Next Article in Journal
Analysis of Position, Pose and Force Decoupling Characteristics of a 4-UPS/1-RPS Parallel Grinding Robot
Previous Article in Journal
Scalable Post-Processing of Large-Scale Numerical Simulations of Turbulent Fluid Flows
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Prediction of Large Second Harmonic Generation in the Metal-Oxide/Organic Hybrid Compound CuMoO3(p2c)

1
College of Chemistry, Fuzhou University, Fuzhou 350116, China
2
State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter (FJIRSM), Chinese Academy of Sciences (CAS), Fuzhou 350002, China
3
Fujian Science and Technology Innovation Laboratory for Optoelectronic Information of China, Fuzhou 350108, China
4
Department of Chemistry, North Carolina State University, Raleigh, NC 27695-8204, USA
5
College of Physics and Energy, Fujian Normal University, Fuzhou 350117, China
*
Authors to whom correspondence should be addressed.
Symmetry 2022, 14(4), 824; https://doi.org/10.3390/sym14040824
Submission received: 22 March 2022 / Revised: 8 April 2022 / Accepted: 11 April 2022 / Published: 14 April 2022
(This article belongs to the Section Chemistry: Symmetry/Asymmetry)

Abstract

:
Noncentrosymmetric hybrid framework (HF) materials are an important system in discovering new practical second-order nonlinear optical materials. We calculated the second harmonic generation (SHG) response of a noncentrosymmetric (NCS) organic–inorganic HF compound, CuMoO3(p2c) (p2c = pyrazine-2-carboxylate) to find that it exhibits the largest SHG response among all known NCS HF materials with one-dimensional helical chains. Further atom response theory analysis revealed that the metal atoms Cu and Mo contribute much more strongly than do nonmetal atoms in determining the strength of the SHG response, which is a novel example in nonlinear optical materials known to date.

1. Introduction

Nonlinear optical (NLO) crystals play a vital role in modern laser technologies and sciences due to their ability to convert the frequency of an incident laser beam through the second harmonic generation (SHG) process [1,2,3,4,5]. A noncentrosymmetric (NCS) arrangement of atoms is a prerequisite for the generation of non-zero second-order NLO properties in bulk materials. The search for new NCS structures with excellent SHG properties remains a hot scientific challenge. Although there exist a number of extensively studied and commercially available inorganic NLO crystals, there has been tremendous interest in finding new NLO materials in other systems such as organic molecular crystals, inorganic–organic hybrid nanocomposites, self-assembled chromophoric superlattices and inorganic–organic hybrid framework materials (including both nanoporous metal–organic frameworks (MOFs) and dense inorganic–organic frameworks (IOFs)) [6,7,8,9,10]. Hybrid framework materials are constructed by employing modular synthetic procedures with metal ions (i.e., Zn2+, Cd2+, Mn2+, Ag+, Cu+) or inorganic nodes (i.e., metal-oxide, metal-fluoride, metal-chalcogenides and metal-halides) that are covalently and/or coordinately bonded to various organic linkers. Because of the advantages over conventional inorganic materials (e.g., structural tunability at a molecular level, chemical stability, and ease of synthesis on a large bulk scale), NCS hybrid framework materials are expected to form a new class of potentiality practical NLO materials [11,12,13].
Numerous hybrid framework materials have been reported to be SHG-active. According to the dimensionality of the coordination networks, they can be roughly classified into three groups: (a) three-dimensional (3D) networks (e.g., diamondoid, octupolar), (b) two-dimensional (2D) grid structures, and (c) one-dimensional (1D) chains (e.g., helical chains). Lin et al. suggested that 3D and 2D networks with a high degree of controllability and predictability are better platforms for the synthesis of NCS MOFs [6,8], and they reported a number of MOFs with a large SHG response, which includes Cd{3-[2-(4-pyridyl)ethenyl]benzoate}2 (powder SHG intensity I powder 2 ω of 800 × α-SiO2, Fdd2) [14] and other systems [15]. For hybrid framework materials, the molecular origin of SHG is believed to arise from the electronic asymmetry (push–pull effect or acceptor–donor effect) induced by the polarization of ligand electron density or the metal-to-ligand/ligand-to-metal charge transfer [6,9,10,16]. The alignment of NLO chromophores may be necessary for NCS hybrid framework materials to have a large SHG response. For example, Ye et al. obtained a zinc MOF based on a 2D square grid network, in which the ligands involve an aligned two-center acceptor–donor system, which displays a very strong SHG response ( I powder 2 ω of 50 × urea and 500 × KH2PO4 (KDP), respectively) [16]. Yu et al. encapsulated a matching ordered organic dipolar chromophore in the pores of porous MOFs forming a highly NLO active anionic MOF (ZJU-28 DPASD, I powder 2 ω of 18.3 × α-SiO2) [17]. However, the NCS hybrid framework materials with 1D helical chains are not known to have large SHG responses, although it may be related to synthetic difficulty and unpredictability. Maggard et al. used short organic ligands to bridge polar or chiral inorganic basic building units (i.e., MoO2F42−) to obtain two isomorphic SHG-active NCS solids with helical chains [18,19], A(pyz)(H2O)2MoO2F4 (A = Zn, Cd, pyz = pyrazine, I powder 2 ω of ~0.28–1 × α-SiO2, P3221). Other examples with 1D helical chains include Cu(pzc)2AgReO4 (pzc = pyrazinecarboxylate, I powder 2 ω of ~0.5–0.7 × α-SiO2, P43212) [20], [Zn(mpz)3]2[MoO2F4]2 (mpz = 3-methylpyrazole, I powder 2 ω of ~10 × α-SiO2, Pna21) [21], and [(S)-C5H14N2][(MoO3)3(SO4)]·H2O ( I powder 2 ω of ~5 × α-SiO2, Pna21) [22]. The cancellation or the absence of a net dipole moment in these structures is usually considered to be responsible for their low SHG response [19]. Finding hybrid framework materials with 1D helical chains possessing a large SHG response remains a challenge. It should be pointed out that the SHG phenomenon, as a second-order polarization, involves both the occupied and the unoccupied states of a material, while the vector sum of the dipole moments of the groups within a unit cell, as a zero-order polarization, associated with only the occupied states. The non-zero vector sum of the dipole moments lead to a polar structure which guarantees the presence of an NCS structure, which is a necessary condition for the occurrence of the SHG phenomenon, but it does not influence the magnitude of the second-order NLO properties. It is desirable to understand the SHG responses for the NCS hybrid framework compounds at the electronic and the atomic levels.
Recently, Luo et al. [23] reported a new NCS metal-oxide/organic hybrid compound CuMoO3(p2c) (p2c = pyrazine-2-carboxylate) with 1D helical chains. However, so far neither theoretical nor experimental work on the linear and the nonlinear optical properties of this compound has been reported. Based on our first-principles calculation, we found CuMoO3(p2c) exhibits a very strong SHG response with a static effective SHG value d e f f p calculated to be 30.6 pm/V. This value is about 92.7 times that of KDP (~927 × α-SiO2, 3.12 × AgGaS2 (AGS)), which ranks the SHG response of CuMoO3(p2c) to be the highest among all the reported hybrid framework materials with 1D helical chains. Besides, our atom response theory (ART) analysis [24] shows that the SHG response of CuMoO3(p2c) originates largely from the states of the metal ions, Cu+ and Mo6+, i.e., approximately 56% of the total SHG response. Such a large contribution from metal ions also makes CuMoO3(p2c) a quite special example in NLO materials.

