Two Indium Sulfate Tellurites: Centrosymmetric In2(SO4)(TeO3)(OH)2

Jul 31, 2019 - In2(SO4)(TeO3)(OH)2(H2O) crystallized in centrosymmetric (CS) space group ... Simulated and experimental XRD powder patterns, IR and ...
0 downloads 0 Views 6MB Size
Article pubs.acs.org/IC

Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

Two Indium Sulfate Tellurites: Centrosymmetric In2(SO4)(TeO3)(OH)2(H2O) and Non-centrosymmetric In3(SO4)(TeO3)2F3(H2O) Ya-Ping Gong,† Yun-Xiang Ma,† Shao-Ming Ying,‡ Jiang-Gao Mao,† and Fang Kong*,†

Downloaded via NOTTINGHAM TRENT UNIV on August 1, 2019 at 00:39:23 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, People’s Republic of China ‡ Fujian Provincial Key Laboratory of Featured Materials in Biochemical Industry, Ningde Normal University, Ningde 352100, People’s Republic of China S Supporting Information *

ABSTRACT: Two new indium sulfate tellurites, namely, In2(SO4)(TeO3)(OH)2(H2O) and In3(SO4)(TeO3)2F3(H2O), were synthesized by hydrothermal method in a one-pot reaction. Their pure phase yields have been successfully optimized to 76% and 21%, respectively. In2(SO4)(TeO3)(OH)2(H2O) crystallized in centrosymmetric (CS) space group P21/n, while In3(SO4)(TeO3)2F3(H2O) formed a non-centrosymmetric (NCS) and chiral space group P212121. The CS compound features a 2D layered structure composed of 2D indium oxide layers decorated by sulfate tetrahedra and tellurite groups. The NCS compound displays a 3D network consisting of indium tellurite layers bridged by sulfate tetrahedra. Powder second harmonic generation measurements disclosed that In3(SO4)(TeO3)2F3(H2O) exhibits a weak frequency-doubling efficiency about 11% of the commercial KDP. Its powder laser damage threshold quantity was estimated to be 79.6 MW/cm2, which is about 36 times that of AGS. The two samples present wide optical band gaps of 4.86 and 4.10 eV, respectively, which were determined by Te, In, and O atoms based on density functional theory calculations.



INTRODUCTION Nonlinear optical (NLO) material, as an important branch of optical materials, has received broad and increasing research efforts, of which frequency-doubling crystals attract the utmost attention in application and scientific studies.1−3 Frequency doubling, also called second harmonic generation (SHG) or second-order nonlinearity, is only permitted in a substance without inversion symmetry.4−7 So, design and synthesis of non-centrosymmetric (NCS) structures are the fundamental problem to obtain new crystals with SHG properties. However, the chance of getting centrosymmetric (CS) structures is almost five times more than that of NCS compounds based on the results of ICSD (Inorganic Crystal Structure Database). Considerable efforts have been devoted to promote the probability of the formation of NCS structures.8−12 Some special building blocks were regarded as the functional units which are apt at constructing NCS and SHG active compounds, such as π-conjugated groups, cations containing a lone-pair electron (LPE), tetrahedral units, and d0 transition metals, etc.13−15 Most of the reported SHG oxides contain at least one kind of the above functional units, although sometimes they may not be the main contributions of their SHG origins. It is reported that combination of two different functional units to one compound is an effective method to create novel NCS and SHG materials.16−19 Additionally, different building © XXXX American Chemical Society

blocks are suitable for producing brand new anion frameworks and enrich the structure chemistry of the inorganic compounds.20−25 We have been interested in assembling lone-pair electrons and tetrahedral groups together to make new NCS architectures and SHG materials for decades.26−32 The LPE groups were focused on selenite or tellurite containing Se(IV) or Te(IV) cations. We created the first boroselenite B2Se2O7, which is composed of SeO3 triangle pyramids and BO4 tetrahedra and crystallized in the chiral space group of P212121.26 In the isomeric gallium tellurites Ga2(TeO3)3, α phase, consisting of TeO3 triangle pyramids and BO4 tetrahedra, is crystallized in the NCS space group of I4̅3d, while the β phase, formed by TeO3 groups and GaO6 octahedra, is centrosymmetric.29 When we introduce PO4 tetrahedra into d10 transition metal selenite systems, three new cadmium/zinc phosphate selenites with large optical band gap were isolated.32 Recently, we have been working on a metal−SO4−TeO3 system to expand the tetrahedral groups to sulfates, which were reported as the promising near-ultraviolet (UV) even deep-UV SHG materials.33−35 The known sulfate tellurites were mainly converged on lanthanide compounds, which often crystallized in the CS space group because of their large coordination Received: June 11, 2019

A

DOI: 10.1021/acs.inorgchem.9b01730 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. SEM images of In2(SO4)(TeO3)(OH)2(H2O)(a), In3(SO4)(TeO3)2F3(H2O) (b), and their elemental distribution maps.

