Abstract

Imidazo[1,2-a]pyrazines are cyclic amidine-type compounds composed of α-amino acid residues. A full structural identification of these molecules constitutes an analytical challenge, especially when imidazo[1,2-a]pyrazines are obtained from physical processes (e.g., sublimation and pyrolysis of amino acids). A valuable source of molecular information can be obtained from absorption spectroscopies and related techniques encompassing the use of metallic substrates. The aim of this study is to provide new knowledge and insights into the noncovalent intermolecular interactions between imidazo[1,2-a]pyrazines and two Agn (n = 4 and 20) clusters using density functional theory (DFT) methods. Semiempirical DFT dispersion (DFT-D) corrections were addressed using Grimme’s dispersion (GD2) and Austin–Petersson–Frisch (APF) functionals in conjunction with the 6-31+G(d,p) + LANL2DZ mixed basis set. These DFT-D methods describe strong interactions; besides, in all cases, the APF dispersion (APF-D) energies of interaction appear to be consistently overestimated. In comparison with B3LYP calculations, the mean values for the difference in the energies of interaction calculated are 2.25 (GD2) and 6.24 (APF-D) kcal mol–1 for Ag4–molecules, and 2.30 (GD2) and 8.53 (APF-D) kcal mol–1 for Ag20–molecules. The effect of applying GD2 and APF-D corrections to the noncovalent complexes is nuanced in the intermolecular distances calculated, mainly in the Ag···N(amidine) bonding, which appears to play the most important role for the adsorptive process. Selective enhancement and considerable red shifts for Raman vibrations suggest strong interactions, whereas a charge redistribution involving the metallic substrate and the absorbate leads to a significant rearrangement of frontier molecular orbitals mainly in the Ag20–molecule complexes. Finally, time-dependent DFT calculations were carried out to access the orbital contributions to each of the transitions observed in the absorption spectrum. The corresponding UV–vis spectra involve transitions in the visible region at around 400 and 550 nm for the Ag4–molecule and the Ag20–molecule complexes, respectively.
Introduction
Imidazo[1,2-a]pyrazines are heterocyclic structures composed of α-amino acid residues contained into a five-membered imidazole fused to a six-membered pyrazine ring with a bridgehead nitrogen atom. These unusual molecules are of great interest from the point of view of amino acid–peptide chemistry because this class of amidine-type compounds might manifest themselves as the central compounds for the whole amidine chemistry.1 The natural occurrence of imidazo[1,2-a]pyrazines was discovered in bioluminescence sea organisms2 involving luciferine3 and coelenterazine4,5 systems, which constitute the most representative natural imidazopyrazines compounds. Their ability to luminesce under the action of oxidants has allowed the development of chemiluminescent sensors based on the determination of peroxides,6 glucose,7 and chemiluminescent immunoassay.8 A variety of biological activities (e.g., antibacterial, anti-inflammatory, antiulcer, antibronchospastic, cardiac stimulating, and inotropic properties) and cytotoxic effects on cancer cell lines have been also reported for synthetic imidazo[1,2-a]pyrazines derivatives.9−16
Several methods from wet-chemistry16 to sublimation17 and pyrolysis18−21 of amino acids are used to prepare imidazo[1,2-a]pyrazines. However, one of the major difficulties to study these molecules is lack of an adequate technique for a full structural identification. Combinatorial techniques such as gas-phase Fourier transform infrared spectroscopy coupled to gas chromatography and mass spectrometry17−21 identified some derivatives, but such methods also exhibited disadvantages because in situ decomposition of imidazo[1,2-a]pyrazines is very notable.18−21 The interest to study these cyclic amidine-type molecules lies in the fact that derivatives of glycine have failed to be detected by these techniques. The latter fact can be directly associated with the selective detection, thermochemical stability, or simply because cycle-condensation reactions of amino acids are not able to produce enough amounts of imidazo[1,2-a]pyrazines. Previous quantum chemical calculations22 suggested that the formation of imidazo[1,2-a]pyrazines can proceed overcoming a high reaction barrier of about 49 kcal mol–1 and only supported by a silica-catalyzed process. This has opened up an area of great dubiety about the formation and the spectroscopic properties23−25 of imidazo[1,2-a]pyrazines, which deserves more definitive answers and novel approaches for experimental identification.
Vibrational spectroscopy (e.g. Raman) can be the technique of choice for such a molecular identification. In particular, surface-enhanced Raman spectroscopy (SERS)26,27 dramatically can improve the limit of detection using metallic substrates, resulting in detection even at a single-molecule-level concentration.28 Under such a scenario, it is important to study the adsorptive processes to attempt surface-enhanced Raman detection. Of particular interest is the study of noncovalent intermolecular interactions with metal clusters29 because no mechanism was put forward to account for the adsorption of imidazo[1,2-a]pyrazines onto silver clusters. In the present study, the molecular properties of the noncovalent complexes formed with imidazo[1,2-a]pyrazines and Agn (n = 4 and 20) clusters were studied using density functional theory (DFT) methods. The chosen Ag clusters represent to simple models as much as the physical and chemical properties of metallic substrates are of great relevance in SERS.26−29 In particular, semiempirical dispersion corrections from Grimme (GD2)30 and Austin–Petersson–Frisch (APF-D)31 applied over the exchange–correlation B3LYP32,33 functional can provide theoretical insights into the strength of the noncovalent intermolecular interactions for these complexes. In addition, DFT calculations were carried out to predict static Raman activities, which further provide a measure of the chemical enhancements that arise during the adsorption onto the metallic substrates. Finally, time-dependent DFT (TD-DFT) calculations were carried out to access the orbital contributions to each of the transitions observed in the absorption spectrum, and molecular excitation energies were obtained for both isolated molecule and metallic cluster-organic compounds. Arising from sublimation of α-amino-acids (e.g., glycine), bi and tricyclic imidazo[1,2-a]pyrazines prepared require the formation of a first intermediate.22 The current series of molecules studied correspond to glycine (Gly, 1), piperazine-2,5-dione (DKP, 2), hexahydroimidazo[1,2-a]pyrazine-3,6-dione (BCA, 3), and hexahydroimidazo[1,2-a]imidazo[1,2-d]pyrazine-3,8-dione (TCA, 4), which all are depicted in Scheme 1.
