Next Article in Journal
Mycobiota of Wheat Seeds with Signs of “Black Point” under Conditions of Forest-Steppe and Forest Zones of Ukraine
Previous Article in Journal
Asymmetric Transfer Hydrogenation of Aryl Heteroaryl Ketones and o-Hydroxyphenyl Ketones Using Noyori-Ikariya Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Investigation of Interactions of ortho- and para-N-Aryl-Substituted 2-Trifluoromethylcinnamanilides †

1
Department of Analytical Chemistry, Faculty of Natural Sciences, Comenius University, Ilkovicova 6, 842 15 Bratislava, Slovakia
2
Department of Chemical Drugs, Faculty of Pharmacy, Masaryk University, Palackeho 1946/1, 612 00 Brno, Czech Republic
3
Regional Centre of Advanced Technologies and Materials, Czech Advanced Technology and Research Institute, Palacky University, Slechtitelu 27, 783 71 Olomouc, Czech Republic
4
Department of Biochemistry, Faculty of Medicine, Masaryk University, Kamenice 5, 625 00 Brno, Czech Republic
5
Institute of Neuroimmunology, Slovak Academy of Sciences, Dubravska Cesta 9, 845 10 Bratislava, Slovakia
*
Author to whom correspondence should be addressed.
Presented at the 25th International Electronic Conference on Synthetic Organic Chemistry, 15–30 November 2021; Available online: https://ecsoc-25.sciforum.net/.
Chem. Proc. 2022, 8(1), 100; https://doi.org/10.3390/ecsoc-25-11651
Published: 13 November 2021

Abstract

:
Unsubstituted (2E)-N-phenyl-3-[2-(trifluoromethyl)phenyl]prop-2-enamide and six other ortho- or para-halogen-substituted anilides of 2-(trifluoromethyl)cinnamic acid were prepared. As the benzene nucleus of cinnamic acid itself is substituted in C(2) position with a trifluoromethyl moiety that is spatially close to both the amide bond and the halogen (F, Cl, CF3) ortho-substitution of the anilide ring, interesting intramolecular interactions can be expected. Other derivatives are substituted at the para-position of the anilide ring, so that intermolecular interactions can be expected. Thus, it can be assumed that predicted properties, especially lipophilicity, will differ significantly from experimentally determined values. All the discussed compounds were analyzed using the reversed-phase high-performance liquid chromatography method. The procedure was performed under isocratic conditions with methanol as an organic modifier in the mobile phase using an end-capped non-polar C18 stationary reversed-phase column. In the present study, the structure–lipophilicity relationships of the studied compounds are discussed.

1. Introduction

Permeability, solubility, and clearance, i.e., lipophilicity-dependent parameters, affect the bioavailability of drugs. More lipophilic drugs pass better through membranes by passive processes; on the other hand, they are less soluble in water, bind more to components of plasma, and are more extensively metabolized (i.e., faster eliminated) or, conversely, are increasingly accumulated in adipose tissues. Thus, lipophilicity is an extremely important physicochemical parameter that crucially affects the absorption, distribution, metabolism, excretion, and toxicity of any biologically active compound. It should be noted that pesticides tend to have higher lipophilicity due to the need to penetrate more lipophilic barriers on the surfaces of plants, fungi, and insects, but in principle, the same laws apply to this category of bioactive agents. Studies show that the optimal range of lipophilicity (expressed as a logarithm of partition coefficient n-octanol-water) log P 0–3 is recommended for optimal gastrointestinal absorption by passive diffusion permeability after oral administration, as there is a good balance between permeability and solubility in this range. As mentioned above, the high lipophilicity of compounds leads to their limited solubility, toxicity, rapid metabolism, and overall inappropriate pharmacokinetic profile; so there is a need to monitor and control the lipophilic properties of drugs [1,2,3,4,5].
Lipophilicity reflects the primary backbone/scaffold of the molecule but is strongly affected by the subsequent substitution of this scaffold with lipophilic/hydrophilic or even ionizable substituents. In addition, substituents affect interactions of the molecule with the environment, i.e., the solvent, other small molecules, and biomolecules (lipid/glycolipid structures, enzymes, and target proteins). Weak intra- and intermolecular interactions of molecules with the environment affect the final shape of the molecule, and thus the ability/ease of binding to receptors/active sites of specific shapes [1,2,3,4,6].
Since lipophilicity can be understood as a physicochemical property of fundamental importance in medicinal chemistry, the lipophilic and hydrophilic properties of newly prepared cinnamic acid derivatives were extensively studied both by prediction using chemical software and liquid chromatography, and it was found that compound retention in the reversed-phase column is affected by their lipophilicity and shows a significant correlation with the n-octanol/water partition coefficient [1,6,7,8].
The studied anilides of 2-(trifluoromethyl)cinnamic acid are substituted by the CF3 group (which is spatially close to the amide bond –CONH–) in position C(2), and, at the same time, the compounds are substituted by groups (F, Cl, CF3) capable of forming weak interactions in the anilide part, either in the ortho (C(2)’) or para (C(4)’) position, so differences between in silico predicted and experimental results are expected.