Table 1. Crystal Data and Structural Refinements for the Title Compounds formula fw cryst syst space group a/Å b/Å c/Å α/deg β/deg γ/deg V/Å3 Z Dc/(g·cm−3) μ(Mo Kα)/mm−1 GOF on F2 Flack factor R1, wR2 [I > 2σ(I)]a R1, wR2 (all data)

In2(SO4)(TeO3)(OH)2(H2O) 553.33 monoclinic P21/n 8.8430(8) 6.7033(5) 13.7144(13) 90 90.260(8) 90 812.94(12) 4 4.521 9.486 1.079 0.0280, 0.0650 0.0350, 0.0701

In3(SO4)(TeO3)2F3(H2O) 866.74 orthorhombic P212121 8.3115(6) 9.4341(6) 14.8068(10) 90 90 90 1161.02(14) 4 4.959 11.113 1.063 0.02(5) 0.0280, 0.0663 0.0290, 0.0670

R1 = ∑||Fo| - |Fc||/∑|Fo|, wR2 = {∑w[(Fo)2 - (Fc)2]2/∑w[(Fo)2]2}1/2

a

resolution of 2 cm−1 at room temperature. The UV−vis−NIR diffuse reflectance spectrum was measured with a PE Lambda 900 UV−vis− NIR spectrophotometer at 200−2500 nm at room temperature. Thermogravimetric analyses (TGAs) were performed on Netzsch STA 449C with a heating rate of 10 °C/min under N2 atmosphere between 20 and 1000 °C. The powder frequency-doubling efficiency was studied by the method reported.46 The fundamental wavelength is 1064 nm produced by a Q-switched Nd:YAG laser. Sieved crystals of KDP (70−100 mesh) were used as the references. The powder laser damage thresholds (LDTs) of In2(SO4)(TeO3)(OH)2(H2O) and In3(SO4)(TeO3)2F3(H2O) in the particle size of 70−100 mesh were determined with single-pulse measurement method.47 The average input lasers’ energy density of single pulses was 25.0 mJ when the sample was damaged by the laser with a flat-top laser beam distribution (λ = 1064 nm and τ = 10 ns in a 1 Hz repetition). The area of the laser spot focused on the sample is 3.14 mm2. Preparations of In 2 (SO 4 )(TeO 3 )(OH) 2 (H 2 O) and In 3 (SO 4 )(TeO3)2F3(H2O). The single crystals of In2(SO4)(TeO3)(OH)2(H2O) and In3(SO4)(TeO3)2F3(H2O) were obtained simultaneously by hydrothermal reaction of In2O3, Al2(SO4)3·18(H2O), TeO2, LiF, and HF. The ingredients are as follows: In2O3 (0.277 g, 1.0 mmol), Al2(SO4)3·18(H2O) (0.103 g, 0.3 mmol), LiF (0.104 g, 4.0 mmol), TeO2 (0.159 g, 1.0 mmol), and HF (0.2 mL). The blend was mixed with 2.5 mL of deionized water and loaded into an autoclave equipped with a Teflon liner (20 mL), which was heated at 200 °C for

number.36−45 Only one structure of Al2(OH)2TeO3SO4 with group 13 elements was reported.43 Our efforts in the In(III)− SO4−TeO3 system afford two new indium sulfate tellurites, namely, centrosymmetric In2(SO4)(TeO3)(OH)2(H2O) and non-centrosymmetric In3(SO4)(TeO3)2F3(H2O). We will report their syntheses, crystal structures, thermal stabilities, and optical properties. Theoretical calculations were also performed.



EXPERIMENT SECTION

Caution! Hydrof luoric acid is toxic and corrosive. Avoid inhalation and contact with eyes and skin. Reagents and Instruments. The chemical reagents were bought from commercial sources: In2O3 (Aladdin, 99.99%), TeO2 (Aldrich, 99.9%), sulfuric acid (Sinopharm, 40%+), Al2(SO4)3·18H2O (Shanghai Jinsha Xingta Chemical Factory, 99.9%), hydrofluoric acid (Macklin, 40%+), and LiF (Aladdin, 99.9%). Powder X-ray diffraction (PXRD) was performed on a Rigaku MiniFlex II diffractometer using Cu Kα radiation in the angular range of 2θ = 10−70° with a step size of 0.02°. Elemental distribution maps were measured on a fieldemission scanning electron microscope (FESEM, JSM6700F) equipped with an energy-dispersive X-ray spectroscope (EDS, Oxford INCA). IR spectrum analysis was carried out on a Magna 750 FT-IR spectrometer using KBr as the diluent in 4000−400 cm−1 with a B

DOI: 10.1021/acs.inorgchem.9b01730 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry 2 days. Colorless plate and block shaped crystals of In2(SO4)(TeO3)(OH)2(H2O) and In3(SO4)(TeO3)2F3(H2O) were achieved in one pot (Supporting Information Figure S1). The pure phase of In2(SO4)(TeO3)(OH)2(H2O) with a yield of 76% (based on Te) was obtained by reacting In2O3 (1.0 mmol), TeO2 (1.0 mmol), and H2SO4 (0.2 mmol) in 3 mL of DI water. Pure In3(SO4)(TeO3)2F3(H2O) with a yield of 21% (based on Te) was isolated from hydrothermal reaction of In2O3 (1.5 mmol), TeO2 (1.0 mmol), HF (0.2 mmol), and H2SO4 (0.2 mL). Their purities were checked by PXRD, which are coincidence with its calculated pattern (Figure S2). As shown in Figure 1, elements of In, S, and Te are evenly dispersed in both crystals, while F element only exists in the NCS compound, which is consistent with the results determined from their crystal structures. Crystal Structure Determination. Diffraction data were collected on an Agilent Technologies SuperNova dual wavelength CCD diffractometer with graphite-monochromated Mo Kα radiation (λ = 0.71073 Å) at 293 K. Data reduction was performed by CrysAlisPro, and mutiscan method was applied in absorption corrections.48 The structure was solved by the direct method and refined by full matrix least-squares fitting on F2 by SHELX-97.49 Crystallographic data are summarized in Table 1, and important bond lengths are itemized in Table S1.

