Designing promising ultraviolet (UV) birefringent crystals with different hydrogen-bonded phosphate frameworks

Miao-Bin Xu a, Jin Chen *ab, Huai-Yu Wu a, Jia-Jia Li a, Ning Yu a, Mo-Fan Zhuo a, Fei-Fei Mao *bc and Ke-Zhao Du a
aCollege of Chemistry and Materials Science, Fujian Key Laboratory of Polymer Materials, Fujian Normal University, Fuzhou, 350002 China. E-mail: cj2015@fjnu.edu.cn
bState Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou, 350002, P. R. China
cCollege of Science, Nanjing Agricultural University, Nanjing 210095, P. R. China

Received 15th May 2024 , Accepted 8th June 2024

First published on 10th June 2024


Abstract

In this study, we computationally identified using first-principles calculations two pyridine derivatives with high polarizability anisotropy: (C6H7N2O)+ (protonated nicotinamide, abbreviated as 3AP+) and (C6H6NO2)+ (protonated nicotinic acid, abbreviated as 3CP+). Subsequently, we synthesized two novel semiorganic crystals: (3AP) (H2PO4) (I) and (3CP) (H2PO4) (II). Both crystals incorporate hydrogen-bonded phosphate frameworks (HPFs), which are constituted by H2PO4 anions linked through O–H⋯O hydrogen bonds. Compound I (Fdd2) features one-dimensional (1D) hydrogen-bonded phosphate frameworks (HPFs) that interconnect with 2D [(3AP)(H2PO4)] layers, forming a complex 3D network. In contrast, compound II (Pbca) possesses 2D HPFs that are integrated with 3CP+ cations to form a 2D layered network. Compound I exhibits a moderate second-harmonic generation (SHG) effect (1 × KH2PO4), and a significant birefringence (Δncal.: 0.191@1064 nm). Furthermore, compound II exhibits both a broad bandgap (4.24 eV) but also an exceptional birefringence (Δnexp.: 0.284@546 nm), which is the highest value reported among all semiorganic phosphates. This suggests that II could be a promising candidate for use as an outstanding ultraviolet (UV) birefringent material.


Introduction

Birefringent crystals are key materials used in the production of optical devices, including polarizers, optical isolators, circulators, and phase retarders.1–7 Inorganic phosphate crystals, known for their short ultraviolet cutoff edges and good crystal growth habits, represent a classic optoelectronic material.8–13 KDP (KH2PO4) and KTP (KTiPO4) are quintessential examples of phosphates, boasting significant application as nonlinear optical (NLO) materials. However, the high symmetry and weak anisotropy of the PO43− tetrahedron often limit inorganic phosphates, resulting in low birefringence and phase mismatch.14–18 Therefore, the search for phosphates with high birefringence remains an active area of research.

To date, researchers have obtained some metal phosphates with high birefringence (Δn > 0.05) and proposed several effective strategies, including: (1) introducing lone pair electron cations (such as Sn2+, I5+) or d0 transition metal cations (such as Mo6+, W6+).19–25 These cations can form the building units with high structural distortion and strong polarizability anisotropy, which is beneficial for enhancing the birefringence of phosphates, such as Sn2PO4I (Δncal. = 0.664@546 nm).21 (2) Partial substitution of O atoms with X groups (including F, S, or –NH3) can break the high symmetry of PO43− groups, thereby constructing distorted POmX4−m tetrahedra with strong polarizability anisotropy, leading to enhanced birefringence, for example, NaNH4PO3F·H2O (Δnexp. = 0.053@589.3 nm),26 KH2PO3S (Δncal. = 0.1@1 μm),27 and NaPO3NH3nexp. = 0.062@546.1 nm).28 (3) Combining PO43− with traditional planar π-conjugated birefringent functional groups (such as CO32− and BO33−). This method can increase the birefringence of compounds (Δn > 0.1) while also having a wide bandgap (Eg > 6.2 eV), such as Sr3Y(PO4)(CO3)3ncal. = 0.121@532 nm and Eg = 6.9 eV).29 Notably, Rb[PO2(NH)3(CO)2]·0.5H2O,30 KPO2(NHCONH2)2,31 and NaPO2(NH)3(CO)2[thin space (1/6-em)]32 can be regarded as products designed using both methods (2) and (3) concurrently. Of these, NaPO2(NH)3(CO)2 exhibits very large birefringence and a wide bandgap (Δnexp. = 0.280@550 nm; Eg ≥ 6.5 eV).32 (4) The “zipper” arrangement of the PO43− tetrahedron with large-angle deviation can also lead to the enhancement of birefringence of metal phosphates. For example, α-YSc(PO4)2 exhibits deep ultraviolet transmission and a high birefringence (Δncal. = 0.102@1064 nm).33

Recently, several semiorganic cations with planar π-conjugated configurations, including [C(NH2)3]+, (C5H6ON)+, (C3H7N6)+, etc., have emerged as high-performance anisotropic functional moieties.34–37 The combination of π-conjugated organic cations with phosphate tetrahedra has led to the discovery of numerous novel crystals with high birefringence, such as [C(NH2)3]3PO4·2H2O (Δnexp. = 0.055@546 nm),38 [C(NH2)3]6(PO4)·3H2O (Δnexp. = 0.078@546 nm),39 (C3H7N6)6(H2PO4)4(HPO4)·4H2O (Δncal. = 0.220@1064 nm),40 and (C5H6ON)+(H2PO4)ncal. = 0.25@1064 nm).41 Clearly, as the π-conjugation in organic cations expands, their contribution to birefringence similarly increases significantly. Therefore, we identified two novel birefringent functional units, characterized by higher π-conjugation and stronger polarizability anisotropy, which are (C6H6N2O)+ (protonated 3-carboxamide pyridine, 3AP+) and (C6H6NO2)+ (protonated 3-carboxypyridine, 3CP+) (Scheme 1). And then, our efforts have resulted in the discovery of two novel semiorganic phosphates, namely (3AP) (H2PO4) (I) and (3CP)(H2PO4) (II). Experimental characterization and theoretical calculations have revealed that I exhibits large birefringence and moderate SHG effect (Δnexp. = 0.196@546 nm; 1.0 × KDP). II can serve an excellent UV birefringent crystal with the bandgap of 4.20 eV and birefringence of 0.284@546 nm, the highest value among known semiorganic phosphates.


image file: d4qi01220h-s1.tif
Scheme 1 The HOMO and LUMO maps (a) as well as polarizability anisotropy (b) of related units.

