(6) pm (average value of 136.8 pm). The average B−O distance exhibiting 137 pm for this coordination agrees very well with this value.37 The atoms B1 and B3 exhibit a tetrahedral coordination (four oxygen atoms). The B−O distances in these polyhedra vary between 145.7(6) and 151.9(6) pm, in good agreement with the mean value of 147.6 pm.37,38 In the [BO4]5− group, the bond angles O−B−O exhibit a mean value of 109.5°, which fits to the typical tetrahedral angle (109.47°). The three different coordination environments are exhibited by the three crystallographically unique Sn atoms (see Figure 3), whereby the Sn12+ and Sn32+
Figure 3. Coordination environments of the three Sn2+ cations in Sn3[B3O7]I.
coordinate to four O atoms. The Sn12+ is connected to the O atoms O1, O2, O3, and O4 with distances of 208.1(3)− 277.1(3) pm (mean value: 230.1 pm). The Sn3 cation shows a coordination to the O atoms O1, O2, O3, and O5 with distances Sn−O of 219.4(3)−281.0(3) pm with a mean value of 235.5 pm. The iodide atom coordinates to the Sn2 cation and to the O7 oxygen atom with a Sn2−O distance of 219.4(4) pm. The interatomic distance to I1 is 293.3(1) pm corresponding to a Pb−I single bond (262.3(5) pm).2 A stereochemically active 5s2 lone electron pair is clearly observed in all of the Sn2+ cations that is oriented toward the open flank of the coordination polyhedron (Figure 3). According to the CHARDI (charge distribution in solids, ΣQ) and the BLBS (bond-length/bond-strength, ΣV) concepts,39−43 the charge distribution was calculated. In Table 2, the results are listed, being consistent with the standard valence states of the anions and cations. At this point, a short comparison with the recently published compound Sn3[B3O7]F,20 which seems very similar at first sight, will be given in the following. The comparison of the lattice parameters of the single crystals of Sn3[B3O7]I and Sn3[B3O7]F are given in Table S5 in the Supporting C
DOI: 10.1021/acs.inorgchem.9b00377 Inorg. Chem. XXXX, XXX, XXX−XXX
Article
Inorganic Chemistry
Table 2. Charge Distribution in Sn3[B3O7]I (Space Group Pbca (no. 61)) Calculated with the Bond-Length/Bond-Strength Concept (ΣV)38,39 and the CHARDI Concept (ΣQ) Sn1
Sn2
Sn3
B1
B2
B3
O1
ΣV ΣQ
+2.08 +1.95 O2
+1.82 +1.95 O3
+1.66 +1.88 O4
+2.97 +3.06 O5
+3.02 +3.21 O6
+3.02 +2.99 O7
−2.15 −2.21 I1
ΣV ΣQ
−1.96 −1.93
−1.92 −2.08
−1.76 −1.78
−2.06 −2.11
−1.80 −1.81
−2.27 −2.05
−0.83 −1.03
Information. The compound Sn3[B3O7]F shows SHG properties and crystallizes in the orthorhombic noncentrosymmetric space group Pna21 (no. 33). The compound Sn3[B3O7]I crystallizes in the centrosymmetric space group Pbca (no. 61), showing no SHG properties. Further differences can be found in the FBB, in the crystal structures, and in the coordination environment of the three tin cations. B3O7 groups are found in the structure of Sn3[B3O7]F displaying the FBB. This FBB is built up of one BO4 and two BO3 groups represented by ⟨2Δ□⟩37 (see Figure 4a). Sn3[B3O7]I contains B3O8 groups, representing the FBB and consisting of one BO3 unit and two BO4 tetrahedra, formulated as ⟨Δ2□⟩ (see Figure 4c). In Sn3[B3O7]F, the tin cations Sn2 and Sn3 show a coordination to three oxygen atoms and one fluoride anion. The Sn1 cation shows a coordination to three oxygen atoms. Interestingly, in Sn 3 [B 3O 7 ]I, the tin cations show a slightly different coordination. Only the Sn2 cation coordinates to the iodide anion and one oxygen atom, and Sn1 and Sn3 show coordinations to four oxygen atoms. Figure 4b and d show a comparison of the unit cells of Sn3[B3O7]F and Sn3[B3O7]I. 3.2. Vibrational Spectroscopy. Figure 5 shows the experimental IR spectrum of Sn3[B3O7]I recorded from a sample (powder) between 400 and 4000 cm−1. As mentioned in the analysis of the powder pattern, a small fraction