Skip to main content
Erschienen in: Forensic Toxicology 2/2019

Open Access 01.03.2019 | Original Article

Methylone and pentylone: structural analysis of new psychoactive substances

verfasst von: Dita Spálovská, Tereza Maříková, Michal Kohout, František Králík, Martin Kuchař, Vladimír Setnička

Erschienen in: Forensic Toxicology | Ausgabe 2/2019

Abstract

Purpose

The growing availability of new psychoactive substances with unknown toxicity is alarming all over the world. The simplicity of acquiring these drugs of abuse from internet markets has caused an increase in the total number of seizures. We present the first systematic spectroscopic study of methylone and pentylone, which are in the class of synthetic cathinones.

Methods

High performance liquid chromatography (HPLC) was used for the enantioseparation of methylone and pentylone. The substances were further analysed by the conventional methods of ultraviolet (UV) and infrared (IR) absorption and chiroptical methods, specifically electronic circular dichroism (ECD) and vibrational circular dichroism (VCD). The obtained data were supported with the density functional theory (DFT) calculations using the B3LYP or B3PW91 functional and 6-311++G(d,p) basis sets, including the solvent effects.

Results

The 3D structure of methylone and pentylone in solution was revealed. Moreover, the chiral separation of both studied substances was achieved and the absolute configuration of the respective enantiomers was determined.

Conclusion

Vibrational circular dichroism spectroscopy seems to be the best of the methods employed to distinguish structurally similar chiral substances, especially in combination with quantum chemical calculations. According to the structural analysis, 5 and 9 stable conformers of methylone and pentylone, respectively, were revealed in an aqueous solution. Finally, very good agreement between the experimental and simulated spectra was achieved.
Hinweise

Electronic supplementary material

The online version of this article (https://​doi.​org/​10.​1007/​s11419-019-00468-z) contains supplementary material, which is available to authorized users.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Introduction

New psychoactive substances (NPS), compounds structurally similar to already legislatively controlled illicit drugs recently appearing on the drug scene with increasing frequency, pose an increasing global problem. The EU Early Warning System for NPS currently monitors over 670 substances, with 51 reported in 2017 for the first time and 101 in 2014 [1, 2]. In addition, the total number of NPS all over the world monitored by the United Nations Office on Drugs and Crime is 803 individual NPS [3]. In 2016, about 71,000 seizures of NPS were reported through the EU Early Warning and almost one-third of these seizures are of synthetic cathinones [1].
New psychoactive substances are often considered by users as a safe alternative to illicit narcotics and psychotropic substances, despite the fact that they possess similar structures and psychoactive effects. Structural modification of the parent NPS may often lead to the enhancement of hallucinogenic or psychostimulative effects, e.g. in case of synthetic cannabinoids (THC analogues) and cathinones (cathinone analogues) [4, 5]. Synthetic cathinones (often sold as “legal highs”, “bath salts” or “research chemicals”) [6] are the second most frequently intercepted NPS. Their effects include empathy, openness, increased energy and libido of consumers, but also dangerous neurological, cardiac and psychiatric signs and symptoms [1, 7]. Methylone and pentylone (Fig. 1) studied here are homologues in the class of synthetic cathinones. Unlike natural cathinone, methylone and pentylone both possess the methylenedioxy functional group on the phenyl ring, similarly to the well-known psychoactive substance 3,4-methylenedioxyamphetamine (MDMA), better known as ecstasy. Due to this structural similarity, methylone and pentylone combine the effects of cathinone and MDMA [8, 9]. Their mechanism of action is based primarily on inhibiting the uptake of norepinephrine, dopamine and serotonin by monoamine neurotransmitter transporters [10]. However, the full metabolism in the human body is not yet completely understood.
Methylone and pentylone differ in the length of the α-alkyl chain; methylone contains a methyl group and pentylone a propyl group. According to pharmacokinetic studies on rats [9, 11], a longer α-alkyl chain causes higher lipophilicity, hence, pentylone exhibits a higher plasma concentration than methylone after dosing the same amount of the pure substance. However, safety data concerning their toxicity on humans as well as information about the side effects of long-term usage are unavailable [3].
The current trends in the control of NPS demand a fast and reliable analysis. Moreover, methods for a detailed description of their structure are required. However, most studies [1214] are focused on the detection of NPS in biological samples (blood, plasma, urine, hair) performed by high-performance liquid chromatography (HPLC) or gas chromatography (GC) with mass spectrometry (MS). For studies on identification of real NPS samples, vibrational spectroscopy techniques such as infrared (IR) or Raman spectroscopy are often used. The advantage of this approach is a fast and reliable analysis, which can even be performed by portable spectrometers [15, 16].
Molecular spectroscopy is one of the effective tools for the identification of the NPS. Since synthetic cathinones are chiral, chiroptical methods [17] can be used for their detailed study. The chiroptical methods, specifically vibrational circular dichroism (VCD), electronic circular dichroism (ECD) and Raman optical activity (ROA), in combination with theoretical quantum chemical predictions can provide further 3D structural information on NPS in solution, which is crucial, e.g. for the understanding of their biological activity and toxicity [1719]. The comparative analysis of experimental chiroptical properties and DFT calculations also allow the determination of the absolute configuration of individual separated enantiomers [20, 21]. Analyses performed by different independent chiroptical spectroscopic methods provide more reliable structural characterizations of the studied molecules [2224].
The precise structural characterization of enantiomers, enabled by the combination of chiroptical spectroscopic methods and DFT calculations, may be helpful for understanding differences in their toxicity. These methods can also be used to determine the presence of a particular drug enantiomer in a mixture of different compounds—a typical feature of synthetic cathinone products available on the Internet [7]. There are many examples that enantiomers of licit as well as illicit drugs have different properties and toxicity. Apart from the widely known Contergan affair [25], it has already been shown in literature that enantiomers of some widely used drugs of abuse, e.g. ketamine and deschloroketamine possess significantly different properties in terms of pharmacological effect and toxicity [26, 27]. With the increasing number of synthetic cathinones, the need for detailed information on the structure of individual enantiomers and structure-properties relationships is essential. This is due to the fact that naturally occurring active substance is (S)-cathinone, and it has already been documented that (S)-enantiomers of synthetic cathinones are in many cases more potent/toxic than (R)-enantiomers [2830]. It is therefore evident that not only simple identification of a substance by analytical methods, but also more in depth structural information is required, and in this case, DFT calculations represent an important tool.
In the continuation of our systematic study of NPS [27, 31], we introduce a comparison of two NPS homologues—methylone and pentylone—in an aqueous solution. Employing the methods of VCD, ECD, IR spectroscopy and ultraviolet (UV) absorption spectroscopy supported by the DFT calculations at B3LYP/6-311++G(d,p) or B3PW91/6 311++G(d,p) levels, including solvent effects, we elucidated in detail the 3D structure of both drugs and compared their structural features. Moreover, we performed chiral separation of the studied substances and determined the absolute configuration of the respective enantiomers.