Experimental

Materials and synthesis

C6H6N2O (99%) and H3PO4 (85%) were used as purchased from Adamas Beta. Crystals of I were synthesized using a simple evaporation technique in aqueous solution. The raw reactants of C6H6N2O (8.2 mmol, 1.00 g) and H3PO4 (0.5 mL) were mixed together in deionized water (3 mL) in a glass beaker. The solution was allowed to reach room temperature and then slowly evaporated, resulting in white flaky crystals that precipitated. For the preparation of II, the starting materials consisted of C6H6N2O (2 mmol, 244.25 mg), H3PO4 (0.5 mL), and H2O (1 mL). A mixture of the starting materials was placed into Teflon pouches (23 mL) and sealed inside an autoclave, which was heated to 110 °C for 3 days, and then cooled to 30 °C at a rate of 1.6 °C h−1. Colorless flake-like crystals I and II were obtained with yields of about 80% and 76% (based on C6H6N2O), respectively.

Single crystal structure determination

Single-crystal X-ray diffraction data for the title compounds were collected using a Rigaku XtaLAB Synergy-DW dual-wavelength CCD diffractometer, equipped with Cu Kα radiation (λ = 1.54184 Å) at 293 K. Data reduction was performed using CrysAlisPro, and absorption correction, based on the multi-scan method, was applied.42 The structures of I and II were determined using direct methods and refined by full-matrix least-squares fitting on F2 utilizing SHELXL-2014.43 Anisotropic thermal parameters were applied to refine all non-hydrogen atoms. The structure underwent a check for missing symmetry elements using PLATON, and none were found.44 Crystallographic data and structural refinements of the compounds are listed in Tables S1–S11.

Powder X-ray diffraction

Powder X-ray diffraction (PXRD) patterns were recorded on a Rigaku Ultima IV diffractometer with graphite-monochromated Cu Kα radiation over a 2θ range of 10° to 70° with a step size of 0.02°.

Thermal analysis

Thermogravimetric analysis (TGA) and differential thermal analysis (DTA) data were obtained, via using a Rigaku TG-DTA 8121 unit under an argon (Ar) atmosphere, at a heating rate of 10 °C min−1, ranging from 30 °C to 800 °C.

Optical measurements

Infrared (IR) spectra were recorded on a Thermo Fisher Scientific Nicolet 5700 FT-IR spectrometer using KBr pellets over a range from 4000 to 400 cm−1.

Ultraviolet–visible (UV-vis) spectra, ranging from 200 to 800 nm, were recorded on a PerkinElmer Lambda 750 UV-vis spectrophotometer. Reflectance spectra were converted to an absorption spectrum using the Kubelka–Munk function.45

Second harmonic generation measurements

Powder SHG measurements were taken using a Q-switched Nd:YAG laser generating radiation at 1064 nm according to the method of Kurtz and Perry.46 Crystalline samples of I were sieved into distinct particle-size ranges: 50–70, 70–100, 100–140, 140–200, 200–250, and 250–325 μm. Sieved KDP (KH2PO4) samples, within the same particle-size ranges, were used as references.

Birefringence

The optical path difference of the title compounds were characterized using a polarizing microscope (Nikon LV1000) equipped with a Berek compensator.47 The average wavelength of the light source was 546 nm. Under orthogonal polarization, the two beams of polarized light pass through the crystal at different speeds, producing interference color phenomena after passing through the analyzer. Polarized light with vibration directions parallel to K1, K2, K1′, K1′′, K2′, K2′′ represents light with different vibration directions. Polarizers are used to filter out light whose vibration directions are not parallel to the analyzer (AA) or polarizer (PP).

The birefringence (Δn) of microcrystals are calculated with the formula: R = d·(nsnf) = T·Δn. Herein, R is the optical path difference. The ns represents the refractive index of the slow light, and nf represents the refractive index of the fast light. Here, T represents the thickness of the tested single crystal.35

Elemental analysis

The elemental content was measured on Vario EL Cube elemental analyzer from Elementar Analysensysteme GmbH, Germany. The combustion temperature was 800 °C.

Energy-dispersive X-ray spectroscope

Microprobe elemental analyses and elemental distribution maps were measured using a field-emission scanning electron microscope (Phenom LE) equipped with an energy-dispersive X-ray spectrometer (EDS, Phenom LE).

Computational method

The electronic structures and optical properties of the compounds were computed using the plane-wave pseudopotential method within the framework of density functional theory (DFT), as implemented in the total energy code Cambridge Sequential Total Energy Package (CASTEP).48,49 For the exchange–correlation functional, we selected the Perdew–Burke–Ernzerhof (PBE) formula within the generalized gradient approximation (GGA).50 The interactions between the ionic cores and the valence electrons were modeled using norm-conserving pseudopotentials.51 The following electrons were considered as valence: C-2s22p2, N-2s22p3, O-2s22p4, P-3s23p3, and H-1s1. The basis set was constructed with plane waves up to a cutoff energy of 750 eV. For both compounds, the self-consistent field (SCF) and optical-property calculations utilized a k-point separation of 0.04 Å−1 for the numerical integration over the Brillouin zone. The k-point samplings employed were 4 × 4 × 4 for compound I and 2 × 1 × 1 for compound II.48,49