of a second unknown phase is present. However, in the structure of Sn3[B3O7]I, the existence of BO3 and BO4 groups can be clearly confirmed. The bending vibrations of BO4 and BO3 groups can be assigned to the bands in the range of 500 and 700 cm−1.44 In the range between 1450 and 850 cm−1, strong bands of B−O stretching vibrations are found. Bands of [BO3]3− vibrations lie in the range of 1450 and 1200 cm−1,45−47 and the absorption peaks between 1025 and 900 cm−1 correspond to the asymmetric and symmetric stretching modes of the BO4 polyhedra.46−48 The weak band at 3500 cm−1, which is normally assigned to O−H stretching vibrations, could come from surface-absorbed water or from the mentioned impurity. Since the unknown impurity is not known, the weak band at 3500 cm−1 can also originate from this phase. 3.4. Thermal Behavior. The thermoanalytical investigations were performed for studying the high-temperature stability of Sn3[B3O7]I. The results from the differential thermal analysis (DTA), the thermogravimetric analysis (TG), and the differential thermal gravimetry (DTG), which were recorded from ambient temperature to 800 °C, are shown in Figure 6. The compound Sn3[B3O7]I shows stability up to a temperature of about 380 °C. The DTA curve (see Figure 6, green line) shows two peaks, one small endothermic peak at 423 °C and a larger exothermic peak at 577 °C. The exothermic reaction starts at a temperature of 490 °C (see Figure 6, blue line) and shows a maximum at 577 °C and is accompanied by a mass increase of about 0.42% (see Figure 6 red line), which is probably due to the transformation into a
Figure 4. (a) Isolated three membered ring ⟨□2Δ⟩, presenting the FBB in the compound Sn3[B3O7]F. (b) The unit cell along the b axis of the compound Sn3[B3O7]F. (c) FBB in the new compound Sn3[B3O7]I forming a three membered ring ⟨Δ2□⟩. (d) The unit cell of Sn3[B3O7]I along the b axis. The oxygen atom O5 (encircled in red in c) shows the connection to further FBBs.
new phase. The exothermic reaction is accompanied by a mass loss of 5%. Above 577 °C, no further exothermic or endothermic peaks can be detected. D
DOI: 10.1021/acs.inorgchem.9b00377 Inorg. Chem. XXXX, XXX, XXX−XXX
Article
Inorganic Chemistry
Figure 5. FT-IR absorbance spectrum of the powder sample of Sn3[B3O7]I.
Figure 7 shows the HT-PXRD at different temperatures. Up to 380 °C, no changes are evident. At 390 °C, there is an additional reflection at 11.7° (2θ) (see Figure 7, pattern at 390 °C (encircled in red)), which cannot be assigned to the original, still existing compound. This reflection shows a maximum at 440 °C (see Figure 7, 440 °C) and disappears at 500 °C (see Figure 7, 500 °C). This additional reflection may originate from a new phase, which is formed briefly in this temperature range. The isolation and characterization of this phase of the thermal decomposition was not successful. In the HT-PXRD measurements at 500 and 520 °C, the formation of SnO2 can be observed. Above 560 °C, the HT-PXRD shows pure SnO2 (see Figure 7, 560, 580, and 700 °C (dark red circles)). 3.5. Electronic Structure Calculations. The structural parameters obtained by a full relaxation of the crystal structure are slightly overestimated with the two applied functionals PBE and HSE06 (see Table S6). The calculated cell volume
Figure 7. XRD patterns (Cu Kα radiation, λ = 1.54178 Å) of Sn3[B3O7]I and its thermal decomposition at 25, 380, 390, 440, 500, 520, 560, 580, and 700 °C. The reflections marked with a red asterisk (25 °C) derive from an unknown impurity. The reflection encircled in red in the pattern at 390 and 440 °C cannot be assigned to the original compound and occurs and disappears in the range between 390 and 500 °C. The reflections circled in red in the patterns at 500 °C and above 500 °C stem from SnO2.
