Next Article in Journal
Catalase Detection via Membrane-Based Pressure Sensors
Previous Article in Journal
The Monitoring and Cell Imaging of Fe3+ Using a Chromone-Based Fluorescence Probe
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring Geometrical, Electronic and Spectroscopic Properties of 2-Nitroimidazole-Based Radiopharmaceuticals via Computational Chemistry Methods

1
Faculty of Physics, Babeș-Bolyai University, Str. M. Kogălniceanu 1, RO-400084 Cluj-Napoca, Romania
2
Department of Nuclear Medicine, County Clinical Hospital, Clinicilor 3-5, RO-400006 Cluj-Napoca, Romania
3
Institute for Research, Development and Innovation in Applied Natural Sciences, Babeș-Bolyai University, Str. Fântânele 30, RO-400327 Cluj-Napoca, Romania
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(7), 1505; https://doi.org/10.3390/molecules29071505
Submission received: 14 February 2024 / Revised: 22 March 2024 / Accepted: 25 March 2024 / Published: 28 March 2024
(This article belongs to the Section Computational and Theoretical Chemistry)

Abstract

:
Tumor hypoxia plays an important role in the clinical management and treatment planning of various cancers. The use of 2-nitroimidazole-based radiopharmaceuticals has been the most successful for positron emission tomography (PET) and single photon emission computed tomography (SPECT) imaging probes, offering noninvasive means to assess tumor hypoxia. In this study we performed detailed computational investigations of the most used compounds for PET imaging, focusing on those derived from 2-nitroimidazole: fluoromisonidazole (FMISO), fluoroazomycin arabinoside (FAZA), fluoroetanidazole (FETA), fluoroerythronitroimidazole (FETNIM) and 2-(2-nitroimidazol-1-yl)-N-(2,2,3,3,3-pentafluoropropyl)acetamide (EF5). Conformational analysis, structural parameters, vibrational IR and Raman properties (within both harmonic and anharmonic approximations), as well as the NMR shielding tensors and spin-spin coupling constants were obtained by density functional theory (DFT) calculations and then correlated with experimental findings, where available. Furthermore, time-dependent DFT computations reveal insight into the excited states of the compounds. Our results predict a significant change in the conformational landscape of most of the investigated compounds when transitioning from the gas phase to aqueous solution. According to computational data, the 2-nitroimidazole moiety determines to a large extent the spectroscopic properties of its derivatives. Due to the limited structural information available in the current literature for the investigated compounds, the findings presented herein deepen the current understanding of the electronic structures of these five radiopharmaceuticals.

1. Introduction

Hypoxia is a common feature of the microenvironment of solid tumors and is associated with metastasis, recurrence and the development of resistance to radiotherapy, chemotherapy and immunotherapy [1,2,3]. There is a continuous effort to develop methods able to facilitate proper oncological treatment [2]. Among these methods, molecular imaging techniques like PET and SPECT, using nitroimidazole derivatives, have shown promising results [4,5]. Their effectiveness is due to their ability to undergo oxygen dependent reduction specifically in hypoxic cells. A significant number of 18F-labelled radiopharmaceuticals for hypoxia imaging have been successfully developed and evaluated in clinical trials [6]. Fluromisonidazole (FMISO) was the first 2-nitroimidazole (2nim) derivative to undergo clinical testing and remains the preferred standard in clinical imaging. However, its high lipophilicity results in low uptake and low-contrast images, thereby prompting the development of more optimized radiopharmaceuticals such as fluoroazomycin arabinoside (FAZA), fluroetanidazole (FETA), fluoroerythronitroimidazole (FETNIM) and 2-(2-nitroimidazol-1-yl)-N-(2,2,3,3,3-pentafluoropropyl) acetamide (EF5) (see Figure 1) [5].
The current understanding of the molecular mechanism of nitroimidazole compounds in hypoxia cells is limited. A recent study by Rashed et al. [7] investigated the mechanism of action of two nitroimidazoles (IAZA and FAZA) in a head and neck cancer model. They concluded that covalent binding of the investigated compounds to proteins can inhibit catalytic functions of critical cellular enzymes and they plan on further analyzing the proteins targeted by the nitroimidazoles. Once the targets are identified, studies of receptor–ligand binding could further advance the understanding of the mechanism of action and would require detailed structural (conformational) information of the nitroimidazole derivatives.
Thus, we consider that our results related to the structure and conformations of the investigated compounds could be useful for docking studies and ligand–receptor interactions [8,9]. The predicted spectroscopic properties might also be of help for drug monitoring [10], for drug detection or for pharmaceutical research and development [11].
Although extensively utilized in nuclear medicine, detailed information on their molecular and electronic configurations remains limited. Therefore, in this study, we present a computational structural analysis of the aforementioned radiopharmaceuticals. Despite their development beginning in the late 1980s, a review of the current literature reveals a lack of data regarding the electronic structures of these compounds. According to the Web of Science database, only 64 papers were found when searching for the keywords “radiopharmaceuticals” AND “DFT”.
Our aim was to provide a characterization of their conformational landscape and to determine the optimal structures of possible conformers, as well as to analyze their vibrational, NMR and electronic spectroscopic properties using density functional theory calculations. All these compounds show favorable properties for hypoxia imaging; however, there are still limitations in their implementation in the clinical setting [5]. Furthermore, compounds such as FMISO will act as a benchmark when evaluating the performance of new imaging agents. To this extent, we consider that a thorough structural characterization of these compounds could provide information that can be useful when optimizing and developing new agents.

2. Results and Discussion

2.1. Conformers and Boltzmann Population Analysis

The conformers of the five compounds in the gas phase were generated from the 2D relaxed potential energy surfaces (PES) defined by the dihedral angles shown in Table 1. In the case of FAZA, because of its low structural flexibility we investigated only the 1D PES generated by scanning the C4-N3-C9-C10 dihedral angle. The scanned dihedrals were incremented 24 times in steps of 15° over the 360° interval.
We identified 9 local minima for FMISO, 11 for FETNIM, 3 for FAZA and 2 for FETA and EF5. The geometries of the minima for each compound were further fully optimized at the B3LYP/6-311+G(d,p) level of theory, both in gas phase and water. These new geometry optimizations revealed a lower number of unique conformations for FMISO and FETNIM (5 and 7, respectively). The relative Boltzmann populations were calculated using the following formula:
P i = e G i k B T i e G i k B T
where ∆Gi are the relative (Gibbs) free energies, kB is the Boltzmann constant, and T is the temperature (T = 298 K) [12,13,14]. The Gibbs free energy was used for Boltzmann population calculations as recommended in ref. [12].
The structures of the most stable conformers in the gas phase and water are shown in Figure S1, while Boltzmann populations are presented in Table 2. Geometry re-optimizations in water show small changes in the structures of the conformers. On the other hand, as can be seen in Table 2, important changes are noted between the Boltzmann populations of the conformers when passing from the gas phase to the liquid state. Regarding the energetic order of the conformations for the five investigated compounds, the following matchings were observed:
  • FMISO: 1g ↔ 1w, 2g ↔ 3w, 3g ↔ 5w, 4g ↔ 2w, 5g ↔ 4w
  • FETNIM: 1g ↔ 1w, 2g ↔ 3w, 3g ↔ 2w, 4g ↔ 4w, 5g ↔ 5w, 6g ↔ 6w, 7g ↔ 7w
  • FAZA: 1g ↔ 1w, 2g ↔ 3w, 3g ↔ 2w
  • FETA: 1g ↔ 2w, 2g ↔ 1w
  • EF5: 1g ↔ 1w, 2g ↔ 2w
As an example, the above mentioned correspondences for FMISO means that the conformer 1g is the most stable in the gas phase. It exhibits structural similarity with the conformer 1w, which attains maximal stability in an aqueous environment. The 3g conformer, which ranks as the third most stable in the gas phase, experiences a decrease in relative stability in water, where it is ranked as the fifth most stable, denoted as 5w.
According to data collected in Table 2, in the case of FMISO, the most stable conformer in the gas phase remains the most stable in water as well, with a similar relative population 62.15% vs. 65.91%. The conformers 2g and 3g suffer a destabilization due to solvation and they transform into the 3w and 5w conformers in water, with the relative population of 7.59% and 2.33%, respectively. On the other hand, the conformer 4g shows an increase in relative population to 20.56% in water, becoming the second most stabilized.
For FETNIM, most conformers suffer a slight destabilization due to solvation effects; however, the conformer 3g shows an increase in relative population in water from 8.22% to 23.16%, becoming the second most stabilized in the aqueous environment.
Significant solvent destabilization effects can be observed for FAZA 1g and FETA 1g, while a large increase in relative population can be seen for the conformer FAZA 3g from 0.29% in gas phase to 28.46% in water. This change from the gas phase to water comes with a structural reconfiguration, mainly in the dihedral angles related to the oxolane group.
The two conformers of EF5 show negligible solvation effects and retain the same energetic order in the gas phase and water.

