DOI:
10.1039/D4QI01203H
(Research Article)
Inorg. Chem. Front., 2024,
11, 5648-5656
[Sn3OF]PO4vs. [Sn3F3]PO4: enhancing birefringence by breaking the R3 symmetry and realigning lone pairs†
Received
14th May 2024
, Accepted 25th June 2024
First published on 9th July 2024
Abstract
Tin(II)-containing phosphates are promising candidates for designing and studying materials with enhanced birefringence. Herein, we present a novel compound [Sn3OF]PO4 that exhibits significant birefringence (Δnobv. = 0.117 at 546 nm), a five-fold improvement compared to that of the related compound [Sn3F3]PO4 discovered nearly 40 years ago. The remarkable enhancement in birefringence is primarily attributed to the anisotropic [SnO3F]5− structure building unit. By simple regulation of the F/O ratio, the undesirable R3 symmetry in [Sn3F3]PO4 is disrupted. This regulation also realigns the lone pairs, resulting in substantial optical anisotropy.
10th anniversary statement
Congratulations on the 10th anniversary of ICF! These 10 years of ICF coincide with the second decade of my personal research career. As a founding member of the ICF editorial board, I have been actively involved in the forums organized by ICF.
Over the past decade, my collaborator Li-Ming Wu and I at BNU have made significant progress in the field of inorganic solid-state chemistry. We have explored the design and synthesis of novel compounds, studied the relationship between properties and crystal structures (focusing primarily on nonlinear optical and thermoelectric properties), and found ways to modify and enhance these properties.
Thank you for being with me throughout this journey!
|
Introduction
Birefringence (Δn) is a phenomenon that manifests in anisotropic media, particularly in non-cubic crystals.1–4 It reflects the optical anisotropy of a crystal, which is one of the essential properties of optoelectronic functional crystals. A crystal with an appropriate birefringence is a crucial component of all-solid-state lasers, as it can modulate the polarized light for a single wavelength laser output. In fact, the birefringence phenomenon originates from the structural anisotropy of crystals. Therefore, effective and significant approaches are employed for enhancing birefringence to increase the anisotropy of the structure building units and optimize the arrangement in the crystal structure.
At present, there are several main methods to increase birefringence.5–8 Firstly, introducing anisotropic building units (ABUs) with large birefringence is an effective approach for designing birefringent crystals, such as [BO3]3−, [NO3]−, [CO3]2−, [C4O4]2−, and [C3N3O3]3−.5,6 Secondly, to broaden the range of available ABUs, a novel approach has recently emerged: designing ABUs by encompassing diverse chemical bonds (ABUCBs), which has garnered interest among researchers.9 The incorporation of ABUCBs into the crystal structure is likely to induce anisotropy in both the crystal structure and the spatial electron distribution. The latest experimental advancements indicate, for instance, that the introduction of P–F or B–F bonds enhances microscopic polarization anisotropy and hyperpolarization, eventually resulting in improved macroscopic properties, such as Δn.10–18 Thirdly, the utilization of stereochemically active lone pairs could effectively increase the structural anisotropy, leading to a substantial Δn enhancement.19–23 For instance, cations such as Pb2+, Sn2+, Sb3+, and Bi3+, serving as coordination centres, constitute the ABUs. For example, BaSn2(PO4)2 exhibits a high Δncal. of 0.071,19 while Sn2B5O9Cl demonstrates a Δnobv. of 0.168.20
Tin-based phosphates are excellent candidates for studying crystals with large birefringence. For example, Sn2PO4I24 exhibits the largest Δnobv. of ≥0.664 at 546 nm among the reported borates and phosphates to date. However, the presence of Sn2+ with stereochemically active lone pairs does not always lead to an increase in Δn. In particular, in high-symmetry crystal structures, Δn values are typically very small, as observed in compounds like KSn4(PO4)3
25 (R3c, 0.013), NaSn4(PO4)3
26 (R3c, 0.005) and [Sn3F3]PO4
27 (R3, 0.011).9,28–30 In these compounds, the Sn2+ ions are located along the C3 or 31 axes, forming a highly symmetric ABU arrangement that does not exhibit large anisotropy. We consider that in Sn2+-containing phosphates, the orientation of lone pairs plays a crucial role in determining the birefringence. However, controlling this orientation remains a challenge.
Herein, we report a novel compound, [Sn3OF]PO4, obtained by a conventional hydrothermal synthesis method. Interestingly, this compound, together with [Sn3F3]PO4
27 known for about 40 years, provides a compelling example demonstrating that a straightforward regulation of the F/O ratio leads to a remarkable five-time enhancement of Δnobv. (0.117 vs. 0.027 at 546 nm). Furthermore, we delve deeper into how this simple regulation effectively disrupts the R3 high symmetry of the latter compound, causing all lone pairs to align consistently along the dielectric axis. As a result, a significant enhancement in birefringence is achieved. We believe that this strategy may offer new experimental guidelines for searching and designing effective birefringent crystals.
