Deep-ultraviolet nonlinear-optical crystals LiBePO4 and BeP2O6 synthesized by ionic potential modulation towards uniform arrangement of PO4 groups

Xia Zhanga, Jian-Qiang Shena, Hong-Xiao Lvb, Peng-Hui Guoa, Yi-Gang Chen*a, Chun-Li Hu*c and Xian-Ming Zhang*ab
aKey Laboratory of Magnetic Molecules and Magnetic Information Material of Ministry of Education, School of Chemistry and Material Science, Shanxi Normal University, Taiyuan 030031, China. E-mail: yg_chen80@sina.com; zhangxm@dns.sxnu.edu.cn
bCollege of Chemistry & Chemical Engineering, Key Laboratory of Interface Science and Engineering in Advanced Material, Ministry of Education, Taiyuan University of Technology, Taiyuan, Shanxi 030024, P. R. China
cState Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, China. E-mail: clhu@fjirsm.ac.cn

Received 17th March 2025 , Accepted 24th April 2025

First published on 28th April 2025


Abstract

The uniform arrangement of functional groups is a key factor in improving nonlinear properties in nonlinear-optical (NLO) materials, but currently there is no feasible and guiding strategy to modulate the uniform arrangement. Herein, we first apply the ionic potential concept to deep-ultraviolet (DUV, λ < 200 nm) NLO phosphates for a uniform arrangement of PO4 tetrahedral functional groups. Adopting Cs4LiBe4P7O24 with a non-uniform arrangement of PO4 as a structural model, by removing low ionic potential Cs+ and Li+ successively, two DUV NLO crystals LiBePO4 and BeP2O6 were synthesized. LiBePO4 features a [Be3P3O18] six-membered ring constructed by alternate connection of BeO4 and PO4, while BeP2O6 exhibits two kinds of [PO3] helical chains bridged by BeO4. Remarkably, the arrangement of the PO4 in LiBePO4 and BeP2O6 exhibits uniform evolution. As a result, LiBePO4 exhibits an enhanced second-harmonic-generation (SHG) effect up to 4.3 times that of Cs4LiBe4P7O24, while BeP2O6 shows an even more enhanced SHG effect, reaching 7.0 that of Cs4LiBe4P7O24 (2.1 × KDP). Moreover, BeP2O6 exhibits a short DUV absorption edge below 175 nm and the shortest SHG phase-matching output wavelength down to 211 nm. The universality of the new ionic potential modulation strategy is supported through analyzing known NLO materials containing alkali/alkaline-earth metal cations.


Introduction

The exploration of deep-ultraviolet (DUV, λ < 200 nm or band-gap > 6.2 eV) nonlinear-optical (NLO) crystals with large second-harmonic-generation (SHG) effects is of tremendous academic interest,1,2 as a large SHG effect is beneficial for high laser frequency-conversion efficiency, which can be widely applied in a series of optoelectronic fields, e.g., material micromachining, generation of entangled photon pairs, photoelectron spectroscopy, etc.3–6 In particular, those crystals with both a large SHG effect and short SHG phase-matching output wavelength (preferably in the short-wave UV (λ ≤ 266 nm), and even DUV region) are popular research areas because of essential applications in frontier technology. However, a large SHG effect and wide band-gap are naturally incompatible due to the fact that the SHG effect is inversely proportional to the band-gap in principle. Over the past few decades, a common strategy for achieving a large SHG effect has been to introduce SHG-active π-conjugated groups such as [BO3] and [CO3] to construct DUV NLO crystals, and this strategy has produced many crystals, such as KBe2BO3F2 (KBBF),7 NH4Be2BO3F2,8 β-Rb2Al2B2O7,9 NH4B4O6F,10 and LiZn(OH)CO3.11

Recent studies show that non-π-conjugated tetrahedral groups such as [SiO4], [PO4], and [SO4] with large HOMO–LUMO gaps can achieve high transmittance in the DUV region, and a series of tetrahedral DUV NLO crystals were thus synthesized, e.g. Li2SrSiO4,12 Rb6Si10O23,13 Ba3P3O10Cl,14 Cs6Mg6(PO3)18,15 RbNaMgP2O7,16 Ba5P6O20,17 (NH4)2Na3Li9(SO4)7,18 etc. However, most of these DUV NLO crystals exhibit a small SHG effect because the constituent tetrahedral groups show nearly nonpolar high symmetry and therefore exhibit weak SHG activity. In order to improve the SHG performance, some strategies have been proposed, mainly including the introduction of (i) second-order Jahn–Teller effect d0 cations (e.g. Mo6+, Ti4+) or ns2 lone-pair cations (e.g. Pb2+, Sn2+),19–22 (ii) anisotropic tetrahedral groups involving partially substituted oxygen atoms (e.g. [BO4−xFx], [PO4−xFx], [SO4−x(NH2)x]),23–26 and (iii) π-conjugated groups (e.g. [BO3]).27,28 However, the strategies all face seemingly insurmountable difficulties, such as an evident red-shift of the optical absorption edge, low thermal stability or returning to π-conjugated-characteristic NLO materials. Notably, that alkali/alkaline-earth metal cations with large HOMO–LUMO gaps have been only thought of as counterions for charge balance in these acentric structures, and their electrostatic interactions with surrounding tetrahedral groups are always overlooked. However, the interaction could affect the arrangement of tetrahedral groups, while a uniform arrangement can drive the generation of a large SHG effect according to anionic group theory.