To investigate the polarizability anisotropy and electronic structure of selected birefringent units, a systematic computational approach was undertaken using the Gaussian 09 software suite.52 The hybrid B3LYP functional was employed at the 6-31G(d,p) level of theory for these calculations. Subsequent analysis of the computational results was performed using the Multiwfn 3.8 software package.53 The polarizability anisotropy was quantified based on the static polarizability values.54

Results and discussion

Compound I, akin to many previously reported semiorganic phosphates, was synthesized by dissolving a specific stoichiometric ratio of nicotinamide (C6H6N2O) and phosphoric acid (H3PO4) in an aqueous solution, followed by slow evaporation. It should be noted that nicotinamide is easily hydrolyzed to nicotinic acid in aqueous solutions (Fig. S1), a process that can be accelerated by heating.55–57 This hydrolysis facilitated the hydrothermal synthesis of compound IIvia the reaction involving nicotinamide and H3PO4 at 110 °C. The transparent single crystals of II, with dimensions of approximately 13 × 7 × 3 mm3, were observed (Fig. 1). Powder X-ray diffraction (PXRD) analysis confirmed the purity of the synthesized samples (Fig. S2). Field-emission scanning electron microscopy (FESEM) analyses of both I and II detected the presence of C, N, O and P elements (Fig. S3). Elemental analysis provided weight ratios of C[thin space (1/6-em)]:[thin space (1/6-em)]H[thin space (1/6-em)]:[thin space (1/6-em)]N in I and II as 5.97[thin space (1/6-em)]:[thin space (1/6-em)]1.95[thin space (1/6-em)]:[thin space (1/6-em)]8.76 and 5.61[thin space (1/6-em)]:[thin space (1/6-em)]0.99[thin space (1/6-em)]:[thin space (1/6-em)]7.90, respectively, which are congruent with the crystal structure analysis of 6[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]9 and 6[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]8, respectively (Table S12). This agreement corroborates the presence of (C6H7N2O)+ (3AP+) and (C6H6NO2)+ (3CP+) as the organic cations in the structures of I and II, respectively. The infrared (IR) spectra for I and II are presented in Fig. S4, with the detailed assignments of the absorption peaks provided in Table S13, showing consistency with those of previously reported semiorganic phosphates.39–41 Thermal analysis demonstrated that the thermal stability of both compounds decompose above 150 °C (Fig. S5), comparable to other reported crystals, such as (C5H6ON)+(H2PO4) (166 °C),41 (C3H7N6)6(H2PO4)4(HPO4)·4H2O (120 °C),40 [C(NH2)3]6(PO4)·3H2O and (100 °C).39
image file: d4qi01220h-f1.tif
Fig. 1 As-grown crystals of I (a) and II (b).

Crystal structure

Compound I crystallizes in the non-centrosymmetric space group Fdd2, with each unit cell housing 16 asymmetric units (Table S1). Each asymmetric unit consists of one (C6H6N2O)+ (3AP+) cation and one (H2PO4) anion (Fig. 2a). Within the structure of I, each 3AP+ cation is interconnected with three (H2PO4) anions through three N–H⋯O and one O–H⋯O hydrogen bonds (Fig. S6a). Conversely, each (H2PO4) anion is coordinated with three 3AP+ cations via three N–H⋯O hydrogen bonds and is additionally linked to two axial (H2PO4) anions through two O–H⋯O hydrogen bonds (Fig. S7a). Consequently, adjacent (H2PO4) anions are interconnected via O–H⋯O hydrogen bonds, resulting in the formation of one-dimensional [H2PO4] chains, denote as 1D hydrogen-bonded phosphate frameworks (1D HPFs) (Fig. 2b). Furthermore, neighboring 3AP+ cations and (H2PO4) anions are connected through hydrogen bonds, creating a 2D neutral layer of [(3AP)(H2PO4)] parallel to the ac plane (Fig. 2c). Thus, the overall three-dimensional (3D) architecture of I can be envisioned as a network of 2D [(3AP)(H2PO4)] layers interconnected by 1D HPFs (Fig. 2d).
image file: d4qi01220h-f2.tif
Fig. 2 The asymmetric unit (a), 1D HPFs (b), 2D neutral layer (c) along b direction, and 3D overall structure along c direction (d).

Compound II crystallizes in the centrosymmetric space group Pbca, with each unit cell comprising 8 asymmetric units (Table S1). Each asymmetric unit consists of one (C6H6NO2)+ (3CP+) cation and one (H2PO4) anion (Fig. 3a). Within the structure of II, pairs of (H2PO4) anions are interconnected via O–H⋯O hydrogen bonds, forming a (H4P2O8)2− dimer (Fig. S7b). These dimers further associate through O–H⋯O hydrogen bonds to create two-dimensional hydrogen-bonded phosphate frameworks (2D HPFs), which are aligned parallel to the ab plane (Fig. 3c). Additionally, adjacent 3CP+ cations are linked into one-dimensional [C6H6NO2]+ chains through weaker C–H⋯O hydrogen bonds (Fig. 3b and S6b). These chains are then integrated with the 2D HPFs through N–H⋯O and O–H⋯O hydrogen bonds (Fig. S7b). Consequently, each pair of cationic 1D [C6H6NO2]+ chains is sequentially embedded on either side of the 2D HPFs, generating a neutral 2D [(3CP)(H2PO4)] layer. These 2D layers interlace with each other along the c-axis, constituting the overall architecture of compound II (Fig. 3d).


image file: d4qi01220h-f3.tif
Fig. 3 The asymmetric unit (a), 1D [C6H6NO2]+ chain (b), 2D HPFs (c) along c direction, and overall structure along b direction (d).