Experimental section
Synthesis of compounds
Reagents including NH4HF2 (Aladdin, 98%), CO(NH)2 (Aladdin, 98%), SnF2 (Bidepharm, 99%), H3PO4 (Aladdin, 99%), P2O5 (Aladdin, 99%) and B2O3 (Aladdin, 99.9%) were used as purchased.
Colourless block-like crystals of [Sn3OF]PO4 are obtained by hydrothermal synthesis. The raw materials, (NH4)2PO3F, SnF2 and B2O3, were well ground in a ratio of 1
:
2
:
4 and then put into PTFE pouches. The pouches were then enclosed in 1 mL autoclaves with PTFE inner linings, with addition of a little deionized water, and heated at 190 °C for 4 days. The pouches were cooled to room temperature at 0.5 °C min−1. After cleaning with deionized water and alcohol, colorless block-like crystals were obtained. The yield of the [Sn3OF]PO4 product could not be calculated, because it is obtained by hydrothermal synthesis along with impurities, which cannot be separated.
The polycrystalline powder of [Sn3OF]PO4 is prepared by solid state synthesis. The raw materials, SnO, SnF2 and P2O5, were well ground in a ratio of 5
:
1
:
1 and then loaded into a tidy silica tube (Φ 10 mm × 100 mm) washed with deionized water and dried at a high temperature to remove the impure phases, and then the tube was flame-sealed under 10−3 Pa. The tube was heated to 400 °C in 3 h and held at this temperature for 4 days. Afterwards, it was cooled slowly to 250 °C at a rate of 1.0 °C h−1 and then cooled to room temperature quickly. Finally, the polycrystalline powder of [Sn3OF]PO4 is obtained and the yield is 100% without impurity.
In order to compare the birefringence and other optical performance, we have also synthesized [Sn3F3]PO4 according to the literature,27 and block-like colorless crystals and polycrystalline powders were obtained.
Single crystal X-ray diffraction and powder XRD analysis
A Bruker D8 Quest diffractometer with Mo Kα radiation (λ = 0.71073 Å) was used to collect the single-crystal X-ray diffraction (XRD) data at room temperature, and the data were integrated using the SAINT program.31 The direct methods and the SHELXTL system32 were used to solve and refine the crystal structure, respectively. All the atom positions were refined using full-matrix least-squares techniques. The purity of the sample was confirmed by powder XRD, which was performed on a Bruker Model D8 Advance diffractometer equipped with Cu Kα radiation at room temperature. The diffraction patterns were recorded with the 2θ range from 10 to 70°, the scanning step width was 0.02° and the scanning rate was 1 s per step. The simulated and experimental powder X-ray diffraction peaks of [Sn3F3]PO4 and [Sn3OF]PO4 are consistent, ensuring the purity of the sample (Fig. S1†).
UV diffuse-reflectance spectroscopy
The UV-Vis-NIR diffuse reflectance spectra for [Sn3F3]PO4 and [Sn3OF]PO4 were characterized using a Shimadzu UV-3600i Plus UV-Vis-NIR spectrophotometer at room temperature. The reflectance spectrum was transformed to absorbance using the Kubelka–Munk function.33
Infrared (IR) spectroscopy
Infrared (IR) spectra were recorded on a Thermo Scientific Nicolet iS50 FT-IR spectrophotometer in the range from 500 to 2000 cm−1 with a resolution of 1 cm−1. The sample was mixed thoroughly with 500 mg of dried KBr.
Birefringence measurement
The refractive index differences of [Sn3F3]PO4 and [Sn3OF]PO4 were characterized using a polarizing microscope (ZEISS Axis Scope. 5 pol) equipped with a Berek compensator. The average wavelength of the light source was 546 nm. The formula for calculating the birefringence is as follows: | R = |Ng − Np| × d = Δn × d | (1) |
where R represents the optical path difference; Ng, Np and Δn mean the refractive index of fast light, that of slow light, and the difference value of the refractive index, respectively; and d denotes the thickness of the crystal.
Thermal analysis
A NETZSCH STA 449F5 instrument was used to examine the thermal stability of [Sn3F3]PO4 and [Sn3OF]PO4 under static N2 conditions. The sample and reference (Al2O3) were encapsulated in a Pt crucible, which was heated up to 800 °C at a rate of 10 °C min−1 and then cooled to room temperature.
DFT calculations
Density functional theory34,35 (DFT) calculations on the electronic structures and linear optical properties of [Sn3F3]PO4 and [Sn3OF]PO4 were carried out using the Vienna Ab initio Simulation (VASP) Package.36,37 The projector augmented wave (PAW) method was used to treat the valence–core interactions. The 5s and 5p states of Sn atoms, the 3s and 3p states of P atoms, the 2s and 2p states of O atoms, and the 2s and 2p states of F atoms were treated as valence states. The kinetic energy cutoff of the plane wave basis was set to be 500 eV. The Perdew–Burke–Ernzerhof38,39 (PBE) functional within the generalized gradient approximation (GGA) was chosen to account for exchange–correlation effects. The Brillouin zone was sampled using a Gamma-centered mesh with a minimum spacing of 0.4 Å−1 between k points. The convergence criterion for self-consistent field (SCF) iterations and structural relaxation was set to be 10−6 and 10−5 eV, respectively. The frequency-dependent dielectric matrix was calculated using the summation-over-states approach, and then linear optical properties, including the refractive index n and the birefringence Δn, can be assessed accordingly. Recognizing the inherent underestimation of band gaps by GGA, a scissor value was applied to align calculated and experimental band gaps. The refractive indices along the crystallographic axes na, nb and nc, as well as nx, ny and nz along the dielectric axes, were calculated using the ellipsoid formula, with the aid of included angles between the dielectric and crystallographic axes.
