Next Article in Journal
Preparation of a First 18F-Labeled Agonist for M1 Muscarinic Acetylcholine Receptors
Next Article in Special Issue
Noncovalent Complexes of Cyclodextrin with Small Organic Molecules: Applications and Insights into Host–Guest Interactions in the Gas Phase and Condensed Phase
Previous Article in Journal
Opening up ZSM-5 Hierarchical Zeolite’s Porosity through Sequential Treatments for Improved Low-Density Polyethylene Cracking
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enantioseparation and Determination of Mephedrone and Its Metabolites by Capillary Electrophoresis Using Cyclodextrins as Chiral Selectors

1
Department of Analytical Chemistry, Faculty of Chemical Engineering, University of Chemistry and Technology, Prague, Technická 5, 166 28 Prague 6, Czech Republic
2
Department of Chemistry of Natural Compounds, Forensic Laboratory of Biologically Active Substances, Faculty of Food and Biochemical Technology, University of Chemistry and Technology, Prague, Technická 5, 166 28 Prague 6, Czech Republic
3
Department of Organic Chemistry, Faculty of Chemical Technology, University of Chemistry and Technology, Prague, Technická 5, 166 28 Prague 6, Czech Republic
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(12), 2879; https://doi.org/10.3390/molecules25122879
Submission received: 25 May 2020 / Revised: 15 June 2020 / Accepted: 19 June 2020 / Published: 23 June 2020
(This article belongs to the Special Issue Cyclodextrins: Recent Advances in Chemistry and Applications)

Abstract

:
Mephedrone, a psychoactive compound derived from cathinone, is widely used as a designer drug. The determination of mephedrone and its metabolites is important for understanding its possible use in medicine. In this work, a method of capillary electrophoresis for the chiral separation of mephedrone and its metabolites was developed. Carboxymethylated β-cyclodextrin was selected as the most effective chiral selector from seven tested cyclodextrin derivates. Based on the simplex method, the optimal composition of the background electrolyte was determined: at pH 2.75 and 7.5 mmol·L−1 carboxymethylated β-cyclodextrin the highest total resolution of a mixture of analytes was achieved. For mephedrone and its metabolites, calibration curves were constructed in a calibration range from 0.2 to 5 mmol·L−1; limits of detection, limits of quantification, precision, and repeatability were calculated, and according to Mandel’s fitting test, the linear calibration ranges were determined.

Graphical Abstract

1. Introduction

Mephedrone, a psychostimulant drug first synthesized in the 1920s, is classified as a synthetic cathinone [1]. According to the European Monitoring Centre for Drugs and Drug Addiction [2], synthetic cathinones, together with synthetic cannabinoids, are the most abundant substances in the group of so-called new psychoactive substances. In most cases, these drugs have stimulant entactogenic effects; however, the biological activity of such substances is difficult to predict and some of these artificially prepared substances have been shown to have potential therapeutic uses [3], for example, bupropion is used to treat depression and smoking cessation [4] and diethylpropion is prescribed as an anti-obesity drug [5]. On the other hand, pyrovalerone has been prescribed for the treatment of chronic obesity and lethargy but has been withdrawn for its abuse by patients [4,6]. These new psychoactive substances can also mimic prescription therapeutic psychoactive drugs; for example, phenmetrazine, modafinil, and methylphenidate, and hence are more available even for healthy people because they are available on the Internet [7]. So, an intensive study of these compounds is needed.
The most commonly used techniques for isolating new psychoactive substances from biological samples are liquid-to-liquid extraction (LLE) and solid-phase extraction (SPE) [8]. LLE is a simple method, but its problems are easy contamination and matrix influence. On the contrary, the SPE is very selective but time-consuming [9]. A popular method is the QuEChERS method (quick, easy, cheap, effective, rugged, and safe) modified in 2012 for the extraction of psychoactive substances from biological samples [10].
Simple colorimetric tests are used for rapid orientation detection of drugs. However, the identification is very indicative, detecting only the presence of certain structural motifs [11]. However, immunological methods that are commercially available are more commonly used for rapid drug detection. However, there is no sufficiently effective immunoassay on the market for the detection of cathinones that does not show cross-reactivity between analogs [9,12,13].
Of the advanced chromatographic techniques, gas and liquid chromatography with a mass spectrometer (GC-MS and LC-MS) are the most suitable for the analysis of synthetic cathinones. In general, the GC-MS technique is used more for drug analysis, mainly due to shorter elution times [14,15,16,17,18,19]. Chiral separation is most often performed in an achiral environment after the previous conversion of enantiomers to diastereoisomers [20], which, however, brings complications for the quantification itself [21]. It should be also noted that cathinones are thermodegradable substances and due to the thermal conditions necessary for separation by GC, their partial decomposition occurs [5]. Therefore, LC-MS [22,23,24] and LC-UV [25,26] using a chiral stationary or mobile phase is often used. Rapid separation of cathinones is also possible by supercritical fluid chromatography (SFC) [27,28]. Carnes et al. compared the separation of cathinones by ultrahigh efficiency SFC (UHPSFC), ultrahigh efficiency LC (UHPLC) with a non-polar column, UHPLC with a hydrophilic column, and GC with a weakly polar column. The combination of GC and UHPSFC, which is the fastest of the tested methods, seems to be ideal [29].
The first enantioseparation of cathinone derivatives by capillary electrophoresis (CE) was performed by Mohr et al. in 2012 using sulfated β-cyclodextrin as a chiral selector [30]. As with HPLC, CE-UV and CE-MS techniques are used [31,32,33]. For example, to study the incorporation of cathinones into hair, a CE method was developed in which extraction from hair at an elevated temperature and pressure into ammonium hydroxide solution was first performed. Subsequently, the extract was concentrated on a solid phase that was part of the CE capillary (inline coupling of SPE-CE), and analysis was performed using various cyclodextrins as chiral selectors [34].
Mephedrone is referred to as a catecholamine reuptake inhibitor [35,36,37], but it also positively affects the release of catecholamines into the synaptic cleft [13,38,39,40,41]. Therefore, it seems to have a mixed effect [42,43,44]. In both cases, the result is an increase in the concentration of catecholamines in the synaptic cleft, resulting in the typical action of psychostimulants. Acute intoxication with mephedrone, usually in combination with other drugs and alcohol, has already resulted in hundreds of deaths across Europe [9].
Mephedrone has a chiral center on the α carbon and exists in two enantiomeric forms (Figure 1). Both enantiomers have a similar affinity for dopaminergic transporters, but (S)-mephedrone is about 50 times more potent as a serotonergic mediator. Nevertheless, the (R)-enantiomer is primarily responsible for the euphoric effects [45].
In phase I biotransformation, mephedrone can undergo several reactions: (i) oxidative N-demethylation, (ii) oxidation of the 4-methyl group, (iii) ω-oxidation at the 3′ position, and (iv) reduction of the carbonyl group (Figure 2) [46].
The major metabolites of mephedrone (1; 4-methyl methcathinone) are 4-methylcathinone (2; 4-MC) and 4-hydroxymethyl methcathinone (3; 4-OH-MMC). These, together with 4-methylephedrine (4; 4-ME), 4-carboxymethylcathinone (6; 4-CMC), 4-methylnorephedrine (7; 4-MNE), and normephedrone-ω-carboxylic acid (3′-OOH-4-MC), were identified in the urine of users by ultra-high-performance liquid chromatography‒mass spectrometry (UHPLC-MS) [47]. Other metabolites, namely 4-carboxyephedrine (5; 4-CE), 4-MNE (7), 4-hydroxymethylcathinone (8; 4-OH-MC) [17], 4-carboxynorephedrine (9; 4-CNE), 4-carboxycathinone (4-CC), and 4-hydroxymethylnorephedrine (4-OH-MNE) [46] were found in the urine of rats by GC-MS and LC-MS.
The aim of this work was to find the optimal conditions for the separation of mephedrone and its metabolites by capillary electrophoresis (for details about this method, see, for example, Bernardo-Bermejo et al. [48]), using cyclodextrins (CDs) [49] as chiral selectors. The optimization of the background electrolyte (BGE) was carried out by using the simplex procedure, which was thoroughly described by Catai and Carrilho [50]. Mephedrone and all its available metabolites, selected on the basis of current knowledge on the biotransformation of mephedrone, are chiral substances that may undergo different metabolic pathways and thus have different biological effects for individual enantiomers.

