Matan Oliel* and
Yitzhak Mastai
Department of Chemistry and Institute of Nanotechnology, Bar-Ilan University, 5290002 Ramat-Gan, Israel. E-mail: matanoliel5@gmail.com; Yitzhak.Mastai@biu.ac.il
First published on 10th July 2025
Lithium niobate is an achiral crystal that intriguingly exhibits chiral selectivity, as well as properties typically associated with chiral crystals such as nonlinear polarization and piezoelectricity. Here, we explore L-alanine and D-alanine as chiral inducers during the hydrothermal synthesis of LiNbO3 nanocrystals. Structural and morphological analyses using X-ray diffraction, Fourier-transform infrared spectroscopy, high-resolution scanning electron microscopy, and transmission electron microscopy reveal clear differences between pristine and induced crystals. Low-frequency Raman and circular dichroism spectroscopy, high-performance liquid chromatography, and isothermal titration calorimetry were employed to investigate the chirality and enantiomeric selectivity. L-Ala-induced LiNbO3 crystals preferentially adsorbed L-enantiomers, while D-Ala-induced crystals favored D-enantiomers. This study sheds light on the mechanisms of chiral induction in inorganic crystals, highlighting the potential of LiNbO3 for chiral sensing, enantioselective catalysis and biomolecular recognition.
There is surprisingly little research on the chirality of chiral inorganic crystals due to complicated synthesis requiring very high temperature and pressure that are difficult to reach. However, the surface of chiral inorganic crystals shows great potential for several uses including chiral discrimination,2–4 chiral sensing,5 and enantioselective catalysis.6,7 As chiral selectivity occurs mostly on the material surface, it is essential to comprehend how chiral induction affects surface chirality and to manage chirality in these materials.
Numerous studies were conducted on chiral discrimination on metal surfaces, in particular with metals that are cubic close-packed (CCP).8,9 Remarkably, chiral properties are shown in the crystal planes of Cu (643), even though chirality is absent in the bulk metal. The mirror image of the Cu(6, 4, 3) surface – Cu(−6, −4, −3) – is obtained by kinks on the (6, 4, 3) surface plane that make it chiral, denoted by Cu(6, 4, 3)R and Cu(6, 4, 3)S. Their chiral selectivity is proven by desorption of R/S-propylene oxide from those surfaces. For chiral surfaces, adsorption of even a tiny chiral molecule can lead to an outstanding enantiomeric excess (ee)4,10 due to the mechanism of its adsorption onto ligands connected to the surface or even onto the surface itself.
Many crystals investigated over time possess chiral characteristics, including minerals like quartz (SiO2),11 a non-chiral inorganic material that crystallizes as a conglomerate and exists in two variations, right- and left-handed. Moreover, quartz has three crystal structures that are very different from each other but are all chiral owing to their silicate skeletons that form right- or left-handed SiO4 helices. Like quartz, there are many other chiral crystals that crystallize in non-centrosymmetric space groups.12
Calcite (CaCO3),13 on the other hand, is an achiral mineral that displays mirror-related faces, and is the source of many studies dealing with surface chirality of minerals. Calcite formation in the presence of D or L aspartic acid shows a different morphology of the (1, 0, 4) plane, indicating chiral recognition. The formation of gypsum (CaSO4·2H2O) in the presence of 28 different chiral organic molecules, among them amino acids, resulted in crystals with asymmetric morphology.14
Wulfingite (ε-Zn(OH)2)15 (space group P212121) and α-HgS (ref. 16 and 17) (space group P3121) can be influenced by chiral molecules. By utilizing chiral surfactant molecules, such as penicillamine, in the synthesis of α-HgS nanoparticles, or amino acids, in the synthesis of (ε-Zn(OH)2), a chiral crystal phase with a significant ee can be achieved.