Results and discussion
Crystal structure description and comparison
[Sn3OF]PO4 is first reported herein and crystallizes in the monoclinic space group P21/c (no. 14), with a = 4.8865(2) Å, b = 11.4484(5) Å, c = 11.9947(5) Å, β = 97.286(1)°, V = 665.60(5) Å3, and Z = 4 (Fig. 1 and Tables S1–S3†). Another related yet different compound established nearly 40 years ago,27 [Sn3F3]PO4, crystallizes in a higher symmetry trigonal space group R3 (no. 146), with a = 11.8537(3) Å, c = 4.6271(2) Å, V = 563.05 (4) Å3, and Z = 3. Herein, the crystallographic asymmetric unit consists of one unique 9b site occupied by Sn, one unique 3a site occupied by P, one unique 9b site and one unique 3a site occupied by O and one unique 9b site occupied by F in [Sn3F3]PO4, respectively. Because the P(1) atom is located on the triad axis, an independent P(1) and O(1) and O(2) atoms form a C3 [PO4]3− tetrahedron. There are three same P(1)–O(1) bonds with a length of 1.523 Å. Meanwhile, a longer P(1)–O(2) bond with a length of 1.557 Å can be found along the triad axis. All the P–O bond lengths are within a reasonable range (Fig. S2†).
 |
| Fig. 1 The structures of [Sn3F3]PO4 and [Sn3OF]PO4F. Tin: gray, phosphorus: dark blue, oxygen: red, fluorine: dusty blue, and [PO4]3− dusty blue. (a) The spiral chain [SnO2F]3−∞. (b) The 3D structure of [Sn3F3]PO4. (c) The [Sn3O4F6]8− polyhedron with the attached [PO4]3−. (d) The 3D structure of [Sn3OF]PO4F. (e) The [Sn3O5F]5−∞ layer. (f) [Sn3O6F2]8−. | |
In addition, there is one unique crystallographic Sn atom that coordinates with O(1), O(2) and two F(1) atoms to form a distorted [SnO2F2]4− tetrahedron in [Sn3F3]PO4 (Fig. 1c). In the [SnO2F2]4− tetrahedron, the lengths of Sn–O bonds are 2.098 and 2.567 Å and Sn–F bonds are 2.058 and 2.264 Å within a reasonable bond length range. Furthermore, each [SnO2F2]4− tetrahedron is linked with the adjacent one by a shared F atom to form a 31 counterclockwise spiral chain [SnO2F]3−∞ which is along the c direction (Fig. 1a). The spiral chains are connected by shared O atoms to form a 3D network structure, within which the [PO4]3− tetrahedra are incorporated. Observing this structure, a hole with a diameter of 2.35 Å can be found. However, it's smaller than the diameter 2.75 Å of a H2O molecule reported in the literature,40 and it's assumed that the H2O molecule couldn't be filled in. This observation is in line with the experimental result that regardless of changing the experimental conditions, the crystalline hydrate of [Sn3F3]PO4 can't be obtained.
Besides, the [Sn3OF]PO4 compound can be transformed from the [Sn3F3]PO4 compound by substituting two F− with one O2−. To be specific, the change in the crystal structure begins with the conversion from the [Sn3O4F6]8− polyhedron in [Sn3F3]PO4 to the [Sn3O6F2]8− polyhedron in [Sn3OF]PO4. As shown in Fig. 1c, the three same [SnO2F2]4− polyhedra in the [Sn3F3]PO4 compound form a [Sn3O4F6]8− trimer by sharing the O(2) atom. Moreover, it is found that the [Sn3O4F6]8− polyhedron is similar to the [Sn3O6F2]8− polyhedron in the [Sn3OF]PO4 compound. Therefore, what can force the structure transformation? Actually, as illustrated in Fig. 1c and f, a [SnO2F2]4− polyhedron remains and the other two [SnO2F2]4− polyhedra rotate along the Sn(1)–O(2) bonds. The rotation ultimately leads to a change in the polyhedral structure from [Sn3O4F6]8− in [Sn3F3]PO4 to [Sn3O6F2]8− in [Sn3OF]PO4. Then, through the substitution of O and F atoms, the transformation can be achieved.