increased by 6.3% (PBE) and 3.7% (HSE06). This is a wellknown effect for exchange correlation functionals based on the generalized gradient approximation. The electronic band structure reveals an indirect band gap with the conduction band minimum located at the Γ-Z line for both methods
Figure 6. Simultaneous thermal analysis (STA) of Sn3[B3O7]I, with the thermogravimetric curve (TG) in red, the differential thermogravimetric curve (DTG) in blue, and the differential thermal analysis curve (DTA) in green. E
DOI: 10.1021/acs.inorgchem.9b00377 Inorg. Chem. XXXX, XXX, XXX−XXX
Article
Inorganic Chemistry
Figure 8. (a) Isosurfaces of the electron localization function (η = 0.85) representing the O lone pairs, covalent B−O bonds, the stereochemically active Sn2+ lone pairs, and I lone pairs. Color codes: yellow is the isosurface, the boron in red, the oxygen in blue, the tin cation in gray, and the iodine in violet. (b) Electronic band structure as obtained with PBE (black) and HSE06 (yellow); the inserted arrows highlight the indirect nature of the band gap. Note that only the 12 highest occupied bands and eight lowest unoccupied bands are shown for both functionals.
(Figure 8b), while the valence band maximum is at Γ with HSE06 and at X with PBE. The size of the band gap as obtained with PBE (2.37 eV) seems more plausible in the context of the yellow-greenish color of Sn3[B3O7]I, whereas the hybrid functional HSE06 seems to overestimate the gap (3.07 eV). A real space bonding analysis via the electron localization function (ELF) displays the covalent B−O bonds and the lone pairs located at the oxygen atoms in the borate network. Further, very large areas around the attractors are obtained for the lone pairs of I (see Figure 8a). Isosurfaces of stereochemically active lone pairs of the Sn atoms represent a further coordination partner resulting in a 5-fold coordination for Sn1 and Sn3 and a 4-fold coordination sphere around Sn2. Note that the lone pair of Sn2 is not visible with the chosen value of the ELF as shown in Figure 8; it will become visible at η = 0.6.
Accession Codes
CCDC 1893973 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing
[email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
■
Corresponding Author
*Fax: +43 512 50757099. E-mail:
[email protected]. at. ORCID
Hubert Huppertz: 0000-0002-2098-6087 Notes
The authors declare no competing financial interest.
■
4. CONCLUSION In this work, we present the first tin iodide borate Sn3[B3O7]I synthesized by a simple hydrothermal synthesis. B3O8 rings are connected to chains along the b axis. Between them, the Sn2 cations are located like layers, and the Sn12+ and Sn32+ are in a row with the chains. This new compound shows a thermal stability up to 380 °C. HT-PXRD measurements indicate the formation of an additional reflection at 11.7° (2θ) and 390 °C, which cannot be assigned to the original compound. This reflection shows a maximum at 440 °C and disappears at 500 °C and may be due to a new phase being formed briefly in this temperature range. Furthermore, the Sn3[B3O7]I phase still exists at this temperatures.
■
AUTHOR INFORMATION
ACKNOWLEDGMENTS F.P. thanks Prof. Bettina Lotsch, Dr. Ulrich Wedig, and the Computer Service group from the Max-Planck Institute for Solid State Research (Stuttgart, Germany) for access to CRYSTAL17 and computational facilities.