In 2020, we reported two isostructural DUV NLO crystals M4LiBe4P7O24 (M = Cs, Rb) by combining alkali and alkaline-earth metal cations. Unfortunately, both crystals exhibit a weak SHG effect (∼0.3 × KDP).29 Structurally, the weak SHG effect is mainly caused by a non-uniform arrangement of PO4 groups (Fig. S1). Additionally, it is found that the characteristic differences between the three metal cations Cs+, Li+ and Be2+ in Cs4LiBe4P7O24 (as an example) are size and charge number. The influence of the size and charge number on the electrostatic interactions of the respective ions is consistent with the ionic potential concept. In this regard, we take Cs4LiBe4P7O24 as a structural model and put forward an ionic potential modulation strategy towards uniform arrangement of PO4. By removing low ionic potential Cs+ and Li+, two NLO-active crystals LiBePO4 and BeP2O6 were synthesized successively, and the arrangement of the PO4 exhibits uniform evolution. An increased SHG effect from Cs4LiBe4P7O24 to LiBePO4 to BeP2O6 is thus displayed. Notably, BeP2O6 exhibits a large SHG effect of 7.0 × Cs4LiBe4P7O24 (2.1 × KDP) and the shortest SHG phase-matching output down to the short-wave UV region of 211 nm. Furthermore, the universality of the ionic potential modulation strategy is supported through investigating known alkali/alkaline-earth metal NLO structures.

Experimental

Reagents

BeO (99.99%), NH4H2PO4 (99.9%), Li2CO3 (99.99%), MoO3 (99.9%) and Li2MoO4 (99.9%) were commercially available and used as received. Caution: due to the high toxicity of BeO upon inhalation, all experiments were performed with sufficient ventilation.

Synthesis

A polycrystalline sample of LiBePO4 was synthesized by a solid-state reaction method. Stoichiometric amounts of Li2CO3, BeO, and NH4H2PO4 were mixed, ground completely, put in a platinum crucible, and heated in a muffle furnace at 350 °C for 10 h. Then, the material was reground thoroughly, pressed into a pellet, and heated at 850 °C for 3d with several intermittent grindings.

Crystal growth

We used a high-temperature solution method to grow the crystals of LiBePO4 and BeP2O6. For LiBePO4, a mixture of LiBePO4 polycrystalline powder (0.5 mmol, 0.055 g), MoO3 (4 mmol, 0.576 g), and Li2MoO4 (3 mmol, 0.522 g) was melted at 850 °C, then quickly cooled to 750 °C and held for 10 h. Then, the melt was cooled to 550 °C at a rate of 3 °C h−1 and finally cooled to 350 °C at a rate of 20 °C h−1, and we switched off the furnace. Colorless millimeter-crystals were obtained after dissolving the material in water.

For BeP2O6, a mixture of Li2CO3 (0.5 mmol, 0.037 g), BeO (1 mmol, 0.025 g) and NH4H2PO4 (3 mmol, 0.345 g) was melted at 850 °C and held for 12 h. Then, the melt was cooled down to 400 °C at a rate of 1 °C h−1. Finally, the melt was cooled down to 350 °C at a rate of 10 °C h−1, and we switched off the furnace. Colorless millimeter-crystals were obtained after dissolving the material in water.

Single-crystal X-ray diffraction (XRD)

Single-crystal XRD data for LiBePO4 and BeP2O6 were collected at 293 K on a Bruker D8 Venture diffractometer equipped with a graphite monochromator using Mo Kα radiation (λ = 0.71073 Å). Empirical absorption correction using spherical harmonics was performed using the SCALE3 ABSPACK scaling algorithm. The structure was solved by direct methods with the program SHELXS and refined by the full-matrix least-squares program SHELXL.30 The structure was carefully checked using the program PLATON and no higher symmetries were found.31 The details of relevant crystallographic data are summarized in Table 1. The atomic coordinates and equivalent isotropic displacement parameters are listed in Table S1. The selected bond distances and angles are shown in Table S2.
Table 1 Crystal data and structure refinements for LiBePO4 and BeP2O6
a R1 = ∑||Fo|–|Fc||/∑|Fo|, and wR2 = [w(Fo2Fc2)2/w(Fo2)2]1/2.
Empirical formula LiBePO4 BeP2O6
Formula weight 221.84 166.95
Temperature/K 293(2) K 293(2) K
Crystal system Monoclinic Monoclinic
Space group Cc P21
a 16.331(1) 7.084(2)
b 9.151(1) 8.586(3)
c 15.965(1) 14.092(7)
α 90 90
β 111.62(1) 94.06(3)
γ 90 90
Volume/Å3 2218.34(9) 855.0(6)
Z 16 8
ρcalc/g cm−3 2.117 2.594
μ/mm−1 0.788 0.953
F(000) 1728 656
Goodness-of-fit on F2 1.03 1.036
R1, wR2 [I ≥ 2 σ(I)]a 0.0261, 0.0674 0.0328, 0.0793
R1, wR2 (all data)a 0.0316, 0.0717 0.0381, 0.0823
Flack parameter 0.02(3) −0.01(5)


Powder X-ray diffraction (XRD)

Powder XRD data of LiBePO4 and BeP2O6 were collected using a Rigaku MiniFlex600 powder diffractometer equipped with Cu Kα radiation (λ = 1.5418 Å) in the 2θ angular range of 5–80° with a scan step width of 0.01° at 293 K. These experimental XRD patterns were in good agreement with the calculated patterns (Fig. S2 and S3).

Element analysis

Element analysis was performed by using a PerkinElmer Optima 8000 DV inductively coupled plasma optical emission spectrometer (ICP-OES). Two crystal samples were dissolved in perchloric acid at the boiling point for 0.5 h. ICP-OES measurements give molar ratios of 1.08(Li)[thin space (1/6-em)]:[thin space (1/6-em)]0.95(Be)[thin space (1/6-em)]:[thin space (1/6-em)]1.00(P) for LiBePO4 and 1.00(Be)[thin space (1/6-em)]:[thin space (1/6-em)]2.11(P) for BeP2O6, which are consistent with their formulae.

X-ray photoelectron spectroscopy (XPS)

The chemical states and surface compositions of LiBePO4 and BeP2O6 were analyzed by a Thermo Scientific K-Alpha+ X-ray photoelectron spectrometer with a monochromatic Al Kα X-ray source (1486.6 eV) operating at 72 W (12 kV, 6 mA). All of the XPS spectra underwent background subtraction and were fitted using mixed Gaussian-Lorentzian peak shapes.