Herein, [Sn3O6F2]8− as the Sn-centered structure framework unit forms a [Sn3O5F]5−∞ 2D layer (Fig. 1e) through the edge-sharing O(1) atom along the b direction and the corner-sharing F(1) atom along the c direction. Linked with the [PO4]3− tetrahedra along the a direction, the adjacent layers are connected to form a 3D [Sn3OF]PO4 network structure (Fig. 1d). In [Sn3OF]PO4, each crystallographic symmetric unit consists of three unique 4e sites occupied by Sn, one unique 4e site occupied by P, five unique 4e sites occupied by O and one unique 4e site occupied by F. All the bonds are within a reasonable range, of which the Sn–O bond is from 2.067 Å to 2.740 Å, the Sn–F bond is from 2.259 Å to 2.283 Å and the P–O bond is from 1.532 Å to 1.559 Å (Fig. 1f and S2†).
Because of the transformation of the crystal structure from [Sn3F3]PO4 to [Sn3OF]PO4, the space group changes from R3 to P21/c, that is to say, from a high symmetry to a low symmetry. What causes it? From Fig. 2a viewed along the ab plane, it can be observed that the [Sn3F3]PO4 compound has two kinds of symmetric axes, C3 and 31 axes. For instance, the [PO4]3− polyhedron is located on the C3 axis, and the [SnO2F]3−∞ chain is parallel to the 31 axis. However, due to the abovementioned rotation (Sn(1)–O(2)) in Sn-centered structure framework units and the assistance from the O atom which results in asymmetric components from [Sn3F3]PO4 to [Sn3OF]PO4, the higher symmetric 31 and C3 axes are destroyed, resulting in a transformation from a high symmetry (R3) to a low symmetry (P21/c).
 |
| Fig. 2 (a) [Sn3F3]PO4 viewed along the ab plane and the [001] projection of the symmetry elements of the space group R3. (b) [Sn3OF]PO4F viewed along the ac and bc planes and the [010] projection of the symmetry elements of the space group P21/c. (c and e) The orientation of Sn-based polyhedra in [Sn3F3]PO4 and [Sn3OF]PO4F, respectively. (d and f) The calculated birefringence and the observed birefringence at 546 nm for [Sn3F3]PO4 and [Sn3OF]PO4F, respectively. | |
Optical and thermal properties
The UV/Vis-NIR diffuse reflection spectrum was obtained and is shown in Fig. S3.† It presents the UV cutoff edges of [Sn3F3]PO4 and [Sn3OF]PO4, which are 258 nm and 306 nm, respectively. The converted absorption pattern indicates that the corresponding band gaps of [Sn3F3]PO4 and [Sn3OF]PO4 are 4.31 eV and 3.62 eV, respectively, and the distinction derives from that [Sn3F3]PO4 has more F atoms that are more electronegative than the O atom.41
Combining the IR absorption spectrum in Fig. S4,† it could be indicated that 1058, 997, and 940 cm−1 for [Sn3F3]PO4 and 1039, 981, and 948 cm−1 for [Sn3OF]PO4 are from the asymmetric stretching vibrations of [PO4]3− groups, while 586 and 555 cm−1 for [Sn3F3]PO4 and 610, 579, and 533 cm−1 for [Sn3OF]PO4 are from the asymmetric bending vibrations of [PO4]3− groups. This demonstrates that there are [PO4]3− groups in the title compounds.
The second harmonic generation (SHG) signal of [Sn3F3]PO4 at a 1064 nm light source with a different particle size is shown in Fig. S5.† The result illustrates that [Sn3F3]PO4 is phase-matchable at 1064 nm and features a weak SHG response, about 0.3 × KDP. Besides, energy dispersive spectroscopy (EDS) confirmed the existence of Sn, P, O and F elements in both [Sn3F3]PO4 and [Sn3OF]PO4 (Fig. S6†). The thermogravimetric analysis (TG) and differential scanning calorimetry (DSC) curves indicated that [Sn3F3]PO4 and [Sn3OF]PO4 are relatively stable before 300 °C and gradually decompose after that (Fig. S7†).
In addition, for linear optical properties of [Sn3F3]PO4 and [Sn3OF]PO4, according to crystal thicknesses d (38.6 and 31.0 μm, respectively) and delay values for full extinction R at 546 nm (1.06 and 3.62 μm, respectively) (Fig. S8†), the birefringences of [Sn3F3]PO4 and [Sn3OF]PO4 are larger than 0.027 and 0.117 at 546 nm based on formula (1), respectively. The birefringence of [Sn3OF]PO4 exceeds that of many tin(II) phosphates, such as Sn3(PO4)2 (0.052 at 1064 nm), Rb(Sn3O)2(PO4)3 (0.081 at 1064 nm), and NH4(Sn3O)2(PO4)3 (0.082 at 1064 nm). Moreover, the calculated birefringences of [Sn3F3]PO4 and [Sn3OF]PO4 are 0.025 and 0.104 at 546 nm, respectively, which are in good agreement with the experimental data (Fig. 2d and f). Furthermore, the birefringence of [Sn3OF]PO4 is almost five times that of [Sn3F3]PO4. But why is the optical anisotropy so different for the title compounds that contain the same element and why is there such a large birefringence gain?