Materials and methods

Chemicals, reagents

Racemic standards of methylone hydrochloride and pentylone hydrochloride with purity exceeding 98% were obtained from Alfarma (Černošice, Czech Republic). Deuterium oxide (D2O; 99.9% D) for the dissolution of methylone and pentylone was purchased from ISOSAR GmbH, Germany. HPLC grade solvents (hexane, heptane, propan-2-ol IPA) were obtained from Labicom (Olomouc, Czech Republic) and ethanol was purchased from Merck (Darmstadt, Germany). Diethylamine, used as a basic additive, was obtained from Sigma-Aldrich (St. Louis, MO, USA).

Enantioseparation

Both the analytical and the preparative chiral separation of the drugs were performed using the Waters AutoPurification System (Milford, MA, USA) equipped with a UV–Vis detector (detection wavelength of 254 nm). The analytical separation was performed on a chiral polysaccharide column ChiralArt Amylose-SA (250 × 4.6 mm i.d., 5 μm, YMC Europe GmbH, Dinslaken, Germany) with an analyte concentration of 1 mg ml−1 and a flow rate of 1 ml min−1 at ambient temperature. For the preparative chromatography, an analogous polysaccharide column available in our laboratory—Chiralpak IA (250 × 20 mm i.d., 5 μm, Chiral Technologies Europe, Illkirch, France)—was used. For methylone, the sample concentration was 17.5 mg ml−1, while in case of pentylone, for which a different mobile phase was required (vide infra), the sample concentration was 4 mg ml−1. In both cases, at flow rate of 15 ml min−1 and ambient temperature conditions were employed. For analytical chromatography screening, commercially available drugs were dissolved in the mobile phase containing 0.1% of diethylamine. The sample for preparative chromatography was dissolved in a mixture of 1% diethylamine in propan-2-ol (0.45 ml) and a total volume of 1 ml was adjusted by the addition of heptane. In both cases, the addition of diethylamine ensured the liberation of a free base of the drug from the corresponding hydrochloric acid salt.

Electronic circular dichroism

The experimental ECD spectra were acquired using a J-815 spectrometer (JASCO Corporation, Tokyo, Japan) purged with nitrogen gas (purity 99.99%, Siad, Prague, Czech Republic) during the measurement. To measure the ECD spectra, solutions of individual enantiomers of a concentration of 15 mg l−1 in demineralized water (UCT Prague, Czech Republic) were prepared. These sample solutions (3 ml) were pipetted into a quartz cuvette (Hellma, Müllheim, Germany) with a pathlength of 1 cm and measured in a spectral range of 185–380 nm at ambient temperature with a 20 nm min−1 scanning speed, 8-s response time, 3 accumulations, and 1 nm band width. The baseline was corrected by subtracting the spectra of demineralized water obtained under identical experimental conditions. The corresponding UV absorption was calculated from the detector HT voltage using the Spectra Analysis module of the Spectra Manager software (JASCO Corporation, Tokyo, Japan).

