Pb@Pb8 Basket-like-Cluster-Based Lead Tellurate–Nitrate Kleinman

Jun 26, 2017 - Two Phosphates: Noncentrosymmetric Cs6Mg6(PO3)18 and Centrosymmetric Cs2MgZn2(P2O7)2. Inorganic Chemistry. Chen, Xing, Liu, Guo, ...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/IC

Pb@Pb8 Basket-like-Cluster-Based Lead Tellurate−Nitrate KleinmanForbidden Nonlinear-Optical Crystal: Pb9Te2O13(OH)(NO3)3 Yi-Gang Chen,† Nan Yang,† Xing-Xing Jiang,‡ Yao Guo,*,§ and Xian-Ming Zhang*,† †

School of Chemistry & Material Science, Shanxi Normal University, Linfen, Shanxi 041004, China Beijing Center for Crystal Research and Development, Key Laboratory of Functional Crystals and Laser Technology, Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Beijing 100190, China § Department of Chemical and Environmental Engineering, Anyang Institute of Technology, Anyang 455000, China ‡

S Supporting Information *

ABSTRACT: The first lead tellurate−nitrate nonlinear-optical (NLO) crystal, Pb9Te2O13(OH)(NO3)3, featuring a 3D anionic diamondlike toplogical structural motif formulated as [Pb9O13]8− constructed by unique nonanuclear basket-shaped Pb@Pb8O14 clusters has been hydrothermally achieved, in which the Pb@ Pb8O14 cluster shows a core−shell structure with one PbO10 polyhedron enclosed by eight PbOx (x = 3 and 4) polyhedera via edge and face sharing. The hexavalent Te in TeO5(OH) octahedra is linked to a [Pb9O13]8− cluster to form a rare [Pb9Te2O13(OH)]3+ cationic framework. The compound demonstrates a powder second-harmonic-generation (SHG) response of about 1.2 times that of KH2PO4 (KDP) as well as a wide transparency range. Calculations on the local dipole moment and SHG coefficient reveal that the net polarization is zero because of a Kleinmanforbidden point group. Further analysis shows that the SHG response results from structure-induced variations by thermal vibrations of the lattice rather than the intrinsic dipole moment, which offers another insight on the design of new NLO materials.



INTRODUCTION Nonlinear-optical (NLO) crystals are one of the important optoelectronic information materials that have been widely applied to laser frequency conversion, optical parametric oscillators, and other optical devices.1 Up to now, numerous NLO materials have been discovered, and some have extensively been used including borates such as β-BaB2O42 and LiB3O5,3 phosphates such as KH2PO4 (KDP)4 and KTiOPO4,5 niobates such as LiNbO3,6 and chalcogenides such as AgGaS27 and AgGaSe2.8 The general strategies to enhance NLO crystals include utilization of the asymmetric structural units with local polarity (i.e., π-conjugated [BO3]3−, [CO3]2−, [NO3]−, etc.),9 second-order Jahn−Teller distorted stereoactive lone-pair cations (Tl+, Sn2+, Sb3+, Te4+, I5+, etc.)10 and MO6 octahedra (M = d0 transition metals, Ti4+, V5+, Mo6+, etc.).11 These asymmetric structural units are required to be aligned properly in noncentrosymmetric (NCS) or polar space groups so as to generate large macroscale polarization. Therefore, the design and packing of asymmetric structural units are crucial to the crystal engineering of NLO materials. Recently, it has been found that the combination of two or more types of asymmetric structural units is more likely to produce NCS structure and result in materials with secondharmonic-generation (SHG) properties.12 For example, the combination of two kinds of lone-pair cations has resulted in a series of NLO materials, such as Pb2TiOF(SeO3)2Cl,13 TeSeO4,14 Pb3SeO5,15 and Pb3Mg3TeP2O14;16 the combination of lone-pair cations and π-conjugated systems has also brought © XXXX American Chemical Society

about another class of oxide materials, such as Pb 16 (OH) 16 (NO 3 ) 16 , 17 Pb 2 Ba 3 (BO 3 ) 3 Cl, 18 Pb 2 B 5 O 9 I, 19 (Pb 4 O)Pb 2 B 6 O 14 , 20 RbPbCO 3 F, 21 CsPbCO 3 F, 22 Pb 7 O(OH)3(CO3)3(BO3),23 and Pb2(BO3)(NO3).24 The electron configuration of PbII contains the 6s2 electron pair, which can be either stereochemically active or inactive in PbII compounds. The large ionic radius of the Pb2+ cation as well as the lone-pair electron gives very variable coordination numbers from 2 to 10. Different PbOx (x = 2−10) units can aggregate to form interesting Pb cluster building units such as α-PbO-like “slabs”,17 a Pb7O(OH)3 cubane,23 [Pb4(OH)4]4+, [Pb6O(OH)4]6+, and [Pb6O4(OH)]3+ cuboids,25−27 a [Pb13O8(OH)6]4+ cluster,28 a [Pb3O2(OH)]+ sheet,29 a [Pb20O6(OH)16]12+ chain,30 and a [Pb2O(OH)]+ layer,31 and their effect on the material properties has not received additional consideration. Thus, Pb2+ and Te4+ cations as well as the [NO3]− unit with a π-conjugated system are introduced into the reaction system to explore new NLO materials. In this paper, we report the first example of a threechromophore lead tellurate−nitrate hydroxide NLO crystal Pb9Te2O13(OH)(NO3)3 synthesized through a hydrothermal reaction. The compound Pb9Te2O13(OH)(NO3)3 belonging to the NCS chiral space group P43212 contains an unusual 3D anionic diamondlike toplogical structural motif formulated as [Pb9O13]8− constructed by a unique nonanuclear basket-shaped Received: March 11, 2017