2. Results and Discussion

Following the previously published ring-substituted arylcinnamanilides/arylcinnamates, which showed a wide range of biological properties [9,10,11,12,13], new derivatives were prepared by microwave synthesis. Briefly, 2-(trifluoromethyl)cinnamic acid dissolved in dry chlorobenzene in the presence of phosphorus trichloride and the appropriate aniline in a microwave reactor provided the desired N-arylcinnamamides 17, see Scheme 1.
The lipophilicities (log P/Clog P data) of all seven compounds were calculated by means of commercially available programs ACD/Percepta ver. 2012 and ChemBioDraw Ultra 13.0. In addition, the lipophilicity of the prepared compounds was studied using reversed-phase high-performance liquid chromatography (RP-HPLC). The procedure is used to measure the retention time under isocratic conditions with methanol as the organic modifier in the mobile phase using end-capped non-polar C18 stationary RP columns and then calculate the logarithm of the capacity factor k [7,8,9,12]. Furthermore, distribution coefficients D at pH 7.4 and 6.5 were determined, and their logarithms were calculated. The distribution coefficient, which takes into account possible ionization, is a more reliable expression of lipophilicity at physiological pH, and log D7.4 values (at pH 7.4) are particularly important because they resemble actual physiological values. Likewise, from the point of view of absorption after oral administration, the partition coefficient at pH 6.5 (log D6.5) is important because this is the pH in the small intestine [1,2,14,15]. All the results are shown in Table 1.
Log P values calculated by the ChemBioDraw software for individual anilide positional isomers are not distinguished; therefore, these values are listed only in Table 1 without further discussion. On the other hand, the predicted log P (ACD/Percepta) and Clog P (ChemBioDraw) values of compounds 17 are distinguished for the individual ortho and para positional isomers.
The graphs of Figure 1 show the agreement of the dependences of the experimentally determined values of lipophilicity (log k, log D7.4, and log D6.5) on log P values. It is evident from the individual graphs that the correlation coefficients R2 (n = 7) are low (range 0.5297–0.5376), indicating significant interactions of the compounds in the aqueous medium and/or with the aqueous medium, which cannot be captured by this prediction program. These observations are completely different from previous experiments with anilides of unsubstituted cinnamic acid [9,12], 3,4-dichlorocinnamic acid [16], 3-(trifluoromethyl)cinnamic acid, and 4-(trifluoromethyl)cinnamic acid [17], where consensus expressed by correlation coefficients was approximately R2 = 0.90, and thus, it was possible to state that the log P values predicted by ACD/Percepta recognized the hydro-lipophilic properties in good agreement with the experimentally determined values [9,12,16,17]. However, in the case of the anilides of 2-(trifluoromethyl)cinnamic acid, this program failed.
Clog P values reflect the presence of intra- and intermolecular interactions much better. Clog P is the logarithm of n-octanol/water partition coefficient based on established chemical interactions. The dependences of the experimentally obtained data (log k, log D7.4, and log D6.5) on the predicted Clog P data are shown in the graphs of Figure 2. The mutual consensus is considerably higher, as expressed by the correlation coefficients in the range 0.9004–0.9038. However, the most significant correlations are shown in the graphs of Figure 3, where the experimental values of log k are compared with log D. There, it is possible to observe correlation coefficients of 0.99.
The order of lipophilicity of individual derivatives 17 is shown in Table 2. It can be seen that unsubstituted compound 1 has the lowest lipophilicity, and the ortho-substituted derivatives 2, 4, and the lipophilicity of compound 6 are lower than that of para-substituted compounds 3, 5, and 7. The unexpected fact that derivative 6 (R = 2-CF3) is less lipophilic than compound 4 (R = 2-Cl) is very interesting, while for para-substituted derivatives 5 and 7, the order is exactly opposite; this order is logical and expected.
Based on all these observed differences between the predicted and experimentally obtained values in comparison with other previously described cinnamic acid derivatives [9,12,16,17], it can be concluded that mainly fluorinated substituents cause significant interactions of the investigated compounds with the aqueous environment. These interactions are not taken into account in ACD/Percepta, and so this software cannot be used to predict physicochemical properties. The interactions affect the observed properties, and it is possible to assume the effect of these interactions on the value of biological activities and structure–lipophilicity relationships, which will be investigated in detail.