Scheme 1. Chemical Structures of Glycine (1), Piperazine-2,5-dione (DKP, 2), Imidazo[1,2-a]pyrazine-3,6-dione (BCA, 3), and Imidazo[1,2-a]imidazo[1,2-d]pyrazine-3,8-dione (TCA, 4).

Results and Discussion
In this study, the noncovalent intermolecular interactions occurring on the metallic substrates were addressed using two sizes for the silver clusters, namely, Agn (n = 4 and 20). Similarity and differences for these Ag clusters were reported elsewhere.34−39 In particular, the Ag4 and Ag20 clusters were chosen because they are the most stable structures in the potential energy surface, respectively, for small-sized clusters (n < 10) and medium-sized clusters (12 < n < 22).34 Substantial differences with other clusters is the relatively lower symmetry that can lead to metastable isomers. Several DFT methods35 predicted the global minimum for the neutral Ag4 and Ag20 clusters to be rhombic (D2h) and tetrahedral (Td) structures, respectively. Ag20 (Td) is a relaxed fragment of the face-centered cubic lattice of bulk silver, whereas Ag4 (D2h) is a subdomain located at the vertex of Ag20 cluster. Because Ag4 (D2h) represents the local minima with respect to other isomers (e.g. Td, C2v, see the Supporting Information), this structure was used as a structural criterion for a respective comparison between both clusters. The energy gap between highest occupied molecular orbital (HOMO) and lowest occupied molecular orbital (LUMO), and the Fermi energy level were calculated to characterize the metallic substrates. The HOMO–LUMO gap energies, ΔGAP, (see Table 1), were calculated to be about 1.8 and 2.5 eV, respectively, for the isolated Ag4 and Ag20 clusters. It is well known that the large energy gap corresponds to the higher stability for close-shell electron configurations; therefore, Ag20 is more stable than Ag4 (see Table 1). Empirical dispersion does not affect the properties of the wavefunction31 and thus GD2 and APF-D calculations result in similar values for HOMO–LUMO gap energies. The Fermi level (EF, see the Supporting Information) was calculated to be about −4.36 and −4.44 eV for Ag4 and Ag20, respectively. Such a difference can be related with a different work function for transferring one quasi-electron from the Fermi level to the vacuum state. For the large size cluster, the value calculated is approximately similar for the bulk metal as obtained from photoelectric measurements on the (111) surface of silver metal surfaces.40
Table 1. Orbital Interaction Energies (in kcal mol–1) for the Noncovalent Intermolecular Interactions between Molecules (1–4) and Agn (n = 4 and 20) Clusters As Calculated Using B3LYP and Semiempirical Dispersion Correction with Grimme’s D2 and APF-D Functionals in Conjunction with the 6-31+G(d,p) + LANL2DZ Mixed Basis Seta.
| B3LYP |
Grimme-D2 |
APF-D |
||||
|---|---|---|---|---|---|---|
| complexes | ΔENC | ΔGAP | ΔENC | ΔGAP | ΔENC | ΔGAP |
| Ag4 | 1.77 | 1.85 | 1.86 | |||
| Ag4–Gly | –9.69 | 1.48 | –11.62 | 1.48 | –15.51 | 1.48 |
| Ag4–DKP | –7.03 | 1.56 | –10.02 | 1.52 | –12.78 | 1.52 |
| Ag4–BCA | –9.05 | 1.51 | –14.07 | 1.50 | –15.66 | 1.50 |
| Ag4–TCA | –8.20 | 1.48 | –14.25 | 1.57 | –15.00 | 1.57 |
| Ag20 | 2.49 | 2.49 | 2.50 | |||
| Ag20–Gly | –14.40 | 2.93 | –18.75 | 2.43 | –23.47 | 2.44 |
| Ag20–DKP | –11.97 | 2.45 | –18.43 | 2.43 | –20.16 | 2.43 |
| Ag20–BCA | –14.55 | 2.43 | –21.81 | 2.43 | –23.10 | 2.44 |
| Ag20–TCA | –13.44 | 2.33 | –20.29 | 2.27 | –21.77 | 2.28 |
HOMO–LUMO gap energies are reported in eV for each level of theory.
Figure 1 shows the geometries at the stationary points as obtained using DFT-D methods. It was assumed that molecules (1–4) interacted with an ad-atom site located at the vertex of the silver clusters. Previous studies37 showed that this orientation is more stable (∼2.3 kcal mol–1) that on-top binding onto one of the four faces (111) of the Ag20 cluster. From Figure 1, it becomes evident that mainly two bonds stabilize the adsorptive processes, in which the noncovalent interaction occurs along the Ag···N (BCA and TCA) or Ag···O (DKP) bonds. As a result, one can observe that molecules (1–4) can be adsorbed mainly by cooperative interactions with the participation of NH2 and COOH groups. The unique exception is Ag20–Gly, in which the interactions arise effectively from both amine and carboxyl groups. For the Ag4–Gly complex, the calculated bond distance between the N atom and the closest silver atom is 2.382 Å (DFT-D values), whereas this distance is slightly smaller (2.353 Å) for the Ag20–Gly complex. In this complex, it is also observed that NH2 and COOH groups are coordinated to the metal cluster, in which both O and N atoms lead to form a three-membered ring. Such a bidentate interaction set up the bonding between the O atom and the closest Ag atom (2.423 Å). On other hand, the cyclic derivative DKP (2) forms complexes with Ag clusters via the Ag···O(amide) bond because its N(amide) atom is covalently bounded to the H atom. The resulting Ag···O separation is 2.298 Å for the Ag4–DKP complex, whereas the intermolecular distance is shortest (2.246 Å) for Ag20–DKP. For BCA (3), the Ag···N separation is longer (2.226 Å) in Ag4–BCA as compared to Ag20–BCA (2.191 Å). Finally, for the Ag4–TCA complex the interaction occurred through the imidazole ring, in which the Ag···N separation (2.314 Å) is longer as comparted to the Ag20–TCA complex (2.191 Å). In the case of BCA and TCA, these molecules are adsorbed with an edge-on orientation and the imidazo ring almost lying perpendicular to the Ag clusters. The optimized structures suggested that the interaction of BCA and TCA is governed mainly by the anchoring Ag···N(amine) bonds. Mulliken population analysis indicates that N(amine) is rich in electrons and can be able to donate electron density to the empty 4d and 5s orbitals of silver. The charges on Ag(vertex) and N(amide) atoms calculated upon the formation of the complexes are Ag(−0.101)···N(−0.709), Ag(−0.169)···N(−0.302), and Ag(−0.155)···N(−0.292) for Ag20–Gly, Ag20–BCA and Ag20–TCA complexes, respectively. Therefore, Ag clusters can behave as electron acceptors when charges arise from nitrogen (e.g., Gly, BCA, and TCA). On the other hand, the Ag···O(amine) bond gives rise to a weak interaction, mainly, in the Ag–DKP complexes. The charges on the Ag20–DKP complex are Ag(−0.346)···O(−0.380). The full list of Mulliken population are in the Supporting Information. A reasonable explanation to this is that nitrogen is less electronegative than oxygen, making it a better nucleophile. Finally, the optimized Ag···N(amine) bond lengths are notably shorter in Ag20 clusters than the ones in the Ag4 cluster. The intermolecular distances for Ag20–molecule complexes are in the order of those calculated in the interaction of amino acids (and short peptides) with small silver clusters.29
Figure 1.