2. Materials and Methods

Computational Details

VASP calculations. The structural and the electronic properties of CuMoO3(p2c) were calculated within the framework of density functional theory (DFT) [25,26] by using the Vienna ab-initio simulation package (VASP) [27,28,29] with the projector augmented wave (PAW) method [30]. The generalized gradient approximation (GGA) within the Perdew–Burke–Ernzerhof (PBE)-type exchange-correlation potentials [31] was used throughout this work. The employed PAW–PBE pseudopotentials [32] with 11 (3d104s1), 14 (4s24p64d55s1), 6 (2s22p4), 5 (2s22p3), 4 (2s22p2) and 1 (1s1) valence electrons for Cu, Mo, O, N, C, and H were used to describe the electron−ion interactions, respectively. The plane wave cutoff energy for the expansion of wave functions was set at 600 eV and the tetrahedron method with Blöchl corrections was used for integrations. The numerical integrations in the Brillouin zones were performed by utilizing 7 × 7 × 4 Monkhorst–Pack k-point mesh, which showed an excellent convergence of the energy differences (0.005 eV) and stress tensors (0.001 eV/Å). The quasi-Newton algorithm as implemented in the VASP code was used in all structural relaxations. In this work, both the cell volume and the atomic positions were all allowed to relax to minimize the internal forces. The optimized lattice parameters, a = 7.791 Å and c = 11.229 Å, are slightly overestimated with respect to the experimental, which were measured at room temperature as usual with the PBE functional. Details of the optimized structure and the agreement with the experimental values are given in Table S1 of the Supporting Information.
In our calculations for the linear and the nonlinear optical properties, we employed the sum over states (SOS) method [33,34,35] using the code we developed [24] based on the calculated electronic structures from the VASP optical module. The SOS formalism for second-order susceptibility was derived by Aversa and Sipe [33] and later modified by Rashkeev et al. [34,36] and Sharma et al. [35,37]. To gain insight into the origin of SHG response, the contributions Aτ of the individual atoms τ to a specific component, e.g., the largest, of the total SHG response tensor, were determined by performing atom response theory (ART) analysis [24,38] for the CuMoO3(p2c) structure. Note that, the chirality may lead to higher-order interactions beyond the electric dipole approximation, for which one needs to use model studies instead of the first-principles method [39].
Partial response functional (PRF) method. The contribution of a certain occupied energy region between EB and valence band maximum (VBM), ζ V ( E B ) , to each SHG coefficient d i l = 1 2 χ i j k ( 2 ) ,   l = 1 ,   2 ,   3 ,   4 ,   5 ,   6 is determined by considering only those excitations from all occupied states between EB and VBM to all the unoccupied states of the conduction bands (CBs) and the contribution, δ ζ V ( E B ) , of specific occupied states of energy EB to each d i l by the excitations from that energy to all unoccupied states of the CBs.
δ ζ V ( E B ) = d ζ V ( E B ) d E B
Similarly, the contribution,   ζ C ( E B ) , of a certain unoccupied region between the conduction band minimum (CBM) and EB to each d i l is determined by the excitations from all occupied states of the VBs only to all unoccupied states between CBM and EB, and the contribution, δ ζ C ( E B ) , of specific unoccupied states of energy EB to each dil by the excitations from all occupied states of the VBs only to that energy.
δ ζ C ( E B ) = d ζ C ( E B ) d E B
Atom response theory (ART) analysis. To evaluate the individual atom contributions to the SHG components, d i l , it is computationally more convenient to express the corresponding PRFs in terms of the band index IB, ζ ( I B ) , where the band index I B runs from 1 to Ntot (i.e., the total number of band orbitals) with increasing energy, EB, from E m i n to E m a x . Here, ζ V ( I B ) and ζ C ( I B ) are denoted as ζ V B j and ζ C B j , respectively, with IB replaced by a subscript j.
Suppose that a specific atom τ has L atomic orbitals with a coefficient C V B L τ k j in the valence band j at a wave vector   k . The total contribution A V B τ of an atom τ makes to the SHG coefficient from all the VB bands j is written as
A V B τ = Ω ( 2 π ) 3   d k · L , j ζ V B j | C V B L τ k j | 2
where Ω is the unit cell volume, ζ V B j is the corresponding PRFs in terms of the band index j. Similarly, the total contribution A C B τ of an atom τ makes to the SHG coefficient from all the CB bands j is written as
A C B τ = Ω ( 2 π ) 3   d k · L , j ζ C B j | C C B L τ k j | 2
in which we assumed that the atom has L atomic orbitals with coefficient C C B L τ k j in the conduction band j at a wave vector k . To calculate the actual contribution of each constituent atom in a unit cell to the total SHG response, one needs to consider the signs of ζ V B j and ζ C B j .
The total contribution, A τ , each individual atom makes to the SHG response from both the VBs and the CBs (i.e., from all the bands) is given by
A τ = ( A V B τ + A C B τ ) 2
where the factor of 1/2 is applied to remove the double-counting of each excitation.