2.2. Vibrational Properties

2.2.1. 2-Nitroimidazole

The theoretical IR and Raman spectra of 2nim in the harmonic (scaled frequencies) and anharmonic approximations in the gas phase and water are presented in Figure 2. The assignments of the most representative bands are given in Table 3.
In the fingerprint region of the IR spectrum, we note the most intense calculated bands at 1330 and 1546 cm−1, both involving the vibrations of the nitro group, assigned to the ν(C-N) + νsym(NO2) and νasym(NO2) normal modes. On the other hand, the most Raman active predicted mode (ν(C-N) + ν(C=C)) is calculated at 1318 and 1345 cm−1 by the harmonic and anharmonic approximations, respectively (see Figure 2b). This mode, together with those calculated at 1059/1206 and 1143/1162 cm−1 form the set of characteristic Raman bands for 2nim.
At high wavenumbers, the three observed bands were assigned to a ν(NH) mode at 3512 cm−1 and two ν(CH) modes at 3165 and 3141 cm−1.
Experimental data for the IR spectrum of 2nim were obtained from references [15,16]. As seen in Table 3, there is a good agreement between the experimental and computed IR wavenumbers. Expectedly, the experimental bands are generally better described by the anharmonic approximation with the exception of some observed bands at 623, 944, 3146 and 3164 cm−1 which are more closely reproduced by the scaled harmonic data.

2.2.2. 2-Nitroimidazole Derivatives

The theoretical harmonic IR (gas-phase) and Raman (water) spectra of the five nitroimidazole derivatives are presented in Figure 3 (spectra of 2nim are included for comparison purposes). The assignments of the most representative bands for each compound are available in the Supplementary Materials (Tables S1–S5).
We can distinctly observe characteristic spectral bands associated with 2nim in the spectra of all five compounds. Such calculated bands include the ν(C-N)imidazole + νsym(NO2) mode at ~1330 cm−1, the ν(C-N)imidazole + β(NH) + β(CH) mode at ~1456 cm−1, νasym(NO2) at ~1546 cm−1 and two ν(CH) modes at ~ 3141 and 3165 cm−1. The specific wavenumbers for these bands can be found in Table 3. It is noteworthy that the first three bands exhibit a red-shift, whereas the second ν(CH) mode in FAZA experiences a blue-shift of 23 cm−1. Notably, the spectral bands at 997 and 1056 cm−1 do not have a corresponding mode in the calculations for the 2nim derivatives.
The Boltzmann-weighted spectra presented in Figure 3 take into account the conformers with a Boltzmann population higher that 5%. Comparative figures illustrating the spectral variances among different conformers for each compound can be found in the Supplementary Materials (Figures S2–S6). We observe that the composite spectra are following the spectral landscape of the spectrum for the most stable conformer. For FMISO and FETNIM, significant differences in the spectral bands of the predicted conformers are observed, particularly in the lower fingerprint region. However, as the Boltzmann population for the most stable conformer is much larger, its contribution to the composite spectra outweighs that of the other conformers. A notable exception is evident in the composite spectra of FMISO where one can observe two distinct ν(OH) bands: at 3696 cm−1 originating from the conformers 1g and 2g, and at 3621 cm−1 attributed to the contribution of the conformer 3g (IR harmonic gas phase). In the case of FETA, both conformers exhibit similar vibrational patterns, without any obvious individual contributions. Finally, for EF5, where the Boltzmann populations of the two most stable conformers are very close, three doublets are discerned in the composite spectra which are due to the contributions of both conformers: 383 (1g)/371 (2g), 651 (1g)/675 (2g) and 3418 (1g)/3444 (2g) cm−1.
Conformer analysis and Boltzmann population analysis show an energetic reordering of the most stable conformers for each compound when considering water as a solvent. Radiopharmaceuticals are usually administered to patients through intravenous injections of aqueous solutions. Figure 3b presents the calculated harmonic Raman spectra for 2nim and the Boltzmann-weighted averages of the most stable conformers for the five derivatives in water. In the case of FETNIM, FETA and FAZA, no specific conformer contributions can be observed due the similarities of the corresponding spectra for the conformers with the highest predicted populations. For EF5, one doublet can be discerned at 3469 (1w)/3446 (2w) and two conformer-specific bands at 203 (2w) and at 180 (1w) cm−1. In the case of FMISO, calculations predict significant conformer-specific characteristics in the averaged Raman spectrum weighted by the three most stable conformers in water: two doublets at 3693 (1w)/3650 (2w) and at 1272 (1w)/1253 (2w + 3w) cm−1, as well as the band at 1065 cm−1 arising due to the contributions of the conformers 1w and 3w and the band at 632 cm−1 due to the conformers 2w and 3w.
There are no significantly intense overtones or combination modes calculated for the anharmonic spectra. Figures presenting the comparison between the IR and Raman harmonic and anharmonic spectra of the most stable conformer in the gas phase are available in the Supplementary Materials for all compounds (Figures S7–S11). Calculated anharmonic wavenumbers are shown in the assignment tables in the Supplementary Materials Tables S1–S5.
In Table S1, we provide a tentative comparison with experimental IR data from Ref. [17] for FMISO. It is important to note that the reference only provides IR wavenumbers, without IR intensity. Therefore, our assignment in Table S1 is based solely on the positions of the bands in the spectrum. We observe a good agreement with the calculated anharmonic spectra of FMISO for most of the bands, particularly in the fingerprint region. However, the harmonic bands at 786, 1121 and 1535 cm−1 were found to be in better agreement with the experimental band at 795, 1116 and 1534 cm−1. Such discrepancies can be due to intermolecular interactions but also to differences in the anharmonicity of the normal modes.
Next, we analyzed the calculated IR and Raman spectra (Figures S12–S16) for the five compounds in which the 19F atom was replaced by the 18F radioisotope. The calculations were performed for the most stable conformers in each case. The harmonic spectra for FMISO, FETNIM, FETA and FAZA exhibit a distinct ν(CF) mode at 977, 992, 997 and 991 cm−1. In all four cases, this band suffers a blue-shift of 5, 6, 11 and 9 cm−1 in the 18F spectra with respect to the 19F spectra. When comparing anharmonic frequencies, this difference decreases for FETNIM (3 cm−1), remains the same for FETA and increases for FMISO (6 cm−1) and FAZA (13 cm−1). No other significant differences were observed in the calculated harmonic bands of the IR and Raman spectra of these compounds.
In the case of EF5, spectroscopic calculations reveal the presence of multiple bands involving ν(CF) and ν(CF2) vibrations. Based on the results of the conformational analysis, we considered both conformers as the relative Boltzmann populations were very similar 55.05% for 1g and 44.95% for 2g. Upon substituting one fluorine atom with the 18F isotope we have not noticed any significant shifts in the 18F harmonic spectra for these bands in either conformer, as all differences were less than 5 cm−1.
To conclude this section, it is worth noting that all five investigated radiopharmaceuticals maintain the vibrational characteristics of the 2nim template, both in terms of band positions and intensities. Furthermore, the computed anharmonic frequencies generally exhibit better agreement with experimental data. However, in some instances, the scaled harmonic frequencies have also proven to be useful for reliable assignments of vibrational spectra.