few examples of seven-coordinated indium atoms have been illustrated in metal selenites or tellurites. In addition to the pentagonal bipyramid, the only octahedron with face monocapped geometry was found in compound In2MoTe2O10 reported by our group in 2009 (Figure S3b).50 Eightcoordinated indium cube was realized in a complicated polyoxopalladate Na8H7[Pd12(SeO3)8(InO8)]3·24H2O (Figure S3c).52 Evidently, the coordinate modes of In3+ cations are rich. The tellurium atom in In2(SO4)(TeO3)(OH)2(H2O) shows a 3 + 1 coordination mode with three regular [1.855(4)−1.895(4) Å] and one extended [2.513(5) Å] Te− O bonds, which was also reported in a compound of In2MoTe2O10. The sulfur atom is linked with four oxygen atoms to a SO4 tetrahedron with the S−O bond lengths ranging from 1.464(5) to 1.477(5) Å, which bears a resemblance to those in metal sulfates.33−35 Bond valence calculations revealed that the valence state of In, Te, and S atoms should be +3, +4, and +6, respectively. The detailed total bond valences for In(1), In(2), Te(1), and S(1) are 3.017, 3.027, 4.145, and 6.016, respectively (Table S4).53,54 The In(1)O5(OH)2 pentagonal bipyramids are interconnected to a 1D linear chain via edge-sharing of O(1) and O(3) atoms (Figure S4a). Two In(2)O3(OH)2(H2O) octahedra are edge-shared to a In2O4(OH)4(H2O)2 dimer (Figure S4b). The In(2) dimers bridged the 1D In(1) chains into a 2D indium oxide layer with In(1)4In(2)4 eight member polyhedral rings (MPRs; Figure 3a). Sulfate is a bidentate ligand and bridges with one In(1) and one In(2) atom, while tellurite is a pentadentate anion group, chelating bidentately with one In(1) atom and bridging with two In(1) and two In(2) atoms (Figure 3b,c). The SO4 tetrahedra are locating at the 8-MPRs of the indium oxide layers. However, the TeO3 groups are capping on the indium polyhedra to strengthen the connection of the 2D layers (Figure 3d). The sulfate and tellurite groups are further related by the elongated [2.513(5) Å] Te−O weak bonds (Figure 3b,c). The interlamellar distance is about 8.16 Å



RESULTS AND DISCUSSION The single crystals of the two title compounds were achieved by hydrothermal method in one-pot synthesis. They feature different crystal morphologies and structures (Figure S1). In2(SO4)(TeO3)(OH)2(H2O) displays long transparent sticks with wedge ends, while In3(SO4)(TeO3)2F3(H2O) presents clustered blocks, which are corresponding to their individual structural features. Structure of In2(SO4)(TeO3)(OH)2 (H2O). In2(SO4)(TeO3)(OH)2(H2O) features a 2D layered structure composed of 2D indium oxide layers decorated by sulfate tetrahedra and tellurite groups (Figure 2). The asymmetric

calculated by ( a 2 + c 2 )/2. We evaluated the hydrogen bonds between water molecules or protonated oxygen with oxygen atoms in this compound (Table S2). Figure S5 displayed that the strong hydrogen bonds between O(8) and O(6) linked the adjacent 2D layers to a pseudo-3D framework while the hydrogen bonds observed between O(1w) and O(6) only reinforced the inner layer construction. Structure of In 3 (SO 4 )(TeO 3 ) 2 F 3 (H 2 O). In 3 (SO 4 )(TeO3)2F3(H2O) presents a 3D network composed of indium tellurite layers bridged by sulfate tetrahedral (Figure 4). Its asymmetric unit includes three Indium, two tellurium, one sulfur, 11 oxygen, and three fluorine atoms in general sites. Three indium atoms were six-coordinated in the octahedral geometry (Figure S6). In(1) and In(3) are linked with two fluorine and four oxygen atoms, while In(2) is coordinated with one fluorine, one water molecule, and four oxygen atoms. The In−O and In−F bond distances are in the ranges of 2.131(6)−2.249(7) and 2.077(6)−2.147(7) Å, respectively. The fluorine atoms in In(1)F2O4 octahedron are adjacent to each other in cis mode, while the two F ligands in In(3)F2O4 octahedron are on opposite sides of the indium atom in trans mode. Te(1) and Te(2) are three-coordinated with oxygen ligands in triangle pyramids with the Te−O bonds within range of 1.855(7)−1.902(6) Å. S(1) is linked with four oxygen atoms in a tetrahedral geometry with the S−O distances in the range of 1.442(7)−1.491(7) Å, which is consistent with those

Figure 2. View of the structure of In2(SO4)(TeO3)(OH)2(H2O) along the b axis.

unit of In2(SO4)(TeO3)(OH)2(H2O) contains two indium, one tellurium, one sulfur, and 10 oxygen atoms in general sites. In(1) is coordinated with five O2− anions in an equatorial plane and two protonated oxygen atoms in vertical axis to a InO5(OH)2 pentagonal bipyramid (Figure S3a). In(2) is connected with three O2− anions, two protonated oxygen atoms and one water molecule to a InO3(OH)2(H2O) octahedral geometry. The In−O bond distances fall in the range of 2.108(4)−2.401(4) Å, which is consistent with those in other indium tellurites.50−52 It should be mentioned that C

DOI: 10.1021/acs.inorgchem.9b01730 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. Indium oxide layer (a), connectivity schemes of sulfate (b) and tellurite (c), and 2D layer of In2(SO4)(TeO3)(OH)2(H2O).