3. Results and Discussion

3.1. Structure

The Cu/Mo-oxide CuMoO3(p2c) crystallizes in the NCS space group P32 (Figure 1 and Table S1). Within each unit cell, Cu and Mo atoms each have one crystallographically unique site, whereas the O, N, C, and H atoms occupy five, two, five and three independent crystallographic positions, respectively. All atoms are at special Wyckoff positions of 3a. The Mo atoms are coordinated with five O atoms (O1, O3, O4, O4, O5) and one N1 atom, forming a distorted MoO5N octahedron with Mo-O and Mo-N bond lengths around 1.92 and 2.37 Å, respectively. Neighbouring Mo-centered octahedra are connected by sharing the O4 atom with alternating short (1.82 Å) and long (2.12 Å) Mo-O4 bond distances as well as alternating small ( Mo-O4-Mo = 145.64°) and large ( O4-Mo-O4 = 162.57°) bond angles forming 1D zigzag -Mo-O-Mo-O- chains along the c-axis (Figure 1a). Besides, this zigzag chain propagates along the 32 screw axis, that is, the MoO5N octahedron could duplicate itself by rotating 120° within the ab plane and gliding 2c/3 along the c axis. In addition, the Cu atoms form a distorted CuO3N tetrahedron with the surrounding three O atoms (O2, O3, O5) and one N2 atom with the Cu-O and Cu-N distances of ~2.03 Å and ~1.99 Å, respectively. Each CuO3N tetrahedra is corner shared with two neighbouring MoO5N octahedra through the O3 and O5 atoms, forming the inorganic CuMoO3 helical chains. Further, the N1 and O1 atoms from one p2c ligand simultaneously coordinate with the Mo atom, while the O2 and N2 atoms from two different p2c ligands connect the Cu atoms (Figure 1b). Thus, the 3D inorganic–organic hybrid network is constructed through the bridging p2c ligands. Note that, the Mo-O1 and the Cu-O2 distances (2.16 and 2.08 Å) connecting to the p2c ligands are slightly longer than the other Mo-O and Cu-O bonding lengths (1.86 and 2.01 Å) within each MoO5N octahedra and CuO3N tetrahedra, respectively. This fact reflects that the metal/ligand interaction is relatively weak compared with the interaction within the inorganic framework.

3.2. Electronic Structures

Our electronic calculations reveal that CuMoO3(p2c) has a small indirect bandgap (EgPBE) of 0.743 eV, which is smaller than the experimentally measured (Egexp = 1.32 eV). In computing optical properties, this deficiency of the DFT [26] is often corrected empirically by employing the scissor operation [40] in which the conduction bands (CBs) are shifted in energy to have the experimental bandgap [41,42]. The calculated density of states (DOS) and electronic band structure (Figure 2a and Figures S1–S4) show that the top portion of valence bands (VBs) (−1.0 eV to EF) is dominated by the Cu-3d orbitals, while the O-2p orbitals occupy the relatively lower energy range (−6.0 to −1.8 eV). The bottom part of the conduction bands (CBs) (Eg-7 eV) is primarily made up of the Mo-4d, C-2p, and N-2p states, while the antibonding states involving the 2p states of the O, C, N atoms are found in the energy range between 7 to 25 eV.
The crystal orbital Hamilton population (COHP) [43,44] analysis (Figure 2b) shows that the frontier orbital states (−4.4 eV to EF) are described mainly by nonbonding states made up of Cu-3d and O-2p orbitals with a small mixture of weak Cu-O/N d-p antibonding states (Figure S2). Generally, these filled nonbonding and antibonding states near EF are highly polarizable and hence are important for the optical properties. Besides, some O-2p and N-2p states make weak bonding interaction with Mo-4d and Cu-3d states as well as C-2p orbitals around −9.0 to −4.4 eV, leading to a dispersive orbitals feature. Strong Cu-O and Mo-O s-p bonding interaction can be found at around −18 eV. The covalent character of the C-O/N bonds is much stronger than that of the Mo-O/N and the Cu-O/N bonds (Figure 2b).