2.3. NMR Data

Given the lack of structural information for the radiopharmaceuticals, we compiled in this section the available experimental NMR data for the investigated compounds and compared the observed values with those predicted by quantum chemical calculations carried out at the PCM-B3LYP/6-311+G(d,p) level of theory (see Table 4, Table 5, Table 6, Table 7 and Table 8).
For all compounds the geometries have been fully optimized prior to NMR calculation at the same level of theory. Atom numbering schemes for all compounds are given in Figure 1. The computed NMR data were determined using the appropriate solvent specified in the respective available experimental works. Tetramethylsilane (TMS) was used as an NMR reference standard for FMISO (both for 1H and 13C spectra), FAZA (both for 1H and 13C spectra), FETA (for 1H spectrum), EF5 (for 1H spectrum) and for 2nim (both for 1H and 13C spectra), whereas the CFCl3 or CF3COOH standards have been employed for 19F nuclei, according to the experimental conditions. The calculated spin-spin couplings are reported merely where experimental values are available. No scaling procedure was applied for any type of nuclei.
First, we have to note that in the case of the FETNIM compound, NMR experimental data have only been documented for one of its precursors, as reported by Yang et al. [18]. To our knowledge, NMR data for the compound itself are not available and it will be excluded from the following discussion. However, we have provided computational data for the chemical shifts and spin-spin couplings in Table S6 in the Supplementary Material for future reference. On the other hand, for the FMISO and FAZA compounds, a more substantial body of experimental NMR data has been amassed for the 1H, 13C and 19F nuclei and therefore allow for a more detailed analysis.
As seen in Table 4, Table 5, Table 6, Table 7 and Table 8, the overall agreement between the experimental and calculated NMR data is good, both for chemical shifts and J coupling constants. Significant discrepancies are evident for the chemical shifts associated with the 13C nuclei, which is a common behavior (see, for instance, ref. [19]) and it has not prevented the assignments for these nuclei. For the chemical shifts of the 19F nuclei, large deviations between experiment and theory have been observed (e.g., −230.70 ppm experimental vs. −266.16 ppm calculated, in DMSO for FMISO). In order to mitigate systematic errors for these nuclei, we employed a linear regression approach, a common practice found in the literature [20,21,22]. Thus, we fitted the calculated shifts against the experimental values (see Figure S17), resulting in the following equation for the scaled 19F chemical shifts at the B3LYP/6-311+G(d,p) level of theory:
δ s c a l e d ( ppm ) = 0.90977 × δ p r e d i c t e d ( ppm ) + 10.71500
The goodness of fit is supported by the value of R2 = 0.9992, as well as the statistical parameters of mean absolute error (MAE = 1.486 ppm) and root mean square deviation (RMSD = 1.713 ppm).
Based on this scaling procedure, we can predict the chemical shifts for 19F nuclei in FETA and FETNIM at −225.08 and −232.69 ppm, respectively.
For FMISO, one can note a much lower value of the calculated chemical shift for the H19 nucleus with respect to its experimental counterpart (see Table 4). This discrepancy may be attributed to intermolecular hydrogen bonding (HB) interactions involving this particular atom. We note that the optimized conformer with intramolecular HB O12-H19-O7 still does not replicate successfully the observed chemical shift.
For each compound, J couplings have been predicted by calculations at least in qualitative agreement with experimental results. Largest deviations between experiment and theory are observed for 1JC11H21 and 3JF13H18 in case of FMISO, and 2JF19C12 and 1JF19C18 for FAZA.
Table 4. Experimental and calculated NMR data for FMISO.
Table 4. Experimental and calculated NMR data for FMISO.
Nucleus aExperimental Data bCalculated Data d
Chemical Shift (ppm)J Coupling (Hz)Chemical Shift (ppm)J Coupling (Hz)
H(14) H(15)7.64, 7.22
7.20, 7.15
7.44, 7.35-
H(16)4.00–4.10
4.304.26
n.a. c3.90-
H(17)4.65
4.77
13.70, 3.60
13.00, 3.00
4.762JH17H16 = 12.70
4JH17H19 = 2.32
H(18)4.48–4.58
4.544.57
n.a.4.55-
H(19)5.64
5.30
n.a.2.17-
H(20) H(21)4.32–4.46
4.324.52
n.a.4.94, 4.89-
C(2)144.91n.a.154.60-
C(4) C(5)128.52, 127.35n.a.140.81, 136.10-
C(9)51.417.6055.403JC9H19 = 8.26
C(10)67.7019.3076.193JC10F13 = 15.15
C(11)84.58169.1093.081JC11H21 = 147.9
F(13)−230.7046.00 19.00−266.162JF13H20 = 46.46
2JF13H21 = 49.90
3JF13H18 = 12.79
a Atom numbering can be seen in Figure 1. b from ref. [23] (in DMSO) and ref. [17] (in CDCl3). Experimental values in italics for 1H are from ref. [17]. c not available. d solvent: DMSO or CHCl3; reference standard: TMS for 1H and 13C, CFCl3 for 19F.
Table 5. Experimental and calculated NMR data for FAZA.
Table 5. Experimental and calculated NMR data for FAZA.
Nucleus aExperimental Data bCalculated Data d
Chemical Shift (ppm)J Coupling (Hz)Chemical Shift (ppm)J Coupling (Hz)
H(14)7.79/7.681.00/1.007.723JH14H15 = 0.86
H(15)7.28/7.141.00/1.007.433JH15H14 = 0.86
H(20)6.48/6.441.00/1.506.403JH20H21 = 1.92
H(21)4.13/4.662.10/n.a. c4.393JH21H22 = 3.48
H(22)4.62/4.50n.a./6.00, 2.00
1.50
4.793JH22H23 = 8.06
3JH22H21 = 4.13
4F16H22 = 2.50
H(23)4.56/4.60n.a./n.a.4.30-
H(26)4.66/4.20n.a./57.005.09JF19H26 = 51.17
H(27)4.66/4.20n.a./57.004.87JF19H27 = 47.22
C(4)123.80/125.22n.a./125.22130.62-
C(5)126.90/128.25n.a./128.25137.12-
C(9)95.50/96.80n.a./n.a.104.38-
C(10)82.20/83.63n.a./n.a.94.92-
C(11)75.80/77.215.60/4.7082.283JF19C11 = 6.54
C(12)87.70/89.0420.60/20.6088.672JF19C12 = 13.61
C(18)82.20/83.63170.00/169.7085.301JF19C18 = −212.43
F(19)−227.20 d/
−147.72 e
n.a./n.a.−273.48 d/
−180.55 e
-
a Atom numbering can be seen in Figure 1. b from refs. [24] (in CD3OD) and [25] (in CD3OD); Experimental values in italics for 1H and 13C are from ref. [25]. c not available. d solvent: methanol; reference standards: TMS for 1H and 13C, CFCl3 and CF3COOH for 19F; e Values in italics for 19F refer to CF3COOH reference standard.
Table 6. Experimental and calculated NMR data for FETA.
Table 6. Experimental and calculated NMR data for FETA.
Nucleus aExperimental Data bCalculated Data d
Chemical Shift (ppm)J Coupling (Hz)Chemical Shift (ppm)J Coupling (Hz)
H(14)7.401.307.423JH14H15 = 0.81
H(15)7.141.307.433JH15H14 = 0.81
H(18)4.83n.a. c<5.07>-
H(19)
H(21)3.5026.75<3.75><3JFH>trans = 27.09
H(22)4.80<3JHH>cis = 5.71
H(23)4.4447.50<4.70><2JFH> = 49.25
H(24)4.80<3JHH> = 5.71
a Atom numbering can be seen in Figure 1. b from ref. [26]; c not available; d solvent: DMSO; reference standard: TMS, < > represents average values.
Table 7. Experimental and calculated NMR data for EF5.
Table 7. Experimental and calculated NMR data for EF5.
Nucleus aExperimental Data bCalculated Data d
Chemical Shift (ppm)J Coupling (Hz)Chemical Shift (ppm)J Coupling (Hz)
H(14)7.54n.a. c7.63-
H(15)7.15n.a.7.43-
H(23)5.37n.a.<5.22>-
H(24)
H(25)8.22n.a.7.14-
H(26)4.06n.a.<3.97>-
H(27)
F(21,22)−81.70n.a.−102.33-
F(18,19)−118.7616.00−145.503JFF = 19.15
a Atom numbering can be seen in Figure 1. b from ref. [27]; c not available; d solvent: water, reference standards: TMS for 1H, CHCl3 for 19F; < > represents average values.
Table 8. Experimental and calculated NMR data for 2nim.
Table 8. Experimental and calculated NMR data for 2nim.
Nucleus aExperimental Data bCalculated Data d
Chemical Shift (ppm)J Coupling (Hz)Chemical Shift (ppm)J Coupling (Hz)
H(9)n.a.n.a. c10.14-
H(10) H117.41
7.26
n.a.7.54-
C(2)146.10n.a.154.95-
C(4) C(5)126.35n.a<134.78>-
a Atom numbering can be seen in Figure 1. b from refs. [15,28]; Experimental values in italics for 1H are from ref. [28]. c not available; d solvent: DMSO; reference standard: TMS, < > represents average values.