We conducted a detailed analysis to discern the differences in π–π and dipole–dipole interactions stemming from the dimensional variation of hydrogen-bonded phosphate frameworks (HPFs). In compound I, a π–π interaction along the 1D HPFs is evident, characterized by a plane-to-plane distance of 4.5654(15) Å and a dihedral angle of 15.10(13)° (Fig. S8). Moreover, adjacent 3AP+ cations bordering the 1D HPFs exhibit larger dihedral angles of 24.295°. In contrast, compound II displays a π–π interaction along one side of the 2D HPFs with a plane-to-plane distance of 5.0811(13) Å and a dihedral angle of 9.06(11)° (Fig. S9). Additionally, neighboring 3CP+ cations from distinct 2D [(3CP)(H2PO4)] layers are nearly antiparallel, with a dihedral angle of 0.12(11)° and a plane-to-plane distance of 5.1088(13) Å (Fig. S9). Notably, this π–π interactions act as the binding force for the stacking of the 2D neutral layers in II. Consequently, compound II demonstrates stronger dipole–dipole interactions, which may contribute to an enhanced birefringence performance.

Transmittance and bandgaps

Ultraviolet–visible (UV-Vis) absorption spectra indicate that compounds I and II both exhibit bandgaps of 4.24 eV, with cutoff wavelengths at 265 nm and 270 nm, respectively (Fig. 4a and b). Electronic structure calculations substantiate that the two compounds possess indirect and direct bandgaps, with theoretical values of 2.279 eV and 1.919 eV, respectively (Fig. S10). To mitigate the systematic underestimation inherent in the Generalized Gradient Approximation (GGA) method when calculating optical properties, scissor operators of 1.921 eV and 2.281 eV are applied. Density of states (DOS) analysis reveals that for both compounds, the top of the valence band predominantly originates from O-2p orbitals, while the conduction band minimum is largely composed of C-2p orbitals, with a minor contribution from N-2p and O-2p orbitals (Fig. 4c and d). This suggests that the bandgaps are mainly shaped by the organic constituents, a finding consistent with previous research on semiorganic phosphates.39–41
image file: d4qi01220h-f4.tif
Fig. 4 UV–vis diffuse reflectance spectra (the inset show the bandgaps and F(R) is absorption coefficient/scattering coefficient.) for I (a) and II (b), as well as the density of states for I (c) and II (d).

SHG property

Given that compound I crystallizes in a polar space group, we assessed its SHG effect. Upon 1064 nm laser excitation, the SHG intensity of I was found to be on par with that of KH2PO4 (KDP), utilizing type I phase matching (Fig. 5). Compound I demonstrates a superior SHG response compared to certain reported semiorganic phosphates, such as (C3H7N6)6(H2PO4)4(HPO4)·4H2O (0.1 × KDP)40 and (C5H12NO)H2PO4 (0.15 × KDP),58 and is comparable to [C(NH2)3]2HPO4·H2O (1.2 × KDP)59 and [C(NH2)3]2PO3F (1.0 × KDP).60 However, it underperforms in SHG effect relative to [C(NH2)3]PO4·2H2O (1.5 × KDP),38 [C(NH2)3]6(PO4)·3H2O (3.8 × KDP),39 and (C5H6ON)+(H2PO4) (3.0 × KDP).41 Additionally, theoretical NLO coefficients for compound I were calculated, identifying five non-zero independent tensor components: d31 = 0.262 pm V−1, d15 = 0.262 pm V−1, d32 = 0.087 pm V−1, d24 = 0.087 pm V−1, and d33 = −0.426 pm V−1. Since the space group of II is Fdd2, in the mm2 point group, its effective tensor should be d31, which is comparable with the measured result.61–63
image file: d4qi01220h-f5.tif
Fig. 5 Phase-matching curve of I with 1064 nm laser radiation (insert is oscilloscope traces of the SHG signals for powders of I and KDP (150–210 μm) with 1064 nm laser radiation).

Birefringence property

Using a polarization microscope equipped with a Berek compensator, we measured the optical path difference (R). This measurement is required to determine the transition of the crystals from orthogonal polarization to complete extinction. By applying the formula R = Δn × T, where Δn represents the birefringence and T the sample thickness, we obtained the experimental birefringence values for compounds I and II (Fig. 6). Using a 546 nm laser source, we observed optical path difference of 1007.41 nm for I and 1139.46 nm for II, corresponding to sample thicknesses of 5.13 μm and 4.01 μm, respectively (Fig. S11). The resulting birefringence (Δnexp.) were measured as 0.196 at 546 nm for I and 0.284 at 546 nm for II. The enhanced birefringence of II is attributed to the intensified π–π and dipole–dipole interactions stemming from the smaller dihedral angles between neighboring 3CP+ cations within its structure. Notably, the birefringence exhibited by I and II significantly exceeds the experimental values of recently reported semiorganic phosphates, such as [C(NH2)3]3PO4·2H2O (0.055@546 nm),38 [C(NH2)3]6(PO4)·3H2O (0.078@546 nm),39 [C(NH2)3]2Sb3F3(HPO3)4 (0.027@546 nm),64 [C(NH2)3]SbFPO4·H2O (0.151@546 nm),64 NaIn(C2O4)(HPO4)(H2O)5 (0.098@546 nm),65 [Te(C6H5)2][PO3(OH)]n (0.133@550 nm)66 (Fig. 7 and Table S14). The increased birefringence of I and II is primarily due to the enhanced π-conjugation of the 3AP+ or 3CP+ cations, which introduces greater anisotropy compared to guanidinium or oxalate. Moreover, the experimental birefringence of II surpasses that of commercial birefringent crystals, including α-BaB2O4 (0.122@532 nm), CaCO3 (0.172@532 nm), and YVO4 (0.204@532 nm), highlighting its potential as a novel high-performance UV birefringent material.
image file: d4qi01220h-f6.tif
Fig. 6 Original crystal, crystals achieving complete extinction and triaxial ellipsoid of three refractive indices along crystallographic axes for I (a–c) and II (d–f).

image file: d4qi01220h-f7.tif
Fig. 7 Birefringence comparison between title compounds and reported phosphates or sulfates containing organic groups.