Origin of birefringence
Generally speaking, a rigid tetrahedron, such as [PO4]3−, with small polarizability anisotropy contributes little to birefringence.42–44 Hence, the optical anisotropy of [Sn3F3]PO4 and [Sn3OF]PO4 mainly hinges on the geometry of the asymmetric Sn-centered polyhedron ([SnO2F2]4− for [Sn3F3]PO4 and [SnO3F]5− for [Sn3OF]PO4) with stereochemically active lone pairs as well as the orientation. The orientation of lone pairs is consistent with the direction from the center of the polyhedron ([SnO2F2]4− in [Sn3F3]PO4 and [SnO3F]5− in [Sn3OF]PO4) to the Sn atom (Fig. 2c and e).45 In fact, optical anisotropy originates from the different contributions of stereochemically active lone pairs towards the dielectric axis. Large contribution differences often correspond to large optical anisotropy.
On the one hand, since the [Sn3F3]PO4 compound crystallizes in the higher symmetric space group R3, there are angles of 59.3° between all the lone pairs in Sn-centered polyhedra and the c-axis in the unit cell (Fig. 2c). Then the orientation of the lone pairs in Sn-based polyhedra can be projected on the ac plane and perpendicular to the ac plane, respectively. It is speculated that if the angle is 45° between the lone pairs and the c-axis, it can result in minimum anisotropy because there are the two approximate contributions of stereochemically active lone pairs to the optical anisotropy along the ac plane and perpendicular to the ac plane. Consequently, the contributions of stereochemically active lone pairs of [SnO2F2]4− in [Sn3F3]PO4 to the optical anisotropy along the ac plane and perpendicular to the ac plane are 5.46 and 4.29, respectively (detailed calculations are provided in Table S4†). Therefore, it demonstrates that there is a little difference in the contribution, which results in small optical anisotropy.
On the other hand, [Sn3OF]PO4 belongs to a lower symmetry (P21/c), and the dielectric axes are shown in Fig. 2e. The contributions of three unique crystallographic Sn-based polyhedra in [Sn3OF]PO4 are different, as for [Sn(1)O3F]5−: 0.56, 0.48, and 3.44, [Sn(2)O3F]5−: 0.16, 3.44, and 0.60, and [Sn(3)O3F]5−: 2.08, 2.24, and 2.20 along the dielectric axis (nx, ny, and nz), respectively. Compared with Sn(3), [Sn(1)O3F]5− and [Sn(2)O3F]5− have a greater distinction in contributions on the three axes, and it could be inferred that Sn(1) and Sn(2) may have a greater effect on optical anisotropy. At the same time, the total contributions are 2.80, 6.16 and 6.24 along the dielectric axis (nx, ny, and nz), respectively (detailed calculations are provided in Table S4†). This agrees well with the relationship among the calculated refractive indices nx, ny and nz, where nz and ny are far larger than nx. Hence, a relatively large difference in the contribution drives large optical anisotropy, agreeing with the experimental birefringence.
Additionally, considering the aligned cooperativity of the asymmetric Sn-centered polyhedron, defined
(θi represents the ith angle between the polyhedra; n represents the number of angles per unit),45 it is observed that asymmetric Sn-centered polyhedra in [Sn3OF]PO4 have more aligned cooperativity than those in [Sn3F3]PO4 (0.54 vs. 0.47) (detailed calculations are provided in Tables S4 and S5†).
Thus, it can be inferred that a large birefringence enhancement is achieved by breaking the symmetry with disadvantage for optical anisotropy and reconstructing the aligned arrangement of the lone pairs. Combined with the above view, the explanation is represented for KSn4(PO4)3 (R3c, 0.013 at 1064 nm) and NaSn4(PO4)3 (R3c, 0.005 at 1064 nm) with high symmetry and small birefringence. The reason is that [SnO3]4− polyhedra in the two isostructural compounds, Sn-centered structure framework units, are coincidentally located on the C3 and 31 axes, as shown in Fig. S9.† It results in a small distinction in contributions of optical anisotropy.
Theoretical calculations
Both the geometric configuration and electronic structure are significant factors affecting the optical properties. The partial density of states (PDOS) of [Sn3F3]PO4 and [Sn3OF]PO4 are illustrated in Fig. 3(a) and (d), respectively. For both [Sn3F3]PO4 and [Sn3OF]PO4, the CBM (Conduction Band Minimum) is mainly composed of the 5p states of Sn atoms, while the VBM (Valence Band Maximum) is mainly composed of the 5s states of Sn atoms and the 2p states of O atoms. However, the contribution of O-2p states around the VBM in [Sn3OF]PO4 is larger than that in [Sn3F3]PO4. This is consistent with the larger band gap of [Sn3F3]PO4 (band gaps calculated using PBE are shown in Fig. S10†). To deeply understand the bonding feature of [Sn3F3]PO4 and [Sn3OF]PO4, the negative partial crystal orbital Hamiltonian population46–48 (pCOHP) is presented in Fig. 3(b) and (e). Below the Fermi level, a partition of the valence states appears as antibonding states for both Sn–O and Sn–F bonds. However, the antibonding interactions in [Sn3F3]PO4 are smaller than those in [Sn3OF]PO4. Thus, the Sn–O and Sn–F bonds in [Sn3F3]PO4 are stronger than those in [Sn3OF]PO4. Furthermore, it can be visualized from the electron localization function (ELF) diagrams (Fig. 3c and f) that Sn2+ has an asymmetric cloud of lone pairs, indicating that both [Sn3F3]PO4 and [Sn3OF]PO4 show considerable strength of lone pairs under the same conditions. Therefore, there is no obvious difference in the electronic structure compared to the large observed birefringence difference between [Sn3F3]PO4 and [Sn3OF]PO4. Therefore, the effect of the electronic structure on birefringence is relatively small. Hence, it can be inferred that the geometries rather than electronic structures play an important role in contributing to the optical properties.