2. Results and Discussion

2.1. Cyclodextrin Optimization

The first optimization step was to select a suitable CD. Based on our previous experience with CDs [51], analyses were performed in the presence of CDs at a concentration of 10 mmol·L−1, either in a 50 mmol·L−1 phosphate buffer at pH 2.5 (Figure 3) or in a 50 mmol·L−1 acetate buffer at pH 5 (Figure 4). From the measured electropherograms, it was concluded that the separation proceeds better in the phosphate buffer at a lower pH value. From the selected CDs (Figure 5), i.e., β-CD, carboxymethylated β-cyclodextrin (CM-β-CD), heptakis(2,3,6-tri-O-methyl)-β-cyclodextrin (Me-β-CD), 2-hydroxypropylated β-cyclodextrin (HP-β-CD), sulfated β-cyclodextrin (S-β-CD), γ-CD, and carboxymethylated γ-cyclodextrin (CM-γ-CD), CM-β-CD, which is often used for the separation of basic nitrogenous substances [52,53], seems to be the most suitable. This may be due to the interaction of the carboxyl group of CM-β-CD with the NH group of the analyte and also due to the appropriate size of the β-CD cavity. The inappropriateness of S-β-CD may be due to too many sulfate groups (7–11) compared to three carboxylic groups of CM-β-CD, see Section 3.1. In its presence, 18 diastereomers were separated within 30 min, so it was used in all the other work. The advantage of CM-β-CD over β-CD is its higher solubility in water (50 mg·mL−1 compared to 19 mg·mL−1). Thus, in the search for optimal separation conditions, it is possible to achieve higher concentrations if necessary.