■
REFERENCES
(1) Huang, Y.-Z.; Wu, L.-M.; Wu, X.-T.; Li, L.-H.; Chen, L.; Zhang, Y.-F. Pb2B5O9I: An Iodide Borate with Strong Second Harmonic Generation. J. Am. Chem. Soc. 2010, 132, 12788−12789. (2) Yu, H.; Koocher, N. Z.; Rondinelli, J. M.; Halasyamani, P. S. Pb2BO3I: A Borate Iodide with the Largest Second-Harmonic Generation (SHG) Response in the KBe2BO3F2 (KBBF) Family of Nonlinear Optical (NLO) Materials. Angew. Chem., Int. Ed. 2018, 57, 6100−6103. (3) Li, H.; Wu, H.; Su, X.; Yu, H.; Pan, S.; Yang, Z.; Lu, Y.; Han, J.; Poeppelmeier, K. R. Pb3B6O11F2: the first non-centrosymmetric lead borate fluoride with a large second harmonic generation response. J. Mater. Chem. C 2014, 2, 1704−1710. (4) Plachinda, P. A.; Dolgikh, V. A.; Stefanovich, S. Y.; Berdonosov, P. S. Nonlinear-optical susceptibility of hilgardite-like borates M2B5O9X (M = Pb,Ca,Sr,Ba; X = Cl,Br). Solid State Sci. 2005, 7, 1194−1200. (5) Egorova, B. V.; Olenev, A. V.; Berdonosov, P. S.; Kuznetsov, A. N.; Stefanovich, S. Y.; Dolgikh, V. A.; Mahenthirarajah, T.; Lightfoot, P. Lead−strontium borate halides with hilgardite-type structure and their SHG properties. J. Solid State Chem. 2008, 181, 1891−1898.
ASSOCIATED CONTENT
S Supporting Information *
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.9b00377. Wyckoff positions, atomic coordinates, anisotropic displacement parameters, interatomic distances and bond angles, and DFT optimized lattice parameters as well as a comparison of Sn3[B3O7]I and Sn3[B3O7]F (PDF) F
DOI: 10.1021/acs.inorgchem.9b00377 Inorg. Chem. XXXX, XXX, XXX−XXX
Article
Inorganic Chemistry (6) Belokoneva, E. L.; Al-Ama, A. G.; Stefanovich, S. Y.; Plachinda, P. A. Crystal structure of the lead bromo-borate Pb2[B5O9]Br from precision single-crystal X-ray diffraction data and the problem of optical nonlinearity of hilgardites. Crystallogr. Rep. 2007, 52, 795− 800. (7) Wu, Y.; Sasaki, T.; Nakai, S.; Yokotani, A.; Tang, H.; Chen, C. CsB3O5: A new nonlinear optical crystal. Appl. Phys. Lett. 1993, 62, 2614−2615. (8) Chen, C.; Wu, B.; Jiang, A.; You, G. A new type ultraviolet SHG crystal p-BaB204. Sci. Sin. (Engl. Ed.) 1985, B28, 235−243. (9) Tu, J.-M.; Keszler, D. A. CsLiB6O10 A noncentrosymmetric polyborate. Mater. Res. Bull. 1995, 30, 209−215. (10) Mori, Y.; Kuroda, I.; Nakajima, S.; Sasaki, T.; Nakai, S. New nonlinear optical crystal: Cesium lithium borate. Appl. Phys. Lett. 1995, 67, 1818−1820. (11) Furusawa, S.-i.; Chikagawa, O.; Tange, S.; Ishidate, T.; Orihara, H.; Ishibashi, Y.; Miwa, K. Second Harmonic Generation in Li2B4O7. J. Phys. Soc. Jpn. 1991, 60, 2691−2693. (12) Lu, Z.; Zhang, F.; Tudi, A.; Zhang, Z.; Yang, Z.; Pan, S. Sn2B7O12F with a 2∞[B14O24]−6 layer constructed by the unprecedented [B7O16]−11 fundamental building block. Inorg. Chem. Front. 