As seen in Fig. S4, XPS spectra of LiBePO4 and BeP2O6 demonstrate the presence of Li (for LiBePO4), Be, P, and O (Fig. S4a and f). For LiBePO4, a peak is observed at 55.38 eV for Li 1s, ascribed to the Li(I) oxidation state (Fig. S4b). The Be 1s XPS spectrum (Fig. S4c) shows a peak at 113.50 eV, which is attributed to the Be(II) oxidation state. The peak at 133.46 eV (Fig. S4d) corresponds to the P 2p binding energy for the +5 oxidation state. The O 1s XPS spectrum exhibits a peak at 531.54 eV (O2−) (Fig. S4e). For BeP2O6, the Be 1s XPS spectrum (Fig. S4g) shows a peak at 114.20 eV, which is attributed to the Be(II) oxidation state. The peak at 133.56 eV (Fig. S4h) corresponds to the P 2p binding energy for the +5 oxidation state. The O 1s XPS spectrum exhibits a peak at 531.74 eV (O2−) (Fig. S4i).

Thermal behavior

Thermogravimetry (TG) and differential scanning calorimetry (DSC) analysis of LiBePO4 and BeP2O6 were carried out on a NETZSCH STA 449 F5 Jupiter simultaneous analysis system under flowing N2. About 15 mg of polycrystalline samples of LiBePO4 and BeP2O6 were separately placed in Pt crucibles, and heated from room temperature to 1300 °C (for LiBePO4) or 1100 °C (for BeP2O6) at a rate of 10 °C min−1 under a flowing N2 atmosphere.

TG and DSC curves of LiBePO4 and BeP2O6 (Fig. S5) show that weight losses in the TG curves were not obviously observed up to 1300 °C for LiBePO4 and 1100 °C for BeP2O6. An endothermic peak at 1263 °C in the DSC curve for LiBePO4 was detected, while the endothermic peak of BeP2O6 was not observed until 1068 °C, indicating the excellent thermostability. Further, polycrystalline samples of LiBePO4 and BeP2O6 were separately calcined at 1300 °C and 1100 °C for 0.5 h under an air atmosphere. Powder XRD of the residues (Fig. S2 and S3) suggests that LiBePO4 samples before and after melting are consistent, showing that it is a congruently melting compound, while BeP2O6 samples before and after melting are inconsistent. According to the powder XRD pattern of the residue, BeP2O6 after melting is transformed into another known polymorph BeP2O6 and an unknown phase (Fig. S3b), demonstrating that BeP2O6 is an incongruently melting compound.

Infrared (IR) spectroscopy

Fourier transform infrared (FTIR) spectra of LiBePO4 and BeP2O6 were recorded from KBr pellets in the range 4000–400 cm−1 on a Nicolet Model 5DX spectrometer. As shown in Fig. S6, the IR spectrum of LiBePO4 indicates that the peak at 1055 cm−1 is attributed to the P–O asymmetric stretching vibration of PO4. The peaks at 789, 715, and 611 cm−1 are attributed to the P–O symmetric stretching vibrations of PO4. The peaks at 577 and 508 cm−1 are attributed to the P–O bending vibrations of PO4. The IR spectrum of BeP2O6 indicates that the peaks at 1287, 1174, 1109, 1030, and 995 cm−1 are attributed to the P–O asymmetric stretching vibrations of PO4. The peak at 735 cm−1 is attributed to the P–O symmetric stretching vibration of PO4. The peaks at 518 and 438 cm−1 are attributed to the P–O bending vibrations of PO4. The IR spectra further indicate the presence of P–O units of LiBePO4 and BeP2O6.

SHG measurement

Powder SHG measurements of LiBePO4 and BeP2O6 were performed by the Kurtz–Perry method at 298 K. The measurement was carried out using a Q-switched Nd:YAG solid-state laser at 1064 nm and 532 nm, and the intensity of the frequency-doubled output emitted from the samples was measured using a photomultiplier tube. Polycrystalline samples LiBePO4 and BeP2O6 were sieved into a series of distinct size ranges: 20–40 μm, 40–60 μm, 60–80 μm, 80–120 μm, 120–150 μm, 150–200 μm, and 200–300 μm. Polycrystalline KH2PO4 (KDP) and β-BaB2O4 (BBO) with the same particle size ranges were used as references. The SHG signal intensity of the LiBePO4/KDP and BeP2O6/KDP (or BBO) ratio was identified based on the same size (150–200 μm).

Optical transmission spectrum

Optical transmission spectra of LiBePO4 and BeP2O6 were measured using a Shimadzu SolidSpec-3700 DUV spectrometer in the 175–800 nm region. Two polished single-crystals with sizes of 2.0 × 1.2 × 0.3 mm3 for LiBePO4 and 1.8 × 1.1 × 0.3 mm3 for BeP2O6 were separately used to perform the measurements.

Birefringence measurement

Birefringences of LiBePO4 and BeP2O6 were measured on a CNOPTEC cross-polarizing microscope BK-POLR with an LED light filter. The LED light source has an optical wavelength of 590 ± 3 nm. The birefringence was calculated by using the following equation: R = d × |NeN0| = d × Δn, where R, d, and Δn represent the optical retardation, thickness of the crystal, and birefringence, respectively.

Interference colors of the selected LiBePO4 and BeP2O6 single-crystal samples are second-order yellow and second-order green based on a Michel–Levy diagram, and complete extinction can be achieved (Fig. S7a and S7b). The tested crystals have retardations (R) of 810 and 710 nm, with thicknesses (d) of 45.13 and 25.23 μm, respectively. Accordingly, the measured birefringences are 0.018 and 0.028 for LiBePO4 and BeP2O6, respectively.

Bond-valence-sum (BVS) calculation

BVS calculations for LiBePO4 and BeP2O6 were performed to further validate the structural model, and the bond valence sums were calculated using the formula Vi = ∑Sij = ∑exp[(r0rij)/B], where Sij is the bond valence associated with the bond length rij, and r0 and B (usually 0.37) are empirically determined parameters.