2.4. TD-DFT Data

2.4.1. 2-Nitroimidazole

In the following, we were interested in the electronic absorption properties of the compounds under examination. Particularly, we focused on calculating the absorption wavelengths and the nature of the involved transitions, as well as the electronic density changes triggered by the electronic excitation of these radiopharmaceuticals. For the following discussion, we considered only the electronic transitions with λabs ≥ 200 nm and f ≥ 0.01.
The three singlet-singlet electronic transitions of 2nim in water, with oscillator strengths larger than 0.01, are listed in Table 9.
According to quantum chemical calculations, the transition to the first singlet excited state has a very low oscillator strength and the λmax band in the UV–Vis, predicted at 315 nm, is mainly due to the HOMO → LUMO excitation with a contribution of 99.12%. The other two significantly intense bands appear as a result of S0 → S5 and S0 → S5 transitions (see Table 9).
The low probability of the S0 → S1 excitation is due to a poor overlap of the orbitals involved in this electronic transition, and consequently, a low transition dipole moment. The corresponding natural transition orbitals (NTOs) [29] HOTO and LUTO for this transition are depicted in Figure S18. On the other hand, for the transition to the second excited state the HOTO and LUTO NTOs suggest a good overlap, particularly on the nitro group, which confirms the high intensity of the S0 → S2 excitation.
It is worth mentioning that the transition to the lowest excited state does not involve the HOMO → LUMO excitation for 2nim, while for its derivatives, the same excitation contributes to a lesser extent. Such behaviors have been noted very recently by Kimber and Plasser [30]. These authors concluded that the interplay of the terms corresponding to the dynamic electron−hole binding energy, exchange repulsion between electron and hole and/or to the response of the exchange-correlation potential govern the different types of excited states like locally excited or charge-transfer character and influence the pairs of occupied and unoccupied molecular orbitals involved in particular transitions.
To the best of our knowledge, there are no experimental UV–Vis data for 2nim, but Yu and Bernstein [31] reported such data for the 4nim isomer.
To validate the computed data for 2nim, we also calculated the absorption spectrum of 4nim, being able to quantitatively reproduce the experimental data. That is, the three main peaks observed at 311, 287 and 203 nm were calculated at 300, 274 and 201 nm, respectively. Our calculated data for 4nim agree very well with the computationally predicted values at the CASSCF(10,7)/6-31G(d) level of theory [31]. Based on the agreement between experiment and theory for 4nim, we estimate that the calculated data for 2nim will also demonstrate a high degree of reliability, thus providing confidence in the validity of our computational results for this compound.
The PBE0 [32] and ωB97XD [33] density functionals are also frequently employed for the calculation of electronic transitions. Both predict UV–Vis spectra shapes similar to that obtained by using the B3LYP functional, with zero intensity for the S0 → S1 transition. Within the 200–350 nm wavelength range, these functionals yield band positions at 306 nm, 234 nm and 200 nm for PBE0 and 293 nm, 222 nm and 197 nm for ωB97XD, respectively.

2.4.2. 2-Nitroimidazole Derivatives

As shown in Figure 4, the calculated λmax bands for the derivatives display remarkable similarity, with each of them closely following the pattern observed for 2nim.
Table 10 lists the calculated excitation energies of the most stable conformers in water for the five 2-nitroimidazole derivatives. In contrast to the results obtained for the 2nim molecule, the transition to the S1 excited singlet state is allowed for all five compounds. Furthermore, the largest contributions to the S1 excited state are HOMO-2 → LUMO (60%) predicted at 331 nm for FMISO, HOMO-3 → LUMO (59%) at 332 nm for FETNIM, HOMO-4 → LUMO (58%) at 328 nm for FAZA, HOMO-1 → LUMO (78%) at 328 nm for FETA and HOMO-3 → LUMO (41%) at 335 nm for EF5. The contribution of the HOMO → LUMO transition to the S1 excited state is less than 10% in each case.
As seen in Table 10, the largest contribution for the S2 singlet excited state, corresponding to the most intense band, is due to the HOMO → LUMO transitions for all five compounds: 89% for FMISO predicted at 315 nm, 76% for FETNIM at 318 nm, 73% for FAZA at 317 nm, 60% for FETA at 323 nm and 54% for EF5 at 322 nm.
For FMISO, the band predicted at 316 nm appears as a result of the joint contributions of the transitions to the S1 and S2 excited states. For FETNIM and FAZA, the λmax bands at 319 nm are again attributed to the transitions to the S1 and S2 states. However, for FETNIM we observe an additional band at 286 nm, while for FAZA a shoulder is seen at 297 nm. They are assigned to the transitions to the S3 excited states with the main contributions being HOMO-1 → LUMO, 92.78% for FETNIM and HOMO-2 → LUMO, 68.72%, for FAZA (see Table 10).
The band at 314 nm in the spectrum for FETA is due to the joint transitions to the S2, S3 and S4 states. In contrast to the other investigated compounds where the S2 transition was calculated with the highest oscillator strength, the transition to the S4 excited state of FETA, predicted at 312 nm, has the largest value for the oscillator strength and is mainly due to HOMO → LUMO (54.27%) and HOMO-1 → LUMO (37.03%) transitions.
A distinct characteristic for the UV–Vis spectrum of EF5 is the broadness of the 317 nm band. The main contributions to it comes from transitions to the S2 and S3 states, predicted at 322 and 307 nm, respectively, and mainly due to the HOMO → LUMO (for S2) and HOMO-1 → LUMO (for S3) transitions.
An elegant and useful way to describe how the electronic density changes as a result of excitation is by plotting the difference between the electronic densities of the two states involved in the transition. In Figure 5 are illustrated the electronic density differences (EDDs) between the state of interest for 2nim and the five derivatives. In the case of 2nim, we note a charge transfer from the imidazole ring to the nitro group during the transition to the excited state. All the other derivatives show basically the same pattern for EDDs, excepting the FETA and EF5 compounds for which the carbonyl group acts together with the imidazole group as a donor of electronic density to the nitro group.