Optimized structures as obtained at the APFD-B3LYP/MBS level of theory for the gas-phase formation of Gly, DKP, BCA, and TCA complexes with Ag4 and Ag20 clusters. Selected interatomic bond distances in angstroms are included for comparison between dispersive (values in parenthesis) and nondispersive calculations.
Table 1 shows the energies of interaction for the noncovalent complexes, ΔENC, calculated in terms of an adiabatic two-state process using B3LYP, GD2 and APF-D functionals (absolute energies are reported in the Supporting Information). In particular, both GD2 and APF-D values account for empirical correction on dispersion through either electron pair bond formation, charge transfer, or polarization mechanisms,30,31 whereas B3LYP values represent the level of interaction when dispersion corrections are not taken into account. Therefore, GD2 and APF-D interaction energies were benchmarked against B3LYP energies. The mean values calculated for the Ag4–molecule complexes are 8.5, 12.5, and 14.7 kcal mol–1, respectively, for B3LYP, GD2, and APF-D, whereas the mean value calculated for the Ag20–molecule complexes are 13.6, 19.8, and 22.1 kcal mol–1, respectively, for B3LYP, GD2, and APF-D. It is observed that empirical dispersion correction using the APF-D functional resulted in severe strong interactions because in all cases such energies appear to be consistently overestimated in comparison with B3LYP calculations. Differences for the energies of interaction calculated with respect to B3LYP are 2.25 (GD2) and 6.24 (APF-D) kcal mol–1 for Ag4–molecules, and 2.30 (GD2) and 8.53 (APF-D) kcal mol–1 for Ag20–molecules. According to other studies, the dispersion term in the APF-D functional should probably be re-evaluated,38 because deviations can arise from some long-range dispersion that were double-counted by the APF function and the spherical atom model. The HOMO–LUMO gap energies (ΔGAP, see Table 1) are very similar for GD2 and APF-D calculations, and slightly close with those values calculated without dispersion (i.e., B3LYP). However, the HOMO–LUMO gap energies becomes smaller after complexation with molecules, and this is noticeable for those noncovalent complexes formed with the Ag4 cluster (in about 0.34 eV). Figure 2 shows the iso-surfaces for HOMO and LUMO (at 0.02 a.u.) depicting the frontier orbitals for the Ag–molecule complexes together with a simplified orbital energy diagram. In the case of Ag4–molecule complexes (Figure 2A), a small extended distribution occurred in HOMO orbitals, whereas in the Ag20–molecule complexes (Figure 2B) a large portion of HOMO is delocalized, suggesting the formation of electrostatic field gradients around the Ag20 cluster that can induce dipoles in the absorbates because of the changes in the surface charge density. In particular, a significant rearrangement of the frontier molecular orbitals occur in the Ag20–BCA and Ag20–TCA complexes. As regards spatial distribution of frontier orbitals, the following situations can be observed:
-
(1)
HOMO is found exclusively on Ag metal clusters.
-
(2)
Both HOMO and LUMO appear on the Ag4 metal cluster only.
-
(3)
Traces of LUMO appear on BCA and TCA.
-
(4)
LUMO is mainly localized in the imidazole ring of BCA and TCA.
Figure 2.

Schematic of the frontier orbital bonding HOMO and LUMO for the complexes formed between Gly and its related imidazo[1,2-a]pyrazines (DKP, BCA, and TCA) when adsorbed onto metallic clusters. (A) Interaction of molecules with the Ag4 cluster (A) and interaction of molecules with the Ag20 cluster (B).
According to these situations, one can suggest that imidazole rings are responsible for binding imidazo[1,2-a]pyrazines (i.e., BCA and TCA) to Ag clusters.
TDDFT calculations were carried out on 15 electronic transitions for the ground-state complexes. Table 2 shows the most important excitation energies (i.e., those with the largest oscillator strength) as calculated for vertical excitation at using DFT-D methods. It is observed that the Ag4–molecules complexes exhibited the highest oscillator strengths, whereas very small oscillator strengths were calculated for the Ag20–molecules complexes. On one hand, the excitations for electronic transitions in Ag4–molecule systems are found at energies in the proximity of the absorption maxima (i.e., 386 nm relative to the Ag4 cluster), that is, Ag4–Gly, 408 nm (f = 0.795); Ag4–DKP, 362 nm (f = 0.639); Ag4–BCA, 379 nm (f = 0.879); and Ag4–TCA, 393 nm (f = 0.601). Therefore, these metal-molecule transitions should occur from orbitals below the HOMO (HOMO – N, with N = 1 and 2) of Ag4 to the LUMO of each molecule. On the other hand, the calculated absorption spectrum for Ag20–molecule complexes involves excitations in the visible region and they are very close to the absorption maxima (i.e., 537 nm, relative to the Ag20 cluster); that is, Ag20–Gly, 554 nm (f = 0.021); Ag20–DKP, 543 nm (f = 0.025); Ag20–BCA, 547 nm (f = 0.024); and Ag20–TCA, 647 nm (f = 0.019). Figure 3 shows the simulated absorptive spectrum of the studied complexes. The absorption showed changes as a function of both the molecule and the cluster size. For the Ag4–molecules complexes, the corresponding spectra involve transitions in the visible region that might occur at about 400 nm, which are blue-shifted regarding the transitions for the Ag20–molecule complexes occurring at around 550 nm.