Vibrational circular dichroism

The VCD and IR absorption spectra were acquired using the FTIR IFS 66/S spectrometer equipped with a VCD/IRRAS PMA 37 module (Bruker, Bremen, Germany), a ZnSe beam splitter (Hinds Instrument, Hillsboro, OR, USA), a BaF2 polarizer, and an MCT detector (InfraRed Associates, Stuart, FL, USA) [32]. The solutions of individual enantiomers of methylone hydrochloride and pentylone hydrochloride in D2O (100 mg ml−1) were placed individually into a BioCell cuvette with CaF2 windows (BioTools, Inc., Jupiter, FL, USA) and an optical pathlength of 27.3 μm. The spectra were measured at ambient temperature in a spectral range of 1750 to 1250 cm−1 with a resolution of 8 cm−1. The VCD spectra were averaged from 9 to 12 blocks, each measured for 20 min and containing 3680 interferometric records. The baseline of each enantiomer was corrected by the subtraction of the solvent (D2O) spectra measured under identical experimental conditions. The VCD noise spectra were offset from zero for clarity.

Quantum chemical calculations

The starting geometries of (R)-methylone and (R)-pentylone hydrochlorides were optimized using the DFT method at the B3LYP/6-31G(d) level. Reoptimization of the final lowest energy conformers, considering their Gibbs free energies, was performed by the DFT method at several higher levels of theory (B3LYP/6-311++G(d,p), B3PW91/aug-cc-pVDZ, B3PW91/6-311++G(d,p), CAM-B3LYP/6-311++G(d,p), CAM-B3LYP/aug-cc-pVDZ, wB97XD/6-311++G (d,p), wB97XD/aug-cc-pVDZ, and wB97XD/TZVP, Electronic Supplementary Material, Table S1 and S2) using the Gaussian 09 program package [33]. In this work, only the results providing the best level of agreement with the experimental spectra are presented in detail (B3LYP/6-311++G(d,p), B3PW91/6-311++G(d,p)). All the calculations were carried out on a supercomputer Altix UV 2000 (UCT Prague, Czech Republic). In general, 12 processors with total memory of 64 GB were used for individual computation tasks, which took 40–48 h.
The D2O solvation effect was considered via the conductor-like polarizable continuum model (CPCM). The DFT calculations of VCD spectra included the effect of deuteration of the amine hydrogens. The population weighted average VCD, ECD, IR and UV spectra were calculated using the Boltzmann distribution based on Gibbs free energy at a temperature of 298 K. The similarity overlaps of the experimental and calculated spectra were undertaken using the CDSpecTech program, based on the similarity of the dissymmetry factor to determine quantitatively the agreement between both spectra and to choose the optimal scaling factor [34, 35]. A scaling factor of 1.0 marks the maximum similarity.
For visualization of the simulated VCD and IR spectra, a Lorentz profile function with a 10 cm−1 bandwidth was assumed. For ECD and UV absorption, the 30 lowest singlet states of each conformer were calculated and Gaussian band profiles with a 12 nm bandwidth were assumed.

Results and discussion

Enantioseparation

Since methylone and pentylone are in the cathinone family, they can be enantiomerically resolved by several different methods including, for example, capillary electrophoresis, supercritical fluid chromatography, and HPLC [3638]. However, only some methods from this list can provide pure enantiomers in preparative amounts of the milligram to gram scale. The most prominent method for the separation of chiral substances nowadays is preparative HPLC. We have already employed this method in our previous study focused on the determination of the absolute configuration of butylone [31], thus HPLC was selected as the method of choice in this study.
Starting from the previously optimized conditions, we have found that although they are in the same family of NPS, methylone, butylone and pentylone exhibit significantly different chromatographic behaviours. Therefore, different methods of chiral separation for methylone and pentylone have been developed; more detailed information on the screening of the mobile phase conditions can be found in the Electronic Supplementary Material. Methylone, as the more polar substance, was resolved using a mobile phase composed of a mixture of hexane/EtOH (4/1, v/v) with diethylamine (0.1%) (Fig. 2), whereas the more lipophilic pentylone was successfully resolved using a mixture of heptane/IPA (95/5, v/v) with diethylamine (0.1%) (Fig. 3).
In both cases, automatic fraction collection was set to ensure high purity of the collected enantiomers. In order to assure the stability of the separated enantiomers and to avoid a cyclization reaction occurring for the cathinone free base [39], the separated substances were transformed to hydrochlorides by adding an etheric solution of hydrogen chloride. Both enantiomers (methylone 1 and methylone 2) were found to be of comparable optical purity with the enantiomeric excess of > 96.3%. This value was still acceptable for the subsequent spectroscopic measurements. For pentylone, even better values of > 99.4% and > 98.9% for the first and the second eluting enantiomers, respectively, were found.