3.3. Optical Properties

Calculations of the refractive indices (no and ne) as a function of wavelength (Figure 3) reveal that CuMoO3(p2c) can meet the Type-I phase matching condition at λ = 1.88 μm, a value in the IR range. Besides, the calculated birefringence value Δn for CuMoO3(p2c) is 0.21 at 1910 nm (Figure S5). Such a large birefringence reflects a strong optical anisotropy of CuMoO3(p2c). The value of Δn in a uniaxial optical material is the difference between no and ne, Δn = |none|. As no > ne in CuMoO3(p2c), it is a negative uniaxial crystal.
Due to the point group of 3, the SHG tensor of CuMoO3(p2c) has 13 non-zero components, in which 5 of them are independent, i.e., d11 = −d12 = −d25, d22 = −d21 = −d16, d31 = d32 = d24 = d15, d14 = −d25 and d33, as presented in Table S2. As the Kleinman symmetry, i.e., d14 = −d25 = 0, is not followed in CuMoO3(p2c), it was not enforced in calculating the NLO properties in this work. The effective   d e f f p , an average SHG coefficient over all possible orientations of the powder crystals, is estimated from the formula derived by Kurtz-Perry [45] and Cyvin et al. [46] based on the calculated non-zero SHG tensors. The static effective SHG value of CuMoO3(p2c) is calculated to be 30.6 pm/V, which is about 3.12 times that of commercial AGS (static d eff AGS = 9.8 pm/V) and 92.7 times that of KDP (static d eff KDP = 0.33 pm/V). At the wavelength of 1910 nm (~0.65 eV), the d e f f p value of CuMoO3(p2c) is predicted to be 121.35 pm/V, i.e., ~8.20 × d eff AGS ( d eff AGS = 14.8 pm/V at 1910 nm). We have tried to estimate the deff for large crystals using Midwinter et al.’s method [47] and obtained a maximum value of ~168.7 pm/V at θ = 0.0° and φ = 45.0° at ω = 1910 nm, which is larger than the powder d e f f p . Using the deff and the refractive indices, one can get a much larger figure of merit value [48] ( F O M = d e f f 2 / ( n ( ω ) 2 n ( 2 ω ) ) , ~2058.7 (pm/V)2 than that 150 (pm/V)2, of LiNbO3, which suggests a great potential value of CuMoO3(p2c)—though the large crystal has not yet been obtained. We also calculated the SHG responses for several NCS hybrid framework compounds with helical chains, as presented in Table S2, which also lists the available experimental SHG values. Clearly, CuMoO3(p2c) has the highest d e f f p value among all reported NCS hybrid framework materials with helical chains. This remarkably strong SHG response is larger than those found in the majority of hybrid framework materials reported to date. Only a few are exceptional: [Cd(L-N3)2(H2O)2]n ( I powder 2 ω of 80 × urea) [49], Zn((E)-4-pyv-3-bza)2 ( I powder 2 ω of 1000 ×α-SiO2) [50], and [(CN4-C6H4-C12H7N-C5H4N)2Zn]·1.5H2O ( I powder 2 ω of 50 × urea) [16]. Besides, the SHG response of CuMoO3(p2c) is even larger than several newly reported inorganic IR-NLO crystals, e.g., Li[LiCs2Cl] [Ga3S6] ( I powder 2 ω of 0.7 × AGS) [51], Ba2SnS5 ( I powder 2 ω of 1.1 × AGS) [52], Na2MSn2Se6 (M = Zn/Cd) (( I powder 2 ω of 3 and 2.2 × AGS) [53], and δ-Ga2Se3 ( I powder 2 ω of 2.3 × AGS) [54]. These facts indicate that CuMoO3(p2c) is a promising IR hybrid NLO crystal material.

3.4. Atom Response Theory Analyses

We investigate the origin of the SHG responses further by employing the ART analysis [24]. Shown in Figure 2c is the partial response functionals (PRFs), ζ V ( E B ) and ζ C ( E B ) as well as their derivatives (Figures S6 and S7), δ ζ V ( E B ) and δ ζ C ( E B ) , for CuMoO3(p2c). The ζ V ( E B ) functional increases in magnitude with decreasing EB from EF to −4.4 eV indicating that the nonbonding Cu 3d and O 2p states are the dominating contributors to the SHG response in the VB part. From the rising amplitude of the ζ V ( E B ) functional, it is clear that the contribution from the Cu 3d states (EF to −1.0 eV) is greater than that from the O 2p states (−1.8 to −4.4 eV). This is so because the completely filled d orbitals of each Cu+ (d10) possess more electrons than the completely filled 2p orbitals of each O2− (p6). However, in the energy range of −4.4 to −9.0 eV where the C-O/N, Mo-O/N and Cu-O/N bonding interaction occurs, the functional shows small variation, showing that these bonding states contribute little to the SHG response. A dramatical change from the CB minimum to 7 eV and the steady increase at the higher energy range of the ζ C ( E B ) functional reveal the contribution from the unoccupied Mo-4d and C/O/N-2p states in the CB part.
The quantitative contribution of an individual atom τ, Aτ (in %), to the strongest SHG coefficient d11 for CuMoO3(p2c) is obtained based on the PRFs. As presented in Table 1 and Table S3, the Aτ of Cu (~11.7%) is nearly 1.7 times that of Mo (~7.0%) and 7.3 times that of O (~1.6%) and N (~1.2%). Besides, the Aτ value of C and H are negligibly small, <0.8%. These results reflect that the metal atoms Cu and Mo serve as the NLO-active centers at the atomic scale. Considering the number of atoms of each element in the unit cell, the total contributions of Cu, Mo, O, N, C, and H are 35.2, 20.9, 24.1, 7.4, 11.1, and 1.4%. Although the uneven stoichiometry effect is included, the metal atom Cu is still the leading contributor to the SHG response, while the contributions from the Mo and O atoms are comparable. This further reflects the important contribution from the metal atoms since the number of O in a unit cell is five times that of Cu and Mo. The relative atom contributions decrease in the order Cu >> Mo > O > N > C> H in the VB contributions (Table S3), and in the order Mo >> Cu > N > C > O> H in the CB contributions (Table S3). These findings show that the SHG of CuMoO3(p2c) is governed largely by the occupied states of Cu 3d, Mo 4p, and O 2p, and by the unoccupied states of Mo 4d, Cu 3d, and N-2p. The metal atoms Cu and Mo contribute much more strongly than do the nonmetal atoms in determining the strength of the SHG response in CuMoO3(p2c), which is quite special among the NLO materials known to date.
According to the individual atomic contribution to the SHG response, the contribution of an atomic group can be calculated by summing the contributions of the center atom and those of its ligands. In this work, we partition the contribution of an anion (O2− and N3−) equally to all the atomic groups it belongs to. The contribution of an atomic group can be calculated by summing the contributions of the center atom and those of its coordinated atoms [55]. Considering the coordination number for each O and N atom (2 and 3, respectively, Figure S8), the group CuO3N can be rewritten as CuO21/2O31/2O51/2N21/3. Therefore, the group contribution of CuO3N to SHG response is calculated as follows,
χ [ CuO 3 N ] ( 2 ) = χ Cu ( 2 ) + 1 × ( χ O 2 ( 2 ) 2 ) + 1 × ( χ O 3 ( 2 ) 2 ) + 1 × ( χ O 5 ( 2 ) 2 ) + 1 × ( χ N 2 ( 2 ) 3 )
The group with anion coordination numbers and the total group contributions for the largest SHG component d11 of CuMoO3(p2c) are given in Figure 4, which shows that the metal-centered group [CuO3N] and [MoO5N] contribute much more strongly to the SHG response than does the organic p2c ligand. That is, the total contribution of the inorganic part is ~79.1%, which far surpasses that of the organic part (i.e., ~20.9%). Our results reflect that the inorganic part contributes dominantly to the SHG response, while the organic part is important in the stabilization of the crystal structure of CuMoO3(p2c). It is worth mentioning that we did not separately calculate the hyperpolarizability tensor βijk by cutting out the ligand or groups from the structure of the compound. Such an approach will unavoidably lead to uncontrolled errors; thus, they are not used in our calculations.