 |
| Fig. 3 (a and d) Partial density of states (PDOS), (b and e) negative crystal orbital Hamiltonian population (–pCOHP), and (c and f) electron localization function (ELF) diagrams of [Sn3F3]PO4 and [Sn3OF]PO4, respectively. | |
Conclusions
In conclusion, a novel compound [Sn3OF]PO4 is discovered, exhibiting significant birefringence (Δnobv. = 0.117 at 546 nm), a five-fold improvement compared to that of a related compound [Sn3F3]PO4. The remarkable enhancement in birefringence is primarily attributed to the anisotropic [SnO3F]5− structure building unit. By simple regulation of the F/O ratio, the undesirable R3 symmetry in [Sn3F3]PO4 is disrupted. This regulation also realigns the lone pairs, resulting in substantial optical anisotropy. These strategies for designing compounds and optimizing performance could serve as an experimental guideline for discovering and optimizing effective birefringent crystals.
Author contributions
Y. H. H., J. Y. G. and L. M. W. conceived, designed and performed all the experimental work. X. X. and J. Y. H. performed the theoretical calculations. Y. H. H. and L. C. wrote the draft paper. All the authors discussed the results and commented on the manuscript. All authors contributed to the general discussion.
Data availability
The data supporting this article have been included as part of the ESI.†
Crystallographic data have been deposited at CCDC 2332572 and 2334179 for [Sn3OF]PO4 and [Sn3F3]PO4, respectively.†
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (22205024 and 22193043), Start-up Funding from Beijing Normal University (310432105) and the National Key R&D Program of China (2022YFB3803902). This work was also supported by the Interdisciplinary Intelligence SuperComputer Center of Beijing Normal University, Zhuhai.
References
- S. J. Han, A. Tudi, W. B. Zhang, X. L. Hou, Z. H. Yang and S. L. Pan, Recent development of SnII, SbIII-based birefringent material: crystal chemistry and investigation of birefringence, Angew. Chem., Int. Ed., 2023, 62, e202302025 CrossRef CAS PubMed.
- K. M. Ok, Toward the rational design of novel noncentrosymmetric materials: factors influencing the framework structures, Acc. Chem. Res., 2016, 49, 2774–2785 CrossRef CAS.
- C. Wu, X. X. Jiang, Z. J. Wang, L. Lin, Z. S. Lin, Z. P. Huang, X. F. Long, M. G. Humphrey and C. Zhang, Giant optical anisotropy in the UV-transparent 2D nonlinear optical material Sc(IO3)2(NO3), Angew. Chem., Int. Ed., 2021, 60, 3464–3468 CrossRef CAS.
- X. D. Chai, M. Z. Li, S. J. Lin, W. F. Chen, X. M. Jiang, B. W. Liu and G. C. Guo, Cs4Zn5P6S18I2: the largest birefringence in chalcohalide achieved by highly polarizable nonlinear optical functional motifs, Small, 2023, 19, 2303847 CrossRef CAS.
- Y. Lin, C. L. Hu, Z. Fang, J. Chen, W. J. Xie, Y. Chen, J. P. Wang and J. G. Mao, KRE(CO3)2 (RE = Eu, Gd, Tb): new mixed metal carbonates with strong photoluminescence and large birefringence, Inorg. Chem. Front., 2022, 9, 5645–5652 RSC.
- G. Q. Zhou, J. Xu, X. D. Chen, H. Y. Zhong, S. T. Wang, K. Xu, P. Z. Deng and F. X. Gan, Growth and spectrum of a novel birefringent α-BaB2O4 crystal, J. Cryst. Growth, 1998, 191, 517–519 CrossRef CAS.
- A. Tudi, S. J. Han, Z. H. Yang and S. L. Pan, Potential optical functional crystals with large birefringence: recent advances and future prospects, Coord. Chem. Rev., 2022, 459, 214380 CrossRef CAS.
- J. Lu, Y. K. Lian, L. Xiong, Q. R. Wu, M. Zhao, K. X. Shi, L. Chen and L.-M. Wu, How to maximize birefringence and nonlinearity of π-conjugated cyanurates, J. Am. Chem. Soc., 2019, 141, 16151–16159 CrossRef CAS.
- X. Liu, Y. C. Yang, M. Y. Li, L. Chen and L.-M. Wu, Anisotropic structure building unit involving diverse chemical bonds: a new opportunity for high-performance second-order NLO materials, Chem. Soc. Rev., 2023, 52, 8699–8720 RSC.