3. Experimental

3.1. General

All reagents were purchased from Merck (Sigma-Aldrich, St. Louis, MO, USA) and Alfa (Alfa-Aesar, Ward Hill, MA, USA). Reactions were performed using an Anton-Paar Monowave 50 microwave reactor (Graz, Austria). All 1H- and 13C-NMR spectra were recorded on a JEOL JNM-ECA 600II device (600 MHz for 1H and 150 MHz for 13C, JEOL, Tokyo, Japan) in dimethyl sulfoxide-d6 (DMSO-d6). 1H and 13C chemical shifts (δ) are reported in ppm. High-resolution mass spectra were measured using a high-performance liquid chromatograph Dionex UltiMate® 3000 (Thermo Scientific, West Palm Beach, FL, USA) coupled with an LTQ Orbitrap XLTM Hybrid Ion Trap-Orbitrap Fourier Transform Mass Spectrometer (Thermo Scientific) equipped with an HESI II (heated electrospray ionization) source in the positive mode.

3.2. Synthesis

General procedure for synthesis of target compounds17: 2-(Trifluoromethyl)cinnamic acid (1 mM) was suspended in dry chlorobenzene (6 mL) at ambient temperature, and phosphorus trichloride (0.5 mM, 0.5 eq.) and the corresponding substituted aniline (1 mM, 1 eq.) were added dropwise. The reaction mixture was transferred to the microwave reactor, where the synthesis was performed (50 min, 130 °C). Then, the mixture was cooled to 40 °C, and then the solvent was removed to dryness under reduced pressure. The residue was washed with hydrochloride acid and water. The crude product was recrystallized from ethanol.
(2E)-N-Phenyl-3-[2-(trifluoromethyl)phenyl]prop-2-enamide (1). Yield 64%; 1H-NMR (DMSO-d6) δ: 10.35 (s, 1H), 7.91–7.81 (m, 3H), 7.78 (t, J = 7.5 Hz, 1H), 7.72–7.69 (m, 2H), 7.63 (t, J = 7.8 Hz, 1H), 7.37–7.32 (m, 2H), 7.11–7.07 (m, 1H), 6.91 (d, J = 15.6 Hz, 1H); 13C-NMR (DMSO-d6), δ: 162.66, 138.96, 134.76 (m), 133.24, 133.16, 129.82, 128.86, 127.87, 126.91, 126.91 (q, J = 29.9 Hz), 126.20 (q, J = 4.8 Hz), 124.18 (q, J = 273.6 Hz), 123.67, 119.31; HR-MS: for C16H13ONF3 [M + H]+ calculated 292.0944 m/z, found 292.0937 m/z.
(2E)-N-(2-Fluorophenyl)-3-[2-(trifluoromethyl)phenyl]prop-2-enamide (2). Yield 74%; 1H-NMR (DMSO-d6), δ: 10.11 (s, 1H) 8.14–8.11 (m, 1H), 7.91–7.77 (m, 4H), 7.64 (t, J = 7.3 Hz, 1H), 7.32–7.27 (m, 1H), 7.23–7.14 (m, 3H). 13C-NMR (DMSO-d6), δ: 163.14, 153.32 (d, J = 245.7 Hz), 153.19 (m), 133.17 (m), 129.94, 127.89, 126.97 (q, J = 28.9 Hz), 126.45, 126.23 (q, J = 5.8 Hz), 126.12 (d, J = 10.6 Hz), 125.34 (m), 124.49 (d, J = 3.9 Hz), 124.18 (q, J = 274.6 Hz), 123.61, 115.60, 115.41. HR-MS: for C16H12ONF4 [M + H]+ calculated 310.0850 m/z, found 310.0842 m/z.