Conformational analysis

The individual (R)-enantiomers of methylone and pentylone hydrochloride were selected for the conformational analysis. Energetically preferred geometries of these molecules were determined by three or five dihedral angles α1, α2, α3, and optionally α4 and α5 (Fig. 1) for methylone and pentylone, respectively. The dihedral angle α1 was systematically changed by 180° rotation about a single bond and the remaining angles α2, α3 and possibly α4 and α5 by 120° using the MCM software [40]. As a result, 18 and 162 starting geometries for methylone and pentylone, respectively, were obtained. Those geometries were subsequently optimized by the DFT method at the B3LYP/6-31G(d) level of theory. Since most starting structures were converging to the same geometry or their population was too low, conformational analysis revealed all the relevant stable conformers (relative energy < 2 kcal mol−1): 5 and 9 for methylone and pentylone, respectively. These stable conformers were reoptimized at a more refined level of theory: B3LYP/6-311++G(d,p) or B3PW91/6-311 ++G(d,p) to provide the final structures (Fig. 4) and their relative abundances based on the Boltzmann distribution (Table 1).
Table 1
Stable conformers of methylone and pentylone with their relative Gibbs free energies, relative abundances and dihedral angles calculated at the B3LYP/6-311++G(d,p) level
Conformer
ΔG (kJ mol−1)
Relative abundance (%)
Dihedral angle (°)
α 1
α 2
α 3
α 4
α 5
Methylone I
0
85
− 173
− 96
− 166
Methylone II
6.3
7
8
− 97
67
Methylone III
7.9
3
− 160
− 117
− 87
Methylone IV
8.2
3
15
− 115
− 88
Methylone V
9.5
2
− 172
− 98
66
Pentylone I
0
33
5
− 98
87
− 164
− 180
Pentylone II
0.6
26
− 171
− 82
75
− 70
− 177
Pentylone III
1.3
19
− 174
− 100
92
− 164
− 179
Pentylone IV
2.5
12
5
− 100
84
56
180
Pentylone V
5.5
4
− 173
− 100
83
57
− 180
Pentylone VI
6.6
2
5
− 97
− 169
57
− 180
Pentylone VII
6.8
2
− 173
− 99
− 167
56
179
Pentylone VIII
8.0
1
7
− 80
− 170
− 63
− 176
Pentylone IX
8.0
1
−173
− 99
90
− 169
66
All the conformers were stabilized by the hydrogen bonding between the oxygen atom of the carbonyl group and the hydrogen atom of the amino group. From all the conformers of methylone, this interaction most stabilized the structure of the conformer M I, probably resulting in its high relative abundance (85%). The distance of hydrogen and oxygen in the case of the conformer M I was ~ 2.1 Å, for other conformers of methylone this distance was 2.2–2.5 Å. As for the individual dihedral angles (Table 1), the angle α1 occupied predominantly two positions that differed by ~ 180°. The total number of stable conformers of pentylone was higher than methylone due to the elongation of the alkyl chain and the associated increase in the number of dihedral angles in the molecule. All 9 conformers of pentylone were also stabilized by the hydrogen bonding between the oxygen of the carbonyl group and the hydrogen of the amino group, with the distance of these atoms being in the range of 2.0–2.2 Å for all the stable conformers. Relatively small differences in the length of this stabilizing interaction may be the reason for the more uniform relative abundances of conformers (33, 26 and 19%) than in the case of methylone.

Electronic spectroscopy

Since the experimental ECD spectra of the first and the second eluting compounds (Fig. 5) were mirror images, the enantiomeric character of the separated products was confirmed. The comparison of the experimental and simulated spectra determined the absolute configuration of the individual enantiomers: methylone 1 and 2 were determined as (R)- and (S)-enantiomer, respectively. On the contrary, pentylone 1 and 2 were determined as (S)- and (R)-enantiomer, respectively. These results could be confirmed further using the similarity overlap plots (Electronic Supplementary Material, Figs. S3 and S4). For the correct assignment of the bands in the experimental and simulated spectra, the simulated spectra of methylone were scaled by a factor of 0.96 in order to reach the maximum value of the similarity index. After scaling, a similarity index of 0.85 was achieved in this case. For pentylone, a scaling factor of 0.97 was applied and the corresponding index of similarity reached 0.75. Both results were more than sufficient to reliably determine the absolute configuration of individual enantiomers, which was subsequently confirmed by a detailed VCD spectra analysis.
In the experimental ECD spectra of methylone and pentylone hydrochloride (Fig. 5), six very similar bands reflecting the combination of electron transitions π → π *, n → π * and n → σ * were observed. The bands at 330 nm and partially at 278 nm were mainly an expression of an electron transition from the highest occupied to the lowest unoccupied molecular orbitals (HOMO → LUMO, Fig. 6). Three positive and three negative bands in the spectra of (R)-enantiomers were also present in the simulated weighted average spectra of (R)-enantiomers. The spectra of individual conformers of the (R)-enantiomers differed considerably from one another (Fig. 5, top), indicating the exceptional sensitivity of ECD to the 3D structure of the molecules compared to non-polarized methods (UV, IR). However, the experimental ECD spectra did not allow reliable differentiation between the two studied compounds.
The simulated UV spectra of (R)-methylone and (R)-pentylone hydrochloride correctly predicted four significant bands of the experimental spectra (Fig. 7). The simulated spectra were scaled by a factor of 0.97 and 0.98 for methylone and pentylone, respectively, and the final similarity overlap reached 0.82 and 0.78 for methylone and pentylone, respectively (Electronic Supplementary Material, Figs. S3 and S4). However, this method does not seem to be suitable either for the differentiation of substances differing in the number of –CH2 groups (since the experimental spectra of both studied compounds did not significantly differ) nor for the studies of the 3D structural changes of the individual conformers. The shapes of the partial spectra (Fig. 7, top) were very similar, with slight differences up to 10 nm observed in the change of the position of the band maximum at 330 nm. Differences between the intensities of band at ~ 234 nm in the simulated spectra of (R)-pentylone were caused mainly by the change of dihedral angle α1: if its value was positive, the intensity of the band was lower. In contrast, the intensity of the band at 270 nm was higher for positive values of α1.