2.2. BGE Optimization

The enantioselective separation of ionic analytes with ionic CDs is controlled by at least two mechanisms. The first is the interaction of the hydrophobic cavity of the CD derivative with the analyte; the second is the electrostatic interaction between the ionic groups of the CD derivative and the analyte [54]. The approximate value of the dissociation constant pKa CM-β-CD is 3.0–3.5 [55]. Cathinones are basic drugs and have high values of dissociation constants; for example, the pKa of mephedrone is 8.77 [56], for methylated cathinones it is 8.4–9.5 [57], for phenylethylamine it is 10.3 [58], and for methamphetamine it is 9.9 [59]. In addition to pH, the separation is also affected by the concentration of cyclodextrin, which affects the amount of complexed analyte and thus affects its overall electrophoretic mobility [49]. The separation efficiency can be increased at higher voltages, but this also increases the temperature difference in the middle and at the edge of the capillary, which leads to a decrease in efficiency. Based on our experience with a similar BGE [51], we have chosen a voltage (see Section 3.3) such that the current is around 50 μA. In addition to voltage, the current also depends on the ionic strength of the BGE. Again, we used our previous experience here, and in most cases used a buffer with a concentration of 50 mmol·L−1.
Thus, the optimization step was to find a suitable concentration of CM-β-CD and a suitable pH of BGE. In order to monitor both parameters simultaneously, the simplex method was used to find the optimal conditions [60]. The first step was to construct an initial simplex. The first two points correspond to the conditions: (1) pH 2.5 and 5 mmolv·L−1 CM-β-CD; (2) pH 3.5 and 5 mmol·L−1 CM-β-CD (Figure 6A). The third point was calculated on the basis of the properties of an equilateral triangle, i.e., pH3 = 2.5 + 0.5·1 = 3, where 2.5 corresponds to the first point and value 1 to the length of the simplex edge on the pH axis, and c3 = 5 + √3⁄2·5.774 = 10 mmol·L−1, where 5 is the initial concentration and 5.774 is the length of the simplex edge corresponding to the c-axis, which has been chosen so that the CM-β-CD concentration is an integer.
A mixture of analytes 1–9 (Figure 2) with a concentration of 1 mmol·L−1 was measured twice under each pair of conditions. Based on the resolution of the peaks of the individual enantiomers, the average criterion Rkr was calculated (see above), which served as a compared response for the simplex procedure. Simplex followed the rules of the equilateral simplex procedure. After calculating the coordinates of point (8), the simplex looped. Thus, the chiral separation proceeded best under the initial conditions chosen. The simplex procedure is summarized in Figure 6A.
A half-simplex was designed to more accurately determine the optimal chiral separation conditions (Figure 6B). The first point corresponds to the first determined optimal conditions: (1) pH 2.5 and 5 mmol·L−1 CM-β-CD. The other two points correspond to: (2) pH2 = 2.5 − 0.5·0.5 = 2.25 and c2 = 5 + √3⁄2·2.887 = 7.5; (3) pH3 = 2.5 + 0.5·0.5 = 2.75 and c3 = c2, where the values 0.5 and 2.887 correspond to half the sizes of the edges of the first simplex.
Furthermore, the simplex proceeded according to the rules of the equilateral simplex procedure. Since points (3) and (4) provide the same response, Rkr = 12.5, it is possible in a situation where the simplex forms points (3), (4), and (5), to proceed in two directions, namely to point (6) and point (6′), which gives a small response. Points (3) and (4) also provide the maximum value of the criterion Rkr. Point (3) was chosen as the point meeting the optimal conditions, as in the second case the analysis was extended to more than 100 min. The simplex procedure is summarized in Figure 6B. The resulting optimum was found at pH 2.75 and 7.5 mmol·L−1 CM-β-CD (Figure 7).
At pH 2.75, approximately 65–85% of CM-β-CD is in a protonated form. At a low pH, most of the silanol groups on the capillary wall are protonated, leading to low electroosmotic flow, and CM-β-CD carries an approximately one-fold negative charge. At pH 2.75, the amine groups of all analytes are practically 100% protonated. Metabolites 5, 6, and 9 carry a carboxyl group on the aromatic nucleus in the para position. Upon approximation with benzoic acid, whose pKa is 4.21, it can be assumed that at pH 2.75 the carboxyl group is 4% deprotonated. Thus, it is clear that pH plays an important role in the separation. A lower or higher pH than 2.75 affects the degree of deprotonation of the used CM-β-CD, and thus the separation efficiency.
The concentration of CM-β-CD is also an important factor as it affects how much analyte will be complexed with CD. For a given racemic mixture, there is an optimal concentration at which the individual enantiomers will be separated as best as possible [49,61]. This concentration depends on the stability constants of the individual enantiomers and the given chiral selector. In the case of separation of a mixture of racemic substances, as in our case, it is necessary to choose a compromise between these concentrations.

2.3. Calibration

Calibration solutions were measured under the optimized BGE composition, namely at 7.5 mmol·L−1 CM-β-CD and pH 2.75. Reduced peak areas (area/migration time) were subjected to the Dean–Dixon test to exclude outliers, which is adapted to small datasets with unknown distributions.
Furthermore, the regression parameters of the calibration curves were calculated. Based on the test of the significance of the parameter, we assessed whether it is possible to set the intercept a in the regression equation y = bx + a. In the case of analytes 1, 2, 5, and 7, the size of the intercept is insignificant according to the test. Subsequently, the confidence intervals for the intercept a and the slope b of all calibration equations were calculated. If the significance test of the parameter did not allow for ignoring the intercept, the number of degrees of freedom for calculating tcrit was equal to n − 2, and in the case of a zero intercept n − 1. Ignoring the intercept improved the values of the coefficient of determination R2 and thus allowed us to reduce the confidence intervals (Table 1).
According to Mandel’s test, linear concentration ranges were determined and subsequently the LOD and LOQ of individual analytes were calculated (Table 2). Precision for all studied analytes was <4.9% and repeatability <4.5%.

3. Materials and Methods

3.1. Chemicals and Materials

The following substances were used: ortho-phosphoric acid (50%), β-cyclodextrin (β-CD, 97%), carboxymethylated β-cyclodextrin sodium salt (CM-β-CD, average degree of substitution ~3, 95%), 2-hydroxypropylated β-cyclodextrin (HP-β-CD, average degree of substitution ~0.5–1.3, 95%), sulfated β-cyclodextrin (S-β-CD, average degree of substitution ~7–11, 95%), acetic acid (99%), thiourea (99%) (all Sigma-Aldrich, Praha, Czech Republic), ultrapure water (Milli-Q grade, Millipore, Molsheim, France), 1 mol·L−1 sodium hydroxide (Tripur, Merck, Darmstadt, Germany), hydrochloric acid (30%)(Suprapur, Merck, Darmstadt, Germany), γ-cyclodextrin (γ-CD, 97%, FUJIFILM Wako, Richmond, VA, USA), carboxymethylated γ-cyclodextrin sodium salt (CM-γ-CD, average degree of substitution ~3–6, 95%, AraChem, Kuala Lumpur, Malaysia), and heptakis(2,3,6-tri-O-methyl)-β-cyclodextrin (Me-β-CD, 98%, Fluka, Munich, Germany). Psychoactive amines 1–9 (Figure 2) in the form of hydrochloride salts were prepared previously [46]. The purity of the salts studied was determined by UHPLC/MS Agilent 6460 (Waldbronn, Germany) and was higher than 95%.

3.2. Equipment

CE separations were performed with an Agilent CE instrument 7100 (Agilent 3D HPCE, Waldbronn, Germany) equipped with a UV-Vis diode-array detector. A bare fused-silica capillary of 375/75 μm od/id and 58.5/50 cm total/effective length was obtained from Polymicro Technologies (Phoenix, AZ, USA) was used.

