Next Article in Journal
Study on Anti-Penetration Performance of Semi-Cylindrical Ceramic Composite Armor against 12.7 mm API Projectile
Next Article in Special Issue
Study on Precipitation Processes and Phase Transformation Kinetics of Iron Phosphate Dihydrate
Previous Article in Journal
Unambiguous Identification of Crystal Plasticity Parameters from Spherical Indentation
Previous Article in Special Issue
Solution-Mediated Polymorphic Transformation of L-Carnosine from Form II to Form I
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Croconic Acid Doped Glycine Single Crystals: Growth, Crystal Structure, UV-Vis, FTIR, Raman and Photoluminescence Spectroscopy

1
Ioffe Institute, Politechnicheskaya 26, 194021 Saint Petersburg, Russia
2
Institute of Earth Sciences, Saint-Petersburg State University, Universitetskaya Nab. 7/9, 199034 Saint-Petersburg, Russia
3
Functional Materials and Acoustooptic Instruments Institute, School of Instrument Science and Engineering, Harbin Institute of Technology, Harbin 150080, China
4
Scool of Materials Sciences and Engineering, Harbin Institute of Technology, Harbin 150080, China
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(10), 1342; https://doi.org/10.3390/cryst12101342
Submission received: 19 August 2022 / Revised: 17 September 2022 / Accepted: 17 September 2022 / Published: 22 September 2022
(This article belongs to the Special Issue Polymorphism in Crystals)

Abstract

:
Glycine (Gly) single crystals doped with croconic acid (CA) were grown by evaporation from aqueous solutions. Depending on the weight ratio of Gly and CA in solutions, the crystals take on a plate or pyramidal shape. Both powder and single crystal XRD analyses indicate that the crystal lattices of plates (α-Gly:CA) and pyramides (γ-Gly:CA) correspond to the lattices of pure α-Gly and γ-Gly polymorphs, respectively. Raman and FTIR spectra of Gly:CA crystals are very close to the spectra of undoped crystals, but include bands associated with CA impurity. Analysis of UV-Vis absorption spectra indicates that doping does not remarkably change bandgap value Eg~5.2 eV but results in appearance of strong absorption bands in the transparency region of pure glycine crystals, which result from local electronic transitions. Incorporation of CA molecules in Gly creates strong green photoluminescence in a wide spectral range 1.6–3.6 eV. Comparison of the optical spectra of Gly:CA and previously studied TGS:CA crystals indicates that in both cases, the modifications of the optical spectra induced by CA doping are practically identical and are related to the interaction between CA molecules located in the pores of the host Gly crystals and neighboring Gly molecules.

1. Introduction

Protein amino acid glycine (Gly) NH2-CH2-COOH was discovered in 1820 by the chemist, botanist, and pharmacist Henri Braconnot. It was the first case wherein a pure amino acid was obtained from a protein [1]. Since glycine is produced in the human body during metabolism, it is safe and is widely used in medicine as a metabolic regulator that normalizes and activates protective inhibition processes in the central nervous system and reduces psycho-emotional stress.
Piezoelectricity is in increasing demand as it provides diverse entries into electronic, electromechanical, optical, and optoelectronic applications. A very high polarization response, which is promising for capacitors, memories, and piezoelectric applications, has been discovered mostly in inorganic oxides containing toxic lead or rare elements. Purely organic ferroelectrics, which are expected to be used as key materials in organic, printable, and bendable electronic device applications, are being pursued as possible alternatives. Recently, it was demonstrated that a basic component of biological structure, γ-glycine, presents nanoscale ferroelectricity, with an exhibited technologically significant piezoelectric response. Glycine is one of the simplest and smallest biological molecules. It therefore presents a starting point for the design of functional materials as it can be synthetically modified to optimize properties.
Crystallization of glycine is characterized by polymorphism. Currently, α-, β-, γ-, δ-, and ε-Gly modifications of glycine crystal structure are known. The most stable among them are α- and γ-modifications. It is believed that the α modification is the most stable one because γ- and β-glycine are transformed to α-glycine at certain conditions. However, there exists an opposite perspective in the literature, which relates the γ-modification to the most stable one [2]. Γ-Gly transforms in α-Gly at T ≈ 460 K [3,4]. Β-Gly transforms in α- Gly or γ-Gly at ambient conditions [2]. Similarly, under high pressure, phase transitions of β-Gly to δ-Gly and γ-Gly to ε-Gly occur [5].
α-Gly crystallizes in centrosymmetric crystal structure [space group P21/n (14)]. In contrast, β- and γ-Gly crystallize in non-centrosymmetric structure of monoclinic [sp.gr. P21 (4)], and trigonal [sp.gr. P32 (145) or P31 (144)] symmetry, respectively [2,3,4,5,6,7]. Nominally pure glycine single crystals are transparent in a wide spectral range. Strong increase of absorption due to band-to-band electronic transitions begins in ultraviolet (UV) region for the light wavelength λ shorter than cut-off wavelength λgap. In both α-Gly [8] and γ-Gly [9,10,11,12] λgap ≈ 240 nm. Estimations of the bandgap Eg are based on use of Tauc plot and provide the values of Eg = 5.11 ± 0.02 eV in α-Gly [8] and Eg = 5.09–5.3 eV in γ-Gly [11,12,13]. A much higher value of Eg = 6.2 eV has been reported in Ref. [14].
Non-centrosymmetric γ-Gly attracts much attention due to remarkable piezoelectric, pyroelectric, and non-linear optical (NLO) properties. The longitudinal piezoelectric coefficient d33 value for the γ-Gly crystals was discovered to be d33 = 7.37 pC/N [13], piezoelectric strain constant ~9.93 pm∙V−1 [15], pyroelectric coefficient p~13–21.4 μC/m2K [16,17]. The large value of piezoelectric coefficient makes γ-Gly an attractive material for design of flexible amino acid-based energy harvesting [18]. The glycine polycrystalline sensors indicate the effective piezoelectric constants of d33 = 0.9 pC/N and of g33 = 60 mV∙m/N, respectively. Note that the latter one exceeds that of commercial piezoelectric lead zirconium titanate (PZT) [18].
Since glycine crystals grow from aqueous solutions, the addition of various substances to solutions makes it possible to significantly influence the process of glycine crystallization and obtain doped glycine crystals with modified physical properties. In particular, the type of additive and its amount can determine what polymorph of glycine will grow from solution. It was demonstrated in [19] that the addition to the solution of a small amount of zinc sulphate ZnSO4 with a molar concentration (0.2–0.4 M) is accompanied by growth of α-Gly, and an increase in concentration (0.6–1 M) leads to the appearance of the γ-polymorph. In [20], the effects of seven common salts on primary nucleation of the metastable α-Gly and the stable γ-Gly were studied. It was demonstrated that addition of (NH4)2SO4, NaCl and KNO3, in general, results in primary nucleation of γ-Gly simultaneously inhibiting α-Gly primary nucleation. Addition of Ca (NO3)2 and MgSO4 also promotes γ-Gly and inhibits α-Gly primary nucleation but not sufficiently to induce γ-Gly. Na2SO4 and K2SO4 promote not only γ-Gly but also α-Gly primary nucleation.
Polymorphic control of glycine by using the microdroplet technique for NaCl additive was investigated in [21]. Analysis of thermodynamics and nucleation kinetics of glycine polymorphs in microdroplets indicates that NaCl reduces the nucleation energy barrier of the γ-form of glycine, and for the same form of glycine, crystallization by the conventional method has a higher interfacial tension than in microdroplets.
Comparison of the crystallization process of glycine polimorphs and its salts/co-crystals in the presence of organic carboxylic acids through implementation of several different preparative techniques was studied in [6]. It was discovered that for the outcome of crystallization in “glycine– carboxylic acid” systems upon slow evaporation of aqueous solutions, both the choice of the acid and its concentration are of importance.
Additives can also influence a morphology of glycine crystals. As indicated in [22], the γ-Gly crystals grown from the solution containing glycine and sodium fluoride, sodium hydroxide, or sodium acetate had only three well developed faces (101), (0–11), and (−111) with the remaining faces developed in the form of a pyramid. In contrast, adding to the solution of sodium nitrate resulted in a different morphology with four well-developed faces.
It is important that additions can strongly affect a nonlinear properties of glycine crystals. It was determined that the addition of alkali metal nitrides, bromides, chlorides, or acetates to a solution can significantly change the second-harmonic generation (SHG) efficiency in γ-Gly [11,23,24,25], reaching a maximum value for Cs chloride 5.06 times higher than that in a potassium dihydrogen phosphate (KDP) crystal [26].
From aqueous solutions of glycine with various inorganic and organic acids and salts, many new crystals were obtained [6,7], particularly the well-known ferroelectric triglycine sulfate (TGS), which has record values of the pyroelectric coefficient [27], and other ferroelectric crystals based on glycine and methylated glycine [7,28,29].
The effect of organic croconic acid (CA) additives on the glycine crystallization process has not yet been studied. Croconic acid crystals are a promising material in which CA molecules bonded by O-H∙∙∙O hydrogen bonds provide excellent ferroelectricity at room temperature (RT) [30]. However, CA is corrosive and requires another acid in the solution to crystallize. In this regard, it is of considerable interest to study the properties of mixed Gly-CA crystals. The CA molecule, C5H2O5, belongs to a series of planar monocyclic CnH2On oxocarbon acids. It has highly symmetric (D5h) plane topology and large value of molecular dipole momentum d~ (9–10) D [30]. CA molecule is small and may be incorporated into the host lattice as an impurity in a wide concentration range.
It was recently discovered that doping above RT ferroelectric TGS crystals with CA molecules results in appearance of a strong green luminescence band in region of photon energy Eph = 1.6–3.5 eV with the band maximum at Eph = 2.55 eV. The doping also leads to the appearance of absorption bands related to local optical transitions in the transparency region of nominally pure TGS crystals and change of dielectric hysteresis loops [31]. CA doping does not change significantly TGS crystal structure, does not result in appearance of new phases, and does not influence frequencies of molecular vibrations measured by Raman and Fourier-transform infrared spectroscopy (FTIR).
Since, unlike TGS (which includes ordinary glycinium ions, zwitterions, and sulfuric acid molecules), glycine crystals are formed only by glycine zwitterions; the crystal structure of glycine can be either centrosymmetric (α-Gly) or non-centrosymmetric (γ-Gly). It is interesting to compare the changes in the crystal structure, luminescence, and optical absorption caused by doping TGS and Gly crystals with croconic acid. For these reasons, in this work Gly single crystals were grown from aqueous solutions of glycine and croconic acid at the same relative ratios (80:20, 90:10, 98:2) as in previous studies of TGS:CA crystals.
Particularly, the primary questions of interest for the present study were the following: how does CA doping affect the crystal structure, crystal morphology, unit cell parameters of various glycine isomorphs, the amplitudes and frequencies of glycine molecular vibrations, the electronic band structure, in particular, the value of the band gap Eg, optical properties including absorption and luminescence spectroscopy of doped crystals. Generally, the aim of this research was the synthesis of CA–doped α- and γ-Gly single crystals and the study of their structural and optical properties. Preliminary results on Raman and photoluminescence spectroscopy were presented in our conference paper [32].