Vibrational spectroscopy

The shape of the experimental IR absorption spectra of methylone and pentylone in the spectral range 1750–1250 cm−1 was very similar with a total number of 11 and 9 bands for methylone and pentylone, respectively (Fig. 8). For the correct assignment of the individual bands and their respective vibrational modes (Tables 2, 3), the spectra were scaled by a factor of 0.99 and 0.98 and resulted in a very convincing index of spectra similarity: 0.85 and 0.91 for methylone and pentylone, respectively (Electronic Supplementary Material, Figs. S3 and S4). Looking at the individual spectra of the conformers (Fig. 6, top), it is worth noting the large influence of the dihedral angle α1 on the spectral shape. Band 1 was shifted to the lower wavenumbers in the spectra of all the conformers with a positive value of angle α1 (~ 5°) compared with the spectra of the other conformers. Band 2 was apparent only in the spectra of the conformers with the positive value of angle α1 (~ 5°), hence the intensity of this band was lower in the averaged spectra and appeared only as a shoulder. In contrast, bands 9 in the spectra of methylone and 8 in the spectra of pentylone were not present in the spectra of those conformers. The bands in this area reflected the bending vibration of the C–H group, including the umbrella bending vibration of –CH3. Band 7 in the spectra of methylone and band 6 in the spectra of pentylone were more intense in the case of the conformers with positive value of α1. Band 10 in the spectra of methylone reflected the bending vibration of the C–H group in CH–CH3 and was correctly predicted by the DFT calculations. This group was not present in the structure of pentylone, therefore we did not expect a band reflecting the bending vibration of the C–H group in CH–CH3 to occur in its IR spectra. This was confirmed by the experimental spectra as well. Furthermore, slight differences in the spectral shapes were observed between the experimental IR absorption spectra of methylone and pentylone hydrochloride (Fig. 8, bottom), particularly in the spectral range of 1420–1280 cm−1. Therefore, this method can be used to distinguish structurally similar substances, but, unfortunately, with low sensitivity.
Table 2
Individual IR vibrational modes in the experimental and calculated (B3LYP/6-311++G(d,p)) spectra of (R)-methylone hydrochloride
Band number
Experimental spectrum (cm−1)
Calculated spectrum (cm−1)
Vibrational mode
1
1675
1676
ν (C=O)
2
1621
1628
ν (C=C), Ar
3
1606
1608
δ (N–H), ν (C=C), Ar
4
1506
1514
δ (C–H), O–CH2–O
5
1494
1500
δ (C–H)
6
1472
1469
δ (C–H), –CH3
7
1452
1450
δ (C–H), δ (C=C), Ar
8
1397
1399
δ (C–H), –CH3
9
1356
1359
δ (C–H), δ (C=C), Ar
10
1303
1309
δ (C–H), –CH–CH3
11
1262
1261
δ (C=C), Ar, δ (C–H), –CH–
ν stretching vibration, δ bending vibration
Table 3
Individual IR vibrational modes in the experimental and calculated (B3LYP/6-311++G(d,p)) spectra of (R)-pentylone hydrochloride
Band number
Experimental spectrum (cm−1)
Calculated spectrum (cm−1)
Vibrational mode
1
1671
1669
ν (C=O)
2
1621
1623
ν (C=C), Ar
3
1607
1604
δ (N–H)
4
1506
1514
δ (C–H), O–CH2–O
5
1494
1494
δ (C–H)
6
1452
1450
δ (C–H), δ (C=C), Ar
7
1388
1387
δ (C–H), –CH3
8
1356
1356
δ (C–H), δ (C=C), Ar
9
1263
1265
δ (C=C), Ar
ν stretching vibration, δ bending vibration
The experimental VCD spectra (Fig. 9, bottom) were presented in the spectral range of 1750–1263 cm−1 because of the high level of VCD noise caused by overabsorption at lower wavenumbers. The simulated VCD spectra were scaled by a factor of 0.98 and 0.99 for methylone and pentylone, respectively, for the correct assignment of the individual bands (Tables 4, 5). The resulting index of spectra overlap reached convincing values of 0.86 and 0.70 for methylone and pentylone, respectively (Electronic Supplementary Material, Figs. S3 and S4). In the experimental VCD spectra of (R)-methylone and (R)-pentylone hydrochloride (Fig. 9, bottom), there were several areas in which individual substances could be reliably distinguished, mainly in a spectral range of 1400–1263 cm−1, where the molecules not only had a different number of bands but also different spectral shapes. Moreover, that was predicted correctly by the quantum chemical simulations (Fig. 9, middle). In addition, in the spectral range of 1700–1580 cm−1, there were different intensities of bands 1 and 2, which was also correctly predicted by the simulations. Our results demonstrating a higher propensity of distinguishing between two structurally similar compounds by VCD than ECD are in very good agreement with the literature [22]. The conformers of methylone and pentylone with a positive value of dihedral angle α1 again had a distinct spectral shape different from the rest of the conformers, e.g. band 3 in the VCD spectra of both (R)-methylone and (R)-pentylone as well as bands 5, 6 and 7.
Table 4
Experimental and calculated (B3LYP/6-311++G(d,p)) band frequencies and signs in the VCD spectra of (R)-methylone hydrochloride
Band number
Experimental spectrum (cm−1)
Calculated spectrum (cm−1)
Band sign
1
1675
1666
+
2
1605
1602
+
3
1465
1474
4
1447
1442
5
1394
1390
+
6
1374
1374
7
1355
1350
+
8
1304
1303
+
9
1260
1253
Table 5
Experimental and calculated (B3LYP/6-311++G(d,p)) band frequencies and signs in the VCD spectra of (R)-pentylone hydrochloride
Band number
Experimental spectrum (cm−1)
Calculated spectrum (cm−1)
Band sign
1
1670
1677
+
2
1605
1606
+
3
1493
1496
+
4
1458
1469
5
1396
1398
+
6
1378
1383
7
1364
1364
+
8
1327
1330
+
9
1310
1313
10
1282
1278
+
11
1257
1251