A

DOI: 10.1021/acs.inorgchem.7b00647 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Powder X-ray Diffraction (PXRD). The PXRD pattern of Pb9Te2O13(OH)(NO3)3 was collected on a Rigaku MiniFlex II X-ray diffractometer using Cu Kα radiation (λ = 1.540598 Å) in the 2θ angular range 4−70° at 293 K with a step size of 0.02°. The measured PXRD pattern is consonant with the calculated XRD pattern on the basis of the single crystal (Figure S2). Elemental Analysis. Microprobe elemental analysis on several crystals was fulfilled by field-emission scanning electron microscopy (JSM6700F) using energy-dispersive X-ray spectroscopy (EDS; Oxford INCA). The average atomic ratio of Pb/Te was 4.51:1, in agreement with the determination from single-crystal X-ray structure analyses (Table S3 and Figure S3). IR Spectroscopy. An IR spectrum of the sample was recorded on a Nicolet 5DX spectrometer in the range 400−2500 cm−1 at room temperature. The sample and KBr were pressed into a pellet after mixing and grinding. UV−Vis−Near-IR Diffuse-Reflectance Spectroscopy. A Varian Cary 5000 Scan spectrophotometer was used to measure the spectra of a powder over the 200−2500 nm wavelength range at room temperature. A powdered BaSO4 sample was used as a standard (100% reflectance). Reflectance spectra were converted to absorbance by the Kubelka−Munk function: F(R) = (1 − R)2/2R = K/S, where R is the reflectance, K is the absorption, and S is the scattering. SHG Measurements. SHG measurements on the basis of the Kurtz−Perry method at 298 K were performed using a Q-switched Nd:YAG laser at a wavelength of 1064 nm. Polycrystalline Pb9Te2O13(OH)(NO3)3 was ground and sieved into distinct particle size ranges: 25−40, 40−63, 63−80, 80−125, 125−150, 150−200, and 200−300 μm. Polycrystalline KDP as the reference was ground and sieved into the same particle size ranges to make relevant comparisons. The SHG signal intensity of the Pb9Te2O13(OH)(NO3)3/KDP ratio was identified based on the same size (125−150 μm). Thermogravimetric Analysis (TGA). TGA and differential scanning calorimetry (DSC) analyses were carried out on a Netzsch STA 449 F3 unit. The sample of 9.9 mg was enclosed in a platinum crucible and heated from 30 to 800 °C at a rate of 10 °C min−1 under a N2 atmosphere. Computational Method. First-principle calculations were performed using density functional theory (DFT) employing pseudopotentials and plane-wave basis sets, as implemented in the MedeA-VASP software package.32 In all calculations, a large unit cell (160 atoms in total) without any symmetry restrictions (space group P1) was developed under periodic boundary conditions. The electron−ion interaction was described using the projector-augmented-wave method. The energy cutoff for the plane-wave basis set was kept fixed at 400 eV. The reciprocal space was sampled by a 2 × 2 × 2 Monkhorst−Pack k-point mesh and integrated with a Gaussian smearing (σ = 0.2 eV). The generalized gradient approximation of Perdew−Burke−Ernzerhof33 was employed in geometrical optimization calculations to describe the exchange-correlation energies. To improve the accuracy of DFT calculations for the electronic properties, the hybrid functionals of the Becke three-parameter Lee−Yang−Parr (B3LYP)34 and Heyd−Scuseria−Ernzerhof (HSE06)35 were further employed in the electronic band structure and density of states (DOS) calculations. Electron localization function (ELF) analysis was performed to understand the bonding properties. Visualization of the atomic structures was obtained using the VESTA program.36

Pb@Pb8O14 cluster in which the Pb@Pb8O14 cluster, with a core−shell structure with one PbO10 polyhedron, is enclosed by eight PbOx (x = 3 and 4) polyhdera via edge and face sharing. The compound exhibits a non-phase-matchable powder SHG response of about 1.2 times that of KDP and a wide transparency range (0.34−2.5 μm). Interestingly, structureinduced variations by thermal vibrations of the lattice rather than the intrinsic dipole moment are supposed to cause nonzero net polarization and SHG. Herein, the synthesis, structure, optical properties, and thermal behavior of Pb9Te2O13(OH)(NO3)3 were investigated, and its SHG effect was studied. Furthermore, first-principle calculations and local dipole moment analysis were fulfilled to explore the relationship between the electronic structure and optical properties.



EXPERIMENTAL SECTION

Reagents. TeO2 (99.99%) was purchased from Aladdin Chemistry Co. Ltd., Pb(NO3)2 (99.0%) from Damao Chemical Factory, and NaOH (96.0%) from Xilong Chemical Factory. The reagents were used as received. Synthesis. Pb9Te2O13(OH)(NO3)3 was achieved by hydrothermal synthesis. Pb(NO3)2 (0.994 g, 4.0 mmol), NaOH (0.020 g, 4.0 mmol), and TeO2 (0.079 g, 0.5 mmol) were put in deionized water of 4.0 mL and sealed in a stainless steel bomb equipped with a Teflon liner (10 mL). The mixture was heated at 230 °C for 4 days. After the mixture had cooled to 30 °C at a rate of 4 °C h−1, the transparent block crystals of Pb9Te2O13(OH)(NO3)3 were collected and washed with ethanol (Figure S1). Single-Crystal X-ray Diffraction (XRD). Single-crystal XRD data were carried out on an Agilent Technologies Gemini EOS diffractometer with an EOS CCD detector at 293 K using Mo Kα radiation (λ = 0.71073 Å) and integrated with the CrysAlisPro program. The structures were solved by direct methods with the program SHELXS and refined by the full-matrix least-squares program SHELXL. All non-H atoms were refined with anisotropic thermal parameters. H atomic positions of the H+ groups were identified by fixed-lattice-constant optimization obtained from the X-ray crystal structure. The space structure was inspected for probable missing symmetry through the program PLATON. Details of crythe stallographic data and structural refinements are summarized in Table 1. The atomic positions, thermal parameters, and important bond distances and angles are listed in Tables S1 and S2.