Subsequently, we computationally determined the dielectric function and refractive indices for compounds I and II, from which we calculated the theoretical birefringence (Δncal.) using the relationships ε(ω) = ε1(ω) + 2(ω) and n2(ω) = ε(ω). Both I (Fdd2) and II (Pbca) are characterized as biaxial crystals, with their refractive indices ordered as nX > nY > nZ. Specifically, for I and II, the refractive indices are arranged such that n[010] > n[001] > n[100]. To align with the standard crystallographic axes, we performed the transformations [010] → X, [001] → Y, and [100] → Z (Fig. 6 and S12). At a wavelength of 546 nm, the static refractive indices are as follows: for I, n[100] = 1.14, n[010] = 1.35, n[001] = 1.32; and for II, n[100] = 1.41, n[010] = 1.69, n[001] = 1.66. The calculated birefringence (Δncal.), derived using the formula Δncal. = nmaxnmin, is 0.209 at 546 nm for I and 0.284 at 546 nm for II, which are consistent with their experimental counterparts. Additionally, the calculated birefringence values at 1064 nm are 0.191 for I and 0.264 for II (Fig. S12). Importantly, the calculated birefringence of IIncal. = 0.286@532 nm and 0.264 nm@1064 nm) exceeds that of currently reported semiorganic phosphates, such as (C5H6ON)+(H2PO4) (0.25@1064 nm),41 [C(NH2)3]2PO3F (0.039@532 nm),60 (C3H7N6)6(H2PO4)4(HPO4)·4H2O (0.220@1064 nm),40 and (N2H6)[HPO3F]2 (0.077@1064 nm)67 as well as most semiorganic sulfates, including [C5H6O2N3][HSO4]·H2O (0.25@1064 nm),68 C(NH2)3SO3F (0.133@1064 nm)69 and Te(CS(NH2)2)4SO4·2H2O (0.210@546.1 nm)70 (Fig. 7 and Table S14).

To elucidate the mechanism behind the exceptional birefringence observed in compound II, we undertook a computational analysis of the Highest Occupied Molecular Orbital (HOMO) and Lowest Unoccupied Molecular Orbital (LUMO) distributions, complemented by electron localization function (ELF) mapping, as shown in Fig. S13 and Fig. 8. The HOMO is predominantly constituted by O-2p orbitals from the (H2PO4) group, while the LUMO is significantly influenced by the vacant π-conjugated orbitals of the 3CP+ cation, a finding that aligns with the DOS plots. The 2D ELF maps, depicted in Fig. 8a and b, are oriented parallel and perpendicular to the planar 3CP+ cation, respectively. These maps reveal significant variations in electron cloud density around the 3CP+ cation across different crystallographic planes, in contrast to the relatively uniform electron cloud density observed around the (H2PO4) anion. By synthesizing the information derived from the LUMO and ELF analyses, we infer that the in-plane and inter-plane anisotropy is predominantly due to the π-conjugated bonds within the 3CP+ organic ligands, which are aligned parallel to the (100) crystal plane. This alignment leads to a substantial difference in refractive indices, with nin-planenout-plane, consistent with the observed refractive index profile (n[010] ≈ n[001] ≫ n[100]). Therefore, the 3CP+ organic ligands are identified as the key contributors to the outstanding birefringence properties of compound II.


image file: d4qi01220h-f8.tif
Fig. 8 2D ELF sections projected on crystal planes parallel (a) to and perpendicular (b) to 3CP+ cation.

Conclusions

In summary, through the strategic modification of organic cations, we have successfully synthesized two novel semiorganic crystals: (C6H7N2O)+(H2PO4) (I) with 1D hydrogen-bonded phosphate frameworks (HPFs) and (C6H6NO2)+(H2PO4) (II) with 2D HPFs. Both compounds exhibit an expansive bandgap of 4.24 eV and remarkable birefringence (Δnexp. = 0.196 and 0.284@546 nm for I and II, respectively). These characteristics position them as prospective candidates for applications in UV birefringent crystals. Structural and computational analyses indicate that the 2D HPFs present in compound II augment dipole–dipole interactions and minimize the dihedral angles among the (C6H6NO2)+ cations, subsequently enhancing the birefringence. Given the structural diversity achievable with inorganic frameworks and the compositional flexibility inherent in organic ligands within hybrid organic–inorganic compounds, this study paves the way for further exploration and development of novel high-performance birefringent crystals.

Conflicts of interest

The authors declare no competing financial interest.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (No. 22205037 and 22031009) and Natural Science Foundation of Fujian Province (2023J01498).