Through a protracted process of chiral discrimination, some chiral minerals are thought to have had a role in the formation of homochirality in biological molecules.18,19 Moreover, we can even see that chiral discrimination affects chemical elements like tellurium which is naturally chiral, and its chirality can be controlled by using chiral agents in the formation of tellurium nanocrystals.20–22
It is possible to uncover the underlying natural processes governing the essence of chiral biomolecules by exploring the complexities of chiral induction and simulating the natural crystallization conditions of chiral minerals in the presence of chiral amino acids, to achieve chiral discrimination in these systems. However, while some chiral minerals can be easily crystallized in their chiral phase using simple techniques, other chiral minerals are difficult to crystallize. Berlinite is a chiral polymorph of AlPO4 that resembles quartz structurally. High temperatures are necessary for it to crystallize into the chiral phase during annealing.23,24 Analyzing chiral induction processes is significantly hampered by these extreme circumstances.
An interesting phenomenon is the chiral induction of achiral crystals, which assume properties that exist in chiral crystals. Ag2CO3 and CuO are good examples: the former (space group P121/m) shows chiral selectivity when induced by arginine, while the latter (space group C2/c) shows helical symmetry when induced by both sodium dodecyl sulfate (SDS) as a structure-directing agent and amino alcohol as a symmetry breaking agent.25 These examples increase the variety of possibilities that exist today for the chirality of crystals when properties that characterize chiral crystals also appear in achiral crystals thanks to chiral induction.
LiNbO3 is an achiral crystal that belongs to space group R3c, which displays piezoelectricity and nonlinear polarization of light that characterize chiral crystals.26–29 Therefore, we assumed that chiral selectivity can be induced and even controlled for LiNbO3 crystals, similar to the induced selectivity of Ag2CO3 and CuO. Owing to its properties, LiNbO3 is a very interesting system; its chirality and chiral selectivity could be highly significant for biology and biochemistry.
Here, we investigate the effects of chiral induction by alanine enantiomers, while crystallizing LiNbO3 by a hydrothermal method, on chirality and chiral discrimination, seeking to control nanocrystal orientation and chiral selectivity. The morphology and structural changes were evaluated by X-ray diffraction (XRD), Fourier transform infrared (FTIR) spectroscopy, high-resolution scanning electron microscopy (HR-SEM), and transmission electron microscopy (TEM). The chiral properties were analyzed by direct power measurement using low-frequency Raman (LFR), chiral adsorption using circular dichroism (CD) spectroscopy, isothermal titration calorimetry (ITC) and high-performance liquid chromatography (HPLC) (Scheme 1).
FTIR was performed using a Thermo Scientific Nicolet iS10 spectrometer equipped with a Smart iTR attenuated total reflectance sampler with a single bounce diamond crystal. Data were collected in the 530–4000 cm−1 range at a spectral resolution of 4 cm−1 and analyzed using OMNIC software.
HR-SEM images were taken using a field-emission FEI (Helios 600) instrument. Samples were sputtered with gold.
Transmission electron microscopy (TEM) images were taken using a JEOL JEM-1400 microscope operated at 120 kV. Images were taken using a digital micrograph with a multiscan camera model 794 (Gatan) in different resolutions.
HPLC chiral adsorption was measured by adding a 5 mg mL−1 freshly prepared solution of DL-Ala in DDW to 10 mg mL−1 of L- or D-Ala-induced LiNbO3 crystals. The crystals were kept suspended in the solution overnight using a rotary suspension mixer and filtered out using a 0.2 mm filter. The filtered solutions were subjected to HPLC measurement and compared to the control solution using an Astec Chirobiotic TAG HPLC column (5 μm particle size, 15 cm × 4.6 mm ID). The mobile phase was MeOH:
H2O 60
:
40 and the injected sample volume was 2 μL. HPLC measurements were done using a BETA 10 Plus Gradient Pump (Ecom, Czech Republic), a SAPPHIRE 800 UV-vis Variable Wavelength DETECTOR at a wavelength of 210 nm, and a DEGASSEX MODEL DG-4400 degasser (Phenomenex, Torrance, CA, USA).