member polyhedral rings (Figure 5a, S7). Pentadentate Te(1)O3 and Te(2)O3 groups are capping on the In(1)In(2)2In(3)2 and In(1)2In(2)In(3)2 polyhedral rings, respectively, forming a 2D indium tellurite layer parallel to the (0 0 1) plane (Figure 5b,c). The interlamellar space is half of the length of c of about 7.40 Å. Similarly in In2(SO4)(TeO3)(OH)2(H2O), the S(1)O4 tetrahedron is a bidentate ligand, bridging with one In(1) and one In(2) atom (Figure 5d). The bidentate sulfates bridge the indium tellurite layers into a 3D framework with 10 and 6 MPR spiral tunnels along the a axis (Figure 4). The hydrogen bonding in In 3 (SO 4 )(TeO3)2F3(H2O) is very weak; only feeble interactions between water molecules and O(1) atoms coming from the tellurite groups were found (Table S2). The only reported group 13 sulfate tellurite Al2(SO4)(TeO3)(OH)2 features a 2D wavy layer consisting of an indium oxide layer decorated with TeO3 and SO4 anions (Figure S8).43 In these three structures, the counter cations of In3+ or Al3+ polyhedra were edge and corner shared into a 2D metal oxide layer with five, six, or even eight MPRs. The multidentate tellurite groups were capping on these polyhedral rings. The LPEs of tellurites were pointed to the opposite directions, and the polarizations were canceled out. Interestingly, the tetrahedral sulfates present two different modes in these structures. In CS In2(SO4)(TeO3)(OH)2(H2O) and Al2(SO4)(TeO3)(OH)2, the sulfates were attaching on the metal oxide layers while the SO4 tetrahedra in NCS

Figure 4. 3D framework of In3(SO4)(TeO3)2F3(H2O).

in In2(SO4)(TeO3)(OH)2(H2O). The oxidation state of In, Te, and S atoms was proved to be +3, +4, and +6, respectively. The calculated total bond valences for In(1), In(2), In(3), Te(1), Te(2), and S(1) are 2.813, 2.949, 2.863, 3.919, 3.857, and 6.066, respectively (Table S4).53,54 In(2)FO4(OH2) and In(3)F2O4 octahedra are interconnected into a 1D wavy chain via corner-sharing and further joined together by In(1)F2O4 octahedra to a 2D layer with In(1)In(2)2In(3)2 and In(1)2In(2)In(3)2, two different five D

DOI: 10.1021/acs.inorgchem.9b01730 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 5. Indium oxide layer (a), connectivity scheme of tellurites (b), indium tellurite layer (c), and connectivity scheme of sulfate (d) in In3(SO4)(TeO3)2F3(H2O).

Some of the reported tellurites with large band gaps were listed in Table S5. From the table we can find that metal tellurites with Eg value larger than 4.0 eV are not common. Thermal Gravimetric Analysis. The TGA curves of the two compounds disclosed their similar decomposition process (Figure 6). They can remain stable up to 350 °C and then undergo four steps of weight losses. The first weight loss corresponds to the release of one molecule coordination water. The second step occurred at 477−586 and 482−570 °C belonging to the loss of one molecule constitution water and partial fluorine in In 2 (SO 4 )(TeO 3 )(OH) 2 (H 2 O) and In3(SO4)(TeO3)2F3(H2O), respectively. The evaporation of sulfate molecule occurred at the third step, and then TeO2 began to decompose. The tellurites have not been eliminated completely at the end of the experiments. Figure 6 shows that the observed values are very close to the calculated weight losses. Powder SHG and LDT measurements. Qualitative SHG detection was performed on the crystalline sample of In3(SO4)(TeO3)2F3(H2O) since it is crystallized in the NCS space group. On the basis of the UV−vis−NIR spectrum, this compound is transparent between 500 and 1500 nm, so a Qswitched Nd:YAG 1064 nm laser was chose as the fundamental radiation. Clear green signals were observed by naked eyes. To obtain more accurate results, a sieved sample (70−100 mesh) was used to assess its second-order susceptibility coefficient.

In3(SO4)(TeO3)2F3(H2O) bridged the adjacent layers to a 3D framework. The arrangements of the trivalent metals play an important role in the symmetry of these structures. Even member polyhedral rings prefer CS configuration, while odd rings support NCS conformation. The introduction of fluorine should be the vital contributor to the asymmetric permutation pattern. In3(SO4)(TeO3)2F3(H2O) is the first fluoride sulfate tellurite compound and the first NCS sulfate tellurite structure too. IR and UV−Vis−NIR Spectra. The IR spectra of In2(SO4)(TeO3)(OH)2(H2O) and In3(SO4)(TeO3)2F3(H2O) present the charcteristic frequencies of sulfate, tellurite, and aqua ligands (Figure S9).55 The stretching and bending vibrations of the coordination waters in two compounds appear at about 3300 and 1600 cm−1, respectively. Figure S9a shows the stretching vibrations of hydroxy groups in In2(SO4)(TeO3)(OH)2(H2O) at 3554 cm−1. The strong absorptions range from 1200 to 900 cm−1 can be assigned to (SO4)2− tetrahedra. The charcteristic peaks for (TeO3)2− groups can be seen between 700 and 600 cm−1. UV−vis− NIR diffuse reflectance spectra indicate that the two samples are transparent between 500 and 1500 nm (Figure S10).56,57 The minor broad bands after 1500 nm can be ascribed to the water molecules. The insets revealed that In2(SO4)(TeO3)(OH)2(H2O) and In3(SO4)(TeO3)2F3(H2O) represent a wide optical band gap with values of 4.86 and 4.10 eV, respectively. E

DOI: 10.1021/acs.inorgchem.9b01730 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 7. Oscilloscope traces of the SHG signals for the sieved samples (70−100 mesh) of In3(SO4)(TeO3)2F3(H2O) and KDP.

Figure 6. TGA curves of In2(SO4)(TeO3)(OH)2(H2O) (a) and In3(SO4)(TeO3)2F3(H2O) (b).