3. Materials and Methods

For this work, we used state-of-the-art, yet confident, density functional theory (DFT) methods to obtain the calculated properties of the investigated compounds. All calculations were performed with the Gaussian 16, revision C.01 software package [34] by using the B3LYP hybrid exchange-correlation functional [35,36,37,38] in conjunction with Pople’s style split-valence basis sets [39].
Tight and very tight criteria were employed to define the convergence of the molecular geometries and SCF convergence. The ultrafine grid was applied for the numerical integration of the electronic density. All geometry optimizations were followed by vibrational frequency calculations in order to check and validate the minima on the potential energy surfaces (PES) of the investigated molecules. Gibbs free energies were calculated using the standard thermochemistry model (T = 298.15 K).
Relaxed PESs in the gas phase were calculated at the B3LYP/3-21G level of theory. Geometry optimization and IR and Raman spectra calculations were performed at the B3LYP/6-311+G(d,p) level of theory for all the minima found on the PES. Solvent effects in water have been accounted for by using the implicit polarizable continuum model [40] within the integral equation formalism [41]. We chose water as a solvent because radiopharmaceuticals are usually administered intravenously to patients in an aqueous saline solution.
Both harmonic and anharmonic approximations were used to calculate the IR and Raman spectra. The harmonic spectra were scaled by a factor of 0.9668 [42], and the anharmonic wavenumbers were calculated using the vibrational second-order perturbation theory (VPT2) [43], as implemented in the Gaussian software package. The anharmonic IR and Raman spectra were simulated using the computed anharmonic wavenumbers. The IR intensities and Raman activities were those obtained within the harmonic approximation. Because of their very low calculated intensities, the overtones and the combination modes were not included in the simulation.
This approach was chosen because the anharmonic calculations can result in spurious very large activities [44,45]. Pure Lorentzian band shapes with a half-width at half maximum of 4 cm−1 were used for the plots of the calculated IR and Raman spectra.
NMR spectra calculations were performed using the GIAO (gauge-including atomic orbital) method [46,47] as implemented in the Gaussian software package, at the B3LYP/6-311+G(d,p) level of theory. The computed NMR shielding tensors were converted into chemical shifts using the standard referencing method to tetramethylsilane (TMS) for 1H and 13C, and to trichloro-fluoro-methane (CFCl3) for 19F. The TMS and CFCl3 shielding tensors were calculated at the same level of theory as that used for the 2nim derivatives, after geometry optimization of the reference molecules.
To assess the absorption properties of the compounds, time-dependent density functional theory (TD-DFT) methods [48] were employed, as implemented in the Gaussian software package. Single-point calculations on the excitation energies to singlet excited states were carried out at the TD-B3LYP/6-311+G(d,p) level of theory. The simulated UV–Vis spectra were obtained by convoluting the calculated vertical transitions with Gaussian functions with a half-width at half maximum of 0.15 eV.

4. Conclusions

The present study offers a detailed structural investigation of the most successful PET imaging agents for evaluating tumor hypoxia based on 2nim. To our knowledge, this is one of the very few works available in the current literature that provide a thorough characterization of the electronic structures and spectroscopic properties of radiopharmaceuticals.
Conformational analysis, vibrational and NMR spectra were obtained by means of quantum chemical DFT calculations. We have determined the structures of all unique conformers in the gas phase and water. Apart from the EF5 compound, important changes are observed in the conformers’ stabilities when going from the gas phase to the liquid state for all the other compounds. The available experimental data were accurately reproduced by the computational results obtained in this work. Anharmonic calculations proved to be more consistent in predicting the experimental IR spectrum for 2nim. Both calculated chemical shifts and spin-spin couplings were in good agreement with the available experimental data.
The photophysical properties of the compounds were predicted by means of TD-DFT calculations. UV–Vis spectra were obtained for all investigated compounds; however no such experimental data were available for validation. The migration of the electronic density during transitions was explained based on the electronic density differences between the ground and the excited states.
Our findings contribute to a deeper understanding of the spectroscopic properties of the five investigated radiopharmaceuticals. Moreover, the obtained results in this study can prove useful in further understanding the receptor binding mechanism of current and prospective 2-nitroimidazole based radiopharmaceuticals. Computational data indicate that the electronic excitation energies of all examined compounds fall outside the NIR spectral range, rendering them unsuitable as optical imaging probes.
In the development of multimodal agents based on 2-nitroimidazole, particularly for applications in nuclear medicine and optical imaging/photodynamic therapy, it would be imperative to incorporate a functional group that is compatible with both radiofluorination and the requisite absorption/emission characteristics. This dual functionality is essential to ensure the efficacy and versatility of these agents in their respective diagnostic and therapeutic applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29071505/s1, Figure S1: optimized structures of the conformers for FMISO, FETNIM, FAZA, FETA and EF5; Figures S2–S6: Harmonic IR and Raman spectra for the most stable conformers; Figure S7–S11: Harmonic and anharmonic IR and Raman spectra for the most stable conformers, Figures S12–S16 Harmonic IR and Raman spectra for the 18F variant of the derivatives; Figure S17: Fit of the calculated vs. experimental values for the chemical shifts of the 19F nuclei; Figure S18: NTOs corresponding to the first to excited singlet states of 2nim; Tables S1–S5: Assignment of vibrational modes for the 5 derivatives; Table S6: Calculated NMR data for FETNIM; Tables S7–S12: Cartesian coordinates for the optimized structures of 2nim, FMISO, FETNIM, FETA, FAZA and EF5.

