3. Results and Discussion
MS is in
Supplementary Information (SI); Figure S1 showed the relative abundance of the fragmented molecular ion versus
ratio peaked at
m/
z 73.37, 75.2, 73.32, and 70.29, corresponding to molecular weight (M.wt) 75.07 g mol
−1 of glycine. The last peaks correspond to M.w. 614.32, 662.38, 758.57, 834.65 g mol
−1 for glycine–copper sulphate (GCS), glycine–copper sulphate doped 10 ppm SNPs (GCSN1), glycine–copper sulphate doped 20 ppm SNPs (GCSN2), and glycine–copper sulphate doped 30 ppm SNPs (GCSN3) crystals, respectively. An increasing concentration of SNPs increased chelation of glycine organic ligand to Cu(II) ion in CuSO
4.
EDX spectra and SEM micrographs:
Figure 2a–d show SEM-EDX analysis of GCS, GCSN1, GCSN2, and GCSN3 crystals, respectively. These spectra were produced as a focused electron beam on the sample ejected electrons from the inner-most electron atoms in the crystal leaving holes filled by ejected electrons from higher level emissions of X-ray [
14,
15].
EDX spectra confirmed that SNPs improved self-assembly of GCS from
Figure 2a–d. Perfect crystallization is attained in
Figure 2d.
The data in
Table 1 indicated that oxygen, carbon, and nitrogen have maximum weight %, confirming that glycine is the base matrix of these single crystals. EDX spectrums of the doped crystals confirm the entry of both Cu(II) ion and Ag(I) ions into glycine crystal lattice.
The crystals have molar stoichiometric formula [CH2NH2COOH]0.95, [CuSO4·5H2O]0.05 in the absence and presence of SNPs.
The vibrational band in FTIR spectra, as shown in
Figure S2, is assigned to the function groups in the crystals,
Table 2.
IR spectra of the GCS crystal showed a strong vibrational band at 509.21 cm
−1 due to Cu–N stretching [
16], NH stretching band at 3811.34 cm
−1, And medium peak at 1111.00 cm
−1 for CH
2 rocking [
17]. There is a strong peak at 1334.79 cm
−1 due to CH
2 wagging. Intense peak C=O asymmetric stretching occurs at 1604.77 cm
−1 [
18], symmetric stretching COO
– at 1411.68 cm
−1, intense band asymmetric stretching COO
– at 1519.91 cm
−1 [
19], and medium peak at 1033.85 cm
−1 for CCN asymmetric stretching deformation [
20]. There is strong band SO
4− stretching at 894.97 cm
−1, an intense band due to bending COO
− at 694.37 cm
−1, medium peak wagging COO
− at 609.51 cm
−1 [
13], and NH
2 asymmetric stretching at 2823.79 cm
−1 [
21]. FTIR spectra of the samples GCSN1, GCSN2, and GCSN3 have small shift compared to that of GCS observed, which suggests the incorporation of SNPs into the crystals lattice.
Figure 3a–d showed indexed pXRD profile versus Rietveld refined PXRD patterns for GCS, GCSN1, GCSN2, and GCSN3 crystals. All pXRD patterns showed a prominent sharp diffraction peak at 30°. SNPs increased the peaks’ intensity and modified crystal structure and lattice planes [
22,
23,
24]. The crystals’ structure and geometry agreed with Crystallography Opened Database, COD files. GCS and GCSN1 have monoclinic unit cell alpha glycine. Triclinic GCSN2 and GCSN3 have gamma glycine. pXRD patterns are refined using Full prof Suit software using CIF files containing crystal information. Peak patterns are refined following pseudo-Voigt profile analytical function [
25]. Background and peak shapes are modeled with linear fitting by applying least-squares cycles and six background (polynomial 6th grade parameters) at the wavelength of Cu-detector and neglecting instrument contribution [
26]. The crystallinity followed the trend:
Doping GCS with SNPs improved crystallinity, hence purity and crystal engineering.