4. Conclusions

Although a number of NCS HF materials with 3D and 2D frameworks have been reported with large SHG response, only low SHG intensities have been measured for HF materials with 1D helical chains. Our first-principles calculations predict that the recently synthesized CuMoO3(p2c) exhibits the largest SHG response among all the NCS HF materials with 1D helical chains. Its static effective SHG response, 30.6 pm/V, is about 3.12 times greater than that of commercial AGS, and 92.7 times greater than that of KDP. This value also exceeds those of most NCS hybrid framework materials reported so far. Our ART analysis shows that the SHG of CuMoO3(p2c) is determined largely by the occupied states composed of Cu 3d, Mo 4p, and O 2p, and by the unoccupied states composed of Mo 4d, Cu 3d, and N 2p. The metal atoms Cu and Mo contribute much more strongly than do the nonmetal atoms in determining the strength of the SHG response in CuMoO3(p2c). The latter is quite special in the NLO materials known to date. Our work based on the quantitative calculations at the electronic and the atomic level reveals the importance of the contribution from metal atoms and the metal-centered inorganic groups for the SHG response of CuMoO3(p2c).

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/sym14040824/s1, Figure S1: Calculated band structure for P32-CuMoO3(p2c). Figure S2: HOMO (M1) and LUMO (A1) for CuMoO3(p2c). Figure S3: Calculated partial DOS for P32-CuMoO3(p2c). Figure S4: Calculated partial DOS of O and N atoms at independent crystallographic positions. Figure S5: Frequency-dependent refractive indices n (left) and birefringence Δn (right) of CuMoO3(p2c). Figure S6: (a) δ ζ V ( E B ) -vs- E B plot, and (b) δ ζ C ( E B ) -vs-EB plot calculated for the SHG coefficient d11 of CuMoO3(p2c). The values of the functions are in pm/V. Figure S7: Plots of (a) ζ V ( I B ) -vs- I B , (b) ζ C ( I B ) -vs- I B , (c) δ ζ V ( I B ) -vs- I B and (d) δ ζ C ( I B ) -vs-IB calculated for the SHG coefficient d11 of CuMoO3(p2c). The values of the functions are in pm/V. Figure S8: Coordination environment for each inequivalent O and N atom of CuMoO3(p2c). Table S1: The optimized crystal structure data for P32-CuMoO3(p2c). Table S2: Calculated SHG tensors dil and d e f f p for CuMoO3(p2c) and several NCS HF compounds with helical chains. For the compounds with no available experimental band gap (Eg), the calculated Eg based on HSE06 with mixing parameter α = 0.3 is applied. The available experimental measured SHG responses are also presented. In this work, Kleinman symmetry is not enforced in calculating the NLO properties. Table S3: Contributions of the individual atoms to the SHG component d11 of CuMoO3(p2c). WA refers to the number of the same type of atoms (on the same Wyckoff site) in a unit cell. Aτ is the contribution (in %) from a single atom τ, and CA from all atoms of the same type. VBAτ, is the contribution (in %) the VBs, and CBAτ from the CBs. The contributions from the s, p, and d states of the atom τ to of VBAτ and A C B τ are also shown [56,57].