Table 2. Principal Molecular–Cluster Complexes’ Excitation Energies As Obtained from TDDFT Calculations.
| complex | MOsa | energy/wavelengthb | fc | wave function |
|---|---|---|---|---|
| Ag4–Gly | 58 → 61 | 3.039/408 | 0.7952 | 0.521 |H → L + 2⟩ |
| 58 → 65 | 3.609/344 | 0.2350 | 0.569 |H → L + 6⟩ | |
| 57 → 59 | 3.776/328 | 0.3328 | 0.456 |H – 1 → L⟩ | |
| Ag4–DKP | 68 → 75 | 3.405/364 | 0.3103 | 0.400 |H → L + 6⟩ |
| 68 → 75 | 3.421/362 | 0.6388 | 0.506 |H → L + 6⟩ | |
| 66 → 69 | 3.886/319 | 0.2111 | 0.419 |H – 2 → L⟩ | |
| Ag4–BCA | 78 → 84 | 3.267/379 | 0.8792 | 0.478 |H → L + 5⟩ |
| 78 → 88 | 3.811/325 | 0.1773 | 0.654 |H → L + 9⟩ | |
| 77 → 79 | 3.847/322 | 0.1541 | 0.387 |H – 1 → L⟩ | |
| Ag4–TCA | 88 → 95 | 3.157/393 | 0.6007 | 0.474 |H → L + 6⟩ |
| 88 → 96 | 3.521/352 | 0.2121 | 0.508 |H → L + 7⟩ | |
| 87 → 90 | 3.712/334 | 0.1105 | 0.490 |H – 1 → L + 1⟩ | |
| Ag20–Gly | 208 → 213 | 2.179/569 | 0.0137 | 0.624 |H – 2 → L + 2⟩ |
| 206 → 212 | 2.232/556 | 0.0187 | 0.466 |H – 4 → L + 1⟩ | |
| 206 → 211 | 2.236/554 | 0.0209 | 0.518 |H – 4 → L⟩ | |
| 207 → 212 | 2.253/550 | 0.0181 | 0.602 |H – 3 → L + 2⟩ | |
| 208 → 214 | 2.295/540 | 0.0155 | 0.545 |H – 2 → L + 3⟩ | |
| Ag20–DKP | 216 → 221 | 2.229/556 | 0.0132 | 0.594 |H – 4 → L⟩ |
| 216 → 222 | 2.258/549 | 0.0109 | 0.461 |H – 4 → L + 1⟩ | |
| 218 → 223 | 2.285/543 | 0.0250 | 0.486 | H – 2 → L + 2⟩ | |
| Ag20–BCA | 229 → 231 | 1.939/639 | 0.0138 | 0.486 |H – 1 → L⟩ |
| 226 → 232 | 2.234/555 | 0.0170 | 0.417 |H – 4 → L + 1⟩ | |
| 226 → 231 | 2.241/553 | 0.0136 | 0.346 |H–4 → L⟩ | |
| 227 → 232 | 2.261/548 | 0.0133 | 0.330 |H – 3 → L + 1⟩ | |
| 229 → 233 | 2.265/547 | 0.0238 | 0.385 |H – 1 → L + 2⟩ | |
| Ag20–TCA | 239 → 241 | 1.844/672 | 0.0107 | 0.545 |H – 1 → L⟩ |
| 240 → 242 | 1.916/647 | 0.0192 | 0.593 |H → L + 1⟩ | |
| 236 → 241 | 2.166/572 | 0.0129 | 0.486 |H – 4 → L⟩ | |
| 236 → 242 | 2.210/561 | 0.0163 | 0.474 |H – 4 → L + 1⟩ | |
| 237 → 242 | 2.235/555 | 0.0135 | 0.529 |H – 3 → L + 1⟩ |
Molecular orbitals.
In eV/nm.
Oscillator strength.
Figure 3.
Absorption spectra of the complexes formed between Gly, DKP, BCA, and TCA molecules with metallic clusters. (A) Predicted spectra using the Ag4 cluster and (B) predicted cluster using the Ag20 cluster. Spectra are broadened by a Lorentzian having a width of 12 nm.
Theoretical Raman spectra for neutral Gly (1), DKP (2), BCA (3,) and TCA (4) molecules obtained using DFT-D methods are shown in Figure 4. Frequencies at the gas-phase with the corresponding band assignments and the absolute Raman cross sections are respectively listed in Tables 3–6. Raman spectra were calculated only for neutral molecules, as it was demonstrated that Raman spectra of amino acids do not depend on the pH values.41 In detail, Raman spectral profiles contain bands grouped in two regions, namely, bands between 200 and 1000 cm–1 and bands within the range of 1000–2000 cm–1. However, peak intensities at the second range are slightly smaller than the peaks located at the first range. Taking into account the basis set effect on fundamental frequencies, the wavenumber scaled was used for the full assignment on the vibrational modes. In particular, the Raman spectra profile of Gly (1) shows bands associated with COO– bend + CH2 bend at 494 and 586 cm–1, stretching-type modes of CH2 at 1190, 1398, and 1462 cm–1, and stretching of C=O at 1674 cm–1. In-plane deformation bending δ(C=O) was at 807 cm–1, and δ(NH2) and wagging ρw(NH2) at 1092 and 1591 cm–1, respectively. A number of additional bands because of CH2 and C=N and C=O vibration modes are evident in the case of BCA (3) and TCA (4) molecules, whereas the spectra of DKP (2) showed the presence of NH bands that remained as the spectral characteristics of Gly (1). The fingerprint region (i.e., 200 and 800 cm–1) for DKP, BCA, and TCA molecules is constituted by peaks corresponding to the bands that all belong to deformation modes involving ring distortions (i.e., deformations of pyrazine and imidazo rings) as well as in-plane and out-of-plane bending of C=O moieties. The strongest bands because of rings’ breathing of DKP, BCA, and TCA molecules are located at 755, 707, and 702 cm–1, respectively.
Figure 4.

Simulated normal Raman scattering spectra of Gly, DKP, BCA, and TCA molecules as calculated at the APFD-B3LYP/MBS level of theory. Differential cross sections of molecules in units of 10–31 cm2/sr were calculated at an incident wavelength of 514.5 nm based on static polarizability derivatives. Bands are given a Lorentzian width of 20 cm–1. Frequencies were scaled with a 0.959 factor.