2. Materials and Methods

2.1. Crystal Synthesis

Figure 1 presents images of crystals used in this research. Nominally pure single crystals of α-Gly were grown by evaporation from saturated aqueous solutions of aminoacetic acid (α-Gly) (Figure 1a). Colorless “pure” γ-Gly crystals were grown from aqueous solution of α-Gly crystals with adding 5% of citric acid C6H8O7 and may contain a small admixture of the acid (Figure 1b). Croconic acid crystals (Figure 1c) were prepared from the CA reagent (Alfa Aesar B21809, 98% purity) according to the procedure described in [33].
Croconic acid doped Gly crystals were grown from aqueous solutions with different weight rations of Gly and CA—80:20, 90:10, 98:2. Gly:CA crystals grown from 80:20 and 98:2 solutions have a pyramidal shape with a hexagonal base (Figure 1d,f) analogous to nominally pure γ-Gly (Figure 1b). Crystals grown from 90:10 solution have plate shape like pure α-Gly (Figure 1a).

2.2. Experimental Details of Single Crystal XRD

Single crystal X-ray diffraction (XRD) studies of Gly:CA (90:10) and Gly:CA (80:20) were performed in the Harbin Institute of Technology, China, by a Bruker D8 Venture setup (50 kV, 30 mA) at RT using Mo-Kα radiation, and Gly:CA (80:20) by Rigaku XtaLAB Synergy diffractometer at 100 K using Cu-Kα radiation (50 kV, 1 mA) in the Resource Center “X-ray Diffraction Methods” of St. Petersburg State University. The structures have been solved by the direct methods by means of the SHELX program [34] incorporated in the OLEX2 program package [35]. The carbon and nitrogen-bound H atoms were placed in calculated positions and were included in the refinement in the ‘riding’ model approximation. Empirical absorption correction was applied in CrysAlisPro program complex [CrysAlisPro, Version 1.171.36.32, Agilent Technologies Inc., Santa Clara, CA, USA] using spherical harmonics, implemented in SCALE3 ABSPACK scaling algorithm [for Gly:CA (80:20)] and using the Bruker software SADABS-2016/2 [Bruker, 2016/2] [for Gly:CA (90:10)]. Supplementary crystallographic data have been deposited in the Cambridge Crystallographic Data Centre (CCDC) (2194293-2194294) and can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif (accessed on 8 April 2022). The drawings of the Gly:CA (90:10) and Gly:CA (80:20) crystal structures were made using the Vesta program [36].

2.3. Experimental Details of Powder XRD

To prepare the samples for powder XRD investigations, the grown crystals were grinded in a corundum mortar. Low-background holders in the form of polished single-crystal Si (119) plate were used to prepare powder samples for XRD measurements.
Powder XRD measurements were performed utilizing a D2 Phaser X-ray powder diffractometer (Bruker AXS, Karlsruhe, Germany) equipped with an X-ray tube with a copper anode and a Ni Kβ filter (Cu-Kα radiation, λ = 1.54184 Å). The XRD patterns were recorded in a symmetric scanning θ-2θ mode using a semiconductor position-sensitive linear X-ray detector LYNXEYE (Bruker AXS, Karlsruhe, Germany). During the XRD measurements, temperature in sample chamber was 313 ± 1 K. Additional XRD measurements of the powder samples mixed with internal XRD powder standard Si640f (NIST, Gaithersburg, MD, USA) were performed to win the angular corrections of the observed Bragg angles of the XRD reflections to zero shift and displacement.
After introducing the angular corrections to the Bragg angles 2θB of XRD reflections with attributed Miller indices hkl, at first stage of analysis, unit cell parameters of the crystalline phases were calculated using the program Celsiz [37], utilizing the least-square technique. Microstructural parameters (average size D of crystallites and absolute values of mean microstrain εs in them) for crystalline phases were estimated from parameters of XRD reflections [Bragg angles 2θB, full width at half maximum FWHMobs, maximum (Imax) and integral (Iint) intensities] by graphical methods of Williamson–Hall plot (WHP) [38] and size-strain plot (SSP) [39] adapted for pseudo-Voigt (pV) type of XRD reflections observed in the XRD patterns. All microstructure calculations [including the correction of the FWHMobs to instrumental broadening for pV reflections (resulting in FWHMcorr)] were performed using the program SizeCr [40]. Detailed description of the formalism of microstructural calculations can be found elsewhere [40,41].
To confirm the reliability of the crystal structure model and the results of single-crystal XRD analysis, to obtain more accurate values of the unit cell and microstructure parameters, and to verify the single-phase nature of powder samples, the powder XRD patterns were studied by the Le Bail (LB) and Rietveld methods. The TOPAS program [42] was used for LB and Rietveld fittings. For LB fitting the XRD patterns, a structure model is not required, but only an approximate value of the unit cell parameters and symmetry space group knowledge [43]. Rietveld fitting is used for structure model refinement, including atomic coordinates and temperature parameters [44].
The XRD patterns of both samples are characterized by a noticeable influence of preferential orientation effects. Unlike the Rietveld method in the LB fitting, the influence of preferred orientation effects to XRD pattern is corrected automatically. When fitted by the Rietveld method, considering the effect of preferential orientation within the March–Dollase model [45] led to a decrease in the weighted profile factor Rwp, characterizing the quality of the fit by ~2.5 and ~4 times for α-Gly:CA (90:10) and γ-Gly:CA (80:20), respectively. Additional consideration of the remaining effects of preferential orientation in the framework of the 8th-order spherical harmonics model [46] led to a further decrease in Rwp by ~1% for both samples.
Understated due to serial correlations [47], estimated standard deviations (e.s.d.s) of the refined parameters obtained using LB or Rietveld refinement were corrected by multiplying by the coefficient me.s.d. calculated using the RietESD program [48], based on the formalism developed in [47,49].
Other parameters of powder XRD measurements, WHP and SSP analysis, LB, and Rietveld fittings are the same as detailed in [31,41,50] and briefly discussed in Section S2 of the Supplementary Materials.

2.4. Raman Scattering, FTIR, UV Vis Absorption, and Photoluminescence

Raman and micro-photoluminescence (μ-PL) measurements in the Gly:CA single crystals were carried out using LabRAM HREvo UV-VIS-NIR-Open spectrometer (Horiba Jobin–Yvon, Lille, France) equipped with a confocal microscope and a silicon CCD cooled by liquid nitrogen. Polarized micro-Raman measurements were performed at RT, in the spectral range 5–4000 cm−1 at different backscattering geometries: −Z (XX)Z, −Z (YY)Z, and −Z (XY)Z. Here, the Z-axis is oriented normally to the crystal surface (100), and X and Y are along c*- and b-crystal axes, respectively. The line at λ = 532 nm (2.33 eV) of Nd:YAG laser (Torus, Laser Quantum, Inc., Edinburg, UK) was used as the excitation source. The laser power on the samples was as low as ~25–80 µW with a spot size of ~1 µm in diameter, to avoid sample heating. We used 1800 lines/mm grating and 100× (NA = 0.90) objective lens to measure Raman spectra. In the low frequency spectral region, the Rayleigh line was suppressed using three BragGrate notch filters (OptiGrate Corp., Oviedo, FL, USA) with an OD = 4 and a spectral bandwidth <0.3 nm.
The µ-PL measurements were performed in microscope stage Linkam THMS600 (Linkam Sci. Inst. Ltd., Salfords, Surrey, UK). The line at λ = 325 nm (2.81 eV) of HeCd laser (Plasma JSC, Ryazan, Russia) was used for continuous wave (CW) excitation. We used 600 lines/mm grating and a large working distance lens [Mitutoyo 50× UV (NA = 0.40)] with a spot size of ~2 µm and power density of 6 kWt/cm−2 on a sample was used to measure µ-PL.
Measurements of the infrared (IR) absorption spectra (FTIR spectra) were performed using an IR-Fourier spectrophotometer IRPrestige-21 with an IR microscope AIM-8000 (Shimadzu Corp., Kyoto, Japan), both in the specular reflection mode and in the transmission mode, followed by the Kramers–Kronig transformation. The results were then converted to absorbance. The measured spectral range was from 650 to 5000 cm−1.
Optical absorption spectra of Gly:CA in the UV-visible (UV-Vis) spectral range were obtained using a UV-3600i Plus UV-Vis-NIR spectrophotometer (Shimadzu Corp., Kyoto, Japan) operating at RT in the wavelength range of 200–2000 nm. The test was conducted in reflection mode using an integrating sphere. BaSO4 was used as a reference sample.

3. Results and Discussion

3.1. Single Crystal XRD Analysis

Single crystal XRD analysis indicates that the pyramid-like samples of Gly:CA (80:20) crystallize in a lattice typical to γ-Gly (space group P31 (144)) (γ-Gly:CA), and the plate-like crystals of Gly:CA (90:10) crystallize in α-Gly-like structure (space group P21/n (14)) (α-Gly:CA) (Table 1). Structures of α-Gly:CA (90:10) and γ-Gly:CA (80:20) are drawn in Figure 2 using the data of Table 1 and refined atomic coordinates (CCDC 2194293-2194294). Relative coordinates of atoms (x/a, y/b, z/c), their isotropic temperature factors U, and atomic displacement parameters Uij in α-Gly:CA (90:10) and γ-Gly:CA (80:20) are presented in Supplementary Materials (Tables S1–S4). Figure S1 portrays the structure of glycine molecule in (a) α-Gly:CA (90:10) and (b) γ-Gly:CA (80:20) with atoms represented by thermal ellipsoids by means of program ORTEP [51].
The crystal structures of all glycine phases are built from chains of glycine molecules linked by N1-H3···O2 hydrogen bonds and directed along the crystal c axis of all polymorphic modifications. In α-Gly, somewhat longer hydrogen bonds involving N1-H4···O1 atoms form the second set of hydrogen bond chains running along the a crystal axis. These chains form a layer parallel to the (010) plane. The layers are arranged perpendicular to the b crystal axis. Due to the strong hydrogen bonds N1-H5···O1, a two-layer structure is formed, in which each pair of layers is connected to each other by weak hydrogen bonds of the C1-H1···O1 and C1-H1···O2.
γ-Glycine is trigonal, crystallizing in the chiral space groups P31 (144). Its crystal structure does not consist of layers. Molecules form helices around the 31 screw-axis coinciding with the c crystal axis, successive molecules being linked by N1-H5···O2 type hydrogen bonds. Hydrogen chains N1-H3···O2 type along the c axis, like alpha-glycine, are formed between each molecule and symmetry-equivalent molecules away along the helix, i.e., connected by one repetition of the lattice along the c crystal axis.
Analysis of single crystal XRD data indicates that the best refinement of the structure of γ-Gly:CA crystals can be achieved by considering twinning (twin components are related by screw axis 31). The crystals of α-Gly:CA do not demonstrate twinning. Comparison with the literature data indicates (Table 2) that in centrosymmetric α-Gly:CA (90:10), the unit cell volume Vcell = 309.43 (5) Å3 is little more than in pure α-Gly [Vcell = 309.00 (3) Å3] observed in [52], and smaller than Vcell = 310.10 (4) Å3 in [53]. In the magnitude of lattice parameters, CA doping manifests itself most clearly in decrease of c parameter and β angle in α-Gly:CA and a and c parameters in γ-Gly:CA (Table 2 and Table 3).
Comparison of bond lengths and angles of glycine molecules in CA-doped and pure α- and γ-glycine crystals indicates that glycine zwitterions (+NH3CH2CO2) in these crystals have close magnitudes of corresponding bond angles and bond lengths (Table 4 and Table 5). Nevertheless, some small differences, which can be attributed to CA doping, may be indicated. CA doping of α- and γ-Gly causes small increase of angle O1-C1-O2 and decrease of C2–C1–O1 that means the change of O1 atom position. However, doping does not influence other angles C2–C1–O2 and C1–C2–N1 in glycine molecules (Table 4 and Table 5).
Thus, CA doping of α- and γ-modifications of glycine single crystals does not lead to a change in symmetry and noticeable changes in the lattice parameters of the corresponding nominally pure glycine isomorphs. Like TGS:CA crystals [31], the incorporation of CA molecules into glycine crystals leads only to a slight modification of the unit cell parameters. In contrast to TGS:CA, the presence of twins is observed in noncentrosymmetric γ-Gly:CA crystals.
Note that both in TGS:CA and Gly:CA crystals, changes in bond angles and bond lengths were observed in glycine zwitterions. In TGS:CA, all bond angles in glycine zwitterions (+NH3CH2CO2) are changed to bring it closer to the N+H3CH2COOH [31] glycinium ion. Conversely, only a slight change in the position of O1 atoms in the carboxyl group and N atoms in the amino group of zwitterions is observed in Gly:CA, which indicates the formation of O…H–O and O…H–N types of hydrogen bonds between glycine zwitterions and croconic acid molecules.