References

  1. X.-H. Meng, W.-L. Yin and M.-J. Xia, Cyanurates consisting of intrinsic planar π-conjugated 6-membered rings: An emerging source of optical functional materials, Coord. Chem. Rev., 2021, 439, 213916 CrossRef CAS.
  2. A. Tudi, S. Han, Z. Yang and S. Pan, Potential optical functional crystals with large birefringence: Recent advances and future prospects, Coord. Chem. Rev., 2022, 459, 214380 CrossRef CAS.
  3. T. Wu, X. Jiang, K. Duanmu, C. Wu, Z. Lin, Z. Huang, M. G. Humphrey and C. Zhang, Giant Optical Anisotropy in a Covalent Molybdenum Tellurite via Oxyanion Polymerization, Adv. Sci., 2024, 11, 2306670 CrossRef CAS PubMed.
  4. P.-F. Li, C.-L. Hu, Y.-F. Li, J.-G. Mao and F. Kong, Hg4(Te2O5)(SO4): A Giant Birefringent Sulfate Crystal Triggered by a Highly Selective Cation, J. Am. Chem. Soc., 2024, 146, 7868–7874 CrossRef CAS PubMed.
  5. J.-H. Wu, C.-L. Hu, T.-K. Jiang, J.-G. Mao and F. Kong, Highly Birefringent d0 Transition Metal Fluoroantimonite in the Mid Infrared Band: Order–Disorder Regulation by Cationic Size, J. Am. Chem. Soc., 2023, 145, 24416–24424 CrossRef CAS PubMed.
  6. W.-F. Chen, J.-Y. Lu, J.-J. Li, Y.-Z. Lan, J.-W. Cheng and G.-Y. Yang, Sr2[B5O8(OH)]2·[B(OH)3]·H2O: A Strontium Borate That Shows Deep-Ultraviolet-Transparent Nonlinear Optical Properties, Chem. – Eur. J., 2024, 30, e202400739 CrossRef CAS PubMed.
  7. Y.-N. Zhang, Q.-F. Li, B.-B. Chen, Y.-Z. Lan, J.-W. Cheng and G.-Y. Yang, Na3B6O10(HCOO): an ultraviolet nonlinear optical sodium borate-formate, Inorg. Chem. Front., 2022, 9, 5032–5038 RSC.
  8. Y. Shang, J. Xu, H. Sha, Z. Wang, C. He, R. Su, X. Yang and X. Long, Nonlinear optical inorganic sulfates: The improvement of the phase matching ability driven by the structural modulation, Coord. Chem. Rev., 2023, 494, 215345 CrossRef CAS.
  9. 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.
  10. H. Fan, N. Ye and M. Luo, New Functional Groups Design toward High Performance Ultraviolet Nonlinear Optical Materials, Acc. Chem. Res., 2023, 56, 3099–3109 CrossRef CAS PubMed.
  11. Y. Li, J. Luo and S. Zhao, Local Polarity-Induced Assembly of Second-Order Nonlinear Optical Materials, Acc. Chem. Res., 2022, 55, 3460–3469 CrossRef CAS PubMed.
  12. M. Mutailipu, Z. Yang and S. Pan, Toward the Enhancement of Critical Performance for Deep-Ultraviolet Frequency-Doubling Crystals Utilizing Covalent Tetrahedra, Acc. Mater. Res., 2021, 2, 282–291 CrossRef CAS.
  13. X. Chen, Y. Li, J. Luo and S. Zhao, Recent advances in non-π-conjugated nonlinear optical sulfates with deep-UV absorption edge, Chin. J. Struct. Chem., 2023, 42, 100044 CrossRef CAS.
  14. B. L. Wu, C. L. Hu, F. F. Mao, R. L. Tang and J. G. Mao, Highly Polarizable Hg2+ Induced a Strong Second Harmonic Generation Signal and Large Birefringence in LiHgPO4, J. Am. Chem. Soc., 2019, 141, 10188–10192 CrossRef CAS PubMed.
  15. S. Zhao, X. Yang, Y. Yang, X. Kuang, F. Lu, P. Shan, Z. Sun, Z. Lin, M. Hong and J. Luo, Non-Centrosymmetric RbNaMgP2O7 with Unprecedented Thermo-Induced Enhancement of Second Harmonic Generation, J. Am. Chem. Soc., 2018, 140, 1592–1595 CrossRef CAS PubMed.
  16. Z. Fang, W.-H. Ma, Q.-Y. Chen, X.-T. Zhu, X.-M. Zeng, P.-B. Li, Q.-F. Zhou, T.-T. Song and M.-H. Duan, From (NH4)3[Zr(PO4)2F] to (NH4)3[Sn2(PO4)2]Cl: the rational design of a tin-based short-wave ultraviolet phosphate with large optical anisotropy, Inorg. Chem. Front., 2024, 11, 1775–1780 RSC.
  17. T. Zheng, Q. Wang, J.-X. Ren, L.-L. Cao, L. Huang, D.-J. Gao, J. Bi and G.-H. Zou, Halogen regulation triggers structural transformation from centrosymmetric to noncentrosymmetric switches in tin phosphate halides Sn2PO4X (X = F, Cl), Inorg. Chem. Front., 2022, 9, 4705–4713 RSC.
  18. H.-Y. Wu, C.-L. Hu, M.-B. Xu, Q.-Q. Chen, N. Ma, X.-Y. Huang, K.-Z. Du and J. Chen, From H12C4N2CdI4 to H11C4N2CdI3: a highly polarizable CdNI3 tetrahedron induced a sharp enhancement of second harmonic generation response and birefringence, Chem. Sci., 2023, 14, 9533–9542 RSC.
  19. X.-H. Zhang, B.-P. Yang, J. Chen, C.-L. Hu, Z. Fang, Z. 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.
  20. M. Ji, C. Hu, Z. Fang, Y. Chen and J. Mao, Tin(II)-Induced Large Birefringence Enhancement in Metal Phosphates, Inorg. Chem., 2021, 60, 15744–15750 CrossRef CAS PubMed.
  21. J. Guo, A. Tudi, S. Han, Z. Yang and S. 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 PubMed.
  22. X. Lu, Z. Chen, X. Shi, Q. Jing and M. H. Lee, Two Pyrophosphates with Large Birefringences and Second-Harmonic Responses as Ultraviolet Nonlinear Optical Materials, Angew. Chem., Int. Ed., 2020, 59, 17648–17656 CrossRef CAS PubMed.
  23. Y. Deng, L. Huang, X. Dong, L. Wang, K. M. Ok, H. Zeng, Z. Lin and G. Zou, K2Sb(P2O7)F: Cairo Pentagonal Layer with Bifunctional Genes Reveal Optical Performance, Angew. Chem., Int. Ed., 2020, 59, 21151–21156 CrossRef CAS PubMed.