3.3. CE Conditions

The BGE used for the optimizing of the chiral selector consisted of 50 mmol·L−1 sodium phosphate buffer, pH 2.5 (50 mmol·L−1 orthophosphoric acid adjusted to appropriate pH with 1 mol·L−1 NaOH) or acetate buffer, pH 5.0 (50 mmol·L−1 acetic acid adjusted to appropriate pH with 1 mol·L−1 NaOH), and 10 mmol·L−1 of the studied CDs (β-CD, γ-CD, CM-β-CD, CM-γ-CD, HP-β-CD, 2-Me-β-CD, or S-β-CD). The background electrolyte used for the optimizing of the chiral selector concentration and pH consisted of 46.6 mmol·L−1 hydrochloric acid (pH 1.5) or 50 mmol·L−1 phosphate buffer (pH 2; 2.25; 2.5; 2.75; 3; 3.25; 3.5) and 0–10 mmol·L−1 of CM-β-CD.
A new fused-silica capillary was first rinsed with 1 mol·L−1 NaOH for 30 min, then with H2O for 30 min. Between the runs, the capillary was rinsed at 99.4 kPa first with 0.1 mol·L−1 NaOH for 2 min, then with H2O also for 2 min, and finally with the running buffer again for 2 min (for the capillary washing, a different buffer solution than for the subsequent analysis was used). The analytes were injected hydrodynamically at a pressure of 1.5 kPa for 5 s. Separations were performed at 10 kV, 20 kV (anode at the injection capillary end), or −20 kV with a voltage ramp time of 12 s. Detection was carried out at 207 nm during the optimization step and 258 nm for analytes 1, 2, 3, and 6; 236 nm for analyte 5; and 214 nm for analytes 4 and 7 (i.e., at wavelengths corresponding to the absorption maxima) during the calibration step. The capillary was thermostated at 25 °C during the analyses.

3.4. Sample Preparation

Individual analytes were dissolved in water at a 20 mmol·L−1 concentration. For optimizing the chiral selector, the analyte solutions were further mixed and diluted with water to a final concentration of 1 mmol·L−1. For calibration, seven different mixtures of analytes 1–7 (Figure 2) at concentration ranges from 0.2 to 5 mmol·L−1 (the sum of all analytes’ concentrations was 10 mmol·L−1) were prepared and each mixture was analyzed five times.

3.5. Simplex Method

For optimizing the BGE composition, the two-dimensional simplex method was used [60]. The optimizing parameters were CM-β-CD concentration and electrolyte pH. As an evaluation criterion, Rcr was introduced and calculated as follows. For resolution between two neighborhood peaks i and j, if (Rs,ij) < 0.5 then Rcr,ij = 0, if 0.5 < Rs,ij < 1.5 then Rcr,ij = 0.5, and if Rs,ij > 1.5 then Rcr,ij = 1. The sum of Rcr,ij = Rcr.

4. Conclusions

In this work, the composition of the basic electrolyte for the chiral separation of mephedrone and its selected metabolites was optimized. A total of seven cyclodextrin derivatives were selected as potential chiral selectors. From the seven tested cyclodextrins, carboxymethylated β-cyclodextrin was selected as the most suitable. Cyclodextrin concentration and pH were determined based on the simplex method for two parameters. The optimized pH was found to be 2.75 and the concentration of carboxymethylated β-cyclodextrin was 7.5 mmol·L−1. A total of nine analytes were present in the mixture, and 18 distinguishable peaks were found. Thus, all analytes were separated.
Furthermore, the data for the construction of calibration dependences were measured. According to the Dean‒Dixon test, outliers were excluded, and regression parameters of calibration lines were calculated. Based on the test of significance of the parameter, the intercept on the y-axis was neglected in some cases. Finally, linear concentration ranges according to Mandel’s test and detection limits, limits of determination, precision, and repeatability were determined.