Powder SHG examination revealed that In 3 (SO 4 )(TeO3)2F3(H2O) displays a weak frequency-doubling efficiency about 11% of commercial KDP (Figure 7). The weak SHG signal can be predicted by the opposite arrangements of the TeO3 triangular pyramids in the structure of In3(SO4)(TeO3)2F3(H2O) (Figure 4). Figure S10 indicates that In2(SO4)(TeO3)(OH)2(H2O) and In3(SO4)(TeO3)2F3(H2O) present wide optical band gaps of 4.86 and 4.10 eV, which are favorable for the large laser damage thresholds (LDTs). So it is worth estimating the LDTs quantity of two compounds preliminarily. A Q-switched pulse laser with flat-top laser beam distribution was employed to evaluate its powder LDTs. Sieved crystals of AGS were used as the reference. The powder LDT of In2(SO 4)(TeO3)(OH)2(H2O) (79.6 MW/cm2) and In3(SO4)(TeO3)2F3(H2O) (88.2 MW/cm2) are about 40 and 36 times that of AGS (2.2 MW/cm2), respectively, which are comparable with that of monoclinic Ga2S3 (30 × AGS).47 Theoretical Studies. The electronic properties of the title compounds were calculated by CASTEP based on DFT methods. Figure 8 presents their band structures along high symmetry points of the first Brillouin zone. From the curves we can find that the two structures displayed different features.

Figure 8. Calculated band structures of In 2 (SO 4 )(TeO 3 )(OH)2(H2O) (a) and In3(SO4)(TeO3)2F3(H2O) (b).

The top of the valence bands (VBs) for In 3 (SO 4 )(TeO3)2F3(H2O) are flat and sparse, while they are uneven and dense for In2(SO4)(TeO3)(OH)2(H2O). The bottom of conduction bands (CBs) for the two compounds fluctuated F

DOI: 10.1021/acs.inorgchem.9b01730 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

are constituted by Te 5s and H 1s mainly. The CBs in the higher energy from 7.0 to 16.0 eV are composed with In 5p, S 3p, In 5s, O 2p, and H 1s states primarily. To understand more about the atoms’ interaction, we concentrate our attention on the VBs and CBs close to the Fermi level (from −7.8 to 7.0 eV), which accounts for the majority of the bonding character. In these areas, we can find that the O 2p states match perfectly with Te 5p and S 3p states, which points out the definite Te− O and S−O covalent bonding. Furthermore, the PDOS of the two compounds disclosed that the top of the VBs is dominated by O 2p nonbonding states and that the bottom of the CBs is from the empty Te 5p and In 5s orbitals largely. Consequently, the band gaps of the title compounds are determined by Te, In, and O atoms. In3(SO4)(TeO3)2F3(H2O) crystallizes in the chiral space group P212121, belonging to class 222. According to Kleinman symmetry, In3(SO4)(TeO3)2F3(H2O) has one independent SHG tensor d14 (d14 = d25 = d36), which is calculated to be 7.81 × 10−10 esu in the static limit. The results are consistent with the weak SHG signal in the experimental measurement. To better interpret the SHG origin, the SHG density of d14 has been analyzed (Figure 10). In VB, the positive contribution to

moderately. The state energies of the lowest CB and highest VB of the structures are shown in Table S3. For In2(SO4)(TeO3)(OH)2(H2O), the top of VBs and the bottom of CBs are situated at the same point (G) with band gap of 3.79 eV, indicating it is a direct band gap compound. For In3(SO4)(TeO3)2F3(H2O), the top of the VBs is located at G point while the bottom of CBs is placed at X point with the band gap of 2.79 eV, which shows it is an indirect band gap material. The calculated band gaps are much smaller than the experimental results due to the limitation of the DFT methods.58 The total and partial density of states (TDOS and PDOS) were shown in Figure 9. The TDOS and PDOS of the two structures are very similar except for the F orbitals in compound In3(SO4)(TeO3)2F3(H2O). Specifically, the VBs in the lower energy from −23.4 to −15.8 eV are mostly originated from O 2s and F 2s in In3(SO4)(TeO3)2F3(H2O)], S 3s3p, Te 5s5p, and H 1s states. The peaks around −10.0 eV

Figure 10. SHG density of d14 in the VB (a) and the CB (b) for In3(SO4)(TeO3)2F3(H2O).

SHG mainly comes from O atoms in SO4 units, but other O and F atoms contribute negatively, causing the weakening of the SHG effect. In CB, Te 5p and O 2p in Te(1)O3 groups make the major contributions. The contribution percentages of SO42−, InO4F2/InO5F, TeO3, and H2O groups are 77.95%, 6.18%, 13.36%, and 2.17%, respectively. Hence, SO4 groups are the main source of the SHG effect in In 3 (SO 4 )(TeO3)2F3(H2O).



CONCLUSION The first indium sulfate tellurites, namely, In2(SO4)(TeO3)(OH)2(H2O) and In3(SO4)(TeO3)2F3(H2O), were successfully synthesized by hydrothermal method. Centrosymmetric In2(SO4)(TeO3)(OH)2(H2O) exhibits a 2D layered structure composed of 2D indium oxide layers decorated by sulfate tetrahedra and tellurite triangular pyramids, while NCS In3(SO4)(TeO3)2F3(H2O) features a 3D network consisting of indium tellurite layers bridged by sulfate tetrahedra. The coordination mode of sulfate plays a key role in the structural dimension. Attached sulfate tetrahedra tend to form a 2D layered structure, while bridged sulfates prefer a 3D framework. The introduction of fluorine elements changed the coordination geometries of the balance cations, which further influenced the arrangements of the polyhedra in structure. Crystal structures with even member polyhedral rings are in favor of CS configuration, while those with odd rings would make for NCS conformation. The permutation of the predistorted tellurite groups is the dominant factor for the SHG efficiency of the NCS samples. In In 3 (SO 4 )(TeO3)2F3(H2O), the opposite arrangements of the TeO3

Figure 9. Full and partial density of states for In2(SO4)(TeO3)(OH)2(H2O) (a) and In3(SO4)(TeO3)2F3(H2O) (b). G