Table 1. Crystal Data and Structure Refinement for Pb9Te2O13(OH)(NO3)3 empirical formula fw cryst syst space group a (Å) c (Å) α = β = γ (deg) V (Å3) Z Dcalc (g cm−3) μ(Mo Kα) (mm−1) cryst size (mm3) Flack parameter GOF on F2 R1, wR2 [I ≥ 2σ(I)]a R1, wR2 (all data) largest diff peak/hole (e Å−3) a

Pb9Te2O13(OH)(NO3)3 2530.25(13) tetragonal P43212 12.7268(4) 14.5103(5) 90.00 2350.25(17) 4 7.153 66.755 0.25 × 0.20 × 0.15 0.000(11) 1.105 0.0338, 0.0849 0.0345, 0.0853 2.94/−1.78



RESULTS AND DISCUSSION Crystal Structures. Pb9Te2O13(OH)(NO3)3 crystallizes in the NCS chiral space group P43212 (No. 96). The asymmetric unit is composed of 5 crystallogarphically independent Pb atoms, 1 Te atom, 2 N atoms, 12 O atoms, and 1 H atom (Figure 1a). The Pb2 and N2 atoms are located in a special position (Wyckoff letter 4a) with an occupancy of 0.5. In terms of Davidovicha,37 the primary coordination sphere of the PbII atom can reach a Pb−O distance of 2.70 Å, whereas the secondary coordination varies from 2.70 to 3.30 Å.

R1 = ∑||Fo| − |Fc||/∑|Fo|. bwR2 = [w(Fo2 − Fc2)2/w(Fo2)2]1/2. B

DOI: 10.1021/acs.inorgchem.7b00647 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

atoms (Figure 1b). The Pb2−O bond lengths are in the range of 2.592(9)−2.789(2)Å. Pb1, Pb3, and Pb4 have similar coordination geometries and are ligated by four O atoms in the primary coordination sphere to give a (4 + E)ψ-trianglular biprism. The Pb−O bond lengths for Pb1, Pb3, and Pb4 in the primary coordination sphere are 2.198(10)−2.64(3) Å. In addition, there is a Pb3−O8 bond of 2.716(14) Å, slightly longer than the limit of 2.70 Å proposed by Davidovicha. Pb5 is coordinated by three O atoms in the primary coordination sphere to give a (3 + E)ψ tetrahedron. The Pb5−O bond lengths are in the range of 2.332(9)− 2.395(9) Å. All of the PbOn polyhedra are seriously distorted as a result of the effect of lone-pair electrons, and the distances for the Pb−O bonds are in the range 2.197(10)−2.789(2) Å, which are close to those reported in the literature.25−31 Each Te6+ is coordinated to six O atoms with Te−O bond lengths of 1.899(9)− 1.937(10) Å,38 which shows that the TeO6 octahedra are a little distorted along a local C3 direction (Figure S4). Calculation of the bond valence sum (BVS)39 gives the total bond valences of Pb9Te2O13(OH)(NO3)3 for the Pb, Te, and N atoms, and they are 1.733−2.098, 5.922, and 4.850−5.382, respectively. The O6 atom with a calculated bond valence of 0.963 is attributed to the hydroxyl group. The N1 atom is localized on a 2-fold axis, and the N1O3− nitrate is dually disordered (Figure S5a). Another nitrate, N2O3− (Figure S5b), is coordinated to Pb atoms in μ2 mode with Pb−O distances of 2.716(14) and 2.645(3) Å. It should be noted that there are relatively strong bonds between the Pb atoms and two O atoms of the planar N2O3− nitrate [2.645(3) and 2.716(14) Å], while the linkages between the Pb atoms and two O atoms of the disordered N1O3− ion are weak bonds [2.887(11) and 3.134(12) Å]. The compound possesses a 21 helical axis along the a-axis direction and a 43 helical axis along the c-axis direction. IR Spectroscopy. According to the characteristic peaks of nitrates, the IR absorption bands at 804, 1134, and 1385 cm−1 can be ascribed to N−O bending and stretching vibrations, while the peaks at 450 and 587 cm−1 can be attributed to the Te−O vibration (Figure S6), which are in conformity with other reported metal nitrates and tellurates.30,40 The IR spectrum further verifies the existence of NO and TeO groups. Thermal Behavior. TGA and DSC studies demonstrate that Pb9Te2O13(OH)(NO3)3 is thermally stable up to about 400 °C (Figure S7). The process of weight loss corresponds to the loss of nitrate groups. The corresponding weight loss of about 7.8% is close to the calculated loss of 8.1%. UV−Vis−IR Diffuse-Reflectance Spectroscopy. UV absorption spectra of Pb9Te2O13(OH)(NO3)3 show little absorption in the range of 340−2500 nm (0.34−2.5 μm) at room temperature. Optical diffuse-reflectance spectroscopy reveals that Pb9Te2O13(OH)(NO3)3 is a wide-band-gap semiconductor with an optical band gap of approximately 3.62 eV (Figure 3). Theoretical Calculation. To better interpret the crystal structure and optical property relationship, the electronic band structure and DOS based on DFT methods were calculated. The band structure calculations show that Pb9Te2O13(OH)(NO3)3 is an indirect-band-gap compound with a calculated band-gap value of 3.20 eV (Figure 4a), a little smaller than that of the experimental result (3.62 eV) because of the limitation of the DFT method. The total and partial DOSs of Pb9Te2O13(OH)(NO3)3 are displayed in Figure 4b. Because the localized inner electrons with lower energy states basically do not contribute to electron transfer, the upper regions of the