Table 3. Selected Gas-Phase Normal Raman Active Modes, Scattering Factors (Si), Absolute Raman Cross Sections (dσ/dΩ) at an Incident Wavelength of 514.5 nm Based on Static Polarizability Derivatives, Together with Their Corresponding Band Assignments for Gly (1).
| νia | Sib | dσ/dΩc× 10–31 | band assignment |
|---|---|---|---|
| 66 | 0.31 | 0.21 | molecule deform |
| 162 | 3.32 | 0.46 | NH deform out-of-plane |
| 376 | 2.28 | 0.09 | molecule deform |
| 452 | 1.70 | 0.07 | COOH bending |
| 494 | 3.82 | 0.11 | COO– bend + CH2 bend |
| 586 | 3.62 | 0.09 | C–C stretch |
| 807 | 9.60 | 0.13 | COOH bending |
| 1092 | 6.48 | 0.08 | NH deform out-of-plane |
| 1190 | 9.64 | 0.11 | C–H bending |
| 1266 | 5.37 | 0.05 | C–H bending |
| 1398 | 3.91 | 0.03 | C–H bending |
| 1462 | 15.72 | 0.10 | C–H bending |
| 1591 | 3.76 | 0.02 | NH bending |
| 1674 | 6.16 | 0.03 | C=O stretch |
In cm–1.
In Å4/amu.
In cm2/sr.
Table 6. Selected Normal Raman Active Modes, Scattering Factors (Si), Absolute Raman Cross Sections (dσ/dΩ) at an Incident Wavelength of 514.5 nm Based on Static Polarizability Derivatives, Together with Their Corresponding Band Assignments for Hexahydroimidazo[1,2-a]imidazo[1,2-d]pyrazine-3,8-dione (4).
| νia | Sib | dσ/dΩc× 10–31 | band assignment |
|---|---|---|---|
| 44 | 1.11 | 0.17 | pyrazine ring deform |
| 197 | 3.05 | 0.52 | rings deform |
| 236 | 3.04 | 0.39 | rings deform |
| 401 | 5.99 | 0.24 | pyrazine ring deform |
| 509 | 3.23 | 0.20 | rings deform |
| 612 | 11.51 | 0.32 | rings deform |
| 702 | 15.97 | 0.48 | rings breath |
| 1019 | 7.28 | 0.10 | CH2 rock |
| 1134 | 10.29 | 0.13 | CH2 wagging |
| 1179 | 5.15 | 0.09 | CH2 wagging |
| 1235 | 6.50 | 0.14 | CH2 wagging |
| 1270 | 4.29 | 0.10 | CH2 wagging |
| 1325 | 15.83 | 0.17 | CH2 wagging |
| 1407 | 35.42 | 0.33 | CH2 scissors |
| 1469 | 15.01 | 0.17 | CH2 scissors |
| 1624 | 32.29 | 0.25 | C=N stretch |
| 1669 | 20.85 | 0.16 | C=O stretch |
In cm–1.
In Å4/amu.
In cm2/sr.
Table 4. Selected Normal Raman Active Modes, Scattering Factors (Si), Absolute Raman Cross Sections (dσ/dΩ) at an Incident Wavelength of 514.5 nm Based on Static Polarizability Derivatives, Together with Their Corresponding Band Assignments for Piperazine-2,5-dione (2).
| νia | Sib | dσ/dΩc× 10–31 | band assignment |
|---|---|---|---|
| 60 | 0.69 | 0.56 | ring deform |
| 174 | 1.31 | 0.20 | ring deform |
| 428 | 4.23 | 0.18 | ring deform |
| 506 | 4.34 | 0.19 | CH2 stretch |
| 577 | 6.31 | 0.16 | ring stretch |
| 755 | 18.32 | 0.28 | ring breath |
| 1125 | 3.39 | 0.04 | ring deform |
| 1217 | 13.83 | 0.13 | CH2 rocking |
| 1296 | 11.65 | 0.12 | CH2 wagging |
| 1346 | 13.21 | 0.11 | CH2 bending, NH bending |
| 1427 | 8.59 | 0.09 | CH2 scissors, NH bending |
| 1477 | 11.77 | 0.10 | CH2 scissors |
| 1642 | 15.43 | 0.08 | C=O stretch |
In cm–1.
In Å4/amu.
In cm2/sr.
Table 5. Selected Normal Raman Active Modes, Scattering Factors (Si), Absolute Raman Cross Sections (dσ/dΩ) at an Incident Wavelength of 514.5 nm Based on Static Polarizability Derivatives Together with Their Corresponding Band Assignments for Hexahydroimidazo[1,2-a]pyrazine-3,6-dione (3).
| νia | Sib | dσ/dΩc× 10–31 | band assignment |
|---|---|---|---|
| 55 | 0.86 | 0.83 | rings deform |
| 167 | 1.54 | 0.27 | rings deform |
| 202 | 1.24 | 0.22 | rings deform, CH2 rock |
| 233 | 1.62 | 0.20 | rings deform, CH2 rock |
| 416 | 6.56 | 0.26 | pyrazine ring deform |
| 521 | 3.29 | 0.16 | CH2 rock, NH deform out-of-plane |
| 569 | 8.37 | 0.26 | CH2 rock, NH deform out-of-plane |
| 678 | 8.59 | 0.30 | CH2 rock, NH deform out-of-plane |
| 707 | 11.28 | 0.27 | rings breath |
| 823 | 2.31 | 0.05 | pyrazine ring deform |
| 1017 | 2.71 | 0.05 | CH2 rock, NH deform out-of-plane |
| 1133 | 5.70 | 0.07 | CH2 rock |
| 1211 | 7.47 | 0.13 | pyrazine ring deform |
| 1305 | 11.39 | 0.14 | CH2 wagging |
| 1360 | 6.29 | 0.10 | NH bending |
| 1472 | 16.24 | 0.16 | CH2 scissors |
| 1627 | 19.89 | 0.15 | C=N stretch |
| 1669 | 10.41 | 0.09 | C=O stretch |
In cm–1.
In Å4/amu.
In cm2/sr.