CD chiral adsorption was measured by preparing 2.5 mM solutions of L- and D-Arg in DDW and adding them to 10 mg mL−1 of L or D-Ala-induced LiNbO3 crystals. The crystals were kept suspended in solution overnight using a rotary suspension mixer and were filtered out using a 0.2 mm filter. The filtered solutions were subjected to CD measurement and compared to the control L- and D-Arg solutions at RT using a 5 mL quartz cuvette and Chirascan spectrometer (Applied Photophysics, UK).
ITC chiral adsorption was measured using 5 mM solutions of L- and D-Arg in DDW for the syringe (80 μL) and 10 mg mL−1 of L- or D-Ala-induced LiNbO3 crystal suspension in DDW after 10 min of sonication for the sample cell (1.4 mL). The reference cells were filled with DDW (1.4 mL). Each consecutive 5 μL injection lasted 8.5 s, with an interval of 400 s between injections. The cell was stirred constantly at a rate of 300 rpm, and the reference power was 5 μcal s−1. Prior to measurement, the solutions were degassed under vacuum using a ThermoVac (MicroCal) for 5 min at 25 °C to remove air bubbles. The results were analyzed using Origin software.
![]() | ||
Fig. 1 XRD of the control and chiral-induced crystals at 2θ = 10–80° showing the diffraction planes of trigonal LiNbO3. |
The Scherrer equation was applied to the five most intense XRD reflections to calculate the average particle size: K is a dimensionless shape factor (0.89), λ is the X-ray wavelength (0.15406 nm), β is the net line broadening (subtracting the instrumental broadening) at half the maximum intensity, and θ is the Bragg angle (both in radians). The calculations show an increase in size as a result of the chiral induction, by 10–12% (75 vs. 84 nm for L-Ala and 82 nm for D-Ala, see Table 1). Similar reflection ratios indicate no change of preference to different planes during crystallization as a result of the chiral induction. Moreover, the weaker intensities of the D-Ala-induced LiNbO3 crystals compared to those of the control LiNbO3 and L-Ala-induced LiNbO3 crystals result from the lower sample quality.
LiNbO3 | Plane | Intensity | 2θ | FWHM | Size (nm) | Average size (nm) |
---|---|---|---|---|---|---|
(0, 1, 2) | 23![]() |
23.77 | 0.22613 | 77 | 75 | |
(1, 0, 4) | 8191.5 | 32.77 | 0.25414 | 74 | ||
(1, 1, 0) | 4815.328 | 34.89 | 0.24523 | 79 | ||
(0, 2, 4) | 3042.3 | 48.59 | 0.33732 | 71 | ||
(1, 1, 6) | 4203.106 | 53.31 | 0.31865 | 83 |
L-LiNbO3 | Plane | Intensity | 2θ | FWHM | Size (nm) | Average size (nm) |
---|---|---|---|---|---|---|
(0, 1, 2) | 27![]() |
23.74 | 0.20416 | 85 | 84 | |
(1, 0, 4) | 8376.02 | 32.74 | 0.23661 | 80 | ||
(1, 1, 0) | 5143.255 | 34.86 | 0.22577 | 86 | ||
(0, 2, 4) | 3376.629 | 48.56 | 0.29764 | 81 | ||
(1, 1, 6) | 4332.208 | 53.28 | 0.30725 | 86 |
D-LiNbO3 | Plane | Intensity | 2θ | FWHM | Size (nm) | Average size (nm) |
---|---|---|---|---|---|---|
(0, 1, 2) | 18![]() |
23.78 | 0.20102 | 86 | 82 | |
(1, 0, 4) | 4526.962 | 32.78 | 0.26023 | 73 | ||
(1, 1, 0) | 2624.061 | 34.89 | 0.23178 | 84 | ||
(0, 2, 4) | 1559.819 | 48.6 | 0.26676 | 90 | ||
(1, 1, 6) | 2467.739 | 53.32 | 0.33537 | 79 |
FTIR spectroscopy was used to characterize the as-synthesized LiNbO3 (Fig. S2†). The spectra for pure and D/L-Ala-induced LiNbO3 were the same, showing the desired Nb–O vibration band at 562 cm−1. No other peaks are observed in the spectra, indicating that the product is clean from any traces of alanine.