Computational method

Single-crystal structural data of LiBePO4 and BeP2O6 were used for the theoretical calculations. The electronic structures and optical properties were calculated using a plane-wave pseudopotential method within density functional theory (DFT) implemented in the total energy code CASTEP.32,33 Generalized gradient approximation (GGA) Perdew–Burke–Ernzerhof (PBE) was used as the exchange and correlation functional.34 The norm-conserving pseudopotential was applied to present the core–electron interactions.35 Li 2s1, Be 2s2, O 2s2 2p4, and p 3s2 2p3 orbital electrons were considered as valence electrons. A cutoff energy of 850 eV determined the number of plane waves included in the basis sets. Monkhorst–Pack k-point sampling of 2 × 1 × 4 was used to perform the numerical integration of the Brillouin zone. During the optical property calculations, about 540 empty bands were set to ensure the convergence of SHG coefficients. Birefringence is calculated by subtracting the minimum refractive index from the maximum refractive index. The theoretical birefringences of LiBePO4 and BeP2O6 are 0.016 and 0.024, respectively, at 590 nm, which are close to the experimental results. The calculations of second-order NLO susceptibilities were based on length-gauge formalism within the independent particle approximation.36,37

Results and discussion

Crystal structures

LiBePO4 and BeP2O6 single-crystals were synthesized through high-temperature molten reactions, and both crystal structures were characterized by single-crystal XRD. LiBePO4 crystallizes in acentric and polar monoclinic space group Cc (no. 9), and contains eight crystallographically independent Li atoms, eight Be atoms and eight PO4 groups, all of which are localized on crystallographically general positions with site occupancies of 1. Basic tetrahedral BeO4 and PO4 show typical bond distances of 1.606(3)–1.645(3) Å for Be–O and 1.525(1)–1.542(1) Å for P–O bonds. The material features a [Be3P3O18] six-membered ring constructed by alternate connection of BeO4 and PO4. The six-membered rings are fused by shared oxygen to generate [BePO4] two-dimensional (2D) pseudo layers in the ac plane (Fig. 1a). The pseudo layers are interconnected through Be–O–P bonds to form a 3D structure with pores (Fig. 1b). It should be noted that c glide planes in the structure cause polarization cancellation along the b-axis and net polarization in the ac plane. Small four-coordinated Li+ fills in pores to balance charge (Fig. 1b). In comparison, in Cs4LiBe4P7O24, large high-coordinated Cs+ requires large coordination space that induces four-coordinated Li+ on large [LiBe4P7O24] twelve-membered rings of framework layers (Fig. S1).
image file: d5qi00779h-f1.tif
Fig. 1 Crystal structures of LiBePO4 and BeP2O6. LiBePO4: (a) [BePO4] 2D pseudo layer formed by [Be3P3O18] six-membered rings. (b) 3D stacking of [BePO4] pseudo layers viewed along the a-axis. BeP2O6: (c) two kinds of [PO3] chains with 21 helices along the b-axis. (d) 3D stacking of [PO3] chains bridged by BeO4 viewed along the b-axis.

BeP2O6 crystallizes in acentric and polar monoclinic space group P21 (no. 4). There are four crystallographically independent Be atoms and eight PO4 groups, all of which are localized on crystallographically general positions with Wyckoff letter 2a. These PO4 by corner-sharing generate two kinds of [PO3] helical chains with a 21 axis along the b-axis (Fig. 1c). Adjacent [PO3] chains are bridged by BeO4 to produce a 3D structure. Net polarization of the structure is along the b-axis, as deduced from its P21 space group. Viewed along the b-axis (Fig. 1d), two [PO3] chains both exhibit almost uniform stacking. In all BeO4 and PO4 unis, Be–O (1.579(6)–1.633(6) Å) and P–O (1.434(4)–1.603(3) Å) bond lengths both vary in reasonable ranges. Validity of both structure models is checked through BVS calculations and XPS measurements, revealing their constituent elements in rational oxide states (Table S1 and Fig. S4).

Ionic potential modulation model of PO4 arrangement

Arrangement changes of the PO4 from Cs4LiBe4P7O24 to LiBePO4 to BeP2O6 are related to the respective counter-balance cations Cs+, Li+ and Be2+. The Pauling electronegativity of these elements increases in turn from Cs (0.79) to Li (0.98) to Be (1.57). In Cs4LiBe4P7O24, Cs+ has the strongest ionic bonding character and thus is regarded as the counter-balance cation of the structure, and the remaining ions Li+ and Be2+ mainly act as local charge balancing cations in the structure. Similarly, in LiBePO4, Li+ is counter-balance cation of the structure and Be2+ mainly acts as a local charge balancing cation. The size and charge number are the main differences between these cations. Their effect on PO4 arrangement is studied using an ionic potential modulation model. According to the equation φ = z/r (φ, ionic potential; z, ionic charge number; r, ionic radius), z values of Cs+, Li+ and Be2+ are 1, 1 and 2, respectively, and r values (effective ionic radii) of these cations in Cs4LiBe4P7O24, LiBePO4 and BeP2O6 are 0.18, 0.059 and 0.027 nm, respectively. φLi+ and φBe2+ are calculated to be 16.95 and 74.07, 3.1 and 13.4 times greater, respectively, than 5.52 for φCs+. The impacts of the alkali/alkaline-earth metal cations on the respective PO4 groups are depicted in the radar maps in Fig. 2. As shown in Fig. 2, with increasing φ from φCs+ to φLi+ to φBe2+, the arrangement of these PO4 groups around the cations exhibits uniform evolution. The uniform evolution can bring the following advantages: (i) the number of PO4 groups per unit volume increases, which is confirmed by calculating the packing density (ρ) of PO4, showing that ρ[PO4] increases from 1.2 × 10−2 to 1.5 × 10−2 to 1.9 × 10−2 per Å3; (ii) favorable superposition of microscopic hyperpolarizability of PO4 that is beneficial for increasing the overall SHG effect is generated, and it is verified by calculating the dipole moment (μ) of PO4, revealing that μ[PO4] increases from 0, 0.7 × 10−3 to 4.5 × 10−3 Debye (D) Å−3. Essentially, ionic potential modulation is mainly due to electrostatic interactions between alkali/alkaline-earth metal cations and surrounding ions. The electrostatic interaction can be estimated using Coulomb's law based on a simple point-charge model, revealing that the electrostatic interactions between Li+ and P5+ in LiBePO4 and between Be2+ and P5+ in BeP2O6 are 1.8 and 3.7 times that between Cs+ and P5+ in Cs4LiBe4P7O24 (average distances of Cs and P, Li and P, and Be and P are 3.99, 3.00 and 2.93 Å according to Fig. 2). In addition, within the same structure, the ionic potential concept remains effective and applicable. In Cs4LiBe4P7O24 and LiBePO4, the arrangement uniformity of the PO4 surrounding Cs+/Li+/Be2+ and Li+/Be2+ is improved in turn (Fig. S8 and S9).
image file: d5qi00779h-f2.tif
Fig. 2 Ionic potential (φ) modulation model of PO4 arrangement. In the radar maps, the numbers and arrows represent the distance (Å) between the cations and dipole moment direction of PO4. Only PO4 tetrahedra are displayed for clarity.