Figure 1. (a) Coordination environment of the Pb and Te atoms. (b) Pb2O10 polyhedron. (c) Basket-shaped core−shell Pb@Pb8O14 cluster. (d) Basket-shaped Pb@Pb8 cluster showing the core Pb2 atom with space-filling mode. (e) Each Pb@Pb8O14 cluster linked to four of its neighbors.

The compound features an unusual 3D anionic diamond topological framework, Pb9O138−, constructed by a unique basket-like Pb9O14 cluster, in which the Pb9O14 cluster shows a Pb@Pb8O14 core−shell structure with one PbO10 polyhedron enclosed by eight PbOx (x = 3 and 4) polyhedera via edge and face sharing (Figure 1b−e). The eight PbOx polyhedra (two each of Pb1, Pb3, Pb4, and Pb5) are linked via corner, edge, and face sharing into an empty Pb8O16 basket, which is filled by the Pb2 atom to form an unprecedented Pb@Pb8O16 secondary building unit with C2 symmetry along the (110) direction (Figure 1c,d). Because of repulsion of the lone-pair electrons, the lone pair of Pb atoms points outward. Each Pb@Pb8O16 cluster is connected to four of its neighbors via the sharing O4 atoms to generate an anionic diamondlike network (Figures 1e and 2a). The distance between adjacent Pb@Pb8O16 clusters is 7.45 Å. Such an anionic framework has voids and is connected with TeO6 to form a [Pb9Te2O13(OH)]3+ ∞ cationic framework that still has space for accommodating nitrates (Figure 2a,b). Specifically, the Pb2 atom shows a distorted bicapped antitetragonal prism with C2 symmetry, coordinated by 10 O

Figure 2. (a) Simplified 3D diamondlike structure with the Pb@ Pb8O14 cluster, TeO6 octahedra, and nitrates shown as large purple, middle pale-blue, and small blue balls, respectively. (b) 3D structure of Pb9Te2O13(OH)(NO3)3 along the c axis. C

DOI: 10.1021/acs.inorgchem.7b00647 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

5a). On the basis of the anionic group theory,41 geometrical superposition of the second-order susceptibility results in the macroscale SHG response of a crystal. Therefore, in Pb9Te2O13(OH)(NO3)3, the packing of the TeO6 and NO3 structural groups as well as the PbOn polyhedra with stereochemically active lone pairs (SCALPs) can affect the macroscopic SHG coeffcients. To further probe the SHG origin of the compound, the bond-valence method17,42 was employed to calculate the direction and magnitude of the dipole moments, which had confirmed the validity of this method in the calculation of NO3, MoO6, TeOn, PbOn, and BaOn polar polyhedra.17,42 For the PbOn polyhedra, the lone pair is considered to have a charge of 2− and localized 0.86 Å from the Pb2+ cation.43 Furthermore, to further visualize the SCALPs of Pb2+ cations, the ELF map, including the unit cell (Figure S8), was examined and clearly shows highly asymmetric lobes on the Pb2+ cations as a result of SCALPs. Table 2

Figure 3. UV−vis−IR spectra of Pb9Te2O13(OH)(NO3)3.

Table 2. Dipole Moments of PbOn Polyhedra, TeO6 Octahedra, and NO3 Triangles and Total Polarization of the Asymmetric Unit dipole moment (D) x

y

z

total magnitude

∑k = 1 (PbOn)k

−0.0976 −6.2190 −4.7989 −6.5572 −0.3736 −0.4511 −18.3998

0.1174 −3.5097 −4.7989 4.13054 −8.2640 −4.7737 −17.2158

−1.0027 2.3018 0 −4.0782 −2.3884 4.7689 0.6042

1.0143 7.5028 6.7867 8.7573 8.6103 6.7627 25.2052

N1O3 N2O3 asymmetric unit unit cell

6.7102 −0.3257 −12.1130 0

6.7102 −0.6888 −11.0771 0

0 0.7276 0.3291 0

9.4896 1.0536 16.4175 0

polar unit TeO6 Pb1O4 Pb2O10 Pb3O5 Pb4O4 Pb5O4

Figure 4. Band structure (a) and DOS (b) of Pb9Te2O13(OH)(NO3)3.

valence band (VB) and conduction band (CB) (about −7 to 8 eV) are mostly focused. It is evident that the VB consists mainly of O 2p orbital mixing with a small amount of Te 5p and Pb 6s and 6p orbitals, whereas the CB is composed of mainly Pb 6p, Te 5s and 5p, and N 2p orbitals. In the vicinity of the Fermi level, O 2p orbitals overlapping a few Pb 6s and 6p orbitals occupy the top region of the VB, when Pb 6p, N 2p, and Te 5s and 5p orbitals occupy the bottom region of the CB. Consequently, the bonding interactions between O and Pb, Te, and N determine the band gap of Pb9Te2O13(OH)(NO3)3 and therefore are dedicated predominantly to the optical properties. NLO Response. As Pb9Te2O13(OH)(NO3)3 crystallizes in a NCS space group, its detailed NLO properties were studied. Powder SHG measurements at 1064 nm reveal that Pb9Te2O13(OH)(NO3)3 is non-phase-matchable (Figure 5) and the SHG intensity is about 1.2 times that of the KDP standard with the same grain size (shown in the inset of Figure