Author Contributions

Conceptualization, P.Ř. and M.K.; methodology, P.Ř. and D.M.; validation, D.M.; formal analysis, D.M.; investigation, D.M.; resources, P.Ř., R.J., and M.H.; writing—original draft preparation, P.Ř.; writing—review and editing, P.Ř. and M.K.; visualization, P.Ř. and D.M.; supervision, P.Ř. and M.K.; project administration, P.Ř.; funding acquisition, P.Ř. and M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Interior of the Czech Republic (Grant number VI20172020056).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sanchez:, S. Sur un homologue de l’ephedrine. Bull. Soc. Chim. Fr. 1929, 45, 284–286. [Google Scholar]
  2. European Monitoring Centre for Drugs and Drug Addiction. Available online: http://www.emcdda.europa.eu/publications/edr/trends-developments/2019_en (accessed on 5 May 2020).
  3. Busardo, F.P.; Di Trana, A.; Montanari, E.; Mauloni, S.; Tagliabracci, A.; Giorgetti, R. Is etizolam a safe medication? Effects on psychomotor perfomance at therapeutic dosages of a newly abused psychoactive substance. Forensic Sci. Int. 2019, 301, 137–141. [Google Scholar] [CrossRef] [PubMed]
  4. Prosser, J.M.; Nelson, L.S. The toxicology of bath salts: A review of synthetic cathinones. J. Med. Toxicol. 2012, 8, 33–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kerrigan, S.; Savage, M.; Cavazos, C.; Bella, P. Thermal degradation of synthetic cathinones: Implications for forensic toxicology. J. Anal. Toxicol. 2016, 40, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Goldberg, J.; Gardos, G.; Cole, J.O. A controlled evaluation of pyrovalerone in chronically fatigued volunteers. Int. Pharmacopsychiatry 1973, 8, 60–69. [Google Scholar] [CrossRef] [PubMed]
  7. Batisse, A.; Eiden, C.; Peyriere, H.; Djezzar, S. Use of new psychoactive substances to mimic prescription drugs: The trend in France. NeuroToxicology 2020, 79, 20–24. [Google Scholar] [CrossRef]
  8. Alremeithi, R.; Meetani, M.A.; Alaidaros, A.A.; Lanjawi, A.; Alsumaiti, K. Simultaneous Quantitative Determination of Synthetic Cathinone Enantiomers in Urine and Plasma Using GC-NCI-MS. J. Anal. Methods Chem. 2018, 2018, 1–11. [Google Scholar] [CrossRef] [Green Version]
  9. Majchrzak, M.; Celinski, R.; Kus, P.; Kowalska, T.; Sajewicz, M. The newest cathinone derivatives as designer drugs: An analytical and toxicological review. Forensic Toxicol. 2018, 1–18. [Google Scholar] [CrossRef] [Green Version]
  10. Usui, K.; Hayashizaki, Y.; Hashiyada, M.; Funayama, M. Rapid drug extraction from human whole blood using a modified QuEChERS extraction method. Leg. Med. 2012, 14, 286–296. [Google Scholar] [CrossRef]
  11. Toole, K.E.; Fu, S.; Shimmon, R.G.; Kraymen, N.; Taflaga, S.; Forensic, A.F. Color tests for the preliminary identification of methcathinone and analogues of methcathinone. Microgram J. 2012, 9, 27–32. [Google Scholar]
  12. Ellefsen, K.N.; Anizan, S.; Castaneto, M.S.; Desrosiers, N.A.; Martin, T.M.; Klette, K.L.; Huestis, M.A. Validation of the only commercially available immunoassay for synthetic cathinones in urine: Randox Drugs of Abuse V Biochip Array Technology. Drug Test. Anal. 2014, 6, 728–738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Paillet-Loilier, M.; Cesbron, A.; Bourgine, J.; Le, B.R.; Debruyne, D. Emerging drugs of abuse: Current perspectives on substituted cathinones. Subst. Abuse Rehabil. 2014, 5, 37–52. [Google Scholar] [PubMed] [Green Version]
  14. Daeid, N.N.; Savage, K.A.; Ramsay, D.; Holland, C.; Sutcliffe, O.B. Development of gas chromatography–mass spectrometry (GC–MS) and other rapid screening methods for the analysis of 16 ‘legal high’cathinone derivatives. Sci. Justice 2014, 54, 22–31. [Google Scholar] [CrossRef] [PubMed]
  15. Camilleri, A.; Johnston, M.R.; Brennan, M.; Davis, S.; Caldicott, D.G. Chemical analysis of four capsules containing the controlled substance analogues 4-methylmethcathinone, 2-fluoromethamphetamine, alpha-phthalimidopropiophenone and N-ethylcathinone. Forensic Sci. Int. 2010, 197, 59–66. [Google Scholar] [CrossRef]
  16. Fujii, H.; Hara, K.; Kageura, M.; Kashiwagi, M.; Matsusue, A.; Kubo, S. High throughput chiral analysis of urinary amphetamines by GC-MS using a short narrow-bore capillary column. Forensic Toxicol. 2009, 27, 75–80. [Google Scholar] [CrossRef]
  17. Meyer, M.R.; Wilhelm, J.; Peters, F.T.; Maurer, H.H. Beta-keto amphetamines: Studies on the metabolism of the designer drug mephedrone and toxicological detection of mephedrone, butylone, and methylone in urine using gas chromatography-mass spectrometry. Anal. Bioanal. Chem. 2010, 397, 1225–1233. [Google Scholar] [CrossRef]
  18. Olesti, E.; Pujadas, M.; Papaseit, E.; Perez-Mana, C.; Pozo, O.J.; Farre, M.; de la Torre, R. GC–MS quantification method for mephedrone in plasma and urine: Application to human pharmacokinetics. J. Anal. Toxicol. 2017, 41, 100–106. [Google Scholar] [CrossRef] [Green Version]
  19. Rasmussen, L.B.; Olsen, K.H.; Johansen, S.S. Chiral separation and quantification of R/S-amphetamine, R/S-methamphetamine, R/S-MDA, R/S-MDMA, and R/S-MDEA in whole blood by GC-EI-MS. J. Chromatogr. B 2006, 842, 136–141. [Google Scholar] [CrossRef]
  20. Ribeiro, C.; Santos, C.; Gonçalves, V.; Ramos, A.; Afonso, C.; Tiritan, M.E. Chiral Drug Analysis in Forensic Chemistry: An Overview. Molecules 2018, 23, 262. [Google Scholar] [CrossRef] [Green Version]