SHG effect analysis

SHG effects based on the Kurtz–Perry method were measured at laser wavelengths 1064 and 532 nm, showing that the SHG effect of LiBePO4 is 1.3 × benchmark KDP (KH2PO4) at 1064 nm (Fig. 3a), and 4.3 times that of Cs4LiBe4P7O24. The SHG effect of BeP2O6 further increases to 2.1 × KDP (Fig. 3a), reaching 7.0 times that of Cs4LiBe4P7O24, and meanwhile, BeP2O6 exhibits an SHG effect of 0.4 × BBO (β-BaB2O4) at 532 nm (Fig. 3c). The large SHG effect of BeP2O6 is compared not only with reported non-π-conjugated tetrahedral DUV NLO materials, such as Ba2NaClP2O7 (0.9 × KDP),38 KMg(H2O)PO4 (1.14 × KDP),39 BPO4 (2.0 × KDP),40 Li2SrSiO4 (0.3 × KDP),13 Li3AlSiO5 (0.8 × KDP),41 Li2NaNH4(SO4)2 (1.1 × KDP),18 and LiKSO4 (1.2× KDP),42 but also with well-known π-conjugated NLO materials, such as KBBF (1.21 × KDP),7 LiB3O5 (LBO, 3.0 × KDP),43 CsB3O5 (CBO, 2.7 × KDP),44 and CsLiB6O10 (CLBO, 2.2 × KDP).45 Furthermore, LiBePO4 reveals phase-matchable behavior in the visible region, whereas BeP2O6 exhibits phase-matchable behavior in both visible and ultraviolet (UV) regions, as shown by the increasing SHG intensities with increasing particle size (Fig. 3b and d).
image file: d5qi00779h-f3.tif
Fig. 3 (a) SHG signals and (b) SHG signal intensity vs. particle size for LiBePO4, BeP2O6, and benchmarks Cs4LiBe4P7O24 and KDP at 1064 nm. (c) SHG signals and (d) SHG signal intensity vs. particle size for BeP2O6 and benchmark BBO at 532 nm.

The phase-matching capacities of both structures are studied through experimental and theoretical methods. Calculated refractive index dispersions (Fig. 4a and b) reveal that the shortest SHG phase-matching output wavelengths are 317 and 211 nm for LiBePO4 and BeP2O6, respectively, consistent with the corresponding experimental observations of phase-matching in the 1064-to-532 nm and 532-to-266 nm second-harmonic processes (Fig. 3b and d), confirming their phase-matching capacity. In addition, calculated birefringences of LiBePO4 and BeP2O6 are 0.016 and 0.024, respectively, at 590 nm, and experimental birefringences determined using a cross-polarizing microscope at 590 ± 3 nm through an optical retardation method show that the tested birefringences are 0.018 for LiBePO4 and 0.028 for BeP2O6 (Fig. S7), in general agreement with the calculated results. Remarkably, the shortest SHG phase-matching output of BeP2O6 is down to the short-wave UV region of 211 nm, much better than non-phase-matching BPO4, 450 nm of LiCs2PO4,46 and well-known borates LBO (277 nm), CBO (273 nm) and CLBO (236 nm) (Table 2). BeP2O6 may be applied to frequency quadruple and quintuple the Nd:YAG laser output, producing 266 nm and 213 nm coherent light. In addition, optical transmission spectra show that LiBePO4 and BeP2O6 both have short DUV absorption edges below 175 nm (Fig. 4c and d), comparable to that of Cs4LiBe4P7O24 (<190 nm).29


image file: d5qi00779h-f4.tif
Fig. 4 Refractive index dispersions of (a) LiBePO4 and (b) BeP2O6. Single-crystal transmission spectra of (c) LiBePO4 and (d) BeP2O6, with single-crystal photographs in the insets.
Table 2 Comparison of DUV absorption edges (λDUV), SHG effects at 1064 nm, and SHG phase-matching cutoff wavelengths (λPM) of well-known crystals
Material λDUV (nm) SHG effect (×KDP) λPM (nm)
BPO4 134 2.0 Nonphase-matching
LiCs2PO4 174 2.6 450
LBO 158 3.0 277
CBO 167 2.7 273
CLBO 180 2.2 236
BeP2O6 <175 2.1 211


Theoretical calculation

To gain deep insight into the origin of the good optical properties of LiBePO4 and BeP2O6, first-principles calculations were implemented. Band structures calculated using the hybrid exchange–correlation functional of HSE06 (Fig. 5a and c) indicate that LiBePO4 and BeP2O6 have very wide band-gaps of 8.33 and 8.20 eV, corresponding to DUV absorption edges of 149 and 151 nm, respectively, in good support of the experimental results. Partial densities of states (Fig. 5b and d) show that the upper part of the valence-band consists mainly of P 3p, O 2p, and Be 2p electronic states, and the bottom part of the conduction-band is mainly composed of unoccupied P 3s 3p, O 2p, Be 2s 2p, and Li 2s (for LiBePO4) orbitals, revealing that all the chemical bonding interactions (including P–O, Be–O and Li–O) make contributions to the optical transitions of LiBePO4 and BeP2O6.
image file: d5qi00779h-f5.tif
Fig. 5 Electronic band structures of (a) LiBePO4 and (c) BeP2O6. Density of states diagrams for (b) LiBePO4 and (d) BeP2O6.