  24. T. Baiheti, S. Han, A. Tudi, Z. Yang and S. Pan, Polar polymorphism: α- and β-KCsWP2O9 nonlinear optical materials with a strong second harmonic generation response, J. Mater. Chem. C, 2020, 8, 11441–11448 RSC.
  25. T. Baiheti, S. Han, A. Tudi, Z. Yang and S. Pan, Alignment of Polar Moieties Leading to Strong Second Harmonic Response in KCsMoP2O9, Chem. Mater., 2020, 32, 3297–3303 CrossRef CAS.
  26. 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.
  27. X. Zhang, L. Kang, P. Gong, Z. Lin and Y. 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.
  28. L. Wu, H. Tian, C. Lin, X. Zhao, H. Fan, P. Dong, S. Yang, N. Ye and M. Luo, Optimized arrangement of non-π-conjugated PO3NH3 units leads to enhanced ultraviolet optical nonlinearity in NaPO3NH3, Inorg. Chem. Front., 2024, 11, 1145–1152 RSC.
  29. 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 PubMed.
  30. X. Song, Z. Du, B. Ahmed, Y. Li, Y. Zhou, Y. Song, W. Huang, J. Zheng, J. Luo and S. Zhao, A UV solar-blind nonlinear optical crystal with confined π-conjugated groups, Inorg. Chem. Front., 2023, 10, 5462–5467 RSC.
  31. F. Chen, F. Mo, H. Chen, M.-J. Lin and Y. Chen, KPO2(NHCONH2)2: A Promising Deep-Ultraviolet Nonlinear Optical Phosphate Containing Polar [PO2(NHCONH2)2] Tetrahedra, Chem. Mater., 2024, 36, 2985–2992 CrossRef CAS.
  32. Y. Zhou, X. Zhang, M. Hong, J. Luo and S. Zhao, Achieving effective balance between bandgap and birefringence by confining π-conjugation in an optically anisotropic crystal, Sci. Bull., 2022, 67, 2276–2279 CrossRef CAS PubMed.
  33. B.-H. Lei, Z. Yang, H. Yu, C. Cao, Z. Li, C. Hu, K. R. Poeppelmeier and S. Pan, Module-Guided Design Scheme for Deep-Ultraviolet Nonlinear Optical Materials, J. Am. Chem. Soc., 2018, 140, 10726–10733 CrossRef CAS PubMed.
  34. M. Mutailipu, J. Han, Z. Li, F. Li, J. Li, F. Zhang, X. Long, Z. Yang and S. 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.
  35. W.-Q. Huang, X. Zhang, Y.-Q. Li, Y. Zhou, X. Chen, X.-Q. Li, F.-F. Wu, M.-C. Hong, J.-H. Luo and S.-G. Zhao, A Hybrid Halide Perovskite Birefringent Crystal, Angew. Chem., Int. Ed., 2022, 61, e202202746 CrossRef CAS PubMed.
  36. Q. Q. Chen, C. L. Hu, J. Chen, Y. L. Li, B. X. Li and J. G. Mao, [o-C5H4NHOH]2[I7O18(OH)]·3H2O: An Organic-Inorganic Hybrid SHG Material Featuring an [I7O18(OH)] infinity 2 - Branched Polyiodate Chain, Angew. Chem., Int. Ed., 2021, 60, 17426–17429 CrossRef CAS PubMed.
  37. Z.-P. Zhang, X. Liu, X. Liu, Z.-W. Lu, X. Sui, B.-Y. Zhen, Z. Lin, L. Chen and L.-M. Wu, Driving Nonlinear Optical Activity with Dipolar 2-Aminopyrimidinium Cations in (C4H6N3)+(H2PO3), Chem. Mater., 2022, 34, 1976–1984 CrossRef CAS.
  38. X. Wen, C. Lin, M. Luo, H. Fan, K. Chen and N. Ye, [C(NH2)3]3PO4·2H2O: A new metal-free ultraviolet nonlinear optical phosphate with large birefringence and second-harmonic generation response, Sci. China Mater., 2021, 64, 2008–2016 CrossRef CAS.
  39. C. Wu, X. Jiang, Z. Wang, H. Sha, Z. Lin, Z. Huang, X. Long, M. G. Humphrey and C. Zhang, UV Solar-Blind-Region Phase-Matchable Optical Nonlinearity and Anisotropy in a π-Conjugated Cation-Containing Phosphate, Angew. Chem., Int. Ed., 2021, 60, 14806–14810 CrossRef CAS PubMed.
  40. S.-F. Li, L. Hu, Y. Ma, F.-F. Mao, J. Zheng, X.-D. Zhang and D. Yan, Noncentrosymmetric (C3H7N6)6(H2PO4)4(HPO4)·4H2O and Centrosymmetric (C3H7N6)2SO4·2H2O: Exploration of Acentric Structure by Combining Planar and Tetrahedral Motifs via Hydrogen Bonds, Inorg. Chem., 2022, 61, 10182–10189 CrossRef CAS PubMed.
  41. J. Lu, X. Liu, M. Zhao, X.-B. Deng, K.-X. Shi, Q.-R. Wu, L. Chen and L.-M. Wu, Discovery of NLO Semiorganic (C5H6ON)+(H2PO4): Dipole Moment Modulation and Superior Synergy in Solar-Blind UV Region, J. Am. Chem. Soc., 2021, 143, 3647–3654 CrossRef CAS PubMed.
  42. R. H. Blessing, An empirical correction for absorption anisotropy, Acta Crystallogr., Sect. A: Found. Crystallogr., 1995, 51, 33–38 CrossRef PubMed.
  43. G. M. Sheldrick, Crystal structure refinement with SHELXL, Acta Crystallogr., Sect. C: Struct. Chem., 2015, 71, 3–8 Search PubMed.
  44. A. Spek, Single-crystal structure validation with the program PLATON, J. Appl. Crystallogr., 2003, 36, 7–13 CrossRef CAS.
  45. P. Kubelka and F. Munk, An article on optics of paint layers, Z. Tech. Phys., 1931, 12, 259–274 Search PubMed.
  46. S. K. Kurtz and T. T. Perry, A Powder Technique for the Evaluation of Nonlinear Optical Materials, J. Appl. Phys., 1968, 39, 3798–3813 CrossRef CAS.
  47. L. Cao, G. Peng, W. Liao, T. Yan, X. Long and N. Ye, A microcrystal method for the measurement of birefringence, CrystEngComm, 2020, 22, 1956–1961 RSC.