  21. Mohr, S.; Weiß, J.A.; Spreitz, J.; Schmid, M.G. Chiral separation of new cathinone-and amphetamine-related designer drugs by gas chromatography–mass spectrometry using trifluoroacetyl-l-prolyl chloride as chiral derivatization reagent. J. Chromatogr. A 2012, 1269, 352–359. [Google Scholar] [CrossRef]
  22. Pedersen, A.J.; Dalsgaard, P.W.; Rode, A.J.; Rasmussen, B.S.; Müller, I.B.; Johansen, S.S.; Linnet, K. Screening for illicit and medicinal drugs in whole blood using fully automated SPE and ultra-high-performance liquid chromatography with TOF-MS with data-independent acquisition. J. Sep. Sci. 2013, 36, 2081–2089. [Google Scholar] [CrossRef]
  23. Paul, M.; Ippisch, J.; Herrmann, C.; Guber, S.; Schultis, W. Analysis of new designer drugs and common drugs of abuse in urine by a combined targeted and untargeted LC-HR-QTOFMS approach. Anal. Bioanal. Chem. 2014, 406, 4425–4441. [Google Scholar] [CrossRef]
  24. Concheiro, M.; Castaneto, M.; Kronstrand, R.; Huestis, M.A. Simultaneous determination of 40 novel psychoactive stimulants in urine by liquid chromatography–high resolution mass spectrometry and library matching. J. Chromatogr. A 2015, 1397, 32–42. [Google Scholar] [CrossRef] [Green Version]
  25. Mohr, S.; Taschwer, M.; Schmid, M.G. Chiral separation of cathinone derivatives used as recreational drugs by HPLC-UV using a CHIRALPAK® AS-H column as stationary phase. Chirality 2012, 24, 486–492. [Google Scholar] [CrossRef]
  26. Taschwer, M.; Seidl, Y.; Mohr, S.; Schmid, M.G. Chiral Separation of Cathinone and Amphetamine Derivatives by HPLC/UV Using Sulfated ß-Cyclodextrin as Chiral Mobile Phase Additive. Chirality 2014, 26, 411–418. [Google Scholar] [CrossRef]
  27. Pauk, V.; Zihlová, V.; Borovcova, L.; Havlicek, V.; Schug, K.; Lemr, K. Fast separation of selected cathinones and phenylethylamines by supercritical fluid chromatography. J. Chromatogr. A 2015, 1423, 169–176. [Google Scholar] [CrossRef]
  28. Geryk, R.; Kaliková, K.; Schmid, M.G.; Tesarová, E. Enantioselective separation of biologically active basic compounds in ultra-performance supercritical fluid chromatography. Anal. Chim. Acta 2016, 932, 98–105. [Google Scholar] [CrossRef]
  29. Carnes, S.; O’Brien, S.; Szewczak, A.; Tremeau-Cayel, L.; Rowe, W.F.; McCord, B.; Lurie, I.S. Comparison of ultra high performance supercritical fluid chromatography, ultra high performance liquid chromatography, and gas chromatography for the separation of synthetic cathinones. J. Sep. Sci. 2017, 40, 3545–3556. [Google Scholar] [CrossRef]
  30. Mohr, S.; Pilaj, S.; Schmid, M.G. Chiral separation of cathinone derivatives used as recreational drugs by cyclodextrin-modified capillary electrophoresis. Electrophoresis 2012, 33, 1624–1630. [Google Scholar] [CrossRef]
  31. Merola, G.; Fu, H.; Tagliaro, F.; Macchia, T.; McCord, B.R. Chiral separation of 12 cathinone analogs by cyclodextrin-assisted capillary electrophoresis with UV and mass spectrometry detection. Electrophoresis 2014, 35, 3231–3241. [Google Scholar] [CrossRef]
  32. Taschwer, M.; Weiß, J.A.; Kunert, O.; Schmid, M.G. Analysis and characterization of the novel psychoactive drug 4-chloromethcathinone (clephedrone). Forensic Sci. Int. 2014, 244, 56–59. [Google Scholar] [CrossRef]
  33. Nowak, P.M.; Olesek, K.; Wozniakiewicz, M.; Koscielniak, P. Simultaneous enantioseparation of methcathinone and two isomeric methylmethcathinones using capillary electrophoresis assisted by 2-hydroxyethyl-β-cyclodextrin. Electrophoresis 2018, 39, 2406–2409. [Google Scholar] [CrossRef] [PubMed]
  34. Baciu, T.; Borrull, F.; Calull, M.; Aguilar, C. Enantioselective determination of cathinone derivatives in human hair by capillary electrophoresis combined in-line with solid-phase extraction. Electrophoresis 2016, 37, 2352–2362. [Google Scholar] [CrossRef] [PubMed]
  35. Simmler, L.D.; Buser, T.A.; Donzelli, M.; Schramm, Y.; Dieu, L.H.; Huwyler, J.; Chaboz, S.; Hoener, M.C.; Liechti, M.E. Pharmacological characterization of designer cathinones in vitro. Br. J. Pharmacol. 2013, 168, 458–470. [Google Scholar] [CrossRef] [Green Version]
  36. Martinez-Clemente, J.; Escubedo, E.; Pubill, D.; Camarasa, J. Interaction of mephedrone with dopamine and serotonin targets in rats. Eur. Neuropsychopharmacol. 2012, 22, 231–236. [Google Scholar] [CrossRef] [PubMed]
  37. Lopez-Arnau, R.; Martinez-Clemente, J.; Pubill, D.; Escubedo, E.; Camarasa, J. Comparative neuropharmacology of three psychostimulant cathinone derivatives: Butylone, mephedrone and methylone. Br. J. Pharmacol. 2012, 167, 407–420. [Google Scholar] [CrossRef] [Green Version]
  38. Chavant, F.; Boucher, A.; Le, B.R.; Debruyne, D.; Deheul, S. New synthetic drugs in addictovigilance. Therapie 2015, 70, 167–189. [Google Scholar] [CrossRef]
  39. Philogene-Khalid, H.L.; Hicks, C.; Reitz, A.B.; Liu-Chen, L.Y.; Rawls, S.M. Synthetic cathinones and stereochemistry: S enantiomer of mephedrone reduces anxiety- and depressant-like effects in cocaine- or MDPV-abstinent rats. Drug Alcohol Depend. 2017, 178, 119–125. [Google Scholar] [CrossRef]
  40. Schifano, F.; Albanese, A.; Fergus, S.; Stair, J.L.; Deluca, P.; Corazza, O.; Davey, Z.; Corkery, J.; Siemann, H.; Scherbaum, N.; et al. Mephedrone (4-methylmethcathinone; meow meow’): Chemical, pharmacological and clinical issues. Psychopharmacology (Heidelb. Ger.) 2011, 214, 593–602. [Google Scholar] [CrossRef] [Green Version]
  41. Ciechomska, M.; Wozniakiewicz, M.; Wietecha-Posluszny, R. Activity and biotransformation of three synthetic legal highs: Mephedrone, methylone and 3,4-methylenodioxypyrovalerone. Z Zagadnien Nauk Sadowych 2012, 89, 71–85. [Google Scholar]