Calculated SHG coefficients show that the maximum tensors d15 for LiBePO4 and d23 for BeP2O6 are 0.42 and 0.57 pm V−1, respectively, close to experimentally measured SHG effects. To intuitively determine the main SHG-contributed electronic orbitals, SHG-density analyses for both maximum SHG tensors were carried out. For LiBePO4 and BeP2O6, in the valence-band (Fig. 6a and c), the dominant SHG-contributed orbitals are O 2p nonbonding states, while in the conduction-band (Fig. 6b and d), unoccupied P 3s 3p and O 2p orbitals have a crucial impact on SHG effects, while the unoccupied Be 2p orbitals make a small contribution to the SHG effects. Furthermore, SHG contributions of different groups are summed based on the SHG-density analysis results shown in Fig. 6e, showing that PO4 is in both cases the leading group producing SHG effects, accounting for 62.1% in LiBePO4 and 83.4% in BeP2O6. The BeO4 groups make small contributions to the SHG effects, accounting for 25.6% in LiBePO4 and 16.6% in BeP2O6, respectively. The applicability of the ionic potential modulation strategy is investigated for known NLO structures containing alkali/alkaline-earth metal cations (Fig. 6f and Table S3), and it is found that the strategy can be applied to these NLO materials well.


image file: d5qi00779h-f6.tif
Fig. 6 SHG density maps of the valence band (VB) and conduction band (CB) for (a and b) LiBePO4 and (c and d) BeP2O6. (e) SHG contribution of different groups in LiBePO4 and BeP2O6. (f) Applicability of the ionic potential modulation strategy in known alkali/alkaline-earth metal NLO materials. Blue represents cations with low ionic potential and their corresponding compounds, and red represents cations with high ionic potential and their corresponding compounds. a[thin space (1/6-em)]Represents the SHG effect based on benchmark AGS (AgGaS2).

Conclusions

An ionic potential modulation strategy is applied to tetrahedral DUV NLO phosphates that have long suffered from a weak SHG effect. Taking Cs4LiBe4P7O24 as a structural prototype, through continuous removal of low ionic potential Cs+ and Li+, two DUV NLO crystals LiBePO4 and BeP2O6 were successfully synthesized. LiBePO4 exhibits a 3D structure constructed from [Be3P3O18] six-membered rings composed of alternate BeO4 and PO4, while BeP2O6 features a 3D structure constructed from two kinds of [PO3] helical chains bridged by BeO4. Notably, the arrangement of PO4 in LiBePO4 and BeP2O6 exhibits uniform evolution. Thus, an increased SHG effect from Cs4LiBe4P7O24 to LiBePO4 to BeP2O6 is generated. Furthermore, BeP2O6 exhibits a large SHG effect of 7.0 × Cs4LiBe4P7O24 (2.1 × KDP), the shortest SHG phase-matching output down to the short-wave UV region of 211 nm, and a short DUV absorption edge below 175 nm, showing that BeP2O6 is a potential short-wave UV NLO crystal. Moreover, the new modulation strategy is universally supported through known alkali/alkaline-earth metal NLO materials. This work may not only be used to update the possible roles of alkali/alkaline-earth metal cations in NLO materials but also provide new opportunities for the precise design of material structures and properties.

Author contributions

All authors have given approval to the final version of the manuscript.

Data availability

The detailed crystallographic information and ESI figures and tables can be found in the ESI. Crystallographic data are available via the Cambridge Crystallographic Data Centre (CCDC): 2384142 and 2384141 for LiBePO4 and BeP2O6. The data supporting this article have been included as part of the ESI.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the NSFC (21971155, 5207021329), Fundamental Research Program of Shanxi Province (202303021221153).