During refinement, the number and order of crystalline planes and diffraction peaks increased in the same order. Many iteration cycles and all noise data are neglected; too long iteration time is consumed for GCSN2 and GCSN3 due to extra high crystallinity, long cartesian coordinates, and different angles in the triclinic unit cell. Intense peaks shifted to lower 2-theta, indicating a pillared crystal structure. Small peak absence and no polycrystallinity regions are observed in the perfect GCSN3 crystal, which confirmed good surface. There is good fitting of pXRD spectra (calculated intensity of the diffraction peaks are close to each other, resulting in a very negligible difference between the observed and calculated intensities (Yobs.-Ycal.)) with less than zero in arbitrary unit.
Both observed and calculated profiles are closely coincided to each other in a nonlinearly fit.
Figure 3.
Observed and refined pXRD patterns of crystals respectively: (a): GCS, (b): GCSN1, (c): GCSN2 and (d): GCSN3.
Figure 3.
Observed and refined pXRD patterns of crystals respectively: (a): GCS, (b): GCSN1, (c): GCSN2 and (d): GCSN3.
Sharp intense pXRD patterns with a dominating diffraction peak in crystals confirmed good crystallinity. Intensity changes and a slight shift in peak positions of GCS by SNPs reflected modified crystalline planes [
23]. Diffraction peaks at 44.4° and 64.7° in GCSN2 and GCSN3 confirmed doping impurities [
24].
Miller indices h k
of the crystal planes, full width at half maximum (FWHM) of XRD patterns, peak position, and inter planar distance (d
cal.), are collected in
Table S1. The crystallinity and geometry of single crystals are deduced from pXRD by matching these diffraction patterns to pdf cards of similar crystals in Crystallography Opened Database using FWHM that characterized different material properties and surface integrity features [
25,
26]. The unit cell parameters are collected in
Table 3.
UV-Vis. absorbance spectrum of crystals at wavelength range 190–1100 nm are shown in
Figure 4. UV-Vis. absorbance curve showed cut-off of wavelength λ
cut off is lower than that of glycine [
27].
Doping by SNPs decreased λ
cut off, i.e., increased band gas for UV–electronic transition. SNPs enhanced transparency of crystals to UV radiation-enabled deposition as thin film on glass for protection against UV radiation [
28,
29]:
Band gap-controlled UV absorption coefficient depends on the energy of the incident photon and is estimated using Equation (2) [
30,
31].
where ν frequency of the incident radiation is inversely proportional to the wavelength of absorbance (λ), A is constant, and the exponent r depends on the nature of electronic transition. r = 2 for indirect transition, and r = ½ for allowed direct transition, r = ½. Since all crystals are blue colored, allowed direct transition is considered [
15,
30]. The optical gaps, Eg, are calculated from plots (αhν)
2 as a function of photon energy (hν), as shown in
Figure 5, and the observed values Eg are given in
Table 4 along with the values of all reported glycine single crystals.
λ
cut off 287 nm for GCS decreased to 283, 276, and 280 nm for GCSN1, GCSN2, and GCSN3, respectively, which indicated an increasing band gap. Blue shift of λ
cut off to lower values indicated SNPs’ improved polarizability of electron density on the single crystals. This finding suggested the suitability of GCS-doped SNPs single crystals for applications in optoelectronic devices such as frequency multiplier, sum-, difference-, and blue laser frequency generators, etc. [
34]. Optical band gaps’ eV are 4.58, 4.61, 4.65, and 4.67 for GCS, GCSN1, GCSN2, and GCSN3 crystals, respectively. SNPs increased Eg and enhanced optical properties of GCS. High Eg indicated the decrease in the localized energy states on doping by SNPs due to extrinsic defects or disorders in GCS caused by interstitial doped SNPs [
35].
Good optical properties of the GCS crystals doped by SNPs are confirmed by extinction coefficient K calculated using Equation (3); α and refractive index (n) reflect dissipated incident radiation by absorption and scattering [
36,
37].
Reflectance R and K depend on photon energy,
Figure S3: The reflectance is calculated using Equation (4): [
36]
where t is the sample thickness.
Refractive index (n) depends on wavelength, and photon energy is calculated using mathematical Equation (5): [
36,
37,
38],
Figure 6.