Author Contributions

Conceptualization, V.C. and G.C.; methodology, V.C.; formal analysis, V.C. Ș.S. and G.C.; investigation, V.C., Ș.S. and G.C.; writing—original draft preparation, G.C.; writing—review and editing, V.C., Ș.S. and G.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Emami Nejad, A.; Najafgholian, S.; Rostami, A.; Sistani, A.; Shojaeifar, S.; Esparvarinha, M.; Nedaeinia, R.; Haghjooy Javanmard, S.; Taherian, M.; Ahmadlou, M.; et al. The Role of Hypoxia in the Tumor Microenvironment and Development of Cancer Stem Cell: A Novel Approach to Developing Treatment. Cancer Cell. Int. 2021, 21, 62. [Google Scholar] [CrossRef]
  2. Sørensen, B.S.; Horsman, M.R. Tumor Hypoxia: Impact on Radiation Therapy and Molecular Pathways. Front. Oncol. 2020, 10, 562. [Google Scholar] [CrossRef]
  3. Hompland, T.; Fjeldbo, C.S.; Lyng, H. Tumor Hypoxia as a Barrier in Cancer Therapy: Why Levels Matter. Cancers 2021, 13, 499. [Google Scholar] [CrossRef] [PubMed]
  4. Mittal, S.; Mallia, M.B. Molecular Imaging of Tumor Hypoxia: Evolution of Nitroimidazole Radiopharmaceuticals and Insights for Future Development. Bioorg. Chem. 2023, 139, 106687. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, Y.; Fan, J.; Li, Y.; Fu, S.; Chen, Y.; Wu, J. Imaging of Tumor Hypoxia With Radionuclide-Labeled Tracers for PET. Front. Oncol. 2021, 11, 731503. [Google Scholar] [CrossRef]
  6. Crișan, G.; Moldovean-Cioroianu, N.S.; Timaru, D.-G.; Andrieș, G.; Căinap, C.; Chiș, V. Radiopharmaceuticals for PET and SPECT Imaging: A Literature Review over the Last Decade. Int. J. Mol. Sci. 2022, 23, 5023. [Google Scholar] [CrossRef]
  7. Rashed, F.B.; Diaz-Dussan, D.; Mashayekhi, F.; Macdonald, D.; Nation, P.N.; Yang, X.-H.; Sokhi, S.; Stoica, A.C.; El-Saidi, H.; Ricardo, C.; et al. Cellular mechanism of action of 2-nitroimidazoles as hypoxia-selective therapeutic agents. Redox Biol. 2022, 52, 102300. [Google Scholar] [CrossRef] [PubMed]
  8. Hsieh, C.-J.; Giannakoulias, S.; Petersson, E.J.; Mach, R.H. Computational Chemistry for the Identification of Lead Compounds for Radiotracer Development. Pharmaceuticals 2023, 16, 317. [Google Scholar] [CrossRef]
  9. Moldovean, S.N.; Timaru, D.-G.; Chiş, V. All-Atom Molecular Dynamics Investigations on the Interactions between D2 Subunit Dopamine Receptors and Three 11C-Labeled Radiopharmaceutical Ligands. Int. J. Mol. Sci. 2022, 23, 2005. [Google Scholar] [CrossRef]
  10. Feng, J.; Zhou, P.; Qin, C.; Chen, R.; Chen, Q.; Li, L.; Chen, J.; Cheng, H.; Huang, W.; Cao, J. Magnetic solid-phase extraction-based surface-enhanced Raman spectroscopy for label-free therapeutic drug monitoring of carbamazepine and clozapine in human serum. Spectrochim. Acta A 2024, 310, 123924. [Google Scholar] [CrossRef]
  11. Jurina, T.; Cvetnić, T.S.; Šalić, A.; Benković, M.; Valinger, D.; Kljusurić, J.G.; Zelić, B.; Jurinjak Tušek, A. Application of Spectroscopy Techniques for Monitoring (Bio)Catalytic Processes in Continuously Operated Microreactor Systems. Catalysts 2023, 13, 690. [Google Scholar] [CrossRef]
  12. Willoughby, P.H.; Jansma, M.J.; Hoye, T.R. A Guide to Small-Molecule Structure Assignment through Computation of (1H and 13C) NMR Chemical Shifts. Nat. Protoc. 2014, 9, 643–660. [Google Scholar] [CrossRef] [PubMed]
  13. Lopes Jesus, A.J.; Jarmelo, S.; Fausto, R.; Reva, I. Conformational Preferences of 3,4-Dihydroxyphenylacetic Acid (DOPAC). Spectrochim. Acta A 2015, 140, 54–64. [Google Scholar] [CrossRef] [PubMed]
  14. Ducasse, L.; Castet, F.; Fritsch, A.; Huc, I.; Buffeteau, T. Density Functional Theory Calculations and Vibrational Circular Dichroism of Aromatic Foldamers. J. Phys. Chem. A 2007, 111, 5092–5098. [Google Scholar] [CrossRef] [PubMed]
  15. AIST:Spectral Database for Organic Compounds, SDBS. Available online: https://sdbs.db.aist.go.jp/sdbs/cgi-bin/direct_frame_disp.cgi?sdbsno=13553 (accessed on 8 September 2023).
  16. SpectraBase. Available online: https://spectrabase.com/spectrum/KWTAHQsUXFH (accessed on 8 September 2023).
  17. Lim, J.-L.; Berridge, M.S. An Efficient Radiosynthesis of [18F]Fluoromisonidazole. Appl. Radiat. Isot. 1993, 44, 1085–1091. [Google Scholar] [CrossRef] [PubMed]
  18. Yang, D.J.; Wallace, S.; Cherif, A.; Li, C.; Gretzer, M.B.; Kim, E.E.; Podoloff, D.A. Development of F-18-Labeled Fluoroerythronitroimidazole as a PET Agent for Imaging Tumor Hypoxia. Radiology 1995, 194, 795–800. [Google Scholar] [CrossRef] [PubMed]
  19. Castillo, A.E.; Pérez-Gutiérrez, E.; Ceballos, P.; Venkatesan, P.; Thamotharan, S.; Siegler, M.A.; Percino, M.J. Non-Covalent Interactions towards 2-(4-(2,2-Dicyanovinyl) Benzylidene)malononitrile Packing Polymorphism Due to Solvent Effect. Experimental and Theoretical Spectroscopy Approach. J. Mol. Struct. 2023, 1275, 134674. [Google Scholar] [CrossRef]
  20. Isley, W.C.; Urick, A.K.; Pomerantz, W.C.K.; Cramer, C.J. Prediction of 19F NMR Chemical Shifts in Labeled Proteins: Computational Protocol and Case Study. Mol. Pharm. 2016, 13, 2376–2386. [Google Scholar] [CrossRef] [PubMed]
  21. Saunders, C.; Khaled, M.B.; Weaver, J.D.; Tantillo, D.J. Prediction of 19F NMR Chemical Shifts for Fluorinated Aromatic Compounds. J. Org. Chem. 2018, 83, 3220–3225. [Google Scholar] [CrossRef]
  22. Benassi, E. An Inexpensive Density Functional Theory-based Protocol to Predict Accurate 19F-NMR Chemical Shifts. J. Comput. Chem. 2022, 43, 170–183. [Google Scholar] [CrossRef]
  23. Grierson, J.R.; Link, J.M.; Mathis, C.A.; Rasey, J.S.; Krohn, K.A. A Radiosynthesis of Fluorine-18 Fluoromisonidazole. J. Nucl. Med. 1989, 30, 343–350. [Google Scholar] [PubMed]
  24. Zanato, C.; Testa, A.; Zanda, M. Improved Synthesis of the Hypoxia Probe 5-Deutero-5-Fluoro-5-Deoxy-Azomycin Arabinoside (FAZA) as a Model Process for Tritium Radiolabeling. J. Fluor. Chem. 2013, 155, 110–117. [Google Scholar] [CrossRef]
  25. Kumar, P.; Stypinski, D.; Xia, H.; McEwan, A.J.B.; Machulla, H.-J.; Wiebe, L.I. Fluoroazomycin Arabinoside (FAZA): Synthesis,2H and3H-Labelling and Preliminary Biological Evaluation of a Novel 2-Nitroimidazole Marker of Tissue Hypoxia. J. Label. Compd. Radiopharm. 1999, 42, 3–16. [Google Scholar] [CrossRef]
  26. Tewson, T.J. Synthesis of [18F]Fluoroetanidazole: A Potential New Tracer for Imaging Hypoxia. Nucl. Med. Biol. 1997, 24, 755–760. [Google Scholar] [CrossRef] [PubMed]