5

demonstrates the direction and magnitude of the dipole moments for the TeO6, PbOn in the range 2.197(10)− 2.789(2) Å, and NO3 structural groups in Pb9Te2O13(OH)(NO3)3. As shown in Table 2, the magnitude of the dipole moment of the TeO6 octahedron is only 1.0143 D, which is in agreement with a slightly distorted octahedral coordination. As a result of the SCALPs of Pb2+ cations, the magnitude of the dipole moment for PbOn varies between 6.7627 and 8.7573 D, and the total dipole moment reaches 25.2052 D, which is far larger than that of the NO3 groups (total: 8.8061 D). Also, dipole moment analysis provides concrete evidence of threechromophore participation. However, because the equivalent microscopic groups are arranged in the opposite direction to that shown in Figure 6 and their dipole moments totally balance, the net polarization in the unit cell is zero. At first glance, the nonzero SHG response of Pb9 Te2 O13(OH)(NO3 ) 3 is very attractive. Because the compound is located at space group P43212, belonging to class 422, the point-group symmetry claims two nonvanishing tensors of second-order susceptibility to satisfy the equation d14 = −d25. Nonetheless, on the basis of the restriction of Kleinman symmetry,44 d14 must equal d25. As a result, the two tensors have to be zero. Thus, because of Kleinman symmetry, the SHG response for any compound in the point group 422 is forbidden. In order to further confirm this, the length-gauge formalism by Aversa and Sipe45 was used to calculate the SHG

Figure 5. SHG intensity curve for Pb9Te2O13(OH)(NO3)3 (a) and KDP (b). Oscilloscope traces of the SHG signals for Pb9Te2O13(OH)(NO3)3 and KDP samples (125−150 μm) in the inset in part a. The curve is to guide the eye and is not a fit to the data. D

DOI: 10.1021/acs.inorgchem.7b00647 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

calculated XRD pattern data, coordination of tellurium and nitrogen, IR spectra, DSC curves, and ELF for Pb9Te2O13(OH)(NO3)3 (PDF) Accession Codes

CCDC 1516143 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 data_ [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



Figure 6. Direction of the dipole moments of groups.

*E-mail: [email protected]. Fax: +86 372 2909732. *E-mail: [email protected]. Fax: +86 357 2051402.

coefficient, and a null static SHG coefficient was then achieved. However, according to ab initio molecular dynamics simulations using the Nosé algorithm (Ln4InSbS9, 422 point group),46 the strong SHG effect of Pb9Te2O13(OH)(NO3)3 may originate from structure-induced variations by thermal vibrations of the lattice because the SHG wavelength (532 nm or 2.33 eV) deviates appreciably from its optical band gap (3.62 eV). It is worth noting that, although similar phenomena (zero net polarization) emerge in MgTeMoO6 (P21212), CdTeMoO6 (P4̅21m), and ZnTeMoO6 (P21212),47 the difference is that the three compounds exhibit nonzero tensors under the restriction of Kleinman symmetry, and induced dipole oscillations by an external optical electric field bring about nonzero net polarization and SHG response.48 So, it may be deduced that the net polarization in Pb9Te2O13(OH)(NO3)3 results mainly from the three types of NLO-active units (TeO6, NO3, and PbOn) by thermal vibrations of the lattice.

ORCID

Yao Guo: 0000-0002-3964-8002 Xian-Ming Zhang: 0000-0002-8809-3402 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was endowed financially by the Plan for 10000 Talents in China, National Science Fund for Distinguished Young Scholars (Grant 20925101), and Shanxi Province Foundation for Key Subject.



REFERENCES

(1) (a) Burland, D. M.; Miller, R. D.; Walsh, C. A. Second-order nonlinearity in poled-polymer systems. Chem. Rev. 1994, 94, 31−75. (b) Rao, C. N. R.; Behera, J. N.; Dan, M. Organically-templated metal sulfates, selenites and selenates. Chem. Soc. Rev. 2006, 35, 375. (c) Ruck, M.; Locherer, F. Reprint of “Coordination chemistry of homoatomic ligands of bismuth, selenium and tellurium. Coord. Chem. Rev. 2015, 297-298, 208−217. (d) Rondinelli, J. M.; Kioupakis, E. Predicting and Designing Optical Properties of Inorganic Materials. Annu. Rev. Mater. Res. 2015, 45, 491−518. (2) Lin, J.; Lee, M.-H.; Liu, Z.-P.; Chen, C.; Pickard, C. J. Mechanism for linear and nonlinear optical effects in β−BaB2O4 crystals. Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 60, 13380−13389. (3) Chen, C.; Wu, Y.; Jiang, A.; Wu, B.; You, G.; Li, R.; Lin, S. New nonlinear-optical crystal: LiB3O5. J. Opt. Soc. Am. B 1989, 6, 616−621. (4) De Yoreo, J. J.; Burnham, A. K.; Whitman, P. K. Developing KH2PO4 and KD2PO4 crystals for the world’s most power laser. Int. Mater. Rev. 2002, 47, 113−152. (5) Driscoll, T. A.; Hoffman, H. J.; Stone, R. E.; Perkins, P. E. Efficient second-harmonic generation in KTP crystals. J. Opt. Soc. Am. B 1986, 3, 683−686. (6) Boyd, G. D.; Miller, R. C.; Nassau, K.; Bond, W. L.; Savage, A. Electro-Optic Properties of LiNbO3. Appl. Phys. Lett. 1964, 5, 234− 236. (7) Okorogu, A. O.; Mirov, S. B.; Lee, W.; Crouthamel, D. I.; Jenkins, N.; Dergachev, A. Y.; Vodopyanov, K. L.; Badikov, V. Tunable middle infrared downconversion in GaSe and AgGaS2. Opt. Commun. 1998, 155, 307−312. (8) Boyd, G.; Kasper, H.; McFee, J.; Storz, F. Linear and nonlinear optical properties of some ternary selenides. IEEE J. Quantum Electron. 1972, 8, 900−908. (9) (a) Zhao, S.; Gong, P.; Bai, L.; Xu, X.; Zhang, S.; Sun, Z.; Lin, Z.; Hong, M.; Chen, C.; Luo, J. Beryllium-free Li4Sr(BO3)2 for deepultraviolet nonlinear optical applications. Nat. Commun. 2014, 5, 1−7. (b) Narsimha Rao, E.; Appalakondaiah, S.; Yedukondalu, N.; Vaitheeswaran, G. Structural, electronic and optical properties of novel carbonate fluorides ABCO3F (A = K, Rb, Cs; B = Ca, Sr). J. Solid State Chem. 2014, 212, 171−179. (c) Chen, Y. G.; Xing, M. L.; Liu, P.