Conclusions

The enantioseparation of two cathinone homologues—methylone and pentylone—by HPLC with UV–Vis detection has been successfully carried out using different methods of chiral separation for individual substances. Furthermore, a detailed structural analysis of both drugs of abuse has been achieved by the methods of electronic and vibrational spectroscopy, specifically conventional UV and IR absorption with ECD and VCD in combination with DFT calculations at the B3LYP/6-311++G(d,p) or B3PW91/6-311++G(d,p) levels of theory, including solvent effects. According to this study, these homologues can be difficult to distinguish solely by ECD or UV absorption. VCD spectroscopy appeared to be the best of the employed methods for distinguishing structurally similar chiral substances, not only in a small spectral range such as IR absorption spectroscopy provides, but essentially across the spectrum, especially in combination with quantum chemical calculations. Structural analysis suggested the existence of 5 and 9 stable conformers of methylone and pentylone, respectively, in an aqueous solution. Finally, very good agreement between the experimental and the corresponding simulated spectra was achieved. Thus, we were able not only to determine the absolute configuration of the individual conformers of methylone and pentylone, but also to describe the molecular structure of both compounds. In conclusion, chiroptical spectroscopic methods confirmed the high potential for structural analysis of molecules significant in the forensic sciences. These data can be helpful in considering the forensic evidence in criminal proceedings. Moreover, the structural analysis performed by the quantum chemical calculations can be used for the understanding of the biological activity and toxicity of the NPS. A detailed knowledge of the 3D structure of the substances may allow not only the detailed study of the binding properties, metabolism, transport or distribution of these substances in the human organism, but also the development of antagonist drugs for overdoses, addictions or poisoning.

Acknowledgements

The work was supported by the Ministry of the Interior of the Czech Republic (MV0/VI20172020056). Partial support was provided by the Operational Program Prague—Competitiveness (CZ.2.16/3.1.00/24503 and CZ.2.16/3.1.00/21537), the National Program of Sustainability (NPU MSMT—LO1601; 43760/2015), and the Specific University Research (MSMT No. 21-SVV/2018 and 21-SVV/2019).

Compliance with ethical standards

Conflict of interest

The authors declare that they have no conflict of interest.

Ethical approval

This article does not contain any studies with human participants or animals performed by any of the authors.
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