(2E)-N-(4-Fluorophenyl)-3-[2-(trifluoromethyl)phenyl]prop-2-enamide (3). Yield 69%; 1H-NMR (DMSO-d6) δ: 10.41 (s, 1H), 7.91–7.76 (m, 4H), 7.74–7.70 (m, 2H), 7.63 (t, J = 7.3 Hz, 1H), 7.21–7.17 (m, 2H), 6.87 (d, J = 15.6 Hz, 1H); 13C-NMR (DMSO-d6), δ: 162.57, 158.22 (d, J = 239.9 Hz), 135.38 (d, J = 2.9 Hz), 134.83, 133.17, 129.85, 127.89, 126.92 (q, J = 28.9 Hz), 126.71, 126.21 (q, J = 5.8 Hz), 124.17 (q, J = 274.6 Hz), 121.07 (d, J = 8.7 Hz), 115.55, 115.39; HR-MS: for C16H12ONF4 [M + H]+ calculated 310.0850 m/z, found 310.0842 m/z.
(2E)-N-(2-Chlorophenyl)-3-[2-(trifluoromethyl)phenyl]prop-2-enamide (4). Yield 70%; 1H-NMR (DMSO-d6), δ: 9.84 (s, 1H), 7.95–7.93 (m, 2H), 7.86 (dd, J = 15.1 Hz, J = 2.1 Hz, 1H), 7.83 (d, J = 7.6 Hz, 1H), 7.79 (t, J = 7.6 Hz, 1H), 7.64 (t, J = 7.6 Hz, 1H), 7.54–7.53 (m, 1H), 7.38–7.35 (m, 1H), 7.23–7.19 (m, 2H); 13C-NMR (DMSO-d6), δ: 163.15, 153.31, 134.71, 133.15, 133.10, 129.96, 129.54, 127.95, 127.50, 126.97 (q, J = 28.9 Hz), 126.36 (m), 126.22 (q, J = 5.8 Hz), 125.77 (m), 125.55, 124.17 (q, J = 274.6 Hz); HR-MS: for C16H12ONClF3 [M + H]+ calculated 326.0554 m/z, found 326.0546 m/z.
(2E)-N-(4-Chlorophenyl)-3-[2-(trifluoromethyl)phenyl]prop-2-enamide (5). Yield 78%; NMR (DMSO-d6) δ: 10.50 (s, 1H), 7.91–7.82 (m, 3H), 7.78 (t, J = 7.6 Hz, 1H), 7.75–7.72 (m, 2H), 7.64 (t, J = 7.6 Hz, 1H), 7.43-7.39 (m, 2H), 6.88 (d, J = 15.1 Hz, 1H); 13C-NMR (DMSO-d6), δ: 162.79, 137.92, 135.06 (m), 133.20, 129.94, 128.80, 128.53, 127.91, 127.26, 126.95 (q, J = 28.9 Hz), 126.58, 126.24 (q, J = 5.8 Hz), 124.17 (q, J = 273.6 Hz), 120.87; HR-MS: for C16H12ONClF3 [M + H]+ calculated 326.0554 m/z, found 326.0545 m/z.
(2E)-N,3-bis [2-(Trifluoromethyl)phenyl]prop-2-enamide (6). Yield 75%; 1H-NMR (DMSO-d6), δ: 9.89 (s, 1H), 7.93 (d, J = 7.8 Hz, 1H), 7.87–7.77 (m, 4H), 7.74–7.62 (m, 3H), 7.48 (t, J = 7.3 Hz, 1H), 7.09 (d, J = 15.6 Hz, 1H); 13C-NMR (DMSO-d6), δ: 163.84, 135.33 (q, J = 1.9 Hz), 135.00 (q, J = 1.9 Hz), 133.17, 133.05, 133.01 (q, J = 1.9 Hz), 129.98, 129.72, 127.94, 126.99 (q, J = 29.9 Hz), 126.81, 126.28 (m), 125.94, 124.29 (q, J = 29.9 Hz), 124.17 (q, J = 273.6 Hz), 123.60 (q, J = 273.6 Hz). HR-MS: for C17H12ONF6 [M + H]+ calculated 360.0818 m/z, found 360.0809 m/z.
(2E)-3-[2-(Trifluoromethyl)phenyl]-N-[4-(trifluoromethyl)phenyl]prop-2-enamide (7). Yield 66%; 1H-NMR (DMSO-d6) δ: 10.72 (s, 1H), 7.92–7.86 (m, 4H), 7.83 (d, J = 7.8 Hz, 1H), 7.79 (t, J = 7.5 Hz, 1H), 7.72 (d, J = 8.7 Hz, 2H), 7.65 (t, J = 7.3 Hz, 1H), 6.91 (d, J = 15.1 Hz, 1H); 13C-NMR (DMSO-d6), δ: 163.21, 142.51, 135.61 (q, J = 1.9 Hz), 133.23, 132.97 (m), 130.06, 127.96, 127.02 (q, J = 29.9 Hz), 126.37, 126.21 (m), 124.35 (q, J = 270.7 Hz), 124.16 (q, J = 274.6 Hz), 123.63 (q, J = 31.8 Hz), 119.29. HR-MS: for C17H12ONF6 [M + H]+ calculated 360.0818 m/z, found 360.0809 m/z.