- G. Q. Shi, Y. Wang, F. F. Zhang, B. B. Zhang, Z. H. Yang, X. L. Hou, S. L. Pan and K. R. Poeppelmeier, Finding the next deep-ultraviolet nonlinear optical material: NH4B4O6F, J. Am. Chem. Soc., 2017, 139, 10645–10648 CrossRef CAS.
- S. J. Han, B. B. Zhang, Z. H. Yang and S. L. Pan, From LiB3O5 to NaRbB6O9F2: fluorine-directed evolution of structural chemistry, Chem. – Eur. J., 2018, 24, 10022–10027 CrossRef CAS PubMed.
- J. Lu, J. N. Yue, L. Xiong, W. K. Zhang, L. Chen and L.-M. Wu, Uniform alignment of non-π-conjugated species enhances deep ultraviolet optical nonlinearity, J. Am. Chem. Soc., 2019, 141, 8093–8097 CrossRef CAS PubMed.
- S. J. Han, M. Mutailipu, A. Tudi, Z. H. Yang and S. L. Pan, PbB5O7F3: a high-performing short-wavelength nonlinear optical material, Chem. Mater., 2020, 32, 2172–2179 CrossRef CAS.
- M. Mutailipu, M. Zhang, B. B. Zhang, L. Y. Wang, Z. H. Yang, X. Zhou and S. L. Pan, SrB5O7F3 functionalized with [B5O9F3]6− chromophores: accelerating the rational design of deep-ultraviolet nonlinear optical materials, Angew. Chem., Int. Ed., 2018, 57, 6095–6099 CrossRef PubMed.
- M. Mutailipu, J. Han, Z. Li, F. M. Li, J. J. Li, F. F. Zhang, X. F. Long, Z. H. Yang and S. L. Pan, Achieving the full-wavelength phase-matching for efficient nonlinear optical frequency conversion in C(NH2)3BF4, Nat. Photonics, 2023, 17, 694–701 CrossRef CAS.
- Y. Wang, B. B. Zhang, Z. H. Yang and S. L. Pan, Cation-tuned synthesis of fluorooxoborates: towards optimal deep-ultraviolet nonlinear optical materials, Angew. Chem., Int. Ed., 2018, 57, 2150–2154 CrossRef CAS.
- B. B. Zhang, G. P. Han, Y. Wang, X. L. Chen, Z. H. Yang and S. L. Pan, Expanding frontiers of ultraviolet nonlinear optical materials with fluorophosphates, Chem. Mater., 2018, 30, 5397–5403 CrossRef CAS.
- L. Xiong, J. Chen, J. Lu, C.-Y. Pan and L.-M. Wu, Monofluorophosphates: a new source of deep-ultraviolet nonlinear optical materials, Chem. Mater., 2018, 30, 7823–7830 CrossRef CAS.
- Y. Yang, Y. Qiu, P. F. Gong, L. Kang, G. M. Song, X. M. Liu, J. L. Sun and Z. S. Lin, Lone-Pair enhanced birefringence in an alkaline-earth metal tin(II) phosphate BaSn2(PO4)2, Chem. – Eur. J., 2019, 25, 5648–5651 CrossRef CAS.
- J. Y. Guo, A. Tudi, S. J. Han, Z. H. Yang and S. L. Pan, Sn2B5O9Cl: a material with large birefringence enhancement activated prepared via alkaline-earth-metal substitution by tin, Angew. Chem., Int. Ed., 2019, 58, 17675–17678 CrossRef CAS PubMed.
- Y. C. Liu, X. M. Liu, S. Liu, Q. R. Ding, Y. Q. Li, L. N. Li, S. G. Zhao, Z. S. Lin, J. H. Luo and M. C. Hong, An unprecedented antimony(III) borate with strong linear and nonlinear optical responses, Angew. Chem., Int. Ed., 2020, 59, 7793–7796 CrossRef CAS PubMed.
- J. Y. Guo, A. Tudi, S. J. Han, Z. H. Yang and S. L. Pan, α-SnF2: a UV birefringent material with large birefringence and easy crystal growth, Angew. Chem., Int. Ed., 2021, 60, 3540–3544 CrossRef CAS.
- J. Y. Guo, S. C. Cheng, S. J. Han, Z. H. Yang and S. L. Pan, Sn2B5O9Br as an outstanding bifunctional material with strong second-harmonic generation effect and large birefringence, Adv. Opt. Mater., 2021, 9, 2001734 CrossRef CAS.
- J. Y. Guo, A. Tudi, S. J. Han, Z. H. Yang and S. L. Pan, Sn2PO4I: an excellent birefringent material with giant optical anisotropy in non π-conjugated phosphate, Angew. Chem., Int. Ed., 2021, 60, 24901–24904 CrossRef CAS.
- J. F. Deng, Y. J. Kang, J. X. Mi, M. R. Li, J. T. Zhao and S. Y. Mao, KSn4(PO4)3, Acta Crystallogr., 2004, 60, i116–i117 CAS.
- S. Y. Mao, J. F. Deng, M. R. Li, J. X. Mi, H. H. Chen and J. T. Zhao, Crystal structure of sodium tetratin(II) triphosphate, NaSn4(PO4)3, Z. Kristallogr. – New Cryst. Struct., 2004, 219, 205–206 CAS.