CONCLUSIONS We identified the first nonpolar lead tellurate−nitrate NLO crystal, Pb9Te2O13(OH)(NO3)3, derived from a convenient hydrothermal method. The compound features a novel 3D anionic diamond topological framework, Pb9O138−, constructed by a unique basket-like Pb9O14 cluster that shows a Pb@Pb8O14 core−shell structure. The anionic cluster is connected with TeO6 to form a [Pb9Te2O13(OH)]3+ ∞ cationic framework that accommodates nitrates. Pb9Te2O13(OH)(NO3)3 is thermally stable up to 400 °C. Powder NLO measurement indicates that this compound has a wide transparency range and is non-phasematchable (type I) with a SHG response of about 1.2 times that of KDP at 1064 nm. Because NLO-active units (TeO6, NO3, and PbOn) are arranged in the opposite direction, the total polarization in the unit cell cancels out completely. Further calculation on the SHG coefficients reveals that the static SHG coefficient was null because of Kleinman-forbidden symmetry. It is believed that structure-induced variations by thermal vibrations of the lattice result in nonzero net polarization and SHG, which will give another insight on the further exploration of new NLO materials.



AUTHOR INFORMATION

Corresponding Authors

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b00647. Atomic coordinates and equivalent isotropic displacement parameters, selected bond lengths and angles, photographs of crystals, EDS plots, experimental and E

DOI: 10.1021/acs.inorgchem.7b00647 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry F.; Guo, Y.; Yang, N.; Zhang, X. M. Two Phosphates: Noncentrosymmetric Cs 6 Mg 6 (PO 3 ) 1 8 and Centrosymmetric Cs2MgZn2(P2O7). Inorg. Chem. 2017, 56, 845−851. (10) (a) Phanon, D.; Gautier-Luneau, I. Promising Material for Infrared Nonlinear Optics: NaI3O8 Salt Containing an Octaoxotriiodate (V) Anion Formed from Condensation of [IO3]− Ions. Angew. Chem., Int. Ed. 2007, 46, 8488−8491. (b) Almond, P. M.; McKee, M. L.; Albrecht-Schmitt, T. E. Unusual Uranyl Tellurites Containing [Te2O6]4− Ions and Three-Dimensional Networks. Angew. Chem., Int. Ed. 2002, 41, 3426−3429. (c) 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. (11) (a) Jin, C. G.; Li, Z.; Huang, L. X.; He, M. Z. Top-seeded solution growth and characterization of a novel nonlinear opticalcrystal MnTeMoO6. J. Cryst. Growth 2013, 369, 43−46. (b) Hubbard, D. J.; Johnston, A. R.; Casalongue, H. S.; Sarjeant, A. N.; Norquist, A. J. Synthetic Approaches for Noncentrosymmetric Molybdates. Inorg. Chem. 2008, 47, 8518−8525. (c) Liu, Z.; Hu, Y.-Y.; Dunstan, M. T.; Huo, H.; Hao, X.; Zou, H.; Zhong, G.; Yang, Y.; Grey, C. P. Local Structure and Dynamics in the Na Ion Battery Positive Electrode Material Na3V2(PO4)2F3. Chem. Mater. 2014, 26, 2513−2521. (12) (a) 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. (b) Zhao, S.; Yang, Y.; Shen, Y.; Zhao, B.; Li, L.; Ji, C.; Wu, Z.; Yuan, D.; Lin, Z.; Hong, M.; Luo, J. Cooperation of Three Chromophores Generates the Water-Resistant Nitrate Nonlinear Optical Material Bi3TeO6OH(NO3)2. Angew. Chem., Int. Ed. 2017, 56, 540−544. (c) Daub, M.; Krummer, M.; Hoffmann, A.; Bayarjargal, L.; Hillebrecht, H. Synthesis, Crystal Structure, and Properties of Bi3TeBO9 or Bi3(TeO6)(BO3): A Non-Centrosymmetric Borate− Tellurate (VI) of Bismuth. Chem.Eur. J. 2016, 22, 1−8. (13) Cao, X. L.; Hu, C. L.; Xu, X.; Kong, F.; Mao, J. G. Pb2TiOF(SeO3)2Cl and Pb2NbO2(SeO3)2Cl: small changes in structure induced a very large SHG enhancement. Chem. Commun. 2013, 49, 9965−9967. (14) Porter, Y.; Bhuvanesh, N. S. P.; Halasyamani, P. S. Synthesis and Characterization of Te2SeO7: A Powder Second-Harmonic-Generating Study of TeO2, Te2SeO7, Te2O5, and TeSeO4. Inorg. Chem. 2001, 40, 1172−1175. (15) Kim, S.-H.; Yeon, J.; Halasyamani, P. S. Noncentrosymmetric Polar Oxide Material, Pb3SeO5: Synthesis, Characterization, Electronic Structure Calculations, and Structure-Property Relationships. Chem. Mater. 2009, 21, 5335−5342. (16) Yu, H.; Young, J.; Wu, H.; Zhang, W.; Rondinelli, J. M.; Halasyamani, P. S. Electronic, Crystal Chemistry, and Nonlinear Optical Property Relationships in the Dugganite A3B3CD2O14 Family (A = Sr, Ba or Pb; B = Mg or Zn; C = Te or W, and D = P or V). J. Am. Chem. Soc. 2016, 138, 4984−4989. (17) Chang, L.; Wang, L.; Su, X.; Pan, S.; Hailili, R.; Yu, H.; Yang, Z. A Nitrate Nonlinear Optical Crystal Pb16(OH)16(NO3)16 with a Large Second-Harmonic Generation Response. Inorg. Chem. 2014, 53, 3320−3325. (18) Dong, X.; Jing, Q.; Shi, Y.; Yang, Z.; Pan, S.; Poeppelmeier, K. R.; Young, J.; Rondinelli, J. M. Pb2Ba3(BO3)3Cl: A Material with Large SHG Enhancement Activated by Pb-Chelated BO3 Groups. J. Am. Chem. Soc. 2015, 137, 9417−9422. (19) 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. (20) Zhang, F.; Zhang, F.; Lei, B.-H.; Yang, Z.; Pan, S. Synthesis, Characterization, and Theoretical Studies of (Pb4O)Pb2B6O14: A New Lead(II) Borate with Isolated Oxygen-Centered Pb4O Tetrahedra and Large Second Harmonic Generation Response. J. Phys. Chem. C 2016, 120, 12757−12764. (21) Tran, T. T.; Halasyamani, P. S.; Rondinelli, J. M. Role of Acentric Displacements on the Crystal Structure and Second-