Positive refractive index (n) 1.6–1.8 indicated dispersion of incident radiation on the crystals is inversely proportional to the photon energy. Refractive index (n) is decreased by increasing photon energy and concentration of SNPs. High photon energy enables passing through the crystal lattice with low dispersion. SNPs decreased dispersion of incident radiation by improving transparency to UV radiation. High transmission and low absorbance of UV radiation and low refractive index suggest the single crystals are suitable for antireflection coating in solar thermal devices and NLO applications [
38].
Figure 6.
Refractive index with wavelength and photon energy.
Figure 6.
Refractive index with wavelength and photon energy.
Figure 7 shows the nonlinear variation of the electrical susceptibility (χ) with the photon energy.
GSC crystals showed large noise scattering in electrical susceptibility. This scattering disappeared on doping by SNPs that elevated electrical susceptibility (χ) at high photon energy following the order:
GCSN3 showed lower χ than GCSN2 due to extra high crystallinity; GCSN3 decreased the mean free path for electron charge transfer.
The calculated electric susceptibility is plotted with photon energy;
Figure 7 shows electric susceptibility (χ) is about 0.25 for all crystals at low photon energy.
Melting point, crystallization, phase transition, enthalpy, and entropy changes are clarified using DSC thermograms [
39],
Figure 8a–d. A sharp endothermic peak at decomposition temperature reflects good crystallinity. A weak broad peak at 195.80 and 206.29 °C in GCSN2 and GCSN3 characterized phase change from
-glycine to α-glycine [
40]. Melting point (m.p.) is 253.83, 262.37, 265.64, 268.25 °C for GCS, GCSN1, GCSN2, and GCSN3, respectively. Increased m.p. on doping by SNPs is attributed to high thermal stability.
Heat capacity “Cp” is the ratio between heat flow to heating rate and depends on temperature: [
41,
42,
43,
44].
where parameters a and b are calculated from intercept (a) and slope b of straight-line Cp-T plot. Cp of all prepared single crystals showed nonlinear variation with the absolute temperature,
Figures S4–S7.
Thermal lattice coefficient α and electronic heat capacities γ are obtained from linear
T
2 plot, (coefficient, R
2 above 0.99)
Figures S4a,b and S7a,b.
Table 5 shows linear fits parameters of DCS.
Heat capacity at constant pressure, Cp, is the heat required to raise the temperature of the crystal sample by 1 °C and represents the variation of the heat content of the crystal sample on heating. The variation of α, γ coefficients approved Cp variation with temperature. This finding indicated that the Cp amount of thermal heat absorbed by the crystals increases on heating, enabling the application of a heat shielding coating on thin film glasses.
Thermograms are shown in
Figure 8a–d. DTG showed one peak confirmed one thermal decomposition step. GCS showed weight loss (wt. loss) 74.3 wt. % at temperature range 200.65–787.67 °C, DTG peak at 249.7 °C. Glycine decomposed into CO
2, NH
3 [
45] leaving 25.7% residue, 72.31%. GCSN1 decomposed at 208.27–788.29 °C and 27.69% residue. DTG peak was at 257.88 °C. In addition, 72.93 wt. % of GCSN2 decomposed at 217.41–788.84 °C, 27.07 wt. % residue, DTG peaked at 262.28 °C. Additionally, 72.86 wt. % of GCSN3 decomposed at 221.54–787.96 °C, gave 27.14 wt. % residue, and peak at 264.07 °C. SNPs decreased wt. loss on thermal decomposition of the crystals. Higher residue left after thermal decomposition confirmed improved crystallinity on doping GCS by SNPs.
Crystals contain no water of crystallization as wt. loss at 100 °C is absent [
46]. TGA thermograms confirmed thermal stability up to 200.65, 208.27, 217.41, and 221.54 °C for GCS, GCSN1, GCSN2, and GCSN3, which enabled laser applications [
47]. Peak temperature Tm increased from 249.71 to 257.88, 262.28, 264.07 °C on increasing SNPs’ concentration.
ESR spectra of powder sample crystals are shown in
Figure 9. Anisotropy g-factor for crystals confirmed low symmetry. Spin Hamiltonian parameters g and A tensors revealed rhombic symmetry crystal field around Cu(II) ion split ground state. Degeneracy of ground state energy level is lifted giving static Jahn–Teller distortion [
48,
49].