3.2. Powder XRD Analysis

Phase analysis has led to a conclusion that both the measured XRD patterns of α-Gly:CA (90:10) and γ-Gly:CA (80:20) contain the XRD reflections attributed to one phase expected according to preparation, namely, α-Gly-like or γ-Gly-like phases (Table 1). However, a thorough inspection revealed that the reflections of the α-Gly:CA (90:10) and γ-Gly:CA (80:20) has an asymmetric shape, especially clearly seen for reflections non-overlapped with neighboring ones (see examples in small insets of Figure 3).
Recently, a similar asymmetric shape of powder XRD reflections of organic ferroelectric 2-methylbenzimidazole (MBI) has been successfully described in the framework of models of two or three crystalline phases with the same MBI structure but with slightly different unit cell parameters and differing microstructure parameters. The formation of the MBI modifications has been associated with penetration of solvent molecules (ethanol, acetone, deuterated acetone, or water) into the voids of the MBI crystal structure.
The structures of α-Gly:CA (90:10) and γ-Gly:CA (80:20) are also characterized by rather large pores with sizes of ~0.4 nm–0.6 nm (Figure 2). The CA molecules used during the synthesis of samples have a flat shape and a maximum size of ~0.5 nm. Based on these geometric observations, it can be assumed that CA molecules occupy the pores of structures during the sample synthesis. If CA molecules do not fill the lattice voids homogeneously, but only some regions, then the filled lattice regions will stretch, forming modifications of the same structure with slightly larger lattice parameters. The average size of such regions with approximately the same content of CA molecules will determine the size of the crystallites of these different modifications. Simultaneously, due to the entry of CA molecules into the pores and the neighborhood of crystallites with slightly different lattice parameters, the presence of local microstrains in the crystallites of different modifications is possible.
To verify these assumptions, which appear reasonable, the WHP and SSP methods at first stage were used to perform a profile analysis of XRD reflections (XRD Line profile analysis, LPA) observed in XRD patterns, assuming two phases with close parameters of unit cells (correspondingly, α1 and α2 modifications or γ1 and γ1 ones for α-Gly:CA (90:10) and γ- Gly:CA (80:20) XRD patterns). To do this, each observed reflection was separated into two reflections with the same known Miller indices hkl [for α-Gly:CA (90:10)] or Miller–Bravais indices hkil [for γ-Gly:CA (80:20)], where possible (some reflections, especially in the high-angle region of 2θ > ~45°, are highly overlapped, and it is difficult to extract them). The resulting WHP and SSP graphs are presented in the Figures S2 and S3 of Supplementary Materials. Table 6 presents the results of Celsiz and LPA calculations.
Based on the results of LPA, to perform the tasks set out in Section 2.3, the LB and Rietveld fittings of XRD patterns were performed in the next steps assuming the existence of two close-phase modifications in the samples (α1 and α2 in α-Gly:CA (90:10) and γ1 and γ2 in γ-Gly:CA (80:20)).
Final graphic results of Rietveld fitting of the simulated XRD patterns to experimental ones (with atomic coordinates of the both phase modifications fixed to the vales obtained in single-crystal refinement) for investigated powder samples are presented in Figure 3. The graphic results of LB fitting are provided in Figure S4 of Supplementary Materials.
Final values of agreement factors characterizing the quality of fitting simulated XRD patterns to experimental ones [weighted profile (Rwp) and profile (Rp) factors and their analogues after subtracting the background contribution (cRwp and cRp)] and the degree of correspondence of the structure model to the observed integral intensities of reflections attributed to crystalline phases (Bragg factors RB for crystalline phases) for LB and Rietveld fitting are summarized in Table 7. Definitions of the agreement factors can be found, for example, in Ref. [55].
When only one phase is considered to exist, and all other things during fittings being equal, the weight profile factors Rwp increase by ~1.7 times, and the profile factors Rp by more than 2 times compared to the factors provided in Table 7. In turn, the Bragg factor RB increases by ~3 times. Thus, the fitting of XRD patterns by LB and Rietveld methods unambiguously demonstrated that the assumption of only one phase does not allow us to describe XRD patterns and individual XRD reflections well enough. The powder samples contain at least two phases with a small difference in the unit cell parameters [α1 and α2 in α-Gly:CA (90:10) and γ1 and γ2 in γ-Gly:CA (80:20)] at least.
As noted above, the coordinates of the atoms in the Rietveld fitting were not refined. In the case of γ-Gly:CA (80:20), in addition, a structural model was obtained by measuring a single crystal at 100 K, while the powder XRD pattern of this sample was taken at 313 K. Nevertheless, as can be observed from Figure 3 and Figure S4 (Supplementary Materials), the quality of powder XRD pattern fittings is quite high not only by the LB method, but also by the Rietveld method. The high quality of the fittings of powder XRD patterns indicates the high quality of the results of the determination (refinement) of the structures of the samples performed on single crystals (which is expected), in addition to the absence of crystalline phases in the powder with the structures other than α-Gly-like and γ-Gly-like in α-Gly:CA (90:10) and γ-Gly:CA (80:20), respectively.
An attempt to refine the atomic coordinates for modifications of the α-Gly:CA (90:10) does not reveal a noticeable improvement in fitting quality (see Table 7 for agreement factors and Figure S5 of the Supplementary Materials). The values of the parameters of the unit cells of the modifications α-Gly:CA (90:10) and γ-Gly:CA (80:20) phases and the values of their microstructural parameters obtained in LB and Rietveld refinements practically coincide within one e.s.d. and are close (within two to three e.s.d.s) with values obtained independently by LPA methods (Table 6). Thus, for comparison with the results of single-crystal experiments (Table 2 and Table 3), it is possible to consider the results with more precise values, for example, ones obtained in LB fitting (Table 6).
The similarity of the values of the parameters and volume of the unit cells of α1 phase in powder and α-Gly:CA (90:10) single-crystal phase with the values for pure α-Gly according to the literature data [52,53] (cf. Table 2 and Table 6) indicates that in the phase α1 of powder and in the single-crystal α-Gly:CA (90:10), likely only a small number of CA molecules have penetrated into the voids of the structures. The same conclusion can be drawn from the comparison with the literature data [53,54] for pure γ-Gly, apparently about the powder phase γ1 and the single-crystal γ-Gly:CA (80:20) (cf. RT and low-temperature data in Table 3 and Table 6 for γ1 powder, γ-Gly:CA (80:20) single crystal and pure γ-Gly). Simultaneously, likely as a result of a larger number of CA molecules embedded in the pores of α2 and γ2 modifications, their parameters and volumes of unit cells are noticeably larger than, respectively, in α1 and γ1 (cf. Vcell = 311.31 (5) Å3 in α2 and 236.99 (10) Å3 in γ2, in comparison to Vcell = 309.85 (9) Å3 in α1 and 235.40 (12) Å3 in γ1 (LB fit)).
According to the refinement by the Rietveld method, the weight content of phases α1 and γ1 with a small number of embedded CA molecules in powders is significantly less than phases α2 and γ2 with an increased volume due to the introduction of a higher number of CA molecules [correspondingly, ~24 wt.% and ~33 wt.% in α1 and γ1 in comparison to ~76 wt.% and ~67 wt.% in α2 and γ2 (Table 7)]. Nevertheless, as noted above, studies of single crystals have revealed precisely phases that are close in terms of unit cell parameters α1 and γ1 (cf. Table 1, Table 2, Table 3 and Table 6). Apparently, this fact can be explained by the results of the determination of microstructural parameters (Table 6). According to the results of LPA and refinements by LB and Rietveld methods, modifications α1 and γ1 are characterized by large values of crystallite sizes in comparison with α2 and γ2 (cf. D~95 nm in α1 and γ1 and D~40 nm in α2 and γ2). As a result, the reflections of α1 and γ1 modifications are much narrower, and those of α2 and γ2 phases are more blurred (see small insets in Figure 3a,b). It is likely that single crystals with unit cell parameters close to α2 and γ2 were also characterized by more diffuse reflections and were rejected by quality when selected for single-crystal study.
Likely due to the embedding of CA molecules into structural pores and the neighborhood of crystallites of different modifications, microstrains are observed in the crystallites of α1 and γ1 phases, which are approximately the same size (εs~0.05 % (Table 6)). Simultaneously, in α2 crystallites with a more expanded crystal lattice, the absolute value of the average microstrain increases to εs~0.10%. However, microstrains were not detected in the crystallites of modifications γ2 also characterized by an increased volume of the unit cell, which requires further investigation.
Thus, XRD studies of α-Gly:CA (90:10) and γ-Gly:CA (80:20) powders revealed the presence in them of at least two crystalline modifications with the same α-Gly or γ-Gly type structure, respectively, but with slightly different unit cell parameters and noticeably different sizes of crystallites. The formation of two-phase modifications is associated with the embedding of different amounts of CA molecules into the pores of the α-Gly and γ-Gly structures.