DOI: 10.1021/acs.inorgchem.9b01730 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

(9) Liang, M. L.; Hu, C. L.; Kong, F.; Mao, J. G. BiFSeO3: An Excellent SHG Material Designed by Aliovalent Substitution. J. Am. Chem. Soc. 2016, 138, 9433−9436. (10) 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. 2017, 129, 14307−14311. (11) Wang, Q.; Yang, F.; Wang, X.; Zhou, J.; Ju, J.; Huang, L.; Gao, D.; Bi, J.; Zou, G. Deep-Ultraviolet Mixed-Alkali-Metal Borates with Induced Enlarged Birefringence Derived from the Structure Rearrangement of the LiB3O5. Inorg. Chem. 2019, 58, 5949−5955. (12) Dong, X.; Huang, L.; Liu, Q.; Zeng, H.; Lin, Z.; Xu, D.; Zou, G. Perfect balance harmony in Ba2NO3(OH)3: a beryllium-free nitrate as a UV nonlinear optical material. Chem. Commun. 2018, 54, 5792− 5795. (13) Halasyamani, P. S. Asymmetric Cation Coordination in Oxide Materials: Influence of Lone-Pair Cations on the Intra-octahedral Distortion in d0 Transition Metals. Chem. Mater. 2004, 16, 3586− 3592. (14) Kong, F.; Sun, C. F.; Yang, B. P.; Mao, J. G. Second-Order Nonlinear Optical Materials Based on Metal Iodates, Selenites, and Tellurites. Structure-Property Relationships in Non-Linear Optical Cyrstals I; Structure and Bonding; Springer: Berlin, 2012; Vol. 144, pp 43−104, DOI: 10.1007/430_2011_65. (15) Zou, G.; Lin, Z.; Zeng, H.; Jo, H.; Lim, S. J.; You, T. S.; Ok, K. M. Cs3VO(O2) 2CO3: an exceptionally thermostable carbonatoperoxovanadate with an extremely large second-harmonic generation response. Chem. Sci. 2018, 9, 8957−8961. (16) Liu, Y. S.; Dentz, D.; Belt, R. High-Average-Power Intracavity Second-Harmonic Generation Using KTiOPO4 in an AcoustoOptically Q-Switched Nd:YAG Laser Oscillator at 5 kHz. Opt. Lett. 1984, 9, 76−78. (17) Sykora, R. E.; Ok, K. M.; Halasyamani, P. S.; Albrecht-Schmitt, T. E. Structural Modulation of Molybdenyl Iodate Architectures by Alkali Metal Cations in AMoO3(IO3) (A = K, Rb, Cs): A Facile Route to New Polar Materials with Large SHG Responses. J. Am. Chem. Soc. 2002, 124, 1951−1957. (18) Zou, G.; Jo, H.; Lim, S. J.; You, T. S.; Ok, K. M. Rb3VO(O2)2CO3: A Four-in-One Carbonatoperoxovanadate Exhibiting an Extremely Strong Second-Harmonic Generation Response. Angew. Chem., Int. Ed. 2018, 57, 8619−8622. (19) Xia, M. J.; Jiang, X. X.; Lin, Z. S.; Li, R. K. All-Three-in-One”: A New Bismuth-Tellurium-Borate Bi3TeBO9 Exhibiting Strong Second Harmonic Generation Response. J. Am. Chem. Soc. 2016, 138, 14190−14193. (20) Gong, Y. P.; Hu, C. L.; Kong, F.; Mao, J. G. Exploration of New Birefringent Crystals in Bismuth d0 Transition Metal Selenites. Chem. - Eur. J. 2019, 25, 3685−3694. (21) Lu, W. Q.; Gao, Z. L.; Wu, Q.; Tian, X. X.; Sun, Y. X.; Liu, Y.; Tao, X. T. Tailored fabrication of a prospective acousto-optic crystal TiTe3O8 endowed with high performance. J. Mater. Chem. C 2018, 6, 2443−2451. (22) Oh, S. J.; Lim, S. J.; You, T. S.; Ok, K. M. From a Metastable Layer to a Stable Ring: A Kinetic Study for Transformation Reactions of Li2Mo3TeO12 to Polyoxometalates. Chem. - Eur. J. 2018, 24, 6712− 6716. (23) Wu, Q.; Gao, Z. L.; Wang, H. Y.; Li, C. G.; Lu, W. Q.; Tian, X. X.; Sun, Y. X.; Lin, N.; Xia, S. Q.; Tao, X. T. Polymorphic Smooth Interfaces Formation Based on the Biphasic BaTeMo2O9 Using Top Multi-Seeded Growth. Cryst. Growth Des. 2018, 18, 5054−5062. (24) Li, C. G.; Tian, X. X.; Gao, Z. L.; Wu, Q.; Zhao, P.; Sun, Y. X.; Zhang, C. Q.; Zhang, S. J.; Cui, D. L.; Tao, X. T. Controlled Growth of Layered Acentric CdTeMoO6 Single Crystals with Linear and Nonlinear Optical Properties. Cryst. Growth Des. 2018, 18, 3376− 3384. (25) Feng, X. X.; Zhang, J. J.; Gao, Z. L.; Zhang, S. J.; Sun, Y. X.; Tao, X. T. Investigation of the second-order nonlinear optical properties of Cs2TeMo3O12 single crystal. Appl. Phys. Lett. 2014, 104, 081912.

triangular pyramids result in the weak SHG signal of about 11% of commercial KDP. Next, we will make our efforts in fluoride structures to synthesize new NCS sulfate tellurites with strong SHG efficiency.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.9b01730. Simulated and experimental XRD powder patterns, IR and UV−vis−NIR spectra, and computational method (PDF) Accession Codes

CCDC 1919983 and 1919984 contain 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.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Jiang-Gao Mao: 0000-0002-5101-8898 Fang Kong: 0000-0001-8538-5226 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (Grant Nos. 21773244 and 21875248), the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDB20000000), and the NSF of Fujian Province (Grant No. 2018J01025).