Anhänge

Electronic supplementary material

Below is the link to the electronic supplementary material.
Literatur
1.
Zurück zum Zitat European Monitoring Centre for Drugs and Drug Addiction (2018) European drug report 2018: trends and developments. Publications Office of the European Union, Luxembourg European Monitoring Centre for Drugs and Drug Addiction (2018) European drug report 2018: trends and developments. Publications Office of the European Union, Luxembourg
2.
Zurück zum Zitat European Monitoring Centre for Drugs and Drug Addiction (2014) European drug report 2014: trends and developments. Publications Office of the European Union, Luxembourg European Monitoring Centre for Drugs and Drug Addiction (2014) European drug report 2014: trends and developments. Publications Office of the European Union, Luxembourg
3.
Zurück zum Zitat United Nations Office on Drugs and Crime (2018) World drug report 2018. United Nations publication, ViennaCrossRef United Nations Office on Drugs and Crime (2018) World drug report 2018. United Nations publication, ViennaCrossRef
4.
Zurück zum Zitat Barceloux DG (2012) Medical toxicology of drug abuse: synthesized chemicals and psychoactive plants. Wiley, HobokenCrossRef Barceloux DG (2012) Medical toxicology of drug abuse: synthesized chemicals and psychoactive plants. Wiley, HobokenCrossRef
6.
Zurück zum Zitat European Monitoring Centre for Drugs and Drug Addiction (2016) The internet and drug markets. Publications Office of the European Union, Luxembourg European Monitoring Centre for Drugs and Drug Addiction (2016) The internet and drug markets. Publications Office of the European Union, Luxembourg
12.
Zurück zum Zitat Montesano C, Vannutelli G, Gregori A, Ripani L, Compagnone D, Curini R, Sergi M (2016) Broad screening and identification of novel psychoactive substances in plasma by high-performance liquid chromatography–high-resolution mass spectrometry and post-run library matching. J Anal Toxicol 40:519–528. https://doi.org/10.1093/jat/bkw043 CrossRefPubMed Montesano C, Vannutelli G, Gregori A, Ripani L, Compagnone D, Curini R, Sergi M (2016) Broad screening and identification of novel psychoactive substances in plasma by high-performance liquid chromatography–high-resolution mass spectrometry and post-run library matching. J Anal Toxicol 40:519–528. https://​doi.​org/​10.​1093/​jat/​bkw043 CrossRefPubMed
17.
Zurück zum Zitat Berova N, Polavarapu PL, Nakanishi K, Woody RW (2012) Comprehensive chiroptical spectroscopy: applications in stereochemical analysis of synthetic compounds, natural products, and biomolecules. Wiley, HobokenCrossRef Berova N, Polavarapu PL, Nakanishi K, Woody RW (2012) Comprehensive chiroptical spectroscopy: applications in stereochemical analysis of synthetic compounds, natural products, and biomolecules. Wiley, HobokenCrossRef
18.
Zurück zum Zitat Nafie LA (2011) Vibrational optical activity: principles and applications. Wiley, ChichesterCrossRef Nafie LA (2011) Vibrational optical activity: principles and applications. Wiley, ChichesterCrossRef
21.
Zurück zum Zitat Stephens PJ, Pan JJ, Devlin FJ, Urbanová M, Julínek O, Hájíček J (2008) Determination of the absolute configurations of natural products via density functional theory calculations of vibrational circular dichroism, electronic circular dichroism, and optical rotation: the iso-schizozygane alkaloids isoschizogaline and isoschizogamine. Chirality 20:454–470. https://doi.org/10.1002/chir.20466 CrossRefPubMed Stephens PJ, Pan JJ, Devlin FJ, Urbanová M, Julínek O, Hájíček J (2008) Determination of the absolute configurations of natural products via density functional theory calculations of vibrational circular dichroism, electronic circular dichroism, and optical rotation: the iso-schizozygane alkaloids isoschizogaline and isoschizogamine. Chirality 20:454–470. https://​doi.​org/​10.​1002/​chir.​20466 CrossRefPubMed
22.
Zurück zum Zitat Polavarapu PL, Donahue EA, Shanmugam G, Scalmani G, Hawkins EK, Rizzo C, Ibnusaud I, Thomas G, Habel D, Sebastian D (2011) A single chiroptical spectroscopic method may not be able to establish the absolute configurations of diastereomers: dimethylesters of hibiscus and garcinia acids. J Phys Chem A 115:5665–5673. https://doi.org/10.1021/jp202501y CrossRefPubMedPubMedCentral Polavarapu PL, Donahue EA, Shanmugam G, Scalmani G, Hawkins EK, Rizzo C, Ibnusaud I, Thomas G, Habel D, Sebastian D (2011) A single chiroptical spectroscopic method may not be able to establish the absolute configurations of diastereomers: dimethylesters of hibiscus and garcinia acids. J Phys Chem A 115:5665–5673. https://​doi.​org/​10.​1021/​jp202501y CrossRefPubMedPubMedCentral
24.
Zurück zum Zitat Kohout M, Vandenbussche J, Roller A, Tůma J, Bogaerts J, Bultinck P, Herrebout W, Lindner W (2016) Absolute configuration of the antimalarial erythro-mefloquine—vibrational circular dichroism and X-ray diffraction studies of mefloquine and its thiourea derivative. RSC Adv 6:81461–81465. https://doi.org/10.1039/c6ra19367f CrossRef Kohout M, Vandenbussche J, Roller A, Tůma J, Bogaerts J, Bultinck P, Herrebout W, Lindner W (2016) Absolute configuration of the antimalarial erythro-mefloquine—vibrational circular dichroism and X-ray diffraction studies of mefloquine and its thiourea derivative. RSC Adv 6:81461–81465. https://​doi.​org/​10.​1039/​c6ra19367f CrossRef
27.
Zurück zum Zitat Jurásek B, Králík F, Rimpelová S, Čejka J, Setnička V, Ruml T, Kuchař M, Kohout M (2018) Synthesis, absolute configuration and in vitro cytotoxicity of deschloroketamine enantiomers: rediscovered and abused dissociative anaesthetic. New J Chem 42:19360–19368. https://doi.org/10.1039/c8nj03107j CrossRef Jurásek B, Králík F, Rimpelová S, Čejka J, Setnička V, Ruml T, Kuchař M, Kohout M (2018) Synthesis, absolute configuration and in vitro cytotoxicity of deschloroketamine enantiomers: rediscovered and abused dissociative anaesthetic. New J Chem 42:19360–19368. https://​doi.​org/​10.​1039/​c8nj03107j CrossRef
31.
33.
Zurück zum Zitat Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Scalmani G, Barone V, Mennucci B, Petersson GA, Nakatsuji H, Caricato M, Li X, Hratchian HP, Izmaylov AF, Bloino J, Zheng G, Sonnenberg JL, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Vreven T, Montgomery JA Jr, Peralta JE, Ogliaro F, Bearpark M, Heyd JJ, Brothers E, Kudin KN, Staroverov VN, Keith T, Kobayashi R, Normand J, Raghavachari K, Rendell A, Burant JC, Iyengar SS, Tomasi J, Cossi M, Rega N, Millam JM, Klene M, Knox JE, Cross JB, Bakken V, Adamo C, Jaramillo J, Gomperts R, Stratmann RE, Yazyev O, Austin AJ, Cammi R, Pomelli C, Ochterski JW, Martin RL, Morokuma K, Zakrzewski VG, Voth GA, Salvador P, Dannenberg JJ, Dapprich S, Daniels AD, Farkas O, Foresman JB, Ortiz JV, Cioslowski J, Fox DJ (2013) Gaussian 09 citation. Gaussian Inc, Wallingford Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Scalmani G, Barone V, Mennucci B, Petersson GA, Nakatsuji H, Caricato M, Li X, Hratchian HP, Izmaylov AF, Bloino J, Zheng G, Sonnenberg JL, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Vreven T, Montgomery JA Jr, Peralta JE, Ogliaro F, Bearpark M, Heyd JJ, Brothers E, Kudin KN, Staroverov VN, Keith T, Kobayashi R, Normand J, Raghavachari K, Rendell A, Burant JC, Iyengar SS, Tomasi J, Cossi M, Rega N, Millam JM, Klene M, Knox JE, Cross JB, Bakken V, Adamo C, Jaramillo J, Gomperts R, Stratmann RE, Yazyev O, Austin AJ, Cammi R, Pomelli C, Ochterski JW, Martin RL, Morokuma K, Zakrzewski VG, Voth GA, Salvador P, Dannenberg JJ, Dapprich S, Daniels AD, Farkas O, Foresman JB, Ortiz JV, Cioslowski J, Fox DJ (2013) Gaussian 09 citation. Gaussian Inc, Wallingford
39.
40.
Zurück zum Zitat Bouř P, Maloň P (1995–2009) The MCM program. Academy of Sciences Prague, Czech Republic Bouř P, Maloň P (1995–2009) The MCM program. Academy of Sciences Prague, Czech Republic
Metadaten
Titel
Methylone and pentylone: structural analysis of new psychoactive substances
verfasst von
Dita Spálovská
Tereza Maříková
Michal Kohout
František Králík
Martin Kuchař
Vladimír Setnička
Publikationsdatum
01.03.2019
Verlag
Springer Singapore
Erschienen in
Forensic Toxicology / Ausgabe 2/2019
Print ISSN: 1860-8965
Elektronische ISSN: 1860-8973
DOI
https://doi.org/10.1007/s11419-019-00468-z