3.3. Raman and FTIR Spectroscopies

Polarized Raman spectra measured in Gly:CA crystals of both pyramidal (Figure 4a) and plate-like shapes (Figure 4b) are like those of nominally pure α- and γ-Gly [56,57,58]. They consist of many lines, originated predominantly from intramolecular vibrations of different types. Analysis of the Raman and FTIR spectra makes it possible to ascertain the chemical structure and polymorphous modification of the molecular crystals studied. Table 8 and Table 9 portray the position of the lines in the polarized Raman and FTIR absorption spectra in pyramidal [γ-Gly:CA (80:20)] and plate-like [α-Gly:CA (90:10)] Gly:CA crystals, correspondingly. The positions of the lines in pyramidal Gly:CA crystals coincide with the positions of most lines in both nominally pure γ-Gly, and in plate-like α-Gly:CA crystals [56,57,58]. The shifts in the positions of high-frequency spectral lines at high frequency in each pair of crystals (α-Gly, α-Gly:CA) and (γ-Gly, γ-Gly:CA) do not exceed several cm−1. This indicates that the types and frequencies of molecular modes active in in Raman spectra of α- and γ-Gly doped with CA are close to those in pure Gly crystals. This is in agreement with the results of XRD analysis of Gly:CA single crystals and powders, which demonstrate very small changes in the crystal structure and position of atoms caused by doping with CA. Weak lines observed in α-Gly:CA spectrum portrayed in Figure 4c by red arrows are absent in pure Gly crystals and can be attributed to CA ions [59] (see Table 9).
The difference in polarized Raman spectra of γ- and α-Gly is most evident in the position of strong lines in the region of low (0–300 cm−1) and high (2800–3100 cm−1) frequencies of the spectrum. This difference is clearly observed in the Raman spectra of pyramidal and plate-like Gly:CA crystals (Figure 4a,b). At high frequencies (2800–3100 cm−1) the stretching symmetric and antisymmetric vibrations of CH2 group have been observed in α-Gly:CA at 2973 cm−1 and 3007 cm−1 and in γ-Gly:CA at 2963 cm−1 and 2999 cm−1 (Figure 5b), respectively, which corresponds to literature data related to pure α- and γ-Gly [58]. At low frequencies (0–300 cm−1), positions of lines in α- and γ-Gly:CA spectra are slightly shifted in compared to pure Gly crystals [56,57] (Tables S5 and S7 in Supplementary Materials), which may be attributed to CA doping.
As in the case of Raman spectra, a comparison of the FTIR absorption spectra of doped CA and nominally pure Gly indicates their close similarity (Table 8 and Table 9). However, in the case of α-Gly:CA (90:10) plates, additional lines in the photon energy range Eph = 0.5–0.6 eV (4000–5000 cm−1) were observed in the FTIR absorption spectra measured in transmission geometry. Note that similar lines have been observed in the FTIR spectra of TGS crystals doped with CA [31] (Figure 6). No additional lines were observed in the reflection geometry, most likely due to their low intensity. Note that we did not measure FTIR absorption spectra of γ-Gly crystals in transmission geometry because of their pyramidal shape.
So, polarized Raman and FTIR absorption spectra of Gly:CA single crystals indicates that addition of croconic acid to the aqueous solution which was used to grow Gly:CA crystals does not significantly affect the polymorphic modification of the crystals grown. It has been established that γ-Gly forms pyramidal crystals, while α-Gly forms plate-like crystals. The incorporation of CA molecules in a glycine crystal does not noticeably change the spectral positions of Raman lines in the high-energy region of the spectrum (2900–3100 cm−1), but leads to modest shift of several low-energy lines, which characterizes the effect of CA doping on the lattice vibration spectrum [32].

3.4. UV-Vis Absorption

RT absorption spectra of α-Gly:CA (90:10) and γ-Gly:CA (80:20), are portrayed in Figure 7. Colorless α- and γ-Gly crystals are transparent in wide spectral range. The energy band gaps of different polymorphic modifications of Gly crystals practically coincide and are about 5.1 eV [8,11,12]. Significant optical absorption is observed only for photons with energies Eph > 5 eV, which manifests itself in a sharp increase of absorbance at λ < 240 nm. The doping of Gly with CA is accompanied by the appearance of absorption in the transparent region of pure Gly. This absorption manifests itself in the form of a structured long-wavelength tail adjacent to the edge of the fundamental absorption of a Gly crystal. The traits of the absorption are clearly observed when considering the absorption in Tauc coordinates (αEph)2, Eph (Figure 7c,d) (where α is the absorption coefficient of the crystal). The shortest wavelength part of the absorption spectrum in region Eph = (5.8–6.2) eV in Figure 7c,d is well approximated by the linear dependence (solid black line), which can be regarded as the Tauc plot related to direct allowed interband transitions in crystals. The Tauc plot yields an estimation of the value of the optical bandgap as ~5.3 eV, which is close to the Eg value in pure Gly (~5.1 eV) (Figure 7c,d). Note that in Tauc coordinates, the shape of the absorption edge of undoped Gly is indeed well approximated by a straight line [8]. The presence of the additional absorption in the Gly:CA spectra (as compared with undoped Gly) can be associated with electronic transitions involving defects in the Gly lattice induced by the introduction of CA molecules. The contribution of these transitions to the absorption spectrum can be estimated as the difference between the absorption spectrum of Gly:CA (line 1 in the insets in Figure 7e,f) and the fragment of its fundamental absorption spectrum modeled by the Tauc plot (line 2 in the insets in Figure 7e,f). The absorption spectra thus obtained are portrayed in Figure 7e,f, which includes at least two absorption bands with maxima at ~5.2 and ~3.9 eV. It should be noted that similar structure is observed in the absorption spectrum of CA solution in ethanol (olive line in Figure 7e,f) or in water [60]. In Gly:CA crystals the absorption bands can presumably be associated with the creation of excitons localized at the lattice defects, including impurity CA molecules or neighboring Gly ones.
Table 10 presents a comparison of the Eg values, and high and low frequency local bands in α-Gly:CA (90:10), γ-Gly:CA (80:20), and TGS:CA (80:20) [31]. It is observed that CA doping results in practically the same changes of absorption spectra in these crystals.
Thus, doping of crystalline Gly with CA significantly affects the shape of the absorption edge of Gly, which manifests itself in smearing of the sharp absorption edge and the appearance of noticeable absorption at frequencies < Eg. Note that the same changes of absorption spectrum have been observed earlier in CA-doped triglycine sulfate crystals [31].

3.5. Photoluminescence

The photoluminescence (PL) of “pure” glycine crystals is very weak. As a result, in the emission spectrum of “pure” γ-Gly narrow Raman lines are clearly observed, whose intensity is comparable to that of the luminescence band (Figure 8a). Like pure glycine, solid croconic acid is also practically non-luminescent [61]. However, doping glycine crystals with croconic acid leads to a dramatic increase in the emission intensity, while the intensity of Raman scattering does not change (Figure 8b). Thus, the relatively bright luminescence of γ-Gly:CA is a distinctive characteristic of the doped crystals. A similar situation also occurs in the case of α- Gly:CA.
The RT PL spectra of α- and γ-Gly:CA crystals excited by light with λexc = 325 nm are portrayed in Figure 8b. The spectra of both γ- and α-modifications consist of broad structured emission bands with their maxima at ~2.50 eV (γ-Gly:CA) and ~2.56 eV (α-Gly:CA), respectively. Comparison of the PL spectra of pure and doped Gly crystals (Figure 8) portrays that the shapes of the PL bands are practically identical, but doping Gly with croconic acid leads mainly to a dramatic increase in the PL intensity. Considering that the band gap of γ-glycine crystals is about 5 eV [8,11,12,13] the optical electron transitions responsible for the emission involve localized electron states whose energy levels are located within the crystal band gap.
Figure 8c,d portrays the decomposition of the emission band profile into four Gaussian components both in the case of α-Gly:CA and γ-Gly:CA. The positions of the Gaussian components and their relative intensity in the PL spectra of α-Gly:CA (90:10), γ-Gly:CA (80:20), and TGS:CA (90:10) [31] are presented in Table 11. The observed emission-band structure implies the presence of at least four different emitting centers that contribute to the emission of Gly:CA crystals. Note that the expansion of the TGS:CA (90:10) PL spectrum in terms of the Gaussian component presented in our previous paper [31] contains only three components. Here, we present the results of more accurate calculations that include four Gaussian components which provide a better fit. Both in Gly:CA crystals and in TGS:CA crystals, the positions and normalized intensities of the Gaussian components in both Gly:CA and TGS:CA crystals are close, which demonstrates that in all the crystals compared the PL originates from the same emission centres. The similar PL behavior in CA-doped Gly and TGS crystals clearly indicates that the appearance of PL is not associated with the SO4 groups, which are present in TGS but absent in Gly crystals. The PL is probably the result of the interaction of CA with glycine ions and, most likely, with glycine zwitterions, since, as XRD phase analysis demonstrates, the structural parameters of these ions are affected by CA doping.
From Table 11 we observe that the normalized intensities of Gauss components may depend on crystal symmetry because bands 3 and 4 in non-centrosymmetric γ-Gly, and TGS:CA are more intensive than in centrosymmetric α-Gly:CA and band 1 has higher amplitude in α-Gly:CA.
Although the specific structure of these centers is currently not clear, it can be assumed that they include molecules of both glycine and croconic acid (glycine and CA crystals are not luminescent by themselves). The presence of an analogous emission band in the luminescence spectrum of “pure” γ-Gly (Figure 8a) with a small admixture of citric acid demonstrates that not only CA but also other organic acids can promote the formation of emitting centers (glycine lattice defects) with a similar structure of radiative transitions.