References

  1. E. Goulielmakis, M. Schultze, M. Hofstetter, V. S. Yakovlev, J. Gagnon, M. Uiberacker, A. L. Aquila, E. Gullikson, D. T. Attwood and R. Kienberger, Single-cycle nonlinear optics, Science, 2008, 320, 1614–1617 Search PubMed.
  2. P. S. Halasyamani and J. M. Rondinelli, The must-have and nice-to-have experimental and computational requirements for functional frequency doubling deep-UV crystals, Nat. Commun., 2018, 9, 2972 Search PubMed.
  3. C. Wu, X. X. Jiang, Y. L. Hu, C. B. Jiang, T. H. Wu, Z. S. Lin, Z. P. Huang, M. G. Humphrey and C. Zhang, A lanthanum ammonium sulfate double salt with a strong SHG response and wide deep–UV transparency, Angew. Chem., 2022, 134, e202115855 CrossRef.
  4. J. X. Zhang, Q. G. Yue, S. H. Zhou, X. T. Wu, H. Lin and Q. L. Zhu, Screening strategy identifies an overlooked deep–ultraviolet transparent nonlinear optical crystal, Angew. Chem., Int. Ed., 2024, 63, e202413276 Search PubMed.
  5. Z.-S. Feng, Z.-H. Kang, F.-G. Wu, J.-Y. Gao, Y. Jiang, H.-Z. Zhang, Y. M. Andreev, G. V. Lanskii, V. V. Atuchin and T. A. Gavrilova, SHG in doped GaSe: In crystals, Opt. Express, 2008, 16, 9978–9985 Search PubMed.
  6. B. M. Oxley, J. B. Cho, A. K. Iyer, M. J. Waters, J. He, N. C. Smith, C. Wolverton, V. Gopalan, J. M. Rondinelli and J. I. Jang, Heteroanionic control of exemplary second-harmonic generation and phase matchability in 1D LiAsS2−xSex, J. Am. Chem. Soc., 2022, 144, 13903–13912 Search PubMed.
  7. C. T. Chen, Y. B. Wang, Y. N. Xia, B. C. Wu, D. Y. Tang, K. Wu, Z. Wenrong, L. H. Yu and L. F. Mei, New development of nonlinear optical crystals for the ultraviolet region with molecular engineering approach, J. Appl. Phys., 1995, 77, 2268–2272 Search PubMed.
  8. G. Peng, N. Ye, Z. S. Lin, L. Kang, S. L. Pan, M. Zhang, C. S. Lin, X. F. Long, M. Luo and Y. Chen, NH4Be2BO3F2 and γ–Be2BO3F: Overcoming the layering habit in KBe2BO3F2 for the next–generation deep–ultraviolet nonlinear optical materials, Angew. Chem., 2018, 130, 9106–9110 Search PubMed.
  9. T. T. Tran, N. Z. Koocher, J. M. Rondinelli and P. S. Halasyamani, Beryllium–free β–Rb2Al2B2O7 as a possible deep–ultraviolet nonlinear optical material replacement for KBe2BO3F2, Angew. Chem., Int. Ed., 2017, 129, 3015–3019 Search PubMed.
  10. 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 Search PubMed.
  11. X. M. Liu, L. Kang, P. F. Gong and Z. S. Lin, LiZn(OH)CO3: A deep–ultraviolet nonlinear optical hydroxycarbonate designed from a diamond–like structure, Angew. Chem., Int. Ed., 2021, 60, 13574–13578 Search PubMed.
  12. H. P. Wu, B. B. Zhang, H. W. Yu, Z. G. Hu, J. Y. Wang, Y. C. Wu and P. S. Halasyamani, Designing silicates as deep–UV nonlinear optical (NLO) materials using edge–sharing tetrahedra, Angew. Chem., 2020, 132, 9007–9011 CrossRef.
  13. S. Z. Huang, Y. Yang, J. B. Chen, W. Q. Jin, S. C. Cheng, Z. H. Yang and S. L. Pan, “Removing center”–An effective structure design strategy for nonlinear optical crystals, Chem. Mater., 2022, 34, 2429–2438 CrossRef CAS.
  14. 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 PubMed.
  15. Y.-G. Chen, M.-L. Xing, P.-F. Liu, Y. Guo, N. Yang and X.-M. Zhang, Two phosphates: Noncentrosymmetric Cs6Mg6(PO3)18 and centrosymmetric Cs2MgZn2(P2O7)2, Inorg. Chem., 2017, 56, 845–851 Search PubMed.
  16. S. G. Zhao, X. Y. Yang, Y. Yang, X. J. Kuang, F. Q. Lu, P. Shan, Z. H. Sun, Z. S. Lin, M. C. Hong and J. H. Luo, Non-centrosymmetric RbNaMgP2O7 with unprecedented thermo-induced enhancement of second harmonic generation, J. Am. Chem. Soc., 2018, 140, 1592–1595 Search PubMed.
  17. S. G. Zhao, P. F. Gong, S. Y. Luo, L. Bai, Z. S. Lin, Y. Y. Tang, Y. L. Zhou, M. C. Hong and J. H. Luo, Tailored synthesis of a nonlinear optical phosphate with a short absorption edge, Angew. Chem., Int. Ed., 2015, 54, 4217–4221 Search PubMed.
  18. T. Baiheti, S. J. Han, A. Tudi, Z. H. Yang and S. L. Pan, Alignment of polar moieties leading to strong second harmonic response in KCsMoP2O9, Chem. Mater., 2020, 32, 3297–3303 Search PubMed.
  19. Y. Q. Li, F. Liang, S. G. Zhao, L. N. Li, Z. Y. Wu, Q. R. Ding, S. Liu, Z. S. Lin, M. C. Hong and J. H. Luo, Two non-π-conjugated deep-UV nonlinear optical sulfates, J. Am. Chem. Soc., 2019, 141, 3833–3837 CrossRef CAS PubMed.
  20. T.-L. Chao, W.-J. Chang, S.-H. Wen, Y.-Q. Lin, B.-C. Chang and K.-H. Lii, Titanosilicates with strong phase-matched second harmonic generation responses, J. Am. Chem. Soc., 2016, 138, 9061–9064 CrossRef CAS PubMed.
  21. Y. Y. Yang, Y. Xiao, B. X. Li, Y.-G. Chen, P. H. Guo, B. B. Zhang and X.-M. Zhang, Stereochemically active lone-pair containing metal substitution in polar axis toward a giant phase-matchable optical nonlinear silicate crystal Li3(OH)PbSiO4, J. Am. Chem. Soc., 2023, 145, 22577–22583 CrossRef CAS PubMed.
  22. 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 Search PubMed.
  23. M. Mutailipu and S. L. Pan, Emergent deep–ultraviolet nonlinear optical candidates, Angew. Chem., Int. Ed., 2020, 59, 20302–20317 Search PubMed.