REFERENCES

(1) Chen, C.; Liu, G. Recent Advances in Nonlinear Optical and Electro-optical Materials. Annu. Rev. Mater. Sci. 1986, 16, 203−243. (2) Sekikawa, T.; Kosuge, A.; Kanai, T.; Watanabe, S. Nonlinear optics in the extreme ultraviolet. Nature 2004, 432, 605−608. (3) Mutailipu, M.; Zhang, M.; Yang, Z. H.; Pan, S. L. Targeting the Next Generation of Deep-Ultraviolet Nonlinear Optical Materials: Expanding from Borates to Borate Fluorides to Fluorooxoborates. Acc. Chem. Res. 2019, 52, 791−801. (4) Ok, K. M.; Chi, E. O.; Halasyamani, P. S. Bulk Characterization Methods for Non-Centrosymmetric Materials: Second-Harmonic Generation, Piezoelectricity, Pyroelectricity and Ferroelectricity. Chem. Soc. Rev. 2006, 35, 710−717. (5) Hu, C. L.; Mao, J. G. Recent Advances on Second-order NLO Materials based on Metal Iodates. Coord. Chem. Rev. 2015, 288, 1−17. (6) Li, Z.; Liu, Q.; Wang, Y.; Iitaka, T.; Su, H.; Tohyama, T.; Yang, Z.; Pan, S. Second-Harmonic Generation in Noncentrosymmetric Phosphates. Phys. Rev. B: Condens. Matter Mater. Phys. 2017, 96, 035205. (7) Pan, Y.; Guo, S. P.; Liu, B. W.; Xue, H. G.; Guo, G. C. Secondorder nonlinear optical crystals with mixed anions. Coord. Chem. Rev. 2018, 374, 464−496. (8) Song, J. L.; Hu, C. L.; Xu, X.; Kong, F.; Mao, J. G. A Facile Synthetic Route to a New SHG Material with Two Types of Parallel π-Conjugated Planar Triangular Units. Angew. Chem., Int. Ed. 2015, 54, 3679−3682. H

DOI: 10.1021/acs.inorgchem.9b01730 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (26) Kong, F.; Huang, S. P.; Sun, Z. M.; Mao, J. G.; Cheng, W. D. Se2(B2O7): A New Type of Second-Order NLO Material. J. Am. Chem. Soc. 2006, 128, 7750−7751. (27) Kong, F.; Jiang, H. L.; Mao, J. G. La4(Si5.2Ge2.8O18)(TeO3)4 and La2(Si6O13)(TeO3)2: Intergrowth of the lanthanum(III) tellurite layer with the XO4 (X = Si/Ge) tetrahedral layer. J. Solid State Chem. 2008, 181, 263−268. (28) Jennene Boukharrata, N.; Laval, J. P.; Thomas, P. Two new types of oxyfluorotellurates(IV): ScTeO3F and InTeO3F. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2008, 64, i57−i61. (29) Kong, F.; Xu, X.; Mao, J. G. A Series of New Ternary and Quaternary Compounds in the LiI-GaIII-TeIV-O System. Inorg. Chem. 2010, 49, 11573−11580. (30) Zhang, J. H.; Kong, F.; Yang, B. P.; Mao, J. G. A series of boroselenite-based open frameworks mediated by the cationic sizes of the alkali metals. CrystEngComm 2012, 14, 8727−8733. (31) Qian, C.; Kong, F.; Mao, J. G. Syntheses, structures and characterizations of three novel vanadium selenites with organically bonded copper/nickel complex. J. Solid State Chem. 2016, 238, 1−8. (32) Ma, Y. X.; Gong, Y. P.; Hu, C.; Mao, J. G.; Kong, F. Three New d10 Transition Metal Selenites Containing PO4 Tetrahedron: Cd 7 (HPO 4 ) 2 (PO 4 ) 2 (SeO 3 ) 2 , Cd 6 (PO 4 ) 1 . 3 4 (SeO 3 ) 4 . 6 6 and Zn3(HPO4)(SeO3)2(H2O). J. Solid State Chem. 2018, 262, 320−326. (33) Dong, X. H.; Huang, L.; Hu, C. F.; Zeng, H. M.; Lin, Z. E.; Wang, X.; Ok, K. M.; Zou, G. H. CsSbF2SO4: An Excellent Ultraviolet Nonlinear Optical Sulfate with a KTiOPO4 (KTP)-Type Structure. Angew. Chem., Int. Ed. 2019, 58, 6528−6534. (34) Li, Y. Q.; Liang, F.; Zhao, S. G.; Li, L. N.; Wu, Z. Y.; Ding, Q. R.; Liu, S.; Lin, Z. S.; Hong, M. C.; Luo, J. H. Two Non-π-Conjugated Deep-UV Nonlinear Optical Sulfates. J. Am. Chem. Soc. 2019, 141, 3833−3837. (35) He, F. F.; Wang, Q.; Hu, C. F.; He, W.; Luo, X. Y.; Huang, L.; Gao, D. J.; Bi, J.; Wang, X.; Zou, G. H. Centrosymmetric (NH4)2SbCl(SO4)2 and Non-centrosymmetric (NH4)SbCl2(SO4): Synergistic Effect of Hydrogen-Bonding Interactions and Lone-Pair Cations on the Framework Structures and Macroscopic Centricities. Cryst. Growth Des. 2018, 18, 6239−6247. (36) Weil, M.; Shirkhanlou, M. Incorporation of Sulfate or Selenate Groups into Oxotellurates(IV): I. Calcium, Cadmium, and Strontium Compounds. Z. Anorg. Allg. Chem. 2017, 643, 330−339. (37) Guo, W. B.; Tang, Y. Y.; Wang, J. F.; He, Z. Z. Layered Cu7(TeO3)2(SO4)2(OH)6 with Diluted Kagomé Net Containing Frustrated Corner-Sharing Triangles. Inorg. Chem. 2017, 56, 1830− 1834. (38) Poe, T.; White, F.; Proust, V.; Villa, E.; Polinski, M. [Ag2M(Te2O5)2]SO4 (M = CeIV or ThIV): A New Purely Inorganic d/f -Heterometallic Cationic Material. Inorg. Chem. 2018, 57, 4816− 4819. (39) Frost, R.; Keeffe, E. Raman spectroscopic study of the tellurite mineral: poughite Fe23+SO4(TeO3)2·3H2O-a multi-anion mineral. J. Raman Spectrosc. 2008, 39, 1794−1798. (40) Lin, J.; Diefenbach, K.; Silver, M.; Dalal, N.; Albrecht-Schmitt, T. Structure-Property Correlations in the Heterobimetallic 4f/3d Materials Ln2M(TeO3)2(SO4) (Ln = Y, Nd, Sm, Eu, Gd, Tb, Dy, Ho, Er, Tm, Yb, or Lu; M = Co or Zn). Cryst. Growth Des. 2015, 15, 4606−4615. (41) Tang, Y. Y.; He, Z. Z.; Guo, W. B.; Zhang, S. Y.; Yang, M. Syntheses and Magnetic Properties of New Tellurite-Sulfate Compounds M2(TeO3)(SO4)·H2O (M = Co, Mn) with a Layer Structure Showing a Distorted Honeycomb Spin-Lattice. Inorg. Chem. 2014, 53, 5862−5868. (42) Tang, H. X.; Zhang, Y. X.; Zhuo, C.; Fu, R. B.; Lin, H.; Ma, Z. J.; Wu, X. T. A Niobium Oxyiodate Sulfate with a Strong SecondHarmonic Generation Response Built by Rational Multi-Component Design. Angew. Chem., Int. Ed. 2019, 58, 3824−3828. (43) Johansson, G. B.; Lindqvist, O. The Crystal Structure of Al2(OH)2TeO3SO4. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1976, 32, 407−411.