4. Conclusions

Crystals of α- and γ-glycine doped with croconic acid were grown from aqueous solutions of glycine and croconic acid. α-Gly crystals grown from a solution with a relative weight content Gly and CA of 90:10 have a plate shape. Pyramid-shaped γ-Gly crystals were obtained from a solution with a relative weight content of 98:2 and 80:20. A study of the structural and optical properties of the grown crystals has been performed. The study indicates that the crystal symmetry and morphology of CA-doped-doped glycine single crystals grown from aqueous solutions with different components weight ratios correspond well to well-known α- and γ-isomorphs of nominally pure glycine. Incorporation of CA molecules in the host glycine crystal structures does not change remarkably the parameters of the unit cells, but results in small changes of unit cell volumes, and bond angles and lengths of glycine zwitterions.
In many aspects, this situation is quite similar to that observed in CA-doped TGS single crystals [31]. CA doping both Gly and TGS crystals does not significantly change the frequencies of molecular vibrations in Raman and FTIR spectra. Also, it does not remarkably influence the magnitude of forbidden gap Eg, caused by direct allowed electronic transitions.
The most significant changes caused by CA doping glycine (Gly:CA), as in the case TGS crystals, are associated with absorption and photoluminescence spectra. In the UV Vis region CA doping results in appearance of a strong absorption band caused by local electronic transitions near the band gap (Eph ≈ 5.2–5.4 eV), and a broad band of lower intensity with a maximum in the region of Eph ≈ 3.7–3.9 eV. Weak absorption bands were detected by FTIR in TGS:CA and Gly:CA in the region of Eph ≈ 0.5–0.7 eV.
Strong PL spectra in α-, γ-Gly:CA, and TGS:CA crystals demonstrate the contributions of four emission centers emitting photons of close energies (~2.2 eV, 2.57 eV, 2.86 eV, 3.18 eV) and intensity ratios. It can be assumed that the appearance of these centers is due to the incorporation of small CA ions into the pores of the host crystal structure and the formation of hydrogen bonds between the CA molecule and glycine zwitterions.
In α- and γ-Gly:CA, in contrast to TGS:CA, different phases (regions) with a crystal structure of the same symmetry but slightly different unit cell parameters arising from the inhomogeneous doping distribution of the CA molecule in the host crystal were revealed by powder X-ray diffraction methods. The presence of twins in non-centrosymmetric γ-Gly:CA was detected using single-crystal XRD. These crystal structure details do not appear in optical spectroscopy experiments due to the integral nature of these methods.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst12101342/s1. Table S1. Results of refinement of the α-Gly:CA (90:10) structure (space group P21/n (14)) using single crystal XRD data. Table S2. Atomic displacement parameters for the α-Gly:CA (90:10) (Å2). Table S3. Results of refinement of the γ-Gly:CA (80:20) structure (space group P31 (N 144)) using single crystal XRD data. Table S4. Atomic displacement parameters for the γ-Gly:CA (80:20) (Å2). Figure S1. Structures of (a) α-Gly:CA (90:10) and (b) γ-Gly:CA (80:20) with atoms represented by thermal ellipsoids by means of program ORTEP using, correspondingly, the data presented in Tables S1–S4. Figure S2. (a), (c) WHP and (b), (d), (f) SSP of α1 and α2 modifications of α-Gly. Figure S3. (a), (c) WHP and (b), (d), (f) SSP of γ1 and γ2 modifications of γ-Gly. Figure S4. LB fit results for (a) α-Gly:CA (90:10 and (b) γ-Gly:CA (80:20). Figure S5. Rietveld fit results for α-Gly:CA (90:10). Table S5. Low frequency intermolecular modes and symmetry modes in Gly:CA plate in different experimental geometries, and in nominally pure α-Gly [5]. Table S6. High frequency intramolecular modes in Gly:CA plate in different experimental geometries, and in nominally pure α-Gly [6]. Table S7. Low frequency intermolecular modes in Gly:CA pyramid in different experimental geometries, and in nominally pure γ- Gly [7]. Table S8. High frequency intramolecular modes in Gly:CA pyramid.