3.3. Lipophilicity Determination by HPLC

An HPLC separation module Waters Alliance 2695 XE equipped with a Waters Dual Absorbance Detector 2486 (Waters Corp., Milford, MA, USA) was used. A chromatographic column Symmetry® C18 5 μm, 4.6 × 250 mm, Part No. W21751W016 (Waters Corp., Milford, MA, USA) was used. The HPLC separation process was monitored by Empower® 3 Chromatography Manager Software (Waters Corp.). Isocratic elution by a mixture of MeOH p.a. (72%) and H2O-HPLC Mili-Q grade (28%) as a mobile phase was used for the determination of capacity factor k. Isocratic elution by a mixture of MeOH p.a. (72%) and acetate buffered saline (pH 7.4 and pH 6.5) (28%) as a mobile phase was used for the determination of distribution coefficient expressed as D7.4 and D6.5. The total flow of the column was 1.0 mL/min, injection 20 μL, column temperature 40 °C, and sample temperature 10 °C. The detection wavelength of 210 nm was chosen. A KI methanolic solution was used for the determination of dead times (td). Retention times (tr) were measured in minutes. Capacity factors k were calculated according to the formula k = (trtd)/td, where tR is the retention time of the solute, and td is the dead time obtained using an unretained analyte. Distribution coefficients DpH were calculated according to the formula DpH = (trtd)/td. Each experiment was repeated three times. The log k values of individual compounds are shown in Table 1.