The morphology and surface of the crystals were examined by HR-SEM as presented in Fig. 2. Both non-induced and L-chiral-induced LiNbO3 crystals show a structure of mesocrystals30,31 made up of nm-sized building blocks, arranged in a regular, repeating pattern. Mesocrystals are constructed of small, organized nanocrystals that can be joined in many ways, as opposed to typical crystals, made up of a continuous lattice of atoms or molecules. The pristine LiNbO3 mesocrystals are arranged by small nanometric crystals on top of larger ones (both rhombic) in a way that resembles terraces with an average size of ∼300 nm, and sharp vertex crystal blocks (I). The addition of the amino acid as a chiral inducer conserved the arrangement of the mesocrystalline structure, but minor differences in morphology are indicated. Addition of L-Ala to LiNbO3 crystals cause small deformation in the rhombic crystals which is reflected in less pointed vertices compared to the un-induced LiNbO3 crystals, although no change in crystal size was detected (average of ∼300 nm) (II). D-Chiral-induced LiNbO3 rhombic crystals show similar deformation of crystal morphology like the L-chiral-induced LiNbO3 crystals. Moreover, differences in arrangement were indicated which are reflected in a matrix arrangement, and without size change (average of ∼300 nm) (III).
![]() | ||
Fig. 2 HR-SEM images of the control (I/II), L-Ala-induced (III/IV) and D-Ala-induced (V/VI) LiNbO3 crystals. |
After learning about the induced crystalline and morphological changes, we examined the chirality and chiral selectivity. Some techniques to examine the chirality and selectivity of inorganic systems involve direct measurements, while others are indirect. Direct measurements include methods like LFR, solid and powder CD spectroscopy, and even second-harmonic generation (SHG),32–34 which provide solid proof for the chirality of the crystals; however, there is great difficulty finding the conditions for such measurement. Indirect methods like adsorption and crystallization experiments allow checking the selectivity of the chiral inorganic systems by techniques such as polarimetry, CD, ITC and HPLC; however, these methods do not provide information about the crystals themselves.
To study our chiral-induced LiNbO3, we used the LFR method to measure its chirality directly. The LFR method is a typical Raman measurement that allows the measurement of the chirality of crystals at low frequencies in the addition of an optical polarizer.35–38 In this measurement, polarized light hits the sample, and the collected excited light passes through the optical polarizer to the detector to give Raman signals. In chirality measurements, chiral molecules and chiral crystals give different signal intensities when they are perpendicularly/parallelly polarized (S-Pol/P-Pol). To prove the chirality, the enantiomer should give stronger intensity for S-Pol, while its co-enantiomer should show the opposite phenomenon – stronger signals for P-Pol. The crystals are excited, and the collected light passes through an optical polarization at S-Pol and P-Pol.