  27. Dolbier, W.R.; Li, A.-R.; Koch, C.J.; Shiue, C.-Y.; Kachur, A.V. [18F]-EF5, a Marker for PET Detection of Hypoxia: Synthesis of Precursor and a New Fluorination Procedure. Appl. Radiat. Isot. 2001, 54, 73–80. [Google Scholar] [CrossRef] [PubMed]
  28. Orsi, A.; Price, D.J.; Kahr, J.; Pillai, R.S.; Sneddon, S.; Cao, S.; Benoit, V.; Łozińska, M.M.; Cordes, D.B.; Slawin, A.M.Z.; et al. Porous Zinc and Cobalt 2-Nitroimidazolate Frameworks with Six-Membered Ring Windows and a Layered Cobalt 2-Nitroimidazolate Polymorph. CrystEngComm 2017, 19, 1377–1388. [Google Scholar] [CrossRef]
  29. Martin, R.L. Natural transition orbitals. J. Chem. Phys. 2003, 118, 4775–4777. [Google Scholar] [CrossRef]
  30. Kimber, P.; Plasser, F. Energy Component Analysis for Electronically Excited States of Molecules: Why the Lowest Excited State Is Not Always the HOMO/LUMO Transition. J. Chem. Theory Comput. 2023, 19, 2340–2352. [Google Scholar] [CrossRef] [PubMed]
  31. Yu, Z.; Bernstein, E.R. Experimental and Theoretical Studies of the Decomposition of New Imidazole Based Energetic Materials: Model Systems. J. Chem. Phys. 2012, 137, 114303. [Google Scholar] [CrossRef]
  32. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6169. [Google Scholar] [CrossRef]
  33. Chai, J.-D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. [Google Scholar] [CrossRef] [PubMed]
  34. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  35. Becke, A.D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  36. 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. [Google Scholar] [CrossRef] [PubMed]
  37. Vosko, S.H.; Wilk, L.; Nusair, M. Accurate Spin-Dependent Electron Liquid Correlation Energies for Local Spin Density Calculations: A Critical Analysis. Can. J. Phys. 1980, 58, 1200–1211. [Google Scholar] [CrossRef]
  38. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  39. Frisch, M.J.; Pople, J.A.; Binkley, J.S. Self-Consistent Molecular Orbital Methods 25. Supplementary Functions for Gaussian Basis Sets. J. Chem. Phys. 1984, 80, 3265–3269. [Google Scholar] [CrossRef]
  40. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
  41. Scalmani, G.; Frisch, M.J. Continuous Surface Charge Polarizable Continuum Models of Solvation. I. General Formalism. J. Chem. Phys. 2010, 132, 114110. [Google Scholar] [CrossRef]
  42. NIST. Computational Chemistry Comparison and Benchmark Database; NIST Standard Reference Database Number 101 Release 22; Johnson, R.D., III, Ed.; National Institute of Standards and Technology: Gaithersburg, MD, USA, May 2022. Available online: http://cccbdb.nist.gov/ (accessed on 2 September 2023).
  43. Barone, V. Anharmonic Vibrational Properties by a Fully Automated Second-Order Perturbative Approach. J. Chem. Phys. 2005, 122, 014108. [Google Scholar] [CrossRef]
  44. Bloino, J.; Baiardi, A.; Biczysko, M. Aiming at an Accurate Prediction of Vibrational and Electronic Spectra for Medium-to-large Molecules: An Overview. Int. J. Quantum. Chem. 2016, 116, 1543–1574. [Google Scholar] [CrossRef]
  45. Crișan, G.; Macea, A.-M.; Andrieș, G.; Chiș, V. Experimental and Computational Raman Spectroscopies Applied to 2-Methoxy-2-Methylpropylisonitrile (MIBI) Ligand of the 99mTc-Sestamibi Radiopharmaceutical. J. Mol. Struct. 2021, 1246, 131159. [Google Scholar] [CrossRef]
  46. Ditchfield, R. Self-Consistent Perturbation Theory of Diamagnetism: I. A Gauge-Invariant LCAO Method for N.M.R. Chemical Shifts. Mol. Phys. 1974, 27, 789–807. [Google Scholar] [CrossRef]
  47. Wolinski, K.; Hinton, J.F.; Pulay, P. Efficient Implementation of the Gauge-Independent Atomic Orbital Method for NMR Chemical Shift Calculations. J. Am. Chem. Soc. 1990, 112, 8251–8260. [Google Scholar] [CrossRef]
  48. Casida, M.E.; Jamorski, C.; Casida, K.C.; Salahub, D.R. Molecular Excitation Energies to High-Lying Bound States from Time-Dependent Density-Functional Response Theory: Characterization and Correction of the Time-Dependent Local Density Approximation Ionization Threshold. J. Chem. Phys. 1998, 108, 4439–4449. [Google Scholar] [CrossRef]
Figure 1. B3LYP/6-311+G(d,p) optimized structures of the investigated molecules, with atom numbering schemes (a) 2nim (b) FMISO; (c) FAZA; (d) FETA; (e) FETNIM; (f) EF5. Atom color scheme: gray—C, light gray—H, light purple—N, red—O and light green—F.
Figure 1. B3LYP/6-311+G(d,p) optimized structures of the investigated molecules, with atom numbering schemes (a) 2nim (b) FMISO; (c) FAZA; (d) FETA; (e) FETNIM; (f) EF5. Atom color scheme: gray—C, light gray—H, light purple—N, red—O and light green—F.
Molecules 29 01505 g001
Figure 2. Theoretical IR spectrum in the gas phase (a) and Raman spectrum in water (b) for 2nim in the harmonic (red) and anharmonic (black) approximation, at the B3LYP/6-311+G(d,p) level of theory.
Figure 2. Theoretical IR spectrum in the gas phase (a) and Raman spectrum in water (b) for 2nim in the harmonic (red) and anharmonic (black) approximation, at the B3LYP/6-311+G(d,p) level of theory.
Molecules 29 01505 g002
Figure 3. The harmonic (scaled) IR (gas-phase) (a) and Raman (water) (b) spectra for 2nim and the five derivatives. The spectra for the derivatives are the Boltzmann-weighted averages of the most stable conformers.
Figure 3. The harmonic (scaled) IR (gas-phase) (a) and Raman (water) (b) spectra for 2nim and the five derivatives. The spectra for the derivatives are the Boltzmann-weighted averages of the most stable conformers.
Molecules 29 01505 g003
Figure 4. Calculated UV–Vis spectra for 2nim and the five derivatives at the PCM(H2O)-B3LYP/6-311+G(d,p) level of theory.
Figure 4. Calculated UV–Vis spectra for 2nim and the five derivatives at the PCM(H2O)-B3LYP/6-311+G(d,p) level of theory.
Molecules 29 01505 g004
Figure 5. The total SCF electronic density isosurfaces (0.02 a.u.) mapped with electronic density differences between the S2 singlet excited state and the ground state for 2nim (a) and between the S1 singlet excited state and the ground state for FMISO (b), FETNIM (c), FAZA (d), FETA (e) and EF5 (f), calculated at the B3LYP/6-311+G(d,p) level of theory in water. The blue (red) areas denote an increase (depletion) in electronic density in the excited state.