Weitere Artikel der Ausgabe 2/2019

Forensic Toxicology 2/2019 Zur Ausgabe

Neu im Fachgebiet Rechtsmedizin

Molekularpathologische Untersuchungen im Wandel der Zeit

Open Access Biomarker Leitthema

Um auch an kleinen Gewebeproben zuverlässige und reproduzierbare Ergebnisse zu gewährleisten ist eine strenge Qualitätskontrolle in jedem Schritt des Arbeitsablaufs erforderlich. Eine nicht ordnungsgemäße Prüfung oder Behandlung des …

Vergleichende Pathologie in der onkologischen Forschung

Pathologie Leitthema

Die vergleichende experimentelle Pathologie („comparative experimental pathology“) ist ein Fachbereich an der Schnittstelle von Human- und Veterinärmedizin. Sie widmet sich der vergleichenden Erforschung von Gemeinsamkeiten und Unterschieden von …

Gastrointestinale Stromatumoren

Open Access GIST CME-Artikel

Gastrointestinale Stromatumoren (GIST) stellen seit über 20 Jahren ein Paradigma für die zielgerichtete Therapie mit Tyrosinkinaseinhibitoren dar. Eine elementare Voraussetzung für eine mögliche neoadjuvante oder adjuvante Behandlung bei …

Personalisierte Medizin in der Onkologie

Aufgrund des erheblichen technologischen Fortschritts in der molekularen und genetischen Diagnostik sowie zunehmender Erkenntnisse über die molekulare Pathogenese von Krankheiten hat in den letzten zwei Jahrzehnten ein grundlegender …