The LFR spectrum of the LiNbO3 crystal shows typical peaks at 151 cm−1 corresponding to the ETO1 mode, at 237 cm−1 corresponding to the ETO2 mode (the strongest vibration mode), and at 365 cm−1 corresponding to the ETO3 mode.39–41 In general, ETO modes are transverse optical vibrations in the xy plane of the crystal. The ETO1 and ETO2 modes describe deformations/titling of oxygen octahedra combined by cationic displacements in the xy plane and ETO3 mode describe tilting of the oxygen octahedra.42 In our measurements, we focused on the 237 cm−1 frequency because it is the most intense peak of the LiNbO3 LFR scattering, and the P/S-polarization effect is the strongest at this mode. For un-induced LiNbO3 crystals, no difference in intensities are indicated when the light has been collected at S-Pol/P-Pol throughout the spectrum with 99.6% overlap. When the light was collected at S-Pol, L-Ala-induced LiNbO3 crystals showed a higher intensity than when the light was collected at P-Pol throughout the spectrum with a maximum intensity ratio of 1.25 at the strongest vibration mode (237 cm−1, Fig. 3II). As expected, the co-enantiomer D-Ala-induced LiNbO3 crystals showed the opposite phenomenon – higher intensity with light collected at P-Pol throughout the spectrum with a maximum intensity ratio of 1.28 (237 cm−1, Fig. 3III).
![]() | ||
Fig. 3 LFR spectra of un-induced (I), L-Ala-induced (II) and D-Ala-induced (III) LiNbO3. Black/red curves present signals excited using S/P-Pol. |
LFR represents long-range interactions, where polarizable interactions are orientation-dependent. These long-range interactions are particularly sensitive to changes in the symmetry of the excitation cone induced by the polarizer. Therefore, in a measurement performed in this spectral region, observation of the phenomenon, in which the intensity of light collected from a chiral material varies as a function of the collection angle (S-Pol/P-Pol), allows us to distinguish and prove the chirality of LiNbO3.
CD allows the checking of the chirality of solutions and powders by using circular polarization.43–45 In CD measurements, the chirality of the molecule is determined by measuring the absorbance difference between left- and right-handed circular light, providing the molar CD. The measurements help to prove chiral selectivity using adsorption experiments by determining the CD ratio between the enantiomer adsorbed on the crystal surface and the pure enantiomer. As the concentration is related to peak intensity, the disparity was taken to represent the quantity of chiral molecules adsorbed onto the crystal surface.
Fig. 4I presents the CD of D/L-Arg adsorbed onto un-induced LiNbO3 and onto chiral-induced LiNbO3 at a concentration of 10 mg mL−1. As expected, the L- and D-Arg CD after adsorption onto un-induced LiNbO3 doesn't present any chiral selectivity between the enantiomers compared to the pure L- and D-Arg (Fig. 4II). In contrast, L-Ala-induced LiNbO3 crystals show better adsorption for L-Arg (62%, red curve) compared to pure L-Arg (black), while D-Arg shows absorbance of 44% onto L-Ala-induced LiNbO3 (cyan) compared to pure D-Arg (pink). The arginine enantiomers show selectivity to the L-induced crystals, with better selectivity to L-Arg at all concentrations with a maximal selectivity of 18% (Fig. 4III). On the other hand, the D-Ala-induced LiNbO3 crystals show a better adsorption of 54% (green) for D-Arg and 41% for L-Arg compared to the pure enantiomers. The selectivity of the Arg enantiomers towards D-Ala-LiNbO3 reveals a notable preference for D-Arg at all concentrations, with a maximum selectivity of 13% (Fig. 4IV). Moreover, the difference in selectivity between the enantiomers changes only slightly in both L- and D-Ala-induced LiNbO3 crystals at increasing crystal concentrations.
For the chiral HPLC measurements, we checked the induced LiNbO3 chiral selectivity adsorption in the DL-Ala racemic mixture by using a chiral column. HPLC allows separating and quantifying different materials in a mixture. In chiral selectivity measurements, a chiral column can separate the D and L enantiomers in racemic mixtures (the enantiomers exit the column at different times), allowing us to track the changes in the ratio between the quantities of each enantiomer, which is related to their peak area. In adsorption experiments, if one enantiomer is adsorbed more to the crystal than its co-enantiomer, its relative quantity in the remaining solution would be lower, changing the ratio between the enantiomers.