  48. M. D. Segall, P. J. D. Lindan, M. J. Probert, C. J. Pickard, P. J. Hasnip, S. J. Clark and M. C. Payne, First-principles simulation: ideas, illustrations and the CASTEP code, J. Phys.: Condens. Matter, 2002, 14, 2717 CrossRef CAS.
  49. V. Milman, B. Winkler, J. A. White, C. J. Pickard, M. C. Payne, E. V. Akhmatskaya and R. H. Nobes, Electronic structure, properties, and phase stability of inorganic crystals: A pseudopotential plane-wave study, J. Quantum Chem., 2000, 77, 895–910 CrossRef CAS.
  50. J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  51. J. S. Lin, A. Qteish, M. C. Payne and V. Heine, Optimized and transferable nonlocal separable ab initio pseudopotentials, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 4174 CrossRef PubMed.
  52. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci and G. A. Petersson, Gaussian 09, Revision B.01, 2010 Search PubMed.
  53. T. Lu and F. Chen, Multiwfn: A multifunctional wavefunction analyzer, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed.
  54. A. Alparone, Linear and nonlinear optical properties of nucleic acid bases, Chem. Phys., 2013, 410, 90–98 CrossRef CAS.
  55. S. Mahesh, K. C. Tang and M. Raj, Amide Bond Activation of Biological Molecules, Molecules, 2018, 23, 2615 CrossRef PubMed.
  56. G. Li, P. Lei and M. Szostak, Transition-Metal-Free Esterification of Amides via Selective N-C Cleavage under Mild Conditions, Org. Lett., 2018, 20, 5622–5625 CrossRef CAS PubMed.
  57. J. I. Mujika, J. M. Mercero and X. Lopez, Water-Promoted Hydrolysis of a Highly Twisted Amide:Rate Acceleration Caused by the Twist of the Amide Bond, J. Am. Chem. Soc., 2005, 127, 4445–4453 CrossRef CAS PubMed.
  58. D. Wang, X. Meng, N. Zhang, D. Sun, R. Hou, H. Chen and X. Liu, Characteristics of structure, thermal stability and optical properties for a novel NLO crystal (C5H12NO)H2PO4, Opt. Mater., 2023, 135, 113225 CrossRef CAS.
  59. I. Němec, I. Matulková, P. Held, J. Kroupa, P. Němec, D. Li, L. Bohatý and P. Becker, Crystal growth, crystal structure, vibrational spectroscopy, linear and nonlinear optical properties of guanidinium phosphates, Opt. Mater., 2017, 69, 420–431 CrossRef.
  60. 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.
  61. V. G. Dmitriev and D. N. Nikogosyan, Effective nonlinearity coefficients for three-wave interactions in biaxial crystal of mm2 point group symmetry, Opt. Commun., 1993, 95, 173–182 CrossRef CAS.
  62. Z. Chen, Z. Zhang, X. Dong, Y. Shi, Y. Liu and Q. Jing, Li3VO4: A Promising Mid-Infrared Nonlinear Optical Material with Large Laser Damage Threshold, Cryst. Growth Des., 2017, 17, 2792–2800 CrossRef CAS.
  63. T. Abudouwufu, M. Zhang, S.-C. Cheng, Z.-H. Yang and S.-L. Pan, Ce(IO3)2F2·H2O: The First Rare-Earth-Metal Iodate Fluoride with Large Second Harmonic Generation Response, Chem. – Eur. J., 2018, 25, 1221–1226 CrossRef PubMed.
  64. X. Dong, Y. Long, L. Huang, L. Cao, D. Gao, J. Bi and G. Zou, Large optical anisotropy differentiation induced by the anion-directed regulation of structures, Inorg. Chem. Front., 2022, 9, 6441–6447 RSC.
  65. X. Zhang, Y. Xue, L. Yang, C. Wu, Z. Huang, J. Xu, M. G. Humphrey and C. Zhang, NaIn(C2O4)(HPO4)(H2O)5: An Ultraviolet Transparent Nonlinear Optical Oxalate-Phosphate with Large Birefringence, SSRN, 2022, 253, 4056766 Search PubMed.
  66. M. Xue, L. Zhang, X. Wang, Q. Dong, Z. Zhu, X. Wang, Q. Gu, F. Kang, X.-X. Li and Q. Zhang, A Metal-Free Helical Covalent Inorganic Polymer: Preparation, Crystal Structure and Optical Properties, Angew. Chem., Int. Ed., 2024, 63, e202315338 CrossRef CAS PubMed.
  67. H. Qiu, F. Li, C. Jin, J. Lu, Z. Yang, S. Pan and M. Mutailipu, (N2H6)[HPO3F]2: maximizing the optical anisotropy of deep-ultraviolet fluorophosphates, Chem. Commun., 2022, 58, 5594–5597 RSC.
  68. L. Zhang, X. Zhang, F. Liang, Z. Hu and Y. Wu, Rational Design of Noncentrosymmetric Organic–Inorganic Hybrids with a π-Conjugated Pyridium-Type Cation for High Nonlinear-Optical Performance, Inorg. Chem., 2023, 62, 14518–14522 CrossRef CAS PubMed.
  69. M. Luo, C. Lin, D. Lin and N. Ye, Rational Design of the Metal-Free KBe2BO3F2 (KBBF) Family Member C(NH2)3SO3F with Ultraviolet Optical Nonlinearity, Angew. Chem., Int. Ed., 2020, 59, 15978–15981 CrossRef CAS PubMed.
  70. X. Weng, C. Lin, G. Peng, H. Fan, X. Zhao, K. Chen, M. Luo and N. Ye, Te(CS(NH2)2)4SO4·2H2O: A Three-in-One Semiorganic Nonlinear Optical Crystal with an Unusual Quadrilateral (TeS4)6− Chromophore, Cryst. Growth Des., 2021, 21, 2596–2601 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. CCDC 2343554 and 2343555 for (3AP) (H2PO4) (I) and (3CP) (H2PO4) (II). For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4qi01220h

This journal is © the Partner Organisations 2024
Click here to see how this site uses Cookies. View our privacy policy here.