Harmonic Generating Properties of RbPbCO3F and CsPbCO3F. Inorg. Chem. 2014, 53, 6241−6251. (22) Zou, G. H.; Huang, L.; Ye, N.; Lin, C. S.; Cheng, W. D.; Huang, H. CsPbCO3F: A Strong Second-Harmonic Generation Material Derived from Enhancement via p−π Interaction. J. Am. Chem. Soc. 2013, 135, 18560−18566. (23) Abudoureheman, M.; Wang, L.; Zhang, X.; Yu, H.; Yang, Z.; Lei, C.; Han, J.; Pan, S. Pb7O(OH)3(CO3)3(BO3): First Mixed Borate and Carbonate Nonlinear Optical Material Exhibiting Large SecondHarmonic Generation Response. Inorg. Chem. 2015, 54, 4138−4142. (24) 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 pConjugated Planar Triangular Units. Angew. Chem., Int. Ed. 2015, 54, 3679−3682. (25) Grimes, S. M.; Johnston, S. R.; Abrahams, I. Characterisation of the predominant low-pH lead (II)-hydroxo cation, [Pb4(OH)4]4+; crystal structure of [Pb4(OH)4][NO3]4 and the implications of basic salt formation on the transport of lead in the aqueous environment. J. Chem. Soc., Dalton Trans. 1995, 12, 2081−2086. (26) Spiro, T. G.; Templeton, D. H.; Zalkin, A. The Crystal Structure of a Hexanuclear Basic Lead (II) Perchlorate Hydrate: Pb6O(OH)6(ClO4)4·H2O. Inorg. Chem. 1969, 8, 856−861. (27) Li, Y.; Krivovichev, S. V.; Burns, P. C. Crystal Chemistry of Lead Oxide Hydroxide Nitrates I. The Crystal Structure of [Pb6O4](OH)(NO3)(CO3). J. Solid State Chem. 2000, 153, 365−370. (28) Li, Y.; Krivovichev, S. V.; Burns, P. C. Crystal Chemistry of Lead Oxide Hydroxide Nitrates: II. The Crystal Structure of Pb13O8(OH)6(NO3)4. J. Solid State Chem. 2001, 158, 74−77. (29) Krivovichev, S. V.; Li, Y.; Burns, P. C. Crystal Chemistry of Lead Oxide Hydroxide Nitrates III. The Crystal Structure of Pb3O2(OH)(NO3). J. Solid State Chem. 2001, 158, 78−81. (30) Wang, G.; Luo, M.; Ye, N.; Lin, C.; Cheng, W. Series of Lead Oxide Hydroxide Nitrates Obtained by Adjusting the pH Values of the Reaction Systems. Inorg. Chem. 2014, 53, 5222−5228. (31) Welch, M. D.; Criddle, A. J.; Symes, R. F. Mereheadite Pb2O(OH)Cl; a new litharge-related oxychloride from Merehead Quarry, Cranmore, Somerset. Mineral. Mag. 1998, 62, 387−393. (32) Guo, Y.; Li, C.; Tian, D.; Niu, Y.; Hou, S. Structural and electronic characterization of GaN on MgAl2O4 (111) substrates. Phys. Status Solidi B 2016, 253, 1715−1720. (33) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865−3868. (34) Becke, A. D. Density-functional thermochemistry. J. Chem. Phys. 1993, 98, 5648−5652. (35) Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207− 2815. (36) Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272−1276. (37) Davidovich, R. L.; Stavila, V.; Whitmire, K. H. Stereochemistry of lead (II) complexes containing sulfur and selenium donor atom ligands. Coord. Chem. Rev. 2010, 254, 2193−2226. (38) Bae, S.-w.; Yoo, J.; Lee, S.; Choi, K. Y.; Ok, K. M. Synthesis and Characterization of Three New Layered Vanadium Tellurites, MVTe2O8 (M = Al, Ga, and Mn): Three-Dimensional (3-D) Antiferromagnetic Behavior of MnVTe2O8 with a Zigzag S = 2 Spin Chain. Inorg. Chem. 2016, 55, 1347−1353. (39) (a) Yan, D.; Hu, C. L.; Mao, J. G. An enhanced extended hook method to realize tetranuclear metal clusters embedded in energetic metal−organic framework channels. CrystEngComm 2016, 18, 1655− 1664. (b) Christensen, N.; Chevallier, M.-A. J.; Skibsted, B. B.; Iversen, B. B. Synthesis and characterization of basic bismuth (III) nitrates. J. Chem. Soc., Dalton Trans. 2000, 3, 265−270. (40) Ziegler, P.; Grigoraviciute, I.; Gibson, K.; Glaser, J.; Kareiva, A.; Meyer, H. J. On the characterization of BiMO2NO3 (M = Pb, Ca, Sr, Ba) materials related with the Sillén X1 structure. J. Solid State Chem. 2004, 177, 3610−3615. F