3.4. Lipophilicity Calculations

Log P, i.e., the logarithm of the partition coefficient for n-octanol/water, was calculated using the programs ACD/Percepta (Advanced Chemistry Development. Inc., Toronto, ON, Canada, 2012) and ChemBioDraw Ultra 13.0 (CambridgeSoft, PerkinElmer Inc., MA, USA). Clog P values were calculated using ChemBioDraw Ultra 13.0 (CambridgeSoft) software. The results are shown in Table 1.

Author Contributions

Conceptualization, J.K. and J.J.; methodology, D.P., L.V., T.S., J.K. and J.J.; investigation, D.P., L.V. and T.S.; writing, D.P., J.K. and J.J.; funding acquisition, D.P. and J.J. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by a grant project of the Comenius University in Bratislava, Slovakia (UK/228/2021) and by the Slovak Research and Development Agency (APVV-17-0373).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kerns, E.H.; Di, L. Drug-like Properties: Concepts. Structure Design and Methods: From ADME to Toxicity Optimization; Academic Press: San Diego, CA, USA, 2008. [Google Scholar]
  2. Wermuth, C.; Aldous, D.; Raboisson, P.; Rognan, D. The Practice of Medicinal Chemistry, 4th ed.; Academic Press: San Diego, CA, USA, 2015. [Google Scholar]
  3. Savjani, K.T.; Gajjar, A.K.; Savjani, J.K. Drug solubility: Importance and enhancement techniques. Int. Sch. Res. Not. 2012, 2012, 195727. [Google Scholar] [CrossRef] [PubMed]
  4. Arnot, J.A.; Planey, S.L. The influence of lipophilicity in drug discovery and design. Expert Opin. Drug Discov. 2012, 7, 863–875. [Google Scholar] [CrossRef] [PubMed]
  5. Jampilek, J. Potential of agricultural fungicides for antifungal drug discovery. Expert Opin. Drug Dis. 2016, 11, 1–9. [Google Scholar] [CrossRef] [PubMed]
  6. Daze, K.; Hof, F. Molecular interaction and recognition molecular designs and syntheses. In Encyclopedia of Physical Organic Chemistry; Wang, Z., Ed.; John Wiley & Sons: Hoboken, NJ, USA, 2017; Available online: https://onlinelibrary.wiley.com/doi/epdf/10.1002/9781118468586.epoc3001 (accessed on 13 September 2021).
  7. Pliska, V.; Testa, B.; van der Waterbeemd, H. Lipophilicity in Drug Action and Toxicology; Wiley-VCH: Weinheim, Germany, 1996. [Google Scholar]
  8. Rutkowska, E.; Pajak, K.; Jozwiak, K. Lipophilicity–methods of determination and its role in medicinal chemistry. Acta Pol. Pharm. 2013, 70, 3–18. [Google Scholar] [PubMed]
  9. Pospisilova, S.; Kos, J.; Michnova, H.; Kapustikova, I.; Strharsky, T.; Oravec, M.; Moricz, A.M.; Bakonyi, J.; Kauerova, T.; Kollar, P.; et al. Synthesis and spectrum of biological activities of novel N-arylcinnamamides. Int. J. Mol. Sci. 2018, 19, 2318. [Google Scholar] [CrossRef]
  10. Pospisilova, S.; Kos, J.; Michnova, H.; Strharsky, T.; Cizek, A.; Jampilek, J. N-Arylcinnamamides as antistaphylococcal agents. In Proceedings of the 4th International Electronic Conference on Medicinal Chemistry (ECMC-4), Virtual, 1–30 November 2018; p. 5576. Available online: https://sciforum.net/manuscripts/5576/slides.pdf (accessed on 13 September 2021).