Fig. 5 presents the adsorption of the DL-Ala racemic mixture onto un-induced and L- or D-Ala-induced LiNbO3. The control measurement was a racemic solution of DL-Ala with no crystals. The solution contained 52.77% L-Ala (217.5 mAU s) and 47.23% D-Ala (194.7 mAU s). The sample that was suspended with un-induced LiNbO3 showed a peak area decrease of 10.62% for L-Ala (194.4 mAU s) and 11.04% for D-Ala (173.2 mAU s). The solution contained 52.88% L-Ala (47.12% D-Ala) which means that although there is adsorption of alanine to the LiNbO3 crystals, the adsorption isn't selective at all.
![]() | ||
Fig. 5 Selective chiral adsorption of alanine on un-induced and chiral-induced LiNbO3 crystals measured by HPLC equipped with a chiral column. |
In the sample suspended with L-Ala-induced LiNbO3, the solution contained 48.32% L-Ala (51.68% D-Ala). The peak area change in the L-Ala-induced LiNbO3 sample was −22.12% for L-Ala and −7.18% for D-Ala, a chiral selectivity of ∼14.9% in favor of L-Ala.
The sample suspended with the D-Ala-induced LiNbO3 solution contained 58.08% L-Ala (41.92% D-Ala). The peak area change is thus −6.91% for L-Ala and −25.64% for D-Ala – a chiral selectivity of ∼18.7% in favor of D-Ala (Fig. 5). The HPLC chromatograms of DL-Ala before and after adsorption onto un-induced and L- or D-Ala-induced LiNbO3 are shown in Fig. S3.†
The chiral selectivity adsorption of the L- or D-Ala-induced LiNbO3 crystals was examined by ITC, selecting arginine for the adsorption. ITC allows us to measure enthalpy changes at good resolution by titration. The instrument includes a syringe containing the titrant, a sample cell for the titrand, and a reference cell for the solvent. While titrating, a reaction occurs, heat is usually released or consumed, and the temperature in the sample cell changes. The thermostat applies an electric voltage equal to the heat change to keep both cells at the same temperature, and the feedback appears as a peak signal in the ITC curve.
In chirality measurements, small enantiomer molecules titrate into a suspension of chiral crystals, and the enthalpy change (ΔH) differences between both enantiomers indicate selective adsorption onto the crystal surface.44–53 Fig. 6 presents the total enthalpy per injection after integration, normalizing from the raw measurement graph (Fig. S4†). The results in Fig. 6I present the enthalpy for the titration of L- and D-Arg into un-induced LiNbO3, showing very low enthalpy change values and no chiral selectivity at all. When we titrated L- and D-Arg into L-Ala-induced LiNbO3 the L-Arg shows significantly stronger adsorption (average = −1.926 kcal mol−1) compared to D-Arg (−1.587 kcal mol−1). The energy difference between the enantiomers is 0.339 kcal mol−1, a selectivity of 17.6% in favor of L-Arg (Fig. 6II). As expected, Fig. 6III presents the opposite trend: D-Arg adsorbs much more strongly onto D-Ala-induced LiNbO3 crystals (−3.498 kcal mol−1) compared to L-Arg (−2.166 kcal mol−1), a highly significant energy difference (the difference between both enantiomers is 1.332 kcal mol−1), corresponding to a selectivity of 38.1% in favor of D-Arg. The adsorption is thus much better, with more than two-fold greater selectivity compared to the L-Ala-induced LiNbO3 crystals.
![]() | ||
Fig. 6 Calculated selective chiral adsorption of Arg on induced LiNbO3 – enthalpy changes of L- and D-Arg onto crystals: un-induced (I) and induced by L-Ala (II) and D-Ala (III). |
Footnote |
† Electronic supplementary information (ESI) available: SAED, FTIR spectra, HPLC and ITC raw data of L- and D-Ala-induced LiNbO3 (PDF). See DOI: https://doi.org/10.1039/d5ce00522a |
This journal is © The Royal Society of Chemistry 2025 |