2019, DOI: 10.1039/C9QI00120D. (13) Zou, G.; Zhang, L.; Ye, N. Synthesis, structure, and characterization of a new promising nonlinear optical crystal: Cd5(BO3)3F. CrystEngComm 2013, 15, 2422−2427. (14) Yu, H.; Wu, H.; Pan, S.; Yang, Z.; Su, X.; Zhang, F. A novel deep UV nonlinear optical crystal Ba3B6O11F2, with a new fundamental building block, B6O14 group. J. Mater. Chem. 2012, 22, 9665−9670. (15) Zhang, Z.; Wang, Y.; Zhang, B.; Yang, Z.; Pan, S. Polar Fluorooxoborate, NaB4O6F: A Promising Material for Ionic Conduction and Nonlinear Optics. Angew. Chem., Int. Ed. 2018, 57, 6577−6581. (16) Li, L.; Li, G.; Wang, Y.; Liao, F.; Lin, J. Bismuth Borates: OneDimensional Borate Chains and Nonlinear Optical Properties. Chem. Mater. 2005, 17, 4174−4180. (17) Jantz, S. G.; Dialer, M.; Bayarjargal, L.; Winkler, B.; van Wüllen, L.; Pielnhofer, F.; Brgoch, J.; Weihrich, R.; Hö ppe, H. A. Sn[B2O3F2]The First Tin Fluorooxoborate as Possible NLO Material. Adv. Opt. Mater. 2018, 6, 1800497. (18) Wang, X.; Wang, Y.; Zhang, B.; Zhang, F.; Yang, Z.; Pan, S. CsB4O6F: A Congruent-Melting Deep-Ultraviolet Nonlinear Optical Material by Combining Superior Functional Units. Angew. Chem., Int. Ed. 2017, 56, 14119−14123. (19) Mutailipu, M.; Zhang, M.; Zhang, B.; Wang, L.; Yang, Z.; Zhou, X.; Pan, S. SrB5O7F3 Functionalized with [B5O9F3]6− Chromophores: Accelerating the Rational Design of Deep-Ultraviolet Nonlinear Optical Materials. Angew. Chem., Int. Ed. 2018, 57, 6095−6099. (20) Schönegger, S.; Jantz, S. G.; Saxer, A.; Bayarjargal, L.; Winkler, B.; Pielnhofer, F.; Höppe, H. A.; Huppertz, H. Synthesis and Characterization of the First Tin Fluoride Borate Sn3[B3O7]F with Second Harmonic Generation Response. Chem. - Eur. J. 2018, 24, 16036−16043. (21) SADABS, v2014/5; Bruker AXS Inc.: Madison, WI, 2001. (22) Sheldrick, G. M. SHELXL-2013/1, Program Suite for the Solution and Refinement of Crystal Structures; University of Göttingen: Göttingen, Germany, 2013. (23) Sheldrick, G. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, C71, 3−8. (24) Opus, v7.2; Bruker: Billerica, MA, 2012. (25) Dovesi, R.; Saunders, V. R.; Roetti, C.; Orlando, R.; ZicovichWilson, C. M.; Pascale, F.; Civalleri, B.; Doll, K.; Harrison, N. M.; Bush, I. J.; D’Arco, P.; Llunell, M.; Causa, M.; Noel, Y. CRYSTAL14 User’s Manual; University of Torino: Torino, Italy, 2014. (26) Dovesi, R.; Orlando, R.; Erba, A.; Zicovich-Wilson, C. M.; Civalleri, B.; Casassa, S.; Maschio, L.; Ferrabone, M.; De La Pierre, M.; D’Arco, P.; Noel, Y.; Causa, M.; Rerat, M.; Kirtman, B. CRYSTAL14: A program for the ab initio investigation of crystalline solids. Int. J. Quantum Chem. 2014, 114, 1287.