Figure 5 shows the effect of using Ag clusters on the Raman spectra of molecules (1–4). As observed, evident spectral differences resulted in bands that appeared at different wavenumbers. In particular, selective enhancement and considerable red shifts for several Raman vibrations suggest that for some signal enhancements such a result effectively corresponds to the presence of local interactions occurring between the active site of the adsorbate (i.e., imidazole ring) and the effective ad-atom of the metal surface. In particular, for the Ag4–Gly complex, the bands observed are δ(NH2), ρw(CH2), δ(NH2), and ν(C=O) located at 862, 1277, 1598, and 1672 cm–1, respectively. Tenorio et al.41 calculated similar Raman spectra for glycine but using a silver eight-atom cluster. For the Ag4–DKP complex, the most characteristic absorption bands present intense peaks at 764, 1306, 1354, 1454, and 1641 cm–1 assigned to ν(NH), δ(CH2), ν(NH), ν(N=C), and ν(C=O) vibrations, respectively. For both Ag4–BCA and Ag4–TCA complexes, the most intense bands were observed in a particular region between 1000 and 1500 cm–1. Absorption bands were predicted corresponding to δ(CH2) at 996 cm–1, rings’ deformations at 1013 cm–1, ρw(CH2) at 1208 cm–1, pyrazine ring-deformation at 1277 cm–1, bending δ(NH) at 1362 cm–1, imidazo ring-deformation at 1424 cm–1, bending δ(N=C) at 1613 cm–1, and bending δ(C=O) at 1638 cm–1. Finally, the Raman spectra of the molecules adsorbed onto the Ag20 cluster are much more similar than those of isolated molecules. This is the case of the Ag20–Gly complex, in which the most intense bands are δ(CCO), δ(NH2), and δ(CH2), δ(CH2), ω(CH2), δ(NH2), and δ(C=O) located at 629, 1147, 1386, 1435, and 1628 cm–1, respectively. The predicted absorptions for the interval between δ(NH2) and δ(C=O) stretching is 193 cm–1, whereas in the case of Ag4–Gly, it was 74 cm–1. In Ag20–DKP, the principal absorption bands correspond to ring breathing (764 and 889 cm–1), which is more evident than that in the Ag4 complex. The bands δ(CH2), ν(NH), and ν(N=C) were also observed at 1217, 1356, and 1429 cm–1, respectively. However, C=O bands are much weaker and thus they cannot be distinguished in the predicted spectrum. Because BCA and TCA are molecules with the same chemical nature, the interaction energies are similar, and form noncovalent complexes through the same active site, the predicted Raman spectra for both Ag20–BCA and Ag20–TCA complexes showed similar bands in the 800–1500 cm–1 region. According to the interaction energies calculated (∼20 kcal mol–1), it is expected that such values overpass a threshold required to observe the SERS effect.41 However, further studies should address other variables such as the effect of pH, charge of Ag clusters and molecules, different metallic substrates, as well as calculations of Raman activities using different laser wavelengths.
Figure 5.
(A) Raman spectra using the Ag4 cluster and (B) Raman spectra using the Ag20 cluster. For the simulated static normal Raman spectra of the Gly, DKP, BCA, and TCA complexes with a metallic cluster the differential cross sections of molecules (in units of 10–31 cm2/sr) were calculated on the basis of static polarizability derivatives and using an incident wavelength of 514.5 nm. The bands showed are given by a Lorentzian width of 20 cm–1.
Conclusions
Theoretical calculations of the noncovalent intermolecular interactions between imidazo[1,2-a]pyrazines and silver clusters were reported in the framework of dispersion-corrected DFT methods. The main conclusions can be summarized as follows:
In terms of a computational modeling that compensates resource and time, the Ag20 cluster is a very useful model to study the interactions of Ag nanoparticles with imidazo[1,2-a]pyrazines. In terms of an accurate prediction, dispersion-corrected DFT methods are required to estimate interaction energies in the noncovalent complexes studied.
The mean values for the difference in the interaction energies calculated appeared to be consistently overestimated with APF-D, especially when complexes are formed with the Ag20 cluster.
Mulliken population analysis suggests a transfer of charge across the Ag···N(amine), red shifts for Raman vibrations suggest strong interactions, and charge redistribution involving HOMO and LUMO orbitals give rise to the formation of stable noncovalent complexes that can surpass a dissociation limit.
These simulations constitute a useful and valuable first element for further experimental characterization of these unusual classes of cyclic amidines.
Computational Methodology
To account for dispersive interactions in the electron density of the noncovalent molecule-metal cluster systems, all the electronic structure calculations were studied through DFT using the Grimme30 and the APF31 dispersion correction methods over the B3LYP functional. A mixed basis set (MBS) was set using the 6-31G+(d,p) basis set for C, H, O, and N atoms plus the LANL2DZ basis set for Ag. Details of relative interaction energies, normal modes frequencies, Kohn–Sham HOMO and LUMO orbitals, vertical excitation energies, Raman intensities calculations, basis-set superposition error, and the details of the computational methodology adopted for the present work are given in the Supporting Information.
Acknowledgments
This work was partially supported by the National Council of Science and Technology of Mexico, CONACyT (grant 269399) and the School of Science and Engineering, Tecnológico de Monterrey. The author would like to express sincere gratitude to Prof. V.A. Basiuk for his encouragment in the realization of this study.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.9b03127.
Details of computational methods; representative structures for Ag4 clusters; total density-of-states calculated for Ag4 (a) and Ag20 (b) clusters; frontier orbitals HOMO and LUMO for molecules (1–4); absolute electronic energies (in Ha) for the isolated molecules (1–4) and metallic cluster-organic molecule systems HOMO and LUMO energies calculated for the isolated molecules; and Mulliken atomic charges for the Ag20–molecule complexes (PDF)
The author declares no competing financial interest.