  24. C.-Y. Pan, X.-R. Yang, L. Xiong, Z.-W. Lu, B.-Y. Zhen, X. Sui, X.-B. Deng, L. Chen and L.-M. Wu, Solid-state nonlinear optical switch with the widest switching temperature range owing to its continuously tunable Tc, J. Am. Chem. Soc., 2020, 142, 6423–6431 Search PubMed.
  25. H. Tian, N. Ye and M. Luo, Sulfamide: A Promising deep–ultraviolet nonlinear optical crystal assembled from polar covalent [SO2(NH2)2] tetrahedra, Angew. Chem., Int. Ed., 2022, 61, e202200395 Search PubMed.
  26. X. F. Wang, X. Leng, Y. Kuk, J. Lee, Q. Jing and K. M. Ok, Deep–ultraviolet transparent mixed metal sulfamates with enhanced nonlinear optical properties and birefringence, Angew. Chem., 2024, 136, e202315434 Search PubMed.
  27. Z. Y. Zhou, Y. Qiu, F. Liang, L. S. Palatinus, M. Poupon, T. Yang, R. H. Cong, Z. S. Lin and J. L. Sun, CsSiB3O7: A beryllium-free deep-ultraviolet nonlinear optical material discovered by the combination of electron diffraction and first-principles calculations, Chem. Mater., 2018, 30, 2203–2207 Search PubMed.
  28. H. N. Liu, H. P. Wu, Z. G. Hu, J. Y. Wang, Y. C. Wu and H. W. Yu, Cs3 [(BOP)2(B3O7)3]: A deep-ultraviolet nonlinear optical crystal designed by optimizing matching of cation and anion groups, J. Am. Chem. Soc., 2023, 145, 12691–12700 Search PubMed.
  29. Y.-G. Chen, C. Y. Yang, F. Wang, Y. Guo, X. X. Jiang and X.-M. Zhang, M4LiBe4P7O24 and M4Li(Li3P)P7O24 (M = Cs, Rb): Deep-ultraviolet nonlinear-optical phosphates with a tetrahedra-substituted paracelsian-like framework, Chem. Commun., 2020, 56, 8639 Search PubMed.
  30. G. M. Sheldrick, A short history of SHELX, Acta Crystallogr. Sect. A: Found. Crystallogr., 2008, 64, 112–122 CrossRef CAS PubMed.
  31. A. L. Spek, Single-crystal structure validation with the program PLATON, Acta Crystallogr. Sect. A: Found. Crystallogr., 2003, 36, 7–13 CAS.
  32. 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, Int. J. Quantum Chem., 2000, 77, 895–910 Search PubMed.
  33. 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, 2717–2744 Search PubMed.
  34. J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77, 3865–3868 Search PubMed.
  35. J. S. Lin, A. Qteish, M. C. Payne and V. Heine, Optimized and transferable nonlocal separable ab initio pseudopoten tials, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 4174–4180 Search PubMed.
  36. J. E. Sipe and E. Ghahramani, Nonlinear optical response of semiconductors in the independent-particle approxi mation, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 48, 11705–11722 Search PubMed.
  37. S. Sharma, J. K. Dewhurst and C. Ambrosch-Draxl, Linear and second-order optical response of III-V monolayer superlattices, Phys. Rev. B: Condens. Matter Mater. Phys., 2003, 67, 165332 Search PubMed.
  38. J. Chen, L. Xiong, L. Chen and L.-M. Wu, Ba2NaClP2O7: Unprecedented phase matchability induced by symmetry breaking and its unique fresnoite-type structure, J. Am. Chem. Soc., 2018, 140, 14082–14086 Search PubMed.
  39. Z. Y. Bai, C.-L. Hu, L. H. Liu, L. Z. Zhang, Y. S. Huang, F. F. Yuan and Z. B. Lin, KMg(H2O)PO4: A deep-ultraviolet transparent nonlinear optical material derived from KTiOPO4, Chem. Mater., 2019, 31, 9540–9545 Search PubMed.
  40. Z. H. Li, Z. S. Lin, Y. C. Wu, P. Z. Fu, Z. Z. Wang and C. T. Chen, Crystal growth, optical properties measurement, and theoretical calculation of BPO4, Chem. Mater., 2004, 16, 2906–2908 CrossRef CAS.
  41. X. L. Chen, F. F. Zhang, L. L. Liu, B.-H. Lei, X. Y. Dong, Z. H. Yang, H. Y. Li and S. L. Pan, Li3AlSiO5: The first aluminosilicate as a potential deep-ultraviolet nonlinear optical crystal with the quaternary diamond-like structure, Phys. Chem. Chem. Phys., 2015, 18, 4362 Search PubMed.
  42. P. F. Zhu, P. F. Gong, Z. Y. Wang, H. T. Jiang, J. F. Zhao, C. Li, Z. S. Lin, X. L. Duan and F. P. Yu, Enlargement of bandgap and birefringence in nonlinear optical alkali-metal sulfate crystals by the substitution of asymmetrical non-π-conjugated cation, Adv. Opt. Mater., 2023, 11, 2301152 Search PubMed.
  43. C. T. Chen, Y. C. Wu, A. Jiang, B. Wu, G. You, R. Li and S. J. Lin, New nonlinear-optical crystal: LiB3O5, J. Opt. Soc. Am., 1989, 6, 616–621 Search PubMed.
  44. Y. C. Wu, T. Sasaki, S. Nakai, A. Yokotani, H. G. Tang and C. T. Chen, CsB3O5: A new nonlinear optical crystal, Appl. Phys. Lett., 1993, 62, 2614–2615 CrossRef CAS.
  45. J.-M. Tu and D. A. Keszler, CsLiB6O10: A noncentrosymmetric polyborate, Mater. Res. Bull., 1995, 30, 209–215 Search PubMed.
  46. L. Li, Y. Wang, B.-H. Lei, S. J. Han, Z. H. Yang, K. R. Poeppelmeier and S. L. Pan, A new deep-ultraviolet transparent orthophosphate LiCs2PO4 with large second harmonic generation response, J. Am. Chem. Soc., 2016, 138, 9101–9104 Search PubMed.

Footnote

Electronic supplementary information (ESI) available: Crystallographic data, XPS spectra, powder XRD, TG and DSC curves, IR spectra. CCDC 2384142 and 2384141 for LiBePO4 and BeP2O6. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5qi00779h

This journal is © the Partner Organisations 2025
Click here to see how this site uses Cookies. View our privacy policy here.