- S. B. Etcheverry, G. E. Narda, M. C. Apella and E. J. Baran, Hydrolytic properties of Sn3PO4F3 (short communication), Caries Res., 1986, 20, 120–122 CrossRef CAS.
- P. Yu, L.-M. Wu, L. J. Zhou and L. Chen, Deep-ultraviolet nonlinear optical crystals: Ba3P3O10X (X = Cl, Br), J. Am. Chem. Soc., 2014, 136, 480–487 CrossRef CAS.
- H. W. Yu, W. G. Zhang, J. Young, J. M. Rondinelli and P. S. Halasyamani, Design and synthesis of the beryllium-free deep-ultraviolet nonlinear optical material Ba3(ZnB5O10)PO4, Adv. Mater., 2015, 27, 7380–7385 CrossRef CAS PubMed.
- J. Chen, L. Xiong, L. Chen and L.-M. Wu, Ba2NaCIP2O7: unprecedented phase matchability induced by symmetry breaking and its unique fresnoite-type structure, J. Am. Chem. Soc., 2018, 140, 14082–14086 CrossRef CAS PubMed.
-
SAINT, version 7.60A, Bruker Analytical X-ray Instruments, Inc., Madison, WI, 2013 Search PubMed.
-
SHELXTL, version 6.1, Bruker Analytical X-ray Instruments, Inc., Madison, WI, 2013 Search PubMed.
- P. Kubelka and F. Munk, An article on optics of paint layers, Tech. Phys., 1931, 12, 593–609 Search PubMed.
- P. Hohenberg and W. Kohn, Density functional theory (DFT), Phys. Rev., 1964, 136, B864–B871 CrossRef.
- W. Kohn and L. J. Sham, Self-consistent equations including exchange and correlation effects, Phys. Rev., 1965, 140, A1133–A1138 CrossRef.
- G. Kresse and J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 558–561 CrossRef CAS.
- G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS.
- J. P. Perdew, K. A. Jackson, M. R. Pederson, D. J. Singh and C. Fiolhais, Erratum: atoms, molecules, solids, and surfaces: applications of the generalized gradient approximation for exchange and correlation, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 46, 4978 CrossRef.
- J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS.
- M. Caurie, The unimolecular character of the classical Brunauer, Emmett and Teller adsorption equation and moisture adsorption, Int. J. Food Sci. Technol., 2005, 40, 283–293 CrossRef CAS.
- W. Q. Jin, C. W. Xie, X. L. Hou, M. Cheng, E. Tikhonov, M. F. Wu, S. L. Pan and Z. H. Yang, From monofluorophosphates A2PO3F to difluorophosphates APO2F2 (A = alkali metal): design of new potential deep-ultraviolet nonlinear optical system with shortened phase-matching wavelength, Chem. Mater., 2023, 35, 5281–5290 CrossRef CAS.
- X. H. Zhang, B. P. Yang, J. Chen, C. L. Hu, Z. Fang, Z. J. Wang and J. G. Mao, A new iodate-phosphate Pb2(IO3)(PO4) achieving great improvement in birefringence activated by (IO3)- groups, Chem. Commun., 2020, 56, 635–638 RSC.
- L. Xiong, L.-M. Wu and L. Chen, A general principle for DUV NLO materials: π-conjugated confinement enlarges band gap, Angew. Chem., Int. Ed., 2021, 60, 25063–25067 CrossRef CAS.
- X. Y. Zhang, L. Kang, P. F. Gong, Z. S. Lin and Y. C. Wu, Nonlinear optical oxythiophosphate approaching the good balance with wide ultraviolet transparency, strong second harmonic effect, and large birefringence, Angew. Chem., Int. Ed., 2021, 60, 6386–6390 CrossRef CAS PubMed.
- J. Y. Guo, J. B. Huang, A. Tudi, X. L. Hou, S. J. Han, Z. H. Yang and S. L. Pan, Birefringence regulation by clarifying the relationship between stereochemically active lone pairs and optical anisotropy in tin-based ternary halides, Angew. Chem., Int. Ed., 2023, 62, e202304238 CrossRef CAS PubMed.
- A. Walsh and G. W. Watson, Influence of the anion on lone pair formation in Sn(II) monochalcogenides: a DFT study, J. Phys. Chem. B, 2005, 109, 18868–18875 CrossRef CAS PubMed.
- A. Walsh, D. J. Payne, R. G. Egdell and G. W. Watson, Stereochemistry of post-transition metal oxides: revision of the classical lone pair model, Chem. Soc. Rev., 2011, 40, 4455–4463 RSC.
- V. L. Deringer, A. L. Tchougréeff and R. Dronskowski, Crystal orbital hamilton population (COHP) analysis as projected from plane-wave basis sets, J. Phys. Chem. A, 2011, 115, 5461–5466 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available: Experimental sections, computational methods, crystallographic data tables, powder XRD patterns, and optical property measurements. CCDC 2332572 ([Sn3OF]PO4) and 2334179 ([Sn3F3]PO4). For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4qi01203h |
|
This journal is © the Partner Organisations 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.