Author Contributions

Formal analysis, E.B., A.Z., A.A.L., B.K., V.D., A.S. (Alexander Smirnov), A.S. (Anatoly Starukhin), S.P., H.Z., F.L. and H.K.; Investigation, E.B., A.Z., A.A.L., V.D., A.S. (Alexander Smirnov), A.S. (Anatoly Starukhin), S.P., H.Z., F.L. and H.K.; Writing—original draft, E.B., A.Z., A.A.L., A.S. (Anatoly Starukhin), B.K., H.Z., F.L. and H.K.; Writing—review and editing, E.B., B.K. and H.K.; Funding Crystals 2022, 12, 679 23 of 25 acquisition, E.B. and H.K.; Conceptualization, E.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by RFBR (project number 21-52-53015) and NSFC (project number 52111530040).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The X-ray experiment for TGS:CA (90:10) was carried out using facilities of the X-ray diffraction Resource Centers of St. Petersburg University. The XRD powder characterizations were performed using equipment and software of the Joint Research Center “Material science and characterization in advanced technology” (Ioffe Institute, St.-Petersburg, Russia).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vickery, H.B.; Schmid, C.A. The history of the discovery of the amino acids. Chem. Rev. 1931, 9, 169–318. [Google Scholar] [CrossRef]
  2. Towler, C.S.; Davey, R.J.; Lancaster, R.W.; Price, C.J. Impact of Molecular Speciation on Crystal Nucleation in Polymorphic Systems: The Conundrum of γ Glycine and Molecular ‘Self Poisoning’. J. Am. Chem. Soc. 2004, 126, 13347–13353. [Google Scholar] [CrossRef]
  3. Sakai, H.; Hosogai, H.; Kawakita, T. Transformation of α-glycine to γ-glycine. J. Cryst. Growth 1992, 116, 421–426. [Google Scholar] [CrossRef]
  4. Tylczyński, Z.; Busz, P. Transformation from γ to α modification in glycine crystal. Ph. Transit. 2014, 87, 1157–1164. [Google Scholar] [CrossRef]
  5. Dawson, A.; Allan, D.R.; Belmonte, S.A.; Clark, S.J.; David, W.I.F.; McGregor, P.A.; Parsons, S.; Pulham, C.R.; Sawyer, L. Effect of High Pressure on the Crystal Structures of Polymorphs of Glycine. Cryst. Growth Des. 2005, 5, 1415–1427. [Google Scholar] [CrossRef]
  6. Losev, E.A.; Mikhailenko, M.A.; Achkasov, A.F.; Boldyreva, E.V. The effect of carboxylic acids on glycine polymorphism, salt and co-crystal formation. A comparison of different crystallisation techniques. New J. Chem. 2013, 37, 1973–1981. [Google Scholar] [CrossRef]
  7. Fleck, M.; Petrosyan, A.M. Salts of Amino Acids. Crystallization, Structure and Properties; Springer International Publication: Cham, Switzerland, 2014. [Google Scholar]
  8. Flores, M.Z.S.; Freire, V.N.; dos Santos, R.P.; Farias, G.A. Optical absorption and electronic band structure first-principles calculations of α-glycine crystals. Phys. Rev. B 2008, 77, 115104. [Google Scholar] [CrossRef]
  9. Peter, M.E.; Ramasamy, P. Growth of gamma glycine crystal and its characterization. Spectrochim. Acta A 2010, 75, 1417–1421. [Google Scholar] [CrossRef]
  10. Ahamed, Z.A.S.; Dillip, G.R.; Raghavaiah, P.; Mallikarjuna, K.; Raju, B.D.P. Spectroscopic and thermal studies of γ-glycine crystal grown from potassium bromide for optoelectronic applications. Arab. J. Chem. 2013, 6, 429–433. [Google Scholar] [CrossRef]
  11. Dillip, G.R.; Raghavaiah, P.; Mallikarjuna, K.; Reddy, C.M.; Bhagavannarayana, G.; Kumar, V.R.; Raju, B.D.P. Crystal growth and characterization of γ-glycine grown from potassium fluoride for photonic applications. Spectrochim. Acta A 2011, 79, 1123–1127. [Google Scholar] [CrossRef]
  12. Rodríguez, J.S.; Costa, G.; da Silva, M.B.; Silva, B.P.; Honório, L.J.; de Lima-Neto, P.; Santos, R.C.R.; Caetano, E.W.S.; Alves, H.W.L.; Freire, V.N. Structural and Optoelectronic Properties of the α-, β-, and γ-Glycine Polymorphs and the Glycine Dihydrate Crystal: A DFT Study. Cryst. Growth Des. 2019, 19, 5204. [Google Scholar] [CrossRef]
  13. Kumar, R.A.; Vizhi, R.E.; Vijayan, N.; Babu, D.R. Structural, dielectric and piezoelectric properties of nonlinear optical γ-glycine single crystals. Phys. B 2011, 406, 2594–2600. [Google Scholar] [CrossRef]
  14. Justin, P.; Anitha, K. Influence of formic acid on optical and electrical properties of glycine crystal. Mater. Res. Express 2017, 4, 115101. [Google Scholar] [CrossRef]
  15. Tylczyński, Z.; Busz, P. Low-temperature phase transition in γ-glycine single crystal. Pyroelectric, piezoelectric, dielectric and elastic properties. Mater. Chem. Phys. 2016, 183, 254–262. [Google Scholar] [CrossRef]
  16. Guerin, S.; Stapleton, A.; Chovan, D.; Mouras, R.; Gleeson, M.; McKeown, C.; Noor, M.R.; Silien, C.; Rhen, F.M.F.; Kholkin, A.L.; et al. Control of piezoelectricity in amino acids by supramolecular packing. Nat. Mater. 2018, 17, 180–186. [Google Scholar] [CrossRef] [PubMed]
  17. Lemanov, V.V. Piezoelectric and pyroelectric properties of protein amino acids as basic materials of Soft State Physics. Ferroelectrics 2000, 238, 211–218. [Google Scholar] [CrossRef]
  18. Okosun, F.; Guerin, S.; Celikin, M.; Pakrashi, V. Flexible amino acid-based energy harvesting for structural health monitoring of water pipes. Cell Rep. Phys. Sci. 2021, 2, 100434. [Google Scholar] [CrossRef]
  19. Vijayalakshmia, V.; Dhanasekaran, P. Growth and characterization study of γ-glycine crystal grown using different mole concentrations of zinc sulphate as structure-directing agents. J. Cryst. Growth 2018, 498, 372–376. [Google Scholar] [CrossRef]
  20. Han, G.; Chow, P.S.; Tan, R.B.H. Understanding the Salt-Dependent Outcome of Glycine Polymorphic Nucleation. Pharmaceutics 2021, 13, 262. [Google Scholar] [CrossRef]
  21. Li, D.; Zong, S.; Dang, L.; Wang, Z.; Wei, H. Effects of inorganic additives on polymorphs of glycine in microdroplets. CrystEngComm 2018, 20, 164–172. [Google Scholar] [CrossRef]
  22. Bhat, M.N.; Dharmaprakash, S.M. Effect of solvents on the growth morphology and physical characteristics of nonlinear optical g-glycine crystals. J. Cryst. Growth 2002, 242, 245–252. [Google Scholar] [CrossRef]
  23. Balakrishnan, T.; Babu, R.R.; Ramamurthi, K. Growth, structural, optical and thermal properties of γ-glycine crystal. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2008, 69, 1114–1118. [Google Scholar] [CrossRef] [PubMed]
  24. Dhanaraj, P.V.; Rajesh, N.P. Growth and characterization of nonlinear optical γ-glycine single crystal from lithium acetate as solvent. Mater. Chem. Phys. 2009, 115, 413–417. [Google Scholar] [CrossRef]
  25. Azhagan, S.A.C.; Ganesan, S. Crystal growth, structural, optical, thermal and NLO studies of γ-glycine single crystals. Optik 2013, 124, 6456–6460. [Google Scholar] [CrossRef]
  26. Anis, M.; Ramteke, S.P.; Shirsat, M.D.; Muley, G.G.; Baig, M.I. Novel report on g-glycine crystal yielding high second harmonic generation efficiency. Opt. Mater. 2017, 72, 590–595. [Google Scholar] [CrossRef]
  27. Whatmore, R.W. Pyroelectric devices and materials. Rep. Prog. Phys. 1986, 49, 1335–1386. [Google Scholar] [CrossRef]
  28. Szafran, M.; Tylczyński, Z.; Wiesner, M.; Czarnecki, P.; Ghazaryan, V.V.; Petrosyan, A.M. Above-room-temperature ferroelectricity and piezoelectric activity of dimethylglycinium-dimethylglycine chloride. Mater. Des. 2022, 220, 110893. [Google Scholar] [CrossRef]
  29. Balashova, E.V.; Lemanov, V.V. Dielectric and Acoustic Properties of Some Betaine and Glycine Compounds. Ferroelectrics 2003, 285, 179–205. [Google Scholar] [CrossRef]
  30. Horiuchi, S.; Tokunaga, Y.; Giovannetti, G.; Picozzi, S.; Itoh, H.; Shimano, R.; Kumai, R.; Tokura, Y. Above-room-temperature ferroelectricity in a single-component molecular crystal. Nature 2010, 463, 789–792. [Google Scholar] [CrossRef]
  31. Balashova, E.; Levin, A.A.; Davydov, V.; Smirnov, A.; Starukhin, A.; Pavlov, S.; Krichevtsov, B.; Zolotarev, A.; Zhang, H.; Li, F.; et al. Croconic Acid Doped Triglycine Sulfate: Crystal Structure, UV-Vis, FTIR, Raman, Photoluminescence Spectroscopy, and Dielectric Properties. Crystals 2022, 12, 679. [Google Scholar] [CrossRef]
  32. Balashova, E.V.; Smirnov, A.N.; Davydov, V.Y.; Krichevtsov, B.B.; Starukhin, A.N. Raman scattering and luminescence in single crystals of the amino acid glycine C2H5NO2 with an admixture of croconic acid C5H2O5. J. Phys. Conf. Ser. 2021, 2103, 012070. [Google Scholar] [CrossRef]
  33. Braga, D.; Maini, L.; Grepioni, F. Crystallization from hydrochloric acid affords the solid-state structure of croconic acid (175 years after its discovery) and a novel hydrogen-bonded network. CrystEngComm. 2001, 3, 27. [Google Scholar] [CrossRef]
  34. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. A 2008, 64, 112–122. [Google Scholar] [CrossRef]
  35. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  36. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  37. Maunders, C.; Etheridge, J.; Wright, N.; Whitfield, H.J. Structure and microstructure of hexagonal Ba3Ti2RuO9 by electron diffraction and microscopy. Acta Crystallogr. B 2005, 61, 154–159. [Google Scholar] [CrossRef]
  38. Terlan, B.; Levin, A.A.; Börrnert, F.; Simon, F.; Oschatz, M.; Schmidt, M.; Cardoso-Gil, R.; Lorenz, T.; Baburin, I.A.; Joswig, J.-O.; et al. Effect of Surface Properties on the Microstructure, Thermal, and Colloidal Stability of VB2 Nanoparticles. Chem. Mater. 2015, 27, 5106–5115. [Google Scholar] [CrossRef]
  39. Terlan, B.; Levin, A.A.; Börrnert, F.; Zeisner, J.; Kataev, V.; Schmidt, M.; Eychmüller, A. A Size-Dependent Analysis of the Structural, Surface, Colloidal, and Thermal Properties of Ti1–xB2 (x = 0.03–0.08) Nanoparticles. Eur. J. Inorg. Chem. 2016, 6, 3460–3468. [Google Scholar] [CrossRef]
  40. Levin, A.A. Program SizeCr for Calculation of the Microstructure parameters from X-ray Diffraction Data. Preprint. 2022. Available online: https://www.researchgate.net/profile/Alexander-Levin-6/publication/359402043 (accessed on 5 June 2022). [CrossRef]
  41. Balashova, E.; Levin, A.A.; Fokin, A.; Redkov, A.; Krichevtsov, B. Structural Properties and Dielectric Hysteresis of Molecular Organic Ferroelectric Grown from Different Solvents. Crystals 2021, 11, 1278. [Google Scholar] [CrossRef]
  42. Brucker AXS GmbH. TOPAS; Version 5, Technical reference; Bruker AXS GmbH: Karlsruhe, Germany, 2014. [Google Scholar]
  43. le Bail, A.; Duroy, H.; Fourquet, J.L. Ab-initio structure determination of LiSbWO6 by X-ray powder diffraction. Mater. Res. Bull. 1988, 23, 447–452. [Google Scholar] [CrossRef]
  44. Rietveld, H.M. Line profiles of neutron powder-diffraction peaks for structure Refinement. Acta Crystallogr. 1967, 22, 151–152. [Google Scholar] [CrossRef]
  45. Dollase, W.A. Correction of Intensities for Preferred Orientation in Powder Diffractometry: Application of the March Model. J. Appl. Crystallogr. 1986, 19, 267–272. [Google Scholar] [CrossRef]
  46. Järvinen, M. Application of symmetrized harmonics expansion to correction of the preferred orientation effect. J. Appl. Cryst. 1993, 26, 525–531. [Google Scholar] [CrossRef]