Author Contributions

Conceptualization, X.C., M.-H.W. and S.D.; methodology, X.C.; software, X.C. and C.L.; validation, X.C., T.Y. and X.H.; formal analysis, X.C., T.Y., X.H., M.-H.W. and S.D.; investigation, X.C., C.L. and P.A.M.; resources, P.A.M. and S.D.; data curation, X.C., T.Y and X.H; writing—original draft preparation, X.C.; writing—review and editing, X.C., T.Y., X.H., C.L., P.A.M., M.-H.W. and S.D.; visualization, X.C.; supervision, M.-H.W. and S.D.; project administration, M.-H.W. and S.D.; funding acquisition, X.C., P.A.M., M.-H.W. and S.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation (NSF) of China (22031009, 21921001, 22075282, 61874122); the National Key Research and Development Program of China (2021YFB3601501); the NSF of Fujian Province (2019J05151, 2019J01121); the Youth Innovation Promotion of CAS (2019302); the Fujian Science & Technology Innovation Laboratory for Optoelectronic Information of China (2021ZR123); President’s International Fellowship Initiative of CAS (2019VMA0049). P.A.M. acknowledges support from the National Science Foundation (DMR-2004455).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article and supplementary material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shen, Y.R. The Principles of Nonlinear Optics; Wiley-Interscience: New York, NY, USA, 1984. [Google Scholar]
  2. Chen, C.T.; Sasaki, T.; Li, R.; Wu, Y.; Lin, Z.; Mori, Y.; Kaneda, Y. Nonlinear Optical Borate Crystals; Wiley-VCH: Weinheim, Germany, 2012. [Google Scholar]
  3. Chung, I.; Kanatzidis, M.G. Metal Chalcogenides: A Rich Source of Nonlinear Optical Materials. Chem. Mater. 2014, 26, 849–869. [Google Scholar] [CrossRef]
  4. Hu, C.-L.; Mao, J.-G. Recent Advances on Second-Order NLO Materials Based on Metal Iodates. Coor. Chem. Rev. 2015, 288, 1–17. [Google Scholar] [CrossRef]
  5. Ok, K.M. Toward the Rational Design of Novel Noncentrosymmetric Materials: Factors Influencing the Framework Structures. Acc. Chem. Res. 2016, 49, 2774–2785. [Google Scholar] [CrossRef] [PubMed]
  6. Evans, O.R.; Lin, W. Crystal Engineering of NLO Materials Based on Metal-Organic Coordination Networks. Acc. Chem. Res. 2002, 35, 511–522. [Google Scholar] [CrossRef]
  7. Maury, O.; Bozec, H.L. Molecular Engineering of Octupolar NLO Molecules and Materials Based on Bipyridyl Metal Complexes. Acc. Chem. Res. 2005, 38, 691–704. [Google Scholar] [CrossRef]
  8. Wang, C.; Zhang, T.; Lin, W. Rational Synthesis of Noncentrosymmetric Metal-Organic Frameworks for Second-Order Nonlinear Optics. Chem. Rev. 2012, 112, 1084–1104. [Google Scholar] [CrossRef]
  9. Mingabudinova, L.R.; Vinogradov, V.V.; Milichko, V.A.; Hey-Hawkins, E.; Vinogradov, A.V. Metal-Organic Frameworks as Competitive Materials for Non-Linear Optics. Chem. Soc. Rev. 2016, 45, 5408–5431. [Google Scholar] [CrossRef] [Green Version]
  10. Medishetty, R.; Zareba, J.K.; Mayer, D.; Samoc, M.; Fischer, R.A. Nonlinear Optical Properties, Upconversion and Lasing in Metal-Organic Frameworks. Chem. Soc. Rev. 2017, 46, 4976–5004. [Google Scholar] [CrossRef] [Green Version]
  11. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef] [Green Version]
  12. Zhou, H.C.; Kitagawa, S. Metal-Organic Frameworks (MOFs). Chem. Soc. Rev. 2014, 43, 5415–5418. [Google Scholar] [CrossRef] [Green Version]
  13. Cui, Y.; Li, B.; He, H.; Zhou, W.; Chen, B.; Qian, G. Metal-Organic Frameworks as Platforms for Functional Materials. Acc. Chem. Res. 2016, 49, 483–493. [Google Scholar] [CrossRef] [PubMed]
  14. Evans, O.R.; Lin, W. Rational Design of Nonlinear Optical Materials Based on 2D Coordination Networks. Chem. Mater. 2001, 13, 3009–3017. [Google Scholar] [CrossRef]
  15. Lin, W.; Wang, Z.; Ma, L. A Novel Octupolar Metalorganic NLO Material Based on a Chiral 2D Coordination Network. J. Am. Chem. Soc. 1999, 121, 11249–11250. [Google Scholar] [CrossRef]
  16. Ye, Q.; Li, Y.H.; Song, Y.M.; Huang, X.F.; Xiong, R.G.; Xue, Z. A Second-Order Nonlinear Optical Material Prepared Through in situ Hydrothermal Ligand Synthesis. Inorg. Chem. 2005, 44, 3618–3625. [Google Scholar] [CrossRef] [PubMed]
  17. Yu, J.; Cui, Y.; Wu, C.; Yang, Y.; Wang, Z.; O’Keeffe, M.; Chen, B.; Qian, G. Second-Order Nonlinear Optical Activity Induced by Ordered Dipolar Chromophores Confined in the Pores of an Anionic Metal-Organic Framework. Angew. Chem. Int. Ed. 2012, 51, 10542–10545. [Google Scholar] [CrossRef]
  18. Maggard, P.A.; Stern, C.L.; Poeppelmeier, K.R. Understanding the Role of Helical Chains in the Formation of Noncentrosymmetric Solids. J. Am. Chem. Soc. 2001, 123, 7742–7743. [Google Scholar] [CrossRef]
  19. Maggard, P.A.; Kopf, A.L.; Stern, C.L.; Poeppelmeier, K.R. From Linear Inorganic Chains to Helices: Chirality in the M(pyz)(H2O)2MoO2F4 (M)=Zn, Cd) Compounds. Inorg. Chem. 2002, 41, 4852–4858. [Google Scholar] [CrossRef]
  20. Yan, B.; Capracotta, M.D.; Maggard, P.A. Structural Origin of Chirality and Properties of a Remarkable Helically Pillared Solid. Inorg. Chem. 2005, 44, 6509–6511. [Google Scholar] [CrossRef]
  21. Ahmed, B.; Jo, H.; Oh, S.J.; Ok, K.M. Variable Asymmetric Chains in Transition Metal Oxyfluorides: Structure-Second-Harmonic-Generation Property Relationships. Inorg. Chem. 2018, 57, 6702–6709. [Google Scholar] [CrossRef]
  22. Muller, E.A.; Cannon, R.J.; Sarjeant, A.N.; Ok, K.M.; Halasyamani, P.S.; Norquist, A.J. Directed Synthesis of Noncentrosymmetric Molybdates. Cryst. Growth Des. 2005, 5, 1913–1917. [Google Scholar] [CrossRef]