  42. Karila, L.; Megarbane, B.; Cottencin, O.; Lejoyeux, M. Synthetic Cathinones: A New Public Health Problem. Curr. Neuropharmacol. 2015, 13, 12–20. [Google Scholar] [CrossRef] [Green Version]
  43. Papaseit, E.; Perez-Mana, C.; Mateus, J.A.; Pujadas, M.; Fonseca, F.; Torrens, M.; Olesti, E.; de la Torre, R.; Farre, M. Human pharmacology of mephedrone in comparison with MDMA. Neuropsychopharmacology 2016, 41, 2704. [Google Scholar] [CrossRef] [Green Version]
  44. Sichova, K.; Pinterova, N.; Horsley, R.R.; Lhotkova, E.; Stefkova, K.; Vejmola, C.; Uttl, L.; Kuchar, M.; Palenicek, T.; Pinterova, N.; et al. Mephedrone (4-Methylmethcathinone): Acute Behavioral Effects, Hyperthermic, and Pharmacokinetic Profile in Rats. Front. Psychiatry 2017, 8, 306. [Google Scholar] [CrossRef] [Green Version]
  45. Gregg, R.A.; Baumann, M.H.; Partilla, J.S.; Bonano, J.S.; Vouga, A.; Tallarida, C.S.; Velvadapu, V.; Smith, G.R.; Peet, M.M.; Reitz, A.B. Stereochemistry of mephedrone neuropharmacology: Enantiomer-specific behavioural and neurochemical effects in rats. Br. J. Pharmacol. 2015, 172, 883–894. [Google Scholar] [CrossRef] [Green Version]
  46. Linhart, I.; Himl, M.; Zidkova, M.; Balikova, M.; Lhotkova, E.; Palenicek, T. Metabolic profile of mephedrone: Identification of nor-mephedrone conjugates with dicarboxylic acids as a new type of xenobiotic phase II metabolites. Toxicol. Lett. 2016, 240, 114–121. [Google Scholar] [CrossRef]
  47. Pozo, O.J.; Ibanez, M.; Sancho, J.V.; Lahoz-Beneytez, J.; Farre, M.; Papaseit, E.; de la Torre, R.; Hernandez, F. Mass spectrometric evaluation of mephedrone in vivo human metabolism: Identification of phase I and phase II metabolites, including a novel succinyl conjugate. Drug Metab. Dispos. 2015, 43, 248–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Bernardo-Bermejo, S.; Sánchez-López, E.; Castro-Puyana, M.; Marina, M.L. Chiral capillary electrophoresis. TrAC Trends Anal. Chem. 2020, 124, 115807. [Google Scholar] [CrossRef]
  49. Rezanka, P.; Navratilova, K.; Rezanka, M.; Kral, V.; Sykora, D. Application of cyclodextrins in chiral capillary electrophoresis. Electrophoresis 2014, 35, 2701–2721. [Google Scholar] [CrossRef]
  50. Catai, J.R.; Carrilho, E. Simplex optimization of electrokinetic injection of DNA in capillary electrophoresis using dilute polymer solution. Electrophoresis 2003, 24, 648–654. [Google Scholar] [CrossRef]
  51. Rezanka, P.; Navratilova, K.; Rysava, H.; Masek, J.; Rokosova, L.; Blahova, M.; Rezanka, M.; Jindrich, J.; Sykora, D.; Kral, V. Influence of substituent position and cavity size of all regioisomers of monocarboxymethyl-α-, β-, and γ-cyclodextrins on their apparent stability constants of their complexes with both enantiomers of Tröger’s base. J. Sep. Sci. 2016, 39, 980–985. [Google Scholar] [CrossRef]
  52. Řezanková, K.; Kohoutová, R.; Kuchař, M.; Král, V.; Řezanka, P. Enantioseparation of novel psychoactive chiral amines and their mixture by capillary electrophoresis using cyclodextrins as chiral selectors. Chem. Pap. 2018, 72, 2737–2743. [Google Scholar] [CrossRef]
  53. Jurasek, B.; Himl, M.; Jurok, R.; Hajkova, K.; Vobinuskova, A.; Rezanka, P.; Kuchar, M. Synthesis of methoxetamine, its metabolites and deuterium labelled analog as analytical standards and their HPLC and chiral capillary electrophoresis separation. RSC Adv. 2017, 7, 56691–56696. [Google Scholar] [CrossRef] [Green Version]
  54. Szeman, J.; Ganzler, K.; Salgo, A.; Szejtli, J. Effect of the degree of substitution of cyclodextrin derivatives on chiral separations by high-performance liquid chromatography and capillary electrophoresis. J. Chromatogr. A 1996, 728, 423–431. [Google Scholar] [CrossRef]
  55. Sabbah, S.; Scriba, G.K.E. Separation of dipeptide and tripeptide enantiomers in capillary electrophoresis using carboxymethyl-β-cyclodextrin and succinyl-β-cyclodextrin: Influence of the amino acid sequence, nature of the cyclodextrin and pH. Electrophoresis 2001, 22, 1385–1393. [Google Scholar] [CrossRef]
  56. Nowak, P.M.; Olesek, K.; Wozniakiewicz, M.; Mitoraj, M.; Sagan, F.; Koscielniak, P. Cyclodextrin-induced acidity modification of substituted cathinones studied by capillary electrophoresis supported by density functional theory calculations. J. Chromatogr. A 2018, 1580, 142–151. [Google Scholar] [CrossRef] [PubMed]
  57. Gibbons, S.; Zloh, M. An analysis of the ‘legal high’mephedrone. Bioorg. Med. Chem. Lett. 2010, 20, 4135–4139. [Google Scholar] [CrossRef] [PubMed]
  58. Wan, H.; Holmen, A.; Nagard, M.; Lindberg, W. Rapid screening of pKa values of pharmaceuticals by pressure-assisted capillary electrophoresis combined with short-end injection. J. Chromatogr. A 2002, 979, 369–377. [Google Scholar] [CrossRef]
  59. Glicksberg, L.; Winecker, R.; Miller, C.; Kerrigan, S. Postmortem distribution and redistribution of synthetic cathinones. Forensic Toxicol. 2018, 36, 291–303. [Google Scholar] [CrossRef]
  60. Spendley, W.; Hext, G.R.; Himsworth, F.R. Sequential application of simplex designs in optimisation and evolutionary operation. Technometrics 1962, 4, 441–461. [Google Scholar] [CrossRef]
  61. Rezanka, P.; Rysava, H.; Havlik, M.; Jakubek, M.; Sykora, D.; Kral, V. Enantioseparation of Troeger’s Base Derivatives by Capillary Electrophoresis Using Cyclodextrins as Chiral Selectors. Chirality 2013, 25, 379–383. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the compounds 19 (Figure 2) are available from the authors.