  47. Bérar, J.-F.; Lelann, P.J. E.s.d.’s and Estimated Probable Error Obtained in Rietveld Refinements with Local Correlations. J. Appl. Crystallogr. 1991, 24, 1–5. [Google Scholar] [CrossRef]
  48. Levin, A.A. Program RietESD for Correction of Estimated Standard Deviations Obtained in Rietveld-Refinement Program. Preprint. 2022. Available online: https://www.researchgate.net/profile/Alexander-Levin-6/publication/359342753 (accessed on 5 June 2022). [CrossRef]
  49. Andreev, Y.G. The Use of the Serial-Correlations Concept in the Figure-of-Merit Function for Powder Diffraction Profile Fitting. J. Appl. Crystallogr. 1994, 27, 288–297. [Google Scholar] [CrossRef]
  50. Balashova, E.V.; Svinarev, F.B.; Zolotarev, A.A.; Levin, A.A.; Brunkov, P.N.; Davydov, V.Y.; Smirnov, A.N.; Redkov, A.V.; Pankova, G.A.; Krichevtsov, B.B. Crystal structure, Raman spectroscopy and dielectric properties of new semi-organic crystals based on 2-methylbenzimidazole. Crystals 2019, 9, 573. [Google Scholar] [CrossRef]
  51. Farrugia, L.J. ORTEP-3 for Windows - a version of ORTEP-III with a Graphical User Interface (GUI). J. Appl. Crystallogr. 1997, 30, 565. [Google Scholar] [CrossRef]
  52. Langan, P.; Mason, S.A.; Myles, D.; Schoenborn, B.P. Structural characterization of crystals of α-glycine during anomalous electrical behaviour. Acta Crystallogr. B 2002, B58, 728–733. [Google Scholar] [CrossRef] [Green Version]
  53. Boldyreva, E.V.; Drebushchak, T.N.; Shutova, E.S. Structural distortion of the a, b, and g polymorphs of glycine on cooling. Z. Kristallogr. Cryst. Mater. 2003, 218, 366–376. [Google Scholar] [CrossRef]
  54. Aree, T.; Bürgi, H.-B.; Chernyshov, D.; Törnroos, K.W. Dynamics and Thermodynamics of Crystalline Polymorphs. 3. γ-Glycine, Analysis of Variable Temperature Atomic Displacement Parameters, and Comparison of Polymorph Stabilities. J. Phys. Chem. A 2014, 118, 9951–9959. [Google Scholar] [CrossRef]
  55. Hill, R.J.; Fischer, R.X. Profile agreement indices in Rietveld and pattern-fitting analysis. J. Appl. Crystallogr. 1990, 23, 462–468. [Google Scholar] [CrossRef]
  56. Surovtsev, N.V.; Malinovsky, V.K.; Boldyreva, E.V. Raman study of low-frequency modes in three glycine polymorphs. J. Chem. Phys. 2011, 134, 045102. [Google Scholar] [CrossRef] [PubMed]
  57. Andrews, B.; Torrie, B.H.; Powell, B.M. Intermolecular potentials for alpha-glycine from Raman and infrared scattering measurements. Biophys. J. 1983, 41, 293–298. [Google Scholar] [CrossRef]
  58. Baran, J.; Ratajczak, H. Polarised IR and Raman spectra of the γ-glycine single crystal. Spectrochim. Acta A 2005, 61, 1611–1626. [Google Scholar] [CrossRef]
  59. Colmenero, F.; Escribano, R. Thermodynamic, Raman Spectroscopic, and UV–Visible Optical Characterization of the Deltic, Squaric, and Croconic Cyclic Oxocarbon Acids. J. Phys. Chem. A 2019, 123, 4241–4261. [Google Scholar] [CrossRef] [PubMed]
  60. Yamada, K.; Mizuno, N.; Hirata, Y. Structure of Croconic Acid. Bull. Chem. Soc. Jap. 1958, 31, 543–549. [Google Scholar] [CrossRef]
  61. Feng, S.; Yang, H.; Jiang, X.; Wang, Y.; Zhu, M. A New 3D Silver(I) Coordination Polymer with Croconate Ligand Displaying Green Luminescent. J. Mol. Struct. 2015, 1081, 1–5. [Google Scholar] [CrossRef]
Figure 1. Images of crystals used in this research. α-Gly crystals grown from aqueous solution (a). γ-Gly crystal grown from aqueous solution with adding 5% of citric acid C6H8O7 (b). Croconic acid crystal (c). Gly:CA crystals grown from solutions with relative weight ratios of Gly and CA 80:20 (d), 90:10 (e), 98:2 (f).
Figure 1. Images of crystals used in this research. α-Gly crystals grown from aqueous solution (a). γ-Gly crystal grown from aqueous solution with adding 5% of citric acid C6H8O7 (b). Croconic acid crystal (c). Gly:CA crystals grown from solutions with relative weight ratios of Gly and CA 80:20 (d), 90:10 (e), 98:2 (f).
Crystals 12 01342 g001
Figure 2. Crystal structures of (a) α-Gly:CA (90:10) and (b) γ-Gly:CA (80:20). Selected interatomic distances are marked in (a) for α-Gly:CA (90:10) (1: 0.44 nm, 2: 0.53 nm, 3: 0.45 nm, 4: 0.49 nm, 5: 0.50 nm) and in (b) for γ-Gly:CA (80:20) (1: 0.60 nm, 2: 0.55 nm; 3: 0.54 nm; 4: 0.69 nm). Maximum size of flat CA molecule is ~0.5 nm.
Figure 2. Crystal structures of (a) α-Gly:CA (90:10) and (b) γ-Gly:CA (80:20). Selected interatomic distances are marked in (a) for α-Gly:CA (90:10) (1: 0.44 nm, 2: 0.53 nm, 3: 0.45 nm, 4: 0.49 nm, 5: 0.50 nm) and in (b) for γ-Gly:CA (80:20) (1: 0.60 nm, 2: 0.55 nm; 3: 0.54 nm; 4: 0.69 nm). Maximum size of flat CA molecule is ~0.5 nm.
Crystals 12 01342 g002
Figure 3. Rietveld fit results for (a) α-Gly:CA (90:10) and (b) γ-Gly:CA (80:20). During the Rietveld fit the atomic coordinates of the phases were fixed to values determined in single-crystal investigation of this work and not refined. The Miller indices hkl and Miller–Bravais indices hkil of some selected reflections are indicated in (a) and (b), respectively. Large insets in (a,b) portray the quality of fitting the high 2θ angle regions of the XRD patterns in a larger scale. Small insets present examples of Rietveld fitting the asymmetric non-overlapping reflections.
Figure 3. Rietveld fit results for (a) α-Gly:CA (90:10) and (b) γ-Gly:CA (80:20). During the Rietveld fit the atomic coordinates of the phases were fixed to values determined in single-crystal investigation of this work and not refined. The Miller indices hkl and Miller–Bravais indices hkil of some selected reflections are indicated in (a) and (b), respectively. Large insets in (a,b) portray the quality of fitting the high 2θ angle regions of the XRD patterns in a larger scale. Small insets present examples of Rietveld fitting the asymmetric non-overlapping reflections.
Crystals 12 01342 g003
Figure 4. Polarized Raman spectra of pyramidal (a) and plate-like (b) Gly:CA crystals after subtracting the photoluminescence background. Fragment of the plate-like Gly:CA spectrum on an enlarged scale (c). For pyramidal crystals, the Z-axis is parallel to the polar axis (the height of the pyramid); for plate-like crystals, the Y-axis is parallel to the C2 axis. The red arrows in (c) indicate positions of lines observed in CA crystals. λexc = 532 nm.
Figure 4. Polarized Raman spectra of pyramidal (a) and plate-like (b) Gly:CA crystals after subtracting the photoluminescence background. Fragment of the plate-like Gly:CA spectrum on an enlarged scale (c). For pyramidal crystals, the Z-axis is parallel to the polar axis (the height of the pyramid); for plate-like crystals, the Y-axis is parallel to the C2 axis. The red arrows in (c) indicate positions of lines observed in CA crystals. λexc = 532 nm.
Crystals 12 01342 g004
Figure 5. Comparison of Raman spectra of Gly:CA crystals of pyramidal (blue line) and plate-like (red line) shapes in low-energy (a) and high-energy (b) spectral regions.
Figure 5. Comparison of Raman spectra of Gly:CA crystals of pyramidal (blue line) and plate-like (red line) shapes in low-energy (a) and high-energy (b) spectral regions.
Crystals 12 01342 g005
Figure 6. FTIR absorption spectra in spectral region Eph = 0.5–0.6 eV in Gly:CA (90:10) measured in this work and in previously studied TGS:CA (90:10) (see text).
Figure 6. FTIR absorption spectra in spectral region Eph = 0.5–0.6 eV in Gly:CA (90:10) measured in this work and in previously studied TGS:CA (90:10) (see text).
Crystals 12 01342 g006
Figure 7. Absorbance spectra of (a) α-Gly:CA (90:10), (b) γ-Gly:CA (80:20), (c) α-Gly:CA (90:10), and (d) γ-Gly:CA (80:20) in Tauc coordinates. (e,f) Change in the Gly crystal absorption caused by incorporating CA molecules into the Gly crystals. The olive lines represent the absorption spectrum of CA solution in ethanol [60]. The inset portrays the absorption spectra of doped and pure Gly crystals (see text).
Figure 7. Absorbance spectra of (a) α-Gly:CA (90:10), (b) γ-Gly:CA (80:20), (c) α-Gly:CA (90:10), and (d) γ-Gly:CA (80:20) in Tauc coordinates. (e,f) Change in the Gly crystal absorption caused by incorporating CA molecules into the Gly crystals. The olive lines represent the absorption spectrum of CA solution in ethanol [60]. The inset portrays the absorption spectra of doped and pure Gly crystals (see text).
Crystals 12 01342 g007
Figure 8. (a) Emission spectrum of “pure” γ-glycine wih a small admixture of citric acid. Here, R labels the Raman lines. (b) Emission spectra of γ- and α-glycine crystals doped with croconic acid. (c) Decomposition of the α-Gly:CA emission band profile into four Gaussian components. exc = 3.8 eV. T = 300 K. (d) Decomposition of the γ-Gly:CA emission band profile into four Gaussian components. The Gaussian components in (c,d) are shown by different colors.
Figure 8. (a) Emission spectrum of “pure” γ-glycine wih a small admixture of citric acid. Here, R labels the Raman lines. (b) Emission spectra of γ- and α-glycine crystals doped with croconic acid. (c) Decomposition of the α-Gly:CA emission band profile into four Gaussian components. exc = 3.8 eV. T = 300 K. (d) Decomposition of the γ-Gly:CA emission band profile into four Gaussian components. The Gaussian components in (c,d) are shown by different colors.
Crystals 12 01342 g008
Table 1. Details of crystal data and structure refinement of α-Gly:CA (90:10) and γ-Gly:CA (80:20) single crystals.
Table 1. Details of crystal data and structure refinement of α-Gly:CA (90:10) and γ-Gly:CA (80:20) single crystals.
α-Gly:CA (90:10)γ-Gly:CA (80:20)
Chemical formulaC2H5NO2C2H5NO2
Formula weight, Da75.0775.07
Space groupP21/n (14)P31 (144)
a, Å5.1004 (5)6.9848 (3)
b, Å11.9664 (9)6.9848 (3)
c, Å5.4570 (5)5.4744 (2)
β, °111.714 (3)90
Vcell, Å3309.43 (5)231.30 (2)
Z43
Dcalc, g∙cm−31.6111.617
F (000)160.0120.0
μ, mm−10.1431.25
Radiation (λ, Å)Mo-Kα (0.71073)Cu-Kα (1.54184)
Θ max, °26.3577.18
h, k, l max6, 14, 68, 8, 6
Reflections collected32403120
Independent reflections629645
Data/restraints/parameters629/0/47645/1/48
GOOF1.1581.081
Final R indexes [Reflections I ≥ 2σ (I)]R1 = 0.0345, wR2 = 0.1109 [629]R1 = 0.0299, wR2 = 0.0735 [645]
Final R indexes [Reflections all]R1 = 0.0416, wR2 = 0.1173 [538]R1 = 0.0299, wR2 = 0.0735 [645]
Largest difference peak/hole,
e∙Å−3
0.19/−0.190.16/−0.20
Temperature of measurements, K298100
Table 2. Comparison of unit cell parameters a, b, c, angle β, and unit cell volume Vcell in α-Gly:CA (90:10) and nominally pure α-Gly crystals from [52,53].
Table 2. Comparison of unit cell parameters a, b, c, angle β, and unit cell volume Vcell in α-Gly:CA (90:10) and nominally pure α-Gly crystals from [52,53].
α-Gly:CA (90:10)
298 K
α-Gly [53]
294 K
α-Gly [52]
301 K
a, Å5.1004 (5)5.1047 (3)5.0999 (3)
b, Å11.9664 (9)11.9720 (14)11.9516 (6)
c, Å5.4570 (5)5.4631 (3)5.4594 (3)