(27) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868. (28) Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207−8215. (29) Orlando, R.; Dovesi, R.; Roetti, C.; Saunders, V. R. Ab initio Hartree-Fock calculations for periodic compounds: application to semiconductors. J. Phys.: Condens. Matter 1990, 2, 7769. (30) Scaranto, J.; Giorgianni, S. A quantum-mechanical study of CO adsorbed on TiO2: A comparison of the Lewis acidity of the rutile (110) and the anatase (101) surfaces. J. Mol. Struct.: THEOCHEM 2008, 858, 72−76. (31) Nada, R.; Catlow, C. R. A.; Pisani, C.; Orlando, R. An ab-initio Hartree-Fock perturbed-cluster study of neutral defects in LiF. Modell. Simul. Mater. Sci. Eng. 1993, 1, 165. (32) Doll, K.; Stoll, H. Ground-state properties of heavy alkali halides. Phys. Rev. B: Condens. Matter Mater. Phys. 1998, B57, 4327− 4331. (33) Becke, A. D.; Edgecombe, K. E. A simple measure of electron localization in atomic and molecular systems. J. Chem. Phys. 1990, 92, 5397−5403. (34) Silvi, B.; Savin, A. Classification of chemical bonds based on topological analysis of electron localization functions. Nature 1994, 371, 683. (35) Gatti, C.; Cassassa, S. Topond14 User’s Manual; University of Torino: Torino, Italy, 2016. http://www.crystal.unito.it/topond/ topond.pdf. (36) Kokalj, A. Computer graphics and graphical user interfaces as tools in simulations of matter at the atomic scale. Comput. Mater. Sci. 2003, 28, 155−168. (37) Burns, P. C.; Grice, F. C.; Hawthorne, F. C. Hydrogen bonding in borcarite, an unusual borate-carbonate mineral. Mineral. Mag. 1995, 59, 297−304. (38) Zobetz, E. Geometry and simple model calculations for borate (BO4) groups. Z. Kristallogr. 1990, 191, 45−57. (39) Pauling, L. Atomic Radii and Interatomic Distances in Metals. J. Am. Chem. Soc. 1947, 69, 542. (40) Byström, A.; Wilhelmi, K.-A.; Nesland, A. The Crystal Structure of (NH4)2Cr2O7. With a Discussion of the Relation between Bond Number and Interatomic Distances. Acta Chem. Scand. 1951, 5, 1003. (41) Brown, I. D.; Altermatt, D. Bond-valence parameters obtained from a systematic analysis of the Inorganic Crystal Structure Database. Acta Crystallogr., Sect. B: Struct. Sci. 1985, 41 (B41), 244−247. (42) Brese, N. E.; O'Keeffe, M. O. Bond-valence parameters for solids. Acta Crystallogr., Sect. B: Struct. Sci. 1991, 47, 192−197. (43) O’Keeffe, M. Struct. Bonding (Berlin); Springer: Berlin, 1989; vol 71, pp 162−189. (44) Xia, H. R.; Li, L. X.; Teng, B.; Zheng, W. Q.; Lu, G. W.; Jiang, H. D.; Wang, J. Y. Raman scattering from bismuth triborate. J. Raman Spectrosc. 2002, 33, 278−282. (45) Hoppe, R.; Voigt, S.; Glaum, H.; Kissel, J.; Müller, H. P.; Bernet, K. J. A new route to charge distributions in ionic solids. J. LessCommon Met. 1989, 156, 105−122. (46) Machida, K.; Hata, H.; Okuno, K.; Adachi, G.; Shiokawa, J. Synthesis and characterization of divalent-europium (Eu2+) compounds, EuB4O7, EuB2O4 and Eu2B2O5. J. Inorg. Nucl. Chem. 1979, 41, 1425−1430. (47) Laperches, J. P.; Tarte, P. Spectres d’absorption infrarouge de borates de terres rares. Spectrochim. Acta 1966, 22, 1201−1210. (48) Ross, S. D. The vibrational spectra of some minerals containing tetrahedrally co-ordinated boron. Spectrochim. Acta 1972, 28, 1555− 1561.
G
DOI: 10.1021/acs.inorgchem.9b00377 Inorg. Chem. XXXX, XXX, XXX−XXX