Supplementary Material
References
- Basiuk V. A. Imidazo[1,2-a]pyrazines. Russ. Chem. Rev. 1997, 66, 187–204. 10.1070/rc1997v066n03abeh000325. [DOI] [Google Scholar]
- Harvey E. N. Studies on bioluminescence. Am. J. Physiol. 1917, 42, 318–341. 10.1152/ajplegacy.1917.42.2.318. [DOI] [Google Scholar]
- Kishi T.; Goto T.; Hirata Y.; Shimomura O.; Johnson F. H. Cypridina bioluminescence I. Structure of Cypridina luciferin. Tetrahedron Lett. 1966, 7, 3427–3436. 10.1016/s0040-4039(01)82806-9. [DOI] [Google Scholar]
- Goto T.; Iio H.; Inoue S.; Kakoi H. Squid bioluminescence I. Structure of Watasenia oxyluciferin, a possible light-emitter in the bioluminescence of Watasenia scintillans. Tetrahedron Lett. 1974, 15, 2321–2324. 10.1016/s0040-4039(01)92245-2. [DOI] [Google Scholar]
- Shimomura O.; Johnson F. H. Peroxidized coelenterazine, the active group in the photoprotein aequorin. Proc. Natl. Acad. Sci. U.S.A. 1978, 75, 2611–2615. 10.1073/pnas.75.6.2611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mitani M.; Yokoyama Y.; Ichikawa S.; Sawada H.; Matsumoto T.; Fujimori K.; Kosugi M. Determination of horserdish peroxidase concentration using the chemiluminescence of Cypridina luciferin analogue, 2 methyl-6-(p-methyoxyphenyl)-3,7-dihydroimidazo[1,2-a]pyrazin-3-one. J. Biolumin. Chemilumin. 1994, 9, 355–361. 10.1002/bio.1170090602. [DOI] [PubMed] [Google Scholar]
- Mitani M.; Koinuma Y.; Kosugi M. Determination of glucose oxidase by means of the chemiluminescence of Cypridina luciferin analogue. Anal. Sci. 1994, 10, 267–269. 10.2116/analsci.10.267. [DOI] [Google Scholar]
- Kobayashi T.; Saga K.; Shimizu S.; Goto T. The application of chemiluminescence of a Cypridina luciferin analog to immobilized enzyme sensors. Agric. Biol. Chem. 1981, 45, 1403–1408. 10.1080/00021369.1981.10864704. [DOI] [Google Scholar]
- Abignente E.; Arena F.; De Caprariis P.; Nuzzetti R.; Marmo E.; Lampa E.; Rosatti F.; Ottavo R. Research on heterocyclic compounds. X. Imidazo[l,2-a]pyrazine derivatives: synthesis and antiinflammatory activity. Farm. Ed. Sci. 1981, 36, 61–80. [PubMed] [Google Scholar]
- Sablayrolles C.; Milhavet J. C.; Rechenq E.; Chapat J. P.; Cros G. H.; Boucard M.; Serrano J. J.; McNeill J. H. Synthesis of imidazo[1,2-a]pyrazine derivatives with uterine-relaxing, antibronchospastic, and cardiac-stimulating properties. J. Med. Chem. 1984, 27, 206–212. 10.1021/jm00368a018. [DOI] [PubMed] [Google Scholar]
- Kaminski J. J.; Hilbert J. M.; Pramanik B. N.; Solomon D. M.; Conn D. J.; Rizvi R. K.; Elliott A. J.; Guzik H.; Lovey R. G.; Domalski M. S.; Wong S.-C.; Puchalski C.; Gold E. H.; Long J. F.; Chiu P. J. S.; McPhail A. T. Antiulcer agents. 2. Gastric antisecretory, cytoprotective, and metabolic properties of substituted imidazo[1,2-a]pyridines and analogs. J. Med. Chem. 1987, 30, 2031–2046. 10.1021/jm00394a018. [DOI] [PubMed] [Google Scholar]
- Spitzer W. A.; Victor F.; Pollock G. D.; Hayes J. S. Imidazo[1,2-a]pyrimidines and imidazo[1,2-a]pyrazines: the role of nitrogen position in inotropic activity. J. Med. Chem. 1988, 31, 1590–1595. 10.1021/jm00403a018. [DOI] [PubMed] [Google Scholar]
- Bonnet P. A.; Michel A.; Laurent F.; Sablayrolles C.; Rechencq E.; Mani J. C.; Boucard M.; Chapat J. P. Synthesis and antibronchospastic activity of 8-alkoxy- and 8-(alkylamino)imidazo[1,2-a]pyrazines. J. Med. Chem. 1992, 35, 3353–3358. 10.1021/jm00096a008. [DOI] [PubMed] [Google Scholar]
- Meurer L. C.; Tolman R. L.; Chapin E. W.; Saperstein R.; Vicario P. P.; Zrada M. M.; MacCoss M. Synthesis and hypoglycemic activity of substituted 8-(1-piperazinyl)imidazo[1,2-a]pyrazines. J. Med. Chem. 1992, 35, 3845–3857. 10.1021/jm00099a012. [DOI] [PubMed] [Google Scholar]
- Myadaraboina S.; Alla M.; Saddanapu V.; Bommena V. R.; Addlagatta A. Structure activity relationship studies of imidazo[1,2-a]pyrazine derivatives against cancer cell lines. Eur. J. Med. Chem. 2010, 45, 5208–5216. 10.1016/j.ejmech.2010.08.035. [DOI] [PubMed] [Google Scholar]
- Goel R.; Luxami V.; Paul K. Recent advances in development of imidazo[1,2-a]pyrazines: synthesis, reactivity and their biological applications. Org. Biomol. Chem. 2015, 13, 3525–3555. 10.1039/c4ob01380h. [DOI] [PubMed] [Google Scholar]
- Basiuk V. A.; Navarro-González R. Identification of hexahydroimidazo[1,2-a]pyrazine-3,6-diones and hexahydroimidazo[1,2-a]imidazo[1,2-d]pyrazine-3,8-diones, unusual products of silica-catalyzed amino acid thermal condensation and products of their thermal decomposition using coupled high-performance liquid chromatography–particle beam mass spectrometry and gas chromatography–Fourier transform infrared spectroscopy–mass spectrometry. J. Chromatogr. A 1997, 776, 255–273. 10.1016/s0021-9673(97)00331-2. [DOI] [Google Scholar]
- Basiuk V. A.; Navarro-González R.; Basiuk E. V. Pyrolysis of alanine and α-aminoisobutyric acid: identification of less-volatile products using gas chromatography/Fourier transform infrared spectroscopy/mass spectrometry. J. Anal. Appl. Pyrolysis 1998, 45, 89–102. 10.1016/s0165-2370(98)00057-6. [DOI] [Google Scholar]