β, °111.714 (3)111.740 (5)111.781 (2)
Vcell, Å3309.43 (5)310.10 (4)309.00 (3)
Table 3. Comparison of unit cell parameters a, c, and unit cell volume Vcell in γ-Gly:CA (80:20) and nominally pure γ-Gly crystals from [53,54].
Table 3. Comparison of unit cell parameters a, c, and unit cell volume Vcell in γ-Gly:CA (80:20) and nominally pure γ-Gly crystals from [53,54].
γ-Gly:CA (80:20)
100 K
γ-Gly [53] 150 Kγ-Gly [54]
100 K
γ-Gly [53]
294 K
γ-Gly [54]
300 K
a, Å6.9848 (3)6.998 (16)6.9869 (1)7.0383 (7)7.0402 (1)
c, Å5.4744 (2)5.4784 (18)5.4768 (1)5.4813 (8)5.4813 (1)
Vcell, Å3231.30 (2)232.5 (1)231.540 (6)235.15 (3)235.280 (6)
Table 4. Bond angles and bond lengths C1–O1, C1–O2, C2–N1, and C1–C2 of glycine molecules in CA-doped and nominally pure α-Gly crystals [52,53].
Table 4. Bond angles and bond lengths C1–O1, C1–O2, C2–N1, and C1–C2 of glycine molecules in CA-doped and nominally pure α-Gly crystals [52,53].
α-Gly:CA (90:10)α-Gly [53]α-Gly [52]
bond angles, °
O1–C1–O2125.86 (15)125.62 (7)125.45 (7)
C2–C1–O1116.69 (15)116.93 (7)117.16 (7)
C2–C1–O2117.44 (14)117.44 (6)117.38 (6)
C1–C2–N1111.78 (13)111.70 (6)111.83 (5)
bond lengths, Å
C1–O11.251 (2)1.2549 (9)1.2512 (10)
C1–O21.247 (2)1.2517 (9)1.2480 (10)
C2–N11.475 (2)1.4777 (10)1.4743 (8)
C1–C21.526 (2)1.5268 (9)1.5252 (9)
Table 5. Bond angles and bond lengths C1–O1, C1–O2, C2–N1, and C1–C2 of glycine molecule in CA-doped and nominally pure γ-Gly crystals [53,54].
Table 5. Bond angles and bond lengths C1–O1, C1–O2, C2–N1, and C1–C2 of glycine molecule in CA-doped and nominally pure γ-Gly crystals [53,54].
γ-Gly:CA (80:20)
100K
γ-Gly [53] 150 Kγ-Gly [54]
100 K
bond angles, °
O1–C1–O2126.0 (3)125.3 (2)125.70 (7)
C2–C1–O1116.4 (2)116.8 (2)116.61 (7)
C2–C1–O2117.6 (3)117.8 (2)117.63 (6)
C1–C2–N1111.6 (2)111.2 (2)111.55 (6)
bond lengths, Å
C1–O11.260 (4)1.263 (3)1.2572 (9)
C1–O21.248 (4)1.251 (3)1.2487 (9)
C2–N11.479 (4)1.480 (3)1.4719 (10)
C1–C21.524 (4)1.529 (3)1.5249 (9)
Table 6. Results of LPA (WHP and SSP) and LB and Rietveld fitting of the α-Gly:CA (90:10) and γ-Gly:CA (80:20) XRD powder patterns measured at 313 ± 1 K.
Table 6. Results of LPA (WHP and SSP) and LB and Rietveld fitting of the α-Gly:CA (90:10) and γ-Gly:CA (80:20) XRD powder patterns measured at 313 ± 1 K.
a, Å
b, Å
c, Å
β, °
Vcell, ÅD (nm)
εs, %
α-Gly:CA (90:10), LPA
phase α1
5.1050 (16) a
11.9771 (40) a
5.4614 (17) a
111.74 (2) a
310.18 (18)52 (9)/78 (14) b
0.07 (6)/0.16 (4) b
phase α2
5.1216 (20) a
12.0064 (64) a
5.4715 (22) a
111.70 (3) a
312.61 (25)44 (3)/54 (3) b
0.06 (4)/0.14 (2) b
α-Gly:CA (90:10), LB fitting
phase α1
5.1002 (10) c
11.9709 (6) c
5.4613 (10) c
111.68 (2) c
309.85 (9)96 (4) c,d
0.050 (6) c,d
phase α2
5.1107 (5) c
12.0102 (8) d
5.4694 (6) c
111.70 (1) c
311.31 (5)37.4 (2) c,d
0.11 (2) c,d
α-Gly:CA (90:10), Rietveld fitting (without refinement of atomic coordinates)
phase α1
5.1000 (16) c
11.9704 (8) c
5.4613 (9) c
111.71 (2) c
399.76 (12)98 (5) c,d
0.050 (6) c,d
phase α2
5.1107 (5) c
12.0102 (8) d
5.4694 (6) c
111.70 (1) c
311.31 (5)38.4 (6) c,d
0.11 (1) c, d
α-Gly:CA (90:10), Rietveld fitting (with refinement of atomic coordinates)
phase α1
5.0999 (10) c
11.9701 (7) c
5.4613 (10) c
111.71 (2) c
309.74 (10)98 (5) c,d
0.050 (6) c,d
phase α2
5.1109 (7) c
12.0102 (10) d
5.4684 (7) c
111.69 (2) c
311.90 (8)38.9 (5) c,d
0.12 (1) c,d
γ-Gly:CA (80:20), LPA
phase γ1
7.0388 (6) a
a
5.4797 (9) a
90
235.12 (4)70 (12)/82 (7) b
0/0.057 (16) b
phase γ2
7.0541 (14) a
a
5.4890 (18) a
90
236.54 (9)48 (8)/48 (8) b
0/0b
γ-Gly:CA (80:20), LB fitting
phase γ1
7.0404 (4) c
a
5.4838 (28) c
90
235.40 (12)92 (3) c,d
0.044 (16) c,d
phase γ2
7.0600 (7) c
a
5.4903 (21) c
90
236.99 (10)35.7 (7) c,d
0c
γ-Gly:CA (80:20), Rietveld fitting (without refinement of atomic coordinates)
phase γ1
7.0402 (5) c
a
5.4817 (35) c
90
235.30 (15)92 (4) c,d
0.051 (17) c,d
phase γ2
7.0602 (9) c
a
5.4903 (9) c
90
237.01 (5)35.7 (9) c,d
0 c
a calculated by program Celsiz. b calculated by WHP/SSP techniques using program SizeCr. c Parameters portrayed are corrected by multiplication of the e.s.d. values obtained in LB/Rietveld fitting refinement on me.s.d. factor (Table 7). d D = Lvol_FWHM and εs (%) = 2∙e0∙100%, where Lvol_FWHM and e0 are, respectively, crystallite size and microstrain parameters refined by TOPAS (see thorough explanation with examples in Refs. [40,41]).
Table 7. Agreement factors obtained in Rietveld and LB fitting (refinement), weight content Wt of crystalline phases p1 (α1 or γ1) and p2 (α2 or γ2) according to results of Rietveld fitting and the factor me.s.d. for correction of e.s.d.s of refined parameters calculated by RietESD.
Table 7. Agreement factors obtained in Rietveld and LB fitting (refinement), weight content Wt of crystalline phases p1 (α1 or γ1) and p2 (α2 or γ2) according to results of Rietveld fitting and the factor me.s.d. for correction of e.s.d.s of refined parameters calculated by RietESD.
Rwp, %
Rp, %
cRwp, % a
cRp, % a
RB, % p1
RB, % p2
Wt, wt.% p1
Wt, wt.% p2
me.s.d.a
α-Gly:CA (90:10), LB fitting
4.82
3.09
7.04
5.24
-
-
-
-
5.05
α-Gly:CA (90:10), Rietveld fitting (without refinement of atomic coordinates)
6.35
4.17
9.42
7.36
0.98
1.13
24.4 (1) b
75.6 (1) b
5.24
α-Gly:CA (90:10), Rietveld fitting (with refinement of atomic coordinates)
5.77
3.87
8.54
6.82
0.71
0.58
23.4 (1) b
76.6 (1) b
5.33
γ- Gly:CA (80:20), LB fitting
9.96
6.21
13.14
9.58
-
-
-
-
6.97
γ-Gly:CA (80:20), Rietveld fitting (without refinement of atomic coordinates)
11.36
7.40
14.97
11.31
1.94
2.92
33.2 (5) b
66.8 (5) b
7.09
a Parameters calculated by program RietESD (see Supplementary Materials). b E.s.d. of Wt presented is corrected by multiplication of the e.s.d. value obtained in Rietveld fitting refinement on me.s.d. factor.
Table 8. Positions of lines (ν, cm−1) observed in polarized Raman and FTIR spectra of γ-Gly:CA (80:20) crystals and assignments of analogous lines in nominally pure γ-Gly from Ref. [58].
Table 8. Positions of lines (ν, cm−1) observed in polarized Raman and FTIR spectra of γ-Gly:CA (80:20) crystals and assignments of analogous lines in nominally pure γ-Gly from Ref. [58].
Raman (ν, cm−1)FTIR
(ν, cm−1)
Assignments [57]
X(zz)XX(yy)XX(yz)X
-3095.4 m-3100 mνNH (3)···O (1)
-2999.1 s2999.1 s-νaCH2
2963.7 s2962.6 vs2964.8 m2963 wνsCH2
2849.6 w----
2744.5 vw---νNH (1)···O (1)
2695.7 vw---νNH (1)···O (1)
2611.6 w---νNH (1)···O (1)
---1736 w-
-1676.1 w-1684m-
1659.5 w--1665 wδaNH3+
---1630 sδaNH3+
---1601 vsνaCOO
--1589.8 w--
-1576.5 m--νaCOO (?)
---1557 m-
-1506.7 m-1506 m-
--1502.3 vw1499 sδsNH3+
1483.5 vw---δsNH3+
1438.1 w1437 s1440.3 m1439 wδCH2
1389.4 vw1393.8 s-1398 mνsCOO
-1335.2 s-1344 mωCH2
--1321.9 m1321 wτCH2
-1154.7 w1156.9 vw1153 mρNH3+
-1135.9 m1130.3 w-ρNH3+
---1126 sρNH3+
-1047.3 w1048.4 m1049 vwνaCCN
924.47 sh931.11 w931.6 sρCH2
892.37 s892.37 s892.37 w889.2 mνCC
749.55 vw----
683.13 m688.67 vw687.56 w692.4 mσCOO
--̶667.37 m-
607.85 w-607.85 w--
560.24 vw-558.03 w--
--517.07 w-ωCOO
502.68 s---τCOO
--500.46 m-ρCOO
354.33 m354.33 w359.86 w-ρCCN
-213.,73 m215.94 w--
-171.66 sh-ω or τ CCN
174.98 w---Lattice modes
-165.02 sh---“-
-149.52 s---“-
--138.45 s--“-
-127.38 sh---“-
108.56 vw101.92 m103.02 m--“-
90.9 vw86.42 vs87.528 m--“-
Table 9. Positions of lines (ν, cm−1) observed in polarized Raman and FTIR spectra of α-Gly:CA (90:10) crystals and assignments of analogous lines in nominally pure α-Gly from Ref. [58].
Table 9. Positions of lines (ν, cm−1) observed in polarized Raman and FTIR spectra of α-Gly:CA (90:10) crystals and assignments of analogous lines in nominally pure α-Gly from Ref. [58].
X(ZZ)XX(YY)XX(YZ)XFTIRAssignments [58]
3171 sνNH···O
-3143 m--νNH···O
3008 vs3007 s3009 s3005 mνaCH2
2972 vs2973 vs2973 s2974 mνsCH2
-2903 w-2918 wνNH···O
2883 vw2880 w---“-
-2818 vw---“-
2787 vw2746 vw---“-
---2718 s-“-
2608 vw2633 vw-2627 s-“-
---2525 m-“-
---2363 m-
1723 vw1722 vw--CA [59]: ν (CO)
--1673 m1647 vsδaNH3+
---1611 sδaNH3+
---1555 mνaCOO
---1541 m-
1502 m1506 vw-1508 mδsNH3+
1483 w--1474 wδsNH3+
--1455 s1458δCH2
-1439 m-1445 mδCH2
1411 s1410 m1416 w1421 sνsCOO
---1341 mω or τCH2
1327 s1325 s1325 m1314 mτCH2
1138 m1137 w-1134 wρNH3+
-1106 w-1115 wρNH3+
1037 w--1034 wνaCCN
---914 sρCH2
---897 m-
892.8 s892.4 s892.4 m-νCC
802 vw---CA [59]: δ (CO)
695.9 m---σCOO
643 vw639 vw--CA [59]: Ring breath.
-600. vw602.3 m-̶
496.9 w-483.9 m-ρCOO
357.2 m-̶-ρCCN
196.3 m197.1 m̶-Lattice modes
--178.3 m--“-
-161.7 w163.9 sh--“-
154 s----“-
139.2m----“-
--107.5 s--“-
98.95 s99.71 s---“-
73.54 s78.67 sh---“-
--70.92 s--“-
50.26 s51 s53.21 w--“-
Table 10. Energy gap Eg, eV, photon energy Eph of high- and low-energy local bands and normalized amplitude of low energy band in absorption spectra of α-Gly:CA (90:10), γ-Gly:CA (80:20) and TGS:CA (80:20) [31].
Table 10. Energy gap Eg, eV, photon energy Eph of high- and low-energy local bands and normalized amplitude of low energy band in absorption spectra of α-Gly:CA (90:10), γ-Gly:CA (80:20) and TGS:CA (80:20) [31].
Eg, eVHigh Energy Local band Eph, eVLow Energy Local Band Eph, eVNormalized Amplitude of Low Energy Band
α-Gly:CA (90:10)5.25.253.90.16
γ-Gly:CA (80:20)5.45.333.70.07
TGS:CA (80:20)5.25.393.90.2
Table 11. Resonance photon energies Eph, eV, normalized intensity Ai/A2 obtained after decomposition of PL spectra in α-Gly:CA, γ-Gly:CA and TGS:CA in four Gauss components.
Table 11. Resonance photon energies Eph, eV, normalized intensity Ai/A2 obtained after decomposition of PL spectra in α-Gly:CA, γ-Gly:CA and TGS:CA in four Gauss components.
α-Gly:CAγ-Gly:CATGS:CA
Band, iEph, eVAi/A2Eph, eVAi/A2Eph, eVAi/A2
12.200.662.180.472.200.59
22.5412.5712.571
32.860.202.880.242.860.33
43.180.133.170.223.190.22
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Balashova, E.; Levin, A.A.; Davydov, V.; Smirnov, A.; Starukhin, A.; Pavlov, S.; Krichevtsov, B.; Zolotarev, A.; Zhang, H.; Li, F.; et al. Croconic Acid Doped Glycine Single Crystals: Growth, Crystal Structure, UV-Vis, FTIR, Raman and Photoluminescence Spectroscopy. Crystals 2022, 12, 1342. https://doi.org/10.3390/cryst12101342

AMA Style

Balashova E, Levin AA, Davydov V, Smirnov A, Starukhin A, Pavlov S, Krichevtsov B, Zolotarev A, Zhang H, Li F, et al. Croconic Acid Doped Glycine Single Crystals: Growth, Crystal Structure, UV-Vis, FTIR, Raman and Photoluminescence Spectroscopy. Crystals. 2022; 12(10):1342. https://doi.org/10.3390/cryst12101342

Chicago/Turabian Style

Balashova, Elena, Aleksandr A. Levin, Valery Davydov, Alexander Smirnov, Anatoly Starukhin, Sergey Pavlov, Boris Krichevtsov, Andrey Zolotarev, Hongjun Zhang, Fangzhe Li, and et al. 2022. "Croconic Acid Doped Glycine Single Crystals: Growth, Crystal Structure, UV-Vis, FTIR, Raman and Photoluminescence Spectroscopy" Crystals 12, no. 10: 1342. https://doi.org/10.3390/cryst12101342

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