(44) Lin, J.; Cross, J.; Diwu, J.; Meredith, N.; Albrecht-Schmitt, T. Comparisons of Plutonium, Thorium, and Cerium Tellurite Sulfates. Inorg. Chem. 2013, 52, 4277−4281. (45) Lin, J.; Chai, P.; Diefenbach, K.; Shatruk, M.; Albrecht-Schmitt, T. Challenges in the Search for Magnetic Coupling in 3d/4f Materials: Syntheses, Structures, and Magnetic Properties of the Lanthanide Copper Heterobimetallic Compounds, RE 2 Cu(TeO3)2(SO4)2. Chem. Mater. 2014, 26, 2187−2194. (46) Kurtz, S. K.; Perry, T. T. A Powder Technique for the Evaluation of Nonlinear Optical Materials. J. Appl. Phys. 1968, 39, 3798−3812. (47) Zhang, M. J.; Jiang, X. M.; Zhou, L. J.; Guo, G. C. Two phases of Ga2S3: promising infrared second-order nonlinear optical materials with very high laser induced damage thresholds. J. Mater. Chem. C 2013, 1, 4754−4760. (48) Crystal Clear, version 1.3.5; Rigaku.: Woodlands, TX, USA, 1999. (49) Sheldrick, G. M. SHELXTL-97, Crystallographic Software Package, version 5.1; Bruker-AXS: Madison, WI, USA, 1998. (50) Kong, F.; Hu, C. L.; Hu, T.; Zhou, Y.; Mao, J. G. Explorations of new phases in the GaIII/InIII−MoVI−SeIV/TeIV−O systems. Dalton Trans 2009, 25, 4962−4970. (51) Bae, S. W.; Kim, C. Y.; Lee, D. W.; Ok, K. M. Cooperative Effects of Cation Size and Variable Coordination Modes of Te4+ on the Frameworks of New Alkali Metal Indium Tellurites, NaIn(TeO3)2, KIn(TeO3)2, RbInTe3O8, and CsInTe3O8. Inorg. Chem. 2014, 53, 11328−11334. (52) Lin, Z. G.; Wang, B.; Cao, J.; Chen, B. K.; Xu, C.; Huang, X. Q.; Fan, Y. X.; Hu, C. W. Controlled synthesis of polyoxopalladates, and their gas-phase fragmentation study by electrospray ionization tandem mass spectrometry. Eur. J. Inorg. Chem. 2013, 2013, 3442− 3442. (53) Brese, N. E.; O’Keeffe, M. Bond-valence Parameters for Solids. Acta Crystallogr., Sect. B: Struct. Sci. 1991, 47, 192−197. (54) 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, 244−247. (55) Nyquist, R. A.; Kagel, R. O. Infrared Spectra of Inorganic Compounds; Academic Press: New York and London, 1971. (56) Kubelka, P.; Munk, F. An Article on Optics of Paint Layers. Z. Technol. Phys. 1931, 12, 593−609. (57) Wendlandt, W. W.; Hecht, H. G. Reflectance Spectroscopy; Interscience: New York, 1966. (58) Okoye, C. M. I. Theoretical study of the electronic structure, chemical bonding and optical properties of KNbO3 in the paraelectric cubic phase. J. Phys.: Condens. Matter 2003, 15, 5945−5958.

I

DOI: 10.1021/acs.inorgchem.9b01730 Inorg. Chem. XXXX, XXX, XXX−XXX