Value g
⏊ and unpaired electronic state R are given by Equations (7) and (8): [
50,
51].
Table 6 includes g factor, hyper fine constants A, and R for the crystals.
Values A and g factor have no axial symmetry in the crystal lattice (no dynamic Jahn–Teller) [
52]. R = 0.1805, 0.1224, 0.1673, 0.1418 for GCS, GCSN1, GCSN2, and GCSN3, respectively; less than unity indicated d
x2−y2 ground state for unpaired electron [
53].
A
x, A
y equals
are lower than A
z; g-parallel is greater than perpendicular g
⏊ and confirmed d
x2−y2 ground state [
52].
value is less than 2.3, indicating strong covalent copper–glycine bond [
54].
σ bond coefficient or covalence parameter α
2 for unpaired electron density on Cu(II) ion is estimated using Equation (9) [
54,
55],
Table 7.
where
The parameter α
2 is less than unity, indicating covalent bonding between Cu(II) and glycine ligand [
56]. SNPs decreased covalence parameter (α
2) of glycine-Cu(II) bond, except that GCSN2 showed abnormally high α
2, which confirmed its highest electrical susceptibility.
Figure 10 showed UV of thin film coating of GCSN3 on the aluminum (Al) foil sample. An aqueous solution of GCSN3 was evaporated under ultra-high vacuum conditions onto the Al foil where carboxylate COOH of glycine zwitterion amino acid is chemically adsorbed on the aluminum surface. Absorption at long λ 900–1100 nm for this crystal near IR region indicates absorption of thermal energy of IR radiation. Phonon bands at 900 nm originate from vibrational modes of harmonic and unharmonic oscillators in the crystal lattice. Absorbed IR radiation causes thermal vibrations of atoms or molecules and creates thermal phonon waves that propagate in the crystal lattice, dissipating thermal IR energy.
Absorptivity of GCSN3 near the IR region of electromagnetic radiation indicated that crystals can shield thermal heat of IR radiation on the coating as dispersed thin film on alumetal.
The sun provides thousands W·m
−2 energy on the earth’s surface daily. Total solar energy in the upper atmosphere contains 50% IR radiation, 40% Vis. Light, and 10% UV radiation. IR radiation causes vibrations that heats earth’s surface [
57]. Attenuation of thermal energy can be achieved by painting glass windows with these blue color crystals transparent to UV radiation, filtering, and that dissipates IR radiation.
High thermal conductivity of crystals equals: 1.10, 1.21, 1.54, and 1.6 W·m−1K−1 for GCS, GCSN1, GCSN2, and GCSN3 confirmed rapid attenuation of many incident EM waves by dielectric components and rapidly dissipated as heat.
Figure 11 showed electrical conductivity of GCSN3 increased on heating as a typical semiconductor behavior due to thermally activated charge carriers’ mobility [
58].
Figure 12 showed variation of impedance of GCSN3 with applied frequency.
Impedance plots confirmed super conductivity. A plateau region at low 0.1Hz frequency region represents total conductivity of grain boundary. A high-frequency region at 100 kHz represents the contribution of grains to total conductivity. An intermediate frequency region at 1 kHz is due to charges trapped between grain boundaries and grains [
15]. AC conductivity confirmed the dielectric nature of a single crystal can dissipate heat rapidly. The high-frequency dielectric constant is 4.49.
Dielectric study of GCSN3 crystal response of charges to applied electric field showed dielectric constants at 100 Hz, 1 kHz, 10 kHz, and 100 kHz and a temperature range of 200–550 K. Dielectric constant was calculated using equation [
59]:
where
is free space permittivity, C and d are capacitance and thickness of pellet, and A is electrode area.
Real ε′ and imaginary ε″ components of ε represented equals [
60]:
Figure 13 showed ε′ varied with temperature at a different frequency and decreased with increasing frequency, indicating an ability to dissipate incident IR radiation. ε′ decreased until it reached glass transition Tg at 380 K, then became limited up to 470 K. Peak at Curie TC represented phase transition from ferroelectric to paraelectric behavior.