  23. Luo, L.; Chen, Y.; Maggard, P.A. Rare Example of Chiral and Achiral Polymorphs of a Metal-Oxide/Organic Hybrid Compound. J. Solid State Chem. 2020, 287, 121358. [Google Scholar] [CrossRef]
  24. Cheng, X.; Whangbo, M.-H.; Guo, G.-C.; Hong, M.; Deng, S. Large Second Harmonic Generation of LiCs2PO4 Caused by the Metal-Cation-Centered Groups. Angew. Chem. Int. Ed. 2018, 57, 3933–3937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef] [Green Version]
  26. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef] [Green Version]
  27. Kresse, G.; Hafner, J. Ab initio Molecular Dynamics for Liquid Metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef]
  28. Kresse, G.; Furthmüller, J. Efficiency of ab-initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mat. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  29. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for ab initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  30. Blöchl, P.E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [Green Version]
  31. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  32. Perdew, J.P.; Wang, Y. Accurate and Simple Analytic Representation of the Electron-Gas Correlation Energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef]
  33. Aversa, C.; Sipe, J.E. Nonlinear Optical Susceptibilities of Semiconductors: Results with a Length-Gauge Analysis. Phys. Rev. B 1995, 52, 14636–14645. [Google Scholar] [CrossRef] [PubMed]
  34. Rashkeev, S.N.; Lambrecht, W.R.; Segall, B. Efficient ab initio Method for the Calculation of Frequency-Dependent Second-Order Optical Response in Semiconductors. Phys. Rev. B 1998, 57, 3905–3919. [Google Scholar] [CrossRef]
  35. Sharma, S.; Ambrosch-Draxl, C. Second-Harmonic Optical Response from First Principles. Phys. Scripta 2004, T109, 128–134. [Google Scholar] [CrossRef] [Green Version]
  36. Rashkeev, S.N.; Lambrecht, W.R.L. Second-Harmonic Generation of I-III-VI2 Chalcopyrite Semiconductors: Effects of Chemical Substitutions. Phys. Rev. B 2001, 63, 165212. [Google Scholar] [CrossRef]
  37. Sharma, S.; Dewhurst, J.K.; Ambrosch-Draxl, C. Linear and Second-Order Optical Response of III-V Monolayer Superlattices. Phys. Rev. B 2003, 67, 165332. [Google Scholar] [CrossRef] [Green Version]
  38. Cheng, X.; Whangbo, M.-H.; Hong, M.; Deng, S. Atom Response Theory of Nonlinear Optical Responses and Its Applications. Chinese J. Struct. Chem. 2020, 39, 2172–2181. [Google Scholar] [CrossRef]
  39. Sioncke, S.; Verbiest, T.; Persoons, A. Second-Order Nonlinear Optical Properties of Chiral Materials. Mater. Sci. Eng. R 2003, 42, 115–155. [Google Scholar] [CrossRef]
  40. Gonze, X.; Lee, C. Dynamical Matrices, Born Effective Charges, Dielectric Permittivity Tensors, and Interatomic Force Constants from Density-Functional Perturbation Theory. Phys. Rev. B 1997, 55, 10355–10368. [Google Scholar] [CrossRef]
  41. Veithen, M.; Gonze, X.; Ghosez, P. Nonlinear Optical Susceptibilities, Raman Efficiencies, and Electro-Optic Tensors from First-Principles Density Functional Perturbation Theory. Phys. Rev. B 2005, 71, 125107. [Google Scholar] [CrossRef] [Green Version]
  42. Djani, H.; Hermet, P.; Ghosez, P. First-Principles Characterization of the P21ab Ferroelectric Phase of Aurivillius Bi2WO6. J. Phys. Chem. C 2014, 118, 13514–13524. [Google Scholar] [CrossRef]
  43. Dronskowski, R.; Bloechl, P.E. Crystal orbital Hamilton populations (COHP): Energy-resolved visualization of chemical bonding in solids based on density-functional calculations. J. Phys. Chem. 1993, 97, 8617–8624. [Google Scholar] [CrossRef]
  44. Deringer, V.L.; Tchougreeff, A.L.; Dronskowski, R. Crystal Orbital Hamilton Population (COHP) Analysis as Projected from Plane-Wave Basis Sets. J. Phys. Chem. A 2011, 115, 5461–5466. [Google Scholar] [CrossRef] [PubMed]
  45. Kurtz, S.K.; Perry, T.T. A Powder Technique for the Evaluation of Nonlinear Optical Materials. J. Appl. Phys. 1968, 39, 3798–3813. [Google Scholar] [CrossRef]
  46. Cyvin, S.J.; Rauch, J.E.; Decius, J.C. Theory of Hyper-Raman Effects (Nonlinear Inelastic Light Scattering): Selection Rules and Depolarization Ratios for the Second-Order Polarizability. J. Chem. Phys. 1965, 43, 4083–4095. [Google Scholar] [CrossRef]
  47. Midwinter, J.E.; Warner, J. The Effects of Phase Matching Method and of Uniaxial Crystal Symmetry on the Polar Distribution of Second-Order Non-Linear Optical Polarization. Br. J. Appl. Phys. 1965, 16, 1135–1142. [Google Scholar] [CrossRef]
  48. Astill, A.G. Material Figures of Merit for Non-Linear Optics. Thin Solid Films 1991, 204, 1–17. [Google Scholar] [CrossRef]
  49. Ye, Q.; Tang, Y.Z.; Wang, X.S.; Xiong, R.G. Strong Enhancement of Second-Harmonic Generation (SHG) Response Through Multi-Chiral Centers and Metal-Coordination. Dalton Trans. 2005, 9, 1570–1573. [Google Scholar] [CrossRef]
  50. Lin, W.; Evans, O.R.; Xiong, R.G.; Wang, Z. Supramolecular Engineering of Chiral and Acentric 2d Networks. Synthesis, Structures, and Second-Order Nonlinear Optical Properties of Bis(nicotinato)zinc and Bis{3-[2-(4-pyridyl)ethenyl]benzoatocadmium. J. Am. Chem. Soc. 1998, 120, 13272–13273. [Google Scholar] [CrossRef]
  51. Liu, B.W.; Jiang, X.M.; Li, B.X.; Zeng, H.Y.; Guo, G.C. Li[LiCs2Cl][Ga3S6]: A Nanoporous Framework of GaS4 Tetrahedra with Excellent Nonlinear Optical Performance. Angew. Chem. Int. Ed. 2020, 59, 4856–4859. [Google Scholar] [CrossRef]