Figure 5. The total SCF electronic density isosurfaces (0.02 a.u.) mapped with electronic density differences between the S2 singlet excited state and the ground state for 2nim (a) and between the S1 singlet excited state and the ground state for FMISO (b), FETNIM (c), FAZA (d), FETA (e) and EF5 (f), calculated at the B3LYP/6-311+G(d,p) level of theory in water. The blue (red) areas denote an increase (depletion) in electronic density in the excited state.
Molecules 29 01505 g005
Table 1. Dihedral angles used to generate the relaxed PESs for the five investigated molecules.
Table 1. Dihedral angles used to generate the relaxed PESs for the five investigated molecules.
CompoundDihedral Angle (°)
FMISOC4-N3-C9-C10
N3-C9-C10-C11
FETNIMC4-N3-C9-C10
N3-C9-C10-C11
FAZAC4-N3-C9-C10
FETAC4-N3-C9-C10
N3-C9-C10-N11
EF5C4-N3-C9-C10
N3-C9-C10-N11
Table 2. Calculated relative free energies and Boltzmann populations of conformers in the gas phase and water, at the B3LYP/6-311+G(d,p) level of theory.
Table 2. Calculated relative free energies and Boltzmann populations of conformers in the gas phase and water, at the B3LYP/6-311+G(d,p) level of theory.
CompoundGas Phase ConformerΔG (kcal∙mol−1)Relative Boltzmann Population (%)Water
Conformer
ΔG (kcal∙mol−1)Relative Boltzmann Population (%)
FMISO1g0.00062.151w0.00065.91
2g0.63021.452w0.69020.56
3g1.03010.923w1.2807.59
4g1.8002.974w1.7203.61
5g1.9002.515w1.9802.33
FETNIM1g0.00071.361w0.00068.26
2g1.03012.532w0.64023.16
3g1.2808.223w1.3706.75
4g1.6604.334w2.7800.62
5g1.9902.485w2.8100.59
6g2.4801.086w2.8200.58
7g5.3900.017w4.8500.02
FAZA1g0.00099.251w0.00069.66
2g3.1900.452w0.53028.46
3g3.4500.293w2.1401.88
FETA1g0.00050.841w0.00063.97
2g0.02049.162w0.34036.03
EF51g0.00054.211w0.00055.05
2g0.10045.792w0.12044.95
Table 3. B3LYP/6-311+G(d,p) calculated wavenumbers of 2nim in the gas phase alongside the corresponding wavenumbers of the same modes in the five derivative compounds.
Table 3. B3LYP/6-311+G(d,p) calculated wavenumbers of 2nim in the gas phase alongside the corresponding wavenumbers of the same modes in the five derivative compounds.
ModeTheoretical Wavenumbers in Gas Phase (cm−1)Assignments c
2nimFMISO
1g
FETNIM
1g
FETA
1g
FAZA
1g
EF5
1g
HarmonicAnharmonicExperimental a
Q1201199n.a. b216187174200195oop. def. (2nim)
Q2242240n.a.244298308285306ρ(2nim)
Q3548539n.a.566562540600546δ(CNO)
Q4600572623638635625631628γ(imidazole)
Q5660645649660660660661660γ(imidazole)
Q6777772798785785789798788γ(CH)imidazole (ip.)
Q7840830820855854857866857δ(NO2) + δ(CN2)imidazole
Q8890872n.a.888886893893894γ(CH)
Q9929916944932926924934922δ(CNC) + δ(NCC)
Q10997984985-----δ(NCN) + δ ν(C-NO2) + (CH)
Q1110561067n.a.-----β(CH) + β(NH)
Q1210831102110310681061105910471167δ(CH)
Q1311501158115511561154115111561153δ(CN2) + β(CH)
Q1412171223126912701256126012241260β(CH) + νsym(NO2) + ip. def.(imidazole)
Q1513301342134813251323132313241318ν(C-N)imidazole + δ(NO2) + β(CH, NH)
Q1613331349137613311332132913371331ν(C-NO2) + β(CH) + δ(NO2)
Q1713861393142513841380137713801377ν(C-N)imidazole + β(NH) + β(CH)
Q1814561467149414641463145914451459ν(C-N)imidazole + β(NH) + β(CH)
Q1914941513152014761474147314671474ν(C=N)imidazole + ν(C=C) + β(NH) + β(CH)
Q2015461559155415351533153615301538νasym(NO2)
Q2131413123314631393138314031403140νasym(CH)
Q2231653147316431663164316031893161νsym(CH)
Q23351234693423-----ν(NH)
a from references [15,16]; b not available. c oop.—out-of-plane; ip.—in plane; def.—deformation; ρ—rocking; ν—stretching; β—in plane bending XYH angles; γ—out-of-plane bending; δ—in-plane bending; sym.—symmetric; asym—antisymmetric.
Table 9. Theoretical UV–Vis absorption spectral data for 2nim in water at the PCM(H2O)-TD-B3LYP/6-311+G(d,p) level of theory.
Table 9. Theoretical UV–Vis absorption spectral data for 2nim in water at the PCM(H2O)-TD-B3LYP/6-311+G(d,p) level of theory.
CompoundExcited State * λ a b s ( n m ) fTransitionsContributions
(%)
S23150.3287H→L99.12
S52450.0239H-2→L90.08
S62040.1247H-4→L87.57
* Only singlet-singlet transitions with λabs ≥ 200 nm and f ≥ 0.01; H—HOMO, L—LUMO.
Table 10. Theoretical UV–Vis absorption spectral data for FMISO, FETNIM, FAZA, FETA and EF5 in water at the PCM(H2O)-B3LYP/6-311+G(d,p) level of theory.
Table 10. Theoretical UV–Vis absorption spectral data for FMISO, FETNIM, FAZA, FETA and EF5 in water at the PCM(H2O)-B3LYP/6-311+G(d,p) level of theory.
CompoundExcited State * λ a b s ( n m ) fTransitionsContributions (%)
FMISO 1wS13310.0221H-4 → L17.34
H-2 → L59.61
H-1 → L10.33
S23150.2708H → L88.58
S52700.0126H-4 → L61.04
H-1 → L22.37
S62600.0332H-3 → L80.51
S82040.1194H-5 → L78.21
FETNIM 1wS13320.0399H-5 → L14.87
H-3 → L59.22
H → L18.96
S23180.2172H-3 → L17.28
H → L76.24
S32860.0784H-1 → L92.78
S92050.1166H-6 → L82.94
FAZA 1wS13280.0536H-4 → L58.31
H → L22.26
S23170.2195H-4 → L21.16
H → L73.38
S32970.0248H-4 → L22.66
H-2 → L68.72
S42860.0238H-4 → L54.06
H-3 → L15.42
H-2 → L20.91
S102040.1169H-9 → L23.39
H-7 → L64.87
FETA 1wS23230.0541H-1 → L59.96
H → L30.57
S33140.0508H-2 → L84.74
H → L13.82
S43120.1961H → L54.27
H-1 → L37.03
S72510.0312H-4 → L92.54
S102060.1221H-6 → L84.89
EF5 1wS13350.0525H-5 → L12.02
H-3 → L41.36
H → L27.89
S23220.1329H-3 → L23.75
H → L54.16
S33070.0874H-1 → L72.79
H → L12.56
S43020.0547H-2 → L75.39
S102050.1299H-7 → L11.53
H-6 → L77.38
* Only singlet-singlet transitions with λabs ≥ 200 nm and f ≥ 0.01; H—HOMO, L—LUMO.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Crișan, G.; Stan, Ș.; Chiș, V. Exploring Geometrical, Electronic and Spectroscopic Properties of 2-Nitroimidazole-Based Radiopharmaceuticals via Computational Chemistry Methods. Molecules 2024, 29, 1505. https://doi.org/10.3390/molecules29071505

AMA Style

Crișan G, Stan Ș, Chiș V. Exploring Geometrical, Electronic and Spectroscopic Properties of 2-Nitroimidazole-Based Radiopharmaceuticals via Computational Chemistry Methods. Molecules. 2024; 29(7):1505. https://doi.org/10.3390/molecules29071505

Chicago/Turabian Style

Crișan, George, Ștefan Stan, and Vasile Chiș. 2024. "Exploring Geometrical, Electronic and Spectroscopic Properties of 2-Nitroimidazole-Based Radiopharmaceuticals via Computational Chemistry Methods" Molecules 29, no. 7: 1505. https://doi.org/10.3390/molecules29071505

Article Metrics

Back to TopTop