  11. Hosek, J.; Kos, J.; Strharsky, T.; Cerna, L.; Starha, P.; Vanco, J.; Travnicek, Z.; Devinsky, F.; Jampilek, J. Investigation of anti-inflammatory potential of n-arylcinnamamide derivatives. Molecules 2019, 24, 4531. [Google Scholar] [CrossRef]
  12. Kos, J.; Bak, A.; Kozik, V.; Jankech, T.; Strharsky, T.; Swietlicka, A.; Michnova, H.; Hosek, J.; Smolinski, A.; Oravec, M.; et al. Biological activities and ADMET-related properties of novel set of cinnamanilides. Molecules 2020, 25, 4121. [Google Scholar] [CrossRef] [PubMed]
  13. Kos, J.; Strharsky, T.; Stepankova, S.; Svrckova, K.; Oravec, M.; Hosek, J.; Imramovsky, A.; Jampilek, J. Trimethoxycinnamates and their cholinesterase inhibitory activity. Appl. Sci. 2021, 11, 4691. [Google Scholar] [CrossRef]
  14. Caron, G.; Vallaro, M.; Ermondi, G. Log P as a tool in intramolecular hydrogen bond considerations. Drug Discov. Today 2018, 27, 65–70. [Google Scholar] [CrossRef] [PubMed]
  15. Andres, A.; Roses, M.; Rafols, C.; Bosch, E.; Espinosa, S.; Segarra, V.; Huerta, J.M. Setup and validation of shake-flask procedures for the determination of partition coefficients (log D) from low drug amounts. Eur. J. Pharm. Sci. 2015, 76, 181–191. [Google Scholar] [CrossRef] [PubMed]
  16. Pindjakova, D.; Strharsky, T.; Kos, J.; Vrablova, L.; Hutta, M.; Jampilek, J. Study of ADMET descriptors of novel chlorinated N-arylcinnamamides. Chem. Proc. 2021, 3, 121. [Google Scholar]
  17. Vrablova, L.; Pindjakova, D.; Strharsky, T.; Kos, J.; Jampilek, J. Evaluation of lipophilicity of selected cinnamic acid derivatives by RP-HPLC. In Proceedings of the Student Scientific Conference of Faculty of Natural Sciences of Comenius University 2021, Faculty of Natural Sciences, Bratislava, Slovakia, 21 April 2021; Comenius University: Bratislava, Slovakia, 2021; pp. 813–818. [Google Scholar]
Scheme 1. Synthesis of ring-substituted (2E)-N-aryl-3-[2-(trifluoromethyl)phenyl]prop-2-enamides 17. Reagents and conditions: (a) PCl3, chlorobenzene, MW, 130 °C, 50 min [9,12].
Scheme 1. Synthesis of ring-substituted (2E)-N-aryl-3-[2-(trifluoromethyl)phenyl]prop-2-enamides 17. Reagents and conditions: (a) PCl3, chlorobenzene, MW, 130 °C, 50 min [9,12].
Chemproc 08 00100 sch001
Figure 1. Comparison of predicted log P (ACD/Percepta) values with experimentally found log k (A), log D7.4 (B), and log D6.5 (C) values of ring-substituted (2E)-N-aryl-3-[2-(trifluoromethyl)- phenyl]prop-2-enamides 17.
Figure 1. Comparison of predicted log P (ACD/Percepta) values with experimentally found log k (A), log D7.4 (B), and log D6.5 (C) values of ring-substituted (2E)-N-aryl-3-[2-(trifluoromethyl)- phenyl]prop-2-enamides 17.
Chemproc 08 00100 g001aChemproc 08 00100 g001b