- Basiuk V. A. Pyrolysis of valine and leucine at 500°C: identification of less-volatile products using gas chromatography-Fourier transform infrared spectroscopy-mass spectrometry. J. Anal. Appl. Pyrolysis 1998, 47, 127–143. 10.1016/s0165-2370(98)00086-2. [DOI] [Google Scholar]
- Basiuk V. A.; Douda J. Pyrolysis of poly-glycine and poly-l-alanine: analysis of less-volatile products by gas chromatography/Fourier transform infrared spectroscopy/mass spectrometry. J. Anal. Appl. Pyrolysis 2000, 55, 235–246. 10.1016/s0165-2370(00)00072-3. [DOI] [Google Scholar]
- Basiuk V. A.; Douda J. Analysis of less-volatile products of poly-L-valine pyrolysis by gas chromatography/Fourier transform infrared spectroscopy/mass spectrometry. J. Anal. Appl. Pyrolysis 2001, 60, 27–40. 10.1016/s0165-2370(00)00104-2. [DOI] [Google Scholar]
- Contreras-Torres F. F.; Basiuk V. A. Imidazo[1,2-a]pyrazine-3,6-diones derived from α-amino acids: a theoretical mechanistic study of their formation via pyrolysis and silica-catalyzed process. J. Phys. Chem. A 2006, 110, 7431–7440. 10.1021/jp061331m. [DOI] [PubMed] [Google Scholar]
- Basiuk V. A. Calculated gas-phase infrared spectra of 2,2,5,5,8,8-hexamethylhexahydroimidazo[1,2-a]-pyrazine-3,6-dione: relative performance of Hartree–Fock and hybrid density functional theory methods. Spectrochim. Acta, Part A 2001, 57, 1271–1282. 10.1016/s1386-1425(00)00476-5. [DOI] [PubMed] [Google Scholar]
- Basiuk V. A. Calculated gas-phase infrared spectra of imidazo [1,2-a]pyrazinediones derived from alanine. Spectrochim. Acta, Part A 2003, 59, 1867–1879. 10.1016/s1386-1425(02)00422-5. [DOI] [PubMed] [Google Scholar]
- Contreras-Torres F. F.; Basiuk V. A. Theoretical prediction of gas-phase infrared spectra of imidazo[1,2-a] pyrazinediones and imidazo[1,2-a]imidazo[1,2-d] pyrazinediones derived from glycine. Spectrochim. Acta, Part A 2005, 61, 2560–2575. 10.1016/j.saa.2004.09.022. [DOI] [PubMed] [Google Scholar]
- Moskovits M. Surface-enhanced spectroscopy. Rev. Mod. Phys. 1985, 57, 783–826. 10.1103/revmodphys.57.783. [DOI] [Google Scholar]
- Moskovits M. Surface-enhanced Raman spectroscopy: a brief retrospective. J. Raman Spectrosc. 2005, 36, 485–496. 10.1002/jrs.1362. [DOI] [Google Scholar]
- Morton S. M.; Jensen L. Understanding the molecule–surface chemical coupling in SERS. J. Am. Chem. Soc. 2009, 131, 4090–4098. 10.1021/ja809143c. [DOI] [PubMed] [Google Scholar]
- Pakiari A. H.; Jamshidi Z. Interaction of amino acids with gold and silver clusters. J. Phys. Chem. A 2007, 111, 4391–4396. 10.1021/jp070306t. [DOI] [PubMed] [Google Scholar]
- Grimme S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. 10.1002/jcc.20495. [DOI] [PubMed] [Google Scholar]
- Austin A.; Petersson G. A.; Frisch M. J.; Dobek F. J.; Scalmani G.; Throssell K. A density functional with spherical atom dispersion terms. J. Chem. Theory Comput. 2012, 8, 4989–5007. 10.1021/ct300778e. [DOI] [PubMed] [Google Scholar]
- Becke A. D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. 10.1103/physreva.38.3098. [DOI] [PubMed] [Google Scholar]
- Lee C.; Yang W.; Parr R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. 10.1103/physrevb.37.785. [DOI] [PubMed] [Google Scholar]
- McKee M. L.; Samokhvalov A. Density functional study of neutral and charged silver clusters Agn with n=2–22. Evolution of properties and structure. J. Phys. Chem. A 2017, 121, 5018–5028. 10.1021/acs.jpca.7b03905. [DOI] [PubMed] [Google Scholar]
- Zhao S.; Li Z.-H.; Wang W.-N.; Liu Z.-P.; Fan K.-N.; Xie Y.; Schaefer H. F. the uniform electron gas limit important for small Ag clusters? Assessment of different density functionals for Agn (n ≤ 4). J. Chem. Phys. 2006, 124, 184102. 10.1063/1.2193512. [DOI] [PubMed] [Google Scholar]
- Wang J.; Wang G.; Zhao J. Structures and electronic properties of Cu20, Ag20, and Au20 clusters with density functional method. Chem. Phys. Lett. 2003, 380, 716–720. 10.1016/j.cplett.2003.09.062. [DOI] [Google Scholar]
- Zhao L.; Jensen L.; Schatz G. C. Pyridine-Ag20 cluster: a model system for studying surface-enhanced Raman scattering. J. Am. Chem. Soc. 2006, 128, 2911–2919. 10.1021/ja0556326. [DOI] [PubMed] [Google Scholar]
- Minenkov Y.; Cavallo L. Ground-state gas-phase structures of inorganic molecules predicted by density functional theory methods. ACS Omega 2017, 2, 8373–8387. 10.1021/acsomega.7b01203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oliveira L. F. L.; Tarrat N.; Cuny J.; Morillo J.; Lemoine D.; Spiegelman F.; Rapacioli M. Benchmarking density functional based tight-binding for silver and gold materials: from small clusters to bulk. J. Phys. Chem. A 2016, 120, 8469–8483. 10.1021/acs.jpca.6b09292. [DOI] [PubMed] [Google Scholar]
- Chelvayohan M.; Mee C. H. B. Work function measurements on (110), (100) and (111) surfaces of silver. J. Phys. C: Solid State Phys. 1982, 15, 2305–2312. 10.1088/0022-3719/15/10/029. [DOI] [Google Scholar]
- Tenorio F. J.; Sato-Berrú R.; Saniger J. M.; Martínez A. Is the donor–acceptor electronegativity a good indicator for the surface enhanced Raman scattering (SERS)?. Int. J. Quantum Chem. 2012, 112, 3516–3524. 10.1002/qua.24153. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.