DOI: 10.1021/acs.inorgchem.7b00647 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (41) (a) Chen, C. T.; Wu, Y. C.; Li, R. K. The anionic group theory of the non-linear optical effect and its applications in the development of new high-quality NLO crystals in the borate series. Int. Rev. Phys. Chem. 1989, 8, 65−91. (42) (a) Goodey, J.; Broussard, J.; Halasyamani, P. S. Synthesis, structure, and characterization of a new second-harmonic-generating tellurite: Na2TeW2O9. Chem. Mater. 2002, 14, 3174−3180. (b) Maggard, P. A.; Nault, T. S.; Stern, C. L.; Poeppelmeier, K. R. Alignment of acentric MoO3F33− anions in a polar material: (Ag3MoO3F3)(Ag3MoO4)Cl. J. Solid State Chem. 2003, 175, 27−33. (c) Brown, I. D. 14-The bond-valence method: An empirical approach to chemical structure and bonding. Ind. Chem. Libr. 1981, 2, 1−30. (d) Yu, H. W.; Wu, H. P.; Pan, S. L.; Yang, Z. H.; Su, X.; Zhang, F. F. Novel Deep UV Nonlinear Optical Crystal Ba3B6O11F2, with a New Fundamental Building Block, B6O14 Group. J. Mater. Chem. 2012, 22, 9665−9670. (43) Galy, J.; Meunier, G.; Andersson, S.; Astrom, A. Stéréochimie des eléments comportant des paires non liées: Ge (II), As (III), Se (IV), Br (V), Sn (II), Sb (III), Te (IV), I (V), Xe (VI), Tl (I), Pb (II), et Bi (III) (oxydes, fluorures et oxyfluorures). J. Solid State Chem. 1975, 13, 142−159. (44) Kleinman, D. A. Nonlinear Dielectric Polarization in Optical Media. Phys. Rev. 1962, 126, 1977−1979. (45) (a) Aversa, C.; Sipe, J. E. Nonlinear optical susceptibilities of semiconductors: Results with a length-gauge analysis. Phys. Rev. B: Condens. Matter Mater. Phys. 1995, 52, 14636−14645. (b) Rashkeev, S. N.; Lambrecht, W. R. L.; Segall, B. Efficient ab initio method for the calculation of frequency-dependent second-order optical response in semiconductors. Phys. Rev. B: Condens. Matter Mater. Phys. 1998, 57, 3905−3919. (46) Zhao, H. − J.; Zhang, Y. − F.; Chen, L. Strong KleinmanForbidden Second Harmonic Generation in Chiral Sulfide: La4InSbS9. J. Am. Chem. Soc. 2012, 134, 1993−1995. (47) (a) Zhao, S. G.; Jiang, X. X.; He, R.; Zhang, S. Q.; Sun, Z. H.; Luo, J. H.; Lin, Z. S.; Hong, M. C. A combination of multiple chromophores enhances second-harmonic generation in a nonpolar noncentrosymmetric oxide: CdTeMoO6. J. Mater. Chem. C 2013, 1, 2906−2912. (b) Zhang, J.; Zhang, Z.; Sun, Y.; Zhang, C.; Zhang, S.; Liu, Y.; Tao, X. MgTeMoO6: A neutral layered material showing strong second-harmonic generation. J. Mater. Chem. 2012, 22, 9921− 9927. (c) Zhao, S.; Luo, J.; Zhou, P.; Zhang, S.-q.; Sun, Z.; Hong, M. ZnTeMoO6: a strong second-harmonic generation material originating from three types of asymmetric building units. RSC Adv. 2013, 3, 14000−14006. (48) Jiang, X.; Zhao, S.; Lin, Z.; Luo, J.; Bristowe, P. D.; Guan, X.; Chen, C. The role of dipole moment in determining the nonlinear optical behavior of materials: ab initio studies on quaternary molybdenum tellurite crystals. J. Mater. Chem. C 2014, 2, 530−537.

G

DOI: 10.1021/acs.inorgchem.7b00647 Inorg. Chem. XXXX, XXX, XXX−XXX