Figure 2. Comparison of predicted Clog P (ChemBioDraw) values with experimentally found log k (A), log D7.4 (B), and log D6.5 (C) values of ring-substituted (2E)-N-aryl-3-[2-(trifluoromethyl)- phenyl]prop-2-enamides 17.
Figure 2. Comparison of predicted Clog P (ChemBioDraw) values with experimentally found log k (A), log D7.4 (B), and log D6.5 (C) values of ring-substituted (2E)-N-aryl-3-[2-(trifluoromethyl)- phenyl]prop-2-enamides 17.
Chemproc 08 00100 g002
Figure 3. Comparison of experimentally found log k values with log D7.4 (A) and log D6.5 (B) values and log D7.4 with log D6.5 (C) of discussed compounds 17.
Figure 3. Comparison of experimentally found log k values with log D7.4 (A) and log D6.5 (B) values and log D7.4 with log D6.5 (C) of discussed compounds 17.
Chemproc 08 00100 g003
Table 1. Structure of ring-substituted (2E)-N-aryl-3-[2-(trifluoromethyl)phenyl]prop-2-enamides 17, calculated lipophilicities (log P/Clog P), and experimentally determined log k, log D7.4, and log D6.5 values of investigated compounds.
Table 1. Structure of ring-substituted (2E)-N-aryl-3-[2-(trifluoromethyl)phenyl]prop-2-enamides 17, calculated lipophilicities (log P/Clog P), and experimentally determined log k, log D7.4, and log D6.5 values of investigated compounds.
Chemproc 08 00100 i001
Comp.Rlog P alog P/Clog P blog klog D7.4log D6.5
1H3.964.10/4.54700.38970.34700.3457
22-F3.874.26/4.34760.40010.36070.3570
34-F3.794.26/4.94760.44250.40550.4009
42-Cl4.604.66/4.66760.51000.47690.4708
54-Cl4.704.66/5.51760.66510.63040.6250
62-CF34.465.02/4.43080.42470.38740.3814
74-CF34.645.02/5.88080.79480.76030.7532
a ACD/Percepta ver. 2012, b ChemBioDraw Ultra 13.0.
Table 2. Discussed compounds ordered according to increasing lipophilicity values of individual derivatives.
Table 2. Discussed compounds ordered according to increasing lipophilicity values of individual derivatives.
Log P4-F< 2-F< H< 2-CF3< 2-Cl< 4-CF3< 4-Cl
Clog P2-F< 2-CF3< H< 2-Cl< 4-F< 4-Cl< 4-CF3
Log kH< 2-F< 2-CF3< 4-F< 2-Cl< 4-Cl< 4-CF3
Log D7.4H< 2-F< 2-CF3< 4-F< 2-Cl< 4-Cl< 4-CF3
Log D6.5H< 2-F< 2-CF3< 4-F< 2-Cl< 4-Cl< 4-CF3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pindjakova, D.; Vrablova, L.; Strharsky, T.; Kos, J.; Jampilek, J. Investigation of Interactions of ortho- and para-N-Aryl-Substituted 2-Trifluoromethylcinnamanilides. Chem. Proc. 2022, 8, 100. https://doi.org/10.3390/ecsoc-25-11651

AMA Style

Pindjakova D, Vrablova L, Strharsky T, Kos J, Jampilek J. Investigation of Interactions of ortho- and para-N-Aryl-Substituted 2-Trifluoromethylcinnamanilides. Chemistry Proceedings. 2022; 8(1):100. https://doi.org/10.3390/ecsoc-25-11651

Chicago/Turabian Style

Pindjakova, Dominika, Lucia Vrablova, Tomas Strharsky, Jiri Kos, and Josef Jampilek. 2022. "Investigation of Interactions of ortho- and para-N-Aryl-Substituted 2-Trifluoromethylcinnamanilides" Chemistry Proceedings 8, no. 1: 100. https://doi.org/10.3390/ecsoc-25-11651

Article Metrics

Back to TopTop