Figure 1. Structures of (S)- and (R)-mephedrone.
Figure 1. Structures of (S)- and (R)-mephedrone.
Molecules 25 02879 g001
Figure 2. Scheme of phase I biotransformation of mephedrone: (i) oxidative N-demethylation, (ii) oxidation of the 4-methyl group, (iii) ω-oxidation at the 3′ position, and (iv) reduction of the carbonyl group.
Figure 2. Scheme of phase I biotransformation of mephedrone: (i) oxidative N-demethylation, (ii) oxidation of the 4-methyl group, (iii) ω-oxidation at the 3′ position, and (iv) reduction of the carbonyl group.
Molecules 25 02879 g002
Figure 3. Electropherograms of the analyzed mixture of the studied analytes 1–9 (Figure 2) obtained at different CDs used; fused-silica capillary: od/id = 375/75 μm, total/effective length: 58.5/50.0 cm; BGE: 50 mmol·L−1 phosphate buffer, pH 2.5, 10 mmol·L−1 CD; voltage: 20 kV (−20 kV for S-β-CD); temperature: 25 °C.
Figure 3. Electropherograms of the analyzed mixture of the studied analytes 1–9 (Figure 2) obtained at different CDs used; fused-silica capillary: od/id = 375/75 μm, total/effective length: 58.5/50.0 cm; BGE: 50 mmol·L−1 phosphate buffer, pH 2.5, 10 mmol·L−1 CD; voltage: 20 kV (−20 kV for S-β-CD); temperature: 25 °C.
Molecules 25 02879 g003
Figure 4. Electropherograms of the analyzed mixture of the studied analytes 1–9 (Figure 2) obtained at different CDs used; fused-silica capillary: od/id = 375/75 μm, total/effective length: 58.5/50.0 cm; BGE: 50 mmol·L−1 acetate buffer, pH 5, 10 mmol·L−1 CD; voltage: 20 kV; temperature: 25 °C.
Figure 4. Electropherograms of the analyzed mixture of the studied analytes 1–9 (Figure 2) obtained at different CDs used; fused-silica capillary: od/id = 375/75 μm, total/effective length: 58.5/50.0 cm; BGE: 50 mmol·L−1 acetate buffer, pH 5, 10 mmol·L−1 CD; voltage: 20 kV; temperature: 25 °C.
Molecules 25 02879 g004
Figure 5. Structures of CDs derivatives used as chiral selectors.
Figure 5. Structures of CDs derivatives used as chiral selectors.
Molecules 25 02879 g005
Figure 6. Simplex procedures in optimization of separation conditions; (A) full-simplex; (B) half-simplex; fused-silica capillary: od/id = 375/75 μm, total/effective length: 58.5/50.0 cm; BGE: 50 mmol·L−1 phosphate buffer (46.6 mmol·L−1 HCl at pH 1.5), variable CM-β-CD concentration, variable pH; voltage: 20 kV (10 kV for pH 1.5); temperature: 25 °C.
Figure 6. Simplex procedures in optimization of separation conditions; (A) full-simplex; (B) half-simplex; fused-silica capillary: od/id = 375/75 μm, total/effective length: 58.5/50.0 cm; BGE: 50 mmol·L−1 phosphate buffer (46.6 mmol·L−1 HCl at pH 1.5), variable CM-β-CD concentration, variable pH; voltage: 20 kV (10 kV for pH 1.5); temperature: 25 °C.
Molecules 25 02879 g006
Figure 7. Electropherogram of the analyzed mixture of the studied analytes 1–9 (Figure 2) obtained; fused-silica capillary: od/id = 375/75 μm, total/effective length: 58.5/50.0 cm; BGE: 50 mmol·L−1 phosphate buffer, pH 2.75, 7.5 mmol·L−1 CM-β-CD; voltage: 20 kV; temperature: 25 °C.
Figure 7. Electropherogram of the analyzed mixture of the studied analytes 1–9 (Figure 2) obtained; fused-silica capillary: od/id = 375/75 μm, total/effective length: 58.5/50.0 cm; BGE: 50 mmol·L−1 phosphate buffer, pH 2.75, 7.5 mmol·L−1 CM-β-CD; voltage: 20 kV; temperature: 25 °C.
Molecules 25 02879 g007
Table 1. Calibration dependences of separation of analytes 1–7 (Figure 2) measured under optimal conditions, where (a ± L (a)) is the intercept and (b ± L (b)) is the slope with a confidence interval at the significance level α = 0.05; R2 is the coefficient of determination.
Table 1. Calibration dependences of separation of analytes 1–7 (Figure 2) measured under optimal conditions, where (a ± L (a)) is the intercept and (b ± L (b)) is the slope with a confidence interval at the significance level α = 0.05; R2 is the coefficient of determination.
Analytea L(a)b L(b)R2
mmol·L−1 mmol·L−1
10 32.60±0.270.991
20 27.76±0.160.995
3−5.47±1.3123.54±1.140.807
4−5.07±0.6519.68±0.290.967
50 26.14±0.210.989
6−3.55±1.0736.01±0.830.959
70 21.04±0.200.991
Table 2. LOD and LOQ with a linear concentration range of calibration dependences of individual analytes 1–7 (Figure 2) according to Mandel’s test.
Table 2. LOD and LOQ with a linear concentration range of calibration dependences of individual analytes 1–7 (Figure 2) according to Mandel’s test.
AnalyteLOD mmol·L−1Linear Concentration Range
LOQmmol·L−1
10.351.175.00
20.280.925.00
30.612.022.20
40.531.765.00
50.461.535.00
60.311.042.20
70.210.712.20

Share and Cite

MDPI and ACS Style

Řezanka, P.; Macková, D.; Jurok, R.; Himl, M.; Kuchař, M. Enantioseparation and Determination of Mephedrone and Its Metabolites by Capillary Electrophoresis Using Cyclodextrins as Chiral Selectors. Molecules 2020, 25, 2879. https://doi.org/10.3390/molecules25122879

AMA Style

Řezanka P, Macková D, Jurok R, Himl M, Kuchař M. Enantioseparation and Determination of Mephedrone and Its Metabolites by Capillary Electrophoresis Using Cyclodextrins as Chiral Selectors. Molecules. 2020; 25(12):2879. https://doi.org/10.3390/molecules25122879

Chicago/Turabian Style

Řezanka, Pavel, Denisa Macková, Radek Jurok, Michal Himl, and Martin Kuchař. 2020. "Enantioseparation and Determination of Mephedrone and Its Metabolites by Capillary Electrophoresis Using Cyclodextrins as Chiral Selectors" Molecules 25, no. 12: 2879. https://doi.org/10.3390/molecules25122879

Article Metrics

Back to TopTop