  52. Li, R.A.; Zhou, Z.; Lian, Y.K.; Jia, F.; Jiang, X.; Tang, M.C.; Wu, L.M.; Sun, J.; Chen, L. A2SnS5: A Structural Incommensurate Modulation Exhibiting Strong Second-Harmonic Generation and a High Laser-Induced Damage Threshold (A = Ba, Sr). Angew. Chem. Int. Ed. 2020, 59, 11861–11865. [Google Scholar] [CrossRef]
  53. Ye, R.; Cheng, X.; Liu, B.-W.; Jiang, X.-M.; Yang, L.-Q.; Deng, S.; Guo, G.-C. Strong Nonlinear Optical Effect Attained by Atom-Response-Theory Aided Design in the Na2MIIMIV2Q6 (MII = Zn, Cd; MIV = Ge, Sn; Q = S, Se) Chalcogenide System. J. Mater. Chem. C 2020, 8, 1244–1247. [Google Scholar] [CrossRef]
  54. Guo, S.-P.; Cheng, X.Y.; Sun, Z.D.; Chi, Y.; Liu, B.W.; Jiang, X.M.; Li, S.F.; Xue, H.G.; Deng, S.; Duppel, V.; et al. Large SHG Effect and High LIDT Observed Coexisting in Gallium Selenide: A Simple but Perfect Case. Angew. Chem. Int. Ed. 2019, 131, 8171–8175. [Google Scholar] [CrossRef]
  55. Cai, Z.; Cheng, X.; Whangbo, M.-H.; Hong, M.; Deng, S. The Partition Principles for Atom Scale Structure and its Physical Property: Application to the Nonlinear Optical Crystal Material KBe2BO3F2. Phys. Chem. Chem. Phys. 2020, 22, 19299–19306. [Google Scholar] [CrossRef]
  56. Lin, H.; Maggard, P.A. Copper(i)-rhenate hybrids: Syntheses, structures, and optical properties. Inorg. Chem. 2007, 46, 1283–1290. [Google Scholar] [CrossRef] [PubMed]
  57. Gautier, R.; Poeppelmeier, K. Packing of Helices: Is Chirality the Highest Crystallographic Symmetry? Crystals 2016, 6, 106. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) Crystal structure of CuMoO3(p2c) illustrating the CuMoO3 helical chain. (b) Local coordination environment between the p2c ligand and the inorganic framework of CuMoO3(p2c).
Figure 1. (a) Crystal structure of CuMoO3(p2c) illustrating the CuMoO3 helical chain. (b) Local coordination environment between the p2c ligand and the inorganic framework of CuMoO3(p2c).
Symmetry 14 00824 g001
Figure 2. (a) Calculated total DOS of CuMoO3(p2c). (b) –pCOHP plots describing the average chemical bonding interactions. (c) ζ V ( E B ) -vs- E B plot (left) and ζ C ( E B ) -vs- E B plot (right) calculated for the largest SHG component d11 of CuMoO3(p2c). The values are in pm/V.
Figure 2. (a) Calculated total DOS of CuMoO3(p2c). (b) –pCOHP plots describing the average chemical bonding interactions. (c) ζ V ( E B ) -vs- E B plot (left) and ζ C ( E B ) -vs- E B plot (right) calculated for the largest SHG component d11 of CuMoO3(p2c). The values are in pm/V.
Symmetry 14 00824 g002
Figure 3. Calculated refractive indices of the fundamental and second harmonic wavelengths for CuMoO3(p2c). The pink point represents the meet of Type I phase-matching condition (ne(2ω) = no(ω)) for negative uniaxial crystals.
Figure 3. Calculated refractive indices of the fundamental and second harmonic wavelengths for CuMoO3(p2c). The pink point represents the meet of Type I phase-matching condition (ne(2ω) = no(ω)) for negative uniaxial crystals.
Symmetry 14 00824 g003
Figure 4. The cause for the large second harmonic generation (SHG) of the metal-oxide/organic hybrid compound CuMoO3(p2c) is shown to be the inorganic metal-cation-centered groups [CuO3N] and [MoO5N] rather than the organic nonmetal-cation-centered groups [C5O2N2H3].
Figure 4. The cause for the large second harmonic generation (SHG) of the metal-oxide/organic hybrid compound CuMoO3(p2c) is shown to be the inorganic metal-cation-centered groups [CuO3N] and [MoO5N] rather than the organic nonmetal-cation-centered groups [C5O2N2H3].
Symmetry 14 00824 g004
Table 1. Contributions of the atoms to the largest components of the SHG tensors d11 of CuMoO3(p2c). NA refers to the number of the atom type (on the same Wyckoff site) in a unit cell; A τ , NA A τ , A V B τ , A C B τ refer to the contributions (in %) from a single atom, the total atoms of the same type, all VBs for a single atom, and all CBs for a single atom, respectively. d τ denotes the actual value of the contribution (in pm/V) to the SHG for a single atom.
Table 1. Contributions of the atoms to the largest components of the SHG tensors d11 of CuMoO3(p2c). NA refers to the number of the atom type (on the same Wyckoff site) in a unit cell; A τ , NA A τ , A V B τ , A C B τ refer to the contributions (in %) from a single atom, the total atoms of the same type, all VBs for a single atom, and all CBs for a single atom, respectively. d τ denotes the actual value of the contribution (in pm/V) to the SHG for a single atom.
AtomNA A τ A V B τ A C B τ N A A τ d τ
Cu311.7210.691.0335.23.99
Mo36.972.854.1220.92.38
O151.61.140.4624.10.55
N61.230.510.727.40.42
C150.740.270.4711.10.25
H90.150.050.11.40.05
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, T.; Huang, X.; Cheng, X.; Maggard, P.A.; Whangbo, M.-H.; Luan, C.; Deng, S. Prediction of Large Second Harmonic Generation in the Metal-Oxide/Organic Hybrid Compound CuMoO3(p2c). Symmetry 2022, 14, 824. https://doi.org/10.3390/sym14040824

AMA Style

Yang T, Huang X, Cheng X, Maggard PA, Whangbo M-H, Luan C, Deng S. Prediction of Large Second Harmonic Generation in the Metal-Oxide/Organic Hybrid Compound CuMoO3(p2c). Symmetry. 2022; 14(4):824. https://doi.org/10.3390/sym14040824

Chicago/Turabian Style

Yang, Tingting, Xueli Huang, Xiyue Cheng, Paul A. Maggard, Myung-Hwan Whangbo, Chengkai Luan, and Shuiquan Deng. 2022. "Prediction of Large Second Harmonic Generation in the Metal-Oxide/Organic Hybrid Compound CuMoO3(p2c)" Symmetry 14, no. 4: 824. https://doi.org/10.3390/sym14040824

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop